Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Emerging environmentally friendly bio-based nanocomposites for the efficient removal of dyes and micropollutants from wastewater by adsorption: a comprehensive review

Wafa Al-Gethamia, Muhammad Azam Qamar*b, Mohammad Shariq*c, Abdel-Nasser M. A. Alaghazd, Ahmad Farhane, Ashwaq A. Areshif and M. Hisham Alnasirg
aChemistry Department, Faculty of Science, Taif University, Al-Hawiah, PO Box 11099, Taif City, Saudi Arabia
bDepartment of Chemistry, School of Science, University of Management and Technology, Lahore 54770, Pakistan. E-mail: qamariub@gmail.com
cDepartment of Physics, College of Science, Jazan University, Jazan 45142, Saudi Arabia
dDepartment of Chemistry, College of Science, Jazan University, Jazan 45142, Saudi Arabia
eDepartment of Chemistry, University of Agriculture Faisalabad, Faisalabad 38040, Pakistan
fSamtah General Hospital, Ministry of Health, Jazan, 86735, Saudi Arabia
gDepartment of Physics, RIPHAH International University, Islamabad, 44000, Pakistan

Received 23rd September 2023 , Accepted 19th December 2023

First published on 17th January 2024


Abstract

Water scarcity will worsen due to population growth, urbanization, and climate change. Addressing this issue requires developing energy-efficient and cost-effective water purification technologies. One approach is to use biomass to make bio-based materials (BBMs) with valuable attributes. This aligns with the goal of environmental conservation and waste management. Furthermore, the use of biomass is advantageous because it is readily available, economical, and has minimal secondary environmental impact. Biomass materials are ideal for water purification because they are abundant and contain important functional groups like hydroxyl, carboxyl, and amino groups. Functional groups are important for modifying and absorbing contaminants in water. Single-sourced biomass has limitations such as weak mechanical strength, limited adsorption capacity, and chemical instability. Investing in research and development is crucial for the development of efficient methods to produce BBMs and establish suitable water purification application models. This review covers BBM production, modification, functionalization, and their applications in wastewater treatment. These applications include oil–water separation, membrane filtration, micropollutant removal, and organic pollutant elimination. This review explores the production processes and properties of BBMs from biopolymers, highlighting their potential for water treatment applications. Furthermore, this review discusses the future prospects and challenges of developing BBMs for water treatment and usage. Finally, this review highlights the importance of BBMs in solving water purification challenges and encourages innovative solutions in this field.


1. Introduction

Rapid economic development and fast-growing industrialization have resulted in an increase in the volume of contaminants in the water supply from various processes, including electroplating, leather tanning, mining, paper, batteries, and pesticides.1 Because of this, there is an urgent need for specialized materials and technology for both water recycling and waste water treatment. The ideal properties of the material would be a high adsorption capacity, reusability, low cost, and the ability to adjust the porosity. With the ability to optimize different qualities, including mechanical strength, surface area, porosity, hydrophobicity, and dispersibility, bio-based materials are one of the potential possibilities made from diverse polymeric materials for water purification. Dyes, pharmaceutical wastes, heavy metals, and other pollutants have been accumulated in water sources in the last several years due to increased industrialization.2 For example, the presence of lead (Pb(II)) ions in ponceau red, a synthetic red azo dye made from tar that causes hypersensitive allergic responses and led to the growth of colon, stomach, and rectal associated malignancies, has led to its classification as a possible human urine carcinogen. High concentrations of suspended particles, dyes, and chemicals, as well as a high chemical oxygen demand (COD), may be found in water bodies polluted by dye effluents. The textile industry uses more than 8000 chemicals in the development of 3600 dyes.3 This is a severe threat to human health, as ingestion of dyes can have extremely carcinogenic or mutagenic effects.4 Thus, before their disposal in any water bodies, the dyes and chemicals must be treated using extremely efficient adsorbent materials. Similarly, exposure to oil spills may cause irritation of the skin and eyes as well as neurological and respiratory issues. Although the long-term consequences of oil spills are unknown, it is nevertheless necessary to clean them up as soon as possible.5

As a result of the high demand for potable and industrial water, wastewater treatment has become an essential process in recent times. Complexation,6 precipitation,7 ion exchange,8 advanced oxidation processes (AOPs),9 membrane filtration,10 reverse osmosis,11 activated carbon adsorption,9 ultraviolet (UV) photolysis/photocatalysis,12 and electrodialysis13 are all examples of conventional technologies used to treat wastewater. Clean water can be produced sustainably via wastewater processing and desalination.14 Current research studies are focused on enhancing the existing wastewater purification systems, and creating new energy-efficient and cost-effective technologies. Protein nanofibrils,15,16 covalent organic frameworks (COFs),17 metal–organic frameworks (MOFs),18,19 graphene,20 carbon nanotubes,21 fluorous oligoamide nanorings and artificial water channels (AWCs)22,23 are just some of the advanced materials used in water treatment. These materials, however, have significant production costs and complicated synthesis processes.

Owing to their abundance in the environment and use as a bio-source in the fabrication of bio-based nanofibers, biopolymers have recently attracted significant attention. They have unique properties, including being hydrophilic to avoid fouling, being mechanically and chemically stable, and being easily modifiable with chemicals. Biopolymers have been studied for their possible use in removing a wide variety of contaminants from water.24 There has been a lot of research into various polymers for this purpose, including cellulose, chitosan, hemicellulose, polylactic acid, cellulosic biopolymer, and alginic acid.25 Due to their biocompatibility, disinfection capability, non-toxicity, and adsorption behavior, chitosan and cellulose have received the greatest research attention as biopolymers for use in water purification processes, like membrane filtration, micropollutant removal, degradation of dyes and oil–water separation.26

BBMs offer many significant benefits over more traditional water treatment materials, such as metal oxides, activated carbon, clays, metal sulfides, carbon-based nanomaterials, zeolites and mesoporous silica.27,28 BBMs have the following vital benefits: (i) naturally occurring in large amounts, (ii) able to selectively remove trace impurities like heavy metals, (iii) readily biodegradable, and (iv) inexpensive and efficient in terms of both power consumption and heat production. These qualities make them excellent candidates for use as water filtration materials and adsorbents.29 Although bio-based materials have many benefits, they have a few problems that need to be addressed. The cost of BBMs cannot be brought down to a reasonable level for high-volume manufacturing because of the insufficient efficiency and effective exploitation of widely accessible inexpensive feedstocks, i.e., lignocellulosic bio-based materials. The inherent features of the resulting BBMs need to be tailored to the target chemicals. However, suitable production and modification procedures have not yet been discovered.30 The nature of the contaminants, kinetics, and removal mechanism, the green synthesis techniques of BBMs and their efficient and effective applications, and bio-material recyclability are all important concepts and criteria for studying these materials.31–34

Bio-based nanomaterials like chitosan, cellulose, and other plant-based nanofibers are discussed in this study as potential solutions for the remediation of various pollutants such as dyes, heavy metals, pharmaceutical toxins and organic pollutants. Micropollutants may be removed using a variety of traditional techniques, i.e., adsorption, membrane filtering, coagulation, ion exchange, flocculation, and chemical precipitation. Electrospinning, chemical modification and phase inversion are one of the many processes for generating natural nanofibers, and it has the benefits of being cost-effective, highly reproducible, and efficient.35 Functionalization of plant-based materials with other groups results in enhanced properties, which is comprehensively discussed. Therefore, we conduct an in-depth evaluation of the many bio-based materials currently on the market, as well as bio-based nanofibers made by various manufacturing techniques, surface changes, and adsorption behavior for use in water purification. In this review, we have provided an overview of the capabilities and structural features of BBMs. In order to remove diverse contaminants from water systems, this article discusses the construction and modification procedures of BBMs to permit distinct chemical and physical features. The most recent developments in the design of BBMs for use in water purification are highlighted in this article. The reason for using BBMs is comprehensively evaluated. The impact of various functionalizations on the water purification process is also the main focus of this review. In conclusion, we review the key ideas and obstacles encountered while working to improve the BBMs with superior water treatment capabilities. Our study not only adds to the growing body of knowledge on BBMs, but also serves as a comprehensive theoretical foundation upon which to build more sophisticated BBMs.

2. Polysaccharides as a precursor of biomaterials

The harmful effects of using fossil fuels and petrochemicals are increasingly becoming a growing concern. It is predicted that by 2050, synthetic polymers will be manufactured using 20% of all fossil fuels.27,36,37 Therefore, it is critical to discover replacements for petroleum-based goods, moving from hydrocarbon energy to a more desirable renewable fuel. Fossil-based materials, often acquired through environmentally deteriorating processes, might be replaced with bio-based feedstock. Biopolymers may be made from renewable resources in three different ways:

(1) Through fractionation and extraction, as with cellulose, starch, lignin, protein, chitin, and sodium alginate.38

(2) Polymerization of bio-monomers derived from renewable biomass sources, as with polylactic acid.39

(3) Through direct production by microorganisms, as with polyhydroxyalkanoates (PHAs).40

Different hydroxyl, carboxyl, amino, acetal, and ester functional groups in these biopolymers allow for various chemical and physical properties in bio-based materials (Fig. 1).41 BBMs have shown promising commercial applications in water treatment in recent years after being developed into a wide range of products, including fibers, membranes, hydrogels, foams, particles, and powders, via various modification and fabrication routes.


image file: d3ra06501d-f1.tif
Fig. 1 Fundamental physicochemical characteristics of bio-based materials.

2.1 Lignin

Naturally occurring lignin is a complicated heterogeneous bio-polymer made up of 3 different kinds of monomeric phenolic substructures, syringyl (S) units, guaiacyl (G) units, and p-hydroxyphenyl (H) units, that originate from coniferyl alcohol, coumaryl alcohol, sinapyle alcohol, and monolignols.42 Several different forms of ether linkages and carbon–carbon bonds, including biphenyl (5-5′), aryl ether (−O-4′), phenylcoumaran (−5′), and resinol, are used to randomly join the three subunits during biosynthesis. S and G units, together with less H units, make up the bulk of hardwood lignin, whereas G units make up the vast majority (92–95%) of softwood lignin. All three components are grass lignin. The chemical and physical properties of lignin, i.e., its reactivity, functionality, and hydrophilicity, are analyzed by the presence of different functional molecules within the macromolecules of lignin, including phenolic and aliphatic hydroxyls, carbonyl, methoxyl groups, and carboxylic.43–46

As an alternative to costly commercial sorbents, lignin-based biosorbents have been recently created using crosslinking, grafting, copolymerization, and hybridization modification procedures. Gao et al.47 devised and produced a variety of lignin-based adsorbents that used lignin as grafted and essential poly(acrylic acid) (PAA) as the tentacles. Better adsorption selectivity correlates with a larger PAA branched chain distribution density, as predicted by a binary non-linear model based on Pearson correlation analysis and phenomenological theory. Theoretical and fundamental advice for creating and designing innovative high-efficiency lignin-based biosorbents is provided, with improved structural connections of graft-modified surface adsorbents in binary pollutants. Using cyanuric chloride as a chemoselective crosslinker, lignin was chemically bonded to amine-functionalized magnetic nanoparticles.48 Good selectivity of Pb(II), regeneration, and outstanding stability characterizes the lignin-based adsorbent synthesized using a temperature-controlled synthetic approach. Lignin-based magnetic materials have the ability to be used as cheap and ecological adsorbents in water decontamination due to their efficient removal and quick recovery capabilities. A nano-trap made of a bendable lignin-based material was developed to sequester metal ions (M+) and inhibit bacteria growth. The lignin-based nano-trap (LBNT) was effective because of the robust M+ binding on the surface of the functionalized lignin. Interestingly, Ag loaded with LBNTs was employed as a novel antimicrobial material, and it showed extraordinary efficacy against E. coli and S. aureus (99.68% and 99.76%, respectively).48,49 To get rid of the As(V) in the water, lignin was treated with triethylenetetramine (TETA) using the Mannich procedure. The As(V) adsorption capacity was best for activated lignin at 25 °C and 9.0 pH, with As(V) > Cr(VI) > P being the order of selectivity (V). Biosorbents made from lignin show promise for widespread use in the industry.50,51

2.2. Cellulose

Cellulose is primarily made from lignocellulosic biomass, which is the most common and widespread source of cellulose in nature. The linear biopolymer comprises –D-glucopyranose units connected by –1,4 glucosidic linkages. The cellulose degree of polymerization (DP) ranges from around 100 to over 10[thin space (1/6-em)]000, depending on the plant type.52 An anhydroglucose cellulose unit has three hydroxyl groups (–OH), one each at carbons 2, 3, and 6. These OH groups establish intra- and intermolecular hydrogen bonds inside the macromolecules, giving them their stiffness and crystallinity. For cellulose to be effective as a biosorbent, it must be in its most basic form, both small and porous, with plenty of active binding sites.53 Because of the abundance of OH groups, cellulose has been the subject of extensive study.54–58

There are two primary classes of cellulose modification techniques. Sulfation,59 etherification,60 carbanilation,61 silylation,62 esterification,63 and amination64 are all examples of functional groups that are linked directly to the OH radical of cellulose molecules, a process known as a direct modification. The other class contains free radical polymerization, ring-opening polymerization, and controlled radical polymerizations to graft monomers onto cellulose chains in homogeneous and heterogeneous solutions.65 The cellulose components that undergo these modifications significantly aid water decontamination and the elimination of harmful organic and metal chemicals.66

As a consequence of their high aspect ratio, high hydrophilicity, low density, and remarkable mechanical properties, natural cellulose-based resources such as wood and cotton-generated cellulose nanocrystals (CNCs) are good nanostructures in membrane processes for ecological sciences and environmental engineering fields. Boosting the antifouling capabilities of CNCs membranes is made easier by removing cellulose at the nanoscale, eliminating major flaws related to the layered structures. Thin film composite nanofiltration membranes made from CNCs were developed by Huang et al. for increased water flow and chlorine resistance.67 A significant improvement in salt removal of MgSO4 (96.1%) and Na2SO4 (98.3%) and water penetration flow (106.9 L m−2 h−1) was seen in CNCs thin film composite (TFC) membranes compared to TFC membranes without CNCs. Bai et al.68 inserted cellulose nanocrystals into a polyamide (PA) layer while manufacturing a TFC membrane to improve antifouling and separation performance (Fig. 2a). Compared to PA TFC without CNCs, the penetration of synthesized CNC-TFC membranes was 60.0% greater with just 0.02 wt% CNCs. Because of their exceptional hydrophilicity and permeability, CNC-TFC membranes show promising desalination and purifying applications for water. CNCs membranes offer more fouling resistance and recovery capabilities than commercial membranes (Fig. 2b).


image file: d3ra06501d-f2.tif
Fig. 2 (a) Synthetic scheme of the CNC-TFC membrane. (b) Adsorption mechanism of CNC-TFC membranes for metal ions. Reproduced with permission.68 Copyright (2018), ACS.

2.3. Chitosan

Partial deacetylation of chitin is the principal source of chitosan, the only naturally occurring cationic polysaccharides. The hydroxyl group and free amino on the CS chains achieve an effective chelation effect or electrostatic adsorption of anionic organic and metal ion contaminants. Chitosan-based materials display various limitations that prevent their broad use in water treatment, including acidic solubility, poor mechanical strength, and insufficient absorption capacity, despite their major benefits in adsorbents.69

Crosslinking, grafting, carboxymethylation, and acetylation are just a few chemical modifications used during the last several decades to boost adsorption capacity and chemical stability. A water purification strategy using a heterostructure chitosan multilayer membrane by layer-by-layer self-assembly design route was modeled after the layered structure of wood. The sieve effect, hydrophobic and strong van der Waals forces effect shown by the biomimetic membrane provided the required removal of diverse contaminants. Meanwhile, this multilayer membrane has demonstrated great durability under constant use and high recycling rates when getting rid of oil droplets.70 The adsorption capacity of Chitosan (CS)-based absorbents for industrial dyes and heavy metal ions has been improved by incorporating different nanoparticles into CS matrices, which has increased their specific surface areas. By combining electrospraying and freeze-casting, radially-oriented microchannel and honeycomb-cobweb designs were used to construct graphene oxide-aerogel microspheres by Yu et al. Higher adsorption capacities of the aerogel for the anionic dyes eosin Y and rhodamine B, the cationic dyes methylene blue and methyl orange (MB), and micropollutants result from the creation of a variety of interactions between the contaminants, including complexation, conjugation, coordination and chelation.71 The developed CS-based nanofiber membrane has a penetration flux of 1533.26 L m−2 h−1, and can efficiently filter out dyes and heavy metals from pure water. Green chemistry and sustainable development are supported by the fact that no VOCs were utilized to manufacture the nanofiber membrane. The research might lead to a more efficient and cost-effective method for creating high-tech membranes for water filtration in the near future.

2.4 Starch

Starch is an adaptable and common natural polymer that has gained a lot of interest as a potentially useful ingredient in the pharmaceutical, nutritional, and chemical industries. Starch has two different types of polysaccharides, amylopectin and amylose.72 Twenty to thirty percent of starch is amylose, which is insoluble in water and consists of a linear chain of D-glucose components with –1,6 glycoside linkages (1%) and 1.4 glycosidic connections (99%). Amylopectin is mostly composed of a branching chain of D-glucose components joined by –1,4 glycosidic connections (70–80%) and −1,6 glycosidic linkages (20–50%). Amylopectin is two to three times as large as amylose, with a molecular weight of 1–2 million daltons.73,74 Amylopectin is actually the water-soluble component of starch, which makes up around 80% of starch.75 Starch has many benefits, including its biocompatibility, low cost, availability, and biodegradability. Nevertheless, natural starch has low adsorption capacity because it lacks necessary functional groups like carboxyl, amino, or ester. Chemical techniques such as crosslinking, grafting, oxidation, irradiation, and esterification have been used to alter starch and enhance its utility beyond these parameters.76 The ability of this hydrogel to soak up anionic and cationic dyes suggests it might be used in wastewater treatment. By reacting starch with (3-chloro-2-hydroxypropyl trimethyl) ammonium chloride, cationic starch-containing ammonium groups were formed, which were then used to make starch-based nanocomposites for cationic dyes adsorption.77 The produced quaternary starch has the ability to absorb the nitrate group from wastewater with a maximum adsorption capacity of 205 mg g−1. This effect was sustained at 78.5% after eight sorption–desorption cycles. There are several studies on removing heavy metals and dyes using modified starch, but far less research on the removal of ammonia and phenol. New methods of modifying starch with a high affinity for these poisonous chemicals need to be investigated in future studies.78

2.5 Alginate

Natural polysaccharide alginate is harvested from brown seaweeds and is a sustainable resource of bio-based materials. It is composed of –L-guluronic acid (S unit) and –D-mannuronic acid (M Unit) in the form of heteropolymeric (GMGMGMGM blocks) and homopolymeric (GGGG or MMMM blocks) sequences, respectively. As a polymer, alginate has many carboxyl and hydroxyl groups, giving it a high affinity for radionuclides and metals.79 However, scientists are very interested in learning more about synthesizing alginate derivatives via surface functionalization to advance their stability and mechanical strength for use in environmentally friendly applications. In most cases, you can functionalize alginate in one of two ways.80 One method involves chemical functionalization, whereby covalent connections are formed between alginate and functional groups via processes, i.e., oxidation, esterification, copolymerization, amination, and phosphorylation. Alginate may also be physically functionalized via ionic crosslinking or blending, which generates relatively weak interactions, i.e., hydrogen bonds, electrostatic contacts, van der Waals forces, and coordination bonds.81 As a means of increasing the adsorption capacity of U(VI) ions and enhancing the material's mechanical, thermal, and radiation resistance, Khajavi et al.82 utilized crosslinking and grafting procedures to effectively construct the new amidoximated modified calcium alginate beads with entrapped functionalized SiO2 nanoparticles for the removal of U(VI) ions from aqueous solutions (Fig. 3a). To improve the mechanical, thermal, and radiative resistance of alginate and the uranium adsorption capacity, we have introduced modified SiO2 nanoparticles with a –SH group into the calcium-alginate gel beads for the first time. Adsorption capability of U(VI) ions onto the produced beads was studied as a function of SiO2 (Fig. 3b) and TMPTMS (Fig. 3c) weight percentages. At 5 weight percent SiO2 and 30 weight percent TMPTMS (relative to the alginate weight), the alginate/SiO2/TMPTMS/PAO beads had 3.6 times the adsorption capacity of the blank one (alginate/PAO) for U(VI) ions. The findings demonstrated that U(VI) ions may be effectively removed from aqueous solutions by using a composite adsorbent composed of alginate, SiO2, TMPTMS, and PAO. Nanoparticles, in addition to SiO2, may be used to raise the alginate resistance, and various functional groups, in addition to amidoxime and –SH, can be used to boost the adsorption capacity of uranium ions, all of which can contribute to future progress in the field of uranium removal from aqueous solutions.
image file: d3ra06501d-f3.tif
Fig. 3 (a) The adsorbent (alginate/SiO2/3-mercaptopropyltrimethoxysilane/amidoxime) fabrication scheme (a–e). (b) The adsorption of U(VI) ions at pH 4 with an initial uranium solution concentration of 120 mg L−1 and an adsorbent dose of 1 g L−1 is affected by the SiO2 content of alginate/SiO2/PAO beads. (c) The adsorption of U(VI) ions at pH 4 with an initial uranium solution concentration of 120 mg L−1 and adsorbent dose of 1 g L−1 was affected by the amount of TMPTMS present in the alginate/5% SiO2/TMPTMS/PAO mixture. Reproduced with permission.82 Copyright (2021), Elsevier.

To create a dispersive magnetic solid-phase extraction adsorbent, Orachorn et al.83 integrated metal–organic frameworks (MOFs), graphene oxide (GOx), and iron oxide (Fe3O4) into alginate fiber. Hydrogen bonding and hydrophobic interactions facilitated by combining GOx with MOF improved the phthalate ester extraction. Ca2+ is the most common cation employed during the alginate hydrogel synthesis. Ca2+ forms a three-dimensional network with the G-block regions through coordinated interactions. By constructing FeCl3-activated seaweed carbon and then crosslinking with calcium chloride,84 researchers were able to produce an alginate hydrogel composite for the effective removal of blue (BB) dye, plasticizer, and bisphenol A (BPA). These studies show that alginate may be used as a long-term absorbent to filter out pollutants in industrial wastewater, which has motivated scientists to investigate this polymer further to evaluate its efficiency.

2.6 Gelatin

Collagen is an important protein found in many different types of biological connective tissue, including bone, cartilage, skin, and tendon. Its partial hydrolysis yields the water-soluble biopolymer known as gelatin. Gelatin has many functional groups (i.e., COOH, NH2, and OH) in its structure, which might operate as binding points to adsorb contaminants in wastewater channels through ionic or polar interactions. Nevertheless, the limited application of gelatin is due to its limitations, such as its quick degradation and low mechanical strength in moist environments. Therefore, nanomaterials (CNT, Fe3O4, MnO2, and TiO2) may improve the material, degradation rate and mechanical stability. Moreover, chemical modification was used to create constituents with high elimination efficiency of pollutants from contaminated water by putting different functional groups onto the gelation molecular chains.85

Ethylene glycol di-glycidyl ether has been used to manufacture xylan and gelatin-based hydrogels. Good methylene blue adsorption and shear thinning behavior were observed, and the gel was stable under high strain and temperature variations without losing its crosslinked structure (MB). To remove anionic and cationic dyes from an aqueous medium, magnetic nanoparticles of iron oxide and multi-walled carbon nanotubes were treated with carboxylic acid to create a novel adsorbent developed by Samaneh et al.86 A hybrid composite of amine-grafted magnetic gelatin has been produced by Kumar et al. that can efficiently remove phosphate (PO43−) and nitrate (NO3−) from wastewater. The gelatin hybrid composite interacted with NO3−/PO43− through electrostatic attraction and a surface complexation process.87 By combining TiO2 mixing, polyethyleneimine (PEI) crosslinking, and freeze-drying, Jiang et al. developed a multipurpose organic–inorganic composite aerogel based on gelatin. Because of their super amphiphilic surface and hierarchical porous structure, the resultant aerogels showed remarkable selectivity for oil/water separation in both aqueous and O/W combinations.88 Adsorbent composites based on gelation show significant potential for use in water purification. Gelation composites with high mechanical strength are thus being developed by scientists for applications, such as the adsorptive elimination of multi-component and pharmaceutical solutions.

2.7 Polyhydroxyalkanoates

Polyhydroxyalkanoates (PHAs), a type of hydrophobic aliphatic polyester manufactured by microbes, have material qualities similar to those of typical plastic goods while providing improved biocompatibility and biodegradability. Since PHAs may be entirely decomposed into innocuous compounds in their natural environment, they are important emerging biopolymers to replace traditional petrochemical plastics. Two distinct classes of PHAs have been discovered: short-chain PHAs (with three to five C-atoms in the monomer) and medium-length PHAs (with 6 to 14 C-atoms in the monomers).89 Changing the feed, the strain, and the fermentation process allows for simple manipulation of the PHA composition. This structural variety makes PHA useful for various applications since they may be rigid bodies or elastomers, two classes of biomaterials currently unavailable.90

Mannina et al.91 presented a new procedure for removing polyhydroxyalkanoates from a contaminated mixed microbial culture to improve the economic and environmental sustainability in wastewater treatment. The underlying premise that scaling up the process from a laboratory to a large-scale ability within the plant might make (mixed microbial cultures) MMC-PHA manufacturing economically and ecologically viable has led to the expansion of the process from the laboratory to the pilot-scale facility. The pilot-scale production of PHAs as final high-value bio-based material yields from municipal solid waste (MSW) and sewage sludge mixes was studied by Valentino et al.92 utilizing a three-step MMC method. This method streamlines the conversion of municipal bio-waste to PHAs using a single basic technology. It significantly reduces the waste that must be disposed of in the end. Samori et al.93 used a combined thermochemical and biotransformation process to extract PHAs and crotonic acid from anaerobically digested sludge, maximizing the sludge's content and chemical energy utilization.

Through a transdisciplinary four-step process (Fig. 4a), the potential of this waste to be used as feedstock for the coproduction of biopolymers and platform chemicals has been evaluated by combining two “tool-boxes,” one based on thermochemical treatments and the other on biological conversions. Fig. 4b shows the microbial fermentation microscale laboratory experiment design. The feedstock's (then its COD's) chemical energy must be transformed into a particular class of organics, namely VFA, excellent substrates for MMC growth/adaptation during feast and famine cycles, in order to exploit the MMC system for PHA generation. Fig. 4c shows the acidogenic fermentation experiments performed on HTCap obtained at 150, 200, and 250 °C using the same ADSS as the acidogenic inoculum. To measure the HTC fermentation's durability over time, a larger-scale semicontinuous test was conducted. The test was conducted on HTC slurry created at 200 °C without any solid–liquid separation (Fig. 4d), which was chosen to maximize the VFA yield and process dependability (less severe conditions are simpler to accomplish and cost less energy). The approach is to reduce the waste volumes and eliminate material and energy losses by managing the anaerobically digested sewage sludge to acquire high-value-added commodities and chemicals. High manufacturing costs are now the biggest barrier to the mass production and broad use of PHAs. Thus, it is important to investigate other low-cost and readily accessible raw materials for PHAs synthesis. Second, PHAs need to have their qualities, such as machinability and hydrophilicity, increased to fulfill the demands of various applications.


image file: d3ra06501d-f4.tif
Fig. 4 (a) From wastewater sludge to biopolymers and platform chemicals: a transdisciplinary four-step process. (b) Acidogenic microbial fermentation microscale laboratory test depiction. (c) After fermenting HTC ap at 150, 200, and 250 °C for 60 minutes, the relative VFA content (percent). (d) Anaerobic fermentation in semicontinuous test yields of biogas, VFA, and other fermentable compounds (percent, based on total COD, CODt). Reproduced with permission from.93 Copyright (2019), ACS.

2.8 Polylactic acid

Starch feedstock from natural bio-based resources like cassava, corn, and sugarcane is the primary source for the commercially promising biopolymer polylactic acid (PLA). High-purity lactic acid (LA) is produced by microbial fermentation.94 Lactic acid is a starting material for the chemical synthesis of polylactic acid (PLA) of a predetermined molecular weight from abundant and renewable raw materials like starch. Moreover, PLA is produced in a zero-waste manner and can be broken down entirely by microorganisms in the natural environment after its use, effectively recycling itself in the process.95,96

Numerous methods are available today for porous PLA, including foaming, electrospinning, freeze drying, thermally induced phase separation, and 3D printing. Using in situ oxidative polymerization, researchers have synthesized PLA electrospun nanofibers covered with chloride-doped polyaniline. Weak electrostatic interactions between the adsorbent and adsorbate was facilitated by combination of the conductive polyaniline (PANI) coating with the porous PLA, leading to good adsorption characteristics of the membranes.97 Preparing porous PLA materials presents several challenges that include finding the optimal solvent and process parameters and preventing the porous structure from collapsing, which would lead to shrinkage and cracking. Drying porous materials with supercritical carbon dioxide prevents shrinkage and fragmentation, allowing the materials' high open porosity and original structure to be preserved. Supercritical CO2 drying can produce porous PLA materials with a permeability of up to 73%, which is ideal for treating wastewater efficiently. The structure and morphology of PLA can be adjusted depending on the drying and solvent method.118 Although PLA materials are used in various wastewater applications, most current research focuses on oily and colored wastewater treatment. Future studies will concentrate on developing highly porous PLA materials for disinfecting water tainted with chemicals like pesticides, surfactants, and bacteria.

2.9 Pectin

Pectin is a linear polysaccharide that occurs naturally, and may be obtained from the cell walls of higher plants. It is important because of its flexible, biocompatible, high-molecular weight, nontoxic, and anionic nature. Most plant cells include pectin in the intercellular layer between their major cell walls. Pectin has seen increasing interest in recent years due to its potential uses in the healthcare and biotechnology sectors, among others. Like many other complex polysaccharides, the composition of pectin varies depending on its source and the isolation conditions used. Pectin is mostly composed of chains of d-galacturonic acid molecules connected by a –(1–4) glycosidic bond, with some neutral sugars present along the side chains.98 Certain COOH groups occur naturally as methyl esters, whereas other may react with NH3 to make carboxamide groups. They form complexes with metal ions in solutions, and reduce them to metallic nanoparticles without the use of dangerous reducing/stabilizing chemicals.

Palladium nanoparticles were stabilized on polysaccharide materials including Pd/gelatin, Pd/chitosan, and Pd/starch as the easiest greener ways for the synthesis of organically recyclable Pd nanocatalysts. Because of its versatility, which is achieved by the hybrid inclusion of both organic and inorganic nanocomposites, there is a great deal of interest in these materials.99 Because of its great temperature durability, strong covalent bonding, and highly active surface area, the agar/pectin composite was developed as a stabilizer. Subsequently, under safer, environmentally preferable circumstances, palladium nanoparticles were produced by in situ reduction of palladium ions and immobilized on a membrane surface. The biopolymer composites demonstrated catalytic activity in reduction processes, suggesting their potential use as stabilizers for a wide range of noble metallic NPs. 1,2-Benzenediamine was produced by reduction of O-nitroaniline, and recycling tests showed a drop in reaction yield from 100% to 83% after eight cycles of using the nanocomposite. Hybrid pectin/titanium oxide nanobeads were created by Bok-Badura et al., and utilized as an effective nano-adsorbent for the elimination of ionic heavy metals. This study investigated the impact of pH on the adsorption behavior of copper, cadmium, zinc, and lead ions within different pH ranges. Specifically, pH levels ranging from 1 to 5 were considered for Cu(II) and Pb(II), pH levels ranging from 1 to 6 were considered for Cd(II), and pH levels ranging from 1 to 7 were considered for Zn(II).These hybrid nanobeads were capable of binding 0.83 mmol g−1 of Pb2+ and 0.68 mmol g−1 of Cu2+.100

2.10 Agarose

Red algae produce agarose, 3,6-anhydrous-L-galactose, and a linear polymer made up of D-galactose. It may create hydrogen bonds with hydrogen atoms already present in its structure, or with water molecules because of the presence of several hydroxyl groups distributed throughout the structural unit.101 Typically, the molecular weight is between 80 and 140 kilodaltons. Hydrogels made from agarose have the potential to be controlled, stable, and hysteretic.102 The outcome is a translucent solution formed when the agarose molecules disperse into the surrounding water medium and assume a random coil shape. A tightly organized gel structure typified by a double helix shape is formed at temperatures between 30 and 40 °C as a consequence of hydrogen bonding between agarose molecules. Current research indicates that the global agarose market will continue to show growth over the next five years. By 2022, experts predict that the global agarose market will be worth USD 83.35 million; by 2028, that number is expected to rise to USD 99.35 million. Over the course of the anticipated time frame, this increase is anticipated to occur at a CAGR of 2.97%.103

Felipe Melo Lima Gomes et al. demonstrated that in order to remove Cd(II) and methyl violet from water, an agar-graphene oxide hydrogel was created for this experiment. Scanning electron microscopy and energy-dispersive X-ray spectroscopy were used to create images and data describing the hydrogel's composition and structure. Studies were conducted across several disciplines, including kinetics, equilibrium, and regeneration. Langmuir, Freundlich, and Sips isotherm models were used to match the equilibrium experimental data. With an R2 of 0.968 and χ2[thin space (1/6-em)]= 0.176, the Sips model provided a very good match to the Cd(II) data. With an R2 of 0.993 and χ2[thin space (1/6-em)]= 0.783, the Sips model likewise showed a very good match to the MV data. Maximum adsorption capacities of 76.65 mg g−1 for monovalent mercury ions and 11.70 mg g−1 for divalent cadmium ions were found. With R2 values over 0.90, the pseudo-order models provided a good description of the kinetics of the Cd(II) and MV adsorption. Regeneration tests showed that the adsorbent could be reused after three cycles of adsorption and desorption for both Cd(II) and monovalent mercury, which is quite an impressive feat. The study demonstrated the feasibility of a practical adsorbent for the repeated removal of Cd(II) and MV from water.104

de Araujo et al. analyzed the A-GO biocomposite's morphology, and showed that the lyophilized hydrogel had a three-dimensional porous structure. The Sips (Safranin-O) and Freundlich (Chloroquine) adsorption isotherms showed a good match (R2 > 0.98) in the batch experiments. The kinetic data were successfully modelled using driving force models and Fick's diffusion equation, with good results in both cases. Experiments with batch selective adsorption showed that competitive adsorption occurred in the presence of both components in water. This was demonstrated by a reduction in the adsorptive capabilities of around 10 mg g−1 for each component. Adsorptive capacities for safranin-O were determined to be 41 mg g−1, whereas those for chloroquine were determined to be 31 mg g−1. The fixed-bed adsorption column breakthrough curves showed that the adsorption capabilities for chloroquine and safranin-O were 63 mg g−1 and 100 mg g−1, respectively.105

de Araujo et al. introduced hydrogels made with graphene oxide and agar, and have shown evidence of graphene oxide-polymer interactions in their investigation. An irregularly shaped, three-dimensional material formed as a result of this interaction. Adjusting the pH of the BF and MG solutions had a minor effect on the adsorptive capacity, as shown by the batch testing. A constant increase in MB's adsorptive capacity was seen, whereas NB's action was more prominent until the pH approached 8. For the MG, BF, and MB isotherms, the Freundlich model was the best fit, whereas the Sips model was the best fit for the NB isotherm. Successful fits were found for the kinetic data using the equations of Lattice Diffusion Flux, Quasi-Dynamic Flux, and Fickian Diffusion. The fixed-bed experiments showed that the individual dry adsorption capacities of the dyes were as follows: methylene blue 79.51 mg g−1, malachite green 58.25 mg g−1, and Brilliant blue FCF 38.11 mg g−1, Nile blue 224.46 mg g−1. These results demonstrated the regenerative capability of the adsorbents. Positive findings regarding the adsorbent's selective separation of colors from synthetic textile effluent were found in an evaluation of the effectiveness of an agar-GO hydrogel-packed column.106

3. Synthetic strategy to fabricate bio-based materials

As techniques for creating nanomaterials and fabricating thin films have advanced rapidly, a number of novel BBMs have been recently published. In this article, we focus on some of the most modern, widely applicable, and highly efficient methods for preparing BBMs. Molecular imprinting, phase inversion, electrospinning, and chemical modification are only some of the manufacturing processes detailed here (Fig. 5), along with discussions of their respective benefits and drawbacks.
image file: d3ra06501d-f5.tif
Fig. 5 Various fabrication strategies for bio-based materials.

3.1 Molecular imprinting (MIPs)

The emergence of molecular imprinting has allowed synthetic polymers to be engineered with built-in recognition sites for specific target molecules. Functional cross-linkers and monomers are copolymerized with a template molecule to produce bio-based materials. Extracting the template groups from the polymeric networks requires forming recognition cavities that are equivalent to the template molecules in structure, dimensions, configuration and chemical functional groups. This allows for the selective rebinding and recognition of the original templates from a mixture of closely related compounds. Because of their low cost, excellent selectivity, high stability, physical robustness, and ease of manufacturing, a broad range of molecular imprinting polymers have been developed and used as adsorbents for the efficient and effective separation of carcinogenic organic pollutants and harmful metal ions in wastewater treatment and water purification.107–113

There has been substantial progress in the use of bio-based polymeric MIPs in the wastewater treatment sector due to their various advantages. Among these advantages are their potency in selective adsorption, high affinity, lack of toxicity, ease of manufacture, and cheap cost. A photo-responsive cellulose-based intelligent imprinted adsorbent for the specific adsorption of common pesticides material utilizing surface-initiated atom transfer radical polymerization was investigated and synthesized by a group of researchers. The PR-Cell-binding MIP's capacity was able to reach 11.039 mg g−1 due to its great adsorption specificity, amazing reusability, and outstanding stability for the target 2,4-D. The cellulose adsorbent was easily regenerated by exposure to UV light by incorporating an azobenzene (Azo) function monomer.114 Chitosan (CS) is a commonly known polysaccharide used in molecular imprinting because it has a high concentration of hydroxyl and amino groups. These groups provide higher affinity to diverse chemicals through different sorts of specialized interactions. In order to selectively remove cadmium from aqueous solutions, Rahangdale et al. prepared a dual surface imprinted acrylamide functionalized chitosan-based polymer using epichlorohydrin as a crosslinker, chitosan (CS) as a substrate, 4-hydroxy benzoic acid (4HBA) as a mimic template, and cadmium (Cd) as a template. For Cd and SA, the highest adsorption capabilities in AGDMIP were 53.42 and 45.77 mg g−1, respectively.115 In order to create ion-imprinted cryo-composites from chitosan, Dinu et al. demonstrated an ice-templating and ion-imprinting method using N,N′-methylenebisacrylamide (BAAm) and acrylamide (AAm) as monomers, glutaraldehyde (GA) as a cross-linker and Cu2+ as a template ion. Due to a chelating reaction between the OH and NH2 groups and Cu2+ ions from the CS matrix, the II-CCs showed remarkable specificity adsorption ability for the Cu2+ ions.116 With the right microscopic recognition cavities, bio-based polymeric MIPs can efficiently remove trace pollutants from water in real-world applications. This is because these MIPs rely on and enhance the formation of bonds between individual recognition sites.

3.2 Phase inversion

Currently, phase inversion is the most adaptable, cost-effective, low maintenance and versatile method for fabricating polymeric filtering membranes. This process involves the controlled transition of an initially homogenous polymer solutions to a solid phase. Flat sheets and spinning fibers can be made from membranes with asymmetric and hollow fibers. Different methods of phase inversion include thermally-induced phase separation (TIPS), vapor-induced phase separation (VIPS), and non-solvent-induced phase separation (NIPS). The physical and chemical interactions between the membrane and its environment are influenced by a number of factors, including the additives, solvent interaction, polymer concentration, and bath temperature utilized in the casting solutions. In this method of membrane synthesis, it yields two distinct membrane morphologies and structures via finger-like and a sponge-like structures. Structures resembling fingers or macrovoids form when demixing occurs instantly, while structures resembling sponges form when demixing is delayed.117 Conversely, the sponge-like structure has much greater mechanical strength than the finger-like structure because its tiny pores are surrounded by solid walls, allowing for the production of an entire thin active layer. Conversely, the flow resistance is lowest in the figure-like pores because they are the least twisted. Therefore, optimal support with a thin layer of sponge-like material placed on top of a finger-like sublayer is required to fabricate high-performance filtering membranes.118–121

A considerable quantity of cellulosic membranes was produced using the phase inversion method. Cellulose acetate (CA) has been used in a variety of separation techniques, including nanofiltration, microfiltration, ultrafiltration, and reverse osmosis.122 Asymmetric CA membranes have much promise as a desalination and water purification technique due to their high water porosity and salt retention. The CA-based composite films endured a phase inversion due to the presence of water vapor at room temperature and pressure. With a separation flow of 667 L m−2 h−1 and an efficiency of 99.99%, the resultant composite film successfully separated nano-sized surfactant-stabilized W/O emulsions.123 The TiO2 photocatalytic activity was also responsible for the bio-film potent photocatalytic destruction of organic molecules under UV irradiation, which is encouraging news for the industrial and domestic water treatment sectors.

The toxic solvents formerly used in membrane preparation by phase inversion have been replaced with environmentally friendly alternatives. This category includes solvents like dimethyl sulfoxide (DMSO),124 triethyl phosphate (TEP),125 dimethyl isosorbide (DMI),117 and ionic liquids (ILs).126 Using NIPS and bio-based green solvents like glycerol derivatives and γ-valerolactone (GVL), it is now possible to create permeable membranes. The manufactured membranes with the related γ-valerolactone may be used for microfiltration, ultrafiltration, and nanofiltration if the pore sizes are small enough.127 Additionally, the cellulose acetate nanofiltration membrane was produced by phase inversion using the green solvents, methyl lactate and 2-methyltetrahydrofuran. A highly porous and permeable morphology, characterized by finger-like macrovoids, was generated at low concentrations of CA. Increasing the CA content in the casting solution decreased the permeance from 32.0 to 2.4 L m−2 h−1 bar−1 and enhanced the rhodamine B dye removal from 31.1% to 99.5%.128

Antifouling, permeation flux, and highly specific and selective separation are just a few areas where membrane performance might be improved with the use of nanomaterials into the phase inversion process used to manufacture BBMs nanofiltration membranes. The key limitations of the organic nanoparticles are their lack of thermal stability and deficiency in acid-base tolerance. These limitations may be mitigated by controlling the structure of the organic nanoparticles. The basic limitations of the use of loose nanofiltration membranes include the incompatibility of inorganic nanoparticles with the biopolymer matrices and their tendency to agglomerate, which results in the development and creation of defects and imperfections at the bipolymer interface. Improving the correlation between the inorganic nanoparticles and biopolymers, modifying the surface morphology and structure, and improving the distribution of inorganic nanomaterials are necessary steps in the development and manufacturing of high-performance, reliable and advanced inorganic hybrid loose nanofiltration membranes. The permeability and selectivity of loose nanofiltration membranes are compromised due to the abovementioned reasons. For this reason, researchers are largely focused on the creation of novel matrix materials that allow for high permeation flux without reducing separation fidelity.

3.3 Electrospinning

Producing polymer nanofibers with good structure–property relationships in the submicron range (5–500 nm) inexpensively and in large quantities is now possible by electrospinning. As a consequence of their high permeability, large surface area, high water permeability, and good stability, electrospun polymer nanomaterials have immense capability as filtration materials in the wastewater purification process. Electrospun polymeric nanofibers are created by uniaxially stretching and elongating a viscoelastic polymeric solution in a high voltage field. A syringe pump, a grounded metal collector, a high voltage power supply, and a spinneret are its four components. By applying a voltage between the needle tip and the grounded collector, the polymer solution is transformed during the electrospinning process from a spherical shape into a conical one (known as Taylor cone). When the electric field strength is strong enough, the surface charges' electrostatic repulsion will overcome the droplet's surface tension. The needle tip's Taylor cone will eject the fluid jet towards the grounded collection. The droplet's surface tension will overcome the surface charges' electrostatic repulsion when the electric field strength is low. The solvent has swiftly evaporated, and the jet has stretched and curved to generate nonwoven membranes.129–131 This nanofibrous membrane may function as an adsorbent or as a filter membrane in separation processes, including ultrafiltration (UF), microfiltration (MF), and nanofiltration, due to its high permeability (>90%), densely interrelated pores, and multi-channeled architectures (NF).

Recently, a new chitosan/polyvinyl alcohol/polyvinylpyrrolidone membrane was developed using electrospinning for the elimination of organic pollutants and heavy metal ions in polluted water.132 With pure water permeability of 4518.91 L m−2 h−1 bar−1 and an average diameter of 160 nm, the CS/PVP/PVA nanofiber membrane generated under ideal electrospinning conditions demonstrated permeability and a consistent biocompatible structure. Maximum removal rates for malachite green, methylene blue, Cd(II), Pb(II), Ni(II), and Cu(II) are 94.20, 69.91, 83.33, 90.35, 80.12, and 84.01%, respectively. The large specific surface area and open pore structure of the electrospun cellulose acetate (CA) membranes have made them a promising water filtering material. Using an electrospinning technique, Goetz et al. developed a hydrophilic cellulose nanofibrous membrane with an anti-fouling surface structure property. As a result, CA fibers tend to arrange themselves in a disorderly fashion, with some fibers preferentially oriented vertically, horizontally, or diagonally with sizes between 0.5 and 2.0 m, which results in a nanofiber membrane that has the ability to degrade dyes by 80–99%.133 The use of electrospun chitosan (CS) membranes for heavy metal ion elimination has become widespread due to their superior metal biodegradability, high hydrophilicity, and chelation ability.134

Due to their high flux, improved reduction, low operating pressure, and antifouling capabilities, electrospun nanofiber membranes are potential options for advanced filter media. Nonetheless, there are some drawbacks to electrospinning technology, i.e., a lack of structural stability, needle clogging, high electric field requirements, a time-consuming process, low productivity, and inapplicability to all biopolymers. The development of needleless electrospinning technology has allowed for the resolution of these limitations. High levels of safety and productivity may be achieved in continuous systems via the use of needle-free electrospinning for the manufacturing of nanofiber membranes. While the large-scale production of electrospun nanofiber membranes is still in the research and development phase, a few medium-sized manufacturers already exist. Electrospun nanofibrous membranes are the newest generation of filtration materials, and they bring with them important benefits that will help in the creation of cutting-edge viable filtration systems.135

3.4 Chemical modification

Chemical modification is the process by which the structures of biopolymers are changed chemically to produce derivatives with unique or improved physicochemical features. Chemical modifiers improve chemical stability by covalently attaching to the substrate. By decreasing the impacts of membrane fouling, chemically altering water treatment membranes to enhance their hydrophilicity has the potential to dramatically improve water permeability. Ahn et al. transformed ordinary A4 printer paper into multifunctional cellulose membranes by performing a chemical alteration. Calcium carbonate filler in the A4 paper was dissolved using a hydrochloric acid solution, and then alkoxide functional groups were added to the cellulose fibers. After treating the A4 membrane with acid and base, trichlorooctylsilane (COS) was selected to alter the OH groups on the cellulose network.136 With this chemical modification technology, commercial A4 paper might be utilized in the future as a multifunctional cellulose-based membrane, with a simple procedure and low chemical costs.

Amine is electrophilically substituted by active hydrogen in a process known as the Mannich reaction, which results in aminated lignin. The Mannich approach aminolation of lignin was followed by chelation of Fe(III) onto the aminated lignin to produce novel lignin-based adsorbents for the very efficient phosphate elimination from wastewater. When it comes to phosphate (5 mg L−1), the newly discovered materials have a removal effectiveness of more than 90%. The adsorption mechanism relies on Fe(III) and phosphate complexation on the lignin-based adsorbent. The findings of this work suggest a realistic and cost-effective way for building low-cost BBMs for the treatment of low-phosphate wastewater, reducing the eutrophication of bodies of water.137 Chaudhary et al. created a highly porous chitosan-based aerogel barrier for rescuing potable water from an oil spill in a shipbreaking yard using genipin as a cross-linking agent. Fig. 6a and b graphically depicts the steps required to make a CS aerogel membrane with a superhydrophilic agarose inner wall coating. The shells of crustaceans like shrimp, lobsters, and crabs are mined for the plentiful natural resource chitosan. The CS's –NH2 functionality allows it to be chemically cross-linked, and its fibrous nature means it may be employed in a variety of ways after undergoing chemical and physical modification. Additionally, by regulating the degree of crosslinking, CS may be converted into a robust scaffold-like structure. Agarose was utilized as a gelling agent, and it contributed to the development of a highly porous aerogel membrane that could accommodate a wide range of pore sizes. Interestingly, agarose also played a role in enhancing the hydrophilic property by interacting with chitosan through H-bonding during lyophilization. At 80 degrees Celsius, as shown in Fig. 6c, genipin rapidly crosslinks chitosan, whereas agarose forms a stable gel by H-bonding interaction with the –OH of chitosan, allowing the gel to be used at room temperature. Water with a purity of >99% and a flow rate of 600 L m−2 h−1 bar−1 may be obtained by modifying the genipin cross-linking process, and changing the chitosan-based aerogel into a column-like and stiff scaffold-like structure.138


image file: d3ra06501d-f6.tif
Fig. 6 Aerogel membrane preparation schematics: (a) uncrosslinked chitosan, (b) genipin-crosslinked chitosan, and (c) genipin-crosslinked chitosan with agarose H-bonding to the inner walls of CS. Reproduced with permission from.138 Copyright (2015), ACS.

The location, amount, and spatial conformation of the groups utilized to alter the biopolymer are intricately connected to the physicochemical features of BBMs. In order to understand more about the structure–function link, scientists must perform extensive research on modified biopolymers. Furthermore, if current techniques for chemical modification are developed and the optimal modification conditions are discovered, the structure and stability of BBMs may be determined more accurately.

4. Benefits of eco-friendly green synthesis over conventional chemical synthesis

There are various benefits to green material production versus chemical synthesis.139 Because biological components are already accessible, production costs are dramatically reduced.140 Natural, biocompatible capping and reducing agents found in biology may aid in lowering production costs. Nanoparticles may be made from low-value byproducts of agriculture and industry. Chemical synthesis procedures, on the other hand, need the use of hazardous, dangerous, and/or expensive ingredients that are harmful to the environment.141

Commercially available activated carbon and iron or iron oxide nanoparticles are anticipated to be 50 times more expensive than green produced iron-based nanomaterials. The use of bio-based materials in the synthesis process accounts for the remarkable price drop.142 Because of its quicker kinetics, several investigations favor plant-based bio-based material synthesis over chemical preparation.143 Furthermore, the biomimetic green synthesis technique demonstrates the interface between biomolecules and inorganic materials, as well as the underlying biological processes. Because the biomolecules utilized have well-defined structures, chemical properties, and compositions, the metal nanoparticles generated by this method have chemical purity, well-defined geometry, and dimension.144,145

5. Reasons to remediate the environment with eco-friendly materials

Since the beginning of the industrial age, environmental remediation has been one of the most important tenets of sustainable practice. Notably, the continual discharge of high volumes of synthetic chemicals might surpass natural degradation capability, leaving behind persistent pollutants that have severe effects on the ecosystem and human health.146 Concentrating pollutants, removing them from environmental matrices, and treating them sequentially to eliminate them are standard steps in any remediation process. The materials' adsorption capacities, which enrich pollutants through chemical and physical interactions between adsorbents and adsorbates,147,148 are crucial to the concentration of environmental contaminants. The adsorption and treatment of environmental contaminations have been studied using various materials created by non-biological processes. Natural mineral clays,149 resins,150 metal–organic frameworks,151,152 and mesoporous silica153 are all examples that have been thoroughly explored in the past as potential candidates. However, there are frequently monetary and environmental issues with the production process for these adsorption materials. Researchers in the field of techno-economic analysis have shown that the production of traditional remediation materials is quite labor- and resource-intensive.154 It is also common for processes like pyrolysis used in the synthesis of certain materials to generate corrosive waste products or hazardous gases. Costs and possible secondary environmental risks are further increased by the post-absorption and rejuvenation process.

Biomaterials, particularly those made from lignocellulosic biomass, provide an eco-friendly, renewable alternative. When used in conjunction with various functionalities, these biomaterials may be used for effective absorption and remediation.155 Heavy metals and persistent organic pollutants (POPs) have been examined for their ability to adsorb onto biochar, chitosan, microbial biomass, and a wide variety of agricultural wastes.156 Biomaterial, and specifically biomass-based material for environmental restoration, has the advantages of being inexpensive, renewable, and non-toxic.157 Engineering these biomaterials with the use of various functionalities has resulted in increased adsorption capacity, more effective pollutant degradation, and lessening the risk of secondary contamination from the leaching of organic compounds. The conventional treatment train approach to environmental cleanups may be avoided with the use of modern biodegradable adsorptive materials, which allow for the degradation and removal of the pollutants to be treated in the same system.158

6. Functionalization of bio-based materials

Biopolymer alterations, in addition to heat processing, have resulted in a more sophisticated material structure design and broader spectrum of environmental cleaning capabilities.159 Adsorption biomaterials have recently concentrated on the development of 3D materials, as well as the achievement of a high surface area via biomaterial design (Table 1). A few of the advantages of functional bio-based materials goods include biodegradability, cost-effective, biocompatibility, environmental friendliness, and renewable resources. These newly advanced biomaterials have shown exceptional functional features, such as improved complexation, chelation, flocculation, adsorption and separation.159 The reactive carboxymethyl groups, cellulose, lignin and alginate functional materials in biomass have enhanced its adsorbent ability owing to their chemical reactivity, good solubility, and significant chelating ability.
Table 1 Functionalized bio-based materials for effective water remediation
Bio-based materials Primary component Pollutant Efficiency Removal time Mechanism of removal Synthesis method References
Poly(ethyleneimine)-graft-alkali lignin/lanthanum hydroxide Lignin Phosphate ion 95% 60 min Ligand exchange and surface precipitation Facile fabrication 159
Poly(ethylenimine)-functionalized cellulose microcrystals Cellulose Per-and polyfluoroalkyl substances 80–85% 120 min Physical adsorption and electrostatic interaction Oxidation and ion exchange followed by amine-functionalization 160
Cross-linking cellulose nanocrystals Cellulose Methylene blue 86% Electrostatic attraction Freeze-drying 161
Carbonized loofah/tin(IV) sulfide Loofah plant Cr 99.7% 120 min Physical adsorption and photocatalytic Scalable method 162
Crosslinked carboxymethyl cellulose grafted carboxymethyl polyvinyl alcohol Cellulose Cu+2 95% 240 min Ion exchange Casting method 163
Carboxymethyl cellulose/sodium styrene sulfonate gels Cellulose Cr, Pb, Mn and Fe 68.0%, 41.4%, 35.2% and 33.5% Electrostatic attraction Radiation grafting 164
Lignin sulfonate-based mesoporous materials Lignin Malachite green 97% 240 min Electrostatic interaction Grafting method 165
Carboxycellulose nanofibers Cellulose Cadmium(II) 84% Electrostatic interaction Nitro-oxidation method 166
Graphene oxide/esterified cellulose nanofibers Cellulose Ofloxacin and ciprofloxacin 96.9% and 97.8% 240 min π–π interactions, and hydrogen bonding, electrostatic interaction Chemical treatments followed by esterification 167
Cellulose nanofibers/PVA blend Cellulose Cationic and anionic dyes 84% Electrostatic interaction In situ 168
TEMPO-oxidized cellulose nanofibers Cellulose Cooper ions 100% Electrostatic interaction Grafting method 169


7. Environmental remediation by using bio-based materials by adsorption

Adsorption is widely utilized to remove contaminants from water supplies in the water treatment industry.170 Contaminants are trapped on the surface of an adsorbent material, and then removed from the water by this procedure. Adsorbents might be in the shape of grains, pellets, or powder, all of which provide a large surface area and boost adsorption efficiency. Many types of commercial adsorbents are used in water treatment applications. Activated carbon's high adsorption capacity and ability to remove a wide variety of contaminants, including organic compounds, disagreeable smells, and substances that contribute to poor taste,171 make it a popular adsorbent. Silica gel, zeolites, and clay minerals are a few more examples of commercial adsorbents. These adsorbents feature diverse adsorption characteristics that make them suitable for the removal of various pollutants.172 Despite their shown efficacy, commercial adsorbents are not without their drawbacks. Limitations include the high costs of implementing such systems, which become especially problematic in large-scale implementations. Adsorbent regeneration may be a costly and time-consuming procedure, depending on the specific circumstances.

Significant progress in using bionanocomposites as adsorbents has been made in recent years.173 Hybrid materials known as bionanocomposites combine biopolymers with nanoparticles to enhance their adsorption properties.174 Nanoparticles, with their huge surface area and functional capabilities, are combined with biopolymers, which are biocompatible and renewable, in these composite materials. The scientific community is very interested in chitosan-clay nanocomposites because they are cheap, easy to produce, and effective adsorbents. This class of composites has been found to successfully remove up to 99% of hazardous anionic ions, metals, and dyes from water.175 Additionally, the effectiveness of graphene oxide-potato starch composites in the adsorption of MB dye from industrial effluents has been investigated. The composite material's adsorption capacity was rather high, at around 90%.176 The composite material formed a structure similar to plywood, with nanocages, from the combination of GO nanosheets and polysaccharide long chains. According to the study, this structural configuration greatly enhanced the material's ability to absorb organic colors. Electrostatic interactions between graphene oxide and cationic dyes (a) play a role in the adsorption process, and (b) – stacking interactions between the aromatic part of the dye and the delocalized -electron system of GO play a role in the adsorption process.177

Graphene oxide shows surface complexation with heavy metal ions, enhancing their adsorption. When used in the adsorption treatment of wastewater, graphene oxide-chitosan composites demonstrate remarkable stability and mechanical qualities. Chitosan is able to efficiently coagulate pollutants out of aqueous solutions because it is a positively charged biopolymer with amino and hydroxyl groups. However, challenges such as the high synthesis costs, the necessity for pH adjustment, and decreased efficacy at low concentrations prevent their widespread application. The adsorption properties of the specific heavy metal ions are the primary focus of the current research. However, studies on the treatment of wastewater containing a combination of heavy metal ions are lacking.178,179 Water contamination by a wide variety of organic and inorganic compounds has been effectively adsorbed by AP-g-3D GO composites.180 Adsorption capacities for the dyes rhodamine B and Congo red were found to be particularly high in a chitosan-poly(vinyl alcohol)-graphene oxide cross-linked sponge.181 The adsoption – interaction of the aromatic rings of the dye molecules was enhanced by the 3D porous structures of the biopolymer sponge and the composite sponge.

According to a recent study, the amalgamation of a geopolymer and a biopolymer has been proven to improve the adsorption effectiveness of Ni(II) and Co(II) ions in wastewater. Synergy between lateritic geopolymer and the Tiliaceae family's Grewia biopolymer led to the effective removal of almost 80% of Ni(II) and Co(II) ions from the wastewater samples. Iron, aluminium, and silica oxides are concentrated in the geopolymer produced from laterite clay. The Grewia plant biopolymer, on the other hand, is made up of simple sugars, including mannose, xylose, glucose, glucuronic acid, and arabinose. Kaolinite platelets make up the structure of the laterite-clay geopolymer, whereas open holes of varying sizes are seen in the Grewia biopolymer. The improved adsorption capabilities of the Grewia biopolymer are thought to result from these structural properties.182

An extensive study was done on the use of graphene oxide/chitosan composites in the treatment of wastewater via adsorption. Biodegradable and cationic biopolymer chitosan has been shown to efficiently remove pollutants from liquid solutions via its coagulation properties. However, their use is limited because of the high manufacturing costs, the requirement for pH regulation, and the inefficiency when employed in low concentrations. The research studies presented here give helpful insights into the contaminant adsorption capacities of various graphene oxide/chitosan composites. The adsorption capacity of the GO-cl-potato starch bio-composite for MB dye was remarkable at 500 mg g−1 (90%).176,183

7.1 Dye removal by BBMs

Because of their high specific surface area, three-dimensional network structure, sorption capacity, and recyclability, BBMs have been suggested as very efficient adsorbents for the removal of dangerous chemical substances. Recent research has focused on the use of BBMs to remove organic dyes from wastewater because of their cheap operating cost and high efficiency. Mutations, skin irritation, and allergy in humans are only some of the effects that have been linked to dye pollution in water. Consequently, it is crucial to treat dyes using different processes, including oxidation, adsorption, coagulation, reverse osmosis, biodegradation, and so on. Researchers have been interested in developing BBMs for the degradation of organic dyes without negatively impacting the environment from the production of harmful byproducts. The adsorbents are designed to promote electrostatic contact between the dyes via their ionic charges. Electrospinning generates the ultrapure chitosan nanofibrous membranes used in acid blue-113 dye removal. The optimal conditions for nanofiber structure and morphology were determined to be a reduction in the nanofiber diameter, which led to an increase in the adsorption capacity owing to the increased surface area, and a maximum adsorption capacity of 1338 mg g−1.184 Hydrophilicity, biocompatibility, and biodegradability were all enhanced when chitosan and polyamide-6 were combined and electrospun the resulting nanomaterials. Tests were developed utilizing response surface methods to select the optimal filtering parameters to remove 95% of Polar Yellow GN and 96% of Solophenyl Red 3BL. The purpose of fusing SiO2 nanoparticles onto the polyvinyl alcohol (PVA)/chitosan composite was to make the fiber highly stable and more efficient for a longer period of time. Due to the increased surface area (60.76 m2 g−1), higher compaction resistance, and decreased degradation rate (2.65) brought about by the incorporation of nanoparticles, the electrostatic affinity between the anionic dyes and amine resulted in 98% elimination of the DR23 dye at a flux rate of 1711 L m−2 h−1.185 In a separate investigation, Mahmoodi and colleagues tested the efficacy of PVA-chitosan nanofibers in removing dye from a colored effluent. Decreases in pH improved the dye removal efficiency because protonation was decreased. On the other hand, at high pH, desorption is preferred, with release efficiencies of 85–90%.186 Dye recovery performance was also produced by the functionalization of chitosan fibers with ethylenediamine. The thermogravimetric curves showed that ethylenediamine's low boiling point caused functionalization to lower the degradation temperature. A rise in pH allowed for the recovery of both cationic and anionic dyes on the fibers, with enhanced adsorption of MB and reduced adsorption of coccine owing to the protonation-preferred adsorption of anionic dyes.187 Following this, fibers were electrospun, and chitosan hydrogels were coated to increase their density. Hydrocarbon bonds have a binding energy of 284.32 eV; the carbon–oxygen bond has a binding energy of 286.83 eV; ester groups and alcoholic hydroxyl groups both have a binding energy of 289.31 eV; and the carbon–oxygen bond has a binding energy of 289.31 eV. Numerous carbonyl and alcoholic OH point to the introduction of (multi-walled carbon nanotubes) MWCNTs onto cellulose. Coated materials were able to eliminate 90% of the dye with a molecular mass greater than 600 g mol−1 at an operating pressure (0.5 MPa) and significant high flux rate (150.72 L m−2 h−1).188

Electrospinning between the shoulders is also used to create chitosan nanofibers functionalized with carboxylated CNTs for the removal of methyl orange and methylene blue. Fabrication of clay-based nanofibers for dye removal makes use of montmorillonite's hydrophilicity through a thermally induced sol–gel transition. Water flow (1765 L m−2 h−1), compaction resistance (0.4 bar), dye removal (95% of BB41), resistance to fouling, and reusability potential were all improved by the addition of 2.0 wt% of CNTs.189 Xu et al. decorated clay nanosheets onto the chitin nanofibrous matrix, resulting in the construction of hierarchical architectures. Dyestuffs having a large specific surface area and a high cationic exchange capacity were adsorbed onto the microsphere surface, transported via the chitosan channels, and subsequently interacted with clay sheets. Surprisingly, methylene blue's adsorption behavior showed very minor changes depending on pH. Active adsorption sites were not completely exposed when chitosan fibers were dissolved into the solution of binary solvents.189 Xing et al. created sponges by means of thermally induced phase separation, by utilizing tetrahydrofuran in the water/acetic acid solvent system. THF is not the best solvent, but it did an excellent job of opening up more room for the lean phase to develop and showing off the sponge's porous, interconnected structure. The sample with 1[thin space (1/6-em)]:[thin space (1/6-em)]79[thin space (1/6-em)]:[thin space (1/6-em)]20 (AA[thin space (1/6-em)]:[thin space (1/6-em)]water[thin space (1/6-em)]:[thin space (1/6-em)]THF) and 0.5 wt% chitosan was quenched at 196 °C, and had the highest adsorption capacity of 604.7 mg g−1.190

Hydrogels resistant to disintegration in water were fabricated by scaling up the process of making adsorbents based on the physical entanglement and hydrogen bonding of nanofibrils. Because of their high specific surface area and surface carboxyl groups, aerogels prepared from cellulose nanofibrils through a freezing-thawing process are able to remove 92% of MG at a 10[thin space (1/6-em)]:[thin space (1/6-em)]5 mg mL−1 aerogel/MG w/v ratio, and the adsorbed dye is rapidly released when the ionic strength of the solution is increased up to 200 mm NaCl.

Sajab et al. added Fe(III) ions to graphene oxide-cellulose nanofibril aerogels. As a result of the Fenton oxidation process, methylene blue was degraded in the presence of hydrogen peroxide, and the loaded fibers were able to remove the dye at a rate that was 30.4% greater than that of the unloaded fibers.191

Chemically crosslinked cellulose/graphene oxide nanofibers containing photocatalytic TiO2–NH2 nanoparticles (NPs) have remarkably improved dye-degradation. Titanium dioxide (TiO2), when exposed to ultraviolet light, generates an active oxidation species (hydrogen peroxide), which in turn degrades indigo carmine and methylene blue. Power density, pH, duration, and starting dye concentration all had significant impacts on the degradation efficiency.192 Due to its unique conical shape that is both hydrophobic on the inside and hydrophilic on the outside, the oligosaccharide cyclodextrin is favored. Enhanced adsorption and improved mechanical strength owing to hollow cavity were achieved by incorporating the B-cyclodextrin polymer into poly-caprolactone during electrospinning. Upon the completion of up to eight cycles of readsorption, the maximum efficiency was 24.1 mg g−1 adsorbing 78%.193

Researchers have successfully adsorbed crystal violet and rhodamine B (RB) onto graphene oxide nanosheet-coated dialdehyde starch nanocrystals in aqueous solutions (CV). There was a maximum adsorption capacity of 539 mg g−1 for RB and 318 mg g−1 for CV in aerogels.194 Dye-mediated interaction between the CS-based polyelectrolyte and organo-montmorillonites allowed for the effective and quick removal of cariogenic industrial dyes in an employed CS-based polyelectrolyte/organoclay hybrid. Using a CS/organoclay hybrid with a weight ratio of 10/90, researchers were able to completely remove a wide range of hydrolyzed anionic dyes.195 Polycyclic aromatic hydrocarbons (PAHs) may be extracted from wastewater with the use of magnetic nanoparticles encased in chitosan and pyrolyzed chitosan. The chitosan-based polymer has a good adsorption capacity for both naphthalene and anthracene.196 In their production of a bifunctional alginate-based composite hydrogel, Qian et al. accomplished synergistic photocatalytic and adsorption degradation and decomposition of organic contaminants. Malachite green crystal violet, and methylene blue each had very high adsorption abilities of 3000.08, 993.29, and 1610.34 mg g−1 in the alginate -based hydrogel, respectively.197

Corn starch, 3-chloro-2-hydroxypropyl trimethylammonium chloride, and epichlorohydrin were used in the production of crosslinked cationic starch by Guo et al.198 A maximum adsorption capacity of 208.77 mg g−1 at 308.15 K was obtained by crosslinking cationic starch when it was employed to remove reactive golden yellow dye from wastewater channels. Crosslinking PVA functionalized with chlorosulfonic acid-based vinyl acetate onto crude starch to provide functionality led Pourjavadi et al.199 to create the MNPs@Starch-g-poly(vinyl sulphate) nanomaterial. After five adsorption–desorption cycles, the regenerated adsorbent is able to remove up to 90% of MG and MB dyes from solution, demonstrating its effectiveness against these cationic dyes in water. Adsorption capabilities of 567 and 621 mg g−1 were established initially.

Table 2 provides a quick overview of the many bio-based materials that may be utilized to effectively remove dyes.

Table 2 Bio-based composites for the effective removal of organic dyes
Bio-based composites Pollutant Contact time pH Adsorption capacity Synthesis method References
Hemicellulose-based adsorbent Malachite green (MG) dye 60 min 6.5 456.23 mg g−1 Periodate oxidation and acetylation 200
Cellulose-based flexible carbon aerogels Methylene blue 60 min 8 Hydrothermal 201
Cellulose-based adsorbents Congo red, Eriochrome blue 200 min and 100 min 7 380.084, 349.284 mg g−1 Free radical polymerization method 202
Cellulose and gelatin-based hydrogel Ethidium bromide and eosin 30 min 10 Grafting of poly(acrylic acid) 203
Copolymerizing acrylic acid (AA)-hydroxyethyl methacrylate (HEMA)-sodium alginate Brilliant cresyl blue 100 min 94% Free radical polymerization method 204
Magnetic lignin Titan yellow Congo red 180 min 7 192.51 and 198.24 mg g−1 205
Starch/activated carbon composite Methylene blue 90 min 10.5 90% Carbonization 206
Pd supported on nanocellulose-alginate hydrogel Methylene blue 5 min CNCC and sodium alginate combination solution drops gradually added to CaCl2 solution 207
Bacteria on polycaprolactone/polylactic acid fibers Setazal blue BRF-X 95% Electrospun 208
Polyvinyl alcohol–chitosan Direct red 80 1.5 h 2.1 151 mg g−1 Electrospun 186
Chitosan/zeolite/polyvinyl alcohol MO 14 min 8–11 153 mg g−1 Electrospun 209
Silica-sand/cationized-starch composite Acid green 25, methyl orange 20 min 6.8 and 5.8 912.6, 458.7 mg g−1 One-step etherification reaction 210
Gelatin/activated carbon composite beads Rhodamine B 45 min 4 256.41 mg g−1 Facile method 211
Cellulose acetate/graphene oxide nanofibers Indigo carmine 150 min 2 99.8% Electrospinning 192
Cassava starch-based hydrogel Methylene blue 45 h 2000 mg g−1 212
Chitosan/polyamide Reactive black 5 4 h 1 456.9 mg g−1 Force-spinning technology 213
Chitosan–gelatin hydrogel loaded with ZnO Congo red 8 h 10 Grafting of acrylamide 214
Polyvinyl alcohol/silica/chitosan composite Direct red 80 1.5 h 2 322 mg g−1 Electrospinning 215
Starch-coated Fe3O4 NPs Optilan blue textile dye 2 50 mg L−1 Green synthetic approach 216
Fe3O4 NPs/pectin MO Co-precipitation 217
CNCs grafted with acryloyloxyethyltrimethyl ammonium chloride Neutral reactive blue 19 7 80% Polymerization 218
Starch-doped Fe2O3 nanostructures Methylene blue 120 min 4, 7 Co-precipitation 219
Bacterial cellulose@CdS nanocomposite Methylene blue 180 min 77% An “anchoring-reacting-forming’’ pathway 206
Polyamide-6/chitosan Solophenyl red 3 91% Electrospinning 220
Cellulose aerogel/trimethyl-ammonium chloride Blue dye 19 30 min 7 160 mg g−1 Freeze drying and chemical crosslinking 221


7.2 Micropollutants removal

A variety of procedures have been used to remove micropollutants (heavy metals pharmaceutical and textile effluents) from water supplies, since their accumulation in humans may be very dangerous if it exceeds acceptable limits. BBMs are viewed as beneficial in this context for micropollutant removal due to their malleability, highly selective surface area, and improved adsorption rates (Table 3).
Table 3 Bio-based composites for the effective removal of micropollutants
Bio-based composites Pollutant Contact time pH Adsorption capacity Synthetic method References
Graphene/lignin/sodium-alginate nanocomposite Pb(II) and Cd(II) 90 min 6 226.24 mg g−1 and 79.88 mg g−1 Hydrothermal polymerization 235
Nanocellulose/poly(vinyl alcohol-co-ethylene) Pb(II) 24 h 4 471.55 mg g−1 Melt blending extrusion 236
Graphene oxide microcrystalline cellulose Metformin   4.5, 6.5 and 8.5 132.10 mg g−1 Acid hydrolysis, chemical oxidation cum exfoliation 237
Polyethyleneimine functionalized chitosan–lignin Hg(II) 120 min 5.5 663.5 mg g−1 Crosslinking 238
Dual heteroatom NB-co-doped lignin-based biochar 4-NP 12 h 11 One-pot carbonization 239
Poly vinylalcohol/chitosan/zinc oxide-NH2 Cd(II) 240 min 6 1.239 mmol g−1 Cast and electrospun method 240
γ-Fe2O3@starch As(III) 120 min 9 96 & Co-precipitation 241
Chitosan/PVA Tetracycline 300 h 6 102 mg g−1 Electrospun method 232
Starch, CMC-stabilized nano zero-valent iron Sulfamethazine 60 min 5 & 9 83% Reduction method 242
Phosphate-decorated carboxymethyl and phosphorylated chitosan (CSP) cellulose U(VI) 5 977.54 mg g−1 Crosslinking 243
TiO2 supported on chitosan scaffolds Amoxicillin 180 min 7 50% Sol–gel transition 244
Lignin nanofibers Fluoxetine 120 min 4.5 185 mg g−1 231
Ethylenediamine modified pectins Pb 4 94% 245
Magnetic starch-based composite adsorbent Hg(II) 1.5 h 4–7 324.42 mg g−1 Co-precipitation 246
Thiolated-spherical nanocellulose Hg(II) 20 min 5.6 98.6 mg g−1 Acid hydrolysis 247
Starch/Fe–Mn binary oxide As(V) and As(III) 120 min 7 160.63 and 284.64 mg g−1 Redox and coprecipitation reaction 243
Cellulose acetate/silk fibroin Cu(II) 2 h 75.91 mg g−1 Electrospinning 248
Chitosan/polyethylenimine Pb and Cd 7 341 mg g−1 and 321 mg g−1 249
Crosslinked starch polymer Acetophenone 7 180.2 mg g−1 Facile one-pot synthetic route 250
Sodium alginate/graphene oxide composite beads Ciprofloxacin 48 h 4 86.12 mg g−1 Magnetic stirring 251
3D alginate-based MOF hydrogel Tetracycline 8 364.89 mg g−1 One step method 252
Alginate–protein cryogel beads Immunoglobulin G (lgG) 90 min 5 175 mg g−1 Extrusion dripping method 253
Chitosan/alginate/Fe3O4@SiO2 hydrogel Pb 120 min 6 234.77 mg g−1 254
Tricarboxylic cellulose nanofiber Cu(II) 2 h 92.23 mg g−1 255
f-MWCNTs/FeCl3·4H2O Ibuprofen 120 min 7 1.15 and 11.8 mg g−1 Solvothermal 256
Chitosan, hemicellulose, TiO2 Ni(II) 4 370.4 mg g−1 Self-assembly, sol–gel method and polymerization 257
UV-cured chitosan and gelatin hydrogels Pb2+, As5+ 6 Microwave-assisted method 258


Micropollutant removal from wastewater at a reasonable cost has emerged as a major area of study. Micropollutants may be selectively removed at low concentrations, and the source materials are the most common biopolymers, making BBMs a promising option for separating and recycling these micropollutants and preventing secondary contamination. Specifically, the highest adsorption capacity of pure chitosan nanofibers was found to be 30.8 mg g−1 at acidic pH, which is equivalent to the widely testified CS adsorbents for As(III) elimination. Doping iron (Fe) into chitosan nanofibers improved the specific surface area, exposing a higher number of potential active adsorption sites on the adsorbent substrate and improving As(III) selectivity in the metal ion. It is revealed that the presence of CN(CC) groups on the adsorbent surface facilitates As(III) removal.222 Crosslinking glutaraldehyde and immersing epichlorohydrin resulted in amine grafted on the chitosan nanofiber membranes for electrospinning removal of copper ion from wastewater (Fig. 7a and b). Incorporating amine groups through diethylenetriamine gave the nanofibers an increased stability (with just 6% weight loss after 24 h), which may have been induced by the imine and amine synthesis (Fig. 7c). Fig. 7d shows the loss in weight at varying temperature. Higher quantities of active nitrogen groups in AGNFs result in a stronger chelation adsorption of Cu(II), with maximum adsorption capacities of 166.67 mg g−1 and 166.67 mg g−1 from aqueous solution, respectively.223


image file: d3ra06501d-f7.tif
Fig. 7 (a) Methodology for optimizing the electrospinning parameters and preparing the NFs membrane. (b) Fabrication steps for the synthesis of novel AGNFs. (c) Stability curve of the functionalized membrane. (d) TGA thermograms of the prepared novel membrane.223 Copyright (2018), Elsevier.

Thiol-functionalized cellulose nanocrystals/PVA/chitosan nanofibrous films displayed improved mechanical characteristics for the adsorption of Cu(II) and Pb(II) ions, with recovery efficiencies of the cellulose nanocrystals (CNC)-containing films reaching 90.58% and 90.21%, respectively.224 The high surface area of SNC combined with the selectivity of the thiol groups towards the mercuric ions resulted in an 86% improvement in efficiency when compared to untreated nanocellulose.224 Then, a cellulose acetate nanofibrous membrane hybridized with bio-based silk fibroin was studied for its ability to adsorb heavy metals. Adsorption was generated by the binding of an electronegative COOH group and an amino group in creating a coordination covalent bond with Cu(II) ions. The crosslinked fibers containing 20% CA recovered 75.91 mg g−1 of copper ions after 60 minutes of adsorption.225 It was observed that adding carboxyl groups and MnO2 nanoparticles to the tree-shaped cellulose nanofibers increased the surface area by a factor of 2.3 and enhanced the adsorption of the Cu(II) ions to 399.14 mg g−1. To create nanofibers, waste pulp with a low cellulose and chitosan content was physically disintegrated before being oxidized with TEMPO to enhance the COO groups for superior adsorption. Both pH (up to 6.2) and oxidation level (26%) were shown to promote adsorption. Bio-based nanofibers can recover about 90% of their original mass after five cycles, which is similar to commercially available montmorillonite nanoclay adsorbents (57 mg g−1). Over 94% of the metal ions were captured in the fibers, demonstrating the enormous ability of using CS mixed PVA nanofibers for heavy metal elimination polluted water channels.226 Carbon-based polyacrylonitrile nanofibers had their carboxyl groups functionalized with chitosan and citric acid. Stability was achieved in adding functional groups to the fibers through pretreatment, peroxidation, and hydrothermal carbonization. When citric acid was added in increments of 0.5 g, the adsorption capacity increased until it reached 1.0 g, and then it began to decrease. Therefore, incorporation during fiber preparation is optimal. After 5 cycles of adsorption and desorption, the removal efficiency decreased from 80% to 70%. The adsorption of Pb(II) ions was increased by 500% and that of Cd(II) ions by 3000% due to the larger surface area when BBMs was functionalized with oxolane-2,5-dione. Because of this modification to the surface, the efficiency of removal is greatly enhanced while reusability is not compromised.227

The dumping of chromium ion into water streams from textile industries has been increasing at an alarming pace, demanding the use of highly effective and selective adsorbents. To combat this issue, researchers have developed aerogels made of cellulose nanofibers grafted with quaternary ammonium. The gel capacity (17.66 mg g−1) is very close to that of commercially available quaternary ammonium anion exchangers.227 A synthetic polymeric material, i.e., polyurethane (PU) was utilized as a matrix to bind carboxymethylated cellulose nanofibrils to avoid accumulation and enable the fibrils to freely interrelate with adsorbate, which surprisingly influenced the adsorption rates. Because of the improved malleable strength of the PU combined CMCNFs, the strain-to-failure ratio rose, revealing the interfacial contacts between the matrix and the fibrils. The metal ions Cu(II), Pb(II), and Cd(II) adsorption efficiencies on the PU/CMCNFs were 40.2%, 95.1%, and 61.3%, respectively, while neat-PU adsorption efficiencies were extremely low (1%), demonstrating that the NH2 groups on polyurethane did not contribute to the adsorption.227

Using a spiral wound module to sustain electrospun chitosan nanomaterials improved both membrane surface area to packing density and volume ratio. Adsorption of heavy metal ions is proportional to the density of the chitosan nanofiber input flow rate and the deposition. The flow rate of 0.9 L h−1 is similar to that of commonly available nanofiltration membranes, and the nanofibers were able to retain over 90% of chromium (Cr(VI)). The specific adsorption of chromium(VI) over other heavy metal ions was also shown, with peak desorption occurring in less than 2 minutes in both alkaline and acidic conditions, indicating a high regeneration potential.228 In another study a poly(ethylene oxide)/chitosan complexed permutit was developed for treatment of metal ions. In this case, PEO is meant for enhanced blending, chitosan is meant for chelation, and permutit is meant for adsorption; each polymeric fiber is unique and may provide a combinatory effect. They concluded that the initial chromium concentration and integration of low-quantity permutit are suitable for adsorption. It involves exchanging Na+ in the solution for H+, which prevents the formation of H2CrO4 and the electrostatic interactions of the ionized (NH3+, OH2+) molecules with HCrO4. The BBMs were 90.44% effective in removing water from a solution containing 50 mg L−1.229 The same team of scientists used chitosan, polyvinyl alcohol, and polyethylene glycol to generate several nanofibrous layers. By integrating the aminated Fe3O4 nanomaterial into the BBMs, they improved their water resistance and thermomechanical properties. As the amount of bio-based material increased, the number of active sites inside the nanofibers capable of chelating metal ions for removal subsequently increased. Maximum adsorption capabilities of 525.8 and 509.7 mg g−1 for Pb(II) and Cr(VI) ions were determined, respectively. Regeneration cycles showed the membranes to be highly reusable for industrial adsorption applications.230

To combat these problems, researchers reported on the synthesis of chitosan nanofibers by a thermally-induced phase separation method at low temperatures. Nanofibers were produced at relatively high concentrations of chitosan solutions, and hole growth in the membrane was induced by a solution of water, acetic acid, and ethanol. An ethylenediamine tetraacetic acid (EDTA) solution was used to remove heavy metal chelation by amine groups on the fiber surface. After six washings in water, the chitosan nanofibers still managed to hold on to 90% of the Cu(II) that had been dissolved in the water. The method of production not only improved productivity, but also showed remarkable recyclable potential.230

Biopolymer adsorbent development lessened some of the risks associated with the aforementioned pollutants, but new contaminants continue to be a challenge for water preservation. Human waste was a major factor in the overall pollution problem. Electrospun lignin nanofibers were characterized by the presence of functional groups, such as ethers, alcohols, and aromatic rings by Camera and colleagues, who found that they removed 70% of the pharmaceuticals pollutant fluoxetine from wastewater (van der Waals forces, electrostatic interactions, – stacking and hydrogen bonds).231 Chemical crosslinking by glutaraldehyde successfully removed tetracycline, another significant pharmaceutical contaminant, from chitosan combined with PVA nanofibers. A strong adsorption capacity (102 mg g−1) was discovered when the various concentrations were adjusted to a volumetric ratio of PVA/chitosan (25[thin space (1/6-em)]:[thin space (1/6-em)]75). Ultraviolet (UV) light degrades bisphenol A and diclofenac from TiO2 nanoparticles immobilized on PVA nanofibers/electrospun gum karaya.232 Methane plasma-treated TiO2 membranes had increased sorption and hydrophobicity capacity because of the development of water droplets on their surface, which resulted in the reduction of 20% of diclofenac and 18% of bisphenol A because of the formation of free radicals in response to UV light.233 Water beaded up on cellulose nanofibers treated with trichloro(heptadecafluorodecyl) silane because of the resultant hydrophobicity. This property is the subject of investigation as a possible means of extracting oil from water. Nanofiber membranes that have been modified in this way have a 99% effectiveness rate at separating oil and water merely by utilizing gravity234 Increases in hydrophobicity and penetration fluxes for a broad range of oils were achieved using silver nanoparticles, polydopamine, and -cyclodextrins on PLA nanofibers, with a separation performance of 95% being considered satisfactory.

7.3 Oil spills

Worldwide interest in water–oil separation has been inspired by the serious environmental contamination caused by the discharge and leakage of organic compounds from crude oil. The Deepwater Horizon oil disaster in 2010 released an estimated 5 million barrels of crude oil into the Gulf of Mexico.259 The oil and gas sectors create over 0.250 billion barrels of sludgy sewage per day. In light of this, there has been a lot of focus on developing new, more eco-friendly oil-absorbing materials at lower costs for use in oil spill cleaning. BBMs are the greatest raw material choice for making water–oil separation products that are cheap, efficient, and safe for the environment. Oil–water separation materials developed from BBMs with hydrophobic/hydrophilic interface structures have been intensively researched and used by scientists in recent years, with positive results (Table 4). The dopamine-cellulose surface was modified by fabricating a stimulus-responsive azeobenzene-fluorosilane compound and grafting it onto the surface. The hydrophilic surface of the cellulose-based substance was brought about by the isomerization impact of UV radiation. The biomaterial's ability to resist organic pollutants, bacteria, and fungi, and its impressive capability for adsorbing oil from O/W combinations make it an attractive choice for wastewater treatment. Ejeta et al. developed a cotton-based superwetting material with superoleophilicity (oil-loving) and superhydrophobicity (water contact angle > 150°) to separate water-in-oil emulsions (oil contact angle 5°). Extremely high infiltration fluxes of 10[thin space (1/6-em)]400 L m−2 h−1 and 867[thin space (1/6-em)]500 L m−2 h−1 for W/O emulsion were observed in the resulting SCM after filtering by gravity and external pressure, respectively.260–262 Since they are durable in chemically aggressive conditions and recyclable, superwetting materials for water–oil emulsion separation are a practical choice for widespread industrial application.
Table 4 Bio-based composites for the effective adsorption of oils
Bio-based materials Oil Hydrophobic modification Adsorption capacity Method References
Carboxymethyl chitosan Marine diesel Monochloroacetic acid Partial carboxymethylation 264
H-Oleoyl-carboxymethyl chitosan Wastewater of oil extraction Oleoyl chloride 95% Carboxymethylation and acylation 265
Modified activated carbon aerogel Gasoline, organic solvent Polydimethylsiloxane 4.06–12.31 g g−1 Polymer coating 266
Sodium salt of oleoyl carboxymethyl chitosan Marine diesel Oleoyl chloride 75–85% Carboxymethylation and acylation 267
Chitosan-silica hybrid (nano-sized) Toluene, cyclohexane, n-heptane, chloroform Tetraethyl orthosilicate, 3-(triethoxysilyl)propyl isocyanate 90% Sol–gel encapsulation 268
Cellulose nanocrystals (nano-sized) Marine diesel Choline chloride and oxalic acid dihydrate Deep eutectic solvents 269
Polyvinylpyrrolidone-coated magnetite nanoparticles (nano-sized) MC252 oil ∼100 Hydrothermal method 270
Cellulose nanofibril and polyvinyl alcohol Crude oil Span-80 54–140 g g−1 Emulsification and freeze drying 271
Blends of poly(vinyl alcohol) nanoparticles with chitosan or starch (nano-sized) Toluene, kerosene and hydraulic oil 48.7 g g−1, 39.3 g g−1 and 22.7 g g−1 Emulsion polymerization 272
Polyurethane/chitosan foam Castor oil Chitosan and ricinoleic acid 267.24 mg g−1 Polymerization 273
H-PAA-T Insoluble oil Chitosan and polyacrylic acid 135.9 mg g−1 and 990.1 mg g−1 Thermal cross-linking 274


Epoxidized soybean oil was impregnated into cellulose aerogels (GA) so they could soak up oil (ESO). The water contact angle of the cellulose-based aerogels was increased to 132.6 degrees after hydrophobicity adjustment by ESO. An aerogel that has been adsorbed maintains almost 90% of its adsorption capacity after 30 absorption–desorption cycles, proving its use in W/O separation of industrial wastewater and oil spill removal.263 Recently, researchers working on BBMs for water–oil separation shifted their attention to how to control the surfaces of hydrophilic and hydrophobic interfaces.

7.4 Membrane filtration

The membrane separation technique is well established and effective for wastewater treatment due to its comfortable working conditions, size-sieving ability, and fuel efficiency. Water filtering membranes composed of bio-based nanoparticles have a lot of untapped potential because of their high surface-to-volume ratio, extraordinary mechanical capabilities, low ecological effect, and sustainable development. Polymer matrices, including poly(ether sulfone) (PES), poly(vinyl alcohol) (PVA) and poly(vinylidene fluoride) (PVDF), were often used with BBM nanocomposite-based membranes to increase the membrane characteristics like selectivity, permeability, tensile strength, and anti-biofouling.275–278

Recently, a vacuum filtration approach was employed to synthesize graphene oxide nanocolloid (nanoGO) biohybrids and TEMPO-mediated oxidative cellulose nanofibers for the purpose of developing self-assembled bio-based membrane for water purification. In the biohybrids, TOCNF has a large amount of CCOH groups that facilitated Cu(II) adsorption, while NanoGO acted as nodes and joints, vastly enhancing the BBM membranes' flexibility, mechanical strength, hydrolytic stability and water flux.279 By constructing an ultrathin GO barrier layer on a cellulose nanofiber (CNF) membrane without the use of any crosslinker, high-flux water purification membranes may be made with a water flux rate of 18[thin space (1/6-em)]124 L m−2 h−1 bar−1 and a dye infiltration rate of 90 L m−2 h−1 bar−1.280,281 Meanwhile, 2-methyltetrahydrofuran and glycerol derivatives were used to create asymmetric CA full-bio-based NF membranes via non-solvent induced phase separation. Increasing the concentrations of 2-MeTHF and CA led to the highest performance, with porosity between 12.8 and 5.5 L m−2 h−1 bar−1 and RB removal of >90%.282

The cost-effective membrane developed by BBMs has vast possible applications and promising future growth in the area of wastewater resource recovery. The BBMs membrane has low strength, which results in a limited life, and is readily polluted by waste during the industrial reaction process, which further reduces its usefulness. Therefore, high-strength, green chemistry and antifouling thin film materials should be explored as part of future research into BBMs filtering membrane for industrial wastewater treatment.

8. Conclusion and future perspective

The next wave of readily available resources is undoubtedly going to be made up of natural biopolymer-based resources (especially polysaccharides) produced from various derivatives. Many researchers have been drawn to the study of natural biopolymers because of their many desirable characteristics, including their abundance, nontoxicity, renewability, adaptability, low cytotoxicity, and potential for a wide range of applications. Importantly, biopolymer-supported Ag, Cu, Pd, ZnO, TiO2, Fe3O4, and carbon-based materials (GO, CNTs and nanotubes) are among the most successful techniques in nanotechnologies, and the abovementioned polysaccharides are regarded as substitute renewable bio-resources and excellent supports for the nanocatalysts manufacturing.

Natural or biopolymeric-based nanomaterials, such as the conventional adsorbents, require more research to improve their manufacture in order to be practical and long-lasting for use in industry. It has been proven possible to use nanomaterials based on natural/biopolymeric substances to aid in the flocculation/coagulation of solid compounds in wastewater treatment. Polysaccharides, such as starch, cellulose, chitosan, and pectin, show promise as natural/biopolymers due to their ability to compete with conventional activated carbon in terms of sustainability, reactivity, excellent physicochemical attributes, chemical stability, and significant specificity towards dyes, aromatic compounds and toxic metals.

Concerns about the quality of the environment and the wellbeing of its inhabitants have long made wastewater treatment a top priority for public health officials. The use of evolving nanostructured materials has the potential to improve industrial water remediation, as pollutants can be eliminated while using fewer resources than traditional methods. However, more research is needed in this area, as a major technical barrier is their inflexibility for engineering processes and lack of competitiveness compared to conventionally deployed existing water treatment options. The creation of bio-based nanomaterials using traditional procedures is fairly easy, facile, and innocuous to the environment. Nevertheless, new studies have discovered considerable promise for numerous applications of the biopolymer-based nanomaterials across varied study disciplines. From the perspectives of both ecological remediation and resource management, they thus have the capability to develop as a successful, cost-efficient, and environmentally friendly alternative to current water treatment technologies.

Despite significant progress in the catalytic and synthetic uses of natural biopolymer-based nanomaterials, the following should be given special attention in the next investigations:

• A low-cost, scalable method of producing nanomaterials from polysaccharides.

• Synthesis of biopolymers and its utilization in the wastewater purification process using a wide variety of bio-waste nanomaterials.

• For the production of nano membranes and their use in the wastewater treatment process, it is advantageous to make use of naturally occurring supports such as clays, zeolites, and montmorillonite since they are inexpensive, non-toxic, and readily accessible.

• Processing of animal and agricultural waste, such as bristles, bone, and eggshells, to create biopolymer-based nanomaterials for use in water purification.

• Utilization of greener biological procedures to lower the price of producing bio-based nanomaterials under moderate settings and boost their nano-catalytic efficacy.

• Enhanced magnetic nanoparticles based on polysaccharides for water/wastewater purification.

Conflicts of interest

There are no conflicts to declare.

References

  1. A. Boretti and L. Rosa, Reassessing the projections of the World Water Development Report, npj Clean Water, 2019, 2(1), 15 CrossRef.
  2. R. Li, et al., Porous Biochar Materials for Sustainable Water Treatment: Synthesis, Modification, and Application, Water, 2023, 15(3), 395 CrossRef CAS.
  3. Y. Liu, et al., Fabrication of multifunctional biomass-based aerogel with 3D hierarchical porous structure from waste reed for the synergetic adsorption of dyes and heavy metal ions, Chem. Eng. J., 2023, 451, 138934 CrossRef CAS.
  4. U. Haripriyan, K. Gopinath and J. Arun, Chitosan based nano adsorbents and its types for heavy metal removal: a mini review, Mater. Lett., 2022, 131670 CrossRef CAS.
  5. M. K. Uddin, et al., Pyrolysis of rubber seed pericarp biomass treated with sulfuric acid for the adsorption of crystal violet and methylene green dyes: an optimized process, Int. J. Phytorem., 2023, 25(4), 393–402 CrossRef CAS PubMed.
  6. Z. Pan, et al., Cr (VI) adsorption on engineered iron oxide nanoparticles: exploring complexation processes and water chemistry, Environ. Sci. Technol., 2019, 53(20), 11913–11921 CrossRef CAS PubMed.
  7. L. Hu, et al., Simultaneous recovery of ammonium, potassium and magnesium from produced water by struvite precipitation, Chem. Eng. J., 2020, 382, 123001 CrossRef CAS.
  8. J. A. Korak, A. L. Mungan and L. T. Watts, Critical review of waste brine management strategies for drinking water treatment using strong base ion exchange, J. Hazard. Mater., 2022, 129473 Search PubMed.
  9. J. Wang and S. Wang, Toxicity changes of wastewater during various advanced oxidation processes treatment: an overview, J. Cleaner Prod., 2021, 315, 128202 CrossRef CAS.
  10. S. Zhang, et al., Membrane-confined iron oxychloride nanocatalysts for highly efficient heterogeneous Fenton water treatment, Environ. Sci. Technol., 2021, 55(13), 9266–9275 CrossRef CAS PubMed.
  11. X. Zhang and Y. Liu, Reverse osmosis concentrate: an essential link for closing loop of municipal wastewater reclamation towards urban sustainability, Chem. Eng. J., 2021, 421, 127773 CrossRef CAS.
  12. R. Yin, E. R. Blatchley III and C. Shang, UV photolysis of mono-and dichloramine using UV-LEDs as radiation sources: photodecay rates and radical concentrations, Environ. Sci. Technol., 2020, 54(13), 8420–8429 CrossRef CAS PubMed.
  13. R. Mohammadi, W. Tang and M. Sillanpää, A systematic review and statistical analysis of nutrient recovery from municipal wastewater by electrodialysis, Desalination, 2021, 498, 114626 CrossRef CAS.
  14. S. Hand and R. D. Cusick, Electrochemical disinfection in water and wastewater treatment: identifying impacts of water quality and operating conditions on performance, Environ. Sci. Technol., 2021, 55(6), 3470–3482 CrossRef CAS PubMed.
  15. M. Peydayesh and R. Mezzenga, Protein nanofibrils for next generation sustainable water purification, Nat. Commun., 2021, 12(1), 3248 CrossRef CAS PubMed.
  16. G. Wei, et al., Biomass vs inorganic and plastic-based aerogels: structural design, functional tailoring, resource-efficient applications and sustainability analysis, Prog. Mater. Sci., 2021, 100915 Search PubMed.
  17. Y. Zhang, et al., Molecularly soldered covalent organic frameworks for ultrafast precision sieving, Sci. Adv., 2021, 7(13), eabe8706 CrossRef CAS PubMed.
  18. Y. Wen, et al., Metal-organic framework enables ultraselective polyamide membrane for desalination and water reuse, Sci. Adv., 2022, 8(10), eabm4149 CrossRef CAS PubMed.
  19. Y. L. Ji, et al., Superfast water transport zwitterionic polymeric nanofluidic membrane reinforced by metal–organic frameworks, Adv. Mater., 2021, 33(38), 2102292 CrossRef CAS PubMed.
  20. L. Shen, et al., Highly porous nanofiber-supported monolayer graphene membranes for ultrafast organic solvent nanofiltration, Sci. Adv., 2021, 7(37), eabg6263 CrossRef CAS PubMed.
  21. P. Shao, et al., Potential difference driving electron transfer via defective carbon nanotubes toward selective oxidation of organic micropollutants, Environ. Sci. Technol., 2020, 54(13), 8464–8472 CrossRef CAS PubMed.
  22. W. Song, et al., Artificial water channels enable fast and selective water permeation through water-wire networks, Nat. Nanotechnol., 2020, 15(1), 73–79 CrossRef CAS PubMed.
  23. J. Shen, et al., Aquafoldmer-based aquaporin-like synthetic water channel, J. Am. Chem. Soc., 2020, 142(22), 10050–10058 CrossRef CAS PubMed.
  24. M. S. Sheikh, et al., Fabrication of nano composite membrane filter from graphene oxide (GO) and banana rachis cellulose nano crystal (CNC) for industrial effluent treatment, J. Ind. Eng. Chem., 2023, 128, 196–208 CrossRef.
  25. S. Wadhawan, et al., Role of nanomaterials as adsorbents in heavy metal ion removal from waste water: a review, J. Water Process. Eng., 2020, 33, 101038 CrossRef.
  26. G. Zhang, et al., Multifunctional Bacterial Cellulose/Covalent Organic Framework Composite Membranes with Antifouling and Antibacterial Properties for Dye Separation, ACS Appl. Mater. Interfaces, 2023, 15(27), 32903–32915 CrossRef CAS PubMed.
  27. R. M. Cywar, et al., Bio-based polymers with performance-advantaged properties, Nat. Rev. Mater., 2022, 7(2), 83–103 CrossRef CAS.
  28. M. A. Qamar, et al., Fabrication of g-C3N4/transition metal (Fe, Co, Ni, Mn and Cr)-doped ZnO ternary composites: excellent visible light active photocatalysts for the degradation of organic pollutants from wastewater, Mater. Res. Bull., 2022, 147, 111630 CrossRef CAS.
  29. Y. Yu, et al., Synthesis of eco-friendly lignin-betaine and its application for dye wastewater treatment, Ind. Crops Prod., 2023, 192, 116014 CrossRef CAS.
  30. S. Mansoori, et al., Membranes based on non-synthetic (natural) polymers for wastewater treatment, Polym. Test., 2020, 84, 106381 CrossRef CAS.
  31. D. Cambie, et al., Applications of continuous-flow photochemistry in organic synthesis, material science, and water treatment, Chem. Rev., 2016, 116(17), 10276–10341 CrossRef CAS PubMed.
  32. L. Buglioni, et al., Technological innovations in photochemistry for organic synthesis: flow chemistry, high-throughput experimentation, scale-up, and photoelectrochemistry, Chem. Rev., 2021, 122(2), 2752–2906 CrossRef PubMed.
  33. C. Xu, et al., Nanostructured materials for photocatalysis, Chem. Soc. Rev., 2019, 48(14), 3868–3902 RSC.
  34. L. Candish, et al., Photocatalysis in the life science industry, Chem. Rev., 2021, 122(2), 2907–2980 CrossRef PubMed.
  35. B. Wang, et al., New sustainable Mannich-functionalized lignin flocculants for ultra-efficiently tailoring wastewater purification, J. Cleaner Prod., 2023, 387, 135801 CrossRef CAS.
  36. Y.-N. Zhou, et al., Precision polymer synthesis by controlled radical polymerization: fusing the progress from polymer chemistry and reaction engineering, Prog. Polym. Sci., 2022, 101555 CrossRef CAS.
  37. Y.-M. Tu, et al., Biobased high-performance aromatic–aliphatic polyesters with complete recyclability, J. Am. Chem. Soc., 2021, 143(49), 20591–20597 CrossRef CAS PubMed.
  38. G. Machado, et al., Biopolymers from lignocellulosic biomass: feedstocks, production processes, and applications, Lignocellulosic Biorefining Technologies, 2020, pp. 125–158 Search PubMed.
  39. A. Vinod, et al., Renewable and sustainable biobased materials: an assessment on biofibers, biofilms, biopolymers and biocomposites, J. Cleaner Prod., 2020, 258, 120978 CrossRef CAS.
  40. M. Shahini, B. Ramezanzadeh and H. E. Mohammadloo, Recent advances in biopolymers/carbohydrate polymers as effective corrosion inhibitive macro-molecules: a review study from experimental and theoretical views, J. Mol. Liq., 2021, 325, 115110 CrossRef CAS.
  41. S. Wu, et al., An ultrasensitive electrochemical platform based on imprinted chitosan/gold nanoparticles/graphene nanocomposite for sensing cadmium (II) ions, Microchem. J., 2020, 155, 104710 CrossRef CAS.
  42. C. Song, et al., Influence of structure and functional group of modified kraft lignin on adsorption behavior of dye, Int. J. Biol. Macromol., 2023, 240, 124368 CrossRef CAS PubMed.
  43. Y. Ge and Z. Li, Application of lignin and its derivatives in adsorption of heavy metal ions in water: a review, ACS Sustain. Chem. Eng., 2018, 6(5), 7181–7192 CrossRef CAS.
  44. D. Bajwa, et al., A concise review of current lignin production, applications, products and their environmental impact, Ind. Crops Prod., 2019, 139, 111526 CrossRef CAS.
  45. B. L. Tardy, et al., Deconstruction and reassembly of renewable polymers and biocolloids into next generation structured materials, Chem. Rev., 2021, 121(22), 14088–14188 CrossRef CAS PubMed.
  46. W. Sun, et al., Bioinspired lignin-based loose nanofiltration membrane with excellent acid, fouling, and chlorine resistances toward dye/salt separation, J. Membr. Sci., 2023, 670, 121372 CrossRef CAS.
  47. B. Gao, Q. Chang and H. Yang, Selective adsorption of ofloxacin and ciprofloxacin from a binary system using lignin-based adsorbents: quantitative analysis, adsorption mechanisms, and structure-activity relationship, Sci. Total Environ., 2021, 765, 144427 CrossRef CAS PubMed.
  48. X. Zhou, et al., Functionalized lignin-based magnetic adsorbents with tunable structure for the efficient and selective removal of Pb (II) from aqueous solution, Chem. Eng. J., 2021, 420, 130409 CrossRef CAS.
  49. Y. Liu, et al., Recent advances in lignin-based porous materials for pollutants removal from wastewater, Int. J. Biol. Macromol., 2021, 187, 880–891 CrossRef CAS PubMed.
  50. C. Jin, et al., Lignin-inspired porous polymer networks as high-performance adsorbents for the efficient removal of malachite green dye, Colloids Surf., A, 2022, 643, 128760 CrossRef CAS.
  51. B. Du, et al., pH fractionated lignin for the preparation of lignin-based magnetic nanoparticles for the removal of methylene blue dye, Sep. Purif. Technol., 2022, 295, 121302 CrossRef CAS.
  52. J. Kaur, P. Sengupta and S. Mukhopadhyay, Critical review of bioadsorption on modified cellulose and removal of divalent heavy metals (Cd, Pb, and Cu), Ind. Eng. Chem. Res., 2022, 61(5), 1921–1954 CrossRef CAS.
  53. D. Jyoti, R. Sinha and C. Faggio, Advances in biological methods for the sequestration of heavy metals from water bodies: a review, Environ. Toxicol. Pharmacol., 2022, 103927 CrossRef CAS PubMed.
  54. Z. Dong and L. Zhao, Surface modification of cellulose microsphere with imidazolium-based ionic liquid as adsorbent: effect of anion variation on adsorption ability towards Au (III), Cellulose, 2018, 25, 2205–2216 CrossRef CAS.
  55. T. Yao, et al., Preparation of cellulose-based chromatographic medium for biological separation: a review, J. Chromatogr. A, 2022, 463297 CrossRef CAS PubMed.
  56. C. Ling, et al., Novel High-efficiency adsorbent consisting of magnetic Cellulose-based ionic liquid for removal of anionic dyes, J. Mol. Liq., 2022, 353, 118723 CrossRef CAS.
  57. D. Wen, et al., Aminotriazole isomers modified cellulose microspheres for selective adsorption of U (VI): performance and mechanism investigation, Carbohydr. Polym., 2021, 257, 117666 CrossRef CAS PubMed.
  58. X. Yang, et al., Facile preparation of L-cysteine–modified cellulose microspheres as a low-cost adsorbent for selective and efficient adsorption of Au (III) from the aqueous solution, Environ. Sci. Pollut. Res., 2020, 27, 38334–38343 CrossRef CAS PubMed.
  59. A. D. Rizkyana, et al., Sulfation and characterization of polysaccharides from Oyster mushroom (Pleurotus ostreatus) extracted using subcritical water, J. Supercrit. Fluids, 2022, 179, 105412 CrossRef CAS.
  60. M. M. A. N. Ranjha, et al., Modification in cellulose films through ascent cold plasma treatment and polymerization for food products packaging, Trends Food Sci. Technol., 2023, 134, 162–176 CrossRef CAS.
  61. S. Barthel and T. Heinze, Acylation and carbanilation of cellulose in ionic liquids, Green Chem., 2006, 8(3), 301–306 RSC.
  62. A. Gupta, C. R. Ladino and T. H. Mekonnen, Cationic modification of cellulose as a sustainable and recyclable adsorbent for anionic dyes, Int. J. Biol. Macromol., 2023, 234, 123523 CrossRef CAS PubMed.
  63. L. Tang, et al., Ultrasonication-assisted manufacture of cellulose nanocrystals esterified with acetic acid, Bioresour. Technol., 2013, 127, 100–105 CrossRef CAS PubMed.
  64. F. Azzam, et al., Surface peeling of cellulose nanocrystals resulting from periodate oxidation and reductive amination with water-soluble polymers, Cellulose, 2015, 22, 3701–3714 CrossRef CAS.
  65. G. Dalei, S. Das and M. Pradhan, Dialdehyde cellulose as a niche material for versatile applications: an overview, Cellulose, 2022, 29(10), 5429–5461 CrossRef CAS.
  66. N. H. Sari, et al., A review on cellulose fibers from Eichornia crassipes: synthesis, modification, properties and their composites, J. Nat. Fibers, 2023, 20(1), 2162179 CrossRef.
  67. S. Huang, et al., Polyamide nanofiltration membranes incorporated with cellulose nanocrystals for enhanced water flux and chlorine resistance, ACS Sustain. Chem. Eng., 2019, 7(14), 12315–12322 CAS.
  68. L. Bai, et al., Incorporation of cellulose nanocrystals (CNCs) into the polyamide layer of thin-film composite (TFC) nanofiltration membranes for enhanced separation performance and antifouling properties, Environ. Sci. Technol., 2018, 52(19), 11178–11187 CrossRef CAS PubMed.
  69. K. Mohan, et al., Recent trends on chitosan based hybrid materials for wastewater treatment: a review, Curr. Opin. Environ. Sci. Health, 2023, 100473 CrossRef.
  70. A. R. Alawady, et al., Polysulfone membranes with CNTs/Chitosan biopolymer nanocomposite as selective layer for remarkable heavy metal ions rejection capacity, Chem. Eng. J., 2020, 388, 124267 CrossRef.
  71. A. Balakrishnan, et al., Chitosan-based beads as sustainable adsorbents for wastewater remediation: a review, Environ. Chem. Lett., 2023, 1–25 CAS.
  72. N. A. Abd El-Ghany, et al., Recent Advances in Various Starch Formulation for Wastewater Purification via Adsorption Technique: A Review, J. Polym. Environ., 2023, 1–34 Search PubMed.
  73. P. S. Khoo, et al., Starch-Based Polymer Materials as Advanced Adsorbents for Sustainable Water Treatment: Current Status, Challenges, and Future Perspectives, Polymers, 2023, 15(14), 3114 CrossRef CAS PubMed.
  74. H. Zhang, et al., Preparation of starch-based adsorbing-flocculating bifunctional material St-A/F and its removal of active, direct and disperse dyes from textile printing and dyeing wastewater, Polym. Bull., 2023, 1–24 Search PubMed.
  75. X. Wang, et al., Research advances in chemical modifications of starch for hydrophobicity and its applications: a review, Carbohydr. Polym., 2020, 240, 116292 CrossRef CAS PubMed.
  76. D. Sarmah and N. Karak, Double network hydrophobic starch based amphoteric hydrogel as an effective adsorbent for both cationic and anionic dyes, Carbohydr. Polym., 2020, 242, 116320 CrossRef CAS PubMed.
  77. Y. Luo, et al., Study on Flocculation Performance of New Cationic Starch-Based Sludge Wastewater Flocculant, Starch-Stärke, 2023, 75(5–6), 2200226 CrossRef CAS.
  78. S. Zeng, et al., Transparent, flexible, and multifunctional starch-based double-network hydrogels as high-performance wearable electronics, Carbohydr. Polym., 2021, 267, 118198 CrossRef CAS PubMed.
  79. H. Guo, et al., Modified alginate materials for wastewater treatment: application prospects, Bioresour. Technol., 2023, 129639 CrossRef CAS PubMed.
  80. D.-N. Phan, et al., A review on the fabrication of several carbohydrate polymers into nanofibrous structures using electrospinning for removal of metal ions and dyes, Carbohydr. Polym., 2021, 252, 117175 CrossRef CAS PubMed.
  81. Z. A. Sutirman, M. M. Sanagi and W. I. W. Aini, Alginate-based adsorbents for removal of metal ions and radionuclides from aqueous solutions: a review, Int. J. Biol. Macromol., 2021, 174, 216–228 CrossRef CAS PubMed.
  82. P. Khajavi, A. R. Keshtkar and M. A. Moosavian, The optimization of U (VI) removal by a novel amidoximated modified calcium alginate gel bead with entrapped functionalized SiO2 nanoparticles, Prog. Nucl. Energy, 2021, 140, 103887 CrossRef CAS.
  83. N. Orachorn, P. Klongklaew and O. Bunkoed, A composite of magnetic GOx@ MOF incorporated in alginate hydrogel fiber adsorbent for the extraction of phthalate esters, Microchem. J., 2021, 171, 106827 CrossRef CAS.
  84. F. Marrakchi, et al., Cross-linked FeCl3-activated seaweed carbon/MCM-41/alginate hydrogel composite for effective biosorption of bisphenol A plasticizer and basic dye from aqueous solution, Bioresour. Technol., 2021, 331, 125046 CrossRef CAS PubMed.
  85. H.-A. S. Tohamy, G. Taha and M. Sultan, Dialdehyde cellulose/gelatin hydrogel as a packaging material for manganese oxides adsorbents for wastewater remediation: characterization and performance evaluation, Int. J. Biol. Macromol., 2023, 248, 125931 CrossRef CAS PubMed.
  86. S. Saber-Samandari, et al., Adsorption of anionic and cationic dyes from aqueous solution using gelatin-based magnetic nanocomposite beads comprising carboxylic acid functionalized carbon nanotube, Chem. Eng. J., 2017, 308, 1133–1144 CrossRef CAS.
  87. R. Gusain, N. Kumar and S. S. Ray, Recent advances in carbon nanomaterial-based adsorbents for water purification, Coord. Chem. Rev., 2020, 405, 213111 CrossRef CAS.
  88. J. Jiang, et al., A multifunctional gelatin-based aerogel with superior pollutants adsorption, oil/water separation and photocatalytic properties, Chem. Eng. J., 2019, 358, 1539–1551 CrossRef CAS.
  89. A. Mtibe, et al., Synthetic biopolymers and their composites: advantages and limitations—An overview, Macromol. Rapid Commun., 2021, 42(15), 2100130 CrossRef CAS PubMed.
  90. J. Gracia, et al., Production of Polyhydroxyalkanoates using Volatile Fatty Acids from Municipal Wastewater Treatment Plant Sludge, Chem. Eng. Trans., 2023, 100, 559–564 Search PubMed.
  91. G. Mannina, et al., Bioplastic recovery from wastewater: a new protocol for polyhydroxyalkanoates (PHA) extraction from mixed microbial cultures, Bioresour. Technol., 2019, 282, 361–369 CrossRef CAS PubMed.
  92. F. Valentino, et al., Pilot-scale polyhydroxyalkanoate production from combined treatment of organic fraction of municipal solid waste and sewage sludge, Ind. Eng. Chem. Res., 2019, 58(27), 12149–12158 CrossRef CAS.
  93. C. Samori, et al., Polyhydroxyalkanoates and crotonic acid from anaerobically digested sewage sludge, ACS Sustain. Chem. Eng., 2019, 7(12), 10266–10273 CrossRef CAS.
  94. P. M. Gore, et al., Surface functionalized waste-silk fabric engineered with polylactic acid & activated charcoal for oil/solvent recovery from oily wastewater, J. Indian Chem. Soc., 2023, 100(7), 101022 CrossRef CAS.
  95. T. Krasian, W. Punyodom and P. Worajittiphon, A hybrid of 2D materials (MoS2 and WS2) as an effective performance enhancer for poly (lactic acid) fibrous mats in oil adsorption and oil/water separation, Chem. Eng. J., 2019, 369, 563–575 CrossRef CAS.
  96. J. Xue, et al., Tetradecylamine-MXene functionalized melamine sponge for effective oil/water separation and selective oil adsorption, Sep. Purif. Technol., 2021, 259, 118106 CrossRef CAS.
  97. N. Mohammad and Y. Atassi, Adsorption of methylene blue onto electrospun nanofibrous membranes of polylactic acid and polyacrylonitrile coated with chloride doped polyaniline, Sci. Rep., 2020, 10(1), 13412 CrossRef CAS PubMed.
  98. N. Abbasi, S. A. Khan and T. A. Khan, Statistically optimised sequestration of mefenamic acid from polluted water by acacia gum phthalate/pectin hydrogel: a novel multifunctional adsorbent material synthesised via microwave-assisted process, Chem. Eng. J., 2023, 466, 143296 CrossRef CAS.
  99. R. Rakhshaee and M. Panahandeh, Stabilization of a magnetic nano-adsorbent by extracted pectin to remove methylene blue from aqueous solution: a comparative studying between two kinds of cross-likened pectin, J. Hazard. Mater., 2011, 189(1–2), 158–166 CrossRef CAS PubMed.
  100. J. Bok-Badura, et al., Sorption studies of heavy metal ions on pectin-nano-titanium dioxide composite adsorbent, Sep. Sci. Technol., 2018, 53(7), 1034–1044 CrossRef CAS.
  101. S. Graham, P. F. Marina and A. Blencowe, Thermoresponsive polysaccharides and their thermoreversible physical hydrogel networks, Carbohydr. Polym., 2019, 207, 143–159 CrossRef CAS PubMed.
  102. M. Lahaye, Developments on gelling algal galactans, their structure and physico-chemistry, J. Appl. Phycol., 2001, 13, 173–184 CrossRef CAS.
  103. F. Jiang, et al., Extraction, Modification and Biomedical Application of Agarose Hydrogels: A Review, Mar. Drugs, 2023, 21(5), 299 CrossRef CAS PubMed.
  104. B. Felipe Melo Lima Gomes, et al., Adsorption of Cd (II) ions and methyl violet dye by using an agar-graphene oxide nano-biocomposite, Environ. Technol., 2023, 1–12 CrossRef PubMed.
  105. C. M. B. de Araujo, et al., Continuous removal of pharmaceutical drug chloroquine and Safranin-O dye from water using agar-graphene oxide hydrogel: selective adsorption in batch and fixed-bed experiments, Environ. Res., 2023, 216, 114425 CrossRef PubMed.
  106. C. M. B. de Araujo, et al., Wastewater treatment using recyclable agar-graphene oxide biocomposite hydrogel in batch and fixed-bed adsorption column: bench experiments and modeling for the selective removal of organics, Colloids Surf., A, 2022, 639, 128357 CrossRef CAS.
  107. H. Guo, et al., Surface molecular imprinting on carbon microspheres for fast and selective adsorption of perfluorooctane sulfonate, J. Hazard. Mater., 2018, 348, 29–38 CrossRef CAS PubMed.
  108. D. Zhang, W. Zhang and Y. Liang, Adsorption of perfluoroalkyl and polyfluoroalkyl substances (PFASs) from aqueous solution-A review, Sci. Total Environ., 2019, 694, 133606 CrossRef CAS PubMed.
  109. S. Dou, et al., Fish scale-based biochar with defined pore size and ultrahigh specific surface area for highly efficient adsorption of ciprofloxacin, Chemosphere, 2022, 287, 131962 CrossRef CAS PubMed.
  110. Z. Zhang, et al., Highly efficient and selective removal of trace lead from aqueous solutions by hollow mesoporous silica loaded with molecularly imprinted polymers, J. Hazard. Mater., 2017, 328, 160–169 CrossRef CAS PubMed.
  111. N. A. Samah, et al., Molecularly imprinted polymer for the removal of diclofenac from water: synthesis and characterization, Sci. Total Environ., 2018, 631, 1534–1543 CrossRef PubMed.
  112. L. Fang, et al., Efficient removal of norfloxacin in water using magnetic molecularly imprinted polymer, Chemosphere, 2021, 262, 128032 CrossRef CAS PubMed.
  113. G. Guan, J. H. Pan and Z. Li, Innovative utilization of molecular imprinting technology for selective adsorption and (photo) catalytic eradication of organic pollutants, Chemosphere, 2021, 265, 129077 CrossRef CAS PubMed.
  114. C. Lin, et al., Fabrication of photo-responsive cellulose based intelligent imprinted material and selective adsorption on typical pesticide residue, Chem. Eng. J., 2020, 394, 124841 CrossRef CAS.
  115. D. Rahangdale and A. Kumar, Chitosan as a substrate for simultaneous surface imprinting of salicylic acid and cadmium, Carbohydr. Polym., 2018, 202, 334–344 CrossRef CAS PubMed.
  116. M. V. Dinu, et al., Chitosan-based ion-imprinted cryo-composites with excellent selectivity for copper ions, Carbohydr. Polym., 2018, 186, 140–149 CrossRef CAS PubMed.
  117. F. Russo, et al., Dimethyl isosorbide as a green solvent for sustainable ultrafiltration and microfiltration membrane preparation, ACS Sustain. Chem. Eng., 2019, 8(1), 659–668 CrossRef.
  118. P. J. Espitia, C. A. Fuenmayor and C. G. Otoni, Nanoemulsions: synthesis, characterization, and application in bio-based active food packaging, Compr. Rev. Food Sci. Food Saf., 2019, 18(1), 264–285 CrossRef CAS PubMed.
  119. J.-Y. Zhou, et al., A comprehensive study on phase inversion behavior of a novel polysulfate membrane for high-performance ultrafiltration applications, J. Membr. Sci., 2020, 610, 118404 CrossRef CAS.
  120. J. D. Mota-Morales and E. Morales-Narvaez, Transforming nature into the next generation of bio-based flexible devices: new avenues using deep eutectic systems, Matter, 2021, 4(7), 2141–2162 CrossRef CAS.
  121. B. Doshi, M. Sillanpää and S. Kalliola, A review of bio-based materials for oil spill treatment, Water Res., 2018, 135, 262–277 CrossRef CAS PubMed.
  122. L. C. Battirola, et al., Cellulose acetate/cellulose nanofiber membranes for whey and fruit juice microfiltration, Cellulose, 2017, 24, 5593–5604 CrossRef CAS.
  123. X. Yang, et al., Cellulose acetate-based SiO2/TiO2 hybrid microsphere composite aerogel films for water-in-oil emulsion separation, Appl. Surf. Sci., 2018, 435, 609–616 CrossRef CAS.
  124. W. Xie, et al., Using the green solvent dimethyl sulfoxide to replace traditional solvents partly and fabricating PVC/PVC-g-PEGMA blended ultrafiltration membranes with high permeability and rejection, Ind. Eng. Chem. Res., 2019, 58(16), 6413–6423 CrossRef CAS.
  125. J. Chang, et al., Using green solvent, triethyl phosphate (TEP), to fabricate highly porous PVDF hollow fiber membranes for membrane distillation, J. Membr. Sci., 2017, 539, 295–304 CrossRef CAS.
  126. X. Yan, et al., Ionic liquids combined with membrane separation processes: a review, Sep. Purif. Technol., 2019, 222, 230–253 CrossRef CAS.
  127. M. P. Garralaga, et al., Ecotoxicological study of bio-based deep eutectic solvents formed by glycerol derivatives in two aquatic biomodels, Green Chem., 2022, 24(13), 5228–5241 RSC.
  128. R. Techie-Menson, et al., New bio-based sustainable polymers and polymer composites based on methacrylate derivatives of furfural, solketal and lactic acid, Mater. Today Commun., 2021, 28, 102721 CrossRef CAS.
  129. A. Haider, S. Haider and I.-K. Kang, A comprehensive review summarizing the effect of electrospinning parameters and potential applications of nanofibers in biomedical and biotechnology, Arabian J. Chem., 2018, 11(8), 1165–1188 CrossRef CAS.
  130. M. Rezaei, et al., Wetting phenomena in membrane distillation: mechanisms, reversal, and prevention, Water Res., 2018, 139, 329–352 CrossRef CAS PubMed.
  131. H. Chen, et al., Functionalized electrospun nanofiber membranes for water treatment: a review, Sci. Total Environ., 2020, 739, 139944 CrossRef CAS PubMed.
  132. S. Wu, et al., Preparation and performance evaluation of chitosan/polyvinylpyrrolidone/polyvinyl alcohol electrospun nanofiber membrane for heavy metal ions and organic pollutants removal, Int. J. Biol. Macromol., 2022, 210, 76–84 CrossRef CAS PubMed.
  133. L. A. Goetz, et al., All cellulose electrospun water purification membranes nanotextured using cellulose nanocrystals, Cellulose, 2018, 25, 3011–3023 CrossRef CAS.
  134. D. Yang, et al., Functionalized chitosan electrospun nanofiber membranes for heavy-metal removal, Polymer, 2019, 163, 74–85 CrossRef CAS.
  135. N. Neibolts, et al., Needle-free electrospinning of nanofibrillated cellulose and graphene nanoplatelets based sustainable poly (butylene succinate) nanofibers, Mater. Today Chem., 2020, 17, 100301 CrossRef CAS.
  136. E. Ahn, et al., A4 paper chemistry: synthesis of a versatile and chemically modifiable cellulose membrane, ACS Nano, 2020, 14(5), 6173–6180 CrossRef CAS PubMed.
  137. X. Luo, et al., Interfacial solid-phase chemical modification with Mannich reaction and Fe (III) chelation for designing lignin-based spherical nanoparticle adsorbents for highly efficient removal of low concentration phosphate from water, ACS Sustain. Chem. Eng., 2017, 5(8), 6539–6547 CrossRef CAS.
  138. J. P. Chaudhary, et al., Chitosan-based aerogel membrane for robust oil-in-water emulsion separation, ACS Appl. Mater. Interfaces, 2015, 7(44), 24957–24962 CrossRef CAS PubMed.
  139. T. P. Chau, et al., Green synthesis of Zirconium nanoparticles using Punica granatum (pomegranate) peel extract and their antimicrobial and antioxidant potency, Environ. Res., 2022, 209, 112771 CrossRef CAS PubMed.
  140. K. K. Brar, et al., Green route for recycling of low-cost waste resources for the biosynthesis of nanoparticles (NPs) and nanomaterials (NMs)-A review, Environ. Res., 2022, 207, 112202 CrossRef CAS PubMed.
  141. Y. Duan, et al., Sustainable biorefinery approaches towards circular economy for conversion of biowaste to value added materials and future perspectives, Fuel, 2022, 325, 124846 CrossRef CAS.
  142. M. Arthy and B. Phanikumar, Efficacy of Iron-Based Nanoparticles and Nanobiocomposites in Removal of Cr 3+, J. Hazard., Toxic Radioact. Waste, 2016, 20(3), 04016005 CrossRef.
  143. S. Jadoun, et al., Green synthesis of nanoparticles using plant extracts: a review, Environ. Chem. Lett., 2021, 19, 355–374 CrossRef CAS.
  144. R. R. Nasaruddin, et al., Toward greener synthesis of gold nanomaterials: from biological to biomimetic synthesis, Coord. Chem. Rev., 2021, 426, 213540 CrossRef CAS.
  145. W. B. Fernandes Filho, et al., Pectin hydrogels modified with montmorillonite: case studies for the removal of dyes and herbicides from water, J. Environ. Chem. Eng., 2023, 110846 CrossRef.
  146. P. J. Landrigan, et al., The Lancet Commission on pollution and health, Lancet, 2018, 391(10119), 462–512 CrossRef PubMed.
  147. C. C. Azubuike, C. B. Chikere and G. C. Okpokwasili, Bioremediation techniques–classification based on site of application: principles, advantages, limitations and prospects, World J. Microbiol. Biotechnol., 2016, 32, 1–18 CrossRef CAS PubMed.
  148. R. Kishor, et al., Ecotoxicological and health concerns of persistent coloring pollutants of textile industry wastewater and treatment approaches for environmental safety, J. Environ. Chem. Eng., 2021, 9(2), 105012 CrossRef CAS.
  149. B. Biswas, J. Labille and B. Prelot, Clays and modified clays in remediating environmental pollutants, Environ. Sci. Pollut. Res., 2020, 27, 38381–38383 CrossRef CAS PubMed.
  150. A. M. James, et al., Selective environmental remediation of strontium and cesium using sulfonated hyper-cross-linked polymers (SHCPs), ACS Appl. Mater. Interfaces, 2019, 11(25), 22464–22473 CrossRef CAS PubMed.
  151. A. Kirchon, et al., Effect of isomorphic metal substitution on the fenton and photo-fenton degradation of methylene blue using Fe-based metal–organic frameworks, ACS Appl. Mater. Interfaces, 2020, 12(8), 9292–9299 CrossRef CAS PubMed.
  152. Y. Wen, et al., Integrated photocatalytic reduction and oxidation of perfluorooctanoic acid by metal–organic frameworks: key insights into the degradation mechanisms, J. Am. Chem. Soc., 2022, 144(26), 11840–11850 CrossRef CAS PubMed.
  153. W. Teng, et al., Rapid and efficient removal of microcystins by ordered mesoporous silica, Environ. Sci. Technol., 2013, 47(15), 8633–8641 CrossRef CAS PubMed.
  154. J. Kochanek, et al., Biochar for intensification of plant-related industries to meet productivity, sustainability and economic goals: a review, Resour., Conserv. Recycl., 2022, 179, 106109 CrossRef CAS.
  155. N. N. Rudi, et al., Evolution of adsorption process for manganese removal in water via agricultural waste adsorbents, Heliyon, 2020, 6(9), e05049 CrossRef CAS PubMed.
  156. S. Tamjidi and A. Ameri, A review of the application of sea material shells as low cost and effective bio-adsorbent for removal of heavy metals from wastewater, Environ. Sci. Pollut. Res., 2020, 27, 31105–31119 CrossRef CAS PubMed.
  157. J. Li, et al., Sustainable environmental remediation via biomimetic multifunctional lignocellulosic nano-framework, Nat. Commun., 2022, 13(1), 4368 CrossRef CAS PubMed.
  158. S. Hokkanen, A. Bhatnagar and M. Sillanpää, A review on modification methods to cellulose-based adsorbents to improve adsorption capacity, Water Res., 2016, 91, 156–173 CrossRef CAS PubMed.
  159. E. Zong, et al., A lignin-based nano-adsorbent for superfast and highly selective removal of phosphate, J. Mater. Chem. A, 2018, 6(21), 9971–9983 RSC.
  160. M. Ateia, et al., Rapid removal of poly-and perfluorinated alkyl substances by poly (ethylenimine)-functionalized cellulose microcrystals at environmentally relevant conditions, Environ. Sci. Technol. Lett., 2018, 5(12), 764–769 CrossRef CAS.
  161. L. Liang, et al., Chemically cross-linked cellulose nanocrystal aerogels for effective removal of cation dye, Front. Chem., 2020, 8, 570 CrossRef CAS PubMed.
  162. Y.-L. Zhong, et al., Three Dimensional Functionalized Carbon/Tin(IV) Sulfide Biofoam for Photocatalytical Purification of Chromium(VI)-Containing Wastewater, ACS Sustain. Chem. Eng., 2018, 6(8), 10660–10667 CrossRef CAS.
  163. N. A. Dahlan, et al., Developing of a magnetite film of carboxymethyl cellulose grafted carboxymethyl polyvinyl alcohol (CMC-g-CMPVA) for copper removal, Carbohydr. Polym., 2017, 173, 619–630 CrossRef CAS PubMed.
  164. T. H. Tran, et al., Removal of metal ions from aqueous solutions using carboxymethyl cellulose/sodium styrene sulfonate gels prepared by radiation grafting, Carbohydr. Polym., 2017, 157, 335–343 CrossRef CAS PubMed.
  165. Y. Tang, et al., Preparation of lignin sulfonate-based mesoporous materials for adsorbing malachite green from aqueous solution, J. Environ. Chem. Eng., 2016, 4(3), 2900–2910 CrossRef CAS.
  166. P. R. Sharma, et al., Nanocellulose from Spinifex as an Effective Adsorbent to Remove Cadmium(II) from Water, ACS Sustain. Chem. Eng., 2018, 6(3), 3279–3290 CrossRef CAS.
  167. Komal, et al., Amelioration of adsorptive efficacy by synergistic assemblage of functionalized graphene oxide with esterified cellulose nanofibers for mitigation of pharmaceutical waste, J. Hazard. Mater., 2022, 424, 127541 CrossRef CAS PubMed.
  168. L. M. Sanchez, et al., Cellulose nanofibers/PVA blend polymeric beads containing in-situ prepared magnetic nanorods as dye pollutants adsorbents, Int. J. Biol. Macromol., 2022, 209, 1211–1221 CrossRef CAS PubMed.
  169. P. Liu, K. Oksman and A. P. Mathew, Surface adsorption and self-assembly of Cu(II) ions on TEMPO-oxidized cellulose nanofibers in aqueous media, J. Colloid Interface Sci., 2016, 464, 175–182 CrossRef CAS PubMed.
  170. B. S. Rathi and P. S. Kumar, Application of adsorption process for effective removal of emerging contaminants from water and wastewater, Environ. Pollut., 2021, 280, 116995 CrossRef CAS PubMed.
  171. J. Saleem, et al., Production and applications of activated carbons as adsorbents from olive stones, Biomass Convers. Biorefin., 2019, 9, 775–802 CrossRef CAS.
  172. C. Yu and X. Han, Adsorbent material used in water treatment-a review, in 2015 2nd International Workshop on Materials Engineering and Computer Sciences, Atlantis Press, 2015 Search PubMed.
  173. M. del Mar Orta, et al., Biopolymer-clay nanocomposites as novel and ecofriendly adsorbents for environmental remediation, Appl. Clay Sci., 2020, 198, 105838 CrossRef.
  174. S. Ebnesajjad, Handbook of Biopolymers and Biodegradable Plastics: Properties, Processing and Applications, William Andrew, 2012 Search PubMed.
  175. S. Biswas, et al., Chitosan–clay composites for wastewater treatment: a state-of-the-art review, ACS ES&T Water, 2021, 1(5), 1055–1085 Search PubMed.
  176. A. Bhattacharyya, 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 CrossRef CAS PubMed.
  177. Y. Qi, et al., Natural polysaccharides-modified graphene oxide for adsorption of organic dyes from aqueous solutions, J. Colloid Interface Sci., 2017, 486, 84–96 CrossRef CAS PubMed.
  178. N. Politaeva, et al., Graphene oxide-chitosan composites for water treatment from copper cations, Water, 2022, 14(9), 1430 CrossRef CAS.
  179. T. T. C. Trang and K. Takaomi, Chitosan and its biomass composites in application for water treatment, Curr. Opin. Green Sustainable Chem., 2021, 29, 100429 CrossRef CAS.
  180. X.-R. Zhao, et al., 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 CrossRef CAS PubMed.
  181. T. S. Vo and T. T. B. C. Vo, Organic dye removal and recycling performances of graphene oxide-coated biopolymer sponge, Prog. Nat. Sci.: Mater. Int., 2022, 32(5), 634–642 CrossRef CAS.
  182. K. N. M. Bernard, et al., 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 CrossRef CAS.
  183. H. Hosseinzadeh and S. Ramin, 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 CrossRef CAS PubMed.
  184. X. Zhao, et al., Microwave assisted copolymerization of sodium alginate and dimethyl diallyl ammonium chloride as flocculant for dye removal, Int. J. Biol. Macromol., 2020, 156, 585–590 CrossRef CAS PubMed.
  185. S. A. Hosseini, et al., Efficient dye removal from aqueous solution by high-performance electrospun nanofibrous membranes through incorporation of SiO2 nanoparticles, J. Cleaner Prod., 2018, 183, 1197–1206 CrossRef CAS.
  186. N. M. Mahmoodi and Z. Mokhtari-Shourijeh, Preparation of PVA-chitosan blend nanofiber and its dye removal ability from colored wastewater, Fibers Polym., 2015, 16, 1861–1869 CrossRef CAS.
  187. C. Zhijiang, et al., Calcium alginate-coated electrospun polyhydroxybutyrate/carbon nanotubes composite nanofibers as nanofiltration membrane for dye removal, J. Mater. Sci., 2018, 53, 14801–14820 CrossRef CAS.
  188. C. Zhijiang, et al., Preparation and characterization of a bi-layered nano-filtration membrane from a chitosan hydrogel and bacterial cellulose nanofiber for dye removal, Cellulose, 2018, 25, 5123–5137 CrossRef.
  189. J. Shi, et al., Simultaneous electrospinning and spraying toward branch-like nanofibrous membranes functionalised with carboxylated MWCNTs for dye removal, Mater. Lett., 2016, 166, 26–29 CrossRef CAS.
  190. J. Xing, et al., Preparation of micro-nanofibrous chitosan sponges with ternary solvents for dye adsorption, Carbohydr. Polym., 2018, 198, 69–75 CrossRef CAS PubMed.
  191. M. S. Sajab, et al., Bifunctional graphene oxide–cellulose nanofibril aerogel loaded with Fe (III) for the removal of cationic dye via simultaneous adsorption and Fenton oxidation, RSC Adv., 2016, 6(24), 19819–19825 RSC.
  192. N. M. Aboamera, et al., An effective removal of organic dyes using surface functionalized cellulose acetate/graphene oxide composite nanofibers, Cellulose, 2018, 25, 4155–4166 CrossRef CAS.
  193. R. Guo, et al., Fabrication and highly efficient dye removal characterization of beta-cyclodextrin-based composite polymer fibers by electrospinning, Nanomaterials, 2019, 9(1), 127 CrossRef PubMed.
  194. Y. Chen, G. Dai and Q. Gao, Starch nanoparticles–graphene aerogels with high supercapacitor performance and efficient adsorption, ACS Sustain. Chem. Eng., 2019, 7(16), 14064–14073 CrossRef CAS.
  195. P. Huang, et al., Dye-mediated interactions in chitosan-based polyelectrolyte/organoclay hybrids for enhanced adsorption of industrial dyes, ACS Appl. Mater. Interfaces, 2019, 11(12), 11961–11969 CrossRef CAS PubMed.
  196. R. Nisticò, et al., Chitosan-derived iron oxide systems for magnetically guided and efficient water purification processes from polycyclic aromatic hydrocarbons, ACS Sustain. Chem. Eng., 2017, 5(1), 793–801 CrossRef.
  197. D. Qian, et al., A bifunctional alginate-based composite hydrogel with synergistic pollutant adsorption and photocatalytic degradation performance, Ind. Eng. Chem. Res., 2019, 58(29), 13133–13144 CrossRef CAS.
  198. J. Guo, et al., Optimization of the removal of reactive golden yellow SNE dye by cross-linked cationic starch and its adsorption properties, J. Eng. Fibers Fabr., 2019, 14, 1558925019865260 CAS.
  199. A. Pourjavadi, A. Abedin-Moghanaki and A. Tavakoli, 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(44), 38042–38051 RSC.
  200. D. Gautam, et al., A new hemicellulose-based adsorbent for malachite green, J. Environ. Chem. Eng., 2018, 6(4), 3889–3897 CrossRef CAS.
  201. R. Cui, et al., Cuprous oxide/copper oxide interpenetrated into ordered mesoporous cellulose-based carbon aerogels for efficient photocatalytic degradation of methylene blue, Front. Chem. Sci. Eng., 2023, 1–12 Search PubMed.
  202. Y. Wang, et al., Removal of anionic dyes from aqueous solutions by cellulose-based adsorbents: equilibrium, kinetics, and thermodynamics, J. Chem. Eng. Data, 2016, 61(9), 3266–3276 CrossRef CAS.
  203. V. Kumar, et al., Synthesis and characterization of silver nanoparticle embedded cellulose–gelatin based hybrid hydrogel and its utilization in dye degradation, RSC Adv., 2023, 13(12), 8409–8419 RSC.
  204. B. K. Bhangi and S. K. Ray, Nano silver chloride and alginate incorporated composite copolymer adsorbent for adsorption of a synthetic dye from water in a fixed bed column and its photocatalytic reduction, Int. J. Biol. Macromol., 2020, 144, 801–812 CrossRef CAS PubMed.
  205. L. Hu, et al., Adsorption behavior of dyes from an aqueous solution onto composite magnetic lignin adsorbent, Chemosphere, 2020, 246, 125757 CrossRef CAS PubMed.
  206. F. Z. Benhachem, T. Attar and F. Bouabdallah, Kinetic study of adsorption methylene blue dye from aqueous solutions using activated carbon, Chem. Rev. Lett., 2019, 2(1), 33–39 Search PubMed.
  207. Y. Wang, et al., Novel environmental-friendly nano-composite magnetic attapulgite functionalized by chitosan and EDTA for cadmium (II) removal, J. Alloys Compd., 2020, 817, 153286 CrossRef CAS.
  208. O. F. Sarioglu, et al., Bacteria encapsulated electrospun nanofibrous webs for remediation of methylene blue dye in water, Colloids Surf., B, 2017, 152, 245–251 CrossRef CAS PubMed.
  209. U. Habiba, et al., Degradation of methyl orange and congo red by using chitosan/polyvinyl alcohol/TiO2 electrospun nanofibrous membrane, Int. J. Biol. Macromol., 2019, 131, 821–827 CrossRef CAS PubMed.
  210. P. Li, et al., Highly selective adsorption of dyes and arsenate from their aqueous mixtures using a silica-sand/cationized-starch composite, Microporous Mesoporous Mater., 2018, 263, 210–219 CrossRef CAS.
  211. F. Hayeeye, et al., Kinetics and thermodynamics of Rhodamine B adsorption by gelatin/activated carbon composite beads, Colloids Surf., A, 2017, 513, 259–266 CrossRef CAS.
  212. K. Junlapong, et al., Effective adsorption of methylene blue by biodegradable superabsorbent cassava starch-based hydrogel, Int. J. Biol. Macromol., 2020, 158, 258–264 CrossRef CAS PubMed.
  213. G. Dotto, et al., Chitosan/polyamide nanofibers prepared by Forcespinning® technology: a new adsorbent to remove anionic dyes from aqueous solutions, J. Cleaner Prod., 2017, 144, 120–129 CrossRef CAS.
  214. S. Sethi and S. Thakur, Synthesis and characterization of nanocomposite chitosan-gelatin hydrogel loaded with ZnO and its application in photocatalytic dye degradation, Mater. Today: Proc., 2023, 78, 815–824 CAS.
  215. N. M. Mahmoodi, Z. Mokhtari-Shourijeh and J. Abdi, Preparation of mesoporous polyvinyl alcohol/chitosan/silica composite nanofiber and dye removal from wastewater, Environ. Prog. Sustainable Energy, 2019, 38(s1), S100–S109 CAS.
  216. M. Stan, 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, 65–73 CrossRef CAS.
  217. R. Rakhshaee, M. Giahi and A. Pourahmad, Removal of methyl orange from aqueous solution by Azolla filicoloides: synthesis of Fe3O4 nano-particles and its surface modification by the extracted pectin of Azolla, Chin. Chem. Lett., 2011, 22(4), 501–504 CrossRef CAS.
  218. X. Jiang, et al., Cellulose nanocrystals-based flocculants for high-speed and high-efficiency decolorization of colored effluents, J. Cleaner Prod., 2020, 251, 119749 CrossRef CAS.
  219. M. Ikram, et al., Toward Efficient Bactericidal and Dye Degradation Performance of Strontium-and Starch-Doped Fe2O3 Nanostructures: In Silico Molecular Docking Studies, ACS Omega, 2023, 8(8), 8066–8077 CrossRef CAS PubMed.
  220. F. Khosravi Mohammad Soltan, M. Hajiani and A. Haji, Application of Polymeric Nanofibers for Removal of Dyes, J. Stud. Color World, 2019, 9(1), 23–39 Search PubMed.
  221. W. Maatar and S. Boufi, Microporous cationic nanofibrillar cellulose aerogel as promising adsorbent of acid dyes, Cellulose, 2017, 24, 1001–1015 CrossRef CAS.
  222. T. S. Anirudhan, P. Senan and P. S. Suchithra, Evaluation of iron (III)-coordinated amino-functionalized poly (glycidyl methacrylate)-grafted cellulose for arsenic (V) adsorption from aqueous solutions, Water, Air, Soil Pollut., 2011, 220, 101–116 CrossRef CAS.
  223. S. Haider, et al., Novel route for amine grafting to chitosan electrospun nanofibers membrane for the removal of copper and lead ions from aqueous medium, Carbohydr. Polym., 2018, 199, 406–414 CrossRef CAS PubMed.
  224. D. Wang, et al., Electrospun cellulose nanocrystals/chitosan/polyvinyl alcohol nanofibrous films and their exploration to metal ions adsorption, Polymers, 2018, 10(10), 1046 CrossRef PubMed.
  225. K. Zhang, et al., Tree-like cellulose nanofiber membranes modified by citric acid for heavy metal ion (Cu 2+) removal, Cellulose, 2019, 26, 945–958 CrossRef CAS.
  226. M. Karim and Md. R. Awual, et al.) Composite nanofibers membranes of polyhaHavard (vinyl alcohol)/chitosan for selective lead (II) and cadmium (II) ions removal from wastewater, Ecotoxicol. Environ. Saf., 2019, 169, 479–486 CrossRef CAS PubMed.
  227. J. Li, et al., Shape memory aerogels from nanocellulose and polyethyleneimine as a novel adsorbent for removal of Cu (II) and Pb (II), Carbohydr. Polym., 2018, 196, 376–384 CrossRef CAS PubMed.
  228. L. Li, et al., Removal of Cr (VI) with a spiral wound chitosan nanofiber membrane module via dead-end filtration, J. Membr. Sci., 2017, 544, 333–341 CrossRef CAS.
  229. P. Wang, et al., Adsorption of hexavalent chromium by novel chitosan/poly (ethylene oxide)/permutit electrospun nanofibers, New J. Chem., 2018, 42(21), 17740–17749 RSC.
  230. S. Koushkbaghi, et al., Aminated-Fe3O4 nanoparticles filled chitosan/PVA/PES dual layers nanofibrous membrane for the removal of Cr (VI) and Pb (II) ions from aqueous solutions in adsorption and membrane processes, Chem. Eng. J., 2018, 337, 169–182 CrossRef CAS.
  231. A. Camiré, B. Chabot and A. Lajeunesse, Sorption Capacities of a Lignin-Based Electrospun Nanofibrous Material for Pharmaceutical Residues Remediation in Water, IntechOpen, London, UK, 2019, vol. 25 Search PubMed.
  232. A. Y. Abdolmaleki, et al., Adsorption of tetracycline from water using glutaraldehyde-crosslinked electrospun nanofibers of chitosan/poly (vinyl alcohol), Water Sci. Technol., 2018, 77(5), 1324–1335 CrossRef CAS PubMed.
  233. E. Kudlek, et al., TiO2 immobilised on biopolymer nanofibers for the removal of bisphenol A and diclofenac from water, Ecol. Chem. Eng. S, 2017, 24, 417–429 CAS.
  234. M. M. Bashar, et al., Superhydrophobic surfaces with fluorinated cellulose nanofiber assemblies for oil–water separation, RSC Adv., 2017, 7(59), 37168–37174 RSC.
  235. F. Zhou, et al., High performance of 3D porous graphene/lignin/sodium alginate composite for adsorption of Cd (II) and Pb (II), Environ. Sci. Pollut. Res., 2018, 25, 15651–15661 CrossRef CAS PubMed.
  236. Q. Zhu, et al., Activable carboxylic acid functionalized crystalline nanocellulose/PVA-co-PE composite nanofibrous membrane with enhanced adsorption for heavy metal ions, Sep. Purif. Technol., 2017, 186, 70–77 CrossRef CAS.
  237. K. Balasubramani, et al., Removal of Antidiabetic Pharmaceutical (Metformin) Using Graphene Oxide Microcrystalline Cellulose (GOMCC): Insights to Process Optimization, Equilibrium, Kinetics, And Machine Learning, Ind. Eng. Chem. Res., 2023, 62(11), 4713–4728 CrossRef CAS.
  238. D. Zhang, et al., Novel polyethyleneimine functionalized chitosan–lignin composite sponge with nanowall-network structures for fast and efficient removal of Hg (ii) ions from aqueous solution, Environ. Sci.: Nano, 2020, 7(3), 793–802 RSC.
  239. C.-M. Hung, et al., Nitrogen and boron co-doped lignin biochar for enhancing calcium peroxide activation toward organic micropollutants decontamination in waste activated sludge and related microbial structure dynamics, Bioresour. Technol., 2023, 372, 128673 CrossRef CAS PubMed.
  240. M. Bozorgi, et al., Performance of synthesized cast and electrospun PVA/chitosan/ZnO-NH 2 nano-adsorbents in single and simultaneous adsorption of cadmium and nickel ions from wastewater, Environ. Sci. Pollut. Res., 2018, 25, 17457–17472 CrossRef CAS PubMed.
  241. S. I. Siddiqui, et al., Arsenic removal from water by starch functionalized maghemite nano-adsorbents: thermodynamics and kinetics investigations, Colloid Interface Sci. Commun., 2020, 36, 100263 CrossRef CAS.
  242. H. Dong, et al., A comparative study on the activation of persulfate by bare and surface-stabilized nanoscale zero-valent iron for the removal of sulfamethazine, Sep. Purif. Technol., 2020, 230, 115869 CrossRef CAS.
  243. X. Yan, et al., Arsenic stabilization performance of a novel starch-modified Fe-Mn binary oxide colloid, Sci. Total Environ., 2020, 707, 136064 CrossRef CAS PubMed.
  244. L. Bergamonti, et al., 3D printed chitosan scaffolds: a new TiO2 support for the photocatalytic degradation of amoxicillin in water, Water Res., 2019, 163, 114841 CrossRef CAS PubMed.
  245. R.-H. Liang, et al., Pb2+ adsorption by ethylenediamine-modified pectins and their adsorption mechanisms, Carbohydr. Polym., 2020, 234, 115911 CrossRef CAS PubMed.
  246. C. Geng, et al., Mild routine to prepare Fe-Mn bimetallic nano-cluster (Fe-Mn NCs) and its magnetic starch-based composite adsorbent (Fe-Mn@SCAs) for wide pH range adsorption for Hg(II) sewage, J. Taiwan Inst. Chem. Eng., 2023, 144, 104768 CrossRef CAS.
  247. B. Ram and G. S. Chauhan, New spherical nanocellulose and thiol-based adsorbent for rapid and selective removal of mercuric ions, Chem. Eng. J., 2018, 331, 587–596 CrossRef CAS.
  248. W. Zhou, et al., Preparation of electrospun silk fibroin/Cellulose Acetate blend nanofibers and their applications to heavy metal ions adsorption, Fibers Polym., 2011, 12, 431–437 CrossRef CAS.
  249. B. Li, et al., Environmentally friendly chitosan/PEI-grafted magnetic gelatin for the highly effective removal of heavy metals from drinking water, Sci. Rep., 2017, 7(1), 1–9 CrossRef.
  250. K. Chai, et al., Rapid and selective recovery of acetophenone from petrochemical effluents by crosslinked starch polymer, J. Hazard. Mater., 2018, 348, 20–28 CrossRef CAS PubMed.
  251. Y. Fei, et al., Adsorptive removal of ciprofloxacin by sodium alginate/graphene oxide composite beads from aqueous solution, J. Colloid Interface Sci., 2016, 484, 196–204 CrossRef CAS PubMed.
  252. Y. Zhuang, et al., Novel one step preparation of a 3D alginate based MOF hydrogel for water treatment, New J. Chem., 2019, 43(19), 7202–7208 RSC.
  253. M. Sharma, et al., Hybrid alginate–protein cryogel beads: efficient and sustainable bio-based materials to purify immunoglobulin G antibodies, Green Chem., 2020, 22(7), 2225–2233 RSC.
  254. D. P. Facchi, et al., New magnetic chitosan/alginate/Fe3O4@ SiO2 hydrogel composites applied for removal of Pb (II) ions from aqueous systems, Chem. Eng. J., 2018, 337, 595–608 CrossRef CAS.
  255. R. E. Abou-Zeid, et al., Novel method of preparation of tricarboxylic cellulose nanofiber for efficient removal of heavy metal ions from aqueous solution, Int. J. Biol. Macromol., 2018, 119, 207–214 CrossRef CAS PubMed.
  256. G. Hanbali, et al., Enhanced ibuprofen adsorption and desorption on synthesized functionalized magnetic multiwall carbon nanotubes from aqueous solution, Materials, 2020, 13(15), 3329 CrossRef CAS PubMed.
  257. Y. Zhao, et al., Fabrication of biomaterial/TiO2 composite photocatalysts for the selective removal of trace environmental pollutants, Chin. J. Chem. Eng., 2019, 27(6), 1416–1428 CrossRef CAS.
  258. C. Noè, et al., UV-cured chitosan and gelatin hydrogels for the removal of as (V) and Pb (II) from water, Polymers, 2022, 14(6), 1268 CrossRef PubMed.
  259. J. Beyer, et al., Environmental effects of the Deepwater Horizon oil spill: a review, Mar. Pollut. Bull., 2016, 110(1), 28–51 CrossRef CAS PubMed.
  260. R. H. Kollarigowda, H. M. Bhyrappa and G. Cheng, Stimulus-responsive biopolymeric surface: molecular switches for oil/water separation, ACS Appl. Bio Mater., 2019, 2(10), 4249–4257 CrossRef CAS PubMed.
  261. J. K. Rad, Z. Balzade and A. R. Mahdavian, Spiropyran-based advanced photoswitchable materials: a fascinating pathway to the future stimuli-responsive devices, J. Photochem. Photobiol., C, 2022, 100487 Search PubMed.
  262. D. D. Ejeta, et al., Preparation of superhydrophobic and superoleophilic cotton-based material for extremely high flux water-in-oil emulsion separation, Chem. Eng. J., 2020, 402, 126289 CrossRef.
  263. X. Xu, et al., Preparation and characterization of cellulose grafted with epoxidized soybean oil aerogels for oil-absorbing materials, J. Agric. Food Chem., 2019, 67(2), 637–643 CrossRef CAS PubMed.
  264. B. Doshi, et al., Effectiveness of N, O-carboxymethyl chitosan on destabilization of Marine Diesel, Diesel and Marine-2T oil for oil spill treatment, Carbohydr. Polym., 2017, 167, 326–336 CrossRef CAS PubMed.
  265. G. Z. Sun, et al., Adsorption characteristics of residual oil on amphiphilic chitosan derivative, Water Sci. Technol., 2010, 61(9), 2363–2374 CrossRef CAS PubMed.
  266. Y. Huang, et al., Bio-Based Porous Aerogel with Bionic Structure and Hydrophobic Polymer Coating for Efficient Absorption of Oil/Organic Liquids, Polymers, 2022, 14(21), 4579 CrossRef CAS PubMed.
  267. B. Doshi, et al., Sodium salt of oleoyl carboxymethyl chitosan: a sustainable adsorbent in the oil spill treatment, J. Cleaner Prod., 2018, 170, 339–350 CrossRef CAS.
  268. S. F. Soares, et al., Chitosan-silica hybrid nanosorbents for oil removal from water, Colloids Surf., A, 2017, 532, 305–313 CrossRef CAS.
  269. O. Laitinen, et al., Sustainable stabilization of oil in water emulsions by cellulose nanocrystals synthesized from deep eutectic solvents, Cellulose, 2017, 24, 1679–1689 CrossRef CAS.
  270. S. Mirshahghassemi and J. R. Lead, Oil recovery from water under environmentally relevant conditions using magnetic nanoparticles, Environ. Sci. Technol., 2015, 49(19), 11729–11736 CrossRef CAS PubMed.
  271. T. Zhai, et al., Synthesis of polyvinyl alcohol/cellulose nanofibril hybrid aerogel microspheres and their use as oil/solvent superabsorbents, Carbohydr. Polym., 2016, 148, 300–308 CrossRef CAS PubMed.
  272. R. R. Fouad, H. A. Aljohani and K. R. Shoueir, Biocompatible poly(vinyl alcohol) nanoparticle-based binary blends for oil spill control, Mar. Pollut. Bull., 2016, 112(1), 46–52 CrossRef CAS PubMed.
  273. R. da Rosa Schio, et al., Synthesis of a bio–based polyurethane/chitosan composite foam using ricinoleic acid for the adsorption of Food Red 17 dye, Int. J. Biol. Macromol., 2019, 121, 373–380 CrossRef CAS PubMed.
  274. D. Li, et al., A versatile bio-based material for efficiently removing toxic dyes, heavy metal ions and emulsified oil droplets from water simultaneously, Bioresour. Technol., 2017, 245, 649–655 CrossRef CAS PubMed.
  275. Y.-L. Kang, et al., Full-biobased nanofiber membranes toward decontamination of wastewater containing multiple pollutants, ACS Sustain. Chem. Eng., 2018, 6(9), 11783–11792 CrossRef CAS.
  276. R. E. Abouzeid, et al., Current state and new trends in the use of cellulose nanomaterials for wastewater treatment, Biomacromolecules, 2018, 20(2), 573–597 CrossRef PubMed.
  277. N. Mohammed, N. Grishkewich and K. C. Tam, Cellulose nanomaterials: promising sustainable nanomaterials for application in water/wastewater treatment processes, Environ. Sci.: Nano, 2018, 5(3), 623–658 RSC.
  278. T. Li, et al., Developing fibrillated cellulose as a sustainable technological material, Nature, 2021, 590(7844), 47–56 CrossRef CAS PubMed.
  279. S. Ahmad, W. A. Siddiqi and S. Ahmad, Sustainable Nanocomposite Porous Absorbent and Membrane Sieves: Definition, Classification, History, Properties, Synthesis, Applications, and Future Prospects, J. Environ. Chem. Eng., 2023, 109367 CrossRef CAS.
  280. P. Liu, C. Zhu and A. P. Mathew, Mechanically robust high flux graphene oxide-nanocellulose membranes for dye removal from water, J. Hazard. Mater., 2019, 371, 484–493 CrossRef CAS PubMed.
  281. M. A. Rasool and I. F. Vankelecom, Preparation of full-bio-based nanofiltration membranes, J. Membr. Sci., 2021, 618, 118674 CrossRef CAS.
  282. V. Vatanpour, et al., Cellulose acetate in fabrication of polymeric membranes: a review, Chemosphere, 2022, 133914 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2024