Next Article in Journal
The Origin of KO-KUTANI Porcelain—Part II: The Unearthed Secrets of the Hakuji Shallow Bowl
Previous Article in Journal
Experimental Study on Anti-Crystallization Performance of Tunnel Drainage Pipes Based on Magnetic Powder Effect
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress on the Healing Mechanisms of Self-Healing Superhydrophilic Surfaces

1
College of New Energy & Materials, Northeast Petroleum University, Daqing 163318, China
2
College of Chemistry and Chemical Engineering, Northeast Petroleum University, Daqing 163318, China
*
Author to whom correspondence should be addressed.
Coatings 2025, 15(9), 1006; https://doi.org/10.3390/coatings15091006
Submission received: 21 July 2025 / Revised: 20 August 2025 / Accepted: 27 August 2025 / Published: 31 August 2025
(This article belongs to the Section Surface Characterization, Deposition and Modification)

Abstract

Superhydrophilic surfaces have important applications in fields such as energy, military, and medicine due to their unique wettability. However, the micro-/nano-structures of superhydrophilic surfaces are fragile and prone to damage, which can cause them to lose their superhydrophilicity and reduce their service life, severely limiting their applications. This paper discusses recent research progress and self-healing mechanisms of self-healing superhydrophilic surfaces from the perspectives of composition and structure self-healing. Additionally, it also introduces the research progress of superhydrophilic surfaces healed in air and underwater environments. Finally, the limitations of the self-healing superhydrophilic surfaces are summarized, and perspectives on future development are discussed.

1. Introduction

Superhydrophilic surfaces can improve the energy utilization efficiency of systems; reduce energy loss as a protective barrier and medium for interaction with the external environment; and have important applications in marine pollution prevention, transportation resistance reduction, membrane separation technology, biomedicine, photovoltaic power generation, catalysis, new energy, and other fields [1,2,3,4]. However, the micro-/nano-structures and hydrophilic composition of superhydrophilic surfaces suffer from insufficient mechanical and chemical stability, making it extremely important to improve surface durability. This can not only extend the surface service life but also reduce energy consumption caused by maintenance [5]. However, surface damage is inevitable, such as cracks, fractures, and even local detachment in extreme cases during use, which reduces surface durability and greatly limits its application.
At present, the improvement of superhydrophilic surface durability mainly focuses on two aspects: mechanical durability and chemical durability. Mechanical durability mainly focuses on improving the mechanical properties of surface structure. For example, Zhao et al. [6] prepared a composite micro-/nano-superhydrophilic surface using laser etching method, which has excellent durability characteristics and can withstand a wear distance of more than 6 m. Chemical durability mainly focuses on improving the stability of surface chemical composition [7]. For example, Zhuang et al. [8] deposited hydroxides of Fe, Co, and Ni on amino grafted carbon substrates by electrodeposition to prepare superhydrophilic surfaces with alkaline and salt resistance. However, the superhydrophilicity may decrease once the structure of the surface is damaged by external forces. Therefore, how to heal the damaged surface structure is crucial for extending the service life of superhydrophilic surfaces.
Inspired by the self-healing function of organisms, endowing the surface with a self-healing property can effectively enhance its durability. The preparation of self-healing superhydrophilic surfaces is usually to introduce dynamic reversible bonds or substances capable of restoring damaged areas, enabling the surface to autonomously restore its structure integrity after physical damage. Concurrently, by grafting hydrophilic chemical groups and constructing micro-/nano-scale rough structures, the high surface energy and extremely low water contact angles are ensured, thereby achieving the superhydrophilic effect. Common preparation methods include coating, surface functionalization, grafting modification, layer-by-layer assembly, and in situ synthesis. The preparation of self-healing superhydrophilic surfaces typically combines these methods synergistically to endow the surface with durable anti-fouling, anti-fogging, and self-cleaning properties, significantly extending its service life.
Some researchers have reported studies of self-healing superhydrophilic surfaces, but few papers have reported the progress of superhydrophilic surfaces in integrating fabrication methods, damage categories, ways of healing, and respective advantages and disadvantages. In this review, we summarize the latest research progress in self-healing superhydrophilic surfaces based on the basic principle of wettability, focusing on the composition self-healing and structure self-healing (Figure 1). The advantages and disadvantages of self-healing strategies are also summarized, followed by a look forward to their future development.

2. Thermodynamic Principle of Superhydrophilic Surfaces

Under the idealized assumption of a perfectly smooth and flat surface, Young’s equation [9] could quantitatively describe the contact angle θ at the three-phase interface of a droplet based on the balance of the surface free energy:
cos θ = γ s γ sl γ l
where γs is the surface tension of the solid, γsl is the solid–liquid interfacial tension, and γl is the surface tension of liquid (Figure 2).
Thus, the wettability of the solid surface can be classified into the following four categories based on the contact angle θ:
(1)
When θ = 0°, the liquid completely spreads across the solid surface, achieving a full wetting state.
(2)
When 0° < θ < 90°, the liquid spreads across a finite contact area on the solid surface, presenting partial wetting state.
(3)
When 90° < θ < 180°, the liquid contracts into a bead-like shape on the surface the solid surface, exhibiting a non-wetting state.
(4)
When θ = 180°, the liquid forms a perfect sphere on the solid surface, demonstrating a perfect non-wetting state.
However, the classical Young’s model is only applicable to ideally smooth surfaces. To expand its practical application, in 1936, Wenzel [10] established a quantitative relationship between the apparent contact angle θr and the intrinsic contact angle θ of a rough surface based on the thermodynamic equilibrium principle
cos θ r = rcos θ
where r is the surface roughness factor, which is defined as the ratio of the real solid–liquid contact area to the apparent projection area. This model introduced the surface morphology parameter into the wetting theory system for the first time. In 1944, Cassie and Baxter [11] proposed different modifications to Young’s equation by investigating the non-homogeneous surface wettability of two different rough chemical materials in terms of the ratio of the solid surface to the gas surface at the three-phase junction:
cos θ c = f 1 cos θ + f 2 cos θ g
where f1 and f2 denote the solid–liquid and solid–gas contact area fractions, respectively, which satisfy f1 + f2 = 1. θ is the solid–liquid contact angle, and θg is the solid–gas contact angle. Compared with Wenzel’s formula, Cassie’s formula more accurately describes the real situation of the system interface.
Systematic comparison of the two modified models shows that surface roughening can significantly enhance the intrinsic wetting characteristics of the material [12].
For the design of superhydrophilic and underwater superoleophobic materials, the contact angle of oil in air and the contact angle of oil underwater have been formulated according to Young’s equation:
cos θ o = γ s v γ s o γ o v cos θ ow = γ s w γ s o γ o v
γsv, γso, and γov denote the interfacial tensions at the solid–gas, solid–oil, and oil–gas interfaces, respectively, and γsw and γow correspond to those at the solid–water and oil–water interfaces. Combining the above equations and Young’s equation, the underwater contact angle can also be written as
cos θ ow = γ o v cos θ o γ w v c o s θ w γ o v
As the superhydrophilic surface is immersed in water, water spreads rapidly over it, forming a stable film of water, thus isolating it from the oil phase. Hence, hydrophilic surfaces are usually oleophobic [13].
During the dynamic spreading of the liquid film, the apparent contact angle at the three-phase contact line changes: the contact angle θ gradually evolves from an initial obtuse angle to an acute angle as the free energy of the solid–liquid interface continues to decrease. When θ < 10° [14], the solid–liquid interaction energy significantly exceeds the free energy of the liquid–gas interface, and the surface is superhydrophilic.
The hydrophilicity of the surface is primarily influenced by hydrophilic components and surface micro-/nano-structures. The composition determines whether the surface is hydrophilic, while the micro-/nano-structures enhance surface wettability, such as porous structures and hierarchical structures. Typically, hydrophilic surfaces are influenced by hydrophilic composition and exhibit high surface energy. These hydrophilic components primarily consist of substances containing polar functional groups such as hydroxyl, carboxyl, amino groups, and other partially oxygen-containing functional groups. For example, Wang et al. [15] used ultraviolet-visible light to irradiate ZnO@stearic acid nanoarrays prepared on a Zn substrate to excite electron–hole pairs. It will continuously oxidize and reduce the adsorbed oxygen and water on the surface, form hydroxyl radicals adsorbed at oxygen vacancies on the ZnO surface, create stable hydrophilic sites, and induce the surface to transition from superhydrophobic to superhydrophilic. Additionally, Yu et al. [16] combined NH2–MIL-125(Ti) with reduced graphene oxide (RGO) and modified it with polydopamine (PDA) to prepare a superhydrophilic membrane containing polar functional groups such as carboxyl and amino groups. Based on in-depth studies of hydroxyl and carboxyl groups, it was found that other oxygen-containing functional groups also exhibit hydrophilic properties, such as carbonyl and sulfonic acid groups. Hung et al. [17] functionalized graphene nanoplates (GNPs) with oxygen-containing groups, enriching their surfaces with hydrophilic groups such as hydroxyl, carboxyl, and carbonyl groups, thereby synthesizing a superhydrophilic surface with a water contact angle of 0° and exhibiting ultra-fast water penetration performance. For hydrophilic surfaces, constructing micro-/nano-structures, such as porous structures or micro-nano hierarchical structures, can enhance surface hydrophilicity. For example, Bastami et al. [18] used ZrO2 films as raw materials to investigate how different temperatures affect surface roughness, verifying that increasing surface roughness within a certain range enhances the hydrophilicity of hydrophilic materials. Liu et al. [4] successfully prepared a robust superhydrophilic surface by spraying the mixture of PEI and TiO2 nanoparticles onto an epoxy resin surface to alter surface roughness, thereby effectively enhancing the substrate hydrophilicity and stability.

3. Self-Healing Mechanism of Superhydrophilic Surfaces

It is difficult for superhydrophilic surfaces to withstand long-term physical wear or chemical erosion. To address this challenge, researchers have innovatively developed superhydrophilic surfaces with self-healing capabilities inspired by the healing mechanisms of biological tissues [19]. Upon damage, the pre-designed healing units are activated to facilitate autonomous self-healing via two reversible dynamic processes: (i) composition self-healing based on the composition recovery and (ii) structure self-healing based on the reformation of chemical bonds via polymer chain migration and reorganization. This innovative self-healing mechanism ensures the surface maintains superhydrophilic properties and interfacial function stability, even causing structure defects such as cracks, pores, and so on. Furthermore, the hydrogels are formed by hydrophilic polymer chain networks [20], and they can quickly absorb the liquid droplets added onto their surface, presenting a state of rapid spreading on the surface and having superhydrophilic properties [21]. Therefore, some self-healing hydrogels are classified as superhydrophilic surfaces for introduction in this paper.

3.1. Composition Self-Healing

The self-healing phenomena inherent to biological systems have inspired the development of self-healing materials. By facilitating the regeneration of surface compositions, these materials can restore the composition of the surface, recover functional properties, and reestablish surface wettability, thereby enhancing durability. Srubar et al. [22] significantly enhanced the self-healing performance of active building materials by combining the physical re-crosslinking of hydrogels with the mineralization effects of different microorganisms. When material surfaces are damaged, rapid healing of the damaged areas can be achieved through biomineralization using the CO2 concentration mechanism of photosynthetic microorganisms or through the hydrolysis of urea by urease-producing bacteria to produce CaCO3. To extend the application to underwater environments, Chen et al. [23] obtained a high-strength underwater superoleophobic surface through a synergistic combination of superspreading and biomineralization techniques. The wettability was found that could be recovered by the re-biomineralization process after mechanical wear.
Additionally, by employing the polymer-functionalized microgel spheres as the templates, Chen et al. [24] achieved surface hydrophilicity recovery through volumetric expansion of the microgel matrix coupled with interfacial reaggregation of hydrophilic polymer chains at the liquid/solid interface (Figure 3). Compared with the local self-healing of microgel beads, the self-polishing strategy can effectively address large-scale surface scratches and damage. For example, Kang et al. [25] developed a self-polishing multi-layer coating material based on dextran and chitosan, achieving interlayer bonding through imine bonds between dextran aldehyde and carboxymethyl chitosan. When the surface undergoes biofouling due to bacterial adhesion, it creates a local acidic microenvironment, which promotes the breaking of imine bonds to accelerate surface self-polishing and restore the coating’s anti-fouling and antibacterial properties. Additionally, self-polishing polymer/SiO2 composite surfaces containing hydrolytically cleavable side chains were fabricated by molecular engineering strategies, as reported by Wang et al. [26]. Autonomous wettability restoration was demonstrated via marine environment-triggered hydrolytic cleavage of siloxane ester linkages at the composite–liquid interface.
However, once the surface suffers structure damage, the composition self-healing cannot effectively restore the damaged structure, leading to the loss of surface wettability. Therefore, how to heal surface structure damage is extremely important for maintaining the anti-fouling performance of superhydrophilic surfaces.

3.2. Structure Self-Healing

The exposure of material surfaces to external mechanical stresses can induce structural damage, resulting in diminishing surface hydrophilicity. Therefore, a comprehensive understanding of surface restoration mechanisms and development of effective strategies for surface structure self-healing is of great necessity. The restoration of both micro-structural defects and macrostructural integrity of material surfaces has been demonstrated to effectively enhance its longevity, reducing maintenance frequency while improving structural safety and system energy efficiency [27].
Currently, the structural self-healing mechanisms of superhydrophilic surfaces can be classified into two categories, including extrinsic type and intrinsic type. The difference between these two categories lies in the need of adding external healing agents to complete the self-healing progress. The self-healing microcapsules are the primary type of extrinsic self-healing [28]. The intrinsic type, on the other hand, can be classified into electrostatic interaction reconfiguration, hydrogen bonding reorganization, host–guest recognition, and metal–ligand interactions according to the types of interactions.

3.2.1. Extrinsic Type

The self-healing microcapsule is accomplished by constructing a core–shell microcapsule containing a healing agent, and then the microcapsule ruptures upon reaching a specific environmental trigger such as when the local stress field reaches the critical strength, releasing the healing agent to accomplish autonomous self-healing for the damaged area [29]. For instance, Fu et al. [30] used chitosan and sodium tripolyphosphate as primary materials to fabricate a microcapsule, and the restorative agents of fluorosurfactant FS-60 and dopamine hydrochloride were encapsulated. When the material surface undergoes mechanical abrasion, washing, and chemical erosion, the surface superhydrophilicity can be restored by soaking the material with distilled water followed by heating to induce partial degradation of the microcapsule and release restorative agents. Since this coated fabric exhibits excellent performance in separating oil–water mixtures, it holds broad application prospects in the field of oil–water separation. However, limited by the spatial distribution density of the pre-embedded microcapsules, the healing performance of this method is greatly constrained, leading to a strong confinement on the times of localized self-healing and making it difficult to achieve multiple times of healing at localized areas (Figure 4). In addition, the introduction of larger-sized microcapsules can disrupt the overall homogeneous structure of the material, causing defects and thereby reducing its strength and resistance to impacts [31].

3.2.2. Intrinsic Type

The intrinsic-type self-healing surfaces rely on reversible bonding interactions within the material itself to achieve autonomous healing [32]. Such a type of self-healing strategy is theoretically capable of undergoing multiple healing cycles at the same location, enhancing the durability and longevity of the material [33,34,35]. Compared with extrinsic self-healing strategies, surfaces with intrinsic self-healing properties typically exhibit more homogeneous structure.
Electrostatic Interaction
Electrostatic interactions help the establishment of reversible crosslinks between polymer chains through ion–dipole or dipole–dipole interactions [36] (Figure 5a). When scratches or cracks emerged on the surface, external environmental stimulations can induce the happening of electrostatic effects leading to the reorientation of molecular chain segments or functional groups and repair the damaged structure through breakage-reassembly mechanisms of non-covalent bonds [37].
Hozumi et al. [38] prepared a hydrophilic composite membrane using amine-mineralized nanoclay particles (AMP-NCPs) and polyvinylpyrrolidone (PVP). Upon surface scratching, the electrostatic interactions between the positively charged amino groups on the AMP-NCPs and the negatively charged groups of PVP facilitate the self-healing process in humid environments, restoring the initial membrane performance. Furthermore, combining the synergistic effect of the nanoparticles, internal electrostatic interactions, and hydrogen bonding is also a good strategy to improve the mechanical strength and durability while maintaining surface self-healing capability [39] (Figure 5b).
Figure 5. (a) Schematic diagram of the electrostatic interaction self-healing mechanism [36]. (b) Internal structure of the superhydrophilic coating [39]. (a) Reprinted with permission from ref. [36]. Copyright (2022) (Elsevier). (b) Reprinted with permission from ref. [39]. Copyright (2019) (American Chemical Society).
Figure 5. (a) Schematic diagram of the electrostatic interaction self-healing mechanism [36]. (b) Internal structure of the superhydrophilic coating [39]. (a) Reprinted with permission from ref. [36]. Copyright (2022) (Elsevier). (b) Reprinted with permission from ref. [39]. Copyright (2019) (American Chemical Society).
Coatings 15 01006 g005
Hydrogen Bond
Hydrogen bonding interactions arise from the attraction and repulsion between molecules, including electrostatic, polarization, charge transfer, dispersion, and exchange mutual exclusion [40], mainly resulting from the interaction between hydrogen atoms and electronegative heteroatoms (F, O, and N) that possess lone electron pairs. As a highly efficient supramolecular interaction, it plays a critical role in the self-healing progress of superhydrophilic surfaces, particularly owing to its reversible and reconstructive nature.
Fan et al. [41] used polyvinyl alcohol (PVA) as the matrix material and salicylic acid (SA) as the cross-linking agent to establish a hydrogen-bonding network to achieve self-healing of damage. However, the healing rate of the coating is slow. It requires a small amount of water existing in the damaged location at the temperature of 85 °C for 15 min to complete the healing progress.
To address this issue, Lai et al. [42] fabricated a self-healing superhydrophilic coating with anti-fog and anti-freezing properties by spinning a mixed solution of polyacrylic acid (PAA), carboxymethyl cellulose (CMC), and Fe (III). The interaction between the carboxyl group of PAA and the hydroxyl group of CMC forms a hydrogen bond structure, which can achieve rapid self-healing of scratches within 1 min.
You et al. [43] developed a composite coating by mixing surfactant Triton X-100 into the epoxy resin/amine hardener. This coating exhibits exceptional resistance against a series of environmental challenges, including acids, alkalis, salts, high temperatures, and ultrasonic waves. Moreover, this surface could effectively heal the damages caused by scratches, cuts, and exposure to polar solutions such as tetrahydrofuran (THF), which thus successfully addressed the long-standing limitation of application scenarios for conventional superhydrophilic surfaces. The robust performance is rooted in the hydrogen bonding interactions established between the hydroxyl groups of Triton X-100 and the ether group of the epoxy resin (Figure 6).
Host–Guest Interaction
Generally, the host molecules form inclusion complexes with guest molecules, which include polymers, organic molecules, and biomolecules. The formation of these complexes is typically driven by non-covalent interactions, such hydrophobic interactions, hydrogen bonding, and π–π stacking [44,45]. Zhang et al. [46] used Fe3O4@cucurbituron (CB) nanoparticles as the cross-linking agent to fabricate nanocomposite hydrogels. The gelation happened under the interaction of the CB cavity with the polymer formed by acrylamide and guest molecules of 1-benzyl-3-vinylimidazole. Nie et al. [47] demonstrated that the host–guest interactions between β-cyclodextrin-modified silk fibroin (CD-SF) as the host polymer and adamantane-modified silk fibroin (AD-SF) as the guest polymer, combined with hydrogen bonding interactions involving quaternary ammonium chitosan (QCS), enable rapid self-healing of damaged hydrogels following cutting injuries. Additionally, functionalized MoS2 can be utilized to significantly improve the mechanical properties of hydrogels without losing self-healing properties [48] (Figure 7).
Metal–Ligand Interaction
The formation of coordination bonds typically accompanied the increase in entropy and the decrease in enthalpy. Hence, the metal–ligand interaction is a dynamic noncovalent bond with high efficiency and capable of spontaneous self-healing without any external stimulations [49] (Figure 8). Recent studies have demonstrated that both zinc (II) and nickel (II) ions serve as effective central metals in complexes of self-healing superhydrophilic surfaces. Notably, pyridine analogs exhibit superior coordination abilities with zinc(II) compared with imidazole or histidine, leading to the formation of more stable and more effective self-healing chelating ligands, while the nickel (II) in comparison to zinc (II) has even better performance [50,51]. Furthermore, Pourjavadi et al. [52] developed interpenetrating network (IPN) hydrogels by polymerizing acrylates and acrylamides in gelatin with the addition of the cross-linking agent of Fe (III)). Such hydrogel exhibits superior mechanical strength, along with controlled degradation and sustained self-healing efficiency even after being cut.
The aforementioned four mechanisms could all facilitate the self-healing of damage at the same location for multiple times. However, it should be noted that they exhibit significant differences. Electrostatic interactions respond fast while they are susceptible to the environmental factors. Hydrogen bonding has a simple structure, which allows for easy synergy with a variety of interactions to improve the surface stability, while it is easy to be influenced by water molecules. The metal–ligand interactions are fast but severely affected by other ions to disrupt the healing progress. Part of the host–guest interactions are insensitive to the water, but most of the surfaces with host–guest molecules have some problems such as low mechanical properties due to the low bonding energy; sensitivity to the healing environment in terms of temperature, pH, and so on; a low healing rate limiting the dynamics; a complex synthesis route; and high cost. Table 1 summarizes different researches under various self-healing mechanisms in recent years.
Reversible Dynamic Covalent Bond
Dynamic covalent bonds are a class of chemical bonds that can undergo reversible exchange under certain conditions (e.g., light, heat, humidity stimulation, etc.) [54]. Due to their reversibility and condition-responsive nature, they exhibit significant application potential in fields such as materials science and biomedicine. As classic types of dynamic covalent bonds, the hydrazone bonds, imine bonds, borate ester bonds, and disulfide bonds are widely studied. Zhou et al. [55] utilized the reaction between the aldehyde group of aldehyde hyaluronic acid (AHA) and the imide group of 3,3′-dithiobis (propionyl hydrazide) (DTP) to form dynamic covalent imide bonds, which constitute the multiple types of reversible interactions of the hydrogel network. When the hydrogel is damaged by external forces, the hydrazone bonds break due to mechanical stress, leading to local network disintegration. After removing the external force, the aldehyde groups of AHA and the hydrazone groups of DTP recontact through molecular diffusion, rapidly reforming the hydrazone bonds at room temperature to achieve network healing.
Some dynamic covalent bonds, such as hydrazone bonds, are unstable and can reversibly form and break under acid catalysis. Therefore, in certain environments, such as pH changes or exposure to moisture, unintended bond breakage may occur, potentially affecting the material’s long-term stability and the reliability of self-healing. However, some dynamic covalent bonds are relatively stable, such as disulfide bonds, whose breaking and reformation typically require specific conditions such as redox reactions, thermal activation, or UV light irradiation. Therefore, they exhibit relatively high stability and are unlikely to break unexpectedly in non-catalyzed environments, effectively addressing the instability issue of hydrazone bonds.
For example, Jeon et al. [56] prepared a polysulfide bond hydrogel using 2,3-dithiol-1-propanol, racemic 2,3-dithiolbutanedioic acid, and hydrobromic acid as raw materials. When subjected to cutting damage, the hydrogel can achieve complete fusion at the cutting interface in approximately 5 s in air and approximately 6 s underwater, approaching the initial mechanical properties. Based on this theory, researchers have expanded the application areas of this self-healing hydrogel. For example, Guo et al. [57] used thioglycolic acid containing disulfide bonds as a green cross-linking agent to design and construct an in situ self-healing cross-linked polymer gel system with a dynamic covalent adaptive net-work and dynamic disulfide bonds, endowing the hydrogel with excellent toughness and self-healing performance, with a healing efficiency of up to 88% after cutting damage, making it suitable for high-temperature, high-salinity reservoir environments.
Compared with dynamic non-covalent bond networks based on weak interactions, dynamic covalent network structures connect polymer chains via reversible covalent bonds, showing superior mechanical strength and chemical stability to the material. Additionally, since dynamic covalent bonds and dynamic non-covalent bonds exhibit complementary advantages during the healing process, recent research has focused on combining them together to ensure both mechanical strength and rapid response after damage, significantly improving the material’s healing efficiency.
Lu et al. [58] utilized the synergistic effect of dynamic acylhydrazone bonds and host–guest interactions to achieve rapid healing of damaged hydrogels while also conferring hydrolysis resistance and functional stability for over 30 days. Li et al. [59] utilized the synergistic effect of dynamic disulfide bonds and multiple supramolecular interactions to accelerate the healing process. Disulfide bonds participate in the reconstruction of the network structure via reconnecting of broken disulfide bonds, ensuring the stability of the hydrogel, while supramolecular interactions assisted in driving molecular chains closer and reconnecting in the damaged region, thereby realizing the self-healing process of damaged hydrogels.
In recent years, superhydrophilic surfaces have gradually become a hot topic of research. However, the service life of surfaces is greatly reduced due to issues such as damage caused by external impacts, severely limiting their application. The self-healing superhydrophilic surfaces can effectively address this issue. With their unique wettability and durability, they have broad application prospects in areas such as self-cleaning coatings, biomaterials, oil–water separation, microfluidic devices, textiles, and anti-corrosion coatings.

4. Self-Healing in Air and Water

The environment plays a key role in the self-healing property, even determining whether it will happen or not. Most of the self-healing superhydrophilic surfaces demonstrate remarkable healing efficiency of damage in ambient environments. However, their self-healing performances are found to degrade significantly when placed in water, which leads to inferior interfacial stability and lowering the cyclic numbers of self-healing. Hence, the development of superhydrophilic surfaces with stable self-healing efficiency in water has attracted extensive research interest.

4.1. Self-Healing in Air

The superhydrophilic surface is usually healed in the ambient environments, where the presence of water molecules does not interfere with the reconstruction of the dynamic bonding network, thereby allowing the autonomous self-healing process of damaged surfaces. For example, Fu et al. [60] designed a nano-composite hydrogel based on a dual-network structure. At room temperature, simply keeping the two parts of the slice in contact allows for self-healing within 30 min under the dynamic action of borate bonds. To further reduce the healing time, Wang et al. [61] prepared a self-healing hydrogel based on 3D printing technology. Through the bonding and interaction of dynamic boron–diol bonds and hydrogen bonding interactions, the cut sample can be rapidly repaired within 30 s at room temperature without any external stimulation. Sun et al. [62] utilized the imine bonds between 4-formylphenylboronic acid (FPBA) and lysine (Lys) to construct a dynamic borate–imine–imine–borate bond structure between the PVA chains, which endowed the hydrogel with excellent tensile properties and ultra-fast self-healing capabilities without external stimulation.

4.2. Self-Healing in Water

The self-healing progress of conventional superhydrophilic surfaces faces three key challenges in water environments: (i) the mobility of water molecules limiting the transport of healing motifs; (ii) the imbalance of dynamic hydrogen bonding networks due to competitive hydration effects; and (iii) the presence of metal ions in water, which interferes with the chelation of metal–ligand bonds. With the growing demand for self-healing superhydrophilic surfaces in biomedicine, underwater protection devices, and other fields, researchers are conducting more in-depth studies on underwater self-healing strategies [63,64,65,66].
For example, researchers have found that energy can be delivered to the vicinity of the crack using NIR irradiation, which could melt the surface coating and triggers the movement of the surrounding polymer chains toward the crack [67]. The reconnection of the fracture site can also be achieved by exploiting the attraction of charges of the equal and opposite sign between dipoles [68]. Instead of reversible bonding, strong dipole interactions can effectively avoid the interference of water molecules, such as saturation of hydrogen bonding, coordination of metal cations, and solvation of ions. It could maintain stable performance, even in the flowing water environment, for a long time. Jiang et al. [69] established a dual network structure relying on hydrogen bonding for crosslinking, which maintains excellent self-healing performance in both air and underwater environments. Jia et al. [70], inspired by the hydrophilic shell and hydrophobic core structure in nature, realized rapid self-healing in various environments such as air, water, seawater, and sweat by constructing a hydrophilic–hydrophobic synergistic network. Table 2 compares the contents of several studies on underwater self-healing superhydrophilic surfaces in recent years.

5. Conclusions

This article systematically reviews the latest research progress and self-healing mechanisms of superhydrophilic surfaces from the perspectives of composition self-healing and structure self-healing in recent years. For composition self-healing, superhydrophilic surfaces mainly achieve the restoration of wettability through strategies such as composition regeneration, re-biomineralization, re-enrichment of hydrophilic composition on surfaces, and self-polishing. However, the composition self-healing cannot effectively restore the surface wettability once the structure of the superhydrophilic surface is damaged, and the structure self-healing is highly desired. Structure self-healing of superhydrophilic surfaces is mainly divided into extrinsic and intrinsic types. Extrinsic structure self-healing mainly utilizes microcapsule technology to release the healing agent and restore the surface wettability, while intrinsic structure self-healing mainly utilizes dynamic non-covalent bonds, supramolecular interactions, and dynamic covalent bonds.
Although research in the field of self-healing superhydrophilic surfaces has made some progress, there are still many challenges ahead, with the following three areas being particularly prominent.
First, during the healing process, healing in air is mainly used, and how to heal in water remains a challenge, especially in complex underwater environments such as strong acids, strong alkalis, salts, and low temperatures. This is mainly because the healing conditions are crucial to the composition healing process. Once the healing conditions change, it is likely to lead to the failure of composition self-healing. For structure self-healing, healing conditions are equally crucial. This is mainly because the conventional dynamic healing process mainly relies on the reversible dynamic bonds, which are easily affected by pH, ions, water, etc., resulting in low healing efficiency or even failure. The environmental tolerance of healing can be enhanced by designing strong reversible dynamic bonds with anti-interference properties, such as dipole interactions, π–π interactions, hydrophobic interactions, etc., to reduce the influence of water molecules and various ions on the healing process. Combining multiple interactions such as hydrogen bonds and disulfide bonds to form dynamic crosslinking networks, the self-healing efficiency and stability of the surface in complex environments can be improved.
Second, the healing process is also prone to be limited by spatial conditions, and it is difficult to carry out the healing process in restricted spaces. By using various stimulus-responsive healing systems, such as light, heat, and electric fields, the spatial operability of the healing process can be improved. In addition, designing superhydrophilic surfaces with high strength and self-healing properties is also necessary, which can further enhance the service life of the surface. It will broaden the application of superhydrophilic surfaces in fields such as marine engineering and underwater equipment [71] by exploring multi-stimulus-responsive collaborative healing strategies, realizing the healing in restricted spaces, and achieving stable self-healing under extreme conditions.
Finally, composition self-healing can restore the surface composition, while structure self-healing can heal internal damage within the surface. By combining them together, full-scale surface healing can be achieved, restoring the surface wettability and mechanical properties and extending its service life.

Author Contributions

Conceptualization, Z.L. and F.L.; writing—original draft preparation and visualization, Z.L.; writing—review and editing, supervision and finding acquisition, F.L. All authors have read and agreed to the published version of the manuscript.

Funding

The research was funded by the National Natural Science Foundation of China (22205035).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Asha, A.B.; Chen, Y.J.; Narain, R. Bioinspired dopamine and zwitterionic polymers for non-fouling surface engineering. Chem. Soc. Rev. 2021, 50, 11668–11683. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, L.; Li, R.; Zheng, S.; Zhu, H.; Cao, M.; Li, M.; Hu, Y.; Long, L.; Feng, H.; Tang, C.Y. Hydrogel-embedded vertically aligned metal-organic framework nanosheet membrane for efficient water harvesting. Nat. Commun. 2024, 15, 9738. [Google Scholar] [CrossRef]
  3. Liu, M.; Wang, S.; Jiang, L. Nature-inspired superwettability systems. Nat. Rev. Mater. 2017, 2, 1–17. [Google Scholar] [CrossRef]
  4. Du, H.; Liu, F.; Wang, H. Bio-inspired robust superhydrophilic/underwater superoleophobic coating with lubrication, anti-crude oil fouling and anti-corrosion performances. J. Colloid Interface Sci. 2022, 616, 720–729. [Google Scholar] [CrossRef]
  5. Zhang, H.; Bu, X.; Li, W.; Cui, M.; Ji, X.; Tao, F.; Gai, L.; Jiang, H.; Liu, L.; Wang, Z. A Skin-Inspired Design Integrating Mechano–Chemical–Thermal Robustness into Superhydrophobic Coatings. Adv. Mater. 2022, 34, 2203792. [Google Scholar] [CrossRef]
  6. Wang, R.; Ma, W.; Zhao, J.; Guo, Z.; Cui, Y.; Zhu, Y.; Chen, H. Nanosecond laser etching for enhanced multi-scale layered micro-nano superhydrophilic 304 stainless steel surface. Opt. Laser Technol. 2024, 174, 110645. [Google Scholar] [CrossRef]
  7. Chen, Z.; Zhao, T.; Tan, Q.; Zuo, J.; Xu, S.; Nong, Y.; Wen, X.; Pi, P. A charged superhydrophilic-oleophobic anti-fouling sponges for rapid separation of O/W emulsions. Sep. Purif. Technol. 2024, 335, 126110. [Google Scholar] [CrossRef]
  8. Chen, Q.; Wang, K.; Li, S.; Wang, Y.; Lei, L.; Zhu, M.; Zhuang, L.; Xu, Z. Interfacial amine-assisted electrodeposition of superhydrophilic/superaerophobic metal hydroxides for robust oxygen evolution catalysis. Sci. China Chem. 2025, 68, 2188–2195. [Google Scholar] [CrossRef]
  9. Young, T., III. An essay on the cohesion of fluids. R. Soc. 1805, 95, 65–87. [Google Scholar] [CrossRef]
  10. Wenzel, R.N.J.I.; Chemistry, E. Resistance of solid surfaces to wetting by water. Ind. Eng. Chem. 1936, 28, 988–994. [Google Scholar] [CrossRef]
  11. Cassie, A.; Baxter, S.J.T.O.T.F.S. Wettability of porous surfaces. Trans. Faraday Soc. 1944, 40, 546–551. [Google Scholar] [CrossRef]
  12. Gao, Y.; Zhu, C.; Zuhlke, C.; Alexander, D.; Francisco, J.S.; Zeng, X.C. Turning a Superhydrophilic Surface Weakly Hydrophilic: Topological Wetting States. J. Am. Chem. Soc. 2020, 142, 18491–18502. [Google Scholar] [CrossRef]
  13. Zarghami, S.; Mohammadi, T.; Sadrzadeh, M.; van Der Bruggen, B. Superhydrophilic and underwater superoleophobic membranes-A review of synthesis methods. Prog. Polym. Sci. 2019, 98, 101166. [Google Scholar] [CrossRef]
  14. Yan, H.; Yuanhao, W.; Hongxing, Y. TEOS/silane coupling agent composed double layers structure: A novel super-hydrophilic coating with controllable water contact angle value. Appl. Energy 2017, 185, 2209–2216. [Google Scholar] [CrossRef]
  15. Wang, J.; Pei, X.; Zhang, J.; Liu, S.; Li, Y.; Wang, C. Reversible switch of wettability of ZnO@stearic acid nanoarray through alternative irradiation and heat-treatment. Ceram. Int. 2021, 47, 9164–9168. [Google Scholar] [CrossRef]
  16. Wang, J.; Yu, Z.; Zhu, X.; Xiao, X.; Pang, Y.; Tan, Q.; Liu, Y. A super-hydrophilic NH2-MIL-125 composite film with dopamine-modified graphene oxide is used for water treatment. New J. Chem. 2022, 46, 13303–13314. [Google Scholar] [CrossRef]
  17. Lim, C.S.; Kueh, T.C.; Soh, A.K.; Hung, Y.M. Engineered superhydrophilicity and superhydrophobicity of graphene-nanoplatelet coatings via thermal treatment. Powder Technol. 2020, 364, 88–97. [Google Scholar] [CrossRef]
  18. Bastami, H.; Grayeli, A.; Arman, A.; Matos, R.S.; Ferreira, N.S.; Da Fonseca Filho, H.D.; Ţălu, Ş. Microstructure of ZrO2 films: Hydrophilic properties and optical band gaps. Mater. Today Commun. 2025, 45, 112338. [Google Scholar] [CrossRef]
  19. Zhang, J.; Sun, F.; Xu, J.; Zhao, Z.H.; Fu, J. Research Progress of Human Biomimetic Self-Healing Materials. Small 2024, 21, e2408199. [Google Scholar] [CrossRef] [PubMed]
  20. Schmidt, T.A. Lubricating lipids in hydrogels. Science 2020, 370, 288–289. [Google Scholar] [CrossRef] [PubMed]
  21. Su, X.; Hao, D.; Li, P.; Yang, M.; Guo, X.; Ai, X.; Zhao, T.; Jiang, L. Setaria viridis-inspired hydrogels with multilevel structures for efficient all-day fresh water harvesting. J. Mater. Chem. A 2023, 11, 7702–7710. [Google Scholar] [CrossRef]
  22. Delesky, E.A.; Jones, R.J.; Cook, S.M.; Cameron, J.C.; Hubler, M.H.; Srubar, W.V. Hydrogel-assisted self-healing of biomineralized living building materials. J. Clean. Prod. 2023, 418, 138178. [Google Scholar] [CrossRef]
  23. Chen, W.; Zhang, P.; Zang, R.; Fan, J.; Wang, S.; Wang, B.; Meng, J. Nacre-Inspired Mineralized Films with High Transparency and Mechanically Robust Underwater Superoleophobicity. Adv. Mater. 2020, 32, 1907413. [Google Scholar] [CrossRef]
  24. Chen, K.; Zhou, S.; Wu, L. Self-Healing Underwater Superoleophobic and Antibiofouling Coatings Based on the Assembly of Hierarchical Microgel Spheres. ACS Nano 2016, 10, 1386–1394. [Google Scholar] [CrossRef]
  25. Xu, G.; Liu, P.; Pranantyo, D.; Neoh, K.-G.; Kang, E.-T. Dextran and Chitosan-Based Antifouling, Antimicrobial Adhesion, and Self-Polishing Multilayer Coatings from pH-Responsive Linkages-Enabled Layer-by-Layer Assembly. ACS Sustain. Chem. Eng. 2018, 6, 3916–3926. [Google Scholar] [CrossRef]
  26. Wang, D.; Liu, H.; Yang, J.; Zhou, S. Seawater-Induced Healable Underwater Superoleophobic Antifouling Coatings. ACS Appl. Mater. Interfaces 2019, 11, 1353–1362. [Google Scholar] [CrossRef] [PubMed]
  27. Elashnikov, R.; Khrystonko, O.; Jilková, T.; Rimpelová, S.; Prchal, J.; Khalakhan, I.; Kolská, Z.; Švorčík, V.; Lyutakov, O. High-Strength Self-Healable Supercapacitor Based on Supramolecular Polymer Hydrogel with Upper Critical Solubility Temperature. Adv. Funct. Mater. 2024, 34, 2314420. [Google Scholar] [CrossRef]
  28. Zhao, B.Q.; He, X.L.; Li, W.J.; Zhang, H.; Liu, H.; Qiu, H.; Gu, P.; Chen, K.L. Multi-touch microcapsules/silicone-polyacrylate hybrid hydrogels with precise self-healing ability and their self-powered sensing applications in emergency rescue. Chem. Eng. J. 2024, 500, 157149. [Google Scholar] [CrossRef]
  29. Li, Y.; Yu, J.Y.; Cao, Z.L.; He, P.; Liu, Q.T.; Han, X.B.; Wan, Y. Preparation and application of novel microcapsules ruptured by microwave for self-healing concrete. Constr. Build. Mater. 2021, 304, 124616. [Google Scholar] [CrossRef]
  30. Fu, S.; Li, H.; Liu, H.; Zhao, Y.; Xu, Z. Self-healing superoleophobic and superhydrophilic fabrics for efficient oil/water separation. J. Polym. Eng. 2024, 44, 347–353. [Google Scholar] [CrossRef]
  31. Ni, Z.; Guo, Z.; Lin, Y.H. Synthesis and Self-Healing Mechanism of Self-Healing Epoxy Microcapsule Materials. Key Eng. Mater. 2020, 842, 3–9. [Google Scholar] [CrossRef]
  32. Ratwani, C.R.; Zhao, S.; Huang, Y.; Hadfield, M.; Kamali, A.R.; Abdelkader, A.M. Surface Modification of Transition Metal Dichalcogenide Nanosheets for Intrinsically Self-Healing Hydrogels with Enhanced Mechanical Properties. Small 2023, 19, 2207081. [Google Scholar] [CrossRef]
  33. Ribeiro, M.M.; Simões, M.; Vitorino, C.; Mascarenhas-Melo, F. Physical crosslinking of hydrogels: The potential of dynamic and reversible bonds in burn care. Coord. Chem. Rev. 2025, 542, 216868. [Google Scholar] [CrossRef]
  34. Li, B.; Li, C.; Yan, Z.; Yang, X.; Xiao, W.; Zhang, D.; Liu, Z.; Liao, X. A review of self-healing hydrogels for bone repair and regeneration: Materials, mechanisms, and applications. Int. J. Biol. Macromol. 2025, 287, 138323. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, C.; Shen, T.; Yang, J.; Cao, W.; Wei, J.; Li, W. Room-Temperature Intrinsic Self-Healing Materials: A review. Chem. Eng. J. 2024, 498, 155158. [Google Scholar] [CrossRef]
  36. Cheng, M.; Fu, Q.; Tan, B.; Ma, Y.; Fang, L.; Lu, C.; Xu, Z. Build a bridge from polymeric structure design to engineering application of self-healing coatings: A review. Prog. Org. Coat. 2022, 167, 106790. [Google Scholar] [CrossRef]
  37. Ikura, R.; Park, J.; Osaki, M.; Yamaguchi, H.; Harada, A.; Takashima, Y. Design of self-healing and self-restoring materials utilizing reversible and movable crosslinks. NPG Asia Mater. 2022, 14, 10. [Google Scholar] [CrossRef]
  38. Sato, T.; Wassgren, J.; Nakamura, S.; Hozumi, A. Recent trends on anti-fogging treatments using self-healing hydrophilic/superhydrophilic materials. Mater. Lett. 2024, 368, 136681. [Google Scholar] [CrossRef]
  39. Liang, B.; Zhong, Z.; Jia, E.; Zhang, G.; Su, Z. Transparent and Scratch-Resistant Antifogging Coatings with Rapid Self-Healing Capability. ACS Appl. Mater. Interfaces 2019, 11, 30300–30307. [Google Scholar] [CrossRef]
  40. Fargher, H.A.; Sherbow, T.J.; Haley, M.M.; Johnson, D.W.; Pluth, M.D. C–H⋯S hydrogen bonding interactions. Chem. Soc. Rev. 2022, 51, 1454–1469. [Google Scholar] [CrossRef]
  41. Wang, W.; Lu, P.; Fan, Y.; Tian, L.; Niu, S.; Zhao, J.; Ren, L. A facile antifogging/frost-resistant coating with self-healing ability. Chem. Eng. J. 2019, 378, 122173. [Google Scholar] [CrossRef]
  42. Wang, X.; Li, S.; Huang, J.; Mao, J.; Cheng, Y.; Teng, L.; Chen, Z.; Lai, Y. A multifunctional and environmentally-friendly method to fabricate superhydrophilic and self-healing coatings for sustainable antifogging. Chem. Eng. J. 2021, 409, 128228. [Google Scholar] [CrossRef]
  43. Li, W.; Wu, G.; Tan, J.; Yu, X.; Sun, G.; You, B. Transparent Surfactant/Epoxy Composite Coatings with Self-Healing and Superhydrophilic Properties. Macromol. Mater. Eng. 2019, 304, 1800765. [Google Scholar] [CrossRef]
  44. Li, Y.B.; Wei, J.X.; Wang, J.; Wang, Y.Y.; Yu, P.S.; Chen, Y.; Zhang, Z.J. Covalent organic frameworks as superior adsorbents for the removal of toxic substances. Chem. Soc. Rev. 2025, 54, 2693–2725. [Google Scholar] [CrossRef]
  45. Liu, Y.; Zhao, X.; Chen, Z.; He, Y. Deciphering the Entropy-Driven Host–Guest Interactions within Single Covalent Organic Frameworks for Trapping of 131I. Nano Lett. 2024, 24, 13861–13866. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, S.; Nie, Y.; Huang, Y.; Yang, Y.; Chen, H.; Zhang, X. Self-healing, adhesive, photothermal responsive, stretchable, and strain-sensitive supramolecular nanocomposite hydrogels based on host–guest interactions. Polym. Eng. Sci. 2024, 64, 5627–5636. [Google Scholar] [CrossRef]
  47. Guo, W.; Gao, X.; Ding, X.; Ding, P.; Han, Y.; Guo, Q.; Ma, Y.; Okoro, O.V.; Sun, Y.; Jiang, G.; et al. Self-adhesive and self-healing hydrogel dressings based on quaternary ammonium chitosan and host-guest interacted silk fibroin. Colloids Surf. A Physicochem. Eng. Asp. 2024, 684, 133145. [Google Scholar] [CrossRef]
  48. Zohreband, Z.; Adeli, M.; Zebardasti, A. Self-healable and flexible supramolecular gelatin/MoS2 hydrogels with molecular recognition properties. Int. J. Biol. Macromol. 2021, 182, 2048–2055. [Google Scholar] [CrossRef] [PubMed]
  49. Li, C.H.; Zuo, J.L. Self-Healing Polymers Based on Coordination Bonds. Adv. Mater. 2019, 32, 1903762. [Google Scholar] [CrossRef]
  50. Mugemana, C.; Grysan, P.; Dieden, R.; Ruch, D.; Bruns, N.; Dubois, P. Self-Healing Metallo-Supramolecular Amphiphilic Polymer Conetworks. Macromol. Chem. Phys. 2020, 221, 1900432. [Google Scholar] [CrossRef]
  51. Xu, X.; Jerca, V.V.; Hoogenboom, R. Self-Healing Metallo-Supramolecular Hydrogel Based on Specific Ni2+ Coordination Interactions of Poly(ethylene glycol) with Bistriazole Pyridine Ligands in the Main Chain. Macromol. Rapid Commun. 2020, 41, 1900457. [Google Scholar] [CrossRef]
  52. Pourbadiei, B.; Monghari, M.A.A.; Khorasani, H.M.; Pourjavadi, A. A light-responsive wound dressing hydrogel: Gelatin based self-healing interpenetrated network with metal-ligand interaction by ferric citrate. J. Photochem. Photobiol. B Biol. 2023, 245, 112750. [Google Scholar] [CrossRef]
  53. Wang, Z.; Xie, C.; Yu, C.; Fei, G.; Wang, Z.; Xia, H. A Facile Strategy for Self-Healing Polyurethanes Containing Multiple Metal–Ligand Bonds. Macromol. Rapid Commun. 2018, 39, 1700678. [Google Scholar] [CrossRef]
  54. Zheng, N.; Xu, Y.; Zhao, Q.; Xie, T. Dynamic Covalent Polymer Networks: A Molecular Platform for Designing Functions beyond Chemical Recycling and Self-Healing. Chem. Rev. 2021, 121, 1716–1745. [Google Scholar] [CrossRef]
  55. Yang, K.; Yang, J.; Chen, R.; Dong, Q.; Yang, H.; Gu, S.; Zhou, Y. Antibacterial hyaluronic acid hydrogels with enhanced self-healing properties via multiple dynamic bond crosslinking. Int. J. Biol. Macromol. 2024, 256, 128320. [Google Scholar] [CrossRef]
  56. Tran, V.T.; Mredha, M.T.I.; Na, J.Y.; Seon, J.-K.; Cui, J.; Jeon, I. Multifunctional poly(disulfide) hydrogels with extremely fast self-healing ability and degradability. Chem. Eng. J. 2020, 394, 124941. [Google Scholar] [CrossRef]
  57. Kang, C.; Guo, J.; Kiyingi, W.; Li, J.; Xue, P. A New Self-Healing Green Polymer Gel with Dynamic Networks for Flow Control in Harsh Reservoirs. Adv. Funct. Mater. 2024, 35, 2423892. [Google Scholar] [CrossRef]
  58. Jiang, X.; Zeng, F.; Yang, X.; Jian, C.; Zhang, L.; Yu, A.; Lu, A. Injectable self-healing cellulose hydrogel based on host-guest interactions and acylhydrazone bonds for sustained cancer therapy. Acta Biomater. 2022, 141, 102–113. [Google Scholar] [CrossRef] [PubMed]
  59. Chen, B.; Liu, Z.; Shen, Z.; Gong, H.; Jiang, Y.; Su, Y.; Zhou, J.; Li, Y. Ultrahard and Ultraflexible Supramolecular Hydrogel with Superenergy Dissipating Structure for Biomimetic Skin. ACS Appl. Mater. Interfaces 2024, 16, 70891–70905. [Google Scholar] [CrossRef] [PubMed]
  60. Yan, Q.; Zhou, M.; Fu, H. Study on mussel-inspired tough TA/PANI@CNCs nanocomposite hydrogels with superior self-healing and self-adhesive properties for strain sensors. Compos. Part B Eng. 2020, 201, 108356. [Google Scholar] [CrossRef]
  61. Zhang, J.; Wang, Y.; Wei, Q.; Wang, Y.; Li, M.; Li, D.; Zhang, L. A 3D printable, highly stretchable, self-healing hydrogel-based sensor based on polyvinyl alcohol/sodium tetraborate/sodium alginate for human motion monitoring. Int. J. Biol. Macromol. 2022, 219, 1216–1226. [Google Scholar] [CrossRef]
  62. Ji, N.; Luo, J.; Zhang, W.; Sun, J.; Wang, J.; Qin, C.; Zhuo, Q.; Dai, L. A Novel Polyvinyl Alcohol-Based Hydrogel with Ultra-Fast Self-Healing Ability and Excellent Stretchability Based on Multi Dynamic Covalent Bond Cross-Linking. Macromol. Mater. Eng. 2022, 308, 2200525. [Google Scholar] [CrossRef]
  63. Liu, C.; Mao, Y.; Jiang, L.; Hu, Q.; Zhang, Y.; Zhao, F.; Zhang, E.; Sun, X. Large strain, tissue-like and self-healing conductive double-network hydrogel for underwater information transmission. Chem. Eng. J. 2024, 482, 148863. [Google Scholar] [CrossRef]
  64. Tian, Z.; Luo, S.; Yue, S.; Li, Z.; Yang, D. Self-Powered and Self-Healing Underwater Sensors with Functions of Intelligent Drowning Warning and Rescuing by Using a Nonswelling Hydrogel. ACS Appl. Mater. Interfaces 2025, 17, 26055–26064. [Google Scholar] [CrossRef] [PubMed]
  65. Sun, X.; He, S.; Qin, Z.; Li, J.; Yao, F. Fast self-healing zwitterion nanocomposite hydrogel for underwater sensing. Compos. Commun. 2021, 26, 100784. [Google Scholar] [CrossRef]
  66. Li, M.; Han, X.; Fan, Z.; Zhang, Y.; Li, Q.; Xie, G. Autonomous ultrafast-self-healing hydrogel for application in multiple environments. Colloids Surf. A Physicochem. Eng. Asp. 2021, 631, 127669. [Google Scholar] [CrossRef]
  67. Shang, B.; Zhan, Y.; Chen, M.; Wu, L. NIR triggered healable underwater superoleophobic coating with exceptional anti-biofouling performance. Appl. Surf. Sci. 2020, 528, 146805. [Google Scholar] [CrossRef]
  68. Fu, R.; Guan, Y.; Xiao, C.; Fan, L.; Wang, Z.; Li, Y.; Yu, P.; Tu, L.; Tan, G.; Zhai, J.; et al. Tough and Highly Efficient Underwater Self-Repairing Hydrogels for Soft Electronics. Small Methods 2022, 6, 2101513. [Google Scholar] [CrossRef]
  69. Chen, W.-P.; Hao, D.-Z.; Hao, W.-J.; Guo, X.-L.; Jiang, L. Hydrogel with Ultrafast Self-Healing Property Both in Air and Underwater. ACS Appl. Mater. Interfaces 2017, 10, 1258–1265. [Google Scholar] [CrossRef] [PubMed]
  70. Liu, X.; Zhang, Q.; Jia, F.; Gao, G.H. Underwater flexible mechanoreceptors constructed by anti-swelling self-healable hydrogel. Sci. China Mater. 2021, 64, 3069–3078. [Google Scholar] [CrossRef]
  71. Wang, Z.; Yang, X.; Guan, M.; Li, H.; Guo, J.; Shi, H.; Li, S. Triple-defense design of multifunctional superhydrophilic surface with outstanding underwater superoleophobicity, anti-oil-fouling properties, and high transparency. J. Membr. Sci. 2024, 689, 122161. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of a self-healing superhydrophilic surface.
Figure 1. Schematic diagram of a self-healing superhydrophilic surface.
Coatings 15 01006 g001
Figure 2. Diagram of the gas–liquid–solid three-phase interface on a superhydrophilic surface. γl, γs, and γsl represent the interfacial tensions of gas-liquid, gas-solid, and liquid-solid interfaces, respectively.
Figure 2. Diagram of the gas–liquid–solid three-phase interface on a superhydrophilic surface. γl, γs, and γsl represent the interfacial tensions of gas-liquid, gas-solid, and liquid-solid interfaces, respectively.
Coatings 15 01006 g002
Figure 3. Preparation process of self-healing underwater superhydrophobic and anti-biofouling coatings based on hierarchical microgel sphere assembly [24]. Reprinted with permission from ref. [24]. Copyright (2016) (American Chemical Society).
Figure 3. Preparation process of self-healing underwater superhydrophobic and anti-biofouling coatings based on hierarchical microgel sphere assembly [24]. Reprinted with permission from ref. [24]. Copyright (2016) (American Chemical Society).
Coatings 15 01006 g003
Figure 4. The preparation process of the self-healing superhydrophilic coating. (a) Chemical structure of CS, FS-60, TPP, dopamine, and the process for fabricating self-healing superoleophobic and superhydrophilic fabrics; (b) The formation and healing process of microcapsules [30]. Reprinted with permission from ref. [30]. Copyright (2024) (Walter de Gruyter).
Figure 4. The preparation process of the self-healing superhydrophilic coating. (a) Chemical structure of CS, FS-60, TPP, dopamine, and the process for fabricating self-healing superoleophobic and superhydrophilic fabrics; (b) The formation and healing process of microcapsules [30]. Reprinted with permission from ref. [30]. Copyright (2024) (Walter de Gruyter).
Coatings 15 01006 g004
Figure 6. Schematic diagram of the healing mechanism of superhydrophilic coatings with hydrogen bonding interactions [43].
Figure 6. Schematic diagram of the healing mechanism of superhydrophilic coatings with hydrogen bonding interactions [43].
Coatings 15 01006 g006
Figure 7. (a) Schematic representation of the functionalization of MoS2 by β-cyclodextrin. (b) Host–guest interaction between CDMoS2 and adamantyl segments [48]. Reprinted with permission from ref. [48]. Copyright (2021) (Elsevier).
Figure 7. (a) Schematic representation of the functionalization of MoS2 by β-cyclodextrin. (b) Host–guest interaction between CDMoS2 and adamantyl segments [48]. Reprinted with permission from ref. [48]. Copyright (2021) (Elsevier).
Coatings 15 01006 g007
Figure 8. Schematic diagram of self-healing of metal–ligand compounds after damage [53]. Reprinted with permission from ref. [53]. Copyright (2018) (Wiley).
Figure 8. Schematic diagram of self-healing of metal–ligand compounds after damage [53]. Reprinted with permission from ref. [53]. Copyright (2018) (Wiley).
Coatings 15 01006 g008
Table 1. Comparison of healing performance of materials with different healing mechanisms.
Table 1. Comparison of healing performance of materials with different healing mechanisms.
TypeRaw MaterialsTypes of DamageTriggering MechanismHealing TimeHealing EfficiencyReferences
Self-healing microcapsulesCS, sodium tripolyphosphate, FS-60AbrasionHeat distilled water to 100 °C after moistening30 minBasically fully recovered[30]
Electrostatic interactionAMP-NCPs, PVPSevere damage of 20 micronsPlaced in a humid environment36–48 h>80%[38]
Electrostatic interaction and hydrogen bondSulfobetaine methacrylate and 2-hydroxyethyl methacrylateShallow grooves formed after scrapingSoak in deionized water for 2 s, then place in a room temperature environment2 min90%[39]
Hydrogen bondPAA, CMC, and Fe (III)CutWater vapor wettingWithin 1 minFull restoration of transparency[42]
Epoxy resin, Triton X-100Scratches, cuts, and THF solutionHeat separately at 60 °C, 140 °C, and 100 °C3 min, 3 min, and 30 minBasically restored to its original appearance[43]
Host–guest interactionβ-Cyclodextrin,
MoS2, and adamantyl segments
Cut30 °C10 minFully healed[48]
Metal–ligand interactionPyridine and Zinc (II)-based complexesScratch damage with a width of 10 ± 3 μm and a depth of 100 μmHeat 120 °C16 hHealed but with scars remaining[50]
2,6-diethynyl pyridine,
PEG-diazide (L2), Ni2+
CutContinuous contact2 minFully healed[51]
Table 2. Comparison of healing performance and stability of underwater self-healing superhydrophilic surfaces.
Table 2. Comparison of healing performance and stability of underwater self-healing superhydrophilic surfaces.
Main IngredientsTypes of DamageSelf-Healing PropertiesDurability and Self-Healing PropertiesStabilityReferences
SiO2@PDA@Ag particlesScratchNIR (2.00 W) irradiation for 15 s can almost completely healCan exist in strong acids, strong alkalis, or salt solutions of varying concentrationsThe coating maintains stable hydrophilic and oil-repellent properties[67]
Acrylonitrile, acrylamide, methyl acrylateCutCompletely re-bonded after 60 min of contact, no scratchesSelf-healing in air, underwater, and in seawaterAll maintain stable high healing performance[68]
Agarose, PVACutContact 10 s tensile stress recovery 94.7%Ultrafast self-healing in air and underwaterAgarose networks improve the strength and stability of hydrogels[69]
MethacrylamideCutPlace in water and gently squeeze for 8 s. Once healed, both ends can withstand strong tensile forces.Self-healing in deionized water, air, seawater, sweat, alkaline, and acidic aqueous solutionsExcellent healing efficiency in air and underwater but significantly reduced healing efficiency in seawater and acid–alkaline solutions[70]
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

Liu, Z.; Liu, F. Recent Progress on the Healing Mechanisms of Self-Healing Superhydrophilic Surfaces. Coatings 2025, 15, 1006. https://doi.org/10.3390/coatings15091006

AMA Style

Liu Z, Liu F. Recent Progress on the Healing Mechanisms of Self-Healing Superhydrophilic Surfaces. Coatings. 2025; 15(9):1006. https://doi.org/10.3390/coatings15091006

Chicago/Turabian Style

Liu, Zhimeng, and Fatang Liu. 2025. "Recent Progress on the Healing Mechanisms of Self-Healing Superhydrophilic Surfaces" Coatings 15, no. 9: 1006. https://doi.org/10.3390/coatings15091006

APA Style

Liu, Z., & Liu, F. (2025). Recent Progress on the Healing Mechanisms of Self-Healing Superhydrophilic Surfaces. Coatings, 15(9), 1006. https://doi.org/10.3390/coatings15091006

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