Open Access Article
Jieyuan
Zheng†
ab,
Xingfen
Huang†
ab,
Pingwei
Liu
ab,
Wenjun
Wang
*ab and
Hong
Fan
*bc
aInstitute of Zhejiang University – Quzhou, 99 Zheda Road, Quzhou 324000, China
bCollege of Chemical and Biological Engineering, Zhejiang University, Hangzhou 310027, China. E-mail: hfan@zju.edu
cWenzhou Advanced Materials International Sci-Tech Innovation Center, Institute of Wenzhou, Zhejiang University, Wenzhou 325006, China
First published on 16th September 2025
To meet sustainable development demands, reducing non-renewable resource consumption while promoting material recycling has become imperative. For thermosetting resins, the commercialization of high-performance bio-based alternatives remains challenging due to biomass cost constraints and processing limitations. Herein, we demonstrate a facile method for large-scale preparation of a eugenol-derived thermoset monomer (EUEP) characterized by high yield, high purity, and low waste generation. Through acrylation modification, a fully bio-based resin system with dual UV-thermal curing capability was developed. The material's initial shape stabilization was rapidly achieved via UV-initiated radical polymerization, while subsequent thermally induced transesterification allowed precise modulation of covalent crosslinking network density. This sequential curing approach enabled controllable tuning of the resin's mechanical properties, achieving tensile strength ranging from 2 to 23 MPa with the corresponding elongation at break varying between 5% and 18%. Leveraging dynamic ester bond exchange at interfaces, the thermoset exhibited intrinsic crack-repairing and self-adhesion functionalities. This methodology establishes a novel strategy for developing fully bio-based reworkable resin systems, demonstrating significant potential for enhancing 3D printing process efficiency.
Dynamic covalent bonds (DCBs) have emerged as a promising strategy for designing recyclable thermosetting polymers, particularly in photopolymer systems.8,9 These bonds undergo reversible cleavage and reformation under external stimuli such as thermal, photonic, or chemical activation,10–12 enabling topological reorganization of polymer networks. This unique behavior effectively converts conventional thermosets into materials exhibiting thermoplastic-like processability. Various DCBs have been explored for recycling thermosets, such as ester,13 imine,14 Diels–Alder (DA),15 disulfide,16 and boronic ester DCBs.17 Notably, DCBs can maintain network connectivity while exhibiting Arrhenius-like flow behavior during exchange reactions, a characteristic that defines them as “vitrimers”.18
Within photopolymer systems, acrylate-based resins constitute a dominant commercial technology in both ink formulations and additive manufacturing, primarily owing to their rapid processing kinetics and adjustable mechanical performance. The ester functionalities within these polymers serve as critical sites for dynamic covalent reorganization, enabling the transformation of thermosetting photopolymers into vitrimeric materials. Zinc(II) salts, particularly zinc acetate (Zn(OAc)2), have been widely adopted as effective transesterification catalysts. Seminal work by Leibler's group19 established concentration-dependent control of exchange kinetics in ester-based networks through Zn(OAc)2 modulation. Subsequent studies by Tournilhac et al.20 revealed that Zn2+ coordination with ester groups enhances carbonyl electrophilicity while stabilizing nucleophilic alkoxide intermediates. Zhang and coworkers21 pioneered the development of Zn(acac)2-catalyzed photopolymers with 3D-printable, self-healing capabilities via intercatenary transesterification. This dynamic bonding mechanism not only increased network crosslink density but also enhanced cured resin tensile strength by 40% relative to conventional systems. Nevertheless, absolute tensile strengths remain constrained (15 MPa), while persistent dependence on petrochemical monomers continues to hinder sustainable implementation.
Bio-based feedstocks are emerging as pivotal components in sustainable material innovation, motivated by escalating ecological imperatives and finite petroleum reserves. Substantial research progress has been documented for biomass-derived precursors including lignin,22 vanillin,23 eugenol,24 vegetable oils,25 and furan compounds,26 with varying technology readiness levels. While triglyceride derivatives and cardanol-based systems have achieved industrial adoption, their cured networks frequently display compromised mechanical integrity due to structural limitations inherent in their long aliphatic chains. Conversely, rigid bio-based monomers such as eugenol, vanillin derivatives, and furan-based architectures have demonstrated superior performance metrics to petroleum-based analogs like bisphenol A-glycidyl methacrylate (BisGMA) in terms of modulus and thermal stability.27–29 Eugenol, a naturally occurring phenolic compound from clove oil, has attracted significant interest as a thermoset precursor owing to its multiple reactive sites (allyl, aromatic, and phenolic hydroxyl groups) and cost-effectiveness. Our team recently developed a novel synthesis protocol for high-purity eugenol epoxy (EUEP) under solvent-free conditions, demonstrating exceptional performance in UV-curable coating formulations.30,31 Despite these advancements, eugenol-based photopolymer development remains predominantly confined to laboratory-scale investigations, with persistent obstacles in scalable production and unresolved processing complexities hindering industrial implementation.
In this study, we report for the first time a scalable and green synthesis process for a bio-based aromatic epoxy resin. A method was developed for the synthesis of a high-purity eugenol-based epoxy monomer (EUEP), which also enables the efficient recycling of raw materials. Subsequently, a resin monomer capable of UV-thermal dual curing was developed through the acrylation of EUEP. The resin is first cured by UV radiation via its unsaturated double bonds to establish the material's basic shape. During the subsequent thermal curing stage, dynamic transesterification reactions involving the abundant ester and hydroxyl groups are utilized to modulate the cross-linking density, thereby tuning the resin's mechanical properties. Furthermore, the presence of these dynamic ester bonds endows the material with reparability and reprocessability. This has been applied in 3D printing, where the material demonstrates unique welding capabilities.
![]() | ||
| Fig. 1 (a) EUEP pilot test procedure, (b) synthesis of EUEP, and (c) GC spectra of two step products. | ||
1H-NMR: 6.83 (d, 1H, ArH), 6.71 (m, 2H, ArH), 6.40 (m, 1H, CH), 6.13 (m, 1H, CH), 5.91 (m, 1H, CH), 5.79 (m, 1H, CH), 5.06 (m, 2H, CH2), 4.54 (m, 1H, CH), 4.29 (m, H, CH2), 4.10–3.94 (m, 4H, CH2), 3.78 (s, 3H, CH3), 3.31 (m, 2H, CH2).
H-MQ 1H NMR: 4.66 (s, H, SiH), 0.01 (s, CH3).
H-MQ (20 g), EUEP (24 g), toluene (10 g), and Karstedt catalyst (2 wt% platinum, 0.025 g) were introduced into a glass flask fitted with a reflux condenser. The resulting solution was reacted at 70 °C for twelve hours. After that, toluene was removed by rotary evaporation, yielding EUEP-MQ, 43 g.
EUEP-MQ 1H NMR: 6.71 (d, 1H, ArH), 6.56 (m, 2H, ArH), 4.06 (m, H, CH), 3.88 (m, H, CH), 3.72 (s, 3H, CH3), 3.24 (s, H, CH), 2.75 (m, H, CH), 2.59 (m, H, CH), 2.43 (q, 2H, CH2), 1.49 (t, 2H, CH2), 0.50 (m, 2H, CH2), 0.01 (m, CH3).
EUEP-MQ (20 g) was transferred into a glass flask fitted with a reflux condenser. A mixture of triphenylphosphine (0.7 g, 2.6 mmol), acrylic acid (14.4 g, 0.22 mol), and p-methoxyphenol (0.67 g, 5 mmol) was added into the flask dropwise over 30 minutes. The reaction mixture was agitated at 100 °C until the acid value was below 5 mg KOH per g, to obtain eugenol acrylate MQ silicone resin (EUAC-MQ), 31 g.
EUAC-MQ 1H NMR: 6.73 (d, 1H, ArH), 6.54 (m, 2H, ArH), 6.29 (m, H, CH2), 6.05 (m, H, CH2), 5.73 (m, H, CH), 4.22 (m, H, CH2), 4.10 (m, 2H, CH2), 3.91 (m, 2H, CH2), 3.67 (s, 3H, CH3), 2.41 (m, 2H, CH2), 1.51 (m, 2H, CH2), 0.49 (m, 2H, CH2), 0.01 (m, CH3).
| Sample | EUAC/g | EUAC-1.4MQ/g | Zn(acac)2/g | Photoinitiator/g |
|---|---|---|---|---|
| EUAC | 100 | 0 | 6 | 1 |
| EUAC-MQ-10% | 90 | 10 | 6 | 1 |
| EUAC-MQ-20% | 80 | 20 | 6 | 1 |
To improve the efficiency and environmental sustainability of EUEP synthesis, we optimized the reaction protocol and characterized the products using GC-MS (Fig. 1c and Table 2); the mass spectra corresponding to each chromatographic peak are presented in Fig. S1. In the first stage, eugenol and epichlorohydrin were reacted at a 1
:
5 molar ratio. After etherification, excess epichlorohydrin was recovered via vacuum distillation. GC-MS analysis indicated that 92% of the eugenol was converted into EUEP (retention time: 7.9 min) and EUO (8.4 min). Notably, a portion of eugenol directly formed EUEP while simultaneously generating 1,3-dichloropropanol (4.4 min) during the initial etherification step. Due to the reversible nature of the reaction, EUO reverted to eugenol (6.7 min) upon solvent removal. Under alkaline conditions, 1,3-dichloropropanol underwent cyclization to regenerate epichlorohydrin, which subsequently reacted with residual eugenol. Meanwhile, EUO was converted to EUEP in the second step. This process yielded highly pure EUEP (98.6%) without requiring additional purification, and spontaneous crystallization at 4 °C produced white crystals.
| Procedure | Num. | Time/min | Component | Content/% |
|---|---|---|---|---|
| Step 1 | A | 4.4 | 1,3-Dichloro-2-propanol | 10.6 |
| B | 6.7 | Eugenol | 6.6 | |
| C | 7.9 | EUEP | 34.5 | |
| D | 8.4 | EUO | 44.3 | |
| Step 2 | C | 7.9 | EUEP | 98.6 |
This process significantly improves both the conversion rate and product purity (>98%), ensures high-purity epichlorohydrin recovery, and minimizes byproduct formation (only NaCl and water). Furthermore, it can be extended to other aromatic bio-based epoxy resin syntheses, such as those derived from cardanol, vanillin, and resveratrol. Compared to conventional bisphenol A glycidyl ether (BADGE)-type epoxy resin synthesis, this method offers multiple advantages, including a streamlined procedure, substantial reduction in waste emissions (gaseous, liquid, and solid), high conversion efficiency, and exceptional product purity, rendering it highly suitable for industrial-scale production.
![]() | ||
| Fig. 3 NMR spectra of EUEP and EUAC: (a) 1H NMR; (b) 13C NMR; and (c) 1H NMR spectra of EUAC-MQ; and (d) FT-IR spectra of EUAC, EUAC, and EUAC-MQ-10%. | ||
Fig. 2 outlines the synthesis of the epoxy MQ resin, EUEP-MQ, and its esterified derivative, EUAC-MQ, with each step monitored by NMR analysis (Fig. 3c). Firstly, H-MQ was synthesized via the hydrolytic condensation of TEOS, followed by the capping of HMM with MM. The proton NMR spectrum revealed resonances at 4.6 and 0 ppm, which were attributed to the proton atoms in Si–H and Si–CH3, respectively. Secondly, the allyl groups of EUEP underwent hydrosilylation with H-MQ. As a result, the Si–H signal at 4.6 ppm disappeared, and new resonances appeared at 0.5 and 1.5 ppm, which were allocated to the –Si–CH2– and –Si–CH2–CH2– groups, respectively. Finally, solvent was introduced to reduce system viscosity esterification and to prevent the self-polymerization of acrylic double bonds. The epoxy peaks disappeared, and –CH
CH2 resonances were observed between 5.7 and 6.4 ppm. These data provide strong evidence for the successful synthesis of EUAC-MQ.
CH2, while the peak at approximately 1635 cm−1 was associated with the C
C groups. After EUAC and EUAC-MQ were mixed and cured under UV light, these peaks disappeared, indicating that the double bonds had converted to saturated bonds due to polymerization. The abundant hydroxyl groups, observed at 3500 cm−1, and ester groups, observed at 1740 cm−1, within the crosslinked network are likely to accelerate the transesterification reaction during the subsequent thermosetting process.
The curing mechanism of the eugenol-based photothermal resin is shown in Fig. 4. EUAC was incorporated as the reactive monomer, with EUAC-MQ serving as the crosslinking agent in the prepolymer solution. In Stage I, the photoinitiator was activated by UV light to open the C
C bonds of both the monomer and crosslinking agent, forming permanent covalent bonds within seconds, as indicated by the red network. After photopolymerization, the cured product lacked sufficient strength, deforming under the pressure of a 200 g weight.
In Stage II, the cured resin was heated to a high temperature (180 °C), where the transesterification agent, zinc acetylacetonate, facilitated transesterification within the crosslinked network. During transesterification, ester bonds in EUAC broke and reacted with hydroxyl groups of other EUAC molecules, forming new dynamic covalent ester bonds (orange dots). Simultaneously, free diol monomers (orange stars) were generated, reacting with other ester groups. The continuous cleavage and reformation of ester bonds increased the density of crosslinking nodes while ensuring that the total covalent bond count was preserved, thereby improving the strength of the cured product. After curing and crosslinking, the specimen withstood a 200 g load without deformation.
As illustrated in Fig. S2 and Table S1, the thermal degradation performance of the cured materials with varying proportions of the crosslinking agent was investigated. The incorporation of EUAC-1.4MQ was found to enhance the thermal stability and crosslinking density of the cured materials. Specifically, the initial decomposition temperature (T5%) increased from 223.4 °C to a maximum of 239.3 °C with the addition of 20% EUAC-1.4MQ. The first peak degradation temperature (TP1) was observed in the range of 264.2–291.3 °C, which is attributed to the degradation of free diol monomers (indicated by orange stars). Subsequently, a second degradation peak (T) occurred at approximately 390 °C for all samples, corresponding to the cleavage of C–O bonds. On the other hand, an excessive amount of MQ silicone resin led to a higher final char yield, increasing it from 26.6% to 31.9%. These thermogravimetric analysis (TGA) results demonstrate that the dual photothermal bio-based resin possesses excellent thermal stability, highlighting its potential for practical applications.
| Sample | Tensile strength (MPa) | Elongation at break (%) | Young's modulus (MPa) |
|---|---|---|---|
| EUAC-0h | 1.2 ± 0.2 | 15.6 ± 2.4 | 7.5 ± 0.9 |
| EUAC-2h | 2.5 ± 0.3 | 17.6 ± 2.1 | 15.0 ± 1.6 |
| EUAC-4h | 7.3 ± 0.8 | 12.7 ± 2.0 | 60.7 ± 7.5 |
| EUAC-6h | 14.6 ± 1.3 | 2.4 ± 0.5 | 1089 ± 120 |
| EUAC-MQ-10%-0h | 2.5 ± 0.3 | 18.2 ± 2.8 | 18.4 ± 2.1 |
| EUAC-MQ-10%-2h | 6.1 ± 0.7 | 14.4 ± 2.2 | 58.2 ± 6.1 |
| EUAC-MQ-10%-4h | 12.8 ± 1.2 | 11.7 ± 1.8 | 320.3 ± 35.5 |
| EUAC-MQ-10%-6h | 22.6 ± 2.5 | 4.4 ± 0.8 | 1050 ± 115 |
| EUAC-MQ-20%-0h | 4.5 ± 0.5 | 20.2 ± 3.1 | 37.3 ± 4.0 |
| EUAC-MQ-20%-2h | 9.5 ± 1.1 | 17.2 ± 2.5 | 171.5 ± 19.2 |
| EUAC-MQ-20%-4h | 12.6 ± 1.3 | 12.8 ± 2.1 | 346.2 ± 38.1 |
| EUAC-MQ-20%-6h | 17.1 ± 1.9 | 8.5 ± 1.4 | 582.9 ± 62.3 |
![]() | ||
| Fig. 5 Tensile curves of specimens cured at 180 °C for different times: (a) EUAC; (b) EUAC-MQ-10%; and (c) EUAC-MQ-20%; and (d) literature-reported bio-based UV-cured resin tensile properties. | ||
Adding 10 wt% and 20 wt% EUAC-MQ increased the tensile strength of the cured material to 2.5 MPa and 4.4 MPa, respectively, while improving the elongation from 16% to 18% and 20%. Each EUAC-MQ molecule possesses more than four functional groups, which promote the creation of a crosslinked network and improve tensile strength. Additionally, prior research has highlighted that the unique cage structure of MQ silicone resin strengthens the interfacial bonds within the crosslinked network, thereby enhancing toughness.34
After heating the samples at 180 °C for 2, 4, and 6 hours, the tensile strength of all three specimens increased significantly with prolonged heating time. During the thermal stage, ester and hydroxyl groups underwent reversible reactions within the crosslinked network, forming new dynamic bonds in the covalent structure. It is noteworthy that during the UV curing stage, due to the absence of a crosslinking agent, the EUAC forms only a linear network. When subjected to external force, the molecular chains readily undergo relative slippage, leading to plastic deformation or failure of the material under relatively low stress. However, after heating for 2 hours, both the strength and elongation at break increase simultaneously, indicating that the network with low crosslinking density still exhibits significant mobility. Under external force, the molecular segments and the entire network can fully extend, orient, and move, thereby resulting in a large elastic deformation of 2.5 MPa and an elongation at break of 17.6%. After 6 hours of heating, the tensile strength of EUAC reached 14.6 MPa with a reduced elongation at break of 2.4%. With 10% EUAC-MQ incorporated, the tensile strength of EUAC-MQ-10% rose to 22.6 MPa, while its elongation at break improved to 4.6%. Although MQ silicone resin does not take part in transesterification, its rigid cage structure and ability to undergo plastic deformation contribute to the improved strength and toughness of the cured product. Increasing the EUAC-MQ content to 20% further raised the elongation at break to 8.6% after 6 hours of heating, but the tensile strength decreased to 17.2 MPa. Higher MQ silicone resin content could further enhance toughness, but the lower EUAC content reduced crosslinking ester exchange bonds, leading to decreased strength. Among the cured products, EUACQ-10% exhibited the best performance and will be used in subsequent testing and discussion.
As depicted in Fig. 5d, compared with literature-reported bio-based UV-cured resins including epoxidized soybean oil,35–39 eugenol,40 cardanol,41 vanillin,42–44 and malic acid,45 the resin developed in this work exhibits a relatively low elongation at break due to the absence of flexible segments. However, it demonstrates superior tensile strength (22.6 MPa) owing to its densely crosslinked network containing abundant rigid structures and dynamic crosslinking sites. Notably, the EUAC-MQ-10% system enables wide-range tunability of mechanical properties (tensile strength: 2.5–22.6 MPa) through simple adjustment of the heating duration, a distinctive feature not achievable in other reported resin systems. This remarkable property modulation capability provides unprecedented flexibility for tailoring material performance according to specific application requirements.
δ peaks decreased from 0.65 to 0.45, indicating reduced damping and increased rigidity in the crosslinked network.
| Sample | E′, 25 °C (MPa) | T g (°C) | v c (mol m−3) |
|---|---|---|---|
| EUAC-MQ-10%-0h | 797.6 | 18.3 | 1268 |
| EUAC-MQ-10%-2h | 1635.3 | 25.5 | 1610 |
| EUAC-MQ-10%-4h | 3221.2 | 39.1 | 2022 |
| EUAC-MQ-10%-6h | 3339.9 | 46.7 | 2037 |
| EUAC-MQ-10%-8h | 3491.7 | 47.1 | 2049 |
The storage modulus progression mirrored tan
δ behavior, plateauing after 6 hours of treatment (Fig. 6b). This stabilization indicates equilibrium attainment in transesterification-driven network reorganization. The crosslink density (νc) is calculated through rubber elasticity theory using eqn (1):
| vc = E′/3RT | (1) |
![]() | (2) |
The stress relaxation time exhibited a strong temperature dependence, fitting well with the Arrhenius equation. As shown in Fig. 6d, an excellent linear correlation was observed. The calculated activation energy (Ea) was 68.2 kJ mol−1, in agreement with previously reported values.46
This approach involved re-embedding the fractured segments. The fractured segments of the tensile specimens were re-embedded in a fresh resin matrix, which was then light-cured to form a monolithic, integrated sample. This newly formed ensemble was subsequently post-heated at 180 °C for 6 hours before tensile testing. After the fractured specimen was reassembled into a single piece, it was photocured and subsequently annealed at 180 °C for 6 hours. The tensile properties of the repaired material were then evaluated. As shown in Fig. 7d, the original tensile strength was 22.7 MPa, while the repaired specimen exhibited a minimal reduction in strength (21.9 MPa), with no significant change in tensile modulus. These results demonstrate that the material retains excellent mechanical integrity after repair. Furthermore, Fig. 7c reveals that fracture did not occur at the repair interface, confirming that the ester-exchange reaction restored interfacial strength to a level comparable to that of the bulk material.
Footnote |
| † These authors contributed equally. |
| This journal is © The Royal Society of Chemistry 2025 |