Open Access Article
Loc Tan
Nguyen
a,
Adithya
Tharayil
a,
Nezha
Badi
a,
Johan M.
Winne
*b and
Filip E.
Du Prez
*a
aPolymer Chemistry Research Group, Centre of Macromolecular Chemistry (CMaC), Department of Organic and Macromolecular Chemistry, Faculty of Sciences, Ghent University, Krijgslaan 291 S4, 9000 Ghent, Belgium. E-mail: Filip.DuPrez@UGent.be
bLaboratory of Organic Synthesis, Department of Organic and Macromolecular Chemistry, Faculty of Sciences, Ghent University, Krijgslaan 291 S4, 9000 Ghent, Belgium. E-mail: Johan.Winne@UGent.be
First published on 5th November 2025
The growing environmental challenge of non-recyclable thermosets underscores the urgent need for sustainable alternatives. Covalent adaptable networks (CANs) containing dynamic β-amino amide moieties have emerged as promising reprocessable polymer networks, combining mechanical robustness with recyclability. In this work, we elucidate the exchange kinetics of both the well-established (retro)-aza-Michael addition and a newly identified transamidation pathway that is operative in β-amino amides. Systematic catalyst screening reveals that acidic catalysts significantly enhance viscoelastic control, thereby improving (re)processing efficiency. Furthermore, we introduce a chemical recycling protocol that enables the recovery of the original amino building blocks with up to 86% purity, demonstrating their direct use as feedstock in material (re)synthesis. These insights advance the fundamental understanding of dynamic bond exchange in β-amino amide based CANs and establish a viable route towards circular thermoset materials for numerous applications.
Covalent adaptable networks (CANs), also referred to as dynamic covalent polymer networks, have emerged in recent decades as a promising approach to reconcile the durability of thermosets with reprocessability.5–7 By incorporating exchangeable (or dynamic) linkages, CANs facilitate reprocessability of thermosets under specific stimuli such as heat or light, while (ideally) retaining the superior properties of polymeric networks.6,8 A wide range of dynamic chemistry platforms have been introduced in CANs, including transesterifications,9,10 and vinylogous urethane,11,12 Diels–Alder,13,14 dithioacetal,15 and imine exchange,16,17 a few of which have progressed toward commercial products.18,19 Despite this progress, their practical adoption is still constrained due to their dynamicity–performance trade-off.20,21 While dynamic bonds endow CANs with recyclability, they are typically labile (low activation energies – Ea), which can lead to premature bond exchange (i.e., flow or creep) under reaction conditions.22 As a consequence, this phenomenon leads to an undesirable loss in the designated performance of CANs, thus often making them unsuitable for replacing conventional thermosets. Addressing this long-standing limitation requires the development of suitable dynamic covalent chemistry (DCC) and/or technologies that combine robustness with controlled exchange reactivity.
Polyamides, long recognized for their outstanding thermal and chemical stability, are particularly attractive for advancing CAN performance.23 However, this inherent stability of amide bonds also implies their relative inertness towards exchange reactions, resulting in their limited exploration as a DCC-platform in CANs to date. In 2021, one of the first amide-based DCCs for CAN was reported by us and others through a dicarboxamide-imide equilibrium, governed by the imide ring size as an entropic factor,24,25 facilitating reprocessing of materials through dissociative transamidation at elevated temperatures (>150 °C) with high Ea ranging from 120 to 200 kJ mol−1. In the meantime, this DCC has shown its industrial potential in providing both creep resistance and recyclability, applicable for a wide range of applications.26 Subsequently, Xu et al. introduced dynamic maleic acid–tertiary amide linkages, where reversible amidation between maleic anhydrides and secondary amines enabled efficient monomer recovery (>94% yield) and closed-loop chemical recycling of cross-linked polymeric materials.27 Although both platforms demonstrate the feasibility of recycling polyamide networks, their reliance on specific monomers (e.g., glutarate ester or secondary amine building blocks) may restrict broader applicability.
Building on our earlier investigation of dynamic β-amino esters (BAEs),28 we recently developed reversible β-amino amides (BAAs) as the next-generation platform for recyclable polyamide networks (Fig. 1A).29 BAA-based CANs (BAANs) and their epoxy-derived materials exhibit excellent creep resistance (up to 120 °C) and hydrolytic stability under both acidic and basic conditions, while maintaining reprocessability.29,30 Our initial study showed that BAAs undergo reversible dissociation into acrylamides and amines above 140 °C. However, their exchange kinetics and the effect of catalysis have not been studied in detail so far. Notably, in the related BAE networks, neighboring amino groups act as internal catalysts that also promote transesterification in the presence of hydroxyl groups in addition to retro-Michael reactions,28,31 raising the question whether an analogous transamidation process could occur in BAA-promoted DCC.
![]() | ||
| Fig. 1 (A) Schematic representation of a previous study on materials including dynamic covalent β-amino amide chemistry.29 (B) Representation of the present work: investigating transamidation, catalytic effects, and chemical recycling of CANs based on β-amino amide groups. | ||
The current study started with an explicit study of the reversible Michael addition kinetics, further investigating the dynamic exchange, influence of catalysts, and the chemical recyclability of BAANs (Fig. 1B). Initially, small-molecule model studies were conducted to elucidate exchange pathways and their kinetics. Then, catalysts of different chemical nature – ranging from acidic to basic nature – were introduced into BAANs, followed by rheological analysis to assess their effect on the viscoelasticity behavior. Based on the newly acquired insights, we also investigated chemical degradation strategies to recover original amino building blocks from the BAANs as recycled feedstock resources. These studies provide mechanistic insights into BAA-based DCC, including the experimental verification of an unusual transamidation pathway, and further identify BAANs as high-performance polyamide networks with the potential for recycling and reprocessing.
Kinetic studies were conducted at varying temperatures from 120 to 180 °C, enabling the determination of reaction constants at each temperature (Fig. S5 and Table S1). Based on these data, an Arrhenius plot was constructed, revealing an activation energy (Ea) of 61.5 kJ mol−1, consistent with reported values for (most) reversible Michael adducts ranging from 50 to 80 kJ mol−1.28,32,33
Subsequently, we uncovered an unsuspected transamidation pathway operating in these reactions by repeating the same exchange experiment with BAA M and an excess of octylamine (10 eq.). Interestingly, alongside the expected (reversible) aza-Michael exchange product, analysis of the reaction mixture by 1H-NMR spectroscopy also revealed new resonances that are characteristic of a transamidation reaction: benzylic proton B at 4.4 ppm (from M) is indeed being converted into methylene proton T (at 3.3 ppm) belonging to the trans-amidated adduct TA (Fig. 2B). The formation of TA was observed at temperatures above 140 °C, and thus showed similar kinetics to the (retro)-aza-Michael pathway. It is worth noting that the asymmetric structure of the model compound M – with aliphatic and benzylic substituents – allowed for the exclusive monitoring of transamidation by 1H-NMR, since the possible (retro)-aza-Michael product (i.e., N-benzyl-3-(octylamino)propanamide) would exhibit identical proton resonances. The formation of the transamidation product TA was additionally confirmed by ESI-MS (Fig. S6). From the integration in the NMR measurements, Arrhenius plots (Fig. S7 and Table S2) revealed an Ea of 47.4 kJ mol−1, which is a bit higher than the one reported for the transesterifications of BAE (35 kJ mol−1).28
Unlike transesterification, such direct transamidation is more challenging and typically requires catalysts, or high temperatures, and is thus unexpected in these systems.34–38 Herein, we propose two possible reasonable mechanistic pathways, which are presented in Fig. 3 and are both related to a neighbouring group effect of the β-amine functionality.
![]() | ||
| Fig. 3 Proposed mechanism of the exchange of β-amino amide: (A) (retro)-aza-Michael addition and (B) transamidation via two possible pathways. | ||
While the dissociation of BAA is linked to the reversible nature of Michael nucleophilic addition via the formation of a zwitterionic intermediate after a proton transfer (Fig. 3A),39 the transamidation can possibly be enhanced by intramolecular hydrogen bonding between a protonated β-amino group and the amide carbonyl (Fig. 3B-top). Although direct transamidation begins with a high-enthalpy nucleophilic attack of an amine on the carbonyl carbon,40 the resulting tetrahedral intermediate is stabilized by an intramolecular hydrogen bond and thus a general acid catalysis, thereby lowering the energy barrier. In β-amino amides, such intramolecular interactions form an entropically favorable, six-membered pseudo-ring via hydrogen bonding,41 potentially accelerating the transamidation exchange. Interestingly, a benchmark exchange reaction performed on an amide lacking this β-amino group (i.e., n-octyloctanamide) also showed a minor conversion to the trans-amidated product after 2 h at 160 °C (Fig. S8). This observation suggests that the effect of hydrogen bonding by a protonated amine moiety, i.e., the general acid catalysis ‘charge relay’ type acylation mechanism between amide and an additional (partially) protonated amine group, can also facilitate transamidation at elevated temperature. This effect is further enhanced by the proximity of neighboring amino groups in β-amino amides (vide supra).
Alternatively, in a scenario that cannot be definitely excluded at this time, the transamidation pathway can possibly also proceed via the formation of a zwitterionic lactam intermediate (Fig. 3B-bottom). While such lactamization is expected to show a high enthalpic barrier due to the involved ring strain, the intramolecular nucleophilic attack of the β-amino group on a carbonyl carbon could be accelerated through a favorable entropic factor, also known as the neighboring group participation effect.42–45 The low activation enthalpy seen in the kinetic plots (Fig. 2B) is thus actually not indicative of such a reversible ring formation scenario, as is the already mentioned slow exchange in simple amides. In an attempt to further probe our mechanistic rationales, we performed studies to force the formation of β-lactam intermediates, but these transient species could not be isolated or observed due to their unstable (short-lived) nature in our various attempts.46
Overall, these model studies reveal that the dynamic behavior of BAANs not only arises from the reversible nature of BAA but also from a significant, kinetically comparable transamidation in the presence of β-amino groups. Although free amines are absent in our designated BAA networks, they could be released during BAA dissociation at elevated temperatures and potentially participate in transamidation exchange (vide supra).
In general, the addition of catalyst(s) linearly shifted the Arrhenius plots to a modified relaxation time (τ), while it maintained the temperature-dependent viscoelasticity (i.e., Ea) comparable to that of the benchmark network (N-Blank) (Fig. S16A). While all acidic catalysts lower the stress relaxation time (τ), the base TBD hampers BAAN's dynamicity, resulting in slower relaxation profiles, indeed indicating a kinetic role of protonated amine intermediates. More specifically, in comparison with the catalyst-free network N-Blank with a τ of 138 s at 200 °C, the addition of 5 mol% TBD into the network resulted in a τ of 194 s (Fig. 4B, Fig. S16B). In contrast, the addition of an identical molar amount of acidic catalyst significantly accelerated stress relaxation, reducing the relaxation time by more than fourfold for the N-MSA network (τ = 29 s). It should be noted here that the most pronounced acceleration was not observed for the network incorporating the stronger acid TfOH (pKa(H2O) ≈ −14, compared to pKa(H2O) ≈ −1.9 for MSA). In fact, this observation is attributed to strong interactions between TfOH and amino moieties in the network, which are known to reduce molecular mobility and thus limit catalytic effectiveness.48
As can be observed from the dissociation pathway of BAA (Fig. 3A), the β-amino group must be protonated to complete the dissociation, although the reversion of BAA is initiated by the deprotonation at the α-carbon position. Therefore, the Brønsted acids here act as a proton source, which favors such reversion. On the other hand, the strongly basic catalyst or additive TBD will compete for protons with the amines, thereby reducing the abundance of protonated amines and hampering the exchange process. Moreover, the amines released during BAA dissociation can enable further transamidation exchange, enhancing the overall network dynamicity and stress relaxation behavior. This increased relaxation ability implies improved processability, such as reduced processing time or lower processing temperature. To examine this, the N-Blank and N-MSA networks were cut into small pieces and subjected to the same compression molding process at variable temperatures ranging from 180 to 160 °C. While both samples were fully processable at 180 °C, cut samples of N-Blank failed to heal completely after 30 min at 170 °C (Fig. 4C). In contrast, N-MSA retained its (re)processability, resulting in homogeneous remolded samples at lower temperatures (160 °C). Notably, although the presence of catalysts might compromise the thermal stability of the materials, no significant impact was observed at the catalyst loading used (5 mol%), as revealed by thermogravimetric analysis of N-MSA and N-Blank (Fig. S17). This demonstrates the enhanced (re)processability of the acid-catalysed CAN, implying beneficial energy savings.
Full dissolution of P-BAAN was observed to be significantly faster when heating was done in ethanolamine (after 2 h at 140 °C) than in octylamine (after 4 h at 140 °C), which could be attributed to the acceleration effect of the (proton-donating) hydroxyl group in ethanolamine (Fig. S20).52 Moreover, a benchmark experiment conducted with DMSO as an ‘inert’ solvent did not affect the integrity of the network under identical conditions, even after 24 h at 140 °C (Fig. S21). To isolate the recovered priamine, the degraded crude product in ethanolamine was further subjected to DCM–water extraction. Although the dissolution was visualized after 2 h, the depolymerization proceeded at a much longer time scale. Specifically, incomplete depolymerization was observed by matrix-assisted laser desorption/ionization time-of-flight analysis, which primarily detected mono- and di-Michael adducts of priamine after 3 days (Fig. S22). The purity of recovered priamine, assuming Michael adducts as impurities, was quantified by 1H-NMR spectroscopy (Fig. 6).
After 3 days, priamine was recovered with 48% purity, which gradually increased over time. After 10 days of depolymerization, a recovery purity of 86% was achieved with a yield of 84 wt% relative to the total priamine used for P-BAAN synthesis, in which a mono-Michael adduct was the main residual impurity (Fig. 5B and Fig. S23). It is recognized that such impurity may potentially influence material properties (e.g., crosslink density), especially over multiple recycling cycles. On the other hand, as a proof of concept, this recovered amino compound was subsequently used as a feedstock to (re)synthesize recycled P-BAAN (Fig. 5C and Table 1).
| Samples |
T
g
(°C) |
T
d5%
(°C) |
m
iso 180, 1 h
(%) |
Swelling degreed (%) | Soluble fractiond (%) | Modulus (MPa)e |
ν
e
(10−3 mol m−3) |
|---|---|---|---|---|---|---|---|
| a Determined from DSC analysis with a heating rate of 10 °C min−1. b TGA onset temperatures after 5% weight loss (Td5%) under an air atmosphere. c Weight loss after isothermal TGA at 200 °C for 1 hour under an air atmosphere. d Swelling degree and soluble fraction obtained from solubility tests in chloroform using at least four samples at room temperature. e Storage modulus and cross-linking density (νe) from dynamic mechanical thermal analysis. νe was calculated using the rubber elasticity equation: E′ = 3νeRT, where R is the gas constant and T is the absolute temperature at Tg + 50 °C. | |||||||
| Pristine P-BAAN | −25 | 320 | 1.4 | 477 ± 09 | 3.2 ± 2.2 | 0.286 | 0.029 |
| Recycled P-BAAN | −26 | 319 | 1.2 | 436 ± 36 | 4.4 ± 1.9 | 0.292 | 0.027 |
While a brownish coloration and slightly increased soluble fraction (attributed to residual impurities) were observed, the recycled P-BAAN showed desirable chemical characteristics comparable to pristine P-BAAN, as shown using FTIR spectroscopy analysis (Fig. S24). Moreover, the chemically recycled P-BAAN largely retained its thermal and thermomechanical properties (Fig. S25–28), highlighting the potential for the chemical recycling of BAA-derived CANs.
The NMR expertise centre (Ghent University) is also acknowledged for providing support and access to the NMR infrastructure funded by FWO (I006920N) and the Bijzonder Onderzoeksfonds (BOF.BAS.2022.0023.01).
C Metathesis Enabled Glassy Vitrimers with High Rigidity, Toughness, and Malleability, J. Am. Chem. Soc., 2024, 146(23), 16112–16118, DOI:10.1021/jacs.4c03503.| This journal is © The Royal Society of Chemistry 2025 |