Next Article in Journal
Remarkably Efficient [4+4] Dimerization of [n]-Cyclacenes
Previous Article in Journal
Morphology and Composition of Brake Wear Particles Ameliorated by an Alumina Coating Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Types of Crosslinkers and Their Applications in Biomaterials and Biomembranes

1
University of Balamand, Faculty of Arts and Sciences, Tripoli P.O. Box 100, Lebanon
2
American University of Beirut, Faculty of Medicine, Beirut, Lebanon
3
Lebanese University, Faculty of Public Health, Tripoli, Lebanon
4
Electrochemistry Consulting & Services (E2CS), Tripoli, Lebanon
*
Author to whom correspondence should be addressed.
Chemistry 2025, 7(2), 61; https://doi.org/10.3390/chemistry7020061
Submission received: 8 March 2025 / Revised: 31 March 2025 / Accepted: 8 April 2025 / Published: 11 April 2025

Abstract

:
Biomaterials and biomembranes play a crucial role in a variety of applications, particularly in the medical field due to their ability to mimic natural biological structures and functions. Crosslinkers play also an important role in enhancing the structural integrity and functionality of biomaterials and in the design of biomembranes. This review article explores the fundamentals of biomaterials and biomembranes, with a particular focus on the role of crosslinkers in biology, chemistry and medicine. We explore the various types of crosslinkers commonly used in biomaterials synthesis, examining their chemical structure, classification, and synthesis methods. Additionally, we analyze the biological properties of crosslinkers and their interactions, highlighting their biological impact, particularly in cellular behavior and cytotoxicity. This article further emphasizes recent advances and innovation, particularly in tissue engineering, drug delivery, and wound healing. Finally, we conclude by addressing current challenges and suggesting potential futures directions for research in this field.

1. Introduction

Biomaterials and biomembranes are materials that can be used in the human body to help heal or replace tissues and organs, or to aid in restoring normal bodily functions. Although closely interconnected, biomaterials and biomembranes serve distinct functions, each playing an important role in various biomedical applications. Biomaterials can be natural or synthetic designed to interact with biological systems for clinical or medical purposes. These materials are commonly used in applications like implants [1,2], tissue engineering [3], prosthetics [4], drug delivery systems [5,6], regenerative medicine [7], and medical devices [8]. The primary characteristic of biomaterials is to support biological functions in the human body while minimizing adverse reactions. Biomaterials include metals, such as titanium and its alloys, stainless steel and cobalt–chromium alloys [9], various natural and synthetic polymers [10], ceramics [11], composites [12], biodegradable and bioactive materials [13], elastomers and rubbers [14], micro and nanoparticles [15], and hydrogels [16].
On the other hand, biomembranes are thin layers of specialized engineered structures designed to mimic or support the function of natural biological membranes. These membranes can separate or protect biological systems from the surrounding environment [17,18]. In biological systems, membranes consist of lipid bilayers, proteins, and carbohydrates that enclose cells and other organelles [19], playing a crucial role in (1) signal transduction through proteins on the plasma membrane which receive external signals, create pathways for signal transduction through lipid rafts where signaling molecules are concentrated, mediate cross-talk between various signaling pathways within the same cell, and aid in cell–cell communication [20]; (2) maintaining cellular integrity by acting as a barrier between the cell’s interior and the external environment, providing structural and mechanical support to the cell that helps maintain the shape and integrity of the cell and resist deformation [21]; (3) communication by enabling interaction between different cells and their surroundings through receptors on the plasma membrane by means of ion channels and thus creating membrane potential, membrane-bound proteins that can facilitate signaling between cells, and junctions that allow cells to exchange signals and information [22]; and (4) energy production, as they are critical in photosynthesis (chloroplast membranes), cellular respiration (mitochondrial membranes), and ion transport (plasma membrane and ion gradients), all of which are essential processes for generating and storing energy [23]. Given their functional versatility and ability to mimic or support natural biological membranes, biomembranes have a wide range of applications, such as in controlled or targeted drug delivery systems [24], tissue engineering [25], regenerative medicine [19] and artificial organs [26], medical implants [27], and diagnostic devices [28].
For biomaterials and biomembranes to perform optimally, crosslinkers must be used as they can dramatically influence the stability, functionality, and mechanical properties of these materials. Crosslinking occurs when polymer chains or molecules chemically bond together, significantly affecting the materials’ properties [29]. This process is critical in tissue engineering, drug delivery, and wound healing [30]. There are two major types of crosslinkers: chemical crosslinkers and physicals crosslinkers. Chemical crosslinkers include formaldehyde and glutaraldehyde (used to crosslink proteins or tissues) [31], ethylene glycol dimethylacrylate or EGDMA (to crosslink hydrogels) [32], isocyanates (for polymer-based biomaterials) [33], and genipin (a biodegradable crosslinker used in tissue engineering) [34]. Physical crosslinkers include ultraviolet (UV) light to crosslink certain polymers [35] or ionic interactions between anionic and cationic groups which create crosslinks in polyelectrolyte-based materials like hydrogels [36].
Therefore, the selection of crosslinker is consequently of high importance for the design of biomaterials and biomembranes since it affects the desired target properties. The crosslinkers enhance the mechanical strength of the crosslinked material, making them more resistant to deformation and tailored to have specific rigidity to exactly mimic natural tissue [37]. Crosslinkers improve material stability against changes such as in temperature or moisture and can be designed to degrade at a controlled rate [38]. They also offer the advantage of controlled and sustained delivery of drugs in response to specific stimuli (like temperature) or deliver to certain target locations under specific conditions only (like the pH level or the presence of a certain enzyme) [39] and are mostly biodegradable or bioabsorbable, which minimizes their long-term toxicity, especially in medical applications [40].
In this article, we review the fundamental properties of biomaterials and biomembranes and how they can enhance human bodily functions. We also investigate crosslinked materials, the biological and chemical properties of crosslinkers, and the various medical applications of crosslinked biomaterials and biomembranes. We investigate the mechanism of crosslinking, the recent advances in crosslinking techniques, and the emergence of new biomaterials and biomembranes.

2. Fundamentals of Biomaterials

Biomaterials are materials designed to interact with biological systems [41]. Biomaterials include a wide variety of compounds, such as metals including titanium and its alloys, stainless steel, and cobalt–chromium alloys, various natural and synthetic polymers, ceramics, composites, biodegradable and bioactive materials, elastomers and rubbers, micro and nanoparticles, and hydrogels. Table 1 summarizes the typical usage of various biomaterials.
Table 1. Typical usage of various biomaterials.
Table 1. Typical usage of various biomaterials.
MetalsPolymersCeramicsCompositesBiodegradable and Bioactive MaterialsElastomers and RubbersMicroparticles and NanoparticlesHydrogels
Implants
Prosthetics
Drug Delivery Systems
Tissue Engineering
Diagnostic Devices
Natural polymers are natural substances composed of large macromolecules formed by the binding of small identical or different molecules called monomers. These polymers can be produced via biological processes like protein and nucleic acid synthesis. Synthetic polymers are produced via artificial chemical reactions to produce a desirable product of industrial importance, e.g., polystyrene, polyvinyl chloride, nylon, silicon, etc. [42]. Polymers are widely used in our normal lives today; however, many polymers are resistant to degradation and contribute to the accumulation of polymer waste. This waste, as it continues to accumulate, becomes a serious threat to the environment and remains undecomposed for a long period. Due to these problems, scientists created biodegradable polymers that are easily degraded by microorganisms, such as aliphatic polyesters, which result in the reduced accumulation of waste and cause less harm and pollution to the environment [43,44] (Figure 1).
Synthetic aliphatic polyesters, derived from lactide, glycolide, caprolactone, and their copolymers, are among the most widely used degradable polymeric biomaterials for medical applications, including implantable devices, drug delivery systems [45], and tissue engineering materials. Despite its biocompatibility, polylactide (PLA) does not meet the optimal properties needed to replace soft tissues like ligaments, tendons, cartilage, or blood vessels [46].
In an attempt to form temporary implantable biomaterials, PLA was modified, improving its mechanical properties by means of crosslinking. This technique al-lows for the formation of structures with identical vulcanized rubbers [47]. The main drawbacks of these materials are that they are difficult to design with complex shapes due to the impossibility of remolding by heating or solvent casting.
In order to improve specific mechanical properties, PLA-based block polymers containing F 127 and Tetronic 1107 were synthesized, and their short-term degradation was studied. The results showed that the modification of the crystallinity of PLA blocks, along with the initial global molecular weight, can lead to a possible modulation of different degradation properties. In fact, an important decrease in molecular weight was detected with 200 kg.mol−1 copolymers with the conservation of high mechanical integrity. In addition, a lower strain at failure was demonstrated in the case of copolymers based on PLA 96 compared to those containing PLA 94. These 200 kg.mol−1 star and linear PLA co-polymers are considered to be promising materials for temporary biomedical applications requiring cell cultures due to their good biocompatibility and degradation rates [48].
On the other hand, many aspects of modern medicine have been affected by the use of smart biomaterials due to their ability to respond to modifications in physiological parameters. In fact, a study of the uses of smart biomaterials in tissue engineering was carried out. This recent study showed the impact of these smart biomaterials in many other medical fields, such as medical devices, drug delivery, and immune engineering, leading to promising therapies for the treatment of many debilitating diseases [49].
A vascular scaffold fabricated by electrospinning poly(ε-caprolactone) (PCL) and collagen was designed, leading to an important adhesion substrate for vascular cells and adequate structural support. However, the small-sized pores were responsible for the decrease in the efficacy of smooth muscle cell (SMC) seeding due to their inability to infiltrate into the scaffolds. The development of a bilayered scaffold that improves EC adhesion onto the luminal surface and the homogenous infiltration of SMCs into the outer layer was studied in order to overcome this challenge. This advanced scaffold was found to be biodegradable and biocompatible, allowing a good ability for infiltration and cell adhesion support. In addition, sufficient mechanical support, along with physiological vascular conditions, were provided. This recent study demonstrated that vessel tissue function could be improved by using these bilayered scaffolds, facilitating endothelialization and smooth muscle maturation [50].
The use of carbon dots (CDs) as biosensors and bioimaging agents due to their particular properties was an important challenge for scientists when first evaluating the osteoconductivity of CDs in poly (ε-caprolactone) (PCL)/polyvinyl alcohol (PVA) [PCL/PVA] nanofibrous scaffolds. In fact, after several studies on the evaluation of eggshell-derived calcium phosphate (TCP3) and its cellular responses, an important proliferation and osteogenic differentiation was observed with these derivatives compared to other scaffolds. As a result, the implementation of CDs and novel eggshell-derived calcium phosphate showed high importance for the fabrication of bioscaffolds used for bone tissue engineering [51].
More research should be implemented to develop the economic and environmental aspects related to the application of these block copolymers for designing nanostructured materials [52].
Figure 1. Examples of the natural polymers and derivatives used for biomaterials design [53].
Figure 1. Examples of the natural polymers and derivatives used for biomaterials design [53].
Chemistry 07 00061 g001

3. Fundamentals of Biomembranes

Defining a biomembrane comprehensively is difficult because of its wide range of characteristics and functionalities. Biomembranes are a specific type of biomaterials used to form membranes that separate two regions and selectively regulate the movement of substances between them. While biomaterials are typically used for structural applications, biomembranes are designed to regulate the movement of substances across them, making them essential for applications such as filtration and drug delivery. Biomembranes can have different forms, such as being homogeneous (the plasma membrane of some bacteria) or heterogeneous (the plasma membrane of some animal cells), symmetrical (lipid bilayers created for model membranes) or asymmetrical (the plasma membrane of eucaryotic cells), solid (lipid rafts) or liquid (the plasma membrane of most cells), and neutral (the outer leaflet of the plasma membrane) or charged (positive, negative, or both) (the inner left of the plasma membrane) [54]. Regardless of the type, all biomembranes share the fundamental property of controlling the passage of substances in a specific way [55]. Biomembranes are versatile, as their properties can be “tailored” or customized to suit specific separation tasks, making them adaptable for different applications (e.g., tissue engineering, drug delivery systems, biosensors, and biomedical devices, etc.).
Biomembrane materials can be broadly classified into three main categories: inorganic (ceramic), synthetic polymers, and natural polymers [56]. Additionally, there are mixed membranes, inorganic/polymeric or polymeric/polymeric materials, used to enhance the properties of the membrane, offering a balance of durability and selectivity for specific applications. Inorganic membranes, such as alumina (Al2O3) and zirconia (ZrO2), are widely used for their thermal and chemical stability. However, their higher production costs can be a limiting factor compared to polymeric and natural membranes [57]. This has led to a growing interest in low-cost raw materials such as clays, ash, and waste products to develop more affordable ceramic membranes. These membranes offer significant advantages, including lower costs and reduced environmental impact. However, the design and fabrication process is complex, requiring careful attention to material composition, additives, and safety concerns, particularly regarding the presence of heavy metals and toxic compounds [58].
Polymers can be divided into two types based on their origin: synthetic polymers, which are artificially created from petroleum-based sources, and natural polymers, which are derived from plants or animals [59,60]. Noticeably, biopolymer production is sustainable, carbon neutral, and renewable because biopolymers are made from sea or land plant materials [61,62].
Chitin is the second most abundant natural polysaccharide, commonly found in crustacean shells, insects, molluscan organs, and fungi. Chitosan is an eco-friendly polymer derived from chitin through N-deacetylation and consists of glucosamine and N-acetylglucosamine [63]. Its amino and hydroxyl groups enable diverse applications in pharmaceuticals, wastewater treatment, fuel cells, and biodegradable food packaging due to their antimicrobial and hydrophilic properties. Chitosan is available as membranes, films, and fibrous mats. Despite its advantages, chitosan has drawbacks that limit its application. Chitosan has low mechanical strength and low electrical conductivity. It is also very brittle due to its high glass transition temperature. These drawbacks are mitigated by blending with other polymers, adding inorganic fillers, or chemical modifications.
Alginate, a water-soluble polysaccharide from brown seaweed, is composed of mannuronic and guluronic acid units [64]. Its advantages include biocompatibility, non-toxicity, biodegradability, and ease of combination with divalent cations, making it useful in fields like drug delivery, tissue engineering, and textiles (Figure 2). Alginate can be manufactured into a variety of forms, such as film, microspheres, and fibers, because of its reversible solubility. However, like chitosan, alginates have certain limits to their applications due to several weaknesses, which include high water solubility and low mechanical strength. Alginate’s six-membered ring backbone makes it challenging to enhance rigidity or compactness, resulting in larger void volumes that allow significant water absorption. Excessive water uptake leads to membrane swelling, reducing selectivity (Table 2). Striking a balance between membrane permeability and selectivity for water or gas is essential. To overcome this limitation, alginate has been improved through methods like covalent crosslinking, ionic crosslinking, and non-bond interactions, which enhance its structural stability and functional performance [65].
Figure 2. SEM images for different biomembranes made using different biomaterials (chitosan, cellulose, alginate, gelatin, etc.) and crosslinked by various cross-linkers (genipin, glutaraldehyde, etc.) to highlight the microstructure organization after the crosslinking process (adapted from [66,67,68,69]). Red arrow indicates interconnected pores. Each yellow scale bar represents 100 μm.
Figure 2. SEM images for different biomembranes made using different biomaterials (chitosan, cellulose, alginate, gelatin, etc.) and crosslinked by various cross-linkers (genipin, glutaraldehyde, etc.) to highlight the microstructure organization after the crosslinking process (adapted from [66,67,68,69]). Red arrow indicates interconnected pores. Each yellow scale bar represents 100 μm.
Chemistry 07 00061 g002
Cellulose stands out as the most abundant natural polymer and is extensively documented in the literature for membrane fabrication. Its extensive use is due to its abundance and remarkable properties. Cellulose, a prevalent polysaccharide, is valued for its abundance, low cost, high hydrophilicity, and biocompatibility, making it ideal for industrial applications like paper, textiles, membranes, and pharmaceuticals. It is found in plants, trees, fungi, algae, and bacteria, with plant fibers being the primary source. Structurally, cellulose is a linear polymer with repeating units of D-glucopyranose [56]. Phase inversion via immersion precipitation is the most commonly used method for fabricating cellulose-based composite membranes. In this process, cellulose or a cellulose-containing polymer mixture is dissolved in a suitable solvent, and the resulting solution is cast onto an appropriate support layer. The cast layer is then submerged in a coagulation bath, often containing pure water as a non-solvent, which leads to the formation of the membrane. Another widely used technique for producing well-interconnected porous membranes is thermally induced phase separation (TIPS) [70]. TIPS involves lowering the temperature to induce phase separation in the dope solution, followed by solvent removal through extraction, evaporation, or freeze-drying. Interfacial polymerization is the predominant technique for fabricating thin-film composite (TFC) membranes [71], which involves coating cellulose-based polymers onto a polymeric substrate. Recent advancements in nanotechnology, coupled with the limitations of traditional methods, have led to the adoption of electrospinning for the production of nanoscale fibers with exceptional properties. Nanofibrous polymeric networks produced through electrospinning offer large surface areas, easy functionalization, and superior mechanical properties. This technique uses an electric field to create highly porous nanofibrous membranes from polymer fibers. Compared to membranes produced by traditional phase inversion methods, cellulose-based nanofibrous membranes exhibit higher flux, reduced fouling, and a nanoscale porous structure (Figure 3).
Figure 3. Summary of natural and synthetic biomembranes [72] (republished from [72] under license number 5998461306754).
Figure 3. Summary of natural and synthetic biomembranes [72] (republished from [72] under license number 5998461306754).
Chemistry 07 00061 g003
Polymer blending is a versatile technique used to enhance material performance. Natural polymers such as cellulose (cellulose acetate, CA) can be blended with synthetic polymers like polyether sulfone (PES) to improve membrane properties [73]. Additionally, blending natural polymers with ceramics, such as adding inorganic fillers like TiO2 [74], Fe3O4, and alumina [75] into CA and CAP (cellulose acetate phthalate), has shown significant improvements in selectivity, hydrophilicity, surface area, and porosity, even with small amounts of filler materials. However, optimizing filler loading is a key challenge, as excessive loading can lead to poor dispersion, causing aggregation and mechanical instability. The toxicity of certain inorganic nanoparticles is also a concern, as their leakage into the aqueous environment can pose health and environmental risks. To address these issues, further purification stages are necessary.
There is a growing interest in blends composed entirely of bio-based materials. For instance, combining cellulose with biopolymers like chitosan has led to materials with unique properties, including antibacterial activity [76], metal ion adsorption, and self-healing characteristics [77].
Other natural polymers have been used in research, though they are not as widely applied as the ones mentioned earlier. One such polymer is poly(lactic acid) (PLA), a thermoplastic aliphatic polyester derived from renewable sources such as corn, sugar beets, or rice. Silk fibroin (SF), another natural biopolymer, has also gained attention due to its non-toxicity, excellent biocompatibility, and biodegradability. Silk fibroin is typically extracted from silkworm cocoons by removing the outer sericin layer. The polymer’s structure, which contains numerous amino (–NH2) and carboxyl (–COOH) groups, makes it an effective adsorbent for heavy metal ions [78].
Table 2. Comparison of membrane types: origin, properties, applications, and limitations.
Table 2. Comparison of membrane types: origin, properties, applications, and limitations.
Membrane TypeOriginPropertiesApplicationsLimitations
Inorganic membrane [58]Ceramics (alumina, zirconia, clay), metalsHigh mechanical strength
Thermal resistance
Chemical resistance
Gas separation
Water treatment
Pharmaceutical
Filtration and desalination application
High production cost
Fragility
limited flexibility
Synthetic polymers (Petrochemical-derived) [79,80] PolysulfoneResistance to extreme pH, high temperatures
Good mechanical properties
Microfiltration, ultrafiltration, reverse osmosisExpensive
Not eco-friendly
Poor stability
Poly (vinylidene fluoride) (PVDF)High mechanical strength
Chemical resistance
Aging resistance
Excellent thermal stability
Filtration, water treatment, pharmaceutical separations
Natural polymers (Plant or animal-derived)
+
their derivatives
[61,62]
Chitosan (from crustacean shells, fungi, insects) acetylglucosamine [63]Abundant
Biocompatible
Biodegradable
Hydrophilic
Less cost
Drug delivery
Sensors
Wound healing
Tissue engineering
Pharmaceutical applications
water filtration
Low mechanical strength
Low thermal stability
Alginate (from seaweed) [64]
Cellulose (from plants) [56]
Mixed matrix membranesCombination of inorganic and polymeric materialsEnhanced properties, combining the best of both materials (e.g., durability, selectivity)Water treatment
Gas separation
Industrial filtration
Challenges in achieving optimal dispersion of fillers, potential instability
Biopolymer blendChitosan–alginate blend
Chitosan–cellulose blend
Alginate–cellulose blend
Increased mechanical strength
High stability
Synergistic interactions
Drug delivery
Water treatment
Food packaging
Biomedical applications
Complex preparation process
Limited scalability
potential instability under certain conditions

4. Crosslinkers: Chemistry and Classification

Crosslinking is an important phenomenon in polymer chemistry; it enhances the properties of materials by forming connections between polymer chains. Crosslinkers create bonds between polymer chains, leading to a three-dimensional network structure. The nature of these bonds determines the properties and applications of the resulting material. Crosslinking can occur through different types of bonding: covalent, ionic, or hydrogen bonds.

4.1. Covalent Crosslinking

Covalent crosslinking is observed when stable covalent bonds are formed between polymer chains. These bonds are irreversible, creating durable materials with high mechanical strength and thermal stability. Covalent crosslinking is common in elastomers and hydrogels, where strong bonds are required to resist significant stress or deformation. Many covalent crosslinkers have been used: glutaraldehyde [81], formaldehyde, ethyleneglycol diglycidyl ether (EGDGE) [82], N,N’-methylenebisacrylamide (MBAA) [83], the isocyanate-based crosslinker [84], and others [85].
An example of covalent crosslinking is the use of glutaraldehyde as a crosslinker for collagen-based hydrogels. Glutaraldehyde forms covalent bonds with amino groups on collagen, resulting in a highly stable, crosslinked network used in biomedical applications such as tissue engineering and wound healing [86,87,88]. The general reaction for the covalent crosslinking process is as follows:
Collagen-NH2 + Glutaraldehyde → Collagen-N=CH-R + H2O
Collagen-NH2 represents the amino group on collagen (usually from lysine).
Glutaraldehyde has two aldehyde groups, which can react with the amino groups.
Collagen-N=CH-R is the resulting Schiff base, where “R” represents the remaining part of the glutaraldehyde structure.

4.2. Ionic Crosslinking

Ionic crosslinking is observed when electrostatic interactions occur between oppositely charged groups on polymer chains. These bonds are weaker than covalent bonds, making these materials more flexible and responsive to environmental changes such as pH. Ionic crosslinkers are used in many applications, such as tissue engineering, wound healing, drug delivery systems, food and cosmetic industries, environmental and water treatments, etc. [89,90,91,92]. Many ionic crosslinkers have been used, like calcium ions (Ca2⁺) [93], magnesium ions (Mg2⁺) [94], zinc ions (Zn2⁺) [95], and polyvalent metal cations such as copper or iron [96]. An example of ionic crosslinking is the use of alginate, a naturally derived polysaccharide, which can be crosslinked using calcium ions to form a hydrogel. The divalent calcium ions bridge the negatively charged carboxylate groups on alginate chains, creating a gel matrix widely used in drug delivery systems. The general reaction for the ionic crosslinking process is as follows:
Alginate-COO + Ca2+ → Alginate-Ca2+(-COO) (hydrogel)
Alginate-COO⁻ represents the carboxylate group on the alginate chain.
Ca2⁺ is the calcium ion that crosslinks the chains.
Alginate-Ca2⁺(-COO)⁻ represents the crosslinked structure where calcium ions bridge the carboxylate groups.

4.3. Hydrogen Bonding

Hydrogen bonding crosslinking is observed when there are interactions between hydrogen atoms and electronegative atoms such as oxygen or nitrogen. These types of bonds are weaker than covalent and ionic bonds; they can provide a significant level of crosslinking, particularly in biological materials or polymers, enhancing their elasticity and self-healing properties.
Many hydrogen-bonding crosslinkers have been used such as polyvinyl alcohol (PVA) [97], urethan-based crosslinkers, polyethylene glycol (PEG) [98], carboxymethyl cellulose (CMC) [99], dextrin, starch, chitosan, and others [100]. An example of hydrogen bonding crosslinking is poly(vinyl alcohol) (PVA), which can be used as a hydrogen bonding crosslinker through the formation of intermolecular hydrogen bonds between the hydroxyl groups (-OH) present on the polymer chains or with other chemical species. This crosslinking method forms robust, biocompatible materials suitable for application in cartilage replacement and drug delivery [101]. The general reaction for the hydrogen bonding crosslinker is as follows:
PVA-OH⋅⋅⋅PVA-OH → Hydrogen Bond Formation
Each type of crosslinking covalent, ionic, and hydrogen bonding offers unique advantages that can be tailored to specific applications, from biomedical to industrial fields. The selection of crosslinking chemistry depends on the desired material properties, such as flexibility, strength, and responsiveness, providing a versatile toolkit for material science and engineering.
Here, we review some of the most common crosslinkers used in research and industry, including glutaraldehyde, genipin, EDC/NHS, and epoxies, each with unique mechanisms and applications. Each crosslinking agent has its advantages and limitations. The choice of crosslinker depends on the material requirements, desired properties, and application, providing a range of options for diverse scientific and industrial needs.

4.4. Glutaraldehyde

Glutaraldehyde is a widely used crosslinker, especially in the formation of covalent bonds with biomaterials containing amino groups such as chitosan, collagen, and gelatin [102]. Glutaraldehyde reacts with amine groups to form stable covalent linkages, making it ideal for applications that require durable and stable crosslinks. However, glutaraldehyde, like many other aldehydes, is toxic, limiting its use in some biomedical applications unless thorough post-processing is conducted (neutralization, washing, dehydration, etc.). Another strategy to overcome this toxicity issue is to add the glutaraldehyde in a limited quantity. Due to its high reactivity (crosslinking in a few seconds), the added glutaraldehyde quantity is totally consumed (forming the covalent linkages), which guarantees the complete consumption of the added crosslinker quantity. Generally, this strategy allows us to avoid the post-processing treatment or to minimize the necessary treatment time. Glutaraldehyde-crosslinked collagen is often used in tissue engineering and medical devices [103].

4.5. Genipin

Genipin, a natural crosslinker derived from the gardenia plant, is considered a safer alternative to synthetic agents like glutaraldehyde [104,105]. Genipin reacts with primary amines to form blue-colored crosslinked materials, making it especially useful in biomedical applications where biocompatibility is essential, such as tissue engineering and drug delivery. Genipin-crosslinked materials exhibit excellent biocompatibility and low cytotoxicity, making it an ideal crosslinker for hydrogels and collagen-based materials [106].

4.6. EDC/NHS

1-Ethyl-3-(3-dimethylaminopropyl) carbodiimide (EDC) is often used with N-hydroxysuccinimide (NHS) to activate carboxyl groups on polymers, allowing them to react with amine groups to form amide bonds [107]. The EDC/NHS system is widely used in biomaterial functionalization because it forms stable, covalent amide linkages without leaving residual toxic by-products. EDC/NHS crosslinking is commonly applied in the conjugation of peptides, proteins, and polysaccharides, particularly in drug delivery and tissue engineering applications where biocompatibility is critical [108].

4.7. Epoxies

Epoxy-based crosslinkers are synthetic agents commonly used for forming covalent bonds in polymers due to their high reactivity and ability to form durable, chemically resistant networks [109]. Epoxies are versatile, used in a variety of applications, from adhesives to coatings, and are particularly valued in the automotive, aerospace, and construction industries for their toughness and thermal stability. Epoxies are also applied in biomedical applications, though biocompatibility must be considered depending on the specific context [110].

4.8. Other Crosslinkers

Several other crosslinking agents, such as poly (ethylene glycol) diglycidyl ether (PEGDE) and formaldehyde, are also commonly used based on application-specific needs [111]. PEGDE, for example, is often used to create hydrogels with enhanced water retention and flexibility, making it ideal for soft tissue engineering and wound care applications [112]. Formaldehyde is used with caution due to its known toxicity and is primarily applied in industrial materials rather than biomedical contexts [113].

4.9. Mechanisms of Crosslinking: Reaction Pathways and Effects on Properties

Crosslinking in polymer chemistry involves various reaction pathways depending on the crosslinking agent and the type of bonds formed. Crosslinkers can bond through covalent, ionic, or hydrogen bonding mechanisms, each influencing the properties of the resulting polymer network. Here, we discuss several common crosslinking mechanisms.

4.9.1. Step-by-Step Mechanism of Crosslinking with Glutaraldehyde [114,115,116]

  • Activation of the aldehyde group
Glutaraldehyde has two reactive aldehyde groups (-CHO) at each end of its molecule. Each aldehyde group can form a covalent bond with a nucleophilic site, such as an amino group (–NH2), typically found on lysine residues in proteins.
  • Nucleophilic attack by the amino group
The amino group of a protein or other nucleophilic molecules attack the carbonyl carbon of the glutaraldehyde aldehyde group. Carbonyl carbon is electrophilic, making it susceptible to nucleophilic attack. This reaction forms a hemiketal intermediate, which is unstable.
  • Formation of a Schiff base (Imine bond)
The hemiketal intermediate undergoes dehydration (loss of water) to form a stable imine (Schiff base) linkage. This reaction involves the formation of a double bond between the nitrogen of the amino group and the carbon of the carbonyl group, resulting in the loss of water:
R-NH2+ CHO−(CH2)3CHO → RNH−CH=(CH2)3CH=O + H2O
where R-NH2 represents an amino group (from a lysine residue or similar) and CHO−(CH2)3 CHO- represents glutaraldehyde.
  • Crosslinking
The imine group formed in the previous step can further react with another aldehyde group from a second glutaraldehyde molecule or from another lysine residue, leading to the formation of a covalent crosslink between two molecules. This creates a three-dimensional network of crosslinked molecules.
This is a key step in biological sample fixation (e.g., tissues), as glutaraldehyde forms highly stable crosslinks between protein molecules, which helps to preserve their structure.
  • Potential hydrolysis
The imine bond (Schiff base) formed between the lysine residue and glutaraldehyde is not always stable over time. It can undergo hydrolysis, especially in aqueous conditions, to regenerate the aldehyde group and the amine group. This can lead to some degree of “reversibility”, but under controlled conditions, the crosslinking can be quite stable.
  • Hydrolysis reaction:
RNH=CH2CH2CHO + H2O → RNH2 + CHO−(CH2)3CH2OH
This step is generally avoided or minimized by controlling the environmental conditions (e.g., reducing the pH or adding stabilizers) (Figure 4a).

4.9.2. Step-by-Step Mechanism of Enzymatic Crosslinking with Genipin [117,118,119,120]

  • Genipin activation by transglutaminase (TGase)
The enzyme transglutaminase (TGase), which catalyzes the formation of covalent bonds between glutamine and lysine residues in proteins, plays a significant role in facilitating the activation of genipin. TGase activates the genipin molecule by catalyzing the hydrolysis of the epoxide ring, promoting the formation of reactive intermediates (such as a quinone methide or aldehyde group). These intermediates are highly electrophilic, are prone to nucleophilic attack, and can participate in crosslinking.
Genipin (C9H12O5) + TGase + Enzyme → Reactive intermediate (aldehyde/quinone methide) + H2O
  • Crosslinking with amino groups
Once the genipin molecule is activated into a reactive intermediate, the electrophilic species (such as the quinone methide or aldehyde) can react with nucleophilic groups, like amine groups (–NH2) found in proteins or collagen. This forms a covalent bond through a Schiff base formation or direct nucleophilic attack.
Protein-NH2 + Reactive Genipin intermediate → Protein-NH-Genipin linkage + H2O
The reaction typically forms a Schiff base (an imine bond, –C=N–), which stabilizes the linkage between the protein and the genipin molecule.
  • Polymerization and crosslinking
If multiple reactive intermediates from genipin are present, they can further react with other amino groups (either from the same protein or other proteins), leading to the formation of a 3D-crosslinked network.
Protein-NH-Genipin linkage + Protein-NH2 → 3D Crosslinked Protein Network
The final product is a stable, covalently crosslinked matrix, where proteins or biomolecules are connected via genipin-induced linkages, providing structural stability and mechanical strength to the material (Figure 4b) [121].

4.9.3. Step-by-Step Mechanism of Crosslinking Carbodiimide-Mediated with EDC/NHS [121,122,123,124]

The carbodiimide crosslinking system, using 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide (EDC) in the presence of N-hydroxysuccinimide (NHS), activates carboxyl groups, allowing them to react with primary amines to form covalent amide bonds. This reaction proceeds without leaving cytotoxic by-products, making it especially suitable for biological applications like drug delivery and tissue engineering applications [125].
  • Activation of the carboxyl group by EDC
EDC is a carbodiimide compound that activates carboxyl groups by reacting with them. It interacts with the carboxyl group (–COOH) of a protein or other molecules containing a carboxyl functional group.
R-COOH + EDC → R-COO⋅EDC
The resulting intermediate is highly reactive but unstable.
  • The formation of a more stable NHS-activated ester
N-Hydroxysuccinimide (NHS) is added to the reaction mixture. NHS reacts with the intermediate formed in the previous step, stabilizing it and converting it into an NHS ester.
R-COO⋅EDC + NHS →R-COO-NHS + EDC
The NHS ester is much more stable than the previous intermediate, allowing for easier handling and more controlled reaction conditions.
  • Nucleophilic attack by amine group
The NHS ester, which is now reactive, can be attacked by nucleophilic amino groups (–NH2) from another molecule (such as a protein or peptide). The nitrogen from the amine group performs a nucleophilic attack on the carbonyl carbon of the NHS ester, displacing the NHS group.
R-COO-NHS + R’-NH2 → R-CONH-R’ + NHS
This results in the formation of a covalent amide bond between the carboxyl group and the amine group, and the NHS group is released.
  • Crosslinking between two molecules
If both reactants (molecules with amine or carboxyl groups) are involved in the reaction, this can lead to crosslinking between the two biomolecules. The process can result in the formation of dimers, trimers, or higher-order crosslinked products depending on the concentration of the reagents and reaction conditions (Figure 4c).

4.9.4. Step-by-Step Mechanism of Epoxy Crosslinking [126,127,128,129,130]

Epoxy crosslinking is a process where epoxy resins (which contain epoxy groups, typically oxirane rings) are crosslinked with curing agents, such as polyamines, polyamides, or anhydrides, to form a three-dimensional polymer network bonded through covalent bonds. This reaction opens the epoxide ring, leading to a highly crosslinked polymer with excellent resistance to thermal and mechanical stress. Due to these properties, epoxy crosslinked materials are extensively used in coatings, adhesives, and structural applications, particularly in high-stress environments like automotive and aerospace industries [131]. Here is a step-by-step breakdown of the use of an amine group.
  • Nucleophilic Attack
A diamine has two nucleophilic amine groups (-NH2). One of the amines attacks the electrophilic carbon in the epoxide ring, opening the ring and forming a covalent bond between the epoxy and the amine group.
R-CH2-O-CH2-C + Diamine → R-CH2-NH-CH2-C
  • Hydroxyl Group Formation
After the nucleophilic attack, the oxygen that was part of the epoxide ring now holds a hydrogen atom from the reaction. The oxygen is converted into a hydroxyl group (-OH).
  • Crosslinking
The second amine group from diamine attacks another epoxide group, causing further crosslinking and forming a complex network.
R-CH2-NH-CH2-C → R-CH2-NH-CH2-C-NH-CH2-C
  • Curing and Hardening
As the crosslinking continues, a rigid structure is formed where the epoxy molecules are bound together by the curing agent.

4.9.5. Step-by-Step Mechanism of Ionic Crosslinking with Calcium Ions (e.g., Alginate) [132,133,134,135]

Ionic crosslinking occurs when divalent or multivalent cations, such as calcium ions, interact with negatively charged groups on polymers. For example, calcium ions can crosslink alginate, a natural polysaccharide derived from brown algae containing repeated units of M (mannuronic acid) and G (guluronic acid) that have carboxyl groups (-COOH). Calcium ions will bind to the carboxylate groups, forming an “egg-box” structure. This pathway produces materials that are flexible and reversible, making them suitable for applications such as controlled drug release [136].
  • Interaction of Calcium Ions with Alginate
The ionic crosslinking process starts when calcium ions (Ca2⁺) are introduced to a solution of alginate. The calcium ions interact with the carboxyl groups on the alginate chains, particularly those located in the guluronic acid (G-block) regions. This is an electrostatic interaction, where the positively charged calcium ion interacts with the negatively charged carboxylate groups (-COO⁻) on the alginate.
R-COO + Ca2+ → R-COO-Ca2+
This binding occurs because the calcium ion has a high charge density and can coordinate with two carboxylate groups from adjacent alginate chains.
  • Formation of the “Egg-box” Structure
The key feature of the ionic crosslinking mechanism is the formation of the “egg-box” model. In this model, the calcium ion (Ca2⁺) forms a bridge between two alginate chains, specifically between the carboxylate groups of guluronic acid residues.
This creates a network where multiple chains of alginate are crosslinked through calcium ions, leading to the formation of a gel or a network structure.
(R-COO-Ca-COO-R) (egg-box structure)
This ionic crosslinking forms a gel that is stable as long as the calcium ions remain present.
  • Gelation and Network Formation
As more calcium ions are added, more alginate chains are crosslinked, resulting in the formation of a three-dimensional network. The strength of the gel depends on the number of crosslinks formed between the alginate chains. High concentrations of calcium ions lead to more crosslinks, resulting in a more rigid gel. In contrast, low calcium ion concentrations produce weaker gels with fewer crosslinks.
  • Reversibility and Stability
The formed gel is reversible and can be disassembled if the calcium ions are removed, typically by washing with a chelating agent like EDTA or by changing the pH. A chelating agent like EDTA can bind to the calcium ions and remove them from the alginate network, breaking the ionic bonds and leading to gel dissolution (Figure 4d).
R-COO-Ca-COO-R + EDTA → R-COOH + Ca-EDTA complex
Altering the pH of the solution can also influence the stability of the gel. At a low pH, the carboxylate groups on the alginate can protonate to form -COOH, reducing the ionic interaction with calcium ions and causing the gel to break apart. Each crosslinking mechanism offers distinct benefits and limitations, impacting the thermal stability, mechanical strength, elasticity, and biocompatibility of the resulting material. The choice of crosslinking chemistry is critical in tailoring material properties for specific applications, from biomedicine to industrial settings.
Figure 4. (a) The crosslinking process of glucose oxidase into the chitosan polymer using the glutaraldehyde crosslinker [100]. (b) The mechanism of chitosan and gelatin crosslinking using genipin [137]. (c) The mechanism of chitosan crosslinking using EDC/NHS [126]. (d) The crosslinking of sodium alginate with calcium chloride [138].
Figure 4. (a) The crosslinking process of glucose oxidase into the chitosan polymer using the glutaraldehyde crosslinker [100]. (b) The mechanism of chitosan and gelatin crosslinking using genipin [137]. (c) The mechanism of chitosan crosslinking using EDC/NHS [126]. (d) The crosslinking of sodium alginate with calcium chloride [138].
Chemistry 07 00061 g004

5. Biological Aspects of Crosslinkers

A wide selection of chemical crosslinkers, either natural or synthetic, has been already explored in various tissue engineering applications. Their biocompatibility and safety have been studied using both in vitro and in vivo methods. The choice of the crosslinker should be based on its ability to improve the properties of the scaffold without inducing any adverse effects on the tissues [139].
Glutaraldehyde (GTA) is one of the most frequently used synthetic crosslinkers in the generation of scaffold materials, including chitosan, gelatin, and collagen. It is highly stable, easily accessible, and inexpensive [140]. However, the presence of the aldehyde group has limited its use since it has induced cytotoxicity and inflammation [141,142,143,144]. To reduce these toxic effects, the aldehyde group can be washed off with a glycine or citric acid solution [145], and the concentration of GTA used could be limited to less than 8%, which has been determined to be biocompatible for the cells [146]. The use of GTA has been efficient in producing bioprostheses for heart valve replacement [145,147]. Several reports have developed GTA-crosslinked scaffolds and shown applications for their potential use in bone tissue engineering. These studies have been conducted on established cell lines, and some have been applied in vivo in mouse, rabbit, or rat models. Interestingly, a higher tensile strength and lower immunogenicity were shown in these GTA-crosslinked bone tissue scaffolds [139].
Other synthetic crosslinkers include carbodiimide agents. These have been used for generating scaffolds that show good biocompatibility for ophthalmic applications [148]. EDC/NHS crosslinked matrices are shown to be effective for bone and neural repair but in some cases induce low cell–ECM attachment [149,150,151]. Thus, it is highly advisable to reduce the concentration of EDC/NHS by up to 10-fold to obtain higher stability and mechanical strength [152]. Several studies have used EDC/NHS as crosslinker in the formation of bone matrices, with concentrations of 50 mM and 25 mM, respectively [139]. Dual crosslinking with GTA and EDC in decellularized human amniotic membranes and collagen scaffolds improves mechanical resistance with high biocompatibility, including stemness, adhesion, and survival of cells, leading to potential applications in tissue engineering [153,154].
Similar to GTA, epoxy compounds like BDDGE can exhibit some cytotoxicity. However, BDDGE is being used for skin regeneration by crosslinking gelatin nano-fibers with no apparent cytotoxicity [155]. In addition, collagen and chitosan crosslinked with epoxy compounds have demonstrated high mechanical and thermal stability in corneal progenitor cells and dermal cells [156,157].
To avoid the disadvantages of synthetic crosslinkers, the natural chemical genipin is being implemented in several studies for higher efficiency and biocompatibility. Compared to GTA, genipin has very low toxicity as demonstrated by MTT and proliferation assays, with 10000 times less cytotoxicity in 3T3 fibroblast cells [158,159]. With genipin, these cells can produce more collagen and have higher metabolic activity while modulating the growth and polarization of immune cells, leading to decreased immunogenicity [160]. The use of genipin has also shown good mechanical strength, higher cell uniformity, tissue growth, the secretion of glycosaminoglycans, and resistance to collagenase when compared to matrices crosslinked with GTA [161,162].
Genipin has been applied for the successful preparation of corneal lens implants [163], the repair of intervertebral discs using either fibrin hydrogel with silk scaffold or collagen scaffold [164,165], and for halting and repairing tendon tears [166,167]. Genipin crosslinked with elastin has also been effective in treating knee injuries in rabbit models [168].
Chitosan hydrogels crosslinked with genipin have shown high mechanical properties, improved swelling, and resistance to water. The biocompatibility of these hydrogels has been evaluated in vitro in 3T3 fibroblasts, macrophages, and dendritic cells and has revealed high cell viability and no cytotoxicity showing high potential for clinical applications [169,170,171]. Another study has used a chitosan–gelatin–polysaccharide scaffold crosslinked with genipin and has shown high neural cell proliferation and migration in vitro, indicating potential applications in neural regeneration to treat traumatic brain injuries [172]. Similarly, a spinal cord ECM matrix has been developed through genipin crosslinking for treating spinal cord injuries [173]. In addition, using genipin can increase stiffness in a collagen-based scaffold that is able to induce chondrogenesis [174]. Other studies have demonstrated the efficiency and biocompatibility of genipin-crosslinked scaffolds for applications in corneal tissue engineering [175], hepatic tissue engineering [176], and wound healing [177].
Genipin has also been used as a crosslinker in a hydroxyapatite–chitosan–L-arginine scaffold, leading to a 3-dimensional porous hydrogel that can be effectively applied for bone regeneration with high cell viability, as demonstrated by biocompatibility assays [178]. Various studies have evaluated the use of genipin-crosslinked scaffolds for applications in bone tissue engineering [139]. To ensure biocompatibility, data recommend keeping the dose below 0.5 mM for non-cytotoxic effects on cells [139,167,179]. Most of these scaffolds were based on chitosan, collagen, gelatin, and hydroxyapatite, and were tested on different types of established cell lines. Mouse and rabbit models have also been used to test the effectiveness of these scaffolds in vivo (reviewed in [139]). The concentrations of genipin used in these studies were variable, indicating that there is a need to optimize the concentration for tissue-specific applications. Together, the analysis of all data indicates that genipin use is considered effective and biocompatible when used at the optimal minimum concentration.
A range of other natural crosslinkers has been promising for use in tissue repair. Crosslinking using citric acid is able to promote the osteogenic differentiation of mesenchymal stem cells for bone regeneration [180] and improve mechanical strength and durability in methylcellulose hydrogel for applications in cell sheet engineering [181]. Tannic acid has been shown to be effective in wound healing applications with no adverse effects on cell function [182]. In addition, dual crosslinking of procyanidins with GTA has allowed the development of a vascular scaffold with higher mechanical strength and lower calcification for application in the tissue engineering of blood vessels [183]. The use of either cinnamaldehyde or epigallocatechin gallate crosslinkers increases the biomechanical strength of collagen scaffold in dental pulp cells, leading to improved cell proliferation, differentiation, and adhesion [184,185]. Epigallocatechin gallate has also been used to crosslink gelatin sponges, leading to improved bone formation [186].
In conclusion, with the continuous demand in regenerative medicine for optimal scaffolds for tissue engineering, it is important for scientists to keep on developing novel matrices based on crosslinked material that can match the microenvironment of the surrounding tissues while maintaining negligible cytotoxicity (Table 3).
Table 3. The main chemical crosslinkers used in tissue engineering applications, with some limitations and possible solutions.
Table 3. The main chemical crosslinkers used in tissue engineering applications, with some limitations and possible solutions.
CrosslinkerLimitations and Possible SolutionsApplicationsReferences
Glutaraldehyde
-
Cell toxicity and inflammation in vivo
-
Needs washing
-
Concentration used needs to be less than 8%
Bioprostheses for heart valve replacement
Bone tissue engineering
[145,147]
[139,145,147]
Carbodiimide
-
Expensive
-
May induce low cell adhesion
-
Concentration reduced 10-fold for biocompatibility
Ophthalmic applications
Bone tissue engineering
Wound healing
[148]
[139]
[139,148,187]
Epoxy Compounds
-
Cell toxicity
Bone tissue engineering
Skin regeneration
Ophthalmic applications
Wound healing
[139]
[155]
[156]
[139,155,156,157]
Genipin
-
Expensive
-
Concentration used is advised to be less than 0.5 mM
-
Need to optimize concentration for specific tissue applications
Corneal tissue engineering
In vertebral disc repair
Bone tissue engineering
Tendon tear repair
Knee injury treatment
Neural and spinal cord regeneration
Chondrogenesis
Hepatic tissue engineering
wound healing
[163,175]
[164,165]
[139,178]
[166,167]
[168]
[172,173]
[174]
[176]
[139,163,164,165,166,167,168,172,173,174,175,176,177,178]
Citric Acid
-
Might need catalyst
-
Might need high temperature for crosslinking
Bone tissue engineering
Cell sheet engineering
[180]
[181]
Tannic Acid
-
Need to optimize concentration, pH, and temperature conditions
Wound healing[182]

6. Chemical Aspects of Crosslinkers

The development and refinement of chemical crosslinkers have become the center of attention in materials science, polymer chemistry, and biomedical research. Recent examinations have introduced novel approaches to improve crosslinking effectiveness, strengthen structural durability, and encourage eco-friendly solutions (Figure 5). Many tools and methods are commonly used to measure crosslinking and crosslinking density in polymers: swelling tests, differential scanning calorimetry (DSC), infrared spectroscopy (FTIR), dynamic mechanical analysis (DMA), nuclear magnetic resonance (NMR), mechanical testing (tensile strength, compression set, etc.), and electron spin resonance (ESR). Often, a combination of techniques is used to obtain a comprehensive understanding of the material’s crosslinked network.
This section explores the major developments in sustainable crosslinking techniques, innovative hydrogel compositions, and advanced chromatographic materials while focusing on the strategies and optimization methods utilized.
A significant study in this area, conducted by Zhang and colleagues in 2024, explores advancements in sustainable dual-crosslinking polymer networks (DCPNs) [188]. The research explores the combination of dynamic covalent bonds, such as imine, disulfide, and ester bonds, with non-covalent forces, such as hydrogen bonding and coordination interactions, to enhance both mechanical performance and recyclability. The study optimizes crosslinking efficiency through molecular dynamics simulations and rheological assessments. Researchers were able to achieve a customizable balance between elasticity and rigidity by adjusting the proportion of covalent and non-covalent crosslinking elements, thus making these materials highly relevant for applications in flexible electronics and biomedical scaffolds.
The improvement of hydrogel technology has taken a significant leap forward with the introduction of formulations that do not require crosslinkers. A study by Alsaafeen et al. [189] in 2023 outlines a streamlined, single-step method for creating hydrogels using a combination of gelatin, chitosan, and glycerol. Instead of depending on conventional chemical crosslinkers, this approach exploits hydrogen bonding and hydrophobic interactions to achieve gel formation. Through techniques like differential scanning calorimetry (DSC) and rheometry, the researchers fine-tuned the gelation time and temperature, showing that a controlled cooling process strengthens the hydrogel network and enhances its mechanical properties. These insights are particularly valuable for biomedical fields such as tissue engineering and wound care, where eliminating synthetic crosslinkers reduces potential cytotoxic effects.
The effectiveness of linking proteins and polysaccharides depends significantly on the specific reaction conditions and type of crosslinking agents used. A study by Alavarse et al. [190] explored different crosslinking strategies, including enzymatic, chemical, and physical approaches. The researchers applied analytical techniques such as Fourier-transform infrared spectroscopy (FTIR) and nuclear magnetic resonance (NMR) spectroscopy to measure bond formation to evaluate crosslinking density. By adjusting factors like pH, ionic strength, and temperature, they enhanced gel strength and improved the resistance of protein–polysaccharide complexes to degradation. Such insights contribute to the advancement of biocompatible hydrogels for applications in drug delivery and tissue regeneration.
Achieving efficient crosslinking in chromatographic stationary phases is essential in the field of analytical chemistry. A study by Zhang et al. [191] investigated strategies to enhance the synthesis of hyper-crosslinked stationary phases tailored for ultra-fast, high-temperature, two-dimensional liquid chromatography. The work employed high-performance liquid chromatography (HPLC) to assess how varying crosslinker concentrations impact retention times and separation efficiency. By incorporating a modified Friedel–Crafts alkylation technique, they strengthened the polymer network, thus enabling it to function at elevated temperatures without deterioration. Adjustments to reaction durations and catalyst levels have further refined chromatographic resolution, highlighting its relevance for pharmaceutical and environmental applications.
In nanomedicine, researchers have made advancements in crosslinking techniques, where a recent study by Siddoway et al. (2023) introduced an innovative pentablock copolymer tailored for vaccine delivery [192]. The work utilized dynamic light scattering (DLS) and transmission electron microscopy (TEM) to analyze the size and structural characteristics of the nanoparticles. By employing a factorial experimental approach, the team optimized the balance between hydrophobic and hydrophilic segments, resulting in an enhanced immune response with minimal toxicity. The copolymer’s crosslinked core played a crucial role in prolonging antigen retention and enabled controlled release, thus representing a notable breakthrough in vaccine development.
In the field of tissue engineering, genipin, a naturally occurring compound known for its ability to facilitate crosslinking, has been extensively investigated. A study conducted by Pugliese et al. [193] in 2019 examined the formation of peptide-based hydrogels, utilizing genipin as a crosslinking agent. Through rheological assessments and analysis of the swelling behavior, the researchers refined the crosslinking parameters. By modifying the concentration of genipin and the duration of the reaction, they developed hydrogels with improved mechanical strength and controlled degradation rates. These optimized hydrogels demonstrated strong cell attachment and proliferation, making them promising for use in regenerative therapies and drug release systems. Recent developments and refinement of crosslinkers have led to pronounced enhancements in material durability, compatibility with biological systems, and eco-friendliness. Techniques such as computational modeling, experimental design strategies, and sophisticated spectroscopic methods have allowed for greater precision in regulating crosslinking mechanisms. Accordingly, research efforts should prioritize eco-friendly crosslinkers (i.e., the production of safe, biodegradable alternatives to conventional synthetic crosslinking agents), optimized multi-scale approaches (i.e., combining simulation techniques with experimental data to better predict and fine-tune crosslinking processes), and adaptive polymer networks (i.e., investigating crosslinkers that respond to environmental triggers like pH shifts, temperature variations, or external stimuli for versatile material applications). By promoting innovation through interdisciplinary research, scientists can continue to improve material performance and broaden the functional scope of next-generation polymer systems.
Crosslinkers are vital in shaping the mechanical properties, stability, and degradation patterns of polymers, especially in biomedical and structural contexts. By altering the concentration, form, and chemical composition of crosslinks, scientists can adjust the material’s durability, flexibility, and rate of degradation. Various studies have examined these influences, offering valuable knowledge for creating advanced materials with specific and desired performance traits.
Mechanical strength is a crucial characteristic of polymer networks since it defines their capacity to tolerate external stressors without deforming or failing. The mechanical behavior of crosslinked materials is primarily governed by factors such as crosslinker density, bond strength, and the structure of the polymer chains. In a study by Jiang et al. [194] in 2013, it was shown that varying crosslinking methods impacted the stiffness and toughness of polymer networks. Their work indicated that networks with a high degree of crosslinking exhibited greater stiffness due to a more compact polymer framework, which improved their ability to withstand more load. However, excessively dense crosslinking may lead to brittleness, making the material less capable of absorbing mechanical energy and more prone to cracking under stress. Further research emphasized that adjusting crosslinker density in hydrogels allows for fine-tuning of their mechanical properties. The work on degradable hydrogels revealed that increasing crosslink density boosted the material’s initial strength but could also slow down its degradation rate, which might affect its long-term effectiveness in biomedical applications [195,196].
Polymeric materials must retain their structural integrity over time, even when exposed to various mechanical, chemical, or environmental strains. A method to enhance the stability of chemically crosslinked hydrogels was developed by adjusting the chemistry of the crosslinkers. The findings revealed that hydrogels with covalent crosslinks resisted mechanical and hydrolytic breakdown more effectively, making them well suited for applications like tissue engineering scaffolds and drug delivery systems [197]. Similarly, Carberry et al. [198] explored how the chemical properties of crosslinkers influenced stability through their interactions with the surrounding environment. Hydrogels crosslinked with thioesters, for example, exhibited adjustable stability depending on factors such as thiol concentration and the pKa of the crosslinkers. These insights are valuable for designing biomaterials that should remain stable for a set period of time before undergoing controlled degradation.
The rate at which polymeric materials degrade plays a key role in biomedical fields, particularly in drug delivery systems and tissue regeneration. Ideally, degradation should occur in a controlled manner, thus ensuring that the material preserves structural integrity while gradually breaking down in response to specific environmental factors. In a 2007 study, the degradation behavior of chemically crosslinked hyaluronic acid hydrogels was examined, revealing that the choice of crosslinker significantly influences the degradation rate. The research highlighted that the breakdown of hydrogels can be regulated by selecting crosslinkers with certain chemical properties, such as a hydrolysable ester or disulfide bonds [199]. Likewise, another study developed an innovative biodegradable polymer crosslinker that enabled the independent regulation of stiffness, toughness, and the degradation rate. This finding facilitated the creation of materials that combined mechanical durability with predictable degradation, making them ideal for applications in soft tissue engineering and wound healing [200].
The research discussed above underlines the significant impact of crosslinker chemistry on the mechanical properties, stability, and degradation of polymeric materials. By carefully choosing the right crosslinkers and adjusting their concentrations, scientists can create materials with specific desired characteristics, making them suitable for use in fields like biomedical engineering, drug delivery, and advanced structural applications. The ability to adjust these properties provides new possibilities in the development of smart materials, where crosslinking methods are employed to design adaptable and responsive systems for specialized uses.
Figure 5. Approaches of different crosslinking processes providing different mechanical strengths and properties (A) bonding groups: thermal gelation, hydrogen bonding and ionic crosslinking and (B) crosslinking via radiation, enzymes or photocrosslinking [201].
Figure 5. Approaches of different crosslinking processes providing different mechanical strengths and properties (A) bonding groups: thermal gelation, hydrogen bonding and ionic crosslinking and (B) crosslinking via radiation, enzymes or photocrosslinking [201].
Chemistry 07 00061 g005

7. Medical Applications of Crosslinked Biomaterials and Biomembranes

The use of chemical crosslinkers has gained immense interest in the development of suitable scaffolds for tissue engineering and regenerative medicine and has been significantly implemented in advanced drug delivery technologies. These applications are being explored to overcome the use of traditional organ transplantation techniques, helping in replacing damaged tissues due to diseases, aging, or accidental injuries. Crosslinkers help in enhancing the biomechanical strength, physical properties, and the structural integrity of the scaffold materials, leading to improved biocompatibility and biodegradation. Crosslinked biomaterials have been especially employed in the fields of obstetrics and gynecology. These materials enable localized, controlled, and sustained therapeutic delivery, exemplified by drug-releasing intrauterine devices (IUDs). These devices offer the long-term delivery of contraceptive hormones, anti-inflammatory drugs, or antibiotics, reducing the reliance on systemic treatments and improving adherence to therapy [202]. For instance, crosslinked polymer matrices have been pivotal in the prolonged release of levonorgestrel, a widely used contraceptive hormone, providing an effective and low-maintenance solution for years [203,204]. Current innovations are focusing on biodegradable crosslinked materials that dissolve after releasing their drug payload, eliminating the need for device removal and enhancing patient convenience (Table 4).
Crosslinked hydrogels are increasingly being used for targeted drug delivery in the treatment of uterine disorders such as endometriosis and adenomyosis. These systems deliver anti-inflammatory and anti-fibrotic drugs directly to the uterine lining, ensuring precise treatment with minimal systemic effects. In managing postpartum hemorrhage, crosslinked systems are utilized to deliver hemostatic agents directly to the source of bleeding, improving treatment effectiveness [205]. In gynecological oncology, crosslinked nanoparticles are emerging as precision drug carriers, particularly in ovarian and endometrial cancer treatments. These nanoparticles can target specific cancer cell receptors, allowing for highly specific drug delivery with reduced toxicity [206]. Stimulus-responsive systems, which release drugs based on environmental triggers like pH or temperature, further enhance the precision and control of drug delivery.
Microneedle patches made from crosslinked polymers are being developed as non-invasive solutions for the transdermal delivery of hormones and vaccines, including contraceptives and HPV vaccines, offering a more patient-friendly alternative to injections [207]. Additionally, crosslinked hydrogels have been used to deliver tocolytic agents locally at the cervix to prevent preterm labor, improving outcomes for both mothers and infants while minimizing systemic exposure [208,209].
On the other hand, crosslinked biomaterials play a crucial role in enhancing wound healing and tissue regeneration after obstetric and gynecological surgeries. Crosslinked hydrogels and membranes serve as effective wound dressings by maintaining moisture, promoting cell growth, and protecting against infections [210,211]. For example, hydrogels made from chitosan have demonstrated benefits in cesarean wound care by accelerating tissue repair, minimizing scarring, and providing antimicrobial properties [212,213].
Postoperative adhesion prevention is another critical application. Crosslinked biomembranes act as physical barriers to reduce adhesion formation, a common complication that can lead to chronic pain, infertility, or bowel issues. Biodegradable membranes, such as those made from hyaluronic acid, have shown excellent results in clinical use for minimizing adhesions [214,215].
In reconstructive surgeries, particularly after gynecological cancer treatments, crosslinked scaffolds like silk fibroin promote cell growth and vascularization, making them ideal for vaginal reconstruction [216,217]. Similarly, alginate-based systems support tissue regeneration by encouraging angiogenesis, providing solutions for extensive reconstructive needs [218]. For conditions like Asherman’s syndrome, bioengineered uterine scaffolds made from crosslinked biomaterials help regenerate tissue, restore uterine function, and prevent adhesions [219].
Crosslinked biomaterials have enabled the creation of advanced implants for gynecological use. Biodegradable vaginal stents made from crosslinked polymers are used following reconstructive or prolapse surgeries, offering improved comfort and eliminating the need for removal, unlike traditional silicone stents [220]. Pelvic organ prolapse treatments now incorporate crosslinked collagen scaffolds to reinforce weakened tissues, reducing complications like erosion and infection associated with synthetic meshes [221,222]. Injectable hydrogels with crosslinked structures provide minimally invasive options for stress urinary incontinence, offering structural support and long-lasting results [209].
Crosslinked polymer meshes, enhanced with antimicrobial agents or growth factors, improve biocompatibility and reduce inflammation in urogynecological surgeries. Crosslinked elastomers are also being explored for cervical cerclage devices, offering better mechanical support during high-risk pregnancies [223].
In fertility preservation, crosslinked hydrogels encapsulate ovarian tissue during cryopreservation to maintain viability [224,225]. Artificial oocyte creation using crosslinked matrices is another area of research with the potential to improve outcomes for individuals undergoing fertility-compromising treatments [226]. Additionally, hydrogels serve as carriers during embryo transfer, providing a protective environment and optimizing implantation success rates [225,227].
Crosslinked biomaterials are transforming in vitro fertilization (IVF) by addressing challenges like oocyte preservation, embryo culture, and implantation. Hydrogels protect oocytes and embryos during cryopreservation, preserving their structure and viability, especially for women undergoing fertility-threatening treatments [225,228]. Encapsulation techniques improve post-thaw survival and developmental potential [229].
Three-dimensional scaffolds made from crosslinked polymers replicate the uterine environment, enhancing embryo growth and implantation success rates compared to traditional culture methods [230,231]. Crosslinked hydrogels are also used to deliver hormones and growth factors to the uterine lining, creating a favorable environment for embryo implantation [232,233].
Bioengineered uterine scaffolds address implantation challenges, such as scarring, by mimicking healthy uterine tissue and promoting regeneration [234,235]. Localized hormone delivery via crosslinked hydrogels further supports implantation while minimizing side effects [206].
Research into artificial gametes and ovarian tissue preservation using crosslinked biomaterials offers new hope for patients with diminished ovarian reserves. These advancements represent a significant shift in reproductive medicine, providing innovative solutions to overcome infertility [236].
Table 4. Examples of crosslinked biomaterials in obstetrics and gynecology applications.
Table 4. Examples of crosslinked biomaterials in obstetrics and gynecology applications.
ApplicationMaterialFunctionReferences
Drug-Eluting IUDsCrosslinked polymersSustained release of contraceptive hormones[237]
Postpartum Hemorrhage ControlCrosslinked hydrogelsLocalized delivery of hemostatic agents[238,239]
Cesarean Section Wound DressingChitosan hydrogelsEnhanced healing and antimicrobial activity[211]
Adhesion Prevention MembranesHyaluronic acid membranesPrevention of postoperative adhesions[240]
Vaginal StentsCrosslinked biodegradable polymersStructural support post-surgery[241]
Injectable Hydrogels for IncontinenceCrosslinked hydrogelsMinimally invasive structural support[209]
Cancer Drug Delivery SystemsCrosslinked nanoparticlesTargeted therapy for gynecological cancers[206]
Reconstructive Surgery ScaffoldsCrosslinked silk fibroinSupport tissue regeneration[216,217]
Cervical Cerclage DevicesCrosslinked elastomersMechanical support during pregnancy[241,242]
Many crosslinked materials have been FDA-approved, especially in the biomedical and drug delivery feed. The most important ones are as follows:
-
Restasis (Cyclosporine A): A formulation for treating dry eye disease using hydrogel crosslinked based systems for sustained drug release.
-
Vicryl (Polyglactin 910): A crosslinked polymer suture material used in both absorbable and non-absorbable forms.
-
Monocryl (Poliglecaprone 25): A monofilament absorbable suture made from a crosslinked polymer, designed for fast absorption and minimal tissue irritation.
-
Duramorph (morphine sulfate): A sustained-release form of morphine delivered via a crosslinked polymeric matrix used in certain pain management systems.
-
Lupron Depot (leuprolide acetate): A crosslinked PLGA-based depot formulation for the controlled release of leuprolide in the treatment of prostate cancer and endometriosis.
-
Depo-Provera (medroxyprogesterone acetate): An injectable contraceptive formulation using crosslinked PLGA for sustained release over months.
-
Aquacel (hydrocolloid dressing): Made from crosslinked sodium carboxymethylcellulose, it helps in absorbing exudates and promoting healing in chronic wounds.
-
Marqibo (vincristine sulfate liposome): A liposomal formulation of vincristine for cancer treatment.
-
Cypher (sirolimus-eluting stent): A drug-eluting stent that uses crosslinked polymer coatings to release sirolimus for the prevention of restenosis.
-
Tisseel (fibrin sealant): A crosslinked fibrin-based adhesive used in surgery to help control bleeding and promote tissue healing.

8. Recent Advances and Innovations

The design and the development of biomaterials and biomembranes have seen significant advancements through innovative crosslinking techniques. Among these, light-induced and enzyme-mediated crosslinking techniques have gained more attention due to their precision, biocompatibility, and adaptability to diverse biomedical and bioengineering applications.
Recently, light-crosslinking techniques have been used to design hydrogels for cartilage regeneration applications because of their capacity to control the location of crosslinking and the timing under physiological environments [243]. Light is a clean energy source that can be used specifically with regard to wavelength (in the ultraviolet UV to infrared IR region) [244]. In this technique, two polymer chains are covalently bounded after exposure to light irradiation in the presence of photo-initiators. The light irradiation offers fast crosslinking within seconds to minutes, making it ideal for use in 3D-bioprinting, dental materials, and rapid polymerization. However, biological changes caused by UV radiation have long been considered harmful, such as in cases caused by chemical agents, but have also been recognized as a benefit [245,246]. Many studies reported in the literature have shown the potential dose–response cytotoxicity of some photoinitiators, and there are several biosafety concerns associated with the application of UV light when it comes to using living cells and tissues. Long exposure times and high light intensities damage the cell’s DNA and may generate reactive oxygen species that induce accelerated tissue aging and/or immunosuppression [247]. This challenge was overcome by the use of longer wavelengths, including visible light, in conjunction with a photoinitiator that can be activated by those wavelengths [248].
Recent efforts have focused on the search for mild crosslinking methods that can take place in the presence of cells by means of non-cytotoxic reactions. Enzymatic crosslinking is the most cytocompatible process that preserves cell viability and bioactivity since this technique relies on the use of a biocatalyst that drives the formation of covalent bonds between the biomolecules and non-exogenous reagents that are used. As enzymes naturally exist in the body, the biocompatibility of enzymatically crosslinked hydrogels is higher than that of chemical hydrogels crosslinked via other methods, such as photo, thermal, or chemical–catalyst crosslinking (Figure 6). Enzyme-mediated crosslinking offers a “greener” approach to covalent bond formation in biology and biotechnology. This could include the use of aqueous solvents in mild conditions, at room temperature, and with heavy metal free reagents, making it ideal for biomedical applications, food science, and tissue engineering [245,249].
Figure 6. Enzyme-mediated crosslinking processes using (a) Transglutaminase, (b) Tyrosinase, (c) Horseradish peroxidase and (d) Lysyl oxidase (reproduced from [243] under the license number 5998490451810).
Figure 6. Enzyme-mediated crosslinking processes using (a) Transglutaminase, (b) Tyrosinase, (c) Horseradish peroxidase and (d) Lysyl oxidase (reproduced from [243] under the license number 5998490451810).
Chemistry 07 00061 g006
Enzyme-mediated crosslinking using tyrosinases, transferases, or peroxidases have proven its efficiency and has attracted increasing attention for application in polymer hydrogel synthesis due to the environmentally friendly process and the possibility of obtaining biomaterials [250].

9. Challenges and Future Directions

The increasing demand for biomaterials, the development of new biomembranes, and the growth rate of their markets make it crucial to develop new crosslinkers that better suit the applications of advanced biomaterials and biomembranes while addressing the issues related to biocompatibility and production scalability. Among the list of new crosslinkers, click chemistry-based crosslinkers, photo-crosslinkable crosslinkers, enzyme-based crosslinkers, and self-healing crosslinkers are included.
Click chemistry-based crosslinkers are considered as highly efficient and sensitive crosslinkers [251]. They retained attention due to their specificity, minimizing by-products, and their ability to function under mild conditions. Key types of click chemistry-based reactions for crosslinking are as follows: copper-catalyzed azide–alkyne cycloaddition (CuAAC) [252], copper-free alkyne cycloaddition (SPAAC) [253,254], the Diels–Alder (DA) reaction [255], thiol–ene reactions [256], and tetrazine–norbornene ligation [257]. The main difference between these click chemistry-crosslinkers remains in the functions that are involved in the bonding and the formation of the new structure (the reaction between an azide group (−N3) and an alkyne group (−C≡C) with or without a catalyst (copper), cycloaddition between a diene group and a dienophile, the reaction between a thiol group (−SH) and an ene group (C=C), and the reaction between tetrazines and strained norbornene derivatives). However, the main advantage includes the biocompatibility issue, particularly for those functioning as copper-free crosslinkers (without catalyst), making them tolerated in biological systems [258,259].
The photo-crosslinkable crosslinkers refer to chemicals that can be activated by ultraviolet (UV) light, leading to the formation of covalent bonds between molecules. Typically, this type of crosslinker requires an activator to initiate the bonding and is inactive without a light inducer [260]. The use of light eliminates the need for high temperatures or harsh chemicals, making the process of crosslinking safe and more environmentally friendly. Some types of chemical compounds are necessary to allow the covalent bonds formation, e.g., cinnamate groups [261], azide groups [262], diazirines, and benzoin derivates [263]. Depending on the target application, the chemical compound is selected to create the bonds while achieving mechanical strength and the required stability. Cinnamate groups are mainly used in creating photo-responsive hydrogels, coatings for medical devices, or controlled drug delivery systems [264], while azide-based crosslinkers are generally utilized in bioconjugation and biomaterial engineering. On the other hand, diazirine-based and benzoin derivatives are used for studying protein–ligan interactions [265], in creating photo-cross-linked biopolymers, and in dental materials and 3D-printing materials that can be cured layer-by-layer using light [266]. Finally, the main challenges and considerations that should be taken into account while using photo-crosslinkable crosslinkers are the following: First, the degree of light penetration, where the ability of crosslinking depends on the ability of light to penetrate in deep homogenous materials. Second, the precision of the crosslinking process to ensure uniform crosslinking is sometimes challenging in complex geometries and when dealing with biological systems. The last challenge is the toxicity of some compounds like azides or diazirines that require special attention when used in some applications.
Enzyme-based crosslinkers seem to be the safest in the list of crosslinkers since they rely on the use of natural biological agents, enzymes to catalyze the formation of covalent bonds between molecules [249]. The main interest in using this type of crosslinker is the high level of biocompatibility and the specificity of the bonding formation. In addition, enzyme-based crosslinking occurs at neutral pH in the absence of harsh chemicals (mild conditions) and produces fewer by-products than the typical chemical crosslinker process. The common enzyme-catalyzed crosslinking reaction includes three types of reactions: (1) oxidative crosslinking using oxidase enzymes (e.g., laccases, peroxidases, etc.) [267,268], (2) transamination using transaminases or aminotransferases (e.g., transglutaminases) [269,270], and (3) esterification or amide bond formation using esterases [271,272] or transglutaminases enzymes [269,273]. Based on the target application (biomaterials and tissue engineering, pharmaceutical applications, food industry [274], environmental remediation, etc.), the enzyme is selected to create the bonds while achieving mechanical strength and the required stability [275]. However, the main limitations of these processes are related to the loss of enzymatic activity at extreme conditions or in the presence of high concentrations of salts and the cost that can be generated for scalability while using purified enzymes in large-scale industrial applications.
Finally, self-healing crosslinkers refer to agents or materials that can restore their original structures and properties after damage by physical stress or environmental factors [276]. Even if they are relatively less used than other crosslinkers agents for biological applications, they may offer innovative solutions for enhancing the durability, longevity, and functionality of materials for industrial applications [277]. They usually work through a combination of dynamic covalent bonds, non-covalent interactions, or reversible chemical reactions [278,279]. Self-healing crosslinkers involve Diels–Alder reactions, ionic bonding in polymers, reversible hydrogen bonding, microencapsulation and vascular networks, and polymer networks with reversible crosslinks. On the other hand, it is challenging and complex to design materials with self-healing properties, especially in terms of achieving controlled and reproducible healing under various conditions. The cost of the self-healing materials can also be expensive due to the complexity of manufacturing, which takes this factor into consideration.
Future directions on the role of crosslinkers in biology, chemistry, and medicine will focus on improving the crosslinking efficiency and selectivity. At the same time, addressing biocompatibility issues will be important to develop more eco-friendly and less cytotoxic crosslinkers in Figure 7 [280]. This will enable the use of more biomaterials and the development of innovative bio-membranes while maintaining affordable fabrication and distribution costs.
Figure 7. Using click chemistry to crosslink refillable depots. Left to middle: azide–alginate strands are combined and injected into target tissues and crosslinked in situ. Middle to right: intravascular administration of cyclooctyne-conjugated therapeutics allows selective capture and display of drug at gel site (reproduced with permission from [280] under license number 5998491504493).
Figure 7. Using click chemistry to crosslink refillable depots. Left to middle: azide–alginate strands are combined and injected into target tissues and crosslinked in situ. Middle to right: intravascular administration of cyclooctyne-conjugated therapeutics allows selective capture and display of drug at gel site (reproduced with permission from [280] under license number 5998491504493).
Chemistry 07 00061 g007

10. Conclusions

In conclusion, biomaterials and biomembranes enhanced by the various types of crosslinkers continue to revolutionize different fields (chemistry, biology, and medicine). Crosslinkers play an important role in improving the structural integrity and functionality of biomaterials and in biomembranes design, leading to significant advancements in applications such as tissue engineering, drug delivery, and wound healing. Despite the promising progress, the toxicity of many crosslinkers and the need for more specificity in crosslinking mechanisms remain a challenging need in order to achieve a biocompatible and eco-friendly process. Future research should focus on developing more biocompatible crosslinkers, exploring more synthesis methods, and optimizing their interaction with biological systems to express their full potential. The integration of crosslinkers into biomaterials design will be key to achieving effective, personalized, and sustainable medical solutions.

Author Contributions

Conceptualization, P.Y. and A.C.; methodology, P.Y. and A.C.; validation, A.C.; investigation, A.C.; resources, A.C.; writing—original draft preparation, P.Y., A.E.S., R.K., P.J.O., Z.N., T.T., N.S.-C., A.M. and A.C.; writing—review and editing, P.Y. and A.C.; funding acquisition, H.E.-N. All authors have read and agreed to the published version of the manuscript.

Funding

The authors received no financial support for the research, authorship, and/or publication of this article.

Conflicts of Interest

Author Ayman Chmayssem is the founder of the company Electrochemistry Consulting & Services (E2CS). The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest.

References

  1. Saini, M.; Singh, Y.; Arora, P.; Arora, V.; Jain, K. Implant biomaterials: A comprehensive review. World J. Clin. Cases WJCC 2015, 3, 52. [Google Scholar] [CrossRef]
  2. Silva-López, M.S.; Alcántara-Quintana, L.E. The Era of Biomaterials: Smart Implants? ACS Appl. Bio Mater. 2023, 6, 2982–2994. [Google Scholar] [CrossRef] [PubMed]
  3. Lee, E.J.; Kasper, F.K.; Mikos, A.G. Biomaterials for Tissue Engineering. Ann. Biomed. Eng. 2013, 42, 323. [Google Scholar] [CrossRef]
  4. Filip, N.; Radu, I.; Veliceasa, B.; Filip, C.; Pertea, M.; Clim, A.; Pinzariu, A.C.; Drochioi, I.C.; Hilitanu, R.L.; Serban, I.L. Biomaterials in Orthopedic Devices: Current Issues and Future Perspectives. Coatings 2022, 12, 1544. [Google Scholar] [CrossRef]
  5. Fenton, O.S.; Olafson, K.N.; Pillai, P.S.; Mitchell, M.J.; Langer, R. Advances in Biomaterials for Drug Delivery. Adv. Mater. 2018, 30, e1705328. [Google Scholar] [CrossRef] [PubMed]
  6. Trucillo, P. Biomaterials for Drug Delivery and Human Applications. Materials 2024, 17, 456. [Google Scholar] [CrossRef]
  7. Jia, B.; Huang, H.; Dong, Z.; Ren, X.; Lu, Y.; Wang, W.; Zhou, S.; Zhao, X.; Guo, B. Degradable biomedical elastomers: Paving the future of tissue repair and regenerative medicine. Chem. Soc. Rev. 2024, 53, 4086–4153. [Google Scholar] [CrossRef]
  8. Festas, A.J.; Ramos, A.; Davim, J.P. Medical devices biomaterials—A review. J. Mater. Des. Appl. 2019, 234, 218–228. [Google Scholar] [CrossRef]
  9. Al Minnath, M. Metals and alloys for biomedical applications. In Fundamental Biomaterials: Metals; Elsevier: Amsterdam, The Netherlands, 2018; pp. 167–174. [Google Scholar] [CrossRef]
  10. Satchanska, G.; Davidova, S.; Petrov, P.D. Natural and Synthetic Polymers for Biomedical and Environmental Applications. Polymers 2024, 16, 1159. [Google Scholar] [CrossRef]
  11. Punj, S.; Singh, J.; Singh, K. Ceramic biomaterials: Properties, state of the art and future prospectives. Ceram. Int. 2021, 47, 28059–28074. [Google Scholar] [CrossRef]
  12. Egbo, M.K. A fundamental review on composite materials and some of their applications in biomedical engineering. J. King Saud Univ.-Eng. Sci. 2021, 33, 557–568. [Google Scholar] [CrossRef]
  13. Sharma, S.; Sudhakara, P.; Singh, J.; Ilyas, R.A.; Asyraf, M.R.M.; Razman, M.R. Critical Review of Biodegradable and Bioactive Polymer Composites for Bone Tissue Engineering and Drug Delivery Applications. Polymers 2021, 13, 2623. [Google Scholar] [CrossRef] [PubMed]
  14. McMillin, C.R. Biomedical Applications of Rubbers and Elastomers. Rubber Chem. Technol. 2006, 79, 500–519. [Google Scholar] [CrossRef]
  15. Caldorera-Moore, M.; Peppas, N.A. Micro- and nanotechnologies for intelligent and responsive biomaterial-based medical systems. Adv. Drug Deliv. Rev. 2009, 61, 1391–1401. [Google Scholar] [CrossRef] [PubMed]
  16. Long, S.; Xie, C.; Lu, X. Natural polymer-based adhesive hydrogel for biomedical applications. Biosurf. Biotribol. 2022, 8, 69–94. [Google Scholar] [CrossRef]
  17. Farago, O. A Beginner’s Short Guide to Membrane Biophysics. In Modeling Biomaterials; Necas Center Series; Part F1669; Birkhäuser: Cham, Switzerland, 2021; pp. 1–41. [Google Scholar] [CrossRef]
  18. Marsh, D. Biomembranes. In Supramolecular Structure and Function; Springer: Boston, MA, USA, 1983; pp. 127–178. [Google Scholar] [CrossRef]
  19. Obeid, P.J.; Yammine, P.; El-Nakat, H.; Kassab, R.; Tannous, T.; Nasr, Z.; Maarawi, T.; Dahdah, N.; El Safadi, A.; Mansour, A.; et al. Organ-On-A-Chip Devices: Technology Progress and Challenges. ChemBioChem 2024, 25, e202400580. [Google Scholar] [CrossRef]
  20. Schumann, J. Molecular Mechanism of Cellular Membranes for Signal Transduction. Membranes 2022, 12, 748. [Google Scholar] [CrossRef]
  21. Biomembranes: Molecular Structure and Function—Robert B. Gennis—Google Books. Available online: https://books.google.com.lb/books?hl=en&lr=&id=RJbkBwAAQBAJ&oi=fnd&pg=PA1&dq=biomembranes+and+mechanical+support+scholarly+articles&ots=FZws4d4hR5&sig=FJwmWhMZyzUvZl19ytVD32a8pbI&redir_esc=y#v=onepage&q&f=false (accessed on 19 January 2025).
  22. Gatenby, R.A. The Role of Cell Membrane Information Reception, Processing, and Communication in the Structure and Function of Multicellular Tissue. Int. J. Mol. Sci. 2019, 20, 3609. [Google Scholar] [CrossRef]
  23. Watson, H. Biological membranes. Essays Biochem. 2015, 59, 43. [Google Scholar] [CrossRef]
  24. Gu, Y.; Du, L.; Wu, Y.; Qin, J.; Gu, X.; Guo, Z.; Li, Y. Biomembrane-Modified Biomimetic Nanodrug Delivery Systems: Frontier Platforms for Cardiovascular Disease Treatment. Biomolecules 2024, 14, 960. [Google Scholar] [CrossRef]
  25. Chmayssem, A.; Monsalve-Grijalba, K.; Alias, M.; Mourier, V.; Vignoud, S.; Scomazzon, L.; Muller, C.; Barthes, J.; Vrana, N.E.; Mailley, P. Reference method for off-line analysis of nitrogen oxides in cell culture media by an ozone-based chemiluminescence detector. Anal. Bioanal. Chem. 2021, 413, 1383–1393. [Google Scholar] [CrossRef]
  26. Membrane Systems: For Bioartificial Organs and Regenerative Medicine—Loredana De Bartolo, Efrem Curcio, Enrico Drioli—Google Books. Available online: https://books.google.com.lb/books?hl=en&lr=&id=b34oDwAAQBAJ&oi=fnd&pg=PR5&dq=biomembranes+tissue+engineering+and+regenerative+medicine&ots=Hv2U_p8_ZL&sig=W15GrrVdle5Hn8h0oaljB035Vkg&redir_esc=y#v=onepage&q&f=false (accessed on 19 January 2025).
  27. Silva, M.J.; Gonçalves, C.P.; Galvão, K.M.; D’Alpino, P.H.P.; Nascimento, F.D. Synthesis and Characterizations of a Collagen-Rich Biomembrane with Potential for Tissue-Guided Regeneration. Eur. J. Dent. 2019, 13, 295–302. [Google Scholar] [CrossRef]
  28. Selivanovitch, E.; Ostwalt, A.; Chao, Z.; Daniel, S. Emerging Designs and Applications for Biomembrane Biosensors. Annu. Rev. Anal. Chem. 2024, 17, 339–366. [Google Scholar] [CrossRef] [PubMed]
  29. Mattson, G.; Conklin, E.; Desai, S.; Nielander, G.; Savage, M.D.; Morgensen, S. A practical approach to crosslinking. Mol. Biol. Rep. 1993, 17, 167–183. [Google Scholar] [CrossRef] [PubMed]
  30. Krishnakumar, G.S.; Sampath, S.; Muthusamy, S.; John, M.A. Importance of crosslinking strategies in designing smart biomaterials for bone tissue engineering: A systematic review. Mater. Sci. Eng. C 2019, 96, 941–954. [Google Scholar] [CrossRef]
  31. Ma, B.; Wang, X.; Wu, C.; Chang, J. Crosslinking strategies for preparation of extracellular matrix-derived cardiovascular scaffolds. Regen. Biomater. 2014, 1, 81–99. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Z.; Wu, J.; Shi, X.; Song, F.; Gao, W.; Liu, S. Stereocomplexation of Poly(lactic acid) and Chemical Crosslinking of Ethylene Glycol Dimethacrylate (EGDMA) Double-Crosslinked Temperature/pH Dual Responsive Hydrogels. Polymers 2020, 12, 2204. [Google Scholar] [CrossRef]
  33. Nowatzki, P.J.; Tirrell, D.A. Physical properties of artificial extracellular matrix protein films prepared by isocyanate crosslinking. Biomaterials 2004, 25, 1261–1267. [Google Scholar] [CrossRef]
  34. Bigi, A.; Cojazzi, G.; Panzavolta, S.; Roveri, N.; Rubini, K. Stabilization of gelatin films by crosslinking with genipin. Biomaterials 2002, 23, 4827–4832. [Google Scholar] [CrossRef]
  35. Decker, C. Photoinitiated crosslinking polymerisation. Prog. Polym. Sci. 1996, 21, 593–650. [Google Scholar] [CrossRef]
  36. Mirtič, J.; Ilaš, J.; Kristl, J. Influence of different classes of crosslinkers on alginate polyelectrolyte nanoparticle formation, thermodynamics and characteristics. Carbohydr. Polym. 2018, 181, 93–102. [Google Scholar] [CrossRef] [PubMed]
  37. Moghaddam, S.Y.Z.; Biazar, E.; Esmaeili, J.; Kheilnezhad, B.; Goleij, F.; Heidari, S. Tannic Acid as a Green Cross-linker for Biomaterial Applications. Mini-Rev. Med. Chem. 2022, 23, 1320–1340. [Google Scholar] [CrossRef]
  38. Wong, S.S.; Wong, L.J.C. Chemical crosslinking and the stabilization of proteins and enzymes. Enzyme Microb. Technol. 1992, 14, 866–874. [Google Scholar] [CrossRef]
  39. Manickam, B.; Sreedharan, R.; Elumalai, M. ‘Genipin’—The Natural Water Soluble Cross-linking Agent and Its Importance in the Modified Drug Delivery Systems: An Overview. Curr. Drug Deliv. 2014, 11, 139–145. [Google Scholar] [CrossRef]
  40. Cha, C.; Kohman, R.E.; Kong, H. Biodegradable Polymer Crosslinker: Independent Control of Stiffness, Toughness, and Hydrogel Degradation Rate. Adv. Funct. Mater. 2009, 19, 3056–3062. [Google Scholar] [CrossRef]
  41. Malhotra, B.D.; Ali, M.A. Nanostructured Biomaterials for In Vivo Biosensors. In Nanomaterials for Biosensors; Elsevier: Amsterdam, The Netherlands, 2018; pp. 183–219. [Google Scholar] [CrossRef]
  42. Alizadeh-Osgouei, M.; Li, Y.; Wen, C. A comprehensive review of biodegradable synthetic polymer-ceramic composites and their manufacture for biomedical applications. Bioact. Mater. 2019, 4, 22–36. [Google Scholar] [CrossRef]
  43. Stepanova, M.; Korzhikova-Vlakh, E. Modification of Cellulose Micro- and Nanomaterials to Improve Properties of Aliphatic Polyesters/Cellulose Composites: A Review. Polymers 2022, 14, 1477. [Google Scholar] [CrossRef]
  44. Nakayama, A.; Yamano, N.; Kawasaki, N. Biodegradation in seawater of aliphatic polyesters. Polym. Degrad. Stab. 2019, 166, 290–299. [Google Scholar] [CrossRef]
  45. Janib, S.M.; Moses, A.S.; MacKay, J.A. Imaging and drug delivery using theranostic nanoparticles. Adv. Drug Deliv. Rev. 2010, 62, 1052–1063. [Google Scholar] [CrossRef]
  46. Chen, B.K.; Shen, C.H.; Chen, S.C.; Chen, A.F. Ductile PLA modified with methacryloyloxyalkyl isocyanate improves mechanical properties. Polymer 2010, 51, 4667–4672. [Google Scholar] [CrossRef]
  47. Liu, Q.; Jiang, L.; Shi, R.; Zhang, L.Z. Synthesis, preparation, in vitro degradation, and application of novel degradable bioelastomers—A review. Prog. Polym. Sci. 2012, 37, 715–765. [Google Scholar] [CrossRef]
  48. Leroy, A.; Pinese, C.; Bony, C.; Garric, X.; Noël, D.; Nottelet, B.; Coudane, J. Investigation on the properties of linear PLA-poloxamer and star PLA-poloxamine copolymers for temporary biomedical applications. Mater. Sci. Eng. C 2013, 33, 4133–4139. [Google Scholar] [CrossRef] [PubMed]
  49. Kowalski, P.S.; Bhattacharya, C.; Afewerki, S.; Langer, R. Smart Biomaterials: Recent Advances and Future Directions. ACS Biomater. Sci. Eng. 2018, 4, 3809–3817. [Google Scholar] [CrossRef]
  50. Ju, Y.M.; Choi, J.S.; Atala, A.; Yoo, J.J.; Lee, S.J. Bilayered scaffold for engineering cellularized blood vessels. Biomaterials 2010, 31, 4313–4321. [Google Scholar] [CrossRef]
  51. Shafiei, S.; Omidi, M.; Nasehi, F.; Golzar, H.; Mohammadrezaei, D.; Rezai Rad, M.; Khojasteh, A. Egg shell-derived calcium phosphate/carbon dot nanofibrous scaffolds for bone tissue engineering: Fabrication and characterization. Mater. Sci. Eng. C 2019, 100, 564–575. [Google Scholar] [CrossRef] [PubMed]
  52. Nisticò, R. Block copolymers for designing nanostructured porous coatings. Beilstein J. Nanotechnol. 2018, 9, 2332–2344. [Google Scholar] [CrossRef]
  53. Silva, A.C.Q.; Silvestre, A.J.D.; Vilela, C.; Freire, C.S.R. Natural polymers-based materials: A contribution to a greener future. Molecules 2022, 27, 94. [Google Scholar] [CrossRef]
  54. Marrink, S.J.; Corradi, V.; Souza, P.C.T.; Ingólfsson, H.I.; Tieleman, D.P.; Sansom, M.S.P. Computational Modeling of Realistic Cell Membranes. Chem. Rev. 2019, 119, 6184–6226. [Google Scholar] [CrossRef]
  55. Strathmann, H. Synthetic Membranes and Their Preparation. In Synthetic Membranes: Science, Engineering and Applications; Springer: Dordrecht, The Netherlands, 1986. [Google Scholar]
  56. Mansoori, S.; Davarnejad, R.; Matsuura, T.; Ismail, A.F. Membranes Based on Non-Synthetic (Natural) Polymers for Wastewater Treatment; Elsevier Ltd.: Amsterdam, The Netherlands, 2020. [Google Scholar] [CrossRef]
  57. Dong, Y.; Wu, H.; Yang, F.; Gray, S. Cost and efficiency perspectives of ceramic membranes for water treatment. Water Res. 2022, 220, 118629. [Google Scholar] [CrossRef]
  58. Abdullayev, A.; Bekheet, M.F.; Hanaor, D.A.H.; Gurlo, A. Materials and applications for low-cost ceramic membranes. Membranes 2019, 9, 105. [Google Scholar] [CrossRef]
  59. Vroman, I.; Tighzert, L. Biodegradable Polymers. Materials 2009, 2, 307–344. [Google Scholar] [CrossRef]
  60. Galiano, F.; Briceño, K.; Marino, T.; Molino, A.; Christensen, K.V.; Figoli, A. Advances in biopolymer-based membrane preparation and applications. J. Membr. Sci. 2018, 564, 562–586. [Google Scholar] [CrossRef]
  61. Shaari, N.; Kamarudin, S.K. Chitosan and alginate types of bio-membrane in fuel cell application: An overview. J. Power Sources 2015, 289, 71–80. [Google Scholar] [CrossRef]
  62. Wu, P.; Imai, M. Novel Biopolymer Composite Membrane Involved with Selective Mass Transfer and Excellent Water Permeability. In Advancing Desalination; InTech: London, UK, 2012. [Google Scholar] [CrossRef]
  63. Ma, J.; Sahai, Y. Chitosan biopolymer for fuel cell applications. Carbohydr. Polym. 2013, 92, 955–975. [Google Scholar] [CrossRef]
  64. Pawar, S.N.; Edgar, K.J. Alginate derivatization: A review of chemistry, properties and applications. Biomaterials 2012, 33, 3279–3305. [Google Scholar] [CrossRef]
  65. Li, G.; Li, W.; Deng, H.; Du, Y. Structure and properties of chitin/alginate blend membranes from NaOH/urea aqueous solution. Int. J. Biol. Macromol. 2012, 51, 1121–1126. [Google Scholar] [CrossRef]
  66. Ng, W.C.; Lokanathan, Y.; Fauzi, M.B.; Baki, M.M.; Zainuddin, A.A.; Phang, S.J.; Azman, M. In vitro evaluation of genipin-crosslinked gelatin hydrogels for vocal fold injection. Sci. Rep. 2023, 13, 5128. [Google Scholar] [CrossRef]
  67. Thi, H.N.; Nguyen, D.H.; Vu, M.T.; Tran, H.N.; Pham Tran, L.P.; Nguyen-Thi, N.T.; Le, N.T.T.; Le Minh Tri, N. Fabrication process and characterization of AgNPs/PVA/cellulose as a SERS platform for in-situ detection of residual pesticides in fruit. Mater. Res. Express 2020, 7, 035019. [Google Scholar] [CrossRef]
  68. Maikovych, O.; Pasetto, P.; Nosova, N.; Kudina, O.; Ostapiv, D.; Samaryk, V.; Varvarenko, S. Functional Properties of Gelatin–Alginate Hydrogels for Use in Chronic Wound Healing Applications. Gels 2025, 11, 174. [Google Scholar] [CrossRef]
  69. Oliveira, J.R.; Lopes, D.S.; Barbosa, M.C.S.; Silva, H.N.; Fook, M.V.L.; Silva, S.M.L.; Delgado, J.M.P.Q.; Lima, A.G.B. Chitosan Nanoparticulate System Loaded with Cannabidiol: A Topical Formulation for Potential Alopecia Management. Processes 2025, 13, 617. [Google Scholar] [CrossRef]
  70. Matsuyama, H.; Ohga, K.; Maki, T.; Tearamoto, M.; Nakatsuka, S. Porous cellulose acetate membrane prepared by thermally induced phase separation. J. Appl. Polym. Sci. 2003, 89, 3951–3955. [Google Scholar] [CrossRef]
  71. Puspasari, T.; Peinemann, K.-V. Application of thin film cellulose composite membrane for dye wastewater reuse. J. Water Process Eng. 2016, 13, 176–182. [Google Scholar] [CrossRef]
  72. Jiang, X.; Wu, L.; Zhang, M.; Zhang, T.; Chen, C.; Wu, Y.; Yin, C.; Gao, J. Biomembrane nanostructures: Multifunctional platform to enhance tumor chemoimmunotherapy via effective drug delivery. J. Control. Release 2023, 361, 510–533. [Google Scholar] [CrossRef] [PubMed]
  73. Mansourizadeh, A.; Azad, A.J. Preparation of blend polyethersulfone/cellulose acetate/polyethylene glycol asymmetric membranes for oil-water separation. J. Polym. Res. 2014, 21, 375. [Google Scholar] [CrossRef]
  74. Abedini, R.; Mousavi, S.M.; Aminzadeh, R. A novel cellulose acetate (CA) membrane using TiO2 nanoparticles: Preparation, characterization and permeation study. Desalination 2011, 277, 40–45. [Google Scholar] [CrossRef]
  75. Mukherjee, R.; De, S. Adsorptive removal of phenolic compounds using cellulose acetate phthalate-alumina nanoparticle mixed matrix membrane. J. Hazard Mater. 2014, 265, 8–19. [Google Scholar] [CrossRef]
  76. Wu, Y.B.; Yu, S.H.; Mi, F.L.; Wu, C.W.; Shyu, S.S.; Peng, C.K.; Chao, A.C. Preparation and characterization on mechanical and antibacterial properties of chitsoan/cellulose blends. Carbohydr. Polym. 2004, 57, 435–440. [Google Scholar] [CrossRef]
  77. Duan, J.; Han, C.; Liu, L.; Jiang, J.; Li, J.; Li, Y.; Guan, C. Binding cellulose and chitosan via intermolecular inclusion interaction: Synthesis and characterisation of gel. J. Spectrosc. 2015, 2015, 179258. [Google Scholar] [CrossRef]
  78. Khosa, M.A.; Shah, S.S.; Feng, X. Metal sericin complexation and ultrafiltration of heavy metals from aqueous solution. Chem. Eng. J. 2014, 244, 446–456. [Google Scholar] [CrossRef]
  79. Ciobanu, M.; Marin, L.; Cozan, V.; Bruma, M. Aromatic Polysulfones Used in Sensor Applications. Rev. Adv. Amter. Sci. 2009, 22, 89–96. [Google Scholar]
  80. Zou, D.; Lee, Y.M. Design strategy of poly(vinylidene fluoride) membranes for water treatment. Prog. Polym. Sci. 2022, 128, 101535. [Google Scholar] [CrossRef]
  81. Vajhadin, F.; Mazloum-Ardakani, M.; Raeisi, S.; Hemati, M.; Ebadi, A.; Haghiralsadat, F.; Tofighi, D. Glutaraldehyde crosslinked doxorubicin promotes drug delivery efficiency using cobalt ferrite nanoparticles. Colloids Surf. B Biointerfaces 2022, 220, 112870. [Google Scholar] [CrossRef] [PubMed]
  82. Privar, Y.; Skatova, A.; Golikov, A.; Boroda, A.; Bratskaya, S. Gelation and Cryogelation of Chitosan: Origin of Low Efficiency of Diglycidyl Ethers as Cross-Linkers in Acetic Acid Solutions. Polysaccharides 2024, 5, 731–742. [Google Scholar] [CrossRef]
  83. Nordin, N.; Azman, Z.A.Z.; Adnan, N.A.; Majid, S.R. On the dual crosslinking for functionality enhancement of poly (acrylamide-co-acrylic acid)/chitosan-aluminum (III) ions and its characterization and sensory hydrogel fibers. Int. J. Biol. Macromol. 2024, 274, 133383. [Google Scholar] [CrossRef]
  84. Shi, C.; Li, X.; Zhang, X.; Zou, M. Dual dynamic network structures of recyclable epoxy resins with high strength and toughness via sacrificial hydrogen-bonding clusters and imine bonds: Surpassing the strength-toughness trade-off. Chem. Eng. J. 2024, 493, 152361. [Google Scholar] [CrossRef]
  85. Zhou, Y.; Zeng, G.; Zhang, F.; Li, K.; Li, X.; Luo, J.; Li, J.; Li, J. Design of tough, strong and recyclable plant protein-based adhesive via dynamic covalent crosslinking chemistry. Chem. Eng. J. 2023, 460, 141774. [Google Scholar] [CrossRef]
  86. Pawar, D.R.L.; Jeyapalina, S.; Hafer, K.; Bachus, K.N. Influence of negative pressure wound therapy on peri-prosthetic tissue vascularization and inflammation around porous titanium percutaneous devices. J. Biomed. Mater. Res. B Appl. Biomater. 2019, 107, 2091–2101. [Google Scholar] [CrossRef]
  87. Li, S.; Yang, L.; Zhao, Z.; Yang, X.; Lv, H. A polyurethane-based hydrophilic elastomer with multi-biological functions for small-diameter vascular grafts. Acta Biomater. 2024, 176, 234–249. [Google Scholar] [CrossRef]
  88. Kong, V.A.; Staunton, T.A.; Laaser, J.E. Effect of Cross-Link Homogeneity on the High-Strain Behavior of Elastic Polymer Networks. Macromolecules 2024, 57, 4670–4679. [Google Scholar] [CrossRef]
  89. Bose, S.; Sarkar, N.; Jo, Y. Natural medicine delivery from 3D printed bone substitutes. J. Control. Release 2024, 365, 848–875. [Google Scholar] [CrossRef]
  90. Feijão, T.; Neves, M.I.; Sousa, A.; Torres, A.L.; Bidarra, S.J.; Orge, I.D.; Carvalho, D.T.O.; Barrias, C.C. Engineering injectable vascularized tissues from the bottom-up: Dynamics of in-gel extra-spheroid dermal tissue assembly. Biomaterials 2021, 279, 121222. [Google Scholar] [CrossRef] [PubMed]
  91. Acik, G.; Cakir, N.T.; Altinkok, C. Development of organosoluble, quaternized and naproxen sodium- loaded poly(vinyl alcohol)-based electrospun nanofibers. Eur. Polym. J. 2024, 221, 113565. [Google Scholar] [CrossRef]
  92. Issue Information. J. Biomed. Mater. Res. B Appl. Biomater. 2020, 108, 591–599. [CrossRef]
  93. Daemi, H.; Barikani, M. Synthesis and characterization of calcium alginate nanoparticles, sodium homopolymannuronate salt and its calcium nanoparticles. Sci. Iran. 2012, 19, 2023–2028. [Google Scholar] [CrossRef]
  94. Hu, J.; Huang, G.; Li, L.; Zhan, X.; Zhang, J.; Shao, J.; Hong, S.; Pan, S.-T. A Mg2+-light double crosslinked injectable alginate hydrogel promotes vascularized osteogenesis by establishing a Mg2+-enriched microenvironment. Mater. Today Commun. 2024, 41, 110303. [Google Scholar] [CrossRef]
  95. Cui, S.; Yang, F.; Yu, D.; Shi, C.; Zhao, D.; Chen, L.; Chen, J. Double Network Physical Crosslinked Hydrogel for Healing Skin Wounds: New Formulation Based on Polysaccharides and Zn2+. Int. J. Mol. Sci. 2023, 24, 13042. [Google Scholar] [CrossRef]
  96. Piluso, S.; Hiebl, B.; Gorb, S.N.; Kovalev, A.; Lendlein, A.; Neffe, A.T. Hyaluronic acid-based hydrogels crosslinked by copper-catalyzed azide-alkyne cycloaddition with tailorable mechanical properties. Int. J. Artif. Organs 2011, 34, 192–197. [Google Scholar] [CrossRef]
  97. Xu, X.; Li, L.; Seraji, S.M.; Liu, L.; Jiang, Z.; Xu, Z.; Li, X.; Zhao, S.; Wang, H.; Song, P. Bioinspired, Strong, and Tough Nanostructured Poly(vinyl alcohol)/Inositol Composites: How Hydrogen-Bond Cross-Linking Works? Macromolecules 2021, 54, 9510–9521. [Google Scholar] [CrossRef]
  98. Hwang, U.; Moon, H.; Park, J.; Jung, H.W. Crosslinking and Swelling Properties of pH-Responsive Poly(Ethylene Glycol)/Poly(Acrylic Acid) Interpenetrating Polymer Network Hydrogels. Polymers 2024, 16, 2149. [Google Scholar] [CrossRef]
  99. Cao, Q.; Chen, J.; Wang, M.; Wang, Z.; Wang, W.; Shen, Y.; Xue, Y.; Li, B.; Ma, Y.; Yao, Y.; et al. Superabsorbent carboxymethyl cellulose–based hydrogel fabricated by liquid-metal-induced double crosslinking polymerisation. Carbohydr. Polym. 2024, 331, 121910. [Google Scholar] [CrossRef]
  100. Das, D.; Awan, A.; Kaur, K.; Müller, M.; Schönherr, H.S. Design, Characterization, and Biocompatibility of Modular Biopolymer-Based Single- and Double-Cross-Linked Networks Hydrogels. ACS Appl. Polym. Mater. 2024, 6, 13158–13170. [Google Scholar] [CrossRef]
  101. Peesan, M.; Supaphol, P.; Rujiravanit, R. Preparation and characterization of hexanoyl chitosan/polylactide blend films. Carbohydr. Polym. 2005, 60, 343–350. [Google Scholar] [CrossRef]
  102. Chmayssem, A.; Shalayel, I.; Marinesco, S.; Zebda, A. Investigation of GOx Stability in a Chitosan Matrix: Applications for Enzymatic Electrodes. Sensors 2023, 23, 465. [Google Scholar] [CrossRef]
  103. Keegan, M.E.; Royce, S.M.; Fahmy, T.; Saltzman, W.M. In vitro evaluation of biodegradable microspheres with surface-bound ligands. J. Control. Release 2006, 110, 574–580. [Google Scholar] [CrossRef] [PubMed]
  104. Yu, Y.; Xu, S.; Li, S.; Pan, H.A. Genipin-cross-linked hydrogels based on biomaterials for drug delivery: A review. Biomater. Sci. 2021, 9, 1583–1597. [Google Scholar] [CrossRef]
  105. Shalayel, I.; Vallee, Y.; Zebda, A. Study of Genipin Behavior in Neutral and Acidic Aqueous Solutions at Elevated Temperatures. ChemistrySelect 2023, 8, e202204705. [Google Scholar] [CrossRef]
  106. Neel, E.A.A.; Chrzanowski, W.; Knowles, J.C. Effect of increasing titanium dioxide content on bulk and surface properties of phosphate-based glasses. Acta Biomater. 2008, 4, 523–534. [Google Scholar] [CrossRef]
  107. Li, S.; Wu, P.; Ji, Z.; Zhang, Y.; Zhang, P.; He, Y.; Shen, Y. In vitro biocompatibility study of EDC/NHS cross-linked silk fibroin scaffold with olfactory ensheathing cells. J. Biomater. Sci. Polym. Ed. 2023, 34, 482–496. [Google Scholar] [CrossRef]
  108. Lee, C.H.; Shin, H.J.; Cho, I.H.; Kang, Y.M.; Kim, I.A.; Park, K.D.; Shin, J.W. Nanofiber alignment and direction of mechanical strain affect the ECM production of human ACL fibroblast. Biomaterials 2005, 26, 1261–1270. [Google Scholar] [CrossRef]
  109. Böhm, P.; Dornbusch, M.; Gutmann, J.S. Phytic acid oligomers as bio-based crosslinkers for epoxy and polyol resins. J. Coat. Technol. Res. 2024, 21, 355–367. [Google Scholar] [CrossRef]
  110. Meyers, M.A.; Chen, P.Y.; Lopez, M.I.; Seki, Y.; Lin, A.Y.M. Biological materials: A materials science approach. J. Mech. Behav. Biomed. Mater. 2011, 4, 626–657. [Google Scholar] [CrossRef]
  111. Jeong, C.H.; Kim, D.H.; Yune, J.H.; Kwon, H.C.; Shin, D.M.; Sohn, H.; Lee, K.H.; Choi, B.; Kim, E.S.; Kang, J.H.; et al. In vitro toxicity assessment of crosslinking agents used in hyaluronic acid dermal filler. Toxicol. Vitr. 2021, 70, 105034. [Google Scholar] [CrossRef]
  112. Ahmed, M.M.; O’Doherty, G.A. De novo asymmetric syntheses of C-4-substituted sugars via an iterative dihydroxylation strategy. Carbohydr. Res. 2006, 341, 1505–1521. [Google Scholar] [CrossRef] [PubMed]
  113. Yapo, B.M.; Lerouge, P.; Thibault, J.F.; Ralet, M.C. Pectins from citrus peel cell walls contain homogalacturonans homogenous with respect to molar mass, rhamnogalacturonan I and rhamnogalacturonan II. Carbohydr. Polym. 2007, 69, 426–435. [Google Scholar] [CrossRef]
  114. Cheung, D.T.; Nimni, M.E. Mechanism of crosslinking of proteins by glutaraldehyde I: Reaction with model compounds. Connect. Tissue Res. 1982, 10, 187–199. [Google Scholar] [CrossRef] [PubMed]
  115. Wine, Y.; Cohen-Hadar, N.; Freeman, A.; Frolow, F. Elucidation of the mechanism and end products of glutaraldehyde crosslinking reaction by X-ray structure analysis. Biotechnol. Bioeng. 2007, 98, 711–718. [Google Scholar] [CrossRef] [PubMed]
  116. Cheung, D.T.; Perelman, N.; Ko, E.C.; Nimni, M.E. Mechanism of crosslinking of proteins by glutaraldehyde III. Reaction with collagen in tissues. Connect. Tissue Res. 1985, 13, 109–115. [Google Scholar] [CrossRef]
  117. Tacias-Pascacio, V.G.; García-Parra, E.; Vela-Gutiérrez, G.; Virgen-Ortiz, J.J.; Berenguer-Murcia, Á.; Alcántara, A.R.; Fernandez-Lafuente, R. Genipin as An Emergent Tool in the Design of Biocatalysts: Mechanism of Reaction and Applications. Catalysts 2019, 9, 1035. [Google Scholar] [CrossRef]
  118. Mi, F.L.; Sung, H.W.; Shyu, S.S.; Su, C.C.; Peng, C.K. Synthesis and characterization of biodegradable TPP/genipin co-crosslinked chitosan gel beads. Polymer 2003, 44, 6521–6530. [Google Scholar] [CrossRef]
  119. Butler, M.F.; Ng, Y.F.; Pudney, P.D.A. Mechanism and kinetics of the crosslinking reaction between biopolymers containing primary amine groups and genipin. J. Polym. Sci. A Polym. Chem. 2003, 41, 3941–3953. [Google Scholar] [CrossRef]
  120. Wang, Z.; Liu, H.; Luo, W.; Cai, T.; Li, Z.; Liu, Y.; Gao, W.; Wan, Q.; Wang, X.; Wang, J.; et al. Regeneration of skeletal system with genipin crosslinked biomaterials. J. Tissue Eng. 2020, 29, 2041731420974861. [Google Scholar] [CrossRef] [PubMed]
  121. Kondratowicz, I.; Shalayel, I.; Nadolska, M.; Tsujimura, S.; Yamagata, Y.; Itanda, I.; Zebda, A. Correction to: Impact of Lactic Acid and Genipin Concentration on Physicochemical and Mechanical Properties of Chitosan Membranes. J. Polym. Environ. 2023, 31, 2242. [Google Scholar] [CrossRef]
  122. Kim, E.; Kim, M.H.; Song, J.H.; Kang, C.; Park, W.H. Dual crosslinked alginate hydrogels by riboflavin as photoinitiator. Int. J. Biol. Macromol. 2020, 54, 989–998. [Google Scholar] [CrossRef]
  123. Soares, R.M.D.; Maia, G.S.; Rayas-Duarte, P.; Soldi, V. Properties of filmogenic solutions of gliadin crosslinked with 1-(3-dimethyl aminopropyl)-3-ethylcarbodiimidehydrochloride/N-hydroxysuccinimide and cysteine. Food Hydrocoll. 2009, 23, 181–187. [Google Scholar] [CrossRef]
  124. Nair, M.; Johal, R.K.; Hamaia, S.W.; Best, S.M.; Cameron, R.E. Tunable bioactivity and mechanics of collagen-based tissue engineering constructs: A comparison of EDC-NHS, genipin and TG2 crosslinkers. Biomaterials 2020, 254, 120109. [Google Scholar] [CrossRef] [PubMed]
  125. Söderberg, L.; Dyhre, H.; Roth, B.; Björkman, S. The ‘inverted cup’—A novel in vitro release technique for drugs in lipid formulations. J. Control. Release 2006, 113, 80–88. [Google Scholar] [CrossRef]
  126. Hautmann, A.; Kedilaya, D.; Stojanović, S.; Radenković, M.; Marx, C.K.; Najman, S.; Pietzsch, M.; Mano, J.F.; Groth, T. Free-standing multilayer films as growth factor reservoirs for future wound dressing applications. Biomater. Adv. 2022, 142, 213166. [Google Scholar] [CrossRef]
  127. Prolongo, S.G.; Del Rosario, G.; Ureña, A. Comparative study on the adhesive properties of different epoxy resins. Int. J. Adhes. Adhes. 2006, 26, 125–132. [Google Scholar] [CrossRef]
  128. Krauklis, A.E.; Echtermeyer, A.T. Mechanism of yellowing: Carbonyl formation during hygrothermal aging in a common amine epoxy. Polymers 2018, 10, 1017. [Google Scholar] [CrossRef]
  129. Apostolidis, P.; Liu, X.; Erkens, S.; Scarpas, A. Characterization of epoxy-asphalt binders by differential scanning calorimetry. Constr. Build. Mater. 2020, 249, 118800. [Google Scholar] [CrossRef]
  130. Zhang, Y.; Yuan, J.; Hu, J.; Tian, Z.; Feng, W.; Yan, H. Toward understanding the cross-linking from molecular chains to aggregates by engineering terminals of supramolecular hyperbranched polysiloxane. Aggregate 2024, 5, e404. [Google Scholar] [CrossRef]
  131. Peng, R.; Li, Y.; Su, H.; Lu, Y.; Shia, L.; Yun, Y.; Liao, B. The modification of sintering and microwave dielectric properties of Mn2+ doped LiZnPO4 ceramic. J. Mater. Res. Technol. 2020, 9, 4994–5006. [Google Scholar] [CrossRef]
  132. Yan, B.; Hu, X.; Zhao, Y.; Wu, M.; Feng, Y.; He, Z.; Qi, G.; Ren, W.; Liang, Y.; Wang, W.; et al. Research and development of a sodium alginate/calcium ion gel based on in situ cross-linked double-network for controlling spontaneous combustion of coal. Fuel 2022, 322, 124260. [Google Scholar] [CrossRef]
  133. Samorezov, J.E.; Morlock, C.M.; Alsberg, E. Dual Ionic and Photo-Crosslinked Alginate Hydrogels for Micropatterned Spatial Control of Material Properties and Cell Behavior. Bioconjug. Chem. 2015, 26, 1339–1347. [Google Scholar] [CrossRef]
  134. Sardelli, L.; Tunesi, M.; Briatico-Vangosa, F.; Petrini, P. 3D-Reactive printing of engineered alginate inks. Soft Matter 2021, 17, 8105–8117. [Google Scholar] [CrossRef] [PubMed]
  135. Pilipenko, N.; Gonçalves, O.H.; Bona, E.; Fernandes, I.P.; Pinto, J.A.; Sorita, G.D.; Leimann, F.V.; Barreiro, M.F. Tailoring swelling of alginate-gelatin hydrogel microspheres by crosslinking with calcium chloride combined with transglutaminase. Carbohydr. Polym. 2019, 223, 115035. [Google Scholar] [CrossRef]
  136. Kandimalla, V.B.; Tripathi, V.S.; Ju, H. A conductive ormosil encapsulated with ferrocene conjugate and multiwall carbon nanotubes for biosensing application. Biomaterials 2006, 27, 1167–1174. [Google Scholar] [CrossRef]
  137. Roy, S.; Rhim, J.-W. Genipin-Crosslinked Gelatin/Chitosan-Based Functional Films Incorporated with Rosemary Essential Oil and Quercetin. Materials 2022, 15, 3769. [Google Scholar] [CrossRef]
  138. Lu, H.; Butler, J.A.; Britten, N.S.; Venkatraman, P.D.; Rahatekar, S.S. Natural antimicrobial nano composite fibres manufactured from a combination of alginate and oregano essential oil. Nanomaterials 2021, 11, 2062. [Google Scholar] [CrossRef]
  139. Agarwal, T.; Madaan, S.; Mohammed, S.; Patil, N. Cross-links techniques in the manufacturing of smart biomaterials for implanting bone tissue: A systematic review. Multidiscip. Rev. 2024, 6, 2023ss067. [Google Scholar] [CrossRef]
  140. Bigi, A.; Cojazzi, G.; Panzavolta, S.; Rubini, K.; Roveri, N. Mechanical and thermal properties of gelatin films at different degrees of glutaraldehyde crosslinking. Biomaterials 2001, 22, 763–768. [Google Scholar] [CrossRef]
  141. Barbosa, O.; Ortiz, C.; Berenguer-Murcia, Á.; Torres, R.; Rodrigues, R.C.; Fernandez-Lafuente, R. Glutaraldehyde in bio-catalysts design: A useful crosslinker and a versatile tool in enzyme immobilization. RSC Adv. 2014, 4, 1583–1600. [Google Scholar] [CrossRef]
  142. Hoffmann, B.; Seitz, D.; Mencke, A.; Kokott, A.; Ziegler, G. Glutaraldehyde and oxidised dextran as crosslinker reagents for chitosan-based scaffolds for cartilage tissue engineering. J. Mater. Sci. Mater. Med. 2009, 20, 1495–1503. [Google Scholar] [CrossRef] [PubMed]
  143. Liu, S.; Lau, C.-S.; Liang, K.; Wen, F.; Teoh, S.H. Marine collagen scaffolds in tissue engineering. Curr. Opin. Biotechnol. 2022, 74, 92–103. [Google Scholar] [CrossRef] [PubMed]
  144. Casali, D.M.; Yost, M.J.; Matthews, M.A. Eliminating glutaraldehyde from crosslinked collagen films using supercritical CO2. J. Biomed. Mater. Res. A 2018, 106, 86–94. [Google Scholar] [CrossRef]
  145. Gulbins, H.; Goldemund, A.; Anderson, I.; Haas, U.; Uhlig, A.; Meiser, B.; Reichart, B. Preseeding with autologous fibroblasts improves endothelialization of glutaraldehyde-fixed porcine aortic valves. J. Thorac. Cardiovasc. Surg. 2003, 125, 592–601. [Google Scholar] [CrossRef]
  146. Reddy, N.; Reddy, R.; Jiang, Q. Crosslinking biopolymers for biomedical applications. Trends Biotechnol. 2015, 33, 362–369. [Google Scholar] [CrossRef]
  147. Singhal, P.; Luk, A.; Butany, J. Bioprosthetic Heart Valves: Impact of Implantation on Biomaterials. ISRN Biomater. 2013, 2013, 728791. [Google Scholar] [CrossRef]
  148. Chou, S.-F.; Luo, L.-J.; Lai, J.-Y.; Ma, D.H.-K. Role of solvent-mediated carbodiimide cross-linking in fabrication of electrospun gelatin nanofibrous membranes as ophthalmic biomaterials. Mater. Sci. Eng. C 2017, 71, 1145–1155. [Google Scholar] [CrossRef]
  149. Bax, D.V.; Davidenko, N.; Gullberg, D.; Hamaia, S.W.; Farndale, R.W.; Best, S.M.; Cameron, R.E. Fundamental insight into the effect of carbodiimide crosslinking on cellular recognition of collagen-based scaffolds. Acta Biomater. 2017, 49, 218–234. [Google Scholar] [CrossRef]
  150. Zhong, W.; Hu, Y.; Li, J.; Bian, K.; Su, P.; Zhang, Y.; Zhang, P.; Zhang, H.; Zhang, F.; Shen, Y. In vitro biocompatibility study of a water-rinsed biomimetic silk porous scaffold with olfactory ensheathing cells. Int. J. Biol. Marcromol. 2019, 125, 526–533. [Google Scholar] [CrossRef]
  151. Grabska-Zielińska, S.; Sionkowska, A.; Coelho, C.C.; Monteiro, F.J. Silk Fibroin/Collagen/Chitosan Scaffolds Cross-Linked by a Glyoxal Solution as Biomaterials toward Bone Tissue Regeneration. Materials 2020, 13, 3433. [Google Scholar] [CrossRef] [PubMed]
  152. Davidenko, N.; Schuster, C.F.; Bax, D.V.; Raynal, N.; Farndale, R.W.; Best, S.M.; Cameron, R.E. Control of crosslinking for tailoring collagen-based scaffolds stability and mechanics. Acta Biomater. 2015, 25, 131–142. [Google Scholar] [CrossRef]
  153. Alibabaei-Omran, F.; Zabihi, E.; Seifalian, A.M.; Javanmehr, N.; Samadikuchaksaraei, A.; Gholipourmalekabadi, M.; Asghari, M.H.; Nouri, H.R.; Pourbagher, R.; Bouzari, Z.; et al. Bilateral Crosslinking with Glutaraldehyde and 1-Ethyl-3-(3-Dimethylaminopropyl) Carbodiimide: An Optimization Strategy for the Application of Decellularized Human Amniotic Membrane in Tissue Engineering. J. Tissue Eng. Regen. Med. 2024, 2024, 8525930. [Google Scholar] [CrossRef]
  154. Islam, M.M.; AbuSamra, D.B.; Chivu, A.; Argüeso, P.; Dohlman, C.H.; Patra, H.K.; Chodosh, J.; González-Andrades, M. Optimization of Collagen Chemical Crosslinking to Restore Biocompatibility of Tissue-Engineered Scaffolds. Pharmaceutics 2021, 13, 832. [Google Scholar] [CrossRef]
  155. Dias, J.R.; Baptista-Silva, S.; de Oliveira, C.M.T.; Sousa, A.; Oliveira, A.L.; Bártolo, P.J.; Granja, P.L. In situ crosslinked electrospun gelatin nanofibers for skin regeneration. Eur. Polym. J. 2017, 95, 161–173. [Google Scholar] [CrossRef]
  156. Koh, L.B.; Islam, M.M.; Mitra, D.; Noel, C.W.; Merrett, K.; Odorcic, S.; Fagerholm, P.; Jackson, W.B.; Liedberg, B.; Phopase, J.; et al. Epoxy Cross-Linked Collagen and Collagen-Laminin Peptide Hydrogels as Corneal Substitutes. J. Funct. Biomater. 2013, 4, 162–177. [Google Scholar] [CrossRef]
  157. Zheng, X.; Chen, Y.; Dan, N.; Dan, W.; Li, Z. Highly stable collagen scaffolds crosslinked with an epoxidized natural polysaccharide for wound healing. Int. J. Biol. Macromol. 2021, 182, 1994–2002. [Google Scholar] [CrossRef]
  158. Thakur, G.; Naqvi, M.A.; Rousseau, D.; Pal, K.; Mitra, A.; Basak, A. Gelatin-Based Emulsion Gels for Diffusion-Controlled Release Applications. J. Biomater. Sci. Polym. Ed. 2012, 23, 645–661. [Google Scholar] [CrossRef]
  159. Sung, H.-W.; Huang, R.-N.; Huang, L.L.H.; Tsai, C.-C. In vitro evaluation of cytotoxicity of a naturally occurring cross-linking reagent for biological tissue fixation. J. Biomater. Sci. Polym. Ed. 1999, 10, 63–78. [Google Scholar] [CrossRef]
  160. Wang, Y.; Bao, J.; Wu, X.; Wu, Q.; Li, Y.; Zhou, Y.; Li, L.; Bu, H. Genipin crosslinking reduced the immunogenicity of xenogeneic decellularized porcine whole-liver matrices through regulation of immune cell proliferation and polarization. Sci. Rep. 2016, 6, 24779. [Google Scholar] [CrossRef] [PubMed]
  161. Lien, S.-M.; Li, W.-T.; Huang, T.-J. Genipin-crosslinked gelatin scaffolds for articular cartilage tissue engineering with a novel crosslinking method. Mater. Sci. Eng. C 2008, 28, 36–43. [Google Scholar] [CrossRef]
  162. Pinheiro, A.; Cooley, A.; Liao, J.; Prabhu, R.; Elder, S. Comparison of natural crosslinking agents for the stabilization of xenogenic articular cartilage. J. Orthop. Res. 2016, 34, 1037–1046. [Google Scholar] [CrossRef] [PubMed]
  163. Liu, Z.; Zhou, Q.; Zhu, J.; Xiao, J.; Wan, P.; Zhou, C.; Huang, Z.; Qiang, N.; Zhang, W.; Wu, Z.; et al. Using genipin-crosslinked acellular porcine corneal stroma for cosmetic corneal lens implants. Biomaterials 2012, 33, 7336–7346. [Google Scholar] [CrossRef]
  164. Frauchiger, D.; May, R.; Bakirci, E.; Tekari, A.; Chan, S.; Wöltje, M.; Benneker, L.; Gantenbein, B. Genipin-Enhanced Fibrin Hydrogel and Novel Silk for Intervertebral Disc Repair in a Loaded Bovine Organ Culture Model. J. Funct. Biomater. 2018, 9, 40. [Google Scholar] [CrossRef]
  165. Yu, L.; Liu, Y.; Wu, J.; Wang, S.; Yu, J.; Wang, W.; Ye, X. Genipin Cross-Linked Decellularized Nucleus Pulposus Hydrogel-Like Cell Delivery System Induces Differentiation of ADSCs and Retards Intervertebral Disc Degeneration. Front. Bioeng. Biotechnol. 2021, 9, 807883. [Google Scholar] [CrossRef]
  166. Fessel, G.; Gerber, C.; Snedeker, J.G. Potential of collagen cross-linking therapies to mediate tendon mechanical properties. J. Shoulder Elbow Surg. 2012, 21, 209–217. [Google Scholar] [CrossRef]
  167. Fessel, G.; Cadby, J.; Wunderli, S.; van Weeren, R.; Snedeker, J.G. Dose- and time-dependent effects of genipin crosslinking on cell viability and tissue mechanics—Toward clinical application for tendon repair. Acta Biomater. 2014, 10, 1897–1906. [Google Scholar] [CrossRef]
  168. Hrabchak, C.; Rouleau, J.; Moss, I.; Woodhouse, K.; Akens, M.; Bellingham, C.; Keeley, F.; Dennis, M.; Yee, A. Assessment of biocompatibility and initial evaluation of genipin cross-linked elastin-like polypeptides in the treatment of an osteochondral knee defect in rabbits. Acta Biomater. 2010, 6, 2108–2115. [Google Scholar] [CrossRef]
  169. Vo, N.T.N.; Huang, L.; Lemos, H.; Mellor, A.L.; Novakovic, K. Genipin-crosslinked chitosan hydrogels: Preliminary evaluation of the in vitro biocompatibility and biodegradation. J. Appl. Polym. Sci. 2021, 138, 50848. [Google Scholar] [CrossRef]
  170. Berger, J.; Reist, M.; Mayer, J.M.; Felt, O.; Peppas, N.A.; Gurny, R. Structure and interactions in covalently and ionically crosslinked chitosan hydrogels for biomedical applications. Eur. J. Pharm. Biopharm. 2004, 57, 19–34. [Google Scholar] [CrossRef] [PubMed]
  171. Sazhnev, N.A.; Drozdova, M.G.; Rodionov, I.A.; Kil’deeva, N.R.; Balabanova, T.V.; Markvicheva, E.A.; Lozinsky, V.I. Preparation of Chitosan Cryostructurates with Controlled Porous Morphology and Their Use as 3D-Scaffolds for the Cultivation of Animal Cells. Appl. Biochem. Microbiol. 2018, 54, 459–467. [Google Scholar] [CrossRef]
  172. Cassimjee, H.; Kumar, P.; Ubanako, P.; Choonara, Y.E. Genipin-Crosslinked, Proteosaccharide Scaffolds for Potential Neural Tissue Engineering Applications. Pharmaceutics 2022, 14, 441. [Google Scholar] [CrossRef]
  173. Ma, Y.; Shi, H.; Wei, Q.; Deng, Q.; Sun, J.; Liu, Z.; Lai, B.; Li, G.; Ding, Y.; Niu, W.; et al. Developing a mechanically matched decellularized spinal cord scaffold for the in situ matrix-based neural repair of spinal cord injury. Biomaterials 2021, 279, 121192. [Google Scholar] [CrossRef]
  174. Zheng, L.; Liu, S.; Cheng, X.; Qin, Z.; Lu, Z.; Zhang, K.; Zhao, J. Intensified Stiffness and Photodynamic Provocation in a Collagen-Based Composite Hydrogel Drive Chondrogenesis. Adv. Sci. 2019, 6, 1900099. [Google Scholar] [CrossRef] [PubMed]
  175. Zhou, H.; Wang, Z.; Cao, H.; Hu, H.; Luo, Z.; Yang, X.; Cui, M.; Zhou, L. Genipin-crosslinked polyvinyl alcohol/silk fibroin/nano-hydroxyapatite hydrogel for fabrication of artificial cornea scaffolds—A novel approach to corneal tissue engineering. J. Biomater. Sci. Polym. Ed. 2019, 30, 1604–1619. [Google Scholar] [CrossRef]
  176. Wu, X.; Wang, Y.; Wu, Q.; Li, Y.; Li, L.; Tang, J.; Shi, Y.; Bu, H.; Bao, J.; Xie, M. Genipin-crosslinked, immunogen-reduced decellularized porcine liver scaffold for bioengineered hepatic tissue. Tissue Eng. Regen. Med. 2015, 12, 417–426. [Google Scholar] [CrossRef]
  177. Erdagi, S.I.; Ngwabebhoh, F.A.; Yildiz, U. Genipin crosslinked gelatin-diosgenin-nanocellulose hydrogels for potential wound dressing and healing applications. Int. J. Biol. Macromol. 2020, 149, 651–663. [Google Scholar] [CrossRef]
  178. Zafeiris, K.; Brasinika, D.; Karatza, A.; Koumoulos, E.; Karoussis, I.K.; Kyriakidou, K.; Charitidis, C.A. Additive manufacturing of hydroxyapatite–chitosan–genipin composite scaffolds for bone tissue engineering applications. Mater. Sci. Eng. C 2021, 119, 111639. [Google Scholar] [CrossRef]
  179. Wang, C.; Lau, T.T.; Loh, W.L.; Su, K.; Wang, D. Cytocompatibility study of a natural biomaterial crosslinker—Genipin with therapeutic model cells. J. Biomed. Mater. Res. B Appl. Biomater. 2011, 97, 58–65. [Google Scholar] [CrossRef]
  180. Raucci, M.G.; Alvarez-Perez, M.A.; Demitri, C.; Giugliano, D.; De Benedictis, V.; Sannino, A.; Ambrosio, L. Effect of citric acid crosslinking cellulose-based hydrogels on osteogenic differentiation. J. Biomed. Mater. Res. A 2015, 103, 2045–2056. [Google Scholar] [CrossRef] [PubMed]
  181. Bonetti, L.; De Nardo, L.; Farè, S. Chemically Crosslinked Methylcellulose Substrates for Cell Sheet Engineering. Gels 2021, 7, 141. [Google Scholar] [CrossRef] [PubMed]
  182. Natarajan, V.; Krithica, N.; Madhan, B.; Sehgal, P.K. Preparation and properties of tannic acid cross-linked collagen scaffold and its application in wound healing. J. Biomed. Mater. Res. B Appl. Biomater. 2013, 101, 560–567. [Google Scholar] [CrossRef]
  183. Wang, X.; Ma, B.; Chang, J. Preparation of decellularized vascular matrix by co-crosslinking of procyanidins and glutaraldehyde. Biomed. Mater. Eng. 2015, 26, 19–30. [Google Scholar] [CrossRef]
  184. Kwon, Y.S.; Lee, S.H.; Hwang, Y.C.; Rosa, V.; Lee, K.W.; Min, K.S. Behaviour of human dental pulp cells cultured in a collagen hydrogel scaffold cross-linked with cinnamaldehyde. Int. Endod. J. 2017, 50, 58–66. [Google Scholar] [CrossRef]
  185. Kwon, Y.-S.; Kim, H.-J.; Hwang, Y.-C.; Rosa, V.; Yu, M.-K.; Min, K.-S. Effects of Epigallocatechin Gallate, an Antibacterial Cross-linking Agent, on Proliferation and Differentiation of Human Dental Pulp Cells Cultured in Collagen Scaffolds. J. Endod. 2017, 43, 289–296. [Google Scholar] [CrossRef]
  186. Huang, A.; Honda, Y.; Li, P.; Tanaka, T.; Baba, S. Integration of Epigallocatechin Gallate in Gelatin Sponges Attenuates Matrix Metalloproteinase-Dependent Degradation and Increases Bone Formation. Int. J. Mol. Sci. 2019, 20, 6042. [Google Scholar] [CrossRef]
  187. Deng, A.; Yang, Y.; Du, S.; Yang, S. Electrospinning of in situ crosslinked recombinant human collagen peptide/chitosan nanofibers for wound healing. Biomater. Sci. 2018, 6, 2197–2208. [Google Scholar] [CrossRef] [PubMed]
  188. Zhang, M.; Choi, W.; Kim, M.; Choi, J.; Zang, X.; Ren, Y.; Chen, H.; Tsukruk, V.; Peng, J.; Liu, Y.; et al. Recent Advances in Environmentally Friendly Dual-crosslinking Polymer Networks; John Wiley and Sons Inc.: Hoboken, NJ, USA, 2024. [Google Scholar] [CrossRef]
  189. Alsaafeen, N.B.; Bawazir, S.S.; Jena, K.K.; Seitak, A.; Fatma, B.; Pitsalidis, C.; Khandoker, A.; Pappa, A.M. One-Pot Synthesis of a Robust Crosslinker-Free Thermo-Reversible Conducting Hydrogel Electrode for Epidermal Electronics. ACS Appl. Mater. Interfaces 2024, 16, 61435–61445. [Google Scholar] [CrossRef]
  190. Alavarse, A.C.; Frachini, E.C.G.; da Silva, R.L.C.G.; Lima, V.H.; Shavandi, A.; Petri, D.F.S. Crosslinkers for polysaccharides and proteins: Synthesis conditions, mechanisms, and crosslinking efficiency, a review. Int. J. Biol. Macromol. 2022, 202, 558–596. [Google Scholar] [CrossRef]
  191. Zhang, Y.; Huang, Y.; Carr, P.W. Optimization of the synthesis of a hyper-crosslinked stationary phases: A new generation of highly efficient, acid-stable hyper-crosslinked materials for HPLC. J. Sep. Sci. 2011, 34, 1407–1422. [Google Scholar] [CrossRef]
  192. Siddoway, A.C.; White, B.M.; Narasimhan, B.; Mallapragada, S.K. Synthesis and Optimization of Next-Generation Low-Molecular-Weight Pentablock Copolymer Nanoadjuvants. Vaccines 2023, 11, 1572. [Google Scholar] [CrossRef] [PubMed]
  193. Pugliese, R.; Maleki, M.; Zuckermann, R.N.; Gelain, F. Self-assembling peptides cross-linked with genipin: Resilient hydrogels and self-standing electrospun scaffolds for tissue engineering applications. Biomater. Sci. 2019, 7, 76–91. [Google Scholar] [CrossRef]
  194. Jiang, F.; Huang, T.; He, C.; Brown, H.R.; Wang, H. Interactions Affecting the Mechanical Properties of Macromolecular Microsphere Composite Hydrogels. J. Phys. Chem. B 2013, 114, 13679–13687. [Google Scholar] [CrossRef]
  195. Holloway, J.L.; Ma, H.; Rai, R.; Burdick, J.A. Modulating hydrogel crosslink density and degradation to control bone morphogenetic protein delivery and in vivo bone formation. J. Control. Release 2014, 191, 63–70. [Google Scholar] [CrossRef] [PubMed]
  196. Tanimoto, R.; Ebara, M.; Uto, K. Tunable enzymatically degradable hydrogels for controlled cargo release with dynamic mechanical properties. Soft Matter 2023, 19, 6224–6233. [Google Scholar] [CrossRef]
  197. Echalier, C.; Valot, L.; Martinez, J.; Mehdi, A.; Subra, G. Chemical cross-linking methods for cell encapsulation in hydrogels. Mater. Today Commun. 2019, 20, 100536. [Google Scholar] [CrossRef]
  198. Carberry, B.J.; Hernandez, J.J.; Dobson, A.; Bowman, C.N.; Anseth, K.S. Kinetic Analysis of Degradation in Thioester Cross-linked Hydrogels as a Function of Thiol Concentration, pKa, and Presentation. Macromolecules 2022, 55, 2123–2129. [Google Scholar] [CrossRef]
  199. Maleki, A.; Kjøniksen, A.L.; Nyström, B. Characterization of the chemical degradation of hyaluronic acid during chemical gelation in the presence of different cross-linker agents. Carbohydr. Res. 2007, 342, 2776–2792. [Google Scholar] [CrossRef]
  200. Zafar, S.; Hanif, M.; Azeem, M.; Mahmood, K.; Ashfaq, G.S. Role of crosslinkers for synthesizing biocompatible, biodegradable and mechanically strong hydrogels with desired release profile. Polym. Bull. 2022, 79, 9199–9219. [Google Scholar] [CrossRef]
  201. Song, M.; Wang, J.; He, J.; Kan, D.; Chen, K.; Lu, J. Synthesis of Hydrogels and Their Progress in Environmental Remediation and Antimicrobial Application. Gels 2022, 9, 16. [Google Scholar] [CrossRef] [PubMed]
  202. Zurawin, R.K.; Ayensu-Coker, L. Innovations in contraception: A review. Clin. Obstet. Gynecol. 2007, 50, 425–439. [Google Scholar] [CrossRef]
  203. Thonneau, P.F.; Almont, T.E. Contraceptive efficacy of intrauterine devices. Am. J. Obstet. Gynecol. 2008, 198, 248–253. [Google Scholar] [CrossRef]
  204. Jeffey, T.J.; Lukkari-Lax, E.; Schulze, A.; Wahdan, Y.; Serrani, M.; Kroll, R. Contraceptive efficacy and safety of the 52-mg levonorgestrel intrauterine system for up to 8 years: Findings from the Mirena Extension Trial. Googology 2022, 227, 873. [Google Scholar] [CrossRef]
  205. Yu, P.; Zhong, W. Hemostatic materials in wound care. Burn. Trauma 2021, 9, tkab019. [Google Scholar] [CrossRef]
  206. Brannon-Peppas, L.; Blanchette, J.O. Nanoparticle and targeted systems for cancer therapy. Adv. Drug Deliv. Rev. 2004, 56, 1649–1659. [Google Scholar] [CrossRef]
  207. Mudhol, S.; Peddha, M.S. Development of capsaicin loaded nanoparticles based microneedle patch for transdermal drug delivery. J. Drug Deliv. Sci. Technol. 2023, 80, 104120. [Google Scholar] [CrossRef]
  208. Hansson, E.; Burian, P.; Hallberg, H. Comparison of inflammatory response and synovial metaplasia in immediate breast reconstruction with a synthetic and a biological mesh: A randomized controlled clinical trial. J. Plast. Surg. Hand Surg. 2020, 54, 131–136. [Google Scholar] [CrossRef]
  209. Klosterhalfen, B.; Hermanns, B.; Rosch, R.; Junge, K. Biological response to mesh. Eur. Surg.-Acta Chir. Austriaca 2003, 35, 16–20. [Google Scholar] [CrossRef]
  210. Liu, H.; Wang, C.; Li, C.; Qin, Y.; Wang, Z.; Yang, F.; Li, Z.; Wang, J. A functional chitosan-based hydrogel as a wound dressing and drug delivery system in the treatment of wound healing. RSC Adv. 2018, 8, 7533–7549. [Google Scholar] [CrossRef]
  211. Boateng, J.S.; Matthews, K.H.; Stevens, H.N.E.; Eccleston, G.M. Wound healing dressings and drug delivery systems: A review. J. Pharm. Sci. 2008, 97, 2892–2923. [Google Scholar] [CrossRef]
  212. Che, X.; Zhao, T.; Hu, J.; Yang, K.; Ma, N.; Li, A.; Sun, Q.; Ding, C.; Ding, Q. Application of Chitosan-Based Hydrogel in Promoting Wound Healing: A Review. Polymers 2024, 16, 344. [Google Scholar] [CrossRef] [PubMed]
  213. Alhamad, A.A.; Zeghoud, S.; Amor, I.B.; Hemmami, H. Chitosan-based hydrogels for wound healing: Correspondence. Int. J. Surg. 2023, 109, 1821–1822. [Google Scholar] [CrossRef] [PubMed]
  214. Chen, C.H.; Chen, S.H.; Shalumon, K.T.; Chen, J.P. Dual functional core-sheath electrospun hyaluronic acid/polycaprolactone nanofibrous membranes embedded with silver nanoparticles for prevention of peritendinous adhesion. Acta Biomater. 2015, 26, 225–235. [Google Scholar] [CrossRef] [PubMed]
  215. Chen, C.T.; Chen, C.H.; Sheu, C.; Chen, J.P. Ibuprofen-loaded hyaluronic acid nanofibrous membranes for prevention of postoperative tendon adhesion through reduction of inflammation. Int. J. Mol. Sci. 2019, 20, 5038. [Google Scholar] [CrossRef]
  216. Barreiro, D.L.; Yeo, J.; Tarakanova, A.; Martin-Martinez, F.J.; Buehler, M.J. Multiscale Modeling of Silk and Silk-Based Biomaterials—A Review. Macromol. Biosci. 2019, 19, e1800253. [Google Scholar] [CrossRef]
  217. Babu, P.J.; Suamte, L. Applications of silk-based biomaterials in biomedicine and biotechnology. Eng. Regen. 2024, 5, 56–69. [Google Scholar] [CrossRef]
  218. Fernando, I.P.S.; Lee, W.W.; Han, E.J.; Ahn, G. Alginate-based nanomaterials: Fabrication techniques, properties, and applications. Chem. Eng. J. 2020, 391, 123823. [Google Scholar] [CrossRef]
  219. Henderson, T.; Christman, K.L.; Alperin, M. Regenerative Medicine in Urogynecology: Where We Are and Where We Want to Be. Urogynecology 2024, 30, 519–527. [Google Scholar] [CrossRef]
  220. Romanski, P.A.; Bortoletto, P.; Pfeifer, S.M. Creation of a novel inflatable vaginal stent for McIndoe vaginoplasty. Fertil. Steril. 2021, 115, 804–806. [Google Scholar] [CrossRef]
  221. Mironska, E.; Chapple, C.; Macneil, S. Recent advances in pelvic floor repair. F1000Research 2019, 8, 778. [Google Scholar] [CrossRef] [PubMed]
  222. Mangir, N.; Bullock, A.J.; Roman, S.; Osman, N.; Chapple, C.; MacNeil, S. Production of ascorbic acid releasing biomaterials for pelvic floor repair. Acta Biomater. 2016, 29, 188–197. [Google Scholar] [CrossRef]
  223. Kuramoto, G.; Takagi, S.; Ishitani, K.; Shimizu, T.; Okano, T.; Matsui, H. Preventive effect of oral mucosal epithelial cell sheets on intrauterine adhesions. Hum. Reprod. 2015, 30, 406–416. [Google Scholar] [CrossRef] [PubMed]
  224. Smith, G.D.; Takayama, S. Application of microfluidic technologies to human assisted reproduction. Mol. Hum. Reprod. 2017, 23, 257–268. [Google Scholar] [CrossRef]
  225. Nogueira, D.; Sadeu, J.C.; Montagut, J. In vitro oocyte maturation: Current status. Semin. Reprod. Med. 2012, 30, 199–213. [Google Scholar] [CrossRef] [PubMed]
  226. Das, M.; Son, W.Y. In vitro maturation (IVM) of human immature oocytes: Is it still relevant? Reprod. Biol. Endocrinol. 2023, 21, 110. [Google Scholar] [CrossRef]
  227. Johnson, M.T.; Gardner, D.K. Embryo culture in the twenty-first century. In Human Assisted Reproductive Technology; Cambridge University Press: Cambridge, UK, 2011; pp. 232–247. [Google Scholar] [CrossRef]
  228. Vajta, G.; Kuwayama, M. Improving cryopreservation systems. Theriogenology 2006, 65, 236–244. [Google Scholar] [CrossRef]
  229. Rall, W.F.; Fahy, G.M. Ice-free cryopreservation of mouse embryos at −196 °C by vitrification. Nature 1985, 313, 573–575. [Google Scholar] [CrossRef]
  230. Bavister, B.D. Culture of preimplantation embryos: Facts and artifacts. Hum. Reprod. Update 1995, 1, 91–148. [Google Scholar] [CrossRef]
  231. Wang, X.; Wu, D.; Li, W.; Yang, L. Emerging biomaterials for reproductive medicine. Eng. Regen. 2021, 2, 230–245. [Google Scholar] [CrossRef]
  232. Lee, J.C.; Badell, M.L.; Kawwass, J.F. The impact of endometrial preparation for frozen embryo transfer on maternal and neonatal outcomes: A review. Reprod. Biol. Endocrinol. 2022, 20, 40. [Google Scholar] [CrossRef] [PubMed]
  233. Simon, A.; Laufer, N. Assessment and treatment of repeated implantation failure (RIF). J. Assist. Reprod. Genet. 2012, 29, 1227–1239. [Google Scholar] [CrossRef] [PubMed]
  234. Ferenczy, A. Pathophysiology of endometrial bleeding. Maturitas 2003, 45, 1–14. [Google Scholar] [CrossRef]
  235. Ryder, O.A. Cloning advances and challenges for conservation. Trends Biotechnol. 2002, 20, 231–232. [Google Scholar] [CrossRef]
  236. Nargund, G.; Fauser, B.C.J.M.; Macklon, N.S.; Ombelet, W.; Nygren, K.; Frydman, R. The ISMAAR proposal on terminology for ovarian stimulation for IVF. Hum. Reprod. 2007, 22, 2801–2804. [Google Scholar] [CrossRef]
  237. Sivasankaran, S.; Jonnalagadda, S. Advances in controlled release hormonal technologies for contraception: A review of existing devices, underlying mechanisms, and future directions. J. Control. Release 2021, 300, 797–811. [Google Scholar] [CrossRef] [PubMed]
  238. Rath, W.H. Postpartum hemorrhage—Update on problems of definitions and diagnosis. Acta Obstet. Gynecol. Scand. 2011, 90, 421–428. [Google Scholar] [CrossRef]
  239. Chen, C.; Lee, S.M.; Kim, J.W.; Shin, J.H. Recent update of embolization of postpartum hemorrhage. Korean J. Radiol. 2018, 19, 585–596. [Google Scholar] [CrossRef]
  240. Abatangelo, G.; Vindigni, V.; Avruscio, G.; Pandis, L.; Brun, P. Hyaluronic acid: Redefining its role. Cells 2020, 9, 1743. [Google Scholar] [CrossRef]
  241. Antonelli, A.; Veccia, A.; Morena, T.; Furlan, M.; Peroni, A.; Simeone, C. Robot-assisted vesico-vaginal fistula repair: Technical nuances. Int. Braz. J. Urol. 2021, 47, 684–685. [Google Scholar] [CrossRef]
  242. Li, W.; Li, Y.; Zhao, X.; Cheng, C.; Burjoo, A.; Yang, Y.; Xu, D. Diagnosis and treatment of cervical incompetence combined with intrauterine adhesions. Ann. Transl. Med. 2020, 8, 54. [Google Scholar] [CrossRef] [PubMed]
  243. Nezhad-Mokhtari, P.; Ghorbani, M.; Roshangar, L.; Rad, J.S. Chemical gelling of hydrogels-based biological macromolecules for tissue engineering: Photo- and enzymatic-crosslinking methods. Int. J. Biol. Macromol. 2019, 139, 760–772. [Google Scholar] [CrossRef] [PubMed]
  244. Yue, Y.; Norikane, Y.; Azumi, R.; Koyama, E. Light-induced mechanical response in crosslinked liquid-crystalline polymers with photoswitchable glass transition temperatures. Nat. Commun. 2018, 9, 3234. [Google Scholar] [CrossRef]
  245. Rusitov, D.; Deussen, F.; Rühe, J. Fast UV- and Visible Light Induced Polymer Network Formation Using Carbene Mediated C-H-insertion Based Crosslinking (CHic) Via Nitrophenyl Diazo Esters. Adv. Mater. Interfaces 2023, 10, 2300316. [Google Scholar] [CrossRef]
  246. Bende, A.; Farcaş, A.A.; Toşa, V. Theoretical Study of Light-Induced Crosslinking Reaction Between Pyrimidine DNA Bases and Aromatic Amino Acids. Front. Bioeng. Biotechnol. 2022, 9, 806415. [Google Scholar] [CrossRef] [PubMed]
  247. Naranjo-Alcazar, R.; Bendix, S.; Groth, T.; Gallego Ferrer, G. Research Progress in Enzymatically Cross-Linked Hydrogels as Injectable Systems for Bioprinting and Tissue Engineering. Gels 2023, 9, 230. [Google Scholar] [CrossRef]
  248. Sharifi, S.; Sharifi, H.; Akbari, A.; Chodosh, J. Systematic optimization of visible light-induced crosslinking conditions of gelatin methacryloyl (GelMA). Sci. Rep. 2021, 11, 23276. [Google Scholar] [CrossRef]
  249. Maddock, R.M.A.; Pollard, G.J.; Moreau, N.G.; Perry, J.J.; Race, P.R. Enzyme-catalysed polymer cross-linking: Biocatalytic tools for chemical biology, materials science and beyond. Biopolymers 2020, 111, e23390. [Google Scholar] [CrossRef]
  250. Rusu, A.G.; Nita, L.E.; Simionescu, N.; Ghilan, A.; Chiriac, A.P.; Mititelu-Tartau, L. Enzymatically-Crosslinked Gelatin Hydrogels with Nanostructured Architecture and Self-Healing Performance for Potential Use as Wound Dressings. Polymers 2023, 15, 780. [Google Scholar] [CrossRef]
  251. Stadlmeier, M.; Runtsch, L.S.; Streshnev, F.; Wühr, M.; Carell, T. A Click-Chemistry-Based Enrichable Crosslinker for Structural and Protein Interaction Analysis by Mass Spectrometry. ChemBioChem 2020, 21, 103–107. [Google Scholar] [CrossRef]
  252. Paioti, P.H.S.; Lounsbury, K.E.; Romiti, F.; Formica, M.; Bauer, V.; Zandonella, C.; Hackey, M.E.; del Pozo, J.; Hoveyda, A.H. Click processes orthogonal to CuAAC and SuFEx forge selectively modifiable fluorescent linkers. Nat. Chem. 2024, 16, 426–436. [Google Scholar] [CrossRef]
  253. Wang, C.F.; Mäkilä, E.M.; Kaasalainen, M.H.; Liu, D.; Sarparanta, M.P.; Airaksinen, A.J.; Salonen, J.J.; Hirvonen, J.T.; Santos, H.A. Copper-free azide-alkyne cycloaddition of targeting peptides toporous silicon nanoparticles for intracellular drug uptake. Biomaterials 2014, 35, 1257–1266. [Google Scholar] [CrossRef] [PubMed]
  254. Jain, E.; Neal, S.; Graf, H.; Tan, X.; Balasubramaniam, R.; Huebsch, N. Copper-Free Azide-Alkyne Cycloaddition for Peptide Modification of Alginate Hydrogels. ACS Appl Bio. Mater. 2021, 4, 1229–1237. [Google Scholar] [CrossRef]
  255. Wang, J.; Chen, D.; Xing, S.; Ji, B.; Yang, J.; Tang, J. Highly Thermal Stable, Stiff, and Recyclable Self-Healing Epoxy Based on Diels-Alder Reaction. ACS Appl. Polym. Mater. 2024, 6, 466–474. [Google Scholar] [CrossRef]
  256. Lin, T.C.; Mocny, P.; Cvek, M.; Sun, M.; Matyjaszewski, K. Grafting well-defined polymers onto unsaturated PVDF using thiol-ene reactions. Polymer 2024, 297, 126848. [Google Scholar] [CrossRef]
  257. Hillman, A.S.; Hyland, S.N.; Wodzanowski, K.A.; Moore, D.V.L.; Ratna, S.; Jemas, A.; Sandles, L.M.D.; Chaya, T.; Ghosh, A.; Fox, J.M.; et al. Minimalist Tetrazine N-Acetyl Muramic Acid Probes for Rapid and Efficient Labeling of Commensal and Pathogenic Peptidoglycans in Living Bacterial Culture and During Macrophage Invasion. J. Am. Chem. Soc. 2024, 146, 6817–6829. [Google Scholar] [CrossRef]
  258. Nie, L.; Sun, Y.; Okoro, O.V.; Deng, Y.; Jiang, G.; Shavandi, A. Click chemistry for 3D bioprinting. Mater. Horiz. 2023, 10, 2727–2763. [Google Scholar] [CrossRef]
  259. Zhao, L.; Zhong, B.; An, Y.; Zhang, W.; Gao, H.; Zhang, X.; Liang, Z.; Zhang, Y.; Zhao, Q.; Zhang, L. Enhanced protein–protein interaction network construction promoted by in vivo cross-linking with acid-cleavable click-chemistry enrichment. Front. Chem. 2022, 10, 994572. [Google Scholar] [CrossRef]
  260. Zhang, B.; Gao, H.; Gong, Z.; Zhao, L.; Zhong, B.; Sui, Z.; Liang, Z.; Zhang, Y.; Zhao, Q.; Zhang, L. Improved Cross-Linking Coverage for Protein Complexes Containing Low Levels of Lysine by Using an Enrichable Photo-Cross-Linker. Anal. Chem. 2023, 95, 9445–9452. [Google Scholar] [CrossRef]
  261. Kitayama, Y.; Harada, A. Visible Light Induced Interfacial Photo-Cross-Linking for Fabrication of Polymer Capsules Using Photoreactive Polymers with Substituted Cinnamates. ACS Appl. Polym. Mater. 2024, 6, 3805–3813. [Google Scholar] [CrossRef]
  262. Barman, S.; Padhan, J.; Sudhamalla, B. Uncovering the non-histone interactome of the BRPF1 bromodomain using site-specific azide-acetyllysine photochemistry. J. Biol. Chem. 2024, 300, 105551. [Google Scholar] [CrossRef]
  263. Faustino, A.M.; Sharma, P.; Manriquez-Sandoval, E.; Yadav, D.; Fried, S.D. Progress toward Proteome-Wide Photo-Cross-Linking to Enable Residue-Level Visualization of Protein Structures and Networks In Vivo. Anal. Chem. 2023, 95, 10670–10685. [Google Scholar] [CrossRef] [PubMed]
  264. Nozari, N.; Biazar, E.; Kamalvand, M.; Keshel, S.H.; Shirinbakhsh, S. Photo Cross-linkable Biopolymers for Cornea Tissue Healing. Curr. Stem. Cell Res. Ther. 2021, 17, 58–70. [Google Scholar] [CrossRef] [PubMed]
  265. Farell, I.S.; Toroney, R.; Hazen, J.L.; Mehl, R.A.; Chin, J.W. Photo-cross-linking interacting proteins with a genetically encoded benzophenone. Nat. Methods 2005, 2, 377–384. [Google Scholar] [CrossRef]
  266. Xiao, X.; Yang, Y.; Lai, Y.; Huang, Z.; Li, C.; Yang, S.; Niu, C.; Yang, L.; Feng, L. Customization of an Ultrafast Thiol-Norbornene Photo-Cross-Linkable Hyaluronic Acid-Gelatin Bioink for Extrusion-Based 3D Bioprinting. Biomacromolecules 2023, 24, 5414–5427. [Google Scholar] [CrossRef]
  267. Zhou, X.; Wang, Y.; Wang, G.; Li, P.; Zhao, H.; Liu, Y.; Gao, C. Oxidative crosslinking induced self-reinforcing waterborne polyurethane with tunable structure as multi-synergistic antibacterial biomimetic composite coating. Colloids Surf. A Physicochem. Eng. Asp. 2024, 683, 133054. [Google Scholar] [CrossRef]
  268. Li, M.; Karboune, S.; Light, K.; Kermasha, S. Oxidative cross-linking of potato proteins by fungal laccases: Reaction kinetics and effects on the structural and functional properties. Innov. Food Sci. Emerg. Technol. 2021, 71, 102723. [Google Scholar] [CrossRef]
  269. Wang, H.; Ji, Y.; Yuan, Z.; Tian, J.; Zhang, Y.; Lu, F.; Liu, Y. Insights into the mechanism on the high-temperature activity of transglutaminase from Bacillus clausii and its crosslinked mode at protein level. Biochem. Eng. J. 2022, 185, 108544. [Google Scholar] [CrossRef]
  270. Peterson, L.L.; Zettergren, J.G.; Wuepper, K.D. Biochemistry of transglutaminases and cross-linking in the skin. J. Investig. Dermatol. 1983, 81, S95–S100. [Google Scholar] [CrossRef]
  271. Yang, Z.; Li, G.; Hou, Y. Application of Feruloyl Esterase in Wheat Straw Pulp Bleaching. Highlights Sci. Eng. Technol. 2023, 66, 197–200. [Google Scholar] [CrossRef]
  272. Luo, C.; Hu, Y.; Xing, S.; Xie, W.; Li, C.; He, L.; Wang, X.; Zeng, X. Adsorption–precipitation–cross-linking immobilization of GDSL-type esterase from Aspergillus niger GZUF36 by polydopamine-modified magnetic clarity tetroxide nanocouplings and its enzymatic characterization. Int. J. Biol. Macromol. 2023, 245, 125533. [Google Scholar] [CrossRef] [PubMed]
  273. Greenberg, C.S.; Birckbichler, P.J.; Rice, R.H. Transglutaminases: Multifunctional cross-linking enzymes that stabilize tissues. FASEB J. 1991, 5, 1683845. [Google Scholar] [CrossRef]
  274. Duarte, L.; Matte, C.R.; Bizarro, C.V.; Ayub, M.A.Z. Review transglutaminases: Part II—Industrial applications in food, biotechnology, textiles and leather products. World J. Microbiol. Biotechnol. 2020, 36, 11. [Google Scholar] [CrossRef] [PubMed]
  275. Lu, Z.; Liu, Y. A Review of the Mechanism, Properties, and Applications of Hydrogels Prepared by Enzymatic Cross-linking. J. Agric. Food Chem. 2023, 71, 10238–10249. [Google Scholar] [CrossRef]
  276. Du, E.; Li, M.; Xu, B.; Zhang, Y.; Li, Z.; Yu, X.; Fan, X. Self-Healing and Adhesive Supersoft Materials Derived from Dynamically Cross-Linked Bottlebrush Polymers for Flexible Sensors. Macromolecules 2024, 57, 672–681. [Google Scholar] [CrossRef]
  277. Ji, Y.; Kim, T.; Han, D.; Lee, J.B. Self-Healing and Thermal Responsive DNA Bioplastics for On-Demand Degradable Medical Devices. ACS Mater. Lett. 2024, 6, 1277–1287. [Google Scholar] [CrossRef]
  278. Liao, H.; Su, J.; Han, J.; Xiao, T.; Sun, X.; Cui, G.; Duan, X.; Shi, P. An Intrinsic Self-Healable, Anti-Freezable and Ionically Conductive Hydrogel for Soft Ionotronics Induced by Imidazolyl Cross-Linker Molecules Anchored with Dynamic Disulfide Bonds. Macromol. Rapid. Commun. 2024, 45, e2300613. [Google Scholar] [CrossRef]
  279. Kaymazlar, E.; Dikbas, C.; Matar, G.H.; Andac, O.; Andac, M. Self-healable and conductive mussel inspired PVA/borax@PDA–LiTFSI hydrogel-based self-adhesive for human motion sensor. Polym. Bull. 2024, 81, 8751–8764. [Google Scholar] [CrossRef]
  280. Moody, C.T.; Palvai, S.; Brudno, Y. Click cross-linking improves retention and targeting of refillable alginate depots. Acta Biomater. 2020, 112, 112–121. [Google Scholar] [CrossRef]
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

Yammine, P.; El Safadi, A.; Kassab, R.; El-Nakat, H.; Obeid, P.J.; Nasr, Z.; Tannous, T.; Sari-Chmayssem, N.; Mansour, A.; Chmayssem, A. Types of Crosslinkers and Their Applications in Biomaterials and Biomembranes. Chemistry 2025, 7, 61. https://doi.org/10.3390/chemistry7020061

AMA Style

Yammine P, El Safadi A, Kassab R, El-Nakat H, Obeid PJ, Nasr Z, Tannous T, Sari-Chmayssem N, Mansour A, Chmayssem A. Types of Crosslinkers and Their Applications in Biomaterials and Biomembranes. Chemistry. 2025; 7(2):61. https://doi.org/10.3390/chemistry7020061

Chicago/Turabian Style

Yammine, Paolo, Ali El Safadi, Rima Kassab, Hanna El-Nakat, Pierre J. Obeid, Zeina Nasr, Tony Tannous, Nouha Sari-Chmayssem, Agapy Mansour, and Ayman Chmayssem. 2025. "Types of Crosslinkers and Their Applications in Biomaterials and Biomembranes" Chemistry 7, no. 2: 61. https://doi.org/10.3390/chemistry7020061

APA Style

Yammine, P., El Safadi, A., Kassab, R., El-Nakat, H., Obeid, P. J., Nasr, Z., Tannous, T., Sari-Chmayssem, N., Mansour, A., & Chmayssem, A. (2025). Types of Crosslinkers and Their Applications in Biomaterials and Biomembranes. Chemistry, 7(2), 61. https://doi.org/10.3390/chemistry7020061

Article Metrics

Back to TopTop