Biocomposite-Based on Natural Polymers: Comparison
Please note this is a comparison between Version 2 by Sirius Huang and Version 1 by Emilio Bucio.

Biopolymers are materials obtained from renewable resources. Despite the exciting properties of biopolymers, such as biocompatibility and environmental sustainability, they do not present antimicrobial properties (except chitosan). However, this lack of antimicrobial properties can be solved by incorporating or encapsulating antimicrobial agents. Natural polymers possess low stability in aqueous media and limited mechanical strength, which could be improved through cross-linking strategies. Hydrogels are biocompatible materials that can be synthesized from natural polymers, forming a cross-linking material. Alginate, collagen, fibrin, chitosan, gelatin, and hyaluronic acid are some natural polymers used to synthesize hydrogels.

  • natural polymers
  • biocomposites
  • antimicrobial agents
  • antimicrobial activity
  • biofilm

1. Cellulose-Based Composites

Cellulose is the most common polymer on earth because it can be easily obtained from the cell wall of plants and is even produced by some bacteria [70][1]. There is a 9–25% cellulose content in primary cell walls and 40–80% in secondary cell walls [71][2]. Therefore, cellulose is considered a renewable polymer [72][3] and semi-crystal material with high molecular weight homopolymer β-D-glucopyranose units, with a dimer of glucose (cellobiose) as a repeat unit [135,136][4][5]. It could be obtained as different derivatives such as Ethyl cellulose (EC), methylcellulose (ME), Cellulose acetate (CA), Cellulose sulfate (CS), Cellulose nitrate (CN) [73][6], and nanocellulose [137][7]. Cellulose and its derivatives could present high thermal resistance, protection against ultraviolet agents, low cost, biodegradability, and non-toxicity. However, cellulose has limitations such as high-water absorption capacity and insufficient interfacial adhesion [138][8]. Besides, cellulose has been studied as a carrier of antimicrobial agents, either as nanocapsules or nanofiber, as shown in Figure 31. Antimicrobial activity is one of cellulose fibers’ most critical functional properties [139,140,141,142,143][9][10][11][12][13].
Figure 31.
 Biopolymer nanocarriers structures for encapsulation of antimicrobial agents.
The nanocelluloses combine important cellulose properties such as high specific strength, modulus, hydrophilicity [144][14], biodegradability, nontoxic, extremely high surface area, and tunable surface chemistry [145][15]. Nanocellulose-based antimicrobial materials can be synthesized through surface modification with biocidal agents, making them effective against wound infection [78][16]. The antibacterial assays have confirmed the efficient antibacterial activity of several nanocellulose-based to inhibit bacterial growth (in both liquid medium and agar plates). They can kill several logs of microbial cells [146][17].
Nanofibrillated cellulose has been used to carry chitosan capsules [73][6]. Its high antimicrobial activity studied nanofibrillated cellulose (NFC) against gram-positive and gram-negative bacteria. It presents the ability of the polymeric grafts to penetrate the thick cell wall and destabilize the cellular membrane [147][18]. Nanofibrous scaffolds were prepared from polyurethane and CA and could contain reduced graphene oxide/silver nanocomposites (rGO/Ag), curcumin, or both, which can present antimicrobial properties [74][19]. Good antimicrobial properties were obtained from the composite of cellulose/keratin, where there was a comparison between silver nanoparticles (AgNPs) and ionic silver as an active agent. AgNPs present a higher activity against Escherichia coli (E. coli), S. aureusEnterococcus faecalis, and Pseudomonas aeruginosa (P. aeruginosa) [72][3].
Cinnamon essential oil has been studied as an antimicrobial agent in a cellulose polymeric matrix. Cinnamon was more effective against S. aureus than E. coli due to protecting outer membrane proteins or cell walls, which are more resistant to lipophilic substances of gram-negative bacteria [75][20]. Orlando et al. proposed the functionalization of cellulose with glycidyl trimethylammonium chloride and glycidyl hexadecyl ether for antibacterial wound dressing. These active agents were used for the covalent derivatization of the hydroxyl groups of glucose through a heterogeneous reaction in basic aqueous conditions. The resulting material reduces S. aureus and E. coli by 53 and 43%, respectively, within the first 24 h [76][21]. Silva et al. reported the casting method’s synthesis of an active film based on cellulose derivatives (hydroxyethylcellulose and CA), using hydroxypropyl-γ-cyclodextrin as active agents. The antimicrobial activity of these films was evaluated against Campylobacter jejuniCampylobacter coli, and Arcobacter butzleri. The results indicated that this system is a good approach for Campylobacter coli reduction [77][22].

2. Chitosan-Based Composites

Chitosan is the second polysaccharide more abundant in nature. It is structurally a linear polysaccharide made up of arbitrarily distributed β-(1–4)-linked d-glucosamine (deacetylated) and N-acetyl-d-glucosamine (acetylated) [148][23]. The principal source is shrimps shell wastes, crab peritrophic membranes, lobsters, and cocoons of insects [79][24]. Chitosan has a great potential for a wide range of applications due to its biodegradability, biocompatibility, antimicrobial activity, non-toxicity, and versatile chemical and physical properties [149][25]. Nonetheless, chitosan is insoluble at neutral pH or above. Therefore, it is necessary to modify natural polymers such as chitosan to become partially water solubility and enhance their antimicrobial activity [150][26]. The presence of amino and hydroxyl groups on the structure makes the chemical modification of chitosan possible to improve its solubility and electric change [151,152][27][28].
Chitosan exhibits an intrinsic antibacterial activity, inhibiting bacteria and fungi growth [153][29]. A significant number of amino groups on the surface of chitosan aid in generating positive zeta potentials [154][30]. Protonation of these amine groups on chitosan glucosamine monomers is facilitated at pH below 6.5 (pKa of chitosan), thus conferring cationic properties on chitosan [155][31]. Therefore, chitosan can interact with negatively charged cell wall bacteria [156][32]. This interaction reduces microbial cell membranes’ permeability [157][33]. In addition, another mechanism has been proposed in which the chitosan selectively binds with metals, inhibiting various metabolic enzymes of microbial cells by blocking their active centers and reducing microbial growth [158][34]. It has been determined that small-sized chitosan can block RNA and protein synthesis, thus inhibiting bacterial growth [159][35]. Characteristics such as molecular weight, degree of deacetylation, and environmental conditions of experimentation, such as pH, temperature, and ionic strength, also influence the antimicrobial capacity of chitosan [160,161,162,163][36][37][38][39].
Chitosan can serve as a matrix for the deposition of antibacterial agents. For instance, including oil compounds could improve the surface adhesive properties and prolong the safety of foods [84][40]. EOs have been used as antimicrobial agents in a biopolymeric matrix based on chitosan NPs. The antimicrobial activity depends on the volatility, the release rate of EOs, and the matrix. The incorporation of EOs can be directly or by encapsulation. Different EOs such as rosemary, tea, tree, clove, oregano, and eucalypt have been evaluated for food packaging and showed improved functional properties [164][41], as described in Figure 42. Chitosan NPs and nanofibers were studied as nanocarriers in conjunction with nisin, antimicrobial peptide temporin B, and Cinnamaldehyde as antimicrobial agents [80][42].
Figure 42.
 Schematic representation of chitosan-essential oil films preparation for food packing.
Similarly, montmorillonite and rosemary essential oil were incorporated in a chitosan matrix to enhance the antimicrobial properties against Listeria monocytogenes (L. monocytogenes) and Streptococcus agalactiae. The results showed that the antimicrobial films improved by rosemary essential oil incorporation [81][43]. In general, adding EOs into chitosan improves chitosan’s effectiveness against some bacteria commonly found in food.
A higher antimicrobial effect was reported in chitosan films with grape seed extract and carvacrol microcapsules (CMF) than in control samples (CS) or just the chitosan control film (CCF). These samples were tested in salmon packed to increase the shelf-life of refrigerated salmon to 4–7 days of storage [165][44]. Another research conducted the antimicrobial activity evaluated in bio-nanocomposites based on graphene oxide (GO) and chitosan. The studied material decreases the growth of S aureus and E. coli. The composite release reactive oxygen species (ROS), which increase the bactericidal properties of the bio-nanocomposites [82][45]. Ao et al. reported the hydroxypropyltrimethyl ammonium chloride chitosan (HACC) based multilayer modified plasma-sprayed porous titanium coating generated via the layer-by-layer covalent-immobilized method. They determined the inhibition of the colonization and biofilm formation of several bacterial strains, including S. aureus, methicillin-resistant S. aureus (MSRA), and clinical isolates of methicillin-resistant S. epidermidis (MRSE), in vitro [83][46].

3. Starch-Based Composites

Starch is a semi-crystalline polysaccharide natural polymer with a complex structure that consists of two-component polymers: amylose (AM) and amylopectin (AP) [166][47]. Since most plants contain around 25% amylose and 75% amylopectin, and the ratio of these two related polymers directly influences solution, and structural properties, control of this ratio has a tremendous impact on the properties of starches [167][48]. Depending on its botanical origin (potato, maize, rice, etc.) and genetic background, starch has different chemical structures and functional groups, making it a useful natural polymer for different applications [85][49]. However, starch does not present inherent antimicrobial properties and is commonly used as a carrier [80][42].
The interest in starch is accrued from its high molecular weight and film-forming properties. Using a high molecular weight (37,000 kg mol−1) [168][50] polymer as a carrier of antimicrobial polymer eliminates the problem of leaching via entanglement and other interactions with the baseline polymer [169][51]. The starch polymer was studied with tea polyphenol (TP) for active food packaging, which presents inhibition efficiency against S. aureus and E. coli [86][52]. Biopolymer composite based on starch and carboxymethylcellulose (CMC) was used as an antimicrobial agent polymeric matrix of turmeric oil. It was reported that the increase in the film thickness increases the release of antimicrobial agents for low volumes of turmeric oil [170][53]. Starch has been reported in conjunction with poly-hexamethylene guanidine hydrochloride (PHGH). The composite studied showed surfaces with high antimicrobial potency against Bacillus subtilis (B. subtilis) and E. coli [87][54].
Another area of study is dental health. Starch NPs are used as carriers for curcumin, which has antimicrobial characteristics. Their interactions were studied through molecular dynamics simulation software with which molecular docking was obtained. The experimental and simulated studies reported a minimum inhibitory concentration (MIC) of curcumin against Streptococcus mutans [88][55]. Starch encapsulated biogenic AgNPs were tested to study the starch encapsulation effect. Results showed that encapsulation increases antimicrobial activity and reduces the toxicity of the NPs [89][56]. Chitosan NPs were synthesized via ionic gelation and used to prepare starch-based nanocomposite films. The antimicrobial properties of starch/chitosan NPs films were evaluated in vitro and in vivo against B. cereus, S. aureus, E. coli, and Salmonella typhimurium. Those films containing chitosan NPs were more effective than those with starch-based films [91][57]. Do Evangelho and coworkers reported corn starch films containing orange (Citrus sinensis) essential oil through the casting method. These films showed higher antibacterial activity against S. aureus and L. monocytogenes [92][58].
On the other hand, Li et al. [90][59] reported synthesizing ultrafine composites starch/polyvinyl alcohol, using glutaraldehyde as a cross-linking agent. AgNPs were loaded to the composites to provide a better mechanical and antimicrobial performance against E. coli and S. aureus. The development of this composite suggests a new route for producing less costly antibacterial fiber materials.

4. Collagen-Based Composites

Collagen is a naturally occurring matrix polymer highly conserved across species [171][60]. Collagen is a protein with biological properties that make it a suitable biomaterial for biomedical applications. It is the most abundant animal protein, providing mechanical strength to the tissues. At least 29 types of collagen have been identified in vertebrates [172][61]. Based on the structure, collagens can be classified into different groups such as fibrils, beaded filaments, networks, anchoring fibrils, and fibril-associated collagen [173][62]. Collagen I is the most abundant structural protein of connective tissues such as skin, bone, and tendon [93][63], assembled into fibrils.
Collagen-based materials have received significant attention in medical applications ranging from drug/gene delivery to tissue engineering [100][64] because it possesses outstanding properties such as tensile stiffness to resist plastic deformation and rupture, biocompatibility, biodegradability, and cell growth potential [173,174][62][65]. In addition, collagen can be prepared into cross-linked compacted solids or lattice-like gels [175][66]. Collagen gels are flowable, suggesting the possibility of an easily injectable, biocompatible drug delivery matrix [100][64]. However, collagen has no antimicrobial properties despite its properties [99][67]. Therefore, it is necessary to incorporate antimicrobial agents to obtain the desired antibacterial. For instance, AgNPs present interesting antimicrobial effects. Therefore, AgNPs can be incorporated into a collagen matrix providing bactericidal effects [94][68].
The antibacterial activity can also be bestowed to the collagen by adding NPs such as AgNPs. Recently, the development of a collagen-carboxymethylcellulose biocomposite containing AgNPs for wound dressings was proposed by Neacsu et al. [95][69]. During this experiment, the antimicrobial assessments showed the antimicrobial potential against gram-negative (E. coli), gram-positive (S. aureus) bacteria, and yeast (Candida albicans). These results agreed well with the literature that reports the potential of AgNPs as an antimicrobial agent [96][70]. Alvarez et al. developed silica–collagen type I biocomposite hydrogels loaded with gentamicin and rifamycin to prevent infection in chronic wounds. The biocomposites were evaluated against P. aeruginosa and S. aureus. Nonetheless, only gentamycin-loaded hydrogels showed bactericidal activity [97][71]. Vladkova et al. performed new collagen composites, Collagen/(Silver (Ag)/Reduced Graphene Oxide (RGO)) and Collagen/(Ag/RGO/Silicium oxide (SiO2)). These composites were tested against E. coliS. epidermidisB. cereus, and a fungus C. Lusistaniae. The biological activity found for the Collagen/(Ag/RGO/SiO2) composites is better expressed than that of Collagen/(Ag/RGO) composites with the same level of antimicrobial agent loading [98][72]. Thymol is an antimicrobial compound in the composition of thyme and oregano EOs [176][73]. The antimicrobial activity of collagen/thymol films was studied for wound dressing applications by Michalska-Sionkowska et al. The bacterial tests showed the antimicrobial efficiency against E. coli, B. subtilis, Enterobacter aerogenes, Candida albicans, and S. aureus, the latter being the most sensitive microorganisms to thymol action [99][67].

5. Gelatin-Based Composites

Gelatin is a nontoxic natural biomacromolecule comprised of bioactive polypeptides derived from collagen in animal skin, bones, and connective tissues [177][74]. It is a protein obtained through controlled partial hydrolysis of collagen [101][75]. Gelatin has many glycine, proline, and 4-hydroxy proline residues [178][76]. Depending on the process employed, two types of gelatin can be obtained: type A gelatin produced by acid hydrolysis and type B obtained by an alkaline or lime process [101][75]. Gelatin has long attracted interest in food, packaging, pharmaceutical, and photographic industries, because of its physical and functional properties such as the reversible gel-to-sol transition of aqueous solution; viscosity behavior; protective colloid function, biodegradability, and solubility in hot water but insolubility in cold water. The gelatin-based film has a suitable matrix and compatibility that allows it to act as a medium for incorporating antimicrobial and antioxidant agents [179][77]. There is a considerable number of publications on the preparation of gelatin-based films with antimicrobial activity by incorporating naturally occurring and synthetic antimicrobials such as organic acids [180][78], proteins [181][79], enzymes [182][80], chelating agents [183][81], and EOs [184][82]. For instance, the effect of incorporating tannic acid (TA) and cellulose nanocrystals (CNC) on gelatin films was evaluated by Leite et al. [102][83]. The gelatin films containing nonoxidized TA and CNC (G-nTA-CNC) exhibited antimicrobial activity against S. aureus and E. coli due to the incorporation of TA. Moreover, G-nTA-CNC films showed an improvement in the gelatin’s antioxidant capacity antioxidant capacity, UV barrier, tensile strength, and water vapor barrier properties. Thus, the resulting approach is suitable for different applications, particularly food packaging. Developing wound dressing loaded with antimicrobial agents has also received much interest in reducing wound bacterial colonization [107][84]Figure 53 shows a schematic illustration for the design of bioactive agent-loaded gelatin-based materials by electrospinning for the wound healing process. Recently, the design of a copper peroxide-loaded gelatin sponge with pH-controllable •OH delivery and effective antimicrobial activity for wound healing was reported. The experiments showed that the as-prepared wound dressing could release •OH, specifically in the bacterial-infected skin wound. In addition, in vitro experiments revealed that the wound dressing has good bactericidal properties against E. coli, S. aureus, and P. aeruginosa [103][85]. Pereda et al. reported the synthesis of biodegradable composite films based on gelatin and chitosan. Composite obtained showed a uniformity due to a compact structure indicating good compatibility between components, which could interact by strong hydrogen bonding. The researchers tested these films against E. coli and L. monocytogenes strains. However, only E. coli resulted be sensitive to the gelatin-chitosan composite [104][86]. Thongsrikhem and coworkers developed an antibacterial gelatin-bacterial cellulose nanocomposite (GCB) film using cinnamaldehyde as a crosslinker and an antibacterial additive. These films were evaluated using S. aureus and E. coli, resulting in a vigorous antibacterial activity against both bacteria strains [105][87]. In addition, Roy et al. synthesized Gelatin-based multifunctional composite films reinforcing various amounts of copper sulfide nanoparticles (CuSNPs). The gelatin/CuSNP composite film presented effective antibacterial activity against E. coli and some activity against L. monocytogenes, suggesting their use in food packaging [106][88].
Figure 53.
 Representation of functionalized gelatin-based nanofibers preparation for wound treatment applications.

6. Hyaluronic Acid-Based Composites

Hyaluronic acid (HA) is a natural polymeric polysaccharide that contains N-acetyl glucosamine and glucuronic acid groups [185][89]. It is present in nature, mainly in mammalian connective tissues. HA is a highly reactive, biocompatible, biodegradable, no-inflammatory, and non-toxic polymer. However, HA has poor biomechanical properties in its native form, and various chemical modifications have been devised to provide mechanically and chemically robust materials [186][90]. HA can be cross-linked or conjugated with assorted biomacromolecules, and it is optimal to encapsulate different active agents [187][91]. Hyaluronic acid hydrogels are readily fabricated as microspheres, sponges, and fibers depending on the intended application [188][92]. However, unmodified HA has a poor residence time in vivo, which can be tailored via cross-linking reactions [189][93]. Among various polymers tested as antibacterial coatings, HA and some of its composites offer a well-established long-term safety profile and a proven ability to reduce bacterial adhesion and biofilm formation [190][94]. HA can interfere with bacterial adhesion to a cellular substrate concentration-dependent [191][95]. HA is bacteriostatic but not bactericidal and exhibits dose-dependent effects on different microorganisms in the planktonic phase [189][93]. HA and its derivate may offer a solution and long-term safety with a known ability to retard bacterial adhesion and biofilm formation [192][96]. However, some studies have shown that the bacteriostatic effect of soluble HA in vitro may be attributed to the saturation of the bacterial hyaluronidase by an excess of HA in the medium [193][97]. To impart antimicrobial properties, the polymeric matrix is commonly functionalized with antimicrobial agents such as quaternary ammonium compounds (QACs), improving antimicrobial efficiency through a contact killing mechanism [108][98]. The surface-functionalized scheme is present in Figure 64. HA carboxylic acid groups are modified by ester formation, while hydroxyl groups can be modified utilizing glutaraldehyde [194][99]. It is applied in ophthalmic treatments as a visual carrier material in a long-term antibiotic release.
Figure 64.
 Surface functionalization of biopolymer.
Multiple antimicrobial agents have been studied with HA as a composite to improve antimicrobial activity. For instance, ciprofloxacin and vancomycin were used in HA particles against P. aeruginosa, S. aureus, and B. subtilis [109][100]. In addition, it is reported that antimicrobial multilayers based on HA and chitosan developed onto the activated surface of polyethylene terephthalate, obtain polyelectrolyte multilayers (PEMs). Triclosan (TRI) and rifampicin (RIF) were used as bactericidal and antibiotic agents, respectively [195][101]. A composite of hyaluronic acid-oleylamine (HA-OLA) was studied as a promising nano-carrier to delivery agents for the treatment of bacterial/Methicillin-resistant Staphylococcus aureus (MRSA) infections [110][102]. D-α-tocopherol polyethylene glycol 1000 succinate (TPGS)-poly(lactic-co-glycolic acid) (PLGA) with azithromycin are studied to improve the antimicrobial activity against P. aeruginosa [111][103]. Nisin (an antimicrobial peptide) has been attached to HA to obtain an antimicrobial biopolymer under solution or gel, showing a great activity against S. epidermidisS., and P. aeruginosa bacteria [112][104]. Harris and Richard coated titanium surfaces with HA for orthopedic applications, showing a decrease in S. aureus attachment on the surface [113][105]. Felgueiras et al. use HA as an antifouling agent, previously modified with thiol groups (HA-SH), using polydopamine as the binding agent. The polyurethane films were coated with the HA. Octadecyl acrylate was subsequently used to bind thiol groups to attract albumin, allowing the system to selectively bind albumin, a protein responsible for bacterial adhesion, thus granting an effective antimicrobial activity. The results showed that these films decreased the S. aureus adhesion [114][106].

7. Alginates-Based Composites

Alginate (ALG) is a natural polymer comprising β-D-mannuronic acid and α-L- guluronic acid extracted from brown seaweed [115][107]. This biomaterial exhibits several properties, including biocompatibility, gelation capability, low toxicity [196][108], mild gelation conditions, and simple modifications to prepare alginate derivatives with new properties [197][109], suggesting its use in biomedical and food industry applications. Alginate can absorb water and body fluids up to 20 times its weight, resulting in a hydrophilic gel [198][110]. The formed gel is weak, but it maintains a moist wound healing environment [198][110]. Linear Alginate polymer chains contain multiple carboxyl groups that can bind to divalent cations (Ca2+, Ba2+) to promote the formation of cross-linked structures [199][111]. Applications within biotechnology and medicine are mainly based on the temperature-independent sol-gel transition in multivalent cations (e.g., Ca2+), making alginates highly suitable as an immobilization matrix for living cells [200][112]. Several studies have investigated the effectiveness of incorporating antimicrobial agents such as EOs and NPs into alginate-based materials to induce an antimicrobial activity. Ahmed and Boateng reported the development of antimicrobial films for treating bacterial infections [116][113]. The calcium alginate films were loaded with ciprofloxacin and tested against E. coli, S. aureus, and P. aeruginosa. The results indicated a bacterial kill within 24 h and were highly biocompatible with human keratinocyte cells. In another study, biodegradable alginate films were prepared by adding zinc oxide nanoparticles (ZnONPs) and citronella essential oil (CEO) for cheese packaging. The ZnONPs act as a reinforcing agent and arrange the alginate polymer chains around them (Figure 75). Microbiological studies revealed a synergic effect between antibacterial activities of ZnO and CEO against two gram-negative (E. coli and Salmonella typhi) and two gram-positive (B. cereus and S. aureus) bacterial strains. Moreover, alginate/ZnO/CEO films showed better UV light barrier properties and lowered water vapor permeability (WVP) than pure alginate film [117][114]. Very similar research was reported by using spherical AgNPs and lemongrass essential oil (LGO) as antimicrobial agents, and the result indicated the feasibility of using alginate/Ag NPs/LGO films as antibacterial packaging to preserve the color, surface texture, and softness of cheese for 14 days [118][115].
Figure 75. Schematic representation of alginate films with ZnONPs. Reprinted with permission from ref [182], MDPI, 2022.
 Schematic representation of alginate films with ZnONPs. Reprinted with permission from ref [80], MDPI, 2022.
Furthermore, new and affordable alternatives have been proposed to prevent microbial infections. In a recent study, incorporating a low percentage of carbon nanofibers enhanced the antibacterial properties of alginate films against the Gram-positive bacterial model S. aureus. In addition, the films showed no cytotoxicity, suggesting their potential use in antimicrobial biomedical materials [119][116]. Alginate fibers cross-linked with zinc ions have also been proposed for wound dressings [185][89]. Zinc ions may generate immunomodulatory and antimicrobial effects, enhanced keratinocyte migration, and increased endogenous growth factors [120][117]. Asadpoor et al. developed films containing alginate oligosaccharides (AOS) as an alternative for antibiotic treatment. AOS decreases the biofilm of Streptococcus agalactiae and S. aureus strains by determining the MIC [121][118].

8. Fibrin-Based Composites

Fibrin, derived from critical proteins involved in blood clotting (fibrinogen and thrombin), is a self-assembling biopolymer [122][119]. Fibrin is a critical component of the blood clot that accelerates wound healing, prevents hemorrhage, and protects against bacterial infection [123][120]. Fibrin alone, or in combination with other biomaterials, was employed as a biological scaffold to promote stem or primary cells to regenerate [129][121]. In comparison to alginate-only gel-laden constructs, fibrin has an advantage in cytocompatibility due to cell adhesion moieties within the fibrin structure [201][122]. Several studies have reported the antimicrobial effect of leukocyte- and platelet-rich fibrin (L-PRF) against periodontal pathogens. Castro et al. assessed the antimicrobial properties of L-PRF against pathogens grown on agar plates and in planktonic cultures. A potent inhibition was found against Prevotella intermedia, Fusobacterium nucleatumAggregatibacter actinomycetemcomitans, and especially against Porphyromonas gingivalis [128][123]. The research conducted by Venante and co-workers [124] exhibited the effectiveness of fibrin biopolymer incorporating antimicrobial agents such as digluconate chlorhexidine and Punica granatum alcoholic extract to prevent the development of Candida albicans biofilm. In vitro results displayed the inhibition of the growth of C. albicans biofilm on poly(methyl methacrylate) (PMMA) substrates for up to 72 h, which suggests the excellent performance of the modified fibrin biopolymer as a drug delivery system, preventing the formation of denture biofilm. Fibrin sealant was used as a matrix for teicoplanin as an antimicrobial carrier applied externally to control infection sites [125]. Vancomycin impregnated fibrin sealant was developed to measure antibacterial activity and antibiotic release. This study uses fibrin sealant as a topical hemostat for post-operatory treatments in surgical fields [126]. AgNPs were studied in metal/fibrin nanocomposites, recognized as suitable materials for wound healing. AgNPs produce an antimicrobial effect related to the easy oxidation of silver. The action over the bacteria occurs due to the interaction between AgNPs/Ag+ and the cell membrane of the bacteria. The reaction was tested against E. coli and S. aureus [127]. A bioartificial human dermis substitute was developed for the treatment of infected wounds. It was based on a fibrin-agarose matrix with sodium colistimethate (SCM) and amikacin (AMK) as antimicrobial agents [202][128].

References

  1. Brigham, C. Biopolymers. In Green Chemistry, 1st ed.; Török, B., Dransfield, T., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 753–770. ISBN 9780128095492.
  2. Yahya, E.; Jummaat, F.; Amirul, A.; Adnan, A.; Olaiya, N.; Abdullah, C.; Rizal, S.; Mohamad Haafiz, M.; Khalil, H. A Review on Revolutionary Natural Biopolymer-Based Aerogels for Antibacterial Delivery. Antibiotics 2020, 9, 648.
  3. Aravamudhan, A.; Ramos, D.; Nada, A.; Kumbar, S. Natural Polymers. In Natural and Synthetic Biomedical Polymers, 1st ed.; Kumbar, S., Laurencin, C.T., Deng, M., Eds.; Elsevier Science: Amsterdam, The Netherlands, 2014; pp. 67–89. ISBN 9780123972903.
  4. Trejo-González, L.; Rodríguez-Hernández, A.; del Rocío López-Cuellar, M.; Martínez-Juárez, V.; Chavarría-Hernández, N. Antimicrobial pectin-gellan films: Effects on three foodborne pathogens in a meat medium, and selected physical-mechanical properties. CYTA J. Food 2018, 16, 469–476.
  5. Muñoz-Bonilla, A.; Echeverria, C.; Sonseca, Á.; Arrieta, M.; Fernández-García, M. Bio-Based Polymers With Antimicrobial Properties Towards Sustainable Development. Materials 2019, 12, 641.
  6. Carvalho, J.; Silva, A.; Silvestre, A.; Freire, C.; Vilela, C. Spherical Cellulose Micro and Nanoparticles: A Review of Recent Developments and Applications. Nanomaterials 2021, 11, 2744.
  7. Norrrahim, M.; Nurazzi, N.; Jenol, M.; Farid, M.; Janudin, N.; Ujang, F.; Yasim-Anuar, T.; Syed Najmuddin, S.; Ilyas, R. Emerging development of nanocellulose as an antimicrobial material: An overview. Mater. Adv. 2021, 2, 3538–3551.
  8. Liu, Y.; Ahmed, S.; Sameen, D.; Wang, Y.; Lu, R.; Dai, J.; Li, S.; Qin, W. A Review of Cellulose and Its Derivatives in Biopolymer-Based for Food Packaging Application. Trends Food Sci. Technol. 2021, 112, 532–546.
  9. Szostak-Kotowa, J. Biodeterioration of textiles. Int. Biodeterior. Biodegrad. 2004, 53, 165–170.
  10. Vigo, T.L. Protection of Textiles from Biological Attack. In Handbook of Fibre Science and Technology; Lewin, M., Sello, S.B., Eds.; Marcel Dekker, INC.: New City, NY, USA, 1983; Volume 2, pp. 367–426. ISBN 9781351442732.
  11. Schindler, W.; Hauser, P. Chemical Finishing of Textiles, 1st ed.; Schindler, W., Hauser, P., Eds.; Woodhead Publishing: Cambridge, UK, 2004; pp. 165–174. ISBN 9781845690373.
  12. Simončič, B.; Tomšič, B. Recent Concepts of Antimicrobial Textile Finishes. In Textile Finishing: Recent Developments and Future Trends; Mital, K.L., Bahners, T., Eds.; John Wiley & Sons: Hoboken, NJ, USA, 2017; Volume 1, pp. 3–68. ISBN 978-1-119-42685-1.
  13. Malis, D.; Jeršek, B.; Tomšič, B.; Štular, D.; Golja, B.; Kapun, G.; Simončič, B. Antibacterial Activity and Biodegradation of Cellulose Fiber Blends with Incorporated ZnO. Materials 2019, 12, 3399.
  14. Gopakumar, D.A.; Thomas, S.; Grohens, Y. Nanocelluloses as Innovative Polymers for Membrane Applications. In Multifunctional Polymeric Nanocomposites Based on Cellulosic Reinforcements, 1st ed.; Puglia, D., Fortunati, E., Kenny, J.M., Eds.; William Andrew Publishing: Norwich, NY, USA, 2016; pp. 253–275. ISBN 9780323417396.
  15. Kumar, R.; Kumar, G. Nanocellulose: Fascinating and sustainable nanomaterial for papermaking. In Nanotechnology in Paper and Wood Engineering, 1st ed.; Bhat, R., Kumar, A., Nguyen, T.A., Sharma, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 389–407. ISBN 9780323859639.
  16. Alvarado, D.; Argyropoulos, D.; Scholle, F.; Peddinti, B.; Ghiladi, R. A facile strategy for photoactive nanocellulose-based antimicrobial materials. Green Chem. 2019, 21, 3424–3435.
  17. Rashki, S.; Shakour, N.; Yousefi, Z.; Rezaei, M.; Homayoonfal, M.; Khabazian, E.; Atyabi, F.; Aslanbeigi, F.; Safaei Lapavandani, R.; Mazaheri, S.; et al. Cellulose-Based Nanofibril Composite Materials as a New Approach to Fight Bacterial Infections. Front. Bioeng. Biotechnol. 2021, 9, 732461.
  18. Littunen, K.; Snoei de Castro, J.; Samoylenko, A.; Xu, Q.; Quaggin, S.; Vainio, S.; Seppälä, J. Synthesis of cationized nanofibrillated cellulose and its antimicrobial properties. Eur. Polym. J. 2016, 75, 116–124.
  19. Esmaeili, E.; Eslami-Arshaghi, T.; Hosseinzadeh, S.; Elahirad, E.; Jamalpoor, Z.; Hatamie, S.; Soleimani, M. The biomedical potential of cellulose acetate/polyurethane nanofibrous mats containing reduced graphene oxide/silver nanocomposites and curcumin: Antimicrobial performance and cutaneous wound healing. Int. J. Biol. Macromol. 2020, 152, 418–427.
  20. Han, Y.; Yu, M.; Wang, L. Physical and antimicrobial properties of sodium alginate/carboxymethyl cellulose films incorporated with cinnamon essential oil. Food Packag. Shelf Life 2018, 15, 35–42.
  21. Orlando, I.; Basnett, P.; Nigmatullin, R.; Wang, W.; Knowles, J.; Roy, I. Chemical Modification of Bacterial Cellulose for the Development of an Antibacterial Wound Dressing. Front. Bioeng. Biotechnol. 2020, 8, 557885.
  22. Silva, Â.; Duarte, A.; Sousa, S.; Ramos, A.; Domingues, F. Characterization and antimicrobial activity of cellulose derivatives films incorporated with a resveratrol inclusion complex. LWT 2016, 73, 481–489.
  23. Arakere, U.; Jagannath, S.; Krishnamurthy, S.; Chowdappa, S.; Konappa, N. Microbial bio-pesticide as sustainable solution for management of pests. In Biopesticides, 1st ed.; Rakshit, A., Meena, V.S., Abhilash, P.C., Sarma, B.K., Singh, H.B., Fraceto, L., Parihar, M., Kumar, A., Eds.; Woodhead Publishing: Cambridge, UK, 2022; pp. 183–200. ISBN 9780128236147.
  24. Moeini, A.; Pedram, P.; Makvandi, P.; Malinconico, M.; Gomez d’Ayala, G. Wound healing and antimicrobial effect of active secondary metabolites in chitosan-based wound dressings: A review. Carbohydr. Polym. 2020, 233, 115839.
  25. Dutta, P.; Tripathi, S.; Mehrotra, G.; Dutta, J. Perspectives for chitosan based antimicrobial films in food applications. Food Chem. 2009, 114, 1173–1182.
  26. Panda, P.; Yang, J.; Chang, Y.; Su, W. Modification of different molecular weights of chitosan by p-Coumaric acid: Preparation, characterization and effect of molecular weight on its water solubility and antioxidant property. Int. J. Biol. Macromol. 2019, 136, 661–667.
  27. Samadi, F.; Mohammadi, Z.; Yousefi, M.; Majdejabbari, S. Synthesis of raloxifene–chitosan conjugate: A novel chitosan derivative as a potential targeting vehicle. Int. J. Biol. Macromol. 2016, 82, 599–606.
  28. Liakos, E.V.; Lazaridou, M.; Michailidou, G.; Koumentakou, I.; Lambropoulou, D.A.; Bikiaris, D.N.; Kyzas, G.Z. Chitosan Adsorbent Derivatives for Pharmaceuticals Removal from Effluents: A Review. Macromol 2021, 1, 130–154.
  29. Barbosa, M.; Pêgo, A.; Amaral, I. Chitosan. Compr. Biomater. 2011, 2, 221–237.
  30. Nurunnabi, M.; Revuri, V.; Huh, K.; Lee, Y. Polysaccharide based nano/microformulation: An effective and versatile oral drug delivery system. In Nanostructures for Oral Medicine, 1st ed.; Andronescu, E., Grumezescu, A.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 409–433. ISBN 9780323477215.
  31. Jean, M.; Alameh, M.; De Jesus, D.; Thibault, M.; Lavertu, M.; Darras, V.; Nelea, M.; Buschmann, M.; Merzouki, A. Chitosan-based therapeutic nanoparticles for combination gene therapy and gene silencing of in vitro cell lines relevant to type 2 diabetes. Eur. J. Pharm. Sci. 2012, 45, 138–149.
  32. Das, B.; Patra, S. Antimicrobials. In Nanostructures for Antimicrobial Therapy, 1st ed.; Ficai, A., Grumezescu, A.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 1–22. ISBN 9780323461511.
  33. Kaczmarek, M.; Struszczyk-Swita, K.; Li, X.; Szczęsna-Antczak, M.; Daroch, M. Enzymatic Modifications of Chitin, Chitosan, and Chitooligosaccharides. Front. Bioeng. Biotechnol. 2019, 7, 243.
  34. Kumar, S.; Mukherjee, A.; Dutta, J. Chitosan based nanocomposite films and coatings: Emerging antimicrobial food packaging alternatives. Trends Food Sci. Technol. 2020, 97, 196–209.
  35. Tsai, G.; Su, W. Antibacterial Activity of Shrimp Chitosan against Escherichia coli. J. Food Prot. 1999, 62, 239–243.
  36. Guarnieri, A.; Triunfo, M.; Scieuzo, C.; Ianniciello, D.; Tafi, E.; Hahn, T.; Zibek, S.; Salvia, R.; De Bonis, A.; Falabella, P. Antimicrobial properties of chitosan from different developmental stages of the bioconverter insect Hermetia illucens. Sci. Rep. 2022, 12, 8084.
  37. Bakshi, P.; Selvakumar, D.; Kadirvelu, K.; Kumar, N. Chitosan as an environment friendly biomaterial—A review on recent modifications and applications. Int. J. Biol. Macromol. 2020, 150, 1072–1083.
  38. Lopez-Moya, F.; Suarez-Fernandez, M.; Lopez-Llorca, L. Molecular Mechanisms of Chitosan Interactions with Fungi and Plants. Int. J. Mol. Sci. 2019, 20, 332.
  39. Li, J.; Wu, Y.; Zhao, L. Antibacterial activity and mechanism of chitosan with ultra high molecular weight. Carbohydr. Polym. 2016, 148, 200–205.
  40. Moeini, A.; Germann, N.; Malinconico, M.; Santagata, G. Formulation of secondary compounds as additives of biopolymer-based food packaging: A review. Trends Food Sci. Technol. 2021, 114, 342–354.
  41. Varghese, S.; Siengchin, S.; Parameswaranpillai, J. Essential oils as antimicrobial agents in biopolymer-based food packaging—A comprehensive review. Food Biosci. 2020, 38, 100785.
  42. Bahrami, A.; Delshadi, R.; Assadpour, E.; Jafari, S.; Williams, L. Antimicrobial-loaded nanocarriers for food packaging applications. Adv. Colloid Interface Sci. 2020, 278, 102140.
  43. Abdollahi, M.; Rezaei, M.; Farzi, G. A novel active bionanocomposite film incorporating rosemary essential oil and nanoclay into chitosan. J. Food Eng. 2012, 111, 343–350.
  44. Alves, V.; Rico, B.; Cruz, R.; Vicente, A.; Khmelinskii, I.; Vieira, M. Preparation and characterization of a chitosan film with grape seed extract-carvacrol microcapsules and its effect on the shelf-life of refrigerated Salmon (Salmo salar). LWT 2018, 89, 525–534.
  45. Joz Majidi, H.; Babaei, A.; Arab Bafrani, Z.; Shahrampour, D.; Zabihi, E.; Jafari, S. Investigating the best strategy to diminish the toxicity and enhance the antibacterial activity of graphene oxide by chitosan addition. Carbohydr. Polym. 2019, 225, 115220.
  46. Ao, H.; Yang, S.; Nie, B.; Fan, Q.; Zhang, Q.; Zong, J.; Guo, S.; Zheng, X.; Tang, T. Improved antibacterial properties of collagen I/hyaluronic acid/quaternized chitosan multilayer modified titanium coatings with both contact-killing and release-killing functions. J. Mater. Chem. B 2019, 7, 1951–1961.
  47. McKeen, L. The effect of heat aging on the properties of sustainable polymers. In The Effect of Long Term Thermal Exposure on Plastics and Elastomers, 2nd ed.; William Andrew: Norwich, NY, USA, 2021; pp. 313–332. ISBN 978-0-323-85437-5.
  48. Winkler, S.; Kaplan, D. Biosynthesized Materials: Properties and Processing. In Encyclopedia of Materials: Science and Technology, 2nd ed.; Jürgen, K.H., Cahn, R.W., Flemings, M.C., Ilschner, B., Kramer, E., Mahajan, S., Veyssière, P., Eds.; Pergamon: Oxford, UK, 2001; pp. 609–615. ISBN 9780080431529.
  49. Hutmacher, D. Polymers from Biotechnology. In Encyclopedia of Materials: Science and Technology, 2nd ed.; Jürgen, K.H., Cahn, R.W., Flemings, M.C., Ilschner, B., Kramer, E., Mahajan, S., Veyssière, P., Eds.; Pergamon: Oxford, UK, 2001; pp. 7680–7683. ISBN 9780080431529.
  50. Van Soest, J.; Benes, K.; De Wit, D.; Vliegenthart, J. The influence of starch molecular mass on the properties of extruded thermoplastic starch. Polymer 1996, 37, 3543–3552.
  51. Ogunsona, E.; Ojogbo, E.; Mekonnen, T. Advanced material applications of starch and its derivatives. Eur. Polym. J. 2018, 108, 570–581.
  52. Feng, M.; Yu, L.; Zhu, P.; Zhou, X.; Liu, H.; Yang, Y.; Zhou, J.; Gao, C.; Bao, X.; Chen, P. Development and preparation of active starch films carrying tea polyphenol. Carbohydr. Polym. 2018, 196, 162–167.
  53. Mustapha, F.; Jai, J.; Nik Raikhan, N.; Sharif, Z.; Yusof, N. Response surface methodology analysis towards biodegradability and antimicrobial activity of biopolymer film containing turmeric oil against Aspergillus niger. Food Control 2019, 99, 106–113.
  54. Ojogbo, E.; Ward, V.; Mekonnen, T. Functionalized starch microparticles for contact-active antimicrobial polymer surfaces. Carbohydr. Polym. 2020, 229, 115422.
  55. Rezapour, N.; Rasekh, B.; Mofradnia, S.; Yazdian, F.; Rashedi, H.; Tavakoli, Z. Molecular dynamics studies of polysaccharide carrier based on starch in dental cavities. Int. J. Biol. Macromol. 2019, 121, 616–624.
  56. Saravanakumar, K.; Sriram, B.; Sathiyaseelan, A.; Mariadoss, A.; Hu, X.; Han, K.; Vishnupriya, V.; MubarakAli, D.; Wang, M. Synthesis, characterization, and cytotoxicity of starch-encapsulated biogenic silver nanoparticle and its improved anti-bacterial activity. Int. J. Biol. Macromol. 2021, 182, 1409–1418.
  57. Shapi’i, R.; Othman, S.; Nordin, N.; Kadir Basha, R.; Nazli Naim, M. Antimicrobial properties of starch films incorporated with chitosan nanoparticles: In vitro and in vivo evaluation. Carbohydr. Polym. 2020, 230, 115602.
  58. do Evangelho, J.; da Silva Dannenberg, G.; Biduski, B.; el Halal, S.; Kringel, D.; Gularte, M.; Fiorentini, A.; da Rosa Zavareze, E. Antibacterial activity, optical, mechanical, and barrier properties of corn starch films containing orange essential oil. Carbohydr. Polym. 2019, 222, 114981.
  59. Li, X.; Zhang, Y.; Kong, W.; Zhou, J.; Hou, T.; Zhang, X.; Zhou, L.; Sun, M.; Liu, S.; Yang, B. Cross-Linking of Centrifugally Spun Starch/Polyvinyl Alcohol (ST/PVA) Composite Ultrafine Fibers and Antibacterial Activity Loaded with Ag Nanoparticles. ACS Omega 2022, 7, 7706–7714.
  60. Cheema, U.; Ananta, M.; Muder, V. Collagen: Applications of a Natural Polymer in Regenerative Medicine. In Regenerative Medicine and Tissue Engineering—Cells and Biomaterials; Eberli, D., Ed.; IntechOpen: London, UK, 2011; ISBN 978-953-51-4450-2.
  61. Balasubramanian, P.; Prabhakaran, M.; Sireesha, M.; Ramakrishna, S. Collagen in Human Tissues: Structure, Function, and Biomedical Implications from a Tissue Engineering Perspective. In Polymer Composites—Polyolefin Fractionation—Polymeric Peptidomimetics—Collagens, 1st ed.; Abe, A., Kausch, H.H., Möller, M., Pasch, H., Eds.; Springer: Berlin/Heidelberg, Germany, 2012; pp. 173–206. ISBN 978-3-642-34330-8.
  62. Lin, K.; Zhang, D.; Macedo, M.; Cui, W.; Sarmento, B.; Shen, G. Advanced Collagen-Based Biomaterials for Regenerative Biomedicine. Adv. Funct. Mater. 2018, 29, 1804943.
  63. Chung, H.; Steplewski, A.; Chung, K.; Uitto, J.; Fertala, A. Collagen Fibril Formation. Int. J. Biol. Chem. 2008, 283, 25879–25886.
  64. David, G. Collagen-based 3D structures—Versatile, efficient materials for biomedical applications. In Biopolymer-Based Formulations, 1st ed.; Pal, K., Banerjee, I., Sarkar, P., Kim, D., Deng, W., Dubey, N.K., Majumder, K., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 881–906. ISBN 9780128168981.
  65. Subhan, F.; Ikram, M.; Shehzad, A.; Ghafoor, A. Marine Collagen: An Emerging Player in Biomedical applications. J. Food Sci. Technol. 2014, 52, 4703–4707.
  66. Patino, M.; Neiders, M.; Andreana, S.; Noble, B.; Cohen, R. Collagen as an Implantable Material in Medicine and Dentistry. J. Oral Implantol. 2002, 28, 220–225.
  67. Michalska-Sionkowska, M.; Walczak, M.; Sionkowska, A. Antimicrobial activity of collagen material with thymol addition for potential application as wound dressing. Polym. Test. 2017, 63, 360–366.
  68. Ge, L.; Xu, Y.; Li, X.; Yuan, L.; Tan, H.; Li, D.; Mu, C. Fabrication of Antibacterial Collagen-Based Composite Wound Dressing. ACS Sustain. Chem. Eng. 2018, 6, 9153–9166.
  69. Neacsu, I.; Leau, S.; Marin, S.; Holban, A.; Vasile, B.; Nicoara, A.; Ene, V.; Bleotu, C.; Albu Kaya, M.; Ficai, A. Collagen-Carboxymethylcellulose Biocomposite Wound-Dressings with Antimicrobial Activity. Materials 2021, 14, 1153.
  70. Anees Ahmad, S.; Sachi Das, S.; Khatoon, A.; Tahir Ansari, M.; Afzal, M.; Saquib Hasnain, M.; Kumar Nayak, A. Bactericidal activity of silver nanoparticles: A mechanistic review. Mater. Sci. Energy Technol. 2020, 3, 756–769.
  71. Alvarez, G.; Hélary, C.; Mebert, A.; Wang, X.; Coradin, T.; Desimone, M. Antibiotic-loaded silica nanoparticle–collagen composite hydrogels with prolonged antimicrobial activity for wound infection prevention. J. Mater. Chem. B 2014, 2, 4660.
  72. Vladkova, T.; Ivanova, I.; Staneva, A.; Albu-Kaya, M.; Shalaby, A.; Moskova-Doumanova, V.; Kostadinova, A. Preparation and Biological Activity of New Collagen Composites, Part III. Collagen/(Ag/RGO) and Collagen/(Ag/RGO/SiO2) Composites. J. Arch. Mil. Med. 2017, 5, e57454.
  73. Gavaric, N.; Mozina, S.; Kladar, N.; Bozin, B. Chemical Profile, Antioxidant and Antibacterial Activity of Thyme and Oregano Essential Oils, Thymol and Carvacrol and Their Possible Synergism. J. Essent. Oil Bear. Plants 2015, 18, 1013–1021.
  74. Narayanaswamy, R.; Kanagesan, S.; Pandurangan, A.; Padmanabhan, P. Basics to different imaging techniques, different nanobiomaterials for image enhancement. In Nanobiomaterials in Medical Imaging, 1st ed.; Grumezescu, A.M., Ed.; William Andrew: Norwich, NY, USA, 2016; Volume 8, pp. 101–129. ISBN 9780323417389.
  75. Deshmukh, K.; Basheer Ahamed, M.; Deshmukh, R.; Khadheer Pasha, S.; Bhagat, P.; Chidambaram, K. Biopolymer Composites With High Dielectric Performance: Interface Engineering. In Biopolymer Composites in Electronics, 1st ed.; Sadasivuni, K., Ponnamma, D., Kim, J., Cabibihan, J., AlMaadeed, M.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 27–128. ISBN 9780081009741.
  76. Pereda, M.; Ponce, A.; Marcovich, N.; Ruseckaite, R.; Martucci, J. Chitosan-gelatin composites and bi-layer films with potential antimicrobial activity. Food Hydrocoll. 2011, 25, 1372–1381.
  77. Said, N.; Howell, N.; Sarbon, N. A Review on Potential Use of Gelatin-based Film as Active and Smart Biodegradable Films for Food Packaging Application. Food Rev. Int. 2021, 1–23.
  78. Clarke, D.; Molinaro, S.; Tyuftin, A.; Bolton, D.; Fanning, S.; Kerry, J. Incorporation of commercially-derived antimicrobials into gelatin-based films and assessment of their antimicrobial activity and impact on physical film properties. Food Control 2016, 64, 202–211.
  79. Jridi, M.; Hajji, S.; Ayed, H.; Lassoued, I.; Mbarek, A.; Kammoun, M.; Souissi, N.; Nasri, M. Physical, structural, antioxidant and antimicrobial properties of gelatin–chitosan composite edible films. Int. J. Biol. Macromol. 2014, 67, 373–379.
  80. Bower, C.; Avena-Bustillos, R.; Olsen, C.; McHugh, T.; Bechtel, P. Characterization of Fish-Skin Gelatin Gels and Films Containing the Antimicrobial Enzyme Lysozyme. J. Food Sci 2006, 71, M141–M145.
  81. Abarca, R.; Medina, J.; Alvarado, N.; Ortiz, P.; Carrillo López, B. Biodegradable gelatin-based films with nisin and EDTA that inhibit Escherichia coli. PLoS ONE 2022, 17, e0264851.
  82. Martucci, J.; Gende, L.; Neira, L.; Ruseckaite, R. Oregano and lavender essential oils as antioxidant and antimicrobial additives of biogenic gelatin films. Ind. Crop. Prod. 2015, 71, 205–213.
  83. Leite, L.; Pham, C.; Bilatto, S.; Azeredo, H.; Cranston, E.; Moreira, F.; Mattoso, L.; Bras, J. Effect of Tannic Acid and Cellulose Nanocrystals on Antioxidant and Antimicrobial Properties of Gelatin Films. ACS Sustain. Chem. Eng. 2021, 9, 8539–8549.
  84. Khosravimelal, S.; Chizari, M.; Farhadihosseinabadi, B.; Moosazadeh Moghaddam, M.; Gholipourmalekabadi, M. Fabrication and characterization of an antibacterial chitosan/silk fibroin electrospun nanofiber loaded with a cationic peptide for wound-dressing application. J. Mater. Sci. Mater. Med. 2021, 32, 114.
  85. Cui, H.; Liu, M.; Yu, W.; Cao, Y.; Zhou, H.; Yin, J.; Liu, H.; Que, S.; Wang, J.; Huang, C.; et al. Copper Peroxide-Loaded Gelatin Sponges for Wound Dressings with Antimicrobial and Accelerating Healing Properties. ACS Appl. Mater. Interfaces 2021, 13, 26800–26807.
  86. Zeng, W.; Li, Y.; Wang, Y.; Cao, Y. Tissue Engineering of Blood Vessels. In Encyclopedia of Tissue Engineering and Regenerative Medicine, 1st ed.; Reis, R., Ed.; Academic Press: Cambridge, MA, USA, 2019; Volume 1, pp. 413–424. ISBN 9780128137000.
  87. Thongsrikhem, N.; Taokaew, S.; Sriariyanun, M.; Kirdponpattara, S. Antibacterial activity in gelatin-bacterial cellulose composite film by thermally crosslinking with cinnamaldehyde towards food packaging application. Food Packag. Shelf Life 2022, 31, 100766.
  88. Roy, S.; Rhim, J.-W. Gelatin-Based Film Integrated with Copper Sulfide Nanoparticles for Active Packaging Applications. Appl. Sci. 2021, 11, 6307.
  89. Guvendiren, M.; Purcell, B.; Burdick, J. Photopolymerizable Systems. In Polymer Science: A Comprehensive Reference, 1st ed.; Matyjaszewski, K., Möller, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2012; Volume 9, pp. 413–438. ISBN 9780080878621.
  90. Burd, A. Hyaluronan and Scarring. In Chemistry and Biology of Hyaluronan, 1st ed.; Garg, H.G., Hales, C.A., Eds.; Elsevier Science: Amsterdam, The Netherlands, 2004; pp. 367–394. ISBN 9780080472225.
  91. Bayer, I. Hyaluronic Acid and Controlled Release: A Review. Molecules 2020, 25, 2649.
  92. Burdick, J.; Stevens, M. Biomedical hydrogels. Biomaterials, Artificial Organs and Tissue Engineering, 1st ed.; Hench, L., Jones, J., Eds.; Woodhead Publishing: Cambridge, UK, 2005; pp. 107–115. ISBN 9781845690861.
  93. Zamboni, F.; Okoroafor, C.; Ryan, M.; Pembroke, J.; Strozyk, M.; Culebras, M.; Collins, M. On the bacteriostatic activity of hyaluronic acid composite films. Carbohydr. Polym. 2021, 260, 117803.
  94. Romanò, C.; Vecchi, E.; Bortolin, M.; Morelli, I.; Drago, L. Hyaluronic Acid and Its Composites as a Local Antimicrobial/Antiadhesive Barrier. J. Bone Jt. Infect. 2017, 2, 63–72.
  95. Drago, L.; Cappelletti, L.; De Vecchi, E.; Pignataro, L.; Torretta, S.; Mattina, R. Antiadhesive and antibiofilm activity of hyaluronic acid against bacteria responsible for respiratory tract infections. APMIS 2014, 122, 1013–1019.
  96. Zamboni, F.; Wong, C.; Collins, M. Hyaluronic acid association with bacterial, fungal and viral infections: Can hyaluronic acid be used as an antimicrobial polymer for biomedical and pharmaceutical applications? Bioact. Mater. 2022, 19, 458–473.
  97. Carlson, G.; Dragoo, J.; Samimi, B.; Bruckner, D.; Bernard, G.; Hedrick, M.; Benhaim, P. Bacteriostatic properties of biomatrices against common orthopaedic pathogens. Biochem. Biophys. Res. Commun. 2004, 321, 472–478.
  98. Delfi, M.; Ghomi, M.; Zarrabi, A.; Mohammadinejad, R.; Taraghdari, Z.; Ashrafizadeh, M.; Zare, E.; Agarwal, T.; Padil, V.; Mokhtari, B.; et al. Functionalization of Polymers and Nanomaterials for Biomedical Applications: Antimicrobial Platforms and Drug Carriers. Prosthesis 2020, 2, 12.
  99. Schanté, C.; Zuber, G.; Herlin, C.; Vandamme, T. Chemical modifications of hyaluronic acid for the synthesis of derivatives for a broad range of biomedical applications. Carbohydr. Polym. 2011, 85, 469–489.
  100. Zhang, Z.; Suner, S.; Blake, D.; Ayyala, R.; Sahiner, N. Antimicrobial activity and biocompatibility of slow-release hyaluronic acid-antibiotic conjugated particles. Int. J. Pharm. 2020, 576, 119024.
  101. Pérez-Álvarez, L.; Ruiz-Rubio, L.; Azua, I.; Benito, V.; Bilbao, A.; Vilas-Vilela, J. Development of multiactive antibacterial multilayers of hyaluronic acid and chitosan onto poly (ethylene terephthalate). Eur. Polym. J. 2019, 112, 31–37.
  102. Walvekar, P.; Gannimani, R.; Salih, M.; Makhathini, S.; Mocktar, C.; Govender, T. Self-assembled oleylamine grafted hyaluronic acid polymersomes for delivery of vancomycin against methicillin resistant Staphylococcus aureus (MRSA). Colloids Surf. B Biointerfaces. 2019, 182, 110388.
  103. Kłodzińska, S.; Wan, F.; Jumaa, H.; Sternberg, C.; Rades, T.; Nielsen, H. Utilizing nanoparticles for improving anti-biofilm effects of azithromycin: A head-to-head comparison of modified hyaluronic acid nanogels and coated poly (lactic-co-glycolic acid) nanoparticles. J. Colloid Interface Sci. 2019, 555, 595–606.
  104. Lequeux, I.; Ducasse, E.; Jouenne, T.; Thebault, P. Addition of antimicrobial properties to hyaluronic acid by grafting of antimicrobial peptide. Eur. Polym. J. 2014, 51, 182–190.
  105. Harris, L.; Richards, R. Staphylococcus aureus adhesion to different treated titanium surfaces. J. Mater. Sci. Mater. Med. 2004, 15, 311–314.
  106. Felgueiras, H.; Wang, L.; Ren, K.; Querido, M.; Jin, Q.; Barbosa, M.; Ji, J.; Martins, M. Octadecyl Chains Immobilized onto Hyaluronic Acid Coatings by Thiol–ene “Click Chemistry” Increase the Surface Antimicrobial Properties and Prevent Platelet Adhesion and Activation to Polyurethane. ACS Appl. Mater. Interfaces 2017, 9, 7979–7989.
  107. Preman, N.; Jain, S.; Sanjeeva, S.; Johnson, R. Alginate derived nanoassemblies in drug delivery and tissue engineering. In Polysaccharide Nanoparticles, 1st ed.; Venkatesan, J., Kim, S., Anil, S., Rekha, P.D., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 247–280. ISBN 9780128223567.
  108. Ahmad Raus, R.; Wan Nawawi, W.; Nasaruddin, R. Alginate and alginate composites for biomedical applications. Asian J. Pharm. Sci. 2020, 16, 280–306.
  109. Lee, K.; Mooney, D. Alginate: Properties and biomedical applications. Prog. Polym. Sci. 2012, 37, 106–126.
  110. Aramwit, P. Introduction to biomaterials for wound healing. In Wound Healing Biomaterials, 1st ed.; Agren, M., Ed.; Woodhead Publishing: Cambridge, UK, 2016; Volume 2, pp. 3–38. ISBN 9780081006061.
  111. Hamai, R.; Anada, T.; Suzuki, O. Novel scaffold composites containing octacalcium phosphate and their role in bone repair. In Octacalcium Phosphate Biomaterials, 1st ed.; Suzuki, O., Insley, G., Eds.; Woodhead Publishing: Cambridge, UK, 2020; pp. 121–145. ISBN 9780081025123.
  112. Skjåk-Bræk, G.; Draget, K. Alginates. In Polymer Science: A Comprehensive Reference, 1st ed.; Matyjaszewski, K., Möller, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2012; Volume 10, pp. 213–220.
  113. Ahmed, A.; Boateng, J. Calcium alginate-based antimicrobial film dressings for potential healing of infected foot ulcers. Ther. Deliv. 2018, 9, 185–204.
  114. Motelica, L.; Ficai, D.; Oprea, O.; Ficai, A.; Trusca, R.; Andronescu, E.; Holban, A. Biodegradable Alginate Films with ZnO Nanoparticles and Citronella Essential Oil—A Novel Antimicrobial Structure. Pharmaceutics 2021, 13, 1020.
  115. Motelica, L.; Ficai, D.; Oprea, O.; Ficai, A.; Ene, V.; Vasile, B.; Andronescu, E.; Holban, A. Antibacterial Biodegradable Films Based on Alginate with Silver Nanoparticles and Lemongrass Essential Oil–Innovative Packaging for Cheese. Nanomaterials 2021, 11, 2377.
  116. Sanmartín-Santos, I.; Gandía-Llop, S.; Salesa, B.; Martí, M.; Lillelund Aachmann, F.; Serrano-Aroca, Á. Enhancement of Antimicrobial Activity of Alginate Films with a Low Amount of Carbon Nanofibers (0.1% w/w). Appl. Sci. 2021, 11, 2311.
  117. Agren, M. Zinc in Wound Repair. Arch. Dermatol. 1999, 135, 1273–1274.
  118. Asadpoor, M.; Ithakisiou, G.; van Putten, J.; Pieters, R.; Folkerts, G.; Braber, S. Antimicrobial Activities of Alginate and Chitosan Oligosaccharides against Staphylococcus aureus and Group B Streptococcus. Front. Microbiol. 2021, 12, 700605.
  119. Catelas, I. Fibrin. In Comprehensive Biomaterials, 1st ed.; Ducheyne, P., Ed.; Elsevier Science: Amsterdam, The Netherlands, 2011; Volume 2, pp. 303–328. ISBN 9780080552941.
  120. Ahmed, T.; Dare, E.; Hincke, M. Fibrin: A Versatile Scaffold for Tissue Engineering Applications. Tissue Eng. Part B Rev. 2008, 14, 199–215.
  121. Climov, M.; Leavitt, T.; Molnar, J.; Orgill, D. Natural Biomaterials for Skin Tissue Engineering. In Skin Tissue Engineering and Regenerative Medicine, 1st ed.; Albanna, M., Holmes, J.H., Eds.; Academic Press: Cambridge, MA, USA, 2016; pp. 145–161. ISBN 9780128017975.
  122. Carrow, J.; Kerativitayanan, P.; Jaiswal, M.; Lokhande, G.; Gaharwar, A. Polymers for Bioprinting. In Essentials of 3D Biofabrication and Translation, 1st ed.; Atala, A., Yoo, J.J., Eds.; Academic Press: Cambridge, MA, USA, 2015; pp. 229–248. ISBN 9780128010150.
  123. Castro, A.; Herrero, E.; Slomka, V.; Pinto, N.; Teughels, W.; Quirynen, M. Antimicrobial capacity of Leucocyte-and Platelet Rich Fibrin against periodontal pathogens. Sci. Rep. 2019, 9, 8188.
  124. Venante, H.; Chappuis-Chocano, A.; Marcillo-Toala, O.; da Silva, R.; da Costa, R.; Pordeus, M.; Barraviera, B.; Ferreira Junior, R.; Lara, V.; Neppelenbroek, K.; et al. Fibrin Biopolymer Incorporated with Antimicrobial Agents: A Proposal for Coating Denture Bases. Materials 2021, 14, 1618.
  125. Tan, R.; Lee, H.; Ma, H.; Lee, H.; Han, S. Antibacterial Effect of Antibiotic-Saturated Fibrin Sealant; In Vitro Study. J. Wound Manag. Res. 2018, 14, 12–17.
  126. Shin, D.; Sohn, M.; Cho, C.; Koo, H.; Yoon, S. Evaluation of Cumulative and Conditional Antibiotic Release from Vancomycin-Embedded Fibrin Sealant and Its Antibacterial Activity: An In Vitro Study. J. Korean Neurosurg. Soc. 2020, 63, 45–55.
  127. Zahran, M.; Marei, A. Innovative natural polymer metal nanocomposites and their antimicrobial activity. Int. J. Biol. Macromol. 2019, 136, 586–596.
  128. Chato-Astrain, J.; Chato-Astrain, I.; Sánchez-Porras, D.; García-García, Ó.; Bermejo-Casares, F.; Vairo, C.; Villar-Vidal, M.; Gainza, G.; Villullas, S.; Oruezabal, R.; et al. Generation of a novel human dermal substitute functionalized with antibiotic-loaded nanostructured lipid carriers (NLCs) with antimicrobial properties for tissue engineering. J. Nanobiotechnol. 2020, 18, 174.
More
Video Production Service