Next Article in Journal
Studies on Polybenzimidazole and Methanesulfonate Protic-Ionic-Liquids-Based Composite Polymer Electrolyte Membranes
Next Article in Special Issue
Modification of Starches and Flours by Acetylation and Its Dual Modifications: A Review of Impact on Physicochemical Properties and Their Applications
Previous Article in Journal
Refractive Properties of Conjugated Organic Materials Doped with Fullerenes and Other Carbon-Based Nano-Objects
Previous Article in Special Issue
Effect of Spent Coffee Grounds on the Crystallinity and Viscoelastic Behavior of Polylactic Acid Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Chitosan and Chitosan Nanofiber: Preparation, Characterization, and Its Potential Applications

1
Department of Biology, College of Science, Majmaah University, Al-Majmaah 11952, Saudi Arabia
2
Faculty of Women for Arts, Science and Education, Ain Shams University, Cairo 11566, Egypt
3
Department of Chemistry, College of Science, Majmaah University, Al-Majmaah 11952, Saudi Arabia
4
Faculty of Applied Arts, Textile Printing, Dyeing and Finishing Department, Helwan University, Cairo 11795, Egypt
5
Pretreatment and Finishing of Cellulosic Fibers Department, Textile Research and Technology Institute, National Research Centre, Cairo 12622, Egypt
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(13), 2820; https://doi.org/10.3390/polym15132820
Submission received: 25 April 2023 / Revised: 13 June 2023 / Accepted: 15 June 2023 / Published: 26 June 2023

Abstract

:
Chitosan is produced by deacetylating the abundant natural chitin polymer. It has been employed in a variety of applications due to its unique solubility as well as its chemical and biological properties. In addition to being biodegradable and biocompatible, it also possesses a lot of reactive amino side groups that allow for chemical modification and the creation of a wide range of useful derivatives. The physical and chemical characteristics of chitosan, as well as how it is used in the food, environmental, and medical industries, have all been covered in a number of academic publications. Chitosan offers a wide range of possibilities in environmentally friendly textile processes because of its superior absorption and biological characteristics. Chitosan has the ability to give textile fibers and fabrics antibacterial, antiviral, anti-odor, and other biological functions. One of the most well-known and frequently used methods to create nanofibers is electrospinning. This technique is adaptable and effective for creating continuous nanofibers. In the field of biomaterials, new materials include nanofibers made of chitosan. Numerous medications, including antibiotics, chemotherapeutic agents, proteins, and analgesics for inflammatory pain, have been successfully loaded onto electro-spun nanofibers, according to recent investigations. Chitosan nanofibers have several exceptional qualities that make them ideal for use in important pharmaceutical applications, such as tissue engineering, drug delivery systems, wound dressing, and enzyme immobilization. The preparation of chitosan nanofibers, followed by a discussion of the biocompatibility and degradation of chitosan nanofibers, followed by a description of how to load the drug into the nanofibers, are the first issues highlighted by this review of chitosan nanofibers in drug delivery applications. The main uses of chitosan nanofibers in drug delivery systems will be discussed last.

1. Introduction

Henry Braconot (1780–1855) isolated chitin from mushrooms in 1811. Chitin was the second carbohydrate discovered in 1859 [1]. It was hydrolyzed in an alkaline medium in 1894 to produce a new carbohydrate soluble in dilute acids called chitosan. Chitosan and its oligomers drew a lot of attention because of their unique properties [2,3,4]. Chitin represents the second natural polysaccharide in nature, after cellulose. Its chemical name is poly N-acetamido-2-decoxy-β-d-glucose [5]. Figure 1 shows the chemical structure and some sources of chitin and chitosan [6,7,8,9].

2. Structure of Chitin and Chitosan

Chitosan, a white polymer found in crabs, is a nitrogenous polysaccharide that is rigid inelastic and found both inside and outside the exoskeleton of invertebrates. The primary distinction is the solubility of chitosan in diluted acid (at PH = 7). The ocean, lakes, and seas are the sources of chitosan products [10]. Chitosan has potential applications in various fields, such as biomedical applications and the fabrication of protective clothes [10].
Chitin and chitosan, as natural renewable biopolymers, have unique properties, as they are biocompatible, biodegradable, and non-toxic [11]. Because of the NH2 groups that open the structure of chitosan, it is easier to modify than cellulose. As a result, it can be converted into a variety of chitosan derivatives [12]. The main purpose of this modification is to improve its solubility and chemical reactivity [13,14].
Acetylation degree is the presence of acetyl glucose amine to glucose amine in the chitin structure. Which determines the solubility of chitin and chitosan? Chitosan is more accepted than chitin due to its high degree of solubility in dilute acetic acid [15,16].
The presence of acetyl groups in chitin and chitosan makes their solubility problematic, although M.W.T. represents an important later component for chitin and chitosan properties. There are several methods used to estimate the molecular weight (M.W.T.) of chitin and chitosan, such as light scattering and viscosity [15,16].
Chitin is hydrophobic material insoluble in water and most organic solvents. Its solvent is a mixture of 1,2-diehloroethane and trichloroacetic acid (35:65), fresh saturated solution of lithium thiocyanate, chloroalcohols in conjugation with aqueous solutions of mineral acids and Dimethylacetamide containing 5% lithium chloride [10,17].
Chitin and chitosan are biocompatible, biodegradable, and non-toxic. Because they are amino polysaccharides, they exhibit interesting properties in biology, pharmacology, and physiology and have numerous applications such as wound healing, wound dressing, hemostatic agency, and antimicrobial activity, requiring them to be used as a biomedical material [17].
Figure 2 shows chitosan-based materials with different shapes and sizes: chitosan nanoparticles, chitosan nanofibers fabricated by electrostatic spinning technology, chitosan–pectin hydrogel grid scaffold prepared by 3D printing technology, chitosan core-alginate shell microspheres, chitosan-based fibers fabricated by solvent spinning technology, and 3D-printed chitosan porous structures [18].

3. Modification of Chitin and Chitosan

Although chitin and chitosan possess several useful qualities, their application is constrained by their poor solubility, small surface area, and porous makeup. To modify chitin and chitosan physically or chemically, various researchers have tested their theories. The two main benefits of modifying chitin and chitosan are to increase their solubility and improve their ability to absorb metals. The OH at the C3, C6, and NH2 at the C2 groups in chitosan undergo substitution reactions. These changes also improve the capacity of the membrane to swell in water [19,20,21].
In the molecular structure of chitosan, there are three types of active groups: amino groups, and primary and secondary hydroxyl groups at the C-3 and C-6 positions, which allow for chemical modification of chitosan. C6-OH is a main hydroxyl group with little steric hindrance, whereas C3-OH is a secondary hydroxyl group with a lot. As a result, the main hydroxyl group could freely rotate while the secondary hydroxyl group could not. The amino group is usually more reactive than the primary hydroxyl group, and the primary hydroxyl group is usually more reactive than the secondary hydroxyl group [22,23,24]. Chitosan can be chemically modified on the amino, hydroxyl, or both amino and hydroxyl groups to generate N-modified, O-modified, or N, O-modified chitosan derivatives [25] (Figure 3).
The proposed mechanism of chitosan modification can pass through one of the following four mechanisms: (a) free radical-induced conjugation to form a polyphenol chitosan conjugation; (b) carbodiimide chemical mechanism to form Schiff base compounds; (c) functional group conversion strategy, which converts the amino group of chitosan into an azide group, substituted carboxyl group, substituted mercapto group, and where the hydroxyl group can be azidate, aminated, oxidized to an aldehyde or carbonyl group, or further oxidized to a carboxyl group; or (d) conjugation of chitosan with polyphenol via enzymatic assisted coupling reaction, as shown in Figure 4a–d [26].
Chitosan has a low specific surface area (2–30 m2·g−1) [27] and is present in flake, which is unsuitable for many applications, so that it has been modified into beads to increase its value in different fields of application [19,20,21]. Due to the open structure and pores of chitosan, which has poor mechanical properties, shrinkage, and deformation in dry form and is only soluble in weak acids, numerous modifications have been made to it to improve its properties [28]. Chitosan contains amino and hydroxyl groups, so it is a poly-nucleophilic polymer. Nucleophilic substitution occurs. Protonation in NH2 groups results in the formation of NH3+ [29,30], such as N-alkylated chitosan prepared from Schiff base reactions followed by imine reduction by sodium borohydride. Furthermore, positive chitosan charges interacted with polyanion polymers such as alginate, carrageen, and pectin between the COOH and NH3+ groups [31]. Chitosan produce new functionalized derivatives via a grafting reaction [32]. The properties of grafted chitosan are determined by the side chains and the cross-liking agent [33]. Chitin threads are prepared to be used in absorbable suture fabrication, dressing, and biodegradable materials for human skin fiber growth [34].
Chitosan is a hydrophilic material used to impart hydrophilicity to some other polymer in its composites, such as polyacrylonitrile (PAN) [35]. Hydrophilic polymer nanoparticles were prepared by ionic gelation in mild conditions at room temperature, via two phases; one is for polymers such as chitosan polysaccharide (CS) and polyethylene oxide (PEO), and the other phase contains sodium tripolyphosphate (TPP). Calvo et al. prepared chitosan nanoparticles with high protein loading capacity, which were released within a week [36]. The morphology of these nanoparticles is spherical, as shown in Figure 5.
Nanoparticles of chitosan by using a variety of agents, the freeze-drying method increases shelf life. Chitosan nanoparticles of two different types were prepared by Alonso et al. in 1999 [37]. The prepared nanoparticles ranged in size from 300 to 400 nm, had a positive surface charge, and were efficient. These nanoparticles are used to address the nasal absorption of insulin. In addition, chitosan nanoparticles were used as a poly load of the anthracycline drug doxorubicin (DOX) [38,39]. In addition, these nanoparticles were used to load dextran sulphate to enhance its drug loading [40]. With cyclosporine serving as the model drug, chitosan nanoparticles (CSNP) were used to enhance the drug at the ocular surface. The size of these nanoparticles was 29 nm [41].
Chitosan-coated PLGA–Lecithin nanoparticles were prepared by the modified double emulsion method; these nanoparticles were used through the oral or nasal administration route. Figure 6 depicts the process of preparing CS–PLGA nanoparticles, and Table 1 summarizes the properties of the prepared nanoparticles [41].
Chitosan nanoparticles loaded by Mycobacterium uaceae aerivatiuem antineoplastic proteoglycans exhibit wide antimicrobial activity, as repeated by Tian and Groves (1999) [42]. The researchers prepared chitosan nanoparticles with particle sizes of 600–700 nm without using organic solvents and discovered that the two reactants affected the absorption and release of bovine, as well as that the initial Nanoparticles of chitosan by using a variety of agents, the freeze-drying method increases shelf life. Chitosan nanoparticles of two different types were prepared by Alonso et al. in 1999 release was followed by a steady release for 4 h in water [43].
Using self-aggregates of chitosan modified by deoxycholic acid, there is a novel and straightforward method of delivering adriamycin [44].
Deoxycholic acid and chitosan are covalently conjugated through an EDC-mediated reaction, resulting in self-aggregating chitosan nanoparticles. The Adriamycin active ingredient was physically trapped in nanoparticles, and the resulting self-aggregates were assessed using a variety of methods, including spectroscopy, which shows that these self-aggregates are spherical and that the drug concentration affects the shape of the particles [44].
These self-aggregates of chitosan modified by deoxycholic acid were used as DNA carriers by Kim et al. in 2001 [45]. Chitosan is used to overcome the side effects of drugs such as their solubility and hydrophobicity so that it can be used as a drug carrier. For example, commercial chitosan can be used to control body weight.

4. Chitosan as Biomaterial

Chitosan is a semi-crystalline polymer that is partially deacetylated chitin, and the degree of crystallinity is correlated with the degree of deacetylation. Chitin and fully deacetylated chitosan have the highest crystallinity, while intermediate levels of deacetylation have the lowest crystallinity [12]. Chitosan is stable, crystalline, and soluble in aqueous solutions at PH 7, but it is insoluble in weak acids due to the protonation of amino groups. Chitosan solutions could be extruded at higher PH values or in non-solvent baths, e.g., methyl alcohol. Chitosan polymer is used for industrial applications in such forms as fiber or film [46]. Chitosan solution’s cationic nature and high charge density make it potentially useful as a biomaterial. Chitosan forms soluble ionic complexes or complexes with water soluble anionic polymers such as alginates and synthetic polymers such as poly (acrylic acid) due to its charge density [21]. When used for local delivery of biologically active poly anions like GAGs and DNA, for example, ionic complexes release heparin to increase the efficacy of growth factor secreted by inflammatory cells [21,47,48,49,50,51].
Chitosan DNA complex is used to facilitate cellular transfection and prevent plasmid degradation by nucleases. Chitosan can become porous through freezing and lyophilization, making it useful for both tissue regeneration and cell transplantation [52]. Figure 7 illustrates how ice crystals form in the freezing process from a solution, grow from the ice crystal phase, and are then removed by lyophilization from a general porous mother.
Pore size and pore orientation affect the mechanical properties of scaffold-based chitosan, with porous chitosan membranes having less elasticity (0.1–0.5 MPa) than non-porous membranes (5–7 MPa).
The maximum strain of the porous membrane of chitosan is greater than that of the non-porous membrane, and the 100% chemical modification of chitosan introduces new biological activity with modified mechanical properties. The chitosan NH2 group is reactive and capable of introducing side group attachment with several reactions that affect primarily crystallinity disruption with lower stiffness and alter soluble derivatives feasible in chemical reactions, such as alkyl derivatives of chitosan, which have lower solubility than chitosan itself and give aggregates miscible for c > 5. In addition, the basic properties of chitosan are hemostatic, cationic, and insoluble at pH 7, which are completely reversed by the solvation process to give ionic, water-soluble derivatives, and anticoagulant properties, so that chitosan can attack with unlimited side groups and be chosen according to specific needed functions, e.g., biological activity or modified properties, etc.

5. Applications of Chitosan and Chitosan Derivatives

Chitosan is a biodegradable, biocompatible, and non-toxic biomaterial; therefore, it has many applications in areas such as medicine, agriculture, food processing, cosmetics, and treatment of water.

5.1. Agricultural Application

Chitosan can be used safely in agricultural applications because it does not pollute the environment or harm consumers; it is used as a leaf coating, fertilizer, and sea coating [10,53]. The use of chitosan in agricultural fields has increased exponentially, especially for germination improvement, leaf growth, retention of moisture, and fungal and disease reduction [53]. Chitosan boosts photosynthetic efficiency while enhancing plant tolerance to salinity, high temperatures, and drought [54]. Chitosan’s hydrophilic properties encourage water absorption while reducing transpiration [55]. To encourage plant development, chitosan can be employed as an additional carbon source in plant synthesis [56]. In comparison to the control group, seedlings treated with Cu-chitosan nanoparticles (NPs) at concentrations of 0.04 percent and 0.12 percent had greater rates of germination, seedling length, root length, and root number [57]. To encourage the mobilization of protein and starch to boost seedling growth, Cu-chitosan NPs can increase protease and -amylase activities [58].

5.2. Wastewater Treatment Applications

Chitosan contains both OH and NH2 groups as chelating agent groups, which allow it to be used in water treatments from wastes due to the high-power effect of these functional groups to bind with heavy dissolved metals present in wastewater such as Cu, Pb, Hg, and Ur [59]. Furthermore, chitosan can be used to break down food particles, particularly food proteins, as well as remove dyes from wastewater [60].
Cu (II) ion-containing wastewater was examined by Qin et al. [61] using sodium alginate and chitosan as treatment agents. They investigated how various parameters affected this ion removal. According to the findings, chitosan and sodium alginate worked better together than they did separately. Pesticide removal from wastewater was explored by Dwivedi et al. [61] using hydrogel beads made of chitosan and gold nanoparticles. The data obtained indicated that the synthetic sorbent has good pesticide removal capacity.
Heterogeneous catalysis is one technique that heavily relies on adsorption. Purification is one of the earliest documented uses of adsorption. Adsorbents are still used to clarify water [62]. Numerous adsorbents, including activated carbon, low-cost biomass adsorbents, waste sludge, rice husk, sugarcane bagasse, lignite, and chitosan, were used in the adsorption tests. Heavy metal contamination clean-up frequently employs chitosan. In order to study the adsorptions of three metal ions—Cu (II), Zn (II), and Pb (II) ions—in an aqueous solution, cross-linked chitosan was created by the homogeneous reaction of chitosan in an aqueous acetic acid solution with epichlorohydrin [63].

5.3. Food Industry Applications

Because of the high potential for toxicity of chitosan as a chelating agent and its high functional properties, as previously stated, chitosan is used in a variety of applications in the food industry, such as removing specific elements, particles, and undesirable materials, such as dyes and fats. In addition, it is widely used as a natural, safe preservative in the United States to store food [64,65].
There has been an increase in interest in recent years in studying the potential applications of chitosan as films or coatings in food packaging. This is due to their film-forming, antioxidant, and antibacterial properties, as well as their mechanical and barrier properties, which were studied as films. In order to extend the storability and shelf life of perishable commodities, these experiments sought to develop active packaging based on chitosan, either on its own or in combination with other materials. By combining chitosan with other natural antimicrobial agents, it is also possible to make food products that guarantee food safety against a variety of mutating and pathogenic bacteria [66,67,68].
An edible coating or thin edible film made of chitosan can be applied to food to act as active food packaging. The edible coating is a thin layer that is added to a food item and is generated by dipping the item in a chitosan solution or spraying, in which case, the film-forming solution is crushed up using an aerosol spray coating. Even though the chitosan film is a thin, prefabricated layer, once it is produced, it can be deposited on the surface of or in between food ingredients [66,67,69]. Studies using chitosan films and coatings on food products are shown in Table 2 [66].

5.4. Medical Applications

Chitosan is applicable in several medical industries, especially periodontal and orthopedic drug delivery, wound healing, and tissue engineering applications [86]. Surgical sutures, contact lenses, eye fluids, artificial skin, artificial blood vessels, bandages, sponges, burn dressings, blood cholesterol vessels, antitumor, antibacterial, antiviral, bane regenerator, antimicrobial, and hemostatic agents are the most well-known examples of these applications [86,87,88,89,90,91].
Chitosan inhibits tumor cell proliferation, and Liu et al. demonstrated that chitosan induces apoptosis in tumor cells by decreasing Bcl-2 and increasing Caspase-3 expression [92,93]. Carboxymethyl chitosan (CMCS) increases macrophage viability, deeply penetrates the tumor microenvironment, generates cytokines like TNF and IL-1, improves phagocytosis, and increases NO levels. Notably, CM-COS is not significantly toxic to normal liver cells but inhibits the growth of BEL-7402 and sarcoma cells in vivo [94,95]. Additionally, chitosan inhibits the invasion and metastasis of tumor cells. During tumor invasion and metastasis, matrix metalloenzymes (MMP) are involved in the breakdown of extracellular matrix, and MMP-2 can encode an enzyme that breaks down type IV collagen [96].
Chitosan is a great material for creating wound dressings because it can promote wound healing. It exhibits good antibacterial activity due to its alkaline amino groups, which cause the destruction of bacterial cells and protect the wound surface from microbial infection [97]. To improve the antibacterial and coagulation capabilities, Zhang et al. created a nanocomposite hydrogel utilizing zinc oxide (ZnO), chitosan, and aldehydic sodium alginate (SA) [98]. It significantly inhibited the growth of Escherichia coli and Staphylococcus aureus. I-sexual collagen hydrogels sulphated chitosan-doped by Shen et al. improved macrophage polarization from M1 to M2. The IL-4 and TGF-1 secreted by macrophages were stimulated, which resulted in an increase in collagen synthesis, regeneration epithelialization, and neovascularization [99].
Tissue engineering is an emerging interdisciplinary discipline combining material science, engineering mechanics, and biomedicine. The structure and function of damaged tissues and organs are repaired or replaced by cell transplantation combined with bioactive molecules and 3D scaffolds. Chitosan is similar to mucopolysaccharides of the extracellular matrix and is used as a scaffolding material for cartilage tissue engineering [100].
Kaviani et al. used the cryogel approach to construct nano porous scaffolds from chitosan, collagen, and nanohydroxyapatite, which lowered the rate of biodegradation, increased mechanical characteristics, and enabled cell proliferation and adhesion [58]. Porous scaffolds made from chitosan, gelatin, and silk proteins have higher compressive strength and modulus, whereas incidental chondrocytes can produce seed scaffolds and stimulate cartilage tissue regeneration [58]. By mixing gelatin, chitosan, and polyvinyl alcohol with nano-hydroxyapatite, Martino et al. created porous composite scaffolds [58]. Table 3 summarizes the biological application of chitosan and its derivatives.

6. Electro-Spun Nanofibers: Process and Application

6.1. History of Electrospinning

In 1897, Rayleigh discovered electrospinning. The early 1700s saw the discovery of electrostatic effects on water behavior and their impact on the dielectric values of liquid excitation. Around the turn of the twentieth century, Cooley and Morton developed electrospinning technology. The rotatory electrode was then incorporated into electrospinning by Cooley. Formhals created yarns using electrospinning in 1930 without the use of a spinneret, and his technique and apparatus granted him a patent for his creation [119,120,121].
After that, Formhals submitted a patent for a different method of creating electrostatic polymer fibers: composite fibers made of several polymers. In 1969, Taylor conducted research on the composition of the polymer droplet formed at the needle tip by a strong electric field. This proved that the droplet assumed a cone-like shape, with jets emerging from vortices. The “Taylor cone” is the final name given to this cone. Additionally, factors affecting fiber stability, such as the electric field, flow rate, and experimental settings, were examined [122,123].
Compared to conventional spinning, electrospinning produces fibers at a much slower rate. Electrospinning produces yarn at a rate of 30 m per minute, compared to conventional spinning’s 200–1500 m per minute [119,120,124]. So, before 1990, melt spinning was the preferred method for creating fibers from natural and synthetic polymers, and only a small number of businesses were interested in electrospinning for fiber production. Nanometer-scale fibers cannot be produced through melt spinning [120,124].
For the purpose of removing harmful solvents and for use in tissue engineering applications, Dalton et al. applied an electro spun nanofiber web to tissue cells [125]. The use of electro-spun nanofibers today spans a wide range of potential uses [119,124,126]. Surface nanostructures may produce extraordinary phenomena, such as the lotus effect (self-cleaning effect). Since proteins, viruses, and bacteria all have dimensions in this range, the nanoscale is particularly important for biological systems. Electro spun fibers display a strikingly broad range of sizes when compared to the diameters of these things. Figure 8 depicts the dimensions of bacterial cells, proteins, viruses, and nanofibers [119,124,126,127].

6.2. Electrospinning Process

A high electric field (kV) is used to create micro or nanofibers from polymer solutions while maintaining ambient temperature and pressure. There are two primary setups for electrospinning devices: vertical and horizontal [119,124,128].
The major components of the electrospinning device are the power supply (high voltage), the syringe (spinneret), and the collector (electrode) [119,123,129].
High voltage creates electric charges on the surface of the polymer solution during the electrospinning process, and these charges build up on the surface. These charges possess repelling forces in a critical electric field that can dissipate the surface tension of the solution and the unstable charges. A Taylor cone-tip jet ejects a solvent-causing jet [119,123,124,130].
A stable jet created at the spinneret needle then changes into an unstable jet to create electro-spun fibers using a straightforward process. When the applied electric field reaches a critical value, jets shoot out of the cone tip, and the tensile force is transferred to the polymer, creating bending instability in that polymer. After that, a jet moves from the cone’s apex to a collector with opposing charges, which has the power to draw charged fibers to it. Jet travel causes the solvent to evaporate, leaving the dry fiber on the collector [119,122,123,124,131,132].
A typical electrospinning device can be created using a power supply with high voltage power, a syringe, and a collector electrode alone, as demonstrated in the experimental part. Polymeric materials can be electro spun to create continuous nanofibers, and there are a number of variables that affect the nanofibers’ properties. These variables are either processing variables (electric field) or polymer properties (concentration, viscosity, surface tension, and conductivity) (extrusion rate: distance from needle tip to collector) [119,123,124,129,133].
The definition of the electrospinning technique is the use of a strong electric field to spin polymer solutions into micro- to nanofibers at room temperature and atmospheric pressure (kV). Electrospinning devices can be set up in either a vertical or horizontal orientation [119,124,134,135,136], (Figure 9).
The three primary components of an electrospinning device are the power supply (high voltage), the syringe (spinneret), and the collector (electrode) [119,123,137,138]. When a polymer solution is subjected to a high voltage, electric charges accumulate on its surface. These charges repel one another so strongly that they can overcome the surface tension of the solution and form a Taylor cone in a critical electric field. When the electric field stretches the Taylor cone tip further, a charged jet is ejected. The jet eventually transforms into solid fibers due to solvent evaporation [123,139,140,141].
Using a simple process, a stable jet at the spinneret needle is converted to an unstable jet to generate electro-spun fibers. When the applied electric field reaches a critical magnitude, the surface tension is overcome by the charge repulsion force, and jets erupt from the cone tip, transferring the tensile force to the polymer and causing it to bend. The charged fibers are then attracted by jets, with opposing charges moving from the cone apex to the collector. As the solvent evaporates through the jet, it leaves dry fiber on the collector [119,124,131,139,141].
A high-voltage power source, a syringe, and a collector electrode may be used to construct a standard electrospinning apparatus that can be configured vertically or horizontally, as illustrated in Figure 7. Continuous nanofibers may be generated by electrospinning polymeric materials, and the quality of electro-spun nanofibers is determined by some criteria. Polymer physical properties (concentration, viscosity, surface tension, and conductivity) or processing factors (electric field, flow rate, needle tip to collector distance) are examples [119,124,139,141].
Prior to electrospinning, most polymers are dissolved in a range of solvents; when fully dissolved, they create a polymer solution. The polymer solution is then poured into the capillary tube in preparation for electrospinning. However, because some polymers emit unpleasant or even dangerous odors, the procedures should be done in well-ventilated areas [142]. In the electrospinning process, a polymer solution at the capillary tube’s end is exposed to an electric field, which causes an electric charge to form on the liquid’s surface. The repelling electrical forces outweigh the surface tension forces when the applied electric field is strong enough. The solvent evaporates, and a polymer is formed when a charged jet of the solution is eventually released from the Taylor cone’s tip. The jet is unstable and whips quickly in the area between the capillary tip and collector. Just past the spinneret’s tip, where the jet is stable, instability sets in. Consequently, the electrospinning technology makes the process of making a fiber simpler [143].
Solution parameters, process parameters, and environmental or “ambient” parameters are the three main groups into which factors that have an impact on the electrospinning process are divided [119,144]. These factors are gathered in order to produce smooth fibers without beads, so a thorough understanding of these factors is required in order to obtain electro-spun nanofibers that are bead-free.

6.3. Application of Electro-Spun Nanofibers

Nanofiber membranes are created using electrospinning and used in a variety of applications, including biomedicine, security, clothing, and nano-sensors. Nanofibers are used in the biomedical field to focus on tissue engineering, wound dressing, drug delivery systems, and enzyme immobilization because they resemble the majority of organs and tissues, including skin, collagen, cartilage, and bone [145].
Electro-spun nanofibers are unique in that they have a consistent morphology, a high surface area to volume ratio, and inter- and inner porosity. These characteristics make them promising as scaffold biomaterials [146,147]. Additionally, nanofibers improve protein absorption, cell growth, cell differentiation, and cell adhesion [148,149].
Due to their pores, nanofibers are also used in filtration as micro and nano filters based on membrane design and construction, allowing liquids and small particles to pass while arresting larger particle sizes (contaminants), similar to how paper coffee filters prevent undissolved particles from passing through their pores while allowing dissolved ones to do so [150,151,152,153].
Affinity membranes, which have numerous uses in the biomedical and environmental fields, were also developed to select immobilized targets and remove contaminant targets [146,147,152,153,154]. Although electro-spun nanofibers have many uses, including those for tissue engineering, drug delivery, enzyme immobilization, wound dressing, antibacterial properties, filtration, desalination, and protective clothing, the focus of this article will be on biomedical uses.

6.3.1. Tissue Engineering Applications

Electro-spun nanofibers are used in tissue engineering scaffold construction [155]. The use of biodegradable and biocompatible nanofibers to provide target tissues has increased daily [156]. Due to the similarities between these fibers and the natural extracellular matrix, these fiber scaffolds had an impact on both cell-to-cell and cell-to-matrix communication and produced excellent growth factors [157].

6.3.2. Drug Delivery Applications

Based on the observation that the drug’s rate of dissolution increased as the surface area of both the drug and the carrier increased, nanofibers were used to coat the drug and deliver it to the target site [158,159]. Anticancer medications, antibiotics, proteins, ribonucleic acid (RNA), and deoxyribonucleic acid are all delivered by electro-spun nanofibers (DNA) [160].
Recently, Yang et al. created a composite scaffold using nanofibers made of gelatin and polyvinyl alcohol (PVA) to transport raspberry ketone [161,162]. Additionally, to transport the growth factor calcium hydroxyapatite, Haider et al. prepared PLGA nanofibers [146,147].

6.3.3. Enzymes Immobilization Applications

Immobilization of enzymes onto insoluble material is essential to improve durability and maintain the enzyme properties such as bioprocessing and long duration controls [163]. The immunized material essentials are biocompatible, durable and hydrophobic or hydrophilic [164].
Recently, electro-spun nanofiber prepared from the dual electrospinning process increased the enzyme immobilization [165]. Yet, there are some limitations that hinder this technique, and enzymes encapsulation and the enzyme immobilized on the fibers surface are limited [166], so that some chemical modification of the surface needs to overcome these limitations [167,168].

6.3.4. Wound Dressing Applications

Inhibiting microorganisms, removing exudate, and protecting the wound site are all important functions of wound dressing. In addition to having antimicrobial properties, wound dressings provide a pleasant, moist environment to speed up the healing process [169,170,171]. Consequently, electro-spun wound dressings have more benefits than those made using traditional techniques [172]. These benefits include fiber pores, a large surface area, and the ability to stimulate fibroblast cells, making it suitable for use in cosmetic masks for skin cleansing and healing [142,173]. The electro spun nanofiber matrix incorporates various skin-treatment components [174].
Due to its capacity to create cationic clusters that can bind with anions on red blood cells, chitosan nanofibers have amazing hemostatic capabilities that can speed up platelet and red blood cell aggregation and ultimately reduce blood loss [175]. Additionally, this technique is successful even in individuals with coagulation abnormalities and is independent of the patient’s own clotting system [176]. By electrospinning, Ren and colleagues [177] created a medicinal dressing made of a composite of silk fibroin, chitosan, and halloysite nanotubes. Aluminum silicate-based halloysite nanotubes have a hollow tubular structure, can efficiently bind antibacterial medications, and can provide delayed, sustained drug release [177]. The findings showed that using halloysite nanotubes caused the loaded-drug release time to increase by about 8 days. Additionally, the electro-spun chitosan composite membrane demonstrated a better blood coagulation rate, improved tensile property, and antibacterial activity, all of which support its potential value as a medical dressing.
An anti-fibrinolytic medication known as tranexamic acid (TXA) is frequently used during trauma surgery and has been found to improve wound healing [178]. For hemorrhage control applications, Sasmal and colleagues [178] created TXA-loaded chitosan/PVA electro-spun nanofibers. The findings support the role of chitosan in hemostasis by demonstrating that the total blood clotting time of pure chitosan/PVA nanofibrous membranes decreased from 21,010 s to 1676 s as the amount of chitosan increased. Additionally, clotting time and plasma recalcification time were dramatically shortened after TXA was added to chitosan nanofibers, demonstrating the enormous potential of TXA-loaded chitosan nanofibers for managing civil and military hemostasis. Additionally, by fabricating chitosan within a hydrogel carrier template produced from cyclodextrin through proton exchange and complexation, Leonhardt and colleagues [179] observed the development of nanoscale features in chitosan mats. With nanofiber diameters of 9.23.7 nm and a macroscopic shape resembling a honeycomb, the assembled chitosan was highly entangled. When compared to commercially available absorbable hemostatic dressings, the chitosan-based composite hydrogels result in much less blood loss and faster timeframes for hemostasis.

6.3.5. Antibacterial Applications

Many antibacterial hybrid electro-spun nanofiber scaffolds, including polyacrylonitrile/silver PAN and Ag nanofiber scaffolds, prepared by various research groups, have antibacterial activity against both gram-positive and gram-negative bacteria. Therefore, a wide variety of antimicrobial amidoxime was immobilized using PAN nanofibers. Amidoxime’s antibacterial action is caused by its binding to the Mg2+ and Ca2+ ions, which upset the balance of the bacteria and result in bacterial death [180].

6.4. Electrospinning of Chitosan

Template synthesis, drawing, phase separation, electrospinning, self-assembly, and other techniques have all been used to create nanofibers [181]. For the creation of micro- and nanofibers, electrospinning is one of these techniques that is particularly adaptable [182]. With their high surface area to volume ratio, oxygen permeable porosity, and variety of pore sizes, electro-spun nanofibers act as a wound dressing material by promoting fibroblast growth [183]. Chitosan has a low electro spinnability and needs a strong applied electric field to work properly because of its polycationic nature, which is brought on by the amino groups on its backbone [184,185]. A strong electric field is necessary because the polycationic nature of chitosan results in very viscous solutions with high surface tension. Additionally, the chitosan’s strong hydrogen bond network minimizes molecule exposure to the applied voltage [184,185].
Chitosan, like the majority of other polysaccharides, needs a strong acidic environment to dissolve properly, but this can be dangerous and prevent the use of chitosan fibers in some circumstances. Furthermore, when it comes to the creation of pure chitosan fibers, chitosan with a low molecular weight and low biological activity frequently produces better results. This has a negative impact on the biological functioning of the electro-spun chitosan [184,185,186]. As in all electrospinning techniques, changing parameters cause morphological changes in the spun fibers. The correlations between parameters and morphology are generally as follows: The diameter and length of the end fiber are decreased as the applied voltage is increased. Low chitosan concentrations can lead to fiber formation and breakdown, which increases fiber diameter and leads to morphological flaws [184,187].
Chitosan’s electro spinnability may be improved by using co-spinning polymers, either natural or synthetic. Frequently, chitosan is employed in the following fields: collagen, zein, silk fibroin, PEO, PVA, PLA, and zein [185,188]. Another method for resolving the issues with electrospinning chitosan is to chemically modify it to produce derivatives that are suitable for electrospinning, as with what was done with cellulose. More often than not, these derivatives have improved solubility and electro spinnability. Chitosan derivatives that have been investigated for this purpose include quarternized chitosan, hexanoyl chitosan, N-carboxyethyl chitosan, and others [189].
Chitosan is soluble in diluted aqueous formic, acetic, and lactic acids but insoluble in most mineral acidic media, alkalis, and water solutions. Chitosan dissolves when a small amount of acid is added to mixtures of water, ethanol, methanol, and acetone. Chitosan is a positively charged polyelectrolyte with a pH range of 2–6, which results in its greater solubility when compared to chitin. This characteristic makes chitosan solutions extremely viscous, making electrospinning difficult [190,191]. In addition, the three-dimensional network created by the potent hydrogen bonds prevents the polymeric chains from moving when they are exposed to an electrical field [192,193].
The unique properties of the polymer in solution, such as its polycationic nature, high molecular weight, and broad molecular weight dispersion, make chitosan electrospinning a challenging process. The inner tip–collector gap, the electric field voltage, the molecular weight, and the input velocity are just a few of the variables that have an impact on the quality of the electrospinning process and the finished product. When the electrostatic force in a solution is greater than the solution’s surface tension, the electrospinning process starts. When an electric field is applied, the surface of the polymer solution charges up. A strong electrical charge encourages jet extension and increases the volume of solution the needle can draw. On the other hand, a higher voltage leads to a longer stretch of the solution. This has a significant impact on electro-spun fiber morphology, frequently reducing their diameter, and raising the possibility of bead formation [190,194].
The feed rate, which regulates the number of solutions available, is another crucial component. A solution’s jet velocity and transfer rate are also impacted by its feed rate. For the evaporation of the solvent and the production of solid nanofibers, lower feed rates are preferred [119]. The solution must be removed from the tip at a rate that is substantially higher than the feed rate. Low feeding rates may prevent electrospinning, and high feeding rates may cause beaded large diameter fibers because sufficient solvent evaporation time must pass before the collector is reached [195].
SEM images of Figure 10 and Figure 11 show the effect of tip–collector distance and electrical field voltage, respectively, on the structure of electro-spun nanofibers. As shown in these figures, increasing gap distance and voltage not only decreases and refines nanofibers diameters, but also improves the quality of electro-spun nanofibers.
Another factor that affects the sizes and shapes of the nanofibers is the separation between the tip and the collector. In order to give the fibers enough time to dry before reaching the collector, a minimum distance is necessary; otherwise, beads have been observed when distances are either too close or too far apart [119]. This variable directly affects the strength of the electric field and the duration of the jet’s flight. The reduced tip–collector distance has a nearly identical effect on the circuit as increased voltage [196].
Additionally, rheological and electrical properties like viscosity, surface tension, conductivity, and dielectric strength are significantly influenced by molecular weight. High molecular weight nanofiber solutions produce fibers with a larger average diameter while too low a molecular weight solution tends to produce beads instead of fibers [197]. Polymer molecular weight, which is important to the electrospinning process, reveals how many polymer chains condense in a solution. It has been suggested that if there are sufficient interactions between molecules to compensate for chain entanglements, high molecular weights may not always be required for the electrospinning process. High molecular weight solutions typically result in the production of very long fibers. When a low molecular weight solution is employed, beads rather than fibers are produced [198].

6.5. Applications of Chitosan Nanofibers

6.5.1. Chitosan Nanofibers in Tissue Engineering

Tissue-engineered scaffolds should be biocompatible and capable of mimicking the extracellular matrix to create a microenvironment conducive to cell growth. Natural bone extracellular matrix, for example, has a complex microstructure of multi-layered collagen fibers and calcium deposits with fiber sizes in the nano meter range [199].
Nanofibrous scaffolds, as nanostructured scaffolds, have the distinct characteristics of large surface area, high porosity, and mechanical strength, resulting in extraordinary biological properties such as mimicking the nanoscale properties of the extracellular matrix, promoting cell adhesion and migration, transporting nutrients, and discharging waste [200].
Among these properties are the functional groups of nanofibrous polymers, which can coordinate with other components to promote cell adhesion, proliferation, differentiation, and, eventually, tissue regeneration [201]. Nanofibrous scaffolds may be a viable solution for the synthesis of extracellular matrix substitutes with the required biological functions [201].
Chitosans are polysaccharide polymers that are similar to extracellular matrix components, so they can be metabolized and their degradation products can be stored as proteoglycans in vivo [202]. Because of their excellent biological properties, chitosan nanofibers are popular in tissue engineering research [202].
Nanofibrous scaffolds based on chitosan are widely used in a variety of applications, including nerve, bone, and cartilage engineering, cardiac and vascular tissue engineering, tendon, ligament, and skeletal muscle regeneration, and wound healing (Table 4). Because of their chemical properties, chitin and chitosan can be dissolved in solvents such as acetic acid, trifluoroacetic acid, formic acid, succinic acid, and 1,1,1,3,3,3-Hexafluoroisopropanol (HFIP) [203].
Electrospinning, self-polymerization, and thermally induced phase separation can all be used to create chitin/chitosan nanofibrous scaffolds [204]. Furthermore, chitin/chitosan nanofibers derived from polymer blending can alter the biological and mechanical properties of composite scaffolds [205]. To meet the requirements of tissue engineering, these composite nanofibrous scaffolds can mimic the nanoscale structure and porosity of the extracellular matrix [206,207].
Table 4. Tissue engineering applications of chitosan nanofibers.
Table 4. Tissue engineering applications of chitosan nanofibers.
CompositionNanofiber
Diameter
CellsApplicationReference
Chitosan/montmorillonite/PVA nanofiber composite mesh60–140 nmHuman dental pulp stem cellsNerve tissue engineering[208]
PVA/chitosan nanofiber composite scaffolds94–410 nmPC12 nerve cellsNerve tissue engineering and
repair
[209]
Polycaprolactone/chitosan nanofibers composite110–240 nmSchwann cell (SC)Nerve tissue engineering[210]
Chitosan/PVA/graphene oxide nanofibers composite123–160 nmATDC5 cellsCartilage tissue engineering[211]
Chitosan/polyethylene oxide nanofiber composite140 ± 41 nmC2C12 myoblast cellsTendon tissue engineering[212]
Chitosan nanofiber scaffolds50–450 nmPrimary ventricular cardiomyocytesCardiac tissue engineering[213]
Polycaprolactone/chitosan nanofibers150 ± 2 nmMouse model and sheepVascular tissue engineering[214]
Chitosan/poly (vinyl alcohol) nanofibrous composite scaffold137 nmrMSCSkeletal muscle regeneration[215]
Chitosan nano-/micro fibrous double-layered composite20–300 nmBovine chondrocytesCartilage tissue engineering[216]

6.5.2. Chitosan Nanofibers in Enzyme Immobilization

Nanofibers, in contrast to other nanomaterials, have recently attracted attention for enzyme immobilization not only due to their higher surface-area-to-volume ratio and greater enzyme-loading capacity, but also due to their higher chemical and mechanical protection, which is crucial to ensure physical resistance for the support. The disadvantages of nanoparticles, such as the behavior of aggregation, which affects the stability and activity of enzymes, are not present in nanofibers. They also do not need an additional procedure (such as centrifugation or membranes) to separate nano capsules. The gathered nanofibers form a macrostructure of nanofibers that is simple to separate and repurpose [217,218,219,220]. Therefore, chitosan has been used as a support for enzyme immobilization because of its benefits [221,222,223,224]. In fact, it is important to note that chitosan-based nanofibers should be further investigated as a support for enzyme immobilization due to the characteristics of chitosan combined with the benefits of nanofibrous structures. The use of chitosan-based nanofibers in enzyme immobilization is summarized in Table 5.

6.5.3. Chitosan Nanofibers in Cancer Treatment

There are numerous anticancer medications, including Doxorubicin, Paclitaxel, Berberine, Methotrexate, Adriamycin, Curcumin, Vincristine, Mercaptopurine, Indomethacin, Ibuprofen, Ketoprofen, Dexamethasone, Tetracycline, Gemifloxacin, Tetanus Toxoid, and Folic acid. The quantity of these anticancer medications is loaded onto the carriers of fibers during in vitro and in vivo cancer treatment. The polycationic nature of chitosan makes it a good candidate in this field among other fiber carriers. Table 6 lists several anticancer drug-loaded chitosan fiber systems along with information on how they affected cancer tissues in vitro.

6.5.4. Chitosan Nanofibers in Food Technology

Chitosan’s polycationic nature, biodegradability, nontoxicity, antimicrobial, chelating, mucoadhesive, and gelling properties set it apart. All these properties, combined with the high surface area: volume ratio of nanofibrous structures, make chitosan-based nanofibers suitable for a wide range of applications in food technology that have yet to be fully explored. As a result, the following sessions highlight some plausible studies in this field.
Food waste poses a problem for the food industry. According to the Food and Agriculture Organization of the United Nations [249], one-third of the world’s food is wasted, causing economic and environmental problems. Indeed, food packaging has emerged as a viable option for improving the quality and safety of food products by increasing their biological, physical, and chemical stability. It is necessary to avoid chemical contaminants, oxygen, microorganisms, light, and moisture to achieve these properties. As a result, active packaging with antimicrobial properties has received a lot of attention [250,251,252].
Nanofibers made of chitosan have enormous potential for use in active food packaging. It has been widely reported that chitosan is effective against bacteria, viruses, and fungi [253,254,255,256,257]. In addition, chitosan, a renewable and biodegradable polymer, presents an intriguing real alternative to petroleum-based polymers in the development of green packaging materials.
Arkoun et al. [258] prepared chitosan/poly(ethylene oxide) (PEO) and tested its antibacterial activity against pathogenic microorganisms such as E. coli, Salmonella enterica serves Typhimurium, Staphylococcus aureus, and Listeria innocuous.
The nanofibers had an irreversible antibacterial effect, resulting in a bactericidal rather than bacteriostatic mechanism, according to the authors. Furthermore, bacterial growth was reduced at pH 5.8, which is lower than the pKa of amine groups on chitosan, and as a result, the authors proposed that the nanofibers could be applied to foods such as yoghurt, milk, cheese, meat, and fish, where lactic acid is liberated during the storage period.
Cui et al. [259] formed chitosan/poly(ethylene oxide) loaded with tea tree oil and tested it against Salmonella enteric subsp. enteric serovar Enteritidis and Salmonella typhimurium, two food pathogenic microorganisms. When the concentration of tea tree liposome was increased to 50%, the tensile strength increased by around 350%. Tea tree liposomes improved the antibacterial effect of the chitosan/poly (ethylene oxide) nanofiber as well. Furthermore, the chitosan/poly (ethylene oxide) loaded with liposomes tea tree demonstrated a four-day stable antibiofilm activity in chicken meat samples.
To prevent microbial spoilage of fish fillets, Ceylan et al. [260] created electro spun chitosan/thymol/liquid smoke nanofibers. The electro spun chitosan/thymol/liquid smoke nanofibers effectively reduced nearly 60% of total mesophilic bacteria. Furthermore, the authors stated that the nanofibers were thermostable until 150 °C, which is within the temperature range used in traditional food preservation methods.
The field of food packaging is a developing one that will most likely expand in the coming years. There are still numerous opportunities to investigate the role of chitosan-based nanofibers in other pathogenic microorganisms and different food types. Furthermore, the effect of other biopolymers and bioactive compounds on food packaging properties, such as mechanical properties and water vapor permeability, can be evaluated. Chitosan-based nanofibers could also be used to control food quality by monitoring external and internal conditions.
The majority of bioactive substances, volatile molecules, antioxidants, and flavors are unstable or even degradable [261].
Therefore, in addition to increasing bioavailability, chitosan-based nanofibers could improve the stability of functional food compounds. Chitosan’s mucoadhesive property can be used to deliver bioactive molecules to the body in a manner similar to how drugs are delivered; the mucoadhesion of the functional compounds increases their absorption through the gastrointestinal tract [262,263]. These qualities have increased interest in using chitosan-based nanofibers as a drug delivery vehicle in addition to their non-toxic and biocompatibility benefits [264,265]. Although some research has been done using various nanomaterials as food carriers [266,267], chitosan-based nanofibers have not yet been sufficiently investigated in this regard.
As a delivery vehicle for curcumin, Shekarforoush et al. [268] created electro spun chitosan/xanthan gum nanofiber. The authors discovered that pH 2.2 had a lower release of curcumin from nanofibers than other studied pH values at 6.5 and 7.4. According to one theory, there were stronger electrostatic interactions between chitosan and xanthan at pH levels below the pKa of the amine groups on chitosan, such as pH 2.2, which decreased the swelling behavior of the nanofiber and, as a result, the diffusion of curcumin. As a result, curcumin was able to be transported using the electro spun chitosan/xanthan gum, which improved its stability and bioavailability.
Biocatalysts called enzymes are essential in the food industry. However, due to their poor operational stability, short shelf life, and challenging recovery and reuse, their green chemistry and substrate specificity are compromised [269,270]. These issues might be solved by the enzyme immobilization on chitosan-based nanofibers, though. Due to their high pore-interconnectivity, which improves mass transfer between the substrate and the enzyme and, as a result, the enzyme activity, they stand out as promising supports.
Chitosan’s functional groups can be used to functionalize surfaces, which can then be used to adsorb controlled-release enzymes. Lactose is hydrolyzed using chitosan/poly (vinyl alcohol) nanofibers. According to the authors, at 50 °C, the immobilized enzyme was more thermostable than the free enzyme. Additionally, after 28 days of storage at 4 °C and 25 °C, the immobilized -d-galactosidase retained 77% and 42%, respectively, of its activity. As a result, the thermal and storage stability was increased by immobilizing -d-galactosidase in chitosan/poly (vinyl alcohol) nanofibers.
Nanofibers made of chitosan have been developed for the immobilization of enzymes.
Indeed, they offer promising resources for the creation of hypoallergenic foods as well as a number of dairy products (baking, jam, jellies, wine, beer and juices). The application of chitosan-based nanofibers on the immobilization of enzymes in food processing, including amylase, trypsin, pectinase, protease, tyrosinase, lipase, pectin lyase, pectin, laccase, among many others, should therefore receive significant attention.

7. Limitations and Future Perspectives for Chitosan Application

The use of chitosan has several restrictions in addition to its many benefits. Chitosan’s low solubility at a neutral pH is its most significant drawback. Numerous chemical and physical procedures have been employed to increase its solubility in order to get around this drawback. Chitosan has three types of functional groups: an amino group at C-2, a primary hydroxyl group at C-6, and a secondary hydroxyl group at C-3. Recent research has attempted to modify the three reactive functional groups in order to improve antimicrobial properties as well as solubility. For example, adding CH3 to chitosan increased its solubility and allowed it to be used in a wider pH range. The addition of disaccharides and N-alkylation increased its solubility and antimicrobial activity against E. coli across a wider pH range.
These findings collectively appear to suggest that chitosan can be modified in a variety of ways to increase its solubility and antimicrobial activity. Chitosan and its derivatives have received a lot of attention in recent decades due to their numerous applications in various fields. Several studies have shown that chitosan’s antimicrobial properties are affected by a variety of factors, including pH, temperature, Mw, metal chelation, and microorganism type. In vivo studies have also shown that chitosan and its derivatives can be used to treat microbial infections with no side effects. More research is needed, however, to determine the optimal chitosan conditions.
Chitosan’s nontoxic, biocompatible, biodegradable, and antimicrobial properties indicate that this compound and its derivatives have a wide range of applications, which have been discussed in this review. In the future, chitosan may be used as an alternative to synthetic bactericides for crops because it has already been shown to be effective against the treatment of bacterial infections in animals. In the food industry, it can be used in food packaging materials to extend the shelf-life of food products, as well as in dressings that treat wounds in the pharmaceutical industry.
However, more research is required to determine its mode of action. A more standardized and comprehensive description of procedures, on the other hand, is required to match the results of different investigators. More research is needed to understand the molecular events that underpin chitosan’s antimicrobial action. Finally, more research should be focused on improving chitosan’s antimicrobial activity while maintaining its low toxicity and biodegradability.

8. Conclusions

Chitosan-based polymers can be used in a variety of biological applications. Chitosan is a biomaterial that is both biocompatible and biodegradable. Chemistry increases the utility of chitosan. Chitosan and its derivatives are employed as nano-delivery systems in biotechnology. Electrospinning includes preparation, fiber configuration, material selection, intended applications, and the spinning method. Foundation polymers based on chitosan are anticipated to be used in biological applications more frequently.
The development of biodiversity materials with applications in biomedicine (tissue engineering, wound treatment, drug delivery), as well as environmental protection (air and water filters), is a promising field. Because of the difficulties, the most common method for electrospinning clean chitosan fibers is blending with a co-spinning agent, which has the advantages of easier electrospinning and complementary qualities for specific applications. The result is that chitosan has been combined with a wide range of synthetic polymers, such as poly (ethylene oxide), polyvinyl alcohol, poly (lactic acid), polycaprolactone, polyurethanes, polyamides, polyacrylates, polyethylene terephthalate, polyacrylonitrile, polyaniline, and natural polyproteins (collagen, gelatin, silk fibroin, sericin), as well as polyanionic polysaccharides (hybrid). The systematic analysis of the properties of these blend nanofibers revealed the advantages and disadvantages of each method, the main rules that should be followed when aiming for a specific morphology, and the impact of the co-spinning agents on the fiber properties, which further directs their use. All this foundational knowledge in chitosan electrospinning is useful for the design of materials for real-world applications, which appear to be focused on the fabrication of complex chitosan-based nanofiber blends or composites in order to meet the need for multifunctionality.

Author Contributions

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

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The authors extend their appreciation to the Deanship of scientific research Majmaah university, Saudi Arabia, for funding this research work through the Project No: R-2023-495.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brück, W.M.; Slater, J.W.; Carney, B.F. Chitin and chitosan from marine organisms. In Chitin, Chitosan, Oligosaccharides and Their Derivatives: Biological Activities and Applications; Taylor & Francis: Boca Raton, FL, USA, 2010; pp. 11–19. [Google Scholar]
  2. Zhou, A.; Wei, Y.; Wu, B.; Chen, Q.; Xing, D. Pyropheophorbide A and c (RGDyK) comodified chitosan-wrapped upconversion nanoparticle for targeted near-infrared photodynamic therapy. Mol. Pharm. 2012, 9, 1580–1589. [Google Scholar] [CrossRef] [PubMed]
  3. Vinsova, J.; Vavrikova, E. Chitosan derivatives with antimicrobial, antitumour and antioxidant activities—A review. Curr. Pharm. Des. 2011, 17, 3596–3607. [Google Scholar] [CrossRef] [PubMed]
  4. Philippova, O.; Korchagina, E. Chitosan and its hydrophobic derivatives: Preparation and aggregation in dilute aqueous solutions. Polym. Sci. Ser. A 2012, 54, 552–572. [Google Scholar] [CrossRef] [Green Version]
  5. Dutta, P.K.; Ravikumar, M.; Dutta, J. Chitin and chitosan for versatile applications. J. Macromol. Sci. Part C Polym. Rev. 2002, 42, 307–354. [Google Scholar] [CrossRef]
  6. Salama, R.; Osman, H.; Ibrahim, H.M. Preparation of biocompatible chitosan nanoparticles loaded with Aloe vera extract for use as a novel drug delivery mechanism to improve the antibacterial characteristics of cellulose-based fabrics. Egypt. J. Chem. 2022, 65, 581–595. [Google Scholar] [CrossRef]
  7. Mosaad, R.M.; Alhalafi, M.H.; Emam, E.A.M.; Ibrahim, M.A.; Ibrahim, H. Enhancement of Antimicrobial and Dyeing Properties of Cellulosic Fabrics via Chitosan Nanoparticles. Polymers 2022, 14, 4211. [Google Scholar] [CrossRef]
  8. Mohamed, F.A.; Sheier, M.B.; Reda, M.M.; Ibrahim, H.M. Synthesis, Application, and Antibacterial Activity of New Direct Dyes based on Chromene Derivatives. Curr. Org. Synth. 2022, 19, 757–766. [Google Scholar] [CrossRef]
  9. Ali, M.A.; Bydoon, E.A.; Ibrahim, H.M. Bioactive Composite Nonwoven Surgical Dressing based on Cellulose Coated with Nanofiber Membrane using the layer-by-layer technique. Egypt. J. Chem. 2022, 65, 525–542. [Google Scholar] [CrossRef]
  10. Madni, A.; Kousar, R.; Naeem, N.; Wahid, F. Recent advancements in applications of chitosan-based biomaterials for skin tissue engineering. J. Bioresour. Bioprod. 2021, 6, 11–25. [Google Scholar] [CrossRef]
  11. Hamed, I.; Özogul, F.; Regenstein, J.M. Industrial applications of crustacean by-products (chitin, chitosan, and chitooligosaccharides): A review. Trends Food Sci. Technol. 2016, 48, 40–50. [Google Scholar] [CrossRef]
  12. Wang, J.; Chen, C. Chitosan-based biosorbents: Modification and application for biosorption of heavy metals and radionuclides. Bioresour. Technol. 2014, 160, 129–141. [Google Scholar] [CrossRef]
  13. Zappino, M.; Cacciotti, I.; Benucci, I.; Nanni, F.; Liburdi, K.; Valentini, F.; Esti, M. Bromelain immobilization on microbial and animal source chitosan films, plasticized with glycerol, for application in wine-like medium: Microstructural, mechanical and catalytic characterisations. Food Hydrocoll. 2015, 45, 41–47. [Google Scholar] [CrossRef]
  14. Salmon, S.; Hudson, S.M. Crystal morphology, biosynthesis, and physical assembly of cellulose, chitin, and chitosan. J. Macromol. Sci. Part C Polym. Rev. 1997, 37, 199–276. [Google Scholar] [CrossRef]
  15. Aranaz, I.; Mengíbar, M.; Harris, R.; Paños, I.; Miralles, B.; Acosta, N.; Galed, G.; Heras, Á. Functional characterization of chitin and chitosan. Curr. Chem. Biol. 2009, 3, 203–230. [Google Scholar]
  16. Gupta, K.; Kumar, M.N.R. pH dependent hydrolysis and drug release behavior of chitosan/poly (ethylene glycol) polymer network microspheres. J. Mater. Sci. Mater. Med. 2001, 12, 753–759. [Google Scholar] [CrossRef] [PubMed]
  17. Jayakumar, R.; Menon, D.; Manzoor, K.; Nair, S.; Tamura, H. Biomedical applications of chitin and chitosan based nanomaterials—A short review. Carbohydr. Polym. 2010, 82, 227–232. [Google Scholar] [CrossRef]
  18. Wang, J.; Zhuang, S. Chitosan-based materials: Preparation, modification and application. J. Clean. Prod. 2022, 355, 131825. [Google Scholar] [CrossRef]
  19. Mourya, V.; Inamdar, N.N.; Tiwari, A. Carboxymethyl chitosan and its applications. Adv. Mat. Lett. 2010, 1, 11–33. [Google Scholar] [CrossRef]
  20. Peniche, C.; Argüelles-Monal, W.; Goycoolea, F. Chitin and chitosan: Major sources, properties and applications. Monomers Polym. Compos. Renew. Resour. 2008, 1, 517–542. [Google Scholar]
  21. Aranaz, I.; Harris, R.; Heras, A. Chitosan amphiphilic derivatives. Chemistry and applications. Curr. Org. Chem. 2010, 14, 308. [Google Scholar] [CrossRef]
  22. Wang, W.; Xue, C.; Mao, X. Chitosan: Structural modification, biological activity and application. Int. J. Biol. Macromol. 2020, 164, 4532–4546. [Google Scholar] [CrossRef]
  23. Ahmed, F.; Soliman, F.M.; Adly, M.A.; Soliman, H.A.M.; El-Matbouli, M.; Saleh, M. Recent progress in biomedical applications of chitosan and its nanocomposites in aquaculture: A review. Res. Vet. Sci. 2019, 126, 68–82. [Google Scholar] [CrossRef]
  24. Zhao, D.; Yu, S.; Sun, B.; Gao, S.; Guo, S.; Zhao, K. Biomedical Applications of Chitosan and Its Derivative Nanoparticles. Polymers 2018, 10, 462. [Google Scholar] [CrossRef] [Green Version]
  25. Ibrahim, H.M.; Klingner, A. A review on electrospun polymeric nanofibers: Production parameters and potential applications. Polym. Test. 2020, 90, 106647. [Google Scholar] [CrossRef]
  26. Qin, Y.; Li, P. Antimicrobial Chitosan Conjugates: Current Synthetic Strategies and Potential Applications. Int. J. Mol. Sci. 2020, 21, 499. [Google Scholar] [CrossRef] [Green Version]
  27. Crini, G. Recent developments in polysaccharide-based materials used as adsorbents in wastewater treatment. Prog. Polym. Sci. 2005, 30, 38–70. [Google Scholar] [CrossRef]
  28. Liu, H.; Du, Y.; Wang, X.; Sun, L. Chitosan kills bacteria through cell membrane damage. Int. J. Food Microbiol. 2004, 95, 147–155. [Google Scholar] [CrossRef] [PubMed]
  29. Kenawy, E.; Abdel-Hay, F.; Mohy Eldin, M.; Tamer, T.; Ibrahim, E. Novel Aminated Chitosan-Aromatic Aldehydes Schiff Bases: Synthesis, Characterization and Bio-evaluation. Int. J. Adv. Res 2015, 3, 563–572. [Google Scholar]
  30. Sashiwa, H.; Shigemasa, Y. Chemical modification of chitin and chitosan 2: Preparation and water soluble property of N-acylated or N-alkylated partially deacetylated chitins. Carbohydr. Polym. 1999, 39, 127–138. [Google Scholar] [CrossRef]
  31. Martău, G.A.; Mihai, M.; Vodnar, D.C. The use of chitosan, alginate, and pectin in the biomedical and food sector—biocompatibility, bioadhesiveness, and biodegradability. Polymers 2019, 11, 1837. [Google Scholar] [CrossRef] [Green Version]
  32. Di Martino, A.; Sedlarik, V. Amphiphilic chitosan-grafted-functionalized polylactic acid based nanoparticles as a delivery system for doxorubicin and temozolomide co-therapy. Int. J. Pharm. 2014, 474, 134–145. [Google Scholar] [CrossRef] [PubMed]
  33. Bosquez-Molina, E.; Zavaleta-Avejar, L. New Bioactive Biomaterials Based on Chitosan. In Chitosan in the Preservation of Agricultural Commodities; Academic Press: Cambridge, MA, USA, 2016; pp. 33–64. [Google Scholar]
  34. Furuike, T.; Komoto, D.; Hashimoto, H.; Tamura, H. Preparation of chitosan hydrogel and its solubility in organic acids. Int. J. Biol. Macromol. 2017, 104, 1620–1625. [Google Scholar] [CrossRef] [PubMed]
  35. Schügerl, K. Integrated processing of biotechnology products. Biotechnol. Adv. 2000, 18, 581–599. [Google Scholar] [CrossRef]
  36. Calvo, P.; Remunan-Lopez, C.; Vila-Jato, J.L.; Alonso, M.J. Novel hydrophilic chitosan-polyethylene oxide nanoparticles as protein carriers. J. Appl. Polym. Sci. 1997, 63, 125–132. [Google Scholar] [CrossRef]
  37. Sathiyabama, M.; Parthasarathy, R. Biological preparation of chitosan nanoparticles and its in vitro antifungal efficacy against some phytopathogenic fungi. Carbohydr. Polym. 2016, 151, 321–325. [Google Scholar] [CrossRef]
  38. Ramezani, Z.; Zarei, M.; Raminnejad, N. Comparing the effectiveness of chitosan and nanochitosan coatings on the quality of refrigerated silver carp fillets. Food Control 2015, 51, 43–48. [Google Scholar] [CrossRef]
  39. Janes, K.A.; Fresneau, M.P.; Marazuela, A.; Fabra, A.; Alonso, M.A.J. Chitosan nanoparticles as delivery systems for doxorubicin. J. Control. Release 2001, 73, 255–267. [Google Scholar] [CrossRef]
  40. De Campos, A.M.; Sánchez, A.; Alonso, M.a.J. Chitosan nanoparticles: A new vehicle for the improvement of the delivery of drugs to the ocular surface. Application to cyclosporin A. Int. J. Pharm. 2001, 224, 159–168. [Google Scholar] [CrossRef]
  41. Pan, Y.; Li, Y.-J.; Zhao, H.-Y.; Zheng, J.-M.; Xu, H.; Wei, G.; Hao, J.-S. Bioadhesive polysaccharide in protein delivery system: Chitosan nanoparticles improve the intestinal absorption of insulin in vivo. Int. J. Pharm. 2002, 249, 139–147. [Google Scholar] [CrossRef]
  42. Tian, X.X.; Groves, M.J. Formulation and biological activity of antineoplastic proteoglycans derived from Mycobacterium vaccae in chitosan nanoparticles. J. Pharm. Pharmacol. 1999, 51, 151–157. [Google Scholar] [CrossRef]
  43. Hassan, E.A.; Hassan, M.L.; Abou-Zeid, R.E.; El-Wakil, N.A. Novel nanofibrillated cellulose/chitosan nanoparticles nanocomposites films and their use for paper coating. Ind. Crops Prod. 2016, 93, 219–226. [Google Scholar] [CrossRef]
  44. Lee, Y.-M.; Park, Y.-J.; Lee, S.-J.; Ku, Y.; Han, S.-B.; Klokkevold, P.R.; Choi, S.-M.; Chung, C.-P. Tissue engineered bone formation using chitosan/tricalcium phosphate sponges. J. Periodontol. 2000, 71, 410–417. [Google Scholar] [CrossRef] [PubMed]
  45. Kim, Y.H.; Gihm, S.H.; Park, C.R.; Lee, K.Y.; Kim, T.W.; Kwon, I.C.; Chung, H.; Jeong, S.Y. Structural characteristics of size-controlled self-aggregates of deoxycholic acid-modified chitosan and their application as a DNA delivery carrier. Bioconjugate Chem. 2001, 12, 932–938. [Google Scholar] [CrossRef]
  46. Duan, B.; Dong, C.; Yuan, X.; Yao, K. Electrospinning of chitosan solutions in acetic acid with poly(ethylene oxide). J. Biomater. Sci. Polym. Ed. 2004, 15, 797–811. [Google Scholar] [CrossRef] [PubMed]
  47. Abou-Okeil, A.; Refaei, R.; Khalil, E.M.; Refaei, R.; Moustafa, S.E.; Ibrahim, H.M. Fabrication of novel green magnetic electrospun nanofibers based on Fe3O4 nanoparticles/PVA/chitosan/collagen. Egypt. J. Chem. 2022, 65, 283–299. [Google Scholar] [CrossRef]
  48. Bakr, M.M.; Taha, M.A.; Osman, H.; Ibrahim, H.M. Novel green printing of cotton, wool and polyester fabrics with natural safflower dye nanoparticles. Egypt. J. Chem. 2021, 64, 6221–6230. [Google Scholar] [CrossRef]
  49. Abou-Okeil, A.; Fahmy, H.M.; Fouda, M.M.G.; Aly, A.A.; Ibrahim, H.M. Hyaluronic Acid/Oxidized K-Carrageenan Electrospun Nanofibers Synthesis and Antibacterial Properties. BioNanoScience 2021, 11, 687–695. [Google Scholar] [CrossRef]
  50. Ibrahim, H.M.; Mostafa, M.; Kandile, N.G. Potential use of N-carboxyethylchitosan in biomedical applications: Preparation, characterization, biological properties. Int. J. Biol. Macromol. 2020, 149, 664–671. [Google Scholar] [CrossRef]
  51. Ibrahim, H.; El- Zairy, E.M.R.; Emam, E.A.M.; Adel, E. Combined antimicrobial finishing dyeing properties of cotton, polyester fabrics and their blends with acid and disperse dyes. Egypt. J. Chem. 2019, 62, 965–976. [Google Scholar] [CrossRef]
  52. Aranaz, I.; Gutiérrez, M.C.; Ferrer, M.L.; Del Monte, F. Preparation of chitosan nanocompositeswith a macroporous structure by unidirectional freezing and subsequent freeze-drying. Mar. Drugs 2014, 12, 5619–5642. [Google Scholar] [CrossRef] [Green Version]
  53. de Oliveira, J.L.; Campos, E.V.R.; Bakshi, M.; Abhilash, P.; Fraceto, L.F. Application of nanotechnology for the encapsulation of botanical insecticides for sustainable agriculture: Prospects and promises. Biotechnol. Adv. 2014, 32, 1550–1561. [Google Scholar] [CrossRef] [PubMed]
  54. Pichyangkura, R.; Chadchawan, S. Biostimulant activity of chitosan in horticulture. Sci. Hortic. 2015, 196, 49–65. [Google Scholar] [CrossRef]
  55. Shahrajabian, M.H.; Chaski, C.; Polyzos, N.; Tzortzakis, N.; Petropoulos, S.A. Sustainable Agriculture Systems in Vegetable Production Using Chitin and Chitosan as Plant Biostimulants. Biomolecules 2021, 11, 819. [Google Scholar] [CrossRef] [PubMed]
  56. Ghasemi Pirbalouti, A.; Malekpoor, F.; Salimi, A.; Golparvar, A. Exogenous application of chitosan on biochemical and physiological characteristics, phenolic content and antioxidant activity of two species of basil (Ocimum ciliatum and Ocimum basilicum) under reduced irrigation. Sci. Hortic. 2017, 217, 114–122. [Google Scholar] [CrossRef]
  57. Saharan, V.; Kumaraswamy, R.V.; Choudhary, R.C.; Kumari, S.; Pal, A.; Raliya, R.; Biswas, P. Cu-Chitosan Nanoparticle Mediated Sustainable Approach To Enhance Seedling Growth in Maize by Mobilizing Reserved Food. J. Agric. Food Chem. 2016, 64, 6148–6155. [Google Scholar] [CrossRef]
  58. Zhang, M.; Zhang, F.; Li, C.; An, H.; Wan, T.; Zhang, P. Application of Chitosan and Its Derivative Polymers in Clinical Medicine and Agriculture. Polymers 2022, 14, 958. [Google Scholar] [CrossRef]
  59. Vakili, M.; Rafatullah, M.; Salamatinia, B.; Abdullah, A.Z.; Ibrahim, M.H.; Tan, K.B.; Gholami, Z.; Amouzgar, P. Application of chitosan and its derivatives as adsorbents for dye removal from water and wastewater: A review. Carbohydr. Polym. 2014, 113, 115–130. [Google Scholar] [CrossRef]
  60. Lodhi, G.; Kim, Y.-S.; Hwang, J.-W.; Kim, S.-K.; Jeon, Y.-J.; Je, J.-Y.; Ahn, C.-B.; Moon, S.-H.; Jeon, B.-T.; Park, P.-J. Chitooligosaccharide and its derivatives: Preparation and biological applications. BioMed Res. Int. 2014, 2014, 654913. [Google Scholar] [CrossRef] [Green Version]
  61. Dwivedi, C.; Gupta, A.; Chaudhary, A.; Nandi, C.K. Gold nanoparticle chitosan composite hydrogel beads show efficient removal of methyl parathion from waste water. RSC Adv. 2014, 4, 39830–39838. [Google Scholar] [CrossRef]
  62. Yadla, S.V.; Sridevi, V.; Lakshmi, M.V.V.C. A review on adsorption of heavy metals from aqueous solution. J. Chem. Biol. Phys. Sci. 2012, 2, 1585. [Google Scholar]
  63. Chen, A.-H.; Yang, C.-Y.; Chen, C.-Y.; Chen, C.-Y.; Chen, C.-W. The chemically crosslinked metal-complexed chitosans for comparative adsorptions of Cu (II), Zn (II), Ni (II) and Pb (II) ions in aqueous medium. J. Hazard. Mater. 2009, 163, 1068–1075. [Google Scholar] [CrossRef]
  64. Younes, I.; Rinaudo, M. Chitin and chitosan preparation from marine sources. Structure, properties and applications. Mar. Drugs 2015, 13, 1133–1174. [Google Scholar] [CrossRef] [Green Version]
  65. Kyzas, G.Z.; Bikiaris, D.N. Recent modifications of chitosan for adsorption applications: A critical and systematic review. Mar. Drugs 2015, 13, 312–337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Cazón, P.; Vázquez, M. Applications of chitosan as food packaging materials. In Sustainable Agriculture Reviews 36; Springer: Berlin/Heidelberg, Germany, 2019; pp. 81–123. [Google Scholar]
  67. Falguera, V.; Ceron, J.P.Q.; Jiménez, A.; Muñoz, A. Películas y recubrimientos comestibles: Estructuras, funciones activas y tendencias en su uso. Tend. Cienc. Tecnol. Los Aliment. 2011, 22, 292–303. [Google Scholar]
  68. Aider, M. Chitosan application for active bio-based films production and potential in the food industry. LWT-Food Sci. Technol. 2010, 43, 837–842. [Google Scholar] [CrossRef]
  69. Muxika, A.; Etxabide, A.; Uranga, J.; Guerrero, P.; De La Caba, K. Chitosan as a bioactive polymer: Processing, properties and applications. Int. J. Biol. Macromol. 2017, 105, 1358–1368. [Google Scholar] [CrossRef] [PubMed]
  70. Chien, P.-J.; Sheu, F.; Lin, H.-R. Coating citrus (Murcott tangor) fruit with low molecular weight chitosan increases postharvest quality and shelf life. Food Chem. 2007, 100, 1160–1164. [Google Scholar] [CrossRef]
  71. Chien, P.-J.; Sheu, F.; Yang, F.-H. Effects of edible chitosan coating on quality and shelf life of sliced mango fruit. J. Food Eng. 2007, 78, 225–229. [Google Scholar] [CrossRef]
  72. Jiang, Y.; Li, Y. Effects of chitosan coating on postharvest life and quality of longan fruit. Food Chem. 2001, 73, 139–143. [Google Scholar] [CrossRef]
  73. Ferreira, L.F.; de Abreu, G.F.; Lago, A.M.T.; Figueiredo, L.P.; Borém, F.M.; Martins, M.A.; Borges, S.V.; Dias, M.V. Development and application of biopolymer coatings to specialty green coffee beans: Influence on water content, color and sensory quality. LWT 2018, 96, 274–280. [Google Scholar] [CrossRef]
  74. Soares, N.M.; Mendes, T.S.; Vicente, A.A. Effect of chitosan-based solutions applied as edible coatings and water glazing on frozen salmon preservation—A pilot-scale study. J. Food Eng. 2013, 119, 316–323. [Google Scholar] [CrossRef] [Green Version]
  75. Chien, P.J.; Chou, C.C. Antifungal activity of chitosan and its application to control post-harvest quality and fungal rotting of Tankan citrus fruit (Citrus tankan Hayata). J. Sci. Food Agric. 2006, 86, 1964–1969. [Google Scholar] [CrossRef]
  76. Choi, W.Y.; Park, H.J.; Ahn, D.J.; Lee, J.; Lee, C.Y. Wettability of chitosan coating solution on ‘Fuji’apple skin. J. Food Sci. 2002, 67, 2668–2672. [Google Scholar] [CrossRef]
  77. Shao, X.F.; Tu, K.; Tu, S.; Tu, J. A combination of heat treatment and chitosan coating delays ripening and reduces decay in “Gala” apple fruit. J. Food Qual. 2012, 35, 83–92. [Google Scholar] [CrossRef]
  78. Eissa, H.A.A. Effect of chitosan coating on shelf life and quality of fresh-cut mushroom. J. Food Qual. 2007, 30, 623–645. [Google Scholar] [CrossRef]
  79. Djioua, T.; Charles, F.; Freire, M., Jr.; Filgueiras, H.; Ducamp-Collin, M.N.; Sallanon, H. Combined effects of postharvest heat treatment and chitosan coating on quality of fresh-cut mangoes (Mangifera indica L.). Int. J. Food Sci. Technol. 2010, 45, 849–855. [Google Scholar] [CrossRef] [Green Version]
  80. Jiang, Y.; Li, J.; Jiang, W. Effects of chitosan coating on shelf life of cold-stored litchi fruit at ambient temperature. LWT-Food Sci. Technol. 2005, 38, 757–761. [Google Scholar] [CrossRef]
  81. Dong, C.; Chen, W.; Liu, C. Flocculation of algal cells by amphoteric chitosan-based flocculant. Bioresour. Technol. 2014, 170, 239–247. [Google Scholar] [CrossRef]
  82. Souza, C.P.; Almeida, B.C.; Colwell, R.R.; Rivera, I.N.G. The Importance of Chitin in the Marine Environment. Mar. Biotechnol. 2011, 13, 823–830. [Google Scholar] [CrossRef]
  83. Ahmed, K.B.M.; Khan, M.M.A.; Siddiqui, H.; Jahan, A. Chitosan and its oligosaccharides, a promising option for sustainable crop production—A review. Carbohydr. Polym. 2020, 227, 115331. [Google Scholar] [CrossRef]
  84. Pen, L.T.; Jiang, Y.M. Effects of chitosan coating on shelf life and quality of fresh-cut Chinese water chestnut. LWT-Food Sci. Technol. 2003, 36, 359–364. [Google Scholar] [CrossRef]
  85. Fan, M.; Hu, Q. Chitosan-LiOH-urea aqueous solution—A novel water-based system for chitosan processing. Carbohydr. Res. 2009, 344, 944–947. [Google Scholar] [CrossRef]
  86. Patrulea, V.; Ostafe, V.; Borchard, G.; Jordan, O. Chitosan as a starting material for wound healing applications. Eur. J. Pharm. Biopharm. 2015, 97, 417–426. [Google Scholar] [CrossRef] [Green Version]
  87. Farag, S.; Ibrahim, H.M.; Amr, A.; Asker, M.S.; El-Shafai, A. Preparation and characterization of ion exchanger based on bacterial cellulose for heavy metal cation removal. Egypt. J. Chem. 2019, 62, 457–466. [Google Scholar] [CrossRef]
  88. Mohamed, F.A.; Ibrahim, H.M.; Aly, A.A.; El-Alfy, E.A. Improvement of dyeability and antibacterial properties of gelatin treated cotton fabrics with synthesized reactive dye. Biosci. Res. 2018, 15, 4403–4408. [Google Scholar]
  89. Mohamed, F.A.; Abd El-Megied, S.A.; Bashandy, M.S.; Ibrahim, H.M. Synthesis, application and antibacterial activity of new reactive dyes based on thiazole moiety. Pigment Resin Technol. 2018, 47, 246–254. [Google Scholar] [CrossRef]
  90. Ibrahim, H.M.; Farid, O.A.; Samir, A.; Mosaad, R.M. Preparation of chitosan antioxidant nanoparticles as drug delivery system for enhancing of anti-cancer drug. In Key Engineering Materials; Trans Tech Publications Ltd.: Wollerau, Switzerland, 2018; Volume 759, pp. 92–97. [Google Scholar]
  91. Mosaad, R.M.; Samir, A.; Ibrahim, H.M. Median lethal dose (LD50) and cytotoxicity of Adriamycin in female albino mice. J. Appl. Pharm. Sci. 2017, 7, 77–80. [Google Scholar] [CrossRef] [Green Version]
  92. Liu, L.; Xin, Y.; Liu, J.; Zhang, E.; Li, W. Inhibitory effect of chitosan oligosaccharide on human hepatoma cells in vitro. Afr. J. Tradit. Complement. Altern. Med. 2017, 14, 272–277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Ngo, D.H.; Ngo, D.N.; Kim, S.-K.; Vo, T.S. Antiproliferative effect of aminoethyl-chitooligosaccharide on human lung A549 cancer cells. Biomolecules 2019, 9, 195. [Google Scholar] [CrossRef] [Green Version]
  94. Jiang, Z.; Wang, S.; Hou, J.; Chi, J.; Wang, S.; Shao, K.; Liu, W.; Sun, R.; Han, B. Effects of carboxymethyl chitosan oligosaccharide on regulating immunologic function and inhibiting tumor growth. Carbohydr. Polym. 2020, 250, 116994. [Google Scholar] [CrossRef]
  95. Zhai, X.; Li, C.; Ren, D.; Wang, J.; Ma, C.; Abd El-Aty, A.M. The impact of chitooligosaccharides and their derivatives on the in vitro and in vivo antitumor activity: A comprehensive review. Carbohydr. Polym. 2021, 266, 118132. [Google Scholar] [CrossRef]
  96. Muanprasat, C.; Chatsudthipong, V. Chitosan oligosaccharide: Biological activities and potential therapeutic applications. Pharmacol. Ther. 2017, 170, 80–97. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, C.-H.; Cherng, J.-H.; Liu, C.-C.; Fang, T.-J.; Hong, Z.-J.; Chang, S.-J.; Fan, G.-Y.; Hsu, S.-D. Procoagulant and antimicrobial effects of chitosan in wound healing. Int. J. Mol. Sci. 2021, 22, 7067. [Google Scholar] [CrossRef]
  98. Zhang, M.; Qiao, X.; Han, W.; Jiang, T.; Liu, F.; Zhao, X. Alginate-chitosan oligosaccharide-ZnO composite hydrogel for accelerating wound healing. Carbohydr. Polym. 2021, 266, 118100. [Google Scholar] [CrossRef] [PubMed]
  99. Shen, T.; Dai, K.; Yu, Y.; Wang, J.; Liu, C. Sulfated chitosan rescues dysfunctional macrophages and accelerates wound healing in diabetic mice. Acta Biomater. 2020, 117, 192–203. [Google Scholar] [CrossRef] [PubMed]
  100. Di Martino, A.; Sittinger, M.; Risbud, M.V. Chitosan: A versatile biopolymer for orthopaedic tissue-engineering. Biomaterials 2005, 26, 5983–5990. [Google Scholar] [CrossRef]
  101. Kaur, S.; Dhillon, G.S. The versatile biopolymer chitosan: Potential sources, evaluation of extraction methods and applications. Crit. Rev. Microbiol. 2014, 40, 155–175. [Google Scholar] [CrossRef]
  102. Wu, L.; Liu, M. Preparation and properties of chitosan-coated NPK compound fertilizer with controlled-release and water-retention. Carbohydr. Polym. 2008, 72, 240–247. [Google Scholar] [CrossRef]
  103. Malerba, M.; Cerana, R. Chitosan Effects on Plant Systems. Int. J. Mol. Sci. 2016, 17, 996. [Google Scholar] [CrossRef]
  104. Abd El-Naby, F.S.; Naiel, M.A.E.; Al-Sagheer, A.A.; Negm, S.S. Dietary chitosan nanoparticles enhance the growth, production performance, and immunity in Oreochromis niloticus. Aquaculture 2019, 501, 82–89. [Google Scholar] [CrossRef]
  105. Abd El-Naby, A.S.; Al-Sagheer, A.A.; Negm, S.S.; Naiel, M.A.E. Dietary combination of chitosan nanoparticle and thymol affects feed utilization, digestive enzymes, antioxidant status, and intestinal morphology of Oreochromis niloticus. Aquaculture 2020, 515, 734577. [Google Scholar] [CrossRef]
  106. Sinha, V.R.; Singla, A.K.; Wadhawan, S.; Kaushik, R.; Kumria, R.; Bansal, K.; Dhawan, S. Chitosan microspheres as a potential carrier for drugs. Int. J. Pharm. 2004, 274, 1–33. [Google Scholar] [CrossRef]
  107. Azad, A.K.; Sermsintham, N.; Chandrkrachang, S.; Stevens, W.F. Chitosan Membrane as a Wound-Healing Dressing: Characterization and Clinical Application. J. Biomed. Mater. Res. Part B Appl. Biomater. 2004, 69, 216–222. [Google Scholar] [CrossRef] [PubMed]
  108. 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]
  109. Fan, L.; Du, Y.; Zhang, B.; Yang, J.; Zhou, J.; Kennedy, J.F. Preparation and properties of alginate/carboxymethyl chitosan blend fibers. Carbohydr. Polym. 2006, 65, 447–452. [Google Scholar] [CrossRef]
  110. Chen, J.-P.; Chang, G.-Y.; Chen, J.-K. Electrospun collagen/chitosan nanofibrous membrane as wound dressing. Colloids Surf. A Physicochem. Eng. Asp. 2008, 313–314, 183–188. [Google Scholar] [CrossRef]
  111. Qasim, S.B.; Najeeb, S.; Delaine-Smith, R.M.; Rawlinson, A.; Ur Rehman, I. Potential of electrospun chitosan fibers as a surface layer in functionally graded GTR membrane for periodontal regeneration. Dent. Mater. 2017, 33, 71–83. [Google Scholar] [CrossRef] [Green Version]
  112. Ma, Y.; Xin, L.; Tan, H.; Fan, M.; Li, J.; Jia, Y.; Ling, Z.; Chen, Y.; Hu, X. Chitosan membrane dressings toughened by glycerol to load antibacterial drugs for wound healing. Mater. Sci. Eng. C 2017, 81, 522–531. [Google Scholar] [CrossRef]
  113. Behera, S.S.; Das, U.; Kumar, A.; Bissoyi, A.; Singh, A.K. Chitosan/TiO2 composite membrane improves proliferation and survival of L929 fibroblast cells: Application in wound dressing and skin regeneration. Int. J. Biol. Macromol. 2017, 98, 329–340. [Google Scholar] [CrossRef] [PubMed]
  114. Ghannam, S.F.; Korayem, H.E.; Farghaly, L.M.; Hosny, S. The effect of chitosan nanosilver dressing versus mesenchymal stem cells on wound healing. J. Afr. Assoc. Physiol. Sci. 2018, 6, 23–31. [Google Scholar]
  115. Denkbaş, E.B.; Öztürk, E.; Özdemir, N.; Keçeci, K.; Agalar, C. Norfloxacin-loaded chitosan sponges as wound dressing material. J. Biomater. Appl. 2004, 18, 291–303. [Google Scholar] [CrossRef] [PubMed]
  116. Deng, C.-M.; He, L.-Z.; Zhao, M.; Yang, D.; Liu, Y. Biological properties of the chitosan-gelatin sponge wound dressing. Carbohydr. Polym. 2007, 69, 583–589. [Google Scholar] [CrossRef]
  117. Obara, K.; Ishihara, M.; Ishizuka, T.; Fujita, M.; Ozeki, Y.; Maehara, T.; Saito, Y.; Yura, H.; Matsui, T.; Hattori, H.; et al. Photocrosslinkable chitosan hydrogel containing fibroblast growth factor-2 stimulates wound healing in healing-impaired db/db mice. Biomaterials 2003, 24, 3437–3444. [Google Scholar] [CrossRef] [PubMed]
  118. Chen, H.; Xing, X.; Tan, H.; Jia, Y.; Zhou, T.; Chen, Y.; Ling, Z.; Hu, X. Covalently antibacterial alginate-chitosan hydrogel dressing integrated gelatin microspheres containing tetracycline hydrochloride for wound healing. Mater. Sci. Eng. C 2017, 70, 287–295. [Google Scholar] [CrossRef] [PubMed]
  119. Bhardwaj, N.; Kundu, S.C. Electrospinning: A fascinating fiber fabrication technique. Biotechnol. Adv. 2010, 28, 325–347. [Google Scholar] [CrossRef]
  120. Teo, W.E.; Ramakrishna, S. A review on electrospinning design and nanofibre assemblies. Nanotechnology 2006, 17, R89. [Google Scholar] [CrossRef]
  121. Celebioglu, A.; Uyar, T. Electrospinning of nanofibers from non-polymeric systems: Electrospun nanofibers from native cyclodextrins. J. Colloid Interface Sci. 2013, 404, 1–7. [Google Scholar] [CrossRef]
  122. Subbiah, T.; Bhat, G.; Tock, R.; Parameswaran, S.; Ramkumar, S. Electrospinning of nanofibers. J. Appl. Polym. Sci. 2005, 96, 557–569. [Google Scholar] [CrossRef]
  123. Ghorani, B.; Tucker, N. Fundamentals of electrospinning as a novel delivery vehicle for bioactive compounds in food nanotechnology. Food Hydrocoll. 2015, 51, 227–240. [Google Scholar] [CrossRef]
  124. Pillay, V.; Dott, C.; Choonara, Y.E.; Tyagi, C.; Tomar, L.; Kumar, P.; du Toit, L.C.; Ndesendo, V.M. A review of the effect of processing variables on the fabrication of electrospun nanofibers for drug delivery applications. J. Nanomater. 2013, 2013, 789289. [Google Scholar] [CrossRef] [Green Version]
  125. Dalton, P.D.; Klinkhammer, K.; Salber, J.; Klee, D.; Möller, M. Direct in vitro electrospinning with polymer melts. Biomacromolecules 2006, 7, 686–690. [Google Scholar] [CrossRef]
  126. Greiner, A.; Wendorff, J.H. Electrospinning: A fascinating method for the preparation of ultrathin fibers. Angew. Chem. Int. Ed. 2007, 46, 5670–5703. [Google Scholar] [CrossRef]
  127. Agarwal, S.; Greiner, A.; Wendorff, J.H. Functional materials by electrospinning of polymers. Prog. Polym. Sci. 2013, 38, 963–991. [Google Scholar] [CrossRef]
  128. Haider, A.; Haider, S.; Kang, I.-K. A comprehensive review summarizing the effect of electrospinning parameters and potential applications of nanofibers in biomedical and biotechnology. Arab. J. Chem. 2018, 11, 1165–1188. [Google Scholar] [CrossRef]
  129. Liang, D.; Hsiao, B.S.; Chu, B. Functional electrospun nanofibrous scaffolds for biomedical applications. Adv. Drug Deliv. Rev. 2007, 59, 1392–1412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Adomavičiūtė, E.; Milašius, R. The influence of applied voltage on poly (vinyl alcohol)(PVA) nanofibre diameter. Fibres Text. East. Eur. 2007, 15, 63. [Google Scholar]
  131. Reneker, D.; Yarin, A.; Zussman, E.; Xu, H. Electrospinning of nanofibers from polymer solutions and melts. Adv. Appl. Mech. 2007, 41, 43–346. [Google Scholar]
  132. Shin, Y.; Hohman, M.; Brenner, M.; Rutledge, G. Electrospinning: A whipping fluid jet generates submicron polymer fibers. Appl. Phys. Lett. 2001, 78, 1149–1151. [Google Scholar] [CrossRef] [Green Version]
  133. Ramakrishna, S.; Fujihara, K.; Teo, W.-E.; Lim, T.-C.; Ma, Z. An Introduction to Electrospinning and Nanofibers; World Scientific: Singapore, 2005; Volume 90. [Google Scholar]
  134. Kidoaki, S.; Kwon, I.K.; Matsuda, T. Mesoscopic spatial designs of nano-and microfiber meshes for tissue-engineering matrix and scaffold based on newly devised multilayering and mixing electrospinning techniques. Biomaterials 2005, 26, 37–46. [Google Scholar] [CrossRef]
  135. Stankus, J.J.; Freytes, D.O.; Badylak, S.F.; Wagner, W.R. Hybrid nanofibrous scaffolds from electrospinning of a synthetic biodegradable elastomer and urinary bladder matrix. J. Biomater. Sci. Polym. Ed. 2008, 19, 635–652. [Google Scholar] [CrossRef] [Green Version]
  136. El-Alfy, E.A.; El-Bisi, M.K.; Taha, G.M.; Ibrahim, H.M. Preparation of biocompatible chitosan nanoparticles loaded by tetracycline, gentamycin and ciprofloxacin as novel drug delivery system for improvement the antibacterial properties of cellulose based fabrics. Int. J. Biol. Macromol. 2020, 161, 1247–1260. [Google Scholar] [CrossRef]
  137. Xue, J.; Wu, T.; Dai, Y.; Xia, Y. Electrospinning and electrospun nanofibers: Methods, materials, and applications. Chem. Rev. 2019, 119, 5298–5415. [Google Scholar] [CrossRef]
  138. Liao, Y.; Loh, C.-H.; Tian, M.; Wang, R.; Fane, A.G. Progress in electrospun polymeric nanofibrous membranes for water treatment: Fabrication, modification and applications. Prog. Polym. Sci. 2018, 77, 69–94. [Google Scholar] [CrossRef]
  139. Senthil, T.; George, G.; Srinivasan, A. Electrospinning: From Fundamentals to Applications. In Advances in Polymer Materials and Technology; CRC Press: Boca Raton, FL, USA, 2016; pp. 171–240. [Google Scholar]
  140. Zagho, M.M.; Elzatahry, A. Recent trends in electrospinning of polymer nanofibers and their applications as templates for metal oxide nanofibers preparation. In Electrospinning: Material, Techniques, and Biomedical Applications; IntechOpen: London, UK, 2016. [Google Scholar]
  141. Bode-Aluko, C.A.; Laatikainen, K.; Pereao, O.; Nechaev, A.; Kochnev, I.; Rossouw, A.; Dobretsov, S.; Branger, C.; Sarbu, A.; Petrik, L. Fabrication and characterisation of novel nanofiltration polymeric membrane. Mater. Today Commun. 2019, 20, 100580. [Google Scholar] [CrossRef]
  142. Huang, Z.-M.; Zhang, Y.-Z.; Kotaki, M.; Ramakrishna, S. A review on polymer nanofibers by electrospinning and their applications in nanocomposites. Compos. Sci. Technol. 2003, 63, 2223–2253. [Google Scholar] [CrossRef]
  143. Mohamed, F.A.; Reda, M.M.; Ibrahim, H. Enhancement of Dyeing and Antimicrobial Properties of Chitosan and Chitosan Nanoparticles-Treated Cotton and Viscose Fabrics with Acid Dyes. Egypt. J. Chem. 2022, 65, 339–344. [Google Scholar]
  144. Teo, W.-E.; Inai, R.; Ramakrishna, S. Technological advances in electrospinning of nanofibers. Sci. Technol. Adv. Mater. 2016, 12, 013002. [Google Scholar] [CrossRef] [Green Version]
  145. Metreveli, G.; Wågberg, L.; Emmoth, E.; Belák, S.; Strømme, M.; Mihranyan, A. A Size-Exclusion Nanocellulose Filter Paper for Virus Removal. Adv. Healthc. Mater. 2014, 3, 1546–1550. [Google Scholar] [CrossRef]
  146. Haider, A.; Gupta, K.C.; Kang, I.-K. Morphological effects of HA on the cell compatibility of electrospun HA/PLGA composite nanofiber scaffolds. BioMed Res. Int. 2014, 2014, 308306. [Google Scholar] [CrossRef] [Green Version]
  147. Haider, A.; Gupta, K.C.; Kang, I.-K. PLGA/nHA hybrid nanofiber scaffold as a nanocargo carrier of insulin for accelerating bone tissue regeneration. Nanoscale Res. Lett. 2014, 9, 314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Tysseling-Mattiace, V.M.; Sahni, V.; Niece, K.L.; Birch, D.; Czeisler, C.; Fehlings, M.G.; Stupp, S.I.; Kessler, J.A. Self-assembling nanofibers inhibit glial scar formation and promote axon elongation after spinal cord injury. J. Neurosci. 2008, 28, 3814–3823. [Google Scholar] [CrossRef] [Green Version]
  149. Woo, K.M.; Chen, V.J.; Ma, P.X. Nano-fibrous scaffolding architecture selectively enhances protein adsorption contributing to cell attachment. J. Biomed. Mater. Res. Part A 2003, 67, 531–537. [Google Scholar] [CrossRef] [Green Version]
  150. Gupta, K.C.; Haider, A.; Choi, Y.-R.; Kang, I.-K. Nanofibrous scaffolds in biomedical applications. Biomater. Res. 2014, 18, 5. [Google Scholar] [CrossRef] [Green Version]
  151. Haider, S.; Binagag, F.F.; Haider, A.; Mahmood, A.; Shah, N.; Al-Masry, W.A.; Khan, S.U.-D.; Ramay, S.M. Adsorption kinetic and isotherm of methylene blue, safranin T and rhodamine B onto electrospun ethylenediamine-grafted-polyacrylonitrile nanofibers membrane. Desalination Water Treat. 2015, 55, 1609–1619. [Google Scholar] [CrossRef]
  152. Haider, S.; Binagag, F.F.; Haider, A.; Mahmood, A.; Al Masry, W.A.; Alhoshan, M.; Khan, S.U.-D. Fabrication of the Diethylenetriamine Grafted Polyacrylonitrile Electrospun Nanofibers Membrane for the Aqueous Removal of Cationic Dyes. Sci. Adv. Mater. 2015, 7, 309–318. [Google Scholar] [CrossRef]
  153. Haider, S.; Haider, A.; Ahmad, A.; Khan, S.U.-D.; Almasry, W.A.; Sarfarz, M. Electrospun nanofibers affinity membranes for water hazards remediation. Nanotechnol. Res. J. 2015, 8, 511. [Google Scholar]
  154. Haider, S.; Al-Zeghayer, Y.; Ali, F.A.A.; Haider, A.; Mahmood, A.; Al-Masry, W.A.; Imran, M.; Aijaz, M.O. Highly aligned narrow diameter chitosan electrospun nanofibers. J. Polym. Res. 2013, 20, 105. [Google Scholar] [CrossRef]
  155. Vasita, R.; Katti, D.S. Nanofibers and their applications in tissue engineering. Int. J. Nanomed. 2006, 1, 15. [Google Scholar] [CrossRef]
  156. Sun, B.; Long, Y.; Zhang, H.; Li, M.; Duvail, J.; Jiang, X.; Yin, H. Advances in three-dimensional nanofibrous macrostructures via electrospinning. Prog. Polym. Sci. 2014, 39, 862–890. [Google Scholar] [CrossRef]
  157. Li, W.J.; Laurencin, C.T.; Caterson, E.J.; Tuan, R.S.; Ko, F.K. Electrospun nanofibrous structure: A novel scaffold for tissue engineering. J. Biomed. Mater. Res. 2002, 60, 613–621. [Google Scholar] [CrossRef]
  158. Kenawy, E.-R.; Bowlin, G.L.; Mansfield, K.; Layman, J.; Simpson, D.G.; Sanders, E.H.; Wnek, G.E. Release of tetracycline hydrochloride from electrospun poly (ethylene-co-vinylacetate), poly (lactic acid), and a blend. J. Control. Release 2002, 81, 57–64. [Google Scholar] [CrossRef] [PubMed]
  159. Liu, W.; Thomopoulos, S.; Xia, Y. Electrospun nanofibers for regenerative medicine. Adv. Healthc. Mater. 2012, 1, 10–25. [Google Scholar] [CrossRef] [PubMed]
  160. Hu, X.; Liu, S.; Zhou, G.; Huang, Y.; Xie, Z.; Jing, X. Electrospinning of polymeric nanofibers for drug delivery applications. J. Control. Release 2014, 185, 12–21. [Google Scholar] [CrossRef] [PubMed]
  161. Kanani, A.G.; Bahrami, S.H. Effect of changing solvents on poly (ε-caprolactone) nanofibrous webs morphology. J. Nanomater. 2011, 2011, 31. [Google Scholar]
  162. Yang, D.; Li, Y.; Nie, J. Preparation of gelatin/PVA nanofibers and their potential application in controlled release of drugs. Carbohydr. Polym. 2007, 69, 538–543. [Google Scholar] [CrossRef]
  163. Xie, J.; Hsieh, Y.-L. Ultra-high surface fibrous membranes from electrospinning of natural proteins: Casein and lipase enzyme. J. Mater. Sci. 2003, 38, 2125–2133. [Google Scholar] [CrossRef]
  164. Ye, P.; Xu, Z.-K.; Che, A.-F.; Wu, J.; Seta, P. Chitosan-tethered poly (acrylonitrile-co-maleic acid) hollow fiber membrane for lipase immobilization. Biomaterials 2005, 26, 6394–6403. [Google Scholar] [CrossRef]
  165. Huang, X.J.; Xu, Z.K.; Wan, L.S.; Innocent, C.; Seta, P. Electrospun nanofibers modified with phospholipid moieties for enzyme immobilization. Macromol. Rapid Commun. 2006, 27, 1341–1345. [Google Scholar] [CrossRef]
  166. Kim, H.S.; Yoo, H.S. MMPs-responsive release of DNA from electrospun nanofibrous matrix for local gene therapy: In vitro and in vivo evaluation. J. Control. Release 2010, 145, 264–271. [Google Scholar] [CrossRef]
  167. Im, J.S.; Yun, J.; Lim, Y.-M.; Kim, H.-I.; Lee, Y.-S. Fluorination of electrospun hydrogel fibers for a controlled release drug delivery system. Acta Biomater. 2010, 6, 102–109. [Google Scholar] [CrossRef]
  168. Theron, S.; Yarin, A.; Zussman, E.; Kroll, E. Multiple jets in electrospinning: Experiment and modeling. Polymer 2005, 46, 2889–2899. [Google Scholar] [CrossRef]
  169. Gallant-Behm, C.L.; Yin, H.Q.; Liu, S.; Heggers, J.P.; Langford, R.E.; Olson, M.E.; Hart, D.A.; Burrell, R.E. Comparison of in vitro disc diffusion and time kill-kinetic assays for the evaluation of antimicrobial wound dressing efficacy. Wound Repair Regen. 2005, 13, 412–421. [Google Scholar] [CrossRef] [PubMed]
  170. Jones, S.A.; Bowler, P.G.; Walker, M.; Parsons, D. Controlling wound bioburden with a novel silver-containing Hydrofiber® dressing. Wound Repair Regen. 2004, 12, 288–294. [Google Scholar] [CrossRef]
  171. Rim, N.G.; Shin, C.S.; Shin, H. Current approaches to electrospun nanofibers for tissue engineering. Biomed. Mater. 2013, 8, 014102. [Google Scholar] [CrossRef]
  172. Gao, Y.; Bach Truong, Y.; Zhu, Y.; Louis Kyratzis, I. Electrospun antibacterial nanofibers: Production, activity, and in vivo applications. J. Appl. Polym. Sci. 2014, 131. [Google Scholar] [CrossRef]
  173. Smith, D.J.; Reneker, D.H.; McManus, A.T.; Schreuder-Gibson, H.L.; Mello, C.; Sennett, M.S. Electrospun Fibers and an Apparatus Therefor. U.S. Patent 6,753,454, 22 June 2004. [Google Scholar]
  174. Si, Y.; Tang, X.; Yu, J.; Ding, B. Electrospun Nanofibers: Solving Global Issues. In Electrospun Nanofibers for Energy and Environmental Applications; Springer: Berlin/Heidelberg, Germany, 2014; pp. 3–38. [Google Scholar]
  175. Nakielski, P.; Pierini, F. Blood interactions with nano- and microfibers: Recent advances, challenges and applications in nano- and microfibrous hemostatic agents. Acta Biomater. 2019, 84, 63–76. [Google Scholar] [CrossRef]
  176. Huang, J.; Cheng, Y.; Wu, Y.; Shi, X.; Du, Y.; Deng, H. Chitosan/tannic acid bilayers layer-by-layer deposited cellulose nanofibrous mats for antibacterial application. Int. J. Biol. Macromol. 2019, 139, 191–198. [Google Scholar] [CrossRef]
  177. Ren, X.; Xu, Z.; Wang, L.; Meng, K.; Wang, H.; Zhao, H. Silk fibroin/chitosan/halloysite composite medical dressing with antibacterial and rapid haemostatic properties. Mater. Res. Express 2019, 6, 125409. [Google Scholar] [CrossRef]
  178. Sasmal, P.; Datta, P. Tranexamic acid-loaded chitosan electrospun nanofibers as drug delivery system for hemorrhage control applications. J. Drug Deliv. Sci. Technol. 2019, 52, 559–567. [Google Scholar] [CrossRef]
  179. Leonhardt, E.E.; Kang, N.; Hamad, M.A.; Wooley, K.L.; Elsabahy, M. Absorbable hemostatic hydrogels comprising composites of sacrificial templates and honeycomb-like nanofibrous mats of chitosan. Nat. Commun. 2019, 10, 2307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Zhang, L.; Luo, J.; Menkhaus, T.J.; Varadaraju, H.; Sun, Y.; Fong, H. Antimicrobial nano-fibrous membranes developed from electrospun polyacrylonitrile nanofibers. J. Membr. Sci. 2011, 369, 499–505. [Google Scholar] [CrossRef]
  181. Subramanian, A.; Krishnan, U.M.; Sethuraman, S. Fabrication of uniaxially aligned 3D electrospun scaffolds for neural regeneration. Biomed. Mater. 2011, 6, 025004. [Google Scholar] [CrossRef] [PubMed]
  182. Jayakumar, R.; Prabaharan, M.; Sudheesh Kumar, P.T.; Nair, S.V.; Tamura, H. Biomaterials based on chitin and chitosan in wound dressing applications. Biotechnol. Adv. 2011, 29, 322–337. [Google Scholar] [CrossRef] [PubMed]
  183. Reneker, D.H.; Yarin, A.L. Electrospinning jets and polymer nanofibers. Polymer 2008, 49, 2387–2425. [Google Scholar] [CrossRef] [Green Version]
  184. Augustine, R.; Rehman, S.R.U.; Ahmed, R.; Zahid, A.A.; Sharifi, M.; Falahati, M.; Hasan, A. Electrospun chitosan membranes containing bioactive and therapeutic agents for enhanced wound healing. Int. J. Biol. Macromol. 2020, 156, 153–170. [Google Scholar] [CrossRef]
  185. Ranjith, R.; Balraj, S.; Ganesh, J.; Milton, M.C.J. Therapeutic agents loaded chitosan-based nanofibrous mats as potential wound dressings: A review. Mater. Today Chem. 2019, 12, 386–395. [Google Scholar] [CrossRef]
  186. Yang, X.; Wang, J.; Guo, H.; Liu, L.; Xu, W.; Duan, G. Structural design toward functional materials by electrospinning: A review. e-Polymers 2020, 20, 682–712. [Google Scholar] [CrossRef]
  187. Guo, H.; Chen, Y.; Li, Y.; Zhou, W.; Xu, W.; Pang, L.; Fan, X.; Jiang, S. Electrospun fibrous materials and their applications for electromagnetic interference shielding: A review. Compos. Part A Appl. Sci. Manuf. 2021, 143, 106309. [Google Scholar] [CrossRef]
  188. Zhao, L.; Duan, G.; Zhang, G.; Yang, H.; He, S.; Jiang, S. Electrospun Functional Materials toward Food Packaging Applications: A Review. Nanomaterials 2020, 10, 150. [Google Scholar] [CrossRef] [Green Version]
  189. Tao, F.; Cheng, Y.; Shi, X.; Zheng, H.; Du, Y.; Xiang, W.; Deng, H. Applications of chitin and chitosan nanofibers in bone regenerative engineering. Carbohydr. Polym. 2020, 230, 115658. [Google Scholar] [CrossRef] [PubMed]
  190. Homayoni, H.; Ravandi, S.A.H.; Valizadeh, M. Electrospinning of chitosan nanofibers: Processing optimization. Carbohydr. Polym. 2009, 77, 656–661. [Google Scholar] [CrossRef]
  191. Liu, L.; Xu, W.; Ding, Y.; Agarwal, S.; Greiner, A.; Duan, G. A review of smart electrospun fibers toward textiles. Compos. Commun. 2020, 22, 100506. [Google Scholar] [CrossRef]
  192. Neamnark, A.; Rujiravanit, R.; Supaphol, P. Electrospinning of hexanoyl chitosan. Carbohydr. Polym. 2006, 66, 298–305. [Google Scholar] [CrossRef]
  193. Li, S.; Guo, H.; He, S.; Yang, H.; Liu, K.; Duan, G.; Jiang, S. Advanced electrospun nanofibers as bifunctional electrocatalysts for flexible metal-air (O2) batteries: Opportunities and challenges. Mater. Des. 2022, 214, 110406. [Google Scholar] [CrossRef]
  194. Geng, X.; Kwon, O.-H.; Jang, J. Electrospinning of chitosan dissolved in concentrated acetic acid solution. Biomaterials 2005, 26, 5427–5432. [Google Scholar] [CrossRef]
  195. Okutan, N.; Terzi, P.; Altay, F. Affecting parameters on electrospinning process and characterization of electrospun gelatin nanofibers. Food Hydrocoll. 2014, 39, 19–26. [Google Scholar] [CrossRef]
  196. Frenot, A.; Chronakis, I.S. Polymer nanofibers assembled by electrospinning. Curr. Opin. Colloid Interface Sci. 2003, 8, 64–75. [Google Scholar] [CrossRef]
  197. Haghi, A.K.; Akbari, M. Trends in electrospinning of natural nanofibers. Phys. Status Solidi (A) 2007, 204, 1830–1834. [Google Scholar] [CrossRef]
  198. Burger, C.; Hsiao, B.S.; Chu, B. NANOFIBROUS MATERIALS AND THEIR APPLICATIONS. Annu. Rev. Mater. Res. 2006, 36, 333–368. [Google Scholar] [CrossRef]
  199. Su, H.; Liu, K.-Y.; Karydis, A.; Abebe, D.G.; Wu, C.; Anderson, K.M.; Ghadri, N.; Adatrow, P.; Fujiwara, T.; Bumgardner, J.D. In vitro and in vivo evaluations of a novel post-electrospinning treatment to improve the fibrous structure of chitosan membranes for guided bone regeneration. Biomed. Mater. 2016, 12, 015003. [Google Scholar] [CrossRef]
  200. Rasouli, R.; Barhoum, A.; Bechelany, M.; Dufresne, A. Nanofibers for biomedical and healthcare applications. Macromol. Biosci. 2019, 19, 1800256. [Google Scholar] [CrossRef]
  201. Singh, B.N.; Pramanik, K. Development of novel silk fibroin/polyvinyl alcohol/sol–gel bioactive glass composite matrix by modified layer by layer electrospinning method for bone tissue construct generation. Biofabrication 2017, 9, 015028. [Google Scholar] [CrossRef]
  202. Kean, T.; Thanou, M. Biodegradation, biodistribution and toxicity of chitosan. Adv. Drug Deliv. Rev. 2010, 62, 3–11. [Google Scholar] [CrossRef] [PubMed]
  203. Pangon, A.; Saesoo, S.; Saengkrit, N.; Ruktanonchai, U.; Intasanta, V. Hydroxyapatite-hybridized chitosan/chitin whisker bionanocomposite fibers for bone tissue engineering applications. Carbohydr. Polym. 2016, 144, 419–427. [Google Scholar] [CrossRef] [PubMed]
  204. Ding, F.; Deng, H.; Du, Y.; Shi, X.; Wang, Q. Emerging chitin and chitosan nanofibrous materials for biomedical applications. Nanoscale 2014, 6, 9477–9493. [Google Scholar] [CrossRef] [PubMed]
  205. Kalantari, K.; Afifi, A.M.; Jahangirian, H.; Webster, T.J. Biomedical applications of chitosan electrospun nanofibers as a green polymer—Review. Carbohydr. Polym. 2019, 207, 588–600. [Google Scholar] [CrossRef]
  206. Hokmabad, V.R.; Davaran, S.; Aghazadeh, M.; Alizadeh, E.; Salehi, R.; Ramazani, A. A Comparison of the Effects of Silica and Hydroxyapatite Nanoparticles on Poly(ε-caprolactone)-Poly(ethylene glycol)-Poly(ε-caprolactone)/Chitosan Nanofibrous Scaffolds for Bone Tissue Engineering. Tissue Eng. Regen. Med. 2018, 15, 735–750. [Google Scholar] [CrossRef]
  207. Lastra, M.L.; Molinuevo, M.S.; Blaszczyk-Lezak, I.; Mijangos, C.; Cortizo, M.S. Nanostructured fumarate copolymer-chitosan crosslinked scaffold: An in vitro osteochondrogenesis regeneration study. J. Biomed. Mater. Res. Part A 2018, 106, 570–579. [Google Scholar] [CrossRef] [Green Version]
  208. Ghasemi Hamidabadi, H.; Rezvani, Z.; Nazm Bojnordi, M.; Shirinzadeh, H.; Seifalian, A.M.; Joghataei, M.T.; Razaghpour, M.; Alibakhshi, A.; Yazdanpanah, A.; Salimi, M.; et al. Chitosan-Intercalated Montmorillonite/Poly(vinyl alcohol) Nanofibers as a Platform to Guide Neuronlike Differentiation of Human Dental Pulp Stem Cells. ACS Appl. Mater. Interfaces 2017, 9, 11392–11404. [Google Scholar] [CrossRef]
  209. Alhosseini, S.N.; Moztarzadeh, F.; Mozafari, M.; Asgari, S.; Dodel, M.; Samadikuchaksaraei, A.; Kargozar, S.; Jalali, N. Synthesis and characterization of electrospun polyvinyl alcohol nanofibrous scaffolds modified by blending with chitosan for neural tissue engineering. Int. J. Nanomed. 2012, 25–34. [Google Scholar]
  210. Bolaina-Lorenzo, E.; Martínez-Ramos, C.; Monleón-Pradas, M.; Herrera-Kao, W.; Cauich-Rodríguez, J.V.; Cervantes-Uc, J.M. Electrospun polycaprolactone/chitosan scaffolds for nerve tissue engineering: Physicochemical characterization and Schwann cell biocompatibility. Biomed. Mater. 2016, 12, 015008. [Google Scholar] [CrossRef] [PubMed]
  211. Cao, L.; Zhang, F.; Wang, Q.; Wu, X. Fabrication of chitosan/graphene oxide polymer nanofiber and its biocompatibility for cartilage tissue engineering. Mater. Sci. Eng. C 2017, 79, 697–701. [Google Scholar] [CrossRef] [PubMed]
  212. Tonda-Turo, C.; Ruini, F.; Ramella, M.; Boccafoschi, F.; Gentile, P.; Gioffredi, E.; Falvo D’Urso Labate, G.; Ciardelli, G. Non-covalently crosslinked chitosan nanofibrous mats prepared by electrospinning as substrates for soft tissue regeneration. Carbohydr. Polym. 2017, 162, 82–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Hussain, A.; Collins, G.; Yip, D.; Cho, C.H. Functional 3-D cardiac co-culture model using bioactive chitosan nanofiber scaffolds. Biotechnol. Bioeng. 2013, 110, 637–647. [Google Scholar] [CrossRef] [PubMed]
  214. Fukunishi, T.; Best, C.A.; Sugiura, T.; Shoji, T.; Yi, T.; Udelsman, B.; Ohst, D.; Ong, C.S.; Zhang, H.; Shinoka, T. Tissue-engineered small diameter arterial vascular grafts from cell-free nanofiber PCL/chitosan scaffolds in a sheep model. PLoS ONE 2016, 11, e0158555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Kheradmandi, M.; Vasheghani-Farahani, E.; Ghiaseddin, A.; Ganji, F. Skeletal muscle regeneration via engineered tissue culture over electrospun nanofibrous chitosan/PVA scaffold. J. Biomed. Mater. Res. Part A 2016, 104, 1720–1727. [Google Scholar] [CrossRef]
  216. Shim, I.K.; Suh, W.H.; Lee, S.Y.; Lee, S.H.; Heo, S.J.; Lee, M.C.; Lee, S.J. Chitosan nano-/microfibrous double-layered membrane with rolled-up three-dimensional structures for chondrocyte cultivation. J. Biomed. Mater. Res. Part A 2009, 90A, 595–602. [Google Scholar] [CrossRef]
  217. De Farias, B.S.; Sant’Anna Cadaval, T.R., Jr.; de Almeida Pinto, L.A. Chitosan-functionalized nanofibers: A comprehensive review on challenges and prospects for food applications. Int. J. Biol. Macromol. 2019, 123, 210–220. [Google Scholar] [CrossRef]
  218. Sharifi, M.; Sohrabi, M.J.; Hosseinali, S.H.; Hasan, A.; Kani, P.H.; Talaei, A.J.; Karim, A.Y.; Nanakali, N.M.Q.; Salihi, A.; Aziz, F.M.; et al. Enzyme immobilization onto the nanomaterials: Application in enzyme stability and prodrug-activated cancer therapy. Int. J. Biol. Macromol. 2020, 143, 665–676. [Google Scholar] [CrossRef]
  219. Bösiger, P.; Tegl, G.; Richard, I.M.T.; Le Gat, L.; Huber, L.; Stagl, V.; Mensah, A.; Guebitz, G.M.; Rossi, R.M.; Fortunato, G. Enzyme functionalized electrospun chitosan mats for antimicrobial treatment. Carbohydr. Polym. 2018, 181, 551–559. [Google Scholar] [CrossRef]
  220. Ismail, A.R.; Baek, K.-H. Lipase immobilization with support materials, preparation techniques, and applications: Present and future aspects. Int. J. Biol. Macromol. 2020, 163, 1624–1639. [Google Scholar] [CrossRef]
  221. Srivastava, B.; Singh, H.; Khatri, M.; Singh, G.; Arya, S.K. Immobilization of keratinase on chitosan grafted-β-cyclodextrin for the improvement of the enzyme properties and application of free keratinase in the textile industry. Int. J. Biol. Macromol. 2020, 165, 1099–1110. [Google Scholar] [CrossRef]
  222. de Oliveira, R.L.; da Silva, M.F.; da Silva, S.P.; de Araújo, A.C.V.; Cavalcanti, J.V.F.L.; Converti, A.; Porto, T.S. Fructo-oligosaccharides production by an Aspergillus aculeatus commercial enzyme preparation with fructosyltransferase activity covalently immobilized on Fe3O4–chitosan-magnetic nanoparticles. Int. J. Biol. Macromol. 2020, 150, 922–929. [Google Scholar] [CrossRef] [PubMed]
  223. Soares, A.M.B.F.; Gonçalves, L.M.O.; Ferreira, R.D.S.; de Souza, J.M.; Fangueiro, R.; Alves, M.M.M.; Carvalho, F.A.A.; Mendes, A.N.; Cantanhêde, W. Immobilization of papain enzyme on a hybrid support containing zinc oxide nanoparticles and chitosan for clinical applications. Carbohydr. Polym. 2020, 243, 116498. [Google Scholar] [CrossRef] [PubMed]
  224. Tavernini, L.; Ottone, C.; Illanes, A.; Wilson, L. Entrapment of enzyme aggregates in chitosan beads for aroma release in white wines. Int. J. Biol. Macromol. 2020, 154, 1082–1090. [Google Scholar] [CrossRef] [PubMed]
  225. Duru Kamaci, U.; Peksel, A. Fabrication of PVA-chitosan-based nanofibers for phytase immobilization to enhance enzymatic activity. Int. J. Biol. Macromol. 2020, 164, 3315–3322. [Google Scholar] [CrossRef]
  226. Haghju, S.; Bari, M.R.; Khaled-Abad, M.A. Affecting parameters on fabrication of β-d-galactosidase immobilized chitosan/poly (vinyl alcohol) electrospun nanofibers. Carbohydr. Polym. 2018, 200, 137–143. [Google Scholar] [CrossRef]
  227. Kutlu, N.; İspirli Doğaç, Y.; Deveci, İ.; Teke, M. Urease immobilized electrospun PVA/chitosan nanofibers with improved stability and reusability characteristics: An application for removal of urea from artificial blood serum. Prep. Biochem. Biotechnol. 2020, 50, 425–437. [Google Scholar] [CrossRef]
  228. Park, J.-M.; Kim, M.; Park, H.-S.; Jang, A.; Min, J.; Kim, Y.-H. Immobilization of lysozyme-CLEA onto electrospun chitosan nanofiber for effective antibacterial applications. Int. J. Biol. Macromol. 2013, 54, 37–43. [Google Scholar] [CrossRef]
  229. Jhuang, J.-R.; Lou, S.-N.; Lin, S.-B.; Chen, S.H.; Chen, L.-C.; Chen, H.-H. Immobilizing laccase on electrospun chitosan fiber to prepare time-temperature indicator for food quality monitoring. Innov. Food Sci. Emerg. Technol. 2020, 63, 102370. [Google Scholar] [CrossRef]
  230. Srbová, J.; Slováková, M.; Křípalová, Z.; Žárská, M.; Špačková, M.; Stránská, D.; Bílková, Z. Covalent biofunctionalization of chitosan nanofibers with trypsin for high enzyme stability. React. Funct. Polym. 2016, 104, 38–44. [Google Scholar] [CrossRef] [Green Version]
  231. Maryšková, M.; Ardao, I.; García-González, C.A.; Martinová, L.; Rotková, J.; Ševců, A. Polyamide 6/chitosan nanofibers as support for the immobilization of Trametes versicolor laccase for the elimination of endocrine disrupting chemicals. Enzym. Microb. Technol. 2016, 89, 31–38. [Google Scholar] [CrossRef] [PubMed]
  232. Demirkan, E.; Avci, T.; Aykut, Y. Protease immobilization on cellulose monoacetate/chitosan-blended nanofibers. J. Ind. Text. 2017, 47, 2092–2111. [Google Scholar] [CrossRef]
  233. Teepoo, S.; Dawan, P.; Barnthip, N. Electrospun Chitosan-Gelatin Biopolymer Composite Nanofibers for Horseradish Peroxidase Immobilization in a Hydrogen Peroxide Biosensor. Biosensors 2017, 7, 47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Radmansouri, M.; Bahmani, E.; Sarikhani, E.; Rahmani, K.; Sharifianjazi, F.; Irani, M. Doxorubicin hydrochloride-Loaded electrospun chitosan/cobalt ferrite/titanium oxide nanofibers for hyperthermic tumor cell treatment and controlled drug release. Int. J. Biol. Macromol. 2018, 116, 378–384. [Google Scholar] [CrossRef]
  235. Lu, L.; Cao, X.; Shen, Z.; Li, L.; Huo, J.; Chen, W.; Liu, C.; Liu, H. Electrospun nitrogen-doped carbon nanofibers for electrocatalysis. Sustain. Mater. Technol. 2020, 26, e00221. [Google Scholar] [CrossRef]
  236. Jouybari, M.H.; Hosseini, S.; Mahboobnia, K.; Boloursaz, L.A.; Moradi, M.; Irani, M. Simultaneous controlled release of 5-FU, DOX and PTX from chitosan/PLA/5-FU/g-C3N4-DOX/g-C3N4-PTX triaxial nanofibers for breast cancer treatment in vitro. Colloids Surf. B Biointerfaces 2019, 179, 495–504. [Google Scholar] [CrossRef]
  237. Balan, P.; Indrakumar, J.; Murali, P.; Korrapati, P.S. Bi-faceted delivery of phytochemicals through chitosan nanoparticles impregnated nanofibers for cancer therapeutics. Int. J. Biol. Macromol. 2020, 142, 201–211. [Google Scholar] [CrossRef]
  238. Ardeshirzadeh, B.; Anaraki, N.A.; Irani, M.; Rad, L.R.; Shamshiri, S. Controlled release of doxorubicin from electrospun PEO/chitosan/graphene oxide nanocomposite nanofibrous scaffolds. Mater. Sci. Eng. C 2015, 48, 384–390. [Google Scholar] [CrossRef] [PubMed]
  239. Rostami, M.; Ghorbani, M.; Aman Mohammadi, M.; Delavar, M.; Tabibiazar, M.; Ramezani, S. Development of resveratrol loaded chitosan-gellan nanofiber as a novel gastrointestinal delivery system. Int. J. Biol. Macromol. 2019, 135, 698–705. [Google Scholar] [CrossRef]
  240. Yan, E.; Cao, M.; Wang, Y.; Hao, X.; Pei, S.; Gao, J.; Wang, Y.; Zhang, Z.; Zhang, D. Gold nanorods contained polyvinyl alcohol/chitosan nanofiber matrix for cell imaging and drug delivery. Mater. Sci. Eng. C 2016, 58, 1090–1097. [Google Scholar] [CrossRef] [PubMed]
  241. Aggarwal, U.; Goyal, A.K.; Rath, G. Development and characterization of the cisplatin loaded nanofibers for the treatment of cervical cancer. Mater. Sci. Eng. C 2017, 75, 125–132. [Google Scholar] [CrossRef] [PubMed]
  242. Ma, G.; Liu, Y.; Peng, C.; Fang, D.; He, B.; Nie, J. Paclitaxel loaded electrospun porous nanofibers as mat potential application for chemotherapy against prostate cancer. Carbohydr. Polym. 2011, 86, 505–512. [Google Scholar] [CrossRef]
  243. Samadi, S.; Moradkhani, M.; Beheshti, H.; Irani, M.; Aliabadi, M. Fabrication of chitosan/poly(lactic acid)/graphene oxide/TiO2 composite nanofibrous scaffolds for sustained delivery of doxorubicin and treatment of lung cancer. Int. J. Biol. Macromol. 2018, 110, 416–424. [Google Scholar] [CrossRef]
  244. Sedghi, R.; Shaabani, A.; Mohammadi, Z.; Samadi, F.Y.; Isaei, E. Biocompatible electrospinning chitosan nanofibers: A novel delivery system with superior local cancer therapy. Carbohydr. Polym. 2017, 159, 1–10. [Google Scholar] [CrossRef] [PubMed]
  245. Irani, M.; Mir Mohamad Sadeghi, G.; Haririan, I. A novel biocompatible drug delivery system of chitosan/temozolomide nanoparticles loaded PCL-PU nanofibers for sustained delivery of temozolomide. Int. J. Biol. Macromol. 2017, 97, 744–751. [Google Scholar] [CrossRef]
  246. Zhou, X.; Chen, L.; Wang, W.; Jia, Y.; Chang, A.; Mo, X.; Wang, H.; He, C. Electrospun nanofibers incorporating self-decomposable silica nanoparticles as carriers for controlled delivery of anticancer drug. RSC Adv. 2015, 5, 65897–65904. [Google Scholar] [CrossRef]
  247. Zhao, J.; Zhu, Y.; Ye, C.; Chen, Y.; Wang, S.; Zou, D.; Li, Z. Photothermal transforming agent and chemotherapeutic co-loaded electrospun nanofibers for tumor treatment. Int. J. Nanomed. 2019, 3893–3909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Bharathi, D.; Ranjithkumar, R.; Chandarshekar, B.; Bhuvaneshwari, V. Bio-inspired synthesis of chitosan/copper oxide nanocomposite using rutin and their anti-proliferative activity in human lung cancer cells. Int. J. Biol. Macromol. 2019, 141, 476–483. [Google Scholar] [CrossRef]
  249. Ibrahim, H.M.; Reda, M.M.; Klingner, A. Preparation and characterization of green carboxymethylchitosan (CMCS) – Polyvinyl alcohol (PVA) electrospun nanofibers containing gold nanoparticles (AuNPs) and its potential use as biomaterials. Int. J. Biol. Macromol. 2020, 151, 821–829. [Google Scholar] [CrossRef]
  250. Farhoodi, M. Nanocomposite Materials for Food Packaging Applications: Characterization and Safety Evaluation. Food Eng. Rev. 2016, 8, 35–51. [Google Scholar] [CrossRef]
  251. Mousavi Khaneghah, A.; Hashemi, S.M.B.; Limbo, S. Antimicrobial agents and packaging systems in antimicrobial active food packaging: An overview of approaches and interactions. Food Bioprod. Process. 2018, 111, 1–19. [Google Scholar] [CrossRef]
  252. Youssef, A.M.; El-Sayed, S.M. Bionanocomposites materials for food packaging applications: Concepts and future outlook. Carbohydr. Polym. 2018, 193, 19–27. [Google Scholar] [CrossRef] [PubMed]
  253. Calderón, L.; Harris, R.; Cordoba-Diaz, M.; Elorza, M.; Elorza, B.; Lenoir, J.; Adriaens, E.; Remon, J.P.; Heras, A.; Cordoba-Diaz, D. Nano and microparticulate chitosan-based systems for antiviral topical delivery. Eur. J. Pharm. Sci. 2013, 48, 216–222. [Google Scholar] [CrossRef]
  254. Helander, I.M.; Nurmiaho-Lassila, E.L.; Ahvenainen, R.; Rhoades, J.; Roller, S. Chitosan disrupts the barrier properties of the outer membrane of Gram-negative bacteria. Int. J. Food Microbiol. 2001, 71, 235–244. [Google Scholar] [CrossRef]
  255. Ing, L.Y.; Zin, N.M.; Sarwar, A.; Katas, H. Antifungal Activity of Chitosan Nanoparticles and Correlation with Their Physical Properties. Int. J. Biomater. 2012, 2012, 632698. [Google Scholar] [CrossRef] [Green Version]
  256. Jaime, M.D.L.A.; Lopez-Llorca, L.V.; Conesa, A.; Lee, A.Y.; Proctor, M.; Heisler, L.E.; Gebbia, M.; Giaever, G.; Westwood, J.T.; Nislow, C. Identification of yeast genes that confer resistance to chitosan oligosaccharide (COS) using chemogenomics. BMC Genom. 2012, 13, 267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Abdul Khalil, H.P.S.; Banerjee, A.; Saurabh, C.K.; Tye, Y.Y.; Suriani, A.B.; Mohamed, A.; Karim, A.A.; Rizal, S.; Paridah, M.T. Biodegradable Films for Fruits and Vegetables Packaging Application: Preparation and Properties. Food Eng. Rev. 2018, 10, 139–153. [Google Scholar] [CrossRef]
  258. Arkoun, M.; Daigle, F.; Heuzey, M.-C.; Ajji, A. Mechanism of action of electrospun chitosan-based nanofibers against meat spoilage and pathogenic bacteria. Molecules 2017, 22, 585. [Google Scholar] [CrossRef] [Green Version]
  259. Cui, H.; Bai, M.; Li, C.; Liu, R.; Lin, L. Fabrication of chitosan nanofibers containing tea tree oil liposomes against Salmonella spp. in chicken. LWT 2018, 96, 671–678. [Google Scholar] [CrossRef]
  260. Ceylan, Z.; Unal Sengor, G.F.; Yilmaz, M.T. Nanoencapsulation of liquid smoke/thymol combination in chitosan nanofibers to delay microbiological spoilage of sea bass (Dicentrarchus labrax) fillets. J. Food Eng. 2018, 229, 43–49. [Google Scholar] [CrossRef]
  261. He, X.; Hwang, H.-M. Nanotechnology in food science: Functionality, applicability, and safety assessment. J. Food Drug Anal. 2016, 24, 671–681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Cook, S.L.; Bull, S.P.; Methven, L.; Parker, J.K.; Khutoryanskiy, V.V. Mucoadhesion: A food perspective. Food Hydrocoll. 2017, 72, 281–296. [Google Scholar] [CrossRef]
  263. Martins, J.T.; Ramos, Ó.L.; Pinheiro, A.C.; Bourbon, A.I.; Silva, H.D.; Rivera, M.C.; Cerqueira, M.A.; Pastrana, L.; Malcata, F.X.; González-Fernández, Á.; et al. Edible Bio-Based Nanostructures: Delivery, Absorption and Potential Toxicity. Food Eng. Rev. 2015, 7, 491–513. [Google Scholar] [CrossRef] [Green Version]
  264. Ali, A.; Ahmed, S. A review on chitosan and its nanocomposites in drug delivery. Int. J. Biol. Macromol. 2018, 109, 273–286. [Google Scholar] [CrossRef]
  265. Darbasizadeh, B.; Motasadizadeh, H.; Foroughi-Nia, B.; Farhadnejad, H. Tripolyphosphate-crosslinked chitosan/poly (ethylene oxide) electrospun nanofibrous mats as a floating gastro-retentive delivery system for ranitidine hydrochloride. J. Pharm. Biomed. Anal. 2018, 153, 63–75. [Google Scholar] [CrossRef]
  266. Hasani, S.; Ojagh, S.M.; Ghorbani, M. Nanoencapsulation of lemon essential oil in Chitosan-Hicap system. Part 1: Study on its physical and structural characteristics. Int. J. Biol. Macromol. 2018, 115, 143–151. [Google Scholar] [CrossRef]
  267. Omwenga, E.O.; Hensel, A.; Shitandi, A.; Goycoolea, F.M. Chitosan nanoencapsulation of flavonoids enhances their quorum sensing and biofilm formation inhibitory activities against an E.coli Top 10 biosensor. Colloids Surf. B Biointerfaces 2018, 164, 125–133. [Google Scholar] [CrossRef]
  268. Shekarforoush, E.; Ajalloueian, F.; Zeng, G.; Mendes, A.C.; Chronakis, I.S. Electrospun xanthan gum-chitosan nanofibers as delivery carrier of hydrophobic bioactives. Mater. Lett. 2018, 228, 322–326. [Google Scholar] [CrossRef]
  269. Kuddus, M. Enzymes in Food Biotechnology: Production, Applications, and Future Prospects; Elsevier Inc.: Amsterdam, The Netherlands, 2018. [Google Scholar]
  270. Wang, Z.-G.; Wan, L.-S.; Liu, Z.-M.; Huang, X.-J.; Xu, Z.-K. Enzyme immobilization on electrospun polymer nanofibers: An overview. J. Mol. Catal. B Enzym. 2009, 56, 189–195. [Google Scholar] [CrossRef]
Figure 1. Chemical Structure and some sources of chitin and chitosan [5].
Figure 1. Chemical Structure and some sources of chitin and chitosan [5].
Polymers 15 02820 g001
Figure 2. Chitosan-based materials with different shapes and sizes: (a) chitosan nanoparticles; (b) chitosan nanofibers fabricated by electrostatic spinning technology; (c) chitosan–pectin hydrogel grid scaffold prepared by 3D printing technology; (d) chitosan core–alginate shell microspheres (e) chitosan-based fibers fabricated by solvent spinning technology; and (f) 3D-printed chitosan porous structures. Copyright 2023. Reproduced with permission from Elsevier Science Ltd. [18].
Figure 2. Chitosan-based materials with different shapes and sizes: (a) chitosan nanoparticles; (b) chitosan nanofibers fabricated by electrostatic spinning technology; (c) chitosan–pectin hydrogel grid scaffold prepared by 3D printing technology; (d) chitosan core–alginate shell microspheres (e) chitosan-based fibers fabricated by solvent spinning technology; and (f) 3D-printed chitosan porous structures. Copyright 2023. Reproduced with permission from Elsevier Science Ltd. [18].
Polymers 15 02820 g002
Figure 3. Chemical modification of chitosan.
Figure 3. Chemical modification of chitosan.
Polymers 15 02820 g003
Figure 4. Proposed mechanisms (pathway) for chemical modification of chitosan via conjugation (a) free radical-induced conjugation to form a polyphenol chitosan conjugation; (b) carbodiimide chemical mechanism to form Schiff base compounds; (c) functional group conversion strategy, or (d) conjugation of chitosan with polyphenol via enzymatic assisted coupling reaction.
Figure 4. Proposed mechanisms (pathway) for chemical modification of chitosan via conjugation (a) free radical-induced conjugation to form a polyphenol chitosan conjugation; (b) carbodiimide chemical mechanism to form Schiff base compounds; (c) functional group conversion strategy, or (d) conjugation of chitosan with polyphenol via enzymatic assisted coupling reaction.
Polymers 15 02820 g004aPolymers 15 02820 g004bPolymers 15 02820 g004cPolymers 15 02820 g004d
Figure 5. Electron transmission microphotography of: (a) CS nanoparticles; (b) CS/PEO–PPO nanoparticles (concentration of PEO–PPO in the chitosan solution: 10 mg/mL) prepared by Calvo et al. [36], reproduced with permission from John Wiley and Sons.
Figure 5. Electron transmission microphotography of: (a) CS nanoparticles; (b) CS/PEO–PPO nanoparticles (concentration of PEO–PPO in the chitosan solution: 10 mg/mL) prepared by Calvo et al. [36], reproduced with permission from John Wiley and Sons.
Polymers 15 02820 g005
Figure 6. Flowchart showing the creation of chitosan–PLGA particles step-by-step.
Figure 6. Flowchart showing the creation of chitosan–PLGA particles step-by-step.
Polymers 15 02820 g006
Figure 7. SEM micrographs of different hybrid hierarchical structures resulting from freezing hydrogel nanocomposites, with identical CHI and calcium phosphate composition (93.25 and 6.75 wt.%, respectively) at different rates of freezing: (a) 0.7 mm/min; (b) 2.7 mm/min, and (c) 5.7 mm/min. Scale bars are 50 μm. TEM (d) micrographs of ACP nanoclusters forming the ACP/CHI hierarchical structure. Ref. [52] Reproduced with permission from MDPI.
Figure 7. SEM micrographs of different hybrid hierarchical structures resulting from freezing hydrogel nanocomposites, with identical CHI and calcium phosphate composition (93.25 and 6.75 wt.%, respectively) at different rates of freezing: (a) 0.7 mm/min; (b) 2.7 mm/min, and (c) 5.7 mm/min. Scale bars are 50 μm. TEM (d) micrographs of ACP nanoclusters forming the ACP/CHI hierarchical structure. Ref. [52] Reproduced with permission from MDPI.
Polymers 15 02820 g007
Figure 8. Position of nanofibers between protein, bacteria, and viruses [126]. reproduced with permission from John Wiley and Sons.
Figure 8. Position of nanofibers between protein, bacteria, and viruses [126]. reproduced with permission from John Wiley and Sons.
Polymers 15 02820 g008
Figure 9. A diagram of electrospinning equipment with a syringe injecting polymer solution into an electric field generated by a high voltage power source between the spinneret and the grounded collector; a high voltage is applied. (a) A typical vertical arrangement of an electrospinning apparatus. (b) A typical horizontal setup of an electrospinning apparatus [119].
Figure 9. A diagram of electrospinning equipment with a syringe injecting polymer solution into an electric field generated by a high voltage power source between the spinneret and the grounded collector; a high voltage is applied. (a) A typical vertical arrangement of an electrospinning apparatus. (b) A typical horizontal setup of an electrospinning apparatus [119].
Polymers 15 02820 g009
Figure 10. SEM images of 5 wt.% chitosan hydrolyzed 48 h nanofibers in aqueous acetic acid 90%, tip needle–collector distance: 14 cm (a), 15 cm (b), 16 cm (c). [119] Copyright 2023. Reproduced with permission from Elsevier Science Ltd.
Figure 10. SEM images of 5 wt.% chitosan hydrolyzed 48 h nanofibers in aqueous acetic acid 90%, tip needle–collector distance: 14 cm (a), 15 cm (b), 16 cm (c). [119] Copyright 2023. Reproduced with permission from Elsevier Science Ltd.
Polymers 15 02820 g010
Figure 11. SEM images of 5 wt.% chitosan hydrolyzed 48 h nanofibers in aqueous acetic acid 90%, electric field voltage: 14 kV (a), 16 kV (b), 17 kV (c). [196] Copyright 2023. Reproduced with permission from Elsevier Science Ltd.
Figure 11. SEM images of 5 wt.% chitosan hydrolyzed 48 h nanofibers in aqueous acetic acid 90%, electric field voltage: 14 kV (a), 16 kV (b), 17 kV (c). [196] Copyright 2023. Reproduced with permission from Elsevier Science Ltd.
Polymers 15 02820 g011
Table 1. Particle size zeta potential theoretical loading, and encapsulation efficiency values of CS [41].
Table 1. Particle size zeta potential theoretical loading, and encapsulation efficiency values of CS [41].
PolymerProtein LoadedSize (nm)Potential (mV)Theoretical
Lading (%)
Encapsulation
Efficiency (%)
PLATetanus Toxoid192 ± 12−47.9 ± 1.5136.7 ± 0.3
PEG-PLATetanus Toxoid196 ± 20−23.9 ± 1.2131.1 ± 0.5
CS-PLGATetanus Toxoid500 ± 29−21.8 ± 1.1190.0 ± 3.6
CSTetanus Toxoid354 ± 27−37.1 ± 5.91055.1 ± 3.4
CSInsulin337 ± 14−36.9 ± 0.34094.7 ± 2.1
Table 2. Chitosan-based films have been used in a wide variety of food products.
Table 2. Chitosan-based films have been used in a wide variety of food products.
Chitosan Film/CoatedFoodReferences
ChitosanCitrus fruit[70,71]
ChitosanLogan fruit[72]
ChitosanGreen coffee beans[73]
ChitosanFrozen salmon[74]
ChitosanTankan citrus fruit[75]
ChitosanApples[76,77]
ChitosanMushroom[78]
ChitosanMangoes[70,71,79]
ChitosanLitchi fruit[80,81]
ChitosanSalmon fillets[82]
ChitosanTomatoes[83]
ChitosanStrawberries[83]
ChitosanPeach fruit[72]
ChitosanFresh-cut Chinese water chestnut L.T.[84]
ChitosanSilver carp[85]
Table 3. Biological applications of chitosan and its derivatives.
Table 3. Biological applications of chitosan and its derivatives.
Composite/PropertiesStimulatory EffectsReferences
Environmental purposes
FlocculationRemoves a variety of contaminants from wastewater in an efficient manner[101]
Metals and organic compounds adsorptionRemoves pathogens, radioactive materials, heavy metals, colors, organic chemicals, and inorganic nutrients (nitrates and phosphates)[101]
Agricultural purposes
Biocontrol agentSafe alternative to the use of pesticides and agrochemicals as a biocontrol material against many pathogenic microorganisms[102]
Enhance crop production
-
Effectively raises the productivity of various agricultural plants
-
Effectively encourages plant development in a variety of crops
[103]
Aquafeed additivesPositively impacts the growth, digestive enzymes, body composition, intestinal bacterial count, immunological response, and hematological and liver health of commercial freshwater fish[104,105]
Biomedical purposes
Chitosan microspheres
-
Used as a carrier for targeted and prolonged delivery of drugs
-
Increases the bioavailability of degradable substances such as protein
-
Increases uptake of hydrophilic substances across the epithelial layers
[106]
Chitosan mesh membraneDecreases wound healing time and increases the recovery of the granular layer[107]
Chitosan collagen blend membraneIncreases the antibacterial activity against E. coli and S. aureus and decreases the excessive dehydration of the wound[108]
Alginate/carboxymethyl chitosan blend fibersIncreased water-retention and increased antibacterial activity against S. aureus[109]
Composite nanofibrous membranes (NFM) of collagen and chitosanIncreased wound-healing and increased tissue regeneration[110]
Electro spun chitosan fiber with polyethylene oxideUsed effectively as surface layers on the wound site in periodontal disease[111]
Chitosan membranes loaded with Tetracycline hydrochloride or silver sulfadiazineIncreased wettability, decreased swelling rate, water vapor permeability, and tensile strength, and increased antimicrobial activity against E. coli and S. aureus[112]
Chitosan titanium dioxide composite membranesIncreased antimicrobial activity against S. aureus, decreased oxidative stress and apoptosis of fibroblast cell sand, increased proliferation in L929 fibroblast cells[113]
Chitosan nano silver dressingIncreased wound-healing using the non-invasive dressing[114]
Chitosan sponges loaded with norfloxacinThe dressing can conduct the role of normal skin and the antibiotic release is swelling-controlled[115]
Chitosan-gelatin spongeIncreased antimicrobial activity against E. coli K88 over penicillin
Increased antimicrobial activity against S. aureus over cefradine
[116]
Photo cross linkable chitosan hydrogel containing fibroblast growth factor-2Increased wound healing in diabetic and normal mice[117]
Carboxymethyl chitosan alginate hydrogelIncreases Bactericidal properties toward S. aureus and E. coli, and antibiotic continues to be released from the hydrogel[118]
Table 5. Chitosan–based nanofibers in enzyme immobilization.
Table 5. Chitosan–based nanofibers in enzyme immobilization.
CompositionComponent’s RatioEnzymeReferences
Chitosan/poly (vinyl alcohol) composite in 0.1 M Sodium acetate1:5,4Phytase by entrapment[225]
Chitosan/poly (vinyl alcohol) composite in 1% acetic acid1:1B-d-galactosidase by entrapment[226]
Chitosan/poly (vinyl alcohol) composite in water1:6Urease by Cross-linking[227]
Chitosan/poly (vinyl alcohol) composite in 0.5% acetic acid1:10Lysozyme by Cross-linking[228]
Chitosan/poly (vinyl alcohol) composite in 2% acetic acid1:4Laccase by Cross-linking [229]
Chitosan/polyethylene oxide composite in 1% acetic acid22:1Trypsin by Adsorption/Covalent[230]
Chitosan/poly (ethylene oxide) composite in 90% acetic acid95:5Glucose oxidase by Cross-linking[219]
Chitosan/polyamide 6 composite in Formic acid and Acetic acid (2:1 v/v)1:9Laccase by Cross-linking[231]
Chitosan/Cellulose monoacetate composite in acetone 99%1:5, 2:5, 1:1, 7:5Protease by Adsorption/Cross linking[232]
Chitosan/gelatin composite in 60% acetic acid4:6Peroxidase by Cross-linking[233]
Table 6. The chitosan nanofibers as anticancer drug delivery in vitro conditions.
Table 6. The chitosan nanofibers as anticancer drug delivery in vitro conditions.
Composition Anticancer DrugCancer TypeReferences
ChitosanFe3O4/under magnetic fieldHFL1[234]
ChitosanFe2+/under magnetic fieldHFL1[234]
ChitosanGlutaraldehyde/under magnetic fieldHFL1[234]
Chitosan/poly(ε-caprolactone) composite5-FluorouracilB16F10[235]
Chitosan/polyvinyl alcohol-g-C3N4-g-C3N4 composite5-Fluorouracil,
Doxorubicin,
Paclitaxel
MCF-7[236]
Chitosan/polycaprolactone compositeResveratrol, ferulic acidHaCat, A431[237]
Chitosan/polyethylene oxide compositeBerberineHeLa, BT474, MCF-7, MDA-MB-468[237]
Polyethylene oxide/chitosan/graphene oxide compositeDoxorubicinA549[238]
Chitosan/gelatin compositeResveratrolHT29[239]
Polyvinyl alcohol/chitosan/Au compositeDoxorubicinSKOV3[240]
Polycaprolactone/chitosan compositeCisplatinErlich ascites carcinoma[241]
Chitosan/polyethylene oxide/hyaluronic acid compositePaclitaxelDU145[242]
Chitosan/cobalt ferrite/TiO2 compositeDoxorubicinB16F10[234]
Chitosan/poly (lactic acid)/TiO2/graphene oxideDoxorubicinA549[243]
Chitosan/PVA/graphene oxide/Si compositeCurcuminMCF-7[244]
Poly (ε-caprolactone diol)/polyurethane/chitosan/Au TiO2TemozolomideU-87 MG[245]
Graphene oxide/chitosan compositeCurcuminMCF-7, HepG2, L929[244]
Poly (lactic-co-glycolic acid)/chitosan/SiO2 compositeDoxorubicinHeLa[246]
Chitosan/polyvinyl alcohol/MoS2 compositeDoxorubicinHT29, HT29 cell lines[247]
ChitosanCupric oxideA549[248]
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

Ibrahim, M.A.; Alhalafi, M.H.; Emam, E.-A.M.; Ibrahim, H.; Mosaad, R.M. A Review of Chitosan and Chitosan Nanofiber: Preparation, Characterization, and Its Potential Applications. Polymers 2023, 15, 2820. https://doi.org/10.3390/polym15132820

AMA Style

Ibrahim MA, Alhalafi MH, Emam E-AM, Ibrahim H, Mosaad RM. A Review of Chitosan and Chitosan Nanofiber: Preparation, Characterization, and Its Potential Applications. Polymers. 2023; 15(13):2820. https://doi.org/10.3390/polym15132820

Chicago/Turabian Style

Ibrahim, Marwan A., Mona H. Alhalafi, El-Amir M. Emam, Hassan Ibrahim, and Rehab M. Mosaad. 2023. "A Review of Chitosan and Chitosan Nanofiber: Preparation, Characterization, and Its Potential Applications" Polymers 15, no. 13: 2820. https://doi.org/10.3390/polym15132820

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

Article Metrics

Back to TopTop