Next Article in Journal
A Comparative Study of YOLO, SSD, Faster R-CNN, and More for Optimized Eye-Gaze Writing
Previous Article in Journal
Effects of Physicochemical Characteristics of Two Soils on Agro-Morphological Traits of Two Chickpea Varieties (Cicer arietinum L.)
Previous Article in Special Issue
A Review of the State of the Art for the Internet of Medical Things
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Microbial Enzyme Applications for Sustainable Textile Processing and Waste Management

by
Mohd Faheem Khan
UCD School of Agriculture and Food Science, University College Dublin, Belfield, Dublin 4, D04 V1W8 Dublin, Ireland
Submission received: 2 January 2025 / Revised: 26 January 2025 / Accepted: 1 April 2025 / Published: 9 April 2025
(This article belongs to the Special Issue Feature Papers—Multidisciplinary Sciences 2024)

Abstract

:
Microbial enzymes have revolutionised the textile industry by replacing harmful chemicals with eco-friendly alternatives, enhancing processes such as desizing, scouring, dyeing, finishing, and promoting water conservation while reducing pollution. This review explores the role of enzymes like amylases, pectinases, cellulases, catalases, laccases, and peroxidases in sustainable textile processing, focusing on their ability to mitigate environmental pollution from textile effluents. The review also examines the types and characteristics of hazardous textile waste and evaluates traditional waste treatment methods, highlighting sustainable alternatives such as microbial enzyme treatments for effluent treatment. Recent advancements in recombinant enzyme technology, including enzyme engineering and immobilisation techniques to enhance stability, reusability, and catalytic performance, are also explored. Additionally, the potential of extremozymes in textile processing and effluent treatment is explored, emphasising their stability under harsh industrial conditions. Strategies for reducing textile waste through enzyme-based processes are presented, focusing on principles of the circular economy. The review also addresses challenges such as scalability, cost, and process optimisation, while proposing potential solutions and outlining future directions for the widespread adoption of microbial enzymes in sustainable textile production and waste management. This review underscores the transformative potential of microbial enzymes in achieving greener textile manufacturing practices.

1. Introduction

Enzymes are biological macromolecules that catalyse reactions with high specificity and efficiency, making them essential in numerous industrial applications. Microbial enzymes are recognised as superior biocatalysts due to their scalability, cost-effectiveness, and alignment with sustainability goals [1]. These enzymes are widely utilised in fields such as pharmaceuticals, nutraceuticals, diagnostics, detergents, cosmetics, and textile processing [2].
In the textile industry, enzyme-based biotechnology has revolutionised traditional processes by replacing hazardous chemicals with eco-friendly solutions. Enzymes play an integral role in textile manufacturing, facilitating processes such as desizing, scouring, bleaching, biopolishing, and dyeing, while offering advantages including reduced environmental impact, water conservation, and compatibility with mild processing conditions [3,4]. Key microbial sources, such as Aspergillus, Trichoderma, Streptomyces, and Bacillus, serve as platforms for enzyme production. Enzymes like amylase, cellulase, catalase, protease, pectinase, laccase, and lipase are widely employed due to their biodegradability, specificity, and efficacy in enhancing textile quality [5]. These biocatalysts significantly lower energy, water, and time requirements, establishing enzymatic strategies as a scientifically robust and sustainable alternative to conventional chemical treatments in the textile industry [3]. Despite these benefits, widespread adoption is limited by high production costs and the complexity of enzyme isolation and purification [6].
The textile industry remains a significant source of environmental pollution, with effluents containing hazardous chemicals such as volatile organic compounds, heavy metals, dyes, and ionic liquids posing severe risks to human health and ecosystems [7]. Enzymatic treatment of these effluents has emerged as a green technology, with enzymes like laccases and peroxidases effectively degrading toxic pollutants and reducing environmental impact [8]. Bernal et al. explored filamentous fungi from textile wastewater for enzyme production and dye discolouration. They identified Aspergillus sydowii (ITF 30) as the most effective producer of cellulase, amylase, lipase, and laccase. This strain also exhibited significant discolouration of Remazol Brilliant Blue R (RBBR). Another strain, A. sydowii (ITF 27), demonstrated notable dye discolouration capacity [9]. These findings highlight the potential of these enzymes for bioremediation in textile industrial effluents.
While enzymatic textile processing and waste treatment have been extensively studied, this review provides a distinct perspective by focusing on advancements in enzyme technology, including enzyme engineering and immobilisation, to enhance stability, reusability, and catalytic performance. The discussion emphasises their robustness under harsh industrial conditions and explores strategies to overcome scalability and cost barriers. Furthermore, this review highlights enzyme-based processes aimed at reducing textile waste and emphasises the transformative potential of microbial enzymes in achieving greener and more sustainable textile manufacturing practices. By integrating recent innovations and addressing critical challenges, this review highlights pathways for successful enzyme application in eco-friendly textile production.

2. Textile Industrial Processes

Textile processing converts raw materials like cotton, silk, wool, jute, polyester, and nylon into finished products. Natural fibres are sourced from plants and animals, while synthetic fibres are made from petrochemicals [10]. Natural fibres are biodegradable and eco-friendly, whereas synthetic fibres are durable and versatile but contribute to pollution and health risks due to their chemical components.
The textile production process includes five main stages: yarn formation, fabric formation, wet processing, fabrication, and final product development, as depicted in Figure 1. Wet processing is a critical stage that generates most of the wastewater, involving the use of water, dyes, and chemicals, leading to significant environmental concerns [4].

2.1. Yarn Formation

Yarn is a continuous strand of fibres, either natural or synthetic, used to create fabric through knitting or weaving. Fibre preparation removes impurities to improve absorption for dyeing, printing, and finishing. Synthetic fibres generally lack these natural impurities. The main techniques in yarn formation include texturising and spinning. Texturising, mainly for synthetic fibres, adds bulk and elasticity through processes like crimping and heat-setting. Spinning, which twists fibres together, forms strong yarns using methods such as ring, rotor, and air-jet spinning, depending on the desired yarn quality and production speed.

2.2. Fabric Formation

Fabric formation includes processes like warping, slashing, weaving, knitting, and sizing. Warping aligns yarns for uniform tension, while slashing applies a coating to strengthen yarns for weaving. Weaving interlaces yarns to create different fabric types, and knitting loops yarns to form flexible, elastic fabrics. Sizing coats yarns to improve strength and smoothness, enhancing weaving efficiency and reducing yarn breakage. These processes together improve fabric strength, texture, and performance [11].

2.3. Wet Processing

Wet processing involves chemical treatments to improve textile quality [12,13]. Key steps include:
  • Desizing: Removes sizing from fabric to improve absorption for dyeing and printing, often using eco-friendly enzymatic methods;
  • Scouring: Removes impurities like oils and waxes, making fabric more absorbent;
  • Bleaching: Whitens fabric by removing natural colours with chemicals like hydrogen peroxide (H2O2), sodium chlorite (NaClO2), or sodium hypochlorite (NaOCl), but excessive bleaching can weaken fibres;
  • Mercerisation: Treats fabric with sodium hydroxide (NaOH) to improve strength, lustre, and dye uptake;
  • Dyeing and Printing: Colour fabrics, with dyeing applying colour evenly and printing for specific designs. Dyeing uses more water than printing. Table 1 lists the dyes used for various synthetic and natural fabrics;
  • Stone Washing: Fades colour in denim and canvas, often replaced by eco-friendly enzymatic biostoning;
  • Polishing: Enhances fabric texture and brightness, often using cellulase enzymes for biopolishing.

2.4. Fabrication

Garment fabrication involves stages like designing, measuring, cutting fabrics, sewing, and final packing. Advanced technologies such as automated cutting machines and computer-aided design (CAD) are used to improve efficiency and precision in these processes.

2.5. Final Product

The final textile products serve various purposes, including clothing, household items, and technical textiles. Effective marketing strategies help connect manufacturers, retailers, and consumers, ensuring that the products meet fashion trends and consumer needs.

3. Textile Effluents

The textile industry is a major source of water pollution, primarily due to the heavy use of synthetic dyes and chemicals in its manufacturing processes [3,4,5]. The discharge of textile effluents containing toxic dyes and chemicals damages ecosystems and poses significant health risks to humans, aquatic life, and environmental impacts (Table 2).
The textile industry is a significant consumer of water and chemicals, with an average plant using approximately 1.6 million litres of water daily to produce 8000 kg of fabric [4]. The dyeing and finishing processes contribute to 17–20% of the industry’s water pollution [14]. Textile wastewater is burdened with a range of toxic chemicals, including azo dyes like Remazol Brilliant Blue R, which are carcinogenic, cause skin irritation, and are persistent in ecosystems. Other common pollutants include formaldehyde, used in size preservatives and coating processes, which causes skin and eye irritation, and copper, a heavy metal found in textile effluents that can cause respiratory problems [15]. Additionally, per- and polyfluoroalkyl substances (PFASs), including perfluorooctanoic acid (PFOA), used in textile finishing for its water- and stain-repellent properties, are highly persistent and bioaccumulative, posing serious risks to human health and the environment [16,17]. These pollutants are stable in the environment and resistant to conventional treatment methods, emphasising the need for more effective and sustainable treatment technologies. Other toxic substances like tributyltin oxide, sodium hypochlorite, and nonylphenol ethoxylates are commonly found in textile wastewater. These substances pose significant risks to both human health and the environment, with many being difficult to degrade using conventional wastewater treatment methods [17]. This highlights the urgent need for more effective and sustainable treatment solutions to address the growing environmental challenges in the textile industry.

3.1. Microbial Remediation of Textile Effluents

Microbial remediation is a process that uses microorganisms—such as bacteria, fungi, and algae—to degrade or detoxify pollutants in the environment, converting harmful substances into less toxic or non-toxic forms [18]. This approach offers an environmentally friendly and cost-effective alternative to traditional methods of waste treatment, which can often be inefficient, costly, and environmentally harmful [19]. Microorganisms, including bacteria, fungi, and algae, possess natural abilities to degrade various pollutants, including textile dyes [20,21,22]. Table 3 compares the efficiency, advantages, limitations, and applications of fungi, bacteria, and algae in the bioremediation of textile effluents.
These microorganisms can be harnessed in bioreactors or other bioremediation systems to treat textile effluents effectively. The microbial remediation process typically involves the breakdown of chromophoric groups, which are the coloured part of the dye molecules, followed by the mineralisation of the dye into non-toxic by-products [20]. The use of microbial remediation techniques in textile wastewater treatment minimises the need for harmful chemicals, reduces pollution, and promotes sustainability in the textile industry.

3.1.1. Fungal Bioremediation

Fungi are particularly effective in the biodegradation of textile dyes due to their ability to produce extracellular enzymes capable of breaking down complex organic molecules. Species like Phanerochaete chrysosporium, Trichoderma harzianum, Aspergillus oryza, and Cunninghamella elegans have been extensively studied for their potential in decolourising textile effluents [27]. Phanerochaete chrysosporium, for example, produces enzymes like lignin peroxidase, manganese peroxidase, and laccase that can degrade a wide variety of dyes, including polycyclic aromatic hydrocarbons found in wastewater from textile and pulp industries [28,29]. Similarly, Trichoderma harzianum has shown remarkable efficiency in degrading dyes like Congo red and Bromophenol blue, with studies highlighting the role of fungal biomass in absorbing and breaking down toxic dye molecules [30]. Corso et al. demonstrated that Aspergillus oryzae in its pellet form effectively biosorbs and reduces the toxicity of azo dyes like Direct Red 23 and Direct Violet 51, offering a sustainable solution for industrial wastewater treatment [31]. Non-lignin-degrading fungi like Cunnighamella elegans degrade dyes such as malachite green via cytochrome P450 activity and can simultaneously remove hexavalent chromium, achieving 95% removal of dye and metal over 19 cycles in a semi-continuous process [32]. The combination of aerobic and anaerobic microbial treatment has also proven effective in enhancing the degradation of dyes, as seen in the case of Aspergillus strains, which can decolourise dyes like orange 3R under specific conditions [33]. Fungal bioremediation is advantageous due to the fast growth, high biomass production, and wide hyphal networks of many fungal species.

3.1.2. Bacterial Bioremediation

Bacteria are also powerful agents in bioremediation, particularly in the degradation of azo dyes, which are commonly used in the textile industry. Azo dyes contain the characteristic azo group (-N=N-), which can be biodegraded by bacteria through the enzymatic activity of azoreductases, leading to a reduction in azo bonds under anaerobic conditions and the subsequent formation of colourless aromatic amines [34]. These metabolites are further degraded aerobically or anaerobically, making bacterial bioremediation an efficient method for decolourising and detoxifying textile effluents. The degradation of azo dye is often facilitated by specific bacterial strains capable of breaking the azo bonds, such as Klebsiella pneumoniae GM-04, which demonstrated a 95% colour removal efficiency for Disperse Blue 284 (DB 284) dye under anaerobic conditions at pH 7, 37 °C, and an initial dye concentration of 200 mg/L after 24 h [35]. Similarly, Acinetobacter baumannii effectively decolourised 98.8% of Reactive Black 5, 96% of Reactive Red 120, and 96.2% of Reactive Blue 19 under optimal conditions of pH 7, temperature 37 °C, 500 ppm dye concentration, 48 h incubation, and anaerobic conditions [36]. Kishor et al. demonstrated that the lignocellulosic enzyme-producing bacterium Bacillus albus MW407057 effectively biodegraded methylene blue dye, achieving 93.8% colour removal, 86.02% reduction in BOD, and 77.35% reduction in COD within 48 h [37]. Acidithiobacillus thiooxidans removes up to 91.4% of Sulphur Blue 15 dye via biosorption and further biodegrades 50% of the remaining dye under optimised conditions [38]. Pseudomonas stutzeri SPM-1 decolourised Procion Red-H3B dye (50 mg/L) in 20 h under microaerophilic conditions, reducing BOD, COD, and TOC by 85%, 90%, and 65%, respectively. The strain showed significant activities of azoreductase (95%), laccase (76%), and NADH-DCIP reductase (88%), with GC-MS/FT-IR analyses confirming its effective dye degradation, highlighting its potential for textile wastewater bioremediation [39]. Mixed bacterial cultures such as Providencia rettgeri strain HSL1 and Pseudomonas sp. SUK1 achieved 98–99% decolourisation and 62–72% TOC reduction in various azo dyes (Disperse Red 78, Direct Red 81, Reactive Black 5, Reactive Orange 16) under sequential aerobic and microaerophilic or vice versa at 30 °C for 12–30 h, demonstrating their effectiveness for treating complex textile industry effluents [40]. This highlights the advantages of using mixed cultures over isolated single-bacterial strains. Table 4 lists the degradation of textile dyes by various isolated and mixed cultures.
Bacterial degradation offers several advantages, including growing rapidly and adapting to different environmental conditions. Additionally, bacteria can be genetically engineered to enhance their dye-degrading capabilities, improving the efficiency of the bioremediation process [34].

3.1.3. Algal Bioremediation

Algae are known for their extensive surface area and can effectively remove dyes from water by adsorbing pollutants onto their biomass or metabolising dye chromophores into non-toxic products [54]. Alvarez et al. identified three mechanisms underlying algal dye decolourisation: (1) utilisation of chromophores for algal biomass production, releasing CO2 and H2O; (2) conversion of chromophores into non-chromophoric materials; and (3) adsorption of residual chromophores onto algal biomass [55]. For instance, Devi et al. demonstrated the effectiveness of freshwater algae (Spirulina platensis) and seaweed (Gracilaria edulis) in removing Reactive Blue 19 (RB 19) using batch adsorption techniques [56]. Both algae and cyanobacteria exhibit promising potential in textile wastewater treatment due to their ability to degrade recalcitrant azo dyes and other pollutants. While algae like Chlorella pyrenoidosa utilise intracellular laccase enzymes for effective dye degradation [57], cyanobacteria such as Phormidium autumnale UTEX1580 demonstrate complete breakdown of dyes like indigo, producing non-toxic metabolites [58]. Together, these microorganisms offer eco-friendly solutions for the decolourisation and detoxification of industrial effluents.
In conclusion, microbial remediation offers an eco-friendly, cost-effective, and sustainable alternative to conventional methods for treating textile wastewater. With the ability to break down complex and persistent pollutants, microorganisms provide an efficient solution for long-term environmental management. Advancing research in this field underscores its potential to reduce the environmental impact of the textile industry and support sustainability.

4. Enzymatic Innovations in Textile Processing and Effluent Treatment

The textile industry is a significant contributor to global pollution, making the adoption of green alternatives crucial for sustainable practices. The use of enzymes or non-toxic chemical substitutes in textile processing represents a transformative approach, replacing hazardous chemicals and reducing associated health risks [10,59]. Among the various stages of textile production, wet processing and fabric finishing are particularly harsh and benefit greatly from enzymatic interventions. Enzymes such as amylases for desizing, pectinases for bioscouring, cellulase for stonewashing and biopolishing, and catalase for bleaching and bleach clean-up are integral to greener processing methods [5]. These enzymatic treatments minimise environmental impact and enhance the final product’s quality and sustainability [59]. Table 5 outlines green chemical substitutes and enzyme-aided steps as viable substitutes for harmful chemicals in textile processing. Green alternatives to harmful chemicals in the textile industry offer sustainable solutions across processing stages. In sizing, polyvinyl alcohol is replaced with potato starch or carboxymethylcellulose [60]. Desizing substitutes mineral acids with enzymes like amylase and xylanase, while scouring uses pectinase and xylanase instead of sodium hydroxide [12,61]. Stonewashing and polishing shift from nonylphenol–ethylene oxide adducts to fatty alcohol–ethylene oxide adducts or alkyl polyglycosides, with cellulase as an enzymatic alternative [62]. Bleaching avoids chlorine-based chemicals by using hydrogen peroxide, ozone, or enzymes like glucose–oxidase, laccase, and ligninase [63,64]. Catalase replaces thiosulfates for bleach clean-up [65]. Dyeing and printing eliminate harmful carriers like kerosene and formaldehyde, using water-based thickeners, polycarboxylic acids, and enzymatic options like ligninase [5]. For finishing and effluent treatment, silicone softeners and formaldehyde resins are replaced with natural oils, formaldehyde-free resins, and eco-friendly coagulants, supported by enzymes such as proteases, lipases, and laccase [3,64].
The enzymatic alternatives employed at various stages of processing cotton, silk, wool, and synthetic fibres are illustrated in Figure 2. In cotton processing, cellulases enhance fabric smoothness through biopolishing, while amylases remove starch-based sizing agents [12,62]. Protease enzymes are essential in silk degumming, preserving strength and sheen, while in wool processing, they soften fibres and reduce prickle. Lipases are employed to eliminate grease and oils. For synthetic fibres, enzymes break down lubricants, improve dye uptake, and enhance fabric properties [66]. This enzymatic approach not only improves product quality but also reduces energy, water usage, and pollution, supporting sustainable practices in textile manufacturing [10].

4.1. Amylases in Desizing

Amylase is a key biocatalyst that catalyses the hydrolysis of starch into sugars. Bacteria, fungi, and plants are the major producers of amylases. Additionally, amylases are present in the saliva of humans and other mammals, where they initiate the digestion process [67]. All amylases belong to the hydrolase class of enzymes and are broadly classified as glycosidases, as they cleave the α-1,4-glycosidic bonds in starch, particularly in amylose. There are three types of amylases: α-amylases (which act randomly on starch and require Ca2+ for activity), β-amylases (which act from the non-reducing end and also depend on Ca2+), and γ-amylases (which cleave both α-1,4 and α-1,6-glycosidic bonds from the non-reducing end). Among these, α-amylase is the fastest-acting and most widely used type [68]. Due to its industrial potential, amylases are widely employed in industries such as textiles, food, pharmaceuticals, and detergents [69].
In textile processing, fabrics are often coated with starch-based sizing materials to strengthen them during weaving. This sizing must be removed before subsequent processes like dyeing, printing, and polishing. Enzymatic desizing using amylases is the most commonly employed method, producing short-chain sugars such as dextrin, maltose, and glucose as byproducts of starch breakdown. This method improves fabric quality and uniform wettability, essential for further processing. Rehman et al. demonstrated the effective and eco-friendly desizing of textiles using amylase produced by an indigenous Bacillus cereus strain, optimising desizing conditions to achieve complete starch removal within 15 h at 60 °C with minimal fabric damage [70]. Singh et al. identified Aspergillus fumigatus NTCC1222 as an efficient producer of α-type amylase (341.7 U/mL) under optimised solid-state fermentation conditions using wheat bran, with promising potential for cost-effective textile applications [71]. The enzyme α-amylase was evaluated for its desizing efficiency on starch-sized cotton fabrics under the influence of ultrasound. Although ultrasound reduced enzyme activity on soluble starch, it enhanced desizing efficiency through intensified hydrophobic interactions, with simultaneous ultrasound application yielding superior results by improving mixing, substrate modification, catalytic activity, and product removal [72]. Metal-independent α-amylase from Bacillus sp., exhibited optimal activity at 75–80 °C and pH 4.0–6.0, with its activity and thermal stability unaffected by the presence of Ca2+, EDTA, or calcium absence, demonstrating its robustness for industrial applications [73]. The industrial-level production of α-amylase using Bacillus sp. KR-8104 was optimised for simultaneous enzyme production and desizing of cotton fabric. The highest enzyme yield was achieved at 45 °C, pH 5.0–6.0, using media with novel nitrogen sources, such as soya powder and chicken excrement, while desizing efficiency peaked at 40–45 °C, pH 5.0–6.0, over 10–24 h [74]. Similarly, the large-scale production of a highly thermostable α-amylase from Thermotoga petrophila, cloned into E. coli, was optimised for industrial applications as a textile desizer. Maximum enzyme yield (22.08 U mL−1 min−1) was achieved in ZBM medium with lactose (200 mM) as an inducer under specific fermenter conditions, including 70% medium volume, 2.0 vvm air supply, 20% dissolved oxygen, and 200 rpm agitation. The enzyme demonstrated optimal desizing activity at 85 °C, pH 6.5, and 150 U mL−1 concentration, showcasing its potential for efficient textile processing [75]. These fermentative desizing processes, employing inexpensive raw materials, offer a promising, cost-effective, and environmentally sustainable solution for the textile industry.

4.2. Pectinases in Scouring and Degumming

The primary non-cellulosic impurities in fabrics, particularly cotton, consist of pectins, which bind cellulosic and non-cellulosic polymers. The decomposition of pectins is critical for the removal of non-cellulosic substances from fabrics. Traditional scouring methods using sodium hydroxide (NaOH) can damage textiles due to their harsh alkaline treatments. Enzyme-aided scouring offers a safer, eco-friendly alternative for delicate textiles like silk, wool, and cashmere, effectively removing pectins without causing fabric damage, thus meeting the demands of sustainable textile processing [76,77]. Also, non-ionic surfactants are often employed to enhance enzyme penetration by lowering the surface tension of fibres and facilitating the removal of wax and grease during scouring. This step is essential for preparing textiles for further industrial processing [77]. Pectinases, or pectinolytic enzymes, are the biocatalysts responsible for pectin degradation. This group includes polygalacturonase, pectin lyase, pectate lyase, and pectin esterase, which decompose pectins through glycolysis, hydrolysis, and deesterification [78]. These enzymes are primarily produced by saprophytic and plant-pathogenic microorganisms, including bacteria, yeasts, and fungi [79].
The application of scouring and degumming on various fibres using pectinase has been studied under different conditions, with fungal enzymes from Aspergillus niger being commercially significant and widely utilised across industries. For textile applications, a new prescription process for pectinase production by Aspergillus niger using solid fermentation was developed, optimising the culture substrate and temperature treatments, with the optimal formula consisting of rice dextrose, wheat bran, ammonium sulphate, and water, resulting in a pectinase activity of 36.3 IU/g dry mass [80]. Vigneswaran et al. used a commercial pectinase enzyme for scouring cotton at a pH range of 8.5–9, a temperature of 60 °C, and an enzyme dose of 6% for 1 h [81]. In another study by Rajendran et al. (2011), Fusarium sp. pectinase was applied for cotton scouring at a pH of 8, a temperature of 40 °C, and an enzyme dose of 2% for 0.33 h [82]. For degumming of ramie, Guo et al. (2013) applied Bacillus sp. pectate lyase at pH 8.5 for 2.4 h at 50 °C with an enzyme dose of 40 U/g and a fabric-to-moisture ratio of 1:13 [83]. Zheng et al. (2001) applied a combination of pectate lyase, polygalacturonase, xylanase, and cellulase from Bacillus sp. for degumming ramie at pH 10 for 5 h at 40 °C [84].
The use of mixed enzyme systems has proven highly effective in bioscouring, integrating multiple textile processes into a single step. These systems typically combine pectinase with cellulases, xylanases, cutinases, lipases, or proteases. Such combined processes offer significant advantages, including water and energy conservation [85,86]. For example, desizing, bioscouring, and bleaching can be performed in a single bath, minimising water and energy consumption. Alkaline pectinases are particularly compatible with other enzymes, such as α-amylase and glucose oxidase, and are suitable for simultaneous H2O2 bleaching and reactive dyeing [86,87]. Enzymatic scouring not only meets the demand for sustainable textile processing but also enhances fabric quality while reducing environmental impact, making it a preferred choice in modern textile industries.

4.3. Cellulases in Stonewashing and Finishing

Cellulases are enzymes that hydrolyse the β-1,4 linkages of cellulose, resulting in the release of degraded products such as glucose, cellobiose, and oligosaccharides of varying lengths. These enzymes are produced by various microorganisms, including bacteria and fungi, and have significant industrial applications [88]. Microbial cellulases, particularly those from Cellulomonas sp., Clostridium sp., Thermomonospora sp., Aspergillus sp., and Trichoderma sp., are produced in large quantities and are widely used in different commercial sectors [89]. Cellulase enzymes can be classified into three types: endoglucanase (which hydrolyses internal O-glycosidic bonds), exoglucanase (which hydrolyses cellulose from the ends), and β-glucosidase (which hydrolyses non-reducing β-D-glucosyl residues from the terminals) [89,90].
In the textile industry, cellulases are employed in various processes, such as the pretreatment of fibres, biowashing, mixed enzymatic scouring, stone finishing, and biopolishing [91]. Natural fibres like jute, cotton, ramie, and flax typically contain 15–30% non-cellulosic impurities, which can be efficiently removed by pre-treating the fibres with mixed enzymatic systems [92]. As an eco-friendly alternative, cellulase enzymes are used in biostoning to achieve a washed-out look like traditional stonewashing. This method reduces water consumption, waste, and energy use, while enhancing fabric softness and fade effects. Unlike pumice stones, which can damage machinery and contaminate wastewater, enzymatic biostoning offers a sustainable solution, adding value to denim without the environmental impact of conventional stonewashing [93]. Cellulases are extensively utilised in textiles to enhance fabric softness, smoothness, and lustre, while also creating a distinctive stonewashed effect on denim [94]. For effective fibre pretreatment, cellulases work in conjunction with other enzymes (such as hemicellulases, proteases, and pectinases) to target and remove fibre impurities [94,95]. For example, in mixed enzymatic scouring, cellulases penetrate the outer cuticle layer of the textile fibres, creating space for pectinases to act on the primary wall and hydrolyse pectin [96].
Cellulases are extensively utilised in textiles for their ability to enhance fabric softness, smoothness, and lustre, while also imparting a distinctive stonewashed effect on denim. They are most frequently applied in stonewashing or finishing processes, where they partially hydrolyse colours or dyes from the surface of denim and heavy cotton twills without compromising fabric strength [97]. In biopolishing, cellulases eliminate surface microfibrils on fabrics like cotton, linen, and viscose through controlled hydrolysis, creating a fibre-free surface that improves brightness, texture, luminosity, and softness [96]. Additionally, cellulases reduce pilling in viscose and fibrillation in lyocell, though their effects are less pronounced on cellulose acetate. Among fabrics, linen is the most susceptible to enzymatic hydrolysis, followed by viscose, cotton, and lyocell [98]. The benefits of using cellulases in textile processing over traditional chemical treatments are numerous, including improved textile quality, greener pretreatment methods, eco-friendliness, less damage to garments, reduced wastewater generation, enhanced biodegradability of wastewater, and reduced equipment load [99]. These advantages make cellulases an important component in sustainable textile manufacturing processes.

4.4. Glucose–Oxidases in Textile Bleaching

Many studies have demonstrated that hydrogen peroxide (H2O2) is a safer and more environmentally friendly alternative to traditional chlorine-based oxidising agents in textile bleaching. Hydrogen peroxide is particularly attractive because it decomposes into water and oxygen, making it less harmful to the environment compared to chlorine [100]. Glucose oxidase is an enzyme naturally produced by certain fungi (such as Penicillium glaucum and Aspergillus niger) that can be used to enzymatically produce hydrogen peroxide [101]. Glucose oxidase is a highly stable flavin adenine dinucleotide (FAD)-dependent enzyme that catalyses the oxidation of glucose into gluconic acid, simultaneously producing hydrogen peroxide in the presence of oxygen [29]. This hydrogen peroxide can then be used for biobleaching, a more sustainable method of fabric whitening.
Glucose oxidase is most effective in producing hydrogen peroxide at acidic to neutral pH levels and lower temperatures. However, traditional bleaching processes often require alkaline pH and higher temperatures for optimal results [102]. This mismatch between enzyme activity conditions and the bleaching requirements presents a challenge. To address this, Davulcu et al. employed ultrasound energy to enhance the enzymatic reaction, which improved fabric whiteness without the need for harsh alkaline conditions or high temperatures [103]. Sonicated bleaching with glucose oxidase significantly enhanced fabric whiteness and lightness through enzymatic generation of H2O2, while the application of ultrasound power beyond 150 W offered no further improvement in bleaching efficiency and resulted in a reduction in tensile strength [104]. This innovative approach allows for a more energy-efficient and environmentally friendly bleaching process.
Many researchers have explored the use of combined one-bath systems for desizing and bleaching, where the degradation products of starch (such as glucose) are reused to generate hydrogen peroxide [105,106]. For example, studies have shown that glucose derived from starch degradation can be utilised for hydrogen peroxide formation in a combined process [106]. Recently, Mojsov developed an efficient multi-enzyme one-bath system for treating sized cotton fabric. This system uses enzymes such as amylase, amyloglucosidase, pectinase, and glucose oxidase to convert starch into hydrogen peroxide through a series of reactions [107]. This enzymatic approach not only produces hydrogen peroxide for bleaching but also offers significant advantages in terms of reducing waste generation and water consumption in textile processing.

4.5. Catalase in Bleach Cleanup

Catalase plays a pivotal role in the textile industry, particularly in the bleach cleanup process, where it effectively removes residual hydrogen peroxide from fabrics. This enzymatic treatment ensures that textiles are peroxide-free and ready for subsequent processes such as dyeing. The term “bleach cleanup” was originally coined by Novo Nordisk to describe this specific enzymatic application [108]. Hydrogen peroxide is widely used as a bleaching agent due to its effectiveness in whitening fabrics. However, residual peroxide left on the fabric after bleaching can interfere with the dyeing process, causing uneven colour absorption and diminished fabric quality [109]. Catalase addresses this issue by catalysing the decomposition of hydrogen peroxide into water and oxygen, two environmentally benign byproducts [110].
Sooch et al. isolated a thermophilic strain of Geobacillus (BSS-7) that produced exceptionally high levels of intracellular catalase, demonstrating strong potential for hydrogen peroxide (H2O2) degradation. The strain effectively removed H2O2 in a packed bed reactor and has bleach cleanup applications [111]. Czyzewska et al. proposed the chemical immobilisation of catalase on membranes from psychrotolerant bacteria, enhancing its performance under textile conditions through membrane deposition. The catalytic properties of the immobilised catalase demonstrated its efficiency in decomposing hydrogen peroxide waste from textile bleaching processes [112]. Recently, Sheer et al. investigated recombinant manganese catalases from Geobacillus thermopakistaniensis, demonstrating their superior performance in hydrogen peroxide removal for sustainable textile processing. The manganese–catalase showed high activity, thermostability, and alkalinity, offering significant advantages by reducing water consumption, eliminating extensive rinsing, and operating under mild conditions [113,114]. Therefore, catalases are eco-friendly enzymes that leave no harmful residues, preserve fabric integrity, and enhance dye uptake while promoting environmentally sustainable practices in textile manufacturing [115].

4.6. Laccases in Textile Bleaching and Dye Decolourisation

Laccase is one of the most abundant oxidoreductase enzymes in nature, being widely distributed in microorganisms, including bacteria and fungi (e.g., Trametes versicolour, Pycnoporus cinnabarinus, and Bacillus subtilis). It is capable of oxidising phenols and aromatic amines by reducing molecular oxygen (O2) to water (H2O) through the action of multiple copper ions in its catalytic pocket [32]. In addition to the enzyme itself, the entire redox mediator system of laccase has been applied in various industrial processes. Laccase plays a significant role in several commercial sectors, including textiles, cosmetics, food, and the paper and pulp industries [116]. In textile processing, the use of laccase has gained considerable attention due to its two key applications: (i) textile bleaching and (ii) effluent dye decolourisation. Laccases are used in textile bleaching to modify fabric surfaces and enhance the whiteness of textiles, significantly reducing the amount of hydrogen peroxide (H2O2) required during the bleaching process [117]. Kim et al. demonstrated that laccases can effectively bleach textiles while minimising H2O2 dosages [118]. Additionally, Abou-Okeil et al. found that the combined effect of laccase and H2O2, assisted by ultrasonication, resulted in effective bleaching with less fibre damage. Ultrasound treatment helps to de-aggregate the enzyme, improving its activity and facilitating its diffusion into fibres, enhancing the overall bleaching efficiency [104]. Laccase, when combined with glucose oxidase in a dual-enzyme system, effectively removes lignin, enhancing linen bleaching. The combination demonstrated superior performance compared to individual enzymes, improving the surface properties and whiteness of the treated fibres [119].
Laccase plays a significant role in enhancing the whiteness of fabrics by improving the efficiency of hydrogen peroxide bleaching, leading to similar or better results than conventional methods with lower environmental impact [120]. For instance, Pereira et al. isolated a laccase from Trametes hirsuta, which effectively oxidised various flavonoids, improving the whiteness of cotton fabrics [121]. Furthermore, during the finishing stage of textile processing, laccase can transform toxic chemicals or dyes into non-toxic forms in a more eco-friendly and economical manner, helping reduce environmental pollution. For instance, Neifar et al. screened the salt-tolerant bacterium Pseudomonas extremorientalis BU118 for laccase production, achieving high enzymatic activity (7000 U/L) and demonstrating its potential for decolourising the textile azo dye Congo red under optimised conditions [122]. Similarly, the laccase from Streptomyces cyaneus CECT 3335 decolourised azo dyes like methyl orange and Orange II by 90% with acetosyringone as a mediator [123]. The laccase showed excellent colour removal capabilities across various dye and salt concentrations, highlighting its biotechnological potential in textile effluent treatment.
Laccase from Trametes versicolour has been successfully used in denim finishing, where it efficiently degrades indigo dye without the need for a redox mediator [124]. This application has proven to be a highly effective method for environmentally friendly denim processing. In addition to denim, the economic colouring of wool can be achieved using laccase, which requires less dye and demonstrates enhanced fixation capabilities compared to conventional methods. Laccase from Sclerotium rolfsii has also been used for wool dye decolourisation [125]. Overall, laccase is a promising biocatalyst in textile processing, offering an eco-friendly alternative to harmful chemicals and contributing to more sustainable methods due to its versatility and environmentally beneficial properties.

4.7. Peroxidases in Dye Degradation and Effluent Treatment

Peroxidases are a group of enzymes that play a significant role in the textile industry, particularly in addressing environmental challenges associated with dye degradation and effluent treatment [126]. The textile industry generates substantial quantities of wastewater containing synthetic dyes, which are not only aesthetically displeasing but also toxic to aquatic ecosystems. Peroxidases, by catalysing oxidative reactions, offer an eco-friendly and efficient solution to degrading these persistent dyes and reducing their environmental impact [127]. They are derived from various sources, such as plants (e.g., horseradish peroxidase), fungi (e.g., lignin peroxidase), and bacteria, which can catalyse the breakdown of these dyes by generating highly reactive radicals in the presence of hydrogen peroxide. These radicals attack dye molecules, leading to their oxidation and eventual breakdown into less harmful or completely mineralised products like water, carbon dioxide, and simple organic acids [128].
In textile dyeing processes, synthetic dyes such as azo, anthraquinone, and reactive dyes are extensively used. These dyes often resist natural degradation due to their complex aromatic structures, making traditional wastewater treatment methods insufficient [129]. The application of peroxidases in effluent treatment has shown promising results in degrading a wide range of dye pollutants. For example, lignin peroxidase and manganese peroxidase from white-rot fungi (Phanerochaete chrysosporium) are particularly effective against recalcitrant dyes, such as azo and anthraquinone dyes. These enzymes operate under mild conditions and can break down dye molecules that are otherwise resistant to conventional chemical and biological treatments [130]. A dye-decolourising peroxidase from Bacillus amyloliquefaciens (BaDyP) demonstrated significant potential for bioremediation of industrial effluent by degrading Reactive Blue 19 and Reactive Black 5, and reducing COD, colour, and toxicity. Treated effluent showed enhanced seed germination and reduced bacterial inhibition, highlighting BaDyP as an efficient biocatalyst for pollutant mitigation [131]. Recently, Ilić Đurđić et al. improved the degradation of azo dyes (Amido Black 10B, Evans blue and Guinea green) using lignin peroxidase (LiP) from Phanerochaete chrysosporium through mutagenesis near the catalytic pocket and the application of enzyme-coated yeast cell walls. Mutants exhibited up to 13-fold higher catalytic activity, and LiP-coated cell wall fragments demonstrated enhanced stability, retaining activity across multiple reaction cycles [132]. Saha et al. demonstrated that a bacterial consortium (VITPBC6), comprising six Bacillus strains (B. flexus, B. paraflexus, B. megaterium, B. firmus, B. flexus, and B. aryabhattai), efficiently degraded Reactive Orange 16 (RO-16) through enzymatic action involving azoreductase, tyrosinase, laccase, lignin peroxidase, and manganese peroxidase. This enzymatic degradation significantly reduced the dye’s toxicity, showcasing the consortium’s potential for bioremediation of textile effluents. Additionally, peroxidases are capable of treating mixed dye effluents, making them versatile in real-world applications [133]. Hossain et al. demonstrated efficient dye degradation by Bacillus pseudomycoides and Acinetobacter haemolyticus, isolated from industrial effluents, targeting methylene green, basic violet, and acid blue dyes. These bacteria showed up to 82% degradation for mixed dye combinations, driven by peroxidase and azoreductase enzymes essential for breaking azo bonds. Molecular docking and dynamics studies confirmed strong enzyme-dye interactions, highlighting the potential of these strains for large-scale wastewater treatment [134].
One of the key advantages of using peroxidases is their ability to work in conjunction with other treatment processes. For instance, enzymatic dye degradation can be combined with coagulation, adsorption, or membrane filtration to enhance overall effluent treatment efficiency [19]. The enzymatic approach minimises the use of harsh chemicals, reduces sludge generation, and ensures the treated water meets environmental standards for discharge or reuse.
Peroxidases are emerging as powerful biocatalysts for dye degradation and effluent treatment in the textile industry, offering an enzymatic alternative to toxic chemicals and energy-intensive processes. Their ability to degrade synthetic dyes under ambient, environmentally benign conditions, combined with biodegradability, positions them as a sustainable solution for reducing the environmental impact of textile processing [135]. Advances in enzyme technology and process optimisation could further enhance their potential for eco-friendly textile manufacturing.

4.8. Esterases and Lipases in Pre-Treatment of Textile

Esterases and lipases are enzymes classified as carboxyl ester hydrolases, which are distinguished primarily by their substrate specificity. Esterases catalyse the hydrolysis of water-soluble short-chain acyl esters, while lipases are typically active on water-insoluble long-chain triacylglycerols [136]. These enzymes have found significant applications in textile processing, particularly in improving the dye uptake properties of polyester fabrics [137]. Esterases and lipases are used in cotton fabric pre-treatment to remove sizing lubricants, enhancing absorbency and enabling better dye fixation for vibrant colours. A study on lipase (lipolase) showed it effectively scours cotton by removing wax, pectin, and protein, improving hydrophilicity, whiteness, and dyeing performance. Combining lipase with pectinase further enhanced scouring efficiency, reduced treatment time, and produced superior fabric properties comparable to conventional alkali methods [138]. Another study demonstrated that lipases are particularly effective in removing both natural and added fats/greases from cotton fabrics [139]. This was achieved through a mixed enzymatic bioscouring system that included lipase, along with other enzymes such as pectinase, cellulase, and protease. The combined effect of these enzymes in the bioscouring process helped to clean the cotton fabric more efficiently, enhancing its ability to absorb dye and improving the final quality of the textile [140]. Using esterases and lipases in textile pre-treatment offers several benefits, including improved dye uptake, enhanced fabric quality, and reduced environmental impact compared to traditional chemical treatments. These enzymes help ensure that the fabrics are more receptive to dyeing and promote eco-friendly processing by reducing the need for harsh chemicals and minimising water usage.

4.9. Arylesterases in Scouring and Bleaching

Arylesterases are enzymes classified under the hydrolase group, specifically responsible for catalysing the hydrolysis of carboxylic ester bonds [141]. In textile processing, arylesterases play a crucial role in the enzymatic scouring and bleaching of fabrics. Notably, they can catalyse the in situ production of peracetic acid from hydrogen peroxide (H2O2) under mild conditions, such as at neutral pH and a moderate temperature of 65 °C [142]. This ability to generate peracetic acid in situ makes arylesterases a valuable tool in sustainable textile processing, offering an effective alternative to traditional chemical bleaching methods. The application of arylesterases in the textile industry is particularly advantageous for scouring and bleaching processes, where the enzyme contributes to improving fabric characteristics such as absorbency, whiteness, and fibre tenacity. The enzymatic process is a one-step procedure that efficiently removes impurities from fibres, leading to cleaner textiles with enhanced quality. Studies have shown that enzymatic scouring and bleaching using arylesterases not only achieves high levels of whiteness but also results in minimal damage to the fibre structure, preserving the integrity and strength of the fabric [143,144]. Overall, arylesterases provide a sustainable solution for textile processing, improving dye uptake and uniformity while reducing environmental impact through minimised use of harsh chemicals, water, and energy.

4.10. Cutinases in Surface Modification of Synthetic Fibres

Cutinases are enzymes belonging to the hydrolase class, specifically responsible for hydrolysing cutin into cutin monomers. As serine esterases, cutinases contain a catalytic triad consisting of serine (Ser), histidine (His), and aspartic acid (Asp), which are essential for their hydrolytic activity [145]. Cutinases are used in textile processing to modify synthetic fibres by degrading complex polyesters like cutin, improving fabric properties. PET hydrolases, designed to degrade PET plastics, also play a role in recycling and modifying synthetic textiles by breaking down PET into monomers. These enzymes, isolated from various bacterial (Ideonella sakaiensis and Thermobifida fusca) and fungal (Aspergillus oryzae and Fusarium solani) species, offer diverse potential for textile surface modification and bioremediation applications [146]. Beyond their biological role, cutinases have become useful in the surface modification of synthetic fibres such as polyester, polyamide, and polyacrylonitriles [147]. For instance, in a heterogeneous system, the combined use of Candida antarctica lipase A and Fusarium solani pisi cutinase showed different degradation modes of poly(ε-caprolactone) (PCL), with lipase leading to higher mass loss and a significant decrease in crystallinity. While cutinase degradation maintained thermal stability and one-step decomposition, lipase hydrolysis resulted in reduced thermal stability and multi-step degradation [148]. Similarly, Carniel et al. demonstrated that lipase from Candida antarctica (CALB) and cutinase from Humicola insolens (HiC) work synergistically to hydrolyse poly(ethylene terephthalate) (PET) into terephthalic acid (TPA), with CALB converting bis-(hydroxyethyl) terephthalate (BHET) to TPA even at high concentrations. This combination enhanced PET depolymerisation, achieving a 7.7-fold increase in yield and a TPA mole fraction of up to 0.88, with potential applications in the surface modification of synthetic fibres and PET recycling [149]. Cutinases, when combined with other enzymes like pectin lyase, have been used to improve the wetting properties of cotton fibres, enhancing their water absorption [85]. Their application in synthetic fibre surface modification improves wettability, dye uptake, and functionality, offering an eco-friendly alternative to harsh chemical treatments in sustainable textile processing.

4.11. Proteases in Scouring

Proteases are enzymes produced by microorganisms (Bacillus subtilis, Bacillus licheniformis, and Aspergillus niger) that break down proteins into smaller peptides and amino acids. Fungal proteases typically activate in a pH range of about 4 to 8, while bacterial proteases work best at a pH of 7 to 8. In the textile industry, proteases subtilisin and keratinase are used in processes like enzyme-based biopolishing, degumming of silk, wool finishing, and denim stonewashing, enhancing fabric softness, improving dyeing properties, and reducing the need for harsh chemicals and water consumption, making the process more eco-friendly [150,151]. Protease enzymes from Bacillus subtilis 168 E6-5 and commercial enzymes were tested on wool fabric to modify its surface and prevent felting. The results showed that the protease from B. subtilis enhanced the physical properties of wool, including tear strength and whiteness, more effectively than the commercial enzyme [152]. In the degumming of silk, proteases break down the sericin protein, enhancing the fabric’s softness, sheen, and mechanical properties. Enzymatic degumming under optimal conditions improves silk’s physical properties and serves as an eco-friendly alternative to traditional soap treatments [153]. Sreelakshmi et al. demonstrated that acidic proteases from Bacillus licheniformis and Bacillus cereus have potential for scouring cotton fabrics, with B. licheniformis showing superior efficiency in weight loss, absorbency, and impurity removal. Protease-treated samples exhibited reduced tear strength compared to alkali scouring, while FTIR analysis confirmed the removal of protein-related compounds, retaining wax and pectin residues [154]. Proteases, in combination with cellulases, are also used for optimised bioscouring. For instance, a study has shown that mixing proteases (such as those from Bacillus) with cellulase results in successful scouring, dye removal, and improved water absorbency, although some fibre damage can occur with cellulase presence. For denim stonewashing, proteases break down protein-based impurities, achieving the desired faded look without damaging fabric strength. Proteases combined with commercial enzymes have been optimised using artificial neural network (ANN) techniques to enhance absorbency and remove pectin, improving overall fabric quality [155]. Another study revealed that protease treatment of wool fabric at various concentrations (1–5 g/L) improved physical properties such as softness, absorbency, and surface smoothness, while reducing weight loss and enhancing dyeability and colour fastness compared to untreated wool. The 4 g/L protease concentration showed the best results in improving the fabric’s handle, tensile strength, and overall quality [156]. Therefore, proteases in textile processing provide an eco-friendly alternative, enhancing absorbency and cleaning while reducing chemicals and water usage.

5. Improving Enzymes for Textile Applications

This section covers recombinant and engineered enzymes, immobilised enzymes, and extremozymes—robust biocatalysts essential for the harsh processes of textile applications. Improving enzymes for textile applications involves designing recombinant and engineered enzymes with enhanced specificity, activity, and stability tailored to industrial needs [157]. Immobilised enzymes, fixed on solid supports, offer reusability and operational stability, reducing costs and environmental impact in processes like bioscouring and degumming [158,159]. Extremozymes, derived from extremophiles, are robust biocatalysts with resilience to extreme conditions such as high temperature and pH, making them ideal for harsh textile processing environments [160].

5.1. Recombinant and Engineered Enzymes

Recombinant enzymes for textile processing have significantly advanced industrial biotechnology by enabling the production of tailored enzymes with enhanced properties. Native enzymes often fail to meet industrial yield demands, necessitating heterologous expression in mesophilic hosts. Host selection depends on genome compatibility, metabolism, growth rate, and culture conditions. While E. coli is a preferred host due to its ease of manipulation and rapid growth, its limitations include codon bias, inclusion body formation, and lipopolysaccharide contamination [161]. Alternative hosts, such as Bacillus subtilis, Saccharomyces cerevisiae, and Trichoderma reesei, offer efficient systems for protein expression. Codon harmonisation algorithms effectively address codon bias, enhancing protein expression [162,163].
Advances in enzyme production focus on achieving stability under extreme conditions, such as high or low temperatures, salinity, pressure, and pH variations. For instance, cold-adaptive lipases, such as LSK25 from Pseudomonas sp., have been expressed in E. coli BL21 (DE3) as inclusion bodies, which enhance proteolytic resistance and enable industrial-scale production. These enzymes, with optimal activity at 30 °C and pH 6, are valuable for cold washing and other temperature-sensitive industrial processes [164]. Similarly, thermostable xylanases derived from metagenomics have been expressed in E. coli, achieving yields of ~100 mg/L with optimal activity at 80 °C and pH 7. Metal ions such as Ni, Zn, and Mg enhance activity by up to 55% [165]. Recombinant production of other enzymes, such as cold-adaptive endoxylanases in Pichia pastoris and thermophilic endoglucanases from Sulfolobus shibatae, demonstrates the versatility of heterologous systems [166]. These enzymes exhibit optimal activity under industrially relevant conditions. Using hosts like Bacillus subtilis and Aspergillus spp. for extracellular enzyme production eliminates the need for cell disruption, reducing production costs [167]. In another example, Yu et al. optimised recombinant alkaline catalase (KatA) production in E. coli BL21 under ethanol stress, achieving a high yield of 78,762 U/mL with an extracellular ratio of 92.5%. The purified KatA exhibited stability at 50 °C and pH 6–10, efficiently removing residual H2O2 and reducing water, steam, and energy consumption by 25%, 12%, and 16.7%, respectively, without compromising fabric quality [168].
Enzyme engineering has emerged as a keystone of sustainable biotechnology, particularly in textile processing and waste treatment, by enabling the development of highly efficient, environmentally friendly enzymes. Figure 3A illustrates various enzyme engineering strategies. Modern approaches include directed evolution, site-directed mutagenesis, in silico enzyme design, enzyme truncation or fusion, and integrated combined mutagenesis, each playing a vital role in optimising enzyme functionality for industrial applications [169]. Directed evolution, using techniques like error-prone PCR and DNA shuffling, enhances enzyme traits such as stability, efficiency, and substrate specificity, making them suitable for harsh industrial conditions. Site-directed mutagenesis and in silico enzyme design further refine enzyme activity through precise modifications and computational predictions, while truncation and fusion improve adaptability. Combining these approaches produces engineered enzymes with superior thermal stability, pH resistance, and catalytic performance for industrial applications like textile effluent treatment. For industrial bioscouring applications, enzymes must be stable under elevated temperatures and alkaline pH conditions [170]. The efficiency of pectin removal varies depending on the enzyme, with optimal conditions generally involving 0.05–2% pectinase at temperatures of 30–80 °C and pH levels of 3–9 [171]. Solbak et al. reported a novel pectate lyase mutant with eight mutations (A118H, Y190L, A197G, S208K, S263K, N275Y, Y309W, S312V) that improved the enzyme’s thermal stability by 16 °C, enabling efficient high-temperature bioscouring with lower enzyme dosages [172]. Recently, Li et al. enhanced the activity and alkali resistance of alkaline pectate lyase BspPel from Bacillus RN.1 through loop replacement, achieving a 4.4-fold increase in specific activity (139.4 U/mg) and improved stability at pH 11.0 and 60 °C. The modified enzyme, with enhanced affinity for apple pectin, shows significant potential for industrial applications, including bioscouring in textile processing [173]. Khan et al. employed a rational protein engineering approach to enhancing the thermostability and activity of Bacillus subtilis lipase, generating mutants (F41K, W42E, P119E, Q121K, V149K, Q150E) that exhibited improved activity at 60 °C and pH 10, and successfully removing up to 90% fats/grease from the fabric [139]. In another study, Saha et al. compared the desizing efficiency of full-length and C-terminal truncated α-amylases from Bacillus subtilis on starch-coated polystyrene, silk, and cotton fabrics, with truncated α-amylases achieving up to 85% efficiency on polystyrene fabrics, highlighting their potential for textile applications [12]. Therefore, enzyme engineering in textile processing and waste treatment not only enhances efficiency and sustainability but also supports the transition towards greener, more eco-friendly industrial practices, benefiting both production processes and environmental health.

5.2. Immobilised Enzymes

Enzyme immobilisation has transformed industrial biocatalysis by enabling enzymes to retain their catalytic activity while being reused continuously. Immobilisation involves attaching enzymes to support materials, restricting their mobility to create heterogeneous catalysts suitable for diverse applications. The success of immobilisation depends on the choice of support materials and immobilisation methods, both of which critically influence enzyme performance [174]. Enzyme immobilisation methods include surface attachment, encapsulation, and cross-linking (Figure 3B). Surface attachment involves physical adsorption, ionic bonding, or covalent bonding. Physical adsorption is simple and reversible but can result in desorption under changing conditions. Ionic bonding relies on electrostatic interactions, offering moderate binding strength with activity retention, although it is sensitive to pH and ionic strength variations. Covalent bonding forms stable, irreversible links between the enzyme and carrier, enhancing stability but potentially causing activity loss due to structural changes. Encapsulation confines enzymes within semi-permeable matrices such as alginate beads, protecting them from environmental factors while allowing substrate and product diffusion, albeit with potential mass transfer limitations. Cross-linking aggregates enzyme molecules using reagents like glutaraldehyde, creating a carrier-free immobilisation system with enhanced stability but potentially reduced activity due to rigidity [175,176].
Immobilisation affects enzyme properties like activity, stability, pH and temperature optima, substrate specificity, and kinetics. Efficiency depends on yield, specific activity, stability, and reusability. Yield includes bound enzymes and retained activity, while specific activity reflects reaction rate. Stability and reusability are assessed under operational and storage conditions [177].
For example, a catalase from Bacillus sp. was covalently immobilised on silanised alumina, showing higher stability at alkaline pH and temperatures compared to the free enzyme. The immobilised catalase demonstrated improved stability (214 h at pH 11, 30 °C), but its activity was inhibited by anionic stabilisers or surfactants in the hydrogen peroxide substrate solution [178]. Another study by Costa et al. revealed the covalent immobilisation of commercial catalase (Terminox Ultra 50L) on alumina using glutaraldehyde for the recycling of textile bleaching effluents in dyeing processes. The immobilised enzyme retained 44% of its activity at pH 11 and 30 °C, and 90% at 80 °C and pH 7. Furthermore, the half-life of the immobilised catalase was extended to 2 h at pH 12 and 60 °C [179]. The immobilisation of Bacillus subtilis MTCC 2414 laccase using sodium alginate was optimised for dye degradation studies. The immobilised laccase exhibited higher degradation of Yellow GR dye (81.72%) compared to free laccase (74.69%) and culture filtrate (72.16%), with maximum activity at pH 7 and 35 °C [180]. Table 6 presents more examples of immobilised enzymes based on different immobilisation methods used in various textile processing and waste management applications.

5.3. Extremozymes—Robust Biocatalysts

Extremozymes are specialised enzymes produced by extremophiles—microorganisms that thrive in extreme environments such as high/low temperatures, acidic/alkaline conditions, high salinity, or extreme pressures [195]. These enzymes have evolved unique molecular mechanisms that enable them to function under harsh conditions, where standard enzymes from mesophilic organisms would typically be denatured, with their biochemical properties adapted to ensure stability under environmental stress. For example, thermophilic enzymes often have a large hydrophobic core, which minimises flexibility and helps maintain protein stability at high temperatures. The increased number of disulfide bonds and specific amino acid compositions (e.g., more hydrophobic and aromatic residues) contribute to their heat tolerance. Psychrophiles, on the other hand, have enzymes with greater flexibility due to a reduced number of proline residues, which increases protein mobility and allows them to function effectively at low temperatures [160,195].
Extremozymes play a crucial role in both industrial wastewater treatment and textile processing, particularly in the textile industry, where effluents are often rich in salts, dyes, and chemicals. These enzymes are utilised in key processes such as desizing, bioscouring, biostoning, biobleaching, and biopolishing, where they improve fabric quality, reduce energy consumption, and minimise the use of chemicals [196]. Halophilic extremozymes are highly effective in decolourising azo dyes and degrading a range of chemical pollutants found in textile effluents. Strains like Halogeometricum borinquense and Halomonas sp. have demonstrated exceptional ability to break down dyes even under high-salinity conditions, further enhancing the sustainability of textile wastewater treatment [197]. A thermostable, mildly alkalophilic recombinant laccase from Thermus thermophilus, produced in E. coli, exhibited optimal activity at ~90 °C for short reactions and efficiently decolourised up to six industrial dyes (orange dye, green dye, acid red dye, Naphthol Brilliant Blue, Remazol Brilliant Blue, and Congo red), demonstrating potential for various redox applications [198].
A study on thermostable Bacillus licheniformis α-amylase optimised for desizing cotton fabrics focused on reducing sugars, absorbency, flexural rigidity, and weight loss. The optimal conditions for desizing were 3.0 g/L enzyme concentration, pH 6.0, 40 min treatment, and 85 °C, resulting in improved fabric absorbency and lower residual matters compared to acid desizing [199]. Similarly, a highly thermostable and alkaline α-amylase from Bacillus sp. AB68, isolated from Van soda lake, exhibited activity between 20 °C and 90 °C, with optimal activity at pH 10.5 and 50 °C, maintaining over 90% activity after 30 min at 80 °C [200]. Anish et al. demonstrated that an alkali-stable endoglucanase from Thermomonospora sp. effectively reduces back staining while enhancing softness and abrasion in denim biofinishing, offering a sustainable alternative to conventional acid and neutral cellulases [201].
The use of extremozymes in the textile industry improves process efficiency and fabric quality and contributes to more sustainable and cost-effective manufacturing practices. With ongoing research and technological advancements, extremozymes have the potential to revolutionise industrial processes further, offering a greener alternative to traditional chemical methods.

6. Strategies for Reducing Textile Waste

The textile industry is a significant source of environmental pollution, driven by its high consumption of water, energy, and chemicals during production. To mitigate its environmental footprint, a range of strategies has been developed, focusing on different stages of production [202]. These include preparation, dyeing, finishing, and overarching general practices aimed at promoting sustainability (Figure 4).
The preparation stage involves recovery systems, waste steam reuse, chemical substitutions, and alternative processing methods, all of which play a crucial role in reducing waste and enhancing sustainability in textile production. Recent studies highlight the use of recycled fibres from textile waste in composite materials to promote eco-sustainability and circular economy practices in textile manufacturing [203]. Textile waste management currently relies on incineration for energy recovery, reuse, and recycling, with emerging technologies like mechanical recycling, such as SOEX’s process, which transforms 15,000 tons of clothing annually into insulation materials [204]. Recovery systems such as membrane filtration and centrifugation reclaim valuable materials like sizing agents and caustic soda, minimising waste and reducing resource demand. The environmental impact of textile production, the issue of solid textile waste, and strategies for recycling fibres into thermoplastic polymers, resins, natural constituents, and concrete are also discussed [205]. Waste steam from processes like desizing and scouring can be redirected to other heating applications, conserving energy and water and reducing greenhouse gas emissions [203,206]. Chemical substitutions, such as using enzymes or biodegradable surfactants instead of harsh alkalis, can improve wastewater quality by reducing toxicity. Despite these alternatives, textile auxiliaries like alkyl phenol ethoxylates (APEOs), nonylphenol ethoxylates, sodium chloride, formaldehyde, and perfluoroalkyl chains are commonly used, contributing to environmental and health risks due to their toxicity, slow degradation, and hormone disruption. The diverse and often secretive nature of textile chemical formulations complicates efforts to mitigate their impact, even as alternative chemicals and improved practices help reduce harm [203]. Additionally, innovative methods like enzymatic desizing, ultrasonic washing, and plasma treatment offer resource-efficient and eco-friendly alternatives to traditional techniques, advancing sustainability in textile processing [207].
Reconstituting and reusing dye baths conserve water and dyes by purifying and adjusting them for reuse, ensuring consistent colour quality while reducing resource consumption. Replacing synthetic dyes and traditional mordants with biodegradable or plant-based alternatives minimises the release of harmful chemicals into wastewater [208]. Advanced dyeing techniques, such as digital printing, cold pad batch dyeing, and supercritical CO2 dyeing, significantly reduce water use, eliminate toxic substances, and enhance energy efficiency. Synthetic dyes, however, continue to pollute water bodies due to weak fabric binding and untreated effluent discharge, causing toxicity, mutagenicity, and environmental degradation [19]. The biodegradation of dyes by microbes depends on the dye structure, with synthetic azo dyes being more readily degraded due to their azo bonds (-N=N-), which can be reduced or oxidised by aerobic or anaerobic bacteria. In contrast, natural colourants are more complex, making them less susceptible to microbial breakdown [209]. Eco-friendly wastewater treatment technologies and the adoption of biomordants derived from natural sources offer sustainable solutions, though challenges like scalability and cost remain. By embracing these innovative approaches, the textile industry can mitigate environmental impacts while meeting consumer demand for sustainable products [210].
Reusing residual finishing chemicals like softeners and resins through filtration and purification technologies reduces textile waste without compromising fabric quality, while substituting conventional agents with eco-friendly alternatives such as plant-based resins and natural softeners mitigates environmental harm [211]. Sustainable finishing methods, including enzymatic treatments, plasma technology, and UV curing, require fewer chemicals and less energy while maintaining or enhancing fabric properties. These advances align with the biodegradation of textile fibres, where enzymatic processes play a critical role. Natural fibres like cotton and linen degrade faster than synthetic blends, influenced by crystallinity, polymerisation, and microbial activity, with fungi (Aspergillus sp., Fusarium sp., Penicillium sp., and Trichoderma sp.) and bacteria (Bacillus sp., Cellulomonas sp., Cellvibrio sp., and Clostridium sp.) degrading fibres differently [212]. Enzymes such as cutinases and keratinases enhance the breakdown of synthetic fibres like linen/PET (polyethylene terephthalate) and natural proteins like wool keratin, offering innovative solutions for textile waste recycling and promoting sustainable material management [213].
Characterising waste streams is essential for designing targeted waste reduction strategies. By analysing the type and volume of waste generated, including wastewater, solid waste, and emissions, manufacturers can identify opportunities for recovery and reuse [214]. Choosing sustainable raw materials like organic cotton, recycled fibres, or biodegradable synthetics reduces environmental impact and production waste. Textile waste can be repurposed into lower-value materials such as insulation or gypsum composites, or synthesised into products like cellulose acetate and biofuels. Advanced recycling processes extract components like caprolactam for Nylon 6 or bioconvert cotton into value-added products like glucose syrup, though challenges remain in achieving high yields due to cotton’s compact crystalline structure [215]. Managing the fate of processing chemicals is another vital aspect. Closed-loop systems and advanced treatments allow for the recovery and neutralisation of chemicals, preventing their accumulation in the environment [216]. The equalisation of wastewater flows is achieved through specialised tanks that balance effluent loads, ensuring consistent flow and concentration levels. This enhances the efficiency of downstream wastewater treatment processes and contributes to overall waste reduction.
Adopting these approaches across all stages of textile production not only reduces waste but also conserves valuable resources and minimises the industry’s environmental footprint. As sustainability becomes increasingly important, continued innovation and the implementation of these practices will play a key role in achieving a greener textile industry.

7. Challenges and Benefits of Enzymatic Textile Processing

Enzymatic textile processing has emerged as a promising alternative to conventional chemical-based methods, offering numerous advantages, particularly regarding sustainability, fabric quality, and environmental impact [10,77,217]. However, despite these benefits, several challenges must be addressed for the efficacious use of enzymes in textile applications. One significant challenge lies in the extremely specific nature of enzymes. Enzymes are highly selective in their reactions, acting on only particular substrates, with minimal side effects. While advantageous for precise reactions, this specificity limits the enzyme’s versatility and requires careful selection for different textile processes [218].
Enzymes also operate under mild conditions, making them safe to handle and non-corrosive in their applications. They require low energy, and their use results in reduced chemical consumption, leading to lower loads on effluent treatment plants (ETPs). This is due to the fact that enzymes are biodegradable and produce fewer hazardous residues, reducing the chemical discharge and overall environmental impact of textile processing [219]. However, this advantage is tempered by the enzyme’s sensitivity to environmental conditions. While enzymes remain chemically stable under varying pH or temperature, their physical configuration can alter, leading to “denaturation” and a loss of activity. This presents a challenge, as enzymes must be maintained within an optimal temperature and pH range [220]. For example, live steam should never be added directly to an enzyme bath, and enzymes must be pre-diluted to preserve their functionality.
Another critical challenge in enzymatic textile processing is the compatibility of enzymes with surfactants. Enzymes often exhibit limited compatibility with ionic surfactants, which can affect their effectiveness. To ensure high working efficiency and uniform results, nonionic wetting agents with appropriate cloud points must be selected [221]. Additionally, enzymes are highly sensitive to pH fluctuations, heavy metal contamination, and the precise temperature range. These sensitivities require careful monitoring and intense caution during their use in textile processing.
The high cost of enzyme production is another major barrier to their widespread adoption. While technological advancements and large-scale production may reduce costs, feasible cost-control measures must also be explored. These measures include utilising recombinant DNA technology to engineer microorganisms for enhanced enzyme production, optimising fermentation processes, employing cost-effective raw materials as substrates, and recycling immobilised enzymes for multiple uses [222]. Such approaches can help mitigate enzyme costs and enhance their affordability for textile applications.
Despite these challenges, the advantages of enzymatic processes in textiles are clear. Enzymes contribute to a lower discharge of chemicals and wastewater, reducing the handling of hazardous chemicals by textile workers. This results in improved fabric quality, offering consumers more fashion choices and extending the life of garments due to reduced damage to the original fabric. Furthermore, enzymes are an eco-friendly alternative to traditional chemical methods, leading to reduced water and energy consumption, and ultimately promoting sustainability in textile manufacturing [221].
The properties of enzymes used in textiles further highlight their potential. Enzymes accelerate reactions by lowering the activation energy required, acting as catalysts that remain unchanged at the end of the reaction. This makes enzymes efficient and reusable. Additionally, enzymes work best under specific temperature and pH conditions, with their activity peaking at the optimum values. Beyond their efficiency, enzymes provide an alternative to polluting chemicals, as they do not produce toxic waste upon degradation. Furthermore, enzymes exhibit high specificity, catalysing only specific reactions or interacting with specific substrates [223]. For instance, desizing enzymes used in cotton processing do not affect the cellulose, preserving the strength of the fabric [12].
The ease of controlling enzyme activity is another advantage, as their performance can be managed by maintaining optimal conditions. Moreover, enzymes are biodegradable, meaning they do not contribute to pollution when their activity is complete, making them a cleaner, more sustainable choice for textile processing [220]. However, effective enzymatic textile processing requires careful attention to enzyme production, purification, and optimisation. Factors such as optimum temperature, pH, process time, enzyme dosage, ionic strength, and the use of co-substrates must be considered to ensure the highest efficiency and effectiveness [220,224].
Enzymatic textile processing presents a promising solution to reduce environmental impact, improve fabric quality, and promote sustainability. However, addressing the challenges related to enzyme specificity, sensitivity, and compatibility is crucial for maximising their potential. With proper optimisation and careful control, enzymatic processing can revolutionise the textile industry, providing an eco-friendly and efficient alternative to traditional chemical-based methods.

8. Conclusions and Future Perspectives

Enzymatic solutions present a compelling avenue for revolutionising the textile industry, offering a sustainable and eco-friendly alternative to traditional chemical treatments. Enzymes such as amylases, pectinases, cellulases, and oxidoreductases have demonstrated their ability to replace hazardous chemicals in critical stages of textile processing—desizing, scouring, bleaching, and finishing. These biocatalysts not only improve fabric quality by enhancing softness, appearance, and durability but also significantly reduce the environmental impact by minimising chemical usage, water consumption, and effluent pollution. Furthermore, enzymes catalyse specific reactions with high selectivity, which ensures precision in textile processing while preserving the integrity of the fabric.
Despite their potential, enzymatic treatments face challenges including substrate specificity and sensitivity to environmental factors like pH, temperature, and the presence of inhibitors, as well as limitations in compatibility with certain surfactants and other additives. Additionally, the high cost of enzyme production and purification remains a constraint, though ongoing advancements in recombinant DNA technology and protein engineering are progressively lowering these barriers. Optimisation of process parameters, such as enzyme concentration, pH, and temperature, is also critical for enhancing enzyme performance and industrial viability.
Future research in enzymatic textile processing holds significant promise for addressing current challenges and advancing the field. Genetic engineering techniques, such as directed evolution, site-directed mutagenesis, and enzyme truncation, offer opportunities to improve enzyme properties, including enhanced stability, thermostability, pH tolerance, surfactant resistance, and substrate specificity. The development of immobilisation technologies could further extend enzyme longevity, improve reusability, and enhance operational efficiency. Additionally, the integration of enzymatic processes with complementary eco-friendly technologies, such as membrane filtration and advanced oxidation processes, could significantly improve the sustainability of textile manufacturing, particularly in the context of wastewater treatment. These advancements would not only optimise the use of enzymes in textile processing but also contribute to a more sustainable and environmentally friendly industry.
Collaboration between academia, industry, and regulatory bodies, coupled with growing consumer demand for sustainable textile products, will accelerate the widespread adoption of enzyme-based technologies. With sustained innovation in enzyme engineering, process optimisation, and integration with complementary technologies, the textile industry has the potential to shift towards a more sustainable and circular economy, significantly reducing its environmental footprint while improving product quality and performance.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

M.F.K. expresses gratitude to Mohd. Tasleem Khan, Yusra Khan, and Nesrine Sallem for their valuable insights and suggestions during the preparation of this review article. M.F.K. also acknowledge the use of ChatGPT (ChatGPT-4, OpenAI, San Francisco, CA, USA) for language refinement and grammatical editing of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mazotto, A.M.; de Ramos Silva, J.; de Brito, L.A.A.; Rocha, N.U.; de Souza Soares, A. How can microbiology help to improve sustainability in the fashion industry? Environ. Technol. Innov. 2021, 23, 101760. [Google Scholar] [CrossRef]
  2. Singh, R.S.; Singh, T.; Pandey, A. Microbial enzymes—An overview. In Advances in Enzyme Technology; Pandey, A., Ed.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 1–40. [Google Scholar]
  3. Kumar, D.; Bhardwaj, R.; Jassal, S.; Goyal, T.; Khullar, A.; Gupta, N. Application of enzymes for an eco-friendly approach to textile processing. Environ. Sci. Pollut. Res. 2021, 30, 71838–71848. [Google Scholar] [CrossRef]
  4. 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]
  5. Sen, A.; Kapila, R.; Chaudhary, S.; Nigam, A. Biotechnological applications of microbial enzymes to replace chemicals in the textile industry—A review. Text. Assoc. 2021, 82, 68–73. [Google Scholar]
  6. Sujitha, P.; Kavitha, S.; Shakilanishi, S.; Babu, N.K.C.; Shanthi, C. Enzymatic dehairing: A comprehensive review on the mechanistic aspects with emphasis on enzyme specificity. Int. J. Biol. Macromol. 2018, 118, 168–179. [Google Scholar] [CrossRef] [PubMed]
  7. Azanaw, A.; Birlie, B.; Teshome, B.; Jemberie, M. Textile effluent treatment methods and eco-friendly resolution of textile wastewater. Case Stud. Chem. Environ. Eng. 2022, 6, 100230. [Google Scholar] [CrossRef]
  8. Morsi, R.; Bilal, M.; Iqbal, H.M.; Ashraf, S.S. Laccases and peroxidases: The smart, greener and futuristic biocatalytic tools to mitigate recalcitrant emerging pollutants. Sci. Total Environ. 2020, 714, 136572. [Google Scholar] [CrossRef]
  9. Bernal, S.P.; Lira, M.M.; Jean-Baptiste, J.; Garcia, P.E.; Batista, E.; Ottoni, J.R.; Passarini, M.R. Biotechnological potential of microorganisms from textile effluent: Isolation, enzymatic activity, and dye discolouration. An. Acad. Bras. Ciênc. 2021, 93, e20191581. [Google Scholar] [CrossRef] [PubMed]
  10. Shen, J.; Smith, E. Enzymatic treatments for sustainable textile processing. In Sustainable Apparel; Richard, B., Ed.; Elsevier: Amsterdam, The Netherlands, 2015; pp. 119–133. [Google Scholar]
  11. Chowdhury, I.R.; Summerscales, J. Woven fabrics for composite reinforcement: A review. J. Compos. Sci. 2024, 8, 280. [Google Scholar] [CrossRef]
  12. Saha, P.; Khan, M.F.; Patra, S. Truncated α-amylase: An improved candidate for textile processing. Prep. Biochem. Biotechnol. 2018, 48, 635–645. [Google Scholar] [CrossRef]
  13. Catarino, M.L.; Sampaio, F.; Gonçalves, A.L. Sustainable Wet Processing Technologies for the Textile Industry: A Comprehensive Review. Sustainability 2025, 17, 3041. [Google Scholar] [CrossRef]
  14. Singha, K.; Pandit, P.; Maity, S.; Sharma, S.R. Harmful environmental effects of textile chemical dyeing practice. In Green Chemistry for Sustainable Textiles; Muthu, S.S., Ed.; Woodhead Publishing: Cambridge, UK, 2021; pp. 153–164. [Google Scholar]
  15. Al-Tohamy, R.; Ali, S.S.; Li, F.; Okasha, K.M.; Mahmoud, Y.A.G.; Elsamahy, T.; Jiao, H.; Fu, Y.; Sun, J. A critical review on the treatment of dye-containing wastewater: Ecotoxicological and health concerns of textile dyes and possible remediation approaches for environmental safety. Ecotoxicol. Environ. Saf. 2022, 231, 113160. [Google Scholar] [CrossRef]
  16. Khan, M.F.; Murphy, C.D. Fluorotelomer alcohols are efficiently biotransformed by Cunninghamella elegans. Environ. Sci. Pollut. Res. 2023, 30, 23613–23623. [Google Scholar] [CrossRef]
  17. Khan, M.F.; Paul Guin, J.; Thampi, R.K.; Sullivan, J.A.; Murphy, C.D. Enhanced removal of perfluorooctanoic acid with sequential photocatalysis and fungal treatment. Environ. Sci. Pollut. Res. 2023, 30, 91478–91486. [Google Scholar] [CrossRef]
  18. Khan, M.F.; Murphy, C.D. Application of microbial biofilms in biocatalysis and biodegradation. In Enzymes for Pollutant Degradation; Mulla, S.I., Bharagava, R.N., Eds.; Springer Nature Singapore: Singapore, 2022; Volume 30, pp. 93–118. [Google Scholar]
  19. Kuppan, N.; Padman, M.; Mahadeva, M.; Srinivasan, S.; Devarajan, R. A comprehensive review of sustainable bioremediation techniques: Eco-friendly solutions for waste and pollution management. Waste Manag. Bull. 2024, 2, 154–171. [Google Scholar] [CrossRef]
  20. Das, S.; Cherwoo, L.; Singh, R. Decoding dye degradation: Microbial remediation of textile industry effluents. Biotechnol. Notes 2023, 4, 64–76. [Google Scholar] [CrossRef] [PubMed]
  21. Khan, M.F.; Hof, C.; Niemcová, P.; Murphy, C.D. Biotransformation of fluorinated drugs and xenobiotics by the model fungus Cunninghamella elegans. Methods Enzymol. 2024, 696, 251–285. [Google Scholar]
  22. Khan, M.F.; Murphy, C.D. Bacterial degradation of the anti-depressant drug fluoxetine produces trifluoroacetic acid and fluoride ion. Appl. Microbiol. Biotechnol. 2021, 105, 9359–9369. [Google Scholar] [CrossRef]
  23. Pundir, A.; Thakur, M.S.; Prakash, S.; Kumari, N.; Sharma, N.; Parameswari, E.; He, Z.; Nam, S.; Thakur, M.; Puri, S.; et al. Fungi as versatile biocatalytic tool for treatment of textile wastewater effluents. Environ. Sci. Eur. 2024, 36, 185. [Google Scholar] [CrossRef]
  24. Rathour, R.K.; Sharma, D.; Ullah, S.; Mahmoud, E.H.M.; Sharma, N.; Kumar, P.; Bhatt, A.K.; Ahmad, I.; Bhatia, R.K. Bacterial–microalgal consortia for bioremediation of textile industry wastewater and resource recovery for circular economy. Biotechnol. Environ. 2024, 1, 6. [Google Scholar] [CrossRef]
  25. Sornaly, H.H.; Ahmed, S.; Titin, K.F.; Islam, M.N.; Parvin, A.; Islam, M.A.; Faruquee, H.M.; Biswas, K.K.; Islam, R.; Paul, D.K.; et al. The utility of bioremediation approach over physicochemical methods to detoxify dyes discharges from textile effluents: A comprehensive review study. Sustain. Chem. Pharm. 2024, 39, 101538. [Google Scholar] [CrossRef]
  26. Alazaiza, M.Y.; Albahnasawi, A.; Ahmad, Z.; Bashir, M.J.; Al-Wahaibi, T.; Abujazar, M.S.S.; Amr, S.S.A.; Nassani, D.E. Potential use of algae for the bioremediation of different types of wastewater and contaminants: Production of bioproducts and biofuel for green circular economy. J. Environ. Manag. 2022, 324, 116415. [Google Scholar] [CrossRef]
  27. Singh, L.; Singh, V.P. Textile dyes degradation: A microbial approach for biodegradation of pollutants. In Microbial Degradation of Synthetic Dyes in Wastewaters; Singh, S., Ed.; Springer International Publishing: Cham, Switzerland, 2014; pp. 187–204. [Google Scholar]
  28. Latif, W.; Ciniglia, C.; Iovinella, M.; Shafiq, M.; Papa, S. Role of white rot fungi in industrial wastewater treatment: A review. Appl. Sci. 2023, 13, 8318. [Google Scholar] [CrossRef]
  29. Sung, H.J.; Khan, M.F.; Kim, Y.H. Recombinant lignin peroxidase-catalysed decolourization of melanin using in situ generated H₂O₂ for application in whitening cosmetics. Int. J. Biol. Macromol. 2019, 136, 20–26. [Google Scholar] [CrossRef]
  30. Anastasi, A.; Tigini, V.; Varese, G.C. The bioremediation potential of different ecophysiological groups of fungi. In Fungi as Bioremediators; Goltapeh, E., Danesh, Y., Varma, A., Eds.; Springer: Berlin/Heidelberg, Germany, 2012; pp. 29–49. [Google Scholar]
  31. Corso, C.R.; Almeida, E.J.R.; Santos, G.C.; Morão, L.G.; Fabris, G.S.L.; Mitter, E.K. Bioremediation of direct dyes in simulated textile effluents by a paramorphogenic form of Aspergillus oryzae. Water Sci. Technol. 2012, 65, 1490–1495. [Google Scholar] [CrossRef] [PubMed]
  32. Khan, M.F.; Hof, C.; Niemcová, P.; Murphy, C.D. Recent advances in fungal xenobiotic metabolism: Enzymes and applications. World J. Microbiol. Biotechnol. 2023, 39, 296. [Google Scholar] [CrossRef]
  33. Rao, P.; Prathibha, N.; Birawat, K.K.; Kankrej, S.; Nayak, S.; Varsha, N. Decolourisation of synthetic dyes using Aspergillus species. J. Chem. Eng. Res. 2014, 2, 61–68. [Google Scholar]
  34. Aragaw, T.A. A review on biodegradation of textile dye wastewater: Challenges due to wastewater characteristics and the potential of alkaliphiles. J. Hazard. Mater. Adv. 2024, 5, 100117. [Google Scholar] [CrossRef]
  35. Mustafa, G.; Zahid, M.T.; Ali, S.; Abbas, S.Z.; Rafatullah, M. Biodegradation and discolouration of disperse blue-284 textile dye by Klebsiella pneumoniae GM-04 bacterial isolate. J. King Saud Univ.-Sci. 2021, 33, 101442. [Google Scholar] [CrossRef]
  36. Ameenudeen, S.; Unnikrishnan, S.; Ramalingam, K. Statistical optimization for the efficacious degradation of reactive azo dyes using Acinetobacter baumannii JC359. J. Environ. Manag. 2021, 279, 111512. [Google Scholar] [CrossRef]
  37. Kishor, R.; Saratale, G.D.; Saratale, R.G.; Ferreira, L.F.R.; Bilal, M.; Iqbal, H.M.; Bharagava, R.N. Efficient degradation and detoxification of methylene blue dye by a newly isolated ligninolytic enzyme producing bacterium Bacillus albus MW407057. Colloids Surf. B Biointerfaces 2021, 206, 111947. [Google Scholar] [CrossRef] [PubMed]
  38. Nguyen, T.A.; Fu, C.C.; Juang, R.S. Biosorption and biodegradation of a sulfur dye in high-strength dyeing wastewater by Acidithiobacillus thiooxidans. J. Environ. Manag. 2016, 182, 265–271. [Google Scholar] [CrossRef]
  39. Bera, S.P.; Tank, S.K. Microbial degradation of Procion Red by Pseudomonas stutzeri. Sci. Rep. 2021, 11, 3075. [Google Scholar] [CrossRef] [PubMed]
  40. Lade, H.; Kadam, A.; Paul, D.; Govindwar, S. Biodegradation and detoxification of textile azo dyes by bacterial consortium under sequential microaerophilic/aerobic processes. EXCLI J. 2015, 14, 158. [Google Scholar]
  41. HP, J.S.P.; Girish, K.; Agsar, D. Optimization of process conditions for the effective biodegradation of azo orange dye by actinomycetes. Indian J. Nat. Sci. 2015, 5, 5248–5258. [Google Scholar]
  42. Bharagava, R.N.; Mani, S.; Mulla, S.I.; Saratale, G.D. Degradation and decolourization potential of a ligninolytic enzyme producing Aeromonas hydrophila for crystal violet dye and its phytotoxicity evaluation. Ecotoxicol. Environ. Saf. 2018, 156, 166–175. [Google Scholar] [CrossRef]
  43. Sharma, N.; Chatterjee, S.; Bhatnagar, P. Degradation of Direct Red 28 by Alcaligenes sp. TEX S6 isolated from aeration tank of Common Effluent Treatment Plant (CETP), Pali, Rajasthan. Nat. Environ. Pollut. Technol. 2019, 18, 9–20. [Google Scholar]
  44. Oturkar, C.C.; Patole, M.S.; Gawai, K.R.; Madamwar, D. Enzyme-based cleavage strategy of Bacillus lentus BI377 in response to metabolism of azoic recalcitrant. Bioresour. Technol. 2013, 130, 360–365. [Google Scholar] [CrossRef]
  45. Ewida, A.Y.; El-Sesy, M.E.; Abou Zeid, A. Complete degradation of azo dye acid red 337 by Bacillus megaterium KY848339.1 isolated from textile wastewater. Water Sci. 2019, 33, 154–161. [Google Scholar] [CrossRef]
  46. Amin, S.; Rastogi, R.P.; Chaubey, M.G.; Jain, K.; Divecha, J.; Desai, C.; Madamwar, D. Degradation and toxicity analysis of a reactive textile diazo dye-Direct Red 81 by newly isolated Bacillus sp. DMS2. Front. Microbiol. 2020, 11, 576680. [Google Scholar] [CrossRef]
  47. Dixit, S.; Garg, S. Enzymatic degradation of sulphonated azo dye using purified azoreductase from facultative Klebsiella pneumoniae. Folia Microbiol. 2021, 66, 79–85. [Google Scholar] [CrossRef] [PubMed]
  48. Prasad, A.A.; Satyanarayana, V.S.V.; Rao, K.B. Biotransformation of Direct Blue 1 by a moderately halophilic bacterium Marinobacter sp. strain HBRA and toxicity assessment of degraded metabolites. J. Hazard. Mater. 2013, 262, 674–684. [Google Scholar] [CrossRef]
  49. Imran, M.; Arshad, M.; Khalid, A.; Hussain, S.; Mumtaz, M.W.; Crowley, D.E. Decolourization of Reactive Black-5 by Shewanella sp. in the presence of metal ions and salts. Water Environ. Res. 2015, 87, 579–586. [Google Scholar] [CrossRef]
  50. Yu, L.; Zhang, X.Y.; Tang, Q.W.; Li, J.; Xie, T.; Liu, C.; Cao, M.Y.; Zhang, R.C.; Wang, S.; Hu, J.M.; et al. Decolourization characteristics of a newly isolated salt-tolerant Bacillus sp. strain and its application for azo dye-containing wastewater in immobilised form. Appl. Microbiol. Biotechnol. 2015, 99, 9277–9287. [Google Scholar] [CrossRef] [PubMed]
  51. Chen, Y.; Feng, L.; Li, H.; Wang, Y.; Chen, G.; Zhang, Q. Biodegradation and detoxification of Direct Black G textile dye by a newly isolated thermophilic microflora. Bioresour. Technol. 2018, 250, 650–657. [Google Scholar] [CrossRef]
  52. Rathod, J.; Dhebar, S.; Archana, G. Efficient approach to enhance whole cell azo dye decolourization by heterologous overexpression of Enterococcus sp. L2 azoreductase (azoA) and Mycobacterium vaccae formate dehydrogenase (fdh) in different bacterial systems. Int. Biodeterior. Biodegrad. 2017, 124, 91–100. [Google Scholar] [CrossRef]
  53. Balapure, K.; Aghera, P.; Bhatt, N.; Madamwar, D. Community synergism: Degradation of triazine dye reactive black 1 by mixed bacterial cultures KND_PR under microaerophilic and aerobic conditions. Environ. Process. 2019, 6, 713–739. [Google Scholar] [CrossRef]
  54. Touliabah, H.E.S.; El-Sheekh, M.M.; Ismail, M.M.; El-Kassas, H. A review of microalgae- and cyanobacteria-based biodegradation of organic pollutants. Molecules 2022, 27, 1141. [Google Scholar] [CrossRef]
  55. Alvarez, M.S.; Rodriguez, A.; Sanroman, M.A.; Deive, F.J. Microbial adaptation to ionic liquids. RSC Adv. 2015, 5, 17379–17382. [Google Scholar] [CrossRef]
  56. Devi, S.; Murugappan, A.; Kannan, R.R. Sorption of Reactive Blue 19 onto freshwater algae and seaweed. Desalination Water Treat. 2015, 54, 2611–2624. [Google Scholar] [CrossRef]
  57. Thirumagal, J.; Panneerselvam, A. Isolation of azoreductase enzyme in its various forms from Chlorella pyrenoidosa and its immobilisation efficiency for treatment of water. Int. J. Sci. Res. 2016, 5, 2133–2138. [Google Scholar]
  58. Dellamatrice, P.M.; Silva-Stenico, M.E.; Moraes, L.A.B.D.; Fiore, M.F.; Monteiro, R.T.R. Degradation of textile dyes by cyanobacteria. Braz. J. Microbiol. 2017, 48, 25–31. [Google Scholar] [CrossRef] [PubMed]
  59. Tripathi, M.; Sharma, M.; Bala, S.; Thakur, V.K.; Singh, A.; Dashora, K.; Hart, P.; Gupta, V.K. Recent technologies for transforming textile waste into value-added products: A review. Curr. Res. Biotechnol. 2024, 7, 100225. [Google Scholar] [CrossRef]
  60. Ahmed, T.; Mia, R.; Toki, G.F.I.; Jahan, J.; Hasan, M.M.; Tasin, M.A.S.; Farsee, M.S.; Ahmed, S. Evaluation of sizing parameters on cotton using the modified sizing agent. Clean. Eng. Technol. 2021, 5, 100320. [Google Scholar] [CrossRef]
  61. Colombi, B.L.; Valle, R.D.; Valle, J.A.; Andreaus, J. Advances in sustainable enzymatic scouring of cotton textiles: Evaluation of different post-treatments to improve fabric wettability. Clean. Eng. Technol. 2021, 4, 100160. [Google Scholar] [CrossRef]
  62. Islam, M.T.; Huda, S.Z.; Alam, M.S.; Sahariar, M.F. Single-bath-single-stage enzymatic treatment of denim. Results Eng. 2024, 21, 101944. [Google Scholar] [CrossRef]
  63. Farooq, A.; Ali, S.; Abbas, N.; Fatima, G.A.; Ashraf, M.A. Comparative performance evaluation of conventional bleaching and enzymatic bleaching with glucose oxidase on knitted cotton fabric. J. Clean. Prod. 2013, 42, 167–171. [Google Scholar] [CrossRef]
  64. Aragaw, T.A.; Bogale, F.M.; Tesfaye, E.L. Oxidative ligninolytic enzymes and their role in textile dye biodegradation: A comprehensive review. Water Pract. Technol. 2024, 19, 3598–3630. [Google Scholar] [CrossRef]
  65. Paar, A.; Costa, S.; Tzanov, T.; Gudelj, M.; Robra, K.H.; Cavaco-Paulo, A.; Gübitz, G.M. Thermo-alkali-stable catalases from newly isolated Bacillus sp. for the treatment and recycling of textile bleaching effluents. J. Biotechnol. 2001, 89, 147–153. [Google Scholar] [CrossRef]
  66. Gautam, R.L.; Bharadwaj, A.K.; Kumar, S.; Naraian, R. Microbial enzymes for the variable applications of textile industry processing. In Valorization of Biomass to Bioproducts; Vijai, K.G., Ed.; Elsevier: Amsterdam, The Netherlands, 2023; Volume 3, pp. 297–321. [Google Scholar]
  67. Ramasubbu, N.; Paloth, V.; Luo, Y.; Brayer, G.D.; Levine, M.J. Structure of human salivary α-amylase at 1.6 Å resolution: Implications for its role in the oral cavity. Acta Crystallogr. Sect. D Biol. Crystallogr. 1996, 52, 435–446. [Google Scholar] [CrossRef]
  68. Far, B.E.; Ahmadi, Y.; Khosroshahi, A.Y.; Dilmaghani, A. Microbial alpha-amylase production: Progress, challenges and perspectives. Adv. Pharm. Bull. 2020, 10, 350. [Google Scholar]
  69. Ali, Z.; Abdullah, M.; Yasin, M.T.; Amanat, K.; Sultan, M.; Rahim, A.; Sarwar, F. Recent Trends in Production and potential applications of microbial amylases: A comprehensive review. Protein Expr. Purif. 2024, 227, 106640. [Google Scholar] [CrossRef]
  70. Rehman, A.; Saeed, A.; Asad, W.; Khan, I.; Hayat, A.; Rehman, M.U.; Shah, T.A.; Sitotaw, B.; Dawoud, T.M.; Bourhia, M. Eco-friendly textile desizing with indigenously produced amylase from Bacillus cereus AS2. Sci. Rep. 2023, 13, 11991. [Google Scholar] [CrossRef] [PubMed]
  71. Singh, S.; Singh, S.; Bali, V.; Sharma, L.; Mangla, J. Production of fungal amylases using cheap, readily available agri-residues, for potential application in textile industry. BioMed Res. Int. 2014, 2014, 215748. [Google Scholar] [CrossRef]
  72. Hao, L.; Wang, R.; Fang, K.; Liu, J. Ultrasonic effect on the desizing efficiency of α-amylase on starch-sized cotton fabrics. Carbohydr. Polym. 2013, 96, 474–480. [Google Scholar] [CrossRef] [PubMed]
  73. Sajedi, R.H.; Naderi-Manesh, H.; Khajeh, K.; Ahmadvand, R.; Ranjbar, B.; Asoodeh, A.; Moradian, F. A Ca-independent α-amylase that is active and stable at low pH from the Bacillus sp. KR-8104. Enzym. Microb. Technol. 2005, 36, 666–671. [Google Scholar] [CrossRef]
  74. Chand, N.; Sajedi, R.H.; Nateri, A.S.; Khajeh, K.; Rassa, M. Fermentative desizing of cotton fabric using an α-amylase-producing Bacillus strain: Optimization of simultaneous enzyme production and desizing. Process Biochem. 2014, 49, 1884–1888. [Google Scholar] [CrossRef]
  75. Zafar, A.; Aftab, M.N.; Iqbal, I.; ud Din, Z.; Saleem, M.A. Pilot-scale production of a highly thermostable α-amylase enzyme from Thermotoga petrophila cloned into E. coli and its application as a desizer in textile industry. RSC Adv. 2019, 9, 984–992. [Google Scholar] [CrossRef]
  76. Garg, G.; Singh, A.; Kaur, A.; Singh, R.; Kaur, J.; Mahajan, R. Microbial pectinases: An ecofriendly tool of nature for industries. 3 Biotech 2016, 6, 47. [Google Scholar] [CrossRef]
  77. Madhu, A.; Chakraborty, J.N. Developments in application of enzymes for textile processing. J. Clean. Prod. 2017, 145, 114–133. [Google Scholar] [CrossRef]
  78. Haile, S.; Ayele, A. Pectinase from microorganisms and its industrial applications. Sci. World J. 2022, 2022, 1881305. [Google Scholar] [CrossRef] [PubMed]
  79. Amin, F.; Bhatti, H.N.; Bilal, M. Recent advances in the production strategies of microbial pectinases—A review. Int. J. Biol. Macromol. 2019, 122, 1017–1026. [Google Scholar] [CrossRef] [PubMed]
  80. Debing, J.; Peijun, L.; Stagnitti, F.; Xianzhe, X.; Li, L. Pectinase production by solid fermentation from Aspergillus niger by a new prescription experiment. Ecotoxicol. Environ. Saf. 2006, 64, 244–250. [Google Scholar] [CrossRef]
  81. Vigneswaran, C.; Anbumani, N.; Ananthasubramanian, M.; Rajendran, R. Prediction and process optimization of pectinolytic reaction on organic cotton fabrics for bioscouring with alkaline pectinase. Indian J. Fibre Text. Res. 2012, 37, 183–190. [Google Scholar]
  82. Rajendran, R.; Sundaram, S.K.; Radhai, R.; Rajapriya, P. Bioscouring of cotton fabrics using pectinase enzyme its optimization and comparison with conventional scouring process. Pak. J. Biol. Sci. 2011, 14, 519–525. [Google Scholar] [CrossRef]
  83. Guo, F.; Zou, M.; Li, X.; Zhao, J.; Qu, Y. An effective degumming enzyme from Bacillus sp. Y1 and synergistic action of hydrogen peroxide and protease on enzymatic degumming of ramie fibers. BioMed Res. Int. 2013, 2013, 212315. [Google Scholar] [CrossRef]
  84. Zheng, L.; Du, Y.; Zhang, J. Degumming of ramie fibers by alkalophilic bacteria and their polysaccharide-degrading enzymes. Bioresour. Technol. 2001, 78, 89–94. [Google Scholar] [CrossRef]
  85. Degani, O. Synergism between cutinase and pectinase in the hydrolysis of cotton fibers’ cuticle. Catalysts 2021, 11, 84. [Google Scholar] [CrossRef]
  86. Hebeish, A.; Hashem, M.; Shaker, N.; Ramadan, M.; El-Sadek, B.; Hady, M.A. New development for combined bioscouring and bleaching of cotton-based fabrics. Carbohydr. Polym. 2009, 78, 961–972. [Google Scholar] [CrossRef]
  87. Špička, N.; Tavčer, P.F. Complete enzymatic pre-treatment of cotton fabric with incorporated bleach activator. Text. Res. J. 2013, 83, 566–573. [Google Scholar] [CrossRef]
  88. Ejaz, U.; Sohail, M.; Ghanemi, A. Cellulases: From bioactivity to a variety of industrial applications. Biomimetics 2021, 6, 44. [Google Scholar] [CrossRef] [PubMed]
  89. Kuhad, R.C.; Gupta, R.; Singh, A. Microbial cellulases and their industrial applications. Enzym. Res. 2011, 2011, 280696. [Google Scholar] [CrossRef]
  90. Pei, J.; Pang, Q.; Zhao, L.; Fan, S.; Shi, H. Thermoanaerobacterium thermosaccharolyticum β-glucosidase: A glucose-tolerant enzyme with high specific activity for cellobiose. Biotechnol. Biofuels 2012, 5, 31. [Google Scholar] [CrossRef] [PubMed]
  91. Sutaoney, P.; Rai, S.N.; Sinha, S.; Choudhary, R.; Gupta, A.K.; Singh, S.K.; Banerjee, P. Current perspective in research and industrial applications of microbial cellulases. Int. J. Biol. Macromol. 2024, 264, 130639. [Google Scholar] [CrossRef]
  92. Elfaleh, I.; Abbassi, F.; Habibi, M.; Ahmad, F.; Guedri, M.; Nasri, M.; Garnier, C. A comprehensive review of natural fibers and their composites: An eco-friendly alternative to conventional materials. Results Eng. 2023, 19, 101271. [Google Scholar] [CrossRef]
  93. Niyonzima, F.N.; More, V.S.; Nsanganwimana, F.; Rao, A.S.; Nair, A.; Anantharaju, K.S.; More, S.S. Microbial enzymes used in textile industry. In Biotechnology of Microbial Enzymes, 2nd ed.; Academic Press: London, UK, 2023; pp. 649–684. [Google Scholar]
  94. Maryan, A.S.; Montazer, M. A cleaner production of denim garment using one step treatment with amylase/cellulase/laccase. J. Clean. Prod. 2013, 57, 320–326. [Google Scholar] [CrossRef]
  95. Jayasekara, S.; Ratnayake, R. Microbial cellulases: An overview and applications. Cellulose 2019, 22, 10–5772. [Google Scholar]
  96. Choudhury, A.K.R. Various ecofriendly finishes. Principles of Textile Finishing, 2nd ed.; Woodhead Publishing: Cambridge, UK, 2017; pp. 467–525. [Google Scholar]
  97. Lenin, V.; Kandasamy, N.; Karthick, S.; Kanipriya, M. Enzymes in Textile Finishing. Man-Made Text. India 2009, 52, 7. [Google Scholar]
  98. Araujo, R.; Casal, M.; Cavaco-Paulo, A. Application of enzymes for textile fibres processing. Biocatal. Biotransformation 2008, 26, 332–349. [Google Scholar] [CrossRef]
  99. Korsa, G.; Konwarh, R.; Masi, C.; Ayele, A.; Haile, S. Microbial cellulase production and its potential application for textile industries. Ann. Microbiol. 2023, 73, 13. [Google Scholar] [CrossRef]
  100. Walawska, A.; Olak-Kucharczyk, M.; Kaczmarek, A.; Kudzin, M.H. Environmentally friendly bleaching process of the cellulose fibres materials using ozone and hydrogen peroxide in the gas phase. Materials 2024, 17, 1355. [Google Scholar] [CrossRef] [PubMed]
  101. Singh, P.; Kumar, S. Microbial enzyme in food biotechnology. In Enzymes in Food Biotechnology, 2nd ed.; Academic Press: London, UK, 2019; pp. 19–28. [Google Scholar]
  102. Tochetto, G.A.; Aragão, A.M.; de Oliveira, D.; Immich, A.P. Can enzymatic processes transform textile processes? A critical analysis of the industrial application. Process Biochem. 2022, 123, 27–35. [Google Scholar] [CrossRef]
  103. Davulcu, A.; Eren, H.A.; Avinc, O.; Erişmiş, B. Ultrasound assisted biobleaching of cotton. Cellulose 2014, 21, 2973–2981. [Google Scholar] [CrossRef]
  104. Abou-Okeil, A.; El-Shafie, A.; El Zawahry, M.M. Ecofriendly laccase–hydrogen peroxide/ultrasound-assisted bleaching of linen fabrics and its influence on dyeing efficiency. Ultrason. Sonochem. 2010, 17, 383–390. [Google Scholar] [CrossRef] [PubMed]
  105. Saravanan, D.; Sivasaravanan, S.; Sudharshan Prabhu, M.; Vasanthi, N.S.; Senthil Raja, K.; Das, A.; Ramachandran, T. One-step process for desizing and bleaching of cotton fabrics using the combination of amylase and glucose oxidase enzymes. J. Appl. Polym. Sci. 2012, 123, 2445–2450. [Google Scholar] [CrossRef]
  106. Ofoedu, C.E.; You, L.; Osuji, C.M.; Iwouno, J.O.; Kabuo, N.O.; Ojukwu, M.; Agunwah, I.M.; Chacha, J.S.; Muobike, O.P.; Agunbiade, A.O.; et al. Hydrogen peroxide effects on natural-sourced polysaccharides: Free radical formation/production, degradation process, and reaction mechanism—A critical synopsis. Foods 2021, 10, 699. [Google Scholar] [CrossRef]
  107. Mojsov, K. Enzymatic desizing, bioscouring and enzymatic bleaching of cotton fabric with glucose oxidase. J. Text. Inst. 2019, 110, 1032–1041. [Google Scholar] [CrossRef]
  108. Konczewicz, W.; Kozłowski, R.M. Enzymatic treatment of natural fibres. In Handbook of Natural Fibres, 2nd ed.; Woodhead Publishing: Cambridge, UK, 2012; pp. 168–184. [Google Scholar]
  109. Fu, S.; Farrell, M.J.; Ankeny, M.A.; Turner, E.T.; Rizk, V. Hydrogen peroxide bleaching of cationized cotton fabric. AATCC J. Res. 2019, 6, 21–29. [Google Scholar] [CrossRef]
  110. Anwar, S.; Alrumaihi, F.; Sarwar, T.; Babiker, A.Y.; Khan, A.A.; Prabhu, S.V.; Rahmani, A.H. Exploring therapeutic potential of catalase: Strategies in disease prevention and management. Biomolecules 2024, 14, 697. [Google Scholar] [CrossRef]
  111. Sooch, B.S.; Kauldhar, B.S.; Puri, M. Isolation and polyphasic characterization of a novel hyper catalase producing thermophilic bacterium for the degradation of hydrogen peroxide. Bioprocess Biosyst. Eng. 2016, 39, 1759–1773. [Google Scholar] [CrossRef]
  112. Czyzewska, K.; Trusek-Holownia, A.; Dabrowa, M.; Sarmiento, F.; Blamey, J.M. A catalytic membrane used for H2O2 decomposition. Catal. Today 2019, 331, 30–34. [Google Scholar] [CrossRef]
  113. Shaeer, A.; Aslam, M.; Rashid, N. A highly stable manganese catalase from Geobacillus thermopakistaniensis: Molecular cloning and characterization. Extremophiles 2019, 23, 707–718. [Google Scholar] [CrossRef]
  114. Shaeer, A.; Aroob, I.; Aslam, M.; Azim, N.; Rashid, N. Investigating recombinant manganese-catalases from Geobacillus thermopakistaniensis for sustainable and eco-friendly textile processing. Int. J. Environ. Sci. Technol. 2024, 22, 6903–6912. [Google Scholar] [CrossRef]
  115. Lončar, N.; Fraaije, M.W. Catalases as biocatalysts in technical applications: Current state and perspectives. Appl. Microbiol. Biotechnol. 2015, 99, 3351–3357. [Google Scholar] [CrossRef]
  116. Singh, D.; Gupta, N. Microbial laccase: A robust enzyme and its industrial applications. Biologia 2020, 75, 1183–1193. [Google Scholar] [CrossRef]
  117. Mojsov, K. Biotechnological applications of laccases in the textile industry. Adv. Technol. 2014, 3, 76–79. [Google Scholar] [CrossRef]
  118. Kim, S.; Moldes, D.; Cavaco-Paulo, A. Laccases for enzymatic colouration of unbleached cotton. Enzym. Microb. Technol. 2007, 40, 1788–1793. [Google Scholar] [CrossRef]
  119. Ren, X.; Buschle-Diller, G. Oxidoreductases for modification of linen fibers. Colloids Surf. A Physicochem. Eng. Asp. 2007, 299, 15–21. [Google Scholar] [CrossRef]
  120. Tian, L.; Branford-White, C.; Wang, W.; Nie, H.; Zhu, L. Laccase-mediated system pretreatment to enhance the effect of hydrogen peroxide bleaching of cotton fabric. Int. J. Biol. Macromol. 2012, 50, 782–787. [Google Scholar] [CrossRef]
  121. Pereira, L.; Bastos, C.; Tzanov, T.; Cavaco-Paulo, A.; Guebitz, G.M. Environmentally friendly bleaching of cotton using laccases. Environ. Chem. Lett. 2005, 3, 66–69. [Google Scholar] [CrossRef]
  122. Neifar, M.; Chouchane, H.; Mahjoubi, M.; Jaouani, A.; Cherif, A. Pseudomonas extremorientalis BU118: A new salt-tolerant laccase-secreting bacterium with biotechnological potential in textile azo dye decolourization. 3 Biotech 2016, 6, 107. [Google Scholar] [CrossRef] [PubMed]
  123. Moya, R.; Hernández, M.; García-Martín, A.B.; Ball, A.S.; Arias, M.E. Contributions to a better comprehension of redox-mediated decolouration and detoxification of azo dyes by a laccase produced by Streptomyces cyaneus CECT 3335. Bioresour. Technol. 2010, 101, 2224–2229. [Google Scholar] [CrossRef]
  124. Jassal, S.; Warmoota, R.; Sharma, A.; Sheoran, S.; Kumar, D.; Gupta, N. An Economical Industrially Feasible Approach for Denim Biobleaching: Reusing of Laccase and Mediator for Multiple Cycles. Indian J. Microbiol. 2024, 1–13. [Google Scholar] [CrossRef]
  125. Ryan, S.; Schnitzhofer, W.; Tzanov, T.; Cavaco-Paulo, A.; Gübitz, G.M. An acid-stable laccase from Sclerotium rolfsii with potential for wool dye decolourization. Enzym. Microb. Technol. 2003, 33, 766–774. [Google Scholar] [CrossRef]
  126. Sellami, K.; Couvert, A.; Nasrallah, N.; Maachi, R.; Abouseoud, M.; Amrane, A. Peroxidase enzymes as green catalysts for bioremediation and biotechnological applications: A review. Sci. Total Environ. 2022, 806, 150500. [Google Scholar] [CrossRef] [PubMed]
  127. Gan, J.; Bilal, M.; Li, X.; Shah, S.Z.; Mohamed, B.A.; Hadibarata, T.; Cheng, H. Peroxidases-based enticing biotechnological platforms for biodegradation and biotransformation of emerging contaminants. Chemosphere 2022, 307, 136035. [Google Scholar] [CrossRef]
  128. Catucci, G.; Valetti, F.; Sadeghi, S.J.; Gilardi, G. Biochemical features of dye-decolorizing peroxidases: Current impact on lignin degradation. Biotechnol. Appl. Biochem. 2020, 67, 751–759. [Google Scholar] [CrossRef]
  129. Khan, S.; Borah, D. Microbial cell factories in the degradation of azo-dye and their limiting factors: An insight. Clean. Water 2024, 2024, 100034. [Google Scholar] [CrossRef]
  130. Kumar, V.; Pallavi, P.; Sen, S.K.; Raut, S. Harnessing the potential of white rot fungi and ligninolytic enzymes for efficient textile dye degradation: A comprehensive review. Water Environ. Res. 2024, 96, e10959. [Google Scholar] [CrossRef]
  131. Ren, J.; Li, X.; Zhang, W.; Li, Z.; Wang, Q.; Li, S.; Wang, S.; Li, H. Evaluation of application potential of dye-decolourizing peroxidase from Bacillus amyloliquefaciens in bioremediation of paper and pulp mill effluent. Front. Microbiol. 2022, 13, 1031853. [Google Scholar] [CrossRef]
  132. Ilić Đurđić, K.; Ostafe, R.; Prodanović, O.; Đurđević Đelmaš, A.; Popović, N.; Fischer, R.; Schillberg, S.; Prodanović, R. Improved degradation of azo dyes by lignin peroxidase following mutagenesis at two sites near the catalytic pocket and the application of peroxidase-coated yeast cell walls. Front. Environ. Sci. Eng. 2021, 15, 19. [Google Scholar] [CrossRef]
  133. Saha, P.; Sivaramakrishna, A.; Rao, K.V.B. Bioremediation of reactive orange 16 by industrial effluent-adapted bacterial consortium VITPBC6: Process optimization using response surface methodology (RSM), enzyme kinetics, pathway elucidation, and detoxification. Environ. Sci. Pollut. Res. 2023, 30, 35450–35477. [Google Scholar] [CrossRef]
  134. Hossain, M.S.; Paul, G.K.; Mahmud, S.; Saleh, M.A.; Uddin, M.S.; Dutta, A.K.; Roy, A.K.; Saha, A.K.; Sheam, M.M.; Ahmed, S.; et al. Mixed dye degradation by Bacillus pseudomycoides and Acinetobacter haemolyticus isolated from industrial effluents: A combined affirmation with wetlab and in silico studies. Arab. J. Chem. 2022, 15, 104078. [Google Scholar] [CrossRef]
  135. Goswami, D.; Mukherjee, J.; Mondal, C.; Bhunia, B. Bioremediation of azo dye: A review on strategies, toxicity assessment, mechanisms, bottlenecks and prospects. Sci. Total Environ. 2024, 2024, 176426. [Google Scholar] [CrossRef] [PubMed]
  136. Chahiniana, H.; Sarda, L. Distinction between esterases and lipases: Comparative biochemical properties of sequence-related carboxylesterases. Protein Pept. Lett. 2009, 16, 1149–1161. [Google Scholar] [CrossRef]
  137. El-Shemy, N.S.; El-Hawary, N.S.; El-Sayed, H. Basic and reactive-dyeable polyester fabrics using lipase enzymes. J. Chem. Eng. Process Technol. 2016, 7, 1000271. [Google Scholar] [CrossRef]
  138. Kalantzi, S.; Mamma, D.; Kalogeris, E.; Kekos, D. Improved properties of cotton fabrics treated with lipase and its combination with pectinase. Fibres Text. East. Eur. 2010, 18, 86–92. [Google Scholar]
  139. Khan, M.F.; Kundu, D.; Hazra, C.; Patra, S. A strategic approach of enzyme engineering by attribute ranking and enzyme immobilisation on zinc oxide nanoparticles to attain thermostability in mesophilic Bacillus subtilis lipase for detergent formulation. Int. J. Biol. Macromol. 2019, 136, 66–82. [Google Scholar] [CrossRef]
  140. Buchert, J.; Pere, J.; Puolakka, A.; Nousiainen, P. Scouring of cotton with pectinases, proteases, and lipases. Text. Chem. Colour. Am. Dyest. Rep. 2000, 32, 48. [Google Scholar]
  141. Park, Y.J.; Yoon, S.J.; Lee, H.B. A novel thermostable arylesterase from the archaeon Sulfolobus solfataricus P1: Purification, characterization, and expression. J. Bacteriol. 2008, 190, 8086–8095. [Google Scholar] [CrossRef]
  142. Auterinen, A.L.; Prozzo, B.; Redling, E.; Vermeersch, L.; Yoon, M.Y. Enzymatic Textile Bleaching Compositions and Methods of Use Thereof. U.S. Patent Application 14/220,436, 2014. [Google Scholar]
  143. Spicka, N.; Tavcer, P.F. New combined bio-scouring and bio-bleaching process of cotton fabrics. Mater. Technol. 2013, 47, 409–412. [Google Scholar]
  144. Siddiquee, A.B.; Bashar, M.; Sarker, P.; Tohfa, T.T.; Hossan, M.A.; Azad, M.I.; Akhtar, N. Comparative study of conventional and enzymatic pretreatment (scouring and bleaching) of cotton knitted fabrics. Int. J. Eng. Technol. 2014, 3, 37–43. [Google Scholar]
  145. Kawai, F.; Kawabata, T.; Oda, M. Current state and perspectives related to the polyethylene terephthalate hydrolases available for biorecycling. ACS Sustain. Chem. Eng. 2020, 8, 8894–8908. [Google Scholar] [CrossRef]
  146. Carr, C.M.; Clarke, D.J.; Dobson, A.D. Microbial polyethylene terephthalate hydrolases: Current and future perspectives. Front. Microbiol. 2020, 11, 571265. [Google Scholar] [CrossRef]
  147. Al-Ghanayem, A.A.; Joseph, B.; Alhussaini, M.S.; Ramteke, P.W. Current applications and future trends of extremozymes in detergent industries. In Microbial Extremozymes; Kuddus, M., Ed.; Academic Press: Cambridge, MA, USA, 2022; pp. 223–230. [Google Scholar] [CrossRef]
  148. Shi, K.; Jing, J.; Song, L.; Su, T.; Wang, Z. Enzymatic hydrolysis of polyester: Degradation of poly (ε-caprolactone) by Candida antarctica lipase and Fusarium solani cutinase. Int. J. Biol. Macromol. 2020, 144, 183–189. [Google Scholar] [CrossRef] [PubMed]
  149. Carniel, A.; Valoni, É.; Junior, J.N.; da Conceição Gomes, A.; De Castro, A.M. Lipase from Candida antarctica (CALB) and cutinase from Humicola insolens act synergistically for PET hydrolysis to terephthalic acid. Process Biochem. 2017, 59, 84–90. [Google Scholar] [CrossRef]
  150. Lange, L.; Huang, Y.; Busk, P.K. Microbial decomposition of keratin in nature—A new hypothesis of industrial relevance. Appl. Microbiol. Biotechnol. 2016, 100, 2083–2096. [Google Scholar] [CrossRef]
  151. Danilova, I.; Sharipova, M. The practical potential of bacilli and their enzymes for industrial production. Front. Microbiol. 2020, 11, 1782. [Google Scholar] [CrossRef]
  152. Demirkan, E.; Kut, D.; Sevgi, T.; Dogan, M.; Baygin, E. Investigation of effects of protease enzyme produced by Bacillus subtilis 168 E6-5 and commercial enzyme on physical properties of woolen fabric. J. Text. Inst. 2020, 111, 26–35. [Google Scholar] [CrossRef]
  153. El-Sayed, H.; Mowafi, S.; El-Fiky, A.F.; Khalil, E.M. Low temperature water-saving bio-degumming of natural silk using thermophilic protease. Sustain. Chem. Pharm. 2022, 27, 100681. [Google Scholar] [CrossRef]
  154. Sreelakshmi, S.N.; Vasanthi, N.S.; Saravanan, D. Acidic bacterial proteases from Bacillus species as an alternative agent for scouring of cotton fabrics. J. Text. Inst. 2013, 104, 1118–1124. [Google Scholar] [CrossRef]
  155. Vigneswaran, C.; Ananthasubramanian, M.; Anbumani, N.; Kandhavadivu, P. Ecofriendly approach to improve pectinolytic reaction and process optimization of bioscouring of organic cotton textiles. J. Eng. Fibers Fabr. 2013, 8, 121–133. [Google Scholar] [CrossRef]
  156. Pooja, E.S.; Fatima, N. Quality improvement of wool fabric using protease enzyme. Environ. Ecol. Res. 2014, 2, 301–310. [Google Scholar]
  157. Ndochinwa, O.G.; Wang, Q.Y.; Amadi, O.C.; Nwagu, T.N.; Nnamchi, C.I.; Okeke, E.S.; Moneke, A.N. Current status and emerging frontiers in enzyme engineering: An industrial perspective. Heliyon 2024, 10, e32673. [Google Scholar] [CrossRef]
  158. Morshed, M.N.; Behary, N.; Bouazizi, N.; Guan, J.; Nierstrasz, V.A. An overview on biocatalysts immobilisation on textiles: Preparation, progress and application in wastewater treatment. Chemosphere 2021, 279, 130481. [Google Scholar] [CrossRef] [PubMed]
  159. Khan, M.F.; Kundu, D.; Gogoi, M.; Shrestha, A.K.; Karanth, N.G.; Patra, S. Enzyme-responsive and enzyme immobilised nanoplatforms for therapeutic delivery: An overview of research innovations and biomedical applications. Nanopharm. Princ. Appl. 2020, 3, 165–200. [Google Scholar]
  160. Kakkar, P.; Wadhwa, N. Extremozymes used in textile industry. J. Text. Inst. 2021, 113, 2007–2015. [Google Scholar] [CrossRef]
  161. Pouresmaeil, M.; Azizi-Dargahlou, S. Factors involved in heterologous expression of proteins in E. coli host. Arch. Microbiol. 2023, 205, 212. [Google Scholar] [CrossRef]
  162. Vieira Gomes, A.M.; Souza Carmo, T.; Silva Carvalho, L.; Mendonça Bahia, F.; Parachin, N.S. Comparison of yeasts as hosts for recombinant protein production. Microorganisms 2018, 6, 38. [Google Scholar] [CrossRef]
  163. Nevalainen, H.; Peterson, R. Making recombinant proteins in filamentous fungi—Are we expecting too much? Front. Microbiol. 2014, 5, 75. [Google Scholar]
  164. Salwoom, L.; Raja Abd. Rahman, R.N.Z.; Salleh, A.B.; Mohd. Shariff, F.; Convey, P.; Mohamad Ali, M.S. New recombinant cold-adapted and organic solvent tolerant lipase from psychrophilic Pseudomonas sp. LSK25, isolated from Signy Island Antarctica. Int. J. Mol. Sci. 2019, 20, 1264. [Google Scholar] [CrossRef] [PubMed]
  165. Joshi, N.; Sharma, M.; Singh, S.P. Characterization of a novel xylanase from an extreme temperature hot spring metagenome for xylooligosaccharide production. Appl. Microbiol. Biotechnol. 2020, 104, 4889–4901. [Google Scholar] [CrossRef] [PubMed]
  166. Gil-Durán, C.; Ravanal, M.C.; Ubilla, P.; Vaca, I.; Chávez, R. Heterologous expression, purification and characterization of a highly thermolabile endoxylanase from the Antarctic fungus Cladosporium sp. Fungal Biol. 2018, 122, 875–882. [Google Scholar] [CrossRef] [PubMed]
  167. Boyce, A.; Walsh, G. Expression and characterisation of a thermophilic endo-1,4-β-glucanase from Sulfolobus shibatae of potential industrial application. Mol. Biol. Rep. 2018, 45, 2201–2211. [Google Scholar] [CrossRef]
  168. Yu, Z.; Zheng, H.; Zhao, X.; Li, S.; Xu, J.; Song, H. High level extracellular production of a recombinant alkaline catalase in E. coli BL21 under ethanol stress and its application in hydrogen peroxide removal after cotton fabrics bleaching. Bioresour. Technol. 2016, 214, 303–310. [Google Scholar] [CrossRef]
  169. Chatterjee, A.; Puri, S.; Sharma, P.K.; Deepa, P.R.; Chowdhury, S. Nature-inspired enzyme engineering and sustainable catalysis: Biochemical clues from the world of plants and extremophiles. Front. Bioeng. Biotechnol. 2023, 11, 1229300. [Google Scholar] [CrossRef]
  170. Rigoldi, F.; Donini, S.; Redaelli, A.; Parisini, E.; Gautieri, A. Engineering of thermostable enzymes for industrial applications. APL Bioeng. 2018, 2, 011501. [Google Scholar] [CrossRef]
  171. Shrestha, S.; Chio, C.; Khatiwada, J.R.; Mokale Kognou, A.L.; Chen, X.; Qin, W. Optimization of cultural conditions for pectinase production by Streptomyces sp. and characterization of partially purified enzymes. Microb. Physiol. 2023, 33, 12–26. [Google Scholar] [CrossRef]
  172. Solbak, A.I.; Richardson, T.H.; McCann, R.T.; Kline, K.A.; Bartnek, F.; Tomlinson, G.; Tan, X.; Parra-Gessert, L.; Frey, G.J.; Podar, M.; et al. Discovery of pectin-degrading enzymes and directed evolution of a novel pectate lyase for processing cotton fabric. J. Biol. Chem. 2005, 280, 9431–9438. [Google Scholar] [CrossRef]
  173. Li, P.; Wei, X.; Wang, Y.; Liu, H.; Xu, Y.; Zhang, Z.; Li, J.; Wang, J.; Guo, C.; Sui, S.; et al. Improvement of optimum pH and specific activity of pectate lyase from Bacillus RN. 1 using loop replacement. Front. Bioeng. Biotechnol. 2023, 11, 1242123. [Google Scholar] [CrossRef]
  174. Mohamad, N.R.; Marzuki, N.H.C.; Buang, N.A.; Huyop, F.; Wahab, R.A. An overview of technologies for immobilisation of enzymes and surface analysis techniques for immobilised enzymes. Biotechnol. Biotechnol. Equip. 2015, 29, 205–220. [Google Scholar] [CrossRef] [PubMed]
  175. Kundu, D.; Khan, M.F.; Gogoi, M.; Patra, S. Environmental impact and econanotoxicity of engineered nanomaterials. In Nanotoxicology and Nanoecotoxicology; Kumar, V., Guleria, P., Ranjan, S., Dasgupta, N., Lichtfouse, E., Eds.; Springer: Cham, Switzerland, 2021; Volume 1, pp. 287–312. [Google Scholar]
  176. Khan, M.F.; Murphy, C.D. Environmental remediation by novel nanomaterials and fungi with high-degradation capacity of hazardous contaminants. In Bio and Nanoremediation of Hazardous Environmental Pollutants; Fernández-Luqueño, F., López-Valdez, F., Medina-Pérez, G., Eds.; CRC Press: Boca Raton, FL, USA, 2023; pp. 283–310. [Google Scholar]
  177. Guzik, U.; Hupert-Kocurek, K.; Wojcieszyńska, D. Immobilization as a strategy for improving enzyme properties—Application to oxidoreductases. Molecules 2014, 19, 8995–9018. [Google Scholar] [CrossRef]
  178. Costa, S.A.; Tzanov, T.; Paar, A.; Gudelj, M.; Gübitz, G.M.; Cavaco-Paulo, A. Immobilization of catalases from Bacillus SF on alumina for the treatment of textile bleaching effluents. Enzym. Microb. Technol. 2001, 28, 815–819. [Google Scholar] [CrossRef]
  179. Costa, S.A.; Tzanov, T.; Carneiro, F.; Gübitz, G.M.; Cavaco-Paulo, A. Recycling of textile bleaching effluents for dyeing using immobilized catalase. Biotechnol. Lett. 2002, 24, 173–176. [Google Scholar] [CrossRef]
  180. Narayanan, M.P.; Murugan, S.; Eva, A.S.; Devina, S.U.; Kalidass, S. Application of immobilized laccase from Bacillus subtilis MTCC 2414 on decolourization of synthetic dyes. Res. J. Microbiol. 2015, 10, 421–432. [Google Scholar]
  181. Sankarraj, N.; Nallathambi, G. Enzymatic biopolishing of cotton fabric with free/immobilized cellulase. Carbohydr. Polym. 2018, 191, 95–102. [Google Scholar] [CrossRef] [PubMed]
  182. Kumar, V.S.; Meenakshisundaram, S.; Selvakumar, N. Conservation of cellulase enzyme in biopolishing application of cotton fabrics. J. Text. Inst. 2008, 99, 339–346. [Google Scholar] [CrossRef]
  183. Dinçer, A.; Telefoncu, A. Improving the stability of cellulase by immobilization on modified polyvinyl alcohol coated chitosan beads. J. Mol. Catal. B Enzym. 2007, 45, 10–14. [Google Scholar] [CrossRef]
  184. Antecka, K.; Zdarta, J.; Siwińska-Stefańska, K.; Sztuk, G.; Jankowska, E.; Oleskowicz-Popiel, P.; Jesionowski, T. Synergistic degradation of dye wastewaters using binary or ternary oxide systems with immobilized laccase. Catalysts 2018, 8, 402. [Google Scholar] [CrossRef]
  185. de Souza Lima, J.; Immich, A.P.S.; de Araújo, P.H.H.; de Oliveira, D. Cellulase immobilized on kaolin as a potential approach to improve the quality of knitted fabric. Bioprocess Bioproc. Eng. 2022, 45, 679–688. [Google Scholar] [CrossRef]
  186. Cristóvão, R.O.; Silvério, S.C.; Tavares, A.P.; Brígida, A.I.S.; Loureiro, J.M.; Boaventura, R.A.; Macedo, E.A.; Coelho, M.A.Z. Green coconut fiber: A novel carrier for the immobilization of commercial laccase by covalent attachment for textile dyes decolourization. World J. Microbiol. Biotechnol. 2012, 28, 2827–2838. [Google Scholar] [CrossRef]
  187. Sondhi, S.; Sharma, P.; Saini, S.; Puri, N.; Gupta, N. Purification and characterization of an extracellular, thermo-alkali-stable, metal tolerant laccase from Bacillus tequilensis SN4. PLoS ONE 2014, 9, e96951. [Google Scholar] [CrossRef] [PubMed]
  188. Siddeeg, S.M.; Tahoon, M.A.; Mnif, W.; Ben Rebah, F. Iron oxide/chitosan magnetic nanocomposite immobilized manganese peroxidase for decolourization of textile wastewater. Processes 2019, 8, 5. [Google Scholar] [CrossRef]
  189. Silva, C.J.; Gübitz, G.; Cavaco-Paulo, A. Optimisation of a serine protease coupling to Eudragit S-100 by experimental design techniques. J. Chem. Technol. Biotechnol. 2006, 81, 8–16. [Google Scholar] [CrossRef]
  190. Srivastava, B.; Singh, H.; Khatri, M.; Singh, G.; Arya, S.K. Immobilization of keratinase on chitosan grafted-β-cyclodextrin for the improvement of the enzyme properties and application of free keratinase in the textile industry. Int. J. Biol. Macromol. 2020, 165, 1099–1110. [Google Scholar] [CrossRef]
  191. Bayramoglu, G.; Gursel, I.; Yilmaz, M.; Arica, M.Y. Immobilization of laccase on itaconic acid grafted and Cu (II) ion chelated chitosan membrane for bioremediation of hazardous materials. J. Chem. Technol. Biotechnol. 2012, 87, 530–539. [Google Scholar] [CrossRef]
  192. Bilal, M.; Asgher, M.; Iqbal, M.; Hu, H.; Zhang, X. Chitosan beads immobilized manganese peroxidase catalytic potential for detoxification and decolourization of textile effluent. Int. J. Biol. Macromol. 2016, 89, 181–189. [Google Scholar] [CrossRef]
  193. Sondhi, S.; Kaur, R.; Kaur, S.; Kaur, P.S. Immobilization of laccase-ABTS system for the development of a continuous flow packed bed bioreactor for decolourization of textile effluent. Int. J. Biol. Macromol. 2018, 117, 1093–1100. [Google Scholar] [CrossRef]
  194. Yu, Y.; Yuan, J.; Wang, Q.; Fan, X.; Wang, P.; Sun, X. Immobilization of cellulases on the reversibly soluble polymer Eudragit S-100 for cotton treatment. Eng. Life Sci. 2013, 13, 194–200. [Google Scholar] [CrossRef]
  195. Chakravorty, D.; Khan, M.F.; Patra, S. Multifactorial level of extremostability of proteins: Can they be exploited for protein engineering? Extremophiles 2017, 21, 419–444. [Google Scholar] [CrossRef]
  196. El-Sayed, H.; El-Fiky, A.F.; Mowafi, S. Extremozymes as future appropriate benign elements for eco-friendly wet processing of wool and silk. J. Nat. Fibers 2022, 19, 15035–15044. [Google Scholar] [CrossRef]
  197. Kiadehi, M.S.H.; Amoozegar, M.A.; Asad, S.; Siroosi, M. Exploring the potential of halophilic archaea for the decolourization of azo dyes. Water Sci. Technol. 2018, 77, 1602–1611. [Google Scholar] [CrossRef]
  198. Kumari, A.; Kishor, N.; Guptasarma, P. Characterization of a mildly alkalophilic and thermostable recombinant Thermus thermophilus laccase with applications in decolourization of dyes. Biotechnol. Lett. 2018, 40, 285–295. [Google Scholar] [CrossRef]
  199. Saravanan, D.; Prakash, A.A.; Jagadeeshwaran, D.; Nalankilli, G.; Ramachandran, T.; Prabakaran, C. Optimization of thermophile Bacillus licheniformis -amylase desizing of cotton fabrics. Indian J. Fibre Text. Res. 2011, 36, 253–258. [Google Scholar]
  200. Aygan, A.; Arikan, B.; Korkmaz, H.; Dinçer, S.; Çolak, Ö. Highly thermostable and alkaline α-amylase from a halotolerant-alkaliphilic Bacillus sp. AB68. Braz. J. Microbiol. 2008, 39, 547–553. [Google Scholar] [CrossRef] [PubMed]
  201. Anish, R.; Rahman, M.S.; Rao, M. Application of cellulases from an alkalothermophilic Thermonospora sp. in biopolishing of denims. Biotechnol. Bioeng. 2007, 96, 48–56. [Google Scholar] [CrossRef]
  202. Wiedemann, S.G.; Clarke, S.J.; Nguyen, Q.V.; Cheah, Z.X.; Simmons, A.T. Strategies to reduce environmental impacts from textiles: Extending clothing wear life compared to fibre displacement assessed using consequential LCA. Resour. Conserv. Recycl. 2023, 198, 107119. [Google Scholar] [CrossRef]
  203. Toprak, T.; Anis, P. Textile industry’s environmental effects and approaching cleaner production and sustainability, an overview. J. Text. Eng. Fash. Technol. 2017, 2, 429–442. [Google Scholar] [CrossRef]
  204. Patti, A.; Cicala, G.; Acierno, D. Eco-sustainability of the textile production: Waste recovery and current recycling in the composites world. Polymers 2020, 13, 134. [Google Scholar] [CrossRef]
  205. Frondel, M.; Horbach, J.; Rennings, K. End-of-pipe or cleaner production? An empirical comparison of environmental innovation decisions across OECD countries. Bus. Strategy Environ. 2007, 16, 571–584. [Google Scholar] [CrossRef]
  206. Kumari, P.; Singh, S.J.; Rose, N.M. Eco–textiles: For sustainable development. Int. J. Sci. Eng. Res. 2013, 4, 1379–1390. [Google Scholar]
  207. Panda, S.K.B.C.; Sen, K.; Mukhopadhyay, S. Sustainable pretreatments in textile wet processing. J. Clean. Prod. 2021, 329, 129725. [Google Scholar] [CrossRef]
  208. Islam, S.; Jalil, M.A.; Belowar, S.; Saeed, M.A.; Hossain, S.; Rahamatolla, M.; Ali, S. Role of mordants in natural fabric dyeing and their environmental impacts. Environ. Sci. Pollut. Res. 2024, 32, 452–468. [Google Scholar] [CrossRef] [PubMed]
  209. Varjani, S.; Rakholiya, P.; Ng, H.Y.; You, S.; Teixeira, J.A. Microbial degradation of dyes: An overview. Bioresour. Technol. 2020, 314, 123728. [Google Scholar] [CrossRef]
  210. Pranta, A.D.; Rahaman, M.T. Extraction of eco-friendly natural dyes and biomordants for textile colorations: A critical review. Nano Struct. Nano Objects 2024, 39, 101243. [Google Scholar] [CrossRef]
  211. Asmelash, F.; Ayele, M. Beneficiation of Commiphora africana plant: Extraction and application of green softener on cotton fabric. J. Eng. 2021, 2021, 9910707. [Google Scholar] [CrossRef]
  212. Szostak-Kotowa, J. Biodeterioration of textiles. Int. Biodeterior. Biodegrad. 2004, 53, 165–170. [Google Scholar] [CrossRef]
  213. Egan, J.; Salmon, S. Strategies and progress in synthetic textile fiber biodegradability. SN Appl. Sci. 2022, 4, 22. [Google Scholar] [CrossRef]
  214. Wojnowska-Baryła, I.; Bernat, K.; Zaborowska, M. Strategies of recovery and organic recycling used in textile waste management. Int. J. Environ. Res. Public Health 2022, 19, 5859. [Google Scholar] [CrossRef]
  215. Shirvanimoghaddam, K.; Motamed, B.; Ramakrishna, S.; Naebe, M. Death by waste: Fashion and textile circular economy case. Sci. Total Environ. 2020, 718, 137317. [Google Scholar] [CrossRef]
  216. Suen, D.W.S.; Chan, E.M.H.; Lau, Y.Y.; Lee, R.H.P.; Tsang, P.W.K.; Ouyang, S.; Tsang, C.W. Sustainable textile raw materials: Review on bioprocessing of textile waste via electrospinning. Sustainability 2023, 15, 11638. [Google Scholar] [CrossRef]
  217. Liu, X.; Sathishkumar, K.; Zhang, H.; Saxena, K.K.; Zhang, F.; Naraginiti, S.; Anbarasu, K.; Rajendiran, R.; Aruliah, R.; Guo, X. Frontiers in environmental cleanup: Recent advances in remediation of emerging pollutants from soil and water. J. Hazard. Mater. Adv. 2024, 16, 100461. [Google Scholar] [CrossRef]
  218. Kabir, S.M.M.; Koh, J. Sustainable textile processing by enzyme applications. In Biodegradation Technology of Organic and Inorganic Pollutants; Mendes, K.F., de Sousa, R.N., Mielke, K.C., Eds.; IntechOpen: London, UK, 2021. [Google Scholar] [CrossRef]
  219. Chapman, J.; Ismail, A.E.; Dinu, C.Z. Industrial applications of enzymes: Recent advances, techniques, and outlooks. Catalysts 2018, 8, 238. [Google Scholar] [CrossRef]
  220. Choudhury, A.K.R. Introduction to enzymes. In Sustainable Technologies for Fashion and Textiles; Woodhead Publishing: Cambridge, UK, 2020; pp. 75–90. [Google Scholar]
  221. Vigneswaran, C.; Ananthasubramanian, M.; Kandhavadivu, P. Bioprocessing of Textiles; Woodhead Publishing India Pvt Limited: New Delhi, India, 2014. [Google Scholar]
  222. Fasim, A.; More, V.S.; More, S.S. Large-scale production of enzymes for biotechnology uses. Curr. Opin. Biotechnol. 2021, 69, 68–76. [Google Scholar] [CrossRef] [PubMed]
  223. Garcia, V.S.G.; Tominaga, F.K.; Rosa, J.M.; Borrely, S.I. Emerging pollutants in textile wastewater: An ecotoxicological assessment focusing on surfactants. Environ. Sci. Pollut. Res. 2024, 31, 27817–27828. [Google Scholar] [CrossRef]
  224. Robinson, P.K. Enzymes: Principles and biotechnological applications. Essays Biochem. 2015, 59, 1. [Google Scholar] [CrossRef]
Figure 1. Steps involved in textile processing.
Figure 1. Steps involved in textile processing.
Sci 07 00046 g001
Figure 2. Enzymes are involved in the processing of cotton, silk, wool, and synthetic fibre.
Figure 2. Enzymes are involved in the processing of cotton, silk, wool, and synthetic fibre.
Sci 07 00046 g002
Figure 3. (A) Enzyme engineering strategies; (B) enzyme immobilisation methods.
Figure 3. (A) Enzyme engineering strategies; (B) enzyme immobilisation methods.
Sci 07 00046 g003
Figure 4. Various approaches are used to reduce textile waste.
Figure 4. Various approaches are used to reduce textile waste.
Sci 07 00046 g004
Table 1. Application of various dye types for synthetic and natural fabrics.
Table 1. Application of various dye types for synthetic and natural fabrics.
Fabric TypeExampleDye type Used for Dyeing *
Synthetic fibres
PolyesterDacron, TeryleneDisperse, pigment
PolyamideNylon, Perlon, RilsanAcid, reactive, disperse, mordant, pigment
PolyacrylonitrileAcrilan, Courtelle, OrlonBasic, disperse, pigment
PolyolefinesMeraklon, ProleneDisperse
Polyvinyl chlorideEnvilon, ThermovylBasic, disperse
ElastomersGlospan, LycraAcid, disperse, reactive (wool), vat
Natural fibres
Silk-Azoic, basic, direct, oxidation, reactive, mordant, sulphur, vat
Wool and wool blendsWool–cotton, wool–viscose, etc.Acid, basic, reactive, mordant, solubilised vat
Cotton-Azoic, basic, direct, oxidation, reactive, mordant, sulphur, vat
Modified cellulose fibresViscose, secondary acetate, triacetateDisperse, direct, pigment, reactive, mordant, sulphur, vat, solubilised vat
BastLinen, flax, ramie, hemp, jute Acid, direct, reactive, disperse, vat, solubilised vat
* Textile dye types are classified by their chemical properties and fibre interactions. Acid dyes bond with cationic sites in protein fibres, while basic dyes bind to anionic sites in synthetic fibres, both offering vivid hues. Azoic dyes form directly in fibres for brilliant cellulose colours. Disperse dyes penetrate synthetic fibres through hydrophobic interactions, and direct dyes attach to cellulose via hydrogen bonding. Mordant dyes use metal salts for colour fastness, and pigments adhere with binders for surface applications. Reactive dyes form covalent bonds, sulphur dyes offer dark shades, vat dyes ensure durability, and solubilised vat dyes simplify application with enhanced fastness.
Table 2. Hazardous chemicals used in textile processes cause significant risks to both human health and the environment.
Table 2. Hazardous chemicals used in textile processes cause significant risks to both human health and the environment.
Textile ProcessToxic ChemicalsCAS NumberHazardous Causes
Yarn sizingPolyvinyl alcohol9002-89-5Eye irritation; discomfort in inhalation
Size preservativePentachlorophenol87-86-5Eye irritation; adverse neurological, blood, and liver effects
Formaldehyde50-00-0Skin and eye irritation; coughing; wheezing; nausea; burning sensations in the nose and throat
Coating and degreasingSynthetic non-biodegradable surfactants + solvents-Dermatological compatibility, toxicity, and biodegradability
Anti-caking agent in saltCyanide57-12-5Eyes, nose, and throat irritation; headache; prolonged exposure can cause coma and even death
Biocide on hosiery and fabricsTributyltin oxide56-35-9Extremely hazardous to aquatic lives
BleachingCalcium hypochlorite7778-54-3Coughing and breath shortness
Sodium hypochlorite7681-52-9Skin irritation; coughing; stomach and abdominal pain
Peroxide stabiliserSodium silicate1344-09-8Skin and eye irritation
Phosphorous-based compounds (Tris(2,4-di-tert-butylphenyl)phosphite)31570-04-4Causing explosion hazards
Detergents and emulsifiersNon-ionic surfactants (octylphenol ethoxylate)9036-19-5Skin irritation; environmental risks
Nonylphenol–ethylene oxide adducts (APEOs)26027-38-3Extremely toxic to aquatic organisms
Stain removersCarbon tetrachloride56-23-5Harmful effect on the liver, central nervous system, and kidneys; extended exposure can cause coma and fatality
SoftenerSilicones and amino silicones with APEO emulsifier63148-62-0; 63148-62-9Chronic respiratory effects; extremely toxic to aquatic organisms
DyesAzo dyes (Remazol Brilliant Blue R)12222-67-8Carcinogenic, skin irritation, aquatic toxicity, persistent in ecosystems
Reactive dyes (Procion blue)12236-36-4Skin irritation, allergic reactions, aquatic toxicity
Acid dyes (Cibacron fuchsia)6105-59-9Skin irritation, heavy metal contamination, water pollution
Basic dyes (crystal violet)548-62-9Dermatitis, respiratory irritation, aquatic toxicity
Disperse dyes (Disperse Blue 1)12222-72-5Allergic dermatitis, suspected carcinogen, water pollution
Sulphur dyes (sulphur black)1326-97-6Skin irritation, respiratory issues, aquatic toxicity
Vat dyes (indigo)482-89-3Respiratory irritation, water toxicity, persistent pollutants
Pigments (titanium dioxide)13463-67-7Heavy metal toxicity, carcinogenic, non-biodegradable
Heavy metalsCopper (Cu)7440-50-8Inhalation and respiratory problems; fever; nausea; vomiting
Lead (Pb)7439-92-1Affects the central nervous system, causing coma, convulsions, and death
Mercury (Hg)7439-97-6Adverse effects on the nervous, digestive, and immune systems
Cadmium (Cd)7440-43-9Harmful to lungs, bones, and kidneys
Chromium (Cr)7440-47-3Pulmonary sensitisation; causes lung, nasal, and sinus cancer; severe dermatitis and painless skin ulcers
Arsenic (As)7440-38-2Affects organs such as the eyes, skin, liver, kidneys, and lungs; causes cancer
Carriers in dyeingDichlorobenzene25321-22-6Kidney and liver cancer
Trichlorobenzene12002-48-1Skin, eye, nose, and throat irritation
Oxidation in dyeingSodium dichromate10588-01-9Causes asthma; damages the liver and kidneys
Pigment printing and dye-fixingKerosene8008-20-6Dizziness, headache, and vomiting
Formaldehyde50-00-0Skin and eye irritation; coughing; wheezing; nausea; burning sensations in the nose and throat
FinishingCationic surfactants (cetyltrimethylammonium bromide)57-09-0Skin irritation
Functional synthetic finish (perfluorooctanoic acid)335-67-1Environmental risks; toxic to aquatic organisms
Table 3. Comparison of fungi, bacteria, and algae in terms of efficiency, advantages, limitations, and applications for textile effluent bioremediation.
Table 3. Comparison of fungi, bacteria, and algae in terms of efficiency, advantages, limitations, and applications for textile effluent bioremediation.
Textile Effluent RemediationFungiBacteriaAlgae
Efficiency- High ability to degrade a wide range of dyes
- Effective in removing heavy metals such as chromium and copper
- High tolerance to toxic compounds
- Efficient in decolourising dyes and removing textile pollutants
- Faster growth rate compared to fungi
- Can be engineered for enhanced degradation
- Effective in removing dye pollutants through biosorption and bioaccumulation
- Can absorb heavy metals from effluent
Advantages- High degradation potential
- Efficient in breaking down complex dyes and pollutants
- Can be used in both solid-state and liquid-state remediation
- Fast growth rate and ability to adapt to changing conditions
- Can be used in both aerobic and anaerobic environments
- Lower nutrient requirements
- Easy to cultivate in large quantities
- Can be used in combination with other microorganisms
- Cost-effective
Limitations- Slow growth rate
- Requires specific conditions
- May need additional treatment to handle residual metabolites
- Sensitive to toxic concentrations of pollutants
- May require genetic modification for enhanced degradation
- Limited to certain types of dyes
- Requires light for photosynthesis
- Growth is sensitive to environmental changes
Applications- Bioremediation of textile wastewater
- Removal of toxic dyes and heavy metals from effluents
- Biodegradation of textile dyes and chemicals
- Treatment of wastewater with high concentrations of pollutants
- Bioremediation of dyes and heavy metals
- Integrated treatment systems with fungi and bacteria
Reference[21,23][24,25][24,26]
Table 4. Bacterial bioremediation of textile dyes.
Table 4. Bacterial bioremediation of textile dyes.
Bacterial SpeciesDye% DegradationDegradation Conditions (Dye conc., pH, Temp., Inc. Time)Enzyme InvolvedReference
Isolated single culture
Actinomycetes strainsOrange dye8550 mg/L, pH 7.2, 37 °C, 48 hNot identified[41]
Aeromonas hydrophilaCrystal violet99100 mg/L, pH 7, 35 °C, 8 hLaccase, lignin peroxidase[42]
Alcaligenes sp. TEX S6Direct Red 2886150 mg/L, pH 7, 37 °C, 48 hNot identified[43]
Bacillus lentus BI377Reactive Red 141, Reactive Red 299.11500 mg/L, pH 8, 40 °C, 6 hSuperoxide dismutase, peroxidase[44]
Bacillus megaterium KY848339Acid Red 33791500 mg/L, pH 7, 30 °C, 24 hAzoreductase[45]
Bacillus sp.Direct Red 81-100 mg/L, pH 7, 30 °C, 24 hAzoreductase, laccase[46]
Klebsiella pneumoniaeMethyl orange8320 μM, pH 8, 40 °C, 10 minAzoreductase[47]
Marinobacter sp. HBRADirect Blue 1100100 mg/L, pH 8, 37 °C, 6 hOxidases (not identified)[48]
Shewanella sp.Reactive Black-5, Direct Red-81, Acid Red-8896.9200 mg/L, pH 8.5, 35 °C, 12 hAzoreductase[49]
Bacillus sp. strain CICC 23870Methyl orange, Black 5, Acid Blue 113, methyl red97.8732.7 mg/L, pH 7, 35 °C, 24 hAzoreductase[50]
Mixed culture
Brevibacillus aydinogluensis, Geobacillus thermoleovorans, Anoxybacillus flavithermus, Bacillus thermoamylovoransDirect Black G, Direct Black 38, Congo red, methyl orange97600 mg/L, pH 8, 55 °C, 8 hAzoreductase, laccase, lignin peroxidase, manganese peroxidase[51]
Enterococcus sp. L2 and Mycobacterium vaccaeReactive Violet 5R97.6100 mg/L, pH 6.8, 8 hAzoreductase, formate dehydrogenase[52]
Lysinibacillus sp., Raoultella sp., Enterococcus spp., Citrobacter sp., Lysinibacillus sp.Reactive Black 199100 mg/L, pH 7, 37 °C, 8 hAzoreductase, lignin peroxidase[53]
Table 5. Green alternatives replacing existing harmful chemicals in the textile industry.
Table 5. Green alternatives replacing existing harmful chemicals in the textile industry.
Textile ProcessExisting Harmful ChemicalsGreen Chemical SubstituteEnzymatic AlternativeReference
SizingPolyvinyl alcoholPotato starch; carboxymethylcellulose-[60]
DesizingMineral acids (acid-based desizing)-Amylase; xylanase[12]
ScouringSodium hydroxide (caustic soda)-Pectinase (pectin lyase); xylanase[61]
Stonewashing and polishingNonylphenol–ethylene oxide adducts [alkylphenol polyethoxylates (APEOs)]Fatty alcohol–ethylene oxide adducts; alkyl polyglycosidesCellulase[62]
BleachingCalcium and sodium hypochlorite; other chlorine oxidising chemicalsHydrogen peroxide; ozone at coldGlucose-oxidase; laccase; ligninase (lignin peroxidase); arylesterase[63,64]
Bleach clean-upThiosulfates-Catalase[65]
Dyeing and printingKerosene; formaldehyde; dichlorobenzene; trichlorobenzeneWater-based thickener; polycarboxylic acids; non-formaldehyde products; butyl benzoate; benzoic acidLigninase (lignin peroxidase)[5]
Finishing and effluent treatmentSilicone-based softeners; formaldehyde-based resins; heavy metalsNatural oils (e.g., soybean, castor, or palm oils); formaldehyde-free resins; eco-friendly coagulantsProteases; lipases; laccase[3,64]
Table 6. Examples of immobilised enzymes used for textile applications.
Table 6. Examples of immobilised enzymes used for textile applications.
Immobilisation MethodEnzymeCarrier/SupportImprovementStability/ReusabilityReference
AdsorptionCellulaseCa alginate starch beadLower weight loss, minimal tensile strength reduction (67–98.35%), and improved whiteness index-[181]
CellulaseEpoxy resinMaximum activity was used on cotton fabric for biopolishingReuse up to 6 cycles[182]
CellulasePVA coated chitosanpH optimum of enzyme shifted from 4.0 to 7.0 and showed better stability at neutral pHStability of 52% after 8 cycles[183]
LaccaseTiO2–ZrO2–SiO2100% alizarin red S removal90% stable after 20 days[184]
Adsorption and covalent bondingCellulaseKaolinMinimised tensile strength loss, enzyme reusability, improved product quality, and reduced polishing costsBetter recovery and reuse for 3 cycles[185]
Covalent bindingLaccaseGreen coconut fibreEnhanced thermal stability at 50 °C and high efficiency in continuous reactive dye decolourisationStability retained 45–50% after 4 cycles[186]
LaccaseGlycidyl methacrylate functionalised polyacrylamide alginate55% dye removal50% stability remained after 5 cycles[187]
Manganese peroxidaseFe3O4/chitosan98% Reactive Orange 16 and 96% methylene blue removal86% stable after 5 cycles and 60% after 14 days[188]
Covalent/cross-linkingProtease (esperase)Eudragit S-100Operational stability at 60 °C improved 1.7-foldActivity of 72% after 5 cycles[189]
Cross-linkingKeratinaseChitosan-β-cyclodextrinEnhanced optimal activity at pH 11 and 70–75 °C with superior thermo-stabilityStorage stability of ~53% after 30 days[190]
LaccaseCu(II) ion chelated chitosan43% removal of methyl orange, 69% removal of Cibacron blue and 87% removal of Reactive Black 581% stable after 20 uses[191]
Manganese peroxidaseChitosan beads97% dye removal60% stable after 10 cycles[192]
EntrapmentLaccaseAlginate beads66% dye removal95% stable after 15 days[193]
Non-covalent bindingCellulaseEudragit S-100 and Eudragit L-100Improved fabric softness while minimising weight and tensile strength loss to the surface fibresStability of 51% and 42% after 3 cycles[194]
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

Khan, M.F. Recent Advances in Microbial Enzyme Applications for Sustainable Textile Processing and Waste Management. Sci 2025, 7, 46. https://doi.org/10.3390/sci7020046

AMA Style

Khan MF. Recent Advances in Microbial Enzyme Applications for Sustainable Textile Processing and Waste Management. Sci. 2025; 7(2):46. https://doi.org/10.3390/sci7020046

Chicago/Turabian Style

Khan, Mohd Faheem. 2025. "Recent Advances in Microbial Enzyme Applications for Sustainable Textile Processing and Waste Management" Sci 7, no. 2: 46. https://doi.org/10.3390/sci7020046

APA Style

Khan, M. F. (2025). Recent Advances in Microbial Enzyme Applications for Sustainable Textile Processing and Waste Management. Sci, 7(2), 46. https://doi.org/10.3390/sci7020046

Article Metrics

Back to TopTop