Bhavya Parameswarana,
Tuhin Subhra Pala and
Nikhil K. Singha*ab
aRubber Technology Centre, Indian Institute of Technology Kharagpur, West Bengal 721302, India. E-mail: nks@rtc.iitkgp.ac.in; nks8888@yahoo.com
bSchool of Nano Science and Technology, Indian Institute of Technology Kharagpur, West Bengal 721302, India
First published on 3rd September 2025
Vitrimers represent a promising class of new-generation materials with covalent adaptive networks (CANs) based on an associative exchange mechanism. Herein, we utilised epoxy-functionalized elastomers, like poly(ethylene-co-vinyl acetate-co-glycidyl methacrylate) (EVA-GMA), for designing a dual dynamic network based on β-hydroxyl ester linkage as well as disulfide metathesis reactions, which were enabled by a new crosslinker, succinic anhydride-modified 4-aminophenyl disulfide (SA-APDS), which has a disulfide linkage in the backbone and a –COOH group at the para position. These dynamic linkages are capable of undergoing exchange reactions at elevated temperatures, thereby allowing the rearrangement of the network topology and exhibiting vitrimer-like behaviour. The resultant elastomeric vitrimer exhibits good mechanical performance, including a tensile strength of ∼6.1 MPa and elongation at break up to 1300%, demonstrating super-elastomeric characteristics. Interestingly, the elastomeric vitrimer showed fluorescence behaviour due to the presence of a conjugated system in the new crosslinker. The ability of this material to maintain and reconfigure crosslink density through dual associative mechanisms showed that vitrimer-like materials have self-healing, recyclable, and reprocessable characteristics. Stress relaxation experiments confirmed the vitrimeric behaviour with an activation energy of 46.8 kJ mol−1 and a vitrification temperature of 83 °C. This new vitrimeric elastomer, with fluorescence characteristics, can have potential applications in areas where a unique combination of mechanical and optical properties is necessary.
Incorporating dynamic covalent networks is an interesting approach, as they can undergo reversible covalent bond formation and cleavage, enabling them to adapt their properties and respond to external stimuli. For elastomeric materials, this self-repairing process can take place through either an associative bond exchange or a dissociative bond exchange pathway. Associative exchangeable bonds belong to the class of self-healing elastomers, which can form associative crosslinks after a fracture or damage while maintaining constant crosslinking density throughout the healing process.5 Some of the important examples of associative bond exchange reactions are transesterification,6–8 disulfide metathesis,9,10–12 olefin metathesis,13 transamination,14 etc. Dissociative exchangeable bonds represent a versatile class of dynamic material systems. One of the prominent examples of dissociative exchangeable bonds is the Diels–Alder chemistry.15–17 Reversible networks based on supramolecular interactions rely on non-covalent bonds such as hydrogen bonding,18,19 metal coordination,20 π–π stacking21 or host–guest interactions.22 These dynamic interactions enable the material to respond to external stimuli, making the networks reversible, self-healing, and adaptable.
Vitrimers, as coined by Leibler et al., are an attractive subset of dynamic covalent networks with associative exchangeable bonds and are expected to present a new paradigm for developing reusable elastomeric materials, as they can rearrange the network topology at higher temperatures and maintain the same degree of crosslinking even in a recycled state.23 These materials have an interesting property, that is, the crosslinked material can maintain its structural integrity even after it is reprocessed.7,8,24,25 Leibler et al.23 first demonstrated vitrimers by crosslinking diglycidyl ether of bisphenol A with glutaric anhydride and zinc acetylacetonate as the catalyst. There has been tremendous interest in this new class of materials since the last decade.26,27
Recent literature has extensively explored vitrimers derived from disulfide metathesis and transesterification for the crosslinking of epoxy resins, as well as other low molecular weight compounds containing epoxy groups. Zhang et al. demonstrated rapid stress relaxation in an epoxy resin based on bisphenol A crosslinked by β-hydroxy ester linkages and disulfide metathesis.28 In this case, an aliphatic disulfide with a carboxylic acid at each end was utilised as the crosslinker, which endowed the crosslinked epoxy resin with rapid stress relaxation ranging from 1.5 s (200 °C) to 5500 s (60 °C). Vilanova-Pérez et al.29 synthesised two UV-curable vitrimeric materials based on ethylene glycol phenyl ether methacrylate (EGPMA) and poly(ethylene glycol) methyl ether methacrylate (PEGMA) using an aliphatic compound containing disulfide as well as β-hydroxy ester linkages as the crosslinking agent. Both the materials exhibited near-ambient glass transition temperatures (Tg) and excellent vitrimeric behaviour, with rapid stress relaxation and low activation energies, and showed higher topology freezing temperatures (Tv) than Tg, suggesting potential for shape-memory applications. Wang et al.30 synthesised an extended epoxy resin network based on bisphenol A by reacting it with suberic acid to incorporate the β-hydroxy ester linkages, and was then crosslinked with an aromatic disulfide. These materials showed excellent self-healing properties. Huang et al.31 developed a vitrimer containing epoxy resin based on bisphenol A crosslinked with 2,2′-dithiobenzoic acid, where the dynamic crosslinker had an aromatic disulfide with a –COOH group at the ortho position. The material showed vitrimeric behaviour with potential application as a high-performance adhesive.
Chen et al.32 utilized TEMPO-oxidised cellulose nanocrystals [TOCNS] as a covalent crosslinking agent and a reinforcement in epoxidised natural rubber (ENR). Lin et al.33 developed a recyclable and self-healable ENR-based composite by incorporating a citric acid-modified bentonite composite. Lin et al. reported the vitrimer behaviour of the reactive blend of ENR/XNBR without the aid of any external crosslinker.34 Zhang et al.35 developed a malleable, strong and reversible ENR composite by blending it with carboxyl-functionalised carbon nanodots. Liu et al.6 covalently crosslinked ENR with sebacic acid before grafting N-acetyl glycine to simultaneously produce hydroxy ester linkages and amide groups. Cao et al.36 developed an ENR-based vitrimer-like elastomer with high self-healing efficiency. In that work, the hydroxyl groups on cellulose nanocrystals (CNCs) were converted into carboxyl groups by 2,2,6,6-tetramethylpiperidinyl-1-oxyl radical (TEMPO)-mediated oxidation and incorporated into ENR rubber. Thus, in most of the literature reports, mostly ENR-based vitrimer-like materials have been prepared using a single system of transesterification of β-hydroxy ester linkages.
Ethylene-vinyl acetate (EVA) copolymers are an important class of polymers, which can be thermoplastic or elastomeric depending on the content of vinyl acetate (VA).1 Because of its unique properties, like resistance to UV radiation, oil, heat ageing and excellent flexibility, EVA is widely used in automotive parts, shoes, adhesives, coatings, etc. Recently, in EVA elastomers, a few percent of a third monomer, glycidyl methacrylate (GMA), has been introduced to incorporate the epoxy functionality as a pendant group.1
In this study, we have developed a new elastomeric vitrimer using the epoxy functionality in EVA-GMA based on a dual dynamic network using metathesis reactions of disulfides as well as transesterification of β-hydroxy ester linkages. In this case, we have synthesised a new dual dynamic crosslinker with a disulfide linkage in the middle and dicarboxylic acid as the end group, capable of targeting the epoxy groups in EVA-GMA. The choice of synthesising an aromatic disulfide containing crosslinker with the carboxyl groups at the para position was made due to the more favourable geometry and reduced steric hindrance, as the para configuration allows stable and more commending exchange reactions between the disulfide groups.37 Also, the aromatic disulfides benefit from electron delocalization due to the conjugation between the sulfur atoms and the aromatic ring. The resonance effect in aromatic systems can help stabilise the disulfide bond, making it more thermodynamically favourable for breaking and reformation during exchange reactions.38 Herein, we explored a rather inexpensive route to synthesise the aromatic disulfide containing para-carboxylic acid. To the best of our knowledge, there is no report on the development of a vitrimer-like elastomer based on EVA-GMA, utilising a dual dynamic metathesis mechanism based on disulfide metathesis as well as β-hydroxy ester exchange reactions. The reaction between the epoxy groups and the carboxylic acid moieties results in the formation of β-hydroxy ester linkages, which are known to promote transesterification reactions at elevated temperatures, thereby enabling network topology rearrangement and conferring vitrimer-like properties.
The reversible nature of disulfide bond exchange is well established, enabling the creation of dynamic, responsive materials capable of undergoing structural rearrangements in response to external stimuli. The synergistic effect of the β-hydroxy ester and disulfide linkages endows the crosslinked elastomer with improved mechanical as well as vitrimer-like properties. Additionally, the material exhibited fluorescence behaviour, along with excellent vitrimeric characteristics of self-healing and recyclability.
![]() | (1) |
![]() | (2) |
χ = 0.487 + 0.228VR. | (3) |
Gel fraction (%) = Wf/W0 × 100. | (4) |
1H NMR in DMSO-d6: δ (ppm) = 12.15 (–COO–), 10.12 (–CON
–), 7.69 and 7.49 (phenyl protons), 2.56 and 2.55 (methyl protons) (Fig. S1). The molecular mass obtained from MALDI-TOF mass spectrometry is 448.6 g mol−1, as correlated with the isotopic distribution calculation41 (Fig. S2). FTIR (cm−1) (Fig. S3): 1580 (NH bending of the amide group), 1715 (C
O stretching of the carboxyl group) and 3400 (OH stretching). Thermogravimetric analysis (TGA) and differential thermogravimetry (DTG) plots of the synthesized dual dynamic crosslinker revealed a distinct two-step degradation profile of –COOH and S–S groups at 230–330 °C and C–C linkages at 460 °C, as illustrated in Fig. S4.
SA-APDS acted as a dual dynamic crosslinker through its disulfide linkages, leading to a disulfide metathesis reaction, and through its COOH groups, leading to the β-hydroxy ester metathesis reaction with epoxy groups in EVA-GMA. In general, the epoxy groups are highly reactive due to their strained three-membered oxirane ring structure, which makes them more susceptible to nucleophiles like carboxylic acids. To confirm the occurrence of the epoxy acid reaction and to rule out the plausible transesterification reaction between the –COOH groups in the crosslinker and excess vinyl acetate groups in the EVA-GMA rubber, a model study was carried out by reacting the pristine elastomer with 2-furoic acid (Scheme S2). In this case, we used furoic acid as the model acid for this control study because its aromatic protons will be observed distinctly in the aromatic region for an efficient quantification study.
The opening of epoxy pendants of the EVA-GMA elastomer with 2-furoic acid catalysed by TBD, forming β-hydroxy esters, was confirmed by NMR analysis (Fig. S4). The GMA content (with epoxy pendant groups) in the EVA-GMA elastomer was found to be 2.9 wt%, displaying distinct resonances at δ (ppm) = 2.65 (H1), 2.85 (H1), and 3.21 (H2), which correspond to the protons of the epoxy group in the pristine elastomer (Fig. S4a). Fig. S4b presents the 1H NMR spectrum of EVA-GMA grafted with 2-furoic acid. After modification with 2-furoic acid, the signals corresponding to oxirane (CH2) protons at δ (ppm) 2.65, δ (ppm) 2.85 and δ (ppm) 3.21 completely disappeared. This indicates complete modification of the pendant oxirane protons through grafting with the acid group. The appearance of an OH proton at δ (ppm) = 3.64(H14) in the modified elastomer confirms the formation of hydroxyl groups due to the epoxy-acid reaction, which forms β-hydroxy esters. The methyl protons at δ (ppm) 2.07 (H6) and δ (ppm) 4.95 (H7) corresponding to the vinyl acetate groups remained intact or there is no change in their peak intensity, which confirms the non-participation of vinyl acetate groups in the transesterification reaction.
To evaluate the impact of the dual dynamic crosslinker on EVA-GMA rubber, two control samples were prepared under identical reaction conditions (Scheme 2).
The first control, E2, involved crosslinking EVA-GMA with an equimolar amount of APDS in relation to the GMA content, resulting in a dynamically crosslinked elastomer characterized by reversible disulfide linkages. For the second control, E3, AP was used for crosslinking EVA-GMA under the same conditions, thereby creating an epoxy–amine crosslinking system devoid of reversible linkages.
The correlation between the crosslinking of the pristine elastomer and the incorporation of the dual dynamic crosslinker was thoroughly investigated using FTIR spectroscopy (Fig. 1). Notably, the attenuation of the characteristic epoxy stretching vibrations at 830 and 851 cm−1 in sample E1, compared to the unmodified EVA-GMA elastomer, provides definitive evidence of epoxy ring opening and the subsequent formation of β-hydroxy ester linkages. The observed increase in absorption intensity at 1730 cm−1 is indicative of the formation of β-unsaturated ester functionalities. The C–N stretching vibration, attributed to the crosslinker, is detected at 1498 cm−1.42 Additionally, the N–H bending vibration, corresponding to the formation of amide linkages present in the dual dynamic crosslinker which was formed during the condensation reaction of APDS and SA, is observed at 1530 cm−1 in sample E1, further supporting the successful incorporation of the dual dynamic crosslinker.
Fig. 2 shows the comparative FTIR spectra for samples E1, E2, and E3, which reveal a broad absorption band at 3300–3500 cm−1 in all three samples, which corresponds to the presence of hydroxyl groups formed during the epoxy ring opening by the carboxyl group in E1 as well as due to N–H stretching. This spectral feature confirms the occurrence of transesterification reactions, as evidenced by the epoxy ring opening and the subsequent formation of β-hydroxy ester linkages. Furthermore, the spectra of samples E1, E2, and E3 exhibit N–H stretching vibrations at 3460 cm−1, associated with the presence of amine groups.
![]() | ||
Fig. 3 (a) Photoluminescence spectra of the crosslinked samples; E1 with SAPDS, E2 with APDS, and E3 with AP as the crosslinker (as shown in Scheme 2). (b) The photographic images of E1 under a UV lamp. |
In particular, sample E1, where SA-APDS was used as the crosslinker, demonstrated notable fluorescence characteristics. The benzene ring in SA-APDS attached to the amide group on both sides contributed to the extended conjugation in the crosslinker. Fig. 3a illustrates the photoluminescence spectra of the crosslinked samples. In contrast to E1, samples E2 and E3 do not have amide groups, which led to the lack of extended conjugation and hence the absence of fluorescence behaviour. This reduction in extended conjugation resulted in a less stabilised electronic structure, which was less favourable for fluorescence properties. The photographic image of E1 under the irradiation of a UV lamp is shown in Fig. 3b.
Furthermore, time-resolved photoluminescence (TRPL) spectroscopy was employed to investigate the decay profile of the crosslinked elastomeric vitrimer. As shown in Fig. S7a, all samples displayed a sharp rise followed by an exponential decay after pulsed excitation. Notably, sample E1 exhibited the highest peak intensity and the slowest decay, indicating a longer photoluminescence lifetime. The PL decay profile of the dual dynamic vitrimer was fitted to a double exponential function43 (eqn (5)).
y = A1e−x/t1 + A1e−x/t1 + y0 | (5) |
The average lifetime of the elastomeric vitrimer was calculated to be 28.4 ns (Fig. S7b), which indicates that sample E1 is a suitable candidate for optoelectronic applications.43,44
The decomposition behaviour of the compounds was characterised through TGA and DTG analyses, as shown in Fig. 5a and b, respectively. The EVA-GMA elastomer, which is well known for its excellent resistance to thermal degradation, exhibited a two-step decomposition process. The first stage of decomposition corresponded to the degradation of the side chains, specifically the vinyl acetate (VA) and glycidyl methacrylate (GMA) groups (Tmax1, above 360 °C), and the second stage was attributed to the breakdown of the main polymer chain (–CH2–CH2–) (Tmax2, above 440 °C) (Fig. 5b). The pristine elastomer showed the most resistance to thermal degradation. The thermal properties of the crosslinked elastomers were influenced by both the chemical nature of the bonds and the chain length of the crosslinkers. Sample E3, incorporating a crosslinker with a significantly shorter chain, exhibited the highest thermal stability (Fig. 5a). Although sample E1 contains a longer crosslinker, the presence of thermally stable amide and ester linkages imparts thermal stability comparable to that of the pristine elastomer and E3.46 In contrast, sample E2, which features a medium chain length crosslinker, undergoes faster thermal degradation due to the presence of thermally labile disulfide and amine linkages.
![]() | ||
Fig. 5 (a)TGA curves of the pristine elastomer and the crosslinked elastomer; (b) their corresponding DTG plots. |
![]() | ||
Fig. 6 (a) Tensile properties of the crosslinked elastomer. (b) NINT analysis of the crosslinked samples. |
Guo et al.50 demonstrated that a PDMS-based material containing both hydrogen bonds and disulfide bonds showed high stretchability and excellent self-healing performance, attributed to the synergistic effect of covalent and non-covalent interactions. Similarly, Zhang et al.51 reported a supramolecular polymer system containing disulfide bonds, hydrogen bonds, and iron(III)-carboxylate coordination. Their study highlighted that the hierarchical energy dissipation enabled by multiple dynamic linkages, along with high crosslinking density and chain folding, contributed to enhanced mechanical performance. Sarkar et al.48 reported improved mechanical properties in a dual dynamic system of a Diels–Alder reaction and a disulfide metathesis reaction in carboxylated nitrile rubber (XNBR). Herein, the specific structure of the crosslinker enables a synergistic contribution of multiple interactions, including dynamic disulfide exchange, hydrogen bonding through the amide and –COOH groups, and covalent bonding via carboxylic acid–epoxy ring-opening reactions. The presence of the methylene spacer enhances the flexibility and chain mobility of the elastomer, contributing to improved extensibility and toughness. Specifically, the incorporation of disulfide bonds, β-hydroxyl ester linkages, and hydrogen bonds enables hierarchical energy dissipation and efficient stress relaxation under deformation.51 The presence of hydrogen bonding, introduced via polar functional groups, contributes significantly to mechanical reinforcement by increasing cohesive energy density, facilitating reversible supramolecular interactions, and promoting improved chain orientation during strain. In contrast, the control sample E2 incorporates only disulfide and amine functionalities, which limits the extent of hydrogen bonding and covalent crosslinking, leading to comparatively lower mechanical performance. Sample E3, crosslinked with a short-chain diamine, forms a rigid and brittle network due to the absence of dynamic bonds and flexible segments, resulting in poor mechanical integrity.
A depth-sensing nanoindentation experiment was conducted to evaluate the surface hardness of the film samples. The hardness values, derived from the force–displacement curves in the nanoindentation study, are summarised in Table S2. Among the samples, E1 exhibited the highest hardness and modulus values, recorded at 0.9 GPa and 7.3 MPa, respectively. This superior performance is attributed to the dynamic disulfide linkages, hydrogen bonding, covalent interactions (COOH–epoxy), and a flexible methylene spacer in the crosslinker, which contributed to the material's enhanced rigidity and strength. Sample E3 also demonstrated comparable hardness, although it had lower overall mechanical properties, suggesting that it possessed significant surface hardness due to the tough and brittle nature caused by the short-chain length amine crosslinker. The greater contact depth observed in sample E2, followed by E3 and then E1, further supports the hardness data obtained from the nanoindentation (NINT) analysis. These observations are visually supported by the load–displacement curves presented in Fig. 6b.
The crosslink density in moles per cubic centimetre of the samples was measured using the equilibrium swelling method in toluene and was determined using the Flory–Rehner equation. The details and calculated data are provided in Fig. S6 and Table S3 in the SI. The crosslink density of E3 (1.45 × 10−3 mol cm−3) is higher than that of E1, primarily due to the influence of the crosslinker chain length. In this case, the control sample E3 was crosslinked using AP, which has a significantly shorter chain length compared to the dual dynamic crosslinker (SA-APDS) used in E1. Generally, crosslink density refers to the number of crosslinks per unit volume, and it influences key material properties such as rigidity, thermal stability, and elongation at break.46 A shorter crosslinker typically leads to a higher crosslink density. Accordingly, E3 exhibits a higher crosslink density. However, this increased density results in a more rigid and brittle network, leading to reduced elongation at break and lower tensile strength.
Sol–gel analysis was carried out to obtain the gel fractions of the crosslinked vitrimers (Table S4). The high gel fractions observed for all samples confirm the successful formation of a densely crosslinked network. This also indicates that residual DMF, if any, did not significantly interfere with network integrity or the completeness of the crosslinking reaction.
Stress relaxation is a typical feature of vitrimeric materials, as they can easily release deformation at higher temperatures.54–56 The stress relaxation behaviour of the vitrimers can be explained using the well-known Maxwell model for viscoelastic fluids.21 The model is expressed in eqn (6).
![]() | (6) |
To explain this behaviour of the dual dynamic vitrimer (E1), the stress relaxation experiments were conducted at different temperatures and the decay of stress over time was recorded (Fig. 7a.). At higher temperatures, dynamic covalent bonds in vitrimers undergo rapid exchange reactions, leading to faster stress relaxation over time, and at lower temperatures, the bond exchange reactions slow down, resulting in a reduced rate of stress relaxation. The relaxation time (τ) corresponding to various temperatures is recorded in Table S5.
![]() | ||
Fig. 7 (a) Normalised stress relaxation plots of the dual dynamic vitrimer. (b) Arrhenius plot of ln(τ) versus 1000/T of E1. |
The dual dynamic vitrimer (E1) exhibits faster stress relaxation (∼38 s) at higher temperatures. It implies an effective rearrangement of the network topology and faster exchange reactions of the dynamic covalent networks. Sample E2 with a single dynamic crosslink (disulfide bonds) also exhibited a faster stress relaxation time at higher temperatures but showed a longer relaxation time than E1. The stress relaxation plots of E2 at various temperatures are provided in Fig. S8. In vitrimers, temperature dependency of the relaxation time is generally expressed in terms of Arrhenius plots (Fig. 7b).57 The relaxation time follows an Arrhenius-type behaviour given by eqn (7).
![]() | (7) |
The dual dynamic vitrimer showed Arrhenius-like gradual viscosity variations, as observed in vitreous silica.23 Therefore, stress relaxation was measured at different temperatures to characterize the transesterification reaction. The relaxation time (τ) demonstrated a temperature-dependent behaviour consistent with the Arrhenius law and was accurately described by the Arrhenius equation. The activation energy (Ea) of the dual dynamic vitrimer was calculated to be 46.8 kJ from the Arrhenius plot. Typically, the Ea of epoxy-acid vitrimers ranges from 30 to 165 kJ; moreover, the presence of the phenyl ring followed by amide linkages in the system enhances the dissociation of disulfide bonds, hence lowering the activation energy.58–60 Also, the vitrification temperature was estimated to be 83 °C by extrapolating the Arrhenius plot, where the relaxation time reaches 106 seconds.61 The calculations for the estimation of Ea and Tv are provided in the SI.
![]() | ||
Fig. 8 Dynamic bond recovery test for the crosslinked vitrimers: (a) E1 with SA-APDS; (b) E2 with APDS; (c) E3 with AP as the crosslinker. |
In conventional sulphur- or peroxide-cured samples, polymer chains are chemically crosslinked via irreversible networks, which restrict their mobility. When subjected to strain, the network deforms, but the crosslinks prevent extensive chain slippage or reorganization. Stress relaxation occurs primarily through mechanisms such as recoiling of chain entanglements and segmental motion. Over time, the stress decreases as these mechanisms allow the polymer chains to rearrange into a more relaxed state. This process is relatively slow and often exhibits a characteristic viscoelastic behaviour, where stress relaxes gradually over time. Unlike traditional crosslinked elastomers, vitrimers possess dynamic networks in which the covalent bonds can break and reform reversibly under appropriate conditions. When subjected to strain in an environment where vitrimer chemistry is activated, the dynamic crosslinks in vitrimers allow for more significant rearrangements of the polymer network compared to conventional elastomers. This dynamic rearrangement enables faster stress relaxation in vitrimers compared to normal elastomers, as the network can adapt and reorganize more readily. The stress relaxation in conventional crosslinked elastomers relies on the mobility of polymer chains within a static network, while vitrimers utilize dynamic covalent bonds to facilitate faster and more extensive rearrangements of their polymer networks, leading to quicker stress relaxation and potentially enhanced mechanical properties. To investigate the dynamic behaviour of vitrimeric crosslinks, a strain-controlled experiment was conducted. In this experiment, strain was initially applied to the sample at room temperature. Subsequently, the temperature of the sample was increased to a range where the dynamic bonds are active. At this elevated temperature, the stress in the dynamic bonds was expected to be alleviated, as the new bonds were formed in the strained position. In the following step, the temperature of the sample was decreased back to room temperature while maintaining the applied strain. At room temperature, the dynamic bonds were no longer dynamic, and they remained oriented according to the strained position. To eliminate the strain at room temperature, a negative stress needed to be applied, which reflects the load supported by the dynamic bonds or the quantity of dynamic bonds present.
In physical terms, at elevated temperatures, dynamic bonds break and reform in the direction of the applied strain, and upon cooling, these newly formed bonds lock into place such that, when the strain is released at room temperature, a restoring shear stress is generated that is proportional to the number of dynamic bonds reformed in the strained configuration.
Herein, the stress measured immediately after the application of strain before the application of temperature can be denoted as σinitial. And the stress measured after cooling and strain release can be written as σrecovered. This stress denotes the residual stress that was created at an elevated temperature and locked in after the cooling.
Ideally, in the absence of dynamic bonds in the system,
σrecovered = 0 | (8) |
And if all the strained bonds had undergone topological network rearrangement,
σrecovered = σinitial | (9) |
Hence,
![]() | (10) |
This relationship can be expressed using eqn (9).
![]() | (11) |
The calculation of the percentage of dynamic bonds from the time versus shear stress, shear strain and temperature plots is tabulated in Table S6.
Based on eqn (11), the dual dynamic vitrimer E1 demonstrated the presence of approximately 62% dynamic bonds. In comparison, E2 exhibited about 20% dynamic bonds, while E3 showed only 8%. Notably, complete stress relaxation was observed for samples E1 and E2, as indicated by the stress falling to zero on the Y-axis. This suggests that both E1 and E2 experienced full stress relaxation due to their dynamic linkages. On the other hand, sample E3 did not exhibit complete stress relaxation, which can be attributed to the absence of dynamic linkages in this sample.
The healing efficiency was calculated by comparing the tensile strength of the samples before and after the healing process, using the following formula (eqn (12)).
![]() | (12) |
The stress–strain data of the crosslinked and healed samples are presented in Fig. 9a, and the healing efficiency for each sample is summarized in Table S7. The results demonstrated that sample E1 exhibited good self-healing properties, achieving a healing efficiency of up to 80%. Sample E2 also showed a commendable self-healing efficiency of approximately 75%. However, sample E3 did not exhibit any self-healing behaviour, which can be attributed to the absence of any dynamic bonds. The optical microscopy images confirmed significant healing in sample E1 facilitated by transesterification and disulfide metathesis reactions (Fig. 9c).
The dual dynamic vitrimer material, designated as E1, exhibited impressive recycling efficiency. Remarkably, the material retained up to 84% of its initial tensile strength after the first recycling process and 63% of its original strength after undergoing the second recycling cycle. This retention of mechanical properties underscores the material's durability and resilience, even after multiple recycling processes. The stress–strain plots for the sample following the first and second recycling iterations are presented in Fig. 10, clearly illustrating the material's ability to maintain a significant portion of its mechanical integrity despite repeated processing.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5lp00127g.
This journal is © The Royal Society of Chemistry 2025 |