Next Article in Journal
Synthesis and Characterization of PEG-b-1-Vinyl Imidazole Diblock Copolymers and Their Preliminary Evaluation for Biomedical Applications
Previous Article in Journal
Dynamic Response of EPS Foam in Packaging: Experimental Tests and Constitutive Modeling
Previous Article in Special Issue
A UV-Protective Textile Coating Based on Recycled Poly(vinyl butyral) (PVB): A New Life for a Waste Polymer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polyimine-Based Self-Healing Composites: A Review on Dynamic Covalent Thermosets for Sustainable and High-Performance Applications

School of Mechanical Engineering, Jiangsu University of Science and Technology, Zhenjiang 212000, China
*
Authors to whom correspondence should be addressed.
Polymers 2025, 17(12), 1607; https://doi.org/10.3390/polym17121607
Submission received: 5 May 2025 / Revised: 3 June 2025 / Accepted: 6 June 2025 / Published: 9 June 2025
(This article belongs to the Collection Progress in Polymer Applications)

Abstract

Polyimine-based composites have emerged as a promising class of dynamic covalent thermosets, combining high mechanical strength, thermal stability, self-healing, recyclability, and reprocessability. This review systematically summarizes recent advances in polyimine synthesis, highlighting dynamic covalent chemistry (DCC) strategies such as imine exchange and reversible Schiff base reactions. Structural customization can be achieved by incorporating reinforcing phases such as carbon nanotubes, graphene, and bio-based fibers. Advanced fabrication methods—including solution casting, hot pressing, and interfacial polymerization—enable precise integration of these components while preserving structural integrity and adaptability. Mechanical performance analysis emphasizes the interplay between dynamic bonds, interfacial engineering, and multiscale design strategies. Polyimine composites exhibit outstanding performance characteristics, including a self-healing efficiency exceeding 90%, a tensile strength reaching 96.2 MPa, and remarkable chemical recyclability. Emerging engineering applications encompass sustainable green materials, flexible electronics, energy storage devices, and flame-retardant systems. Key challenges include balancing multifunctionality, enhancing large-scale processability, and developing low-energy recycling strategies. Future efforts should focus on interfacial optimization and network adaptivity to accelerate the industrial translation of polyimine composites, advancing next-generation sustainable materials.

1. Introduction

Polyimine (PI) refers to a family of polymers featuring dynamic imine bonds (C=N), typically generated by condensation reactions between primary amines and aldehydes or ketones [1]. As highlighted by bibliometric analysis (Figure 1), the field’s evolution over the past six years—marked by high-frequency keywords like ‘polymer’, ‘chemistry’, ‘thermoset’, ‘vitrimers’, ‘composites’, ‘covalent adaptable networks (CANs)’, and ‘high performance’—demonstrates a clear trajectory from fundamental chemical exploration towards leveraging dynamic covalent chemistry (DCC) to transform PIs into reprocessable, high-performance composites. The underlying reaction mechanism, known as the Schiff base reaction, involves dehydration–condensation between nucleophilic amines and carbonyl groups, leading to the formation of reversible imine linkages [2,3]. Imine bonds exhibit reversible exchange and cleavage, enabling pH-regulated hydrolysis and reformation in aqueous solutions. These bonds further undergo dissociation or metathesis when catalyzed, as demonstrated in prior studies [4]. Such reversibility allows polyimines to act as dynamic covalent networks (CANs), endowing them with responsiveness to stimuli and the potential for self-healing, recyclability, and reprocessability. Compared to traditional thermosetting polymers such as epoxy resins, PI-based networks not only offer superior thermal and mechanical performance but also enable reversible topological reconfiguration, effectively addressing the irreparable and non-recyclable limitations of conventional thermosets [3].
Recent studies have focused on enhancing the multifunctionality of polyimines by tuning both their molecular design and composite structure. Strategies include the incorporation of organic and inorganic reinforcements, such as carbon nanotubes, graphene, and natural fibers, to significantly improve their mechanical, electrical, and thermal properties. Interfacial engineering, combined with dynamic bond architecture, has been shown to improve stress transfer efficiency within the network, further boosting overall performance. Polyimines have thus become promising candidates for applications in flexible electronics, flame-retardant coatings, and environmentally friendly adhesives. Meanwhile, diverse synthesis techniques such as solution casting, hot pressing, and interfacial polymerization have been developed, offering flexible control over morphology. The integration of bio-based monomers (e.g., lignin and vanillin) into PI systems has further advanced their sustainability, enabling the upcycling of waste into high-performance adhesives and aerogels via closed-loop regeneration processes aligned with circular economy principles.
Building upon previous developments, this review systematically summarizes the synthesis strategies, dynamic bond design, composite fabrication methods, and reinforcement optimization approaches for polyimine-based materials. It highlights the interplay between dynamic covalent chemistry and interfacial engineering in achieving superior mechanical strength, self-healing capability, recyclability, and multifunctionality. By comprehensively analyzing recent advancements across flexible electronics [5], self-healing waterproof coatings [6], and energy storage systems [7], this work identifies critical challenges such as balancing multifunctionality, enhancing large-scale processability, and optimizing low-energy recycling strategies that are driven below 80 °C. Furthermore, this review emphasizes future research directions toward interfacial optimization, dynamic network adaptivity, and sustainable manufacturing pathways, aiming to accelerate the industrial translation of polyimine composites into next-generation high-performance and eco-friendly materials.

2. Synthesis of Polyimines via Dynamic Chemistry Approaches

2.1. Dynamic Imine Chemistry

From a structural perspective, imines are classified as organic compounds in which the oxygen atom of a carbonyl group (aldehyde or ketone) is substituted by a nitrogen atom [8], typically following the general formula RC=NR’, where R and R’ represent hydrocarbon groups or hydrogen atoms, as illustrated in Figure 2 [9,10]. Imine chemistry, often referred to as Schiff base chemistry, involves two primary processes: the condensation/hydrolysis of imine bonds and dynamic imine exchange [11]. These reversible covalent interactions are widely employed in the fabrication of self-healing materials [12]. To illustrate, Taynton et al. [13] demonstrated that malleable polyimine networks could serve as adhesives for woven carbon fiber composites, facilitating efficient closed-loop recycling by allowing complete recovery of both fibers and adhesives [11]. By disrupting the stoichiometric balance of polyimine networks with excess free primary amines, the adhesive dissolves, achieving 100% recycling of carbon fiber composites without additional chemicals or energy consumption. The reversible nature of imine bonds (C=N) not only imparts thermoplasticity to polyimines but also confers self-healing capabilities, allowing materials to exhibit excellent machinability and repair under heat or moisture stimulation.
While conventional imine bonds (Schiff bases) formed from ammonia/primary amines and aldehydes/ketones exhibit poor stability and water sensitivity, limiting their application scope, dynamic covalent chemistry (DCC) provides a robust alternative for synthesizing polyimines [4]. The dynamic chemical synthesis method for polyimine is illustrated in Figure 3. Through the integration of alkyne metathesis with imine chemistry, Zhang et al. [14] successfully fabricated conjugated polymers, porous frameworks, and stretchable covalent network polymers. Covalent adaptive networks (CANs), defined as polymers crosslinked via reversible covalent bonds [15], merge the advantages of thermosets (e.g., dimensional stability, mechanical strength) [16] and thermoplastics (reprocessability) [17].
The incorporation of dynamic covalent bonds has positioned polyimine (PI) as a high-performance smart material. Through reversible bond breaking and reforming, dynamic covalent chemistry enables materials to respond adaptively to environmental triggers, including temperature, pH, and light [12,18]. This property imparts self-healing, shape memory, and other adaptive functions, broadening PI’s application potential. As shown in Figure 4, water can trigger the reversible transition of dynamic covalent bonds. Heat-driven extensibility arises from imine bond exchange reactions, initiated at elevated temperatures (e.g., 80 °C), which relax internal stresses and enhance ductility. Integrating dynamic bonds with bio-based monomers enables the development of sustainable thermosets, reducing reliance on petroleum resources [3]. Taynton et al. [12] reported a catalyst-free, water-driven recyclable polyimine network, achieving green, energy-efficient processing. This material behaves as a thermoset at room temperature and is ideal for load-bearing applications and DIY prototyping. With the continuous in-depth research on the recyclability of polyimines, Wang et al. [19] proposed an innovative ‘aerogel-sol-aerogel (ASA)’ approach to recycle, repair, and reprogram polyimine aerogels. This method decomposes aerogels into oligomer solutions and reconstructs them into new aerogels, offering high repeatability and selectivity for hybrid waste streams.

2.2. The Methods for Adapting Polyimine to Different Molding States

The selection of polyimine fabrication strategies is critically governed by target product morphology and functionality requirements, as systematized in Table 1. Solution casting enables uniform dispersion for films, gels, and composites but faces solvent recovery challenges; hot pressing offers scalability for bulk materials and multilayers, yet risks interfacial weakness; interfacial polymerization achieves ultrathin 2D films and functional coatings, though it requires precise reaction control. This methodological versatility allows precise tuning of structural integrity, interfacial engineering, and dynamic adaptability across diverse applications.

2.2.1. Solution Casting

The selection of polyimine fabrication methods requires optimization based on the target product morphology. Solution casting leverages monomer dissolution in organic solvents to prepare polyimine solutions and derived materials. Precise control over solution concentration and reaction parameters enables directional construction of polymer chain architectures. Lei’s team [26] exemplifies this general solution casting process (Figure 5) to produce robust, transparent polyimine films (~0.15 mm thick). They dissolved diamines (ADH or ODA) and triamine (TREN) with TPA in DMF (aldehyde/amine ratio = 1:1), varied the diamine/triamine ratio, then heated (80 °C, 10 min) to induce gelation via Schiff base reactions, followed by vacuum drying.
It is very common to prepare polyimine cross-linked networks by solution casting method. For example, a mixture of lysine-based amino acid ionic liquids (AAILs), tris(2-aminoethyl)amine (TREN), terephthalaldehyde (TPA), and polyvinylpyrrolidone (PVP), was dissolved in ethanol [27]. Schiff base reactions between amino and aldehyde groups produced imine bonds, forming an IPIN-PVP solution, which upon ethanol evaporation, yielded fluorescent composite films [27] (as shown in Figure 6a). For instance, Tian et al. [28] synthesized polyimine thermosets by dissolving monomers in chloroform, followed by solvent evaporation and curing. Similarly, Yu et al. [29] synthesized semi-aromatic polyimine matrices using 3,4′-ODA and TA monomers, with aliphatic TREN as a crosslinker, and formed composite films via solvent evaporation.

2.2.2. Hot Pressing

Hot pressing combines heat and pressure to induce plastic deformation or chemical crosslinking, enabling efficient fabrication of dense polyimine composites. Generally, researchers first grind the prepared solid material into powder, then place it into a hot press for thermal compression (as shown in Figure 6b). Su et al. [30] prepared Cel-PI films by hot pressing moistened Cel-PI powder between polyimide membranes at 70 °C and 30 MPa for 20 min. Zhang et al. [31] achieved homogeneous wood flour/polyimine composites by blending powders via planetary ball milling and hot pressing at 80 °C and 9 MPa. This method ensures smooth surfaces, strong interfacial adhesion, and enhanced mechanical performance. Similarly, Zhang et al. [32] blended varying amounts of wood flour with bio-based polyimine vitrimers to prepare lignocellulose-based composite films. The mixture was hot-pressed at 120 °C and 10 MPa for 5 min [32]. This method produced polyimine composites with smooth, uniform, and defect-free surfaces. Additionally, it ensured strong interfacial compatibility between reinforcements and the matrix, thereby enhancing overall material performance.
Figure 6. (a) Solution casting of IPIN-PVP membranes [27]; (b) hot pressing of BP/PI composites [33]; (c) SMAIS of PI-2DP films at air/water interfaces: (i) create surfactant monolayer on water surface. (ii) Arrange monomer 1,2 underneath surfactant monolayer. (iii) Carry out polymerization at interface [34].
Figure 6. (a) Solution casting of IPIN-PVP membranes [27]; (b) hot pressing of BP/PI composites [33]; (c) SMAIS of PI-2DP films at air/water interfaces: (i) create surfactant monolayer on water surface. (ii) Arrange monomer 1,2 underneath surfactant monolayer. (iii) Carry out polymerization at interface [34].
Polymers 17 01607 g006

2.2.3. Interfacial Polymerization

Interfacial polymerization occurs at immiscible liquid interfaces (e.g., water–oil) to form stable interfacial layers, improving mechanical and thermal properties. Dai et al. [35] pioneered air/water interfacial synthesis of covalent monolayers using dynamic chemistry, opening new pathways for the synthesis of two-dimensional polymers. Zhang et al. [36] fabricated highly crystalline polyimine films by confining protonated TAPP monomers and DhTPA at water surfaces. Ge et al. [37] developed amorphous 3D-PPIn nanofilms via interfacial oligomer splicing, achieving rapid molecular/ion differentiation. This method is also pivotal in water treatment. Baig et al. [38] synthesized thin-film composite (TFC) membranes through interfacial reactions of aromatic diamines and acyl chlorides. Interfacial polymerization also overcomes the challenge of integrating insoluble covalent organic framework (COF) powders into thin-film electronics. By developing the surfactant monolayer-assisted interfacial synthesis (SMAIS) strategy, Sahabudeen et al. [34] realized the fabrication of PI-2DP films at the air/water interface, leading to the first successful synthesis of 2D COF films (as shown in Figure 6c).

3. Preparation of Polyimine Composites

3.1. Selection of Reinforcements

The choice of reinforcements significantly influences the performance of polyimine composites. Based on their properties, reinforcements can be categorized into organic and inorganic phases. As shown in Table 2, organic reinforcements mainly enhance toughness and processability, while inorganic phases significantly improve strength, conductivity, and thermal stability. Their effects stem from different mechanisms—hydrogen bonding in organics versus interfacial load transfer in inorganics—highlighting the need for application-oriented selection.

3.1.1. Organic Reinforcements

Organic reinforcements, such as polymers and natural fibers, enhance the mechanical properties and processability of composites through their inherent flexibility and toughness. Organic reinforcements play a critical role in polyimine composites. While polyimines inherently exhibit excellent thermal stability and mechanical strength, their processability and toughness remain limited. For example, Zhang et al. [42] synthesized self-healing polyimine networks using peach gum polysaccharide (PGP) and chitosan (CS), achieving a tensile strength of 56.5 MPa and 97.8% healing efficiency. Su et al. [43] filled the pores of cellulose paper with polyimine, functionalized cellulose paper with polyimine covalent adaptive networks (CANs), creating paper–polyimine composites (PPCs) with high mechanical strength (71 MPa), water resistance, and full recyclability. In recent years, supramolecular hydrogen bonding motifs have demonstrated unique advantages in regulating network architectures. Ding et al. [44] synthesized a bio-based polyimine (TMP-IPDA-Si) incorporating acylhydrazide structures using lignin-derived vanillin as a precursor. Molecular dynamics simulations revealed that the introduction of acylhydrazone bonds substantially enhanced intermolecular hydrogen bonding interactions (with a hydrogen bond density of 4 nm−3), endowing the material with an exceptional tensile strength of 85.7 MPa and a high glass transition temperature of 201.4 °C. Additionally, E. Lamm et al. [39] developed polyimine-coated cellulose nanofibrils (CNF-imine) via aqueous synthesis, improving interfacial compatibility in polymer matrices.

3.1.2. Inorganic Reinforcements

The incorporation of inorganic reinforcements such as carbon nanotubes (CNTs), graphene, and metal oxides enhances the thermal stability, electrical conductivity, and mechanical strength of the composites in a synergistic manner. Single-walled carbon nanotubes (SWNTs) [45] and silicon nanoparticles (SiNPs) [7] not only optimize the tensile modulus of composites but also impart unique conductive properties. Magnetic NdFeB microparticles enable directional control over magnetic performance [46]. ZrO2 increased the tensile strength and toughness of polyimine nanocomposites by 40% and 85%, respectively [47]. Furthermore, Mao et al. [48] developed multifunctional composites by incorporating novel inorganic phases (e.g., TiO2@MXene heterostructures and sodium vanadium oxide, NVO) into polyimine matrices, achieving both flame retardancy and electrochemical responsiveness. In practical applications, Yu et al. [18] incorporated carbon fiber sheets into a dynamic covalent fluorinated polyimine matrix, resulting in composites that could be welded and reshaped at 80 °C, with water-driven extensibility and reprocessability. In 2022, Wang et al. [49] fabricated bio-based degradable polyimine vitrimers/carbon fiber (CF) composites. While the mechanical properties of CF-reinforced IPDA vitrimers were marginally lower than those of bisphenol A epoxy (E51), polyimine vitrimers exhibited easy degradability in acidic environments and effectively preserved both the surface morphology and chemical structure of recycled CF, promoting the advancement of eco-friendly CF composites [49]. Despite their advantages (e.g., thermal stability and mechanical strength), poor compatibility between inorganic reinforcements and the matrix may lead to uneven dispersion, compromising final performance. Thus, improving interfacial compatibility is a critical research focus. For instance, Zhang et al. [50] modified silicon carbide whiskers (SiCw) and graphene oxide (GO) with 3-aminopropyltrimethoxysilane (APTMS), achieving superior dispersion and enhanced interfacial interactions. Their experiments also confirmed the significant impact of reinforcement content on mechanical properties.

3.2. Composite Methods

The diversity of preparation methods for polyimine composites provides extensive possibilities for optimizing material performance. Common approaches include physical blending, chemical bonding, and lamination, each tailored to specific application scenarios and performance requirements.

3.2.1. Physical Blending

Physical blending involves mixing reinforcements with a polyimine matrix via mechanical or solution-based methods. This method is relatively simple to operate and suitable for large-scale production. Cui et al. [33] prepared BP/PI composites through the physical blending of bamboo powder (BP) with polyimine precursors, followed by hot pressing, subsequent to Schiff base crosslinking. The resulting composites exhibited high tensile strength (45.2 MPa), bending strength (121.9 MPa), and solvent resistance, alongside closed-loop recyclability and biodegradability (see Figure 6b). Zhu et al. [46] further incorporated NdFeB microparticles into polyimine solutions to create magnetic soft robots with full chemical recyclability and room temperature self-healing capabilities. During the mixing process, ultrasonic treatment or mechanical stirring is typically employed to ensure uniform dispersion of the magnetic particles in the polyimine solution. While simple and cost-effective, physical blending often yields weaker interfacial bonding compared to chemical methods.

3.2.2. Chemical Bonding

Chemical bonding enhances interfacial strength through covalent interactions between reinforcements and the matrix, thereby breaking through the limitations of traditional physical compounding. Zhu et al. [51] developed superhydrophobic coatings by spraying fluorinated epoxy monomers and Fe3O4 SiO2-NH2 nanoparticles onto polyimine films, where epoxy–amine reactions ensured robust adhesion. Wang et al. [52] synthesized Fe3O4 TAPB-Tp core–shell nanocomposites via interfacial Schiff base polymerization, achieving strong covalent linkages between magnetic cores and polyimine shells. TAPB-Tp acts as the reinforcing phase chemically bonded to the surface of Fe3O4, thereby achieving the composite of the substrate and the reinforcing phase. However, the reaction process of this method is relatively complex, requiring precise control of reaction conditions to ensure the quality and performance of the final product.

3.2.3. Lamination

Lamination refers to a composite manufacturing process where distinct material layers, such as fibers, resins, and metals, are stacked using customized techniques to impart integrated properties. Li et al. [53] developed high-performance, recyclable ramie yarn-reinforced polyimine vitrimer composites (RY-PI) by repeatedly hot pressing single-layer RY-PI sheets. Dynamic bond reorganization enabled seamless interlayer integration. Upon completion of polyimine synthesis and carbon fiber (CF) surface modification, Zhang et al. [54] dried the CF fabrics and polyimine (PI) films, then alternately stacked 16 layers of CF fabrics and 17 layers of PI films. The stacked layers were hot-pressed at 100 °C under 5.0 MPa for 10 h, followed by cooling to room temperature, resulting in CF/PI laminated plates (as shown in Figure 7). This approach combines the mechanical advantages of natural fibers (e.g., strength, lightweight) with the reversible nature of dynamically crosslinked polymers, achieving high performance and self-healing capabilities. Innovative chemical/physical recycling strategies further enhance sustainability, positioning lamination as a key technology for replacing traditional plastics and advancing zero-waste goals.

4. Factors Affecting the Mechanical Properties of the Composites

The mechanical properties of polyimines and their composites are crucial for engineering applications, prompting extensive research efforts aimed at performance enhancement. A summary of mechanical properties of polyimine materials is provided in Table 3. Numerous studies have explored strategies to optimize network design and reinforcement compatibility. For example, Wang et al. [55] developed fluorinated bio-based polyimine films with a tensile strength of 96.2 MPa, surpassing that of commercial polycarbonate (~65 MPa), demonstrating the potential of tailored polyimine systems in high-performance applications.

4.1. Correlation Between Dynamic Bonds and Mechanical Properties

The mechanical behavior of polyimines is predominantly governed by the reversible nature of dynamic imine bonds (C=N). Crosslinking density further modulates mechanical performance. Li et al. [56] demonstrated that increasing crosslinking density in polyimine/epoxy networks enhanced elastic modulus and tensile strength, confirming the critical role of covalent interactions in load-bearing capacity.
Unlike conventional covalently crosslinked polymers, polyimines enable network rearrangement through dynamic bond exchange, maintaining high mechanical strength while exhibiting stress relaxation and self-healing capabilities. While traditional networks show gradual stress decay under constant strain, polyimine elastomers (PIEs) with dynamic imine bonds can completely relax stress through reversible network reorganization. Notably, Lei et al. [57] observed a unique ‘stress intensification’ phenomenon in PIEs, where stress rebounded after relaxation due to zwitterionic intermediates formed via transamination/transimination reactions (Figure 8a). This contrasts with the monotonic stress decay in conventional networks. The mechanism involves load-bearing chain rupture during deformation, followed by network reorientation through imine exchange (Figure 8c). After multiple relaxation-strengthening cycles, tensile strength significantly increased (Figure 8b), demonstrating how dynamic bonds enable topology optimization for reconfigurable polymer processing.

4.2. Interfacial Engineering and Multiscale Performance Optimization

Interfacial interactions between reinforcements and the matrix play a pivotal role in the mechanical optimization of polyimine composites. Zhang et al. [41] incorporated graphene into polyimine matrices, achieving composites with high thermal conductivity and electrical conductivity. Graphene nanoplatelets (GNPs) acted as bridges across microcracks and altered crack propagation paths, thereby enhancing fracture energy dissipation. While covalent bonding provides superior stress transfer efficiency compared to physical adsorption, non-covalent strategies—particularly hydrogen bonding and π-π stacking—offer synergistic enhancements. A representative example is the hydrogen-bonded polyimine thermoset (PI-A/T) system [26], where amide-containing ADH monomers introduce reversible hydrogen bonding sites. This dual-network design achieves an optimal balance: hydrogen bonds restrict chain mobility to enhance stiffness, while their dynamic break/reformation dissipates energy under stress, preserving toughness. Further interfacial reinforcement is demonstrated in Lyu’s vanillin-based graphene/polyimine composites [41]. Here, π-π stacking between aromatic rings (abundant in both the polyimine matrix and graphene nanosheets/GnPs) reduces free volume and promotes dense chain packing. The GnPs not only strengthen this stacking network but also suppress phonon scattering, leading to concurrent improvements in mechanical strength and thermal conductivity. Such multi-mechanism interfacial engineering exemplifies how dynamic covalent networks can integrate covalent and non-covalent interactions for tailored performance.
As illustrated in Figure 9, the two composite systems share the same polyimine matrix but differ significantly in how carbon nanotubes are incorporated. In the PI-daMINT system (Figure 9a), a specially designed U-shaped macrocycle acts as a molecular bridge that connects single-walled carbon nanotubes (SWNTs) with the polyimine network. The macrocycle wraps around SWNTs through non-covalent interactions, while its terminal groups enable chemical bonding with the imine matrix. After thermal treatment, the resulting daMINTs integrate covalently into the network, ensuring strong interfacial adhesion [45]. In contrast, the Vitrimer-MWCNT composite (Figure 9b) is formed via physical blending without chemical linkage. Although good dispersion is achieved, the lack of interfacial bonding limits mechanical performance and leads to delamination under repeated stress [40]. This comparison highlights the critical role of covalent integration in enhancing the reinforcement efficiency of nanofillers.

4.3. Impact of Self-Healing Behavior on Mechanical Performance

The self-healing behavior of polyimines, driven by the reversible reorganization of dynamic bonds (e.g., transamination of imine bonds or ionic interactions), significantly restores mechanical properties. For example, Peach Gum—Chitosan based PI networks (PGCS) retained 97.3% strength after two physical reprocessing cycles, while the ionic polyimine network (IPIN) maintained mechanical performance comparable to virgin materials after acidolysis-repolymerization closed-loop recycling [42,58]. According to experimental findings, IPIN-1 achieved a tensile strength of 4.13 MPa and an elongation of 180% after 1 h of self-healing, reaching 4.21 MPa and 241% after 12 h, with a recovery rate over 96%. Cyclic tensile tests showed a low hysteresis and 19.8% residual strain after five cycles, indicating good mechanical stability. During the preparation of IPIN-1, choline lysinate ([Ch] [Lys]) was used, where the hydroxyl group of [Ch]+ participated in hydrogen bonding with iodine. The presence of hydrogen bonds enhances intermolecular interactions, thereby improving the stability and recovery capability of the material during cycling. Additionally, the dynamic reorganization of imine bonds, coupled with electrostatic interactions, constitutes the predominant mechanism responsible for the robust self-healing capability inherent to the material (as shown in Figure 10). The healed material could lift 1000 g, demonstrating excellent load-bearing capacity.
However, current systems face limitations such as toughness reduction after multiple repairs. In the experiments conducted by Yang et al. [59], compared to the original elastomer, after the first, second, and third reprocessing cycles, the tensile strength recovered to 96.0%, 86.5%, and 80.6% of the original values, respectively; the elongation at break reached 96.2%, 89.8%, and 79.3% of the original values, respectively. They systematically studied the self-healing behavior at room temperature. The tensile strength recovered to 0.75 MPa after 6 h, 1.74 MPa after 12 h, and 2.41 MPa after 24 h, corresponding to recovery efficiencies of 30.0%, 69.3%, and 96.0%, respectively. Increasing the healing temperature to 40 °C and 60 °C improved 6 h efficiencies to 61.0% and 84.9%, respectively, confirming the thermal acceleration of dynamic bond reorganization [59]. Future strategies should focus on room temperature triggers (e.g., light/moisture response) and in situ repair technologies for complex components.
Table 3. Mechanical performance of polyimine materials (20 to 25 years).
Table 3. Mechanical performance of polyimine materials (20 to 25 years).
Polyimine
Materials
Physical FormMechanical Property AnalysisKey FindingsRef.
Tensile Strength
(MPa)
Elongation at Break (%)Young’s Modulus (GPa)Thermal Stability
(Td5%, °C)
Self-Healing Efficiency (%)Limiting Oxygen Index (LOI, %)
IPINfilm4.28250->160Self-healing efficiency: 50.04% after 5 min-Exhibits good ductility, self-healing ability, and recyclability. Electrochemical sensors fabricated from IPIN-1 demonstrate high response rates and low detection limits for iodine monitoring.[58]
PE0.5-0.2 vitrimerfilm2.511158--Self-healing at room temperature: 30.0 (6 h), 69.3 (12 h), 96.0 (24 h)-The material was synthesized through a simple two-step one-pot process at room temperature, exhibiting outstanding mechanical properties, self-healing capability, degradability, and reprocessability.[59]
CTM-3film0.56 ± 0.6638.4 ± 0.3-455.0-41.6A catalyst-free vitrimer featuring enhanced flame retardancy, thermal stability, solvent resistance, mechanical strength, and recyclability, outperforming conventional counterparts in reported studies.[60]
PIM-4film94.5 ± 2.66.1 ± 0.93.5 ± 0.2434--Exhibits low water uptake (~0.14–0.15%), excellent mechanical properties minimally affected by absorbed water, chemical resistance, and recyclability, positioning it as a promising alternative to petroleum-based thermosetting resins for harsh environments.[61]
FA-100film28.47 ± 2.017.74 ± 1.240.37 161.32-28.8Elevating D-FA content enhances the network’s crosslink density and mechanical properties. Dynamic imine linkages endow the material with reprocessability and acid-degradability. FA-100 demonstrates superior flame resistance, achieving a limiting oxygen index (LOI) of 28.8%.[62]
CO-PIM-75film62.512.9->242--High thermal stability. Tensile properties are improved by adjusting the 2,4-ODA/6FAPB ratio, achieving performance comparable to PC. Exhibits good hydrolytic and solvent resistance, with negligible deterioration in mechanical properties after recycling.[63]
Cel-PIfilm46.32.22.9---Characterized by dynamic network exchange and an amorphous structure, the material achieves excellent thermal processability, mechanical robustness, water/solvent resistance, thermal stability, and recyclability.[30]
PGCS-100film56.520.60.439227.1Up to 97.8%56.5Demonstrates outstanding mechanical properties, high thermal stability, self-healing, welding ability, shape memory, reprocessability, and chemical recyclability.[42]
PI-A1.0/T1.0film55.2721.7---Incorporating hydrogen bond crosslinking significantly enhances mechanical strength and stiffness without sacrificing ductility, offering an effective strategy for strengthening and toughening dynamic covalent thermosets.[26]
2D Polyimine Filmsfilm-6.5 ± 2.48.6 ± 2.5---In situ TEM tensile testing allows nanoscale observation of structural evolution and fracture dynamics. Crack initiation preferentially occurs along (100)/(010) directions, with chemical structure influencing mechanical failure.[64]
CO-PIM-75film62.512.9->242--High thermal stability. Tensile properties are improved by adjusting the 2,4-ODA/6FAPB ratio, achieving performance comparable to PC. Exhibits good hydrolytic and solvent resistance, with negligible deterioration in mechanical properties after recycling.[63]
Cel-PIfilm46.32.22.9---Characterized by dynamic network exchange and an amorphous structure, the material achieves excellent thermal processability, mechanical robustness, water/solvent resistance, thermal stability, and recyclability.[30]
PGCS-100film56.520.60.439227.1Up to 97.8%56.5Demonstrates outstanding mechanical properties, high thermal stability, self-healing, welding ability, shape memory, reprocessability, and chemical recyclability.[42]
PI-A1.0/T1.0film55.2721.7---Incorporating hydrogen bond crosslinking significantly enhances mechanical strength and stiffness without sacrificing ductility, offering an effective strategy for strengthening and toughening dynamic covalent thermosets.[26]

5. Engineering Applications of Polyimine and Composites

Polyimines and their composites exhibit transformative potential across diverse engineering fields, driven by their unique combination of dynamic covalent adaptability, mechanical robustness, and multifunctional integration. Figure 11 illustrates the various applications of polyimine and its composite materials. Meanwhile, in Table 4, we summarize the applications of polyimines and their composite materials in engineering over the period of 20 to 25 years.

5.1. Green Materials

As global industrialization accelerates, the demand for sustainable alternatives to traditional petroleum-based thermosets has intensified [79]. Therefore, the development of high-performance, recyclable, and biodegradable green materials has become an important research direction in the field of materials science. Polyimines address this need through closed-loop recyclability, biodegradability, and bio-based monomer utilization [80]. Biomass-reinforced recyclable polyimine composites exemplify green materials through dynamic imine chemistry; Figure 12 illustrates this recycling process. It has obvious advantages compared with other traditional materials. For example, although traditional polyimide materials possess excellent thermal and mechanical properties, their high coefficient of friction and wear rate significantly impact service life. Polyimines, through the design of dynamic covalent bonds, exhibit excellent recyclability. Moreover, the introduction of modified fillers and reinforcing phases can significantly enhance their tribological properties.
In the field of green materials, polyimine composites have driven significant advancements in sustainability through innovative applications. Su et al. [81] incorporated imines and amines into polymer networks, establishing an efficient approach for converting biomass into eco-friendly plastics characterized by excellent repairability, renewability, and closed-loop recyclability. These materials accommodate up to 70% wood-derived biomass and display mechanical properties comparable to or exceeding those of traditional plastics [81]. Xiong et al. [21] upcycled non-recyclable polystyrene (PS) into high-performance poly(styrene-imine) aerogels (PSAs) via post-functionalized amination and Schiff base reactions. These aerogels undergo quantitative depolymerization under acidic conditions, enabling infinite material regeneration. Jia et al. [82] developed high-performance bio-based polyimine materials (Bio-Si-PABZs) using a catalyst-free copolymerization strategy. By integrating rigid and flexible structures with dynamic covalent bonds, supramolecular interactions, and hydrogen bonding, these materials demonstrated exceptional properties, including high strain tolerance, tensile strength, thermal stability, self-extinguishing behavior, self-healing capability, and recyclability. Notably, waste from these materials could be repurposed into high-performance adhesives. Li et al. [83] further engineered a series of high-performance polyimine composites by incorporating bio-based monomers and degradable crosslinkers, achieving closed-loop recyclability and biodegradability. Additionally, bio-based resources such as vanillin, bamboo powder, starch, chitosan, and natural peach gum have been widely utilized for polyimine functionalization, further expanding their eco-friendly applications.
Polyimine composites demonstrate outstanding mechanical strength, self-healing, anti-creep, and solvent resistance in coatings. Reprocessability enables heat/pressure-driven restoration of damaged coatings, while self-healing autonomously repairs microcracks. Their dimensional stability under prolonged stress and chemical durability further enhances applicability. To optimize transparency and waterproofing, polydimethylsiloxane (PDMS) has been integrated into polyimine matrices, yielding materials with dense crosslinking, optical clarity, water resistance, UV protection, and self-healing. Xie et al. [6] developed self-healing waterproof coatings by condensing fractionated lignin with PDMS and oleylamine, achieving high crosslinking density and near-invisible post-repair scratches. Similarly, Wang et al. [84] synthesized hyper-crosslinked HHMOP membranes with 91.35% healing efficiency at 25 °C/60% RH within 24 h, showcasing promise for advanced coatings.

5.2. Electronic Applications

In addition to their outstanding green material properties, polyimine-based composites exhibit unique advantages in functional electronics. Their combination of electrical performance and mechanical flexibility offers new pathways for next-generation devices. By doping nanoscale conductive fillers and optimizing structure, PI composites have been developed for wearable systems, energy storage, and intelligent sensors. Their intrinsic network reconstruction and closed-loop recyclability also provide sustainable solutions for electronic waste management, aligning device lifecycles with circular economy principles.
The flexible electronics breakthroughs of PI composites stem from their structure–function synergy. This material system successfully achieves the unification of mechanical deformation adaptability and functional stability through the synergistic effect of dynamic network design and nano-conductive phases. For example, MWCNT-reinforced PI networks enable programmable reconfiguration and electrical percolation, ideal for durable wearables. He et al. [67] fabricated reconfigurable 3D sensors from PI/MWCNT inks, achieving electrothermal-driven self-healing. It is particularly noteworthy that this system exhibits electrothermal response-driven interfacial bond reconfiguration characteristics during the damage repair process. This intelligent repair mode provides an innovative solution for the full lifecycle management of electronic devices. Zhang et al. [5] fabricated monolithic PI/graphene aerogels exhibiting a compressive strength of up to 1.2 MPa, electrical conductivity of 79 S/m, and excellent durability after 3000 compression cycles, emphasizing their potential for sustainable sensing applications.
Polyimine composites have achieved notable advancements in sensor technology. Yang et al. [59] developed room temperature polyimine elastomers with mechanical robustness, self-healing, and recyclability for liquid metal (LM)-based strain sensors. The sensors were fabricated by dip-coating elastomer strips with PVP-coated liquid metal (LM) conductive ink and evaporating ethanol. They exhibited crack-free conductivity under 500% strain, stable electrical responses, and repeatable sensing across strains (Figure 13a). After complete severance, the sensors self-healed at room temperature within 24 h (Figure 13b). The elastomer substrate fully degraded in 0.1 mol/L HCl/DMAc (600 min), enabling LM recovery via centrifugation for closed-loop recycling. In flexible photodetectors, the integration of tellurium nanowires (Te NWs) and MoS2 into polyimine matrices enhanced photoresponse efficiency. Peng et al. [65] fabricated a Te NWs/MoS2/polyimine composite photodetector, achieving a rapid response time of 5 s and a specific detectivity of 1.145 × 1010 Jones under 532 nm illumination, surpassing the performance of conventional counterparts. Additionally, Qi’s team [85] designed high-performance electrochemiluminescence (ECL) sensors by bridging iridium(III)-polyimine complexes (Ir2PD) with Hf-MOFs, achieving enhanced stability and ECL emission through in situ self-assembly on ITO surfaces.
Polyimine composites exhibit remarkable output power and self-healing capabilities in high-performance energy harvesters. Zhu et al. [86] employed dynamic covalent thermosetting PI as a matrix for recyclable hotspot thermoelectric generators (TEGs), achieving post-recovery performance comparable to the original, with excellent mechanical stability under cyclic loading. Figure 14 illustrates the structure and recyclability of PI/GP-based triboelectric nanogenerators (TENGs). In Figure 14a, the composite architecture is shown, where the polyimine matrix, embedded with graphite–polypropylene (GP) powders, enhances surface roughness and dielectric properties [87,88], resulting in high output voltage during mechanical stimulation. Rajabi-Abhari et al. [66] enhanced TENG performance using such PI/GP composites, reaching a power density of 2571 mW/m2 under 15 N force at 6 Hz. Figure 14b demonstrates the dual recycling strategies: physically, the composite can be reprocessed via grinding and hot pressing; chemically, it can be depolymerized in a DCM/diamine solution and reconstituted with fresh components. These closed-loop approaches, enabled by dynamic covalent networks, highlight the material’s excellent reusability and sustainability, making it suitable for next-generation wearable and recyclable energy harvesting devices. Additionally, Zhu et al. [86] used PI to encapsulate 200 pairs of Bi2Te3 and Sb2Te3 thermoelectric legs with liquid metal electrodes in recyclable, self-healing, and stretchable TEGs (RHS-TEGs). Upon damage, the device could be fully depolymerized in a methanol-based solution, enabling easy separation and reuse of components.

5.3. Energy Storage

Polyimine composites offer significant benefits across a range of energy storage technologies, such as lithium-ion batteries, zinc-based batteries, lithium-metal batteries, and metal sulfide/carbon composite systems. Acting as a binder, PI significantly enhances the cycling stability and specific capacity of silicon anodes; as a separator material, it effectively suppresses lithium dendrite growth, thereby improving battery safety and lifespan. Moreover, PI composites achieve high nitrogen doping and three-dimensional superstructures through simple synthetic methods, further enhancing their energy storage performance.
Researchers have utilized the incorporation of nano-silicon into PI matrices as a promising approach to improve the electrochemical performance of silicon anodes. Gao et al. [7] developed a high energy density electrode system by synergistically constructing PI matrices with nano-silicon, using an in situ polymerization approach to form a three-dimensional network. This structure mitigates the volume expansion of silicon, maintaining 80.4% capacity retention after 200 cycles at 95% silicon content, demonstrating excellent structural stability. This interfacial engineering approach offers a novel pathway for advancing the development of high energy density electrodes.
In zinc battery technology, two-dimensional polyimine (2DPM) membranes serve as key interfacial coatings due to their proton-selective transport properties. With dual ion-transport nanochannels and abundant proton-conduction sites, 2DPM promotes selective proton transport and accelerates electrode reaction kinetics. By coating a NaV3O8·1.5H2O cathode (active material load 10 mg/cm2) with an 80 nm thick 2DPM membrane, Guo et al. [89] reported an areal capacity of 4.5 mAh/cm2 and an energy density of 33.8 Wh/m2. As shown in Figure 15, the 2DPM membrane enables H+-dominated intercalation by creating a highly selective interface that effectively suppresses Zn2+ crossover while facilitating fast and directional proton transport. This improves electrochemical kinetics and enhances utilization of the active material, especially in high-mass-loading electrodes. The optimized 2DPM membrane achieved a proton flux higher than 0.9 mol m−2 h−1 and a H+/Zn2+ selectivity ratio of 140.7, markedly improving the cathode/electrolyte interfacial properties.
In lithium-metal batteries, PI aerogels with hierarchical pore architectures and polar chemical environments demonstrate strong dendrite suppression effects. Ding et al. [90] reported that using PI aerogels as functional separators effectively guides uniform lithium deposition via pore structure dispersion and surface functional group interactions. As illustrated in Figure 16, the PI aerogel (PIA) separator enables more homogeneous Li+ diffusion and uniform lithium nucleation compared to conventional polypropylene (PP) separators. This helps mitigate dendrite formation and leads to more stable cycling performance. The schematic comparison highlights the critical role of separator chemistry and morphology in regulating interfacial lithium behavior. Batteries utilizing these separators achieved over twice the Coulombic efficiency compared to conventional polypropylene (PP) separators and maintained dendrite-free morphologies during cycling, offering a promising strategy for safe lithium-metal batteries.
As electrode precursors, PI-based materials provide unique advantages for synthesizing metal sulfide/carbon composites. Their molecular design flexibility enables high nitrogen doping, 3D porous architectures, and uniformly distributed nano-sized metal sulfide phases through simple pyrolysis, significantly simplifying the process route. Chen et al. [91] used a flower-like PI superstructure coordinated with metal ions to prepare composites exhibiting superior sodium storage performance. As a representative case, Co9S8/carbon composites exhibited a reversible capacity of 302 mAh g−1 after 200 cycles at 0.5 A g−1, demonstrating the validity of this strategy [92].
In the field of solid-state batteries, PI incorporation into solid-state electrolytes (SSEs) can effectively inhibit lithium dendrite formation, significantly extending Li/SSE/Li cell lifespans. However, some drawbacks have been identified. Jiang et al. [93] observed that PI particle distribution within LPS matrices can partially block ion transport pathways, reducing ionic conductivity. Smaller PI particles, while more dispersive, exacerbate this issue. Additionally, PI migration between particles over cycling further decreases conductivity.

5.4. Fireproof Materials

Traditional thermosetting materials, such as epoxy and phenolic resins, face significant challenges due to their irrecyclability, limited mechanical performance, poor thermal stability, insufficient fire resistance, environmental incompatibility, and limited dynamic capabilities. These materials typically cannot be reprocessed or recycled through heating, leading to substantial waste generation, resource inefficiency, and environmental pollution [94]. Moreover, traditional thermosets often degrade under high temperature, humidity, or chemically corrosive environments, and combust readily during fires, producing considerable smoke and toxic gases, thereby exacerbating hazards. As summarized in Table 5, traditional thermosets such as epoxy, phenolic, and BMI exhibit high tensile strength and thermal stability but lack recyclability, which limits their long-term sustainability. In contrast, polyimine-based dynamic thermosets combine competitive mechanical and thermal performance with the unique advantage of closed-loop recyclability, positioning them as promising candidates for next-generation green engineering materials.
In contrast, polyimine composites exhibit superior structural integrity at elevated temperatures and leverage reversible dynamic covalent bonds to enable self-healing and recyclability, significantly extending material lifespan and reducing environmental impact. Toldy et al. [76] revealed that polyimine vitrimer systems exhibit superior thermal stability relative to epoxy resins. Flame retardants exhibit more significant reduction in pHRR, with vitreous composites demonstrating superior performance in relative pHRR reduction compared to epoxy composites (as shown in Table 6). Degradation in thermosetting resins is often facilitated by integrating dynamic covalent bonds. Zhang et al. [95] developed a catalyst-free recyclable thermosetting resin (AFD/DCNSA) integrating dual dynamic covalent bonds: pH-sensitive disulfide bonds (-S-S-), which undergo reductive homolytic cleavage to thiyl radicals followed by thiol-disulfide exchange (accelerated at pH > 7), and acetal-type carbon-nitrogen bonds (-C-N-) in hexahydrotriazine rings that degrade via acid-catalyzed reversal of aldoamine condensation. This dual-bond synergy enables three distinct room temperature degradation pathways (acid-triggered, reduction-triggered, or hybrid) while maintaining robust thermal stability (Td5 = 259 °C) and solvent resistance, despite a marginal Tg reduction (116 °C vs. 121 °C for single-bond DDM/DCNSA), overcoming traditional trade-offs between degradability and performance in thermosets. Moreover, Tian et al. [28] designed a polyimine gel (PIT) that can be recycled in acidic solutions and exhibits excellent processability. During combustion, the Schiff base structure undergoes trimerization reactions, forming stable nitrogen-containing hexagonal rings that enhance char layer stability and further improve flame retardancy.
In the selection of reinforcement phases, researchers have gradually shifted from traditional inorganic fillers (such as aluminum hydroxide, magnesium hydroxide, etc.) to nanomaterials (such as MXene, TiO2 nanorods, phosphorus-based aromatic compounds, etc.), which exhibit higher flame-retardant efficiency and synergistic effects. Modern fireproof materials not only need excellent fire resistance but also need to integrate other functionalities, such as self-healing and smart responsiveness. For instance, Mao et al. [48] fabricated a TiO2@MXene/P, N-containing polyimine nanocomposite (TMPNP) featuring self-healing ability and excellent flame retardancy. By enhancing the composite’s flame-retardant properties and increasing its sensitivity to temperature variations, the TiO2@MXene heterostructure demonstrates great potential for application in smart fire warning systems. Additionally, in 2021, Liu et al. [72] developed a polyimine thermoset derived from vanillin-terminated phosphorus-based aromatic monomers (HVP) for fabricating carbon fiber-reinforced composites (CFRPs). Through the introduction of vanillin-terminated phosphorus-based aromatic monomers, Liu et al. achieved simultaneous improvements in flame retardancy and recyclability of CFRPs. Experimental evaluation revealed that the HVP/D230 composite self-extinguished within 1 s after ignition and resisted re-ignition, attributed to the formation of internally combustible carbon layer. Throughout the testing process, no melt dripping was detected. Scanning electron microscopy (SEM) analysis of the residual char morphologies for HVP/D230 and E51/D230 composites indicated that HVP/D230 produced a continuous and uniform surface with distinct stretching tracks, suggesting the formation of a robust, expanded char layer [72]. This well-organized three-dimensional char structure effectively limits combustible gas release, reduces thermal transfer, and significantly enhances the material’s flame-retardant performance.
Researchers have not only improved the flame retardancy of composites but also their mechanical performance, thermal stability, and processability. For instance, Wang et al. (2024) [55] incorporated trifluoromethyl groups into polyimine networks, resulting in a fluorinated bio-based polyimine exhibiting outstanding mechanical strength (tensile strength up to 96.2 MPa), superior flame retardancy (UL-94 V0 rating, LOI of 50%), and excellent reprocessing and solvent-healing capabilities at 160 °C. Yuan et al. (2023) [61] synthesized aromatic polyimine covalent adaptable networks (CANs), significantly enhancing water stability, thermal stability (Td5%: 434–441 °C, Tg: 217–239 °C), and mechanical performance through conjugation effects and aromatic structural stability. These aromatic CANs exhibited very low water absorption (0.14–0.15%) and retained mechanical robustness despite environmental moisture.

5.5. Other Applications

5.5.1. Flexible Magnetic Soft Robotics Based on Polyimine Composites

Polyimine (PI) composites exhibit excellent flexibility, stretchability, and responsiveness, offering significant potential in flexible magnetic soft robotics. To enhance recyclability and self-healing, magnetic particles (e.g., NdFeB) were incorporated into PI matrices. In 2023, Zhu et al. [46] developed a dynamic PI network, achieving deformation and locomotion within 2–3 s by adjusting crosslinking density and hydrogen bonding. The robots demonstrated complete chemical recyclability, room temperature self-healing, and retention of mechanical performance after multiple cycles. Similarly, Li et al. [96] reported a mild approach to fabricate magnetic DVA–ETTA–PI nanospheres. The resulting Fe3O4@PEI@PI nanospheres exhibited fast magnetic response, enriched C-peptides with a detection limit of 0.1 fmol/μL, and maintained performance after at least six reuse cycles, showing potential in chemical sensing and clinical diagnostics.

5.5.2. Applications of Polyimine Composites in Drug Delivery

PI composites, with hierarchical pores and tunable surface chemistry, offer superior drug loading and release properties. Nazanin Mokhtari et al. [97] developed a polyimine-based covalent triazine framework (PI-CTF) to serve as a carrier for sorafenib, attaining an encapsulation efficiency of 98% and a drug loading capacity of 83%. In vitro studies showed pH-dependent release, exhibiting accelerated drug release behavior under acidic conditions (pH 5.3). Cytotoxicity tests confirmed good biocompatibility, indicating PI-CTF’s suitability for targeted drug delivery [97].

5.5.3. Rapid Detection of Volatile Iodine Using Polyimine Composites

Polyimine-based composites show rapid response and high sensitivity in volatile iodine detection. Zhang’s group [27] developed a PI/PVP composite membrane combining the iodine-binding ability of PVP with PI’s selectivity. The sensor responded to trace iodine within 5 s, reducing fluorescence intensity to 62.51% of the initial value, with a detection limit of 4.087 × 10−8 mol/L. It maintained excellent selectivity under interference from gases like ethanol and acetone, supporting applications in nuclear monitoring and medical disinfection.

6. Conclusions

Polyimine-based dynamic networks, featuring reversible imine linkages (C=N), have significantly mitigated the repairability and reprocessability challenges inherent in traditional thermosetting systems. They have emerged as a critical platform for advancing high-performance and sustainable engineering materials.
This review systematically summarizes the fundamental chemical structures, synthesis mechanisms, dynamic bond design strategies, fabrication techniques, and reinforcement approaches of PI-based composites. Their outstanding performances in mechanical strength, thermal stability, self-healing, electrical responsiveness, and recyclability were comprehensively discussed. Diverse processing methods, including solution casting, hot pressing, and interfacial polymerization, offer versatile pathways for morphology construction and structural tuning of PI systems. The incorporation of various organic and inorganic reinforcements, such as carbon nanotubes, graphene, natural fibers, and bio-based fillers, significantly enhanced the rigidity, toughness, thermal conductivity, flame retardancy, and multifunctionality of PI composites.
Mechanical testing revealed that many PI-based composites achieved tensile strengths exceeding 60 MPa and self-healing efficiencies greater than 90%, outperforming conventional thermosets. The introduction of dynamic covalent networks imparted excellent ductility, stress relaxation behavior, and enhanced interfacial stress transfer and crack suppression capabilities, enabling broad applicability in flexible electronics, wearable devices, high-performance sensors, self-healing waterproof coatings, and green adhesives.
Overall, PI composites, with their tunable chemistry, adjustable architecture, and reconfigurable performance, represent a new paradigm bridging thermosetting and thermoplastic materials, demonstrating great potential for sustainable development and intelligent material design. Nevertheless, further research is needed to address critical challenges, including interfacial coupling stability, functional integration complexity, and scalability toward industrial applications. PI composites are currently at a pivotal stage, transitioning from fundamental research to engineering practice, with significant theoretical and practical implications for future materials science and technology.

7. Future Opportunities and Challenges

Substantial advancements have been achieved in the study of polyimine (PI) and its composites, leading to a deeper understanding of their structure–property relationships and application potentials. However, translating laboratory-scale achievements into large-scale engineering applications remains a major challenge. With the continuous evolution of material systems and the advancement of global carbon neutrality strategies, the development of PI composites faces unprecedented opportunities and technological windows.
First, regarding multifunctional integration, PI materials possess inherent potential such as conductivity, self-healing, flame retardancy, and structural reinforcement. Yet, most current studies focus on optimizing individual properties, lacking comprehensive design strategies tailored for complex service environments. Establishing efficient synergy among dynamic networks, reinforcing phases, and external stimuli-responsive systems demands the development of multiscale collaborative models bridging molecular design, structural architecture, and macroscopic configurations.
Second, limitations associated with practical self-healing and closed-loop recycling hinder widespread adoption. The reconfiguration stability of dynamic bonds under real-world conditions remains insufficient. Developing room temperature reversible bonds (e.g., boronate esters [98], hydrazone bonds [44], and synergistic hydrogen bonding systems [58]) or incorporating low-energy stimuli such as light, humidity, or electrical triggers could significantly broaden their applicability in portable electronics, smart sensors, and medical devices.
Finally, from an industrialization perspective, the lack of standardized, scalable manufacturing platforms restricts the engineering translation of PI composites. Current efforts largely remain at the laboratory scale, with few studies addressing continuous and modular production processes for practical forms such as films, coatings, foams, and sheets. Integrating advanced manufacturing techniques such as 3D printing, intelligent spraying, and roll-to-roll coating could enable the establishment of high-throughput, cost-effective PI material production platforms, accelerating their commercialization.
Future developments will require interdisciplinary collaboration, deeply integrating dynamic chemistry, interfacial engineering, sustainable manufacturing, and intelligent responsive systems. Figure 17 provides a forward-looking illustration of the future development landscape of polyimine-based materials. Building a novel, proprietary technology ecosystem for functional materials will be crucial to supporting the advancement toward high-performance, green, and intelligent engineering materials.

Author Contributions

Conceptualization, S.Z. and Y.C.; methodology, X.W. and S.Z.; software, X.W.; validation, X.W., S.Z. and Y.C.; formal analysis, X.W.; investigation, X.W.; resources, S.Z.; data curation, X.W.; writing—original draft preparation, S.Z.; writing—review and editing, Y.C.; visualization, Y.C.; supervision, S.Z.; project administration, S.Z.; funding acquisition, Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PIPolyimine
DCCDynamic Covalent Chemistry
CANsCovalent Adaptive Networks
SWNTsSingle-Walled Carbon Nanotubes
MWCNTMulti-Walled Carbon Nanotube
CNTsCarbon Nanotubes
COFCovalent Organic Framework
CFCarbon Fiber
TENGTriboelectric Nanogenerator
TEGsThermoelectric Generators
SSEsSolid-State Electrolytes
TFCThin-Film Composite
MINTsMechanically Interlocked Nanotube Derivatives
BP/PIBamboo Powder/Polyimine Composite
PVPPolyvinylpyrrolidone
TRENTris(2-aminoethyl)amine
TPATerephthalaldehyde
PGCSPeach Gum Polysaccharide/Chitosan Composite
IPINIonic Polyimine Network
ASAAerogel-Sol-Aerogel Process
TMPNPTiO2@MXene/P, N-containing Polyimine Nanocomposite
CFRPsCarbon Fiber-Reinforced Composites
RY-PIRamie Yarn-Reinforced Polyimine Vitrimer Composites
PPCsPaper–Polyimine Composites
BNNS/PIH TIMBoron Nitride Nanosheets/Polyimine Hybrid Thermal Interface Material
SEMScanning Electron Microscopy
TEMTransmission Electron Microscopy
FTIRFourier Transform Infrared Spectroscopy
TGAThermogravimetric Analysis
DTGDerivative Thermogravimetric
ECLElectrochemiluminescence
VitrimerA polymer combining thermoset properties with dynamic covalent adaptability

References

  1. Cong, F.; Wang, B.; Li, Y.; Liu, P.; Fu, X.; Liu, X.; Huang, Y.; Hu, Z. Waste-free recycling of multifunctional epoxy resins and carbon fiber composites enabled by dynamic aminal bonds. Compos. Sci. Technol. 2025, 267, 111204. [Google Scholar] [CrossRef]
  2. Evans, A.M.; Strauss, M.J.; Corcos, A.R.; Hirani, Z.; Ji, W.; Hamachi, L.S.; Aguilar-Enriquez, X.; Chavez, A.D.; Smith, B.J.; Dichtel, W.R. Two-Dimensional Polymers and Polymerizations. Chem. Rev. 2022, 122, 442–564. [Google Scholar] [CrossRef] [PubMed]
  3. Liguori, A.; Hakkarainen, M. Designed from Biobased Materials for Recycling: Imine-Based Covalent Adaptable Networks. Macromol. Rapid Commun. 2022, 43, 2100816. [Google Scholar] [CrossRef]
  4. Zhao, Z.-H.; Fu, J. Chemical closed-loop recyclable thermosetting polymers based on dynamic covalent bonds. Sci. China-Chem. 2025, 68. [Google Scholar] [CrossRef]
  5. Zhang, J.; Luo, S.; Ma, Y.; Li, R.; Jin, Y.; Qiu, L.; Zhang, W. Monolithic polyimine vitrimer/graphene aerogel composites. Chin. Chem. Lett. 2023, 34, 107363. [Google Scholar] [CrossRef]
  6. Xie, D.; Pu, Y.; Bryant, N.D.; Harper, D.P.; Wang, W.; Ragauskas, A.J.; Li, M. Synthesis of Bio-Based Repairable Polyimines with Tailored Properties by Lignin Fractionation. Acs Sustain. Chem. Eng. 2024, 12, 6606–6618. [Google Scholar] [CrossRef]
  7. Gao, S.; Sun, F.; Brady, A.; Pan, Y.; Erwin, A.; Yang, D.; Tsukruk, V.; Stack, A.G.; Saito, T.; Yang, H.; et al. Ultra-efficient polymer binder for silicon anode in high-capacity lithium-ion batteries. Nano Energy 2020, 73, 104804. [Google Scholar] [CrossRef]
  8. Yang, L.; Wang, J.; Zhao, K.; Fang, Z.; Qiao, H.; Zhai, L.; Mi, L. Photoactive Covalent Organic Frameworks for Catalyzing Organic Reactions. ChemPlusChem 2022, 87, e202200281. [Google Scholar] [CrossRef]
  9. da Silva, C.M.; da Silva, D.L.; Modolo, L.V.; Alves, R.B.; de Resende, M.A.; Martins, C.V.B.; de Fátima, Â. Schiff bases: A short review of their antimicrobial activities. J. Adv. Res. 2011, 2, 1–8. [Google Scholar] [CrossRef]
  10. Fonkui Youmbi, T. Schiff Bases: Synthesis, Characterization and Antimicrobial Properties; University of Johannesburg: Johannesburg, South Africa, 2018. [Google Scholar]
  11. Taynton, P. Development of Polyimine-Based Dynamic Covalent Network: From Malleable Polymers to High-Performance Composites. Ph.D. Thesis, University of Colorado at Boulder, Boulder, CO, USA, 2016. [Google Scholar]
  12. Taynton, P.; Yu, K.; Shoemaker, R.K.; Jin, Y.; Qi, H.J.; Zhang, W. Heat- or Water-Driven Malleability in a Highly Recyclable Covalent Network Polymer. Adv. Mater. 2014, 26, 3938–3942. [Google Scholar] [CrossRef]
  13. Taynton, P.; Ni, H.; Zhu, C.; Yu, K.; Loob, S.; Jin, Y.; Qi, H.J.; Zhang, W. Repairable Woven Carbon Fiber Composites with Full Recyclability Enabled by Malleable Polyimine Networks. Adv. Mater. 2016, 28, 2904–2909. [Google Scholar] [CrossRef] [PubMed]
  14. Jin, Y.; Wang, Q.; Taynton, P.; Zhang, W. Dynamic Covalent Chemistry Approaches Toward Macrocycles, Molecular Cages, and Polymers. Acc. Chem. Res. 2014, 47, 1575–1586. [Google Scholar] [CrossRef] [PubMed]
  15. Kloxin, C.J.; Scott, T.F.; Adzima, B.J.; Bowman, C.N. Covalent Adaptable Networks (CANS): A Unique Paradigm in Cross-Linked Polymers. Macromolecules 2010, 43, 2643–2653. [Google Scholar] [CrossRef] [PubMed]
  16. Hajj, R.; Duval, A.; Dhers, S.; Averous, L. Network Design to Control Polyimine Vitrimer Properties: Physical Versus Chemical Approach. Macromolecules 2020, 53, 3796–3805. [Google Scholar] [CrossRef]
  17. Zhu, C. Structure-Property Relationship Study of Polyimine-Based Dynamic Covalent Networks and Their Applications Towards Functional Composites. Master’s Thesis, University of Colorado at Boulder, Boulder, CO, USA, 2017. [Google Scholar]
  18. Yu, P.; Huang, Q.; Wang, Y.; Peng, W.; Jia, Z.; Wang, H.; Ma, J.; Wang, C.; Yan, X. Water-Driven Malleable, Weldable and Eco-Friendly Recyclable Carbon Fiber Reinforced Dynamic Composites. Chin. J. Chem. 2024, 42, 516–522. [Google Scholar] [CrossRef]
  19. Wang, C.; Eisenreich, F.; Tomovic, Z. Aerogel-To-Sol-To-Aerogel (ASA) Process for Recycling, Repairing, Reprogramming of High-Performance Organic Aerogels. Adv. Funct. Mater. 2024, 34, 2314447. [Google Scholar] [CrossRef]
  20. Hsieh, T.-H.; Wang, Y.-Z.; Ho, K.-S. Cobalt-Based Cathode Catalysts for Oxygen-Reduction Reaction in an Anion Exchange Membrane Fuel Cell. Membranes 2022, 12, 699. [Google Scholar] [CrossRef]
  21. Xiong, Y.; Wang, C.-L.; Dong, L.; Tomovic, Z. Closed-Loop Recyclable High-Performance Aerogels Derived from Polystyrene. Adv. Funct. Mater. 2025, 35, 2420472. [Google Scholar] [CrossRef]
  22. Huang, T.; Pan, X.; Xi, A.; Chen, W.; Zhang, N.; Zeng, Y. Peach gum-based polyimine networks with water resistant, high strength and recycling performances. Polymer 2024, 315, 127845. [Google Scholar] [CrossRef]
  23. Lin, L.; Su, Z.; Zhang, H.; Zhou, G.; Zhou, H.; Ren, J.; Wang, X.; Liu, C.; Wang, X. Thermo-processable chitosan-based plastic substitute with self-adaptiveness and closed-loop recyclability. Carbohydr. Polym. 2022, 291, 119479. [Google Scholar] [CrossRef]
  24. Tiwari, K.; Modak, S.; Sarkar, P.; Ray, S.; Adupa, V.; Reddy, K.A.; Pramanik, S.K.; Das, A.; Karan, S. Interfacial synthesis of large-area ultrathin polyimine nanofilms as molecular separation membrane. Iscience 2022, 25, 104027. [Google Scholar] [CrossRef] [PubMed]
  25. Qi, H.; Sahabudeen, H.; Liang, B.; Polozij, M.; Addicoat, M.A.; Gorelik, T.E.; Hambsch, M.; Mundszinger, M.; Park, S.; Lotsch, B.V.; et al. Near-atomic-scale observation of grain boundaries in a layer-stacked two-dimensional polymer. Sci. Adv. 2020, 6, eabb5976. [Google Scholar] [CrossRef] [PubMed]
  26. Lei, D.; Zhang, Z.; Zhang, W.; Liu, X. Strengthening and Toughening Dynamic Covalent Thermosets via Hydrogen-Bonded Cross-Links. Macromolecules 2024, 57, 1362–1369. [Google Scholar] [CrossRef]
  27. Zhang, G.-H.; Zhu, Q.-H.; Guo, S.-J.; Zhang, L.; Yu, C.; Qin, S.; He, L.; Tao, G.-H. Ionic Polyimine-Based Composite Membrane with Inductive and Complexation Synergistic Effects for Sensitive and On-Site Fluorescent Detection of Volatile Iodine. Adv. Mater. 2024, 36, 2311990. [Google Scholar] [CrossRef]
  28. Tian, T.; Jiang, S.; Lu, B.; Liu, L. Synthesis, degradation Behavior, and recyclability property of novel bio-based Polyimine thermosets. Eur. Polym. J. 2024, 220, 113430. [Google Scholar] [CrossRef]
  29. Yu, P.; Wang, H.; You, C.; Jia, Z.; Huang, Q.; Wang, Y.; Qu, Y.; Dong, X.; Li, R.; Xin, Y.; et al. Reversible Schiff-base chemistry enables thermosetting smart composites with versatile properties. Compos. Commun. 2024, 52, 102153. [Google Scholar] [CrossRef]
  30. Su, Z.; Cui, L.; Zhang, H.; Xiao, L.; Chi, B.; Xu, H.; Ning, L.; Jia, S.; Wang, X. Robust, waterproof, and degradable cellulose-based polyimine vitrimer for plastic replacement. Chem. Eng. J. 2023, 471, 144501. [Google Scholar] [CrossRef]
  31. Zhang, S.; Lv, Y.; Zheng, L.; Li, J.; Liang, S.; Liu, Z.; Ren, L. Bio-inspired Polyimine Copolymers: Facial Integration with High Content Variability and Extremal Transitions of Mechanical Properties. J. Bionic Eng. 2017, 14, 119–129. [Google Scholar] [CrossRef]
  32. Zhang, H.; Pan, M.; Qin, S.; Zheng, Z.; Xu, H.; Ning, L.; Zhang, S.; Jia, S.; Wang, X.; Su, Z. A fully sustainable, flexible, and degradable lignocellulose-based composite film enabled by a bio-based polyimine vitrimer. Int. J. Biol. Macromol. 2025, 307, 141946. [Google Scholar] [CrossRef]
  33. Cui, L.; Pan, M.; Zhou, Y.; Xu, H.; Ning, L.; Jia, S.; Wang, X.; Su, Z. A strong, biodegradable, and closed-loop recyclable bamboo-based plastic substitute enabled by polyimine covalent adaptable networks. Chem. Eng. J. 2023, 477, 146952. [Google Scholar] [CrossRef]
  34. Sahabudeen, H.; Qi, H.; Ballabio, M.; Polozij, M.; Olthof, S.; Shivhare, R.; Jing, Y.; Park, S.; Liu, K.; Zhang, T.; et al. Highly Crystalline and Semiconducting Imine-Based Two-Dimensional Polymers Enabled by Interfacial Synthesis. Angew. Chem. Int. Ed. 2020, 59, 6028–6036. [Google Scholar] [CrossRef] [PubMed]
  35. Dai, W.Y.; Shao, F.; Szczerbinski, J.; McCaffrey, R.; Zenobi, R.; Jin, Y.H.; Schlüter, A.D.; Zhang, W. Synthesis of a Two-Dimensional Covalent Organic Monolayer through Dynamic Imine Chemistry at the Air/Water Interface. Angew. Chem. Int. Ed. 2016, 55, 213–217. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, B.; Liu, X.; Li, W.; Clausner, A.; Conzendorf, S.; Liu, J.; Posseckardt, J.; Jost, B.; Dong, R.; Feng, X.; et al. Patterning damage mechanisms for two-dimensional crystalline polymers and evaluation for a conjugated imine-based polymer. Nanotechnology 2024, 35, 475301. [Google Scholar] [CrossRef] [PubMed]
  37. Ge, L.; Chen, T.; Liu, Z.; Li, Y.; Su, H.; Zhu, J.; Zhang, Y.; van der Bruggen, B. Interfacial oligomer splicing toward 3D silicon-centered polyimine nanofilm for rapid molecule/ion differentiation. J. Membr. Sci. 2025, 713, 123354. [Google Scholar] [CrossRef]
  38. Baig, M.I.; Hardian, R.; Alharthi, F.A.; Fellows, C.M.; Szekely, G. Brine separation with polyamide and polyimine thin film composite nanofiltration membranes obtained from biobased monomers. J. Membr. Sci. 2025, 713, 123324. [Google Scholar] [CrossRef]
  39. Lamm, M.E.; Li, K.; Copenhaver, K.; Kelly, P.V.; Senkum, H.; Tekinalp, H.; Gramlich, W.M.; Ozcan, S. Aqueous-Based Polyimine Functionalization of Cellulose Nanofibrils for Effective Drying and Polymer Composite Reinforcement. ACS Appl. Polym. Mater. 2022, 4, 7674–7684. [Google Scholar] [CrossRef]
  40. Zhang, J.; Lei, Z.; Luo, S.; Jin, Y.; Qiu, L.; Zhang, W. Malleable and Recyclable Conductive MWCNT-Vitrimer Composite for Flexible Electronics. Acs Appl. Nano Mater. 2020, 3, 4845–4850. [Google Scholar] [CrossRef]
  41. Zhang, S.; Xiong, Z.; Zhang, J.; Zhang, X.; Chen, Y.; Chen, Y. Mechanical properties and thermal analysis of graphene nanoplatelets reinforced polyimine composites. E-Polymers 2022, 22, 696–704. [Google Scholar] [CrossRef]
  42. Zhang, N.; Pan, X.; Xi, A.; Chen, W.; Huang, T.; Zeng, Y. Robust, malleable, degradable, self-healable, weldable and recyclable polyimine thermosets from natural peach gum and chitosan. Polym. Chem. 2024, 15, 3287–3299. [Google Scholar] [CrossRef]
  43. Su, Z.; Huang, S.; Wang, Y.; Ling, H.; Yang, X.; Jin, Y.; Wang, X.; Zhang, W. Robust, high-barrier, and fully recyclable cellulose-based plastic replacement enabled by a dynamic imine polymer. J. Mater. Chem. A 2020, 8, 14082–14090. [Google Scholar] [CrossRef]
  44. Ding, H.; Wang, J.; Wang, C.; Wu, L.; Zhang, W.; Liu, L.; Lei, Y.; Sun, N.; Zhou, K.; Yu, B. Intermolecular hydrogen bonding enabling mechanically robust, thermally stable, and solvent-resistance bio-based polyimine networks. Compos. Part A Appl. Sci. Manuf. 2025, 196, 109006. [Google Scholar] [CrossRef]
  45. Isasti, I.; Miranda, S.; Jimenez, D.M.; Parzyszek, S.; Sabanes, N.M.; Pedersen, H.; Perez, E.M. Reinforcement of Polyimine Covalent Adaptable Networks with Mechanically Interlocked Derivatives of SWNTs. Adv. Funct. Mater. 2024, 34, 2408592. [Google Scholar] [CrossRef]
  46. Zhu, G.; Hou, Y.; Xia, N.; Wang, X.; Zhang, C.; Zheng, J.; Jin, D.; Zhang, L. Fully Recyclable, Healable, Soft, and Stretchable Dynamic Polymers for Magnetic Soft Robots. Adv. Funct. Mater. 2023, 33, 2300888. [Google Scholar] [CrossRef]
  47. Zhang, S.; Lv, Y.; Li, J.; Liang, S.; Liu, Z. Mechanical enhancement of zirconia reinforced polyimine nanocomposites. J. Appl. Polym. Sci. 2017, 134, 45183. [Google Scholar] [CrossRef]
  48. Mao, Y.; Shi, S.; Lei, L.; Wang, C.; Wang, D.; Hu, J.; Fu, S. A self-healable and highly flame retardant TiO2@MXene/P, N-containing polyimine nanocomposite for dual-mode fire sensing. Chem. Eng. J. 2024, 479, 147545. [Google Scholar] [CrossRef]
  49. Wang, Y.; Xu, A.; Zhang, L.; Chen, Z.; Qin, R.; Liu, Y.; Jiang, X.; Ye, D.; Liu, Z. Recyclable Carbon Fiber Reinforced Vanillin-Based Polyimine Vitrimers: Degradation and Mechanical Properties Study. Macromol. Mater. Eng. 2022, 307, 2100893. [Google Scholar] [CrossRef]
  50. Zhang, S.; Ji, S.; Wang, Z.; Zhang, J.; Zhao, W.; He, C.; Chen, Y. Mechanical and Recyclable Properties of Polyimine Enhanced by Biomimetic Modification of Graphene Oxide Sheets/Silicon Carbide Nano-Whiskers. Nanomaterials 2022, 12, 4486. [Google Scholar] [CrossRef]
  51. Zhu, K.; Li, Z.; Cheng, F.; Wu, C.; Cai, D.; Zhang, Q.; Zhang, H. Preparation of durable superhydrophobic composite coatings with photothermal conversion precisely targeted configuration self-healability and great degradability. Compos. Sci. Technol. 2021, 213, 108926. [Google Scholar] [CrossRef]
  52. Wang, Y.-F.; Mu, G.-D.; Wang, X.-J.; Zhang, F.; Li, Y.-L.; Lu, D.-J.; Chen, F.-M.; Yang, M.-L.; He, M.-Y.; Liu, T. Fast construction of core-shell structured magnetic covalent organic framework as sorbent for solid-phase extraction of zearalenone and its derivatives prior to their determination by UHPLC-MS/MS. Microchim. Acta 2021, 188, 246. [Google Scholar] [CrossRef]
  53. Li, P.; Hao, C.; Wang, H.; He, T.; Shu, T.; Li, C.; Yu, L.; Yan, N. Eco-friendly recyclable high performance ramie yarn reinforced polyimine vitrimer composites. Chem. Eng. J. 2023, 457, 141341. [Google Scholar] [CrossRef]
  54. Zhang, S.; Wang, Z.; An, C.; Wu, N.; Wang, X.; Chen, Y. Interface and mechanical properties of combined surface-treated carbon fiber fabrics/polyimine laminates: Experiments and numerical simulations. Polym. Compos. 2025, 1–14. [Google Scholar] [CrossRef]
  55. Wang, J.; Ding, H.; Yu, P.; Jia, Z.; Wang, C.; Wang, H.; He, H.; Xing, W.; Yang, W.; Zhang, P.; et al. Ultrastrong, High Fire Resistance, Repairable, and Recyclable Fluorinated Biobased Polyimine Networks. Acs Mater. Lett. 2024, 6, 3899–3908. [Google Scholar] [CrossRef]
  56. Li, B.; Zhu, G.; Hao, Y.; Ren, T. Effect of cross-link density on the performance of polyimine/epoxy vitrimers. Smart Mater. Struct. 2024, 33, 025014. [Google Scholar] [CrossRef]
  57. Lei, Y.; Shan, S.; Lin, Y.; Zhang, A. Network reconfiguration and unusual stress intensification of a dynamic reversible polyimine elastomer. Polymer 2020, 186, 122031. [Google Scholar] [CrossRef]
  58. Zhang, G.-H.; Zhang, L.; Zhu, Q.-H.; Chen, H.; Yuan, W.-L.; Fu, J.; Wang, S.-L.; He, L.; Tao, G.-H. Self-Healable, Malleable, and Flexible Ionic Polyimine as an Environmental Sensor for Portable Exogenous Pollutant Detection. ACS Mater. Lett. 2022, 4, 136–144. [Google Scholar] [CrossRef]
  59. Yang, Z.; Li, H.; Mou, X.; Chen, Z.; Lai, X.; Ding, J.; Zeng, X. Functional and Environmental Friendly Polyimine Elastomer Based on the Dynamic Covalent Network for a Flexible Strain Sensor. Macromolecules 2023, 56, 9766–9777. [Google Scholar] [CrossRef]
  60. Zhang, X.; Eichen, Y.; Miao, Z.; Zhang, S.; Cai, Q.; Liu, W.; Zhao, J.; Wu, Z. Novel Phosphazene-Based flame retardant polyimine vitrimers with Monomer-Recovery and high performances. Chem. Eng. J. 2022, 440, 135806. [Google Scholar] [CrossRef]
  61. Yuan, Y.; Chen, H.; Jia, L.; Lu, X.; Yan, S.; Zhao, J.; Liu, S. Aromatic polyimine covalent adaptable networks with superior water and heat resistances. Eur. Polym. J. 2023, 187, 111912. [Google Scholar] [CrossRef]
  62. Li, P.; Zhang, Q.; Ma, J.; Liao, Z.; Zhang, J.; Xie, H.; Yang, S.; Xu, C.-A.; Hu, Y.; Yang, Z. Synthesis of bio-based polyimine networks with flame-retardancy, acid-degradablility, and reprocessability. Polymer 2024, 302, 127101. [Google Scholar] [CrossRef]
  63. Yu, P.; Wang, H.; Li, T.; Wang, G.; Jia, Z.; Dong, X.; Xu, Y.; Ma, Q.; Zhang, D.; Ding, H.; et al. Mechanically Robust, Recyclable, and Self-Healing Polyimine Networks. Adv. Sci. 2023, 10, e2300958. [Google Scholar] [CrossRef]
  64. Zhang, B.; Liu, X.; Bodesheim, D.; Li, W.; Clausner, A.; Liu, J.; Jost, B.; Dianat, A.; Dong, R.; Feng, X.; et al. Fracture Behavior of a 2D Imine-Based Polymer. Adv. Sci. 2024, 11, e2407017. [Google Scholar] [CrossRef]
  65. Peng, H.; Li, H.; Guo, E.; Zhai, T. High-Performance Te Nanowires/MoS2/Polyimine Nanocomposite-Based Self-Healable, Recyclable and Screen-Printable Flexible Photodetector for Image Sensing. Adv. Funct. Mater. 2024, 34, 2314743. [Google Scholar] [CrossRef]
  66. Rajabi-Abhari, A.; Li, P.; Bagheri, M.H.; Khan, A.A.; Hao, C.; Tanguy, N.R.; Ban, D.; Yu, L.; Yan, N. Self-healable, recyclable, and mechanically robust vitrimer composite for high-performance triboelectric nanogenerators and self-powered wireless electronics. Nano Energy 2024, 131, 110306. [Google Scholar] [CrossRef]
  67. He, X.; Lin, Y.; Ding, Y.; Abdullah, A.M.; Lei, Z.; Han, Y.; Shi, X.; Zhang, W.; Yu, K. Reshapeable, rehealable and recyclable sensor fabricated by direct ink writing of conductive composites based on covalent adaptable network polymers. Int. J. Extrem. Manuf. 2022, 4, 015301. [Google Scholar] [CrossRef]
  68. Liu, W.; Liu, Y.; Zhong, S.; Chen, J.; Li, Z.; Zhang, C.; Jiang, P.; Huang, X. Soft and Damping Thermal Interface Materials with Honeycomb-Board-Mimetic Filler Network for Electronic Heat Dissipation. Small 2024, 20, e2400115. [Google Scholar] [CrossRef]
  69. Wang, S.; Ma, S.; Li, Q.; Xu, X.; Wang, B.; Huang, K.; Liu, Y.; Zhu, J. Facile Preparation of Polyimine Vitrimers with Enhanced Creep Resistance and Thermal and Mechanical Properties via Metal Coordination. Macromolecules 2020, 53, 2919–2931. [Google Scholar] [CrossRef]
  70. Luo, S.; Ma, Y.; Wei, X.; Jin, Y.; Qiu, L.; Zhang, W. Malleable and Recyclable Vitrimer-Graphene Aerogel Composite with High Electrical Conductivity. Acs Appl. Electron. Mater. 2021, 3, 1178–1183. [Google Scholar] [CrossRef]
  71. Zhao, Q.; Liu, X.; Deng, S.; Zheng, X.; Zeng, B.; Liu, X.; Xu, Y.; Yuan, C.; Dai, L. Construction of phosphorus-containing polyimine vitrimer film by incorporating ionic liquid BMIM•PF6 for enhanced mechanical, flame-retardant and reprocessable properties. Chem. Eng. J. 2024, 495, 153283. [Google Scholar] [CrossRef]
  72. Liu, T.; Peng, J.; Liu, J.; Hao, X.; Guo, C.; Ou, R.; Liu, Z.; Wang, Q. Fully recyclable, flame-retardant and high-performance carbon fiber composites based on vanillin-terminated cyclophosphazene polyimine thermosets. Compos. Part B-Eng. 2021, 224, 109188. [Google Scholar] [CrossRef]
  73. Ji, S.; Zhang, S.; Wang, Z.; Li, C.; Cao, W.; Zhu, Y.; He, C.; Chen, Y. High-Impact Performance and Thermal Properties of Polyimine Nanocomposites Reinforced by Silicon Carbide Nano-Whiskers. Materials 2023, 16, 4587. [Google Scholar] [CrossRef]
  74. Lyu, M.; Liu, Y.; Yang, X.; Liang, D.; Wang, Y.; Liang, X.; Hu, Y.; Liang, L.; Zhang, C. Vanillin-based liquid crystalline polyimine thermosets and their composites for recyclable thermal management application. Compos. Part B-Eng. 2023, 250, 110462. [Google Scholar] [CrossRef]
  75. Pomazi, A.; Poor, D.I.; Geier, N.; Toldy, A. Optimising Recycling Processes for Polyimine-Based Vitrimer Carbon Fibre-Reinforced Composites: A Comparative Study on Reinforcement Recovery and Material Properties. Materials 2024, 17, 2372. [Google Scholar] [CrossRef] [PubMed]
  76. Toldy, A.; Poor, D.I.; Szolnoki, B.; Devecser, B.; Geier, N.; Pomazi, A. Comparative study of flame retardancy in polyimine vitrimers and composites: Evaluating additive and reactive flame retardants acting via gas-, solid-, and combined-phase mechanisms. J. Mater. Sci. Technol. 2024, 196, 101–111. [Google Scholar] [CrossRef]
  77. Yu, L.; Lei, Z.; Sun, X.; Ding, P.; Wesche, A.; Jin, Y.; Zhang, W.; Long, R. Rapid Fabrication of Fiber-Reinforced Polyimine Composites with Reprocessability, Repairability, and Recyclability. ACS Appl. Polym. Mater. 2021, 3, 5808–5817. [Google Scholar] [CrossRef]
  78. Ding, C.; Zhang, S.; Pan, M.; Li, M.-C.; Zhang, Y.; Mei, C. Improved processability and high fire safety of wood plastic composites via assembling reversible imine crosslinking network. Chem. Eng. J. 2021, 423, 130295. [Google Scholar] [CrossRef]
  79. Rahimi, A.; García, J.M. Chemical recycling of waste plastics for new materials production. Nat. Rev. Chem. 2017, 1, 0046. [Google Scholar] [CrossRef]
  80. Si, W.-J.; An, X.-P.; Zeng, J.-B.; Chen, Y.-K.; Wang, Y.-Z. Fully bio-based, highly toughened and heat-resistant poly(L-lactide) ternary blends via dynamic vulcanization with poly(D-lactide) and unsaturated bioelastomer. Sci. China Mater. 2017, 60, 1008–1022. [Google Scholar] [CrossRef]
  81. Su, Z.; Hu, Y.; Yang, X.; Long, R.; Jin, Y.; Wang, X.; Zhang, W. Production and closed-loop recycling of biomass-based malleable materials. Sci. China Mater. 2020, 63, 2071–2078. [Google Scholar] [CrossRef]
  82. Jia, Z.; Wang, H.; Yu, P.; He, H.; Huang, Q.; Hong, W.; Liu, C.; Shi, Y.; Wang, J.; Xin, Y.; et al. Soft-Rigid Construction of Mechanically Robust, Thermally Stable, and Self-Healing Polyimine Networks with Strongly Recyclable Adhesion. Small 2024, 20, 2406821. [Google Scholar] [CrossRef]
  83. Li, Z.; Dong, Y.; Li, C.; Guan, H.; Shen, L.; Meng, J.; Guo, K. Sustainable paeonol-derived polyimine-epoxy as a substitute for DGEBA thermoset. React. Funct. Polym. 2024, 194, 105804. [Google Scholar] [CrossRef]
  84. Wang, R.; Lei, Y.; Ke, L.; Bai, L.; Zheng, Y.; Xu, Z.; Li, T.; Chen, S.; Zhang, D. Construction of self-healing multifunctional hyper-crosslinked hyperbranched polymer @ metal-organic polyhedra (HHMOP) membranes. Eur. Polym. J. 2024, 211, 112966. [Google Scholar] [CrossRef]
  85. Qi, H.; Wang, Z.; Li, H.; Li, F. Directionally In Situ Self-Assembled Iridium(III)-Polyimine Complex-Encapsulated Metal-Organic Framework Two-Dimensional Nanosheet Electrode To Boost Electrochemiluminescence Sensing. Anal. Chem. 2023, 95, 12024–12031. [Google Scholar] [CrossRef] [PubMed]
  86. Zhu, P.; Shi, C.; Wang, Y.; Wang, Y.; Yu, Y.; Wang, Y.; Deng, Y.; Xiao, J. Recyclable, Healable, and Stretchable High-Power Thermoelectric Generator. Adv. Energy Mater. 2021, 11, 2100920. [Google Scholar] [CrossRef]
  87. Kuntharin, S.; Harnchana, V.; Klamchuen, A.; Sinthiptharakoon, K.; Thongbai, P.; Amornkitbamrung, V.; Chindaprasirt, P. Boosting the Power Output of a Cement-Based Triboelectric Nanogenerator by Enhancing Dielectric Polarization with Highly Dispersed Carbon Black Nanoparticles toward Large-Scale Energy Harvesting from Human Footsteps. Acs Sustain. Chem. Eng. 2022, 10, 4588–4598. [Google Scholar] [CrossRef]
  88. Tao, X.; Chen, X.; Wang, Z.L. Design and synthesis of triboelectric polymers for high performance triboelectric nanogenerators. Energy Environ. Sci. 2023, 16, 3654–3678. [Google Scholar] [CrossRef]
  89. Guo, Q.; Li, W.; Li, X.; Zhang, J.; Sabaghi, D.; Zhang, J.; Zhang, B.; Li, D.; Du, J.; Chu, X.; et al. Proton-selective coating enables fast-kinetics high-mass-loading cathodes for sustainable zinc batteries. Nat. Commun. 2024, 15, 2139. [Google Scholar] [CrossRef]
  90. Ding, L.; Yue, X.; Zhang, X.; Chen, Y.; Liu, J.; Shi, Z.; Wang, Z.; Yan, X.; Liang, Z. A polyimine aerogel separator with electron cloud design to boost Li- ion transport for stable Li metal batteries. Proc. Natl. Acad. Sci. USA 2023, 120, e2314264120. [Google Scholar] [CrossRef]
  91. Chen, Z.; Huang, J.; Cui, Y.; Fu, R. Precisely tailored morphology of polyimine for simple synthesis of metal sulfide/carbon flower-like superstructures. Carbon 2022, 190, 395–401. [Google Scholar] [CrossRef]
  92. Song, H.-S.; Rumon, M.M.H.; Rahman Khan, M.M.; Jeong, J.-H. Toward Intelligent Materials with the Promise of Self-Healing Hydrogels in Flexible Devices. Polymers 2025, 17, 542. [Google Scholar] [CrossRef]
  93. Jiang, C.-S.; Dunlap, N.; Li, Y.; Guthrey, H.; Liu, P.; Lee, S.-H.; Al-Jassim, M.M. Nonuniform Ionic and Electronic Transport of Ceramic and Polymer/Ceramic Hybrid Electrolyte by Nanometer-Scale Operando Imaging for Solid-State Battery. Adv. Energy Mater. 2020, 10, 2000219. [Google Scholar] [CrossRef]
  94. Long, T.E. Toward Recyclable Thermosets. Science 2014, 344, 706–707. [Google Scholar] [CrossRef] [PubMed]
  95. Zhang, W.; Zhou, Q.; Fang, C.; You, L.; Li, X.; Liu, M.; Dong, X.; Qi, Y.; Wang, B.; Li, W. Dynamic networks for polyimine with disulfide bond to obtain catalyst-free recyclability, multi-degradation and malleability. J. Polym. Res. 2022, 29, 154. [Google Scholar] [CrossRef]
  96. Li, Z.; Zhang, H.; Zhu, Y.; Luo, B.; He, J.; Lan, F.; Wu, Y. Rapid synthesis of magnetic polyimine nanospheres at room temperature for enrichment of endogenous C-peptide. Colloid Interface Sci. Commun. 2021, 42, 100390. [Google Scholar] [CrossRef]
  97. Mokhtari, N.; Taymouri, S.; Mirian, M.; Dinari, M. Covalent triazine-based polyimine framework as a biocompatible pH-dependent sustained-release nanocarrier for sorafenib: An in vitro approach. J. Mol. Liq. 2020, 297, 111898. [Google Scholar] [CrossRef]
  98. Konar, D.; Jakhar, V.K.; Stewart, K.A.; Lester, D.W.; Veige, A.S.; Sumerlin, B.S. Functionalized Cyclic Polymers and Network Gels. Macromolecules 2024, 57, 1779–1787. [Google Scholar] [CrossRef]
Figure 1. Bibliometric statistics for the past six years.
Figure 1. Bibliometric statistics for the past six years.
Polymers 17 01607 g001
Figure 2. Molecular structure of imines.
Figure 2. Molecular structure of imines.
Polymers 17 01607 g002
Figure 3. Synthesis of polyimines via dynamic chemistry approaches.
Figure 3. Synthesis of polyimines via dynamic chemistry approaches.
Polymers 17 01607 g003
Figure 4. Schematic diagram of water-driven dynamic imine bond degradation.
Figure 4. Schematic diagram of water-driven dynamic imine bond degradation.
Polymers 17 01607 g004
Figure 5. General fabrication flowchart of polyimine films (exemplified by PI-T) [26].
Figure 5. General fabrication flowchart of polyimine films (exemplified by PI-T) [26].
Polymers 17 01607 g005
Figure 7. Preparation of CF/PI laminates and schematic representation of interfacial improvement in carbon fiber (CF)/polyimine (PI) laminate [54].
Figure 7. Preparation of CF/PI laminates and schematic representation of interfacial improvement in carbon fiber (CF)/polyimine (PI) laminate [54].
Polymers 17 01607 g007
Figure 8. (a) Stress relaxation and intensification curves of PIE-40% experiencing successive multiple stress relaxation periods, (b) tensile stress–strain curves of pristine and multi-relaxed PIE-40%, and (c) schematic illustration of PIE network reconfiguration adapting to deformation [57].
Figure 8. (a) Stress relaxation and intensification curves of PIE-40% experiencing successive multiple stress relaxation periods, (b) tensile stress–strain curves of pristine and multi-relaxed PIE-40%, and (c) schematic illustration of PIE network reconfiguration adapting to deformation [57].
Polymers 17 01607 g008
Figure 9. (a) Schematic diagram of PI-daMINT chemical ligation composite method [45]; (b) schematic diagram of physical blending method for Vitrimer-MWCNTs [40].
Figure 9. (a) Schematic diagram of PI-daMINT chemical ligation composite method [45]; (b) schematic diagram of physical blending method for Vitrimer-MWCNTs [40].
Polymers 17 01607 g009
Figure 10. (a) Schematic diagram of self-healing principle of IPIN-1. (b) Molecular formula of IPIN-1 [58].
Figure 10. (a) Schematic diagram of self-healing principle of IPIN-1. (b) Molecular formula of IPIN-1 [58].
Polymers 17 01607 g010
Figure 11. Engineering applications of polyimines and composites.
Figure 11. Engineering applications of polyimines and composites.
Polymers 17 01607 g011
Figure 12. Closed-loop recyclable polyimines and composites enabled by dynamic covalent chemistry.
Figure 12. Closed-loop recyclable polyimines and composites enabled by dynamic covalent chemistry.
Polymers 17 01607 g012
Figure 13. (a) A schematic illustration of the sensing mechanism of the sensor; (b) a schematic diagram of the self-healing process of the conductive pathways of the flexible strain sensor [59].
Figure 13. (a) A schematic illustration of the sensing mechanism of the sensor; (b) a schematic diagram of the self-healing process of the conductive pathways of the flexible strain sensor [59].
Polymers 17 01607 g013
Figure 14. (a) Schematic illustration of PI/GP TENG device. Schematic diagram of PI/GP vitrimer composite; (b) schematic diagram of physical recycling strategy for PI/GP vitrimer composite. Schematic illustration of the chemical recycling process for the PI/GP [66].
Figure 14. (a) Schematic illustration of PI/GP TENG device. Schematic diagram of PI/GP vitrimer composite; (b) schematic diagram of physical recycling strategy for PI/GP vitrimer composite. Schematic illustration of the chemical recycling process for the PI/GP [66].
Polymers 17 01607 g014
Figure 15. Schematic illustration showing H+-dominated cathode intercalation chemistry enabled by H+-selective 2DPM coating [89].
Figure 15. Schematic illustration showing H+-dominated cathode intercalation chemistry enabled by H+-selective 2DPM coating [89].
Polymers 17 01607 g015
Figure 16. The interfacial Li+ diffusion and Li nucleation on the Li metal in PP and PIA separator systems [90].
Figure 16. The interfacial Li+ diffusion and Li nucleation on the Li metal in PP and PIA separator systems [90].
Polymers 17 01607 g016
Figure 17. Future outlook schematic: from microscale molecular design to innovations in manufacturing, and ultimately the development trends in the applications of polyimine and its composite materials.
Figure 17. Future outlook schematic: from microscale molecular design to innovations in manufacturing, and ultimately the development trends in the applications of polyimine and its composite materials.
Polymers 17 01607 g017
Table 1. Characteristics of polyimine fabrication methods.
Table 1. Characteristics of polyimine fabrication methods.
MethodProcess StepsAdvantagesLimitationsApplicable FormsRef.
Solution CastingMonomer dissolution → Schiff base reaction → solvent evaporationUniform filler dispersion, film formationHigh energy consumption, solvent recovery challengesFilms, gels, composites[20,21]
Hot PressingPowder compaction → heating/pressure → moldingScalability, simplicityParticle aggregation, weak interfacesBulk materials, multilayers[22,23]
Interfacial PolymerizationMonomer injection → interfacial condensation → film peelingUltrathin films, stable interfacesHarsh reaction conditions, high cost2D films, functional coatings[24,25]
Table 2. Comparison of representative properties of organic and inorganic reinforcing phases in polyimine composites.
Table 2. Comparison of representative properties of organic and inorganic reinforcing phases in polyimine composites.
ReinforcementTypeKey Property ImprovementsPrimary Characteristics EnhancedRef.
Bamboo Powder (BP)Organic
-
Tensile strength: 45.2 MPa (↑17% vs. neat PI)
-
Young’s modulus: 1.46 GPa (↑194% vs. neat PI)
Toughness and stiffness[33]
Cellulose Nanofibrils (CNFs)Organic
-
Tensile strength: ↑25% (30 wt% CNF)
-
Young’s modulus: ↑100% (maximum)
Stiffness[39]
Carbon Nanotubes (CNTs)Inorganic
-
Electrical conductivity: ↑4 orders of magnitude (10 wt% MWCNT)
-
Tensile strength: 74 MPa (↑34% vs. neat PI)
Conductivity and strength[40]
Graphene Nanoplatelets (GnPs)Inorganic
-
Tensile strength: 76.7 MPa (↑7% vs. neat PI)
-
Flexural strength: 98.2 MPa (↑22% vs. neat PI)
Strength and flexural properties[41]
“↑” means “increase”.
Table 4. Engineering applications of polyimine composites (20 to 25 years).
Table 4. Engineering applications of polyimine composites (20 to 25 years).
Abbreviation of CompositeReinforcement PhaseComposite MethodKey Performance ParametersApplication FieldSpecific Engineering Application CasesRef.
MBS-TMPMoS2, Te nanowiresFluid-designed solution shearingStable photocurrent (retaining 94.3% after 50,000 bending cycles); high carrier mobilityFlexible electronicsWearable photodetectors (integrated into textiles for imaging and sensing)[65]
PI/GPGraphite–polypropylenePhysical mixingSelf-healing efficiency of 94.6%; output voltage of 1325 V (15 N, 6 Hz)Energy harvestingSelf-healing triboelectric nanogenerator (TENG) for biomechanical energy collection and wireless transmission[66]
DIWMulti-walled carbon nanotubesPhysical mixing + partial curingElastic modulus of 520 MPa; tunable conductivity3D-printed electronic devicesRecyclable 3D-printed sensors (temperature/strain sensors maintaining performance after repair)[67]
BNNS/PIH TIMBoron nitride nanosheetsHorizontal centrifugal castingIn-plane thermal conductivity of 7.69 W/m·K; low compressive strength (2.16 MPa)Electronic thermal managementHeat dissipation materials for 5G base stations and automotive IGBT modules (reducing temperature fluctuation under vibration)[68]
PETG/CNF-IminePolyimine-coated cellulose nanofibers (CNF-Imine)Melt blending + hot pressingTensile strength increased by 12% (with 30 wt% fibers); strong interfacial bondingGreen materialsIndustrial manufacturing requiring high strength and stiffness[39]
PLA/CNF-IminePolyimine-coated cellulose nanofibers (CNF-Imine)Melt blending + hot pressingSignificant stiffness enhancement (up to 30% across fiber range); moderate increases in storage and loss moduli compared to PETG compositesGreen materialsApplications requiring stiffness with biodegradable materials[39]
PI-MWCNTMulti-walled carbon nanotubesChemical linking97% conductivity recovery; tensile strength of 74 MPa (34% improvement over matrix)Flexible electronicsWearable electronic skins (maintaining conductivity after multiple repairs and recycling)[40]
Polyimine-Metal ComplexMetal ions (Cu2+, Mg2+, Fe3+)One-pot methodTensile strength: 2.51 MPa; elongation at break: 1158%; self-healing efficiency: 96.0%; thermal stability (Td5%) > 278 °CGreen materialsReprocessable and recyclable vitrimers with enhanced thermal, mechanical, solvent, and acid resistance, though poor water resistance[69]
PI-aGAAnnealed graphene aerogelIn situ polymerization + hot pressingTensile strength of 37 MPa; stable conductivity after repeated reshapingFlexible electronicsHigh-conductivity composites for flexible circuits and smart wearable devices[70]
TMPNPTiO2@MXeneChemical linkingLimiting oxygen index (LOI) of 32%; resistance variation rate of −95% (at 70 °C)Fireproof materialsSmart building fireproof coatings (cotton fabrics integrated with wireless fire alarms)[48]
Ionic-PIsBMIM∙PF6Physical mixingTensile strength: 51.8 MPa; elastic modulus: 0.84 GPa; toughness: 1.62 MJ/m3; LOI > 27%Fireproof materialsMaterial recovery rate exceeding 41.9%; regenerated PI-R membranes retain mechanical strength; enhanced flame retardancy via BMIM∙PF6 sacrificial decomposition[71]
HVP/D230-CFCarbon fibersHot pressingTensile strength of 184.4 MPa; LOI of 34.2%Fireproof materialsRecyclable carbon fiber-reinforced composites for aerospace flame-retardant structures[72]
PI-NdFeB Soft RobotMagnetic NdFeB microparticlesPhysical mixing + solvent evaporationTensile strength of 6.3 MPa; elongation at break of 260%Biomedical applicationsMagnetic soft robots designed for minimally invasive surgical procedures and targeted drug delivery applications[46]
CF-PI-daMINTMechanically interlocked nanotube derivatives (MINTs)Planetary ball milling + hot pressingTensile strength: 68 ± 9 MPa; Young’s modulus: 3.2 ± 0.2 GPaGreen materialsEfficient reinforcement of PI CANs; superior mechanical performance of PI-daMINT composites due to enhanced SWNT dispersion and load transfer[45]
CFRFPCarbon fibersPrepreg compression moldingTensile stress of 23.7 MPa; water-driven extensibilityGreen materialsWeldable/self-healing carbon fiber composites for lightweight automotive parts and recyclable drone structures[18]
PI-SiCwSilicon carbide nanowhiskersHot pressingImpact strength improved by 154% (with 2% SiCw); flexural strength of 85.55 MPaGreen materialsHigh-performance electronic packaging materials (high-temperature and impact resistance)[73]
GNPs-PGraphene nanoplateletsHot pressingTensile strength of 73.05 MPa (at 0.5 wt%); enhanced thermal conductivityElectronicsLED heat dissipation substrates (replacing traditional metal heat sinks)[41]
BP/PIBamboo powderHot pressingTensile strength of 45.2 MPa; closed-loop recyclability of 100%Green materialsBamboo-based plastic alternatives for eco-friendly packaging and furniture[33]
RY-PIRamie yarnChemical linking + interlayer compoundingTensile strength: 144 MPa; Young’s modulus: 0.97 GPa; elongation at break: 25%Green materialsHigh-performance and recyclable natural fiber-reinforced plastic composites (NFRPCs)[53]
TDGraphene nanoplatelets (GnPs)Hot pressingTensile strength: 84 MPa; Young’s modulus: 1.6 GPa; thermal conductivity: 1.8 W m−1 K−1ElectronicsTDG-sn composites for LED chip heat dissipation and thermal management applications[74]
CFRPCarbon fibersChemical recyclingTg decreased by 42% in non-woven mat CFRPs; storage modulus (E’) significantly increased; Tg increased by 8% in UD-CFRPGreen materialsFeasibility of recycling fiber-reinforced vitrimer composites, with required process optimizations[75]
VITRIMER TEDAPUnidirectional carbon fibersHot pressingTensile strength: 69 MPa; Tg: 192 °CFireproof materialsEnhanced thermal stability and flame retardancy compared to epoxy systems; significant pHRR reduction in vitrimer composites[76]
PI-FRPs.Carbon fibersPowder compression moldingYoung’s modulus increased to ~17.3 GPa at 190 °C (1.25 MPa, 4 min)Repairable coatingsLow-temperature, mold-free in situ repair of polyimine composites, adapted for curved surface applications[77]
PPCsCellulose paperHot pressingTensile strength: 71 MPa; Young’s modulus: 3.2 GPaGreen materialsPolyimine-filled cellulose papers (PPCs) with excellent mechanical strength, water resistance, gas barrier properties, and recyclability[43]
WPCsWood celluloseHot pressingTransition from brittle to ductile fracture; 13% tensile strength improvement in WPC/25pAPPFireproof materialsImproved interfacial interaction and mechanical strength due to imine bond network and mechanical interlocking[78]
Table 5. Comparison of mechanical, thermal, and recycling properties of high-performance thermosetting polymers.
Table 5. Comparison of mechanical, thermal, and recycling properties of high-performance thermosetting polymers.
PolymerMechanical PropertiesThermal PropertiesRecyclability
Tensile StrengthTgTd5%Recyclable
Epoxy55–85 MPa105–120 °C250–350 °C×
Phenolic Resin30–60 MPa100–150 °C260–350 °C×
Bismaleimide (BMI)60–120 MPa230–350 °C400–500 °C×
Polyimine (PI)30–96 MPa−20–100 °C250–400 °C
Table 6. Comparison of pHRR parameters between polyimine vitrimer and PER epoxy resin matrix (3% phosphorus content) [76].
Table 6. Comparison of pHRR parameters between polyimine vitrimer and PER epoxy resin matrix (3% phosphorus content) [76].
ParameterPer TedapVitrimer AppVitrimer RdpVitrimer Tedap
pHRR (kW/m2)111175290218
Time to pHRR (s)110234207260
ParameterPer CompositeVitrimer CompositeVitrimer App CompositeVitrimer Rdp Composite
pHRR (kW/m2)351289186152
Time to pHRR (s)39140158176
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, X.; Zhang, S.; Chen, Y. Polyimine-Based Self-Healing Composites: A Review on Dynamic Covalent Thermosets for Sustainable and High-Performance Applications. Polymers 2025, 17, 1607. https://doi.org/10.3390/polym17121607

AMA Style

Wang X, Zhang S, Chen Y. Polyimine-Based Self-Healing Composites: A Review on Dynamic Covalent Thermosets for Sustainable and High-Performance Applications. Polymers. 2025; 17(12):1607. https://doi.org/10.3390/polym17121607

Chicago/Turabian Style

Wang, Xiaoxue, Si Zhang, and Yun Chen. 2025. "Polyimine-Based Self-Healing Composites: A Review on Dynamic Covalent Thermosets for Sustainable and High-Performance Applications" Polymers 17, no. 12: 1607. https://doi.org/10.3390/polym17121607

APA Style

Wang, X., Zhang, S., & Chen, Y. (2025). Polyimine-Based Self-Healing Composites: A Review on Dynamic Covalent Thermosets for Sustainable and High-Performance Applications. Polymers, 17(12), 1607. https://doi.org/10.3390/polym17121607

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

Article Metrics

Back to TopTop