Home Journals Progress in Chemistry
Progress in Chemistry

Abbreviation (ISO4): Prog Chem      Editor in chief: Jincai ZHAO

About  /  Aim & scope  /  Editorial board  /  Indexed  /  Contact  / 
Review

Synthesis and Ring-Opening Metathesis Polymerization of Bio-Based Cyclic Olefins

  • Guangyu Pan 1, 3 ,
  • Xin Hu , 2, 3, * ,
  • Jie Yin 4 ,
  • Yihuan Liu 1, 3 ,
  • Kai Guo 1, 3 ,
  • Ning Zhu , 1, 3, *
Expand
  • 1 College of Biotechnology and Pharmaceutical Engineering, Nanjing Tech University, Nanjing 211800, China
  • 2 College of Materials Science and Engineering, Nanjing Tech University, Nanjing 211800, China
  • 3 State Key Laboratory of Materials-Oriented Chemical Engineering, Nanjing 211800, China
  • 4 SINOPEC (Beijing) Research Institute of Chemical Industry Co., Ltd.,, Beijing 100013, China
* e-mail: (Xin Hu);

Received date: 2024-03-22

  Revised date: 2024-05-26

  Online published: 2024-06-30

Supported by

National Key R&D Program of China(2019YFA0905000)

National Natural Science Foundation of China(22278223)

National Natural Science Foundation of China(22278205)

R&D Program of China Petrochemical Corporation (SINOPEC)(30000000-23-ZC0607-0871)

Abstract

The transformations of biomass into bio-based polymeric materials have attracted growing interest from chemistry and material engineering. Ring-opening metathesis polymerizations (ROMP) of cyclic olefins have been identified as the powerful toolbox for synthesis of polyolefins containing double bonds in the polymer mainchains. Recently, a series of novel cyclic olefins are designed by using biomass as the feedstock, and high-performance polyolefins are prepared via ROMP of biomass derived monomers. This review summaries the advances in conversions of cellulose, hemicellulose, lignin, terpenes, vegetable oils, amino acids into norbornene derivatives, oxanorbornene derivatives, cyclooctene derivatives, macrocyclic olefins, etc. Synthesis and properties of bio-based polyolefins via ROMP of biomass derived monomers mentioned above are highlighted. Moreover, the challenges and opportunities are discussed with the aim to promote the development of bio-based polymeric materials.

Contents

1 Introduction

2 Cellulose-based cyclic olefins and ROMP

3 Hemicellulose-based cyclic olefins and ROMP

4 Lignin-based cyclic olefins and ROMP

5 Terpenes-based cyclic olefins and ROMP

6 Vegetable oils-based cyclic olefins and ROMP

7 Amino acids-based cyclic olefins and ROMP

8 Conclusion and outlook

Cite this article

Guangyu Pan , Xin Hu , Jie Yin , Yihuan Liu , Kai Guo , Ning Zhu . Synthesis and Ring-Opening Metathesis Polymerization of Bio-Based Cyclic Olefins[J]. Progress in Chemistry, 2024 , 36(12) : 1956 -1971 . DOI: 10.7536/PC240323

1 Introduction

Polymeric materials, with their rich variety and excellent properties, as well as their large volume and wide application, have made significant contributions to daily life, scientific and technological progress, and economic development. The various monomers for manufacturing polymeric materials mainly come from non-renewable resources such as petroleum. It is projected that by 2050, approximately 20% of fossil raw materials will be used to meet the production demands of polymeric materials[1]. As a renewable resource, biomass has abundant reserves and great potential for utilization. The conversion of biomass into biobased polymeric materials provides a technical solution for addressing resource and environmental issues and achieving the dual carbon strategy, and it has become a research hotspot in fields such as chemistry and materials[2-3].
Lignocellulose is the most abundant biomass resource in nature, with an annual global production of about 18 billion tons, mainly composed of cellulose (30~50 wt%), hemicellulose (20~35 wt%), and lignin (15~30 wt%). It is considered to be the biomass with the greatest industrial application value. Many scholars have summarized the routes for the conversion of lignocellulose into platform compounds[4-12]. Meanwhile, significant progress has also been made in the utilization of terpenes, plant oils, and amino acids as biomass resources[13-19]. The aforementioned biomasses can be converted into monomers such as polyols, polycarboxylic acids, lactones, lactams, epoxides, acrylates, and olefins, which, through stepwise polymerization[20-21], ring-opening polymerization[22-27], radical polymerization[28-30], ionic polymerization[31-32], coordination polymerization[33-35], and ring-opening metathesis polymerization[36-37], produce a series of biobased polymers with different structures and properties, attracting joint attention from academia and industry.
Ring-opening metathesis polymerization (ROMP) is an effective strategy for constructing polyolefins with unsaturated double bonds in the backbone, and commercially available ruthenium[38]and molybdenum[39]metal complexes are widely used ROMP catalysts (see Figure 1a). As shown in Figure 1b, the mechanism involves the formation of a metallacyclobutane intermediate by the metal carbene active center and the carbon-carbon double bond in the cyclic olefin. This intermediate undergoes cleavage through metathesis to form a new metal carbene and a carbon-carbon double bond; this process repeats to form polyolefins with unsaturated double bonds in the backbone[40-41]. Ring-opening metathesis polymerization exhibits living/controlled characteristics, mild reaction conditions, and good functional group tolerance, making it applicable for the synthesis of functionalized polyolefins[42-44], polymers with complex topological structures[45-48], and self-assembled nanomaterials[49-54].
图1 (a) ROMP催化剂结构[49];(b) ROMP反应机理[40]

Fig. 1 (a) Structures of catalysts for ROMP[49] (Copyright © 2020 Elsevier); (b) Mechanism of ROMP[40] (Copyright © 2006 Elsevier)

In recent years, various bio-based cyclic olefin monomers with different structures have been designed and synthesized from biomass or bio-based platform compounds through biological or chemical routes. Ring-opening metathesis polymerization has been used to obtain a series of new bio-based polyolefin materials. This paper categorizes the types of biomass and introduces the synthesis routes for converting cellulose, hemicellulose, lignin, terpenes, plant oils, and amino acids into bio-based cyclic olefin monomers such as norbornene derivatives, oxanorbornene derivatives, cyclooctene derivatives, and macrocyclic olefins (Figure 2). It summarizes the research progress in the preparation, properties, and applications of functional polyolefin materials via the ring-opening metathesis polymerization of cyclic olefin monomers, and discusses and prospects the future and challenges in this field.
图2 生物质转化为环烯烃及其聚合物路线图

Fig. 2 Synthetic routes to cyclic olefins and polyolefins from biomass.

2 Cellulose Route

Cellulose is the most abundant natural polymer, and glucose is the final product of cellulose hydrolysis. Using glucose as the starting material, it is possible to prepare a variety of polymers such as polylactic acid (PLA), poly(β-hydroxybutyrate) (PHB), poly(ethylene 2,5-furandicarboxylate) (PEF), poly(butylene succinate) (PBS), and poly(ethylene terephthalate) (PET)[12]. However, efficient hydrolysis of cellulose (high conversion rate, high conversion speed, high selectivity) remains challenging. Pyrolysis routes, due to their advantages of not requiring acids or bases, being environmentally friendly, and having high energy efficiency, have gained favor among researchers. Levoglucosenone is one of the main products of the pyrolysis route[55]. This section will introduce the synthesis of various biobased cycloolefin monomers from the cellulose route and their ring-opening metathesis polymerization (ROMP) (Figure 3).
图3 纤维素转化制备环烯烃单体

Fig. 3 Conversion of cellulose to cyclic olefin monomers

Sugar substituents typically enhance the hydrophilicity and biocompatibility of polymers, and the sugar units also confer biomimetic recognition functions to the polymers, making them applicable in drug delivery, antibody recognition, bacterial adhesion, bioimaging, and hydrogels[22]. Grubbs et al.[56] used glucosamine as a glycosyl donor and reacted it with 5-norbornene-2-carbonyl chloride through an amidation reaction. They modified the hydroxyl groups of glucose using different protective groups (Figure 4a), synthesizing a series of sugar-substituted norbornene derivative monomers. The polymerization rate and molecular weight distribution of these monomers were highly dependent on the type of protective group. Chaikof et al.[57] started with α-D-glucosamine hydrochloride as the initial material and underwent multiple reactions (Figure 4b), synthesizing a disaccharide-functionalized norbornene derivative with a total yield of 10%. Emulsion polymerization was carried out under the action of the second-generation Grubbs catalyst and dodecyltrimethylammonium bromide (DTAB) (60 ℃, 5 h), producing a disaccharide polymer with a low molecular weight distribution index (Mn=9.0 kg/mol, Mw/Mn=1.17). Nomura et al.[58] began with nitromethane-substituted glucose derivatives, generating in situ glucosylnitrile oxide under the action of toluene diisocyanate and triethylamine, which then underwent [3+2] cycloaddition with norbornadiene (NBD) (Figure 4c), synthesizing isoxazoline-substituted norbornenes containing glucose residues with a yield of 59%. Blanchfield et al.[59] synthesized a series of sugar-functionalized norbornene derivatives via click chemistry between azide-functionalized norbornenes and alkyne-functionalized glucose (Figure 4d), preparing polymers with side-chain structures similar to the bacterial cell membrane component poly-N-acetylglucosamine (PNAG) under the action of the third-generation Grubbs catalyst. Liu et al.[60] also adopted azide-alkyne click chemistry (Figure 4e) to prepare a disubstituted glycosyl norbornene derivative monomer. Under the catalysis of the third-generation Grubbs catalyst at 50 ℃ for 12 h, they synthesized disubstituted glycosyl polymers (Mn = 7.2~7.3 kg/mol, Mw/Mn = 1.04~1.08), whose ability to bind concanavalin A exceeded that of monosubstituted glycosyl polymers by five times. With the development of catalysts, ROMP can be used to prepare well-defined, low-polydispersity sugar-containing polymers, which show great potential in biomedical materials. However, the synthesis of sugar-functionalized norbornene derivative monomers involves many steps, and the separation is challenging, necessitating improvements in synthetic efficiency.
图4 葡萄糖取代降冰片烯的合成[56-60]

Fig. 4 Synthesis of glucose-substituted norbornenes[56-60]

Angelica lactone (AL) is a downstream product of levulinic acid, and the industrial production routes for levulinic acid mainly include the cellulose route (cellulose → hydroxymethylfurfural → levulinic acid) and the hemicellulose route (hemicellulose → furfural → levulinic acid), with the cellulose route being more common[61]. de Vries et al.[62] obtained a β/α-AL mixture (the molar percentage of β-AL confirmed to be 90% by vacuum distillation) through the isomerization reaction of α-AL, and then used the β/α-AL mixture as a substrate to synthesize the monomer Cp/β-AL (yield 82%) via a Diels-Alder reaction with cyclopentadiene. In a system using ethyl acetate as a solvent, this monomer can produce a polymer with a molecular weight close to the theoretical value (Mn=64.1 kg/mol, Mw/Mn=1.85). Compared to polynorbornene, poly(Cp/β-AL) exhibits higher transparency (see Figure 5b) and hydrophilicity (Cp/β-AL contact angle θ=75.7±1.9°, lower than that of polynorbornene θ=83.9±2.3°) due to the presence of the lactone functional group.
图5 (a) Cp/β-AL的合成与ROMP;(b) Cp/β-AL和降冰片烯的聚合物薄膜[62]

Fig. 5 (a) Synthesis and ROMP of Cp/β-AL; (b) Films obtained from poly-Cp/β-AL and polynorbornene[62] (Copyright © The Royal Society of Chemistry 2020)

Microbial fermentation can not only convert biomass into platform compounds but also prepare compounds that are difficult to synthesize through traditional chemical methods. Sophorolipids are extracellular glycolipids produced by yeast during the fermentation of glucose and fatty acids. Gross et al.[63] developed a fermentation process where, through fed-batch feeding, the yield of sophorolipids in each liter of fermentation broth could reach up to 350 g within 7 days (with 100 g of glucose and 400 g of oleic acid as the consumed biomass raw materials). The sophorolipids produced by fermentation are typically a mixture of lactonic and acidic forms; the lactonic sophorolipids were separated and purified by column chromatography for ROMP studies[64]. Unlike monomers with high ring strain, sophorolipids undergo entropy-driven ring-opening metathesis polymerization (ED-ROMP), and various ruthenium-based catalysts showed excellent activity, allowing the preparation of poly(sophorolipid) with Mn=42.2~103 kg/mol and Mw/Mn=1.8~2.2 at 25 ℃ in 5 minutes. Due to the retention of complex structures such as hydroxyl groups, sugar rings, linear fatty chains, esters, and double bonds in the backbone, poly(sophorolipid) exhibits multiple properties including hydrophilicity, hydrophobicity, and degradability (Figure 6).
图6 槐糖脂开环易位聚合[64]

Fig. 6 ROMP of sophorolipid to poly(sophorolipid)[64]

Gross et al.[65] conducted a study on the ROMP reaction kinetics of sophorolipids, performed thermal property tests and crystal structure characterization[66]. The effects of temperature, solvent, type and loading of catalyst on the ring-opening metathesis polymerization (ROMP) of sophorolipids were investigated, providing important references for the ED-ROMP of macrocyclic monomers. Thermal properties and crystal structures showed that poly(sophorolipid) is a semi-crystalline polymer, undergoing a glass transition at 61 ℃, melting at 123 ℃, with an initial thermal decomposition temperature exceeding 200 ℃. The crystalline phase of poly(sophorolipid) is associated with the ordered stacking of aliphatic segments, exhibiting long-range order involving glycan groups (d=2.44 nm). This long-range order persists even after the crystalline phase melts, though the distance is slightly reduced (d=2.27 nm). After annealing at 80 ℃, poly(sophorolipid) can recrystallize.
Erythritol was discovered in yeast-fermented blackstrap molasses in 1950, and Japanese companies began commercial production in the 1990s. Erythritol is a potential source of 1,4-butenediol[67-68], and Buchard et al.[69] synthesized an unsaturated seven-membered cyclic carbonate (Figure 7) using commercial 1,4-butenediol and CO2 as raw materials. The product was purified by column chromatography and recrystallization, with a monomer yield of 51%. This monomer showed good reactivity in both ring-opening metathesis polymerization (ROMP) and ring-opening polymerization (ROP). Polymerization via the ROMP mechanism produced amorphous trans-polymers, while ROP led to the formation of semi-crystalline cis-polymers. A cascade reaction of ROP and ROMP could be achieved by adding the chain transfer agent 1,4-butenediol, leading to the preparation of cis-trans-cis triblock copolymers (Mn=4.6~8.9 kg/mol, Mw/Mn=1.67~2.02).
图7 1,4-丁烯二醇基顺式-反式-顺式三嵌段共聚物的合成[69]

Fig. 7 Synthesis of poly (cis-trans-cis) triblock copolymers by using 1,4-butene glycol as the feedstock[69]

1,4-Butanediol (1,4-BDO) is an important industrial chemical, primarily produced from fossil raw materials. There have been numerous reports on the synthesis of bio-based 1,4-BDO, and the highest yield of 1,4-BDO produced by commercial strains is approximately 140 g/L[70]. 2,3-Dihydrofuran (DHF) is one of the downstream products of 1,4-BDO. Leite et al.[71] developed a route for the cyclization of 1,4-BDO into DHF with a maximum yield of up to 81%. Due to the formation of a stable Fischer carbene structure with Grubbs catalysts, DHF has long been widely used as a quencher in ROMP reactions. Xia et al.[72] conducted bulk polymerization using commercially available DHF as a monomer under the action of 1st or 2nd generation Grubbs catalysts. The reaction reached equilibrium after 4 h at room temperature, and poly(2,3-dihydrofuran) (PDHF) (Mn=6.0~127.7 kg/mol, Mw/Mn=1.35~2.66, Tg =-50 ℃) was obtained for the first time (Figure 8a), which required the use of hydrogen peroxide for quenching, with a monomer conversion rate of about 80%. PDHF exhibits high regioregularity, and its molecular weight can be adjusted by changing the catalyst loading or adding a chain transfer agent such as vinyl ether, but there is significant secondary metathesis during the polymerization process. Through thermodynamic studies, the thermodynamic parameters of DHF were calculated as ΔH(neat)=-5.0 kcal/mol, ΔS(neat)=-19.4 cal/(mol·K), similar to those of cyclopentene. Interestingly, the 2nd generation Grubbs catalyst can catalyze the depolymerization of PDHF at 60 ℃, with a monomer recovery rate exceeding 90% within 2.5 h, or hydrolysis into a mixture of 4-hydroxybutanal and 2-hydroxytetrahydrofuran within 10 min under acidic conditions. The study of PDHF provides a new solution to the problem of the difficulty in degrading traditional ROMP polymers, while also offering a new family of monomers (cyclovinyl ethers) for ROMP. In recent years, research on the preparation of degradable polymers by incorporating DHF has gradually increased, with various functional monomers or macromolecular chain transfer agents terminated with vinyl ether (including polystyrene, polycaprolactone, polylactic acid, polyethylene glycol, etc.) being introduced into PDHF[73-79], and the topological structure of the polymers has evolved from linear structures to more complex star-shaped and cross-linked network structures[80-81].
图8 (a) DHF开环易位聚合;(b) PDHF在酸性条件下水解GPC图;(c) PDHF解聚为DHF[72]

Fig. 8 (a) ROMP of DHF; (b) GPC of PDHF hydrolysis in THF with 135 mM water and 25 mM HCl; (c) Facile depolymerization of PDHF to DHF[72] (Copyright © 2019 American Chemical Society)

Cellulose pyrolysis can produce levoglucosenone (LGO), a chiral cyclic molecule containing a double bond, which is often used for the synthesis of chiral compounds[82-83]. Additionally, LGO is also applied in monomer synthesis and polymerization reactions[84]; however, possibly due to interference from the ketone group near the alkene, LGO is difficult to directly undergo ring-opening metathesis polymerization. Schlaad et al.[85] heated 120 g of acid-pretreated cellulose at 0.1 mbar and 300 ℃ for 20 minutes, then reduced the temperature to 160 ℃ for distillation, obtaining 5 g of pure LGO. The LGO was then subjected to a reduction reaction with NaBH4, and the crude product was purified by sublimation under 0.001 mbar and 40 ℃ conditions, ultimately synthesizing ROMP-reactive levoglucosenol (4.5 g) (Route 1, Figure 9). Despite its bicyclic strained structure, the polymerization of levoglucosenol still proceeds relatively slowly. The ROMP product, thermoplastic polyacetal (Mw=2.3~104 kg/mol, Mw/Mn=1.6~2.3, Tg=100 ℃, Td5%=220 ℃), exhibits good degradability, being completely degraded within 40 days at room temperature in a 1,4-dioxane solution containing p-toluenesulfonic acid/water.
图9 将LGO转化为ROMP单体及其聚合物[85-86]

Fig. 9 Conversion of LGO to cyclic monomers and polymers[85-86]

Allais et al.[86] used commercial LGO as the raw material, which underwent a Diels-Alder reaction with cyclopentadiene at room temperature to produce the norbornene derivative N-LGO (yield 64%) (Figure 9, Route 2). N-LGO was then converted via Baeyer-Villiger oxidation into the monomer N-HBO containing lactone and hydroxyl groups (yield 59%). The absence of other substituents near the double bonds in N-LGO and N-HBO increased their reactivity in ring-opening metathesis polymerization, leading to the formation of functionalized copolymers (Mn=25.9~35.9 kg/mol, Mw/Mn=1.39~1.42, Td5%=310~331 ℃).

3 Hemicellulose Route

Unlike cellulose, hemicellulose consists of a variety of building blocks, including not only glucose but also various pentoses (such as xylose and arabinose) and hexoses (such as mannose and galactose)[8]. These monosaccharides can also serve as functional groups for synthesizing norbornene derivative monomers, with synthesis routes similar to those for glucose[58,60,87]. Among them, furfural derived from xylose is one of the important bio-based chemicals, with an annual global production exceeding 600,000 tons[9]. Furfural and furan-based compounds are often used as diene substrates in Diels-Alder reactions, undergoing DA cycloaddition reactions with dienophiles such as maleic anhydride, maleimide, itaconic anhydride, acrylates, and vinyl compounds, to synthesize oxanorbornene derivative monomers. Ananikov et al.[88] provided a detailed review on this, which will not be reiterated here. This section mainly introduces the synthesis and ring-opening metathesis polymerization (ROMP) of oxanorbornene derivative monomers starting from furfuryl alcohol or furfurylamine (Figure 10).
图10 半纤维素转化制备环烯烃单体

Fig. 10 Conversion of hemicellulose to cyclic olefin monomers

North et al.[89-90] prepared a series of oxanorbornene derivatives with different structures through the Diels-Alder cycloaddition of furfuryl alcohol and itaconic anhydride (Figure 11a). When the substituent was a carboxyl group, ROMP could not be carried out; when the substituent was an ester group, ROMP could proceed, but the reaction time was longer, approximately 72 h. The solubility of the polymer products is closely related to the structure of the ester group, and esters containing more than five carbon atoms have good solubility in common organic solvents. North et al.[91] based on this, attempted to acylate the carboxyl groups (Figure 11b), successfully introducing various amide substituents, including amino acids. Using the second-generation Grubbs catalyst, there was a good linear relationship between the molecular weight of the ROMP product and the feed ratio, and the molecular weight distribution was narrow (Mw/Mn·<1.2).
图11 呋喃基环烯烃的合成与ROMP: (a) endo-酯功能化内酯型三环氧杂降冰片烯[89-90]; (b) endo-酰胺功能化内酯型三环氧杂降冰片烯[91]; (c) 酯功能化内酯型三环氧杂降冰片烯单体[92]; (d) 酯功能化内酰胺型三环氧杂降冰片烯[93-94]

Fig. 11 Synthesis and ROMP of furan-based cyclic olefins (a) endo-tricyclic oxa-norbornene lactone esters[89-90]; (b) endo-tricyclic oxa-norbornene lactone amides[91]; (c) Tricyclic oxa-norbornene lactone esters[92]; (d) Tricyclic oxa-norbornene lactam esters[93-94]

Naguib et al.[92] used maleic anhydride instead of itaconic anhydride to synthesize structurally similar oxanorbornene derivative monomers (Figure 11c), with a total yield from the raw material furfuryl alcohol to the polymerizable monomer of 38% to 46%. Under the action of Grubbs 3rd generation catalyst, narrowly distributed (Mw/Mn <1.2) homopolymers (Mn=21.5~248.8 kg/mol) and copolymers (Mn=99.1~100.2 kg/mol) were synthesized within minutes. North et al.[93] synthesized new oxanorbornene derivative monomers using furfurylamine and maleic anhydride as raw materials, with a total yield of 27% to 61%. Through ROMP, polymers with controllable molecular weight and narrow molecular weight distribution (Mn=4~260 kg/mol, Mw/Mn <1.2, Tg=160~168 ℃, Td10%=340~383 ℃) could be obtained. Naguib et al.[94] also introduced vanillin as an antioxidant component into the monomers, testing their radical scavenging ability and that of their polymers in a solution of 2,2-diphenyl-1-picrylhydrazyl at a fixed concentration, which eliminated 80% and 51% of the radicals after 20 min, respectively, successfully imparting antioxidant properties to the material.

4 Lignin Route

Lignin is composed of phenylpropane units connected through carbon-carbon bonds and ether bonds, featuring a complex chemical structure and diverse types of functional groups[6]. For example, the abundance of aromatic rings gives lignin hydrophobicity, rigidity, and thermal stability, while functional groups such as hydroxyl, double bonds, and methoxy bring good reactivity[11]. This section will introduce the synthesis of lignin-based norbornene or cyclooctene derivatives monomers via Diels-Alder reactions and esterification reactions, and their ring-opening metathesis polymerization (ROMP) (Figure 12).
图12 木质素转化制备环烯烃单体

Fig. 12 Conversion of lignin to cyclic olefin monomers

Fang et al.[95]used eugenol as the raw material and synthesized a fluorinated eugenol-modified norbornene derivative (total yield 33%) through a three-step reaction. Under the action of the 2nd generation Grubbs catalyst, they prepared a fluoropolymer PNBE-TFVE with M n=78 kg/mol and M w/M n=1.81 (Figure 13a). This polymer can further form a crosslinked polymer PNBE-PFCB through [2+2] cycloaddition reactions of trifluoroethenyl ether groups at high temperatures (>150 ℃) (Figure 13b). The coefficient of thermal expansion (CTE) of PNBE-PFCB in the temperature range of 50~300 ℃ is 90.9 ppm/℃, and the storage modulus at 25 ℃ is as high as 1.56 GPa, with an average bonding strength on silicon wafers of about 1.42 GPa. PNBE-PFCB films exhibit good transmittance (>90%) in the wavelength range of 550~1100 nm, and the surface roughness over a 2 μm × 2 μm area is 2.35 nm, indicating that the film has excellent uniformity and flatness. The average dielectric constant, average loss factor, and average surface energy are 2.65, 4.3×10-3, and 26.8 mJ/m2, respectively, all of which are lower than those of PNBE-TFVE, indicating that the dielectric properties and hydrophobicity are enhanced after crosslinking, showing potential application prospects in microelectronics and energy storage fields.
图13 (a) PNBE-TFVE的合成;(b) PNBE-PFCB的合成[95]

Fig. 13 (a) Synthesis of PNBE-TFVE; (b) Synthesis of PNBE- PFCB[95] (Copyright © 2019, American Chemical Society)

Sha et al.[96] used guaiacol as the starting material to synthesize cyclooctene derivative monomers (Figure 14a), with a total yield of 56%. Under the action of Grubbs 2nd generation catalyst, a polymer with guaiacol in the side chain was prepared (Mn=60 kg/mol, Mw/Mn=1.55, Tg=41 ℃, Td5%=344 ℃). According to DFT calculations, after introducing lignin components, the ring strain of the monomer decreased from 8.2 kcal/mol to 5.1 kcal/mol. The lower ring strain prompted entropy increase to become the main driving force for polymerization, leading to a broader molecular weight distribution index. Moreover, the polymer underwent depolymerization within minutes at 50 ℃, with a monomer recovery rate exceeding 90%.
图14 (a) 愈创木酚改性环辛烯单体的合成与ROMP[96];(b) 香兰素基均聚物与ABA型共聚物的合成[97]

Fig. 14 (a) Synthesis and ROMP of guaiacol modified cyclooctene[96]; (b) Synthesis of vanillin-based homopolymer and ABA copolymer[97]

Vanillin is a lignin derivative, and through the esterification reaction of phenolic hydroxyl with 5-norbornene-2-carboxylic acid, Koo et al.[97] converted vanillin into ROMP monomer VN (Figure 14b, yield 90%). The polymer product PVN has good blocking properties for ultraviolet light in the range of 280~400 nm and can be applied to UV-resistant film materials. Further, by copolymerizing VN and n-butyl ester-substituted norbornene (BuN), thermoplastic elastomers P1 (Mn = 32 kg/mol, Mw/Mn = 1.35, Td5% = 323 ℃) and P4 (Mn = 33.1 kg/mol, Mw/Mn = 1.40, Td5% = 259 ℃) were prepared, with hard segment (PVN) contents of 25.2% and 42.1%, respectively. The Young's moduli of P1 and P4 are 28 MPa and 285 MPa, respectively, and the elongation at break are 191% and 59%, respectively, where the Young's modulus of P1 is 7.8 times that of the commercial thermoplastic elastomer SIS.

5 Terpene Pathway

Terpenes, terpenoids, and rosin are the main components of essential oils and natural resins. Terpenes are a general term for a class of olefins composed of isoprene units; terpenoids are modified terpenes containing oxygen functional groups such as hydroxyl and aldehyde groups within their molecules; the main component of rosin is terpene rosin acid with a hydrogenated phenanthrene structure and its derivatives[15]. Over 20,000 types of terpenoids are known in plants[98], but there are few polymers derived from terpenes. Apart from isoprene (the monomer of natural rubber), only high-yield terpenes such as pinene, limonene, and myrcene have been used to synthesize polymers[14,99]. This section will introduce the synthesis and ring-opening metathesis polymerization (ROMP) of terpenes, modified terpenes, and terpene norbornene derivative monomers (Figure 15).
图15 萜转化制备环烯烃单体

Fig. 15 Conversion of terpene to cyclic olefin monomers

Caryophyllene and humulene are terpenes that can directly undergo ROMP, among which polycaryophyllene has attracted more attention due to its unique cyclobutane backbone structure. Mecking and Grau et al.[100] prepared polycaryophyllene for the production of coatings and films using 80% pure caryophyllene catalyzed by the third-generation Grubbs catalyst at room temperature in bulk for 7 days (with Mn=20 kg/mol, Mw/Mn=1.8~1.9, and Tg= -32 ℃). Subsequently, Grau et al.[101] studied the crosslinking process of polycaryophyllene (Figure 16), including thermal crosslinking under organic peroxide or sulfur systems and UV-initiated thiol-ene coupling. By altering the crosslinking method and type of crosslinker, crosslinked polycaryophyllenes with different degrees of crosslinking (0.3~11) and storage moduli (1~100 MPa) were obtained, with the maximum elongation at break reaching up to 360%.
图16 聚石竹烯的合成与交联[101]

Fig. 16 Synthesis and cross-linking of polycaryophyllene[101] (Copyright © 2020, American Chemical Society)

α-Pinene is the most common monoterpene, featuring an unsaturated six-membered ring structure. However, due to the highly congested tri-substituted groups within the molecule, it severely hinders chain initiation and chain growth, making direct ring-opening metathesis polymerization of α-pinene difficult[102]. Thomson et al.[103] started from the commercially available α-pinene derivative myrtenal, heating it with a Pd/BaSO4 mixture in a Dean Stark apparatus to 185 ℃, synthesizing the α-pinene derivative Apopinene (yield 80%, Figure 17a). After reducing the intramolecular substituents, it showed good polymerization activity, allowing for the preparation of polymers with Mn=15.6 kg/mol and Mw/Mn=1.6 after reacting for 1 h at room temperature.
图17 (a) Apopinene的合成与ROMP[103];(b) δ-蒎烯的合成与ROMP[104]

Fig. 17 (a) Synthesis and ROMP of apopinene[103]; (b) synthesis and ROMP of δ-pinene[104]

Kennemur et al.[104] provided another approach (Figure 17b), transforming α-pinene into a δ/α-pinene mixture through a three-step reaction, then selectively modifying α-pinene using chlorosulfonyl isocyanate to convert it into less volatile azetidinone derivatives, and finally obtaining pure δ-pinene via vacuum distillation with an overall yield of 29% (relative to the starting material α-pinene). The isomerization of α-pinene not only reduces the steric hindrance around the double bond but also increases the ring strain from 24.3 kJ/mol to 35 kJ/mol, making it more susceptible to undergo ring-opening metathesis polymerization. Compared to Apopinene, δ-pinene exhibits high regioselectivity, and poly(δ-pinene) has a larger number-average molecular weight and a narrower molecular weight distribution (<italic>M</italic><sub>n</sub> = 9.5~70 kg/mol, <italic>M</italic><sub>w</sub>/<italic>M</italic><sub>n</sub> < 1.2, <italic>T</italic><sub>g</sub> = 100 ℃, <italic>T</italic><sub>d5%</sub> = 330 ℃).
du Prez et al.[105] attempted to develop terpene-based biobased pressure-sensitive adhesives (PSAs) by esterifying tetrahydrolinalool, menthol, and isoborneol with acrylic acid. The resulting acrylates were then subjected to Diels-Alder cycloaddition with cyclopentadiene to produce terpene-based cycloolefin monomers (Figure 18a). After ROMP copolymerization with cyclooctadiene and 5-norbornene-2-carboxylic acid, performance tests showed that the PSA1 derived from tetrahydrolinalool exhibited potential as a high-shear pressure-sensitive adhesive (Mn=59 kg/mol, Mw/Mn=1.98, Tg= -40 ℃, Td5%= 328 ℃), with a peel strength of (5.2 ± 1.0) N/25 mm and a tack strength of (4.1 ± 0.6) N/25 mm. However, there was still a gap compared to commercial PSAs.
图18 (a) 通过Diels-Alder反应合成萜基功能化降冰片烯[105];(b) 通过酯化反应合成萜基功能化降冰片烯[106-108]

Fig. 18 (a) Synthesis of terpene-functionalized norbornene via Diels-Alder reaction[105]; (b) synthesis of terpene- functionalized norbornene via esterification[106-108]

Glycyrrhetinic acid (GA) is a natural triterpenoid compound, commonly used to construct hydrophobic modules in functional materials. Ju and Hu et al.[106] synthesized a norbornene derivative monomer NGA based on GA (Figure 18b). By copolymerizing NGA with oligo(ethylene glycol)-modified norbornene, they prepared block copolymers that can self-assemble in solution (Mn=24.8~42.4 kg/mol, Mw/Mn=1.15~1.27) and random copolymers (Mn=26.3~31.3 kg/mol, Mw/Mn=1.10~1.19). The spherical core-shell micelles formed by these two types of copolymers exhibited good biocompatibility and low cytotoxicity, and possessed dual responsiveness to temperature and pH, showing potential applications in drug delivery and sustained-release carriers.
Rosin is the non-volatile part of pine resin, and its unique hydrogenated phenanthrene structure provides hydrophobicity and rigidity[107]. Tang and Robertson et al.[108] prepared rosin-based monomer M1 (depicted in Figure 18b) from dehydroabietic acid (DHAA) through carboxyl reduction and esterification reactions, with a total yield of 71.4%. The homopolymer of M1 possesses a high glass transition temperature and high thermal stability. By using M1 as the hard segment monomer and epoxidized soybean oil-modified norbornene as the soft segment monomer, a thermoplastic elastomer was synthesized. When the proportion of M1 is 30%, the thermoplastic elastomer P10 (Mn = 70 kg/mol, Mw/Mn = 1.18) exhibits an elongation at break of up to 258.5% ± 26.5%, a tensile strength of (4.7 ± 0.6) MPa, and demonstrates high elastic recovery, with an elastic recovery rate reaching 99% after the third cycle.

6 Plant Oil Route

The main component of vegetable oil is triglyceride, which can provide unsaturated fatty acids with carbon atom numbers ranging from 14 to 22 after hydrolysis, and is widely used in industries such as biodiesel, plasticizers, lubricants, polyurethanes, coatings, and resins[109]. This section will introduce the synthesis and ring-opening metathesis polymerization (ROMP) of vegetable oil-based norbornene derivatives and macrocyclic olefin monomers (Figure 19).
图19 植物油转化制备环烯烃单体

Fig. 19 Conversion of vegetable oil to cyclic olefin monomers

The functionalization of plant oil triglycerides introduces multiple norbornenyl groups, and after ROMP, a cross-linked network is formed. The addition of external cross-linking agents can increase the ROMP rate, so the resulting plant oil-based polymers are mainly thermosetting materials. Figure 20 lists the structures of plant oil-based norbornene derivative monomers and cross-linking agents, with the corresponding polymer material properties summarized in Table 1. Dilulin is a commercially available modified plant oil prepared by the Diels-Alder reaction between linseed oil (Linseed oil) and dicyclopentadiene (DCPD) under high temperature and pressure. Larock et al. [110-111] studied the copolymerization of Dilulin and DCPD. As the DCPD content increased from 0% to 70%, the resin became harder, and the glass transition temperature and thermal decomposition temperature significantly increased. This is due to the increased cross-linking degree brought about by the increased DCPD content, while the decrease in Dilulin content led to a reduction in the plasticizing effect of the fatty chains.
图20 降冰片烯官能化植物油与交联剂的结构

Fig. 20 Structures of norbornene functionalized vegetable oils and cross linkers

表1 通过ROMP合成的植物油基热塑性材料的性能

Table 1 Properties of vegetable oil-based thermosets

Sample Tg
(°C)
E
(MPa)
εb
(%)
E’
(MPa)
Ref
Dil100 -29 / / / 111
Dil60/DCPD40 21 0.91 138 10 110
Dil50/DCPD50 39 27 132 187 110
Dil40/DCPD60 58 68 156 799 110
Dil30/DCPD70 76 525 35 1769 110
NCO100 -17 / / 2.4 112
NCO80/NCA20 -6 / / 5.7 112
NCO60/NCA40 14 11 16.3 27.8 112
NCO40/NCA60 27 25.7 25.2 130 112
NCO20/NCA80 49 166.6 15.6 583.4 112
NCA100 65 407 13 831.9 112
NCA50/DCPD50 102 1699 14 1455* 113
NCA50/CL50 134 2051 8.8 2046* 113
PNESO 66 / 27.9 / 114

E=Young’s modulus; εb=elongation at break; E’=Storage modulus (at 25 oC),* E’ at 30 ℃

NCO is a monomer produced by the reaction of hydroxyl groups on castor oil with 5-norbornene-2-carbonyl chloride, while NCA is a monomer formed by the reaction of fatty alcohols, obtained from the reduction of fatty acids after hydrolysis of castor oil, with 5-norbornene-2-carbonyl chloride. According to 1H NMR analysis, each fatty acid chain of NCO contains an average of 0.8 norbornene functional groups, and each fatty acid chain of NCA contains an average of 1.8 norbornene functional groups. Larock et al.[112] studied the copolymerization of NCO and NCA and the impact of their structure on material properties, finding that increasing the content of NCA can result in higher thermal stability and energy storage density. Kessler et al.[113] used NCA as a monomer to demonstrate that the mechanical properties and thermal stability of thermosetting materials are closely related to the structure and degree of cross-linking of the cross-linking agent. Yang et al.[114] synthesized NESO (46.3 g) containing four norbornene groups using epoxy soybean oil ESO (30 g) and 5-norbornene-2-carboxylic acid (19.4 g) as raw materials. Based on the Grubbs 2nd generation catalyst, NESO was completely converted into degradable polymer PNESO after 15 minutes of polymerization at 120 ℃. In NaOH and KOH aqueous solutions, PNESO fully degraded into small molecules and oligomers.
Ogawa and Hillmyer et al.[115] used castor oil acid (10 g) as the starting material, successfully synthesizing mono-lactone type macrocyclic olefin ML (1.23 g) and di-lactone type macrocyclic olefin DL (3.3 g) through esterification reactions between hydroxyl and carboxyl groups within the molecule (Figure 21). Under the action of the Grubbs 2nd generation catalyst, polyricinoleic acid was prepared (Mn=51~228 kg/mol, Mw/Mn=1.66~3.27, Tg=-65~-67 ℃, Td5%=254~327 ℃), with the molecular weight distribution of the ML polymer being relatively narrow (Mw/Mn=1.66~2.04). After 13C NMR analysis, the ratio of head-head (HH), head-tail (HT), and tail-tail (TT) connections in the polyricinoleic acid was approximately 25:50:25, and different head-tail connection modes may be the cause of the larger Mw/Mn.
图21 来自蓖麻油酸的大环内酯与ROMP[115]

Fig. 21 Synthesis and ROMP of macrolactones derived from ricinoleic acid[115]

7 Amino Acid Pathway

Amino acids, as the basic building blocks of proteins, have attracted significant attention in bionic and medical materials due to the excellent biocompatibility of amino acid polymers[19]. This section will introduce the synthesis and ring-opening metathesis polymerization (ROMP) of amino acid-based norbornene or cyclooctene derivative monomers and macrocyclic olefin monomers (Figure 22).
图22 氨基酸转化制备环烯烃单体

Fig. 22 Conversion of amino acid to cyclic olefin monomers

As early as the end of the 20th century, North, Grubbs, and Masuda, among others, began to study the synthesis of amino acid-functionalized norbornene derivatives. Kempe and Leiske, et al.[19], provided a detailed review of this, with the synthetic methods illustrated as shown in Figure 23. Gibson and North, et al.[116], synthesized alanine (Ala) monofunctionalized norbornene derivatives, using molybdenum-based catalysts to prepare chiral polymers. Subsequently, the range of amino acids expanded to include glycine (Gly), isoleucine (Ile), leucine (Leu), phenylalanine (Phe), and proline (Pro), as well as dipeptides and tripeptides, with most polymer structures exhibiting good controllability (Mn=10~50 kg/mol, Mw/Mn<1.3)[117-119].
图23 氨基酸功能化降冰片烯的合成[19]

Fig. 23 Synthesis of amino acids functionalized norbornenes[19] (Copyright © 2021 Wiley-VCH GmbH)

Masuda et al.[120] synthesized Ala, Leu, and Phe-derived bifunctionalized endo, endo-configured norbornene derivatives, which could not undergo ROMP homopolymerization and had a maximum content of only 9% in copolymerization, with a broad molecular weight distribution (Mw/Mn=2.0~4.7). For the same composition, exo, exo- and endo, exo-configured monomers undergoing ROMP[121], under the same reaction conditions, the exo, exo-configured monomers could be prepared with high conversion rates to produce polymers with Mw/Mn<1.4, while the endo, exo-configured monomers, although capable of homopolymerization, had lower conversion rates, indicating that the polymerization rate and reaction controllability of bifunctionalized monomers are closely related to their stereostructure. In addition, the connection method between amino acids and norbornenes also affects the polymerization activity of the monomers. Sanda and Masuda et al.[122] demonstrated that monomers formed through ester bonds (Figure 23a) are more prone to undergo ROMP than those formed through amide bonds (Figure 23e).
The precise control of polymer structure is a key aspect of polymer synthesis chemistry. 3-substituted cyclooctene derivative monomers can be used to prepare polymers with high "head-to-tail" structural regularity and trans stereo-regularity[123]. Tao et al.[124] designed and synthesized a series of glycine-functionalized 3-substituted and 5-substituted cyclooctene monomers (Figure 24), with the total yield for 3-substituted monomers being 48% to 64%, and for 5-substituted monomers, 47% to 71%, relative to the starting material, methyl glycinate hydrochloride. In the presence of Grubbs 2nd generation catalyst and chain transfer agent cis-4-octene, 3-substituted polymers (Mn=12.3~26.8 kg/mol, Mw/Mn=1.5~2.0) and 5-substituted polymers (Mn=8.9~20.6 kg/mol, Mw/Mn=1.9~2.7) were obtained. Due to steric hindrance, 3-substituted monomers have the lowest free energy barrier for "head-to-tail" connection, thus exhibiting high "head-to-tail" regional regularity and high trans stereo-regularity, whereas 5-substituted monomers do not show selectivity. Additionally, only 3Gly0COE and 3Gly1COE could produce semi-crystalline polymers, with crystallinities of 21% and 35%, respectively, while the rest of the polymers were amorphous. 3-substituted polymers also exhibited a larger contact angle (100o vs 89o) and cloud point (Tc=26 ℃ vs 24.7 ℃) compared to 5-substituted polymers, indicating that the crystallization properties, hydrophilicity, and thermal response behavior of the materials are influenced by the polymer's regional and stereo structures. Tao et al.[125] also synthesized two leucine-functionalized 3-substituted cyclooctene monomers, from which reverse micelles with a diameter of 30 nm were prepared by screening the ratio of comonomers and types of solvents.
图24 3-取代和5-取代氨基酸功能化环辛烯的合成与ROMP[124]

Fig. 24 Synthesis and ROMP of 3-glycine-substituted and 5-glycine-substituted cyclooctenes[124]

Ring-closing metathesis (RCM) is an intramolecular olefin metathesis reaction that can be used to synthesize macrocyclic olefins. Schlaad et al[126] started from tert-butyl (Boc)-protected L-cystine, synthesizing two cystine-derived macrocyclic olefins ( Figure 25 ) with a total yield of 19%~77%. Under the action of the Grubbs 3rd generation catalyst, reacting at 40 ℃ for 30 min, they prepared functionalized polymers containing disulfide bonds in the main chain ( Mn=58.7~ 80 kg/mol, Mw/Mn =1.7~2.2, Tg =2~41 ℃).
图25 L-胱氨酸基大环单体的合成与ROMP[126]

Fig. 25 Synthesis and ROMP of L-cystine-based macrocyclic monomers[126]

8 Conclusions and Future Prospects

The preparation of bio-based polymer materials through biomass conversion is of great significance for the dual carbon strategy and sustainable development. Based on synthesis strategies such as modification of natural cycloalkenes, Diels-Alder reactions of bio-based dienes, cyclization reactions of linear bio-based alkenes, and introduction of bio-based components via norbornene/cyclooctene, biomass sources like cellulose, hemicellulose, lignin, terpenes, plant oils, and amino acids are converted into bio-based cycloalkene monomers, including norbornene derivatives, oxanorbornene derivatives, cyclooctene derivatives, and macrocyclic alkenes. Through ring-opening metathesis polymerization, new functional polyolefin materials with different structures and properties have been successfully developed, achieving phased results. Looking to the future, the synthesis of bio-based cycloalkenes and their ring-opening metathesis polymerization should focus on the following directions: (1) The synthesis efficiency of bio-based cycloalkenes is relatively low; the yield of monomers, separation processes, and monomer purity are crucial for the ring-opening metathesis polymerization reaction, costs, and downstream applications. There is an urgent need to develop efficient monomer synthesis strategies, including new biocatalytic, chemical transformation, and bio-chemical catalysis coupling methods; (2) The rules governing the ring-opening metathesis polymerization of bio-based cycloalkenes are not yet clear. Given the uniqueness of the structure of bio-based monomers, systematic research on catalysts and reactors is needed to elucidate the relationship between material structure and performance; (3) Research on engineering conversion is weak, and the synthesis of bio-based cycloalkenes and their ring-opening metathesis polymerization are still at the basic or applied basic research stage. The practical value and application fields of functional group-containing materials deserve in-depth exploration. It is imperative to strengthen the collaboration among industry, academia, and research, to develop high-performance bio-based polyolefin materials and promote the industrialization of some varieties as soon as possible.
[1]
Cywar R M, Rorrer N A, Hoyt C B, Beckham G T, Chen E Y X. Nat. Rev. Mater., 2022, 7(2): 83.

[2]
Geyer R, Jambeck J R, Law K L. Sci. Adv., 2017, 3(7): e1700782.

[3]
Zhu Y Q, Romain C, Williams C K. Nature, 2016, 540(7633): 354.

[4]
Li T, Chen C J, Brozena A H, Zhu J Y, Xu L X, Driemeier C, Dai J Q, Rojas O J, Isogai A, Wågberg L, Hu L B. Nature, 2021, 590(7844): 47.

[5]
Zhang Z R, Song J L, Han B X. Chem Rev., 2017, 117(10): 6834.

[6]
Upton B M, Kasko A M. Chem. Rev., 2016, 116(4): 2275.

[7]
Jing Y X, Guo Y, Xia Q N, Liu X H, Wang Y Q. Chem, 2019, 5(10): 2520.

[8]
Wang S Y, Cheng A H, Liu F H, Zhang J J, Xia T, Zeng X, Fan W, Zhang Y. Ind. Chem. Mater., 2023, 1(2): 188.

[9]
Gérardy R, Debecker D P, Estager J, Luis P, Monbaliu J C M. Chem. Rev., 2020, 120(15): 7219.

[10]
Delidovich I, Hausoul P J C, Deng L, Pfützenreuter R, Rose M, Palkovits R. Chem. Rev., 2016, 116(3): 1540.

[11]
Ma X Z, Luo Q, Qin D D, Chen J, Zhu J, Yan N. Progress in Chemistry, 2020, 32(5):617.

(马晓振, 罗清, 秦冬冬, 陈景, 朱锦, 颜宁. 化学进展, 2020, 32(5):617.).

[12]
Qiao Y, Teng N, Zhai C K, (Na/nuo) H N, Zhu J. Prog. Chem., 2018, 30(9): 1415

(乔颖, 腾娜, 翟承凯, 那海宁, 朱锦. 化学进展, 2018, 30(9):1415.).

[13]
Mukhtar Gunam Resul M F, Rehman A, Saleem F, Usman M, López Fernández A M, Eze V C, Harvey A P. RSC Adv., 2023, 13(47): 32940.

[14]
Mosquera M E G, Jiménez G, Tabernero V, Vinueza-Vaca J, García-Estrada C, Kosalková K, Sola-Landa A, Monje B, Acosta C, Alonso R, Valera M Á. Sustain. Chem., 2021, 2(3): 467.

[15]
Winnacker M, Rieger B. ChemSusChem, 2015, 8(15): 2455.

[16]
Hao J, Gao Y X, Chen H R, Hu J, Ju Y. Acta Polym. Sin., 2020, 51(3): 239

(郝杰, 高玉霞, 陈厚睿, 胡君, 巨勇. 高分子学报, 2020, 51(3): 239.).

[17]
Zhang C Q, Garrison T F, Madbouly S A, Kessler M R. Prog. Polym. Sci., 2017, 71: 91.

[18]
Miao S D, Wang P, Su Z G, Zhang S P. Acta Biomater., 2014, 10(4): 1692.

[19]
Leiske M N, Kempe K. Macromol. Rapid Commun., 2022, 43(2): 2100615.

[20]
Huang W J, Zhu N, Fang Z, Guo K. Progress in Chemistry, 2018, 30(12):1836.

(黄卫军, 朱宁, 方正, 郭凯. 化学进展, 2018, 30(12):1836.).

[21]
Dourado Fernandes C, Francisco Oechsler B, Sayer C, de Oliveira D, Hermes de Araújo P H. Eur. Polym. J., 2022, 169: 111132.

[22]
Wang J, Zhou J, Ding Y Z, Hu X G, Chen Y W. Polym. Chem., 2023, 14(20): 2414.

[23]
Yuan P J, Hong M. Acta Polymerica Sinica, 2019, 50(4):327.

(袁鹏俊, 洪缪. 高分子学报, 2019, 50(4):327.).

[24]
Balla E, Daniilidis V, Karlioti G, Kalamas T, Stefanidou M, Bikiaris N D, Vlachopoulos A, Koumentakou I, Bikiaris D N. Polymers, 2021, 13(11): 1822.

[25]
Li M S, Tao Y H. Polym. Chem., 2021, 12(10): 1415.

[26]
Qin Y S. J. Funct. Polym., 2019, 32(5): 558.

(秦玉升. 功能高分子学报, 2019, 32(5): 558.).

[27]
Kessaratikoon T, Rehman H U, D’Elia V, Crespy D. Chin. J. Chem., 2024, 42(6): 652.

[28]
Veith C, Diot-Néant F, Miller S A, Allais F. Polym. Chem., 2020, 11(47): 7452.

[29]
Wang J F, Yuan L, Wang Z K, Rahman M A, Huang Y C, Zhu T Y, Wang R B, Cheng J J, Wang C P, Chu F X, Tang C B. Macromolecules, 2016, 49(20): 7709.

[30]
Palà M, Woods S E, Hatton F L, Lligadas G. Macromol. Chem. Phys., 2022, 223(13): 2200005.

[31]
Wahlen C, Frey H. Macromolecules, 2021, 54(16): 7323.

[32]
Wadgaonkar S P, Wagner M, Baptista L A, Cortes-Huerto R, Frey H, Müller A H E. Macromolecules, 2023, 56(2): 664.

[33]
Liu B, Li S H, Wang M Y, Cui D M. Angew. Chem. Int. Ed., 2017, 56(16): 4560.

[34]
You F, Shi W Y, Yan X Y, Wang X Y, Shi X C. Chem., 2022, 17(23).

[35]
Magaña I, López R, Enríquez-Medrano F J, Kumar S, Aguilar-Sanchez A, Handa R, Díaz de León R, Valencia L. J. Mater. Chem. A, 2022, 10(10): 5019.

[36]
Sun H, Ibrahim T, Ritacco A, Durkee K. ACS Macro Lett., 2023, 12(12): 1642.

[37]
Wu Y C, Fan H Z, Zhang W, Wang M Y, Cai Z Z, Zhu J B. Macromolecules, 2022, 55(20): 9232.

[38]
Ogba O M, Warner N C, O’Leary D J, Grubbs R H. Chem. Soc. Rev., 2018, 47(12): 4510.

[39]
Schrock R R. Acc. Chem. Res., 2014, 47(8): 2457.

[40]
Bielawski C W, Grubbs R H. Prog. Polym. Sci., 2007, 32(1): 1.

[41]
Hyatt M G, Walsh D J, Lord R L, Andino Martinez J G, Guironnet D. J. Am. Chem. Soc., 2019, 141(44): 17918.

[42]
Sun H, Liang Y F, Thompson M P, Gianneschi N C. Prog. Polym. Sci., 2021, 120: 101427.

[43]
Xu J X, Hadjichristidis N. Prog. Polym. Sci., 2023, 139: 101656.

[44]
Feng Y C, Jie S Y, Li B G. Prog. Chem., 2015, 27(8): 1074

(冯雨晨, 介素云, 李伯耿. 化学进展, 2015, 27(8):1074.).

[45]
Zhou C L, Hou C P, Chen W, Wang L J, Cheng J H. Acta Chemica Sinica, 2022, 80(02):229-236.

(周楚璐, 侯翠苹, 陈伟, 王立杰, 程建华. 化学学报, 2022, 80(02): 229.).

[46]
Blosch S E, Scannelli S J, Alaboalirat M, Matson J B. Macromolecules, 2022, 55(11): 4200.

[47]
Chen K R, Han W J, Hu X, Liu Y H, Hu Y J, Zhao S F, Zhu N, Fang Z, Guo K. Chem. Eng. J., 2022, 437: 135284.

[48]
Chen K R, Hu X, Qiu J K, Zhu N, Guo K. Prog. Chem., 2020, 32(1): 93

(陈柯睿, 胡欣, 邱江凯, 朱宁, 郭凯. 化学进展, 2020, 32(1):93.).

[49]
Varlas S, Lawrenson S B, Arkinstall L A, O’Reilly R K, Foster J C. Prog. Polym. Sci., 2020, 107: 101278.

[50]
Varlas S, Foster J C, O’Reilly R K. Chem. Commun., 2019, 55(62): 9066.

[51]
Cornel E J, Jiang J H, Chen S, Du J Z. CCS Chem., 2021, 3(4): 2104.

[52]
Liu C, Hong C Y, Pan C Y. Polym. Chem., 2020, 11(22): 3673.

[53]
Zhang H W, Han W J, Liu Y H, Hu X, Zhu N, Guo K. Acta Polymerica Sinica, 2023, 54(11):1663.

(张浩伟, 韩文鉴, 刘一寰, 胡欣, 朱宁, 郭凯. 高分子学报, 2023, 54(11):1663.).

[54]
Tao Y X, Chen L L, Liu Y H, Hu X, Zhu N, Guo K. Acta Polymerica Sinica, 2022, 53(12):1445.

(陶永鑫, 陈蕾蕾, 刘一寰, 胡欣, 朱宁, 郭凯. 高分子学报, 2022, 53(12):1445.).

[55]
Huang X, Ren J, Ran J Y, Qin C L, Yang Z Q, Cao J P. Fuel Process. Technol., 2022, 229: 107175.

[56]
Fraser C, Grubbs R H. Macromolecules, 1995, 28(21): 7248.

[57]
Iyer S, Rele S, Grasa G, Nolan S, Chaikof E L. Chem. Commun., 2003, (13): 1518.

[58]
Murphy J J, Nomura K, Paton R M. Macromolecules, 2006, 39(9): 3147.

[59]
Weaver L G, Singh Y, Burn P L, Blanchfield J T. RSC Adv., 2016, 6(37): 31256.

[60]
Liu Z F, Zhu Y, Ye W L, Wu T, Miao D Y, Deng W, Liu M N. Polym. Chem., 2019, 10(29): 4006.

[61]
Yang J X, Si C L, Liu K, Liu H Y, Li X Y, Liang M. Journal of Forestry Engineering, 2020, 5(5):21.

(杨佳鑫, 司传领, 刘坤, 刘华玉, 李晓云, 梁敏. 林业工程学报, 2020, 5(5):21.).

[62]
Dell’Acqua A, Stadler B M, Kirchhecker S, Tin S, de Vries J G. Green Chem., 2020, 22(16): 5267.

[63]
Guilmanov V, Ballistreri A, Impallomeni G, Gross R A. Biotechnol. Bioeng., 2002, 77(5): 489.

[64]
Gao W, Hagver R, Shah V, Xie W C, Gross R A, Ilker M F, Bell C, Burke K A, Coughlin E B. Macromolecules, 2007, 40(2):145.

[65]
Peng Y F, Decatur J, Meier M A R, Gross R A. Macromolecules, 2013, 46(9):3293.

[66]
Zini E, Gazzano M, Scandola M, Wallner S R, Gross R A. Macromolecules, 2008, 41(20):7463.

[67]
Raju S, Jastrzebski J T B H, Lutz M, Klein Gebbink R J M. ChemSusChem, 2013, 6(9):1673.

[68]
Tshibalonza N N, Monbaliu J C M. Green Chem., 2020, 22(15):4801.

[69]
McGuire T M, Pérale C, Castaing R, Kociok-Köhn G, Buchard A. J. Am. Chem. Soc., 2019, 141(34):13301.

[70]
Cheng J, Li J, Zheng L G. J. Agric. Food Chem., 2021, 69(36):10480.

[71]
Leite L, Stonkus V, Edolfa K, Ilieva L, Plyasova L, Zaikovskii V. Appl.Catal. A Gen., 2006, 311:86.

[72]
Feist J D, Xia Y. J. Am. Chem. Soc., 2020, 142(3): 1186.

[73]
Mandal A, Mandal I, Kilbinger A F M. Macromolecules, 2022, 55(17):7827.

[74]
Mandal A, Kilbinger A F M. Polym. Chem., 2023, 14(23):2797.

[75]
Mandal A, Pal S, Kilbinger A F M. Macromol. Rapid Commun., 2023, 44(18):2300218.

[76]
Mandal A, Mandal I, Kilbinger A F M. Angew. Chem. Int. Ed., 2023, 62(4):e202211842

[77]
Wang W W, Jiang J, An Q L, Shi R, Shao S Y, Wang Z. ACS Sustainable Chem. Eng., 2023, 11(39):14622.

[78]
An T, Ryu H, Choi T L. Angew. Chem., Int. Ed., 2023, 62(47):e202309632.

[79]
Feist J D, Lee D C, Xia Y. Nat. Chem., 2022, 14(1):53.

[80]
Mandal A, Kilbinger A F M. ACS Macro Lett., 2023, 12(10):1372.

[81]
Leguizamon S C, Lyons K, Monk N T, Hochrein M T, Jones B H, Foster J C. ACS Appl. Mater. Interfaces, 2022, 14(45):51301.

[82]
Comba M B, Tsai Y H, Sarotti A M, Mangione M I, Suárez A G, Spanevello R A. Eur. J. Org. Chem., 2018, 2018(5): 590.

[83]
Camp J E, Greatrex B W. Front. Chem., 2022, 10:902239

[84]
Debsharma T, Yagci Y, Schlaad H. Angew. Chem. Int. Ed., 2019, 58(51):18492.

[85]
Debsharma T, Behrendt F N, Laschewsky A, Schlaad H. Angew. Chem. Int. Ed., 2019, 58(20):6718.

[86]
Fadlallah S, Peru A A M, Flourat A L, Allais F. Eur. Polym. J., 2020, 138:109980.

[87]
Nomura K, Schrock R R. Macromolecules, 1996, 29(2):540.

[88]
Galkin K I, Ananikov V P. Int. J. Mol. Sci., 2021, 22(21): 11856.

[89]
Bai Y J, De bruyn M, Clark J H, Dodson J R, Farmer T J, Honoré M, Ingram I D V, Naguib M, Whitwood A C, North M. Green Chem., 2016, 18(14):3945.

[90]
Bai Y J, Clark J H, Farmer T J, Ingram I D V, North M. Polym. Chem., 2017, 8(20):3074.

[91]
Lawrenson S B, Hart S, Ingram I D V, North M, Parker R R, Whitwood A C. ACS Sustainable Chem. Eng., 2018, 6(8):9744.

[92]
Naguib M, Schiller T L, Keddie D J. Macromol. Rapid Commun., 2018, 39(7):1700794.

[93]
Blanpain A, Clark J H, Farmer T J, Guo Y L, Ingram I D V, Kendrick J E, Lawrenson S B, North M, Rodgers G, Whitwood A C. ChemSusChem, 2019, 12(11):2393.

[94]
Naguib M, Yassin M A. ACS Appl. Polym. Mater., 2022, 4(3):2181.

[95]
Fang L X, Zhou J F, Tao Y Q, Wang Y Q, Chen X R, Chen X Y, Hou J R, Sun J, Fang Q. ACS Sustainable Chem. Eng., 2019, 7(4):4078.

[96]
Hu Y, Ran Q Y, Wei S P, Wang C C, Wu Z J, Xu E H, Luo Z Y, Jia P Y, Sha Y. Green Chem., 2023, 25(15):5858.

[97]
Koo B. ACS Appl. Polym. Mater., 2024, 6(3):1653.

[98]
Usman M, Rehman A, Saleem F, Abbas A, Eze V C, Harvey A. RSC Adv., 2023, 13(33):22717.

[99]
Della Monica F, Kleij A W. Polym. Chem., 2020, 11(32): 5109.

[100]
Grau E, Mecking S. Green Chem., 2013, 15(5):1112.

[101]
Medeiros A M M S, Le Coz C, Grau E. ACS Sustainable Chem. Eng., 2020, 8(11):4451.

[102]
Lee J C, Parker K A, Sampson N S. J. Am. Chem. Soc., 2006, 128(14):4578.

[103]
Strick B F, Delferro M, Geiger F M, Thomson R J. ACS Sustainable Chem. Eng., 2015, 3(7):1278.

[104]
Yarolimek M R, Bookbinder H R, Coia B M, Kennemur J G. ACS Macro Lett., 2021, 10(6):760.

[105]
Engelen S, Droesbeke M, Aksakal R, Du Prez F E. ACS Macro Lett., 2022, 11(12):1378.

[106]
Li Y, Gao Y X, Wang B, Hao J, Hu J, Ju Y. Chem., 2018, 13(18):2723.

[107]
Wilbon P A, Chu F X, Tang C B. Macromol. Rapid Commun., 2013, 34(1):1.

[108]
Ganewatta M S, Ding W Y, Rahman M A, Yuan L, Wang Z K, Hamidi N, Robertson M L, Tang C B. Macromolecules, 2016, 49(19):7155.

[109]
Gandini A, Lacerda T M, Carvalho A J F, Trovatti E. Chem. Rev., 2016, 116(3):1637.

[110]
Henna P H, Kessler M R, Larock R C. Macromol. Mater. Eng., 2008, 293(12):979.

[111]
Henna P, Larock R C. J. Appl. Polym. Sci., 2009, 112(3):1788.

[112]
Xia Y, Larock R C. Polymer, 2010, 51(12):2508.

[113]
Ding R, Xia Y, Mauldin T C, Kessler M R. Polymer, 2014, 55(22):5718.

[114]
Li W R, Zhan Q Y, Yang P. ACS Sustainable Chem. Eng., 2023, 11(3):1200.

[115]
Ogawa R, Hillmyer M A. Polym. Chem., 2021, 12(15):2253.

[116]
Coles M P, Gibson V C, Mazzariol L, North M, Teasdale W G, Williams C M, Zamuner D. J. Chem. Soc., Chem. Commun., 1994, (21):2505.

[117]
Biagnini S C G, Coles M P, Gibson V C, Giles M R, Marshall E L, North M. Polymer, 1998, 39(5):1007.

[118]
Biagini S C G, Gareth Davies R, North M, Gibson V C, Giles M R, Marshall E L, Robson D A. Chem. Commun., 1999, (3):235.

[119]
Maynard H D, Okada S Y, Grubbs R H. Macromolecules, 2000, 33(17):6239.

[120]
Sutthasupa S, Terada K, Sanda F, Masuda T. J. Polym. Sci. Part A Polym. Chem., 2006, 44(18):5337.

[121]
Sutthasupa S, Sanda F, Masuda T. Macromol. Chem. Phys., 2008, 209(9):930.

[122]
Sutthasupa S, Terada K, Sanda F, Masuda T. Polymer, 2007, 48(11):3026.

[123]
Kobayashi S, Pitet L M, Hillmyer M A. J. Am. Chem. Soc., 2011, 133(15): 5794.

[124]
Li M S, Cui F C, Li Y Q, Tao Y H, Wang X H. Macromolecules, 2016, 49(24):9415.

[125]
Li M S, Chen J L, Tao Y H. Chin. J. Appl. Chem, 2020, 37(3):280.

(李茂盛, 陈金龙, 陶友华. 应用化学, 2020, 37(3):280.).

[126]
Behrendt F N, Schlaad H. Polym. Chem., 2017, 8(2): 366.

Outlines

/