Next Article in Journal
Physicochemical Characterization, Biocompatibility, and Antibacterial Properties of CMC/PVA/Calendula officinalis Films for Biomedical Applications
Next Article in Special Issue
Preparation of Aloe-Emodin Microcapsules and Its Effect on Antibacterial and Optical Properties of Water-Based Coating
Previous Article in Journal
A Brief Review on Selected Applications of Hybrid Materials Based on Functionalized Cage-like Silsesquioxanes
Previous Article in Special Issue
L-Arginine Grafted Chitosan as Corrosion Inhibitor for Mild Steel Protection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Application Prospects of Chitosan Based Composites for the Metal Contaminated Wastewater Treatment

by
Ashoka Gamage
1,*,
Nepali Jayasinghe
2,
Punniamoorthy Thiviya
3,
M. L. Dilini Wasana
2,4,
Othmane Merah
5,6,
Terrence Madhujith
7 and
Janardhan Reddy Koduru
8,*
1
Chemical and Process Engineering, Faculty of Engineering, University of Peradeniya, Peradeniya 20400, Sri Lanka
2
China Sri Lanka Joint Research and Demonstration Center for Water Technology, Peradeniya 20400, Sri Lanka
3
Postgraduate Institute of Agriculture, University of Peradeniya, Peradeniya 20400, Sri Lanka
4
Department of Agricultural Technology, Faculty of Technology, University of Colombo, Colombo 10200, Sri Lanka
5
Laboratoire de Chimie Agro-Industrielle, LCA, Université de Toulouse, CEDEX 4, 31030 Toulouse, France
6
Département Génie Biologique, IUT Paul Sabatier, Université Paul Sabatier, 32000 Auch, France
7
Department of Food Science and Technology, Faculty of Agriculture, University of Peradeniya, Peradeniya 20400, Sri Lanka
8
Department of Environmental Engineering, Kwangwoon University, Seoul 01897, Republic of Korea
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(6), 1453; https://doi.org/10.3390/polym15061453
Submission received: 6 January 2023 / Revised: 6 March 2023 / Accepted: 12 March 2023 / Published: 14 March 2023

Abstract

:
Heavy metals, known for their toxic nature and ability to accumulate and magnify in the food chain, are a major environmental concern. The use of environmentally friendly adsorbents, such as chitosan (CS)—a biodegradable cationic polysaccharide, has gained attention for removing heavy metals from water. This review discusses the physicochemical properties of CS and its composites and nanocomposites and their potential application in wastewater treatment.

Graphical Abstract

1. Introduction

Various materials, including natural contaminants, fluoride, chloride, nitrate, iron, calcium, magnesium, and other contaminants from byproducts of agriculture and industries such as heavy metals, organic dyes, insecticides and fertilizers, spilled oils, batteries, diesel fuel, household chemicals such as synthetic detergents, and some pathogenic microbes can contaminate the ocean, groundwater and surface water bodies (rivers, lakes, ponds, reservoirs) [1,2,3,4,5,6]. Contaminated water will not be suitable for drinking, habitat, irrigation, recreation, and other industrial activities [4,7].
Heavy metal has gained significant importance in ecotoxicology due to their long persistence, bioaccumulation, and bio-magnification in the food chain [8]. Heavy metal ions are commonly discharged by various industrial activities, including electroplating, battery manufacturing, pesticide production, mining, nuclear power, textile manufacturing, and other similar industries [9,10,11]. The global average content of Cr, Mn, Co, Ni, As and Cd on surface water bodies exceeded the permitted level suggested by WHO and USEPA Guidelines [8]. Heavy metal ions, such as As, Cr, Ni, Co, Hg, Pb, U, and Cd, have become a major threat as they are toxic in nature and accumulate in food chains due to their non-biodegradability and cause various health-related issues even at trace levels. Therefore, in order to safeguard human health and the environment, it is crucial to eliminate heavy metals prior to their release into the surroundings [10,12,13,14,15].
Heavy metal removal can be accomplished through various methods, such as membrane separation, solvent extraction, chemical oxidation-reduction, chemical precipitate, reverse osmosis, flocculation, electrocoagulation, supercritical isothermal treatment, and adsorption [13,16,17]. Among these methods, adsorption has gained wide acceptance due to its high efficiency, simple handling, different adsorbent availability, and cost-effectiveness. In recent years, attention has shifted towards the use of low-cost adsorbents made from renewable resources, considering environmental and cost factors. CS, being a low-cost adsorbent derived from natural sources, has gained interest as a sorbent for heavy metal removal through adsorption in aqueous solutions [13,18]. Adsorbents for heavy metal removal have been widely fabricated from polysaccharides, including chitin/chitosan, cellulose, hemicelluloses, lignin, alginates, carrageenan, granular activated carbon, clays, silica, agricultural waste biomass, metal oxides, various Metal-Organic Frameworks (MOFs), and microbial biomass (probiotics and yeasts) [15,17,19,20,21,22,23].
Chitin can be obtained from shellfish, including shrimp and crabs, and is recognized as the second most abundant polymer in nature, following cellulose [13,24]. Chitin, a linear polymer of N-acetyl-D-glucosamine, is one of the most abundant organic materials produced annually through biosynthesis in animals, especially in crustaceans, mollusks, and insects, as well as some fungi [25,26]. CS (poly[β-(1-4)-2-amino-2-deoxy-d-glucopyranose]) is a derivative of chitin, a polymer of D-glucosamine (Figure 1), which has primary reactive amino groups and hydroxyl groups and can chelate many metal ions. CS is mainly insoluble in water or organic solvents, but at acidic pH, CS becomes a water-soluble cationic polymer as amine gets protonated [25]. Chitin molecule mainly exists in two polymorphic structures: α-and β-chitin. Chitin from crustaceans, including shrimp, prawns, and crab shells, and fungi cell wall consists of an α-crystallographic structure in which the main chains are arranged in an antiparallel fashion with strong intermolecular hydrogen bonding. Chitin is found in diatoms, mollusks and squid pens and has a β-crystallographic structure, in which the chitin chains are arranged in a parallel way with relatively weak intermolecular forces [27,28,29,30]. CS is a natural polyamine made by deacetylating chitin, consisting of poly(β-(1,4)-2-deoxy-2-amino-d-glucose). The physicochemical properties of chitin and CS are affected by several factors, including the degree of acetylation, molecular weight, crystallinity degree, source, and processing method [31].
Today natural materials have emerged as cheaper, sustainable, and more efficient adsorbents in replacement for non-biodegradable synthetic material to treat contaminated water. In this aspect, CS is widely studied as they exhibit a range of characteristics, including biodegradability, non-toxicity, biocompatibility, abundance in nature, low cost, hydrophilicity, high content of functional groups (–NH2, –OH) that can contribute to various chemical modifications, attractive adsorption capacity, and metal ion chelation potential [22,32]. However, drawbacks of CS, including poor mechanical properties, pH sensitivity, low thermal stability, variability of polymer characteristics, poor solubility, low surface area, non-porosity, low adsorption capacity, and lack of reusability, limit their application in wastewater treatment [22,32,33,34]. The physical modification (transform powder form into nanoparticle) and chemical modification (crosslinking, graft modification, etc.) can improve the solubility and physical properties [32,35].
Chitosan has an excellent adsorption capacity for metal ions due to the presence of hydroxyl and amino functional groups, which can form complexes with metal ions via chelation, hydrogen bonding or electrostatic attraction [20]. The adsorption capacity of CS is found to be much higher than that of chitin because of its relatively higher amino groups. Additionally, CS-based composites have been extensively researched for their use in wastewater treatment [36,37,38]. In this context, this review focuses on the application of CS in wastewater treatment, mainly concerning extraction, physicochemical properties of CS, and application of CS on wastewater treatment.

2. Chitin Sources and Composition

To address environmental concerns, the industrial production of chitin and chitosan (CS) must be carried out on a large scale and at a competitive cost [39]. Chitin can be found in various sources, including crustacean exoskeletons (lobsters, shrimps, prawns, crabs, krill, crayfish), mollusks (octopus, cuttlefish, clams, oysters, squids, snails), algae (diatoms, brown algae, green algae), insects cuticles, and fungal cell walls [40]. Presently, crustacean waste, such as shrimp, crabs, prawns, and lobsters, is the primary source of industrial chitin [40,41]. Though, chitin/chitosan from crustaceans has limitations, such as limited raw material supply, a seasonal, higher concentration of CaCO3, requirement of chemical treatment, the possibility of heavy metal contamination, and consume long time. In contrast, chitin/chitosan fungal sources have seasonal independence, superior particle size, uniformity, lower molecular weight, and are free from heavy metals [42]. Table 1 outlines the proximate composition of several chitin sources, with shells comprising chitin, protein, minerals, and pigments [43]. Crustacean (crab, shrimp, and lobster) shell wastes consist of 15–40% chitin, 20–50% CaCO3, and 20–40% protein, along with lipids, pigments, and other minerals in small amounts [44]. Therefore, removing proteins, minerals, and pigments is necessary for chitin production [41].

3. Production of Chitosan from Chitin

The chitin production process can be achieved via four main steps: sample preparation, demineralization (removal of the mineral fraction, mainly CaCO3), deproteinization (removal of protein fraction), and discoloration (removal of carotenoid pigments) (Figure 2). In order to obtain pure chitin without impurities, demineralization, deproteinization, and decolorization are necessary. Demineralization is achieved using acid and deproteinization using alkali, while decolorizing agents such as chloroform, hydrogen peroxide, and acetone are used to remove pigments. It is important to correctly manipulate time, chemical concentration, and temperature to achieve the highest molecular weight of chitin/CS [46]. The deproteinization and demineralization steps can be carried out in reverse order [47], and CS can be produced by deacetylating chitin. A flow diagram for the isolation of chitin and the production of CS is shown in Figure 2.

3.1. Production of Chitin

3.1.1. Preparation of Raw Materials

To remove any residual crustacean flesh, lipids, or other impurities, the shells were initially rinsed with hot tap water or boiling water with stirring using a mechanical stirrer [49]. After that, the shells were washed with hot distilled water and dried in an oven at 60 °C. The dried shells were either ground into small pieces or pulverized and sieved through meshes of size 60–120 µm. Grinding is a crucial step in order to increase the surface area and achieve uniformly-sized particles for the purpose of removing soluble organics and adhered proteins [46,49,50].

3.1.2. Deproteinization

The process of deproteinization is challenging as it involves the breakdown of the chemical bonds that exist between chitin and proteins [51]. The removal of proteins from the grounded crustacean shell waste is achieved through treatment with a hot, weak NaOH solution (1–10%) at a temperature range of 65–100 °C. The deproteinization stage is challenging as it involves breaking down the chemical bonds between chitin and proteins. To achieve deproteinization, one can vary the alkali concentration, time, temperature, and solid-to-solvent ratios [48]. Deproteinization can be accomplished using various reagents that include NaOH, Na2CO3, NaHCO3, KOH, K2CO3, Ca(OH)2, Na2SO3, NaHSO3, CaHSO3, Na3PO4, and Na2S. Nonetheless, NaOH is the most commonly employed reagent for this process [48,52]. The use of chemical treatment can lead to a decrease in the molecular weight of chitin as it may cause partial deacetylation and depolymerization of the biopolymer through hydrolysis [51]. Therefore, chemical treatment is replaced by biological methods, including proteolytic enzymes and organic acids (lactic acid and acetic acid) produced by microbial fermentation. But, considering the possible microbial contamination and difficulty in achieving complete protein removal, chemical treatment is widely used in the deproteinization process [53].

3.1.3. Demineralization

A high content of minerals, i.e., crustaceans’ exoskeleton contains more than 50% of CaCO3, requires a demineralization process [27]. The demineralization process involves the conversion of minerals, mainly CaCO3, into soluble salts using acids (Equation (1)). Dilute HCl is the commonly used acid for demineralization, although other acids such as HNO3, H2SO4, CH3COOH, and HCOOH are also used in this process [48,51].
2 HCl + CaCO3 → CaCl2 + H2O + CO2
Following the production of soluble salts, the chitin solid phase can be filtered and subsequently washed using deionized water, enabling easy separation of the salts [51]. Antifoam can be used to control undesirable foam production due to CO2 generation [54].
The use of mild acids in the demineralization process can help minimize the degradation of chitin, as compared to the use of HCl, which may cause depolymerization and deacetylation of the native chitin [48]. The demineralization process varies based on factors such as the degree of mineralization, extraction time, temperature, particle size, acid concentration, and solute/solvent ratio. The stoichiometric amount of minerals requires two molecules of HCl to convert one molecule of CaCO3 into soluble salts (CaCl2), according to the equation. Due to heterogeneity, the complete removal of minerals is challenging and may require larger amounts or higher concentrations of acids [51]. Demineralization can be carried out by biological method, using acid (e.g., lactic acid) producing bacteria or enzymes, such as Alcalase® [27].

3.1.4. Discoloration

During the processes of deproteinization and demineralization, the resulting chitin product can take on a colored appearance due to the presence of carotenoids found in the exoskeleton of crustaceans [54]. Astaxanthin, astatine, canthaxanthin, lutein, and β-carotene are the main carotenoid components. To produce chitin that is white in color and has added commercial value, the process of discoloration is important. Two steps are involved in the discoloration of shellfish processing discards. The first step involves the use of necessary reagents to extract pigments, while the second step involves the use of appropriate chemicals for bleaching. Organic solvents such as acetone, chloroform, ether, or ethanol are typically employed for pigment extraction. Commonly used bleaching/whitening agents include sodium hypochlorite, potassium permanganate, and hydrogen peroxide [40,48]. After each step (deproteinization, demineralization, and discoloration) sample is collected by filtration, washing to neutralize, and drying [55].

3.2. Production of Chitosan from Chitin (Deacetylation)

Both enzymatic and chemical methods can be employed to convert chitin to CS. However, due to their low cost and suitability for mass production, chemical methods are commonly utilized for the commercial production of CS [51]. Converting chitin to CS can be achieved by partially or completely removing acetyl groups through treatment with concentrated alkali, such as NaOH or KOH (40–50%) at high temperatures (>100 °C) [48]. The use of acids in the deacetylation process is uncommon, as glycosidic bonds are highly susceptible to acid [51]. Finally, the CS can be separated by filtration, washing with distilled water to neutralize, and oven drying [46].
The deacetylation process has an impact on the molecular weight, degree of acetylation, and distribution of acetyl groups in the polymer chain. Several factors that influence the characteristics of CS include the type of alkali reagent used (NaOH being more efficient than KOH), its concentration, temperature, reaction time, repetition of deacetylation steps, atmospheric conditions (nitrogen or air), particle size, source of raw material, chitin/solvent ratio, and the use of a reducing agent (such as sodium borohydride) [51]. Table 2 summarize the degree of deacetylation (or acetylation) at different various process parameters.

4. Characteristics of Chitosan

4.1. Molecular Weight

The N-deacetylation reaction results in molecular weight reduction [60]. Natural chitin typically has a molecular weight exceeding 1000 kDa. Commercially available CS usually ranges from 100–1000 kDa, and its molecular weight is affected by the preparation method and the source of raw materials [61,62]. For example, the molecular weight of CS from insects has reported lower values, ranging from 26–300 kDa [63]. Degradation of CS can be caused by processing factors such as elevated temperature, dissolved oxygen, and shear stress. High temperature or concentrated acids (e.g., HCl, CH3COOH) cause molecular weight changes, while EDTA results in minimal degradation [54]. Techniques such as HPLC and light scattering methods can be used to determine the molecular weight distributions of CS [60].

4.2. Degree of Deacetylation

Chitin is deacetylated by removing the acetyl group, resulting in amino groups (–NH2) that are highly chemically reactive [54]. The degree of deacetylation is a crucial determinant of the physicochemical properties and biodegradation of the material. When the degree of deacetylation of chitin is 50% or higher, it is generally known as CS [64]. The degree of deacetylation is the proportion of glucosamine monomers in the chitosan chain and can affect the solubility and performance of chitosan in various applications [63]. The degree of deacetylation can be determined using different tools, such as infrared (IR) spectroscopy or nuclear magnetic resonance (NMR) spectroscopy [46,60,63].

4.3. Solubility

The strong intra- and intermolecular hydrogen bonding and highly crystalline structure of chitin make it insoluble in water, dilute aqueous salt solutions, and most organic solvents [60,65]. In contrast, CS is a cationic polysaccharide soluble in dilute aqueous acidic solution (below pH 6) [64,65]. Protonation plays an important role in the solubility of CS in acids, such as CH3COOH and HCl, and the pH and the pK of the acid determine the degree of ionization on -NH2 functional group at C-2 of the d-glucosamine repeat unit [66].
Along with the pH, the ionic strength of the solvent determines the degree of protonation or charge density of the amines group, thereby, the solubility of CS. CS is soluble in acidic pH (<pH 6) and insoluble in neutral and basic pH [47]. Organic acids such as acetic, formic, and lactic acids are used for dissolving CS [54]. Table 3 summarizes the frequently used solvents for chitin and CS.
The solubility and applicability of CS depend on two primary factors: the degree of deacetylation and molecular weight. When the degree of acetylation is below 50%, CS obtained from partially deacetylated chitin can dissolve in acidic conditions [46].
The low deacetylated degree of CS (deacetylation degree of 55–70%) is almost completely insoluble in water. Whereas CS with the middle deacetylation degree (70–85%) is partly dissolved in water, and the high deacetylation degree (85–95%) of CS has good solubility in water. Moreover, the ultrahigh deacetylation degree of CS (95–100%) is difficult to obtain [69].
Because of the insolubility of CS in water (at neutral pH), it produces toxic compounds that can pose a threat to the environment. Therefore, CS can be modified to improve the solubility by attaching polar functional groups (-OH and -COOH) to the CS chain and modifying the molecular weight (solubility increased with decreased molecular weight) [70].

5. Neutralization and Chemical and Physical Modification of Chitosan

Chitin and CS act as weak bases and undergo alkaline compound neutralization reactions. In this process, the primary amine group’s nonbonding pair of electrons from the glucosamine units accept protons and become positively charged, as shown in Figure 3 [67,68].
CS is not soluble at neutral and basic pH but can dissolve in acidic pH due to increased polarity and polymer-polymer electrostatic repulsion. In contrast, chitin remains insoluble even in acidic pH because it has fewer amino groups than CS [67].
Due to the presence of a non-bonding electron pair in the primary amino groups, CS functions as a potent nucleophile. Additionally, the primary amine groups at CS exhibit greater nucleophilic characteristics than the primary hydroxyl groups found at the C-6 position [67,68].
Due to the presence of the larger amount of amino (at C-2) and hydroxyl groups (at C-3 and C-6), CS can undergo various chemical modifications, such as acylation, carboxylation and etherification, hydroxylation, phosphorylation, thiolation, methylation, sulphonation, quaternization, crosslinking, graft copolymerization, chelation, oxidation, alkylation etc. This chemical modification generates various CS derivatives with improved solubility (in water or organic solvents), as well as other physicochemical and mechanical properties, which make them a potential material in many applications in various fields [65,68,71].
Furthermore, CS can be modified through various physical modifications into different forms, including solution, powder, flake, fiber, nanofiber, beads, and film, allowing for a wide range of applications across multiple fields, including water treatment (Figure 4) [72,73]. Moreover, to improve adsorption performance, CS has been modified into various forms, such as hydrogel, nanoparticles, and nanofibers, due to low specific area and porosity in powder and flake forms [74].

6. Waste Treatment and Purification of Water

6.1. Chitosan Based Composites to Eliminate Heavy Metal Ions

The non-biodegradability, carcinogenicity, and toxicity at low concentrations make organic dyes and heavy metals the most dangerous water pollutants [75]. Heavy metals, in particular, pose a significant risk to both human health and the environment as they are toxic and can accumulate and magnify in the food chain [12]. To ensure safe water consumption and human activities, it is crucial to eliminate heavy metals from the water before discharging them into the environment [12,75].
To remediate water, different techniques have been employed, such as flotation, coagulation/flocculation, oxidation-reduction, precipitation, membrane techniques (nanofiltration, ultrafiltration, reverse osmosis), ion exchange, electrochemical, photocatalytic degradation, and adsorption [61,76]. Among these methods, adsorption has a high removal rate, easy accessibility, low cost, and reduced secondary pollution [61,77]. Various parameters, such as pH, temperature, contact time, adsorbent dosage, initial metal concentration, and type of sorbent (surface area, porosity, functional group distribution), influence the metal adsorption process [78,79].
In recent years, there has been a growing interest in using CS-based adsorbents for removing heavy metals from water due to their excellent adsorption capacity, reusability, thermal stability, biodegradability, non-toxicity, low cost, and renewability [12].
CS can be used as a chelating agent as well as a flocculant [80]. Due to inter- and intra-molecular hydrogen bonding, CS is poorly soluble in neutral pH. It also has low antioxidant activity due to the absence of H atom donors and is characterized by high hydrophilicity, rigidity, and brittleness. However, these properties can be enhanced through chemical, mechanical/physical, or enzymatic modifications of CS [12].
Several cross-linking agents, including epichlorohydrin, 1,1,3,3-tetra methoxy propane, glycerolpolygglycidyl ether, chloromethyl oxirane, glutaraldehyde, ethylene glycol, diglycidyl ether and tri-polyphosphate, have been utilized [40,81]. The incorporation of new functional groups such as histidine, heparin, succinic anhydride, and N, O carboxymethyl into CS derivatives can improve their metal ion sorption selectivity and adsorption ability [81]. Modified shrimp-based CS, a natural scavenger of heavy metals, exhibited the maximum adsorption capacity of 20.4, 15.9, 7.0, and 6.3 mg/g for Cr, As, Ni, and Co, respectively [56].
Xanthate chitosan cross-linked with epichlorohydrin, which was chemically modified, achieved a maximum adsorption capacity of 43.47 mg/g at 50 °C, as reported in [80]. A CS-based flocculant, CMCTS-g-P(AM-CA), composed of carboxymethyl-CS, acrylamide, and ammonium di-thiocarbamate, was developed for removing heavy metals from wastewater. It exhibited high removal rates of 95.24% and 95.72% for Pb(II) and Cd(II), respectively [82]. Adsorbent fabricated from magnetic thiolated/quaternized-CS composite exhibited high removal efficiency (13.6–235.6 mg/g) for heavy metal ions, including As(V), As(III), Cu(II), Hg(II), Zn(II), Cd(II), and Pb(II), under neutral conditions [77].
Water treatment for heavy metal (As, Pb, Hg, Cd, Cu, Cr) and radionuclide (U, Se, Tc) removal has been accomplished using water-stable adsorbents made from metal-organic frameworks (MOF) and MOF-based composites. These materials possess remarkable porosity and surface area, facilitating adsorption and/or photocatalytic redox processes [83]. Figure 5 depicts CS’s ability to remove heavy metals from wastewater.

6.2. Chitosan Based Nanocomposites to Eliminate Heavy Metal Ions

The outstanding properties of nanomaterials, including high surface areas, numerous active sites, and exceptional sorption capabilities, make them highly valuable for treating wastewater [84,85,86]. In recent years, numerous studies have focused on producing magnetic CS-based nanocomposites for extracting metal ions [87,88,89]. Nano-absorbents made from magnetic CS nanoparticles or nanocomposites have been investigated for their effectiveness in removing metal ions from wastewater. The CS polymer chain contains an increased number of hydroxyl and amino groups, which make it highly selective and capable of strongly chelating metals, enabling efficient sorbent regeneration. As a result, CS nanoparticles have the potential to serve as a reusable absorbent for wastewater treatment [90].
Chitosan-citrate gel beads (CCGBs) containing N, O-carboxymethyl chitosan-coated magnetic nanoparticles (NOCC-MNPs) were developed, exhibiting outstanding adsorption capacity for Cu(II) ions (294.11 mg/g). The chelating ability of NOCC, resulting from the presence of hydroxyl, carboxyl, and amino groups in the carboxymethyl chitosan, was responsible for this result [91].
To enhance the mechanical properties of composite materials, a green and efficient approach involves adding different reinforcing fillers to the polymer matrix [92]. The SiO2/CS composite is capable of adsorbing heavy metal ions in solution, particularly As(V) and Hg(II), with a high performance (198.6 and 204.1 mg/g, respectively). SiO2 possesses stable chemical properties and good mechanical strength and can effectively prevent the adsorbents from agglomerating. In addition, CS NPs coated on SiO2 provide an abundance of functional groups (amino and hydroxyl groups) for heavy metal ions [93]. Incorporating nanofillers such as TiO2 can enhance the adsorption capacity of CS by modifying the polymer’s molecular network and increasing the surface area available for metal adsorption [94]. Razzaz et al. [94] used both entrapped and coating methods to produce chitosan/TiO2 nanofibers, which exhibited high adsorption capacities for Pb(II) and Cu(II) (ranging from 475.50 to 715.70 mg/g).

6.3. Mechanism

Based on the solute affinity, adsorption can occur in three types: physical (physisorption), ion exchange and chemical adsorption (chemisorption). Physical adsorption is the result of mainly Van der Waals attraction due to the electrostatic forces between the adsorbate and the surface of the adsorbent. Ion exchange adsorption depends on electrostatic forces occurring between ions retained on the surface, while chemisorption is the result of covalent bonds occurring via chemical interaction between adsorbate and adsorbent. Chemisorption is highly selective, i.e., it occurs between specific adsorptive and adsorbent species, particularly when the active sites are not blocked by previously adsorbed molecules [79].
CS and its nanocomposites can remove heavy metal ions through various mechanisms, such as electrostatic interaction, ionic exchange, metal chelation, and ion-pair formation [95,96]. The adsorption of CS/Fe3O4 NPs is mainly due to the ion exchange on the hydroxyl groups of CS/Fe3O4 NPs and the chelation action of the abundant amine groups in CS [90]. CS is a natural bio-sorbent for toxic metallic ions due to a large number of hydroxyl and amino groups that serve as sorption sites and the flexibility of the polymer chain structure that enables complexation with metal ions [97]. The chelation capacity of CS depends on the degree of deacetylation, distribution of –NH2 groups, the physical state of CS, the nature of the metal ion, and the pH of the solution [97,98]. Higher degrees of deacetylation result in better chelation [98].
CS is protonated in acidic media and exhibits electrostatic characteristics, and the adsorption occurs via an anion exchange mechanism [99]. To increase the density of sorption sites for metal ion adsorption, CS has been modified through the addition of new functional groups to the CS backbone, resulting in numerous CS derivatives [97].

6.4. Isothermal and Kinetic Model

The equilibrium data of metal ions sorption onto the synthesized CS-based adsorbents can be described using various models, including Freundlich, Langmuir, Redlich–Peterson, Dubinin–Radushkevich (D–R), and Sips isotherm [77,94]. These isotherm models are used to investigate the dominant adsorption mechanism and to calculate adsorption parameters [77].
The isotherm equations are given as follows;
Langmuir : q e = q m K L C e 1 + K L C e
Freundlich : q e = K F C e 1 / n
Sips : q e = q m K S C e 1 / n 1 + K S C e 1 / n
Redlich Peterson : q e = P C e 1 + α C e β
where, qe(mg/g) and Ce (mg/L) are the equilibrium adsorption capacity and the equilibrium concentration of each ion in the solution, qm (mg/g) is the maximum adsorption capacity, KL (L/mg) is Langmuir equilibrium constant. KF [(mg/g)(L/mg)1/n] and n are the Freundlich equilibrium constant, which measures the heterogeneity of surface adsorption sites. The n value ranges from 1–10, indicating that adsorption is favorable. In the Sips isotherm equation, all these parameters keep their meaning. P (L/mg) and α (L/mg) are the isotherm constant of the Redlich–Peterson model, and β is the exponential value ranges from 0–1.
Langmuir and Freundlich are the most commonly adopted model to check the best isotherm fit as they are simple models as compared to the others [100]. Langmuir isotherm model assumes the monolayer surface adsorption on homogenous active sites; each active site can adsorb only a single molecule and reach the maximum adsorption capacity when all sites are occupied [77,91]. The Freundlich isotherm model indicates heterogeneous surface adsorption; both monolayer and multilayer adsorption occurs (Figure 6) [77,94].
Furthermore, adsorption kinetic indicates the rate of adsorption and determines the adsorption reaction mechanism. The pseudo-first-order and the pseudo-second-order kinetic models illustrate the rate of adsorption and its nature, such as chemical or physical adsorption, while intraparticle diffusion. The pseudo-first-order kinetic model suggests that the heavy metals are binding via physical adsorption, whereas the pseudo-second-order kinetic model indicates that the binding is achieved by chemical adsorption. Another model, intraparticle diffusion, gives an idea of whether the mass transfer (diffusion) within the pores limits the adsorption rate or not [55,77,100].
Mi et al. [91] reported that the adsorption of the Cu(II) by magnetic CS-citrate gel beads fitted well with the Freundlich model (R2 = 0.964), which suggests that adsorption occurs on irregular surfaces. Moreover, the adsorption kinetics followed the pseudo-second-order kinetic model, suggesting that Cu(II) is mainly adsorbed by chemisorption [91].
Magnetic graphene oxide/CS [75], Fe3O4/CS/polyethylenimine [101], Fe3O4/CS NPs [90], CS/vanillin CS/ortho-vanillin [97], CS/bentonite [102], and TiO2/CS [94] were well fitted with Langmuir isotherm providing a suggestion of monolayer adsorption of various heavy metals.
CS/EDTA complex was also well fitted to the monolayer Langmuir isotherm, with maximum adsorption capacities of 227.27–370.37 mg/g for Pb(II), Cd(II), and Cu(II). The adsorption followed the pseudo-second-order kinetics and was attributed to the electrostatic interactions between the metals and different functional groups (single bond –OH, –NH2, and –COOH) of the adsorbent and complexation with EDTA [103]. Fe3O4/CS NPs have a strong metal chelating capability, which may be attributed to the strong metal chelating capacity of CS, and the magnetic CS can be efficiently separated the external magnetic field from the media [73,90].
As shown in Table 4, most of the CS-based adsorbents obeyed the pseudo-second-order kinetic model, indicating that the adsorption mechanism of heavy metal is chemisorption, which may be achieved through chelation or electrostatic attraction.

6.5. Reusability or Regeneration

Regeneration and reusability are performed by repeating the adsorption and desorption cycles and are vital criteria in commercial applications, and provide economic and environmental benefits [75,99]. For regeneration or reuse of adsorbents, it can be obtained by washing with various desorption reagents, such as acids (HCl, H2SO4, HNO3, H3PO4), alkalis (NaOH, NH4OH), salts (NaCl, Na2CO3, Na2SO4, KNO3), and chelating (EDTA, Na2EDTA) agents followed by deionized water, and oven drying (60 °C for 1 h). [93,98]. Algethami et al. [104] reported that HCl (91.3%) was the best eluent for the desorption of Cd(II) compared to that of CH3COOH (47.48%) and HNO3 (87.26%).
The removal efficiency of Fe3O4/CS/PEI nanosorbent using NaOH as eluent was reported as above 90.0% after the 5th cycle [101]. The removal efficiency for heavy metals was reported as >90% even after the 5th cycle in Fe3O4/CS NPs [90,101], Magnetic thiolated/quaternized-CS [77], and CS-EDTA [103].
In contrast, a severe collapse also was (44%) observed in removal efficiency with CS/Fe3O4 NPs at the end of the fifth run due to a decline of the NPs mass and lower efficiency of CS modification after several times of washing as well as the irreversible adsorption on the adsorbent active sites or partial regeneration of Cr(IV) by NaOH solution [55]. In another study, compared to TiO2-entrapped CS nanofibers, more than 40% loss of total adsorption was reported after the 5th cycle in TiO2-coated CS nanofibers due to the physical loss of TiO2 NPs by acid cleavage [94]. Sutirman et al. [105] developed crosslinked CS grafted with methyl methacrylate (M-CS) for the removal of Cu(II) ions. The removal efficiency was significantly reduced to 80% after four consecutive adsorption-desorption cycles may be attributed to damage to the structure and the loss of active sites during the reusability processes using HNO3. Shahraki et al. [106] also reported adsorption capacity after 5 cycles as 118 mg/g, which was 51% of the initial value.
Table 4 summarizes the previous studies on CS composites for removing several heavy metals from aqueous solutions.
Table 4. Literature survey on CS composites for the removal of heavy metals from aqueous solutions.
Table 4. Literature survey on CS composites for the removal of heavy metals from aqueous solutions.
AdsorbentAdsorbateKineticsIsothermpHTemperature (°C)Adsorption Capacity (mg/g)Reference
CS/Fe3O4 NPs Cr(VI)PSOL3.37298 K162[55]
Zeolitic imidazolate framework-67 (ZIF-67)/modified bacterial cellulose/CS aerogelCu(II)PSO-625200.6[61]
Cr(VI)152.1
Magnetic graphene oxide/CS (Fe3O4/GO/CS) Ni(II) PFOL 62580.48[75]
Magnetic thiolated/quaternized-CS Pb(II)PSOS730 235.63[77]
As(III)67.69
As(V)66.27
Hg(II) 28.00
Cu(II) 33.99
Fe3O4/CS NPsPb(II)-L6-79.24[90]
Cd(II)36.42
N,O-carboxymethyl chitosan-coated magnetic nanoparticles (NOCC-MNPs)/chitosan-citrate gel beads (CCGBs)Cu(II)PSOF 294.11[91]
SiO2@CSAs(V)PSOF6–7298 K198.6[93]
Hg(II)L204.1
TiO2/CSCu(II)PFOL645526.5–715.7[94]
Pb(II)475.5–579.1
CS/vanillin
CS/ortho-vanillin
Co(II)PSO L4 30 5.899–7.651[97]
Fe3O4/CS/polyethylenimine (PEI)As(III)PSOL6.73077.61[101]
As(V)86.50
Polymer composite (CS-EDTA)Pb(II)PSOL 370.37[103]
Cd(II)243.90
Cu(II)227.27
Dimercaptosuccinic acid-functionalized magnetic CS (Fe3O4@CS@DMSA)Cd(II)PFOL 7.6 314.12[104]
Crosslinked CS grafted with methyl methacrylate (M-CS)Cu(II)PSOL 4 RT 192.31[105]
Magnetically modified CS/3,3-diphenylpropylimine methyl benzaldehyde (PPIMB)Pb(II)PSOL 230.48[106]
Magnetic xanthate-modified CS/polyacrylic acid Cu(II)PSOL 5.530206 [107]
Cd(II)-178
Pb(II)- 168
Co(II)- 140
Glucan/CSCu(II)PSOL725342[108]
Co(II)232
Ni(II)184
Pb(II)395
Cd(II)269
CS/calcium alginate/bentonitePb(II)PSOR5 434.89[109]
Cu(II)115.30
Cd(II)102.38
CS microspheres/sodium alginate hybrid beads Pb(II)PFOL 180[110]
Cr(VI)PSOL 16
CS modified with carboxyl groupsCu(II) L3.525220.5[111]
Zn(II) 124.3
AgNPs/GO/CS nanocompositeMn(II)PSOF6 30 1605[112]
Microfluidically-generated CS microspheresCu(II) PSOL5.5 35 75.52[113]
CS grafted UiO-66-NH2Cu(II)----364.96[114]
Pb(II)555.56
CS-g-acrylamide-orange
peel
Cr(VI)PSOF4 28 178.34[115]
Cu(II)5 28 181.88
PFO, Pseudo-first-order; PSO, Pseudo-second-order; RT, Room temperature; F, Freundlich; L, Langmuir; R, Redlich–Peterson; S, Sips isotherm model.

7. Conclusions and Future Prospective

Heavy metals are a significant water pollutant and a serious threat to human health and the environment because of their non-biodegradability, carcinogenicity, and toxicity, even at low concentrations. Heavy metal pollution in water can result in severe health problems caused by the accumulation of toxic heavy metals through the food chain. Hence, it is necessary to remove heavy metals from the water before discharging it into the environment, ensuring safe water consumption and human activities.
CS-based materials have been extensively studied as eco-friendly and biodegradable adsorbents for water treatment. CS contains hydroxyl and amine groups in the glucosamine unit that facilitate metal adsorption through various mechanisms, such as metal chelation, electrostatic interaction, ionic exchange, and ion-pair formation. CS is soluble in acidic pH and insoluble in neutral and basic pH. The modification of CS through physical, chemical, and biological methods improves its physicochemical properties and metal adsorption capacity, making it a promising material for heavy metal removal from water.
CS-based materials have shown great potential as a sustainable solution for the removal of heavy metals from water. In the future, further research and development are needed to optimize the performance of these materials, such as improving their mechanical properties and enhancing their selectivity for specific metals. There is also a need to explore novel approaches for the synthesis and functionalization of chitosan-based materials, such as using green chemistry and biotechnology methods. Additionally, the scalability and cost-effectiveness of chitosan-based materials need to be addressed for their commercialization and widespread application. Overall, the future prospects of chitosan-based materials for heavy metal removal from water are promising, and continued research and development in this area could lead to a more sustainable and effective solution for water treatment.

Author Contributions

Original draft, A.G., N.J., P.T., M.L.D.W., O.M. and T.M. Writing—review and editing, J.R.K., A.G., N.J., P.T., M.L.D.W., O.M. and T.M. Supervision and Funding Acquisition, J.R.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

J.R.K. would like to gratefully acknowledge the NRF, Korea, for funding this research study through the Ministry of Science and ICT (2021R1F1A106379311).

Conflicts of Interest

The authors declare no conflict of interest. This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

References

  1. Ma, X.; Zhao, S.; Tian, Z.; Duan, G.; Pan, H.; Yue, Y.; Li, S.; Jian, S.; Yang, W.; Liu, K.; et al. MOFs meet wood: Reusable magnetic hydrophilic composites toward efficient water treatment with super-high dye adsorption capacity at high dye concentration. Chem. Eng. J. 2022, 446, 136851. [Google Scholar] [CrossRef]
  2. Mahmoodi, N.M.; Arami, M.; Bahrami, H.; Khorramfar, S. Novel biosorbent (Canola hull): Surface characterization and dye removal ability at different cationic dye concentrations. Desalination 2010, 264, 134–142. [Google Scholar] [CrossRef]
  3. Mahmoodi, N.M.; Arami, M. Modeling and sensitivity analysis of dyes adsorption onto natural adsorbent from colored textile wastewater. J. Appl. Polym. Sci. 2008, 109, 4043–4048. [Google Scholar] [CrossRef]
  4. Schweitzer, L.; Noblet, J. Chapter 3.6—Water Contamination and Pollution. In Green Chemistry; Török, B., Dransfield, T., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 261–290. [Google Scholar] [CrossRef]
  5. Sharma, S.; Bhattacharya, A. Drinking water contamination and treatment techniques. Appl. Water Sci. 2017, 7, 1043–1067. [Google Scholar] [CrossRef]
  6. Yu, H.; Wu, M.; Duan, G.; Gong, X. One-step fabrication of eco-friendly superhydrophobic fabrics for high-efficiency oil/water separation and oil spill cleanup. Nanoscale 2022, 14, 1296–1309. [Google Scholar] [CrossRef]
  7. Zheng, Q.; Li, Z.; Watanabe, M. Production of solid fuels by hydrothermal treatment of wastes of biomass, plastic, and biomass/plastic mixtures: A review. J. Bioresour. Bioprod. 2022, 7, 221–244. [Google Scholar] [CrossRef]
  8. Kumar, V.; Parihar, R.D.; Sharma, A.; Bakshi, P.; Singh Sidhu, G.P.; Bali, A.S.; Karaouzas, I.; Bhardwaj, R.; Thukral, A.K.; Gyasi-Agyei, Y.; et al. Global evaluation of heavy metal content in surface water bodies: A meta-analysis using heavy metal pollution indices and multivariate statistical analyses. Chemosphere 2019, 236, 124364. [Google Scholar] [CrossRef]
  9. Velusamy, S.; Roy, A.; Sundaram, S.; Kumar Mallick, T. A Review on Heavy Metal Ions and Containing Dyes Removal Through Graphene Oxide-Based Adsorption Strategies for Textile Wastewater Treatment. Chem. Rec. 2021, 21, 1570–1610. [Google Scholar] [CrossRef]
  10. Wan Ngah, W.S.; Hanafiah, M.A.K.M. Removal of heavy metal ions from wastewater by chemically modified plant wastes as adsorbents: A review. Bioresour. Technol. 2008, 99, 3935–3948. [Google Scholar] [CrossRef]
  11. Sun, Y.; Liu, R.; Wen, S.; Wang, J.; Chen, L.; Yan, B.; Peng, S.; Ma, C.; Cao, X.; Ma, C.; et al. Antibiofouling Ultrathin Poly(amidoxime) Membrane for Enhanced U(VI) Recovery from Wastewater and Seawater. ACS Appl. Mater. Interfaces 2021, 13, 21272–21285. [Google Scholar] [CrossRef]
  12. Begum, S.; Yuhana, N.Y.; Md Saleh, N.; Kamarudin, N.H.N.; Sulong, A.B. Review of chitosan composite as a heavy metal adsorbent: Material preparation and properties. Carbohydr. Polym. 2021, 259, 117613. [Google Scholar] [CrossRef]
  13. Malayoglu, U. Removal of heavy metals by biopolymer (chitosan)/nanoclay composites. Sep. Sci. Technol. 2018, 53, 2741–2749. [Google Scholar] [CrossRef]
  14. Pooja, G.; Kumar, P.S.; Indraganti, S. Recent advancements in the removal/recovery of toxic metals from aquatic system using flotation techniques. Chemosphere 2022, 287, 132231. [Google Scholar] [CrossRef] [PubMed]
  15. Gao, J.; Yuan, Y.; Yu, Q.; Yan, B.; Qian, Y.; Wen, J.; Ma, C.; Jiang, S.; Wang, X.; Wang, N. Bio-inspired antibacterial cellulose paper–poly(amidoxime) composite hydrogel for highly efficient uranium(VI) capture from seawater. Chem. Commun. 2020, 56, 3935–3938. [Google Scholar] [CrossRef]
  16. Tavakoli, O.; Goodarzi, V.; Saeb, M.R.; Mahmoodi, N.M.; Borja, R. Competitive removal of heavy metal ions from squid oil under isothermal condition by CR11 chelate ion exchanger. J. Hazard. Mater. 2017, 334, 256–266. [Google Scholar] [CrossRef]
  17. Yang, W.; Wang, Y.; Wang, Q.; Wu, J.; Duan, G.; Xu, W.; Jian, S. Magnetically separable and recyclable Fe3O4@PDA covalent grafted by l-cysteine core-shell nanoparticles toward efficient removal of Pb2+. Vacuum 2021, 189, 110229. [Google Scholar] [CrossRef]
  18. Ranjbar-Mohammadi, M.; Arami, M.; Bahrami, H.; Mazaheri, F.; Mahmoodi, N.M. Grafting of chitosan as a biopolymer onto wool fabric using anhydride bridge and its antibacterial property. Colloids Surf. B Biointerfaces 2010, 76, 397–403. [Google Scholar] [CrossRef]
  19. Jjagwe, J.; Olupot, P.W.; Menya, E.; Kalibbala, H.M. Synthesis and Application of Granular Activated Carbon from Biomass Waste Materials for Water Treatment: A Review. J. Bioresour. Bioprod. 2021, 6, 292–322. [Google Scholar] [CrossRef]
  20. Khanniri, E.; Yousefi, M.; Mortazavian, A.M.; Khorshidian, N.; Sohrabvandi, S.; Arab, M.; Koushki, M.R. Effective removal of lead (II) using chitosan and microbial adsorbents: Response surface methodology (RSM). Int. J. Biol. Macromol. 2021, 178, 53–62. [Google Scholar] [CrossRef]
  21. Mahmoodi, N.M.; Taghizadeh, M.; Taghizadeh, A.; Abdi, J.; Hayati, B.; Shekarchi, A.A. Bio-based magnetic metal-organic framework nanocomposite: Ultrasound-assisted synthesis and pollutant (heavy metal and dye) removal from aqueous media. Appl. Surf. Sci. 2019, 480, 288–299. [Google Scholar] [CrossRef]
  22. Sudha, P.N.; Gomathi, T.; Vinodhini, P.A.; Nasreen, K. Chapter Seven—Marine Carbohydrates of Wastewater Treatment. In Advances in Food and Nutrition Research; Kim, S.-K., Ed.; Academic Press: Cambridge, MA, USA, 2014; Volume 73, pp. 103–143. [Google Scholar] [CrossRef]
  23. Zia, Z.; Hartland, A.; Mucalo, M.R. Use of low-cost biopolymers and biopolymeric composite systems for heavy metal removal from water. Int. J. Environ. Sci. Technol. 2020, 17, 4389–4406. [Google Scholar] [CrossRef]
  24. Zahedi, S.; Safaei Ghomi, J.; 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]
  25. Chawla, S.P.; Kanatt, S.R.; Sharma, A.K. Chitosan. In Polysaccharides: Bioactivity and Biotechnology; Ramawat, K.G., Mérillon, J.-M., Eds.; Springer International Publishing: Cham, Switzerland, 2015; pp. 219–246. [Google Scholar] [CrossRef]
  26. Zhao, D.; Huang, W.-C.; Guo, N.; Zhang, S.; Xue, C.; Mao, X. Two-Step Separation of Chitin from Shrimp Shells Using Citric Acid and Deep Eutectic Solvents with the Assistance of Microwave. Polymers 2019, 11, 409. [Google Scholar] [CrossRef] [PubMed]
  27. Bastiaens, L.; Soetemans, L.; D’Hondt, E.; Elst, K. Sources of Chitin and Chitosan and their Isolation. In Chitin and Chitosan; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2019; pp. 1–34. [Google Scholar] [CrossRef]
  28. Hou, J.; Aydemir, B.E.; Dumanli, A.G. Understanding the structural diversity of chitins as a versatile biomaterial. Philos. Trans. R. Soc. Math. Phys. Eng. Sci. 2021, 379, 20200331. [Google Scholar] [CrossRef] [PubMed]
  29. Tolaimate, A.; Desbrières, J.; Rhazi, M.; Alagui, A.; Vincendon, M.; Vottero, P. On the influence of deacetylation process on the physicochemical characteristics of chitosan from squid chitin. Polymer 2000, 41, 2463–2469. [Google Scholar] [CrossRef]
  30. 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]
  31. Birolli, W.G.; Delezuk, J.A.D.M.; Campana-Filho, S.P. Ultrasound-assisted conversion of alpha-chitin into chitosan. Appl. Acoust. 2016, 103, 239–242. [Google Scholar] [CrossRef]
  32. Zhang, Y.; Zhao, M.; Cheng, Q.; Wang, C.; Li, H.; Han, X.; Fan, Z.; Su, G.; Pan, D.; Li, Z. Research progress of adsorption and removal of heavy metals by chitosan and its derivatives: A review. Chemosphere 2021, 279, 130927. [Google Scholar] [CrossRef] [PubMed]
  33. Benettayeb, A.; Ghosh, S.; Usman, M.; Seihoub, F.Z.; Sohoo, I.; Chia, C.H.; Sillanpää, M. Some Well-Known Alginate and Chitosan Modifications Used in Adsorption: A Review. Water 2022, 14, 1353. [Google Scholar] [CrossRef]
  34. Eltaweil, A.S.; Omer, A.M.; El-Aqapa, H.G.; Gaber, N.M.; Attia, N.F.; El-Subruiti, G.M.; Mohy-Eldin, M.S.; Abd El-Monaem, E.M. Chitosan based adsorbents for the removal of phosphate and nitrate: A critical review. Carbohydr. Polym. 2021, 274, 118671. [Google Scholar] [CrossRef]
  35. Davarpanah, S.; Mahmoodi, N.M.; Arami, M.; Bahrami, H.; Mazaheri, F. Environmentally friendly surface modification of silk fiber: Chitosan grafting and dyeing. Appl. Surf. Sci. 2009, 255, 4171–4176. [Google Scholar] [CrossRef]
  36. Omer, A.M.; Dey, R.; Eltaweil, A.S.; Abd El-Monaem, E.M.; Ziora, Z.M. Insights into recent advances of chitosan-based adsorbents for sustainable removal of heavy metals and anions. Arab. J. Chem. 2022, 15, 103543. [Google Scholar] [CrossRef]
  37. Janani, R.; Gurunathan, B.; Sivakumar, K.; Varjani, S.; Ngo, H.H.; Gnansounou, E. Advancements in heavy metals removal from effluents employing nano-adsorbents: Way towards cleaner production. Environ. Res. 2022, 203, 111815. [Google Scholar] [CrossRef]
  38. Weißpflog, J.; Gündel, A.; Vehlow, D.; Steinbach, C.; Müller, M.; Boldt, R.; Schwarz, S.; Schwarz, D. Solubility and Selectivity Effects of the Anion on the Adsorption of Different Heavy Metal Ions onto Chitosan. Molecules 2020, 25, 2482. [Google Scholar] [CrossRef] [PubMed]
  39. Trung, T.S.; Tram, L.H.; 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]
  40. Pellis, A.; Guebitz, G.M.; Nyanhongo, G.S. Chitosan: Sources, Processing and Modification Techniques. Gels 2022, 8, 393. [Google Scholar] [CrossRef] [PubMed]
  41. Tokatlı, K.; Demirdöven, A. Optimization of chitin and chitosan production from shrimp wastes and characterization. J. Food Process. Preserv. 2018, 42, e13494. [Google Scholar] [CrossRef]
  42. Huq, T.; Khan, A.; Brown, D.; Dhayagude, N.; He, Z.; Ni, Y. Sources, production and commercial applications of fungal chitosan: A review. J. Bioresour. Bioprod. 2022, 7, 85–98. [Google Scholar] [CrossRef]
  43. Singh Dhillon, G.; Kaur, S.; Jyoti Sarma, S.; Kaur Brar, S.; Verma, M.; Yadagiri Surampalli, R. Recent Development in Applications of Important Biopolymer Chitosan in Biomedicine, Pharmaceuticals and Personal Care Products. Curr. Tissue Eng. 2013, 2, 20–40. [Google Scholar] [CrossRef]
  44. Yang, H.; Yan, N. Transformation of Seafood Wastes into Chemicals and Materials. In Green Chemistry and Chemical Engineering; Han, B., Wu, T., Eds.; Springer: New York, NY, USA, 2019; pp. 461–482. [Google Scholar] [CrossRef]
  45. El Knidri, H.; Belaabed, R.; Addaou, A.; Laajeb, A.; Lahsini, A. Extraction, Chemical Modification and Characterization of Chitin and Chitosan. Int. J. Biol. Macromol. 2018, 120, 1181–1189. [Google Scholar] [CrossRef]
  46. 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]
  47. Stephen, A.M.; Phillips, G.O. Food Polysaccharides and Their Applications, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2016. [Google Scholar]
  48. No, H.K.; Meyers, S.P. Preparation and characterization of chitin and chitosan—A review. J. Aquat. Food Prod. Technol. 1995, 4, 27–52. [Google Scholar] [CrossRef]
  49. 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]
  50. Shahidi, F. Seafood processing by-products. In Seafoods: Chemistry, Processing Technology and Quality; Springer: Dordrecht, The Netherlands, 1994; pp. 320–334. [Google Scholar] [CrossRef]
  51. 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]
  52. Simpson, B.K.; Haard, N.F. The use of proteolytic enzymes to extract carotenoproteins from shrimp wastes. J. Appl. Biochem. 1985, 7, 212–222. [Google Scholar]
  53. Özel, N.; Elibol, M. A review on the potential uses of deep eutectic solvents in chitin and chitosan related processes. Carbohydr. Polym. 2021, 262, 117942. [Google Scholar] [CrossRef]
  54. Yeul, V.S.; Rayalu, S.S. Unprecedented Chitin and Chitosan: A Chemical Overview. J. Polym. Environ. 2013, 21, 606–614. [Google Scholar] [CrossRef]
  55. Pourmortazavi, S.M.; Sahebi, H.; Zandavar, H.; Mirsadeghi, S. Fabrication of Fe3O4 nanoparticles coated by extracted shrimp peels chitosan as sustainable adsorbents for removal of chromium contaminates from wastewater: The design of experiment. Compos. Part B Eng. 2019, 175, 107130. [Google Scholar] [CrossRef]
  56. Rahman, A.; Haque, M.A.; Ghosh, S.; Shinu, P.; Attimarad, M.; Kobayashi, G. Modified Shrimp-Based Chitosan as an Emerging Adsorbent Removing Heavy Metals (Chromium, Nickel, Arsenic, and Cobalt) from Polluted Water. Sustainability 2023, 15, 2431. [Google Scholar] [CrossRef]
  57. Iber, B.T.; Torsabo, D.; Chik, C.E.N.C.E.; Wahab, F.; Sheikh Abdullah, S.R.; Abu Hassan, H.; Kasan, N.A. Response Surface Methodology (RSM) Approach to Optimization of Coagulation-Flocculation of Aquaculture Wastewater Treatment Using Chitosan from Carapace of Giant Freshwater Prawn Macrobrachium rosenbergii. Polymers 2023, 15, 1058. [Google Scholar] [CrossRef]
  58. Mondal, M.I.H.; Ahmed, F. Cellulosic fibres modified by chitosan and synthesized ecofriendly carboxymethyl chitosan from prawn shell waste. J. Text. Inst. 2020, 111, 49–59. [Google Scholar] [CrossRef]
  59. Sarbon, N.M.; Sandanamsamy, S.; Kamaruzaman, S.F.S.; Ahmad, F. Chitosan extracted from mud crab (Scylla olivicea) shells: Physicochemical and antioxidant properties. J. Food Sci. Technol. 2015, 52, 4266–4275. [Google Scholar] [CrossRef]
  60. Fernández-Saiz, P.; Lagaron, J.M. Chitosan for Film and Coating Applications. In Biopolymers—New Materials for Sustainable Films and Coatings; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2011; pp. 87–105. [Google Scholar] [CrossRef]
  61. Li, D.; Tian, X.; Wang, Z.; Guan, Z.; Li, X.; Qiao, H.; Ke, H.; Luo, L.; Wei, Q. Multifunctional adsorbent based on metal-organic framework modified bacterial cellulose/chitosan composite aerogel for high efficient removal of heavy metal ion and organic pollutant. Chem. Eng. J. 2020, 383, 123127. [Google Scholar] [CrossRef]
  62. Chen, C.S.; Liau, W.Y.; Tsai, G.J. Antibacterial effects of N-sulfonated and N-sulfobenzoyl chitosan and application to oyster preservation. J. Food Prot. 1998, 61, 1124–1128. [Google Scholar] [CrossRef]
  63. Hahn, T.; Tafi, E.; Paul, A.; Salvia, R.; Falabella, P.; Zibek, S. Current state of chitin purification and chitosan production from insects. J. Chem. Technol. Biotechnol. 2020, 95, 2775–2795. [Google Scholar] [CrossRef]
  64. Sikorski, D.; Gzyra-Jagieła, K.; Draczyński, Z. The Kinetics of Chitosan Degradation in Organic Acid Solutions. Mar. Drugs 2021, 19, 236. [Google Scholar] [CrossRef] [PubMed]
  65. Pakizeh, M.; Moradi, A.; Ghassemi, T. Chemical extraction and modification of chitin and chitosan from shrimp shells. Eur. Polym. J. 2021, 159, 110709. [Google Scholar] [CrossRef]
  66. Rinaudo, M. Chitin and chitosan: Properties and applications. Prog. Polym. Sci. 2006, 31, 603–632. [Google Scholar] [CrossRef]
  67. Winterowd, J.G.; Sandford, P.A. Food Polysaccharides and Their Applications; Stephen, A.M., Ed.; Marcel Dekker: New York, NY, USA, 1995. [Google Scholar]
  68. 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]
  69. Lv, S.H. 7—High-performance superplasticizer based on chitosan. In Biopolymers and Biotech Admixtures for Eco-Efficient Construction Materials; Pacheco-Torgal, F., Ivanov, V., Karak, N., Jonkers, H., Eds.; Woodhead Publishing: Cambridge, UK, 2016; pp. 131–150. [Google Scholar]
  70. Panda, P.K.; Yang, J.-M.; Chang, Y.-H.; Su, W.-W. Modification of different molecular weights of chitosan by p-Coumaric acid: Preparation, characterization and effect of molecular weight on its water solubility and antioxidant property. Int. J. Biol. Macromol. 2019, 136, 661–667. [Google Scholar] [CrossRef]
  71. Li, B.; Elango, J.; Wu, W. Recent Advancement of Molecular Structure and Biomaterial Function of Chitosan from Marine Organisms for Pharmaceutical and Nutraceutical Application. Appl. Sci. 2020, 10, 4719. [Google Scholar] [CrossRef]
  72. El-hefian, E.A.; Nasef, M.M.; Yahaya, A.H. Chitosan physical forms: A short review. Aust. J. Basic Appl. Sci. 2011, 5, 670–677. [Google Scholar]
  73. Shaumbwa, V.R.; Liu, D.; Archer, B.; Li, J.; Su, F. Preparation and application of magnetic chitosan in environmental remediation and other fields: A review. J. Appl. Polym. Sci. 2021, 138, 51241. [Google Scholar] [CrossRef]
  74. Upadhyay, U.; Sreedhar, I.; Singh, S.A.; Patel, C.M.; Anitha, K.L. Recent advances in heavy metal removal by chitosan based adsorbents. Carbohydr. Polym. 2021, 251, 117000. [Google Scholar] [CrossRef] [PubMed]
  75. Le, T.T.N.; Le, V.T.; Dao, M.U.; Nguyen, Q.V.; Vu, T.T.; Nguyen, M.H.; Tran, D.L.; Le, H.S. Preparation of magnetic graphene oxide/chitosan composite beads for effective removal of heavy metals and dyes from aqueous solutions. Chem. Eng. Commun. 2019, 206, 1337–1352. [Google Scholar] [CrossRef]
  76. Saleh, T.A.; Mustaqeem, M.; Khaled, M. Water treatment technologies in removing heavy metal ions from wastewater: A review. Environ. Nanotechnol. Monit. Manag. 2022, 17, 100617. [Google Scholar] [CrossRef]
  77. Song, X.; Li, L.; Zhou, L.; Chen, P. Magnetic thiolated/quaternized-chitosan composites design and application for various heavy metal ions removal, including cation and anion. Chem. Eng. Res. Des. 2018, 136, 581–592. [Google Scholar] [CrossRef]
  78. Anderson, A.; Anbarasu, A.; Pasupuleti, R.R.; Manigandan, S.; Praveenkumar, T.R.; Aravind Kumar, J. Treatment of heavy metals containing wastewater using biodegradable adsorbents: A review of mechanism and future trends. Chemosphere 2022, 295, 133724. [Google Scholar] [CrossRef]
  79. Samoila, P.; Humelnicu, A.C.; Ignat, M.; Cojocaru, C.; Harabagiu, V. Chitin and Chitosan for Water Purification. In Chitin and Chitosan; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2019; pp. 429–460. [Google Scholar]
  80. Wu, Z.-B.; Ni, W.-M.; Guan, B.-H. Application of chitosan as flocculant for coprecipitation of Mn(II) and suspended solids from dual-alkali FGD regenerating process. J. Hazard. Mater. 2008, 152, 757–764. [Google Scholar] [CrossRef]
  81. Kannamba, B.; Reddy, K.L.; AppaRao, B.V. Removal of Cu(II) from aqueous solutions using chemically modified chitosan. J. Hazard. Mater. 2010, 175, 939–948. [Google Scholar] [CrossRef]
  82. Sun, Y.; Zhou, S.; Sun, W.; Zhu, S.; Zheng, H. Flocculation activity and evaluation of chitosan-based flocculant CMCTS-g-P(AM-CA) for heavy metal removal. Sep. Purif. Technol. 2020, 241, 116737. [Google Scholar] [CrossRef]
  83. Feng, M.; Zhang, P.; Zhou, H.-C.; Sharma, V.K. Water-stable metal-organic frameworks for aqueous removal of heavy metals and radionuclides: A review. Chemosphere 2018, 209, 783–800. [Google Scholar] [CrossRef] [PubMed]
  84. Xu, P.; Zeng, G.M.; Huang, D.L.; Feng, C.L.; Hu, S.; Zhao, M.H.; Lai, C.; Wei, Z.; Huang, C.; Xie, G.X.; et al. Use of iron oxide nanomaterials in wastewater treatment: A review. Sci. Total Environ. 2012, 424, 1–10. [Google Scholar] [CrossRef] [PubMed]
  85. Ray, P.Z.; Shipley, H.J. Inorganic nano-adsorbents for the removal of heavy metals and arsenic: A review. RSC Adv. 2015, 5, 29885–29907. [Google Scholar] [CrossRef]
  86. Santhosh, C.; Velmurugan, V.; Jacob, G.; Jeong, S.K.; Grace, A.N.; Bhatnagar, A. Role of nanomaterials in water treatment applications: A review. Chem. Eng. J. 2016, 306, 1116–1137. [Google Scholar] [CrossRef]
  87. Charpentier, T.V.J.; Neville, A.; Lanigan, J.L.; Barker, R.; Smith, M.J.; Richardson, T. Preparation of Magnetic Carboxymethylchitosan Nanoparticles for Adsorption of Heavy Metal Ions. ACS Omega 2016, 1, 77–83. [Google Scholar] [CrossRef]
  88. Jiang, Z.; Li, N.; Li, P.-Y.; Liu, B.; Lai, H.-J.; Jin, T. One-Step Preparation of Chitosan-Based Magnetic Adsorbent and Its Application to the Adsorption of Inorganic Arsenic in Water. Molecules 2021, 26, 1785. [Google Scholar] [CrossRef]
  89. Karimi, F.; Ayati, A.; Tanhaei, B.; Sanati, A.L.; Afshar, S.; Kardan, A.; Dabirifar, Z.; Karaman, C. Removal of metal ions using a new magnetic chitosan nano-bio-adsorbent; A powerful approach in water treatment. Environ. Res. 2022, 203, 111753. [Google Scholar] [CrossRef]
  90. Fan, H.L.; Zhou, S.F.; Jiao, W.Z.; Qi, G.S.; Liu, Y.Z. Removal of heavy metal ions by magnetic chitosan nanoparticles prepared continuously via high-gravity reactive precipitation method. Carbohydr. Polym. 2017, 174, 1192–1200. [Google Scholar] [CrossRef]
  91. Mi, F.L.; Wu, S.J.; Chen, Y.C. Combination of carboxymethyl chitosan-coated magnetic nanoparticles and chitosan-citrate complex gel beads as a novel magnetic adsorbent. Carbohydr. Polym. 2015, 131, 255–263. [Google Scholar] [CrossRef]
  92. Zhang, H.; Li, Y.; Shi, R.; Chen, L.; Fan, M. A robust salt-tolerant superoleophobic chitosan/nanofibrillated cellulose aerogel for highly efficient oil/water separation. Carbohydr. Polym. 2018, 200, 611–615. [Google Scholar] [CrossRef] [PubMed]
  93. Liu, J.; Chen, Y.; Han, T.; Cheng, M.; Zhang, W.; Long, J.; Fu, X. A biomimetic SiO2@chitosan composite as highly-efficient adsorbent for removing heavy metal ions in drinking water. Chemosphere 2019, 214, 738–742. [Google Scholar] [CrossRef] [PubMed]
  94. Razzaz, A.; Ghorban, S.; Hosayni, L.; Irani, M.; Aliabadi, M. Chitosan nanofibers functionalized by TiO2 nanoparticles for the removal of heavy metal ions. J. Taiwan Inst. Chem. Eng. 2016, 58, 333–343. [Google Scholar] [CrossRef]
  95. Aizat, M.A.; Aziz, F. 12—Chitosan Nanocomposite Application in Wastewater Treatments. In Nanotechnology in Water and Wastewater Treatment; Ahsan, A., Ismail, A.F., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 243–265. [Google Scholar] [CrossRef]
  96. Olivera, S.; Muralidhara, H.B.; Venkatesh, K.; Guna, V.K.; Gopalakrishna, K.; Kumar K., Y. Potential applications of cellulose and chitosan nanoparticles/composites in wastewater treatment: A review. Carbohydr. Polym. 2016, 153, 600–618. [Google Scholar] [CrossRef]
  97. Al-Shahrani, H.; Alakhras, F.; Al-Abbad, E.; Al-Mazaideh, G.; Hosseini-Bandegharaei, A.; Ouerfelli, N. Sorption of cobalt (II) ions from aqueous solutions using chemically modified chitosan. Glob. Nest J. 2018, 20, 620–627. [Google Scholar] [CrossRef]
  98. Sobahi, T.R.A.; Abdelaal, M.Y.; Makki, M.S.I. Chemical modification of Chitosan for metal ion removal. Arab. J. Chem. 2014, 7, 741–746. [Google Scholar] [CrossRef]
  99. Vakili, M.; Deng, S.; Cagnetta, G.; Wang, W.; Meng, P.; Liu, D.; Yu, G. Regeneration of chitosan-based adsorbents used in heavy metal adsorption: A review. Sep. Purif. Technol. 2019, 224, 373–387. [Google Scholar] [CrossRef]
  100. Sheth, Y.; Dharaskar, S.; Khalid, M.; Sonawane, S. An environment friendly approach for heavy metal removal from industrial wastewater using chitosan based biosorbent: A review. Sustain. Energy Technol. Assess. 2021, 43, 100951. [Google Scholar] [CrossRef]
  101. Alsaiari, N.S.; Alzahrani, F.M.; Katubi, K.M.; Amari, A.; Rebah, F.B.; Tahoon, M.A. Polyethylenimine-Modified Magnetic Chitosan for the Uptake of Arsenic from Water. Appl. Sci. 2021, 11, 5630. [Google Scholar] [CrossRef]
  102. Şenol, Z.M.; Şimşek, S. Insights into Effective Adsorption of Lead ions from Aqueous Solutions by Using Chitosan-Bentonite Composite Beads. J. Polym. Environ. 2022, 30, 3677–3687. [Google Scholar] [CrossRef]
  103. Verma, M.; Ahmad, W.; Park, J.H.; Kumar, V.; Vlaskin, M.S.; Vaya, D.; Kim, H. One-step functionalization of chitosan using EDTA: Kinetics and isotherms modeling for multiple heavy metals adsorption and their mechanism. J. Water Process Eng. 2022, 49, 102989. [Google Scholar] [CrossRef]
  104. Algethami, J.S.; Alqadami, A.A.; Melhi, S.; Alhamami, M.A.M.; Fallatah, A.M.; Rizk, M.A. Sulfhydryl Functionalized Magnetic Chitosan as an Efficient Adsorbent for High-Performance Removal of Cd(II) from Water: Adsorption Isotherms, Kinetic, and Reusability Studies. Adsorpt. Sci. Technol. 2022, 2022, 2248249. [Google Scholar] [CrossRef]
  105. Sutirman, Z.A.; Rahim, E.A.; Sanagi, M.M.; Abd Karim, K.J.; Wan Ibrahim, W.A. New efficient chitosan derivative for Cu(II) ions removal: Characterization and adsorption performance. Int. J. Biol. Macromol. 2020, 153, 513–522. [Google Scholar] [CrossRef]
  106. Shahraki, S.; Delarami, H.S.; Khosravi, F.; Nejat, R. Improving the adsorption potential of chitosan for heavy metal ions using aromatic ring-rich derivatives. J. Colloid Interface Sci. 2020, 576, 79–89. [Google Scholar] [CrossRef] [PubMed]
  107. Dong, L.; Shan, C.; Liu, Y.; Sun, H.; Yao, B.; Gong, G.; Jin, X.; Wang, S. Characterization and Mechanistic Study of Heavy Metal Adsorption by Facile Synthesized Magnetic Xanthate-Modified Chitosan/Polyacrylic Acid Hydrogels. Int. J. Environ. Res. Public. Health 2022, 19, 11123. [Google Scholar] [CrossRef]
  108. Jiang, C.; Wang, X.; Wang, G.; Hao, C.; Li, X.; Li, T. Adsorption performance of a polysaccharide composite hydrogel based on crosslinked glucan/chitosan for heavy metal ions. Compos. Part B Eng. 2019, 169, 45–54. [Google Scholar] [CrossRef]
  109. Lin, Z.; Yang, Y.; Liang, Z.; Zeng, L.; Zhang, A. Preparation of Chitosan/Calcium Alginate/Bentonite Composite Hydrogel and Its Heavy Metal Ions Adsorption Properties. Polymers 2021, 13, 1891. [Google Scholar] [CrossRef]
  110. Ablouh, E.; Hanani, Z.; Eladlani, N.; Rhazi, M.; Taourirte, M. Chitosan microspheres/sodium alginate hybrid beads: An efficient green adsorbent for heavy metals removal from aqueous solutions. Sustain. Environ. Res. 2019, 29, 5. [Google Scholar] [CrossRef]
  111. Abu El-Soad, A.M.; Lazzara, G.; Abd El-Magied, M.O.; Cavallaro, G.; Al-Otaibi, J.S.; Sayyed, M.I.; Kovaleva, E.G. Chitosan Functionalized with Carboxyl Groups as a Recyclable Biomaterial for the Adsorption of Cu (II) and Zn (II) Ions in Aqueous Media. Int. J. Mol. Sci. 2022, 23, 2396. [Google Scholar] [CrossRef]
  112. El Shahawy, A.; Mubarak, M.F.; El Shafie, M.; Abdulla, H.M. Adsorption of Mn(ii) ions from wastewater using an AgNPs/GO/chitosan nanocomposite material. RSC Adv. 2022, 12, 29385–29398. [Google Scholar] [CrossRef] [PubMed]
  113. Wang, B.; Bai, Z.; Jiang, H.; Prinsen, P.; Luque, R.; Zhao, S.; Xuan, J. Selective heavy metal removal and water purification by microfluidically-generated chitosan microspheres: Characteristics, modeling and application. J. Hazard. Mater. 2019, 364, 192–205. [Google Scholar] [CrossRef] [PubMed]
  114. Hu, S.-Z.; Huang, T.; Zhang, N.; Lei, Y.-Z.; Wang, Y. Chitosan-assisted MOFs dispersion via covalent bonding interaction toward highly efficient removal of heavy metal ions from wastewater. Carbohydr. Polym. 2022, 277, 118809. [Google Scholar] [CrossRef] [PubMed]
  115. Pavithra, S.; Thandapani, G.; Sugashini, S.; Sudha, P.N.; Alkhamis, H.H.; Alrefaei, A.F.; Almutairi, M.H. Batch adsorption studies on surface tailored chitosan/orange peel hydrogel composite for the removal of Cr(VI) and Cu(II) ions from synthetic wastewater. Chemosphere 2021, 271, 129415. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structure of (a) chitin and (b) CS (c) α- and β-chitin.
Figure 1. Chemical structure of (a) chitin and (b) CS (c) α- and β-chitin.
Polymers 15 01453 g001
Figure 2. A simplified flow diagram of chitin and CS preparation [47,48].
Figure 2. A simplified flow diagram of chitin and CS preparation [47,48].
Polymers 15 01453 g002
Figure 3. Neutralization reaction of chitin and CS.
Figure 3. Neutralization reaction of chitin and CS.
Polymers 15 01453 g003
Figure 4. Schematic representation of modification of CS and CS-based materials preparation for wastewater treatment.
Figure 4. Schematic representation of modification of CS and CS-based materials preparation for wastewater treatment.
Polymers 15 01453 g004
Figure 5. Chitosan is employed for removing heavy metals from wastewater.
Figure 5. Chitosan is employed for removing heavy metals from wastewater.
Polymers 15 01453 g005
Figure 6. Schematic representation of monolayer and multilayer adsorption of heavy metals. (a) monolayer adsorption, (b) multilayer adsorption.
Figure 6. Schematic representation of monolayer and multilayer adsorption of heavy metals. (a) monolayer adsorption, (b) multilayer adsorption.
Polymers 15 01453 g006
Table 1. Proximate composition of different chitin sources [45].
Table 1. Proximate composition of different chitin sources [45].
SourcesProtein (%)Ash (%)Chitin (%)Moisture (%)Lipid (%)
Shrimp shells32.7732.4636.4345.65-
Shrimp shells
(P. longirostris)
29.2325.0626.983.2515.48
Shrimp shells
(Penaeus durarum)
34.0242.2623.72--
Insect cuticles
(Cicada sloughs)
39.811.736.68.72.7
Crabs’ shells16.6866.5816.73--
Mussel shells9.9923.2523.25--
Squid gladius
(L. vulgaris)
36.522.5731.2-0..32
Adapted from [45].
Table 2. Deacetylation conditions in CS production.
Table 2. Deacetylation conditions in CS production.
SourceDemineralization Deproteinization Decolorization DeacetylationDegree of Acetylation (DA) or Degree of Deacetylation (DD) (%)Reference
Shrimp shells2.5 M HCl (1:10 solid: solvent ratio w/v) for 4 min under MW at 650 W power NaOH (20%) under MW irradiation at 500 W for 8 min.-30% NaOH for 12 min at 500 W23.4% DA[46]
Shrimp shells HCl (7%) at ambient temperature for 24 h.NaOH (10%) at ambient temperature for 24 h.Ethanol for 6 hNaOH (50% w/v) at a boiling temperature in the N2 atmosphere, Repeated twice78% DD[55]
Shrimp shells1.5% HCl (1:30 w/v) for 20 h at room temperatureNaOH (5%) at 90 °C for 24 h (solvent: shell ratio 12:1, v/w).Acetone (99.5%) at room temperature for 24 h.50% NaOH (15%, w/v) at 60 °C for 8 h.-[56]
Giant freshwater prawn carapace 1 M HCl (1:10 solid: liquid ratio) at 60 °C, 250 rpm for 2 h.1 M NaOH (1:10 solid: liquid ratio) at 100 °C, 250 rpm for 2 h.95% ethanol (1:5 mass: volume ratio) for 30 min at ambient temperature60% NaOH (1:10 solid: liquid ratio), at 120 °C, 250 rpm for 2 h.85.2%[57]
Prawn shells1 M HCl (1:16 solid: liquor ratio) at 100 °C for 4 h.1 M NaOH (1:16 solid: liquid ratio) at 100 °C for 4 h. 50% NaOH (1:30 solid: liquid ratio), in presence of ethanol, at 80 °C for 4 h.89% DD[58]
Crab shell2.5% (w/v) HCl at 1:20 (w/v, shell: solution), 20 °C for 6 h. 2% KOH at 1:20 (w/v, shell: solution), 90 °C for 2 hAcetone for 10 min40% NaOH at 1:15 (w/v, chitin: solution) at 105 °C for 2 h.53.42% DD[59]
Table 3. Frequently used solvents for chitin and CS.
Table 3. Frequently used solvents for chitin and CS.
Chitin/CSSolventsReferences
ChitinDimethylformamide + lithium chloride, Diethylformamide + lithium chloride, Hexafluoroisopropanol, Hexafluoroacetone + sequihydrate, 1,2-Chloroethanol + sulphuric acid, high concentrated organic acids (HCl, H2SO4, H3PO4)[65,67,68]
CSAqueous citric acids, acetic acid, lactic acid, formic acid, glutamic acid, HCl acid
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gamage, A.; Jayasinghe, N.; Thiviya, P.; Wasana, M.L.D.; Merah, O.; Madhujith, T.; Koduru, J.R. Recent Application Prospects of Chitosan Based Composites for the Metal Contaminated Wastewater Treatment. Polymers 2023, 15, 1453. https://doi.org/10.3390/polym15061453

AMA Style

Gamage A, Jayasinghe N, Thiviya P, Wasana MLD, Merah O, Madhujith T, Koduru JR. Recent Application Prospects of Chitosan Based Composites for the Metal Contaminated Wastewater Treatment. Polymers. 2023; 15(6):1453. https://doi.org/10.3390/polym15061453

Chicago/Turabian Style

Gamage, Ashoka, Nepali Jayasinghe, Punniamoorthy Thiviya, M. L. Dilini Wasana, Othmane Merah, Terrence Madhujith, and Janardhan Reddy Koduru. 2023. "Recent Application Prospects of Chitosan Based Composites for the Metal Contaminated Wastewater Treatment" Polymers 15, no. 6: 1453. https://doi.org/10.3390/polym15061453

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