Next Article in Journal
Bridging Housing and Climate Needs: Bamboo Construction in the Philippines
Previous Article in Journal
Integrated Value Model for Sustainable Assessment of Modular Residential Towers: Case Study: Ten Degrees Croydon and Apex House in London
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in the Remediation of Textile-Dye-Containing Wastewater: Prioritizing Human Health and Sustainable Wastewater Treatment

by
Aravin Prince Periyasamy
Biomaterial Processing and Products/Textile and Nonwoven Materials, VTT Technical Research Centre of Finland, Tietotie 4E, 02150 Espoo, Finland
Sustainability 2024, 16(2), 495; https://doi.org/10.3390/su16020495
Submission received: 15 November 2023 / Revised: 23 December 2023 / Accepted: 29 December 2023 / Published: 5 January 2024
(This article belongs to the Special Issue Green and Sustainable Textile Materials)

Abstract

:
Water makes up most of the Earth, although just 0.3% is usable for people and animals. The huge oceans, icecaps, and other non-potable water resources make up the remaining 99.7%. Water quality has declined in recent decades due to pollution from population growth, industry, unplanned urbanization, and poor water management. The textile industry has significant global importance, although it also stands as a major contributor to wastewater generation, leading to water depletion and ecotoxicity. This issue arises from the extensive utilization of harmful chemicals, notably dyes. The main aim of this review article is to combine and assess the impacts of textile wastewater that contains dyes and chemicals, and to examine their potential consequences on human health, aquatic health, and the environment. Moreover, the dedicated section presents an in-depth review of various environmentally sustainable approaches for the management and treatment of wastewater in the textile industry. These approaches encompass bio adsorbents, biological methods, membrane technology, ion exchange, advanced oxidation processes, as well as physicochemical and biochemical processes. Furthermore, this study also evaluates the contemporary progressions in this particular domain, taking into account the corresponding advantages and disadvantages. Finally, this article highlights the significance of recovering and reusing dyes, alkalis, and electrolytes in wastewater treatment. Additionally, it emphasizes the necessity of performing technoeconomic analyses and life cycle assessments (LCA) on wastewater treatment plants.

1. Introduction

Currently, ecosystems are primarily experiencing harm due to the exhaustion of natural resources and the deterioration of the environment resulting from industrial expansion and environmental emergencies [1,2]. Water pollution is a significant environmental issue posing significant risks to water, the primary life-sustaining element on Earth, emphasizing its crucial role in supporting life [3,4]. Pollution is primarily caused by the insufficient potable water supply and the harmful exposure to various chemicals and pathogens in the polluted water and food chain. Water pollution is largely defined by two main problems: the lack of safe drinking water and the dangerous exposure to various chemicals and pathogens found in contaminated water and the food chain [5,6]. Water pollution is characterized by the overabundance of harmful substances in water bodies, resulting from both natural and human activities [7,8].
The textile industry is a significant contributor to water pollution [9], and it is also responsible for approximately 20% of global water pollution [10], as the second largest polluter after the oil industry [10]. In comparison to other industrial sectors, the textile industry is known to have the highest water and chemical consumption, with over 8000 species being utilized [11,12,13]. The wastewater generated by this industry is often characterized by a significant amount of unfixed colors and dyeing auxiliaries [14,15,16,17]. Approximately 800,000 tons of dyes are produced annually, with 10–15% of this quantity being lost to the environment [18]. Over 10,000 distinct synthetic dye varieties have been introduced, with 70% of them belonging to the azo type. In general, dyes are classified into various types such as direct, reactive, basic, acidic, disperse, vat, sulfur, metal complex, and mordant dyes [10].
Dyes are a class of organic compounds that possess the ability to impart color to a diverse range of substrates [19,20]. Frequently, these compounds are recognized for their ionization properties and notable water solubility, leading to their facile dissemination into both the surroundings and human physiology [21]. The intricate aromatic structures of these substances pose a challenge for biodegradation and render them inert, thereby rendering their elimination a more arduous and laborious task. In contrast to metal ions, dyes can be classified in various ways. The most prevalent classification method is based on the charge exhibited upon dissolution, which leads to the formation of three distinct groups: anionic (inclusive of reactive, acid, and direct dyes), cationic (encompassing all basic dyes), and non-ionic (comprising disperse dyes). Dyes can be categorized into acid and base based on the various associated groups that dictate the hue of the color. Acid dyes are anionic chemicals containing acid moieties in their molecular structure such as sulfonic SO32− and carboxylic -COO¯, while base dyes are cationic ones presenting quaternary amine groups -NH4+ [22]. Another systematic method of classification is the color index, which is related to the chemical structure of the dye substance; however, due to the complexity of nomenclature from the chemical structure, the classification based on color application is the most preferable [13]. With respect to chemical structure, a variety of groups such as azo, diazo, anthraquinone, nitro, diphenylmethane and triphenylmethane, indigoid and thionindigoid, anthraquinoid, xanthene, phthaleins and metal complex dyes are known (Figure 1). Meanwhile, the mode of application and substrate-based scale classifies them into reactive, acid, base, vat, direct, solvent, disperse, and azoic dyes [23,24].
The classification of colored substances can be divided into two categories: natural and synthetic. Synthetic substances have become the predominant choice in the market due to their vast array of available colors and cost-effectiveness, as noted in [21]. The utilization of synthetic dyes, which are derived from benzene and its derivatives, has supplanted the use of conventional natural dyes, leading to the development of over 10,000 dyes with varying chemical structures and characteristics [25]. The compounds exhibit intricate conjugated architectures that pose challenges in terms of their elimination [25]. Certain dyes, such as azo dyes, possess a high degree of toxicity and carcinogenicity because of their toxic metabolites and aromatic amine byproducts. The removal of anionic and non-ionic dyes through conventional techniques poses a challenge due to their high water solubility and resistance to degradation of non-ionized fused aromatic rings, respectively. In the interim, it has been observed that biological techniques are not entirely effective in the complete elimination of reactive and acidic dyes [7]. In general, azo dyes exhibit a high susceptibility to degradation at their azo N=N linkage, leading to the formation of hazardous aromatic byproducts during treatment. Conversely, other categories of dyes are characterized by a low degradability, which limits the range of viable treatment options. Prior to discharge into the aquatic environment, it is imperative to subject the wastewater generated by the textile sector to appropriate treatment measures. Figure 2 illustrates the relationship between the denim factories and the resultant wastewater, which significantly contributes to the contamination of the Noyal River in Tirupur, India, as well as the adjacent agricultural areas. The production of denim entails the discharge of wastewater containing various pigments (Figure 2a), subsequently leading to the pollution of nearby water bodies (Figure 2b) and ultimately culminating in the contamination of the river (Figure 2c,d). The impact of the situation on agricultural activities is evident in the images, as depicted in Figure 2e. Notably, the groundwater is significantly affected, as seen in Figure 2g,h.
In recent years, there has been a significant focus on the removal of dyes from wastewater owing to their hazardous properties and conspicuousness also now this wastewater also contains the microplastics [27,28]. The presence of dyes in wastewater can have severe consequences, including the disruption of photosynthesis and oxygen deficiency caused by the obstruction of sunlight. These dyes can accumulate in food chains, leading to aesthetic issues in aquatic and soil environments. Moreover, their carcinogenic and mutagenic properties can result in the formation of tumors and mutations. The toxic amines and biologically non-biodegradable byproducts present in dyes can also have immunological and dermal effects. Additionally, the presence of dyes can disrupt seed germination and hinder plant growth, as well as cause chromosomal aberrations and various diseases [23,29]. To mitigate the detrimental impact of dyes and other chemical waste generated by the textile industry on both the environment and human health, it is imperative to extract the contaminated wastewater from the hazardous effluents prior to their release into water bodies. Therefore, the subsequent sections will present a concise review of the impact of health and environmental concerns linked to these effluents. There exists a more profound discussion concerning the significance of employing sustainable methodologies for the removal of effluents using various strategies. The primary objective of this review study is to consolidate and analyze the effects of wastewater containing dyes, as well as their possible implications for human health and the environment. Additionally, the dedicated section delineates a range of ecologically friendly approaches for the management and treatment of wastewater in the textile sector, together with its respective advantages and disadvantages.

2. Common Treatment Methods for Textile Dyes

The primary approaches utilized in the treatment of wastewater may be categorized into three distinct groups: biological, chemical, and physical. Table 1 provides a concise overview of the methodologies and presents a comprehensive analysis of their respective merits and drawbacks. One example of physical techniques is the utilization of membrane technology. On the other hand, chemical methods encompass processes such as oxidation, coagulation, and photochemical oxidation. Additionally, biological approaches include the implementation of anaerobic/aerobic sequential processes [30]. Oxidation is a chemical process that encompasses several techniques, including bleaching, chlorination, and ozonation. These techniques include the utilization of specific chemicals such as hydrogen peroxide, permanganate, chlorine, chlorine dioxide, and ozone (O3), respectively [13,24,31,32].

3. Effluent from the Textile Industry: Human and Environmental Issues

The effluents discharged by the textile industry in their untreated state consist of a wide array of organic contaminants, including unfixed colors, acids, alkalis, and notably, very poisonous dyes [70]. The textile business employs many categories of dyes, with azo dyes being the predominant group utilized, accounting for over 60% of the industry’s usage [71]. Azo dyes are characterized by their structural composition, which includes one or more azo groups. The discharge of unfixed azo dyes into wastewater is attributed to the inefficiency of textile dyeing processes, accounting for a range of 10–50% [29,72,73]. Certain textile manufacturing facilities employ wastewater treatment methods to break down the released free azo dyes in order to mitigate their impact on the environment. Conversely, there are other industries that release untreated industrial effluents straight into water sources, hence presenting significant ecotoxicological risks and causing harmful effects on organisms (see Figure 3). Farmers in various Asian nations, such as India, Bangladesh, Vietnam, and Indonesia, have historically employed the practice of irrigating their agricultural lands with untreated industrial effluents present in wastewater [74,75]. This practice has had detrimental effects on both soil quality and crop germination rates. Furthermore, the presence of toxic chemicals in these effluents has had a significant adverse impact on agricultural productivity, which in turn has had a notable influence on the gross domestic product (GDP) of these countries [76]. The introduction of azo dyes into water bodies has been seen to have detrimental effects on light penetration, hence negatively impacting the growth and productivity of algae and aquatic plants [77]. Additionally, the presence of these colors has been found to hinder the formation of dissolved oxygen (DO) in the water. Moreover, the ingestion of dyes by fish and other creatures can lead to the metabolic conversion of these substances into hazardous intermediates inside their systems, so exerting detrimental effects on the well-being of both the fish and their predators [78]. Azo dyes present in industrial effluents can potentially come into contact with humans and other mammals through two primary routes: oral consumption and direct skin contact [79]. The intestinal microflora present in the human gastrointestinal tract is responsible for the conversion of azo dyes into amino acids that possess toxic properties. These toxic amino acids have detrimental effects on numerous tissues inside the human body [70,80].

3.1. Environmental Consequences

The textile sector disposes of significant quantities of untreated wastewater which largely contains azo dyes and several other organic contaminants (Figure 3). Nevertheless, it should be noted that azo dyes undergo degradation either before or after being disposed of, resulting in treated effluents containing amino acids that are potentially more harmful than the original chemicals [71,81,82,83]. Conversely, untreated wastewater has a diverse array of detrimental effects on aquatic ecosystems and creatures. The introduction of textile dyes into aquatic habitats has been found to have adverse effects on the plants inhabiting these environments. One prominent natural concern associated with dyes is to the phenomenon of sunlight absorption and reflection in water [84,85]. It serves as a barrier that inhibits the penetration of light into the photic zone of the aquatic environment. Consequently, notable ecological ramifications ensue, including alterations in the characteristics of aquatic ecosystems and reduced photosynthetic activity relative to aquatic vegetation. Moreover, the presence of these waste liquids has been found to be associated with adverse health effects in humans, such as allergies, dermatitis, skin irritations, cancers, and mutations. Consequently, these waste liquids can contribute to the degradation of water quality, leading to the development of toxic properties, as evidenced by changes in odor and color [86,87].
As significant primary producers in freshwater and marine environments, microalgae are vital to aquatic ecosystems because they provide food for everyone from tiny zooplanktons to gigantic whales [88,89]. Particularly in aquatic ecosystems, the single-celled tiny algae known as microalgae are essential to the upkeep of the whole food chain [89]. Nonetheless, in aquatic environments, color pollution impedes the development of microalgae and interferes with the trophic transfer of nutrients and energy. The significant quantity of textile dyes discharged into water sources affects algae development. Algae are a great indication of pollution in toxicological studies because they are more vulnerable to pollutants than other aquatic species [88,89,90,91,92].

3.2. Impact of Textile Dyes on Human Health

Fish and other aquatic animals, which are widely acknowledged as a substantial protein source for human consumption, has the potential to ingest dyes through their diet [93,94]. The wastewater generated by the textile industry has significant coloration, a fluctuating pH, and contains various salts, alkalis, and acids, which contribute to elevated levels of biological oxygen demand (BOD), chemical oxygen demand (COD), total organic carbon (TOC), and suspended solids (SS) [95]. In general, the presence of SS hinders the passage of water across the fish’s gills, impeding the exchange of gases and potentially leading to reduced growth or mortality. In addition, prolonged exposure to textile effluents was found to diminish fish feed consumption, thereby leading to a decrease in the growth rate [96]. The genotoxic effects of reactive azo dyes on adult fish involve the promotion of erythrocytic micronuclei development, which is dependent on both the dose and duration of exposure. In fingerlings, the creation of gill micronuclei is also influenced by the duration of exposure to these colors. Fish are prone to a variety of illnesses due to the detrimental impact of hypoxia on their immune system and physiological responses (see Figure 4). Consequently, the presence of contaminated fish exerts a substantial influence on human well-being. Textile dyes are extremely deadly and include aromatic chemicals that have the potential to cause cancer [97,98,99]. They have been connected to a range of disorders in both humans and animals, including dermatitis and issues with the central nervous system [100,101,102]; Figure 4 lists these ailments in humans. Typically, there are two paths, which include textile dye ingestion or inhalation can irritate the skin and eyes [26,70,103], particularly if it occurs in dusty conditions [70]. Persons who work with reactive dyes run the risk of experiencing allergic responses, including occupational asthma, allergic conjunctivitis, and contact dermatitis. Textile dye genotoxicity is the biggest possible long-term risk to human health [97,98,100,104,105]. Certain dyes have the potential to cause mutagenic reactions; one such dye is Disperse Red 1 [106].

4. Sustainable Wastewater Treatment for the Remediation

4.1. Bioadsorbents in Wastewater Treatment

The utilization of conventional chemical coagulation methods results in the generation of sludge, which is then disposed of in landfills. This practice has been found to contribute to the emission of harmful components, including gases that have the potential to contribute to global warming. Additionally, the disposal of this sludge in landfills has risks such as landfill leaching and contamination of groundwater [107]. The introduction section of this article discusses the environmental threat posed by textile effluent containing high levels of color, BOD, COD, TDS, and TSS. Biological treatment is preferred over chemical treatment for sustainable treatment. Generally, the presence of complex groups in dyes, along with the recalcitrance of organic pollutants and their low degradability, restrict the efficacy of biological treatment methods [108]. Therefore, on this occasion, bioadsorbents play a significant role in the dye and heavy metal removal. The exploitation of domestic and agricultural wastes as adsorbents has emerged as a convenient alternative. Numerous adsorbents derived from biomass wastes have been created and utilized as very effective agents for the removal of various pollutants from water and wastewater. These waste materials have been used either in their original form or following suitable modifications. Various agricultural and food waste materials, such as Azolla [109], banana peel [110,111,112,113], cabbage waste [114], chitosan [115,116,117,118], citrus peel [119,120], Citrus limonum leaves [121], corn cob [122], orange peel [123,124], peanut hull [125], rice husk [126,127], sawdust [128], and sugar cane bagasse [129] have demonstrated successful utilization as adsorbents for the purpose of eliminating diverse types of contaminants.
Adsorption generally convert the pollutants from a liquid to a solid phase. This technique has several advantages, including simple, cost-effectiveness, convenience of operation, non-toxicity, and reactive surface atoms. Bioadsorbents are frequently employed for the treatment of textile effluent water owing to their economical, eco-friendly, locally accessible, sustainable, efficient, renewable, and readily disposable characteristics. They surpass commercially available activated carbon in terms of quality, rendering the latter’s high cost unjustifiable. Inexpensive sorbents possess a notable ability to absorb certain dyes, particularly reactive dyes, leading to the accumulation of significant amounts of hydroxylates in wastewater as a result of inadequate fixation of the dyestuff. Adsorption is advantageous over alternative approaches due to its simplicity, cost-effectiveness, ease of operation, non-toxic nature, presence of reactive surface atoms, and large surface area [130]. Currently, a global revolution is underway advocating for the recycling of organic wastes from agriculture, forests, and industries into economically viable products [130]. Some of the commonly used bio adsorbents and their nature of activity in treating textile effluent water are explained below.
The peel of Citrus limetta has been shown to be a cost-effective adsorbent for the removal of various colors [119]. Every year, a significant proportion of citrus fruit (~40% to 60%) is discarded in landfills. Research indicates that the global citrus processing industry generates a substantial amount of trash, estimated at approximately 120 million tons [120], creating serious ecological issue. As an example, orange peels are employed for the removal of 1-naphthyl amine dye from wastewater generated by the textile industry. The findings of the study indicated that the adsorption capacity of the peel waste had a positive correlation with the concentration of dye ions. Additionally, it was observed that the percentage of dye ion removal also rose as the original dye ion concentration increased. Furthermore, the utilization of orange peels in the preparation of activated carbon has proven to be effective as an adsorbent for the removal of MB [131]. Banana fiber is an economically accessible and abundantly available material, owing to its substantial cultivation and extensive presence as a crop, with a global count over 25 billion banana or plantain trees [132]. Banana powder has demonstrated promising potential as a biosorbent for the removal of MB dye. This is attributed to the presence of many functional groups on the surface of banana particles, as well as their uneven morphology [133]. Another study found that banana peel is particularly successful in removing reactive dyes, with 90% of the dyes being removed in 5 min [134]. The utilization of ash derived from banana stem as a potential bio adsorbent for dye removal has promising results. This is evidenced by its ability to achieve a 95% removal efficiency for MB dye [135]. The effectiveness of banana stem ash may be attributed to its diverse array of components and functional groups, as well as its rough and porous surface characteristics. Recent research provides further evidence supporting the removal of 91% of color from the Banana stem [136]. Some of the resent studies confirms that the waste extraction from coffee waste shows promising adsorbents for the dyes [137].
Coconut coir dust refers to a lightweight, porous particle that is separated from the husk during the process of fiber extraction. The weight of coir dust accounts for approximately 35% of the total weight of coconut husk. Coconut coir is comprised of cellulose, lignin, pectin, and hemicellulose. The presence of hydroxyl groups in cellulose and lignin facilitates the adsorption of dyes [138]. Bio chars produced from coconut coir have enhanced dye adsorption capabilities due to their significantly higher specific surface area [139]. The research focuses on investigating the efficacy of coconut shell-activated carbon as a means of removing direct yellow DY-12 dyes. The study demonstrates that the adsorption process is particularly effective under acidic pH conditions. The findings of the study indicate that the process of adsorption exhibits heterogeneity, characterized by the formation of many layers. Furthermore, the adsorption process was seen to be endothermic in nature and occurred spontaneously [130]. Once tea has been prepared, the residual leaves are classified as waste, similar to other forms of biomass. The abundant availability of this waste has led to the increased interest in utilizing discarded tea leaves as an adsorbent [140]. Given the abundance and easy accessibility of this trash, its conversion into an adsorbent is economically viable, offering the added benefit of waste management. The utilization of raw tea waste, as well as its chemically and magnetically modified forms, in conjunction with activated carbon, has been widely employed for the remediation of water contaminated with dyes. In this study, a batch scale reactor was utilized to manufacture and apply tea powder for the purpose of removing MB from an aqueous solution. The effectiveness of adsorption was seen to improve with longer contact time, higher solution pH values, and increasing dose of waste black tea powder [141]. The residual tea waste possesses a significant calorific value, making it suitable for utilization in steam generation within the textile sector following appropriate saturation [142].
The different form of chitosan (i.e., nanoparticles, derivatives, nanofilms, and nanofibers) is employed as a bio adsorbent. This application aims to substitute activated carbon in the pre-treatment of textile effluent, with a specific focus on the removal of metal ions, particularly chromium, as well as colors (Figure 5a). The ability to repeatedly utilize these bio adsorbents with diluted NaOH while maintaining the same level of efficacy is noted, rendering it an intriguing aspect [143]. Cactus juice and aloe vera juice were employed as flocculants for the treatment of textile effluent [144]. The color removal efficiency achieved above 85%. Furthermore, the removal efficiencies for total solids, suspended particles, and dissolved solids were found to be 90% [145]. The efficacy of water chestnut peel in the removal of cationic RhB shows promising results [146] Table 2 provides a comprehensive overview of the prior studies conducted on the treatment of textile wastewater using bioadsorbents. In recent studies, researchers have shown that chitosan modified cellulosic non-woven fabric has superior color removal capabilities for Reactive Red 198 [147]. Additionally, these non-woven fabrics have shown great potential as materials for dye removal in the future [148,149].
Various plant-based waste materials and biomasses have been found to have significant efficacy in the adsorption and retention of dyes (Figure 5b). The primary constituents of plant leaves encompass cellulose, hemicellulose, pectins, and lignin, and additionally it contains many functional groups, such as carboxyl, hydroxyl, carbonyl, amino, and nitro, which can interact with the functional groups of the dyes [150]. The adsorption of Acid Orange 52 (AO-52) dye using Paulownia tomentosa Steud leaves biomass showed promising results [151]. In a separate investigation, the adsorption of Acid Red 27 (AR-27), an anionic dye, was examined utilizing hyacinth leaves [152]. Basic Red 46 (BR-46) dye exhibited strong affinity towards pine tree leaf-based adsorbents [153]. Ashoka leaf powder exhibited interactive behavior towards rhodamine B (RhB), Malachite Green, and Brilliant Green dyes [154]. A novel lignocellulosic biosorbent material, obtained from fully developed leaves of the sour cherry plant (Prunus cerasus L.), has remarkable efficacy in the removal of Methylene Blue and crystal violet dyes [155]. The coffee waste demonstrates a characteristic three-dimensional carbon structure, with a rough surface and a porous system that allows it to function as a promising adsorbent for the removal of anionic CR and RB5 dyes from aqueous solutions [156]. The experimental findings indicate that the utilization of powdered lemon leaves resulted in the removal of Malachite Green up to a maximum efficiency of 82.21%. The highest sorption capacity (qmax) of lemon leaf powders is 8.08 mg/g [157]. In another study, the cationic amino modified banana leaves show the excellent sorption for Congo Red (CR) dyes [158]. Table 2 presents a comprehensive overview of recently studied bio adsorbents, including their respective adsorbent capacities in relation to various dyes.
Figure 5. Different types of bio adsorbents for the dye removal; chitosan (a) (reprinted from [159]); tea waste (b) (reproduced from [160] as distributed by Creative Commons Attribution License).
Figure 5. Different types of bio adsorbents for the dye removal; chitosan (a) (reprinted from [159]); tea waste (b) (reproduced from [160] as distributed by Creative Commons Attribution License).
Sustainability 16 00495 g005
Table 2. Studies on bio adsorbents on textile wastewater treatment.
Table 2. Studies on bio adsorbents on textile wastewater treatment.
Name of AdsorbentsPerformed DyesAdsorption ConditionsRemoval (mg/g)Refs.
Potato peel-based sorbentDirect Blue 71pH 31704[161]
Rice husk ashBrilliant Green dyepH 4–1066[162]
Sunflower stalkBasic Red 9 dye-317[163]
Cane pithBasic Red 22 dyepH 4.1941.7[164]
BagasseBasic Red 22 dyepH 4942[165]
Enosis siliqua shell powder--797[166]
Glutaraldehyde cross-linked magnetic chitosan beadsDirect Red 23pH 41250[167]
Popcorn derived activated carbonMethyl OrangepH 2–112090[168]
Carboxymethyl cellulose-g-poly(2-(dimethylamino) ethyl methacrylate) hydrogelMethyl OrangepH 21825[169]
Non-cross-linked and cross-linked chitosan fibersAcid Orange 7pH 74523[170]
Chitosan grafted with diethylenetriamineAcid Orange 7-2108[171]
Chitosan grafted with poly(methyl methacrylate)Reactive Blue 19pH 31498[172,173]
Chitin nanofiber-/nanowhisker-based hydrogelsReactive Blue 19pH 11331[172]
Cationic cellulose nanocrystals-chitosan film (nanocomposite)Reactive Blue 19pH 31320[174]
Hollow zein nanoparticlesReactive Blue 19pH 91016[175]
Chitosan filmsReactive Blue 19pH 6.8822.4[176]
Template ECH cross-linked chitosan nanoparticlesReactive Black 5pH 32941[177,178,179]
Chitosan beads cross-linked with epichlorohydrinReactive Black 5pH 32043[180,181]
Glutaraldehyde cross-linked chitosan beads/microparticlesReactive Black 5pH 101927[182]
Chitosan cross-linked with sodium edetateReactive Black 5pH 31648[183]
Chitosan hydrogelReactive Black 5-1560[184,185]
Mango bark powderMalachite Green dyepH > 64.22 × 103 mol/g[186]
Calcium-rich biocharMalachite Green dyeNeutral and alkaline pH12,502[187]
Pigments-extracted macro algae derived biocharMethylene Blue-5306.2[188]
Azolla-derived hierarchical nanoporous carbonsMethylene Blue-4448[189]
BananaReactive Blue 235
Methyl Red,
Malachite Green
--[190]
Activated surface of banana and orange peelsReactive Red 24--[191]
Waste tea residueAcid Blue 25--[142]
Palladium nanoparticles synthesized from peel waste of cotton bollToxic azo dye--[192]
Wheat husk wasteTextile effluent water --[193]

4.2. Dye Removal by Biological Methods

Although it is true that certain microorganisms can degrade auxochromes and chromophores found in dyes, hence facilitating the removal of organic materials from textile waste, it is worth noting that some of these microorganisms are also capable of mineralizing colors into carbon dioxide and water (see Figure 6). The rationality of color removal in biological processes, even conventional ones, has not been empirically shown Figure 6. The rate of removal is contingent upon several factors, including the concentration of O2, the ratio of organic load to microorganism load and dye load, and the temperature range [58,194]. Table S1 presents a comprehensive compilation of the merits and demerits associated with diverse biological techniques employed in the elimination of dyes.

4.2.1. Biological Route of Dye Decolorization

Enzymes are employed in the field of biological remediation because to their multifunctionality, effectiveness, and capacity to break down organic waste materials. The effectiveness of oxidative enzymes, such as laccases, peroxidases, and tyrosinases, has been demonstrated in the conversion of hazardous waste into insoluble compounds that may be easily separated [12,57,58,195]. The utilization of microbial enzymes for the degradation of azo dyes has been extensively investigated. Laccases are a class of enzymes that possess copper ions within their active sites, enabling them to facilitate the oxidation reactions of phenolic compounds and aromatic amines. Peroxidases are a class of enzymes that include heme and facilitate the oxidation of organic molecules using hydrogen peroxide as a co-substrate. In contrast, azoreductases are a class of enzymes capable of reducing azo dyes to their respective aromatic amines, resulting in reduced toxicity compared to the original molecules [196,197]. Laccases have demonstrated significant promise in facilitating the oxidation of a wide range of substituted phenolic and non-phenolic chemicals [198]. These organisms are found in several ecosystems, encompassing fungi, plants, and bacteria, exhibiting a broad distribution. Laccases are enzymes that exhibit a notable characteristic of not necessitating peroxidases for their catalytic activity. Instead, they employ molecular oxygen as the primary electron acceptor, rendering them highly prevalent in the enzymatic breakdown of azo dyes. Nevertheless, the economic viability of enzymes has been hindered by their inherent instability, variable activity, and labile characteristics. Enzymatic treatment of effluent water has the potential to achieve a significant reduction in coloration, with removal rates of up to 90% [199].
Bacteria can be employed to remove dye decolorization in specific cases, resulting in a 30% reduction in effluent toxicity [200,201]. Bacteria and fungi exhibit short lifespans and instability, while enzymes, although more effective, are associated with higher costs. However, stability concerns persist in both bacteria and fungi as well [83,202]. In the context of bacterial color biosorption, it has been shown that corynebacterium glutamicum exhibits potential as a biosorbent for Reactive Black-5, with a sorption capacity of 257 mg/g at a pH4 [203]. There exist two prevalent categories of microscopic organisms, namely Gram positive and Gram negative. Microscopic organisms participate in the formation of a dense peptidoglycan layer that is interconnected by amino-acidic linkages. Polyalcohols, which are lipids linked to form lipoteichoic acids known as Gram positive, are present within the cell divider [204]. The degradation of dyes by different bacterial biomass is comprehensively elucidated in the provided Table S2.
Numerous research groups have investigated the ability of microorganisms to digest azo dyes. Pseudomonas bacteria do not readily use azo dyes in aerobic environments. Despite the interruption of metabolic pathways by the intermediates generated during these degradative steps, the trash was not subjected to mineralization. In environments without oxygen, a multitude of microorganisms are known to enzymatically degrade azo dyes through the activity of soluble, unspecified cytoplasmic reductase enzymes, commonly referred to as azoreductases. Enzymes facilitate the production of colorless aromatic amines, which have the potential to exhibit mutagenic, fatal, and potentially carcinogenic effects on organisms. The existing body of research suggests that there are many supplementary methods that may be employed for the reduction of azo dyes. A wide range of microorganisms have the ability to biodegrade both sulfonated and non-sulfonated azo dyes under anaerobic conditions. Furthermore, a multitude of highly charged atomic, polymeric, and sulfonated azo dyes have an inability to traverse the cell membrane. Therefore, the capacity to remove dye is not attributed to the intracellular accessibility of the azo dye [204].

4.2.2. Fungi

Fungal mycelia possess an advantage over unicellular organisms due to their ability to solubilize insoluble substrates, hence producing extracellular enzyme catalysts. Organisms have enhanced enzymatic and physical interactions with their environment because of an increased ratio of cell surface area to volume [205,206]. Several types of fungi, including white-rot fungi, Aspergillus niger [206], Rhizopus arrhizus [207], and Rhizopus oryzae [208], have been found to possess the ability to degrade a wide range of colors.

4.2.3. Algae

Algae possess an abundance of enzymes and other compounds that contribute to the process of dye decolorization in textile effluent. Some algae can metabolize dyes through enzymatic processes, leading to the breakdown and detoxification of these substances. Chlorella vulgaris has been effectively used to remove dyes, including Congo Red, Brilliant Blue R, and Remazol Brilliant Blue R, from wastewater. Algae have unique metabolic abilities that enable them to efficiently remove or break down contaminants, such as dyes [209,210]. Cosmorium sp. has been investigated as a biosorbent for the removal of Malachite green (MG) dye, resulting in a significant removal efficiency of 92% [211]. Several studies have demonstrated the considerable influence of algae on the process of decolorization. For instance, Cosmarium sp. achieved a noteworthy 74% elimination of malachite green [212]. Similarly, Azolla rong pong exhibited decolorization rates of 30% for Acid Green-3 [213] and 43% for Acid Blue-15 [214]. In a separate investigation, Ulva prolifera demonstrated a remarkable 96% efficacy in the removal of Acid Red-274 colorant [215].

4.2.4. Enzymes

Enzymes are frequently employed for effective dye removal from wastewater, with fungi being the primary source of these enzymes. Additionally, laccase is an enzymatic protein that belongs to the class of copper-containing polyphenol oxidases, which are synthesized by many species of bacteria and fungi. The utilization of this technology has been employed for the decolorization of azo dyes. Peroxidase is an enzymatic catalyst responsible for the decomposition of hydrogen peroxide (H2O2) and the promotion of oxidation reactions involving various substrates through the use of molecular oxygen. The utilization of lignin peroxidase has been previously applied for the purpose of removing CR dye. Furthermore, it has been utilized for the purpose of decolorizing Reactive Orange 16 [216,217,218,219].
The mutant laccase enzyme was elevated in Escherichia coli, resulting in the decolorization of indigo carmine by more than 92% (see Figure 7a) [220]. Azoreductases are a distinct class of flavoenzymes that have the potential to facilitate the reduction of azo bonds (–N=N–) present in the aromatic dyes by hydrolysis of azo bonds in the presence of oxygen or in the absence of oxygen, hence participating in the metabolic breakdown of dyes (see Figure 7b) [221]. This process yields aromatic amines, which are subsequently eliminated by microbial enzymes including mono- and di-oxygenases, as well as hydrolases [222]. In addition, azo dyes are capable of reducing the toxicity of nitro-aromatic compounds by reducing their concentration. Azoreductases are typically pH-stable within the range of 5–9, and their activity is at its peak under physiological conditions [197]. In 2019, Dong et al. [223] determined that the use of azoreductases derived from Streptomyces species shown efficacy in the elimination of MR from wastewater. In the same year, Sherifah et al. [224] employed the utilization of Kluyveromyces dobzhanskii bacterial laccase for the purpose of enzymatically degrading MG and MR dyes. In a separate investigation, yeast laccase derived from Yarrowia lipolytica was utilized to enzymatically degrade Bromocresol Purple, Safranin, Bromothymol Blue, and Phenol Red. Additionally, the process of isolating laccase from the edible fungi species Agaricus bisporus was conducted, with the intention of using this enzyme for the purpose of enzymatically degrading Acid Blue [225].

4.2.5. Bacteria

Bacterial isolates provide a viable and ecologically sustainable approach for the breakdown of dyes. Bacterial isolates possess enzymatic systems that enable them to eliminate dyes via degradation or biosorption mechanisms [226]. Bacterial organisms have the advantageous characteristic of exhibiting shorter development durations, in addition to their capability to breakdown and mineralize dyes [227]. Numerous investigations have employed bacterial strains for the purpose of degrading and decolorizing diverse textile effluents that contain azo dyes, such as MO (i.e., color removal of 95%), Reactive Yellow (84%), and Reactive Black 5 (100%) [228,229,230,231,232,233]. Additionally, it shows better color removal with anthraquinone-based dyes like Remazol Brilliant Blue R [234], disperse blue 2BLN (decolorization rate of 93.3%) [235].

4.3. Membrane Separation

Membrane separation technology is commonly employed for the treatment of effluents generated by textile dyeing processes. During the filtering process, the micropores included in the membrane filter effectively separate the organic compounds from the effluent by utilizing selective membrane permeability. The classification of this phenomenon encompasses four distinct categories, namely ultrafiltration (UF), nanofiltration (NF), reverse osmosis (RO), and forward osmosis. The process of separation may be effectively achieved by the utilization of UF, which has shown great potential as a technique. The elimination of dissolved compounds occurs at a reduced transmembrane pressure through the utilization of UF. The utilization of polyelectrolyte complexes, in conjunction with cellulose acetate and inert polymers, is applied in the production of UF membranes that exhibit the capacity to efficiently regulate flow. The normal range for pore size is between 0.001 and 0.02 μm. NF is an intermediate technique between reverse osmosis and ultrafiltration, characterized by the use of membranes with nanometer-scale pores (0.5–10 nm) and operating at pressures of 5–40 bar. NF is a very sophisticated membrane-based technique that demonstrates remarkable efficacy in the removal of heavy metals [236,237,238]. The membranes of NF possess a thin outer layer that is typically non-porous, operating at the nanoscale, and exhibiting a high level of permeability [62]. One of the primary benefits of NF is its reduced energy consumption, which leads to a higher efficiency in the removal of contaminants [239]. Presently, several textile industries employ RO as a means of treating their effluent. RO is categorized as a membrane-based technique. The RO membranes effectively capture suspended particles through their small pores, hence mitigating fouling. The pre-treatment procedure plays a crucial role in the regulation of turbidity levels and fouling tendencies [62]. Figure 8 illustrates the classification of membrane filtering techniques together with their respective advantages and disadvantages.
Figure 9 depicts the design of a textile wastewater treatment plant, presenting a comparative analysis between the conventional approach and the membrane bioreactor-nanofiltration (MBR-NF) technique. The use of a plant-based system that incorporates membrane bioreactor (MBR) and NF technology has promise in reducing the dependence on anaerobic, aerobic, coagulation, and decolorization dosing/sedimentation tanks [240]. The introduction and utilization of an MBR-NF treatment facility has resulted in a substantial decrease in operational costs. Moreover, this sophisticated technology has the benefit of occupying a smaller physical footprint when compared to conventional treatment facilities. Figure S1 illustrates the application of combined coagulation–flotation with forward osmosis technology for wastewater treatment. The removal efficiency in this method is characterized by its high-water flow and high recovery rate. The extent of membrane fouling is rather little; yet it engenders adverse consequences on the environment. The experimental setup employed in the forward osmosis system involved the use of a fabricated forward osmosis membrane, namely a plate-and-frame configuration, with a surface area of 10 cm2 [241]. A spacer-free rectangular canal is installed on both sides of the membranes. In the first stages of forward osmosis, the amount of wastewater is reduced by employing osmosis to extract water from the wastewater, hence increasing the concentration of dye in the remaining solution [242].

4.3.1. Ion Exchange

The term “ion exchange” refers to the reversible process of exchanging ions between a liquid and a solid, without causing any significant alteration to the solid’s structure. Ion exchange is a widely employed method for the widespread removal of inorganic salts and organic anionic constituents, such as antibiotics, amino acids, organic acids, and small compounds [67]. The classification of ion exchange resins is based on the functional groups present, resulting in three main types: anion exchange resin, cation exchange resin, and chelating exchange resin. The substance in question may be categorized as either natural or synthetic, and it possesses several notable advantages, including affordability, minimal equipment requirements, straightforward operation, and the absence of solvents [67]. One significant limitation of this process is the extended duration required for production, as well as the substandard quality of the resulting product. Additionally, the high pH levels and the potential transfer of dirt and contaminants from the effluent to the sludge pose further challenges. Furthermore, the introduction of chemicals for sludge regeneration exacerbates these issues [67,68,243].

4.3.2. Evaporation

During the evaporation process, the concentrated textile effluent is subjected to various evaporator systems and steam within the evaporator. The attainment of the desired salt concentration or specific density is achieved only by the recirculation of the liquid during the evaporation process. The condenser is responsible for gathering the vapor and steam, and the evaporator temperature is subject to variation based on the length of the tubes. Textile effluent is commonly subjected to evaporation through two primary methods, namely solar evaporation and mechanical evaporation [195,236].

4.4. Other Techniques

4.4.1. Granular Activated Carbon (GAC)

Carbon is a non-metallic element that is abundantly present in nature and finds extensive use across many applications in daily human existence. Graphite has a wide range of applications, including as a source of fuel, lubrication, material for pencils, electrodes, and as a means of water filtration [244]. Activated carbon refers to a kind of carbon that has been specifically engineered to possess small, low-volume holes and a significantly increased surface area. This enhanced surface area facilitates the process of adsorption or chemical reactions, hence enabling the purification of both liquids and gases. GAC refers to a specific form of carbon that is capable of being retained in a 50-mesh sieve [245,246,247]. This type of carbon can be obtained from various sources by different extraction procedures and with varying degrees of activation. The substance is offered in many forms, including granules, powder, and pellets. Activated carbon is commonly derived from many sources such as coconut shell, hard and soft wood, peat, olive pits, lignite, and bituminous coal using chemical or steam-based processes. The activated carbon has a surface area of 500 m2/g, indicating its porous characteristics. Various studies have been conducted utilizing a range of biomass materials such as bagasse, coal, rice husk, coconut husk, nutshell, lemongrass, sawdust, cocoa shells, grape peels, and cassava peels. These biomass materials have been subjected to activation processes involving ZnCl2, phosphoric acid, microwave assistance, microwave assistance combined with KOH activation, and steam pressure. The objective of these studies is to investigate the efficacy of these activated biomass materials in the removal of dye from effluent water [245,246,247,248].

4.4.2. The Advanced Oxidation Process (AOP)

The AOP is mostly observed in the field of water purification, but more recently, it has been employed for the remediation of textile effluents. Hydroxyl or sulphate radicals are liberated in sufficient amounts to facilitate the elimination of both organic and inorganic substances, pollutants, and to enhance the water’s biodegradability. In comparison to chlorine and ozone, these substances exhibit superior performance in terms of water decontamination and disinfection. Various categories utilize the hydroxyl radical. Various methods have been employed in the field of environmental remediation, including UV-based processes, ozone treatment, Fenton reactions, and the utilization of sulphate radicals, among others, additionally UV has advantages for disinfection properties. The different advanced oxidation processes are illustrated in Figure 10a. The AOP is well recognized as a prominent technique for the treatment of industrial wastewater, owing to the considerable oxidative potential shown by ozone and the resulting generation of hydroxyl radicals (OH) [249]. Extensive research has been conducted on the application of ozone-based AOPs in both simulated and actual environmental circumstances. The use of auxiliary agents in the dye and their impact on dye degradation, as well as the influence of different salts on the process of ozonation, were investigated through the application of the AOP [250]. The AOP has gained significant popularity in the field of leachate treatment and water reuse [251]. There exist several forms of AOPs, including ozone, ozone/hydrogen peroxide, ozone/UV, UV/TiO2, UV/hydrogen peroxide, Fenton reactions, Photo-Fenton reactions, ultrasonic irradiation, heat/persulfate, UV/persulfate, Fe(II)/persulfate, and OH-/persulfate [252].

4.4.3. Color Removal by Fenton Oxidation

The Fenton oxidation method is a very promising technique for the treatment of textile wastewater due to its cost-effectiveness and ease of implementation [48]. The major objective of Fenton oxidation is the decolorization of the effluent, although it also possesses the ability to degrade organic pollutants (Figure 10b). Hydrogen peroxide can be employed as an oxidizing agent, either in the presence or absence of a catalyst. Notable catalysts that can be utilized include ferrous salts, Al3+, and Cu2+ [49]. The efficacy of Fenton’s reagent has been demonstrated in the treatment of many types of industrial effluent as well as a wide range of dyes. The Fenton process demonstrates a high level of effectiveness in removing color, with an efficiency of 98% achieved at a pH of 3. Similarly, the Fenton process exhibits a significant capability for removing COD, with an efficiency of 85% achieved at a pH of 3 [49,51]. The most efficient decolorization of effluent for all dyestuffs occurs at a pH value of 3, within the range of 2.5–4. The utilization of this reduced value is attributed to the substantial production of OH [51]. When H2O2 and (Fe2+) are combined under these specific pH conditions, hydroxide ions (OH) are generated by a complicated series of interconnected reactions [51,253]. Figure 10c illustrated the decolorization of an indigo-dyed pollutant to colorless.
Figure 10. Different advanced oxidation processes (a) reaction during Fenton and hydrogen peroxide-based oxidation (b) and the color removal with different time spans of the reactions with Fenton oxidation (c) (reprinted [254] with the permission of Springer Publications).
Figure 10. Different advanced oxidation processes (a) reaction during Fenton and hydrogen peroxide-based oxidation (b) and the color removal with different time spans of the reactions with Fenton oxidation (c) (reprinted [254] with the permission of Springer Publications).
Sustainability 16 00495 g010

4.4.4. Color Removal by Peroxide (H2O2)

Hydrogen peroxide has a high degree of efficiency and contains the OH radical, which is accountable for both the chemical breakdown and mineralization of organic molecules, and is generated by the reaction involving another oxidant, H2O2. Furthermore, treatment of halogenated substances results in the generation of non-hazardous halide ions and non-toxic molecules, including carbon dioxide (CO2) and H2O [255]. A notable observation is that the efficiency of H2O2 addition in a recirculated photoreactor is significantly higher when performed in a single-step manner, as opposed to multiple-step addition [255]. Due to its short lifespan, the generation of OH occurs in situ by the reaction induced by UV irradiation, as follows,
H2O2 + UV = 2 OH
The breakdown of organic pollutants is facilitated by the OH radical through four primary routes, radical addition, hydrogen abstraction, electron transfer, and radical combination [256,257]. The application of H2O2-UV results in the degradation of the chromophore configuration of the dye, leading to its decomposition in normal environmental circumstances. This process generates O2, which may be effectively utilized for aerobic treatment [256]. The effectiveness of wastewater decolorization is enhanced in an acidic environment [256,257]. H2O2 is utilized in the oxidation of alkali, resulting in the formation of O2 and H2O. This process generates accessible hydrogen peroxide for the hydroxyl radical (OH). The reduction in radical production leads to a decrease in decolorization efficiency [256].

4.4.5. Ozonation

Ozonation is considered as an environmentally sustainable method for treating wastewater owing to its lack of residue production and absence of chlorinated byproducts, which are known to be toxic. This process effectively oxidizes color, odor, and bacteria without generating any detrimental substances [258,259,260]. The decomposition of organic compounds, detergents, and phenols into smaller molecular components is aided by the process of oxidation, which may be achieved using commercially available sodium hypochlorite. As a potential alternative, the implementation of ozonation might be regarded as a feasible solution to supplant the utilization of hypochlorite [40,261,262,263,264]. Typically, the process is conducted at alkaline conditions, characterized by a pH greater than 9, as the degradation of ozone in water is enhanced under such circumstances. The process of oxidizing inorganic compounds and dissolved organic molecules with ozone involves two distinct processes. The direct reaction of ozone molecules exhibits a higher degree of selectivity, characterized by a relatively slow reaction rate. This reaction is particularly advantageous in acidic conditions. The indirect response exhibited by free radicals, including OH and HOO, is characterized by reduced selectivity and a preference for basic conditions. Another noteworthy characteristic is that the reactivity of the dye is enhanced when it possesses an electron-donating group at its ortho and para locations, as opposed to an electron-withdrawing group [71,265,266]. The ozonation process is impeded by the presence of salts, such as NaCl or Na2SO4. However, it is worth noting that the presence of NaCl is more undesirable compared to Na2SO4. This is since Na2SO4 generates sulfate or peroxysulfate radicals, which might somewhat facilitate the ozonation process [259]. The use of an ozonation membrane biological reactor offers a means to enhance the removal of harmful substances such as pesticides, while simultaneously reducing the reliance on traditional coagulation methods. This innovative approach also lowers the need for additional biological treatment, resulting in a simplified operational procedure [41,267]. In many instances, ozone is employed in conjunction with UV or hydrogen peroxide to achieve enhanced efficacy. The utilization of ozone in conjunction with UV radiation facilitates the activation of ozone molecules, hence facilitating the creation of hydroxyl radicals (OH). UV radiation assists in expediting the oxidation process by aiding in the completion of the process. The exclusive use of O3 may result in incomplete conversion of organic compounds into CO2 and H2O in some cases. In addition, the intermediate process involving H2O2 undergoes photolysis, resulting in the formation of a hydroxyl radical which subsequently decomposes the dye molecules. When H2O2 is employed in conjunction with ozone for oxidation applications, it functions as a catalyst to enhance the generation of hydroxyl radicals (OH) through the breakdown of ozone. The interaction between hydrogen peroxide H2O2 and O3 has a sluggish rate under acidic pH conditions, but it undergoes a significant acceleration the in-reaction rate at elevated pH levels.

4.4.6. Photocatalytic Oxidation

The photocatalytic process is often regarded as the predominant method for treating textile effluent water due to its distinct advantage of optical absorption, a characteristic not shared by other AOPs. Photocatalysts could be stimulated by radiation, resulting in the generation of exceptionally reactive photo-induced charge carriers (free radicals, notably hydroxyl radicals (OH·) that engage in chemical reactions with contaminants [90,91,268]. These free radicals facilitate the oxidation of organic molecules, leading to their full conversion into non-toxic chemicals, such as carbon dioxide CO2 and H2O by the absorption of photons. The mitigation of pollutants carried out under ambient temperature and pressure conditions has the potential to successfully address the issue of excessive energy consumption associated with conventional approaches, such as electrochemical technology. The systematic exploration of photocatalysis research started in the early 1970s. The Honda–Fujishima effect was investigated by Fujishima et al. [269], who employed solar energy and titanium dioxide as photocatalysts to carry out water breakdown and hydrogen reduction activities. The utilization of light as an essential factor in this process allows for the possibility of designing the reactor or photocatalyst in a manner that enables the utilization of sunlight as a cost-effective energy source to drive the reaction [270]. The photocatalysts utilized in this study must possess characteristics that are conducive to their practical application. These characteristics include ready accessibility, reproducibility, photoactivity, non-toxicity, non-corrosiveness, biological or chemical inertness, affordability, and compatibility with near UV or visible light wavelengths.

4.4.7. The Sequencing Batch Reactor (SBR)

The implementation of the Sequencing Batch Reactor (SBR) was undertaken with the objective of mitigating the presence of nitrogen and phosphorus in piggery waste, as well as facilitating the biodegradation of sulfonated azo and diazo reactive colors found in textile effluent [271,272,273]. The study has presented findings on the effectiveness of SBR in the elimination of azo dye. It has been proposed that the inclusion of aerobic bacteria capable of degrading amines in the SBR system might enhance the complete mineralization of reactive azo dye under anaerobic conditions [272,274]. According to reports, it was observed that the decolorization reached a maximum of 97%, while the elimination of COD reached up to 98%. It was also suggested that the rate at which dyes are removed by SBR is influenced by the volumetric dye loading rate [275]. At present, bipolar membrane electrodialysis (BMED) presents itself as an environmentally conscious and sustainable method for commercial implementation. This is achieved by the integration of the traditional electrodialysis process with water dissociation within a bipolar membrane. The process of converting salt into base and acid by BMED from wastewater with elevated salinity and organic content presents a novel approach to the recycling of raw materials and achieving zero liquid discharge [276,277].

4.5. Treatment of Dyes Using Hybrid Technologies

In recent times, there has been a growing interest in hybridized techniques. The importance of a process that combines elements from a blend process measure may be described as “synergistic” and “combinatorial” in nature. These solutions are characterized by their efficiency since they include the utilization of a single container to execute several tasks. The technique shown in Figure 11 is an integrated or blended approach for the treatment of dye waste. Furthermore, it is important to clearly articulate and acknowledge significant modifications in the advantages of hybrid methodologies. The subdivision of the hybrid technique, along with its associated benefits, is presented in Table S3.

4.5.1. Physiochemical Methods

The addition of chemical-physical techniques is a further consideration to be considered in the removal of textile effluents. A photocatalytic/membrane separation (PMS) system consists of a photocatalyst, TiO2, dispersed on a membrane, which is then placed in a photoreactor. Blended submerged membrane photoreactor (sMPR) frameworks exhibit superior photocatalytic removal efficiency compared to PMS due to the utilization of immobilized TiO2 on the membrane surface. The PMS framework facilitates the separation of TiO2 particles for the purpose of reuse. The hybrid process, known as PMS, utilizes a compact reactor and demonstrates energy efficiency while effectively eliminating complex and toxic pollutants [278,279]. This study investigates the photodegradation of Reactive Black 5 (RB5) in a slurry membrane reactor, examining both batch and continuous operational modes. The rate of color removal is higher at lower initial dye concentrations, and it increases as the concentration of ZnO increases up to 1.25 mg/L. The highest rate of RB5 removal, close to 100% within 60 min, was observed at pH 11 due to the combined effects of dye photolysis and the photoactivity of ZnO [280].

4.5.2. Biochemical Methods

Biological approaches represent costly and ecologically sustainable techniques for the removal of intricate azo dyes from waste effluents. However, their efficacy is limited when it comes to the elimination of various dye kinds. In contrast, it has been shown that aromatic amines, namely the molecules of dyes, exhibit a notable level of resistance against the process of biodegradation. The presence of degraded byproducts from azo dye effluents has been observed to impede cell motility and metabolic activity, hence hindering the efficacy of biological treatment approaches [281,282].

4.5.3. Combination-Based Hybrid Chemical–Chemical Scheme

In recent times, there has been a growing interest in the utilization of chemical methodologies in combination. Various AOPs, such as coagulation, utilization of Fenton chemicals, sono-photocatalysis, and the implementation of the photocatalytic hybrid Z-scheme, have demonstrated efficacy in the degradation of diverse hazardous and organic pollutants. Photocatalysis, specifically AOPs, facilitates the acceleration of a photoreaction by the absorption of photons from light. This absorption leads to the generation of electron (e)/hole (h+) pairs, which actively engage in the redox reaction responsible for the degradation of pollutants [283,284,285].

4.5.4. The Z-Scheme Strategy

Despite the development of several wastewater treatment systems, there are still obstacles in the pursuit of efficient approaches to address pollution from complex waste effluents containing organic molecules and heavy metal ions, among other substances. The hybrid Z-scheme system is acquired by the careful selection of two semiconductors based on their suitable bandgap structure, as illustrated in Figure 12a. The design of highly effective Z-scheme heterojunction photocatalysts is a crucial and formidable task in the realm of wastewater purification utilizing solar energy. According to reports, organic pollutants such as dyes, phenol derivatives, and antibiotics can undergo effective degradation by oxidation by photogenerated holes (h+), superoxide radicals (O2), and hydroxyl radicals (OH) [286,287,288,289,290].
When two semiconductors are brought into contact, there is a phenomenon of charge separation that takes place at the interfaces. This separation is primarily caused by the difference in Fermi levels between the two semiconductors, as seen in Figure 12b. The semiconductor denoted as S1, which possesses a relatively low Fermi energy level, undergoes the acceptance of electrons from the semiconductor denoted as S2, which has a higher Fermi energy level. Consequently, an internal electric field is established at the interface, directing the flow of electrons from S2 to S1, as seen in Figure 12b. As depicted in Figure 12c, the interface exhibits an electric field that exerts a propulsive influence on the recombination of non-utilizable photogenerated electrons in the conduction band of S1 and non-utilizable photogenerated holes in the valence band of S2. Additionally, this electric field retains the electrons and holes, thereby enhancing their capacity to engage in the redox reaction within the conduction band of S2 and the valence band of S1, respectively. Consequently, these processes contribute significantly to the photocatalytic activity [291].

4.6. Sustainable Sludge Management

Sludge refers to the leftover, semi-solid substance that remains after the treatment of wastewater generated from textile processes. During the physical-chemical treatment process, the release of heavy metal concentration results in the formation of a chemical sludge. In contrast, impoverished soils might experience an additional advantage through the application of nutrient-rich biological sludge, which contains nitrogen and phosphorus, as well as useful organic matter. The primary issue is in the expeditious and forceful way sludge contaminates water sources. However, it is worth noting that certain locations within developing nations continue to dispose of sludge in inappropriate and environmentally unfriendly ways, such as through land disposal or by releasing it into the sea. To attain sustainable development, it is imperative to employ efficient recycling methods and utilize waste materials properly, rather than resorting to burning or landfilling, which can result in the deposition of hazardous substances such as heavy metals. Practicing prudent management is crucial when it comes to the disposal of sludge, notwithstanding the inherent challenges involved. Various elements play a significant role in the management of sludge, including local and national geographical considerations, agronomic issues, economic aspects, and stakeholder perception [292,293].

Methods of Sludge Treatment

Anaerobic digestion refers to the process of decomposing sludge in an atmosphere devoid of oxygen. The significant characteristics of anaerobic digestion are the reduction in mass, formation of methane, and enhancement of dewatering qualities in the fermented sludge [292,293,294,295,296,297]. A higher level of investment in the digesting chamber is associated with a slower pace of deterioration. To enhance the biodegradability of sludge, several pre-treatment methods may be employed. These include thermal pre-treatment, enzymatic treatment, ozonation, chemical solubilization by acidification or alkaline hydrolysis, as well as mechanical sludge disintegration and ultrasonic pre-treatment [298,299]. The hypothesis posits that the process of anaerobic digestion of textile waste leads to the generation of biogas, as evidenced by the works [300,301,302]. This relationship is depicted in Figure 13.
Aerobic digestion refers to the utilization of microorganisms within an oxygen-rich environment to facilitate the oxidation and decomposition of organic matter sludge. Aerobic sludge digestion is a procedure employed to decrease the levels of organic and inorganic constituents, as well as the overall volume, of sludge. The process under consideration exhibits temperature sensitivity and is susceptible to the presence of heavy metals, among other factors. However, it is noteworthy that despite its significant energy requirements, this process does not generate byproducts such as methane [303,304,305]. The pace of anaerobic digestion is constrained by the hydrolysis of organic materials in sludge, which subsequently leads to an increase in biogas output. This process serves as a pre-treatment technique that enhances the dewatering characteristics of the digested sludge [306]. The process of stabilizing sludge solids involves the application of several chemical treatments to the sludge in diverse manners. The application of polyelectrolytes as a conditioning agent for sludge dewatering operations has gained popularity due to its ability to enhance process yields [306,307,308].

4.7. Roadmap towards ZLD: Focus on Recovery and Reuse

Due to heightened environmental consciousness, escalating expenses related to wastewater treatment, and challenges pertaining to its disposal, there has been a perceptible shift in the public’s perspective on wastewater. Presently, zero liquid discharge (ZLD) is emerging as a prospective preventative measure that plays a significant role in safeguarding the environment against the adverse impacts of industrial activity. The ZLD method in the textile industry focuses on achieving the goal of eliminating any disposal of liquid waste resulting from various waste-generating processes [94,309]. It is transitioning from being perceived as a nuisance that is conveniently ignored to being recognized as a potential avenue for the reclamation of precious resources. This phenomenon is manifesting as a direct consequence of the confluence of these variables. Figure 14 illustrates a graphical depiction of the fundamental factors that drive ZLD, along with its numerous beneficial results.
The practice of reusing and recycling wastewater has the potential to not only mitigate the demand for freshwater resources, but also facilitate the reduction in waste and surplus resources. The textile-processing sector is known for its significant use of colorants, chemicals, and other additives. Consequently, it has significant promise for several approaches to intense chemical recovery and water recycling. The process of recycling has become essential to the industrial sector due to the imposition of constraints on accessible water sources and regulations managing wastewater. Prior to commencing the treatment procedure, all potential avenues for recovery and recycling have been thoroughly explored and utilized to their fullest extent.

4.7.1. Electrolyte Recovery from Reactive Dye Effluent

Electrolytes are frequently employed in the dyeing process within the industry to help with the exhaustion of reactive and direct dyes. In some cases, the concentration of electrolytes employed might reach approximately 90 g/L, leading to substantial increases in the amounts of TDS and chlorides. These substances are known to display resistance to biodegradation. The dye bath and the first rinse bath are responsible for the liberation of approximately 80% of these salts [310]. In the given circumstances, empirical evidence has demonstrated that thermal evaporation treatment is the sole feasible alternative. Thermal evaporation is widely acknowledged as a very efficient technique employed in the wastewater treatment for the textile industry to removing salt and dissolved solids from concentrated effluent, as well as extraction of water. In the context of salt recovery, it is worth noting that the salts obtained following the evaporation process may be effectively employed in dyeing processes (i.e., same color), so contributing to a reduction in the utilization of virgin resources [311,312,313]; however, there are main drawback of this system can provide the mixture of colorants and salts. The partitioning of dye molecules and salts, particularly monovalent salts, can occur using a NF membrane. This approach effectively mitigates the release of detrimental substances into the surrounding ecosystem while concurrently reducing resource use, hence resulting in financial benefits [311,314,315].

4.7.2. Alkali Recovery

Mercerization is a significant textile finishing technique performed on cotton fabric, including the application of a concentrated sodium hydroxide solution (ranging from 20% to 30%) [316,317]. This procedure aims to enhance several characteristics of the fabric, such as its luster, tactile qualities, and other pertinent features. The hygroscopicity of the material is enhanced, resulting in increased strength and improved dye affinity. The recovery of NaOH is crucial when employing a high concentration of alkali. The reduction in effluent load and the recovery of NaOH is a very straightforward operation, as it does not include the presence of additional chemicals, such as dyes. Membrane systems are considered very suitable for the treatment of effluents and the subsequent recovery of NaOH during the initial phases of fiber removal. Polymeric NF membranes allow for the passage of small molecules and ions without the need for mercerization. Nevertheless, several studies employ an ultrafiltration/microfiltration pre-treatment process due to the susceptibility of NF membranes to fouling, which might result in decreased penetration effectiveness [314,318,319,320,321,322]. The proposed approach involves a two-stage combination membrane design, incorporating a ceramic membrane in the first step and a polymeric membrane in the second phase. The recovery of sodium hydroxide from the procedure was found above 90%, recovered NaOH has the potential to be recycled into the process [318]. A separate investigation involved the utilization of microfiltration (MF) membranes with a pore size of 0.2 µm, followed by the implementation of UF membranes with a pore size of 100 kDa [319].

4.7.3. Dye Recovery

The effluent from dyebaths also results in the discharge of significant amounts of unabsorbed dyes, leading to a decrease in dissolved oxygen levels and the generation of high COD [71,323,324,325]. Moreover, the treatment of wastewater containing dyes poses a significant challenge in the field of wastewater management. Mariya et al. [326] aimed to assess the possibility for reusing colors obtained from denim and polyester dyebath effluents. In this investigation, tetraethylammonium bromide was employed as a draw solution. The findings of the study revealed that the forward osmosis membrane exhibited complete rejection of dyes, with a dye reconcentration ranging from 82% to 98%.

4.8. The Need for Technoeconomic Analysis

The discharge of effluent originating from the textile sector is resulting in substantial contamination due to the extensive use of colorants and hazardous chemicals, which are the primary contributors to water pollution and the escalating ecological risks. The escalating expenses associated with dye treatment facilities and the management of waste effluents are generating heightened public concern for environmental sustainability. The effective management and recovery of dyes and other chemical substances can contribute to the attainment of environmental sustainability and foster economic advancement. Furthermore, the substantial consumption of water underscores the imperative to engage in wastewater recycling. This has the potential to reduce the release of harmful compounds and create an atmosphere that is both safe and conducive to good health. The wastewaters under consideration consist of colors that include harmful and hazardous substances, such as pesticides, surfactants, and heavy metals. The cost of wastewater treatment often varies based on factors such as the specific treatment procedure employed, the spatial demands of the equipment, and the catalysts necessary for effectively removing contaminants. Furthermore, the operating expenses of the system encompass both personnel costs and maintenance expenditures. However, the prevailing price is contingent upon the region’s authoritative needs, the availability of feedstock, and the labor force [291,327]. However, it should be noted that relying only on a single treatment procedure may not be a viable approach for effectively degrading extremely contaminated dyes. This is mostly due to the development of intermediates during the treatment, which afterwards necessitate an additional treatment step. Consequently, this supplementary treatment incurs additional costs, thereby impacting the overall feasibility of the process. Hence, the implementation of a plan that combines several approaches will be crucial in addressing efficiency. However, these procedures include the integration of two treatment approaches and offer greater advantages in terms of pollutant degradation.

4.9. Life Cycle Assessment (LCA) in WWTPs

A life cycle assessment (LCA) is a comprehensive approach used to evaluate the total environmental impact of a product or process across its entire life cycle, from raw material extraction to disposal [328,329,330,331]. The primary goal of the LCA is to assess the whole range of environmental impacts resulting from the operation of the textile wastewater treatment facility. It is often essential to define the precise limits of the LCA investigation. The “Gate-to-Gate” methodology is a frequently used strategy in which different stages of the process are designated as gates. For example, we may classify the industry that emits pollutants as the first gate, and the release of cleaned effluent as the second gate. Furthermore, the analysis may also include the use of treated effluent in the textile sector. Moreover, it is crucial to establish a precise demarcation of the extent of the LCA studies carried out on wastewater treatment facilities (WWTPs). This encompasses delineating the parameters of the study, establishing the operational components, summarizing the LCA approach utilized, showcasing the life cycle inventory (LCI) information, and acknowledging the constraints, presumptions, and uncertainties linked to the examination.
LCA serves as a beneficial quantitative ecological assessment approach for examining various prospective functioning scenarios in the context of crucial water area planning. One of the advantages of LCA is in its ability to detect and quantify the impacts and influences of different process sequences, as well as assess the environmental effects of treatment technologies. LCA also facilitates the examination of pollution connections and aids in the achievement of effluent-free product creation. The LCA methodology was employed to conduct a comparative analysis of synthetic colors and natural colors, as well as synthetic finishes and biobased finishes, from an environmental perspective. Additionally, the study investigated the potential impacts of these materials on WWTPs. LCA offers decision makers and policy makers a consistent and transparent means to understand and interpret the ecological performance data of WWTPs in the textile sector. Hence, the utilization of LCA may facilitate the identification of research and development goals and provide guidance for enhancing innovation by mitigating challenges related to waste disposal and the discharge of hazardous chemicals. Figure 15 depicts the typical LCA framework on the WWTPs for the textile industry.

5. Conclusions

The textile industry is responsible for generating effluents that contain significant quantities of hazardous and persistent compounds, including dyes, chemicals, aromatic compounds, formaldehyde, flame retardants, and fluorocarbons. These substances have detrimental impacts on both the environment and human health, as well as on aquatic organisms. This review paper demonstrated the comprehensive literature analysis of several techniques for sustainable treatment of textile wastewater, with a specific focus on bioadsorbents, biological approaches, membrane technology, ion exchange, advanced oxidation processes, as well as physicochemical and biochemical processes. The tabulation of different bioadsorbents and their respective adsorbent capacities has been conducted, including a discussion on energy-efficient and cost-effective membranes and other treatment systems. Furthermore, this paper focused on strategies that result in zero liquid discharge (ZLD) and the subsequent retrieval of significant resources, such as dyes, alkalis, and electrolytes, which are produced in huge quantities during wastewater treatment plants. In recent times, there has been a growing interest in the recovery of dyes from textile waste. Consequently, dye recovery has emerged as a significant subject within the context of textile circular economy and industrial urban symbiosis. Furthermore, this review paper analyzes the management of sludge, conducts a technoeconomic analysis, and highlights their significance within the textile value chain. It is important to perform a life cycle analysis of wastewater treatment plants to ensure the efficient management of their reuse, reduction, and disposal processes.
In this situation, the idea of prevention being more effective than treatments remains valid. The use of biomaterials for functionalization, such as coloring (i.e., mass colorations) and surface changes, has been widely investigated in the literature [73,332,333,334,335]. By incorporating more biobased fibers, including natural and regenerated fibers, as well as other biobased and biodegradable polymers, it is possible to mitigate the overall pollution burden on wastewater treatment plants. Furthermore, the application of mass coloring or mass functionalization in the process of melt extrusion, namely in the case of polylactic acid, or in manmade fibers like viscose, infinna, spinnova, renewcell and ioncel has the potential to decrease the pollution burden associated with both synthetic and regenerate fibers [336,337,338,339].
The textile industry uses effluent treatment to separate water and other substances, with water recovery being easier than salt or residues. Highly polluting effluents, such as dye bath discharge, make up 10% of total effluent discharge, while 90% comes from low-polluting streams like wash water. Sustainable wastewater treatment can be applied to both streams but requires careful selection of highly polluting effluent streams. Rejecting reverse osmosis, nano- and ultrafiltration, the advanced oxidation process and granular activated carbon should be used to treat highly polluting effluent streams. Sustainable wastewater treatment produces less sludge, but it does not comply with norms and standards. In certain cases, using multiple methods could improve pollutant removal efficiency.
Sustainable wastewater treatments have several advantageous attributes, indicating their preference for minimizing chemical use in wastewater treatment plants, reducing energy consumption, and mitigating the carbon impact, among other benefits. The promotion of initiatives aimed at showcasing sustainable wastewater treatments at the industrial level is vital to mitigate the environmental impact produced by wastewater treatment plants.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su16020495/s1, Table S1. Advantages and disadvantages of biological methods for dye removal; Table S2. Dye degradation by various bacterial biomass; Table S3. Advantages of combination-based (hybrid) processes; Figure S1. Schematic representation of the forward osmosis process [47,59,66,340,341,342,343,344,345,346,347,348,349,350,351].

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Rayhan Sarker, M.; Mithun Ali, S.; Kumar Paul, S.; Haque Munim, Z. Measuring Sustainability Performance Using an Integrated Model. Measurement 2021, 184, 109931. [Google Scholar] [CrossRef]
  2. Szilagyi, A.; Mocan, M.; Verniquet, A.; Churican, A.; Rochat, D. Eco-Innovation, a Business Approach towards Sustainable Processes, Products and Services. Procedia Soc. Behav. Sci. 2018, 238, 475–484. [Google Scholar] [CrossRef]
  3. Crystal Gordon, Water: The Prime Essence Of Life. 2019. Available online: Https://Hubpages.Com/Education/Water-Is-the-Essence-of-Life-Itself (accessed on 20 September 2023).
  4. Ajmal, M.; Siddiq, M.; Aktas, N.; Sahiner, N. Magnetic Co–Fe Bimetallic Nanoparticle Containing Modifiable Microgels for the Removal of Heavy Metal Ions, Organic Dyes and Herbicides from Aqueous Media. RSC Adv. 2015, 5, 43873–43884. [Google Scholar] [CrossRef]
  5. Schwarzenbach, R.P.; Egli, T.; Hofstetter, T.B.; von Gunten, U.; Wehrli, B. Global Water Pollution and Human Health. Annu. Rev. Environ. Resour. 2010, 35, 109–136. [Google Scholar] [CrossRef]
  6. Gahlot, R.; Taki, K.; Kumar, M. Efficacy of Nanoclays as the Potential Adsorbent for Dyes and Metal Removal from the Wastewater: A Review. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100339. [Google Scholar] [CrossRef]
  7. Ahmad, A.; Mohd-Setapar, S.H.; Chuong, C.S.; Khatoon, A.; Wani, W.A.; Kumar, R.; Rafatullah, M. Recent Advances in New Generation Dye Removal Technologies: Novel Search for Approaches to Reprocess Wastewater. RSC Adv. 2015, 5, 30801–30818. [Google Scholar] [CrossRef]
  8. Owa, F.D. Water Pollution: Sources, Effects, Control and Management. Mediterr. J. Soc. Sci. 2013, 4, 65–68. [Google Scholar] [CrossRef]
  9. Ahmed, M.J.; Dhedan, S.K. Equilibrium Isotherms and Kinetics Modeling of Methylene Blue Adsorption on Agricultural Wastes-Based Activated Carbons. Fluid. Phase Equilib. 2012, 317, 9–14. [Google Scholar] [CrossRef]
  10. Mohamed, A. Hassaan; Ahmed El Nemr Health and Environmental Impacts of Dyes: Mini Review. Am. J. Environ. Sci. Eng. 2017, 1, 64–67. [Google Scholar]
  11. Ahmad, A.A.; Hameed, B.H.; Aziz, N. Adsorption of Direct Dye on Palm Ash: Kinetic and Equilibrium Modeling. J. Hazard. Mater. 2007, 141, 70–76. [Google Scholar] [CrossRef]
  12. Bhatia, D.; Sharma, N.R.; Singh, J.; Kanwar, R.S. Biological Methods for Textile Dye Removal from Wastewater: A Review. Crit. Rev. Environ. Sci. Technol. 2017, 47, 1836–1876. [Google Scholar] [CrossRef]
  13. Afroze, S.; Sen, T.K. A Review on Heavy Metal Ions and Dye Adsorption from Water by Agricultural Solid Waste Adsorbents. Water Air Soil. Pollut. 2018, 229, 225. [Google Scholar] [CrossRef]
  14. Shukla, S.R. Pollution Abatement and Waste Minimisation in Textile Dyeing. In Environmental Aspects of Textile Dyeing; Christie, R.M., Ed.; Woodhead Publishing Series in Textiles; Woodhead Publishing: Sawston, UK, 2007; pp. 116–148. ISBN 9781845691158. [Google Scholar]
  15. Reddy, N.; Chen, L.; Zhang, Y.; Yang, Y. Reducing Environmental Pollution of the Textile Industry Using Keratin as Alternative Sizing Agent to Poly(Vinyl Alcohol). J. Clean. Prod. 2014, 65, 561–567. [Google Scholar] [CrossRef]
  16. Hasanbeigi, A.; Price, L. A Technical Review of Emerging Technologies for Energy and Water Efficiency and Pollution Reduction in the Textile Industry. J. Clean. Prod. 2015, 95, 30–44. [Google Scholar] [CrossRef]
  17. Periyasamy, A.P.; Tehrani-Bagha, A. A Review on Microplastic Emission from Textile Materials and Its Reduction Techniques. Polym. Degrad. Stab. 2022, 199, 109901. [Google Scholar] [CrossRef]
  18. Klepacz-Smółka, A.; Sójka-Ledakowicz, J.; Paździor, K.; Ledakowicz, S. Application of Anoxic Fixed Film and Aerobic CSTR Bioreactor in Treatment of Nanofiltration Concentrate of Real Textile Wastewater. Chem. Pap. 2010, 64, 230–236. [Google Scholar] [CrossRef]
  19. Karunakaran, G.; Periyasamy, A.P.; Militký, J. Color and Design for Textiles. In Fibrous Structures and Their Impact on Textile Design; Springer Nature: Singapore, 2023; pp. 119–148. [Google Scholar]
  20. Militký, J.; Venkataraman, M.; Periyasamy, A.P. (Eds.) Fibrous Structures and Their Impact on Textile Design; Springer Nature: Singapore, 2023; ISBN 978-981-19-4826-8. [Google Scholar]
  21. Adeyemo, A.A.; Adeoye, I.O.; Bello, O.S. Adsorption of Dyes Using Different Types of Clay: A Review. Appl. Water Sci. 2017, 7, 543–568. [Google Scholar] [CrossRef]
  22. Chisvert, A.; Miralles, P.; Salvador, A. Hair Dyes in Cosmetics. In Analysis of Cosmetic Products; Elsevier: Amsterdam, The Netherlands, 2018; pp. 159–173. [Google Scholar]
  23. Singh, K. Rajani Srivastava an Overview of Textile Dyes and Their Removaltechniques: Indian Perspective. Polution Res. Pap. 2017, 36, 790–797. [Google Scholar]
  24. Robinson, T.; McMullan, G.; Marchant, R.; Nigam, P. Remediation of Dyes in Textile Effluent: A Critical Review on Current Treatment Technologies with a Proposed Alternative. Bioresour. Technol. 2001, 77, 247–255. [Google Scholar] [CrossRef]
  25. Mohd Adnan, M.A.; Muhd Julkapli, N.; Amir, M.N.I.; Maamor, A. Effect on Different TiO2 Photocatalyst Supports on Photodecolorization of Synthetic Dyes: A Review. Int. J. Environ. Sci. Technol. 2019, 16, 547–566. [Google Scholar] [CrossRef]
  26. Periyasamy, A.P.; Periyasami, S. Critical Review on Sustainability in Denim: A Step toward Sustainable Production and Consumption of Denim. ACS Omega 2023. [Google Scholar] [CrossRef] [PubMed]
  27. Pallares Jack Ferre, J.A. A Simple Model to Predict Mass Transfer Rates and Kinetics of Biochemical and Biomedical Michaelis-Menten Surface Reactions. Int. J. Heat. Mass. Transf. 2015, 80, 192–198. [Google Scholar] [CrossRef]
  28. Periyasamy, A.P. Evaluation of Microfiber Release from Jeans: The Impact of Different Washing Conditions. Environ. Sci. Pollut. Res. 2021, 28, 58570–58582. [Google Scholar] [CrossRef] [PubMed]
  29. Sachidhanandham, A.; Periyasamy, A.P. Environmentally Friendly Wastewater Treatment Methods for the Textile Industry. In Handbook of Nanomaterials and Nanocomposites for Energy and Environmental Applications; Springer International Publishing: Cham, Switzerland, 2021; pp. 2269–2307. [Google Scholar]
  30. Samarghandi, M.R.; Hadi, M.; Moayedi, S.; Barjesteh, A.F. Hadi Wo-Parameter Isotherms of Methyl Orange Sorption by Pinecone Derived Activated Carbon. Iran. J. Environ. Health Sci. Eng. 2009, 6, 285–294. [Google Scholar]
  31. Raval, N.P.; Shah, P.U.; Shah, N.K. Malachite Green “a Cationic Dye” and Its Removal from Aqueous Solution by Adsorption. Appl. Water Sci. 2017, 7, 3407–3445. [Google Scholar] [CrossRef]
  32. Roy, S.; Darabdhara, J.; Ahmaruzzaman, M. Recent Advances of Copper- BTC Metal-Organic Frameworks for Efficient Degradation of Organic Dye-Polluted Wastewater: Synthesis, Mechanistic Insights and Future Outlook. J. Hazard. Mater. Lett. 2024, 5, 100094. [Google Scholar] [CrossRef]
  33. Ahsan, A.; Jamil, F.; Rashad, M.A.; Hussain, M.; Inayat, A.; Akhter, P.; Al-Muhtaseb, A.H.; Lin, K.Y.A.; Park, Y.K. Wastewater from the Textile Industry: Review of the Technologies for Wastewater Treatment and Reuse. Korean J. Chem. Eng. 2023, 40, 2060–2081. [Google Scholar] [CrossRef]
  34. Deng, Y.; Zhao, R. Advanced Oxidation Processes (AOPs) in Wastewater Treatment. Curr. Pollut. Rep. 2015, 1, 167–176. [Google Scholar] [CrossRef]
  35. Zaharia, C.; Suteu, D.; Muresan, A.; Muresan, R.; Popescu, A. Textile Wastewater Treatment by Homogeneous Oxidation with Hydrogen Peroxide. Environ. Eng. Manag. J. 2009, 8, 1359–1369. [Google Scholar] [CrossRef]
  36. Hutagalung, S.S.; Rafryanto, A.F.; Sun, W.; Juliasih, N.; Aditia, S.; Jiang, J.; Dipojono, H.K.; Suhardi, S.H.; Rochman, N.T.; Kurniadi, D. Combination of Ozone-Based Advanced Oxidation Process and Nanobubbles Generation toward Textile Wastewater Recovery. Front. Environ. Sci. 2023, 11, 1154739. [Google Scholar] [CrossRef]
  37. Peralta-Zamora, P.; Kunz, A.; De Moraes, S.G.; Pelegrini, R.; De Campos Moleiro, P.; Reyes, J.; Duran, N. Degradation of Reactive Dyes, I. A Comparative Study of Ozonation, Enzymic and Photochemical Processes. Chemosphere 1998, 38, 835–852. [Google Scholar] [CrossRef] [PubMed]
  38. Wang, J.; Chen, H. Catalytic Ozonation for Water and Wastewater Treatment: Recent Advances and Perspective. Sci. Total Environ. 2020, 704, 135249. [Google Scholar] [CrossRef] [PubMed]
  39. Paździor, K.; Wrębiak, J.; Klepacz-Smółka, A.; Gmurek, M.; Bilińska, L.; Kos, L.; Sójka-Ledakowicz, J.; Ledakowicz, S. Influence of Ozonation and Biodegradation on Toxicity of Industrial Textile Wastewater. J. Environ. Manag. 2017, 195, 166–173. [Google Scholar] [CrossRef]
  40. Selcuk, H. Decolorization and Detoxification of Textile Wastewater by Ozonation and Coagulation Processes. Dye. Pigment. 2005, 64, 217–222. [Google Scholar] [CrossRef]
  41. Malik, S.N.; Ghosh, P.C.; Vaidya, A.N.; Mudliar, S.N. Hybrid Ozonation Process for Industrial Wastewater Treatment: Principles and Applications: A Review. J. Water Process Eng. 2020, 35, 101193. [Google Scholar] [CrossRef]
  42. Qu, R.; Xu, B.; Meng, L.; Wang, L.; Wang, Z. Ozonation of Indigo Enhanced by Carboxylated Carbon Nanotubes: Performance Optimization, Degradation Products, Reaction Mechanism and Toxicity Evaluation. Water Res. 2015, 68, 316–327. [Google Scholar] [CrossRef] [PubMed]
  43. Kim, Y.W.; Kim, J.H.; Moon, D.H.; Shin, H.J. Adsorption and Precipitation of Anionic Dye Reactive Red 120 from Aqueous Solution by Aminopropyl Functionalized Magnesium Phyllosilicate. Korean J. Chem. Eng. 2019, 36, 101–108. [Google Scholar] [CrossRef]
  44. Feuzer-Matos, A.J.; Testolin, R.C.; Pimentel-Almeida, W.; Radetski-Silva, R.; Deomar-Simões, M.J.; Poyer-Radetski, L.; Ariente-Neto, R.; Batista-Barwinski, M.J.; Somensi, C.A.; Radetski, C.M. Treatment of Wastewater Containing New and Non-Biodegradable Textile Dyes: Efficacy of Combined Advanced Oxidation and Adsorption Processes. Water Air Soil. Pollut. 2022, 233, 273. [Google Scholar] [CrossRef]
  45. Villaseñor-Basulto, D.L.; Kadier, A.; Singh, R.; Navarro-Mendoza, R.; Bandala, E.; Peralta-Hernández, J.M. Post-Tanning Wastewater Treatment Using Electrocoagulation: Optimization, Kinetics, and Settlement Analysis. Process Saf. Environ. Prot. 2022, 165, 872–886. [Google Scholar] [CrossRef]
  46. Karatas, O.; Khataee, A.; Kobya, M.; Yoon, Y. Electrochemical Oxidation of Perfluorooctanesulfonate (PFOS) from Simulated Soil Leachate and Landfill Leachate Concentrate. J. Water Process Eng. 2023, 56, 104292. [Google Scholar] [CrossRef]
  47. Zhang, J.; Chen, S.; Zhang, Y.; Quan, X.; Zhao, H.; Zhang, Y. Reduction of Acute Toxicity and Genotoxicity of Dye Effluent Using Fenton-Coagulation Process. J. Hazard. Mater. 2014, 274, 198–204. [Google Scholar] [CrossRef] [PubMed]
  48. Pervez, M.N.; Fu, D.; Wang, X.; Bao, Q.; Yu, T.; Naddeo, V.; Tian, H.; Cao, C.; Zhao, Y. A Bifunctional-FeOOH@GCA Nanocomposite for Enhanced Adsorption of Arsenic and Photo Fenton-like Catalytic Conversion of As(III). Environ. Technol. Innov. 2021, 22, 101437. [Google Scholar] [CrossRef]
  49. Patil, A.D.; Raut, P.D. Treatment of Textile Wastewater by Fenton’s Process as a Advanced Oxidation Process. IOSR J. Environ. Sci. Toxicol. Food Technol. 2014, 8, 29–32. [Google Scholar] [CrossRef]
  50. Lucas, M.S.; Dias, A.A.; Sampaio, A.; Amaral, C.; Peres, J.A. Degradation of a Textile Reactive Azo Dye by a Combined Chemical–Biological Process: Fenton’s Reagent-Yeast. Water Res. 2007, 41, 1103–1109. [Google Scholar] [CrossRef] [PubMed]
  51. Ramos, M.D.N.; Santana, C.S.; Velloso, C.C.V.; da Silva, A.H.M.; Magalhães, F.; Aguiar, A. A Review on the Treatment of Textile Industry Effluents through Fenton Processes. Process Saf. Environ. Prot. 2021, 155, 366–386. [Google Scholar] [CrossRef]
  52. Lin, S.H.; Chen, M.L. Purification of Textile Wastewater Effluents by a Combined Fenton Process and Ion Exchange. Desalination 1997, 109, 121–130. [Google Scholar] [CrossRef]
  53. Särkkä, H.; Bhatnagar, A.; Sillanpää, M. Recent Developments of Electro-Oxidation in Water Treatment—A Review. J. Electroanal. Chem. 2015, 754, 46–56. [Google Scholar] [CrossRef]
  54. Verma, A.K.; Dash, R.R.; Bhunia, P. A Review on Chemical Coagulation/Flocculation Technologies for Removal of Colour from Textile Wastewaters. J. Environ. Manag. 2012, 93, 154–168. [Google Scholar] [CrossRef]
  55. Papic, S. Removal of Some Reactive Dyes from Synthetic Wastewater by Combined Al(III) Coagulation/Carbon Adsorption Process. Dye. Pigment. 2004, 62, 291–298. [Google Scholar] [CrossRef]
  56. Hong, J.; Otaki, M. Effects of Photocatalysis on Biological Decolorization Reactor and Biological Activity of Isolated Photosynthetic Bacteria. J. Biosci. Bioeng. 2003, 96, 298–303. [Google Scholar] [CrossRef]
  57. Fu, Z.; Zhang, Y.; Wang, X. Textiles Wastewater Treatment Using Anoxic Filter Bed and Biological Wriggle Bed-Ozone Biological Aerated Filter. Bioresour. Technol. 2011, 102, 3748–3753. [Google Scholar] [CrossRef]
  58. Paz, A.; Carballo, J.; Pérez, M.J.; Domínguez, J.M. Biological Treatment of Model Dyes and Textile Wastewaters. Chemosphere 2017, 181, 168–177. [Google Scholar] [CrossRef]
  59. Martin, I.; Pidou, M.; Soares, A.; Judd, S.; Jefferson, B. Modelling the Energy Demands of Aerobic and Anaerobic Membrane Bioreactors for Wastewater Treatment. Environ. Technol. 2011, 32, 921–932. [Google Scholar] [CrossRef]
  60. Lourenço, N.D.; Franca, R.D.G.; Moreira, M.A.; Gil, F.N.; Viegas, C.A.; Pinheiro, H.M. Comparing Aerobic Granular Sludge and Flocculent Sequencing Batch Reactor Technologies for Textile Wastewater Treatment. Biochem. Eng. J. 2015, 104, 57–63. [Google Scholar] [CrossRef]
  61. Tomei, M.C.; Mosca Angelucci, D.; Daugulis, A.J. Sequential Anaerobic-Aerobic Decolourization of a Real Textile Wastewater in a Two-Phase Partitioning Bioreactor. Sci. Total Environ. 2016, 573, 585–593. [Google Scholar] [CrossRef]
  62. Cinperi, N.C.; Ozturk, E.; Yigit, N.O.; Kitis, M. Treatment of Woolen Textile Wastewater Using Membrane Bioreactor, Nanofiltration and Reverse Osmosis for Reuse in Production Processes. J. Clean. Prod. 2019, 223, 837–848. [Google Scholar] [CrossRef]
  63. Mehmood, C.T.; Qiu, H.; Chen, L.; Achmon, Y.; Zhong, Z. Ceramic Membrane Reactor Integrated with UV/O3/Catalyst Beads for Treating Real Textile Wastewater: Enhanced Effluent Quality, Fouling Control and Molecular Transformations of DOM. J. Environ. Chem. Eng. 2023, 11, 110832. [Google Scholar] [CrossRef]
  64. Guclu, S.; Kizildag, N.; Dizman, B.; Unal, S. Solvent-Based Recovery of High Purity Polysulfone and Polyester from End-of-Life Reverse Osmosis Membranes. Sustain. Mater. Technol. 2022, 31, e00358. [Google Scholar] [CrossRef]
  65. Tang, A.Y.L.; Lee, C.H.; Wang, Y.M.; Kan, C.W. Dyeing Cotton with Reactive Dyes: A Comparison between Conventional Water-Based and Solvent-Assisted PEG-Based Reverse Micellar Dyeing Systems. Cellulose 2019, 26, 1399–1408. [Google Scholar] [CrossRef]
  66. Abebe, B.; Murthy, H.C.A.; Amare, E. Summary on Adsorption and Photocatalysis for Pollutant Remediation: Mini Review. J. Encapsulation Adsorpt. Sci. 2018, 08, 225–255. [Google Scholar] [CrossRef]
  67. Hassan, M.M.; Carr, C.M. A Critical Review on Recent Advancements of the Removal of Reactive Dyes from Dyehouse Effluent by Ion-Exchange Adsorbents. Chemosphere 2018, 209, 201–219. [Google Scholar] [CrossRef]
  68. Raghu, S.; Ahmed Basha, C. Chemical or Electrochemical Techniques, Followed by Ion Exchange, for Recycle of Textile Dye Wastewater. J. Hazard. Mater. 2007, 149, 324–330. [Google Scholar] [CrossRef]
  69. Bozzi, A.; Caronna, T.; Fontana, F.; Marcandalli, B.; Selli, E. Photodecomposition of Substituted 4-Diethylaminoazobenzenes under Visible Light Irradiation in Different Solvents. J. Photochem. Photobiol. A Chem. 2002, 152, 193–197. [Google Scholar] [CrossRef]
  70. Periyasamy, A.P. Microfiber Emissions from Functionalized Textiles: Potential Threat for Human Health and Environmental Risks. Toxics 2023, 11, 406. [Google Scholar] [CrossRef]
  71. Periyasamy, A.P.; Militky, J. Sustainability in Textile Dyeing: Recent Developments. In Sustainability in the Textile and Apparel Industries; Springer Nature: Cham, Switzerland, 2020; pp. 37–79. [Google Scholar]
  72. Periyasamy, A.P.; Rwahwire, S.; Zhao, Y. Environmental Friendly Textile Processing. In Handbook of Ecomaterials; Martínez, L.M.T., Kharissova, O.V., Kharisov, B.I., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 1–38. [Google Scholar]
  73. Periyasamy, A.P.; Ramamoorthy, S.K.; Lavate, S.S. Eco-Friendly Denim Processing. In Handbook of Ecomaterials; Springer: Cham, Switzerland, 2019; Volume 3, pp. 1559–1579. ISBN 9783319682556. [Google Scholar]
  74. Raman, C.D.; Mkandawire, M. GIS Based Spatial and Temporal Investigation of Groundwater and Soil Quality along Noyyal River, Tiruppur, India. J. Geol. Soc. India 2021, 97, 1097–1104. [Google Scholar] [CrossRef]
  75. Ram, S.; Dineshkumar, R.; Pancha, I.; Mishra, S. Prospects and Potential Role of Biological Treatment of Textile Effluent to Restore Water Reservoir. In Cost-efficient Wastewater Treatment Technologies; The Handbook of Environmental Chemistry; Springer: Cham, Switzerland, 2023; Volume 118, pp. 199–212. [Google Scholar]
  76. Indian Textile and Apparel Industry Expected to Reach US $344.1 Billion by 2027. Available online: https://in.apparelresources.com/business-news/trade/indian-textile-apparel-industry-expected-reach-us-344-1-billion-2027/ (accessed on 26 August 2022).
  77. Periyasamy, A.P.; Ramamoorthy, S.K.; Rwawiire, S.; Zhao, Y. Sustainable Wastewater Treatment Methods for Textile Industry. In Sustainable Innovations in Apparel Production; Springer: Singapore, 2018; pp. 21–87. [Google Scholar]
  78. Mehra, S.; Singh, M.; Chadha, P. Adverse Impact of Textile Dyes on the Aquatic Environment as Well as on Human Beings. Toxicol. Int. 2021, 28, 165–176. [Google Scholar] [CrossRef]
  79. Ali, A.E.; Chowdhury, Z.Z.; Devnath, R.; Ahmed, M.M.; Rahman, M.M.; Khalid, K.; Wahab, Y.A.; Badruddin, I.A.; Kamangar, S.; Hussien, M.; et al. Removal of Azo Dyes from Aqueous Effluent Using Bio-Based Activated Carbons: Toxicity Aspects and Environmental Impact. Separations 2023, 10, 506. [Google Scholar] [CrossRef]
  80. Periyasamy, A.P.; Militky, J. Denim Processing and Health Hazards. In Sustainability in Denim; Elsevier: Amsterdam, The Netherlands, 2017; pp. 161–196. [Google Scholar]
  81. Drumond Chequer, F.M.; de Oliveira, G.A.R.; Anastacio Ferraz, E.R.; Carvalho, J.; Boldrin Zanoni, M.V.; de Oliveir, D.P. Textile Dyes: Dyeing Process and Environmental Impact. In Eco-Friendly Textile Dyeing and Finishing; InTech: Isai, Romania, 2013. [Google Scholar]
  82. Blackburn, R. Sustainability Challenges of the Textiles, Dyeing and Finsihing Industries: Opportunities for Innovation. ACS Webinar 2017, 1–33. [Google Scholar]
  83. Khan, R.; Patel, V.; Khan, Z. Bioremediation of Dyes from Textile and Dye Manufacturing Industry Effluent; Elsevier Inc.: Amsterdam, The Netherlands, 2019; ISBN 9780128180952. [Google Scholar]
  84. Ning, X.A.; Liang, J.Y.; Li, R.J.; Hong, Z.; Wang, Y.J.; Chang, K.L.; Zhang, Y.P.; Yang, Z.Y. Aromatic Amine Contents, Component Distributions and Risk Assessment in Sludge from 10 Textile-Dyeing Plants. Chemosphere 2015, 134, 367–373. [Google Scholar] [CrossRef]
  85. Liang, J.; Ning, X.A.; Kong, M.; Liu, D.; Wang, G.; Cai, H.; Sun, J.; Zhang, Y.; Lu, X.; Yuan, Y. Elimination and Ecotoxicity Evaluation of Phthalic Acid Esters from Textile-Dyeing Wastewater. Environ. Pollut. 2017, 231, 115–122. [Google Scholar] [CrossRef]
  86. Soni, V.; Keswani, K.; Bhatt, U.; Kumar, D.; Singh, H. In Vitro Propagation and Analysis of Mixotrophic Potential to Improve Survival Rate of Dolichandra Unguis-Cati under Ex Vitro Conditions. Heliyon 2021, 7, e06101. [Google Scholar] [CrossRef]
  87. Li, F.; Zhao, K.; Ng, T.S.; Dai, Y.; Wang, C.H. Sustainable Production of Bio-Oil and Carbonaceous Materials from Biowaste Co-Pyrolysis. Chem. Eng. J. 2022, 427, 131821. [Google Scholar] [CrossRef]
  88. Khan, A.A.; Gul, J.; Naqvi, S.R.; Ali, I.; Farooq, W.; Liaqat, R.; AlMohamadi, H.; Štěpanec, L.; Juchelková, D. Recent Progress in Microalgae-Derived Biochar for the Treatment of Textile Industry Wastewater. Chemosphere 2022, 306, 135565. [Google Scholar] [CrossRef]
  89. Mishra, V.; Mudgal, N.; Rawat, D.; Poria, P.; Mukherjee, P.; Sharma, U.; Kumria, P.; Pani, B.; Singh, M.; Yadav, A.; et al. Integrating Microalgae into Textile Wastewater Treatment Processes: Advancements and Opportunities. J. Water Process Eng. 2023, 55, 104128. [Google Scholar] [CrossRef]
  90. Guettaï, N.; Ait Amar, H. Photocatalytic Oxidation of Methyl Orange in Presence of Titanium Dioxide in Aqueous Suspension. Part I: Parametric Study. Desalination 2005, 185, 427–437. [Google Scholar] [CrossRef]
  91. Guettaï, N.; Ait Amar, H. Photocatalytic Oxidation of Methyl Orange in Presence of Titanium Dioxide in Aqueous Suspension. Part II: Kinetics Study. Desalination 2005, 185, 439–448. [Google Scholar] [CrossRef]
  92. Cheng, S.; Jessica; Yoshikawa, K.; Cross, J.S. Influence of Synthetic and Natural Microfibers on the Growth, Substance Exchange, Energy Accumulation, and Oxidative Stress of Field-Collected Microalgae Compared with Microplastic Fragment. Sci. Total Environ. 2023, 908, 167936. [Google Scholar] [CrossRef]
  93. Luongo, G. Chemicals in Textiles: A Potential Source for Human Exposure and Environmental Pollution. Ph.D. Thesis, Stockholm University, Faculty of Science, Gothenburg, Sweden, 2015. [Google Scholar]
  94. Kopperi, H.; Hemalatha, M.; Ravi Kiran, B.; Santhosh, J.; Venkata Mohan, S. Sustainable Consideration for Traditional Textile Handloom Cluster/Village in Pollution Abatement—A Case Study. Environ. Pollut. 2023, 324, 121320. [Google Scholar] [CrossRef]
  95. Sela, S.K.; Hossain, A.K.M.N.-U.-; Hussain, S.Z.; Hasan, N. Utilization of Prawn to Reduce the Value of BOD and COD of Textile Wastewater. Clean. Eng. Technol. 2020, 1, 100021. [Google Scholar] [CrossRef]
  96. Chen, Y.-G.; He, X.-L.-S.; Huang, J.-H.; Luo, R.; Ge, H.-Z.; Wołowicz, A.; Wawrzkiewicz, M.; Gładysz-Płaska, A.; Li, B.; Yu, Q.-X.; et al. Impacts of Heavy Metals and Medicinal Crops on Ecological Systems, Environmental Pollution, Cultivation, and Production Processes in China. Ecotoxicol. Environ. Saf. 2021, 219, 112336. [Google Scholar] [CrossRef] [PubMed]
  97. Rendón-Castrillón, L.; Ramírez-Carmona, M.; Ocampo-López, C.; González-López, F.; Cuartas-Uribe, B.; Mendoza-Roca, J.A. Treatment of Water from the Textile Industry Contaminated with Indigo Dye: A Hybrid Approach Combining Bioremediation and Nanofiltration for Sustainable Reuse. Case Stud. Chem. Environ. Eng. 2023, 8, 100498. [Google Scholar] [CrossRef]
  98. Jorge, A.M.S.; Athira, K.K.; Alves, M.B.; Gardas, R.L.; Pereira, J.F.B. Textile Dyes Effluents: A Current Scenario and the Use of Aqueous Biphasic Systems for the Recovery of Dyes. J. Water Process Eng. 2023, 55, 104125. [Google Scholar] [CrossRef]
  99. Mu, B.; Yu, X.; Shao, Y.; McBride, L.; Hidalgo, H.; Yang, Y. Complete Recycling of Polymers and Dyes from Polyester/Cotton Blended Textiles via Cost-Effective and Destruction-Minimized Dissolution, Swelling, Precipitation, and Separation. Resour. Conserv. Recycl. 2023, 199, 107275. [Google Scholar] [CrossRef]
  100. Tamil Selvan, S. Eco-Technological Approaches for Textile Dye Effluent Treatment and Carbon Dioxide (CO2) Capturing Using Green Microalga Chlorella vulgaris. Mater. Today Proc. 2023, in press. [Google Scholar] [CrossRef]
  101. Khan, A.U.; Zahoor, M.; Rehman, M.U.; Ikram, M.; Zhu, D.; Umar, M.N.; Ullah, R.; Ali, E.A. Bioremediation of Azo Dye Brown 703 by Pseudomonas Aeruginosa: An Effective Treatment Technique for Dye-Polluted Wastewater. Microbiol. Res. 2023, 14, 1049–1066. [Google Scholar] [CrossRef]
  102. Agarwal, M.; Maheshwari, K.; Solanki, Y.S. Investigation of Dye Effluent Treatment Using Unmodified and Modified Biobased Sorbent and Its Process Economics. J Hazard Toxic Radioact Waste 2022, 26. [Google Scholar] [CrossRef]
  103. Periyasamy, A.P. Environmentally Friendly Approach to the Reduction of Microplastics during Domestic Washing: Prospects for Machine Vision in Microplastics Reduction. Toxics 2023, 11, 575. [Google Scholar] [CrossRef]
  104. Alghamdi, M.A.; Ayed, L.; Aljarad, M.R.; Altayeb, H.N.; Abbes, S.; Chaieb, K. Whole Genome Sequencing Analysis and Box-Behnken Design for the Optimization of the Decolourization of Mixture Textile Dyes by Halotolerant Microbial Consortium. Microbiol. Res. 2023, 276, 127481. [Google Scholar] [CrossRef]
  105. Mu, B.; Shao, Y.; McBride, L.; Hidalgo, H.; Yang, Y. Rapid Fiber-to-Fiber Recycling of Poly (Ethylene Terephthalate) and Its Dye from Waste Textiles without Damaging Their Chemical Structures. Resour. Conserv. Recycl. 2023, 197, 107102. [Google Scholar] [CrossRef]
  106. Chequer, F.M.D.; Angeli, J.P.F.; Ferraz, E.R.A.; Tsuboy, M.S.; Marcarini, J.C.; Mantovani, M.S.; de Oliveira, D.P. The Azo Dyes Disperse Red 1 and Disperse Orange 1 Increase the Micronuclei Frequencies in Human Lymphocytes and in HepG2 Cells. Mutat. Res./Genet. Toxicol. Environ. Mutagen. 2009, 676, 83–86. [Google Scholar] [CrossRef]
  107. Pattnaik, P.; Dangayach, G.S.; Bhardwaj, A.K. A Review on the Sustainability of Textile Industries Wastewater with and without Treatment Methodologies. Rev. Environ. Health 2018, 33, 163–203. [Google Scholar] [CrossRef] [PubMed]
  108. Bharagava, R.N.; Chowdhary, P. Emerging and Eco-Friendly Approaches for Waste Management. In Emerging and Eco-Friendly Approaches for Waste Management; Springer: Singapore, 2018; pp. 1–435. [Google Scholar] [CrossRef]
  109. Motejadded Emrooz, H.B.; Maleki, M.; Rashidi, A.; Shokouhimehr, M. Adsorption Mechanism of a Cationic Dye on a Biomass-Derived Micro- and Mesoporous Carbon: Structural, Kinetic, and Equilibrium Insight. Biomass Convers. Biorefin 2021, 11, 943–954. [Google Scholar] [CrossRef]
  110. Temesgen, F.; Gabbiye, N.; Sahu, O. Biosorption of Reactive Red Dye (RRD) on Activated Surface of Banana and Orange Peels: Economical Alternative for Textile Effluent. Surf. Interfaces 2018, 12, 151–159. [Google Scholar] [CrossRef]
  111. Oyekanmi, A.A.; Ahmad, A.; Hossain, K.; Rafatullah, M. Adsorption of Rhodamine B Dye from Aqueous Solution onto Acid Treated Banana Peel: Response Surface Methodology, Kinetics and Isotherm Studies. PLoS ONE 2019, 14, e0216878. [Google Scholar] [CrossRef] [PubMed]
  112. Harini, K.; Ramya, K.; Sukumar, M. Extraction of Nano Cellulose Fibers from the Banana Peel and Bract for Production of Acetyl and Lauroyl Cellulose. Carbohydr. Polym. 2018, 201, 329–339. [Google Scholar] [CrossRef]
  113. Munagapati, V.S.; Yarramuthi, V.; Kim, Y.; Lee, K.M.; Kim, D.-S. Removal of Anionic Dyes (Reactive Black 5 and Congo Red) from Aqueous Solutions Using Banana Peel Powder as an Adsorbent. Ecotoxicol. Environ. Saf. 2018, 148, 601–607. [Google Scholar] [CrossRef]
  114. Wekoye, J.N.; Wanyonyi, W.C.; Wangila, P.T.; Tonui, M.K. Kinetic and Equilibrium Studies of Congo Red Dye Adsorption on Cabbage Waste Powder. Environ. Chem. Ecotoxicol. 2020, 2, 24–31. [Google Scholar] [CrossRef]
  115. Mashkoor, F.; Nasar, A. Facile Synthesis of Polypyrrole Decorated Chitosan-Based Magsorbent: Characterizations, Performance, and Applications in Removing Cationic and Anionic Dyes from Aqueous Medium. Int. J. Biol. Macromol. 2020, 161, 88–100. [Google Scholar] [CrossRef]
  116. Zhao, F.; Yang, Z.; Wei, Z.; Spinney, R.; Sillanpää, M.; Tang, J.; Tam, M.; Xiao, R. Polyethylenimine-Modified Chitosan Materials for the Recovery of La(III) from Leachates of Bauxite Residue. Chem. Eng. J. 2020, 388, 124307. [Google Scholar] [CrossRef]
  117. Subedi, N.; Lähde, A.; Abu-Danso, E.; Iqbal, J.; Bhatnagar, A. A Comparative Study of Magnetic Chitosan (Chi@Fe3O4) and Graphene Oxide Modified Magnetic Chitosan (Chi@Fe3O4GO) Nanocomposites for Efficient Removal of Cr(VI) from Water. Int. J. Biol. Macromol. 2019, 137, 948–959. [Google Scholar] [CrossRef]
  118. Crini, G.; Torri, G.; Lichtfouse, E.; Kyzas, G.Z.; Wilson, L.D.; Morin-Crini, N. Dye Removal by Biosorption Using Cross-Linked Chitosan-Based Hydrogels. Environ. Chem. Lett. 2019, 17, 1645–1666. [Google Scholar] [CrossRef]
  119. Shakoor, S.; Nasar, A. Removal of Methylene Blue Dye from Artificially Contaminated Water Using Citrus Limetta Peel Waste as a Very Low Cost Adsorbent. J. Taiwan. Inst. Chem. Eng. 2016, 66, 154–163. [Google Scholar] [CrossRef]
  120. Mahato, N.; Sharma, K.; Sinha, M.; Baral, E.R.; Koteswararao, R.; Dhyani, A.; Hwan Cho, M.; Cho, S. Bio-Sorbents, Industrially Important Chemicals and Novel Materials from Citrus Processing Waste as a Sustainable and Renewable Bioresource: A Review; Cairo University: Giza, Egypt, 2020; Volume 23, ISBN 8201027988476. [Google Scholar]
  121. Tomar, V.; Prasad, S.; Kumar, D. Adsorptive Removal of Fluoride from Aqueous Media Using Citrus limonum (Lemon) Leaf. Microchem. J. 2014, 112, 97–103. [Google Scholar] [CrossRef]
  122. Mohanraj, J.; Durgalakshmi, D.; Balakumar, S.; Aruna, P.; Ganesan, S.; Rajendran, S.; Naushad, M. Low Cost and Quick Time Absorption of Organic Dye Pollutants under Ambient Condition Using Partially Exfoliated Graphite. J. Water Process Eng. 2020, 34, 101078. [Google Scholar] [CrossRef]
  123. Munagapati, V.S.; Kim, D.-S. Adsorption of Anionic Azo Dye Congo Red from Aqueous Solution by Cationic Modified Orange Peel Powder. J. Mol. Liq. 2016, 220, 540–548. [Google Scholar] [CrossRef]
  124. Ahmed, M.; Mashkoor, F.; Nasar, A. Development, Characterization, and Utilization of Magnetized Orange Peel Waste as a Novel Adsorbent for the Confiscation of Crystal Violet Dye from Aqueous Solution. Groundw. Sustain. Dev. 2020, 10, 100322. [Google Scholar] [CrossRef]
  125. Tahir, N.; Bhatti, H.N.; Iqbal, M.; Noreen, S. Biopolymers Composites with Peanut Hull Waste Biomass and Application for Crystal Violet Adsorption. Int. J. Biol. Macromol. 2017, 94, 210–220. [Google Scholar] [CrossRef]
  126. De Azevedo, A.C.N.; Vaz, M.G.; Gomes, R.F.; Pereira, A.G.B.; Fajardo, A.R.; Rodrigues, F.H.A. Starch/Rice Husk Ash Based Superabsorbent Composite: High Methylene Blue Removal Efficiency. Iran. Polym. J. 2017, 26, 93–105. [Google Scholar] [CrossRef]
  127. Ashrafi, S.D.; Kamani, H.; Mahvi, A.H. The Optimization Study of Direct Red 81 and Methylene Blue Adsorption on NaOH-Modified Rice Husk. Desalination Water Treat. 2016, 57, 738–746. [Google Scholar] [CrossRef]
  128. Shakoor, S.; Nasar, A. Adsorptive Decontamination of Synthetic Wastewater Containing Crystal Violet Dye by Employing Terminalia Arjuna Sawdust Waste. Groundw. Sustain. Dev. 2018, 7, 30–38. [Google Scholar] [CrossRef]
  129. Tahir, H.; Sultan, M.; Akhtar, N.; Hameed, U.; Abid, T. Application of Natural and Modified Sugar Cane Bagasse for the Removal of Dye from Aqueous Solution. J. Saudi Chem. Soc. 2016, 20, S115–S121. [Google Scholar] [CrossRef]
  130. Aljeboree, A.M.; Alshirifi, A.N.; Alkaim, A.F. Kinetics and Equilibrium Study for the Adsorption of Textile Dyes on Coconut Shell Activated Carbon. Arab. J. Chem. 2017, 10, S3381–S3393. [Google Scholar] [CrossRef]
  131. Gupta, S.A.; Vishesh, Y.; Sarvshrestha, N.; Bhardwaj, A.S.; Kumar, P.A.; Topare, N.S.; Raut-Jadhav, S.; Bokil, S.A.; Khan, A. Adsorption Isotherm Studies of Methylene Blue Using Activated Carbon of Waste Fruit Peel as an Adsorbent. Mater. Today Proc. 2022, 57, 1500–1508. [Google Scholar] [CrossRef]
  132. Castillo, M.; de Guzman, M.J.K.; Aberilla, J.M. Environmental Sustainability Assessment of Banana Waste Utilization into Food Packaging and Liquid Fertilizer. Sustain. Prod. Consum. 2023, 37, 356–368. [Google Scholar] [CrossRef]
  133. Farias, K.C.S.; Guimarães, R.C.A.; Oliveira, K.R.W.; Nazário, C.E.D.; Ferencz, J.A.P.; Wender, H. Banana Peel Powder Biosorbent for Removal of Hazardous Organic Pollutants from Wastewater. Toxics 2023, 11, 664. [Google Scholar] [CrossRef]
  134. Khaleque, A.; Roy, D.K. Removing Reactive Dyes from Textile Effluent Using Banana Fibre. Int. J. Basic Appl. Sci. 2016, 16, 14–20. [Google Scholar]
  135. Das, E.; Rabha, S.; Talukdar, K.; Goswami, M.; Devi, A. Propensity of a Low-Cost Adsorbent Derived from Agricultural Wastes to Interact with Cationic Dyes in Aqueous Solutions. Environ. Monit. Assess. 2023, 195, 1044. [Google Scholar] [CrossRef]
  136. Akbar, N.A.; Sabri, S.; Abu Bakar, A.A.; Azizan, N.S. Removal of Colour Using Banana Stem Adsorbent in Textile Wastewater. J. Phys. Conf. Ser. 2019, 1349, 12091. [Google Scholar] [CrossRef]
  137. Karunakaran, G.; Periyasamy, A.P.; Tehrani, A. Extraction of Micro, Nanocrystalline Cellulose and Textile Fibers from Coffee Waste. J. Test. Eval. 2023, 51, 20220487. [Google Scholar] [CrossRef]
  138. Malhotra, A.; Lokhande, R.S.; Sahu, R.; Jain, S.K.; Sharma, K.B.; Tripathi, B. Study on Adsorbent Characteristics of Coconut Coir as a Bio Sorbent for Removal of Methylene Blue Dye. Macromol. Symp. 2021, 399, 2100082. [Google Scholar] [CrossRef]
  139. Le, P.T.; Bui, H.T.; Le, D.N.; Nguyen, T.H.; Pham, L.A.; Nguyen, H.N.; Nguyen, Q.S.; Nguyen, T.P.; Bich, N.T.; Duong, T.T.; et al. Preparation and Characterization of Biochar Derived from Agricultural By-Products for Dye Removal. Adsorpt. Sci. Technol. 2021, 2021, 9161904. [Google Scholar] [CrossRef]
  140. Nasar, A. Utilization of Tea Wastes for the Removal of Toxic Dyes from Polluted Water—A Review. Biomass Convers. Biorefin 2023, 13, 1399–1415. [Google Scholar] [CrossRef]
  141. Lin, D.; Wu, F.; Hu, Y.; Zhang, T.; Liu, C.; Hu, Q.; Hu, Y.; Xue, Z.; Han, H.; Ko, T.-H. Adsorption of Dye by Waste Black Tea Powder: Parameters, Kinetic, Equilibrium, and Thermodynamic Studies. J. Chem. 2020, 2020, 5431046. [Google Scholar] [CrossRef]
  142. Jain, S.N.; Tamboli, S.R.; Sutar, D.S.; Jadhav, S.R.; Marathe, J.V.; Shaikh, A.A.; Prajapati, A.A. Batch and Continuous Studies for Adsorption of Anionic Dye onto Waste Tea Residue: Kinetic, Equilibrium, Breakthrough and Reusability Studies. J. Clean. Prod. 2020, 252, 119778. [Google Scholar] [CrossRef]
  143. Kaisri, M.B. Application of Chitosan Derivatives as Promising Adsorbents for Treatment of Textile Wastewater; Elsevier Ltd.: Amsterdam, The Netherlands, 2019; ISBN 9780081024911. [Google Scholar]
  144. Amari, A.; Alalwan, B.; Eldirderi, M.M.; Mnif, W.; Ben Rebah, F. Cactus Material-Based Adsorbents for the Removal of Heavy Metals and Dyes: A Review. Mater. Res. Express 2020, 7, 12002. [Google Scholar] [CrossRef]
  145. Ingole, N.W. Colour Removal from Textile Effluent by Using Biomaterials—An Experimental Evaluation. J. Indian. Water Work. Assoc. 2020, LII. [Google Scholar]
  146. Khan, T.A.; Nazir, M.; Khan, E.A. Adsorptive Removal of Rhodamine B from Textile Wastewater Using Water Chestnut (Trapa natans L.) Peel: Adsorption Dynamics and Kinetic Studies. Toxicol. Environ. Chem. 2013, 95, 919–931. [Google Scholar] [CrossRef]
  147. EL-Ghoul, Y.; Ammar, C.; Alminderej, F.M.; Shafiquzzaman, M. Design and Evaluation of a New Natural Multi-Layered Biopolymeric Adsorbent System-Based Chitosan/Cellulosic Nonwoven Material for the Biosorption of Industrial Textile Effluents. Polymers 2021, 13, 322. [Google Scholar] [CrossRef]
  148. Periyasamy, A.P.; Karunakaran, G.; Rwahwire, S.; Kesari, K. Nonwoven Fabrics Developed from Agriculture and Industrial Waste for Acoustic and Thermal Applications. Cellulose 2023, 30, 7329–7346. [Google Scholar] [CrossRef]
  149. Periyasamy, A.P. Nonwoven Fabrics from Agricultural and Industrial Waste for Acoustic and Thermal Insulation Applications. Textiles 2023, 3, 182–200. [Google Scholar] [CrossRef]
  150. Bulgariu, L.; Escudero, L.B.; Bello, O.S.; Iqbal, M.; Nisar, J.; Adegoke, K.A.; Alakhras, F.; Kornaros, M.; Anastopoulos, I. The Utilization of Leaf-Based Adsorbents for Dyes Removal: A Review. J. Mol. Liq. 2019, 276, 728–747. [Google Scholar] [CrossRef]
  151. Deniz, F.; Saygideger, S.D. Equilibrium, Kinetic and Thermodynamic Studies of Acid Orange 52 Dye Biosorption by Paulownia Tomentosa Steud. Leaf Powder as a Low-Cost Natural Biosorbent. Bioresour. Technol. 2010, 101, 5137–5143. [Google Scholar] [CrossRef] [PubMed]
  152. Guerrero-Coronilla, I.; Morales-Barrera, L.; Cristiani-Urbina, E. Kinetic, Isotherm and Thermodynamic Studies of Amaranth Dye Biosorption from Aqueous Solution onto Water Hyacinth Leaves. J. Environ. Manag. 2015, 152, 99–108. [Google Scholar] [CrossRef] [PubMed]
  153. Deniz, F.; Karaman, S. Removal of Basic Red 46 Dye from Aqueous Solution by Pine Tree Leaves. Chem. Eng. J. 2011, 170, 67–74. [Google Scholar] [CrossRef]
  154. Gupta, N.; Kushwaha, A.K.; Chattopadhyaya, M.C. Adsorption Studies of Cationic Dyes onto Ashoka (Saraca asoca) Leaf Powder. J. Taiwan. Inst. Chem. Eng. 2012, 43, 604–613. [Google Scholar] [CrossRef]
  155. Mosoarca, G.; Vancea, C.; Popa, S.; Dan, M.; Boran, S. A Novel High-Efficiency Natural Biosorbent Material Obtained from Sour Cherry (Prunus cerasus) Leaf Biomass for Cationic Dyes Adsorption. Materials 2023, 16, 4252. [Google Scholar] [CrossRef] [PubMed]
  156. Wong, S.; Ghafar, N.A.; Ngadi, N.; Razmi, F.A.; Inuwa, I.M.; Mat, R.; Amin, N.A.S. Effective Removal of Anionic Textile Dyes Using Adsorbent Synthesized from Coffee Waste. Sci. Rep. 2020, 10, 2928. [Google Scholar] [CrossRef] [PubMed]
  157. Babu, S.B.; Priyanka, S.V. Removal of Malachite Green Dye from Aqueous Solution Using Lemon Leaf Powder as an Adsorbent. J. Solid. Waste Technol. Manag. 2023, 49, 141–151. [Google Scholar] [CrossRef]
  158. Rose, P.K.; Poonia, V.; Kumar, R.; Kataria, N.; Sharma, P.; Lamba, J.; Bhattacharya, P. Congo Red Dye Removal Using Modified Banana Leaves: Adsorption Equilibrium, Kinetics, and Reusability Analysis. Groundw. Sustain. Dev. 2023, 23, 101005. [Google Scholar] [CrossRef]
  159. Ngwabebhoh, F.A.; Gazi, M.; Oladipo, A.A. Adsorptive Removal of Multi-Azo Dye from Aqueous Phase Using a Semi-IPN Superabsorbent Chitosan-Starch Hydrogel. Chem. Eng. Res. Des. 2016, 112, 274–288. [Google Scholar] [CrossRef]
  160. Liu, L.; Fan, S.; Li, Y. Removal Behavior of Methylene Blue from Aqueous Solution by Tea Waste: Kinetics, Isotherms and Mechanism. Int. J. Environ. Res. Public. Health 2018, 15, 1321. [Google Scholar] [CrossRef]
  161. Maleki, A.; Daraei, H.; Khodaei, F.; Aghdam, K.B.; Faez, E. Direct Blue 71 Dye Removal Probing by Potato Peel-Based Sorbent: Applications of Artificial Intelligent Systems. New Pub Balaban 2015, 57, 12281–12286. [Google Scholar] [CrossRef]
  162. Tavlieva, M.P.; Genieva, S.D.; Georgieva, V.G.; Vlaev, L.T. Kinetic Study of Brilliant Green Adsorption from Aqueous Solution onto White Rice Husk Ash. J. Colloid. Interface Sci. 2013, 409, 112–122. [Google Scholar] [CrossRef] [PubMed]
  163. de O. Salomón, Y.L.; Georgin, J.; Franco, D.S.P.; Netto, M.S.; Grassi, P.; Piccilli, D.G.A.; Oliveira, M.L.S.; Dotto, G.L. Powdered Biosorbent from Pecan Pericarp (Carya illinoensis) as an Efficient Material to Uptake Methyl Violet 2B from Effluents in Batch and Column Operations. Adv. Powder Technol. 2020, 31, 2843–2852. [Google Scholar] [CrossRef]
  164. Juang, R.-S.; Tseng, R.-L.; Wu, F.-C. Role of Microporosity of Activated Carbons on Their Adsorption Abilities for Phenols and Dyes. Adsorption 2001, 7, 65–72. [Google Scholar] [CrossRef]
  165. Juang, R.-S.; Wu, F.-C.; Tseng, R.-L. Characterization and Use of Activated Carbons Prepared from Bagasses for Liquid-Phase Adsorption. Colloids Surf. A Physicochem. Eng. Asp. 2002, 201, 191–199. [Google Scholar] [CrossRef]
  166. Hachoumi, I.; el Ouahabi, I.; Slimani, R.; Cagnon, B.; el Haddad, M.; el Antri, S.; Lazar, S. Adsorption Studies with a New Biosorbent Ensis Siliqua Shell Powder for Removal Two Textile Dyes from Aqueous Solution. J. Mater. Environ. Sci. 2017, 8, 1448–1459. [Google Scholar]
  167. Sanlier, S.H.; Ak, G.; Yilmaz, H.; Ozbakir, G.; Cagliyan, O. Removal of Textile Dye, Direct Red 23, with Glutaraldehyde Cross-Linked Magnetic Chitosan Beads. Prep. Biochem. Biotechnol. 2013, 43, 163–176. [Google Scholar] [CrossRef]
  168. Yu, Y.; Qiao, N.; Wang, D.; Zhu, Q.; Fu, F.; Cao, R.; Wang, R.; Liu, W.; Xu, B. Fluffy Honeycomb-like Activated Carbon from Popcorn with High Surface Area and Well-Developed Porosity for Ultra-High Efficiency Adsorption of Organic Dyes. Bioresour. Technol. 2019, 285, 121340. [Google Scholar] [CrossRef]
  169. Salama, A.; Shukry, N.; El-Sakhawy, M. Carboxymethyl Cellulose-g-Poly(2-(Dimethylamino) Ethyl Methacrylate) Hydrogel as Adsorbent for Dye Removal. Int. J. Biol. Macromol. 2015, 73, 72–75. [Google Scholar] [CrossRef]
  170. Yoshida, H.; Okamoto, A.; Kataoka, T. Adsorption of Acid Dye on Cross-Linked Chitosan Fibers: Equilibria. Chem. Eng. Sci. 1993, 48, 2267–2272. [Google Scholar] [CrossRef]
  171. Yan, Y.; Xiang, B.; Li, Y.; Jia, Q. Preparation and Adsorption Properties of Diethylenetriamine-Modified Chitosan Beads for Acid Dyes. J. Appl. Polym. Sci. 2013, 130, 4090–4098. [Google Scholar] [CrossRef]
  172. Liu, L.; Wang, R.; Yu, J.; Hu, L.; Wang, Z.; Fan, Y. Adsorption of Reactive Blue 19 from Aqueous Solution by Chitin Nanofiber-/Nanowhisker-Based Hydrogels. RSC Adv. 2018, 8, 15804–15812. [Google Scholar] [CrossRef] [PubMed]
  173. Jiang, X.; Sun, Y.; Liu, L.; Wang, S.; Tian, X. Adsorption of C.I. Reactive Blue 19 from Aqueous Solutions by Porous Particles of the Grafted Chitosan. Chem. Eng. J. 2014, 235, 151–157. [Google Scholar] [CrossRef]
  174. Tian, X.; Hua, F.; Lou, C.; Jiang, X. Cationic Cellulose Nanocrystals (CCNCs) and Chitosan Nanocomposite Films Filled with CCNCs for Removal of Reactive Dyes from Aqueous Solutions. Cellulose 2018, 25, 3927–3939. [Google Scholar] [CrossRef]
  175. Xu, H.; Zhang, Y.; Jiang, Q.; Reddy, N.; Yang, Y. Biodegradable Hollow Zein Nanoparticles for Removal of Reactive Dyes from Wastewater. J. Environ. Manag. 2013, 125, 33–40. [Google Scholar] [CrossRef] [PubMed]
  176. Nga, N.K.; Chinh, H.D.; Hong, P.T.T.; Huy, T.Q. Facile Preparation of Chitosan Films for High Performance Removal of Reactive Blue 19 Dye from Aqueous Solution. J. Polym. Environ. 2016, 25, 146–155. [Google Scholar] [CrossRef]
  177. Chen, C.Y.; Chang, J.C.; Chen, A.H. Competitive Biosorption of Azo Dyes from Aqueous Solution on the Templated Crosslinked-Chitosan Nanoparticles. J. Hazard. Mater. 2011, 185, 430–441. [Google Scholar] [CrossRef]
  178. Chen, A.H.; Huang, Y.Y. Adsorption of Remazol Black 5 from Aqueous Solution by the Templated Crosslinked-Chitosans. J. Hazard. Mater. 2010, 177, 668–675. [Google Scholar] [CrossRef]
  179. Mohagheghian, A.; Vahidi-Kolur, R.; Pourmohseni, M.; Yang, J.K.; Shirzad-Siboni, M. Application of Scallop Shell-Fe3O4 Nano-Composite for the Removal Azo Dye from Aqueous Solutions. Water Air Soil. Pollut. 2015, 226, 1–16. [Google Scholar] [CrossRef]
  180. Kim, T.Y.; Park, S.S.; Cho, S.Y. Adsorption Characteristics of Reactive Black 5 onto Chitosan Beads Cross-Linked with Epichlorohydrin. J. Ind. Eng. Chem. 2012, 18, 1458–1464. [Google Scholar] [CrossRef]
  181. Auta, M. Batch Adsorption of Reactive Red 120 from Waste Waters Using Activated Carbon from Waste Tea. Int. J. Adv. Eng. Technol. 2012, III, 24–28. [Google Scholar]
  182. Chen, A.H.; Chen, S.M. Biosorption of Azo Dyes from Aqueous Solution by Glutaraldehyde-Crosslinked Chitosans. J. Hazard. Mater. 2009, 172, 1111–1121. [Google Scholar] [CrossRef] [PubMed]
  183. Jóźwiak, T.; Filipkowska, U.; Szymczyk, P.; Kuczajowska-Zadrożna, M.; Mielcarek, A. Application of Chitosan Ionically Crosslinked with Sodium Edetate for Reactive Dyes Removal from Aqueous Solutions. Prog. Chem. Appl. Chitin Deriv. 2015, 20, 82–96. [Google Scholar] [CrossRef]
  184. Jóźwiak, T.; Filipkowska, U.; Szymczyk, P.; Zyśk, M. Effect of the Form and Deacetylation Degree of Chitosan Sorbents on Sorption Effectiveness of Reactive Black 5 from Aqueous Solutions. Int. J. Biol. Macromol. 2017, 95, 1169–1178. [Google Scholar] [CrossRef] [PubMed]
  185. Naushad, M.; Lichtfouse, E. (Eds.) Green Materials for Wastewater Treatment; Environmental Chemistry for a Sustainable World; Springer International Publishing: Cham, Switzerland, 2020; Volume 38, ISBN 978-3-030-17723-2. [Google Scholar]
  186. Srivastava, R.; Rupainwar, D.C. Removal of Hazardous Triphenylmethane Dye through Adsorption over Waste Material-Mango Bark Powder. IJCT 2011, 18, 469–474. [Google Scholar]
  187. Dai, L.; Zhu, W.; He, L.; Tan, F.; Zhu, N.; Zhou, Q.; He, M.; Hu, G. Calcium-Rich Biochar from Crab Shell: An Unexpected Super Adsorbent for Dye Removal. Bioresour. Technol. 2018, 267, 510–516. [Google Scholar] [CrossRef]
  188. Chen, Y.D.; Lin, Y.C.; Ho, S.H.; Zhou, Y.; Ren, N.Q. Highly Efficient Adsorption of Dyes by Biochar Derived from Pigments-Extracted Macroalgae Pyrolyzed at Different Temperature. Bioresour. Technol. 2018, 259, 104–110. [Google Scholar] [CrossRef]
  189. Banna Motejadded Emrooz, H.; Maleki, M.; Rahmani, A. Azolla-Derived Hierarchical Nanoporous Carbons: From Environmental Concerns to Industrial Opportunities. J. Taiwan. Inst. Chem. Eng. 2018, 91, 281–290. [Google Scholar] [CrossRef]
  190. Merah, M.; Boudoukha, C.; Avalos Ramirez, A.; Haroun, M.F.; Maane, S. High Biosorption of Cationic Dye onto a Novel Material Based on Paper Mill Sludge. Sci Rep 2023, 13, 15926. [Google Scholar] [CrossRef]
  191. Perendija, J.; Ljubić, V.; Popović, M.; Milošević, D.; Arsenijević, Z.; Đuriš, M.; Kovač, S.; Cvetković, S. Assessment of Waste Hop (Humulus lupulus) Stems as a Biosorbent for the Removal of Malachite Green, Methylene Blue, and Crystal Violet from Aqueous Solution in Batch and Fixed-Bed Column Systems: Biosorption Process and Mechanism. J. Mol. Liq. 2024, 394, 123770. [Google Scholar] [CrossRef]
  192. Narasaiah, B.P.; Mandal, B.K. Remediation of Azo-Dyes Based Toxicity by Agro-Waste Cotton Boll Peels Mediated Palladium Nanoparticles. J. Saudi Chem. Soc. 2020, 24, 267–281. [Google Scholar] [CrossRef]
  193. Vasu, D.; Kumar, S.; Walia, Y.K. Removal of Dyes Using Wheat Husk Waste as a Low-Cost Adsorbent. Environ. Claims J. 2020, 32, 67–76. [Google Scholar] [CrossRef]
  194. Holkar, C.R.; Jadhav, A.J.; Pinjari, D.V.; Mahamuni, N.M.; Pandit, A.B. A Critical Review on Textile Wastewater Treatments: Possible Approaches. J. Environ. Manag. 2016, 182, 351–366. [Google Scholar] [CrossRef] [PubMed]
  195. Sarayu, K.; Sandhya, S. Current Technologies for Biological Treatment of Textile Wastewater-A Review. Appl. Biochem. Biotechnol. 2012, 167, 645–661. [Google Scholar] [CrossRef] [PubMed]
  196. Campos, R.; Kandelbauer, A.; Robra, K.H.; Cavaco-Paulo, A.; Gübitz, G.M. Indigo Degradation with Purified Laccases from Trametes hirsuta and Sclerotium rolfsii. J. Biotechnol. 2001, 89, 131–139. [Google Scholar] [CrossRef] [PubMed]
  197. Misal, S.A.; Gawai, K.R. Azoreductase: A Key Player of Xenobiotic Metabolism. Bioresour. Bioprocess. 2018, 5, 17. [Google Scholar] [CrossRef]
  198. Ardila-Leal, L.D.; Poutou-Piñales, R.A.; Pedroza-Rodríguez, A.M.; Quevedo-Hidalgo, B.E. A Brief History of Colour, the Environmental Impact of Synthetic Dyes and Removal by Using Laccases. Molecules 2021, 26, 3813. [Google Scholar] [CrossRef]
  199. Bento, R.M.F.; Almeida, M.R.; Bharmoria, P.; Freire, M.G.; Tavares, A.P.M. Improvements in the Enzymatic Degradation of Textile Dyes Using Ionic- Liquid-Based Surfactants. Sep. Purif. Technol. 2020, 235, 116191. [Google Scholar] [CrossRef]
  200. Jadhav, J.P.; Kalyani, D.C.; Telke, A.A.; Phugare, S.S.; Govindwar, S.P. Evaluation of the Efficacy of a Bacterial Consortium for the Removal of Color, Reduction of Heavy Metals, and Toxicity from Textile Dye Effluent. Bioresour. Technol. 2010, 101, 165–173. [Google Scholar] [CrossRef]
  201. Chan, G.F.; Gan, H.M.; Rashid, N.A.A. Genome Sequence of Citrobacter Sp. Strain A1, a Dye-Degrading Bacterium. J. Bacteriol. 2012, 194, 5485–5486. [Google Scholar] [CrossRef] [PubMed]
  202. Sondhi, S. Sustainable Approaches in Effluent Treatment; Elsevier Ltd.: Amsterdam, The Netherlands, 2020; ISBN 9780081028674. [Google Scholar]
  203. Vijayaraghavan, K.; Yun, Y.-S. Chemical Modification and Immobilization of Corynebacterium Glutamicum for Biosorption of Reactive Black 5 from Aqueous Solution. Ind. Eng. Chem. Res. 2007, 46, 608–617. [Google Scholar] [CrossRef]
  204. Szentirmai, É.; Massie, A.R.; Kapás, L. Lipoteichoic Acid, a Cell Wall Component of Gram-Positive Bacteria, Induces Sleep and Fever and Suppresses Feeding. Brain Behav. Immun. 2021, 92, 184–192. [Google Scholar] [CrossRef] [PubMed]
  205. Banat, I.M.; Nigam, P.; Singh, D.; Marchant, R. Microbial Decolorization of Textile-Dyecontaining Effluents: A Review. Bioresour. Technol. 1996, 58, 217–227. [Google Scholar] [CrossRef]
  206. Fu, Y.; Viraraghavan, T. Dye Biosorption Sites in Aspergillus Niger. Bioresour. Technol. 2002, 82, 139–145. [Google Scholar] [CrossRef]
  207. Salvi, N.A.; Chattopadhyay, S. Biosorption of Azo Dyes by Spent Rhizopus Arrhizus Biomass. Appl. Water Sci. 2017, 7, 3041–3054. [Google Scholar] [CrossRef]
  208. Das, S.K.; Bhowal, J.; Das, A.R.; Guha, A.K. Adsorption Behavior of Rhodamine B on Rhizopus Oryzae Biomass. Langmuir 2006, 22, 7265–7272. [Google Scholar] [CrossRef]
  209. Hernández-Zamora, M.; Cristiani-Urbina, E.; Martínez-Jerónimo, F.; Perales-Vela, H.V.; Ponce-Noyola, T.; del Carmen Montes-Horcasitas, M.; Cañizares-Villanueva, R.O. Bioremoval of the Azo Dye Congo Red by the Microalga Chlorella Vulgaris. Environ. Sci. Pollut. Res. 2015, 22, 10811–10823. [Google Scholar] [CrossRef]
  210. Tamil Selvan, S.; Dakshinamoorthi, B.M.; Chandrasekaran, R.; Muthusamy, S.; Ramamurthy, D.; Balasundaram, S. Integrating Eco-Technological Approach for Textile Dye Effluent Treatment and Carbon Dioxide Capturing from Unicellular Microalga Chlorella Vulgaris RDS03: A Synergistic Method. Int. J. Phytoremediation 2023, 25, 466–482. [Google Scholar] [CrossRef]
  211. Khataee, A.R.; Dehghan, G.; Ebadi, A.; Zarei, M.; Pourhassan, M. Biological Treatment of a Dye Solution by Macroalgae Chara Sp.: Effect of Operational Parameters, Intermediates Identification and Artificial Neural Network Modeling. Bioresour. Technol. 2010, 101, 2252–2258. [Google Scholar] [CrossRef]
  212. Daneshvar, N.; Ayazloo, M.; Khataee, A.R.; Pourhassan, M. Biological Decolorization of Dye Solution Containing Malachite Green by Microalgae Cosmarium Sp. Bioresour. Technol. 2007, 98, 1176–1182. [Google Scholar] [CrossRef] [PubMed]
  213. Padmesh, T.V.N.; Vijayaraghavan, K.; Sekaran, G.; Velan, M. Application of Azolla Rongpong on Biosorption of Acid Red 88, Acid Green 3, Acid Orange 7 and Acid Blue 15 from Synthetic Solutions. Chem. Eng. J. 2006, 122, 55–63. [Google Scholar] [CrossRef]
  214. Padmesh, T.V.N.; Vijayaraghavan, K.; Sekaran, G.; Velan, M. Biosorption of Acid Blue 15 Using Fresh Water Macroalga Azolla Filiculoides: Batch and Column Studies. Dye Pigment 2006, 71, 77–82. [Google Scholar] [CrossRef]
  215. Özer, A.; Akkaya, G.; Turabik, M. Biosorption of Acid Red 274 (AR 274) on Enteromorpha Prolifera in a Batch System. J. Hazard. Mater. 2005, 126, 119–127. [Google Scholar] [CrossRef]
  216. Jamee, R.; Siddique, R. Biodegradation of Synthetic Dyes of Textile Effluent by Microorganisms: An Environmentally and Economically Sustainable Approach. Eur. J. Microbiol. Immunol. 2019, 9, 114–118. [Google Scholar] [CrossRef] [PubMed]
  217. El-Kassas, H.Y.; Mohamed, L.A. Bioremediation of the Textile Waste Effluent by Chlorella Vulgaris. Egypt. J. Aquat. Res. 2014, 40, 301–308. [Google Scholar] [CrossRef]
  218. Chin, J.Y.; Chng, L.M.; Leong, S.S.; Yeap, S.P.; Yasin, N.H.M.; Toh, P.Y. Removal of Synthetic Dye by Chlorella Vulgaris Microalgae as Natural Adsorbent. Arab. J. Sci. Eng. 2020, 45, 7385–7395. [Google Scholar] [CrossRef]
  219. Gangola, S.; Bhatt, P.; Chaudhary, P.; Khati, P.; Kumar, N.; Sharma, A. Bioremediation of Industrial Waste Using Microbial Metabolic Diversity. In Microbial Biotechnology in Environmental Monitoring and Cleanup; IGI Global: Pennsylvania, PA, USA, 2018; pp. 1–27. [Google Scholar]
  220. Wang, J.; Lu, L.; Feng, F. Improving the Indigo Carmine Decolorization Ability of a Bacillus Amyloliquefaciens Laccase by Site-Directed Mutagenesis. Catalysts 2017, 7, 275. [Google Scholar] [CrossRef]
  221. Harish, B.S.; Thayumanavan, T.; Nambukrishnan, V.; Sakthishobana, K. Heterogeneous Biocatalytic System for Effective Decolorization of Textile Dye Effluent. 3 Biotech. 2023, 13, 165. [Google Scholar] [CrossRef]
  222. Sarkar, S.; Banerjee, A.; Chakraborty, N.; Soren, K.; Chakraborty, P.; Bandopadhyay, R. Structural-Functional Analyses of Textile Dye Degrading Azoreductase, Laccase and Peroxidase: A Comparative in Silico Study. Electron. J. Biotechnol. 2020, 43, 48–54. [Google Scholar] [CrossRef]
  223. Dong, H.; Guo, T.; Zhang, W.; Ying, H.; Wang, P.; Wang, Y.; Chen, Y. Biochemical Characterization of a Novel Azoreductase from Streptomyces Sp.: Application in Eco-Friendly Decolorization of Azo Dye Wastewater. Int J Biol Macromol 2019, 140, 1037–1046. [Google Scholar] [CrossRef] [PubMed]
  224. Sherifah, M.W.; Seun, A.E.; Kehinde, O.S.; Abiodun, A.O. Decolourization of Synthetic Dyes by Laccase Enzyme Produced by Kluyveromyces Dobzhanskii DW1 and Pichia Manshurica DW2. Afr. J. Biotechnol. 2019, 18, 1–11. [Google Scholar] [CrossRef]
  225. Darvishi, F.; Moradi, M.; Jolivalt, C.; Madzak, C. Laccase Production from Sucrose by Recombinant Yarrowia Lipolytica and Its Application to Decolorization of Environmental Pollutant Dyes. Ecotoxicol. Environ. Saf. 2018, 165, 278–283. [Google Scholar] [CrossRef] [PubMed]
  226. Mustafa, G.; Zahid, M.T.; Ullah, F.; Zia, I.; Younas, A.; Batool, T.; Zahid, I. Bacterial Tools for the Removal and Degradation of Synthetic Dyes from the Wastewater. In Current Developments in Bioengineering and Biotechnology; Elsevier: Amsterdam, The Netherlands, 2023; pp. 339–370. [Google Scholar]
  227. Moyo, S.; Makhanya, B.P.; Zwane, P.E. Use of Bacterial Isolates in the Treatment of Textile Dye Wastewater: A Review. Heliyon 2022, 8, e09632. [Google Scholar] [CrossRef] [PubMed]
  228. Solís, M.; Solís, A.; Pérez, H.I.; Manjarrez, N.; Flores, M. Microbial Decolouration of Azo Dyes: A Review. Process Biochem. 2012, 47, 1723–1748. [Google Scholar] [CrossRef]
  229. Shafqat, M.; Khalid, A.; Mahmood, T.; Siddique, M.T.; Han, J.; Habteselassie, M.Y. Evaluation of Bacteria Isolated from Textile Wastewater and Rhizosphere to Simultaneously Degrade Azo Dyes and Promote Plant Growth. J. Chem. Technol. Biotechnol. 2017, 92, 2760–2768. [Google Scholar] [CrossRef]
  230. Zhuang, M.; Sanganyado, E.; Zhang, X.; Xu, L.; Zhu, J.; Liu, W.; Song, H. Azo Dye Degrading Bacteria Tolerant to Extreme Conditions Inhabit Nearshore Ecosystems: Optimization and Degradation Pathways. J. Environ. Manag. 2020, 261, 110222. [Google Scholar] [CrossRef] [PubMed]
  231. Saravanan, S.; Carolin, C.F.; Kumar, P.S.; Chitra, B.; Rangasamy, G. Biodegradation of Textile Dye Rhodamine-B by Brevundimonas Diminuta and Screening of Their Breakdown Metabolites. Chemosphere 2022, 308, 136266. [Google Scholar] [CrossRef]
  232. Sulejmanović, J.; Kojčin, M.; Grebo, M.; Zahirović, A.; Topčagić, A.; Smječanin, N.; Al-Kahtani, A.A.; Sher, F. Functionalised Mesoporous Biosorbents for Efficient Removal of Hazardous Pollutants from Water Environment. J. Water Process Eng. 2023, 55, 104219. [Google Scholar] [CrossRef]
  233. Al-Tohamy, R.; Ali, S.S.; Xie, R.; Schagerl, M.; Khalil, M.A.; Sun, J. Decolorization of Reactive Azo Dye Using Novel Halotolerant Yeast Consortium HYC and Proposed Degradation Pathway. Ecotoxicol. Environ. Saf. 2023, 263, 115258. [Google Scholar] [CrossRef]
  234. Velayutham, K.; Madhava, A.K.; Pushparaj, M.; Thanarasu, A.; Devaraj, T.; Periyasamy, K.; Subramanian, S. Biodegradation of Remazol Brilliant Blue R Using Isolated Bacterial Culture (Staphylococcus Sp. K2204). Environ. Technol. 2018, 39, 2900–2907. [Google Scholar] [CrossRef] [PubMed]
  235. Pan, H.; Xu, X.; Wen, Z.; Kang, Y.; Wang, X.; Ren, Y.; Huang, D. Decolorization Pathways of Anthraquinone Dye Disperse Blue 2BLN by Aspergillus Sp. XJ-2 CGMCC12963. Bioengineered 2017, 8, 630–641. [Google Scholar] [CrossRef] [PubMed]
  236. Ranganathan, K.; Karunagaran, K.; Sharma, D.C. Recycling of Wastewaters of Textile Dyeing Industries Using Advanced Treatment Technology and Cost Analysis—Case Studies. Resour. Conserv. Recycl. 2007, 50, 306–318. [Google Scholar] [CrossRef]
  237. Balcik-Canbolat, C.; Sengezer, C.; Sakar, H.; Karagunduz, A.; Keskinler, B. Recovery of Real Dye Bath Wastewater Using Integrated Membrane Process: Considering Water Recovery, Membrane Fouling and Reuse Potential of Membranes. Environ. Technol. 2017, 38, 2668–2676. [Google Scholar] [CrossRef] [PubMed]
  238. Van der Bruggen, B.; Canbolat, Ç.B.; Lin, J.; Luis, P. The Potential of Membrane Technology for Treatment of Textile Wastewater. In Sustainable Membrane Technology for Water and Wastewater Treatment; Figoli, A., Criscuoli, A., Eds.; Springer: Singapore, 2017; pp. 349–380. ISBN 978-981-10-5623-9. [Google Scholar]
  239. Tang, C.; Chen, V. Nanofiltration of Textile Wastewater for Water Reuse. Desalination 2002, 143, 11–20. [Google Scholar] [CrossRef]
  240. Li, K.; Liu, Q.; Fang, F.; Wu, X.; Xin, J.; Sun, S.; Wei, Y.; Ruan, R.; Chen, P.; Wang, Y.; et al. Influence of Nanofiltration Concentrate Recirculation on Performance and Economic Feasibility of a Pilot-Scale Membrane Bioreactor-Nanofiltration Hybrid Process for Textile Wastewater Treatment with High Water Recovery. J. Clean. Prod. 2020, 261, 121067. [Google Scholar] [CrossRef]
  241. Bautista, P.; Mohedano, A.F.; Casas, J.A.; Zazo, J.A.; Rodriguez, J.J. An Overview of the Application of Fenton Oxidation to Industrial Wastewaters Treatment. J. Chem. Technol. Biotechnol. 2008, 83, 1323–1338. [Google Scholar] [CrossRef]
  242. Chen, Y.; Truong, V.N.T.; Bu, X.; Xie, G. A Review of Effects and Applications of Ultrasound in Mineral Flotation. Ultrason. Sonochem. 2020, 60, 104739. [Google Scholar] [CrossRef]
  243. Badawy, M.; Ali, M. Fenton’s Peroxidation and Coagulation Processes for the Treatment of Combined Industrial and Domestic Wastewater. J Hazard Mater 2006, 136, 961–966. [Google Scholar] [CrossRef]
  244. Peng, Q.; Tan, X.; Xiong, X.; Wang, Y.; Novotná, J.; Shah, K.V.; Stempień, Z.; Periyasamy, A.P.; Kejzlar, P.; Venkataraman, M.; et al. Insights into the Large-size Graphene Improvement Effect of the Mechanical Properties on the Epoxy/Glass Fabric Composites. Polym. Compos. 2023, 44, 7430–7443. [Google Scholar] [CrossRef]
  245. Hung, Y.-T.; Lo, H.H.; Wang, L.K.; Taricska, J.R.; Li, K.H. Granular Activated Carbon Adsorption. In Physicochemical Treatment Processes; Wang, L.K., Hung, Y.-T., Shammas, N.K., Eds.; Humana Press: Totowa, NJ, USA, 2005; Volume 3, pp. 573–633. ISBN 978-1-59259-820-5. [Google Scholar]
  246. Belaid, K.D.; Kacha, S.; Kameche, M.; Derriche, Z. Adsorption Kinetics of Some Textile Dyes onto Granular Activated Carbon. J. Environ. Chem. Eng. 2013, 1, 496–503. [Google Scholar] [CrossRef]
  247. Abou-Elela, S.I.; Ali, M.E.M.; Ibrahim, H.S. Combined Treatment of Retting Flax Wastewater Using Fenton Oxidation and Granular Activated Carbon. Arab. J. Chem. 2016, 9, 511–517. [Google Scholar] [CrossRef]
  248. Arfin, T.; Varshney, N.; Singh, B. Ionic Liquid Modified Activated Carbon for the Treatment of Textile Wastewater. Green Mater. Wastewater Treat. Environ. Chem. A Sustain. World 2020, 38, 257–275. [Google Scholar] [CrossRef]
  249. Bakht Shokouhi, S.; Dehghanzadeh, R.; Aslani, H.; Shahmahdi, N. Activated Carbon Catalyzed Ozonation (ACCO) of Reactive Blue 194 Azo Dye in Aqueous Saline Solution: Experimental Parameters, Kinetic and Analysis of Activated Carbon Properties. J. Water Process Eng. 2020, 35, 101188. [Google Scholar] [CrossRef]
  250. Bilińska, L.; Gmurek, M.; Ledakowicz, S. Comparison between Industrial and Simulated Textile Wastewater Treatment by AOPs—Biodegradability, Toxicity and Cost Assessment. Chem. Eng. J. 2016, 306, 550–559. [Google Scholar] [CrossRef]
  251. Kumar, P.S.; Yaashikaa, P.R. Sustainable Innovations in Apparel Production; Springer: Singapore, 2018; ISBN 978-981-10-8590-1. [Google Scholar]
  252. Yu, S.; Zhang, W.; Dong, X.; Wang, F.; Yang, W.; Liu, C.; Chen, D. A Review on Recent Advances of Biochar from Agricultural and Forestry Wastes: Preparation, Modification and Applications in Wastewater Treatment. J. Environ. Chem. Eng. 2024, 12, 111638. [Google Scholar] [CrossRef]
  253. Ilhan, F.; Ulucan-Altuntas, K.; Dogan, C.; Kurt, U. Treatability of Raw Textile Wastewater Using Fenton Process and Its Comparison with Chemical Coagulation. Desalination Water Treat. 2019, 162, 142–148. [Google Scholar] [CrossRef]
  254. Sözen, S.; Olmez-Hanci, T.; Hooshmand, M.; Orhon, D. Fenton Oxidation for Effective Removal of Color and Organic Matter from Denim Cotton Wastewater without Biological Treatment. Environ. Chem. Lett. 2020, 18, 207–213. [Google Scholar] [CrossRef]
  255. Asghar, A.; Abdul Raman, A.A.; Wan Daud, W.M.A. Advanced Oxidation Processes for In-Situ Production of Hydrogen Peroxide/Hydroxyl Radical for Textile Wastewater Treatment: A Review. J. Clean. Prod. 2015, 87, 826–838. [Google Scholar] [CrossRef]
  256. Yang, J.; Jia, K.; Lu, S.; Cao, Y.; Boczkaj, G.; Wang, C. Thermally Activated Natural Chalcopyrite for Fenton-like Degradation of Rhodamine B: Catalyst Characterization, Performance Evaluation, and Catalytic Mechanism. J. Environ. Chem. Eng. 2024, 12, 111469. [Google Scholar] [CrossRef]
  257. Bethi, B.; Radhika, G.B.; Sonawane, S.H. Fundamentals of Advanced Oxidation Processes (AOPs) for Wastewater Treatment: Challenges and Opportunities. In Novel Approaches Towards Wastewater Treatment and Resource Recovery Technologies; Elsevier: Amsterdam, The Netherlands, 2022; pp. 209–220. [Google Scholar]
  258. Somensi, C.A.; Simionatto, E.L.; Bertoli, S.L.; Wisniewski, A.; Radetski, C.M. Use of Ozone in a Pilot-Scale Plant for Textile Wastewater Pre-Treatment: Physico-Chemical Efficiency, Degradation by-Products Identification and Environmental Toxicity of Treated Wastewater. J. Hazard. Mater. 2010, 175, 235–240. [Google Scholar] [CrossRef]
  259. Sillanpää, M.E.T.; Agustiono Kurniawan, T.; Lo, W. Degradation of Chelating Agents in Aqueous Solution Using Advanced Oxidation Process (AOP). Chemosphere 2011, 83, 1443–1460. [Google Scholar] [CrossRef] [PubMed]
  260. Sevimli, M.F.; Sarikaya, H.Z. Ozone Treatment of Textile Effluents and Dyes: Effect of Applied Ozone Dose, PH and Dye Concentration. J. Chem. Technol. Biotechnol. 2002, 77, 842–850. [Google Scholar] [CrossRef]
  261. Gähr, F.; Hermanutz, F.; Oppermann, W. Ozonation- An Important Technique to Comly with New German Laws for Textile Waste-Water Treatment. Water Sci. Technol. 1994, 30, 255–263. [Google Scholar] [CrossRef]
  262. Arslan-Alaton, I.; Koba-Ucun, O. Treatment of Reactive Dye Hydrolysates with UV-C- and Ozone-Activated Percarbonate and Persulfate. Int. J. Environ. Res. 2023, 17, 58. [Google Scholar] [CrossRef]
  263. Demirev, A.; Nenov, V. Ozonation of Two Acidic Azo Dyes with Different Substituents. Ozone Sci. Eng. 2005, 27, 475–485. [Google Scholar] [CrossRef]
  264. Fanchiang, J.M.; Tseng, D.H. Degradation of Anthraquinone Dye, C.I. Reactive Blue 19 in Aqueous Solution by Ozonation. Chemosphere 2009, 77, 214–221. [Google Scholar] [CrossRef] [PubMed]
  265. Iyer, S.R.S. Chapter 4—Physical Chemistry of Dyeing: Kinetics, Equilibrium, Dye-Fiber Affinity, and Mechanisms. In The Chemistry of Synthetic Dyes; Venkataraman, K., Ed.; Academic Press: Cambridge, MA, USA, 1974; pp. 115–275. ISBN 978-0-12-717007-7. [Google Scholar]
  266. Shuttleworth, L.; Weaver, M.A. The Chemistry and Application of Dyes; Waring, D.R., Hallas, G., Eds.; Springer: Boston, MA, USA, 1990; pp. 107–163. ISBN 978-1-4684-7715-3. [Google Scholar]
  267. Bai, H.; Liang, L.; Cao, P.; Zhang, H.; Chen, S.; Yu, H.; Quan, X. MgAl2O4 Incorporated Catalytic Ceramic Membrane for Catalytic Ozonation of Organic Pollutants. Appl. Catal. B. 2024, 343, 123527. [Google Scholar] [CrossRef]
  268. Ebrahimi, I.; Parvinzadeh Gashti, M.; Sarafpour, M. Photocatalytic Discoloration of Denim Using Advanced Oxidation Process with H2O2/UV. J. Photochem. Photobiol. A Chem. 2018, 360, 278–288. [Google Scholar] [CrossRef]
  269. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  270. Staehelin, J.; Hoigne, J. Decomposition of Ozone in Water: Rate of Initiation by Hydroxide Ions and Hydrogen Peroxide. Environ. Sci. Technol. 1982, 16, 676–681. [Google Scholar] [CrossRef]
  271. Khosravi, A.; Karimi, M.; Ebrahimi, H.; Fallah, N. Sequencing Batch Reactor/Nanofiltration Hybrid Method for Water Recovery from Textile Wastewater Contained Phthalocyanine Dye and Anionic Surfactant. J. Environ. Chem. Eng. 2020, 8, 103701. [Google Scholar] [CrossRef]
  272. Khouni, I.; Marrot, B.; Amar, R. Ben Treatment of Reconstituted Textile Wastewater Containing a Reactive Dye in an Aerobic Sequencing Batch Reactor Using a Novel Bacterial Consortium. Sep. Purif. Technol. 2012, 87, 110–119. [Google Scholar] [CrossRef]
  273. Mendes, M.; Moreira, I.; Moreira, P.; Pintado, M.; Castro, P. Bioaugmentation of Aerobic Granular Sludge with Dye-Decolorizing Yeast for Textile Industrial Wastewater. Processes 2023, 11, 1654. [Google Scholar] [CrossRef]
  274. Lourenço, N.D.; Novais, J.M.; Pinheiro, H.M. Effect of Some Operational Parameters on Textile Dye Biodegradation in a Sequential Batch Reactor. J. Biotechnol. 2001, 89, 163–174. [Google Scholar] [CrossRef] [PubMed]
  275. Nuid, M.; Aris, A.; Krishnen, R.; Chelliapan, S.; Muda, K. Pineapple Wastewater as Co-Substrate in Treating Real Alkaline, Non-Biodegradable Textile Wastewater Using Biogranulation Technology. J. Environ. Manage. 2023, 344, 118501. [Google Scholar] [CrossRef]
  276. Xu, T.; Huang, C. Electrodialysis-based Separation Technologies: A Critical Review. AIChE J. 2008, 54, 3147–3159. [Google Scholar] [CrossRef]
  277. Lin, J.; Ye, W.; Huang, J.; Ricard, B.; Baltaru, M.-C.; Greydanus, B.; Balta, S.; Shen, J.; Vlad, M.; Sotto, A.; et al. Toward Resource Recovery from Textile Wastewater: Dye Extraction, Water and Base/Acid Regeneration Using a Hybrid NF-BMED Process. ACS Sustain. Chem. Eng. 2015, 3, 1993–2001. [Google Scholar] [CrossRef]
  278. Chakachaka, V.; Tshangana, C.; Mahlangu, O.; Mamba, B.; Muleja, A. Interdependence of Kinetics and Fluid Dynamics in the Design of Photocatalytic Membrane Reactors. Membranes 2022, 12, 745. [Google Scholar] [CrossRef]
  279. Kumar, S.; Bhawna; Sharma, R.; Gupta, A.; Dubey, K.K.; Khan, A.M.; Singhal, R.; Kumar, R.; Bharti, A.; Singh, P.; et al. TiO2 Based Photocatalysis Membranes: An Efficient Strategy for Pharmaceutical Mineralization. Sci. Total Environ. 2022, 845, 157221. [Google Scholar] [CrossRef]
  280. Laohaprapanon, S.; Matahum, J.; Tayo, L.; You, S.-J. Photodegradation of Reactive Black 5 in a ZnO/UV Slurry Membrane Reactor. J. Taiwan Inst. Chem. Eng. 2015, 49, 136–141. [Google Scholar] [CrossRef]
  281. Bakar, B.; Birhanlı, E.; Ulu, A.; Boran, F.; Yeşilada, Ö.; Ateş, B. Immobilization of Trametes Trogii Laccase on Polyvinylpyrrolidone-Coated Magnetic Nanoparticles for Biocatalytic Degradation of Textile Dyes. Biocatal. Biotransformation 2023, 1–18. [Google Scholar] [CrossRef]
  282. Goei, R.; Lim, T.-T. Ag-Decorated TiO2 Photocatalytic Membrane with Hierarchical Architecture: Photocatalytic and Anti-Bacterial Activities. Water Res. 2014, 59, 207–218. [Google Scholar] [CrossRef]
  283. Shafiq, I.; Shafique, S.; Akhter, P.; Yang, W.; Hussain, M. Recent Developments in Alumina Supported Hydrodesulfurization Catalysts for the Production of Sulfur-Free Refinery Products: A Technical Review. Catal. Rev. 2022, 64, 1–86. [Google Scholar] [CrossRef]
  284. Shafiq, I.; Shafique, S.; Akhter, P.; Ishaq, M.; Yang, W.; Hussain, M. Recent Breakthroughs in Deep Aerobic Oxidative Desulfurization of Petroleum Refinery Products. J. Clean. Prod. 2021, 294, 125731. [Google Scholar] [CrossRef]
  285. Shafiq, I.; Hussain, M.; Shafique, S.; Rashid, R.; Akhter, P.; Ahmed, A.; Jeon, J.-K.; Park, Y.-K. Oxidative Desulfurization of Refinery Diesel Pool Fractions Using LaVO4 Photocatalyst. J. Ind. Eng. Chem. 2021, 98, 283–288. [Google Scholar] [CrossRef]
  286. Yu, C.; Wang, K.; Yang, P.; Yang, S.; Lu, C.; Song, Y.; Dong, S.; Sun, J.; Sun, J. One-Pot Facile Synthesis of Bi2S3/SnS2/Bi2O3 Ternary Heterojunction as Advanced Double Z-Scheme Photocatalytic System for Efficient Dye Removal under Sunlight Irradiation. Appl. Surf. Sci. 2017, 420, 233–242. [Google Scholar] [CrossRef]
  287. Cheng, L.; Zhang, Y.; Fan, W.; Ji, Y. Synergistic Adsorption-Photocatalysis for Dyes Removal by a Novel Biochar–Based Z-Scheme Heterojunction BC/2ZIS/WO3: Mechanistic Investigation and Degradation Pathways. Chem. Eng. J. 2022, 445, 136677. [Google Scholar] [CrossRef]
  288. Yang, Y.; Zhang, C.; Lai, C.; Zeng, G.; Huang, D.; Cheng, M.; Wang, J.; Chen, F.; Zhou, C.; Xiong, W. BiOX (X = Cl, Br, I) Photocatalytic Nanomaterials: Applications for Fuels and Environmental Management. Adv. Colloid. Interface Sci. 2018, 254, 76–93. [Google Scholar] [CrossRef]
  289. Wang, H.; Wang, H.; Wang, Z.; Tang, L.; Zeng, G.; Xu, P.; Chen, M.; Xiong, T.; Zhou, C.; Li, X.; et al. Covalent Organic Framework Photocatalysts: Structures and Applications. Chem. Soc. Rev. 2020, 49, 4135–4165. [Google Scholar] [CrossRef]
  290. He, Z.; Liang, R.; Zhou, C.; Yan, G.; Wu, L. Carbon Quantum Dots (CQDs)/Noble Metal Co-Decorated MIL-53(Fe) as Difunctional Photocatalysts for the Simultaneous Removal of Cr(VI) and Dyes. Sep. Purif. Technol. 2021, 255, 117725. [Google Scholar] [CrossRef]
  291. Shabir, M.; Yasin, M.; Hussain, M.; Shafiq, I.; Akhter, P.; Nizami, A.-S.; Jeon, B.-H.; Park, Y.-K. A Review on Recent Advances in the Treatment of Dye-Polluted Wastewater. J. Ind. Eng. Chem. 2022, 112, 1–19. [Google Scholar] [CrossRef]
  292. Lacasse, K.; Baumann, W. Environmental Considerations for Textile Processes and Chemicals. In Textile Chemicals; Lacasse, K., Baumann, W., Eds.; Springer: Berlin/Heidelberg, Germany, 2004; pp. 484–647. ISBN 978-3-642-18898-5. [Google Scholar]
  293. Siddique, K.; Rizwan, M.; Shahid, M.J.; Ali, S.; Ahmad, R.; Rizvi, H. Textile Wastewater Treatment Options: A Critical Review. In Enhancing Cleanup of Environmental Pollutants; Anjum, N.A., Gill, S.S., Tuteja, N., Eds.; Springer International Publishing: Cham, Switzerland, 2017; Volume 2, pp. 183–207. ISBN 9783319554235. [Google Scholar]
  294. Amuda, O.S.; Deng, A.; Alade, A.O.; Hung, Y.-T. Conversion of Sewage Sludge to Biosolids. In Biosolids Engineering and Management; Wang, L.K., Shammas, N.K., Hung, Y.-T., Eds.; Humana Press: Totowa, NJ, USA, 2008; Volume 7, pp. 65–119. ISBN 978-1-59745-174-1. [Google Scholar]
  295. Yadav, A.; Garg, V.K. Industrial Wastes and Sludges Management by Vermicomposting. Rev. Environ. Sci. Biotechnol. 2011, 10, 243–276. [Google Scholar] [CrossRef]
  296. Karn, S.K.; Kumar, A. Hydrolytic Enzyme Protease in Sludge: Recovery and Its Application. Biotechnol. Bioprocess. Eng. 2015, 20, 652–661. [Google Scholar] [CrossRef]
  297. He, L.; Du, P.; Chen, Y.; Lu, H.; Cheng, X.; Chang, B.; Wang, Z. Advances in Microbial Fuel Cells for Wastewater Treatment. Renew. Sustain. Energy Rev. 2017, 71, 388–403. [Google Scholar] [CrossRef]
  298. Zhang, M.; Yang, C.; Jing, Y.; Li, J. Effect of Energy Grass on Methane Production and Heavy Metal Fractionation during Anaerobic Digestion of Sewage Sludge. Waste Manag. 2016, 58, 316–323. [Google Scholar] [CrossRef] [PubMed]
  299. Yang, S.; Hai, F.I.; Price, W.E.; McDonald, J.; Khan, S.J.; Nghiem, L.D. Occurrence of Trace Organic Contaminants in Wastewater Sludge and Their Removals by Anaerobic Digestion. Bioresour. Technol. 2016, 210, 153–159. [Google Scholar] [CrossRef]
  300. Wang, Q. A Roadmap for Achieving Energy-Positive Sewage Treatment Based on Sludge Treatment Using Free Ammonia. ACS Sustain. Chem. Eng. 2017, 5, 9630–9633. [Google Scholar] [CrossRef]
  301. Sultana, S.; Khan, M.D.; Sabir, S.; Gani, K.M.; Oves, M.; Khan, M.Z. Bio-Electro Degradation of Azo-Dye in a Combined Anaerobic–Aerobic Process along with Energy Recovery. New J. Chem. 2015, 39, 9461–9470. [Google Scholar] [CrossRef]
  302. Đurđević, D.; Blecich, P.; Jurić, Ž. Energy Recovery from Sewage Sludge: The Case Study of Croatia. Energies 2019, 12, 1927. [Google Scholar] [CrossRef]
  303. Patinvoh, R.J.; Osadolor, O.A.; Sárvári Horváth, I.; Taherzadeh, M.J. Cost Effective Dry Anaerobic Digestion in Textile Bioreactors: Experimental and Economic Evaluation. Bioresour. Technol. 2017, 245, 549–559. [Google Scholar] [CrossRef] [PubMed]
  304. Bahar, S.; Ciggin, A.S. A Simple Kinetic Modeling Approach for Aerobic Stabilization of Real Waste Activated Sludge. Chem. Eng. J. 2016, 303, 194–201. [Google Scholar] [CrossRef]
  305. Sonai, G.G.; de Souza, S.M.A.G.U.; de Oliveira, D.; de Souza, A.A.U. The Application of Textile Sludge Adsorbents for the Removal of Reactive Red 2 Dye. J. Environ. Manag. 2016, 168, 149–156. [Google Scholar] [CrossRef]
  306. Vigneswaran, C.; Ananthasubramanian, M.; Kandhavadivu, P.; Vigneswaran, C.; Ananthasubramanian, M.; Kandhavadivu, P. 5–Enzymes in Textile Effluents. In Bioprocessing of Textiles; Vigneswaran, C., Ananthasubramanian, M., Kandhavadivu, P.B.T.-B.T., Eds.; Woodhead Publishing: New Delhi, India, 2014; pp. 251–298. ISBN 9789380308425. [Google Scholar]
  307. Faubert, P.; Barnabé, S.; Bouchard, S.; Côté, R.; Villeneuve, C. Pulp and Paper Mill Sludge Management Practices: What Are the Challenges to Assess the Impacts on Greenhouse Gas Emissions? Resour. Conserv. Recycl. 2016, 108, 107–133. [Google Scholar] [CrossRef]
  308. Volmajer Valh, J.; Majcen Le Marechal, A.; Vajnhandl, S.; Jerič, T.; Šimon, E. 4.20—Water in the Textile Industry. In Treatise on Water Science; Wilderer, P.B.T.-T.W.S., Ed.; Elsevier: Oxford, UK, 2011; Volume 4, pp. 685–706. ISBN 9780444531995. [Google Scholar]
  309. Jahan, N.; Tahmid, M.; Shoronika, A.Z.; Fariha, A.; Roy, H.; Pervez, M.N.; Cai, Y.; Naddeo, V.; Islam, M.S. A Comprehensive Review on the Sustainable Treatment of Textile Wastewater: Zero Liquid Discharge and Resource Recovery Perspectives. Sustainability 2022, 14, 15398. [Google Scholar] [CrossRef]
  310. Van der Bruggen, B.; Canbolat, Ç.B.; Lin, J.; Luis, P. The Potential of Membrane Technology for Treatment of Textile Wastewater; Springer: Singapore, 2017. [Google Scholar]
  311. Mamun Kabir, S.M.; Mahmud, H.; Schӧenberger, H. Recovery of Dyes and Salts from Highly Concentrated (Dye and Salt) Mixed Water Using Nano-Filtration Ceramic Membranes. Heliyon 2022, 8, e11543. [Google Scholar] [CrossRef]
  312. Samuchiwal, S.; Bhattacharya, A.; Malik, A. Treatment of Textile Effluent Using an Anaerobic Reactor Integrated with Activated Carbon and Ultrafiltration Unit (AN-ACF-UF Process) Targeting Salt Recovery and Its Reusability Potential in the Pad-Batch Process. J. Water Process Eng. 2021, 40, 101770. [Google Scholar] [CrossRef]
  313. Yildirim, R.; Eskikaya, O.; Keskinler, B.; Karagunduz, A.; Dizge, N.; Balakrishnan, D. Fabric Dyeing Wastewater Treatment and Salt Recovery Using a Pilot Scale System Consisted of Graphite Electrodes Based on Electrooxidation and Nanofiltration. Environ. Res. 2023, 234, 116283. [Google Scholar] [CrossRef]
  314. Ahmad, N.N.R.; Ang, W.L.; Teow, Y.H.; Mohammad, A.W.; Hilal, N. Nanofiltration Membrane Processes for Water Recycling, Reuse and Product Recovery within Various Industries: A Review. J. Water Process Eng. 2022, 45, 102478. [Google Scholar] [CrossRef]
  315. Majewska-Nowak, K.M. Application of Ceramic Membranes for the Separation of Dye Particles. Desalination 2010, 254, 185–191. [Google Scholar] [CrossRef]
  316. Okano, T.; Sarko, A. Mercerization of Cellulose. II. Alkali–Cellulose Intermediates and a Possible Mercerization Mechanism. J. Appl. Polym. Sci. 1985, 30, 325–332. [Google Scholar] [CrossRef]
  317. Nishimura, H.; Sarko, A. Mercerization of Cellulose. III. Changes in Crystallite Sizes. J. Appl. Polym. Sci. 1987, 33, 855–866. [Google Scholar] [CrossRef]
  318. Jang, Y. Recovery of Caustic Soda in Textile Mercerization by Combined Membrane Filtration. In Technical Proceedings of the 2007 Cleantech Conference and Trade Show; CRC Press: London, UK, 2007. [Google Scholar]
  319. Barredo-Damas, S.; Alcaina-Miranda, M.I.; Bes-Piá, A.; Iborra-Clar, M.I.; Iborra-Clar, A.; Mendoza-Roca, J.A. Ceramic Membrane Behavior in Textile Wastewater Ultrafiltration. Desalination 2010, 250, 623–628. [Google Scholar] [CrossRef]
  320. Tahri, N.; Jedidi, I.; Cerneaux, S.; Cretin, M.; Ben Amar, R. Development of an Asymmetric Carbon Microfiltration Membrane: Application to the Treatment of Industrial Textile Wastewater. Sep. Purif. Technol. 2013, 118, 179–187. [Google Scholar] [CrossRef]
  321. Yadav, D.; Karki, S.; Ingole, P.G. Current Advances and Opportunities in the Development of Nanofiltration (NF) Membranes in the Area of Wastewater Treatment, Water Desalination, Biotechnological and Pharmaceutical Applications. J. Environ. Chem. Eng. 2022, 10, 108109. [Google Scholar] [CrossRef]
  322. Weber, R.; Chmiel, H.; Mavrov, V. Characteristics and Application of New Ceramic Nanofiltration Membranes. Desalination 2003, 157, 113–125. [Google Scholar] [CrossRef]
  323. Jing, G.; Ren, S.; Gao, Y.; Sun, W.; Gao, Z. Electrocoagulation: A Promising Method to Treat and Reuse Mineral Processing Wastewater with High COD. Water 2020, 12, 595. [Google Scholar] [CrossRef]
  324. Khatri, A.; White, M. Sustainable Dyeing Technologies. In Sustainable Apparel: Production, Processing and Recycling; Blackburn, R., Ed.; Woodhead Publishing Series in Textiles; Woodhead Publishing: Sawston, UK, 2015; pp. 135–160. ISBN 9781782423577. [Google Scholar]
  325. Periyasamy, A.P. Natural Dyeing of Cellulose Fibers Using Syzygium Cumini Fruit Extracts and a Bio-Mordant: A Step toward Sustainable Dyeing. Sustain. Mater. Technol. 2022, 33, e00472. [Google Scholar] [CrossRef]
  326. Yasmeen, M.; Nawaz, M.S.; Khan, S.J.; Ghaffour, N.; Khan, M.Z. Recovering and Reuse of Textile Dyes from Dyebath Effluent Using Surfactant Driven Forward Osmosis to Achieve Zero Hazardous Chemical Discharge. Water Res. 2023, 230, 119524. [Google Scholar] [CrossRef]
  327. Andooz, A.; Eqbalpour, M.; Kowsari, E.; Ramakrishna, S.; Ansari Cheshmeh, Z. A Comprehensive Review on Pyrolysis from the Circular Economy Point of View and Its Environmental and Social Effects. J. Clean. Prod. 2023, 388, 136021. [Google Scholar] [CrossRef]
  328. Periyasamy, A.P.; Militky, J. Denim and Consumers’ Phase of Life Cycle. In Sustainability in Denim; Muthu, S.S., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 257–282. [Google Scholar]
  329. Periyasamy, A.P.; Wiener, J.; Militky, J. Life-Cycle Assessment of Denim. In Sustainability in Denim; Elsevier: Amsterdam, The Netherlands, 2017; pp. 83–110. [Google Scholar]
  330. Periyasamy, A.P.; Duraisamy, G. Carbon Footprint on Denim Manufacturing. In Handbook of Ecomaterials; Martínez, L.M.T., Kharissova, O.V., Kharisov, B.I., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 1–18. ISBN 978-3-319-48281-1. [Google Scholar]
  331. Periyasamy, A.P. Life Cycle Assessment of Denim: A Review. Colourage 2019, 66, 54–62. [Google Scholar]
  332. Periyasamy, A.P.; Militky, J. Sustainability in Regenerated Textile Fibers. In Sustainability in the Textile and Apparel Industries; Springer: Cham, Switzerland, 2020; pp. 63–95. [Google Scholar]
  333. Venkatesan, H.; Periyasamy, A.P. Eco-Fibers in the Textile Industry. In Handbook of Ecomaterials; Springer International Publishing: Cham, Switzerland, 2019; Volume 3, pp. 1413–1433. ISBN 9783319682556. [Google Scholar]
  334. Periyasamy, A.P.; Venkatesan, H. Eco-Materials in Textile Finishing. In Handbook of Ecomaterials; Springer International Publishing: Cham, Switzerland, 2019; Volume 3, pp. 1461–1482. ISBN 9783319682556. [Google Scholar]
  335. Periyasamy, A.P.; Vikova, M.; Vik, M. A Review of Photochromism in Textiles and Its Measurement. Text. Prog. 2017, 49, 53–136. [Google Scholar] [CrossRef]
  336. Vikova, M.; Sakurai, S.; Periyasamy, A.P.; Yasunaga, H.; Pechočiaková, M.; Ujhelyiová, A. Differential Scanning Calorimetry/Small-Angle X-Ray Scattering Analysis of Ultraviolet Sensible Polypropylene Filaments. Text. Res. J. 2021, 92, 3142–3153. [Google Scholar] [CrossRef]
  337. Periyasamy, A.P.; Vikova, M.; Vik, M. Photochromic Polypropylene Filaments: Impacts of Mechanical Properties on Kinetic Behaviour. Fibres Text. East. Eur. 2019, 27, 19–25. [Google Scholar] [CrossRef]
  338. Viková, M.; Periyasamy, A.P.; Vik, M.; Ujhelyiová, A. Effect of Drawing Ratio on Difference in Optical Density and Mechanical Properties of Mass Colored Photochromic Polypropylene Filaments. J. Text. Inst. 2017, 108, 1365–1370. [Google Scholar] [CrossRef]
  339. Zhang, L.; Sun, W.; Xu, D.; Li, M.; Agbo, C.; Fu, S. Dope Dyeing of Lyocell Fiber with NMMO-Based Carbon Black Dispersion. Carbohydr. Polym. 2017, 174, 32–38. [Google Scholar] [CrossRef] [PubMed]
  340. An, S.-Y.; Min, S.-K.; Cha, I.-H.; Choi, Y.-L.; Cho, Y.-S.; Kim, C.-H.; Lee, Y.-C. Decolorization of Triphenylmethane and Azo Dyes by Citrobacter Sp. Biotechnol. Lett. 2002, 24, 1037–1040. [Google Scholar] [CrossRef]
  341. Das, A.; Dey, A. P-Nitrophenol -Bioremediation Using Potent Pseudomonas Strain from the Textile Dye Industry Effluent. J. Environ. Chem. Eng. 2020, 8, 103830. [Google Scholar] [CrossRef]
  342. Walker, G.M.; Weatherley, L.R. Prediction of Bisolute Dye Adsorption Isotherms on Activated Carbon. Process Saf. Environ. Prot. 2000, 78, 219–223. [Google Scholar] [CrossRef]
  343. Ball, A.S.; Colton, J. Decolorisation of the Polymeric Dye Poly R by Streptomyces Viridosporus T7A. J. Basic Microbiol. 1996, 36, 13–18. [Google Scholar] [CrossRef]
  344. Vijayaraghavan, K.; Yun, Y.S. Utilization of Fermentation Waste (Corynebacterium glutamicum) for Biosorption of Reactive Black 5 from Aqueous Solution. J. Hazard. Mater. 2007, 141, 45–52. [Google Scholar] [CrossRef] [PubMed]
  345. Dionysiou, D.D.; Balasubramanian, G.; Suidan, M.T.; Khodadoust, A.P.; Baudin, I.; Laîné, J.M. Rotating Disk Photocatalytic Reactor: Development, Characterization, and Evaluation for the Destruction of Organic Pollutants in Water. Water Res. 2000, 34, 2927–2940. [Google Scholar] [CrossRef]
  346. Choi, H.; Sofranko, A.C.; Dionysiou, D.D. Nanocrystalline TiO2 Photocatalytic Membranes with a Hierarchical Mesoporous Multilayer Structure: Synthesis, Characterization, and Multifunction. Adv. Funct. Mater. 2006, 16, 1067–1074. [Google Scholar] [CrossRef]
  347. Yu, R.F.; Lin, C.H.; Chen, H.W.; Cheng, W.P.; Kao, M.C. Possible Control Approaches of the Electro-Fenton Process for Textile Wastewater Treatment Using on-Line Monitoring of DO and ORP. Chem. Eng. J. 2013, 218, 341–349. [Google Scholar] [CrossRef]
  348. Rosales, E.; Pazos, M.; Longo, M.A.; Sanromán, M.A. Electro-Fenton Decoloration of Dyes in a Continuous Reactor: A Promising Technology in Colored Wastewater Treatment. Chem. Eng. J. 2009, 155, 62–67. [Google Scholar] [CrossRef]
  349. Liu, Y.; Zhang, Y.; Quan, X.; Zhang, J.; Zhao, H.; Chen, S. Effects of an Electric Field and Zero Valent Iron on Anaerobic Treatment of Azo Dye Wastewater and Microbial Community Structures. Bioresour. Technol. 2011, 102, 2578–2584. [Google Scholar] [CrossRef]
  350. Tao, Y.; Gao, D.W.; Fu, Y.; Wu, W.M.; Ren, N.Q. Impact of Reactor Configuration on Anammox Process Start-up: MBR versus SBR. Bioresour. Technol. 2012, 104, 73–80. [Google Scholar] [CrossRef]
  351. Durán, A.; Monteagudo, J.M.; Sanmartín, I.; Gómez, P. Homogeneous Sonophotolysis of Food Processing Industry Wastewater: Study of Synergistic Effects, Mineralization and Toxicity Removal. Ultrason. Sonochem 2013, 20, 785–791. [Google Scholar] [CrossRef]
Figure 1. Classification of dyes used in the textile industry.
Figure 1. Classification of dyes used in the textile industry.
Sustainability 16 00495 g001
Figure 2. Textile effluent from industry (a) and in the water streams (b) that flow in the Noyal River (ce) in India, and the influence of these effluents on the agricultural land (fh) (pictures taken in April 2022 and reprinted with the permission of ACS [26] licensed under CC-BY 4.0).
Figure 2. Textile effluent from industry (a) and in the water streams (b) that flow in the Noyal River (ce) in India, and the influence of these effluents on the agricultural land (fh) (pictures taken in April 2022 and reprinted with the permission of ACS [26] licensed under CC-BY 4.0).
Sustainability 16 00495 g002
Figure 3. Influence of dye-containing textile wastewater on the environment and health hazardous.
Figure 3. Influence of dye-containing textile wastewater on the environment and health hazardous.
Sustainability 16 00495 g003
Figure 4. Negative impacts of textile dyes containing wastewater and their influence on aquatic life and human health.
Figure 4. Negative impacts of textile dyes containing wastewater and their influence on aquatic life and human health.
Sustainability 16 00495 g004
Figure 6. Treatment of textile wastewater by biological and biosorption methods.
Figure 6. Treatment of textile wastewater by biological and biosorption methods.
Sustainability 16 00495 g006
Figure 7. Site-directed mutagenesis of laccase from B. amyloliquefaciens for improving indigo carmine decolorization (a) (reproduced from [220] as distributed by Creative Commons Attribution License). Azoreductase on azo dyes with the formation of colorless amines (b) (reprinted from [221]).
Figure 7. Site-directed mutagenesis of laccase from B. amyloliquefaciens for improving indigo carmine decolorization (a) (reproduced from [220] as distributed by Creative Commons Attribution License). Azoreductase on azo dyes with the formation of colorless amines (b) (reprinted from [221]).
Sustainability 16 00495 g007
Figure 8. Classification, advantages, and disadvantages of membrane technology.
Figure 8. Classification, advantages, and disadvantages of membrane technology.
Sustainability 16 00495 g008
Figure 9. Process flow of conventional treatment and combination of MBR-NF process (reprinted from [240] with the kind permission of Elsevier).
Figure 9. Process flow of conventional treatment and combination of MBR-NF process (reprinted from [240] with the kind permission of Elsevier).
Sustainability 16 00495 g009
Figure 11. Treatment of dyes using hybrid technologies.
Figure 11. Treatment of dyes using hybrid technologies.
Sustainability 16 00495 g011
Figure 12. Schematic illustration of semiconductor heterojunction (a) before contact; (b) after contact; and (c) hybrid Z-scheme of photogenerated charge carriers (reprinted from [291]).
Figure 12. Schematic illustration of semiconductor heterojunction (a) before contact; (b) after contact; and (c) hybrid Z-scheme of photogenerated charge carriers (reprinted from [291]).
Sustainability 16 00495 g012
Figure 13. The sequence process for energy conversion from the sludge.
Figure 13. The sequence process for energy conversion from the sludge.
Sustainability 16 00495 g013
Figure 14. Benefits of the ZLD concept.
Figure 14. Benefits of the ZLD concept.
Sustainability 16 00495 g014
Figure 15. Life cycle assessment (LCA) framework for the wastewater treatment plants.
Figure 15. Life cycle assessment (LCA) framework for the wastewater treatment plants.
Sustainability 16 00495 g015
Table 1. Advantages and disadvantages of the most common techniques for removal of dyes from wastewater.
Table 1. Advantages and disadvantages of the most common techniques for removal of dyes from wastewater.
TypeTreatment MethodAdvantagesDisadvantagesRef
ChemicalOxidationRapid and effective for both organics and inorganics
Can be used for both soluble and insoluble dyes
No need to use microorganisms
Formation of by products
High energy consumption and running costs
[33,34,35]
OzonationNo alternation in the sample volumeShort half-life and need pretreatment[33,36,37,38,39,40,41,42]
Chemical precipitationLow investment and simple processHigh maintenance and required to dispose the sludge[43,44]
Electro kinetic coagulationEconomic processHigh sludge generation[45]
Electrochemical treatmentModerate metal selectivity
Rapid breakdown
Formation of by products
Require high energy
[46]
Advanced oxidation with Fenton reagents as catalystNo energy input required
Effective for both insoluble and soluble dyes, for wide variety of wastes treatment
Sludge formation
Expensive process
[47,48,49,50,51,52,53]
NaOCLAccelerated azo bond cleavage Toxic aromatic amine release [53]
Photochemical degradation (based on catalyst)Effective oxidation and lab scale applicability
No sludge generation
Formation of by products
Excessive dissolved O2 is required
[37]
Coagulation-Flocculation/SedimentationVariety of coagulants-flocculantsExpensive chemicals and no recycling[54,55]
Biological methodSingle cell organisms such as bacteria, fungi, algae and yeastsGenerally, these are more economical than chemical and physical methods.
For any dye industry and as a preparatory step for removal
Acceptable efficiency for low concentrations and volumes
Highly effective for specific dye species
Requires large land area, less flexible in operation and design and partially to totally non-degrading to dyes[56,57,58]
Aerobic (presence of free DO)Facile
COD removal
Longer detention times[59,60,61]
Anaerobic (absence of DO)Resistance to wide variety of dyes
Steam generation via the produced biogas
Longer acclimatization phase[60,61]
PhysicalMembrane (ultrafiltration, microfiltration, nanofiltration, reverse osmosis)Removes all types of dyesInapplicable for wastewater treatment due to the large pore size[62,63,64,65]
AdsorptionFor all dye industry
Regeneration of adsorbent with low loss
Only soluble dyes
High energy consumption
[66]
Ion exchange For specific applications [67,68]
IrradiationWide range of colorants
Efficient even for low volumes
High dissolved oxygen requirement
Light-resistant colorants cannot be degraded
[69]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Periyasamy, A.P. Recent Advances in the Remediation of Textile-Dye-Containing Wastewater: Prioritizing Human Health and Sustainable Wastewater Treatment. Sustainability 2024, 16, 495. https://doi.org/10.3390/su16020495

AMA Style

Periyasamy AP. Recent Advances in the Remediation of Textile-Dye-Containing Wastewater: Prioritizing Human Health and Sustainable Wastewater Treatment. Sustainability. 2024; 16(2):495. https://doi.org/10.3390/su16020495

Chicago/Turabian Style

Periyasamy, Aravin Prince. 2024. "Recent Advances in the Remediation of Textile-Dye-Containing Wastewater: Prioritizing Human Health and Sustainable Wastewater Treatment" Sustainability 16, no. 2: 495. https://doi.org/10.3390/su16020495

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop