Next Article in Journal
Power Failure of Mitochondria and Oxidative Stress in Neurodegeneration and Its Computational Models
Next Article in Special Issue
Physicochemical, Antioxidant, and Anti-Inflammatory Properties of Rapeseed Lecithin Liposomes Loading a Chia (Salvia hispanica L.) Seed Extract
Previous Article in Journal
Identifying a Role of Red and White Wine Extracts in Counteracting Skin Aging: Effects of Antioxidants on Fibroblast Behavior
Previous Article in Special Issue
Application of a Natural Antioxidant from Grape Pomace Extract in the Development of Bioactive Jute Fibers for Food Packaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Crustacean Waste-Derived Chitosan: Antioxidant Properties and Future Perspective

1
Laboratory of Neo Natural Farming, Chunnampet, Tamil Nadu 603 401, India
2
Department of Environmental Health Sciences, Konkuk University, Seoul 05029, Korea
3
New Prince Shri Bhavani Senior Secondary School, Ullagaram, Chennai 600 091, India
4
Department of Chemistry, Faculty of Science, Jazan University, Jazan P.O. Box 114, Saudi Arabia
5
Department of Bioresources and Food Science, Institute of Natural Science and Agriculture, Konkuk University, 1 Hwayang-dong, Gwangjin-gu, Seoul 05029, Korea
*
Author to whom correspondence should be addressed.
Antioxidants 2021, 10(2), 228; https://doi.org/10.3390/antiox10020228
Submission received: 31 December 2020 / Revised: 16 January 2021 / Accepted: 21 January 2021 / Published: 3 February 2021
(This article belongs to the Special Issue Extraction of Antioxidants from Food Waste II)

Abstract

:
Chitosan is obtained from chitin that in turn is recovered from marine crustacean wastes. The recovery methods and their varying types and the advantages of the recovery methods are briefly discussed. The bioactive properties of chitosan, which emphasize the unequivocal deliverables contained by this biopolymer, have been concisely presented. The variations of chitosan and its derivatives and their unique properties are discussed. The antioxidant properties of chitosan have been presented and the need for more work targeted towards harnessing the antioxidant property of chitosan has been emphasized. Some portions of the crustacean waste are being converted to chitosan; the possibility that all of the waste can be used for harnessing this versatile multifaceted product chitosan is projected in this review. The future of chitosan recovery from marine crustacean wastes and the need to improve in this area of research, through the inclusion of nanotechnological inputs have been listed under future perspective.

1. Introduction

Chitin is the second most abundant (first is cellulose), non-toxic, biodegradable high molecular weight polymer occurring in nature. It exhibits outstanding chemical and biological properties, such as biocompatibility, non-toxicity, biodegradability, and exceptional adsorption properties. These unique properties have enabled chitin to be used in industrial and biomedical applications. Chitin is the major component of cuticles of insects (cockroach, beetle, true fly, and worm), fungal cell walls (Aspergillus niger, Mucor rouxii, Penicillum notatum, yeast) and green algae [1,2,3,4]. Chitin is a constituent of different exoskeletons of marine arthropods such as crustaceans (crab, shrimp, lobster, krill, crayfish, barnacles), cuttlefish, and squid pen. Chitin is made up of a linear chain of acetylglucosamine groups [5]. With a turnover of 10 billion tons annually [6,7], chitin is one of the most abundant biopolymers. Chitin can be readily obtained by simple extraction [8] and the major source of industrial chitin is derived from wastes of marine food production, mainly crustacean shells, e.g., shrimp, crab, or krill shells [9,10,11]. Crustacean shell is composed of 30–40% proteins, 30–50% mineral salts, and 13–42% of chitin occurring in α-, β-, and γ-forms. In the processing of shrimps for human consumption, between 40 and 50% of the total mass is waste and 40% of this waste is chitin. A small part of the waste is usually dried and utilized as chicken feed [11], while the rest is dumped into the sea, which is one of the main pollutants in coastal areas [12,13]. The utilization of shellfish waste has thus been able to solve environmental problems, being an alternative waste disposal method [14,15]. In practice, a polymer where most residues are acetylated is chitin, while the opposite holds well for chitosan. Several processes have been proposed for the use of shellfish waste for extraction of chitin and chitosan. Chitin extraction involves alkaline extraction of organics and acid solubilization/decomposition of minerals, subsequent to protein removal. During such extractions, the chitin molecule also suffers some structural changes, including a moderate degree of deacetylation. Chitosan, is obtained on substantial deacetylation through alkaline treatment of chitin under severe conditions.
Research in the area of renewable marine byproducts has recently been patronized. The current application of crustacean waste mainly focuses on the value-added production of chitin. Chemical methods of chitin extraction use voluminous hazardous chemicals (NaOH and HCl) that are released into the environment. The abundant and renewable marine processing wastes are commercially exploited for the extraction of chitin. However, the traditional chitin extraction processes employ harsh chemicals at elevated temperatures for a prolonged time which can harm chitin’s physico-chemical properties and are also held responsible for the deterioration of environmental health. There are mainly two chitin extraction methods conducted in the industry, those being chemical or biological. Both extraction strategies of chitin consist of two phases, deproteinization with alkaline treatment at high temperatures followed by demineralization predominantly with dilute hydrochloric acid. The sequence of these two phases is interchangeable depending on the source and the proposed use of chitin. During the chemical process, the protein component in the shells is not recovered, however, this is resolved through enzymatic processing methods resulting in the effective recovery of proteins, chitin, and pigments [16,17]. Chitin powder isolated from crustacean sources has a pale pink color, thus necessitating the bleaching process involving hydrogen peroxide, oxalic acid, or potassium permanganate [18,19]. In order to avoid the use of harsh chemicals, green extraction methods are increasingly gaining popularity due to their environmentally friendly nature. Chitin and its derivatives are widely used in innumerable applications ranging from food, agriculture, biomedicine, pharmaceuticals, and cosmetics to environmental sector. Figure 1 gives an overview of the overall process involved in the recovery of chitosan from crustacean shell wastes.
Biotechnological production of chitin offers new perspectives for the production of highly viscous chitosan, with promising inputs in biomedicine and pharmacy [20,21]. Proteolytic enzymes such as, papain, alkalase, chymotrypsintrypsin, pepsin, devolvase and pancreatin have been employed for the extraction and separation of chitin and proteins from shrimp waste [22,23]. Mhamdi et al. reported the use of serine alkaline proteases from Micromonospora chaiyaphumensis S103 for chitin extraction from shrimp shell (Penaeus kerathurus) waste [24]. Other authors have described that the application of crude digestive alkaline proteases from Portunus segnis led to effective extraction of chitin by deproteinization of blue crab (P. segnis) and shrimp (P. kerathurus) [25]. The most used bacterial strains for fermentation are Lactobacillus sp. strain especially L. plantarum, L. paracasei, and L. helveticus. Recently, Castro et al. extracted and purified chitin from Allopetrolisthes punctatus crabs using Lactobacillus plantarum sp.47, a Gram-positive bacterium isolated from Coho salmon that produces high lactic acid concentrations [26]. For chitin recovery with non-lactic acid bacteria, crustacean shell fermentation bacteria and fungi such as Pseudomonas sp., Bacillus sp. and Aspergillus sp. were used as the inoculum source: Ghorbel-Bellaaj et al. isolated a protease bacterium identified as Pseudomonas aeruginosa A2. Most commercial bacterial proteases are mainly produced by Bacillus sp., and Hajji et al. have extracted from the waste of crab shells chitin and fermented-crab supernatants after fermentation using six different strains of Bacillus. The bioextraction of chitin from crustacean shell wastes has been increasingly researched at the laboratory scale, but not at the commercial level. Bioextraction of chitin is thus a greener, cleaner, eco-friendly and economical process. Specifically, microorganisms-mediated fermentation processes are highly desirable due to easy handling, simplicity, rapidity, controllability through optimization of process parameters, ambient temperature, and negligible solvent consumption, thus reducing environmental impact and costs. Driven by reduced energy, wastewater, or solvent, advances in biological extraction of chitin along with valuable byproducts will have high economic and environmental impact.
However, chitosan properties and applications and recovery methods have been already reviewed. The following review highlights the recovery of chitosan from crustacean wastes and emphasizes on the need to seek out “green” extraction methods. The available data on the antioxidant properties of chitosan have been reviewed and the need to expand the research focus in this area of research for harnessing the full potential of the antioxidant properties of chitosan has been proposed. Nanotechnological inputs into areas of chitosan recovery and furtherance of biological activity (especially antioxidant activity) of chitosan have been highlighted as a necessary improvement.

2. Chitin–Chitosan Metamorphosis

The seafood processing industry produces large quantities of byproducts and discards such as heads, tails, skins, scales, viscera, backbones, and shells of marine organisms. Although these are waste residues, they still are an excellent source of lipids, proteins, pigments, and small molecules, and moreso a source of chitinous materials. One of the limitations in the use of chitin on a large-scale is its water insolubility, this is why water-soluble derivatives have been sought after. Chitosan is the most important of these. Chitosan is obtained from chitin by a process called deacetylation, whereby acetyl groups (CH3-CO) are removed resulting in a molecule that is soluble in most diluted acids [27]. The deacetylation process releases amine groups (-NH2) rendering chitosan a cationic nature. Chitosan, a linear polysaccharide is made up of β-(1–4)-linked d-glucosamine and N-acetyl-d-glucosamine moieties [28,29,30,31,32,33,34]. Chitosan is derived from chitin by chemical or enzymatic deacetylations. Although chemical deacetylation is preferentially cheaper and warrants mass production, the major disadvantage is the energy consumption and high environmental pollution risks. Alternatives in the form of enzymatic methods that utilize chitin deacetylases have been explored via enzymatic deacetylation of chitin. Research has identified that selected fungal, viral, and bacterial chitin deacetylases could produce partially acetylated chitosan tetramers with a defined degree of acetylation and a pattern of acetylation [35]. With the recent progress in extraction methodologies and instrumentation sophistication, chitosan extraction from marine crustaceans has also been achieved outside of chemical extraction through autoclave-based methods [36,37,38,39,40,41,42], microwave-based methods [39,43,44,45,46], ultrasonication-based methods [47,48,49], and Graviola extract combined with magnetic stirring on hot plate [50] (Table 1). Table 1 tabulates the published chitosan recovery methods specifically from crustacean wastes, the techniques employed and the recovery variables have been reported. To date, the available recovery options include chemical-, autoclave-, ultrasonication-based methods, and a phytoextract-based methodology using graviola leaf extracts. The predominant recovery method demonstrated is chemical recovery and the least reported is phytoextract-mediated green recovery. This review urges the need to focus research in the area of green extraction.
Chitin and chitosan exhibit several biological properties such as anti-cancer [109], antioxidant [110], antimicrobial [111], and anti-coagulant [112] properties. In addition, they are used as biomaterials in a wide range of applications: for biomedical purposes such as for artificial skin, bones, and cartilage regeneration [113,114], for food preservation such as for edible films [115], and for pharmaceutical purposes such as for drug delivery [116]. Chitin is a versatile, environmentally friendly, modern material [117]. Chitin and chitin derivatives have been used in virtually every significant segment of the economy (e.g., water treatment, pulp and paper industry, biomedical devices and therapies, cosmetics, biotechnology, agriculture, food science, and membrane technology) [118]. Chitin and chitosan are important bioactive materials, with many highly potent activities such as immune function, hemostasis and wound healing, antioxidant action, antimicrobial activity, and heavy metal and other pollutant removal [119]. Therefore, as renewable resources, chitin and its derivatives have a wide range of applications in food and nutrition [120], pharmaceutical [121], biotechnological [122], cosmetic [123], packaging [124], textile, wastewater treatment [125], and agricultural [126] industries.

3. Snap Shot of the Bioactive Properties of Chitosan

Chitosan has three reactive groups, the primary amine group and the primary and secondary hydroxyl groups at C-2, C-3, and C-6 positions, respectively [127]. Among the three groups, the primary amine at the C-2 position is reported to be responsible for the bioactivity of chitosan [128]. The chemical modification of chitosan adds unique functional properties for use towards biological and biomedical applications [129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161]. The biodegradability, biocompatibility, mucoadhesion, hemostatic, analgesic, adsorption enhancer, antimicrobial, anticholesterolemic, and antioxidant attributes of chitosan are those which make it suitable for biomedical applications.
Chitosan has been well established as an alternative to antibiotics, undertaking antimicrobial and antifungal roles. Because of its cationic properties, chitosan is able to lead to membrane-disrupting effects [162,163,164,165], which are higher against Gram-positives than Gram-negatives [165]. The antibacterial activity of chitosan is influenced by the molecular weight of chitosan and allied physicochemical properties. A number of chitosan derivatives with different modifications have shown improved antibacterial activity; in this way, chitosan micro/nanoparticles display unique physical and chemical features [166]. The chitosan nanoparticles penetrate inside the cell, interacting with phosphorus- and sulfur-containing compounds such as DNA and protein, eventually causing damage to the cells [163,164]. Successful experiments were performed using chitosan and reticulated chitosan microparticles against aquaculture related trouble-makers: Lactococcus garvieae (Gram +), Vibrio parahaemolyticus, and Vibrio alginolyticus (Gram −). These microorganisms are the most predominant bacterial pathogens of mariculture industry and are responsible for crucial economic losses in cultured fish and seafood worldwide [167]. The antimicrobial activity of chitin, chitosan, and their derivatives against different groups of microorganisms, such as bacteria, yeast, and fungi, has received considerable attention in recent years [120,168,169].
Traditional chemotherapeutic agents kill actively dividing cells, characteristic of most cancer cells. Cytotoxic drugs continue to play a crucial role in cancer therapy, although side effects such as the destruction of lymphoid and bone marrow cells is inevitable. In this direction, constant efforts to improve cancer therapy-based side effects are sought after. This is why biocompatible anticancer drugs are needed for cancer therapy. The introduction of several groups into chitosan modifies its structure significantly, thereby increasing the biological activity of chitosan. The introduction of sulfates and phenyl groups in carboxymethyl benzylamide dextrans into chitosan, lead to enhanced anticancer activity in breast cancer cells. Sulfated chitosan (SCS) and sulfated benzaldehyde chitosan (SBCS) significantly inhibited cell proliferation, induced apoptosis, and blocked the fibroblast growth factors (FGF)-2-induced phosphorylation of extracellular signal-regulated kinase (ERK) in Middle cranial fossa (MCF)-7 cells [170]. Dialkylaminoalkylation and reductive amination followed by quaternization of chitosan could elicit inhibitory effects on the proliferation of tumor cell lines [171].
Anti-inflammatory refers to the property of reducing inflammation. Theophylline is a drug that reduces the inflammatory effects of allergic asthma but is difficult to administer at an appropriate dosage without causing adverse side effects. The adsorption of theophylline to chitosan nanoparticles modified by the addition of thiol groups has been reported to enhance theophylline absorption by the bronchial epithelium and thereby enhance its anti-inflammatory effects [172]. The beneficial contributions of chitosan and its oligosaccharides include anti-tumor [173], neuroprotective [174], antifungal and antibacterial [175,176], and anti-inflammatory [177] effects. Tissue engineering is applying a combination of cells, engineering and materials methods, and suitable biochemical and physiochemical factors to improve or replace biological functions. Natural and synthetic materials have been used. Chitosan derivatives have been reported for the preparation of several tissue engineering organs, such as skin, bone, liver, nerve, and blood vessels. The chemical modification of chitosan has a profound impact with respect to delivery of different kinds of drugs to a specific place [178]. Chitosan nanoparticles have led to a significant increase in loading capacity and efficacy. The phenolic and polyphenolic compounds with antioxidant effects have been condensed with chitosan to form mutual prodrugs [179,180]. Chitin and chitosan derivatives are ideal candidates as drug carriers in cancer chemotherapy [181]. Apart from these, chitosan has been used in the pharmaceutical industry in drug delivery systems, such as tablets, microspheres, micelles, vaccines, nucleic acids (NAs), hydrogels, and nanoparticles. Chitosan has also been used for pharmaceutical applications, wound healing, and tissue regeneration [182,183]. In tableting, chitosan has been reported as an excipient to delay the release of the active ingredient from the tablets. Chitosan with high molecular weight is more viscous and is used for sustained release of drugs, in order to prolong drug activity, improve therapeutic efficiency, and reduce side effects in oral tablets [183]. Chitosan has also been reported for its use as coating material in drug delivery applications because of its good film-forming and mucoadhesive properties, leading to controlled release of drugs.

4. Antioxidant Activity of Chitosan

Being extracted from crustacean’s exoskeleton and fungi cell walls, chitosan products are biocompatible and biodegradable, and their range of applications include food, wastewater treatment, cell culture, cosmetics, textiles, agrochemicals, and medical devices [30]. Additionally, chitosan also exhibits antioxidant activity [184,185] and can be used as a replacement for synthetic antioxidants such as butylated hydroxytoluene (BHT), butylated hydroxy-anisole (BHA), propyl gallate, and tert-butylhydroquinone (TBHQ) [186]. Reactive oxygen species (ROS) such as H2O2, hydroxyl radicals, and superoxides lead to oxidative stress which is the key behind a wide range of pathologies: cancer [187], cardiovascular disease [188,189], premature aging [190], rheumatoid arthritis, and inflammation [191]. Chitin, similar to vitamin C, exhibits antioxidant effects [192] and can be used as an ingredient for the production of functional foods in order to circumvent age-related and diet-related diseases [193]. Due to oxidation of lipids in food, off-flavors and rancidity manifests, this is why BHT and BHA (synthetic antioxidants) are used. BHT and BHA are well known for their potential health hazards [194], and hence safe and natural alternatives are being sought [195]. The addition of 1% chitosan resulted in 70% decrease in the 2-thiobarbituric acid reactive substance (TBARS) values of frozen meat. Chitosan addition is reported to have lead to chelation of the free iron in heme proteins of meat that are released during processing [196]. In seafoods, oxidation of high concentrations of prooxidants such as hemoglobin and metal ions in fish muscles is also reported [197]. Antioxidant effect of chitosan is reported [198] to be directly proportional to its molecular weight, concentration, and viscosity. Chitosans from crab shell wastes were tested on herring flesh and chitosan with different viscosity were also tested on fish samples. The highest activity was observed with low viscosity chitosan (14 cP) and its action was similar to that of BHA, BHT, and TBHQ. Chitosans are speculated to slowdown lipid oxidation by chelating ferrous ions in fish. This eliminates the prooxidant activity of ferrous ions by preventing their conversion to ferric ion [197]. Kim and Thomas [199] have also reported identical inferences in Atlantic salmon (Salmo salar).
Free radical reaction is connected with several specific human diseases and has gained paramount attention. In the human body, reactive oxygen species (ROS) produced during metabolic process oxidize lipids, proteins, carbohydrates and nucleic material, resulting in oxidative stress [192]. The ROS generated, may activate enzymes that eventually manifest as life-threatening disorders such as cancer, aging, cardiovascular diseases wrinkle formation, rheumatoid arthritis, inflammation, hypertension, dyslipidemia, atherosclerosis, myocardial infraction, angina pectoris, heart failure, and neurodegenerative diseases such as Alzheimer’s, Parkinson’s, and amyotrophic lateral sclerosis [200,201,202,203,204]. The term ROS refers to oxygen-derived free radicals like superoxide, hydroxyl radical, and nitric oxide, and is extended to nonradical oxygen derivatives of high reactivity like singlet oxygen, hydrogen peroxide, peroxynitrite, and hypochlorite [205,206]. Mitochondria in biological cells are responsible for ROS generation [207]. The cellular defense includes enzymes such as catalase, superoxide dismutase, and glutathione peroxidase [200]. When excessive ROS are generated, the defense mechanism is unable to respond appropriately and thus oxidative stress manifests. The antioxidant activity of chitosan has gained paramount importance, with chitosan exhibiting confirmed scavenging activity against various radical species. The degree of deactylation (DDA) and molecular weight (MW) of chitosan determine the scavenging capacity of chitosan [208] and the chitosan NH2 groups are responsible for free radical scavenging effect. Mahdy Samar et al. confirmed that a high rate of DDA and low MW of chitosan exhibits higher antioxidant activity [44]. Hajji et al. studied chitosan obtained from Tunisian marine shrimp (Penaeus kerathurus) waste (DDA: 88%), crab (Carcinus mediterraneus) shells (DDA: 83%), and cuttlefish (Sepia officinalis) bones (DDA: 95%) [74] and tested their antioxidant activities. Cuttlefish with 95% DDA exhibited the highest antioxidant activity. Sun et al. [209], reported that chitosan oligomers exhibited stronger scavenging activity with lower MW. Chang et al. [210] demonstrated that lower MW chitosan (~2.2 kDa) greatly impacts the scavenging ability. Although the antioxidant activity of chitosan has been proven, yet it is limited by the lack of an H-atom donor, to serve as a good chain-breaking antioxidant [211]. The scavenging capacity of free radicals is related to the bond dissociation energy of O–H or N–H and the stability of the formed radicals. Due to the presence of strong intramolecular and intermolecular hydrogen bonds in chitosan molecules, the OH and NH2 groups find it difficult to dissociate and react with hydroxyl radicals [212]. This is why various modifications of chitosan molecules by grafting functional groups into the molecular structure have evolved. Modification of chitosan by grafting polyphenols, has been observed to enhance the antioxidant activity. This has resulted owing to the synergetic effects obtained from both chitosan and polyphenols [213]. Chito-oligosaccharides (COS) are known to be highly promising compounds for use as natural antioxidants in biological systems [62,214]. Li et al. 2018 [215] have recently investigated the preparation and potential free radical scavenging activity of chitosan derivatives with 1,2,3-triazoles and 1,2,3-triazoliums. Their results indicated that all the chitosan derivatives exhibited higher radical scavenging activity than chitosan and the scavenging ability was further enhanced following the N-methylation of 1,2,3-triazole moieties. Other researchers [216,217] have also reported the antioxidant activity of quaternary ammonium groups in chitosan derivatives.
Antioxidant agents like chitosan play a role in scavenging the free radicals and by inhibiting the oxidative damage caused by free radicals (Figure 2). Antioxidant mechanism of chitosan functions by protecting the host against oxidative stress induced damages via interfering with the oxidation chain reaction. The exact mechanism of free radical scavenging activity of chitosan has still not been established. However, Riaz et al., 2019 [218] put forth a plausible theory that the unsteady free radicals may react with the OH group and the amino group at C-2, C-3, and C-6 positions of the pyranose ring to produce a stable macromolecule. This review calls to attention that this (elucidating the mechanism of antioxidant activity of chitosan) is one of the gray areas worth working on with respect to the antioxidant activity of chitosan.

5. The State-of-the-Art Chitosan Trends

With the outputs of nanotechnology having benefitted almost every area of life, inputs from the nano realm in chitosan are no exception. Nanofabrication comprises of top-down approaches and bottom-up approaches that bring about fabrication of nanostructures from native superstructures through successive disintegration or build-up [219]. Chitosan nanofibers and nanowhiskers have been fabricated using methods of disassembly [220,221]. Wet grinding of chitosan flakes, with subsequent high pressure homogenization resulted in 100 to 1000 nm-sized (diameter) chitin nanofibers. Wijesena et al. [222] prepared chitin nanofibrils (diameter of 5 nm) from crab shells applying ultrasonication for mechanical disassembly. Grinding has also been adopted for fabrication of chitin nanofibrils from microfibrils [223]. Chitin nanowhiskers and nanofibers have been prepared using radical assisted oxidation of chitin with 2,2,6,6-tetramethyl piperidine-1-oxyl followed by ultra-sonication [224]. Tanaka et al. [225] reported nanofibrillation of chitin powder under nitrogen gas using ultrasonication. Chitin nanofibers have been reported by top-down approach or by electrospinning [226]. Most of these approaches involve highly acidic or basic solutions, this is why alternative environment-friendly approaches to produce chitin/chitosan nanofibers were sought after. A self-assembly mechanism is one such alternative approach. However, chitosan nanoparticles being insoluble do not self-assemble [227]. Therefore, amphiphilic chitosan derivatives have been constructed for self-assembly [227].
Curcumin-encapsulated chitosan nanoparticles have been reported using the sonication method. The results revealed an improved solubility of curcumin with sustained-release pattern [228]. Chitin nanofibers have been extracted from crab shells [229], from prawn shells [230] including shells of Penaeus monodon (black tiger prawn), Marsupenaeus japonicus (Japanese tiger prawn), and Pandaluseous makarov (Alaskan pink shrimp). Their shells were treated using NaOH and HCl solutions to remove proteins and minerals to yield chitin nanofibers with 10–20 nm uniform width and high aspect ratios. Chitin nanofibers have been reported for use as a drug for inflammatory bowel disease. Moreover, chitosan is also employed as a cell culture media due to its biocompatibility and ability to accelerate growth. In many of these studies, chitosan, polylactic acid (PLA) and polyglycolic acid (PGA) are mixed to form films, porous structures, or beads [231].
Several composites have been prepared, such as chitosan with alginate [139], collagen [232], calcium phosphate [233], hydroxyapatite [234], and polysulfone [235]. Chitosan composite materials have been employed in bone tissue engineering. Bone is mostly composed of collagen and hydroxyapatite; researchers have attempted substituting the function of collagen by chitosan [34,236,237]. Various derivatives of chitosan have also been reported, carboxymethyl-chitosan is an amphoteric polymer and the solubility depends on pH. Hydroxypropyl chitosan was prepared from chitosan and propylene epoxide under alkaline conditions. Other chemical modification of chitosan includes esterification of chitosan and N, O-acylation of chitosan using acetyl chloride in MeSO3H as solvent [238]. It has been confirmed that acetylation of chitosan substantially improves its antifungal activity [239]. There are few methods to obtain phosphorylated derivatives of chitosan. These chitosan derivatives are necessary owing to their unique biological and chemical properties. They could exhibit bactericidal and osteoinductive properties. Phosphorylated chitosan can be prepared by heating chitosan with orthophosphoric acid in N, N-dimethylformamide (DMF). Chitosan nanoparticles have been prepared and put to use for catalytic functions too [240,241].
Chitosans with degrees of polymerization (DPs) <20 and an average molecular weight less than 3900 Da are called chitooligosaccharides (COS) [242]. This is one of the most trending chitosan derivate types. COS are generated by depolymerization of chitin or chitosan using acid hydrolysis, hydrolysis by physical methods, and enzymatic degradation [243]. Recently, COS has gained paramount attention owing to its pharmaceutical and medicinal applications, due to their nontoxicity and high solubility and positive physiological effects. The health benefits of COS include lowering blood cholesterol, lowering high blood pressure, protective effects against infections, controlling arthritis, improvement of calcium uptake, and enhancing antitumor properties.

6. Future Perspective and Conclusions

Disposal in the future will have to cope with more stringent ecological standards, making recovery of valuable byproducts from wastes mandatory. Integral utilization of renewable resources is certainly a goal worth pursuing and can be envisaged as new reclamation technologies are developed. This certainly applies to shellfish processing waste and byproducts. Shellfish waste represents a substantial portion of the raw material. These materials, mainly shells, viscera, heads, and adhered meat, are partly used today to produce fishmeal for animal consumption, the rest being wasted even though it is environmentally offensive. The relative amount of these materials varies with the species and the technology of processing. There is considerable potential for conversion of this waste into value-added products to resolve some of the issues associated with environment pollution and cost of disposal. Chitosan has been extensively expanding and especially the very aspect of recovery of this valuable resource from marine shell waste is by itself a very attractive topic. Given the fact that such a worthwhile recovery on the other hand is also helping through productive use of crustacean wastes, ridding the environment of a bio-pollutant, still the extraction process itself has not been improved much. Only traditional extraction processes, with very minimal use of sophistication in terms of advanced instrumentation have been employed. Green extraction strategies have been minimally attempted and published; more green extraction methods and large scale extraction methods need to be addressed. This is a huge lacuna we observed as this research was reviewed. Muthu et al. have already reported extraction of chitosan from sea waste uisng graviola leaf extracts. This is only the beginning, there are numerous such green aspects that can be probed. This review is expected to draw some response from researchers in this area of research in this specific direction. With chitosan holding huge potentials in industrial, medical, and agricultural applications, there is still need for more chitosan and large-scale recovery units and factories with more advanced technologies. How much crustacean waste are we currently using to extract chitosan? How much more can we utilize this marine waste? Is there a chance that we can perhaps use all of it? Are all questions worth giving a thought or a shot?
Chitosan composite materials with integrated properties are much less reported; this is a crucial area demanding progress. Chitosan derivatives have also not been exploited fully, each derivative has a few reports attached and each derivative has not been extensively assessed of its wholesome properties. COS, which apparently appears to hold a lot of potential, needs to be put forward to various medical applications. Moreover, extraction of COS from crustacean wastes, involving advanced techniques, should be seriously looked into.
Chitosan-based drugs that are employed to treat various cancers are still at the laboratory stages of testing. No clinical trials are reported and most of the medical applications are far from even beginning the journey from benchtop to bedside. This gray area needs to be worked on. There is no doubt that the recovery of chitosan and its derivatives from marine shell wastes is a valuable asset, what remains is how best we can advance the recovery methods and how best we can use the recovered chitosan. Since chitosan has huge potential and given the fact that it can be harnessed from a wasting away resource, everything that it takes to bring chitosan to the forefront needs to be attempted and achieved.
Nanotechnology is a field that has made much room for progress in various areas of research and development. There has been very limited exploitation of this cutting-edge technology in chitosan recovery, in chitosan composites, and in chitosan applications; this is something this review would like to project. Nanotechnology has pushed the limits of various gray areas in various fundamental research, more inclusion of this technology into chitosan research is crucial for progress. When composites have been done with bulk materials such as alginates, collagen, and hydroxyapatites, there is certainly “room for more” making these composites with nano-alginates, nano versions of collagens, and the like. Table 2 presents the nano-chitosan/nano-chitosan composites versions that have been synthesized so far and their deliverables achieved. With such medical applications of nano-chitosan already established [244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266], it is strange that these chitosan nanomaterials have not been assessed for their antioxidant properties. The antioxidant potential of chitosan is far from being harnessed, there has been nothing much reported in the direction of enhancement of the antioxidant potential of chitosan through nanocomposites and a fundamental understanding of antioxidant activity of chitosan (mechanism) is greatly lacking. Moreover, the antioxidant activity of chitosan derivatives has been clearly shown to be higher than that of chitosan, yet other than the existing handful reports there has been low enthusiasm on extending the antioxidant applications to chitosan derivatives.
This review briefly summarized the existing recovery, bioactive, and antioxidant properties and applications of chitosan. The antioxidant property of chitosan has been vaguely and randomly studied and represented by not many authoritative reports, this area needs to be explored more. The most important emphasis here is that since crustacean waste can be turned into a multifaceted beneficial resource; more cutting-edge technologies need to be applied to and improved in chitosan research. This appears to have not been done as we look through what has been achieved, hence we encourage progressive upgradations involving nano aspects and nano-integrated composite approaches as future initiatives.

Author Contributions

M.M., J.G. and I.S. preparation of original draft and revisions; A.J.P.D., S.C. and N.H. participated in review and revisions. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This article was supported by the KU Research Professor Program of Konkuk University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Einbu, A.; Vårum, K.M. Characterization of chitin and its hydrolysis to GlcNAc and GlcN. Biomacromolecules 2008, 9, 1870–1875. [Google Scholar] [CrossRef] [PubMed]
  2. Sato, H.; Mitzutani, S.; Tsuge, S. Determination of the degree of acetylation of chitin/chitosan by pyrolysis-gas chromatography in the presence of oxalic acid. Anal. Chem. 1998, 70, 7–12. [Google Scholar] [CrossRef] [PubMed]
  3. Kikkawa, Y.; Tokuhisa, H.; Shingai, H.; Hiraishi, T.; Houjou, H.; Kanesato, M.; Imanaka, T.; Tanaka, T. Interaction force of chitin-binding domains onto chitin surface. Biomacromolecules 2008, 9, 2126–2131. [Google Scholar] [CrossRef] [PubMed]
  4. Shahidi, F.; Abuzaytoun, R. Chitin, chitosan, and co-products: Chemistry, production, applications, and health effects. Adv. Food Nutr. Res. 2005, 49, 93–135. [Google Scholar] [PubMed]
  5. Wang, X.; Xing, B. Importance of structural makeup of biopolymers for organic contaminant sorption. Environ. Sci. Technol. 2007, 41, 3559–3565. [Google Scholar] [CrossRef] [PubMed]
  6. Zargar, V.; Asghari, M.; Dashti, A. A review on chitin and chitosan polymers: Structure, chemistry, solubility, derivatives, and applications. Chembioeng. Rev. 2015, 2, 204–226. [Google Scholar] [CrossRef]
  7. Yan, N.; Chen, X. Sustainability: Don’t waste seafood waste. Nat. News 2015, 524, 155. [Google Scholar] [CrossRef]
  8. Kjartansson, G.T.; Zivanovic, S.; Kristbergsson, K.; Weiss, J. Sonication-assisted extraction of chitin from North Atlantic shrimps (Pandalus borealis). J. Agric. Food Chem. 2006, 54, 5894–5902. [Google Scholar] [CrossRef]
  9. Chang, K.L.B.; Tsai, G. Response surface optimization and kinetics of isolating chitin from pink shrimp (Solenocera melantho) shell waste. J. Agric. Food Chem. 1997, 45, 1900–1904. [Google Scholar] [CrossRef]
  10. Mojarrad, J.S.; Nemati, M.; Valizadeh, H.; Ansarin, M.; Bourbour, S. Preparation of glucosamine from exoskeleton of shrimp and predicting production yield by response surface. J. Agric. Food Chem. 2007, 55, 2246–2250. [Google Scholar] [CrossRef]
  11. Xu, Y.; Gallert, C.; Winter, J. Chitin purification from shrimp wastes by microbial deproteination and decalcification. Appl. Microbiol. Biotechnol. 2008, 79, 687–697. [Google Scholar] [CrossRef] [PubMed]
  12. Zhai, X.; Hawkins, S.J. Interactions of aquaculture and waste disposal in the coastal zone. J. Ocean. Univ. Qingdao 2002, 1, 8–12. [Google Scholar]
  13. Gimeno, M.; Ramirez-Hernandez, J.Y.; Mártinez-Ibarra, C.; Pacheco, N.; Garcia-Arrazola, R.; Barzana, E.; Shirai, K. One-solvent extraction of astaxanthin from lactic acid fermented shrimp wastes. J. Agric. Food Chem. 2007, 55, 10345–10350. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, S.L.; Chang, T.J.; Liang, T.W. Conversion and degradation of shellfish wastes by Serratia sp. TKU016 fermentation for the production of enzymes and bioactive materials. Biodegradation 2010, 21, 321–333. [Google Scholar] [CrossRef] [PubMed]
  15. Arbia, W.; Arbia, L.; Adour, L.; Amrane, A. Chitin extraction from crustacean shells using biological methods—A review. Food Technol. Biotech. 2013, 51, 12–25. [Google Scholar]
  16. Yadav, M.; Goswami, P.; Paritosh, K.; Kumar, M.; Pareek, N.; Vivekanand, V. Seafood waste: A source for preparation of commercially employable chitin/chitosan materials. Bioresour. Bioprocess. 2019, 6, 8. [Google Scholar] [CrossRef]
  17. Casadidio, C.; Peregrina, D.V.; Gigliobianco, M.R.; Deng, S.; Censi, R.; Di Martino, P. Chitin and chitosans: Characteristics, eco-friendly processes, and applications in cosmetic science. Mar. Drugs 2019, 17, 369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Synowiecki, J.; Al-Khateeb, N.A. Production, properties, and some new applications of chitin and its derivatives. Crit. Rev. Food Sci. Nutr. 2003, 43, 145–171. [Google Scholar] [CrossRef]
  19. Tolaimate, A.; Desbrieres, J.; Rhazi, M.; Alagui, A. Contribution to the preparation of chitins and chitosans with controlled physico-chemical properties. Polymer 2003, 44, 7939–7952. [Google Scholar] [CrossRef]
  20. Das, S.; Ganesh, E.A. Extraction of chitin from trash crabs (Podophthalmus vigil) by an eccentric method. Curr. Res. J. Biol. Sci. 2010, 2, 72–75. [Google Scholar]
  21. Bajaj, M.; Winter, J.; Gallert, C. Effect of deproteination and deacetylation conditions on viscosity of chitin and chitosan extracted from Crangon crangon shrimp waste. Biochem. Eng. J. 2011, 56, 51–62. [Google Scholar] [CrossRef]
  22. Haddar, A.; Hmidet, N.; Ghorbel-Bellaaj, O.; Fakhfakh-Zouari, N.; Sellami-Kamoun, A.; Nasri, M. Alkaline proteases produced by Bacillus licheniformis rp1 grown on shrimp wastes: Application in chitin extraction, chicken feather-degradation and as a dehairing agent. Biotechnol. Bioprocess. Eng. 2011, 16, 669. [Google Scholar] [CrossRef]
  23. Sila, A.; Nasri, R.; Bougatef, A.; Nasri, M. Digestive alkaline proteases from the goby (Zosterisessor ophiocephalus): Characterization and potential application as detergent additive and in the deproteinization of shrimp wastes. J. Aquat. Food Prod. Technol. 2012, 21, 118–133. [Google Scholar] [CrossRef]
  24. Mhamdi, S.; Ktari, N.; Hajji, S.; Nasri, M.; Kamoun, A.S. Alkaline proteases from a newly isolated Micromonospora chaiyaphumensis s103: Characterization and application as a detergent additive and for chitin extraction from shrimp shell waste. Int. J. Biol. Macromol. 2017, 94, 415–422. [Google Scholar] [CrossRef] [PubMed]
  25. Hamdi, M.; Hammami, A.; Hajji, S.; Jridi, M.; Nasri, M.; Nasri, R. Chitin extraction from blue crab (Portunus segnis) and shrimp (Penaeus kerathurus) shells using digestive alkaline proteases from P. segnis viscera. Int. J. Biol. Macromol. 2017, 101, 455–463. [Google Scholar] [CrossRef] [PubMed]
  26. Castro, R.; Guerrero-Legarreta, I.; Bórquez, R. Chitin extraction from Allopetrolisthes punctatus crab using lactic fermentation. Biotechnol. Rep. 2018, 20, e00287. [Google Scholar] [CrossRef] [PubMed]
  27. Du, J.; Tan, E.; Kim, H.J.; Zhang, A.; Bhattacharya, R.; Yarema, K.J. Comparative evaluation of chitosan, cellulose acetate, and polyethersulfone nanofiber scaffolds for neural differentiation. Carbohydr. Polym. 2014, 99, 483–490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Peter, M.G. Applications and Environmental Aspects of Chitin and Chitosan. J. Macromol. Sci. Part A 1995, 32, 629–640. [Google Scholar] [CrossRef]
  29. Kurita, K. Chemistry and application of chitin and chitosan. Polym. Degrad. Stab. 1998, 59, 117–120. [Google Scholar] [CrossRef]
  30. Kumar, R.M.N. A review of chitin and chitosan applications. React. Funct. Polym. 2000, 46, 1–27. [Google Scholar] [CrossRef]
  31. Dutta, P.K.; Dutta, J.; Tripathi, V.S. Chitin and chitosan: Chemistry, properties and applications. J. Sci. Ind. Res. India 2004, 63, 20–31. [Google Scholar]
  32. Rinaudo, M. Chitin and chitosan: Properties and applications. Prog. Polym. Sci. 2006, 31, 603–632. [Google Scholar] [CrossRef]
  33. Pillai, C.K.S.; Paul, W.; Sharma, C.P. Chitin and chitosan polymers: Chemistry, solubility and fiber formation. Prog. Polym. Sci. 2009, 34, 641–678. [Google Scholar] [CrossRef]
  34. Venkatesan, J.; Kim, S.K. Chitosan composites for bone tissue engineering—An overview. Mar. Drugs 2010, 8, 2252–2266. [Google Scholar] [CrossRef] [Green Version]
  35. Hembach, L.; Cord-Landwehr, S.; Moerschbacher, B.M. Enzymatic production of all fourteen partially acetylated chitosan tetramers using different chitin deacetylases acting in forward or reverse mode. Sci. Rep. 2017, 7, 17692. [Google Scholar] [CrossRef] [Green Version]
  36. Abdou, E.S.; Nagy, K.S.A.; Elsabee, M.Z. Extraction and characterization of chitin and chitosan from local sources. Bioresour. Technol. 2008, 99, 1359–1367. [Google Scholar] [CrossRef]
  37. Al-Hassan, A.A. Utilization of waste: Extraction and characterization of chitosan from shrimp byproducts. Civil. Environ. Res. 2016, 8, 117–123. [Google Scholar]
  38. Vilar Junior, J.C.; Ribeaux, D.R.; Alves da Silva, C.A.; Campos-Takaki, G.M. Physicochemical and antibacterial properties of chitosan extracted from waste shrimp shells. Int. J. Microbiol. 2016, 5127515, 1–7. [Google Scholar] [CrossRef] [Green Version]
  39. Sedaghat, F.; Yousefzadi, M.; Toiserkani, H.; Najafipour, S. Bioconversion of shrimp waste Penaeus merguiensis using lactic acid fermentation: An alternative procedure for chemical extraction of chitin and chitosan. Int. J. Biol. Macromol. 2017, 104, 883–888. [Google Scholar] [CrossRef]
  40. Varun, T.K.; Senani, S.; Jayapal, N.; Chikkerur, J.; Roy, S.; Tekulapally, V.B.; Gautam, M.; Kumar, N. Extraction of chitosan and its oligomers from shrimp shell waste, their characterization and antimicrobial effect. Vet. World 2017, 10, 170–175. [Google Scholar] [CrossRef] [Green Version]
  41. Aboudamia, F.Z.; Kharroubi, M.; Neffa, M.; Aatab, F.; Hanoune, S.; Bouchdoug, M.; Jaouad, A. Potential of discarded sardine scales (Sardina pilchardus) as chitosan sources. J. Air Waste Manag. Assoc. 2020, 70, 1186–1197. [Google Scholar] [CrossRef] [PubMed]
  42. Said Al Hoqani, H.A.; AL-Shaqsi, N.; Hossain, M.A.; Al Sibani, M.A. Isolation and optimization of the method for industrial production of chitin and chitosan from Omani shrimp shell. Carbohydr. Res. 2020, 492, 108001. [Google Scholar] [CrossRef] [PubMed]
  43. Sagheer, F.A.A.; Al-Sughayer, M.A.; Muslim, S.; Elsabee, M.Z. Extraction and characterization of chitin and chitosan from marine sources in Arabian Gulf. Carbohydr. Polym. 2009, 77, 410–419. [Google Scholar] [CrossRef]
  44. Samar, M.M.; El-Kalyoubi, M.; Khalaf, M.; El-Razik, M.A. Physicochemical, functional, antioxidant and antibacterial properties of chitosan extracted from shrimp wastes by microwave technique. Ann. Agric. Sci. 2013, 58, 33–41. [Google Scholar] [CrossRef] [Green Version]
  45. EL Knidri, H.; Dahmani, J.; Addaou, A.; Laajeb, A.; Lahsini, A. Rapid and efficient extraction of chitin and chitosan for scale-up production: Effect of process parameters on deacetylation degree and molecular weight. Int. J. Biol. Macromol. 2019, 139, 1092–1102. [Google Scholar] [CrossRef] [PubMed]
  46. Santos, V.P.; Maia, P.; Alencar, N.S.; Farias, L.; Andrade, R.F.S.; Souza, D.; Ribeaux, D.R.; Franco, L.O.; Campos-Takaki, G.M. Recovery of chitin and chitosan from shrimp waste with microwave technique and versatile application. Arq. Inst. Biol. 2019, 86, e0982018. [Google Scholar] [CrossRef]
  47. Hafsa, J.; Smach, M.A.; Charfeddine, B.; Limem, K.; Majdoub, H.; Rouatbi, S. Antioxidant and antimicrobial proprieties of chitin and chitosan extracted from Parapenaeus Longirostris shrimp shell waste. Ann. Pharm. Françaises 2016, 74, 27–33. [Google Scholar] [CrossRef]
  48. Borja-Urzola, A.D.C.; García-Gómez, R.S.; Flores, R.; Durán-Domínguez-de-Bazúa, M.D.C. Chitosan from shrimp residues with a saturated solution of calcium chloride in methanol and water. Carbohydr. Res. 2020, 497, 108116. [Google Scholar] [CrossRef]
  49. Huang, L.; Bi, S.; Pang, J.; Sun, M.; Feng, C.; Chen, X. Preparation and characterization of chitosan from crab shell (Portunus trituberculatus) by NaOH/urea solution freeze-thaw pretreatment procedure. Int. J. Biol. Macromol. 2020, 147, 931–936. [Google Scholar] [CrossRef]
  50. Gopal, J.; Muthu, M.; Dhakshanamurthy, T.; Kim, K.J.; Hasan, N.; Kwon, S.J.; Chun, S. Sustainable ecofriendly phytoextract mediated one pot green recovery of chitosan. Sci. Rep. 2019, 9, 13832. [Google Scholar] [CrossRef] [Green Version]
  51. Methacanon, P.; Prasitsilp, M.; Pothsree, T.; Pattaraarchachai, J. Heterogeneous N-deacetylation of squid chitin in alkaline solution. Carbohydr. Polym. 2003, 52, 119–123. [Google Scholar] [CrossRef]
  52. Chandumpai, A.; Singhpibulporn, N.; Faroongsarng, D.; Sornprasit, P. Preparation and physico-chemical characterization of chitin and chitosan from the pens of the squid species, Loligo lessoniana and Loligo formosana. Carbohydr. Polym. 2004, 58, 467–474. [Google Scholar] [CrossRef]
  53. Fernandez Cerveraa, M.; Heinamaki, J.; Rasanenc, M.; Maunuc, S.L.; Karjalainenb, M.; Nieto Acostaa, O.M.; Iraizoz Colartea, A.; Yliruusib, J. Solid-state characterization of chitosans derived from lobster chitin. Carbohydr. Polym. 2004, 58, 401–408. [Google Scholar] [CrossRef]
  54. Beaney, P.; Jaime, L.-M.; Healy, M. Comparison of chitins produced by chemical and bioprocessing methods. J. Chem. Technol. Biotechnol. 2005, 80, 145–150. [Google Scholar] [CrossRef]
  55. Rao, M.S.; Stevens, W.F. Chitin production by Lactobacillus fermentation of shrimp biowaste in a drum reactor and its chemical conversion to chitosan. J. Chem. Technol. Biotechnol. 2005, 80, 1080–1087. [Google Scholar] [CrossRef]
  56. Sini, T.K.; Santhosh, S.; Mathew, P.T. Study on the production of chitin and chitosan from shrimp shell by using Bacillus subtilis fermentation. Carbohydr. Res. 2007, 342, 2423–2429. [Google Scholar] [CrossRef]
  57. Weska, R.F.; Moura, J.M.; Batista, L.M.; Rizzi, J.; Pinto, L.A.A. Optimization of deacetylation in the production of chitosan from shrimp wastes: Use of response surface methodology. J. Food Eng. 2007, 80, 749–753. [Google Scholar] [CrossRef]
  58. Yen, M.T.; Yang, J.H.; Mau, J.L. Physicochemical characterization of chitin and chitosan from crab shells. Carbohydr. Polym. 2009, 75, 15–21. [Google Scholar] [CrossRef]
  59. Manni, L.; Ghorbel-Bellaaj, O.; Jellouli, K.; Younes, I.; Nasri, M. Extraction and characterization of chitin, chitosan, and protein hydrolysates prepared from shrimp waste by treatment with crude protease from Bacillus cereus SV1. Appl. Biochem. Biotechnol. 2010, 162, 345–357. [Google Scholar] [CrossRef]
  60. Zvezdova, D. Synthesis and characterization of chitosan from marine sources in Black Sea. Annu. Proc. Angel Kanchev Univ. Ruse 2010, 49, 65–69. [Google Scholar]
  61. Kucukgulmez, A.; Celik, M.; Yanar, Y.; Sen, D.; Polat, H.; Kadak, A.E. Physicochemical characterization of chitosan extracted from Metapenaeus stebbingi shells. Food Chem. 2011, 126, 1144–1148. [Google Scholar] [CrossRef] [Green Version]
  62. Benhabiles, M.S.; Salah, R.; Lounici, H.; Drouiche, N.; Goosen, M.F.A.; Mameri, N. Antibacterial activity of chitin, chitosan and its oligomers prepared from shrimp shell waste. Food Hydrocoll. 2012, 29, 48–56. [Google Scholar] [CrossRef]
  63. Prabu, K.; Natarajan, E. In Vitro Antimicrobial and Antioxidant Activity of Chitosan Isolated from Podophthalmus vigil. J. Appl. Pharm. Sci. 2012, 9, 75–82. [Google Scholar] [CrossRef] [Green Version]
  64. Vino, A.B.; Ramasamy, P.; Shanmugam, V.; Shanmugam, A. Extraction, characterization and in vitro antioxidative potential of chitosan and sulfated chitosan from cuttlebone of Sepia aculeata Orbigny, 1848. Asian Pac. J. Trop. Biomed. 2012, 2, 14–18. [Google Scholar] [CrossRef]
  65. Younesa, I.; Ghorbel-Bellaaj, O.; Nasri, R.; Chaabouni, M.; Rinaudo, M.; Nasri, M. Chitin and chitosan preparation from shrimp shells using optimized enzymatic deproteinization. Process. Biochem. 2012, 47, 2032–2039. [Google Scholar] [CrossRef]
  66. Mohammed, M.H.; Williams, P.A.; Tverezovskaya, O. Extraction of chitin from prawn shells and conversion to low molecular mass chitosan. Food Hydrocoll. 2013, 31, 166–171. [Google Scholar] [CrossRef] [Green Version]
  67. Muslim, T.; Rahman, M.H.; Begum, H.A.; Rahman, M.A. Chitosan and carboxymethyl chitosan from fish scales of Labeo rohita. Dhaka Univ. J. Sci. 2013, 61, 145–148. [Google Scholar] [CrossRef] [Green Version]
  68. Dahmane, E.M.; Taourirte, M.; Eladlani, N.; Rhazi, M. Extraction and characterization of chitin and chitosan from Parapenaeus longirostris from Moroccan local sources. Int. J. Polym. Anal. Charact. 2014, 19, 342–351. [Google Scholar] [CrossRef]
  69. Hajji, S.; Younes, I.; Ghorbel-Bellaaj, O.; Hajji, R.; Rinaudo, M.; Nasri, M.; Jellouli, K. Structural differences between chitin and chitosan extracted from three different marine sources. Int. J. Biol. Macromol. 2014, 65, 298–306. [Google Scholar] [CrossRef]
  70. Mohanasrinivasan, V.; Mishra, M.; Paliwal, J.; Singh, S.; Selvarajan, E.; Suganthi, V.; Subathra Devi, C. Studies on heavy metal removal efficiency and antibacterial activity of chitosan prepared from shrimp shell waste. 3 Biotech 2014, 4, 167–175. [Google Scholar] [CrossRef] [Green Version]
  71. Nwe, N.; Furuike, T.; Tamura, H. Chapter One—Isolation and characterization of chitin and chitosan from marine origin. In Advances in Food and Nutrition Research; Se-Kwon, K., Ed.; Academic Press: Waltham, MA, USA, 2014; Volume 72, pp. 1–15. [Google Scholar]
  72. Younes, I.; Hajji, S.; Frachet, V.; Rinaudo, M.; Jellouli, K.; Nasri, M. Chitin extraction from shrimp shell using enzymatic treatment. Antitumor, antioxidant and antimicrobial activities of chitosan. Int. J. Biol. Macromol. 2014, 69, 489–498. [Google Scholar] [CrossRef] [PubMed]
  73. Abdel-Rahman, R.M.; Hrdina, R.; Abdel-Mohsen, A.M.; Fouda, M.M.; Soliman, A.Y.; Mohamed, F.K.; Mohsin, K.; Pinto, T.D. Chitin and chitosan from Brazilian Atlantic Coast: Isolation, characterization and antibacterial activity. Int. J. Biol. Macromol. 2015, 80, 107–120. [Google Scholar] [CrossRef] [PubMed]
  74. Hajji, S.; Younes, I.; Rinaudo, M.; Jellouli, K.; Nasri, M. Characterization and in vitro evaluation of cytotoxicity, antimicrobial and antioxidant activities of chitosans extracted from three different marine sources. Appl. Biochem. Biotech. 2015, 177, 18–35. [Google Scholar] [CrossRef] [PubMed]
  75. Kumari, S.; Rath, P.; Kumar, A.S.H.; Tiwari, T. Extraction and characterization of chitin and chitosan from fishery waste by chemical method. Environ. Technol. Innov. 2015, 3, 77–85. [Google Scholar] [CrossRef]
  76. Ahing, F.A.; Wid, N. Extraction and characterization of chitosan from shrimp shell waste in Sabah. Trans. Sci. Technol. 2016, 3, 227–237. [Google Scholar]
  77. Apetroaei, M.R.; Zgarian, R.G.; Manea, A.M.; Rău, I.; Tihan, G.T.; Schroder, V. New source of chitosan from Black Sea marine Organisms identification. Mol. Cryst. Liq. Cryst. 2016, 628, 102–109. [Google Scholar] [CrossRef]
  78. Boarin-Alcalde, L.; Graciano-Fonseca, G. Alkali process for chitin extraction and chitosan production from Nile tilapia (Oreochromis niloticus) scales. Lat. Am. J. Aquat. Res. 2016, 44, 683–688. [Google Scholar] [CrossRef]
  79. Demir, D.; Öfkeli, F.; Ceylan, S.; Karagülle, N.B. Extraction and characterization of chitin and chitosan from blue crab and synthesis of chitosan cryogel scaffolds. J. Turk. Chem. Soc. Chem. 2016, 33, 131–144. [Google Scholar] [CrossRef] [Green Version]
  80. Duman, F.; Kaya, M. Crayfish chitosan for microencapsulation of coriander (Coriandrum sativum L.) essential oil. Int. J. Biol. Macromol. 2016, 92, 125–133. [Google Scholar] [CrossRef]
  81. He, X.; Xing, R.; Li, K.; Qin, Y.; Zou, P.; Liu, S.; Yu, H.; Li, P. Beta-chitosan extracted from Loligo Japonica for a potential use to inhibit Newcastle disease. Int. J. Biol. Macromol. 2016, 82, 614–620. [Google Scholar] [CrossRef]
  82. Karthik, R.; Manigandan, V.; Saravanan, R.; Rajesh, R.P.; Chandrika, B. Structural characterization and in vitro biomedical activities of sulfated chitosan from Sepia pharaonis. Int. J. Biol. Macromol. 2016, 84, 319–328. [Google Scholar] [CrossRef] [PubMed]
  83. Kumari, S.; Rath, P.; Sri Hari Kumar, A. Chitosan from shrimp shell (Crangon crangon) and fish scales (Labeorohita): Extraction and characterization. Afr. J. Biotechnol. 2016, 15, 1258–1268. [Google Scholar]
  84. Sayari, N.; Sila, A.; Abdelmalek, B.E.; Abdallah, R.B.; Ellouz-Chaabouni, S.; Bougatef, A.; Balti, R. Chitin and chitosan from the Norway lobster byproducts: Antimicrobial and anti-proliferative activities. Int. J. Biol. Macromol. 2016, 87, 163–171. [Google Scholar] [CrossRef] [PubMed]
  85. Abdelmalek, B.E.; Sila, A.; Haddar, A.; Bougatef, A.; Ayadi, M.A. β-Chitin and chitosan from squid gladius: Biological activities of chitosan and its application as clarifying agent for apple juice. Int. J. Biolog. Macromol. 2017, 104, 953–962. [Google Scholar] [CrossRef] [PubMed]
  86. Ben Seghir, B.; Benhamza, M.H. Preparation, optimization and characterization of chitosan polymer from shrimp shells. J. Food Meas. Charact. 2017, 11, 1137–1147. [Google Scholar] [CrossRef]
  87. Kumari, S.; Annamareddy, S.H.K.; Abanti, S.; Rath, P.K. Physicochemical properties and characterization of chitosan synthesized from fish scales, crab and shrimp shells. Int. J. Biol. Macromol. 2017, 104, 1697–1705. [Google Scholar] [CrossRef] [PubMed]
  88. Premasudha, P.; Vanathi, P.; Abirami, M. Extraction and characterization of chitosan from crustacean waste: A constructive waste management approach. Int. J. Sci. Res. 2017, 6, 1194–1198. [Google Scholar]
  89. de Queiroz Antonino, R.S.C.M.; Lia Fook, B.R.P.; de Oliveira Lima, V.A.; de Farias Rached, R.Í.; Lima, E.P.N.; da Silva Lima, R.J.; Peniche Covas, C.A.; Lia Fook, M.V. Preparation and characterization of chitosan obtained from shells of shrimp (Litopenaeus vannamei Boone). Mar. Drugs 2017, 15, 141. [Google Scholar] [CrossRef] [Green Version]
  90. Ramasamy, P.; Subhapradha, N.; Thinesh, T.; Selvin, J.; Selvan, K.M.; Shanmugam, V.; Shanmugam, A. Characterization of bioactive chitosan and sulfated chitosan from Doryteuthis singhalensis (Ortmann, 1891). Int. J. Biol. Macromol. 2017, 99, 682–691. [Google Scholar] [CrossRef]
  91. Takarina, N.D.; Fanani, A.A. Characterization of chitin and chitosan synthesized from red snapper (Lutjanus sp.) scale’s waste. AIP Conf. Proc. 2017, 1862. [Google Scholar] [CrossRef] [Green Version]
  92. Takarina, N.D.; Nasrul, A.A.; Nurmarina, A. Degree of deacetylation of chitosan extracted from white snapper (Lates sp.) scales waste. Int. J. Pharma Med. Biol. Sci. 2017, 6, 16–19. [Google Scholar] [CrossRef]
  93. Hamdi, M.; Hajji, S.; Affes, S.; Taktak, W.; Maâlej, H.; Nasri, M.; Nasri, R. Development of a controlled bioconversion process for the recovery of chitosan from blue crab (Portunus segnis) exoskeleton. Food Hydrocoll. 2018, 77, 534–548. [Google Scholar] [CrossRef]
  94. Huang, C.Y.; Kuo, C.H.; Wu, C.H.; Ku, M.W.; Chen, P.W. Extraction of crude chitosans from squid (Illex argentinus) pen by a compressional puffing-pretreatment process and evaluation of their antibacterial activity. Food Chem. 2018, 254, 217–223. [Google Scholar] [CrossRef] [PubMed]
  95. Muley, A.B.; Chaudhari, S.A.; Mulchandani, K.H.; Singhal, R.S. Extraction and characterization of chitosan from prawn shell waste and its conjugation with cutinase for enhanced thermo-stability. Int. J. Biol. Macromol. 2018, 111, 1047–1058. [Google Scholar] [CrossRef] [PubMed]
  96. Rubini, D.; Farisa Banu, S.; Veda Hari, B.N.; Ramya Devi, D.; Gowrishankar, S.; Karutha Pandian, S.; Nithyanand, P. Chitosan extracted from marine biowaste mitigates staphyloxanthin production and biofilms of Methicillin-resistant Staphylococcus aureus. Food Chem. Toxicol. 2018, 118, 733–744. [Google Scholar] [CrossRef] [PubMed]
  97. Srinivasan, H.; Velayutham, K.; Ravichandran, R. Chitin and chitosan preparation from shrimp shells Penaeus monodon and its human ovarian cancer cell line, PA-1. Int. J. Biol. Macromol. 2018, 107, 662–667. [Google Scholar] [CrossRef]
  98. Tokatli, K.; Demirdöven, A. Optimization of chitin and chitosan production from shrimp wastes and characterization. J. Food Process. Preserv. 2018, 42, e13494. [Google Scholar] [CrossRef]
  99. Arasukumar, B.; Prabakaran, G.; Gunalan, B.; Moovendhan, M. Chemical composition, structural features, surface morphology and bioactivities of chitosan derivatives from lobster (Thenus unimaculatus) shells. Int. J. Biol. Macromol. 2019, 135, 1237–1245. [Google Scholar] [CrossRef]
  100. Boudouaia, N.; Bengharez, Z.; Jellali, S. Preparation and characterization of chitosan extracted from shrimp shells waste and chitosan film: Application for Eriochrome black T removal from aqueous solutions. Appl. Water Sci. 2019, 9, 91. [Google Scholar] [CrossRef] [Green Version]
  101. Metin, C.; Alparslan, Y.; Baygar, T.; Baygar, T. Physicochemical, microstructural and thermal characterization of chitosan from blue crab shell waste and its bioactivity characteristics. J. Polym. Environ. 2019, 27, 2552–2561. [Google Scholar] [CrossRef]
  102. Ni’mah, Y.L.; Harmami, H.; Ulfin, I.; Suprapto, S.; Saleh, C.W. Water-soluble chitosan preparation from marine sources. Mal. J. Fund. Appl. Sci. 2019, 15, 159–163. [Google Scholar] [CrossRef]
  103. Pădurețu, C.C.; Isopescu, R.; Rău, I.; Apetroaei, M.R.; Schröder, V. Influence of the parameters of chitin deacetylation process on the chitosan obtained from crab shell waste. Korean J. Chem. Eng. 2019, 36, 1890–1899. [Google Scholar] [CrossRef]
  104. Pădurețu, C.C.; Isopescu, R.; Gîjiu, C.L.; Rău, I.; Apetroaei, M.R.; Schröder, V. Optimization of chitin extraction procedure from shrimp waste using Taguchi method and chitosan characterization. Mol. Cryst. Liq. Cryst. 2019, 695, 19–28. [Google Scholar] [CrossRef]
  105. Singh, A.; Benjakul, S.; Prodpran, T. Ultrasound-assisted extraction of chitosan from squid pen: Molecular characterization and fat binding capacity. J. Food Sci. 2019, 84, 224–234. [Google Scholar] [CrossRef]
  106. Tolesa, L.D.; Gupta, B.S.; Lee, M.J. Chitin and chitosan production from shrimp shells using ammonium-based ionic liquids. Int. J. Biolog. Macromol. 2019, 130, 818–826. [Google Scholar] [CrossRef]
  107. Aldila, H.; Asmar Fabiani, V.A.; Dalimunthe, D.Y.; Irwanto, R. The effect of deproteinization temperature and NaOH concentration on deacetylation step in optimizing extraction of chitosan from shrimp shells waste. IOP Conf. Ser. Earth Environ. Sci. 2020, 599, 012003. [Google Scholar] [CrossRef]
  108. Trung, T.S.; Van Tan, N.; Van Hoa, N.; Minh, N.C.; Loc, P.T.; Stevens, W.F. Improved method for production of chitin and chitosan from shrimp shells. Carbohydr. Res. 2020, 489, 107913. [Google Scholar] [CrossRef]
  109. Salah, R.; Michaud, P.; Mati, F.; Harrat, Z.; Lounici, H.; Abdi, N.; Drouiche, N.; Mameri, N. Anticancer activity of chemically prepared shrimp low molecular weight chitin evaluation with the human monocyte leukaemia cell line, THP-1. Int. J. Biol. Macromol. 2013, 52, 333–339. [Google Scholar] [CrossRef]
  110. Yen, M.T.; Yang, J.H.; Mau, J.L. Antioxidant properties of chitosan from crab shells. Carbohydr. Polym. 2008, 74, 840–844. [Google Scholar] [CrossRef]
  111. Goy, R.C.; de Britto, D.; Assis, O.B.G. A review of the antimicrobial activity of chitosan. Polim. Cienc. E Tecnol. 2009, 19, 241–247. [Google Scholar] [CrossRef]
  112. Vongchan, P.; Sajomsang, W.; Kasinrerk, W.; Subyen, D.; Kongtawelert, P. Anticoagulant activities of the chitosan polysulfate synthesized from marine crab shell by semi-heterogeneous conditions. Sci. Asia 2003, 29, 115–120. [Google Scholar] [CrossRef]
  113. Dash, M.; Chiellini, F.; Ottenbrite, R.M.; Chiellini, E. Chitosan—A versatile semi-synthetic polymer in biomedical applications. Prog. Polym. Sci. 2011, 36, 981–1014. [Google Scholar] [CrossRef]
  114. Parvez, S.; Rahman, M.M.; Khan, M.A.; Khan, M.A.H.; Islam, J.M.M.; Ahmed, M.; Rahman, M.F.; Ahmed, B. Preparation and characterization of artificial skin using chitosan and gelatin composites for potential biomedical application. Polym. Bull. 2012, 69, 715–731. [Google Scholar] [CrossRef]
  115. Muzzarelli, R.A.; Muzzarelli, C. Chitosan chemistry: Relevance to the biomedical sciences. Adv. Polym. Sci. 2005, 186, 151–209. [Google Scholar]
  116. Riva, R.; Ragelle, H.; des Rieux, A.; Duhem, N.; Jérôme, C.; Préat, V. Chitosan and chitosan derivatives in drug delivery and tissue engineering. Adv. Polym. Sci. 2011, 244, 19–44. [Google Scholar]
  117. Mahmoud, N.S.; Ghaly, A.E.; Arab, F. Unconventional approach for demineralization of deproteinized crustacean shells for chitin production. Am. J. Biochem. Biotechnol. 2007, 3, 1–9. [Google Scholar] [CrossRef]
  118. Li, Q.; Dunn, E.T.; Grandmaison, E.W.; Goosen, M.F.A. Applications and properties of chitin and chitosan. In Applications of Chitin and Chitosan; Mattheus, F.A., Goosen, M.F.A., Eds.; Chitin Recovery; Technomic Publishing Company: Lancaster, PA, USA, 1997; Volume 37, pp. 1359–1366. [Google Scholar]
  119. Khoushab, F.; Yamabhai, M. Chitin research revisited. Mar. Drugs 2010, 8, 1988–2012. [Google Scholar] [CrossRef] [Green Version]
  120. Limam, Z.; Selmi, S.; Sadok, S.; El Abed, A. Extraction and characterization of chitin and chitosan from crustacean byproducts: Biological and physicochemical properties. Afr. J. Biotechnol. 2011, 10, 640–647. [Google Scholar]
  121. Kato, Y.; Onishi, H.; Machida, Y. Application of chitin and chitosan derivatives in the pharmaceutical field. Curr. Pharm. Biotechnol. 2003, 4, 303–309. [Google Scholar] [CrossRef]
  122. Kim, S.K.; Mendis, E. Bioactive compounds from marine processing byproducts—A review. Food Res. Int. 2006, 39, 383–393. [Google Scholar] [CrossRef]
  123. Muzzarelli, R.A.A.; Boudrant, J.; Meyer, D.; Manno, N.; Demarchis, M.; Paoletti, M.G. Current views on fungal chitin/chitosan, human chitinases, food preservation, glucans, pectins and inulin: A tribute to Henri Braconnot, precursor of the carbohydrate polymers science, on the chitin bicentennial. Carbohydr. Polym. 2012, 87, 995–1012. [Google Scholar] [CrossRef]
  124. Leceta, I.; Guerrero, P.; Cabezudo, S.; Caba, K.D.L. Environmental assessment of chitosan-based films. J. Clean. Prod. 2013, 41, 312–318. [Google Scholar] [CrossRef]
  125. Bhatnagar, A.; Sillanpää, M. Applications of chitin-and chitosan-derivatives for the detoxification of water and wastewater-a short review. Adv. Colloid Interface 2009, 152, 26–38. [Google Scholar] [CrossRef] [PubMed]
  126. Jin, R.D.; Suh, J.W.; Park, R.D.; Kim, Y.W.; Krishnan, H.B.; Kim, K.Y. Effect of chitin compost and broth on biological control of Meloidogyne incognita on tomato (Lycopersicon esculentum Mill.). Nematology 2005, 7, 125–132. [Google Scholar] [CrossRef]
  127. Venkatesan, J.; Lowe, B.; Anil, S.; Manivasagan, P.; Kheraif, A.A.A.; Kang, K.H.; Kim, S.K. Seaweed polysaccharides and their potential biomedical applications. Starch 2015, 67, 381–390. [Google Scholar] [CrossRef]
  128. Aranaz, I.; Harris, R.; Heras, A. Chitosan amphiphilic derivatives. Chemistry and applications. Curr. Org. Chem. 2010, 14, 308–330. [Google Scholar] [CrossRef]
  129. Cárdenas, G.; Cabrera, G.; Taboada, E.; Rinaudo, M. Synthesis and characterization of chitosan alkyl phosphates. J. Chil. Chem. Soc. 2006, 51, 815–820. [Google Scholar] [CrossRef]
  130. Jayakumar, R.; Nagahama, H.; Furuike, T.; Tamura, H. Synthesis of phosphorylated chitosan by novel method and its characterization. Int. J. Biol. Macromol. 2008, 42, 335–339. [Google Scholar] [CrossRef]
  131. Jayakumar, R.; Nwe, N.; Tokura, S.; Tamura, H. Sulfated chitin and chitosan as novel biomaterials. Int. J. Biol. Macromol. 2007, 40, 175–181. [Google Scholar] [CrossRef]
  132. Jayakumar, R.; Reis, R.; Mano, J. Chemistry and applications of phosphorylated chitin and chitosan. e-Polymers 2006, 35, 1–16. [Google Scholar] [CrossRef] [Green Version]
  133. Jayakumar, R.; Selvamurugan, N.; Nair, S.; Tokura, S.; Tamura, H. Preparative methods of phosphorylated chitin and chitosan—An overview. Int. J. Biol. Macromol. 2008, 43, 221–225. [Google Scholar] [CrossRef] [PubMed]
  134. Li, B.; Huang, L.; Wang, X.; Ma, J.; Xie, F. Biodegradation and compressive strength of phosphorylated chitosan/chitosan/hydroxyapatite bio-composites. Mater. Des. 2011, 32, 4543–4547. [Google Scholar] [CrossRef]
  135. Li, B.; Huang, L.; Wang, X.; Ma, J.; Xie, F.; Xia, L. Effect of micropores and citric acid on the bioactivity of phosphorylated chitosan/chitosan/hydroxyapatite composites. Ceram. Int. 2013, 39, 3423–3427. [Google Scholar] [CrossRef]
  136. Li, Q.L.; Chen, Z.Q.; Darvell, B.W.; Liu, L.K.; Jiang, H.B.; Zen, Q.; Peng, Q.; Ou, G.M. Chitosan-phosphorylated chitosan polyelectrolyte complex hydrogel as an osteoblast carrier. J. Biomed. Mater. Res. 2007, 82, 481–486. [Google Scholar] [CrossRef] [PubMed]
  137. Li, Q.; Chen, Z.; Darvell, B.; Zeng, Q.; Li, G.; Ou, G.; Wu, M. Biomimetic synthesis of the composites of hydroxyapatite and chitosan–phosphorylated chitosan polyelectrolyte complex. Mater. Lett. 2006, 60, 3533–3536. [Google Scholar] [CrossRef]
  138. Li, Q.; Huang, L.N.; Chen, Z.; Tang, X. Biomimetic synthesis of the nanocomposite of phosphorylatd chitosan and hydroxyapatite and its bioactivity in vitro. Key Eng. Mater. 2007, 330–332, 721–724. [Google Scholar] [CrossRef]
  139. Li, Z.; Ramay, H.R.; Hauch, K.D.; Xiao, D.; Zhang, M. Chitosan-alginate hybrid scaffolds for bone tissue engineering. Biomaterials 2005, 26, 3919–3928. [Google Scholar] [CrossRef]
  140. Chesnutt, B.M.; Yuan, Y.; Brahmandam, N.; Yang, Y.; Ong, J.L.; Haggard, W.O.; Bumgardner, J.D. Characterization of biomimetic calcium phosphate on phosphorylated chitosan films. J. Biomed. Mater. Res. Part A 2007, 82, 343–353. [Google Scholar] [CrossRef]
  141. Cao, Z.; Xu, H.; Tang, R.; Wei, W.; Xu, W. Synthesis, characterization and molluscicidal activity of triphenyltin phosphorylated chitosan. Chem. Bull./Huaxue Tongbao 2008, 71, 528–532. [Google Scholar]
  142. Datta, P.; Dhara, S.; Chatterjee, J. Hydrogels and electrospun nanofibrous scaffolds of N-methylene phosphonic chitosan as bioinspired osteoconductive materials for bone grafting. Carbohydr. Polym. 2012, 87, 1354–1362. [Google Scholar] [CrossRef]
  143. Liu, J.; Zheng, Y.; Wang, W.; Wang, A. Preparation and swelling properties of semi-IPN hydrogels based on chitosan-g-poly(acrylic acid) and phosphorylated polyvinyl alcohol. J. Appl. Polym. Sci. 2009, 114, 643–652. [Google Scholar] [CrossRef]
  144. Srakaew, V.; Ruangsri, P.; Suthin, K.; Thunyakitpisal, P.; Tachaboonyakiat, W. Sodium-phosphorylated chitosan/zinc oxide complexes and evaluation of their cytocompatibility: An approach for periodontal dressing. J. Biomater. Appl. 2012, 27, 403–412. [Google Scholar] [CrossRef] [PubMed]
  145. Yeh, H.Y.; Lin, J.C. Surface phosphorylation for polyelectrolyte complex of chitosan and its sulfonated derivative: Surface analysis, blood compatibility and adipose derived stem cell contact properties. J. Biomater. Sci. Polym. Ed. 2012, 23, 233–250. [Google Scholar] [CrossRef] [Green Version]
  146. Zhao, D.; Xu, J.; Wang, L.; Du, J.; Dong, K.; Wang, C.; Liu, X. Study of two chitosan derivatives phosphorylated at hydroxyl or amino groups for application as flocculants. J. Appl. Polym. Sci. 2012, 125, E299–E305. [Google Scholar] [CrossRef]
  147. Saiki, I.; Murata, J.; Nakajima, M.; Tokura, S.; Azuma, I. Inhibition by sulfated chitin derivatives of invasion through extracellular matrix and enzymatic degradation by metastatic melanoma cells. Cancer Res. 1990, 50, 3631–3637. [Google Scholar]
  148. Ryzhenkov, V.E.; Solovyeva, M.A.; Remesova, O.V.; Okunevich, I.V. Hypolipidemic action of sulfated polysaccharides. Voprosy Meditsinskoj Khimii 1996, 42, 118–119. [Google Scholar]
  149. Kornilaeva, G.V.; Makarova, T.V.; Gamzazade, A.I.; Sklyar, A.M.; Nasibov, S.M.; Karamov, E.V. Sulphated chitosan derivatives as HIV-infection inhibitors. Immunologiya 1995, 1, 13–16. [Google Scholar]
  150. Hirano, S.; Tanaka, Y.; Hasegawa, M.; Tobetto, K.; Nishioka, A. Effect of sulfated derivatives of chitosan on some blood coagulant factors. Carbohydr. Res. 1985, 137, 205–215. [Google Scholar] [CrossRef]
  151. Gamzazade, A.; Sklyar, A.; Nasibov, S.; Sushkov, I.; Shashkov, A.; Knirel, Y. Structural features of sulfated chitosans. Carbohydr. Polym. 1997, 34, 113–116. [Google Scholar] [CrossRef]
  152. Mariappan, M.R.; Alas, E.A.; Williams, J.G.; Prager, M.D. Chitosan and chitosan sulfate have opposing effects on collagen-fibroblast interactions. Wound Repair Regen. 1999, 7, 400–406. [Google Scholar] [CrossRef]
  153. Ma, B.; Huang, W.; Kang, W.; Yan, J. Studies on preparation of sulfated derivatives of chitosan from Mucor rouxianus. Ion Exch. Adsorpt. 2007, 23, 451–458. [Google Scholar]
  154. Vongchan, P.; Sajomasang, W.; Subyen, D.; Kongtawelert, P. Anticoagulant activity of a sulfated chitosan. Carbohydr. Res. 2002, 337, 1239–1242. [Google Scholar] [CrossRef]
  155. Ou, Y.; Wu, Y.; Li, H.; Wang, K.; Zhang, D. Synthesis, characterization and antibacterial activity of hybrid materials of rare earth and chitosan. J. Chin. Rare Earth Soc. 2013, 31, 211–216. [Google Scholar] [CrossRef]
  156. Xing, R.; Liu, S.; Yu, H.; Zhang, Q.; Li, Z.; Li, P. Preparation of low-molecular-weight and highsulfate-content chitosans under microwave radiation and their potential antioxidant activity in vitro. Carbohyd. Res. 2004, 339, 2515–2519. [Google Scholar] [CrossRef] [PubMed]
  157. Xing, R.; Yu, H.; Liu, S.; Zhang, W.; Zhang, Q.; Li, Z. Antioxidative activity of differently regioselective chitosan sulfates in vitro. Bioorg. Med. Chem. 2005, 13, 1387–1392. [Google Scholar] [CrossRef]
  158. Karadeniz, F.; Karagozlu, M.Z.; Pyun, S.Y.; Kim, S.K. Sulfation of chitosan oligomers enhances their anti-adipogenic effect in 3T3-L1 adipocytes. Carbohydr. Polym. 2011, 86, 666–671. [Google Scholar] [CrossRef]
  159. Fan, L.; Wu, P.; Zhang, J.; Gao, S.; Wang, L.; Li, M.; Sha, M.; Xie, W.; Nie, M. Synthesis and anticoagulant activity of the quaternary ammonium chitosan sulfates. Int. J. Biol. Macromol. 2012, 50, 31–37. [Google Scholar] [CrossRef]
  160. Pires, N.R.; Cunha, P.L.; Maciel, J.S.; Angelim, A.L.; Melo, V.M.; De Paula, R.C.; Feitosa, J.P. Sulfated chitosan as tear substitute with no antimicrobial activity. Carbohydr. Polym. 2013, 91, 92–99. [Google Scholar] [CrossRef] [Green Version]
  161. Shelma, R.; Sharma, C.P. Development of lauroyl sulfated chitosan for enhancing hemocompatibility of chitosan. Colloid. Surf. B 2011, 84, 561–570. [Google Scholar] [CrossRef]
  162. Fang, S.W.; Li, C.F.; Shih, D.Y.C. Antifungal activity of chitosan and its preservative effect on low-sugar candied kumwuat. J. Food Prot. 1994, 56, 136–140. [Google Scholar] [CrossRef]
  163. Chung, Y.-C.; Chen, C.-Y. Antibacterial characteristics and activity of acid-soluble chitosan. Bioresour. Technol. 2008, 99, 2806–2814. [Google Scholar] [CrossRef] [PubMed]
  164. Eaton, P.; Fernandes, J.C.; Pereira, E.; Pintado, M.E.; Malcata, F.X. Atomic force microscopy study of the antibacterial effects of chitosans on Escherichia coli and Staphylococcus aureus. Ultramicroscopy 2008, 108, 1128–1134. [Google Scholar] [CrossRef] [PubMed]
  165. 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]
  166. Assis, O.B.G.; Pessoa, J.D.C. Scientific note: Preparation of thin films of chitosan for use as edible coatings to inhibit fungal growth on sliced fruits. Braz. J. Food Sci. Tech. 2004, 7, 17–22. [Google Scholar]
  167. Másson, M.; Holappa, J.; Hjálmarsdóttir, M.; Rúnarsson, Ö.; Nevalainen, T.; Järvinen, T. Antimicrobial activity of piperazine derivatives of chitosan. Carbohydr. Polym. 2008, 74, 566–571. [Google Scholar] [CrossRef]
  168. Khanafari, A.; Marandi, R.; Sanatei, S. Recovery of chitin and chitosan from shrimp waste by chemical and microbal methods. Iran. J. Environ. Health Sci. Eng. 2008, 5, 19–24. [Google Scholar]
  169. Li, Q.; Dunn, E.T.; Grandmaison, E.W.; Goosen, M.F.A. Applications and properties of chitosan. J. Bioact. Compat. Polym. 1992, 7, 370–397. [Google Scholar] [CrossRef]
  170. Farley, J.; Clear, N.; Leroy, B.; Davis, T.; Mcpherson, G. Age, growth and preliminary estimates of maturity of bigeye tuna, Thunnus obesus, in the Australian region. Mar. Freshwater Res. 2006, 57, 713–724. [Google Scholar] [CrossRef]
  171. Lee, J.K.; Lim, H.S.; Kim, J.H. Cytotoxic activity of aminoderivatized cationic chitosan derivatives. J. Bioorgan. Med. Chem. Lett. 2002, 12, 2949–2951. [Google Scholar] [CrossRef]
  172. Lee, D.-W.; Shirley, S.A.; Lockey, R.F.; Mohapatra, S.S. Thiolated chitosan nanoparticles enhance anti-inflammatory effects of intranasally delivered theophylline. Respir. Res. 2006, 7, 112. [Google Scholar] [CrossRef] [Green Version]
  173. Quan, H.; Zhu, F.; Han, X.; Xu, Z.; Zhao, Y.; Miao, Z. Mechanism of anti-angiogenic activities of chitooligosaccharides may be through inhibiting heparanase activity. Med. Hypotheses 2009, 73, 205–206. [Google Scholar] [CrossRef] [PubMed]
  174. Pangestuti, R.; Kim, S.-K. Neuroprotective properties of chitosan and its derivatives. Mar. Drugs 2010, 8, 2117–2128. [Google Scholar] [CrossRef] [PubMed]
  175. Fernandes, J.C.; Tavaria, F.K.; Soares, J.C.; Ramos, O.S.; Monteiro, M.J.; Pintado, M.E.; Xavier Malcata, F. Antimicrobial effects of chitosans and chitooligosaccharides, upon Staphylococcus aureus and Escherichia coli, in food model systems. Food Microbiol. 2008, 25, 922–928. [Google Scholar] [CrossRef] [PubMed]
  176. Wang, Y.; Zhou, P.; Yu, J.; Pan, X.; Wang, P.; Lan, W.; Tao, S. Antimicrobial effect of chitooligosaccharides produced by chitosanase from Pseudomonas CUY8. Asia Pac. J. Clin. Nutr. 2007, 16, 174–177. [Google Scholar] [PubMed]
  177. Yang, E.J.; Kim, J.G.; Kim, J.Y.; Kim, S.C.; Lee, N.H.; Hyun, C.G. Anti-inflammatory effect of chitosan oligosaccharides in RAW 264.7 cells. Cent. Eur. J. Biol. 2010, 5, 95–102. [Google Scholar] [CrossRef]
  178. Sonia, T.A.; Sharma, C.P. Chitosan and Its Derivatives for Drug Delivery Perspective. Adv. Polym. Sci. 2011, 243, 23–54. [Google Scholar]
  179. Mourya, V.K.; Inamdar, N.N. Trimethyl chitosan and its applications in drug delivery. J. Mater. Sci. Mater. Med. 2009, 20, 1057–1079. [Google Scholar] [CrossRef]
  180. Jarmila, V.; Vavríková, E. Chitosan derivatives with antimicrobial, antitumour and antioxidant activities-a review. Curr. Pharm. Des. 2011, 17, 3596–3607. [Google Scholar]
  181. Kato, Y.; Onishi, H.; Machida, Y. Contribution of chitosan and its derivatives to cancer chemotherapy. In Vivo 2005, 19, 301–310. [Google Scholar]
  182. Bayat, A.; Sadeghi, A.M.M.; Avadi, M.R.; Amini, M.; Rafiee-Tehrani, M.; Shafiee, A.; Majlesi, R.; Junginger, H.E. Synthesis of N,N-dimethyl N-ethyl chitosan as a carrier for oral delivery of peptide drugs. J. Bioact. Compat. Polym. 2006, 21, 433–444. [Google Scholar] [CrossRef]
  183. Kofuji, K.; Qian, C.J.; Nishimura, M.; Sugiyama, I.; Murata, Y.; Kawashima, S. Relationship between physicochemical characteristics and functional properties of chitosan. Eur. Polym. J. 2005, 41, 2784–2791. [Google Scholar] [CrossRef]
  184. Cerdá, C.; Sánchez, C.; Climent, B.; Vázquez, A.; Iradi, A.; El Amrani, F.; Bediaga, A.; Sáez, G.T. Oxidative stress and DNA damage in obesity-related tumorigenesis. Adv. Exp. Med. Biol. 2014, 824, 5–17. [Google Scholar] [PubMed]
  185. Wölfle, U.; Seelinger, G.; Bauer, G.; Meinke, M.C.; Lademann, J.; Schempp, C.M. Reactive molecule species and antioxidative mechanisms in normal skin and skin aging. Ski. Pharmacol. Physiol. 2014, 27, 316–332. [Google Scholar] [CrossRef] [PubMed]
  186. Ito, N.; Hirose, M.; Fukushima, H.; Tsuda, T.; Shirai, T.; Tatenatsu, M. Studies on antioxidants: Their carcinogenic and modifying effects on chemical carcinogens. Food Chem. Toxicol. 1986, 24, 1071–1092. [Google Scholar] [CrossRef]
  187. Manda, G.; Nechifor, M.T.; Neagu, T.M. Reactive oxygen species, cancer and anticancer therapies. Curr. Chem. Biol. 2009, 3, 342–366. [Google Scholar] [CrossRef]
  188. Zhang, N.; Bradley, T.A.; Zhang, C. Inflammation and reactive oxygen species in cardiovascular Disease. World J. Cardiol. 2010, 2, 408–410. [Google Scholar] [CrossRef]
  189. Zhang, J.; Xia, W.; Liu, P.; Cheng, Q.; Tahirou, T.; Gu, W.; Li, B. Chitosan modification and pharmaceutical/biomedical applications. Mar. Drugs 2010, 8, 1962–1987. [Google Scholar] [CrossRef] [Green Version]
  190. Cui, H.; Kong, Y.; Zhang, H. Oxidative stress, mitochondrial dysfunction, and aging. J. Signal. Transduct 2012, 2012, 646354. [Google Scholar] [CrossRef] [Green Version]
  191. Filippin, L.I.; Vercelino, R.; Marroni, N.P.; Xavier, R.M. Redox signalling and the inflammatory response in rheumatoid arthritis. Clin. Exp. Immunol. 2008, 152, 415–422. [Google Scholar] [CrossRef]
  192. Park, B.K.; Kim, M.-M. Applications of chitin and its derivatives in biological medicine. Int. J. Mol. Sci. 2010, 11, 5152–5164. [Google Scholar] [CrossRef] [Green Version]
  193. Kerch, G. The potential of chitosan and its derivatives in prevention and treatment of age-related diseases. Mar. Drugs 2015, 13, 2158–2182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Younes, I.; Rinaudo, M. Chitin and chitosan preparation from marine sources. Structure, properties and applications. Mar. Drugs 2015, 13, 1133–1174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Harish Prashanth, K.V.; Tharanathan, R.N. Chitin/chitosan: Modifications and their unlimited application potential—An overview. Trends Food Sci. Technol. 2007, 18, 117–131. [Google Scholar] [CrossRef]
  196. Tharanathan, R.N.; Kittur, F.S. Chitin—The undisputed biomolecule of great potential. Crit. Rev. Food Sci. Nutr. 2003, 43, 61–87. [Google Scholar] [CrossRef]
  197. No, H.K.; Meyers, S.P.; Prinyawiwatkul, W.; Xu, Z. Applications of chitosan for improvement of quality and shelf life of foods: A review. J. Food Sci. 2007, 72, 87–100. [Google Scholar] [CrossRef] [PubMed]
  198. Kamil, J.Y.V.A.; Jeon, Y.J.; Shahidi, F. Antioxidative activity of chitosans of different viscosity in cooked comminuted flesh of herring (Clupea harengus). Food Chem. 2002, 79, 69–77. [Google Scholar] [CrossRef]
  199. Kim, K.W.; Thomas, R.L. Antioxidative activity of chitosans with varying molecular weights. Food Chem. 2007, 101, 308–313. [Google Scholar] [CrossRef]
  200. Pittayapruek, P.; Meephansan, J.; Prapapan, O.; Komine, M.; Ohtsuki, M.; Maki, M. Role of matrix metalloproteinases in photoaging and photocarcinogenesis. Int. J. Mol. Sci. 2016, 17, 868. [Google Scholar] [CrossRef] [Green Version]
  201. Yoo, S.-J.; Go, E.; Kim, Y.-E.; Lee, S.; Kwon, J. Roles of reactive oxygen species in rheumatoid arthritis pathogenesis. J. Rheum. Dis. 2016, 23, 340–347. [Google Scholar] [CrossRef] [Green Version]
  202. Mirshafiey, A.; Mohsenzadegan, M. The role of reactive oxygen species in immunopathogenesis of rheumatoid arthritis. Iran. J. Allergy Asthma Immunol. 2008, 7, 195–202. [Google Scholar]
  203. Rahman, T.; Hosen, I.; Islam, M.M.T.; Shekhar, H.U. Oxidative stress and human health. Adv. Biosci. Biotechnol. 2012, 3, 997–1019. [Google Scholar] [CrossRef] [Green Version]
  204. Aranaz, I.; Mengíbar, M.; Harris, R.; Paños, I.; Miralles, B.; Acosta, N.; Galed, G.; Heras, A. Functional characterization of chitin and chitosan. Curr. Chem. Biol. 2009, 3, 203–230. [Google Scholar]
  205. Bhattacharya, S. Reactive oxygen species and cellular defense system. In Free Radicals in Human Health and Disease; Rani, V., Yadav, U.C.S., Eds.; Springer: New Delhi, India, 2015; pp. 17–29. [Google Scholar]
  206. Lushchak, V.I. Free radicals, reactive oxygen species, oxidative stress and its classification. Chem. Biol. Interact. 2014, 224, 164–175. [Google Scholar] [CrossRef] [PubMed]
  207. Patlevic, P.; Vascova, J.; Svork, P., Jr.; Vasko, L.; Svork, P. Reactive oxygen species and antioxidant defense in human gastrointestinal diseases. Integr. Med. Res. 2016, 5, 250–258. [Google Scholar] [CrossRef] [Green Version]
  208. Bhattacharyya, A.; Chattopadhyay, R.; Mitra, S.; Crowe, S.E. Oxidative stress: An essential factor in the pathogenesis of gastrointestinal mucosal diseases. Physiol. Rev. 2014, 94, 329–354. [Google Scholar] [CrossRef] [Green Version]
  209. Sun, T.; Zhou, D.; Xie, J.; Mao, F. Preparation of chitosan oligomers and their antioxidant activity. Eur. Food Res. Technol. 2007, 225, 451–456. [Google Scholar] [CrossRef]
  210. Chang, S.; Wu, C.; Tsai, G.J. Effects of chitosan molecular weight on its antioxidant and antimutagenic properties. Carbohydr. Polym. 2018, 181, 1026–1032. [Google Scholar] [CrossRef]
  211. Božič, M.; Gorgieva, S.; Kokol, V. Laccase-mediated functionalization of chitosan by caffeic and galic acids for modulating antioxidant and antimicrobial properties. Carbohydr. Polym. 2012, 87, 2388–2398. [Google Scholar] [CrossRef]
  212. Xie, W.; Xu, P.; Liu, Q. Antioxidant activity of water-soluble chitosan derivatives. Bioorg. Med. Chem. Lett. 2001, 11, 1699–1701. [Google Scholar] [CrossRef]
  213. Sousa, F.; Guebitz, G.M.; Kokol, V. Antimicrobial and antioxidant properties of chitosan enzymatically functionalized with flavonoids. Process. Biochem. 2009, 44, 749–756. [Google Scholar] [CrossRef]
  214. Fernandes, J.C.; Eaton, P.; Nascimento, H.; Gião, M.S.; Ramos, Ó.S.; Belo, L.; Santos-Silva, A.; Pintado, M.E.; Malcata, F.X. Antioxidant activity of chitooligosaccharides upon two biological systems: Erythrocytes and bacteriophages. Carbohydr. Polym. 2010, 79, 1101–1106. [Google Scholar] [CrossRef]
  215. Li, Q.; Sun, X.; Gu, G.; Guo, Z. Novel water soluble chitosan derivatives with 1,2,3-triazolium and their free radical-scavenging activity. Mar. Drugs 2018, 16, 107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Liu, J.; Sun, H.; Dong, F.; Xue, Q.; Wang, G.; Qin, S.; Guo, Z. The influence of the cation of quaternized chitosans on antioxidant activity. Carbohydr. Polym. 2009, 78, 439–443. [Google Scholar] [CrossRef]
  217. Wan, A.; Xu, Q.; Sun, Y.; Li, H. Antioxidant activity of high molecular weight chitosan and N,O-quaternized chitosans. J. Agric. Food Chem. 2013, 61, 6921–6928. [Google Scholar] [CrossRef]
  218. Rajoka, M.S.R.; Zhao, L.; Mehwish, H.M.; Wu, Y.; Mahmood, S. Chitosan and its derivatives: Synthesis, biotechnological applications, and future challenges. Appl. Microbiol. Biotechnol. 2019, 103, 1557–1571. [Google Scholar] [CrossRef]
  219. Iqbal, P.; Preece, J.A.; Mendes, P.M. Nanotechnology: The “top-down” and “bottom-up” approaches. In Supramolecular Chemistry; John Wiley & Sons: Chichester, UK, 2012. [Google Scholar]
  220. Liu, D.G.; Chang, P.R.; Chen, M.D.; Wu, Q.L. Chitosan colloidal suspension composed of mechanically disassembled nanofibers. J. Colloid Interface Sci. 2010, 354, 637–643. [Google Scholar] [CrossRef]
  221. Zhou, C.; Wu, Q. A novel polyacrylamide nanocomposite hydrogel reinforced with natural chitosan nanofibers. Colloids Surf. B Biointerfaces 2011, 84, 155–162. [Google Scholar] [CrossRef]
  222. Wijesena, R.; Tissera, N.; Kannankgara, Y.; Lin, Y.; Amaratunga, G.; de Silva, N. A method for top down preparation of chitosan nanoparticles and nanofibers. Carbohydr. Polym. 2015, 117, 731–738. [Google Scholar] [CrossRef]
  223. Ifuku, S.; Shervani, Z.; Saimoto, H. Preparation of chitin nanofibers and their composites. In Biopolymer Nanocomposite; Dufresne, A., Thomas, S., Pothan, L.A., Eds.; Wiley: Hoboken, NJ, USA, 2013; Volume 2, pp. 11–32. [Google Scholar]
  224. Fan, Y.M.; Saito, T.; Isogai, A. TEMPO-mediated oxidation of beta-chitin to prepare individual nanofibrils. Carbohydr. Polym. 2009, 77, 832–838. [Google Scholar] [CrossRef]
  225. Tanaka, K.; Yamamoto, K.; Kadokawa, J. Facile nanofibrillation of chitin derivatives by gas bubbling and ultrasonic treatments in water. Carbohydr. Res. 2014, 398, 25–30. [Google Scholar] [CrossRef]
  226. Kadokawa, J. Fabrication of nanostructured and microstructured chitin materials through gelation with suitable dispersion media. RSC Adv. 2015, 5, 12736–12746. [Google Scholar] [CrossRef]
  227. Li, F.; Zhang, X.; Li, H.; Xiang, L.; Chen, Y. Preparation of self-assembled nanoparticles of chitosan oligosaccharide-graft-polycaprolactone as a carrier of bovine serum albumin drug. Bio-Med. Mater. Eng. 2014, 24, 2041–2048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Raja, M.A.; Zeenat, S.; Arif, M.; Liu, C. Self-assembled nanoparticles based on amphiphilic chitosan derivative and arginine for oral curcumin delivery. Int. J. Nanomed. 2016, 11, 4397–4412. [Google Scholar] [CrossRef] [Green Version]
  229. Raabe, D.; Romano, P.; Sachs, C.; Fabritius, H.; Al-Sawalmih, A.; Yi, S.B.; Servos, G.; Hartwig, H.G. Microstructure and crystallographic texture of the chitin-protein network in the biological composite material of the exoskeleton of the lobster Homarus americanus. Mater. Sci. Eng. 2006, 421, 143–153. [Google Scholar] [CrossRef]
  230. Ifuku, S.; Nogi, M.; Abe, K.; Yoshioka, M.; Morimoto, M.; Saimoto, H.; Yano, H. Simple preparation method of chitin nanofibers with a uniform width of 10–20 nm from prawn shell under neutral conditions. Carbohydr. Polym. 2011, 84, 762–764. [Google Scholar] [CrossRef]
  231. Kim, S.K. Chitin and Chitosan Derivatives: Advances in Drug Discovery and Developments; Kim, S.K., Ed.; CRC Press: Boca Raton, FL, USA, 2013. [Google Scholar]
  232. Tan, W.; Krishnaraj, R.; Desai, T.A. Evaluation of nanostructured composite collagen-chitosan matrices for tissue engineering. Tissue Eng. 2001, 7, 203–210. [Google Scholar] [CrossRef]
  233. Zhang, Y.; Zhang, M. Calcium phosphate/chitosan composite scaffolds for controlled in vitro antibiotic drug release. J. Biomed. Mater. Res. 2002, 62, 378–386. [Google Scholar] [CrossRef]
  234. Kong, L.; Gao, Y.; Cao, W.; Gong, Y.; Zhao, N.; Zhang, X. Preparation and characterization of nano-hydroxyapatite/chitosan composite scaffolds. J. Biomed. Mater. Res. A 2005, 75, 275–282. [Google Scholar] [CrossRef]
  235. Huang, R.Y.M.; Pal, R.; Moon, G.Y. Crosslinked chitosan composite membrane for the pervaporation dehydration of alcohol mixtures and enhancement of structural stability of chitosan/polysulfone composite membranes. J. Membr. Sci. 1999, 160, 17–30. [Google Scholar] [CrossRef]
  236. Martino, A.D.; Sittinger, M.; Risbud, M.V. Chitosan: A versatile biopolymer for orthopaedic tissue-engineering. Biomaterials 2005, 26, 5983–5990. [Google Scholar] [CrossRef]
  237. Thein-Han, W.; Misra, R. Biomimetic chitosan–nanohydroxyapatite composite scaffolds for bone tissue engineering. Acta Biomater. 2009, 5, 1182–1197. [Google Scholar] [CrossRef] [PubMed]
  238. Badawy, M.E.I.; Rabea, E.I.; Rogge, T.M.; Stevens, C.V.; Smagghe, G.; Steurbaut, W.; Hofte, M. Synthesis and fungicidal activity of new N,O-acyl chitosan derivatives. Biomacromolecules 2004, 5, 589–595. [Google Scholar] [CrossRef] [PubMed]
  239. Badawy, M.E.I.; Rabea, E.I.; Rogge, T.M.; Stevens, C.V.; Steurbaut, W.; Höfte, M.; Smagghe, G. Fungicidal and insecticidal activity of O-acyl chitosan derivatives. Polym. Bull. 2005, 54, 279–289. [Google Scholar] [CrossRef]
  240. Ruiz, G.A.M.; Corrales, H.F.Z. Chitosan, Chitosan Derivatives and their Biomedical Applications. In Biological Activities and Application of Marine Polysaccharides; Shalaby, E.A., Ed.; InTech Open: London, UK, 2017. [Google Scholar]
  241. Zahedi, S.; Ghomi, J.S.; Shahbazi-Alavi, H. Preparation of chitosan nanoparticles from shrimp shells and investigation of its catalytic effect in diastereoselective synthesis of dihydropyrroles. Ultrason. Sonochem. 2018, 40, 260–264. [Google Scholar] [CrossRef]
  242. Mourya, V.K.; Inamdar, N.N.; Choudhari, Y.M. Chitooligosaccharides: Synthesis, characterization and applications. Polym. Sci. Ser. A 2011, 53, 583–612. [Google Scholar] [CrossRef]
  243. Aam, B.B.; Heggset, E.B.; Norberg, A.L.; Sørlie, M.; Vårum, K.M.; Eijsink, V.G.H. Production of chitooligosaccharides and their potential applications in medicine. Mar. Drugs 2010, 8, 1482–1517. [Google Scholar] [CrossRef] [Green Version]
  244. Wang, H.; Mukherjee, S.; Yi, J.; Banerjee, P.; Chen, Q.; Zhou, S. Biocompatible chitosan-carbon dot hybrid nanogels for NIR-imaging-guided synergistic photothermal-chemo therapy. ACS Appl. Mater. Interfaces 2017, 9, 18639–18649. [Google Scholar] [CrossRef]
  245. Sahu, P.; Kashaw, S.K.; Jain, S.; Sau, S.; Iyer, A.K. Assessment of penetration potential of pH responsive double walled biodegradable nanogels coated with eucalyptus oil for the controlled delivery of 5-fluorouracil: In vitro and ex vivo studies. J. Control. Release 2017, 253, 122–136. [Google Scholar] [CrossRef]
  246. Belabassi, Y.; Moreau, J.; Gheran, V.; Henoumont, C.; Robert, A.; Callewaert, M.; Rigaux, G.; Cadiou, C.; Vander Elst, L.; Laurent, S.; et al. Synthesis and characterization of PEGylated and fluorinated chitosans: Application to the synthesis of targeted nanoparticles for drug delivery. Biomacromolecules 2017, 18, 2756–2766. [Google Scholar] [CrossRef]
  247. Ju, C.; Mo, R.; Xue, J.; Zhang, L.; Zhao, Z.; Xue, L.; Ping, Q.; Zhang, C. Sequential intra-intercellular nanoparticle delivery system for deep tumor penetration. Angew. Chem. Int. Ed. Engl. 2014, 53, 6253–6258. [Google Scholar] [CrossRef]
  248. Woraphatphadung, T.; Sajomsang, W.; Rojanarata, T.; Ngawhirunpat, T.; Tonglairoum, P.; Opanasopit, P. Development of chitosan-based pH-sensitive polymeric micelles containing curcumin for colon-targeted drug delivery. AAPS Pharm. Sci. Tech. 2018, 19, 991–1000. [Google Scholar] [CrossRef] [PubMed]
  249. Wang, Z.; Sun, J.; Qiu, Y.; Li, W.; Guo, X.; Li, Q.; Zhang, H.; Zhou, J.; Du, Y.; Yuan, H.; et al. Specific photothermal therapy to the tumors with high EphB4 receptor expression. Biomaterials 2015, 68, 32–41. [Google Scholar] [CrossRef] [PubMed]
  250. Shen, L.; Huang, Y.; Chen, D.; Qiu, F.; Ma, C.; Jin, X.; Zhu, X.; Zhou, G.; Zhang, Z. pH-responsive aerobic nanoparticles for effective photodynamic Therapy. Theranostics 2017, 7, 4537–4550. [Google Scholar] [CrossRef] [Green Version]
  251. Wang, G.; Wang, J.J.; Tang, X.J.; Du, L.; Li, F. In vitro and in vivo evaluation of functionalized chitosan-Pluronic micelles loaded with myricetin on glioblastoma cancer. Nanomedicine 2016, 12, 1263–1278. [Google Scholar] [CrossRef] [PubMed]
  252. Shi, S.; Zhang, Z.; Luo, Z.; Yu, J.; Liang, R.; Li, X.; Chen, H. Chitosan grafted methoxy poly(ethylene glycol)-poly(epsilon-caprolactone) nanosuspension for ocular delivery of hydrophobic diclofenac. Sci. Rep. 2015, 5, 11337–11348. [Google Scholar] [CrossRef] [Green Version]
  253. Zhao, X.; Zhou, L.; Li, Q.; Zou, Q.; Du, C. Biomimetic mineralization of carboxymethyl chitosan nanofibers with improved osteogenic activity in vitro and in vivo. Carbohydr. Polym. 2018, 195, 225–234. [Google Scholar] [CrossRef]
  254. Marques, J.; Valle-Delgado, J.J.; Urban, P.; Baro, E.; Prohens, R.; Mayor, A.; Cistero, P.; Delves, M.; Sinden, R.E.; Grandfils, C.; et al. Adaptation of targeted nanocarriers to changing requirements in antimalarial drug delivery. Nanomedicine 2017, 13, 515–525. [Google Scholar] [CrossRef] [Green Version]
  255. Li, F.; Mei, H.; Gao, Y.; Xie, X.; Nie, H.; Li, T.; Zhang, H.; Jia, L. Co-delivery of oxygen and erlotinib by aptamer-modified liposomal complexes to reverse hypoxia-induced drug resistance in lung cancer. Biomaterials 2017, 145, 56–71. [Google Scholar] [CrossRef]
  256. Luo, L.; Bian, Y.; Liu, Y.; Zhang, X.; Wang, M.; Xing, S.; Li, L.; Gao, D. Combined near infrared photothermal therapy and chemotherapy using gold nanoshells coated liposomes to enhance antitumor effect. Small 2016, 12, 4103–4112. [Google Scholar] [CrossRef]
  257. Sun, M.; Li, J.; Zhang, C.; Xie, Y.; Qiao, H.; Su, Z.; Oupicky, D.; Ping, Q. Arginine-modified nanostructured lipid carriers with charge-reversal and pH-sensitive membranolytic properties for anticancer drug delivery. Adv. Healthc. Mater. 2017, 6, 1600693–1600705. [Google Scholar] [CrossRef]
  258. Yan, L.; Crayton, S.H.; Thawani, J.P.; Amirshaghaghi, A.; Tsourkas, A.; Cheng, Z. A pH-responsive drug-delivery platform based on glycol chitosan-coated liposomes. Small 2015, 11, 4870–4874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Lee, J.Y.; Crake, C.; Teo, B.; Carugo, D.; de Saint Victor, M.; Seth, A.; Stride, E. Ultrasound-enhanced siRNA delivery using magnetic nanoparticle-loaded chitosan-deoxycholic acid nanodroplets. Adv. Healthc. Mater. 2017, 6, 1601246–1601254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Kang, M.L.; Kim, J.E.; Im, G.I. Thermoresponsive nanospheres with independent dual drug release profiles for the treatment of osteoarthritis. Acta Biomater. 2016, 39, 65–78. [Google Scholar] [CrossRef] [PubMed]
  261. Jiang, K.; Chi, T.; Li, T.; Zheng, G.; Fan, L.; Liu, Y.; Chen, X.; Chen, S.; Jia, L.; Shao, J. A smart pH-responsive nano-carrier as a drug delivery system for the targeted delivery of ursolic acid: Suppresses cancer growth and metastasis by modulating P53/MMP-9/PTEN/CD44 mediated multiple signaling pathways. Nanoscale 2017, 9, 9428–9439. [Google Scholar] [CrossRef] [PubMed]
  262. Tian, H.; He, Z.; Sun, C.; Yang, C.; Zhao, P.; Liu, L.; Leong, K.W.; Mao, H.Q.; Liu, Z.; Chen, Y. Uniform core-shell nanoparticles with thiolated hyaluronic acid coating to enhance oral delivery of insulin. Adv. Healthc. Mater. 2018, 7, 1800285–1800296. [Google Scholar] [CrossRef] [PubMed]
  263. Chen, G.; Svirskis, D.; Lu, W.; Ying, M.; Huang, Y.; Wen, J. N-trimethyl chitosan nanoparticles and CSKSSDYQC peptide: N-trimethyl chitosan conjugates enhance the oral bioavailability of gemcitabine to treat breast cancer. J. Control. Release 2018, 277, 142–153. [Google Scholar] [CrossRef] [PubMed]
  264. Nascimento, A.V.; Singh, A.; Bousbaa, H.; Ferreira, D.; Sarmento, B.; Amiji, M.M. Overcoming cisplatin resistance in non-small cell lung cancer with Mad2 silencing siRNA delivered systemically using EGFR-targeted chitosan nanoparticles. Acta Biomater. 2017, 47, 71–80. [Google Scholar] [CrossRef] [Green Version]
  265. Lee, J.Y.; Termsarasab, U.; Lee, M.Y.; Kim, D.H.; Lee, S.Y.; Kim, J.S.; Cho, H.J.; Kim, D.D. Chemosensitizing indomethacin-conjugated chitosan oligosaccharide nanoparticles for tumor-targeted drug delivery. Acta Biomater. 2017, 57, 262–273. [Google Scholar] [CrossRef]
  266. Taghavi, S.; Ramezani, M.; Alibolandi, M.; Abnous, K.; Taghdisi, S.M. Chitosan-modified PLGA nanoparticles tagged with 5TR1 aptamer for in vivo tumor-targeted drug delivery. Cancer Lett. 2017, 400, 1–8. [Google Scholar] [CrossRef]
Figure 1. Overview of the recovery of chitosan from crustacean wastes.
Figure 1. Overview of the recovery of chitosan from crustacean wastes.
Antioxidants 10 00228 g001
Figure 2. Schematic diagram showing the mode of action of antioxidant activity of chitosan via interruption of free radical chain reaction.
Figure 2. Schematic diagram showing the mode of action of antioxidant activity of chitosan via interruption of free radical chain reaction.
Antioxidants 10 00228 g002
Table 1. Chitosan recovery methods currently in use and the varying conditions employed for deacetylation.
Table 1. Chitosan recovery methods currently in use and the varying conditions employed for deacetylation.
SourceMethodType of ConditionsDeacetylation Degree (%)References
Squid pens (Loligo formosana)ChemicalChitin was treated with 60% NaOH, 1/10 (w/v) for 60 min at 100 °C.90%Methacanon et al. [51]
Squid pens (Loligo lessoniana and Loligo formosana)ChemicalChitin was treated with 50% NaOH, 1/15 (w/v) for 8 h at 100 °C.92–93%Chandumpai et al. [52]
LobsterChemicalChitin was treated with 45% NaOH, at 130 °C for 30 min.86%Fernandez Cerveraa et al. [53]
ChemicalChitin was treated with 49% NaOH, at 130 °C for 30 min.89%
Dublin Bay prawn shell waste (Nephrops norvegicus)ChemicalChitin (10 g) was added to reaction flasks, 100 cm3 of 500 g kg−1 NaOH solution was added to each flask, heated for 2 h at 100 °C under a nitrogen atmosphere.81.92%Beaney et al. [54]
Shrimp biowaste (heads or shells)ChemicalChitin was added with 50% NaOH, 1:20 (w/v) at 50 °C for 48 h. 81–84%Rao and Stevens [55]
Shrimp shells (Metapenaeopsis dobsoni) ChemicalChitin was boiled with 40% NaOH until it got deacetylated to chitosan.81%Sini et al. [56]
Shrimp wasteChemicalChitin was added with 45% NaOH solution stirring for 90 min and heating at 130 °C.90%Weska et al. [57]
Crabs shellsAutoclaveSteeping in strong NaOH for 24 h before heating. Chitin was treated with 40% NaOH solution autoclaved (at 2-atmosphere pressure) for 2.5–3.0 h. 95%Abdou et al. [36]
Crayfish shells (Procambarus clarkia)95%
Brown shrimp shells (Penaeus aztecus)95%
Pink shrimp shells (Penaeus durarum)92%
Grooved tiger prawn (Penaeus semisulcatus)
Jinga Shrimp (Metapenaeus affinis)
Blue Swimming Crab-Male (Portunus pelagicus)
Blue Swimming Crab-Female (Portunus pelagicus)
Scyllarid Lobster (Thenus orientalis)
Cuttlefish (Sepia spp.)
ChemicalChitin was treated with 45% NaOH, 1/15 (w/v) for 10 h at 110 °C.88–94.4%Sagheer et al. [43]
MicrowaveChitin was added with 45% NaOH solution and irradiated by microwave for 15 min at 600 W.87.5–93.0%
Snow crab (Chionoecetes opilio)ChemicalChitin was treated with 40% NaOH solution at 105 °C for 120 min.93.3%Yen et al. [58]
Shrimp shells
(Metapeneaus monoceros)
ChemicalChitin was treated with 50% NaOH at 80 °C for 4 h.-Manni et al. [59]
Shrimp shells
Crab shells
ChemicalChitin (10 g) was treated with 50% NaOH at 60 °C for 8 h.79.80%
65.89%
Zvezdova [60]
Shrimp shells (Metapenaeus stebbingi)ChemicalChitin was treated with 50% NaOH for 6 h at 120 °C.92.19%Kucukgulmez et al. [61]
Shrimp shells (Parapenaeus longirostris)ChemicalChitin was added in 50% NaOH solution, stirred for 3–5 h at 90–100 °C.80%Benhabiles et al. [62]
Crab shells (Podophthalmus vigil)ChemicalChitin was treated in 40% NaOH for 6 h at 110 °C constant stirring.-Prabu and Natarajan [63]
Cuttlefish (Sepia aculeata)ChemicalChitin was treated in 40% NaOH solution by heating under reflux for 6 h at 110˚C.49.9%Vino et al. [64]
Shrimp shells (Metapenaeus monoceros) ChemicalChitin was treated with 12.5 M NaOH, 1/10 (w/v) at 140 °C for 4 h.78%Younesa et al. [65]
Prawn shells (Litopenaeus vannamei)ChemicalChitin was treated with 50% NaOH, 1/5 (w/v) for 2 h at 100 °C.80%Mohammed et al. [66]
Fish scales (Labeo rohita)ChemicalChitin was added to 40% NaOH, 1/15 (w/v), and refluxed under nitrogen atmosphere for 8 h at 100 °C.78.2%Muslim et al. [67]
Shrimp wasteMicrowaveChitin was treated with 50% NaOH solution and irradiated by microwave for 10 min at 1400 W.95.19%Samar et al. [44]
Shrimp shells (Parapenaeus longirostris)ChemicalChitin was treated with 50% NaOH, 1/60 (w/v) for 5 h at 110 °C.90%Dahmane et al. [68]
Crab shells (Carcinus mediterraneus)ChemicalChitin was treated with 12.5 M NaOH, 1:10 (w/v) for 4 h at 140 °C. 83%Hajji et al. [69]
Cuttlefish bones (Sepia officinalis)95%
Shrimp waste (Penaeus kerathurus)88%
Shrimp shell wasteChemicalChitin was added to 70% NaOH, 1/14 (w/v), and kept room temperature (RT) for 72 h.74.82%Mohanasrinivasan et al. [70]
Squid chitinChemicalChitin was treated with 60% NaOH, 1/10 (w/v) for 60 min at 100 °C.97.3%Nwe et al. [71]
Crab chitinChemicalChitin was treated with 40% NaOH, 1/30 (w/v) for 120 min at 105 °C.93.3%
Shrimp chitinChemicalChitin was treated with 50% NaOH for 20 h at 65 °C.87%
Shrimp shells (Metapenaeus monoceros)ChemicalChitin was treated with 12.5 M NaOH, 1:10 (w/v) for 4 h at 140 °C.81%Younes et al. [72]
Shrimp shellsChemicalChitin was refluxed with 50% NaOH, 1/10 (w/v) at 90 °C for 4 h.95.5%Abdel-Rahman et al. [73]
Crab shells (Carcinus mediterraneus)ChemicalChitin was treated with 12.5 M NaOH, 1/10 (w/v) for 4 h at 140 °C 83%Hajji et al. [74]
Cuttlefish bones (Sepia officinalis)95%
Shrimp waste (Penaeus kerathurus)88%
Fish scales (Labeo rohita)ChemicalSteeping in strong NaOH for 24 h before heating. Chitin was treated with 40–50% NaOH for 5–6 h at 100–160 °C.61%Kumari et al. [75]
Shrimp shell wasteChemicalChitin was treated with 48% NaOH for 48 h at 25 °C.70–85%.Ahing and Wid [76]
Shrimp wasteAutoclaveChitin was added in 50% NaOH, 1/10 (w/v), and autoclaved for 20 min at 15 psi/121 °C.70.9%Al-Hassan [37]
Spawning of veined rapa whelk (Rapana venosa)ChemicalTen grams of the sample was soaked in 4% NaOH, 1/15 (w/v) at 65 °C for 2 h.-Apetroaei et al. [77]
Warty crab shells (Eriphia verrucosa)ChemicalChitin was treated with 45% NaOH, 1:20 (w/v) at 100 °C, for 15 min.-
Fish scales (Oreochromis niloticus)ChemicalChitin was added in 40% NaOH, 1/40 (w/v), stirred for 6 h (300 rpm) at 117 °C.97.5%Boarin-Alcalde and Graciano-Fonseca [78]
Blue crab shell wastes (Callinectes sapidus) ChemicalChitin was added in 50% NaOH, 1/10 (w/v), and heated for 4 h at 150 °C.85%Demir et al. [79]
Crayfish shellsChemicalChitin was treated with 60% NaOH at 100 °C for 4 h. (Pure chitosan was obtained)Duman and Kaya [80]
Shrimp shells (Parapenaeus longirostris)ChemicalChitin was
deacetylated with 15 M NaOH, 1/20 (w/v) at 110 °C under
vacuum and constant stirring for 5 h.
73.68%Hafsa et al. [47]
UltrasonicChitin was
suspended with 15 M NaOH, 1/20 (w/v), irradiated by ultrasonic (v = 50 kHz) for 3 h.
83.55%
Squid pens (Loligo japonica) ChemicalChitin was treated with 40% NaOH at 95 °C for 6 h.91.04%He et al. [81]
Cuttlebone (Sepia pharaonic)ChemicalChitin was treated in hot concentrated NaOH (40–50%) solution to yield chitosan, which was sulfated using chlorosulfonic acid and stirred for 30 min to obtain sulfated chitosan.81%Karthik et al. [82]
Shrimp shell (Crangon crangon)ChemicalChitin was refluxed in NaOH (50% by weight) at 90 to 100 °C, stirred for 6 h.76%Kumari et al. [83]
Fish scale (Labeorohita)80%
Norway lobster (Nephrops norvegicus)ChemicalChitin was treated with 50% NaOH for 4 h at 120 °C.71.59%Sayari et al. [84]
Shrimp shell (Litopenaeus vannamei)AutoclaveExtracted with hot sulfuric acid at 95 °C overnight
and autoclaved for 20 min at 121 °C.
76%Vilar Junior et al. [38]
Chemical Treated with 40–50% NaOH, 1/20 (w/v) at 100–120 °C for 60–720 min.81.7%
Squid gladius (Sepioteuthis
lessoniana)
ChemicalChitin was treated with 50% NaOH at 120 °C for 4 h.71%Abdelmalek et al. [85]
Shrimp shells (Parapenaeus longirostris)ChemicalChitin was treated with 28.6% NaOH at 81.15 °C for 9.55 h.98%Ben Seghir et al. [86]
Crab shells (Crangon crangon)ChemicalChitin was treated with 40% KOH for 6 h at 90 °C.70%Kumari et al. [87]
Fishery waste (Labeo rohita)75%
Shrimp shells (Crangon crangon)78%
Shrimp shells
Crab shells
ChemicalChitin was treated with 65% NaOH for 72 at 30 °C.88.48%
80.12%
Premasudha et al. [88]
Shrimp shells (Litopenaeus vannamei) ChemicalChitin was added with 12.5 M NaOH, 1:15 (w/v) cooled, kept at −83 °C for 24 h, and stirring (250 rpm) for 4 or 6 h at 115 °C.91%de Queiroz Antonino et al. [89]
Squid pin (Doryteuthis singhalensis)ChemicalChitin was treated in 40% NaOH solution
by heating under reflux for 6 h at 110˚C.
83.76%Ramasamy et al. [90]
Shrimp waste (Penaeus merguiensis)AutoclaveChitin was treated with 45% NaOH, 1/15 (w/v), and autoclaved for 30 min at 15 psi/121 °C.88%Sedaghat et al. [39]
MicrowaveChitin was treated with 50% NaOH and irradiated by microwave for 10 min at 1400 W.
ChemicalChitin was treated with 50% NaOH at a 1/5 (w/v) ratio for 2 h at 100 °C.
Shrimp shell wasteAutoclaveOne gram of chitin was added in 50% NaOH and autoclaved for 1 h at 121 °C, 15 psi.-Varun et al. [40]
Fish scales (Red Snapper)ChemicalChitin was treated with 80% NaOH, 1/3 (w/v) for 4 h at 110 °C.90.83%Takarina and Fanani [91]
Fish scales (White Snapper)ChemicalChitin was treated with 80% NaOH for 4 h at 120 °C. 84.05%Takarina et al. [92]
Blue crab shells (Portunus segnis) ChemicalChitin was treated with 12.5 M NaOH, 1/10 (w/v) for 4 h at 140 °C.90.39%Hamdi et al. [93]
Squid pen (Illex argentines) ChemicalChitin was dissolved in 5% acetic acid, filtered, and precipitated with 8 N NaOH solution, washed with reverse osmosis (RO) water until pH reached 7.0.84%Huang et al. [94]
Prawn shellsChemicalChitin was refluxed in 50% NaOH solution for 30–150 min at 100 °C.78.40%Muley et al. [95]
Crab shells (Portunus sanguinolentus)ChemicalChitin was treated with 80% NaOH, 1:20 (w/v) at 90–95 °C for 5h.70.79%Rubini et al. [96]
Shrimp shells (Penaeus monodon)ChemicalOne gram of chitin in 50 mL of 50% NaOH, stirred for 50 min at 90 °C, filtered, and treated with 80% alcohol in 1/30 (w/v) ratio for 24 h at 80 °C.65%Srinivasan et al. [97]
Shrimp wasteChemicalChitosan-1: 40% NaOH 1/20 (w/v) at 120 °C for 300 min.
Chitosan-2: 50% NaOH 1/20 (w/v) at 100 °C for 720 min.
78.2%
84.95%
Tokatli and Demirdöven [98]
Lobster shells (Thenus unimaculatus) ChemicalChitin was added with 40% NaOH, stirred for 6 h at 110 °C, filtered, treated with 10% acetic acid for 12 h.-Arasukumar et al. [99]
Shrimp shells wasteChemicalChitin was treated with 50% NaOH under agitation for 4 h at 90 °C.88%Boudouaia et al. [100]
Shrimp shells (Parapenaeus longirostris)MicrowaveChitin was treated with 40% NaOH, 1:20 (w/v), heated by microwave at 650 W for 12 min.82.8%EL Knidri et al. [45]
Blue crab shell (Callinectes sapidus)ChemicalChitin was treated with 50% NaOH, 1/10 (w/v) at 150 °C for 4 h.71%Metin et al. [101]
Shrimp shellChemicalChitin was treated with 50% NaOH at 60 °C, dry residue was added into 2% (w/w) acetic acid, 30% H2O2
was added and kept for 4 h.
64.18%Ni’mah et al. [102]
Mussel shell35.03%
Squid pen58.04%
Crab shell53.91%
Crab shell wasteChemicalChitin was treated with 50% NaOH, 1/10 (w/v) for 100 min at 100 °C.82%Pădurețu et al. [103]
Shrimp wasteChemicalChitin was added with 50% NaOH, 1/15 (w/v), stirred for 2 h (1 h at RT and 1 h at 100 °C).84%Pădurețu et al. [104]
Squid pens (Loligo formosana)ChemicalChitin was treated with 50% NaOH, 1:50 (w/v) for 8 h at 130 ℃.90%Singh et al. [105]
Shrimp waste (Litopenaeus vannamei)Microwave 1Chitin (16, 32, 60 mesh sizes) was treated 45% NaOH, 1/15 (w/v) irradiated in 6 pulses of 5 min at 600 W.81, 72, 78%Santos et al. [46]
Microwave 2Between each interval, they were stirred for homogenization.81, 92, 89%
Shrimp shellsChemicalChitins were mixed with 40% NaOH, 1:10 (w/v), stirred overnight, and the mixture was heated for 12 h at 100 °C.93%Tolesa et al. [106]
Fish waste (Sardina pilchardus)AutoclaveChitin was added with 40% NaOH and autoclaved for 20 min at 15 psi/121 °C. 87%Aboudamia et al. [41]
Shrimp shell wasteChemicalChitin was added with 50% NaOH, stirred for 1 h at 30 ℃.88.89%Aldila et al. [107]
Shrimp residues (Farfantepenaeus aztecus)UltrasoundTwo grams of shrimp residues powder in 50 mL of CaCl2–MeOH–H2O, stirred for 20 min at 60 °C, ultrasound at 40 kHz 30 min at 60 °C, rest for 48 h at RT.65.87%Borja-Urzola et al. [48]
Crab shell (Portunus trituberculatus)UltrasonicationChitin was treated with 50% NaOH, 1:15 (w/v), at 75 ℃ for 3.5 h with sonication.86.02%Huang et al. [49]
Omani shrimp wasteAutoclaveChitin was treated with 50% NaOH, 1:10 (w/v), autoclaved for 15 min at 121 °C.-Said Al Hoqani et al. [42]
Shrimp shells (Litopenaeus vannamei)ChemicalChitin samples (0–9) were treated with 12.5 M NaOH, 1:5 (w/v) at 65 °C for 12 h.56.10–88.76%Trung et al. [108]
Shrimp shell and Crab ShellsGraviola extract with magnetic stirringShells interacted with graviola leaf extract with magnetic stirring.50.97–94.56%Gopal et al. [50]
Table 2. Various nano-chitosan morphologies and their medical applications.
Table 2. Various nano-chitosan morphologies and their medical applications.
Nano-chitosan MorphologyPreparation MethodApplicationsReferences
NanogelsCovalent cross-linking, ion crosslinking, covalent modificationPhotothermal therapy–chemotheraphy, controlled drug delivery, deep tumor penetration[245,246,247,248]
MicellesCovalent modification/self assembly, ion crosslinkingDrug delivery, photodynamic theraphy, ocular delivery[249,250,251,252,253]
NanofibersElectrospinning processImprove osteogenic activity[254]
LiposomesCovalent modification/self assembly, electrostatic adsorptionAnticancer drug delivery, antimalarial drug delivery, reverse drug resistance, photothermal, and chemotheraphy[255,256,257,258,259]
NanosphereCovalent modification/self assembly, electrostatic adsorption, emulsificationsiRNA delivery, drug delivery, drug release[260,261,262]
NanoparticlesCovalent modification/self assembly, ion crosslinkingOral delivery, siRNA delivery, targeted therapy[263,264,265,266]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Muthu, M.; Gopal, J.; Chun, S.; Devadoss, A.J.P.; Hasan, N.; Sivanesan, I. Crustacean Waste-Derived Chitosan: Antioxidant Properties and Future Perspective. Antioxidants 2021, 10, 228. https://doi.org/10.3390/antiox10020228

AMA Style

Muthu M, Gopal J, Chun S, Devadoss AJP, Hasan N, Sivanesan I. Crustacean Waste-Derived Chitosan: Antioxidant Properties and Future Perspective. Antioxidants. 2021; 10(2):228. https://doi.org/10.3390/antiox10020228

Chicago/Turabian Style

Muthu, Manikandan, Judy Gopal, Sechul Chun, Anna Jacintha Prameela Devadoss, Nazim Hasan, and Iyyakkannu Sivanesan. 2021. "Crustacean Waste-Derived Chitosan: Antioxidant Properties and Future Perspective" Antioxidants 10, no. 2: 228. https://doi.org/10.3390/antiox10020228

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