Next Article in Journal
Fatigue Behavior of Rotary Friction Welding of Acrylonitrile Butadiene Styrene and Polycarbonate Dissimilar Materials
Previous Article in Journal
Cryogenic Mechanical Properties and Stability of Polymer Films for Liquid Oxygen Hoses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Next-Generation Water Treatment: Exploring the Potential of Biopolymer-Based Nanocomposites in Adsorption and Membrane Filtration

Department of Housing Environmental Design, Research Institute of Human Ecology, College of Human Ecology, Jeonbuk National University, Jeonju 54896, Republic of Korea
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(16), 3421; https://doi.org/10.3390/polym15163421
Submission received: 26 June 2023 / Revised: 3 August 2023 / Accepted: 14 August 2023 / Published: 16 August 2023
(This article belongs to the Section Polymer Composites and Nanocomposites)

Abstract

:
This review article focuses on the potential of biopolymer-based nanocomposites incorporating nanoparticles, graphene oxide (GO), carbon nanotubes (CNTs), and nanoclays in adsorption and membrane filtration processes for water treatment. The aim is to explore the effectiveness of these innovative materials in addressing water scarcity and contamination issues. The review highlights the exceptional adsorption capacities and improved membrane performance offered by chitosan, GO, and CNTs, which make them effective in removing heavy metals, organic pollutants, and emerging contaminants from water. It also emphasizes the high surface area and ion exchange capacity of nanoclays, enabling the removal of heavy metals, organic contaminants, and dyes. Integrating magnetic (Fe2O4) adsorbents and membrane filtration technologies is highlighted to enhance adsorption and separation efficiency. The limitations and challenges associated are also discussed. The review concludes by emphasizing the importance of collaboration with industry stakeholders in advancing biopolymer-based nanocomposites for sustainable and comprehensive water treatment solutions.

1. Introduction

Water treatment is a critical global challenge as the demand for clean and safe water rises [1]. With increasing pollution, contamination, and scarcity of freshwater resources, innovative materials and technologies are crucial to addressing these pressing issues [2]. Industrial effluents contain a wide range of pollutants that adversely affect human health when they enter the environment and come into contact with humans through various pathways (in detail in Table 1). Traditionally, water treatment has relied on several methods to ensure the purification of water for various applications, such as coagulation/flocculation [3,4], sedimentation [5,6], filtration [7], disinfection [8], adsorption [9], membrane filtration [10,11]. Although these traditional water treatment methods have specific advantages, they also have certain disadvantages. Coagulation and flocculation [12] effectively reduce turbidity and remove suspended particles, but they may produce a significant amount of toxic sludge, which requires proper disposal. Sedimentation can reduce water treatment costs and the need for additional chemicals but cannot effectively settle smaller particles or remove microbes [13]. Filtration methods, such as sand filtration, may not completely eliminate pathogens and require frequent cleaning and maintenance of filters [14]. Additionally, chlorination is widely used to kill germs, but it is a potent biocide and unsafe [15]. Advanced oxidation techniques offer fast reaction rates and reduced toxicity of organic pollutants, but they are expensive and require careful dose calculations due to their complex chemistry [16]. Ion exchange is an eco-friendly method of neutralizing acidic or basic water, but it faces fouling, bacterial and chlorine contamination, and inadequate regeneration [16]. Anaerobic treatment offers a low-energy and environment-friendly approach but cannot remove inorganic components and generates volatile acids [17]. Microbial treatment is another environment-friendly method used to remove pollutants from water, but it has a short lifespan, is costly, and can result in fouling and low production rates [18].
In contrast, adsorption and membrane filtration stand out for their effectiveness and versatility [19,20]. Adsorption offers high removal efficiency [21], selectivity [22,23], and the potential for regeneration, while membrane filtration provides a reliable method for removing suspended solids, pathogens, and dissolved substances [24]. Therefore, adsorption and membrane filtration technologies have played a crucial role in developing modern water treatment processes, ensuring the delivery of clean and safe drinking water to communities worldwide. Polymers play a crucial role in the enhancing performance and reducing the treatment costs of water treatment methods, particularly in adsorption and membrane filtration processes. Synthetic polymers have traditionally dominated these technologies due to their effectiveness. However, concerns about the biodegradability of synthetic polymers have prompted interest in bio-polymer-based adsorbents and membranes. Bio-polymer-based adsorbents or membranes offer an attractive alternative due to their environmentally friendly nature [25]. Moreover, researchers have recently turned their attention to biopolymeric nanocomposite materials, which are promising for various water treatment applications [26,27]. These materials combine the advantages of biopolymers derived from renewable resources with the unique properties of nanofillers, resulting in enhanced performance and efficiency [28]. Biopolymers, such as chitosan, alginate, cellulose, and starch, offer several advantages in water treatment due to their biocompatibility, biodegradability, and versatility [29,30]. They exhibit excellent adsorption capacities [31], flocculation abilities [32], and antimicrobial properties [33], making them suitable for various water treatment processes [34]. However, incorporating nanofillers into biopolymers has opened up new opportunities for improving their performance and expanding their applications [35]. Nanocomposite materials, composed of biopolymers and nanofillers, possess enhanced mechanical strength, increased surface area, improved stability, and tailored functionalities [36,37]. Nanofillers, such as nanoparticles (e.g., silver, titanium dioxide, zinc oxide, copper oxide, iron oxide, etc.), nanoclays, graphene oxide, and carbon nanotubes, contribute unique properties such as high surface area, catalytic activity, photocatalytic properties, and antimicrobial effects [38,39,40]. These synergistic combinations enable biopolymeric nanocomposites to tackle specific water treatment challenges [41], such as removing heavy metals [42], organic pollutants [43], and microorganisms [44]. The synthesis and characterization of biopolymeric nanocomposites have witnessed significant advancements in recent years [45]. Various methods, including solution casting, in situ polymerization, and electrospinning, have been employed to fabricate nanocomposites with controlled structures and properties [46]. Characterization techniques, such as X-ray diffraction (XRD), scanning electron microscopy (SEM), and thermogravimetric analysis (TGA), provide valuable insights into their morphology, crystallinity, thermal stability, and mechanical properties [47,48].
This review article offers a comprehensive exploration of the emerging field of biopolymeric nanocomposite materials for water treatment applications. It focuses on the synergistic combination of biopolymers and nanofillers, providing insights into their fabrication, characterization, and specific applications in adsorption and membrane filtration. The article also addresses challenges, future perspectives, and the potential to revolutionize water treatment technologies.
Table 1. Details about pollutants, source of pollutants, and effects on human health.
Table 1. Details about pollutants, source of pollutants, and effects on human health.
Sr. No.Type PollutantsChemical Compositions/
Formula
SourcesEffectsRef.
1.Kaolin clayAl2Si2O5 (OH)4Industrial effluents from
  • Porcelain
  • Rubber
  • Paper
  • Paint
  • Many other products
Water with kaolin clay causes
  • Lung cancer
  • Pneumoconiosis
  • Mesothelioma
  • Aquatic-life problems
[49,50]
2.Iron ore slimesFe2O3, Fe3O4, 2Fe2O. 3H2O, FeCO3 and FeS2Washing liquid from
  • Sponge iron industries
  • Steel industries
Water with iron slimes
  • Creates a brown color
  • Damages the reproductive system of aquatic life
  • Degrades the supercoiled DNA
[51,52,53]
3.CoalC, H, O, N, S
  • Coal mining
  • Coal plants
  • Thermal power plants
  • Fly ash
  • Coal bed methane
  • Natural resources
Coal water creates
  • Cancer
  • Lung diseases
  • Heart damage
  • Birth defects
[54,55,56]
4.Silicon dioxideSiO2Wastage from industries of
  • Electrical items
  • Solar cell production
Silica in drinking water causes:
  • Alzheimer’s disease
  • Cognitive impairment
  • Dementia
[57,58,59,60]
5.CopperCuWastewater from industries of
  • Electrical conductors
  • Heat conductors
  • Building materials
  • Jewelry
  • Steel
  • Alloy metals
Copper intoxication
  • Inhibits the growth of phytoplankton
  • Produces hydroxyl radicals
  • Effects on proteins, lipids, central nervous system, and DNA
  • Cognitive impairment
[61,62,63,64,65]
6.NickelNiEffluents from
  • Stainless steel
  • Nickel-processing industries
  • Amalgam-processing industries
Nickel leads to effects such as
  • Disordered lung function
  • Non-remitting bronchitis
  • Neurotoxic
  • Genotoxic
  • Regenerative poisonous
  • Nephrotoxic
  • Embryo toxic
  • Teratogenic
[66,67]
7.ZincZnIndustrial effluents from
  • Brass
  • Corrosion-resistant zinc plating of iron
  • Amalgam steel
  • Zinc processing units
High concertation of zinc causes
  • Sickness
  • Vomiting
  • Epigastric irritation
  • Laziness
  • Toxicity to plants and invertebrates
[68,69]
8.LeadPbWastewater from
  • Lead–acid batteries
  • Atomic power plants
  • Electrodes making
Lead causes mostly
  • Anemia
  • Hypertension
  • Kidney and bone damage
  • Disruption of the nervous system
  • Sperm damage
[70,71]
9.MercuryHgWastage from
  • Coal mining
  • Industrial uses
  • Fossil fuel combustion
  • Mining
Mercury has dangerous effects on
  • The central nervous system
  • Cardiovascular system
  • Kidney and bone
  • Skin
[72]
10.ChromiumCrEffluents from
  • Chromite mineral
  • Stainless steel industries
  • Chemical industries
  • Leather industries
Cr (VI) is dangerous and creates
  • Skin irritation
  • Stomatitis
  • Pneumonitis
  • Inhibits the growth of oceanic life
[73,74]
11.ArsenicAsIndustrial wastewater from
  • Wood preservatives
  • Pesticides
  • Wood processes
  • Agrochemicals
  • Natural resources
Long ingestion of arsenic (III) (>10 ppb) can lead to
  • Arsenicosis
  • Hemolysis
  • Cancer
  • Neurological disorders and painful patches on the hand/feet
[75,76]
12.Malachite green (MG)C23H25Cl2Industrial effluents from
  • Textile dye
  • Wools
  • Plastic materials
  • Calfskin packaging
  • Leather and rubber
  • Paint
High concentration of MG
  • Cuts off sunlight from water
  • Affects the photosynthesis of sea-going plants
  • Disorders the DNA sequence
  • Causes hepatitis
  • Vomiting
[77]
13.Congo red (CR)C32H22N6Na2O6S2Wastewater from
  • Color manufacturer
  • Printing and dyeing machines
  • Paper and plastic productions
High concentration of CR creates
  • Presenile dementia
  • Degenerative disorder diseases
  • Mutagenesis
[78,79]
14.Methylene blue (MB)C16H18ClN3SEffluents from
  • Textile manufacturers
  • Leather and paper manufacturing
  • Color manufacturing industries
MB creates mostly
  • Headache
  • Hypertension
  • Nausea
  • Forgetfulness
  • Dyspepsia
  • Red blood cell breakdown
  • Allergic reactions
  • Central nervous system toxicity
[80,81]
15.Reactive Black 5 (RB5)C26H21N5Na4O19S6Wastewater from
  • Textiles
  • Dye houses
  • Paper printing
Presence of RB5 in water bodies causes
  • Aesthetic problems
  • Obstruction of sun light penetration
  • Oxygen transfer into water
  • Effects on aquatic life
[82,83]
16.Reactive Blue 4 (RB4)C23H14Cl2N6O8S2Effluent from
  • Textile units
  • Plastic manufacturing
  • Printing houses
Presence of RB4 in water bodies causes
  • Pollution in water bodies
  • Effects on aquatic life
  • Mutagenesis
[84,85]
17.Reactive Blue 29 (RB29)C29H15Cl2N5Na2O9S2Discharged wastewater from
  • Textile industries
  • Plastic industries
RB29 affects
  • Esthetic merit
  • Light penetration
  • Carcinogenesis
[86]
18.Reactive Yellow 145 (RY145)C28H20ClN9Na4O16S5Wastewater from
  • Paper printing
  • Textile industries
  • Plastic industries
RY145 toxic towards
  • Humans
  • Aquatic life
[87,88]
19.Pharmaceuticals
n:
(Ibuprofen:
Metformin:
Simvastatin:
Omeprazole:
Fluoxetine:
Ciprofloxacin:
Amlodipine:
Sertraline:
Atorvastatin:)

C8H9NO2
C13H18O2
C4H11N5
C25H38O5
C17H19N3O3S
C17H18F3NO
C17H18FN3O3
C20H25ClN2O5
C17H17Cl2N
C33H35FN2O5
  • Human excretion
  • Improper disposal
  • Pharmaceutical industries
  • Antibiotic resistance
  • Endocrine disruption
  • Allergic reactions
  • Drinking water contamination
[89,90]
20.Personal Care (Products:
Parabens:
Triclosan:
Benzophenone-3:
Phthalates:
Octinoxate:
Formaldehyde:
Sodium Lauryl Sulfate:)


C9H10O3
C12H7Cl3O2
C14H12O3
C8H4O4
C18H26O3
CH2O
C12H25NaO4S
  • Direct disposal into sinks, showers, and toilets
  • Runoff from treated areas (e.g., sunscreen washed off during swimming)
  • Improper disposal and inadequate wastewater treatment
  • Toxic to aquatic life
  • Endocrine disruption
  • Interfere with hormone function
  • Respiratory irritation
[91]
21.Microplastics
(Polyethylene:
Polypropylene:
Polystyrene:)

(C2H4)n
(C3H6)n
(C8H8)n
  • Wastewater discharge from industries and households.
  • Urban runoff carrying plastic litter.
  • Degradation of larger plastic items in the environment.
  • Release of microfibers from washing synthetic clothing.
  • Ingestion
  • Neurotoxicity
  • Damage to the digestive tracts of marine animals
[92,93]

2. Effects of Pollutants on Human Health

The effects of pollutants on human health are of significant concern in environmental health. Table 1 provides an overview of the different sources of pollutants found in industrial effluents and highlights their specific effects on human health. These pollutants include heavy metals, organic compounds, toxins, and other harmful substances. Each pollutant reveals unique health effects depending on its chemical properties and the extent and duration of exposure. The health effects of pollutants can vary extensively, ranging from acute to chronic and long-term health implications. Acute effects may include immediate symptoms such as nausea, vomiting, respiratory distress, or skin irritation. Chronic exposure to pollutants over a prolonged period can lead to various health problems, including respiratory disorders, neurological disorders, cardiovascular diseases, reproductive disorders, and certain cancers. Understanding and monitoring the sources and levels of pollutants in industrial effluents and using improved technologies to minimize human exposure and mitigate potential health risks is crucial.

3. Biopolymers

Biopolymers have expanded significant attention in water treatment due to their unique properties and multipurpose applications [94]. They are derived from renewable sources, making them environmentally friendly alternatives to synthetic polymers. Biopolymers are generally non-toxic, biocompatible, and biodegradable, minimizing the potential for secondary pollution. Moreover, their flexibility allows for modifications and functionalization to improve performance and target specific contaminants [95]. This section will explore the different types of biopolymers commonly used in water treatment and their effectiveness in various processes.
Chitosan: Chitosan [96], derived from chitin (as shown in Figure 1a) [97], is a broadly studied biopolymer with excellent adsorption capabilities. Its high surface area and positive charge effectively remove heavy metals, dyes, and organic pollutants from water through electrostatic attraction and coagulation. Chitosan-based materials can be used as beads, membranes, and hydrogels for applications like adsorption, filtration, and wastewater treatment.
Alginate: Alginate [98], extracted from seaweed (Figure 1b) [99], is a biopolymer known for its gel-forming properties and biocompatibility. It is commonly used in water treatment for applications such as encapsulation, immobilization of enzymes or microorganisms, and removal of metal ions [100,101,102]. Alginate beads and membranes have shown promise in removing metals and dyes through ion exchange and adsorption mechanisms.
Cellulose: Cellulose, the most plentiful biopolymer on Earth (Figure 1c), has garnered attention for its exceptional mechanical strength, hydrophilicity, and biodegradability. Cellulosic materials have been used for water treatment processes such as adsorption, membrane filtration, and separation techniques. Modified cellulose-based materials, including cellulose nanocrystals and cellulose derivatives, offer higher surface area and adsorption properties, effectively removing contaminants [103,104].
Figure 1. Schematic representation of the source of (a) chitosan [97]. Copyright 2022; reproduced with permission from Springer Nature, (b) alginate [99]. Copyright 2020; reproduced with permission from Intechopen Limited, (c) cellulose [103]. Copyright 2021; reproduced with permission from Elsevier B.V. All rights reserved, and (d) starch [105]. Copyright 2022; reproduced with permission from the authors, MDPI, Basel.
Figure 1. Schematic representation of the source of (a) chitosan [97]. Copyright 2022; reproduced with permission from Springer Nature, (b) alginate [99]. Copyright 2020; reproduced with permission from Intechopen Limited, (c) cellulose [103]. Copyright 2021; reproduced with permission from Elsevier B.V. All rights reserved, and (d) starch [105]. Copyright 2022; reproduced with permission from the authors, MDPI, Basel.
Polymers 15 03421 g001
Starch: Starch, derived from crops (Figure 1d) [105], possesses unique properties such as biodegradability, abundance, and low cost. Starch-based materials have been explored for applications such as adsorption, flocculation, and membrane filtration. Modified starches and starch-based hydrogels can potentially remove organic pollutants, dyes, and microorganisms from water [106,107,108].
Poly(hydroxyalkanoates) (PHA): Poly(hydroxyalkanoates) are biodegradable and biocompatible polyesters synthesized by various microorganisms (Figure 2a) [109,110]. PHA-based materials have shown promise in water treatment applications, particularly in removing organic pollutants and heavy metals. PHA films, beads, and membranes have demonstrated efficient adsorption capacities and can be used in adsorption, filtration, and wastewater treatment [111].
Xanthan Gum: Xanthan gum is a polysaccharide produced by the bacterium Xanthomonas campestris [112], as shown in Figure 2d [113]. Its high viscosity and stabilizing properties make it suitable for various water treatment applications [114]. Xanthan gum has been utilized for the flocculation, sedimentation, and stabilization of particles in water and wastewater treatment processes [115].
Figure 2. Schematically presentation of source and formula structures of (a) PHA [109]. Copyright 2022; reproduced with permission from Elsevier B.V. All rights reserved, (b) Xanthan gum [113]. Copyright 2017; reproduced with permission from Wiley-vch Verlag GmbH & Co. KGaA, Weinheim, and (c) lignin [116]. Copyright 2022; reproduced with permission from the Author(s).
Figure 2. Schematically presentation of source and formula structures of (a) PHA [109]. Copyright 2022; reproduced with permission from Elsevier B.V. All rights reserved, (b) Xanthan gum [113]. Copyright 2017; reproduced with permission from Wiley-vch Verlag GmbH & Co. KGaA, Weinheim, and (c) lignin [116]. Copyright 2022; reproduced with permission from the Author(s).
Polymers 15 03421 g002
Lignin: Lignin is a complex and abundant biopolymer found in plant cell walls, as shown in Figure 2e [116]. It is an attractive material for water treatment due to its aromatic structure and adsorption properties [117]. Lignin-based materials, such as lignin derivatives, nanoparticles, and membranes, can potentially remove organic pollutants, dyes, and heavy metals from water [118].
However, it is crucial to note that the effectiveness of biopolymers in water treatment can be influenced by factors such as pH, temperature, concentration, and the nature of contaminants. Optimization of operating conditions and the selection of suitable biopolymers are crucial for achieving desired treatment outcomes.

4. Biopolymer-Based Nanocomposite

Nanocomposite materials, which combine biopolymers with nanofillers, have emerged as promising candidates for water treatment applications [119]. Parallel to their components, these materials offer enhanced performance, enhanced properties, and increased functionality [120]. In this section, we will explore the concept of nanocomposites and the various types of nanofillers combined with biopolymers for water treatment. Nanofillers can be nanoparticles, nanoclays, nanotubes, or other nanoscale materials, and they contribute specific functionalities such as increased surface area, enhanced mechanical strength, catalytic activity, or antimicrobial properties.

4.1. Nano-Fillers for Water Treatment

A wide range of nanofillers can be incorporated into biopolymer matrices for water treatment applications [121]. Here are some commonly used nanofillers:
Nanoparticles: Nanoparticles play a crucial role in water treatment applications, offering various properties and functionalities. Silver nanoparticles (AgNPs) and titanium dioxide nanoparticles (TiO2NPs) are well known for their antimicrobial and photocatalytic properties. AgNPs exhibit strong antibacterial and antiviral effects, while TiO2NPs can induce photocatalytic reactions to degrade organic pollutants in water [122,123,124]. In addition to these, other nanoparticles—such as zinc oxide nanoparticles (ZnONPs) with their photocatalytic and antimicrobial properties [125]; iron oxide nanoparticles (FeOxNPs) for adsorption, catalytic degradation of dyes, and magnetic separation [126,127]; copper nanoparticles (CuNPs) for antimicrobial applications [128]; carbon-based nanoparticles like carbon nanotubes (CNTs) and graphene oxide (GO) for adsorption of contaminants [129]; cerium oxide nanoparticles (CeO2NPs) for catalytic degradation [130]; and hybrid nanoparticles with tailored functionalities—are also being extensively studied [131]. Each type of nanoparticle brings unique advantages to water treatment processes, and ongoing research aims to harness their potential for efficient and sustainable water purification.
Nanoclays: Nanoclays, such as montmorillonite and halloysite nanotubes, have great potential as biopolymer composites for water purification applications [132,133]. These nanoparticles have distinct features that improve the composites’ ability to remove pollutants from water. Incorporating montmorillonite nanoclay into biopolymer matrices such as chitosan or alginate, for example, has been shown to increase the adsorption capacity and selectivity of the composites towards heavy metals, organic contaminants, and colors [134]. Nanoclays’ high surface area and ion exchange capacities allow for effective pollutant adsorption and removal via surface complexation, ion exchange, and intercalation. Furthermore, the layered structure of nanoclays creates a difficult path for pollutants to diffuse, increasing purification efficiency.
On the other hand, halloysite nanotubes have a hollow tubular shape that can be used for the controlled release of antimicrobial agents or active species, providing additional capabilities for water treatment. Combining nanoclays with biopolymers improves adsorption and removal capacities and increases the composite materials’ mechanical strength and stability, making them appropriate for filtering membranes and adsorbent media [135]. Developing nanoclay-based biopolymer composites for water purification has the potential to alleviate water pollution issues while also encouraging sustainable and effective treatment methods. Figure 3a depicts a schematic of integrating nano-filler into the polymer matrix [136].
Carbon Nanotubes: Carbon nanotubes (CNTs) possess unique properties, including exceptional mechanical strength, high electrical conductivity, and excellent thermal stability, making them highly attractive for a variety of applications in water treatment. In the context of nanocomposites for water purification, CNTs have revealed promising results, particularly in membrane technology [137]. When incorporated into polymeric matrices, CNT-based nanocomposites have shown improved performance in water treatment compared to traditional nanomaterials and clay composites. For example, CNT-reinforced polymeric membranes exhibit better permeability, allowing for increased water flux rates while maintaining the practical separation of contaminants. This improved permeability is attributed to the unique structure of CNTs, which provide a network of interconnected nanochannels within the membrane, facilitating the flow of water molecules [138,139].
Moreover, CNTs can enhance the selectivity of membranes by acting as molecular sieves, effectively blocking the passage of specific contaminants based on size or charge. This selectivity is particularly advantageous in removing small organic molecules or heavy metal ions from water sources. Additionally, the high aspect ratio and surface area of CNTs contribute to their adsorption capacity, allowing for the effective removal of organic pollutants and dyes from water [140,141].
It is worth noting that while nanoclays offer their own set of advantages in water treatment, such as ion exchange capacity and high surface area, CNT-based nanocomposites provide distinct benefits. The exceptional mechanical and electrical properties of CNTs contribute to the overall performance and durability of the composite membranes. Moreover, the unique nanochannel structure and fouling resistance properties of CNTs set them apart as promising nanomaterials for membrane technology in water treatment.
Graphene and Graphene Oxide: Graphene and graphene oxide (GO) are highly promising carbon-based materials with excellent properties that make them attractive for various applications in water treatment [40]. These two-dimensional carbon structures possess a large surface area, high mechanical strength, and excellent adsorption capabilities contributing to their effectiveness in water purification processes [142]. In nanocomposites for water treatment, graphene- and GO-based materials have shown significant prospective applications, particularly in removing heavy metal ions, organic pollutants, and emerging contaminants from water sources [143]. Their large surface areas provide ample active sites for adsorption, allowing them to effectively capture and remove contaminants through physical interactions [143]. This adsorption capacity is especially beneficial for removing heavy metal ions, organic compounds, and micropollutants. Graphene-based nanocomposites have established superior adsorption performance compared to traditional nanomaterials, clay composites, and carbon nanotubes (CNTs) [144]. The unique structure of graphene and GO, consisting of a single layer of carbon atoms arranged in a hexagonal lattice, offering a highly accessible and reactive surface for interaction with contaminants. This results in enhanced adsorption efficiency and removal rates.
Moreover, graphene-based nanocomposites reveal exceptional mechanical strength, improving the membranes’ durability and lifespan. Furthermore, the chemical adaptability of graphene and GO allows for functionalization and modification, tailoring their properties for specific water treatment applications. Surface modifications can develop selectivity, improve stability, or introduce additional functionalities to the nanocomposites. This flexibility in functionalization expands the potential applications of graphene-based nanocomposites in water treatment. While nanoclays, CNTs, and other nanomaterials offer their unique advantages in water treatment, graphene- and GO-based nanocomposites stand out due to their exceptional properties, including large surface area, high mechanical strength, excellent adsorption capabilities, and electrical conductivity [144]. These properties collectively contribute to their superior performance in water purification processes. The structures of carbonaceous nanofillers are shown in Figure 3b [145].
Figure 3. Schematic representation of (a) incorporation of nanoclays into the polymer matrix [136]. Copyright 2022; reproduced with permission from the authors, and (b) structures of carbonaceous nanofillers [145]. Copyright 2021; reproduced with permission from MDPI, Basel.
Figure 3. Schematic representation of (a) incorporation of nanoclays into the polymer matrix [136]. Copyright 2022; reproduced with permission from the authors, and (b) structures of carbonaceous nanofillers [145]. Copyright 2021; reproduced with permission from MDPI, Basel.
Polymers 15 03421 g003

4.2. Synthesis and Characterization

The synthesis of biopolymeric nanocomposites embroils the incorporation of nanofillers, such as nanoparticles or nanofibers, into a biopolymer matrix. Various techniques can be employed to disperse the nanofillers within the biopolymer matrix, depending on the preferred structure and properties of the nanocomposite.
Solution casting: One commonly used method is solution casting, in which the biopolymer is dissolved in a suitable solvent and the nanofillers are dispersed in the polymer solution. The mixture is then cast into a mold or onto a substrate and allowed to dry, resulting in a solid nanocomposite material (Figure 4a) [146]. Solution casting allows for specific control over the nanofillers’ dispersion and the nanocomposite’s overall composition.
In situ polymerization: In situ polymerization is another approach to synthesizing biopolymeric nanocomposites. In this method, the biopolymer is synthesized or partially polymerized in the presence of the nanofillers (Figure 4b). The nanofillers act as nucleation sites for the polymerization process, forming a nanocomposite structure [147]. This method offers good interfacial bonding between the nanofillers and the polymer matrix, enhancing mechanical properties and stability.
Melt blending: Melt blending is a technique commonly employed for thermoplastic biopolymers. It involves melting the biopolymer and adding nanofillers, followed by mixing and solidification to obtain a homogeneous nanocomposite material (Figure 5a) [148]. Melt blending is relatively simple and scalable, making it suitable for the large-scale production of biopolymeric nanocomposites.
Electrospinning: Electrospinning is a particular technique used to fabricate nanofiber-based nanocomposites. In this method, a polymer solution containing the nanofillers is electrostatically spun into ultrafine fibers. The resulting nanofibers form a greatly porous structure with a higher surface area, making them suitable for applications such as filtration membranes. Electrospinning allows for precise control over the nanofiber diameter and distribution, resulting in tailored properties for water treatment applications. A schematic of the single electrospinning process and (b) the coaxial electrospinning process is shown in Figure 5b.
Characterization techniques play a crucial role in understanding polymer nanocomposites’ structure, morphology, thermal stability, and mechanical properties [150]. Among the commonly used techniques, X-ray diffraction (XRD) provides valuable information about polymer or polymer nanocomposite materials’ crystalline structure, phase composition, and degree of crystallinity. Furthermore, integrating spectroscopic techniques such as Fourier-transform infrared spectroscopy (FTIR) and Raman spectroscopy provides valuable chemical information about the polymers’ and polymer nanocomposites’ functional groups and molecular interactions. Generally, the Raman intensity varies directly with the size of the crystalline region. The Raman spectrum was also used to evaluate the change in crystal structures of polymers [94]. Examples of FTIR, Raman, and XRD spectra are shown in Figure 6.
Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) are powerful imaging techniques that allow for the visualization of the nanocomposite’s surface and internal structure at different length scales [129]. SEM provides high-resolution images of the material’s surface, enabling the examination of particle distribution, agglomeration, and interfacial interactions (as shown in Figure 7). Conversely, TEM provides an even higher resolution, allowing for a detailed examination of individual nanoparticles and their dispersion within the polymer matrix.
Thermal analysis techniques, such as thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC), provide insights into polymer nanocomposites’ thermal stability and behavior. TGA measures the weight loss of the sample as a function of temperature, providing information about its thermal degradation and decomposition. DSC measures the heat flow associated with phase transitions, melting points, and glass transition temperatures, giving valuable data on the thermal properties and crystallinity of the nanocomposite (as shown in Figure 8).
Advancements in characterization techniques have led to improved resolution, sensitivity, and quantitative analysis capabilities. For example, advanced microscopy techniques such as atomic force microscopy (AFM) and high-resolution TEM (HRTEM) have allowed for more precise characterization of nanoscale structures and interfaces. X-ray photoelectron spectroscopy (XPS), also known as electron spectroscopy for chemical analysis (ESCA), is another crucial characterization technique for analyzing polymer nanocomposites.
XPS analysis is particularly useful for studying the interaction between nanoparticles and the polymer matrix in nanocomposites. It can provide insights into the chemical bonding, charge transfer, and electronic structure at the nanoparticle–polymer interface. This information is crucial for understanding the mechanisms of reinforcement, adhesion, and dispersion of nanoparticles within the polymer matrix (Figure 9).
Combining XPS analysis with other characterization techniques, such as SEM or TEM, allows for a comprehensive understanding of the nanocomposite’s structure, morphology, and surface properties. The information obtained from XPS analysis helps researchers optimize the synthesis parameters, modify the surface chemistry, and tailor the properties of polymer nanocomposites for specific applications.
However, it is essential to exercise caution during polymer nanocomposites’ synthesis and sample preparation to ensure accurate characterization. Factors such as the choice of solvents, processing conditions, and sample handling can influence the dispersion and aggregation of nanoparticles, affecting the final properties of the nanocomposite. Careful control of these parameters is necessary to obtain reliable and reproducible results during characterization.

4.3. Applications in Water Treatment

4.3.1. Adsorption

The adsorption process is widely used in water treatment to remove contaminants from water [153]. It involves the attachment of pollutants to the surface of adsorbent material, leading to their separation from the water. The adsorbent material can be granules, pellets, or powder, providing a large surface area for effective adsorption. Several commercial adsorbents are commonly used in water treatment applications [154]. Activated carbon is one of the most widely used adsorbents due to its high adsorption capacity and versatility in removing many contaminants, including organic compounds, odors, and taste-causing substances [155]. Other commercial adsorbents include zeolites, silica gel, and clay minerals, which have specific adsorption properties for different pollutants [154]. Despite their effectiveness, commercial adsorbents have some drawbacks. One of the main limitations is their high cost, specifically in large-scale applications. The regeneration of adsorbents can also be challenging and costly in some cases.
Bionanocomposites as adsorbents have developed as a promising approach in recent years [119]. Bionanocomposites are hybrid materials that combine biopolymers with nanoparticles, offering enhanced adsorption properties [121]. These composites leverage the high surface area and functional properties of nanoparticles along with the biocompatibility and renewable nature of biopolymers. For example, chitosan–clay nanocomposites have garnered significant attention due to their abundant availability, ease of production, and effectiveness as adsorbents. This specific class of composites has demonstrated the ability to quantitatively remove 99% of dyes, metals, and hazardous negatively charged ions from water [156]. In addition, graphene oxide–potato starch composites have been investigated for their ability to remove MB dye from industrial effluents using adsorption. The composite exhibited a significant adsorption potential of approximately 90% [157]. Incorporating GO nanosheets with polysaccharide long chains in the composite created a plywood-like structure with nanocages, enhancing the adsorption of organic dyes [158]. Two types of interactions contribute to the adsorption mechanism: (a) electrostatic interactions between GO and cationic dyes, and (b) π–π stacking interactions between the aromatic moiety of the dye and the delocalized π-electron system of GO.
Graphene oxide also interacts with heavy metal ions through surface complexation for metal ion adsorption. Graphene oxide–chitosan composites have shown excellent stability and mechanical qualities for wastewater adsorption treatment. Chitosan, a cationic biopolymer with combative amino (―NH2) and hydroxyl (―OH) groups, exhibits efficient coagulative abilities for extracting pollutants from aqueous solutions. However, challenges such as higher synthesis costs, pH maintenance, and inefficacy at low concentrations limit the practical use of these materials. Most current research focuses on the adsorption properties of single heavy metal ions, and limited research is available on treating mixtures of heavy metal ions in wastewater [159,160]. AP-g-3D GO composites have also demonstrated efficient adsorption of various organic and inorganic pollutants from water [161]. Chitosan–poly(vinyl alcohol)–graphene oxide cross-linked sponge showed high adsorption capacities for rhodamine B (RB) and Congo red (CR) dyes [162]. The 3D porous structures of the biopolymer sponge and composite sponge enhanced the π–π interaction with the aromatic rings of the dye molecules. The surface morphology of the biopolymer sponge and 3D composite sponge is shown in Figure 10a–d, and the results of dye removal capacity are shown in Figure 10e–h.
A recent study has revealed that combining a geopolymer with a biopolymer enhances the adsorption efficiency of Ni(II) and Co(II) ions in wastewater. The synergistic effect of lateritic geopolymer and Grewia biopolymer, derived from the Tiliaceae family, resulted in the removal of approximately 80% of Ni(II) and Co(II) ions from the wastewater samples. The laterite-clay-based geopolymer is rich in aluminum, iron, and silica oxides, while the Grewia biopolymer consists of simple carbohydrates such as glucose, mannose, arabinose, xylose, and glucuronic acid. The laterite-clay-based geopolymer exhibits a kaolinite platelet structure, whereas the Grewia biopolymer possesses irregularly shaped open pores (Figure 10i), likely contributing to its superior adsorption capabilities [163].
Graphene oxide/chitosan composites have been extensively studied for wastewater adsorption treatment (Table 2). Chitosan, a cationic biopolymer with hydrophobic and biodegradable properties, possesses efficient coagulative ability in extracting pollutants from aqueous solutions. However, the practical use of these materials is limited due to higher synthesis costs, pH maintenance requirements, and inefficiency at low concentrations. The reported research studies provide valuable insights into the adsorption capacities of different graphene oxide/chitosan composites for various pollutants.
For instance, the GO-cl-potato starch bio-composite exhibited a remarkable adsorption capacity of 500 mg/g (90%) for MB dye [157]. Another composite, St-g-poly(AM-co-GO)/hydroxyapatite, demonstrated 297 mg/g adsorption capacity for MG dye [164]. Furthermore, AP-g-3D GO composites showed adsorption capacities for various pollutants, such as tert-butyl hydroquinone, o-nitrophenol, p-aminophenol, and others [161]. The 3D GO-sodium alginate composite achieved an adsorption capacity of 833.3 mg/g for MB dye [165]. In contrast, GO-calcium/alginate composite exhibited adsorption capacities of 163.93 mg/g and 140.85 mg/g for MB dye at different temperatures [166]. Other composites, such as dextrin-g-poly(m-phenylenediamine)-GO, St-g-3DGO, alginate-chitosan hybrid adsorbent, PAA-chitosan and biochar composite, chitosan-GO-Hap composite, and chitosan–GO composites, demonstrating their adsorption capacities for various pollutants [167,168,169,170,171,172]. Additionally, the sodium alginate-CMC-GO-Gd3O3 composite showed adsorption capacities for Pb, Cr (III), and As (V) [173]. Notably, the chitosan-activated carbon composite achieved a high removal efficiency of 98.7% for Cr (VI) ions [174].
Table 2. Summary of bio-based composite polymers as adsorbents for wastewater treatment.
Table 2. Summary of bio-based composite polymers as adsorbents for wastewater treatment.
Sl. NoAdsorbent/FlocculentPollutantsResults (% or Qmax mg/g)Ref.
1.GO-cl-potato starch bio-compositeMB500 mg/g (90%)[157]
2.St-g-poly(AM-co-GO)/hydroxyapatite composite hydrogelMG297 mg/g[164]
3.AP-g-3D GO compositestert-butyl hydroquinone
o-nitrophenol
p-aminophenol
Hydroquinone
p-nitrophenol
Neutral red
Alizarin red S
Pb (II)
Mn (II)
Cr2O7−2
Cd (II)
Cu (II)
Nd (III)
La (III)
Y (III)
Yb (III)
Yr (III)
22.17 mg/g
36.96 mg/g
116.4 mg/g
16.10 mg/g
36.96 mg/g
44.78 mg/g
39.92 mg/g
84.76 mg/g
7.92 mg/g
13.6 mg/g
17.64 mg/g
30.56 mg/g
25.25 mg/g
12.48 mg/g
16.96 mg/g
23.32 mg/g
30.32 mg/g
[161]
4.3D GO-sodium alginateMB833.3 mg/g, at 303K[165]
5.GO-calcium/alginateMB163.93 mg/g, at 298 K
140.85 mg/g, at 328 K
[166]
6.Dextrin-g-poly(m-phenylenediamine)-GOPb (II)
MB
80 mg/g
76.33 mg/g
[167]
7.St-g-3DGOPb (II)
Cu (II)
Cd (II)
Yb (III)
Nd (III)
108.68 mg/g
32.12 mg/g
46.28 mg/g
41.76 mg/g
38.168 mg/g
[168]
8.Alginate–chitosan hybrid adsorbentPb (II)96.8%, pH 5.0[169]
9.PAA–chitosan and biochar-compositeCu (II), Zn (II), Ni (II), Pb (II), Cd (II), Mn (II), Co (II), and Cr (VI)80%, pH 2–7[170]
10.Chitosan–GO–Hap compositeCR, acid red 1, and reactive red 243.06, 41.32, and 40.03 mg/g, pH 2[171]
11.Chitosan–GO compositesReactive black 5 dye277 mg/g at 25 °C[172]
12.GO–Chitosan compositeCu (II), Pb (II), Cd (II)60.7, 48.7, 32.3 mg/g, pH 1[175]
13.Sodium alginate-CMC-GO-Gd3O3Pb (II), Cr (III) and As (V)29.16, 158.73, and 36.77 mg/g[173]
14.Chitosan-activated carbon compositeCr (VI)98.7%[174]
15.Chitosan–graphene-oxide-dip-coated electrospun nanofiber membraneMB
CR
201 mg/g
152 mg/g
[176]

4.3.2. Magnetic Adsorbents

Magnetic adsorbents are gaining popularity in water treatment due to their ability to facilitate adsorption, minimize adsorbent loss, and simplify separation using an external magnetic field. For instance, the magnetic nanocomposite/activated carbon preparation is illustrated in Figure 11a,b. Compared to activated carbon alone, the resulting composite materials exhibit improved adsorption capacity for dyes such as rhodamine B and methyl orange. The maximum dye adsorption capacities were 182.48 mg/g for rhodamine B and 150.35 mg/g for methyl orange, as shown in Figure 11c–f. The adsorption mechanisms are depicted in Figure 11g–h [177]. The synergistic interactions between the composite and Fe3O4 nanoparticles play a vital role in the dye adsorption strategies of the Fe3O4/activated carbon composite. The utilization of biopolymer-based nanomagnetic adsorbents is on the rise due to their exceptional adsorption capabilities and high surface area. Among the magnetic materials, Fe3O4, Fe0, γ-Fe2O3, and MFe2O4 (M = Cu, Co, Ni and Zn) are well known and commonly employed [178]. Fe3O4 and γ-Fe2O3, in particular, are favored due to their strong superparamagnetic properties, ease of production, and low toxicity. Figure 12A illustrates a simple synthetic process for producing chitosan-based Fe3O4 nanoparticle composites [179].
The magnetism exhibited by these materials plays a crucial role in improving water purification by influencing the physical properties of pollutants. Morphological changes of chitosan-based magnetic nanoparticles are depicted in Figure 12B [179]. Combining these magnetic materials with biopolymers can create composite materials that can be separated magnetically, as shown in Figure 12C [180]. Regeneration tests have been conducted for most magnetic adsorbents, as they can be easily recycled without significant reduction in adsorption capabilities even after 5 to 10 cycles (Figure 12D) [180,181]. Consequently, magnetic-polysaccharide-based adsorbents have garnered significant attention as innovative agents for water purification, offering benefits such as biocompatibility, high performance, low cost, easy separation, and renewability. Combining diverse functionalization and modification techniques enables more efficient removal of mixed pollutants. Organic contaminants are primarily adsorbed through electrostatic interaction, hydrogen bonding, π-π interaction, and ion exchange, while inorganic pollutants are eliminated via electrostatic interaction, complexation, chelation, or ion exchange. The easy separation and reusability of magnetic adsorbents, facilitated by magnetic control, make them highly desirable. Iron oxide nanoparticles, particularly magnetite and maghemite, exhibit superparamagnetic strength and are extensively studied for water purification owing to their large surface area and cost-effectiveness [182].
In contrast, the GO-Fe3O4-chitosan composites have demonstrated exceptional dye removal capabilities, removing 98% (Qmax 249.25 mg/g) of dyes from water [183]. Moreover, the paramagnetic nature of the Fe3O4 intercalated composite allows for effective adsorption in the presence of a magnetic field [183,184]. Another composite, chitosan/poly (methacrylic acid)/Fe3O4-GO, has great wastewater purification potential [185]. This composite exhibited a high adsorption capacity with a Qmax of 2478 mg/g, efficiently removing MB dye from water within 20 min.
Furthermore, nanocomposites based on guar gum and other biopolymers have emerged as promising materials for removing metal ions and dyes from wastewater [186]. Chitosan and composites have also proven highly effective in absorbing various metal ions and organic dye molecules [180]. Additionally, a magnetic nanoparticle-starch-g-poly (vinyl sulfate) nanocomposite exhibited remarkable adsorption capacities for cationic dyes, with 621 mg/g for MB and 567 mg/g for MG [187]. The sulfate groups on the surface of the adsorbent served as active sites for the adsorption of cationic molecules.
Table 3 provides an overview of recently reported iron-oxide-biopolymer-based polymer nanocomposites investigated for water purification applications. These nanocomposites represent a significant advancement in the field, highlighting their potential for addressing water pollution challenges.
However, even with the successful adsorption of contaminants, there is a concern that some residual pollutants may remain, potentially leading to secondary pollution. To address this issue, developing techniques for completely degrading organic contaminants in water has gained significant attention (catalytic degradation). On the other hand, the membrane filtration approach offers promising possibilities to separate fine particles.

5. Membrane Filtration

Membrane filtration techniques are widely used in water treatment to remove contaminants and particles [211]. Two main types of membranes are used in water treatment: symmetric (homogeneous) and asymmetric (heterogeneous) membranes. These membranes differ in terms of pore homogeneity and structure [212,213]. Symmetric membranes comprise a single polymer material, resulting in uniform pore sizes throughout the membrane. On the other hand, asymmetric membranes consist of two distinct layers: a thin outer layer that controls selectivity and a thicker, porous inner layer that provides support [214]. The asymmetric membranes offer higher water flux than symmetric membranes due to their thinner structure, making them suitable for large-scale water purification applications [215]. In addition to symmetric and asymmetric membranes, composite membranes have gained popularity recently. These membranes have a selective layer composed of various inorganic metal oxides. Composite membranes outperform symmetric membranes and even asymmetric membranes in terms of performance. This is mainly attributed to the thin selective layer in composite membranes, which enhances water flux and improves overall membrane efficiency. For example, block copolymers (BCPs) offer great potential for the fabrication of heterogeneous membranes [216]. By manipulating the self-assembled morphologies of BCP membranes, such as their cylindrical and lamellar structures, one- and two-dimensional nanofluidic channels can be readily obtained [217]. The use of macromolecular building blocks in BCPs provides a versatile, engineered, and scalable material system with well-defined pores. In comparison to small molecules, end-tethered polymer brushes offer superior control over ionic transport by increasing the spatial density of functional groups on the channel walls, extending their distribution from two dimensions to three dimensions [217]. A schematic of asymmetric heteraogenious membrane is shown in Figure 13.
On the other hand, membrane filtration techniques, such as microfiltration, nanofiltration, and ultrafiltration, are commonly employed in water treatment processes [218]. The critical distinction among these techniques lies in the pore size of the membranes used [219]. Microfiltration, ultrafiltration, and reverse osmosis remove macroparticles, microparticles, and macromolecules, including inorganic particles, organic colloids, and dissolved natural substances [220]. Nanofiltration, which falls between ultrafiltration and reverses osmosis, operates with a general membrane pore size ranging from 0.5–0.2 nm and 3–8 nm [221]. Additionally, loose nanofiltration (NF) shows great promise for selectively separating dyes and salts, making it a potential candidate for the efficient recovery and cyclic utilization of high-value-added components [222,223]. A schematic of membrane filtration is shown in Figure 14. Membranes for microfiltration and ultrafiltration can be manufactured from both ceramic and polymer materials [224,225]. Ceramic materials offer excellent chemical stability, mechanical strength, ease of cleaning, and long-term durability. However, their brittleness makes large-scale production expensive, which is three times higher than common polymers and challenging. In contrast, polymeric membranes have dominated the market for water purification for a considerable period due to their easy processing and low cost.
One viable solution is the development of ceramic–polymer composite membranes to bridge the significant gap between ceramic and polymeric membranes. These composite membranes are engineered to combine the advantages of both ceramic and polymeric materials, resulting in improved membrane performance, extended lifespan, and additional functionalities [225]. Most membranes used in water treatment, such as cellulose acetate, polyethylene, polytetrafluoroethylene, polypropylene, polyvinylidene fluoride, polyethersulfone, polyacrylonitrile, polysulfone, or other polymers, are made from polymeric materials due to their affordability, good mechanical strength, and operational flexibility [226]. Polyethylene glycol (PEG) and alumina (Al2O3) is commonly employed to enhance the selectivity and hydrophilicity of membranes as well as to act as a pore-forming agent [227].
However, a significant drawback of membrane filtration is fouling, which refers to accumulating unwanted substances on the surface or within the membrane pores. This leads to reduced filtration efficiency and increased energy consumption [228]. Fouling can occur due to the deposition of particulate matter, scaling, biofilm formation, or adsorption of organic compounds onto the membrane surface [229]. It poses a significant challenge in maintaining the long-term performance and lifespan of the membranes. The membrane’s floor morphology and chemistry influence the fouling of polymer membranes. Membrane sheets of polymers with suitable porosity, chemical and mechanical stability, and high hydrophobicity are typically employed. Natural foulants tend to adsorb onto the membrane surface due to their lyophilic (water attraction) nature [218].
Biopolymer nanocomposites have shown promising results in mitigating fouling in membrane filtration processes [221,229,230,231]. For example, incorporating nanocellulose, chitosan nanoparticles, or graphene oxide nanosheets into biopolymer membranes has improved antifouling properties by reducing the adhesion of foulants and enhancing water permeability [232,233,234]. The recent reports on biopolymer-based graphene oxide nanocomposites aim to explore the synergistic features and diverse applications of these materials, with a particular focus on identifying optimal combinations of biopolymers and carbon nanomaterials for various industrial applications [235,236,237]. The review article thoroughly explores the fabrication, application, and bonding mechanisms of biodegradable biopolymers, such as poly(lactic acid), cellulose, starch, chitosan, alginates, polyamides, and other biodegradable materials, with various forms of graphene, including graphene oxide, reduced graphene oxide, graphene nanoplatelets, etc. [238]. Furthermore, an overview of recent studies is presented, focusing on carbon nanotubes modified with natural polymers (biopolymers) like chitosan, cellulose, and cyclodextrin for water treatment applications. The discussion extends to the cost-effectiveness and economic value of using polymeric hybrid materials based on carbon nanotubes as nano-sorbents for water purification [239]. Notably, these nanocomposite membranes have demonstrated higher flux, improved fouling resistance, and prolonged operational lifetime compared to traditional membranes [240,241,242]. Surface modification techniques, such as grafting or chemical modification, can introduce functional groups that repel foulants or enhance the interaction with specific contaminants, improving separation efficiency and reducing fouling propensity [240]. For example, a PVA–silica nanofiber membrane modified with thiol exhibited superior Cu(II) ion adsorption compared to a PVA nanofiber membrane. The removal of toxic metal ions occurred through electrostatic or chelation mechanisms. PVA/PEI membranes have been utilized for the rejection of Cu(II), Cd(II), Pb(II), and Hg(II) metal ions from wastewater, demonstrating selectivity among different metal ions [243,244]. Similarly, a membrane prepared from PVA and poly (vinyl imidazole) exhibited high rejection ratios, particularly for Hg(II) ions at pH 2.5 [245]. A recent notable study demonstrated a facile surface modification of ultrafiltration (UF) membranes using polydopamine with immobilized nanosilver. The process involved initiating dopamine polymerization on the UF membrane and in situ immobilization of silver and dopamine through a redox reaction, resulting in the formation of a loose separating layer on the membrane’s surface. This loose layer exhibited enhanced fractionation capability compared to a compact layer of polydopamine formed without in situ immobilization of nanosilver and dopamine. The modified membrane showed a remarkable rejection rate of 99.9% for Eriochrome Black T (EBT) and an exceptionally high permeation flux of 39.2 LMH [221]. Table 4 summarizes recent research on membrane filtering employing bio nanocomposites materials.
The development of membrane filtration employing the nanocomposite mentioned above holds great promise for effective water treatment. CNTs/chitosan nanocomposites, lignin–cellulose citrate, polylactic acid-polybutylene succinate-polypropylene carbonate-polyhydroxybutyrate (PLA-PBS-PPC-PHB), poly(urethane)/keratin biofiber, PVA and poly (vinyl imidazole), polyvinyl alcohol/SiO2 composites, nanoclay montmorillonite–chitosan. The utilization of CNTs/chitosan nanocomposites demonstrates a high adsorption capacity for metal ions such as Cu(II), Ni(II), Pb(II), Cd(II), and Co(II), hence contributing to the efficient removal of heavy metal pollutants from water [246]. CNTs’ high surface area provides numerous active sites for metal ion adsorption, while chitosan’s amino groups (–NH2) provide binding sites for coordination and chelation. Combining carbon nanotubes and chitosan has a synergistic effect, increasing adsorption efficiency. The nanocomposites’ pH dependency allows for adjustment of adsorption capacity, while their selectivity allows for targeted removal of specific metal ions. Lignin–cellulose-citrate-based membranes have high selectivity for anions and cations, making them suitable for ion separation and purification operations [230]. The functional groups in lignin and cellulose citrate are principally responsible for the membranes’ selectivity. Lignin has phenolic hydroxyl (ߝOH) and carboxyl (ߝCOOH) groups, whereas cellulose citrate has carboxylate (ߝCOO) groups. Electrostatic interactions, hydrogen bonding, and complexation are all ways these functional groups interact with ions. The carboxylate groups in cellulose citrate can effectively connect with positively charged ions via ion exchange reactions in anions. The phenolic hydroxyl groups in lignin can also contribute to anion binding through hydrogen bonding and complexation. This combination of functional groups improves the membrane’s selectivity for anions, allowing for efficient separation and removal.
The oil–grease and total dissolved solids (TDS) removal characteristics of the multi-component polymer mix PLA-PBS-PPC-PHB make it appropriate for treating oily wastewater [150]. Poly(urethane)/keratin biofiber membranes remove Cr(VI) ions well, indicating their potential for treating chromium-contaminated water [247]. PVA and poly (vinyl imidazole) membranes are effective for mercury removal from water sources due to their high Hg(II) removal effectiveness [245]. The high performance of these membranes in mercury removal can be attributed to the affinity of imidazole groups towards metal ions, including Hg(II). Imidazole groups possess lone pair electrons on the nitrogen atom, which can form coordination bonds with metal ions through complexation. These interactions allow the membrane to capture and retain mercury ions effectively. Furthermore, the hydroxyl groups in PVA can also contribute to removing mercury ions through hydrogen bonding and complexation. The hydroxyl groups can form strong interactions with Hg(II), promoting its adsorption and retention on the membrane surface. Polyvinyl alcohol/SiO2 composites exhibit good Cu(II) adsorption capacity, contributing to improved water purification [248]. It is due to the synergistic effect between PVA and SiO2, resulting from their respective chemical properties. Combining the hydroxyl groups in PVA and the adsorption sites provided by SiO2 nanoparticles creates a synergistic effect, enhancing the overall adsorption capacity for Cu(II) ions. The hydroxyl groups of PVA and the silanol groups of SiO2 work cooperatively to capture and bind Cu(II) ions, increasing the efficiency of water purification processes. Integrating nanoclay montmorillonite-chitosan in membranes results in excellent COD removal, demonstrating its potential for treating organic contaminants in wastewater [249]. Chitosan-linked activated carbon–nano-bentonite membranes have a high polyaromatic hydrocarbon (PAH) removal efficiency, making them appropriate for removing organic pollutants. TiO2-loaded fly ash chitosan composite membranes show good dye removal efficacy for Congo red and Methylene blue, suggesting a potential solution for dye-contaminated water [251]. The adsorption process of chitosan successfully catches and maintains the dye molecules on the membrane surface via electrostatic interactions and hydrogen bonding. The synergistic action of TiO2 and chitosan results in increased water filtration performance for dye-contaminated water.

6. Limitation, Challenges, and Opportunities

Biopolymer-based nanocomposites, incorporating adsorbents like graphene oxide, carbon nanotubes, and nanoclays, offer significant potential for water treatment applications [232,233,234]. However, certain challenges need to be addressed to fully realize their benefits. These include scalability and cost-effectiveness in large-scale production, improved nanocomposite stability and reusability for long-term performance, and efficient nanofiller dispersion within the biopolymer matrix [252]. Environmental sustainability requires careful consideration of nanocomposite regeneration and used adsorbent disposal. Despite these obstacles, opportunities exist for innovation through research and development, leading to advanced nanocomposite materials and enhanced synthesis methods [238].
Combining magnetic adsorbents [253] and membrane filtration technology can enhance adsorption and separation efficiency, leading to more effective water treatment solutions [254]. Furthermore, integrating multifunctional and responsive biopolymer-based nanocomposites with current water treatment technologies opens up possibilities for energy-efficient and on-demand water purification systems [252].
Collaboration among researchers and industry stakeholders is crucial for advancing these solutions, ensuring long-term and affordable water treatment. Research on complementary methods and real industrial effluent testing will further enhance water purification effectiveness.

7. Conclusions

Biopolymer-based nanocomposites with adsorbents such as graphene oxide, carbon nanotubes, and nanoclays have much potential for water treatment applications. These nanocomposites offer better adsorption capacities, improved membrane performance, and efficient pollutant removal from water sources. Using nanomaterials and magnetic adsorbents in conjunction with membrane filtration methods yields promising results in improving efficiency and selectivity in water purification operations. However, there are obstacles to solve, such as scalability, cost-effectiveness, stability, and regeneration of nanocomposites and disposal of used adsorbents. Nonetheless, these issues can be overcome with additional research and innovation, creating sustainable and effective water treatment systems. Integrating biopolymer-based nanocomposites with sophisticated water treatment technologies like responsive materials and hybrid systems offers new possibilities for on-demand and energy-efficient water purification systems. Researchers can promote the growth of biopolymer-based nanocomposites and contribute to the development of complete and sustainable water treatment techniques by collaborating across disciplines and involving industry partners.

Author Contributions

H.K.: conceptualization, writing—original draft, writing—review and editing. C.-W.K.: project administration, funding acquisition, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The Basic Science Research Program supported this research through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2019R1I1A3A02059471) and was supported under the framework of an international cooperation program managed by the NRF of Korea (NRF-2020K2A9A2A08000181).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Werber, J.R.; Osuji, C.O.; Elimelech, M. Materials for next-generation desalination and water purification membranes. Nat. Rev. Mater. 2016, 1, 16018. [Google Scholar] [CrossRef]
  2. Coccia, M.; Bontempi, E. New trajectories of technologies for the removal of pollutants and emerging contaminants in the environment. Environ. Res. 2023, 229, 115938. [Google Scholar] [CrossRef] [PubMed]
  3. Kolya, H.; Tripathy, T. Grafted polysaccharides based on acrylamide and N,N-dimethylacrylamide: Preparation and investigation of their flocculation performances. J. Appl. Polym. Sci. 2013, 127, 2786–2795. [Google Scholar] [CrossRef]
  4. Pal, S.; Mal, D.; Singh, R.P. Cationic starch: An effective flocculating agent. Carbohydr. Polym. 2005, 59, 417–423. [Google Scholar] [CrossRef]
  5. Li, Y.; Wang, Y.; Jin, J.; Tian, Z.; Yang, W.; Graham, N.J.D.; Yang, Z. Enhanced removal of trace pesticides and alleviation of membrane fouling using hydrophobic-modified inorganic-organic hybrid flocculants in the flocculation-sedimentation-ultrafiltration process for surface water treatment. Water Res. 2023, 229, 119447. [Google Scholar] [CrossRef] [PubMed]
  6. Raeesh, M.; Devi, T.T.; Hirom, K. Recent Developments on Application of Different Turbulence and Multiphase Models in Sedimentation Tank Modeling—A Review. Water Air Soil Pollut. 2022, 234, 5. [Google Scholar] [CrossRef]
  7. Schelling, M.; Patil, K.; Boving, T.B. Sustainable Water Treatment with Induced Bank Filtration. Water 2023, 15, 361. [Google Scholar] [CrossRef]
  8. Qian, Y.; Shi, Y.; Guo, J.; Chen, Y.; Hanigan, D.; An, D. Molecular characterization of disinfection byproduct precursors in filter backwash water from 10 drinking water treatment plants. Sci. Total Environ. 2023, 856, 159027. [Google Scholar] [CrossRef]
  9. Ghorai, S.; Sarkar, A.; Raoufi, M.; Panda, A.B.; Schönherr, H.; Pal, S. Enhanced Removal of Methylene Blue and Methyl Violet Dyes from Aqueous Solution Using a Nanocomposite of Hydrolyzed Polyacrylamide Grafted Xanthan Gum and Incorporated Nanosilica. ACS Appl. Mater. Interfaces 2014, 6, 4766–4777. [Google Scholar] [CrossRef]
  10. Maraddi, A.; Halakarni, M.; Manohara Halanur, M.; Nataraj, S.K. Fe-MOF induced biopolymer-based sustainable self-cleaning membranes for effective selective separation and wastewater treatment. Sustain. Mater. Technol. 2023, 35, e00537. [Google Scholar] [CrossRef]
  11. Lee, Y.; Park, Y.-J.; Lee, J.; Bae, T.-H. Recent advances and emerging applications of membrane contactors. Chem. Eng. J. 2023, 461, 141948. [Google Scholar] [CrossRef]
  12. Mirza, A.; King, A.; Troakes, C.; Exley, C. Aluminium in brain tissue in familial Alzheimer’s disease. J. Trace Elem. Med. Biol. 2017, 40, 30–36. [Google Scholar] [CrossRef] [PubMed]
  13. Tiehm, A.; Herwig, V.; Neis, U. Particle size analysis for improved sedimentation and filtration in waste water treatment. Water Sci. Technol. 1999, 39, 99–106. [Google Scholar] [CrossRef]
  14. Verma, S.; Daverey, A.; Sharma, A. Slow sand filtration for water and wastewater treatment—A review. Environ. Technol. Rev. 2017, 6, 47–58. [Google Scholar] [CrossRef]
  15. Lin, H.; Ye, C.; Chen, S.; Zhang, S.; Yu, X. Viable but non-culturable E. coli induced by low level chlorination have higher persistence to antibiotics than their culturable counterparts. Environ. Pollut. 2017, 230, 242–249. [Google Scholar] [CrossRef]
  16. Munter, R. Advanced oxidation processes–current status and prospects. Proc. Est. Acad. Sci. Chem. 2001, 50, 59–80. [Google Scholar] [CrossRef]
  17. Nabi, M.; Liang, H.; Cheng, L.; Yang, W.; Gao, D. A comprehensive review on the use of conductive materials to improve anaerobic digestion: Focusing on landfill leachate treatment. J. Environ. Manag. 2022, 309, 114540. [Google Scholar] [CrossRef]
  18. Ping, Q.; Cohen, B.; Dosoretz, C.; He, Z. Long-term investigation of fouling of cation and anion exchange membranes in microbial desalination cells. Desalination 2013, 325, 48–55. [Google Scholar] [CrossRef]
  19. Chai, W.S.; Cheun, J.Y.; Kumar, P.S.; Mubashir, M.; Majeed, Z.; Banat, F.; Ho, S.-H.; Show, P.L. A review on conventional and novel materials towards heavy metal adsorption in wastewater treatment application. J. Clean. Prod. 2021, 296, 126589. [Google Scholar] [CrossRef]
  20. Dharupaneedi, S.P.; Nataraj, S.K.; Nadagouda, M.; Reddy, K.R.; Shukla, S.S.; Aminabhavi, T.M. Membrane-based separation of potential emerging pollutants. Sep. Purif. Technol. 2019, 210, 850–866. [Google Scholar] [CrossRef]
  21. Thurakkal, L.; Porel, M. Removal of cationic and anionic dyes from waste water in five seconds by a facile synthesis of a covalent organic polymer. Eur. Polym. J. 2023, 194, 112169. [Google Scholar] [CrossRef]
  22. Procopio, E.Q.; Fukushima, T.; Barea, E.; Navarro, J.A.R.; Horike, S.; Kitagawa, S. A Soft Copper(II) Porous Coordination Polymer with Unprecedented Aqua Bridge and Selective Adsorption Properties. Chem. A Eur. J. 2012, 18, 13117–13125. [Google Scholar] [CrossRef] [PubMed]
  23. Huo, M.-X.; Jin, Y.-L.; Sun, Z.-F.; Ren, F.; Pei, L.; Ren, P.-G. Facile synthesis of chitosan-based acid-resistant composite films for efficient selective adsorption properties towards anionic dyes. Carbohydr. Polym. 2021, 254, 117473. [Google Scholar] [CrossRef] [PubMed]
  24. Bhuyan, C.; Konwar, A.; Bora, P.; Rajguru, P.; Hazarika, S. Cellulose nanofiber-poly(ethylene terephthalate) nanocomposite membrane from waste materials for treatment of petroleum industry wastewater. J. Hazard. Mater. 2023, 442, 129955. [Google Scholar] [CrossRef]
  25. Kolya, H.; Kang, C.-W. Bio-Based Polymeric Flocculants and Adsorbents for Wastewater Treatment. Sustainability 2023, 15, 9844. [Google Scholar] [CrossRef]
  26. Stephen, D.P.; Palanisamy, S.B. Advances in biopolymer composites and biomaterials for the removal of emerging contaminants. Phys. Sci. Rev. 2021, 000010151520210056. [Google Scholar] [CrossRef]
  27. Arif, Z.U.; Khalid, M.Y.; Sheikh, M.F.; Zolfagharian, A.; Bodaghi, M. Biopolymeric sustainable materials and their emerging applications. J. Environ. Chem. Eng. 2022, 10, 108159. [Google Scholar] [CrossRef]
  28. Joshi, P.; Mehta, S.; Pandey, A.; Khatri, O.P. Inorganic Analogues of Graphene and Their Nanocomposites for Wastewater Treatment BT—Two-Dimensional Materials for Environmental Applications; Kumar, N., Gusain, R., Sinha Ray, S., Eds.; Springer International Publishing: Cham, Switzerland, 2023; pp. 75–102. ISBN 978-3-031-28756-5. [Google Scholar]
  29. Subash, A.; Naebe, M.; Wang, X.; Kandasubramanian, B. Biopolymer—A sustainable and efficacious material system for effluent removal. J. Hazard. Mater. 2023, 443, 130168. [Google Scholar] [CrossRef]
  30. Nawaz, S.; Tabassum, A.; Muslim, S.; Nasreen, T.; Baradoke, A.; Kim, T.H.; Boczkaj, G.; Jesionowski, T.; Bilal, M. Effective assessment of biopolymer-based multifunctional sorbents for the remediation of environmentally hazardous contaminants from aqueous solutions. Chemosphere 2023, 329, 138552. [Google Scholar] [CrossRef]
  31. Saha, A.; Bardhan, S.; Roy, S.; Dutta, S.; Das, S. Biopolymeric Hydrogels: A New Era in Combating Heavy Metal Pollution in Industrial Wastewater BT—Membranes for Water Treatment and Remediation; Nadda, A.K., Banerjee, P., Sharma, S., Nguyen-Tri, P., Eds.; Springer Nature Singapore: Singapore, 2023; pp. 209–226. ISBN 978-981-19-9176-9. [Google Scholar]
  32. Haasler, S.; Christensen, M.L.; Reitzel, K. Synthetic and biopolymers for lake restoration—An evaluation of flocculation mechanism and dewatering performance. J. Environ. Manag. 2023, 331, 117199. [Google Scholar] [CrossRef]
  33. Rocha-Pimienta, J.; Navajas-Preciado, B.; Barraso-Gil, C.; Martillanes, S.; Delgado-Adámez, J. Optimization of the Extraction of Chitosan and Fish Gelatin from Fishery Waste and Their Antimicrobial Potential as Active Biopolymers. Gels 2023, 9, 254. [Google Scholar] [CrossRef] [PubMed]
  34. Romero-Montero, A.; Valencia-Bermúdez, J.L.; Rosas-Meléndez, S.A.; Núñez-Tapia, I.; Piña-Barba, M.C.; Leyva-Gómez, G.; Del Prado-Audelo, M.L. Biopolymeric Fibrous Aerogels: The Sustainable Alternative for Water Remediation. Polymers 2023, 15, 262. [Google Scholar] [CrossRef] [PubMed]
  35. Paul, A.; Sharma, S.S. Biopolymer-Based Nanocomposites BT. In Handbook of Biopolymers; Thomas, S., Ajitha, A.R., Jose Chirayil, C., Thomas, B., Eds.; Springer Nature: Singapore, 2022; pp. 1–28. ISBN 978-981-16-6603-2. [Google Scholar]
  36. Meera, K.; Ramesan, M.T. Tailoring the performance of boehmite nanoparticles reinforced carboxymethyl chitosan/cashew gum blend nanocomposites via green synthesis. Polymer 2023, 268, 125706. [Google Scholar] [CrossRef]
  37. Jafarzadeh, S.; Nooshkam, M.; Zargar, M.; Garavand, F.; Ghosh, S.; Hadidi, M.; Forough, M. Green synthesis of nanomaterials for smart biopolymer packaging: Challenges and outlooks. J. Nanostruct. Chem. 2023, 1–24. [Google Scholar] [CrossRef]
  38. Khan, I.; Khan, I.; Saeed, K.; Ali, N.; Zada, N.; Khan, A.; Ali, F.; Bilal, M.; Akhter, M.S. 7—Polymer nanocomposites: An overview. In Micro and Nano Technologies; Ali, N., Bilal, M., Khan, A., Nguyen, T.A., Gupta, R.K.B.T.-S.P.N., Eds.; Elsevier: Amsterdam, The Netherlands, 2023; pp. 167–184. ISBN 978-0-323-91611-0. [Google Scholar]
  39. Kolya, H.; Kuila, T.; Kim, N.H.; Lee, J.H. 10—Polymer Nanocomposites for Energy-Related Applications; Subramani, N.K., Siddaramaiah, H., Lee, J.H., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 215–248. ISBN 978-0-12-818484-4. [Google Scholar]
  40. Kolya, H.; Mondal, S.; Kang, C.-W.; Nah, C. The use of polymer-graphene composites in catalysis. In Polymer Nanocomposites Containing Graphene; Woodhead Publishing Series in Composites Science and Engineering; Elsevier: Amsterdam, The Netherlands, 2022; pp. 537–556. [Google Scholar]
  41. Baigorria, E.; Souza dos Santos, S.; de Moura, M.R.; Fraceto, L.F. Nanocomposite hydrogels 3D printed for application in water remediation. Mater. Today Chem. 2023, 30, 101559. [Google Scholar] [CrossRef]
  42. Rani Sethy, T.; Biswal, T.; Kumar Sahoo, P. An indigenous tool for the adsorption of rare earth metal ions from the spent magnet e-waste: An eco-friendly chitosan biopolymer nanocomposite hydrogel. Sep. Purif. Technol. 2023, 309, 122935. [Google Scholar] [CrossRef]
  43. Nandanwar, P.; Jugade, R.; Gomase, V.; Shekhawat, A.; Bambal, A.; Saravanan, D.; Pandey, S. Chitosan-Biopolymer-Entrapped Activated Charcoal for Adsorption of Reactive Orange Dye from Aqueous Phase and CO2 from Gaseous Phase. J. Compos. Sci. 2023, 7, 103. [Google Scholar] [CrossRef]
  44. Gupta, P.; Rana, N.; Dash, S.; Gupta, N.; Singh, M.; Kaushik, S. Bionanocomposite Membranes and Adsorbents for Water and Wastewater Treatment. In Bionanotechnology Towards Sustainable Management of Environmental Pollution; CRC Press: Boca Raton, FL, USA, 2023; pp. 141–166. ISBN 1003270956. [Google Scholar]
  45. Das, A.; Ringu, T.; Ghosh, S.; Pramanik, N. A comprehensive review on recent advances in preparation, physicochemical characterization, and bioengineering applications of biopolymers. Polym. Bull. 2023, 80, 7247–7312. [Google Scholar] [CrossRef] [PubMed]
  46. Muhammed Shameem, M.; Sasikanth, S.M.; Annamalai, R.; Ganapathi Raman, R. A brief review on polymer nanocomposites and its applications. Mater. Today Proc. 2021, 45, 2536–2539. [Google Scholar] [CrossRef]
  47. Goyal, R.K. Nanomaterials and Nanocomposites: Synthesis, Properties, Characterization Techniques, and Applications; CRC Press: Boca Raton, FL, USA, 2017; ISBN 1498761674. [Google Scholar]
  48. Mittal, V. Characterization Techniques for Polymer Nanocomposites; John Wiley & Sons: Hoboken, NJ, USA, 2012; ISBN 3527654526. [Google Scholar]
  49. Hu, P.; Yang, H. Insight into the physicochemical aspects of kaolins with different morphologies. Appl. Clay Sci. 2013, 74, 58–65. [Google Scholar] [CrossRef]
  50. Elmore, A.R. Cosmetic Ingredient Review Expert Panel: Final report on the safety assessment of aluminum silicate, calcium silicate, magnesium aluminum silicate, magnesium silicate, magnesium trisilicate, sodium magnesium silicate, zirconium silicate, attapulgite, bentonite, Fuller’s earth, hectorite, kaolin. Int. J. Toxicol. 2003, 22 (Suppl. 1), 37–102. [Google Scholar] [PubMed]
  51. Muwanguzi, A.J.B.; Karasev, A.V.; Byaruhanga, J.K.; Jönsson, P.G. Characterization of Chemical Composition and Microstructure of Natural Iron Ore from Muko Deposits. ISRN Mater. Sci. 2012, 2012, 174803. [Google Scholar] [CrossRef]
  52. Das, S.K.; Kumar, S.; Ramachandrarao, P. Exploitation of iron ore tailing for the development of ceramic tiles. Waste Manag. 2000, 20, 725–729. [Google Scholar] [CrossRef]
  53. Gilmour, P.S.; Brown, D.M.; Lindsay, T.G.; Beswick, P.H.; MacNee, W.; Donaldson, K. Adverse health effects of PM10 particles: Involvement of iron in generation of hydroxyl radical. Occup. Environ. Med. 1996, 53, 817–822. [Google Scholar] [CrossRef] [PubMed]
  54. Vassilev, S.V.; Menendez, R.; Diaz-Somoano, M.; Martinez-Tarazona, M.R. Phase-mineral and chemical composition of coal fly ashes as a basis for their multicomponent utilization. 2. Characterization of ceramic cenosphere and salt concentrates. Fuel 2004, 83, 585–603. [Google Scholar] [CrossRef]
  55. Haibin, L.; Zhenling, L. Recycling utilization patterns of coal mining waste in China. Resour. Conserv. Recycl. 2010, 54, 1331–1340. [Google Scholar] [CrossRef]
  56. DeGraeve, G.M.; Overcast, R.L.; Bergman, H.L. Toxicity of underground coal gasification condenser water and selected constituents to aquatic biota. Arch. Environ. Contam. Toxicol. 1980, 9, 543–555. [Google Scholar] [CrossRef]
  57. Chen, A.; Lo, R. Semiconductor Packaging; CRC Press: Boca Raton, FL, USA, 2011; ISBN 978-1-4398-6205-6. [Google Scholar]
  58. Jäger-Waldau, A. Chapter IC-4—Progress in Chalcopyrite Compound Semiconductor Research for Photovoltaic Applications and Transfer of Results into Actual Solar Cell Production; McEvoy, A., Markvart, T., Castañer, L., Eds.; Academic Press: Boston, MA, USA, 2012; pp. 373–395. ISBN 978-0-12-385934-1. [Google Scholar]
  59. Humborg, C.; Conley, D.J.; Rahm, L.; Wulff, F.; Cociasu, A.; Ittekkot, V. Silicon Retention in River Basins: Far-reaching Effects on Biogeochemistry and Aquatic Food Webs in Coastal Marine Environments. AMBIO A J. Hum. Environ. 2000, 29, 45–50. [Google Scholar] [CrossRef]
  60. Rondeau, V.; Jacqmin-Gadda, H.; Commenges, D.; Helmer, C.; Dartigues, J.-F. Aluminum and Silica in Drinking Water and the Risk of Alzheimer’s Disease or Cognitive Decline: Findings From 15-Year Follow-up of the PAQUID Cohort. Am. J. Epidemiol. 2008, 169, 489–496. [Google Scholar] [CrossRef]
  61. Santore, R.C.; Di Toro, D.M.; Paquin, P.R.; Allen, H.E.; Meyer, J.S. Biotic ligand model of the acute toxicity of metals. 2. Application to acute copper toxicity in freshwater fish and Daphnia. Environ. Toxicol. Chem. 2001, 20, 2397–2402. [Google Scholar] [CrossRef]
  62. Erickson, R.J.; Benoit, D.A.; Mattson, V.R.; Leonard, E.N.; Nelson, H.P., Jr. The effects of water chemistry on the toxicity of copper to fathead minnows. Environ. Toxicol. Chem. 1996, 15, 181–193. [Google Scholar] [CrossRef]
  63. Sunda, W. The Relationship between Cupric Ion Activity and the Toxicity of Copper to Phytoplankton. Ph.D. Thesis, Massachusetts Institute of Technology and Woods Hole Oceanographic Institution, Cambridge, MA, USA, 1976; pp. 511–529. [Google Scholar]
  64. Brewer, G.J. Copper toxicity in Alzheimer’s disease: Cognitive loss from ingestion of inorganic copper. J. Trace Elem. Med. Biol. 2012, 26, 89–92. [Google Scholar] [CrossRef]
  65. Desai, V.; Kaler, S.G. Role of copper in human neurological disorders. Am. J. Clin. Nutr. 2008, 88, 855S–858S. [Google Scholar] [CrossRef] [PubMed]
  66. Das, K.K.; Das, S.N. Nickel, its adverse health effects & oxidative stress—ProQuest. Indian J. Med. Res. 2008, 128, 412–425. [Google Scholar] [CrossRef] [PubMed]
  67. Lin, C.-Y.C.T.-H. Nickel toxicity to human term placenta: In vitro study on lipid peroxidation. J. Toxicol. Environ. Heal. Part A 1998, 54, 37–47. [Google Scholar] [CrossRef]
  68. Fosmire, G.J. Zinc toxicity. Am. J. Clin. Nutr. 1990, 51, 225–227. [Google Scholar] [CrossRef]
  69. Lock, K.; Janssen, C.R. Modeling zinc toxicity for terrestrial invertebrates. Environ. Toxicol. Chem. 2001, 20, 1901–1908. [Google Scholar] [CrossRef]
  70. Gidlow, D.A. Lead toxicity. Occup. Med. 2004, 54, 76–81. [Google Scholar] [CrossRef]
  71. Senthil Kumar, P. Adsorption of lead(II) ions from simulated wastewater using natural waste: A kinetic, thermodynamic and equilibrium study. Environ. Prog. Sustain. Energy 2014, 33, 55–64. [Google Scholar] [CrossRef]
  72. Sigel, A.; Sigel, H. Metal Ions in Biological Systems: Volume 34: Mercury and Its Effects on Environment and Biology; CRC Press: Boca Raton, FL, USA, 1997; ISBN 0824798287. [Google Scholar]
  73. Losi, M.E.; Amrhein, C.; Frankenberger, W.T. Environmental Biochemistry of Chromium BT—Reviews of Environmental Contamination and Toxicology; Ware, G.W., Ed.; Springer: New York, NY, USA, 1994; pp. 91–121. ISBN 978-1-4612-2656-7. [Google Scholar]
  74. Costa, M.; Klein, C.B. Toxicity and Carcinogenicity of Chromium Compounds in Humans. Crit. Rev. Toxicol. 2006, 36, 155–163. [Google Scholar] [CrossRef]
  75. Lee, J.-J.; Kim, Y.-K.; Cho, S.-H.; Park, K.-S.; Chung, I.-J.; Cho, D.; Ryang, D.-W.; Kim, H.-J. Hemolytic Anemia as a Sequela of Arsenic Intoxication Following Long-Term Ingestion of Traditional Chinese Medicine. J. Korean Med. Sci. 2004, 19, 127–129. [Google Scholar] [CrossRef]
  76. Smith, R.; Knight, R.; Fendorf, S. Overpumping leads to California groundwater arsenic threat. Nat. Commun. 2018, 9, 2089. [Google Scholar] [CrossRef] [PubMed]
  77. Cleinmensen, S.; Jensen, J.C.; Jensen, N.J.; Meyer, O.; Olsen, P.; Würtzen, G. Toxicological studies on malachite green: A triphenylmethane dye. Arch. Toxicol. 1984, 56, 43–45. [Google Scholar] [CrossRef]
  78. Lorenzo, A.; Yankner, B.A. Beta-amyloid neurotoxicity requires fibril formation and is inhibited by congo red. Proc. Natl. Acad. Sci. USA 1994, 91, 12243–12247. [Google Scholar] [CrossRef]
  79. Afkhami, A.; Moosavi, R. Adsorptive removal of Congo red, a carcinogenic textile dye, from aqueous solutions by maghemite nanoparticles. J. Hazard. Mater. 2010, 174, 398–403. [Google Scholar] [CrossRef] [PubMed]
  80. Top, W.M.; Gillman, P.K.; de Langen, C.J.; Kooy, A. Fatal methylene blue associated serotonin toxicity. Neth. J. Med. 2014, 72, 179–181. [Google Scholar] [PubMed]
  81. Littlefield, N.A.; Blackwell, B.-N.; Hewitt, C.C.; Gaylor, D.W. Chronic Toxicity and Carcinogenicity Studies of Gentian Violet in Mice. Toxicol. Sci. 1985, 5, 902–912. [Google Scholar] [CrossRef]
  82. Zhang, W.; Liu, W.; Zhang, J.; Zhao, H.; Zhang, Y.; Quan, X.; Jin, Y. Characterisation of acute toxicity, genotoxicity and oxidative stress posed by textile effluent on zebrafish. J. Environ. Sci. 2012, 24, 2019–2027. [Google Scholar] [CrossRef]
  83. Rehman, K.; Shahzad, T.; Sahar, A.; Hussain, S.; Mahmood, F.; Siddique, M.H.; Siddique, M.A.; Rashid, M.I. Effect of Reactive Black 5 azo dye on soil processes related to C and N cycling. PeerJ 2018, 6, e4802. [Google Scholar] [CrossRef]
  84. Epolito, W.J.; Lee, Y.H.; Bottomley, L.A.; Pavlostathis, S.G. Characterization of the textile anthraquinone dye Reactive Blue 4. Dye. Pigment. 2005, 67, 35–46. [Google Scholar] [CrossRef]
  85. Carneiro, P.A.; Osugi, M.E.; Fugivara, C.S.; Boralle, N.; Furlan, M.; Zanoni, M.V.B. Evaluation of different electrochemical methods on the oxidation and degradation of Reactive Blue 4 in aqueous solution. Chemosphere 2005, 59, 431–439. [Google Scholar] [CrossRef] [PubMed]
  86. Dehghani, M.H.; Naghizadeh, A.; Rashidi, A.; Derakhshani, E. Adsorption of reactive blue 29 dye from aqueous solution by multiwall carbon nanotubes. Desalin. Water Treat. 2013, 51, 7655–7662. [Google Scholar] [CrossRef]
  87. Shiva Shankar, Y.; Ankur, K.; Bhushan, P.; Mohan, D. Utilization of Water Treatment Plant (WTP) Sludge for Pretreatment of Dye Wastewater Using Coagulation/Flocculation BT—Advances in Waste Management; Kalamdhad, A.S., Singh, J., Dhamodharan, K., Eds.; Springer: Singapore, 2019; pp. 107–121. [Google Scholar]
  88. Kazi, S.A.; Iqbal, H.H.; Shahid, N.; Shah, G.M.; Jameel, N. Removal of reactive dye yellow 145 by adsorption using white quartz. Bull. Environ. Stud 2016, 1, 43. [Google Scholar]
  89. Nassiri Koopaei, N.; Abdollahi, M. Health risks associated with the pharmaceuticals in wastewater. DARU J. Pharm. Sci. 2017, 25, 9. [Google Scholar] [CrossRef]
  90. Wasi, S.; Tabrez, S.; Ahmad, M. Toxicological effects of major environmental pollutants: An overview. Environ. Monit. Assess. 2013, 185, 2585–2593. [Google Scholar] [CrossRef]
  91. Zhao, Y.; Ye, L.; Zhang, X. Emerging pollutants–part I: Occurrence, fate and transport. Water Environ. Res. 2018, 90, 1301–1322. [Google Scholar] [CrossRef]
  92. Rochman, C.M.; Browne, M.A.; Underwood, A.J.; van Franeker, J.A.; Thompson, R.C.; Amaral-Zettler, L.A. The ecological impacts of marine debris: Unraveling the demonstrated evidence from what is perceived. Ecology 2016, 97, 302–312. [Google Scholar] [CrossRef]
  93. Prata, J.C.; da Costa, J.P.; Lopes, I.; Duarte, A.C.; Rocha-Santos, T. Environmental exposure to microplastics: An overview on possible human health effects. Sci. Total Environ. 2020, 702, 134455. [Google Scholar] [CrossRef]
  94. Kolya, H.; Kang, C.-W. Synthesis of starch-based smart hydrogel derived from rice-cooked wastewater for agricultural use. Int. J. Biol. Macromol. 2023, 226, 1477–1489. [Google Scholar] [CrossRef]
  95. Kolya, H.; Jana, D.; Sasmal, D.; Jana, S.; Tripathy, T. Sulfated katira gum-graft-poly(N-vinyl imidazole): A useful scavenger of mercury(II) ions. J. Appl. Polym. Sci. 2017, 134. [Google Scholar] [CrossRef]
  96. Kumari, S.; Kishor, R. Chapter 1—Chitin and Chitosan: Origin, Properties, and Applications; Gopi, S., Thomas, S., Pius, A., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 1–33. ISBN 978-0-12-817970-3. [Google Scholar]
  97. Ait Hamdan, Y.; El Amerany, F.; Desbrières, J.; Aghrinane, A.; Oudadesse, H.; Rhazi, M. The evolution of the global COVID-19 epidemic in Morocco and understanding the different therapeutic approaches of chitosan in the control of the pandemic. Polym. Bull. 2022, 1–27. [Google Scholar] [CrossRef]
  98. Gacesa, P. Alginates. Carbohydr. Polym. 1988, 8, 161–182. [Google Scholar] [CrossRef]
  99. Pereira, L.; Cotas, J. Introductory Chapter: Alginates—A General Overview; Pereira, L., Ed.; IntechOpen: Rijeka, Croatia, 2020; p. Ch. 1. ISBN 978-1-78985-642-2. [Google Scholar]
  100. Alavinia, S.; Ghorbani-Vaghei, R.; Asadabadi, S.; Atrian, A. Sodium alginate/diethyleneamine-triazine-sulfonamide nanocomposite for adsorptive removal of Pb(II) and methyl violet from aqueous solutions. Mater. Chem. Phys. 2023, 293, 126915. [Google Scholar] [CrossRef]
  101. Kolya, H.; Pal, S.; Pandey, A.; Tripathy, T. Preparation of gold nanoparticles by a novel biodegradable graft copolymer sodium alginate-g-poly (N,N-dimethylacrylamide-co-acrylic acid) with anti micro bacterial application. Eur. Polym. J. 2015, 66, 139–148. [Google Scholar] [CrossRef]
  102. Sen, G.; Singh, R.P.; Pal, S. Microwave-initiated synthesis of polyacrylamide grafted sodium alginate: Synthesis and characterization. J. Appl. Polym. Sci. 2010, 115, 63–71. [Google Scholar] [CrossRef]
  103. Dong, Y.-D.; Zhang, H.; Zhong, G.-J.; Yao, G.; Lai, B. Cellulose/carbon Composites and their Applications in Water Treatment—A Review. Chem. Eng. J. 2021, 405, 126980. [Google Scholar] [CrossRef]
  104. Das, R.; Panda, A.B.; Pal, S. Synthesis and characterization of a novel polymeric hydrogel based on hydroxypropyl methyl cellulose grafted with polyacrylamide. Cellulose 2012, 19, 933–945. [Google Scholar] [CrossRef]
  105. Falua, K.J.; Pokharel, A.; Babaei-Ghazvini, A.; Ai, Y.; Acharya, B. Valorization of Starch to Biobased Materials: A Review. Polymers 2022, 14, 2215. [Google Scholar] [CrossRef]
  106. Kolya, H.; Das, S.; Tripathy, T. Synthesis of Starch-g-Poly-(N-methylacrylamide-co-acrylic acid) and its application for the removal of mercury (II) from aqueous solution by adsorption. Eur. Polym. J. 2014, 58, 1–10. [Google Scholar] [CrossRef]
  107. Kolya, H.; Roy, A.; Tripathy, T. Starch-g-Poly-(N, N-dimethyl acrylamide-co-acrylic acid): An efficient Cr (VI) ion binder. Int. J. Biol. Macromol. 2015, 72, 560–568. [Google Scholar] [CrossRef]
  108. Kolya, H.; Tripathy, T. Preparation, investigation of metal ion removal and flocculation performances of grafted hydroxyethyl starch. Int. J. Biol. Macromol. 2013, 62, 557–564. [Google Scholar] [CrossRef] [PubMed]
  109. Guleria, S.; Singh, H.; Sharma, V.; Bhardwaj, N.; Arya, S.K.; Puri, S.; Khatri, M. Polyhydroxyalkanoates production from domestic waste feedstock: A sustainable approach towards bio-economy. J. Clean. Prod. 2022, 340, 130661. [Google Scholar] [CrossRef]
  110. Che, L.; Jin, W.; Zhou, X.; Han, W.; Chen, Y.; Chen, C.; Jiang, G. Current status and future perspectives on the biological production of polyhydroxyalkanoates. Asia-Pacific J. Chem. Eng. 2023, 18, e2899. [Google Scholar] [CrossRef]
  111. Liu, B.; Wen, Q.; Huang, L.; Chen, Z.; Lin, X.; Liu, S. Insights into integration of polyhydroxyalkanoates (PHAs) production into wastewater treatment: Comparison of different electron acceptors on system function and PHA-producer enrichment. Chem. Eng. J. 2023, 451, 138631. [Google Scholar] [CrossRef]
  112. Palaniraj, A.; Jayaraman, V. Production, recovery and applications of xanthan gum by Xanthomonas campestris. J. Food Eng. 2011, 106, 1–12. [Google Scholar] [CrossRef]
  113. Shekarforoush, E.; Faralli, A.; Ndoni, S.; Mendes, A.C.; Chronakis, I.S. Electrospinning of Xanthan Polysaccharide. Macromol. Mater. Eng. 2017, 302, 1700067. [Google Scholar] [CrossRef]
  114. Maji, B.; Maiti, S. Chemical modification of xanthan gum through graft copolymerization: Tailored properties and potential applications in drug delivery and wastewater treatment. Carbohydr. Polym. 2021, 251, 117095. [Google Scholar] [CrossRef]
  115. Kolya, H.; Tripathy, T.; De, B.R. Flocculation Performance of Grafted Xanthangum: A Comparative Study. J. Phys. Sci. 2012, 16, 221–234. [Google Scholar]
  116. Brethauer, S.; Shahab, R.L.; Studer, M.H. Impacts of biofilms on the conversion of cellulose. Appl. Microbiol. Biotechnol. 2020, 104, 5201–5212. [Google Scholar] [CrossRef]
  117. Wang, T.; Jiang, M.; Yu, X.; Niu, N.; Chen, L. Application of lignin adsorbent in wastewater Treatment: A review. Sep. Purif. Technol. 2022, 302, 122116. [Google Scholar] [CrossRef]
  118. Beck, R.J.; Zhao, Y.; Fong, H.; Menkhaus, T.J. Electrospun lignin carbon nanofiber membranes with large pores for highly efficient adsorptive water treatment applications. J. Water Process Eng. 2017, 16, 240–248. [Google Scholar] [CrossRef]
  119. del Orta, M.M.; Martín, J.; Santos, J.L.; Aparicio, I.; Medina-Carrasco, S.; Alonso, E. Biopolymer-clay nanocomposites as novel and ecofriendly adsorbents for environmental remediation. Appl. Clay Sci. 2020, 198, 105838. [Google Scholar] [CrossRef]
  120. Biswas, S.; Pal, A. Application of biopolymers as a new age sustainable material for surfactant adsorption: A brief review. Carbohydr. Polym. Technol. Appl. 2021, 2, 100145. [Google Scholar] [CrossRef]
  121. Paul, A.; Sharma, S.S. Handbook of Biopolymers; Thomas, S., Ajitha, A.R., Jose Chirayil, C., Thomas, B., Eds.; Springer Nature: Singapore, 2023; pp. 523–550. ISBN 978-981-19-0710-4. [Google Scholar]
  122. Thakur, K.; Kandasubramanian, B. Graphene and graphene oxide-based composites for removal of organic pollutants: A review. J. Chem. Eng. Data 2019, 64, 833–867. [Google Scholar] [CrossRef]
  123. Compton, O.C.; Nguyen, S.T. Graphene Oxide, Highly Reduced Graphene Oxide, and Graphene: Versatile Building Blocks for Carbon-Based Materials. Small 2010, 6, 711–723. [Google Scholar] [CrossRef]
  124. Fahim, I.S. Graphene Reinforced Biopolymer Nanocomposites for Water Filtration Applications BT—Graphene Based Biopolymer Nanocomposites; Sharma, B., Jain, P., Eds.; Springer: Singapore, 2021; pp. 221–232. ISBN 978-981-15-9180-8. [Google Scholar]
  125. Dimapilis, E.A.S.; Hsu, C.-S.; Mendoza, R.M.O.; Lu, M.-C. Zinc oxide nanoparticles for water disinfection. Sustain. Environ. Res. 2018, 28, 47–56. [Google Scholar] [CrossRef]
  126. Aragaw, T.A.; Bogale, F.M.; Aragaw, B.A. Iron-based nanoparticles in wastewater treatment: A review on synthesis methods, applications, and removal mechanisms. J. Saudi Chem. Soc. 2021, 25, 101280. [Google Scholar] [CrossRef]
  127. Kolya, H.; Kang, C.-W. Biogenic Synthesis of Silver-Iron Oxide Nanoparticles Using Kulekhara Leaves Extract for Removing Crystal Violet and Malachite Green Dyes from Water. Sustainability 2022, 14, 15800. [Google Scholar] [CrossRef]
  128. García, A.; Rodríguez, B.; Giraldo, H.; Quintero, Y.; Quezada, R.; Hassan, N.; Estay, H. Copper-Modified Polymeric Membranes for Water Treatment: A Comprehensive Review. Membranes 2021, 11, 93. [Google Scholar] [CrossRef]
  129. Kolya, H.; Kuila, T.; Kim, N.H.; Lee, J.H. Bioinspired silver nanoparticles/reduced graphene oxide nanocomposites for catalytic reduction of 4-nitrophenol, organic dyes and act as energy storage electrode material. Compos. Part B Eng. 2019, 173, 106924. [Google Scholar] [CrossRef]
  130. Safat, S.; Buazar, F.; Albukhaty, S.; Matroodi, S. Enhanced sunlight photocatalytic activity and biosafety of marine-driven synthesized cerium oxide nanoparticles. Sci. Rep. 2021, 11, 14734. [Google Scholar] [CrossRef] [PubMed]
  131. Fatima, H.; Azhar, M.R.; Zhong, Y.; Arafat, Y.; Khiadani, M.; Shao, Z. Rational design of ZnO-zeolite imidazole hybrid nanoparticles with reduced charge recombination for enhanced photocatalysis. J. Colloid Interface Sci. 2022, 614, 538–546. [Google Scholar] [CrossRef] [PubMed]
  132. Aydin, S.; Celik Karakaya, M.; Karakaya, N.; Aydin, M.E. Effective removal of selected pharmaceuticals from sewerage treatment plant effluent using natural clay (Na-montmorillonite). Appl. Water Sci. 2023, 13, 129. [Google Scholar] [CrossRef]
  133. Makaremi, M.; De Silva, R.T.; Pasbakhsh, P. Electrospun Nanofibrous Membranes of Polyacrylonitrile/Halloysite with Superior Water Filtration Ability. J. Phys. Chem. C 2015, 119, 7949–7958. [Google Scholar] [CrossRef]
  134. Zhu, J.; Tian, M.; Zhang, Y.; Zhang, H.; Liu, J. Fabrication of a novel “loose” nanofiltration membrane by facile blending with Chitosan–Montmorillonite nanosheets for dyes purification. Chem. Eng. J. 2015, 265, 184–193. [Google Scholar] [CrossRef]
  135. Pasbakhsh, P.; Churchman, G.J.; Keeling, J.L. Characterisation of properties of various halloysites relevant to their use as nanotubes and microfibre fillers. Appl. Clay Sci. 2013, 74, 47–57. [Google Scholar] [CrossRef]
  136. Huang, J.; Zhou, J.; Liu, M. Interphase in Polymer Nanocomposites. JACS Au 2022, 2, 280–291. [Google Scholar] [CrossRef]
  137. Das, R.; Ali, M.E.; Hamid, S.B.A.; Ramakrishna, S.; Chowdhury, Z.Z. Carbon nanotube membranes for water purification: A bright future in water desalination. Desalination 2014, 336, 97–109. [Google Scholar] [CrossRef]
  138. Aiswarya, R.; Kalaivani, T. Chemically Modified Carbon Nanotubes in Membrane Separation. In Chemically Modified Carbon Nanotubes for Commercial Applications; John Wiley & Sons: Newark, NJ, USA, 2023; pp. 169–195. [Google Scholar]
  139. Noamani, S.; Niroomand, S.; Rastgar, M.; Sadrzadeh, M. Carbon-based polymer nanocomposite membranes for oily wastewater treatment. npj Clean Water 2019, 2, 20. [Google Scholar] [CrossRef]
  140. Kayvani Fard, A.; Mckay, G.; Manawi, Y.; Malaibari, Z.; Hussien, M.A. Outstanding adsorption performance of high aspect ratio and super-hydrophobic carbon nanotubes for oil removal. Chemosphere 2016, 164, 142–155. [Google Scholar] [CrossRef]
  141. Mazrouaa, A.M.; Mousa, A.A.; Mohamed, M.G. Chemically Modified Carbon Nanotubes for Water Purification System. In Chemically Modified Carbon Nanotubes for Commercial Applications; John Wiley & Sons: Newark, NJ, USA, 2023; pp. 197–214. [Google Scholar]
  142. Akpotu, S.O.; Diagboya, P.N.; Lawal, I.A.; Sanni, S.O.; Pholosi, A.; Peleyeju, M.G.; Mtunzi, F.M.; Ofomaja, A.E. Designer composite of montmorillonite-reduced graphene oxide-PEG polymer for water treatment: Enrofloxacin sequestration and cost analysis. Chem. Eng. J. 2023, 453, 139771. [Google Scholar] [CrossRef]
  143. Liu, X.; Ma, R.; Wang, X.; Ma, Y.; Yang, Y.; Zhuang, L.; Zhang, S.; Jehan, R.; Chen, J.; Wang, X. Graphene oxide-based materials for efficient removal of heavy metal ions from aqueous solution: A review. Environ. Pollut. 2019, 252, 62–73. [Google Scholar] [CrossRef] [PubMed]
  144. Bhattacharya, M. Polymer Nanocomposites—A Comparison between Carbon Nanotubes, Graphene, and Clay as Nanofillers. Materials 2016, 9, 262. [Google Scholar] [CrossRef] [PubMed]
  145. Ibrahim, A.; Klopocinska, A.; Horvat, K.; Abdel Hamid, Z. Graphene-Based Nanocomposites: Synthesis, Mechanical Properties, and Characterizations. Polymers 2021, 13, 2869. [Google Scholar] [CrossRef] [PubMed]
  146. Roy, S.; Singha, N.R. Polymeric Nanocomposite Membranes for Next Generation Pervaporation Process: Strategies, Challenges and Future Prospects. Membranes 2017, 7, 53. [Google Scholar] [CrossRef]
  147. Basavegowda, N.; Baek, K.-H. Advances in Functional Biopolymer-Based Nanocomposites for Active Food Packaging Applications. Polymers 2021, 13, 4198. [Google Scholar] [CrossRef]
  148. Gangarapu, S.; Sunku, K.; Babu, P.S.; Sudarsanam, P. Fabrication of Polymer-Graphene Nanocomposites BT—Handbook of Polymer and Ceramic Nanotechnology; Hussain, C.M., Thomas, S., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 1–15. ISBN 978-3-030-10614-0. [Google Scholar]
  149. Poshina, D.; Otsuka, I. Electrospun Polysaccharidic Textiles for Biomedical Applications. Textiles 2021, 1, 152–169. [Google Scholar] [CrossRef]
  150. Ghorbani, M.; Hassan Vakili, M.; Ameri, E. Fabrication and evaluation of a biopolymer-based nanocomposite membrane for oily wastewater treatment. Mater. Today Commun. 2021, 28, 102560. [Google Scholar] [CrossRef]
  151. Valandro, S.R.; Lombardo, P.C.; Poli, A.L.; Horn, M.A., Jr.; Neumann, M.G.; Cavalheiro, C.C.S. Thermal properties of poly (methyl methacrylate)/organomodified montmorillonite nanocomposites obtained by in situ photopolymerization. Mater. Res. 2014, 17, 265–270. [Google Scholar] [CrossRef]
  152. Kolya, H.; Kang, C.W. Polyvinyl acetate/reduced graphene oxide-poly (diallyl dimethylammonium chloride) composite coated wood surface reveals improved hydrophobicity. Prog. Org. Coat. 2021, 156, 106253. [Google Scholar] [CrossRef]
  153. Rathi, B.S.; Kumar, P.S. Application of adsorption process for effective removal of emerging contaminants from water and wastewater. Environ. Pollut. 2021, 280, 116995. [Google Scholar] [CrossRef]
  154. Yu, C.; Han, X. Adsorbent material used in water treatment-a review. In Proceedings of the 2015 2nd International Workshop on Materials Engineering and Computer Sciences, Jinan, China, 10–11 October 2015; Atlantis Press: Amsterdam, The Netherlands, 2015; pp. 286–289. [Google Scholar]
  155. Saleem, J.; Bin Shahid, U.; Hijab, M.; Mackey, H.; McKay, G. Production and applications of activated carbons as adsorbents from olive stones. Biomass Convers. Biorefinery 2019, 9, 775–802. [Google Scholar] [CrossRef]
  156. Biswas, S.; Fatema, J.; Debnath, T.; Rashid, T.U. Chitosan–Clay Composites for Wastewater Treatment: A State-of-the-Art Review. ACS EST Water 2021, 1, 1055–1085. [Google Scholar] [CrossRef]
  157. Bhattacharyya, A.; Banerjee, B.; Ghorai, S.; Rana, D.; Roy, I.; Sarkar, G.; Saha, N.R.; De, S.; Ghosh, T.K.; Sadhukhan, S.; et al. Development of an auto-phase separable and reusable graphene oxide-potato starch based cross-linked bio-composite adsorbent for removal of methylene blue dye. Int. J. Biol. Macromol. 2018, 116, 1037–1048. [Google Scholar] [CrossRef] [PubMed]
  158. Qi, Y.; Yang, M.; Xu, W.; He, S.; Men, Y. Natural polysaccharides-modified graphene oxide for adsorption of organic dyes from aqueous solutions. J. Colloid Interface Sci. 2017, 486, 84–96. [Google Scholar] [CrossRef] [PubMed]
  159. Politaeva, N.; Yakovlev, A.; Yakovleva, E.; Chelysheva, V.; Tarantseva, K.; Efremova, S.; Ilyashenko, S. Graphene Oxide-Chitosan Composites for Water Treatment from Copper. Water 2022, 14, 1430. [Google Scholar] [CrossRef]
  160. Trang, T.T.C.; Takaomi, K. Chitosan and its biomass composites in application for water treatment. Curr. Opin. Green Sustain. Chem. 2021, 29, 100429. [Google Scholar] [CrossRef]
  161. Zhao, X.-R.; Xu, X.; Teng, J.; Zhou, N.; Zhou, Z.; Jiang, X.-Y.; Jiao, F.-P.; Yu, J.-G. Three-dimensional porous graphene oxide-maize amylopectin composites with controllable pore-sizes and good adsorption-desorption properties: Facile fabrication and reutilization, and the adsorption mechanism. Ecotoxicol. Environ. Saf. 2019, 176, 11–19. [Google Scholar] [CrossRef]
  162. Vo, T.S.; Vo, T.T.B.C. Organic dye removal and recycling performances of graphene oxide-coated biopolymer sponge. Prog. Nat. Sci. Mater. Int. 2022, 32, 634–642. [Google Scholar] [CrossRef]
  163. Bernard, K.N.M.; Prakash, O.; Hippargi, G.; Sylvere, N.K.; Joseph, K.G.; Pal, S. Exploring the applicability of a geopolymer and a biopolymer as an environmentally benign treatment option for heavy metals contaminated water. J. Taiwan Inst. Chem. Eng. 2022, 135, 104392. [Google Scholar] [CrossRef]
  164. Hosseinzadeh, H.; Ramin, S. Fabrication of starch-graft-poly(acrylamide)/graphene oxide/hydroxyapatite nanocomposite hydrogel adsorbent for removal of malachite green dye from aqueous solution. Int. J. Biol. Macromol. 2018, 106, 101–115. [Google Scholar] [CrossRef] [PubMed]
  165. Ma, T.; Chang, P.R.; Zheng, P.; Zhao, F.; Ma, X. Fabrication of ultra-light graphene-based gels and their adsorption of methylene blue. Chem. Eng. J. 2014, 240, 595–600. [Google Scholar] [CrossRef]
  166. Li, Y.; Du, Q.; Liu, T.; Peng, X.; Wang, J.; Sun, J.; Wang, Y.; Wu, S.; Wang, Z.; Xia, Y.; et al. Comparative study of methylene blue dye adsorption onto activated carbon, graphene oxide, and carbon nanotubes. Chem. Eng. Res. Des. 2013, 91, 361–368. [Google Scholar] [CrossRef]
  167. Zare, E.N.; Lakouraj, M.M.; Kasirian, N. Development of effective nano-biosorbent based on poly m-phenylenediamine grafted dextrin for removal of Pb (II) and methylene blue from water. Carbohydr. Polym. 2018, 201, 539–548. [Google Scholar] [CrossRef] [PubMed]
  168. Teng, J.; Zeng, X.; Xu, X.; Yu, J.-G. Assembly of a novel porous 3D graphene oxide-starch architecture by a facile hydrothermal method and its adsorption properties toward metal ions. Mater. Lett. 2018, 214, 31–33. [Google Scholar] [CrossRef]
  169. Kumari, N.; Mishra, S. Synthesis, characterization and flocculation efficiency of grafted Moringa gum based derivatives. Carbohydr. Polym. 2022, 281, 119079. [Google Scholar] [CrossRef]
  170. Zhang, L.; Tang, S.; He, F.; Liu, Y.; Mao, W.; Guan, Y. Highly efficient and selective capture of heavy metals by poly(acrylic acid) grafted chitosan and biochar composite for wastewater treatment. Chem. Eng. J. 2019, 378, 122215. [Google Scholar] [CrossRef]
  171. Sirajudheen, P.; Karthikeyan, P.; Ramkumar, K.; Meenakshi, S. Effective removal of organic pollutants by adsorption onto chitosan supported graphene oxide-hydroxyapatite composite: A novel reusable adsorbent. J. Mol. Liq. 2020, 318, 114200. [Google Scholar] [CrossRef]
  172. Travlou, N.A.; Kyzas, G.Z.; Lazaridis, N.K.; Deliyanni, E.A. Graphite oxide/chitosan composite for reactive dye removal. Chem. Eng. J. 2013, 217, 256–265. [Google Scholar] [CrossRef]
  173. Lee, S.; Lingamdinne, L.P.; Yang, J.-K.; Koduru, J.R.; Chang, Y.-Y.; Naushad, M. Biopolymer mixture-entrapped modified graphene oxide for sustainable treatment of heavy metal contaminated real surface water. J. Water Process Eng. 2022, 46, 102631. [Google Scholar] [CrossRef]
  174. Bahador, F.; Foroutan, R.; Esmaeili, H.; Ramavandi, B. Enhancement of the chromium removal behavior of Moringa oleifera activated carbon by chitosan and iron oxide nanoparticles from water. Carbohydr. Polym. 2021, 251, 117085. [Google Scholar] [CrossRef] [PubMed]
  175. Li, L.; Zhao, L.; Ma, J.; Tian, Y. Preparation of graphene oxide/chitosan complex and its adsorption properties for heavy metal ions. Green Process. Synth. 2020, 9, 294–303. [Google Scholar] [CrossRef]
  176. Pathirana, M.A.; Dissanayake, N.S.L.; Wanasekara, N.D.; Mahltig, B.; Nandasiri, G.K. Chitosan-Graphene Oxide Dip-Coated Polyacrylonitrile-Ethylenediamine Electrospun Nanofiber Membrane for Removal of the Dye Stuffs Methylene Blue and Congo Red. Nanomaterials 2023, 13, 498. [Google Scholar] [CrossRef] [PubMed]
  177. Liu, X.; Tian, J.; Li, Y.; Sun, N.; Mi, S.; Xie, Y.; Chen, Z. Enhanced dyes adsorption from wastewater via Fe3O4 nanoparticles functionalized activated carbon. J. Hazard. Mater. 2019, 373, 397–407. [Google Scholar] [CrossRef] [PubMed]
  178. Wang, J.; Zhuang, S. Chitosan-based materials: Preparation, modification and application. J. Clean. Prod. 2022, 355, 131825. [Google Scholar] [CrossRef]
  179. Wang, C.-Y.; Yang, C.-H.; Huang, K.-S.; Yeh, C.-S.; Wang, A.H.-J.; Chen, C.-H. Electrostatic droplets assisted in situ synthesis of superparamagnetic chitosan microparticles for magnetic-responsive controlled drug release and copper ion removal. J. Mater. Chem. B 2013, 1, 2205–2212. [Google Scholar] [CrossRef] [PubMed]
  180. Liu, X.; Zhao, X.; Liu, Y.; Zhang, T. Review on preparation and adsorption properties of chitosan and chitosan composites. Polym. Bull. 2021, 79, 2633–2665. [Google Scholar] [CrossRef]
  181. Mallik, A.K.; Kabir, S.M.F.; Bin Abdur Rahman, F.; Sakib, M.N.; Efty, S.S.; Rahman, M.M. Cu(II) removal from wastewater using chitosan-based adsorbents: A review. J. Environ. Chem. Eng. 2022, 10, 108048. [Google Scholar] [CrossRef]
  182. Wang, M.; You, X. Critical review of magnetic polysaccharide-based adsorbents for water treatment: Synthesis, application and regeneration. J. Clean. Prod. 2021, 323, 129118. [Google Scholar] [CrossRef]
  183. Van Hoa, N.; Khong, T.T.; Thi Hoang Quyen, T.; Si Trung, T. One-step facile synthesis of mesoporous graphene/Fe3O4/chitosan nanocomposite and its adsorption capacity for a textile dye. J. Water Process Eng. 2016, 9, 170–178. [Google Scholar] [CrossRef]
  184. Jazzar, A.; Alamri, H.; Malajati, Y.; Mahfouz, R.; Bouhrara, M.; Fihri, A. Recent advances in the synthesis and applications of magnetic polymer nanocomposites. J. Ind. Eng. Chem. 2021, 99, 1–18. [Google Scholar] [CrossRef]
  185. Sarkar, A.K.; Bediako, J.K.; Choi, J.-W.; Yun, Y.-S. Functionalized magnetic biopolymeric graphene oxide with outstanding performance in water purification. NPG Asia Mater. 2019, 11, 4. [Google Scholar] [CrossRef]
  186. Saya, L.; Malik, V.; Singh, A.; Singh, S.; Gambhir, G.; Singh, W.R.; Chandra, R.; Hooda, S. Guar gum based nanocomposites: Role in water purification through efficient removal of dyes and metal ions. Carbohydr. Polym. 2021, 261, 117851. [Google Scholar] [CrossRef] [PubMed]
  187. Pourjavadi, A.; Abedin-Moghanaki, A.; Tavakoli, A. Efficient removal of cationic dyes using a new magnetic nanocomposite based on starch-g-poly(vinylalcohol) and functionalized with sulfate groups. RSC Adv. 2016, 6, 38042–38051. [Google Scholar] [CrossRef]
  188. Stan, M.; Lung, I.; Soran, M.-L.; Opris, O.; Leostean, C.; Popa, A.; Copaciu, F.; Lazar, M.D.; Kacso, I.; Silipas, T.-D.; et al. Starch-coated green synthesized magnetite nanoparticles for removal of textile dye Optilan Blue from aqueous media. J. Taiwan Inst. Chem. Eng. 2019, 100, 65–75. [Google Scholar] [CrossRef]
  189. Xie, G.; Xi, P.; Liu, H.; Chen, F.; Huang, L.; Shi, Y.; Hou, F.; Zeng, Z.; Shao, C.; Wang, J. A facile chemical method to produce superparamagnetic graphene oxide–Fe3O4 hybrid composite and its application in the removal of dyes from aqueous solution. J. Mater. Chem. 2012, 22, 1033–1039. [Google Scholar] [CrossRef]
  190. Saber-Samandari, S.; Saber-Samandari, S.; Nezafati, N.; Yahya, K. Efficient removal of lead (II) ions and methylene blue from aqueous solution using chitosan/Fe-hydroxyapatite nanocomposite beads. J. Environ. Manag. 2014, 146, 481–490. [Google Scholar] [CrossRef]
  191. Wang, D.; Liu, L.; Jiang, X.; Yu, J.; Chen, X. Adsorption and removal of malachite green from aqueous solution using magnetic β-cyclodextrin-graphene oxide nanocomposites as adsorbents. Colloids Surf. A Physicochem. Eng. Asp. 2015, 466, 166–173. [Google Scholar] [CrossRef]
  192. 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]
  193. He, H.; Meng, X.; Yue, Q.; Yin, W.; Gao, Y.; Fang, P.; Shen, L. Thiol-ene click chemistry synthesis of a novel magnetic mesoporous silica/chitosan composite for selective Hg(II) capture and high catalytic activity of spent Hg(II) adsorbent. Chem. Eng. J. 2021, 405, 126743. [Google Scholar] [CrossRef]
  194. Zheng, C.; Zheng, H.; Hu, C.; Wang, Y.; Wang, Y.; Zhao, C.; Ding, W.; Sun, Q. Structural design of magnetic biosorbents for the removal of ciprofloxacin from water. Bioresour. Technol. 2020, 296, 122288. [Google Scholar] [CrossRef]
  195. Zheng, X.; Zheng, H.; Xiong, Z.; Zhao, R.; Liu, Y.; Zhao, C.; Zheng, C. Novel anionic polyacrylamide-modify-chitosan magnetic composite nanoparticles with excellent adsorption capacity for cationic dyes and pH-independent adsorption capability for metal ions. Chem. Eng. J. 2020, 392, 123706. [Google Scholar] [CrossRef]
  196. Perez, T.; Pasquini, D.; de Faria Lima, A.; Rosa, E.V.; Sousa, M.H.; Cerqueira, D.A.; de Morais, L.C. Efficient removal of lead ions from water by magnetic nanosorbents based on manganese ferrite nanoparticles capped with thin layers of modified biopolymers. J. Environ. Chem. Eng. 2019, 7, 102892. [Google Scholar] [CrossRef]
  197. Wu, S.; Guo, J.; Wang, Y.; Huang, C.; Hu, Y. Facile preparation of magnetic sodium alginate/carboxymethyl cellulose composite hydrogel for removal of heavy metal ions from aqueous solution. J. Mater. Sci. 2021, 56, 13096–13107. [Google Scholar] [CrossRef]
  198. Lin, S.; Zou, C.; Liang, H.; Peng, H.; Liao, Y. The effective removal of nickel ions from aqueous solution onto magnetic multi-walled carbon nanotubes modified by β-cyclodextrin. Colloids Surf. A Physicochem. Eng. Asp. 2021, 619, 126544. [Google Scholar] [CrossRef]
  199. Santhosh, C.; Daneshvar, E.; Tripathi, K.M.; Baltrėnas, P.; Kim, T.; Baltrėnaitė, E.; Bhatnagar, A. Synthesis and characterization of magnetic biochar adsorbents for the removal of Cr(VI) and Acid orange 7 dye from aqueous solution. Environ. Sci. Pollut. Res. 2020, 27, 32874–32887. [Google Scholar] [CrossRef] [PubMed]
  200. Yang, A.; Wang, Z.; Zhu, Y. Facile preparation and highly efficient sorption of magnetic composite graphene oxide/Fe3O4/GC for uranium removal. Sci. Rep. 2021, 11, 8440. [Google Scholar] [CrossRef]
  201. Sun, Y.; Wu, Y.; Fu, Y.; Yang, C.; Jiang, J.; Yan, G.; Hu, J. Rapid and high selective removal of Hg(II) ions using tannic acid cross-linking cellulose/polyethyleneimine functionalized magnetic composite. Int. J. Biol. Macromol. 2021, 182, 1120–1129. [Google Scholar] [CrossRef]
  202. Fan, L.; Luo, C.; Li, X.; Lu, F.; Qiu, H.; Sun, M. Fabrication of novel magnetic chitosan grafted with graphene oxide to enhance adsorption properties for methyl blue. J. Hazard. Mater. 2012, 215–216, 272–279. [Google Scholar] [CrossRef]
  203. Hou, S.; Zhang, Y.; Qin, G.; Song, H.; Shu, C.; Zheng, Y.; Ji, S. Designing degradable lignin-grafted magnetic nano-composite materials for cost-effectively sustainable removal of fluoroquinolone antibiotics from environmental water. J. Clean. Prod. 2022, 360, 132215. [Google Scholar] [CrossRef]
  204. Jiang, M.; Chen, L.; Niu, N. Enhanced adsorption for malachite green by functionalized lignin magnetic composites: Optimization, performance and adsorption mechanism. J. Mol. Struct. 2022, 1260, 132842. [Google Scholar] [CrossRef]
  205. Chen, J.; Zhang, L.; Li, H.; Lu, M.; Xie, J.; Guan, S.; Wang, X.; Liu, X.; Lu, J. Preparation and adsorbability of magnetic composites based on cellulose nanofiber/graphene oxide. Colloids Surf. A Physicochem. Eng. Asp. 2022, 639, 128373. [Google Scholar] [CrossRef]
  206. Ling, C.; Yimin, D.; Qi, L.; Chengqian, F.; Zhiheng, W.; Yaqi, L.; Ling, C.; Bo, L.; Yue-Fei, Z.; Yan, L.; et al. Novel High-efficiency adsorbent consisting of magnetic Cellulose-based ionic liquid for removal of anionic dyes. J. Mol. Liq. 2022, 353, 118723. [Google Scholar] [CrossRef]
  207. Chen, B.; Long, F.; Chen, S.; Cao, Y.; Pan, X. Magnetic chitosan biopolymer as a versatile adsorbent for simultaneous and synergistic removal of different sorts of dyestuffs from simulated wastewater. Chem. Eng. J. 2020, 385, 123926. [Google Scholar] [CrossRef]
  208. Bui, N.T.; Nguyen, V.H.; Le, D.T.; van Tran, T.T.; Bui, T.H. Superparamagnetic cobalt ferric nanoparticles incorporated biopolymers extracted from dragon fruit (hylocereus undatus) peels for nickel(II) removal. Environ. Technol. Innov. 2021, 23, 101773. [Google Scholar] [CrossRef]
  209. Amini Herab, A.; Salari, D.; Ostadrahimi, A.; Olad, A. Preparation of magnetic inulin nanocomposite and its application in the removal of methylene blue and heavy metals from aqueous solution. Mater. Chem. Phys. 2022, 291, 126580. [Google Scholar] [CrossRef]
  210. Jiang, R.; Shen, T.-T.; Zhu, H.-Y.; Fu, Y.-Q.; Jiang, S.-T.; Li, J.-B.; Wang, J.-L. Magnetic Fe3O4 embedded chitosan–crosslinked-polyacrylamide composites with enhanced removal of food dye: Characterization, adsorption and mechanism. Int. J. Biol. Macromol. 2023, 227, 1234–1244. [Google Scholar] [CrossRef]
  211. Cevallos-Mendoza, J.; Amorim, C.G.; Rodríguez-Díaz, J.M.; Montenegro, M.d.C.B.S.M. Removal of Contaminants from Water by Membrane Filtration: A Review. Membranes 2022, 12, 570. [Google Scholar] [CrossRef]
  212. Wu, J.; Wang, N.; Wang, L.; Dong, H.; Zhao, Y.; Jiang, L. Unidirectional water-penetration composite fibrous film via electrospinning. Soft Matter 2012, 8, 5996–5999. [Google Scholar] [CrossRef]
  213. Lee, M.; Wu, Z.; Li, K. 2—Advances in ceramic membranes for water treatment. In Woodhead Publishing Series in Energy; Basile, A., Cassano, A., Rastogi, N.K., Eds.; Woodhead Publishing: Oxford, UK, 2015; pp. 43–82. ISBN 978-1-78242-121-4. [Google Scholar]
  214. Matsuyama, H.; Teramoto, M.; Uesaka, T. Membrane formation and structure development by dry-cast process. J. Memb. Sci. 1997, 135, 271–288. [Google Scholar] [CrossRef]
  215. Zhang, Z.; Kong, X.-Y.; Xiao, K.; Liu, Q.; Xie, G.; Li, P.; Ma, J.; Tian, Y.; Wen, L.; Jiang, L. Engineered Asymmetric Heterogeneous Membrane: A Concentration-Gradient-Driven Energy Harvesting Device. J. Am. Chem. Soc. 2015, 137, 14765–14772. [Google Scholar] [CrossRef] [PubMed]
  216. Schacher, F.H.; Rupar, P.A.; Manners, I. Functional Block Copolymers: Nanostructured Materials with Emerging Applications. Angew. Chemie Int. Ed. 2012, 51, 7898–7921. [Google Scholar] [CrossRef] [PubMed]
  217. Shi, W.; Hamilton, A.L.; Delaney, K.T.; Fredrickson, G.H.; Kramer, E.J.; Ntaras, C.; Avgeropoulos, A.; Lynd, N.A. Creating Extremely Asymmetric Lamellar Structures via Fluctuation-Assisted Unbinding of Miktoarm Star Block Copolymer Alloys. J. Am. Chem. Soc. 2015, 137, 6160–6163. [Google Scholar] [CrossRef] [PubMed]
  218. Barman, S.R.; Roy, U.; Das, P.; Mukhopadhayay, A. Chapter 6—Membrane processes for removal of polyaromatic hydrocarbons from wastewater. In Advances in Green and Sustainable Chemistry; Sharma, S., Ed.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 189–207. ISBN 978-0-12-817742-6. [Google Scholar]
  219. Kamide, K.; Manabe, S. Mechanism of Permselectivity of Porous Polymeric Membranes in Ultrafiltration Process. Polym. J. 1981, 13, 459–479. [Google Scholar] [CrossRef]
  220. Kolya, H.; Singh, V.K.; Kang, C.-W. Polymeric Membranes and Hybrid Techniques for Water Purification Applications BT—Polymer-Based Advanced Functional Materials for Energy and Environmental Applications; Subramani, N.K., Nataraj, S.K., Patel, C., Shivanna, S., Eds.; Springer: Singapore, 2022; pp. 75–91. ISBN 978-981-16-8755-6. [Google Scholar]
  221. Baig, N.; Matin, A.; Faizan, M.; Anand, D.; Ahmad, I.; Khan, S.A. Antifouling low-pressure highly permeable single step produced loose nanofiltration polysulfone membrane for efficient Erichrome Black T/divalent salts fractionation. J. Environ. Chem. Eng. 2022, 10, 108166. [Google Scholar] [CrossRef]
  222. Guo, S.; Wan, Y.; Chen, X.; Luo, J. Loose nanofiltration membrane custom-tailored for resource recovery. Chem. Eng. J. 2021, 409, 127376. [Google Scholar] [CrossRef]
  223. Feng, X.; Peng, D.; Zhu, J.; Wang, Y.; Zhang, Y. Recent advances of loose nanofiltration membranes for dye/salt separation. Sep. Purif. Technol. 2022, 285, 120228. [Google Scholar] [CrossRef]
  224. Hakami, M.W.; Alkhudhiri, A.; Al-Batty, S.; Zacharof, M.-P.; Maddy, J.; Hilal, N. Ceramic Microfiltration Membranes in Wastewater Treatment: Filtration Behavior, Fouling and Prevention. Membranes 2020, 10, 248. [Google Scholar] [CrossRef]
  225. Kotobuki, M.; Gu, Q.; Zhang, L.; Wang, J. Ceramic-Polymer Composite Membranes for Water and Wastewater Treatment: Bridging the Big Gap between Ceramics and Polymers. Molecules 2021, 26, 3331. [Google Scholar] [CrossRef]
  226. Barhoum, A.; Deshmukh, K.; García-Betancourt, M.-L.; Alibakhshi, S.; Mousavi, S.M.; Meftahi, A.; Sabery, M.S.K.; Samyn, P. Nanocelluloses as sustainable membrane materials for separation and filtration technologies: Principles, opportunities, and challenges. Carbohydr. Polym. 2023, 317, 121057. [Google Scholar] [CrossRef]
  227. Garcia-Ivars, J.; Alcaina-Miranda, M.-I.; Iborra-Clar, M.-I.; Mendoza-Roca, J.-A.; Pastor-Alcañiz, L. Enhancement in hydrophilicity of different polymer phase-inversion ultrafiltration membranes by introducing PEG/Al2O3 nanoparticles. Sep. Purif. Technol. 2014, 128, 45–57. [Google Scholar] [CrossRef]
  228. Cai, C.; Sun, W.; He, S.; Zhang, Y.; Wang, X. Ceramic membrane fouling mechanisms and control for water treatment. Front. Environ. Sci. Eng. 2023, 17, 126. [Google Scholar] [CrossRef]
  229. Mpala, T.J.; Etale, A.; Richards, H.; Nthunya, L.N. Biofouling phenomena in membrane distillation: Mechanisms and mitigation strategies. Environ. Sci. Adv. 2023, 2, 39–54. [Google Scholar] [CrossRef]
  230. Manjarrez Nevárez, L.; Ballinas Casarrubias, L.; Canto, O.S.; Celzard, A.; Fierro, V.; Ibarra Gómez, R.; González Sánchez, G. Biopolymers-based nanocomposites: Membranes from propionated lignin and cellulose for water purification. Carbohydr. Polym. 2011, 86, 732–741. [Google Scholar] [CrossRef]
  231. Thakur, S. Gum Based Green Nanocomposites and Their Applications BT—Green-Based Nanocomposite Materials and Applications; Avalos Belmontes, F., González, F.J., López-Manchado, M.Á., Eds.; Springer International Publishing: Cham, Switzerland, 2023; pp. 295–315. ISBN 978-3-031-18428-4. [Google Scholar]
  232. Mukherjee, M.; De, S. Antibacterial polymeric membranes: A short review. Environ. Sci. Water Res. Technol. 2018, 4, 1078–1104. [Google Scholar] [CrossRef]
  233. Cheng, M.; Huang, L.; Wang, Y.; Zhao, Y.; Tang, J.; Wang, Y.; Zhang, Y.; Hedayati, M.; Kipper, M.J.; Wickramasinghe, S.R. Synthesis of graphene oxide/polyacrylamide composite membranes for organic dyes/water separation in water purification. J. Mater. Sci. 2019, 54, 252–264. [Google Scholar] [CrossRef]
  234. Zhang, M.; Cui, J.; Lu, T.; Tang, G.; Wu, S.; Ma, W.; Huang, C. Robust, functionalized reduced graphene-based nanofibrous membrane for contaminated water purification. Chem. Eng. J. 2021, 404, 126347. [Google Scholar] [CrossRef]
  235. Azizi-Lalabadi, M.; Jafari, S.M. Bio-nanocomposites of graphene with biopolymers; fabrication, properties, and applications. Adv. Colloid Interface Sci. 2021, 292, 102416. [Google Scholar] [CrossRef] [PubMed]
  236. Zubair, M.; Ullah, A. Chapter 14—Biopolymers in Environmental Applications: Industrial Wastewater Treatment; Thomas, S., Gopi, S., Amalraj, A.B.T.-B., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 331–349. ISBN 978-0-12-819240-5. [Google Scholar]
  237. Suyambulingam, I.; Gangadhar, L.; Sana, S.S.; Divakaran, D.; Siengchin, S.; Kurup, L.A.; Iyyadurai, J.; Albert Bernad Noble, K.E. Chitosan Biopolymer and Its Nanocomposites: Emerging Material as Adsorbent in Wastewater Treatment. Adv. Mater. Sci. Eng. 2023, 2023, 9387016. [Google Scholar] [CrossRef]
  238. Rouf, T.B.; Kokini, J.L. Biodegradable biopolymer–graphene nanocomposites. J. Mater. Sci. 2016, 51, 9915–9945. [Google Scholar] [CrossRef]
  239. Leudjo Taka, A.; Klink, M.; Yangkou Mbianda, X.; Mtunzi, F.; Bobby Naidoo, E. Carbon Nanotubes Reinforced Polymeric Hybrid Materials for Water Purification BT—Nanohybrid Materials for Water Purification; Swain, S.K., Ed.; Springer Nature: Singapore, 2022; pp. 197–223. ISBN 978-981-19-2332-6. [Google Scholar]
  240. Mostafavi, A.H.; Mishra, A.K.; Gallucci, F.; Kim, J.H.; Ulbricht, M.; Coclite, A.M.; Hosseini, S.S. Advances in surface modification and functionalization for tailoring the characteristics of thin films and membranes via chemical vapor deposition techniques. J. Appl. Polym. Sci. 2023, 140, e53720. [Google Scholar] [CrossRef]
  241. Huang, Z.; Chen, J.; Li, R.; Shi, S.; Gong, Y.-K. Mussel adhesion and cell membrane antifouling mimetic strategies for durable fouling-resistant coating. Prog. Org. Coat. 2023, 182, 107636. [Google Scholar] [CrossRef]
  242. Dongre, R.S.; Sadasivuni, K.K.; Deshmukh, K.; Mehta, A.; Basu, S.; Meshram, J.S.; Al-Maadeed, M.A.A.; Karim, A. Natural polymer based composite membranes for water purification: A review. Polym. Technol. Mater. 2019, 58, 1295–1310. [Google Scholar] [CrossRef]
  243. Bessbousse, H.; Verchère, J.-F.; Lebrun, L. Characterisation of metal-complexing membranes prepared by the semi-interpenetrating polymer networks technique. Application to the removal of heavy metal ions from aqueous solutions. Chem. Eng. J. 2012, 187, 16–28. [Google Scholar] [CrossRef]
  244. Dutta, K.; De, S. Aromatic conjugated polymers for removal of heavy metal ions from wastewater: A short review. Environ. Sci. Water Res. Technol. 2017, 3, 793–805. [Google Scholar] [CrossRef]
  245. Bessbousse, H.; Rhlalou, T.; Verchère, J.-F.; Lebrun, L. Mercury removal from wastewater using a poly(vinylalcohol)/poly(vinylimidazole) complexing membrane. Chem. Eng. J. 2010, 164, 37–48. [Google Scholar] [CrossRef]
  246. Refaat Alawady, A.; Ali Alshahrani, A.; Ali Aouak, T.; Mohamed Alandis, N. Polysulfone membranes with CNTs/Chitosan biopolymer nanocomposite as selective layer for remarkable heavy metal ions rejection capacity. Chem. Eng. J. 2020, 388, 124267. [Google Scholar] [CrossRef]
  247. Saucedo-Rivalcoba, V.; Martínez-Hernández, A.L.; Martínez-Barrera, G.; Velasco-Santos, C.; Rivera-Armenta, J.L.; Castaño, V.M. Removal of Hexavalent Chromium from Water by Polyurethane–Keratin Hybrid Membranes. Water Air Soil Pollut. 2011, 218, 557–571. [Google Scholar] [CrossRef]
  248. Wu, S.; Li, F.; Wang, H.; Fu, L.; Zhang, B.; Li, G. Effects of poly (vinyl alcohol) (PVA) content on preparation of novel thiol-functionalized mesoporous PVA/SiO2 composite nanofiber membranes and their application for adsorption of heavy metal ions from aqueous solution. Polymer 2010, 51, 6203–6211. [Google Scholar] [CrossRef]
  249. Terbish, N.; Popuri, S.R.; Lee, C.-H. Improved performance of organic–inorganic nanocomposite membrane for bioelectricity generation and wastewater treatment in microbial fuel cells. Fuel 2023, 332, 126167. [Google Scholar] [CrossRef]
  250. Barman, S.R.; Banerjee, P.; Mukhopadhayay, A.; Das, P. Biopolymer linked activated carbon-nano-bentonite composite membrane for efficient elimination of PAH mixture from aqueous solutions. Biomass Convers. Biorefinery 2022, 1–15. [Google Scholar] [CrossRef]
  251. Gajera, R.; Patel, R.V.; Yadav, A.; Labhasetwar, P.K. Adsorption of cationic and anionic dyes on photocatalytic flyash/TiO2 modified chitosan biopolymer composite. J. Water Process Eng. 2022, 49, 102993. [Google Scholar] [CrossRef]
  252. Kour, G.; Majhi, P.K.; Bharti, A.; Kothari, R.; Jain, A.; Singh, A.; Tyagi, V.V.; Pathania, D. Biopolymer-Based Nanocomposites and Water Treatment: A Global Outlook. In Biorenewable Nanocomposite Materials, Vol. 2: Desalination and Wastewater Remediation; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2022; Volume 1411, pp. 2–25. ISBN 9780841297807. [Google Scholar]
  253. Osman, A.I.; El-Monaem, E.M.A.; Elgarahy, A.M.; Aniagor, C.O.; Hosny, M.; Farghali, M.; Rashad, E.; Ejimofor, M.I.; López-Maldonado, E.A.; Ihara, I.; et al. Methods to prepare biosorbents and magnetic sorbents for water treatment: A review. Environ. Chem. Lett. 2023, 21, 2337–2398. [Google Scholar] [CrossRef]
  254. Nure, J.F.; Nkambule, T.T.I. The recent advances in adsorption and membrane separation and their hybrid technologies for micropollutants removal from wastewater. J. Ind. Eng. Chem. 2023, 126, 92–114. [Google Scholar] [CrossRef]
Figure 4. Schematic presentation of polymer nanocomposites synthesis process (a) solution casting [146]. Copyright 2017; reproduced with permission from the authors, MDPI, Basel. and (b) In situ polymerization method.
Figure 4. Schematic presentation of polymer nanocomposites synthesis process (a) solution casting [146]. Copyright 2017; reproduced with permission from the authors, MDPI, Basel. and (b) In situ polymerization method.
Polymers 15 03421 g004
Figure 5. Schematic of (a) melt mixing method [148]. Copyright 2020; reproduced with permission from Springer Nature Switzerland AG. and (b) single electrospinning and coaxial electrospinning process [149]. Copyright 2021; reproduced with permission from the authors.
Figure 5. Schematic of (a) melt mixing method [148]. Copyright 2020; reproduced with permission from Springer Nature Switzerland AG. and (b) single electrospinning and coaxial electrospinning process [149]. Copyright 2021; reproduced with permission from the authors.
Polymers 15 03421 g005
Figure 6. Spectra with their background corrections and deconvolution for crystallinity study (a) XRD, (b) ATR-FTIR, and (c) Raman [94]. Copyright 2022; reproduced with permission from Elsevier B.V. All rights reserved.
Figure 6. Spectra with their background corrections and deconvolution for crystallinity study (a) XRD, (b) ATR-FTIR, and (c) Raman [94]. Copyright 2022; reproduced with permission from Elsevier B.V. All rights reserved.
Polymers 15 03421 g006
Figure 7. Micrographs of AgNPs/rGO nanocomposite, (a,b) FESEM, and (c,d) TEM, and (e) SAED pattern image [129]. Copyright 2019; reproduced with permission from Elsevier Ltd. All rights reserved.
Figure 7. Micrographs of AgNPs/rGO nanocomposite, (a,b) FESEM, and (c,d) TEM, and (e) SAED pattern image [129]. Copyright 2019; reproduced with permission from Elsevier Ltd. All rights reserved.
Polymers 15 03421 g007
Figure 8. (a) TGA curves for PMMA, (b) DTGA curves for nanocomposites in acetonitrile medium and (c,d) in ethanol medium [151]. Copyright 2013; reproduced with permission from the authors.
Figure 8. (a) TGA curves for PMMA, (b) DTGA curves for nanocomposites in acetonitrile medium and (c,d) in ethanol medium [151]. Copyright 2013; reproduced with permission from the authors.
Polymers 15 03421 g008
Figure 9. The XPS C1s spectra of GO, rGO, and rGO/PVAc-PDDA are shown in (ac). The XPS O1s spectra of GO, rGO, and rGO/PVAc-PDDA are shown in (df), while the XPS N1s spectra of GO, rGO, and rGO/PVAc-PDDA are presented in (gi) [152]. Copyright 2021; reproduced with permission from Elsevier B.V. All rights reserved.
Figure 9. The XPS C1s spectra of GO, rGO, and rGO/PVAc-PDDA are shown in (ac). The XPS O1s spectra of GO, rGO, and rGO/PVAc-PDDA are shown in (df), while the XPS N1s spectra of GO, rGO, and rGO/PVAc-PDDA are presented in (gi) [152]. Copyright 2021; reproduced with permission from Elsevier B.V. All rights reserved.
Polymers 15 03421 g009
Figure 10. SEM images of biopolymer sponge (a,c) and composite sponge (b,d). UV–Vis spectra of different dye solutions before and after adsorption: RB (e) and CR (f). Dye-adsorbed sponge adsorption studies: adsorptive times (g) and dye concentrations (h) [162]. Copyright 2022; reproduced with permission from the Chinese Materials Research Society. Published by Elsevier B.V. (i) SEM images of laterite-clay-based geopolymer and Grewia biopolymer samples [163]. Copyright 2022; reproduced with permission from the Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
Figure 10. SEM images of biopolymer sponge (a,c) and composite sponge (b,d). UV–Vis spectra of different dye solutions before and after adsorption: RB (e) and CR (f). Dye-adsorbed sponge adsorption studies: adsorptive times (g) and dye concentrations (h) [162]. Copyright 2022; reproduced with permission from the Chinese Materials Research Society. Published by Elsevier B.V. (i) SEM images of laterite-clay-based geopolymer and Grewia biopolymer samples [163]. Copyright 2022; reproduced with permission from the Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
Polymers 15 03421 g010
Figure 11. (a) Synthesis of the composite magnetic–activated carbon composite, (b) removing of RhB and MO from water, (cf) showed UV-Vis spectra of RhB and MO adsorption using composite, (g) adsorption mechanism of RhB and (h) adsorption mechanism of MO [177]. Copyright 2019; reproduced with permission from Elsevier B.V. All rights reserved.
Figure 11. (a) Synthesis of the composite magnetic–activated carbon composite, (b) removing of RhB and MO from water, (cf) showed UV-Vis spectra of RhB and MO adsorption using composite, (g) adsorption mechanism of RhB and (h) adsorption mechanism of MO [177]. Copyright 2019; reproduced with permission from Elsevier B.V. All rights reserved.
Polymers 15 03421 g011
Figure 12. (A) Synthetic process of chitosan-based magnetic nanoparticles composite; (B) surface morphology change: (a–d) the optical microscope images, (e–l) SEM photographs, (a,e,i) chitosan particles, (b,f,J) magnetic chitosan particles at 3.75 kV, (c,g,k) magnetic chitosan particles at 6 kV, (d,h,l) iron oxide-loaded chitosan particles at 9 kV. Scale bar: (a–d) 200 μm, (e–h) 1 mm, and (i–l) 1 μm. All particles were prepared using 10 wt% NaOH solution [179]. Copyright 2013; reproduced with permission from the Royal Society of Chemistry. (C) The magnetic separation process of MCS composites and ion solutions and (D) regeneration of ion-imprinted CS composites [180]. Copyright 2021; reproduced with permission from The Author(s), under exclusive license to Springer-Verlag GmbH Germany, part of Springer Nature.
Figure 12. (A) Synthetic process of chitosan-based magnetic nanoparticles composite; (B) surface morphology change: (a–d) the optical microscope images, (e–l) SEM photographs, (a,e,i) chitosan particles, (b,f,J) magnetic chitosan particles at 3.75 kV, (c,g,k) magnetic chitosan particles at 6 kV, (d,h,l) iron oxide-loaded chitosan particles at 9 kV. Scale bar: (a–d) 200 μm, (e–h) 1 mm, and (i–l) 1 μm. All particles were prepared using 10 wt% NaOH solution [179]. Copyright 2013; reproduced with permission from the Royal Society of Chemistry. (C) The magnetic separation process of MCS composites and ion solutions and (D) regeneration of ion-imprinted CS composites [180]. Copyright 2021; reproduced with permission from The Author(s), under exclusive license to Springer-Verlag GmbH Germany, part of Springer Nature.
Polymers 15 03421 g012
Figure 13. Showcases an engineered asymmetric heterogeneous membrane design. The top layer consists of a pH-responsive porous block copolymer (BCP) membrane (a), while the bottom layer is a pH-responsive porous polyethylene terephthalate (PET) membrane with conical nanochannels. SEM images of the base and tip of the conical nanochannels are displayed in (b,c), respectively, demonstrating a pore density of 107 cm–2. The AFM image in (d) depicts the BCP membrane atop the PET membrane, with a corresponding histogram showcasing the pore size distribution (pore density: 1011 cm–2) and a Gaussian fit (e). The scale bars in all images represent 100 nm [215]. Copyright 2015; reproduced with permission from the American Chemical Society.
Figure 13. Showcases an engineered asymmetric heterogeneous membrane design. The top layer consists of a pH-responsive porous block copolymer (BCP) membrane (a), while the bottom layer is a pH-responsive porous polyethylene terephthalate (PET) membrane with conical nanochannels. SEM images of the base and tip of the conical nanochannels are displayed in (b,c), respectively, demonstrating a pore density of 107 cm–2. The AFM image in (d) depicts the BCP membrane atop the PET membrane, with a corresponding histogram showcasing the pore size distribution (pore density: 1011 cm–2) and a Gaussian fit (e). The scale bars in all images represent 100 nm [215]. Copyright 2015; reproduced with permission from the American Chemical Society.
Polymers 15 03421 g013
Figure 14. A schematic of membrane filtration process.
Figure 14. A schematic of membrane filtration process.
Polymers 15 03421 g014
Table 3. Iron-oxide-biopolymer-based polymer nanocomposites are used for water purification.
Table 3. Iron-oxide-biopolymer-based polymer nanocomposites are used for water purification.
Sl. NoAdsorbent/FlocculentPollutantResults (% or Qmax mg/g)Ref.
1.Fe3O4-Starch-g-PVAMB
MG
621 mg/g
567 mg/g
[187]
2.Starch-coated Fe3O4 NanoparticlesTextile dye Optilan Blue74.05 mg/g[188]
3.GO–Fe3O4 hybrid compositeMB
Natural red
167.2 mg/g
171.3 mg/g
[189]
4.Chitosan/Fe–hydroxyapatite beadsMB1324 mg/g[190]
5.Fe3O4/β-cyclodextrin/GOMG740.7 mg/g[191]
6.Magnetic chitosan nano-adsorbentCu
Cd
Zn
99.9%
95.0%
81.7%
[192]
7.Magnetic mesoporous silica–chitosan compositeHg (II)437.8 mg/g[193]
8.Carboxymethyl-chitosan-based magneticCiprofloxacin527.9 mg/g[194]
9.Anionic-polyacrylamide-modified chitosan magnetic compositeMB1044.1[195]
10.Manganese ferrite nanoparticles covered with carboxymethyl starchPb (II)34–64 mg/g[196]
11.Sodium alginate/CMC/Fe3O4Mn (II), Pb (II), Ni (II)71.83, 89.49, 105.93 mg/g[197]
12.β-cyclodextrin-Fe3O4/MWCNTNi (II)103 mg/g[198]
13.Fe3O4/wood biocharAcid orange and Cr (VI)110.27, 80.96 mg/g[199]
14.GO/Fe3O4/glucoseU (VI)390.70 mg/g[200]
15.Cellulose/polyethyleneimine-Fe3O4Hg (II)247.51 mg/g[201]
16.Chitosan activated carbon-Fe3O4 compositesCr (VI)99.8%[174]
17.Magnetic chitosan−GO nanocompositesMB95.31 mg/g, pH 5.3, 303K[202]
18.GO-Fe3O4-chitosan compositesMB98.0%[183]
19.Chitosan/poly (methacrylic acid)/Fe3O4 -GO compositeMB2478 mg/g[185]
20.Lignin-grafted magnetic nanocompositeFluoroquinolone108.2 mg/g[203]
21.Lignin magnetic compositesMalachite green456.62 mg/g[204]
22.Cellulose nanofiber–GO magnetic compositeMethylene blue83.53%[205]
23.Magnetic-cellulose-based ionic liquidCongo red and methyl blue1299.3 and 1068.1 mg/g[206]
24.Fe3O4–chitosan compositeMethyl orange638.6 mg/g[207]
25.Dragon fruit biopolymer–CoFe2O4Ni (II)88%[208]
26.Inulin–Fe3O4 nanocompositeCo (II), Cu (II), Hg (II)152.5, 167.7, 198.0 mg/g[209]
27.Fe3O4–chitosan–PAMm compositeFood dye359.71 mg/g[210]
Table 4. Recent reports of water purification using membrane filtration.
Table 4. Recent reports of water purification using membrane filtration.
Sl NoMembranePollutantsOptimum Rejection (%)Pressure (bar)pHRef.
1.CNTs/chitosan biopolymer nanocompositeCu(II), Ni(II), Pb (II), Cd(II), and Co(II) ions.71–92.2223.0[246]
2.Lignin–cellulose citrateAnions
Cations
12–42
27–54
-
>7.0[230]
3.Polylactic acid-polybutylene succinate-polypropylene carbonate-polyhydroxybutyrate.Oil–grease
TDS
98.6
89.15
1
4

-
[150]
4.Poly (urethane)/keratin biofiberCr (VI)38.0
11.0
0.7>8–<4
[247]
5.PVA and poly (vinylimidazole)Hg (II)99.43.02.5[245]
6.Polyvinyl alcohol/SiO2 compositesCu (II)93.1-5–6[248]
7.Nanoclay montmorillonite-ChitosanCOD65.7--[249]
8.Chitosan-linked activated carbon–nano-bentonitePolyaromatic hydrocarbons99.3-6.0[250]
9.TiO2 loaded fly ash chitosan compositeCongo red
Methylene blue
98.0
55.7
-1.8[251]
10.Polydopamine-functionalized polysulfone membraneEriochrome
Black T
99.91–3-[221]
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

Kolya, H.; Kang, C.-W. Next-Generation Water Treatment: Exploring the Potential of Biopolymer-Based Nanocomposites in Adsorption and Membrane Filtration. Polymers 2023, 15, 3421. https://doi.org/10.3390/polym15163421

AMA Style

Kolya H, Kang C-W. Next-Generation Water Treatment: Exploring the Potential of Biopolymer-Based Nanocomposites in Adsorption and Membrane Filtration. Polymers. 2023; 15(16):3421. https://doi.org/10.3390/polym15163421

Chicago/Turabian Style

Kolya, Haradhan, and Chun-Won Kang. 2023. "Next-Generation Water Treatment: Exploring the Potential of Biopolymer-Based Nanocomposites in Adsorption and Membrane Filtration" Polymers 15, no. 16: 3421. https://doi.org/10.3390/polym15163421

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