Next Article in Journal
Low Pressure UV Photolysis of the Pharmaceutical Compounds Acetaminophen, Atenolol, Bezafibrate, Diclofenac and Ibuprofen
Next Article in Special Issue
Cu(II) Adsorption from Aqueous Solution onto Poly(Acrylic Acid/Chestnut Shell Pigment) Hydrogel
Previous Article in Journal
Optimization of Cultivation Conditions for Tetraselmis striata and Biomass Quality Evaluation for Fish Feed Production
Previous Article in Special Issue
Acid Mine Drainage Treatment Using a Process Train with Laterite Mine Waste, Concrete Waste, and Limestone as Treatment Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Recent and Systemic Approach Towards Microbial Biodegradation of Dyes from Textile Industries

1
Department of Biotechnology, School of Sciences, P P Savani University, Surat 394125, India
2
Department of Biosciences, School of Liberal Arts & Sciences, Mody University of Science and Technology, Sikar 332311, India
3
Faculty of Science and Technology, Madhyanchal Professional University, Ratibad 462044, India
4
Department of Environmental Sciences, School of Sciences, P P Savani University, Surat 394125, India
5
Centre for Natural Products Discovery, School of Pharmacy and Biomolecular Sciences, Liverpool John Moores University, Byrom Street, Liverpool L3 3AF, UK
6
Department of Chemistry, Faculty of Science, King Khalid University, Abha 61413, Saudi Arabia
7
Department of Physics, Faculty of Sciences, University 20 Août 1955, Skikda 21000, Algeria
8
Department of Clinical Biochemistry, College of Medicine, King Khalid University, Abha 61413, Saudi Arabia
9
Biology Department, Faculty of Science, King Khalid University, Abha 61413, Saudi Arabia
10
Department of Semi Pilot Plant, Nuclear Materials Authority, El Maadi, P.O. Box 530, Cairo 11381, Egypt
11
Department of Civil and Environmental Engineering, Hanyang University, 222-Wangsimni-ro, Seongdong-gu, Seoul 04763, Korea
12
Department of Earth Resources & Environmental Engineering, Hanyang University, 222-Wangsimni-ro, Seongdong-gu, Seoul 04763, Korea
*
Authors to whom correspondence should be addressed.
Water 2022, 14(19), 3163; https://doi.org/10.3390/w14193163
Submission received: 27 August 2022 / Revised: 22 September 2022 / Accepted: 29 September 2022 / Published: 8 October 2022

Abstract

:
The textile industry generated a series of synthetic dyestuffs that threatened environmental protection. Azo dyes, widely utilized in textile, paper, fruit, leathers, cosmeceuticals and pharmaceutical fields, account for most of the dyestuffs made. Since they have colour fastness properties, stability, and susceptibility to oxidation, existing effluent treatment methods cannot entirely strip different dyes from effluents. Under certain environmental factors, bacteria decolourize and degrade dyes. The treatment process is cheap, environmentally safe, and can be used on various dyes. However, textile plant wastewater can produce many polluting chemicals and dyes. Environmental legislation is increasingly being enacted to regulate mainly azo-based dyes in the environment. The potential of the microbes for the decolourization of dyes and metabolizing them is long-known knowledge. The toxic components of dyes challenge a potential threat to all the living forms of life. Though both natural and synthetic dyes are used for the colourization of textiles, only synthetic ones are challenging to decolourize. Microbial-based bioremediation of dyes has been studied and reviewed primarily to accelerate dye degradation. The various piece of the literature revealed that the majority of these dye removal microbes belong to mainly white-rot fungi, a consortium of anaerobic bacteria. In addition to this, there are several (genetically engineered microorganisms) GEMs that remediate dyes efficiently. Here in the current review, the authors have tried to bridge the existing gap in the bioremediation of dyestuff. Moreover, the authors have also tried to provide the latest trend in this field. This study will surely benefit the industries and researchers related to dyestuffs by maintaining eco-friendly approaches.

1. Introduction

Over the last few decades, the utilization of dyes has increased drastically due to the rapid industrialization of dye-based industries and the increase in demand for textiles and clothes [1]. However, due to rapid industrialization, the dyes remained the same synthetic ones, which are very difficult to degrade from the environment. These textile dyes lead to pollution of water and some of them are xenobiotic. Prolonged contamination of water bodies may lead to several diseases among aquatic and other living organisms. Some of the dyes are carcinogenic which will be very severe. Dyes are mainly used for colouring textiles, fabrics, etc. [2]. The classification of dyes could be carried out by various features such as their source of origin, function and chemical makeup, etc. Generally, based on the source of origin dyes could be classified into two categories, i.e., synthetic and natural. Synthetic dyes are generally manufactured by a chemical process in the industries whereas natural dyes are derived from natural materials, such as flowers, leaves, the bark of a tree, etc. [3]. Further, dyes could also be categorized based on their solubility in the aqueous medium for instance acid dyes, basic dyes (cationic), diffuse dyes (non-ionic), and direct and reactive dyes (anionic) [4]. All the dyes are generally soluble in an aqueous medium except non-ionic dyes. The degree of solubility of such dyes varies in the aqueous medium. The types of dyes based on their solubility are shown in Figure 1. Based on solubility, dyes could be broadly categorized into soluble and insoluble dyes. The soluble dyes could be further categorized into anionic and cationic dyes based on their charges. Cationic dyes mainly include basic dyes while anionic dyes could be further categorized into acid, reactive direct, and mordant dyes. Whereas the insoluble dyes can be divided further into azo dyes, disperse, sulphur, vat and solvent dyes. The members present in all these types of dyes are shown in detail in Figure 2.
Dyes commonly contain anthraquinone, indigoid, and azo aromatic structures [6]. Dyes are mainly comprised of two parts, i.e., auxochrome and chromophore. The auxochrome is derived from the Greek word, where ‘auxo’ means ‘to increase’ and ‘chrome’ means to colour. Auxochrome is a group of atoms which imparts different colour to the material only when it is attached to a chromophore but when it is present alone it could not impart that colour. Auxochromes are also called ‘colour helpers’ or ‘colour intensifiers. Chromophores are mainly a part of a dye that when exposed to visible light then it will absorb and reflect a certain colour [7].
Dyes are frequently applied to many substrates in edibles, cosmetics, paper, rubber, and products of textile industries [4]. Day by day the demand for dyes is increasing with the increasing population and industries. Therefore, there is more establishment of dye and textile industries. Due to the increased establishment of the dyes industry, there is a large amount of dye effluent production. These effluents are mainly disposed of in rivers or other water bodies leading to water pollution. Moreover, most of the dyeing process is inefficient at the time of dying of garments, which leads to the loss of the majority of the dyestuff in the water sources [8]. Once these dyes enter the water bodies, they are available in the environment. The number of dyes lost during a dyeing process totally depends on the type of dyes being used. If simple dyes are used, then the loss will be only 2% while if reactive dyes are used then the loss of dyes may reach up to 50% [9]. Therefore, both the process, i.e., disposal of dye effluent in the water bodies and the inefficient dying process deteriorates the quality of water in the environment. Due to these reasons the quality of water in terms of biological oxygen demand, chemical oxygen demand, total suspended solids, and optical features are affected. Due to the dyes, the light could not penetrate into the water systems which affects the growth of flora and fauna of the aquatic system [10]. Dyes in the effluent are highly visible even at low concentrations, and oxidation products of these clothing dyes are often carcinogenic. In addition, these dyes lead to several toxic effects on living organisms including humans. For instance, there are several dyes which are neurotoxic, cytotoxic, genotoxic, mutagenic, hypersensitive and mitochondrial toxic [11]. Due to all these activities, various countries have adopted stricter environmental regulations to deal with the textile industries.
This dye effluent could be treated by various biological, chemical and physical methods. Biological methods include the utilization of microorganisms as such or their enzymes, phytoremediation, activated lagoons, trickling filters etc. [12]. The chemical approaches for dye removal include precipitation, absorption, oxidation, ozonation, electrolysis, reduction, advanced oxidation process, ion exchange etc. [13]. The physical methods include techniques such as ultrasonication, drying, sedimentation, reverse osmosis, and coagulation/flocculation membrane filtration such as microfiltration, nanofiltration, and ultrafiltration are very expensive [9]. Physical methods such as screening, sedimentation, and skimming are used to eliminate floating debris from the wastewater. Figure 3 shows all the biological, physical, and chemical approaches for the decolourization of dyes from wastewater.
Out of the above-mentioned dye effluent treatment methods, majority of them are inefficient, expensive, energy intensive and leading to formation of by-products which are more harmful. Therefore, in all three types of approaches, biological one is comparatively eco-friendly, with no utilization of chemicals during the process and also no harmful metabolites by-products produced. In addition to all these features, biological materials used for dye removal are biodegradable in nature [14].
Even among biological approaches, there are choices to use plant-based remediation or microbial remediation. The selection of the two depends on the availability of the biological material, their cost and time consumption. The remediation by using plants is called phytoremediation while microbial remediation may be further categorized into mycoremediation (fungi) [15], phycoremediation (algae) [16], and bacterial remediation. Out of phytoremediation and microbial remediation, the former is time taking as the plant has to grow firstly in those areas which will depend on the life span of the plant. Therefore, phytoremediation becomes a second choice when it comes to the rate of remediation of dye effluents from the site.
So, due to these beneficial features, developed countries are now switching to microbial-based techniques for the remediation of dyes. The microbes are used either in the dried powder form, or their nanosized form. In addition to this microbe could be used directly as a whole organism or their extracted enzymes/metabolites used for remediation purposes. The dried or nanosized form mainly acts as a biosorbent whereas the whole microbe or its enzyme attacks the various bonds present in the dyes. The former two approaches simply adsorb the dyes (biosorption), whereas the latter two processes mineralize or metabolize the dye molecules and convert them into either mineral or less toxic products (enzymatic degradation) [11]. The microbial-based approaches have been used in either monoculture or a mixture of cultures called microbial consortiums. There are several pieces of literature where monoculture has been used for the remediation or decolourization of dyes. Some of the studies are discussed over here. Khadijah (2009) tested the ability of 1540 bacterial isolates to remove different azo dyes. Kochher and Kumar (2012) investigated the screening of bacteria that could decolourize textile dyes. The findings of this study suggested that bacteria could be used to remediate dye from wastewater. The bacterial population used over here for the remediation of dyes was isolated from waste effluent from the textile dye industry. Investigators have shown that bacteria are naturally adaptable and capable of degrading pollutants. Leena et al., 2008, investigated the remediation of Reactive Black-B dye (RBBD) from textile effluent using effluent-adapted bacteria (EAB) and effluent non-adapted bacteria (ENAB). The investigator found a group of five bacteria, i.e., Alcaligenes sp., Pseudomonas aeruginosa, Bacillus sp., Eubacterium sp., and Arthrobacter sp. For decolourization in ten days, investigators have utilized four ENAB, for say, Kluyvara ascorbate, Bacillus sp., Pseudomonas species, and Pasteurella sp. Effluent-adapted Bacillus sp. reduced colour by 35.68%, while non-adapted isolate of the same genus removed colour by 30.04%. P. aeruginosa adapted to effluent demonstrated 44.2% decolourization, comparable to non-adapted Pseudomonas sps (41.73%). A team of investigators led by Mutafov et al., 2007 revealed the potential of two bacterial species, Alcaligenes faecalis and Rhodococcus erythropolis, to remediate monoazo dye Acid orange (AO) at various initial dye concentrations. A diazo dye Reactive yellow 84A dye was effectively degraded by using a novel bacterial strain, Exiguobacterium sp. [17]. In addition, analytical techniques such as Fourier transform-infrared spectroscopy (FT-IR), gas chromatography-mass spectroscopy (GC-MS), and high-performance liquid chromatography (HPLC) showed that dye degradation resulted in a substantial decrease in phytotoxicity. Furthermore, there are several investigators who have tried to remove malachite green by using Aeromonas hydrophila, where the efficiency was about 96.8% [18], Bacillus pseudomycoides were used by Kumar et al., 2019 to remove acid black 24/40 mg/L where the efficiency was 96% [19]. Ayed et al., 2019 tried to remove brilliant orange 3R/750 mg/L by using Lactobacillus acidophilus with an efficiency of about 99.3% [20]. Similarly, a huge amount of work has been carried out by using monoculture but very less attempts were made by using mixed culture or by using microbial consortium.
Since the microbial consortium has different microbes which produce a variety of enzymes and metabolites so there is more effective removal of dyes with them. There are several reports where dyes remediated were carried out by microbial consortiums. Due to synergistic metabolic behaviour, it has been documented that these diverse microbial cultures can achieve effective dye degradation. Poddar et al., 2022 used a microbial consortium (MC) comprising mainly different Klebsiella sps. The authors have prepared 5 different consortiums from 19 strains. The developed MC was used to remove amaranth, sunset yellow, tartrazine, indigo carmine, and quinoline yellow dyes. The overall efficiency of about 96.5% was achieved for the most potential MC [21].
Algae, yeast, bacteria, and fungi are among the microorganisms that can mineralize and decolourize different dyes [22]. Microalgae have shown huge potential in the bioremediation of dyes from wastewater. It has been used in various forms such as dry powder, live, extracts mostly in lipid-free form and nanoparticles. Mansour et al., 2022 developed nanoparticles of Arthrospira platensis NIOF17/003 microalgae and utilized them for the remediation of methylene blue (MB), from the synthetic aqueous solutions. The developed NPs were analysed by sophisticated instruments for detailed morphological and elemental properties. Further, the adsorption capacity of the Arthrospira platensis NIOF17/003 NPs was analysed against pH, temperature, contact time, and initial dose of MB dye. The authors concluded that the optimum remediation of MB dye was achieved when the NP’s initial dose was about 0.4 g. Moreover, about 93% removal of MB dye was reached within 5 min at a stirring rpm of 150. Further, the authors have also analysed the recyclability of the adsorbent and found that it is possible to achieve MB remediation up to 65.5% after three recycles of the adsorbent [23]. Ashour et al., 2021 used the same strain of alga for the bioremediation of ammonia from aquaculture wastewater effluents. Here the investigators have used complete dry biomass and lipid-free biomass of the above alga. Further after acquiring a detailed investigation of both the adsorbents by sophisticated instruments, it was used to remove ammonium ions from aquaculture wastewater effluents. The investigators removed the ammonium ions from both the conditions, i.e., from the synthetic aquaculture solutions and from real aquaculture effluents. The investigators concluded that the ammonium ions removal percentage was higher in synthetic aquaculture water than in the real one. The removal percentage of ammonium ions by complete dry-biomass of alga was 64.24% and 25.7% from synthetic and real aquaculture wastewater, respectively. While in the case of lipid-free biomass, in synthetic aquaculture wastewater NH4+ ions were removed up to 98.68% and in real aquaculture wastewater was 37.85 [24]. Alprol et al. 2021 used complete dry biomass and lipid-free biomass of the above alga and assessed its potential for the removal of two organic dyes namely Ismate violet 2R, and IV2R from textile effluents. Both the adsorbents were thoroughly characterized by sophisticated instruments for detailed properties. By applying batch adsorption study of the dye removal, it was found that for complete dry biomass it was 75.7% and for lipid-free biomass, it was 61.1%. Therefore, it was found that the complete dry biomass of A. platensis was more effective for the said dye removal at optimized conditions [25].
While there are several enzyme-based dye removals that are found highly reliable for the removal of dyes. For instance, azo-reductase, laccase, lignin peroxidase, manganese peroxidase etc. remediate dyes very efficiently. Azo reductase is one such enzyme which specifically acts on the azo dyes. Azo dyes could be efficiently removed by using azo dye reductases and lactases. Azo-reductases catalyse reactions only in reducing equivalents such as FADH and NADH. The majority of Azo dyes have sulphate groups and have heavy molecular weight, making them difficult to move through [3]. Figure 4 is showing various types of dye biodegradation processes based on the types of microorganisms. The microbial-based dye removal process involves fungi, bacteria as a whole organism, microalgae, microbial enzymes, microbial consortium (MC), and dried or powdered forms of these microbes. While Figure 5 is showing types of microbial-based dye degradation in the presence or absence of oxygen. The oxygen-dependent microbial process generally depends on the nature of the used microorganism, i.e., whether it is aerobic or anaerobic. Therefore, the microbial process could be either aerobic, anaerobic or anoxic.
So, microbial-based bioremediation of dyes effluent from wastewater is the most economical and reliable globally. There are also good numbers of microorganisms that are easy to manage and need no planning. In addition to this periphyton (complex mixture of algae, cyanobacteria, heterotrophic microbes+ detritus or periphytic biofilm, methods may also be utilized for dye remediation. The microbial-based dye remediation has been carried out by various investigators either in natural conditions or in laboratory conditions either in aerobic or anaerobic conditions based on the nature of the microorganism [26].
In the present review, the investigators have emphasized the microbial-based remediation of dye effluents from wastewater. The authors have focused on the existing gap in this field and try to provide a current state of the art. The authors have focused on the different microorganisms used for dye remediation purposes and recent trends. The authors have tried to provide a mechanism of dye removal from wastewater by using a microorganism. Finally, the authors have also focused on microbial enzymes and their role in the remediation of dyes in the environment. In addition to all, one of the aims was to focus on the decolourization and oxidation of dyes by both mixed and axenic bacteria and fungi cultures.

2. Microbial Biodegradation and Decolourization of Textile Dyes

Various microbes can be used to degrade various dyes; they have different structures and pathways for dye degradation. Azo dyes are a valuable family of dyes with the most significant colour diversity. Microorganisms degrade azo dyes in anaerobic conditions with the assistance of azoreductase, resulting in colourless aromatic amines as a by-product. Benzidine is often utilized in the production of direct azo dyes and is reported to be associated with cancer. Natural dyes are low-cost and can be used to colour fabrics, leathers, and papers without the need for pre-treatment. Natural Black 38 (NB-38) is the most commonly used benzidine-based azo dye [27]. Enterococcus gallinarum can be used to degrade Direct Black 38 ink (DB 38) [28]. Active sites for azo-reductase are three azo bonds in Direct Black 38’s structure which were converted to benzidine through metabolic reactions, which are then deaminated to produce 4-amino phenyl. Investigation revealed that those dyes with benzidine as a base are significantly more cancer-causing than dyes without Benzidine [29]. This is due to the contaminants such as 4-amino biphenyl and 2-4, diaminoazo-benzene, which have been linked to cancer. Microbial degradation of dyes involves a common approach towards a particular dye, the only thing which varies is their enzyme interacting with the dyes, intermediate and final products. Some of the microbes perform complete degradation of the dyes while some of them carry out incomplete degradation of dyes. Broadly, dye-degrading microbes could be classified as actinomycetes, bacteria, fungi, and algae. Moreover, based on the nature of microbes and the pathway followed for degradation it could be again aerobic and anaerobic degraders. Generally, bacteria could be classified further as aerobic degraders or anaerobic degraders. Here in Figure 6 a generalized scheme of degradation of dye is shown.

2.1. Bioremediation of Dyes by Actinomycetes

Actinomycetes are fungi-like, filamentous bacteria with high GC content generally present in the soil. Extracellular peroxidases produced by actinomycetes, especially Streptomyces species, are known to play a role in lignin biodegradation [31]. The initial oxidation of lignin is carried out by the peroxidase released by the Streptomyces, into different water-soluble polymeric compounds. It has also been observed that the actinomycetes catalyse hydroxylation, oxidation, and dealkylation reactions in the presence of xenobiotics [32]. Three groups initially looked into the potential of actinomycetes to decolorize and mineralize textile dyes. Ball et al., tested 20 actinomycetes strains from various genera for their potential to bleach polymeric dye Poly R in 1989. Streptomyces badius 252, Thermomonospora fusca MT800, and Streptomyces sp. strain EC22 were the only three strains that greatly decolorized dye [6].
Following that, it embarked on a more extensive screening procedure that looked at the decolorizing ability of 159 actinomycetes. Investigators who utilized real textile effluents in the screening process were excited about this research. Five different effluents were used, each having a single dye of known concentration. Structurally each dye was different from azo compound RR147 to phthalocyanine Reactive Blue 116. Positive findings were obtained for 83 isolates, demonstrating the extensive potential of actinomycetes to cause dye decolourization. Given compounds’ resistance to mineralization by other bacteria under similar conditions, a discovery that actinomycetes can aerobically decolourize and degrade azo dyes was important.
Finally, but perhaps most importantly, a team led by Don Crawford at the University of Idaho began investigating the potential of lignin lytic microbes, including white-rot fungi and Streptomyces, to mineralize and decolourize textile dyes. Initially, the ability of 14 Streptomyces to decolourize two polymeric dyes, Poly B-411 and Poly R-478, as well as azo dye Remazol Brilliant Blue R, was investigated (RBBR). With two dyes, RBBR and Poly B-411, nearly similar findings were reported, indicating a close link between the isolate’s ability to decolourize dyes and its ligninolytic ability. This investigation and the fact that extracellular H2O2 synthesis increased when actinomycetes sps, were allowed to grow in the presence of glucose indicated that peroxidases were involved in the decolourization process. Extracellular peroxidases were previously discovered in Streptomyces bacteria, and enzymes were exhibited to have substrate specificities identical to P. chrysosporium’s Mn (II)-peroxidase [33]. Surprisingly, there was no association between decolorizing activity and ligninolytic activity with the 3rd pigment, Poly R-478 [34]. Enzymatic processes that take place during the decolourization of this dye are unknown. [8]. Recently Blanquez et al., 2019 reported 6–70% removal of AO-63 by using Stp. Ipomoeae CECT 3341, from the textile dyes [35]. Actinomycetes-based removal of various dyes is shown in Table 1, where the removal percentage of dyes varies from 3–100%.

2.2. Bioremediation of Dyes by Aerobic Bacteria

It is very difficult to isolate bacteria that could aerobically decolorize and mineralize dyes, except for actinomycetes. Various studies claim that different azo dyes can be converted aerobically, out of which one study reveals the development of the green color of instant chocolate puddings. including one notable study on the greening of instant chocolate puddings. It has been found that these preadapted aerobic bacteria obtain their energy by utilizing some carboxylated analogues of sulfonated azo compounds as their sole carbon and energy supply [9].
Orange-I azo-reductase [NAD(P)H:1-(4′-sulfophenylazo)-4-naphthol oxidoreductase] and Orange II azo-reductase [NAD(P)H:1-(4′-sulfophenylazo)-2-naphthol oxidoreductase] is distilled and characterized from Pseudomonas. Both of these enzymes are now categorized as the same enzyme, which is called azo-benzene reductase (EC 1.6.6.7). Aside from azo dyes, bacteria’s ability to aerobically metabolize other dye groups has piqued interest, but with little success. Aerobic mineralization of triphenylmethane pigment, MV, by a strain of Pseudomonas mendocina MCM B-402 was recently discovered. Isolate’s sole carbon and energy source was MV, which has several industrial uses in addition to its well-known utilization as a regular bacteriological and histological stain. According to preliminary research, Ps. mendocina degraded dye to phenol by a variety of unspecified metabolites, which then joined the ketoadipic acid pathway [10].
Karim et al., 2018 and Mustafa et al., 2022 reported the decolourization of reactive dyes of textiles by using monoculture and consortium.
John et al., 2020 reported the utilization of a halophilic bacterium (Salinivibrio kushneri HTSP) for the biological decolourization of synthetic dyes (CBB G-250 and Congo red and safranin). Originally this strain was isolated from the saltpan for decolourization and bioremediation of dyes. The investigators obtained about 80% decolorization within 48 h. The further investigator reported that the rate of decolorization by using a particular bacterial strain was in the order of CR > CBBG-25 > Safranin [42].
Aktar et al., 2020 isolated Bacillus sp. and Staphylococcus aureus from the textile dye effluent and applied them for the decolorization and degradation of BY HP-2R, BR-RS, and BR RS 01 (BB) dyes. The removal efficiency of BR and BB reached up to 71% and 83% with Bacillus sps. While with Staph. aureus efficiency of BY dye reached 79% after five days [43].
Shete et al., 2020 isolated Bacillus sps, Klebsiella, and Pseudomonas sps, from textile effluent and applied them for the decolorization of JR, JF, and JGY dyes. The removal or decolorization of dyes was achieved from 45–50% for the following dyes under optimized conditions [44]. Khaled et al., 2022, applied Bacillus cereus (B. Cereus) and Pseudomonas parafulva (Ps. parafulva) for the decolorization of textile azo dyes (T-blue, yellow GR, and orange 3R,) and obtained the removal efficiency of up to 91.69 and 89.21% for orange 3R [45]. Table 2 shows summarized studies of bacterial-based dye removal from the wastewater. From all the studies it was found that in some of the studies bacteria was used directly, while in some cases their dry powder was used while in a few cases a bacterial consortium was also used. In some of the studies, the dye removal was achieved up to 100% either with monoculture or with the consortium. Some of the bacterial consortiums reached up to 97–99% dye removal within a short period of time.

2.3. Mechanism of Bioremediation

The use of biological processes to reduce emissions from atmospheric, aquatic, or terrestrial systems is known as bioremediation. Microorganisms and plants are most often used as biological structures for this purpose. Microorganism-mediated biodegradation is one of the most common bioremediation methods. Microbes may simplify many complex substances to meet their development and energy requirements. Air may or may not be needed for these biodegradation processes. In certain circumstances, the same metabolic pathways that cells utilize for growth and supply of energy are often utilized to simplify the components of pollutants. Microorganisms do not benefit directly in these situations, known as metabolism, but investigators have taken the benefit of this effect and used it for bioremediation. Mineralization, or complete oxidation, results in the production of water and either CO2 or CH4. Incomplete biodegradation can result in the cleaving of the product which is less toxic than the initial pollutant. As a result, bioremediation overcomes the drawbacks of traditional methods by causing physical degradation of specific organic chemicals at a lower rate. As a result, bioremediation has evolved from a nearly unknown technique to one that has been used to clean up a wide variety of toxins over the past two decades [74].

3. Fungi-Based Bioremediation of Dyes

3.1. White Rot Fungi (WRF)

WRF is the most well and extensively studied dye-decolorizing microbe because of its potential to mineralize the intricate polymeric composition of lignin (woody), the group of species that is critical to the global carbon cycle. WRF could mineralize a distinct variety of persistent organic compounds in addition to their natural substrate, which separates them from biodegradative bacteria, which are more substrate-specific. Since their ligninolytic enzymes, Lip, manganese peroxidase (MnP) (EC 1.11.1.13) laccase, are primarily nonspecific, these fungi can mineralize a wide variety of organic compounds. These enzymes and their catalytic properties are not described over here. Still, in a nutshell, LiP catalyses the oxidation of non-phenolic aromatic compounds such as veratryl alcohol, while MnP converts Mn2+ to Mn3+, which could oxidize a vast variety of phenolic compounds [75]. Laccase (benzenediol: oxygen oxidoreductase, EC 1.10.3.2) is a Cu-containing enzyme that catalyses phenolic substrate oxidation by combining it with oxygen reduction in water [76]. Decolourization of dyes by WRF was first documented by researchers who devised a way to quantify Phanerochaete chrysosporium’s ligninolytic behaviour based on the decolorization of a variety of sulfonated polymeric dyes [77]. Pleurotus ostreatus [78] developed an uncommon enzyme capable of Remazol Brilliant Blue R (RBBR) decolorization during solid-state fermentation of wheat straw. This development was distinct from isolate’s MnP, laccase, and manganese-independent peroxidase activities, as well as LiP and veratryl alcohol, oxidize actions, which were not detected in P. ostreatus but are common in other WRF. Since it had a catalytic metal centre and was inhibited by a variety of proven oxygenase inhibitors, activity was dubbed RBBR oxygenase [79].

3.2. Remediation by Non-WRF

Although degradation mechanisms used by non-WRF have not been identified in the literature, they are likely to be close to those found in the metabolism of other aromatic hydrocarbons. One such isolate, the strain of Geotrichum candidum Dec1 isolated from soil and capable of a decolorizing variety of anthraquinone dyes, has been studied in depth. This isolate’s broad substrate specificity led to Kim and revealed an extracellular peroxidase-like enzyme. DSP, glycosylated haeme-based peroxidase isolated from G. candidum with features different from LiP, MnP, horse-radish peroxidase, and other peroxidases, was obtained [80]. The robustness of this isolate in contrast to P. chrysosporium and other WRF makes it potentially useful for dye decolorization. In contrast with the action of P. chrysosporium ligninolytic enzymes, DyP is generated continuously and does not become affected by shear forces in shake flasks or stirred tank reactors. Such characteristics could allow the development of a range of G. candidum-Dec1-based biological reactors for treating a variety of organic contaminants, including dyestuffs. To enable excellent protein development in the safe host, the recently cloning of a gene that encodes DyP into heterologous expression host Aspergillus oryzae. According to molecular research, DSP is different from other peroxidases in the plant peroxidase superfamily because it has a novel haeme-binding area [14]. Table 3 is showing summarized studies of mycoremediation of dyes under different experimental conditions. From all the myco-based remediation of dyes, it was found that some of the fungi have removed the dyes up to 100% from the wastewater. This is so because fungi have a large number of extracellular enzymes. Moreover, the presence of hyphae and mycelium might have been advantageous for the remediation of dyes.

4. Algae-Based Bioremediation of Textile Dyes

There are several reports in the literature where algal systems have been used for the decolorization or bioremediation of textile dyes. Algae also have numerous enzymes and other molecules which play a role in the decolorization of dyes from textile wastewater. Algae are reported to utilize the azo dyes as a source of carbon and energy and degrade them into aromatic amines, followed by their subsequent conversion into simpler organic and inorganic compounds [109]. To date, Chlorella sps, Oscillatoria sps, Synechocystis, and Phormidium sps have been used widely by various investigators for the decolorization of dyes from textile samples [110] and a detailed one is shown in Table 4. From the various studies of phycoremediation of dyes, it was found that the highest efficiency of dye removal was achieved up to 99–99.5%. Though all the studies were carried out with different dyes, different concentrations, and under different reaction conditions, the mixed culture of algae showed maximum removal, i.e., 99.5%, which shows the effectiveness of MC.

5. Enzyme-Based Bioremediation of Dyes

To date mainly two types of enzymes are widely used for the remediation of textile dyes. One is laccase and the other one is hydrogen peroxidases. Both enzymes are efficient and widely used by several investigators for the removal or decolourization of dyes from wastewater. The source of enzymes for these purposes are mainly microbial and even from microbes’ fungi are more widely exploited. Laccase is a type of copper-containing polyphenol oxidase which is produced by several bacteria and fungi. Its potential has been exploited for the decolourization of Azo dyes. While peroxidase is an enzyme that degrades H2O2 and uses oxygen to oxidize other substances. Earlier lignin peroxidase has been used for the removal of Congo red dye. Furthermore, it has also been used for the decolorization of Reactive Orange 16 and remazol Brilliant Blue R (RBBR) [110].
Dong et al., 2019 reported the utilization of azoreductases extracted from Streptomyces sps. for the removal of methyl red from the wastewater [126].
Sherifah et al., 2019 used Kluyveromyces dobzhanskii DW1 and Pichia manshurica DW2 bacterial laccase for the enzymatic degradation of MG and MR [127]. Darvishi et al., 2018 used yeast laccase extracted from Yarrowia lipolytica and applied them for the enzymatic degradation of Bromocresol Purple, Safranin, Bromothymol Blue and Phenol Red [128]. In 2017, Afreen et al., reported the remediation of reactive Blue-4, from wastewater by using algal laccase isolated Spirulina platenis. In another study, a team led by Afreen used algal laccase from Arthrospira maxima and utilized them for the removal of RB-4, RBBR. A team led by Othman et al., 2018 reported the extraction of laccase from the edible mushroom (Agaricus bisporus) and used them for the enzymatic degradation of Acid Blue [129]. Dai et al., 2021 extracted laccase from a novel bacterium (transformed E. coli BL21(DE3) and applied it for the decolorization of acid violet 7, bromophenol blue (BpB), and Coomassie brilliant blue. Investigators obtained a mutant laccase (mut lac 2-9) which was reported to have enhanced decolorization of dyes [130].
Furthermore, these investigators have also used lignin and manganese peroxidase for the enzymatic degradation of dyes. For instance, Zhang et al., 2018 performed enzymatic degradation of MO and BpB by using fungal MnP [131].
Shaheen et al., 2017 and Bouacem et al., 2018 reported the enzymatic degradation of a group of dyes by using fungal lignin peroxidase extracted from Ganoderma lucidum IBL 05 and Bjerkandera adusta CX-9, respectively [132,133].
Here Figure 7 is showing a mechanism of degradation of dyes (indigo dye) by the laccase enzymes. Here, firstly the laccase enzyme attacks the indigo dye and removed the hydrogen. Once the hydrogen is eliminated the indigo dye becomes converted into dehydrindigo. This dehydroindigo goes under several chemical changes leading to the formation of different intermediates. The dehydroindigo becomes converted into isatin which again forms an intermediate structure and finally becomes converted into anthranilic acid which is quite stable in the environment and less harmful than the original dye itself.
Figure 8 is showing image where a genetically modified laccase from Bacillus amyloliquefaciens have been used for the enhanced enzymatic degradation of indigo carmine dye. The image shows reduction of colour intensity of the dye after enzymatic treatment.
Few of the investigators have also reported the utilization of bacterial polysaccharides, i.e., intra and extra both for the removal of dyes from the wastewater [136]. Mustafa et al., 2022, reported the utilization of bacterial polysaccharides mainly exopolysaccharides for the decolorization of dyes from textile water [53].

6. Gaps and Future Prospects and Advantages of Microbial Degradation of Dyes

In order to acquire good efficiency and outcomes in dye biodegradation, further experimentation is required on (a) responsible microorganisms, (b) experimental factor limitations, (c) bioremediation site, and (d) degradation pathways before using microorganisms in the area. It will be critical to find out the type of degradation materials and determine their (none) toxicity to marine or plant species. While dyes have defied many microbial degradation strategies, there is an alternative approach to degrade dyes using genetic engineering, opening up a new domain for investigators working in this area. Advanced molecular biology techniques can investigate genes responsible for dye degradation. Dye degradation can result in the production of by-products, sources of nutrition, and energy, both of which can be utilized as commodities. However, researchers face a difficult task in attaining complete degradation of dye. Various scientific trials must be conducted to apply biodegradation of dye wastewater successfully. Future dye degradation research should focus on reducing factors that inhibit microbial activity. Recent and early promising experiments must be re-examined to increase their effectiveness. The removal of contaminants in the biodegradation process depends on the pathways, rate, and mechanisms of degradation pathways and environmental conditions. It must be essential to ensure that degraded materials do not harm marine organisms or plants. Integration of dye pollution treatment systems is desirable for successful industrial translation. The exploration of bacterial degradation kinetics will benefit from the study of pathways and hypotheses for bacterial degradation of dye wastewater. There are no additives used in this treatment. No toxic substances are released into the atmosphere by the consortium. The procedure, therefore, reduces the need for chemical refining and vapour. No oil is used, resulting in carbon-free operation. Therefore, a carbon credit is also valuable, and no complicated plants are needed.

7. Conclusions

The remediation of dye effluent is possible with all the three approaches, i.e., chemical, physical and biological. Biological approaches especially microbial biodegradation of dyes has shown potential and interest in the recent times. The reliability of microbial-based dye biodegradation is mainly based on its biodegradable nature, eco-friendly, easy to use and faster action. The microbes have been used either as a biosorbent when it is used as a dried powder or has been used for enzymatic degradation as a whole organism. The bacteria, algae, fungi and actinomycetes have shown huge potential for dye removal in the direct form. Indirectly the enzymatic degradation of dyes in the soluble form or the immobilized enzymes has gained a tremendous response. Microorganisms are the warehouse of natural enzymes which mineralizes dyes effectively in the environment without producing a toxic product. The utilization of fungi is more economical and efficient as the enzymes produced by them are extracellular in nature and produce a wide range of enzymes. The mixed culture has shown more potential for dye removal than the monoculture. Some of the microbial-based approaches especially enzymatic degradation has shown 100% removal of dyes. Such approaches for the remediation of dyes from the environment make the technique environment-friendly, and sustainable.

Author Contributions

Conceptualization, H.P., J.-W.P., H.K., M.A. and H.A.I.; data curation, J.-W.P., N.C., M.M.A. and H.K.; methodology, J.-W.P. and S.P.; validation, K.K.Y., H.A.I. and B.-H.J.; formal analysis, A.G., M.M.A. and H.A.I.; resources, K.K.Y., B.-H.J., M.A. and H.A.I.; writing—original draft preparation, V.K.Y., H.P., N.C. and H.K.; writing—review and editing, A.G., N.C., V.K.Y., H.A.I., M.M.A., J.-W.P., S.P., B.-H.J. and H.K.; supervision, V.K.Y., H.P. and B.-H.J.; project administration V.K.Y., M.A. and K.K.Y.; funding acquisition, H.A.I., M.A. and A.G.; investigation, V.K.Y., B.-H.J., N.C. and H.P.; software, S.P., M.A., H.A.I., A.G., M.M.A. and K.K.Y.; visualization, A.G., S.P. and B.-H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was funded by the Deanship of Research, King Khalid University, Saudi Arabia for Small Research Group under the grant number R. G. P. 1/339/1443. This work was supported by the Korea Environment Industry and Technology Institute (KEITI) through Subsurface Environment Management (SEM) Projects, funded by Korea Ministry of Environment (MOE) (No. 2020002480007).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All relevant data are included within the article.

Acknowledgments

H.P. and V.K.Y. are thankful to P P Savani University, Surat, for providing infrastructure and support. V.K.Y. also acknowledges Mody University of Science and Technology, Sikar, Rajasthan for software and resources.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Manzoor, J.; Sharma, M. Impact of Textile Dyes on Human Health and Environment. In Impact of Textile Dyes on Public Health and the Environment; Wani, K.A., Jangid, N.K., Bhat, A.R., Eds.; IGI Global: Hershey, PA, USA, 2020; pp. 162–169. [Google Scholar] [CrossRef]
  2. Halim, A.F.M.F.; Islam, M.T.; Hoque, M.M.U. Chapter 4—Chemistry of sustainable coloration of textile materials. In Green Chemistry for Sustainable Textiles; Ibrahim, N., Hussain, C.M., Eds.; Woodhead Publishing: Sawstone, UK, 2021; pp. 57–67. [Google Scholar] [CrossRef]
  3. Rudakiya, D.M.; Tripathi, A.; Gupte, S.; Gupte, A. Fungal Bioremediation: A Step Towards Cleaner Environment. In Advancing Frontiers in Mycology & Mycotechnology: Basic and Applied Aspects of Fungi; Satyanarayana, T., Deshmukh, S.K., Deshpande, M.V., Eds.; Springer: Singapore, 2019; pp. 229–249. [Google Scholar] [CrossRef]
  4. Kumar, A.; Singh, K.; Gupta, S.P. Structure and Properties of Dyes and Pigments. In Dyes and Pigments; Dixit, U., Ed.; IntechOpen: Rijeka, Croatia, 2021; Chapter 8. [Google Scholar] [CrossRef]
  5. Samsami, S.; Mohamadi, M.; Sarrafzadeh, M.H.; Rene, E.R.; Firoozbahr, M. Recent advances in the treatment of dye-containing wastewater from textile industries: Overview and perspectives. Process Saf. Environ. Prot. 2020, 143, 138–163. [Google Scholar] [CrossRef]
  6. Liu, Y.-Q.; Maulidiany, N.; Zeng, P.; Heo, S. Decolourization of azo, anthraquinone and triphenylmethane dyes using aerobic granules: Acclimatization and long-term stability. Chemosphere 2021, 263, 128312. [Google Scholar] [CrossRef] [PubMed]
  7. Gürses, A.; Açıkyıldız, M.; Güneş, K.; Gürses, M.S. (Eds.) Dyes and Pigments: Their Structure and Properties. In Dyes and Pigments; Springer International Publishing: Cham, Switzerland, 2016; pp. 13–29. [Google Scholar] [CrossRef]
  8. Chequer, F.M.D.; Ferraz, E.R.A.; Cardoso, J.C. Textile Dyes: Dyeing Process and Environmental Impact. In Eco-Friendly Textile Dyeing and Finishing; de Oliveira, G.A.R., Ed.; IntechOpen: Rijeka, Croatia, 2013; Chapter 6. [Google Scholar] [CrossRef] [Green Version]
  9. Soleimani-Gorgani, A.; Taylor, J.A. Dyeing of nylon with reactive dyes. Part 1. The effect of changes in dye structure on the dyeing of nylon with reactive dyes. Dye. Pigment. 2006, 68, 109–117. [Google Scholar] [CrossRef]
  10. Yadav, V.K.; Choudhary, N.; Ali, D.; Kumar, G.; Gnanamoorthy, G.; Khan, A.U.; Kumar, P.; Hari Kumar, S.; Tizazu, B.Z. Determination of Adsorption of Methylene Blue Dye by Incense Stick Ash Waste and Its Toxicity on RTG-2 Cells. Adsorpt. Sci. Technol. 2022, 2022, 8565151. [Google Scholar] [CrossRef]
  11. Bharathi, D.; Nandagopal, J.G.T.; Ranjithkumar, R.; Gupta, P.K.; Djearamane, S. Microbial approaches for sustainable remediation of dye-contaminated wastewater: A review. Arch. Microbiol. 2022, 204, 169. [Google Scholar] [CrossRef]
  12. Yadav, V.K.; Inwati, G.K.; Ali, D.; Gnanamoorthy, G.; Bera, S.P.; Khan, S.H.; Choudhary, N.; Kumar, G.; Chaurasia, T.P.; Basnet, A. Remediation of Azure A Dye from Aqueous Solution by Using Surface-Modified Coal Fly Ash Extracted Ferrospheres by Mineral Acids and Toxicity Assessment. Adsorpt. Sci. Technol. 2022, 2022, 7012889. [Google Scholar] [CrossRef]
  13. Wang, L.K.; Wang, M.-H.S.; Shammas, N.K.; Hahn, H.H. Physicochemical Treatment Consisting of Chemical Coagulation, Precipitation, Sedimentation, and Flotation. In Integrated Natural Resources Research; Wang, L.K., Wang, M.-H.S., Hung, Y.-T., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 265–397. [Google Scholar] [CrossRef]
  14. Singh, D.; Yadav, V.K.; Ali, D.; Soni, S.; Kumar, G.; Dawane, V.; Chaurasia, T.P. Isolation and Characterization of Siderophores Producing Chemolithotrophic Bacteria from the Coal Samples of the Aluminum Industry. Geomicrobiol. J. 2022, 1–7. [Google Scholar] [CrossRef]
  15. Ekanayake, M.S.; Udayanga, D.; Manage, P.M. Mycoremediation of Synthetic Textile Dyes by Fungi Isolated from Textile Wastewater Effluent and Soil. In Mycoremediation Protocols; Udayanga, D., Bhatt, P., Manamgoda, D., Saez, J.M., Eds.; Springer: New York, NY, USA, 2022; pp. 31–43. [Google Scholar] [CrossRef]
  16. Bhardwaj, D.; Bharadvaja, N. Phycoremediation of effluents containing dyes and its prospects for value-added products: A review of opportunities. J. Water Process Eng. 2021, 41, 102080. [Google Scholar] [CrossRef]
  17. Dhanve, R.S.; Kalyani, D.C.; Phugare, S.S.; Jadhav, J.P. Coordinate action of exiguobacterial oxidoreductive enzymes in biodegradation of reactive yellow 84A dye. Biodegradation 2009, 20, 245–255. [Google Scholar] [CrossRef]
  18. Ji, J.; Kulshreshtha, S.; Kakade, A.; Majeed, S.; Li, X.; Liu, P. Bioaugmentation of membrane bioreactor with Aeromonas hydrophila LZ-MG14 for enhanced malachite green and hexavalent chromium removal in textile wastewater. Int. Biodeterior Biodegrad. 2020, 150, 104939. [Google Scholar] [CrossRef]
  19. Kumar, N.; Sinha, S.; Mehrotra, T.; Singh, R.; Tandon, S.; Thakur, I.S. Biodecolorization of azo dye Acid Black 24 by Bacillus pseudomycoides: Process optimization using Box Behnken design model and toxicity assessment. Bioresour. Technol. Rep. 2019, 8, 100311. [Google Scholar] [CrossRef]
  20. Ayed, L.; Zmantar, T.; Bayar, S.; Cheref, A.; Achour, S.; Mansour, H.B.; Mzoughi, R. Potential use of probiotic consortium isolated from kefir for textile azo dye decolorization. J. Microbiol. Biotechnol. 2019, 29, 1629–1635. [Google Scholar] [CrossRef]
  21. Yadav, V.K.; Yadav, K.K.; Gacem, A.; Gnanamoorthy, G.; Ali, I.H.; Khan, S.H.; Jeon, B.-H.; Kamyab, H.; Inwati, G.K.; Choudhary, N. A novel approach for the synthesis of vaterite and calcite from incense sticks ash waste and their potential for remediation of dyes from aqueous solution. Sustain. Chem. Pharm. 2022, 29, 100756. [Google Scholar] [CrossRef]
  22. Patel, Y.; Chhaya, U.; Rudakiya, D.M.; Joshi, S. Biological Decolorization and Degradation of Synthetic Dyes: A Green Step Toward Sustainable Environment. In Microbial Rejuvenation of Polluted Environment; Panpatte, D.G., Jhala, Y.K., Eds.; Springer: Singapore, 2021; Volume 2, pp. 77–110. [Google Scholar] [CrossRef]
  23. Mansour, A.T.; Alprol, A.E.; Abualnaja, K.M.; El-Beltagi, H.S.; Ramadan, K.M.A.; Ashour, M. The Using of Nanoparticles of Microalgae in Remediation of Toxic Dye from Industrial Wastewater: Kinetic and Isotherm Studies. Materials 2022, 15, 3922. [Google Scholar] [CrossRef] [PubMed]
  24. Ashour, M.; Alprol, A.E.; Heneash, A.M.; Saleh, H.; Abualnaja, K.M.; Alhashmialameer, D.; Mansour, A.T. Ammonia bioremediation from aquaculture wastewater effluents using arthrospira platensis niof17/003: Impact of biodiesel residue and potential of ammonia-loaded biomass as rotifer feed. Materials 2021, 14, 5460. [Google Scholar] [CrossRef]
  25. Alprol, A.E.; Heneash, A.M.; Ashour, M.; Abualnaja, K.M.; Alhashmialameer, D.; Mansour, A.T.; Sharawy, Z.Z.; Abu-Saied, M.A.; Abomohra, A.E.F. Potential applications of arthrospira platensis lipid-free biomass in bioremediation of organic dye from industrial textile effluents and its influence on marine rotifer (Brachionus plicatilis). Materials 2021, 14, 4446. [Google Scholar] [CrossRef]
  26. Piaskowski, K.; Świderska-Dąbrowska, R.; Zarzycki, P.K. Dye removal from water and wastewater using various physical, chemical, and biological processes. J. AOAC Int. 2018, 101, 1371–1384. [Google Scholar] [CrossRef]
  27. Bafana, A.; Chakrabarti, T.; Muthal, P.; Kanade, G. Detoxification of benzidine-based azo dye by E. gallinarum: Time-course study. Ecotoxicol. Environ. Saf. 2009, 72, 960–964. [Google Scholar] [CrossRef]
  28. Rane, N.R.; Tapase, S.; Kanojia, A.; Watharkar, A.; Salama, E.S.; Jang, M.; Jeon, B.H. Molecular insights into plant–microbe interactions for sustainable remediation of contaminated environment. Bioresour. Technol. 2022, 344, 126246. [Google Scholar] [CrossRef]
  29. Dapson, R.W. Benzidine-based dyes: Effects of industrial practices, regulations, and world trade on the biological stains market. Biotech. Histochem. 2009, 84, 95–100. [Google Scholar] [CrossRef]
  30. Arun Prasad, A.S.; Kumar, G.; Mary Thomas, D. Microbial Decolorization of Azo Dyes—A Mini Review. Bull. Chem. Pharma Res. 2017, 1, 30–39. [Google Scholar]
  31. Yadav, V.K.; Gupta, N.; Kumar, P.; Dashti, M.G.; Tirth, V.; Khan, S.H.; Yadav, K.K.; Islam, S.; Choudhary, N.; Algahtani, A.; et al. Recent Advances in Synthesis and Degradation of Lignin and Lignin Nanoparticles and Their Emerging Applications in Nanotechnology. Materials 2022, 15, 953. [Google Scholar] [CrossRef] [PubMed]
  32. Mason, M.G.; Ball, A.S.; Reeder, B.J.; Silkstone, G.; Nicholls, P.; Wilson, M.T. Extracellular heme peroxidases in actinomycetes: A case of mistaken identity. Appl. Environ. Microbiol. 2001, 67, 4512–4519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Chater, K.F.; Biró, S.; Lee, K.J.; Palmer, T.; Schrempf, H. The complex extracellular biology of Streptomyces: Review article. FEMS Microbiol. Rev. 2010, 34, 171–198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Fodil, D.; Badis, A.; Jaouadi, B.; Zaraî, N.; Ferradji, F.Z.; Boutoumi, H. Purification and characterization of two extracellular peroxidases from Streptomyces sp. strain AM2, a decolorizing actinomycetes responsible for the biodegradation of natural humic acids. Int. Biodeterior. Biodegrad. 2011, 65, 470–478. [Google Scholar] [CrossRef]
  35. Blánquez, A.; Rodríguez, J.; Brissos, V.; Mendes, S.; Martins, L.O.; Ball, A.S.; Arias, M.E.; Hernández, M. Decolorization and detoxification of textile dyes using a versatile Streptomyces laccase-natural mediator system. Saudi J. Biol. Sci. 2019, 26, 913–920. [Google Scholar] [CrossRef]
  36. Dong, H.; Guo, T.; Zhang, W.; Ying, H.; Wang, P.; Wang, Y.; Chen, Y. Biochemical characterization of a novel azoreductase from Streptomyces sp.: Application in eco-friendly decolorization of azo dye wastewater. Int. J. Biol. Macromol. 2019, 140, 1037–1046. [Google Scholar] [CrossRef]
  37. Adenan, N.H.; Lim, Y.Y.; Ting, A.S.Y. Removal of triphenylmethane dyes by Streptomyces bacillaris: A study on decolorization, enzymatic reactions and toxicity of treated dye solutions. J. Environ. Manag. 2022, 318, 115520. [Google Scholar] [CrossRef]
  38. Kumar, P.; Gacem, A.; Ahmad, M.T.; Yadav, V.K.; Singh, S.; Yadav, K.K.; Alam, M.M.; Dawane, V.; Piplode, S.; Maurya, P.; et al. Environmental and human health implications of metal(loid)s: Source identification, contamination, toxicity, and sustainable clean-up technologies. Front. Environ. Sci. 2022, 1328. [Google Scholar] [CrossRef]
  39. Chakravarthi, B.; Mathakala, V.; Palempalli, U.M.D. Bioremediation of Textile Azo Dyes by Marine Streptomyces. In Recent Trends in Waste Water Treatment and Water Resource Management; Ghosh, S.K., das Saha, P., Di Francesco, M., Eds.; Springer: Singapore, 2020; pp. 129–142. [Google Scholar] [CrossRef]
  40. Zhou, W.; Zimmermann, W. Decolorization of industrial effluents containing reactive dyes by actinomycetes. FEMS Microbiol. Lett. 1993, 107, 157–161. [Google Scholar] [CrossRef] [Green Version]
  41. Adenan, N.H.; Lim, Y.Y.; Ting, A.S.Y. Identification and optimization of triphenylmethane dyes removal by Streptomyces sp. from forest soil. Sustain. Environ. Res. 2021, 31, 1–15. [Google Scholar] [CrossRef]
  42. John, J.; Dineshram, R.; Hemalatha, K.R.; Dhassiah, M.P.; Gopal, D.; Kumar, A. Bio-Decolorization of Synthetic Dyes by a Halophilic Bacterium Salinivibrio sp. Front. Microbiol. 2020, 11, 594011. [Google Scholar] [CrossRef] [PubMed]
  43. Akter, K.; Zerin, T.; Banik, A. Biodegradation of textile dyes by bacteria isolated from textile industry effluents. Stamford J. Microbiol. 2020, 9, 5–8. [Google Scholar] [CrossRef]
  44. Shete, U.R.; Waghmare, V.M.; Dhande, N.M. Isolation and screening of textile dye decolorizing bacteria from textile industrial effluent. Themat. J. Microbiol. 2019, 3. [Google Scholar] [CrossRef]
  45. Khaled, J.M.; Alyahya, S.A.; Govindan, R.; Chelliah, C.K.; Maruthupandy, M.; Alharbi, N.S.; Kadaikunnan, S.; Issac, R.; Murugan, S.; Li, W.J. Laccase producing bacteria influenced the high decolorization of textile azo dyes with advanced study. Environ. Res. 2022, 207, 112211. [Google Scholar] [CrossRef] [PubMed]
  46. Thanavel, M.; Bankole, P.O.; Kadam, S.; Govindwar, S.P.; Sadasivam, S.K. Desulfonation of the textile azo dye Acid Fast Yellow MR by newly isolated Aeromonas hydrophila SK16. Water Resour. Ind. 2019, 22, 100116. [Google Scholar] [CrossRef]
  47. Chen, B.-Y.; Lin, K.-W.; Wang, Y.-M.; Yen, C.-Y. Revealing interactive toxicity of aromatic amines to azo dye decolorizer Aeromonas hydrophila. J. Hazard. Mater. 2009, 166, 187–194. [Google Scholar] [CrossRef]
  48. Jadhav, U.U.; Dawkar, V.v.; Ghodake, G.S.; Govindwar, S.P. Biodegradation of Direct Red 5B, a textile dye by newly isolated Comamonas sp. UVS. J. Hazard. Mater. 2008, 158, 507–516. [Google Scholar] [CrossRef]
  49. Hossen, M.Z.; Hussain, M.E.; Hakim, A.; Islam, K.; Uddin, M.N.; Azad, A.K. Biodegradation of reactive textile dye Novacron Super Black G by free cells of newly isolated Alcaligenes faecalis AZ26 and Bacillus spp obtained from textile effluents. Heliyon 2019, 5, e02068. [Google Scholar] [CrossRef] [Green Version]
  50. Gomare, S.S.; Govindwar, S.P. Brevibacillus laterosporus MTCC 2298: A potential azo dye degrader. J. Appl. Microbiol. 2009, 106, 993–1004. [Google Scholar] [CrossRef]
  51. Fernando, E.; Keshavarz, T.; Kyazze, G. Enhanced bio-decolourisation of acid orange 7 by Shewanella oneidensis through co-metabolism in a microbial fuel cell. Int. Biodeterior. Biodegrad. 2012, 72, 1–9. [Google Scholar] [CrossRef]
  52. Angelova, B.; Avramova, T.; Stefanova, L.; Mutafov, S. Temperature effect on bacterial azo bond reduction kinetics: An Arrhenius plot analysis. Biodegradation 2008, 19, 387–393. [Google Scholar] [CrossRef] [PubMed]
  53. Patil, S.M.; Rane, N.R.; Bankole, P.O.; Krishnaiah, P.; Ahn, Y.; Park, Y.-K.; Yadav, K.K.; Amin, M.A.; Jeon, B.-H. An Assessment of Micro- and Nanoplastics in the Biosphere: A Review of Detection, Monitoring, and Remediation Technology. Chem. Eng. J. 2022, 430, 132913. [Google Scholar] [CrossRef]
  54. Karim, M.E.; Dhar, K.; Hossain, M.T. Decolorization of Textile Reactive Dyes by Bacterial Monoculture and Consortium Screened from Textile Dyeing Effluent. J. Genet. Eng. Biotechnol. 2018, 16, 375–380. [Google Scholar] [CrossRef] [PubMed]
  55. Maurya, P.K.; Ali, S.A.; Alharbi, R.S.; Yadav, K.K.; Alfaisal, F.M.; Ahmad, A.; Ditthakit, P.; Prasad, S.; Jung, Y.-K.; Jeon, B.-H. Impacts of Land Use Change on Water Quality Index in the Upper Ganges River near Haridwar, Uttarakhand: A GIS-Based Analysis. Water 2021, 13, 3572. [Google Scholar] [CrossRef]
  56. Patil, P.S.; Phugare, S.S.; Jadhav, S.B.; Jadhav, J.P. Communal action of microbial cultures for Red HE3B degradation. J. Hazard. Mater. 2010, 181, 263–270. [Google Scholar] [CrossRef]
  57. Patil, N.P.; Bholay, A.D.; Kapadnis, B.P.; Gaikwad, V.B. Biodegradation of model azo dye methyl red and other textile dyes by isolate Bacillus circulans npp1. J. Pure Appl. Microbiol. 2016, 10, 2793–2800. [Google Scholar] [CrossRef]
  58. Kalme, S.; Jadhav, S.; Jadhav, M.; Govindwar, S. Textile dye degrading laccase from Pseudomonas desmolyticum NCIM 2112. Enzyme Microb. Technol. 2009, 44, 65–71. [Google Scholar] [CrossRef]
  59. Ghodake, G.; Jadhav, U.; Tamboli, D.; Kagalkar, A.; Govindwar, S. Decolorization of Textile Dyes and Degradation of Mono-Azo Dye Amaranth by Acinetobacter calcoaceticus NCIM 2890. Indian J. Microbiol. 2011, 51, 501–508. [Google Scholar] [CrossRef] [Green Version]
  60. Ahn, H.J.; Ahn, Y.; Kurade, M.B.; Patil, S.M.; Ha, G.S.; Bankole, P.O.; Khan, M.A.; Chang, S.W.; Abdellattif, M.H.; Yadav, K.K.; et al. The comprehensive effects of aluminum oxide nanoparticles on the physiology of freshwater microalga Scenedesmus obliquus and it’s phycoremediation performance for the removal of sulfacetamide. Environ. Res. 2022, 215, 114314. [Google Scholar] [CrossRef]
  61. Buthelezi, S.P.; Olaniran, A.O.; Pillay, B. Textile dye removal from wastewater effluents using bioflocculants produced by indigenous bacterial isolates. Molecules 2012, 17, 14260–14274. [Google Scholar] [CrossRef] [PubMed]
  62. Aloius Togo, C.; Clifford Zvandada Mutambanengwe, C.; George Whiteley, C. Decolourisation and degradation of textile dyes using a sulphate reducing bacteria (SRB)-biodigester microflora co-culture. Afr. J. Biotechnol. 2008, 7, 114–121. [Google Scholar]
  63. Al-Tohamy, R.; Sun, J.; Fareed, M.F.; Kenawy, E.R.; Ali, S.S. Ecofriendly biodegradation of Reactive Black 5 by newly isolated Sterigmatomyces halophilus SSA1575, valued for textile azo dye wastewater processing and detoxification. Sci. Rep. 2020, 10, 12370. [Google Scholar] [CrossRef] [PubMed]
  64. Choudhary, N.; Yadav, V.K.; Yadav, K.K.; Almohana, A.I.; Almojil, S.F.; Gnanamoorthy, G.; Kim, D.-H.; Islam, S.; Kumar, P.; Jeon, B.-H. Application of Green Synthesized MMT/Ag Nanocomposite for Removal of Methylene Blue from Aqueous Solution. Water 2021, 13, 3206. [Google Scholar] [CrossRef]
  65. Haq, I.; Raj, A.; Markandeya. Biodegradation of Azure-B dye by Serratia liquefaciens and its validation by phytotoxicity, genotoxicity and cytotoxicity studies. Chemosphere 2018, 196, 58–68. [Google Scholar] [CrossRef]
  66. Sheela, T.; Sadasivam, S.K. Dye degradation potential and its degradative enzymes synthesis of bacillus cereus skb12 isolated from a textile industrial effluent. J. Appl. Biol. Biotechnol. 2020, 8, 42–46. [Google Scholar] [CrossRef]
  67. Liu, W.; Liu, C.; Liu, L.; You, Y.; Jiang, J.; Zhou, Z.; Dong, Z. Simultaneous decolorization of sulfonated azo dyes and reduction of hexavalent chromium under high salt condition by a newly isolated salt-tolerant strain Bacillus circulans BWL1061. Ecotoxicol. Environ. Saf. 2017, 141, 9–16. [Google Scholar] [CrossRef]
  68. Mohanty, S.S.; Kumar, A. Enhanced degradation of anthraquinone dyes by microbial monoculture and developed consortium through the production of specific enzymes. Sci. Rep. 2021, 11, 7678. [Google Scholar] [CrossRef]
  69. Schmidt, C.; Berghahn, E.; Ilha, V.; Granada, C.E. Biodegradation potential of Citrobacter cultures for the removal of amaranth and congo red azo dyes. Int. J. Environ. Sci. Technol. 2019, 16, 6863–6872. [Google Scholar] [CrossRef]
  70. Sarim, K.M.; Kukreja, K.; Shah, I.; Choudhary, C.K. Biosorption of direct textile dye Congo red by Bacillus subtilis HAU-KK01. Bioremediat. J. 2019, 23, 185–195. [Google Scholar] [CrossRef]
  71. Garg, N.; Garg, A.; Mukherji, S. Eco-friendly decolorization and degradation of reactive yellow 145 textile dye by Pseudomonas aeruginosa and Thiosphaera pantotropha. J. Environ. Manag. 2020, 263, 110383. [Google Scholar] [CrossRef] [PubMed]
  72. Velayutham, K.; Madhava, A.K.; Pushparaj, M.; Thanarasu, A.; Devaraj, T.; Periyasamy, K.; Subramanian, S. Biodegradation of Remazol Brilliant Blue R using isolated bacterial culture (Staphylococcus sp. K2204). Environ. Technol. 2018, 39, 2900–2907. [Google Scholar] [CrossRef] [PubMed]
  73. Holkar, C.R.; Arora, H.; Halder, D.; Pinjari, D.V. Biodegradation of reactive blue 19 with simultaneous electricity generation by the newly isolated electrogenic Klebsiella sp. C NCIM 5546 bacterium in a microbial fuel cell. Int. Biodeterior. Biodegrad. 2018, 133, 194–201. [Google Scholar] [CrossRef]
  74. Hansda, A.; Kumar, V.; Anshumali. A comparative review towards potential of microbial cells for heavy metal removal with emphasis on biosorption and bioaccumulation. World J. Microbiol. Biotechnol. 2016, 32, 170. [Google Scholar] [CrossRef]
  75. Singh, G.; Dwivedi, S.K.; Mishra, J. Role of Fungal Enzymes in the Removal of Azo Dyes. In Microbial Enzymes: Roles and Applications in Industries; Arora, N.K., Mishra, J., Mishra, V., Eds.; Springer: Singapore, 2020; pp. 231–257. [Google Scholar] [CrossRef]
  76. Moiseenko, K.V.; Vasina, D.V.; Farukshina, K.T.; Savinova, O.S.; Glazunova, O.A.; Fedorova, T.V.; Tyazhelova, T.V. Orchestration of the expression of the laccase multigene family in white-rot basidiomycete Trametes hirsuta 072: Evidences of transcription level subfunctionalization. Fungal Biol. 2018, 122, 353–362. [Google Scholar] [CrossRef]
  77. Kiran, S.; Huma, T.; Jalal, F.; Farooq, T.; Hameed, A.; Gulzar, T.; Bashir, A.; Rahmat, M.; Rahmat, R.; Rafique, M.A. Lignin degrading system of Phanerochaete chrysosporium and its exploitation for degradation of synthetic dyes wastewater. Pol. J. Environ. Stud. 2019, 28, 1749–1757. [Google Scholar] [CrossRef]
  78. Valliyaparambil, P.T.; Kaliyaperumal, K.A.; Gopakumaran, N. Pleurotus ostreatus laccase decolorization of remazol brilliant violet 5R dye: Statistical optimization and toxicity studies on microbes and its kinetics. J. Appl. Biotechnol. Rep. 2019, 6, 88–95. [Google Scholar] [CrossRef] [Green Version]
  79. Palmieri, G.; Cennamo, G.; Sannia, G. Remazol Brilliant Blue R decolourisation by the fungus Pleurotus ostreatus and its oxidative enzymatic system. Enzyme Microb. Technol. 2005, 36, 17–24. [Google Scholar] [CrossRef]
  80. Kim, S.J.; Shoda, M. Decolorization of molasses and a dye by a newly isolated strain of the fungus Geotrichum candidum Dec 1. Biotechnol. Bioeng. 1999, 62, 114–119. [Google Scholar] [CrossRef]
  81. Lan, J.; Huang, X.; Hu, M.; Li, Y.; Qu, Y.; Gao, P.; Wu, D. High efficient degradation of dyes with lignin peroxidase coupled with glucose oxidase. J. Biotechnol. 2006, 123, 483–490. [Google Scholar] [CrossRef]
  82. Svobodová, K.; Erbanová, P.; Sklenář, J.; Novotný, Č. The role of Mn-dependent peroxidase in dye decolorization by static and agitated cultures ofIrpex lacteus. Folia Microbiol. 2006, 51, 573–578. [Google Scholar] [CrossRef] [PubMed]
  83. Cuamatzi-Flores, J.; Esquivel-Naranjo, E.; Nava-Galicia, S.; López-Munguía, A.; Arroyo-Becerra, A.; Villalobos-López, M.A.; Bibbins-Martínez, M. Differential regulation of Pleurotus ostreatus dye peroxidases gene expression in response to dyes and potential application of recombinant Pleos-DyP1 in decolorization. PLoS ONE 2019, 14, e0209711. [Google Scholar] [CrossRef] [Green Version]
  84. Lueangjaroenkit, P.; Teerapatsakul, C.; Sakka, K.; Sakka, M.; Kimura, T.; Kunitake, E.; Chitradon, L. Two Manganese Peroxidases and a Laccase of Trametes polyzona KU-RNW027 with Novel Properties for Dye and Pharmaceutical Product Degradation in Redox Mediator-Free System. Mycobiology 2019, 47, 217–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Verma, P.; Madamwar, D. Production of ligninolytic enzymes for dye decolorization by cocultivation of white-rot fungi Pleurotus ostreatus and Phanerochaete chrysosporium under solid-state fermentation. Appl. Biochem. Biotechnol. 2002, 102, 109–118. [Google Scholar] [CrossRef]
  86. Juárez-Hernández, J.; Castillo-Hernández, D.; Pérez-Parada, C.; Nava-Galicia, S.; Cuervo-Parra, J.A.; Surian-Cruz, E.; Díaz-Godínez, G.; Sánchez, C.; Bibbins-Martínez, M. Isolation of fungi from a textile industry effluent and the screening of their potential to degrade industrial dyes. J. Fungi 2021, 7, 805. [Google Scholar] [CrossRef]
  87. Song, Z.; Song, L.; Shao, Y.; Tan, L. Degradation and detoxification of azo dyes by a salt-tolerant yeast Cyberlindnera samutprakarnensis S4 under high-salt conditions. World J. Microbiol. Biotechnol. 2018, 34, 131. [Google Scholar] [CrossRef]
  88. Sharma, P.; Singh, L.; Dilbaghi, N. Biodegradation of Orange II Dye by Phanerochaete chrysosporium in Simulated Wastewater; CSIR: New Delhi, India, 2009; Volume 68. [Google Scholar]
  89. Harazono, K.; Watanabe, Y.; Nakamura, K. Decolorization of azo dye by the white-rot basidiomycete Phanerochaete sordida and by its manganese peroxidase. J. Biosci. Bioeng. 2003, 95, 455–459. [Google Scholar] [CrossRef]
  90. Yao, J.; Yao, J.; Jia, R.; Zheng, L.; Wang, B. Rapid decolorization of azo dyes by crude manganese peroxidase from Schizophyllum sp. F17 in solid-state fermentation. Biotechnol. Bioprocess Eng. 2013, 18, 868–877. [Google Scholar] [CrossRef]
  91. Lanfranconi, I.; Ceretta, M.B.; Bertola, N.; Wolski, E.A.; Durruty, I. Textile dyeing wastewater treatment by Penicillium chrysogenum: Design of a sustainable process. Water Sci. Technol. 2022, 86, 292–301. [Google Scholar] [CrossRef]
  92. Argumedo-Delira, R.; Gómez-Martínez, M.J.; Uribe-Kaffure, R. Trichoderma biomass as an alternative for removal of congo red and malachite green industrial dyes. Appl. Sci. 2021, 11, 448. [Google Scholar] [CrossRef]
  93. He, X.L.; Song, C.; Li, Y.Y.; Wang, N.; Xu, L.; Han, X.; Wei, D.S. Efficient degradation of Azo dyes by a newly isolated fungus Trichoderma tomentosum under non-sterile conditions. Ecotoxicol. Environ. Saf. 2018, 150, 232–239. [Google Scholar] [CrossRef] [PubMed]
  94. Abd El-Rahim, W.M.; El-Meleigy, M.A.; Ibrahim, E. Bioremoval of Textile Effluent Dye by Aspergillus fumigates. Int. J. Water Resour. Arid. Environ. 2018, 7, 70–78. [Google Scholar]
  95. Sudiana, I.K.; Citrawathi, D.M.; Sastrawidana, I.D.K.; Maryam, S.; Sukarta, I.N.; Wirawan, G.A. Biodegradation of Turquoise Blue Textile Dye by Wood Degrading Local Fungi Isolated from Plantation Area. J. Ecol. Eng. 2022, 23, 205–214. [Google Scholar] [CrossRef]
  96. Rani, B.; Kumar, V.; Singh, J.; Bisht, S.; Teotia, P.; Sharma, S.; Kela, R. Bioremediation of dyes by fungi isolated from contaminated dye effluent sites for bio-usability. Braz. J. Microbiol. 2014, 45, 1055–1063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Almeida, E.J.R.; Corso, C.R. Decolorization and removal of toxicity of textile azo dyes using fungal biomass pelletized. Int. J. Environ. Sci. Technol. 2019, 16, 1319–1328. [Google Scholar] [CrossRef] [Green Version]
  98. Bergsten-Torralba, L.R.; Nishikawa, M.M.; Baptista, D.F.; Magalhães, D.D.P.; Silva, M.D. Decolorization of different textile dyes by Penicillium simplicissimum and toxicity evaluation after fungal treatment. Braz. J. Microbiol. 2009, 40, 808–817. [Google Scholar] [CrossRef] [Green Version]
  99. Chen, S.H.; Cheow, Y.L.; Ng, S.L.; Ting, A.S.Y. Biodegradation of Triphenylmethane Dyes by Non-white Rot Fungus Penicillium simplicissimum: Enzymatic and Toxicity Studies. Int. J. Environ. Res. 2019, 13, 273–282. [Google Scholar] [CrossRef]
  100. Bankole, P.O.; Adekunle, A.A.; Govindwar, S.P. Biodegradation of a monochlorotriazine dye, cibacron brilliant red 3B-A in solid state fermentation by wood-rot fungal consortium, Daldinia concentrica and Xylaria polymorpha: Co-biomass decolorization of cibacron brilliant red 3B-A dye. Int. J. Biol. Macromol. 2018, 120, 19–27. [Google Scholar] [CrossRef]
  101. Mathur, M.; Vijayalakshmi, K.S.; Gola, D.; Singh, K.; Chaudhary, S.; Kaushik, P.; Malik, A. Decolourization of Textile Dyes by Aspergillus lentulus. J. Basic Appl. Eng. Res. 2015, 2, 1469–1473. [Google Scholar]
  102. Sumathi, S.; Manju, B.S. Uptake of reactive textile dyes by Aspergillus foetidus. Enzyme Microb. Technol. 2000, 27, 347–355. [Google Scholar] [CrossRef]
  103. Corso, C.R.; Maganha De Almeida, A.C. Bioremediation of dyes in textile effluents by Aspergillus oryzae. Microb. Ecol. 2009, 57, 384–390. [Google Scholar] [CrossRef] [PubMed]
  104. Esmaeili, A.; Kalantari, M. Bioremoval of an azo textile dye, Reactive Red 198, by Aspergillus flavus. World J. Microbiol. Biotechnol. 2012, 28, 1125–1131. [Google Scholar] [CrossRef]
  105. De Castro, K.C.; Leme, V.F.C.; Souza, F.H.M.; Costa, G.O.B.; Santos, G.E.; Litordi, L.R.V.; Andrade, G.S.S. Performance of inactivated Aspergillus oryzae cells on dye removal in aqueous solutions. Environ. Technol. Innov. 2021, 24, 101828. [Google Scholar] [CrossRef]
  106. Singh, L. Microbial decolourization of textile dyes by the fungus Trichoderma harzianum Characterization of Dye Degrading Fungi and Metabolites produced by some Endophytic Fungi View project. J. Pure Appl. Microbiol. 2012, 6, 1829–1833. [Google Scholar]
  107. Mahmoud, M.S. Decolorization of certain reactive dye from aqueous solution using Baker’s Yeast (Saccharomyces cerevisiae) strain. HBRC J. 2016, 12, 88–98. [Google Scholar] [CrossRef] [Green Version]
  108. Azin, E.; Moghimi, H. Efficient mycosorption of anionic azo dyes by Mucor circinelloides: Surface functional groups and removal mechanism study. J. Environ. Chem. Eng. 2018, 6, 4114–4123. [Google Scholar] [CrossRef]
  109. Jamee, R.; Siddique, R. Biodegradation of synthetic dyes of textile effluent by microorganisms: An environmentally and economically sustainable approach. Eur. J. Microbiol. Immunol. 2019, 9, 114–118. [Google Scholar] [CrossRef]
  110. Gangola, S.; Bhatt, P.; Chaudhary, P.; Khati, P.; Kumar, N.; Sharma, A. Bioremediation of Industrial Waste Using Microbial Metabolic Diversity. In Microbial Biotechnology in Environmental Monitoring and Cleanup; IGI Global: Hershey, PA, USA, 2018; pp. 1–27. [Google Scholar] [CrossRef]
  111. Chin, J.Y.; Chng, L.M.; Leong, S.S.; Yeap, S.P.; Yasin, N.H.M.; Toh, P.Y. Removal of Synthetic Dye by Chlorella vulgaris Microalgae as Natural Adsorbent. Arab. J. Sci. Eng. 2020, 45, 7385–7395. [Google Scholar] [CrossRef]
  112. El-Kassas, H.Y.; Mohamed, L.A. Bioremediation of the textile waste effluent by Chlorella vulgaris. Egypt J. Aquat. Res. 2014, 40, 301–308. [Google Scholar] [CrossRef] [Green Version]
  113. Lim, S.-L.; Chu, W.-L.; Phang, S.-M. Use of Chlorella vulgaris for bioremediation of textile wastewater. Bioresour. Technol. 2010, 101, 7314–7322. [Google Scholar] [CrossRef]
  114. Chu, W.-L.; See, Y.-C.; Phang, S.-M. Use of immobilised Chlorella vulgaris for the removal of colour from textile dyes. J. Appl. Phycol. 2008, 21, 641. [Google Scholar] [CrossRef]
  115. Boulkhessaim, S.; Gacem, A.; Khan, S.H.; Amari, A.; Yadav, V.K.; Harharah, H.N.; Elkhaleefa, A.M.; Yadav, K.K.; Rather, S.-u.; Ahn, H.-J.; et al. Emerging Trends in the Remediation of Persistent Organic Pollutants Using Nanomaterials and Related Processes: A Review. Nanomaterials 2022, 12, 2148. [Google Scholar] [CrossRef] [PubMed]
  116. Ishchi, T.; Sibi, G. Azo Dye Degradation by Chlorella vulgaris: Optimization and Kinetics. Int. J. Biol. Chem. 2019, 14, 1–7. [Google Scholar] [CrossRef]
  117. Dellamatrice, P.M.; Silva-Stenico, M.E.; de Moraes, L.A.B.; Fiore, M.F.; Monteiro, R.T.R. Degradation of textile dyes by cyanobacteria. Braz. J. Microbiol. 2017, 48, 25–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Mahalakshmi, S.; Lakshmi, D.; Menaga, U. Biodegradation of Different Concentration of dye (Congo red dye) by using Green and Blue Green Algae. Int. J. Environ. Res. 2015, 9, 735–744. [Google Scholar]
  119. Shah, V.; Garg, N.; Madamwar, D. An integrated process of textile dye removal and hydrogen evolution using cyanobacterium, Phormidium valderianum. World J. Microbiol. Biotechnol. 2001, 17, 499–504. [Google Scholar] [CrossRef]
  120. Pathak, V.V.; Kothari, R.; Chopra, A.K.; Singh, D.P. Experimental and kinetic studies for phycoremediation and dye removal by Chlorella pyrenoidosa from textile wastewater. J. Environ. Manag. 2015, 163, 270–277. [Google Scholar] [CrossRef]
  121. Parikh, A.; Madamwar, D. Textile dye decolorization using cyanobacteria. Biotechnol. Lett. 2005, 27, 323–326. [Google Scholar] [CrossRef]
  122. Kalasariya, H.S.; Patel, N.B.; Gacem, A.; Alsufyani, T.; Reece, L.M.; Yadav, V.K.; Awwad, N.S.; Ibrahium, H.A.; Ahn, Y.; Yadav, K.K.; et al. Marine Alga Ulva fasciata-Derived Molecules for the Potential Treatment of SARS-CoV-2: An In Silico Approach. Mar. Drugs 2022, 20, 586. [Google Scholar] [CrossRef]
  123. Ghazal, F.M.; Mahdy, E.S.M.; El-Fattah, M.S.A.; EL-Sadany, A.E.G.Y.; Elg, Y.; Doha, N.M.E. The Use of Microalgae in Bioremediation of the Textile Wastewater Effluent. Nat. Sci. 2018, 16, 98–104. [Google Scholar] [CrossRef]
  124. Maruthanayagam, A.; Mani, P.; Kaliappan, K.; Chinnappan, S. In vitro and In silico Studies on the Removal of Methyl Orange from Aqueous Solution Using Oedogonium subplagiostomum AP1. Water Air Soil Pollut. 2020, 231, 232. [Google Scholar] [CrossRef]
  125. Bijekar, S.; Padariya, H.D.; Yadav, V.K.; Gacem, A.; Hasan, M.A.; Awwad, N.S.; Yadav, K.K.; Islam, S.; Park, S.; Jeon, B.-H. The State of the Art and Emerging Trends in the Wastewater Treatment in Developing Nations. Water 2022, 14, 2537. [Google Scholar] [CrossRef]
  126. Gacem, A.; Modi, S.; Yadav, V.K.; Islam, S.; Patel, A.; Dawane, V.; Jameel, M.; Inwati, G.K.; Piplode, S.; Solanki, V.S.; et al. Recent Advances in Methods for Synthesis of Carbon Nanotubes and Carbon Nanocomposite and their Emerging Applications: A Descriptive Review. J. Nanomater. 2022, 2022, 7238602. [Google Scholar] [CrossRef]
  127. Sherifah, M.W.; Seun, A.E.; Kehinde, O.S.; Abiodun, A.O. Decolourization of synthetic dyes by laccase enzyme produced by Kluyveromyces dobzhanskii DW1 and Pichia manshurica DW2. Afr. J. Biotechnol. 2019, 18, 1–11. [Google Scholar] [CrossRef]
  128. Darvishi, F.; Moradi, M.; Jolivalt, C.; Madzak, C. Laccase production from sucrose by recombinant Yarrowia lipolytica and its application to decolorization of environmental pollutant dyes. Ecotoxicol. Environ. Saf. 2018, 165, 278–283. [Google Scholar] [CrossRef]
  129. Othman, A.M.; Elsayed, M.A.; Elshafei, A.M.; Hassan, M.M. Purification and biochemical characterization of two isolated laccase isoforms from Agaricus bisporus CU13 and their potency in dye decolorization. Int. J. Biol. Macromol. 2018, 113, 1142–1148. [Google Scholar] [CrossRef]
  130. Dai, S.; Yao, Q.; Yu, G.; Liu, S.; Yun, J.; Xiao, X.; Deng, Z.; Li, H. Biochemical Characterization of a Novel Bacterial Laccase and Improvement of Its Efficiency by Directed Evolution on Dye Degradation. Front. Microbiol. 2021, 12, 633004. [Google Scholar] [CrossRef]
  131. Zhang, H.; Zhang, X.; Geng, A. Expression of a novel manganese peroxidase from Cerrena unicolor BBP6 in Pichia pastoris and its application in dye decolorization and PAH degradation. Biochem. Eng. J. 2020, 153, 107402. [Google Scholar] [CrossRef]
  132. Bouacem, K.; Rekik, H.; Jaouadi, N.Z.; Zenati, B.; Kourdali, S.; El Hattab, M.; Badis, A.; Annane, R.; Bejar, S.; Hacene, H.; et al. Purification and characterization of two novel peroxidases from the dye-decolorizing fungus Bjerkandera adusta strain CX-9. Int. J. Biol. Macromol. 2018, 106, 636–646. [Google Scholar] [CrossRef]
  133. Shaheen, R.; Asgher, M.; Hussain, F.; Bhatti, H.N. Immobilized lignin peroxidase from Ganoderma lucidum IBL-05 with improved dye decolorization and cytotoxicity reduction properties. Int. J. Biol. Macromol. 2017, 103, 57–64. [Google Scholar] [CrossRef]
  134. Campos, R.; Kandelbauer, A.; Robra, K.H.; Cavaco-Paulo, A.; Gübitz, G.M. Indigo degradation with purified laccases from Trametes hirsuta and Sclerotium rolfsii. J. Biotechnol. 2001, 89, 131–139. [Google Scholar] [CrossRef] [Green Version]
  135. Wang, J.; Lu, L.; Feng, F. Improving the indigo carmine decolorization ability of a bacillus amyloliquefaciens laccase by site-directed mutagenesis. Catalysts 2017, 7, 275. [Google Scholar] [CrossRef] [Green Version]
  136. Xiong, J.Q.; Cui, P.; Ru, S.; Kurade, M.B.; Patil, S.M.; Yadav, K.K.; Fallatah, A.M.; Cabral-Pinto, M.M.S.; Jeon, B.H. A comprehensive review on the effects of engineered nanoparticles on microalgal treatment of pollutants from wastewater. J. Clean. Prod. 2022, 344, 131121. [Google Scholar] [CrossRef]
Figure 1. Types of dyes based on their adopted from Shamsani et al., 2020 [5] with permission.
Figure 1. Types of dyes based on their adopted from Shamsani et al., 2020 [5] with permission.
Water 14 03163 g001
Figure 2. Dyes along with their representative members.
Figure 2. Dyes along with their representative members.
Water 14 03163 g002
Figure 3. Various methods of treatment of textile dyes.
Figure 3. Various methods of treatment of textile dyes.
Water 14 03163 g003
Figure 4. Various types of microbial methods for remediation of textile dyes.
Figure 4. Various types of microbial methods for remediation of textile dyes.
Water 14 03163 g004
Figure 5. Various types of microbial degradation of dyes based on oxygen environment.
Figure 5. Various types of microbial degradation of dyes based on oxygen environment.
Water 14 03163 g005
Figure 6. Basic mechanism for microbial biodegradation of azo dyes [30].
Figure 6. Basic mechanism for microbial biodegradation of azo dyes [30].
Water 14 03163 g006
Figure 7. Mechanism of degradation of Indigo dye by using laccase enzyme adopted from [134].
Figure 7. Mechanism of degradation of Indigo dye by using laccase enzyme adopted from [134].
Water 14 03163 g007
Figure 8. Genetically modified laccase from B. amyloliquefaciens for improving indigo carmine decolorization [135].
Figure 8. Genetically modified laccase from B. amyloliquefaciens for improving indigo carmine decolorization [135].
Water 14 03163 g008
Table 1. Degradation of dyes by using actinomycetes.
Table 1. Degradation of dyes by using actinomycetes.
ActinomycetesDyes UsedEfficiencyReferences
Streptomyces sp. S27MR99%[36]
Streptomyces bacillarisTriphenylmethane dyes: MG,
MV, CV, and CB
MG (94.7%), MV (91.8%),
CV (86.6%), CB (68.4%)
[37]
Streptomyces ipomoeae
CECT 3341
AO636–70%[38]
Streptomyces sviceus (marine)CR-21 [39]
Streptomyces spsRR 147 (azo)
RR 171
RB 114
RB 209
RB 116
3–100%[40]
Streptomyces spsMG,
MV,
CV
CB
(MG 95%, MV 92%, CV 87%,
CB 68%).
[41]
Table 2. Microbial remediation of dyes by using aerobic and aerobic bacteria.
Table 2. Microbial remediation of dyes by using aerobic and aerobic bacteria.
BacteriaDyes UsedEfficiencyReferences
Aeromonas hydrophila
A.   Hydrophila SK16
RR141
MG
Acid fast yellow MR
70–80%
96.8%
91.25%
[18,46,47]
Comamonas sps UVSDB GL, DB5B [48]
Bacillus sp. and
Staph. aureus
BY HP-2R, BS-RS, and B. blue RS 01 (BB)71–83%[43]
Alcaligenes faecalis AZ26,
Bacillus cereus AZ27 and Bacillus sp.
NSB-G90%[49]
Brevibacillus laterosporusDB-MR,
MO, Blue-2B,
Golden yellow (GY)
Brilliant blue
100%[50]
Shewanella oneidensis (MFC)Acid orange (AO 7)80.4[51]
Alcaligenes faecalis and Rhodococcus erythropolisMonoazo dye Acid orange80–95[52]
Consortium of
Ps. aeruginosa, B. pumilis, B. thuringiensis, Enterococcus faecium in different combinations
NR,
N Black,
N Blue dk,
N Navy,
N Yellow
82–97%[53]
Neisseria sp., Vibrio sp., Bacillus sp., Bacillus sp. and Aeromonas spsN Orange FN-R, N Brilliant Blue FN-R, N Super Black G, BY S8-G, and BR S2-B60–90%[54]
Acinetobacter (ST16.16/164 and Klebsiella (ST16.16/034)Monoazo dye RO 16 and diazo dye RG-19More than 80%[55]
Salinivibrio kushneri HTSPCBB G-250 and CR and safraninMore than 80%[42]
Pseudomonas sp. (S3, S11 and S12), Klebisiella sp. (S4). and Aeromonas sp. (S2)JR, JB and JGY45–50%[44]
Bacillus odyssey SUK3, Morganella morganii SUK5 and Proteus sps SUK 7Red HE3B, Reactive Blue 5997–99%[56]
B. circulans NPP1MR and other dyes98%[57]
Ps. desmolyticum NCIM 2112DB 6, Red HEB7, and Green HE4B70, 90%[58]
Acinetobacter calcoaceticusMR, MO90–98%[59]
Enterobacter sp. CV–S1CV [60]
B. subtilis (E1),
Exiguobacterium acetylicum (D1),
Klebsiella terrigena (R2),
S. aureus (A22),
Ps. pseudoalcaligenes (A17), and
Ps. plecoglossicida (A14)
Whale,
mediblue,
fawn and
mixed dye,
97.04%, 80.61%, 94.93% and 81.64%[61]
Sulfate-reducing bacteria (SRB)Orange II95%[62]
Sterigmatomyces halophilus SSA1 575RB-5, RB-5100%[63]
Anoxybacillus ayderensis SK3-4DG 6100%[64]
A.   hydrophila SK 16AFY MRMore than 70[46]
Serratia liquifaciensAzure B90%[65]
Bacillus cereus SKB12RB-588.7% [66]
B. cereus
and
Ps. parafulva
Azo dyes: T-blue, yellow GR, and orange 3R91.69 and 89.21%[45]
B.   circulans BWLI 061MO99.22%[67]
Bacterial consortium-Bacillus flexus TS8 (BF),
Proteus mirabilis PMS (PM), and
Ps. aeruginosa NCH (PA)
Indanthrene Blue RS99%[68]
Bacillus pseudomycoidesMG96.8%[19]
Citrobacter sp.CR93%[69]
Bacillus subtilis HAU-KK01CR92.8%[70]
Thiosphaera pantotrophaRY100%[71]
Staphylococcus sp. K2204RBB100%[72]
Staphylococcus sp. Reactive blue97%[73]
Table 3. Dye degradation by using white rot and non-white rot fungi.
Table 3. Dye degradation by using white rot and non-white rot fungi.
FungiDyes UsedEfficiencyReferences
P. chrysosporiumRhodamine-B91%[81]
P. chrysosporiumBromophenol blue99.3%[82]
Pleurotus ostreatusAcid Blue 12977%[83]
Trametes polyzona KU-RNW027RB-B100%[84]
P. chrysosporiumRanocid Fast Blue83%[85]
Emmia latemarginata (MAP03, MAP04, and MAP05) and Mucor circinelloides (MAP01, MAP02, and MAP06)Azo, indigo, and anthraquinone dyes (Acetyl yellow G) [86]
Cyberlindnera samutprakarnensis S4Acid Red97%[87]
P. chrysosporiumOrange-II85%[88]
Phanerochaete sordidaRR-12090.6[89]
Schizophyllum sp. F17Orange IV76%[90]
Penicillium chrysogenumDB22 [91]
Trichoderma (T. virens, T. viride)CR and MG81–99%[92]
Trichoderma tomentosumAR-3R99.2%[93]
A. fumigatusDirect Violet dye51.38–93.74%[94]
Trametes hirsuta, Microporus xanthopus, and Ganoderma applanatumturquoise blue textile dye82.17%
78.50 and 85.84%
[95]
Aspergillus niger, and Phanerochaete chrysosporium,BF (81.85%),
Nigrosin (77.47%), MG (72.77%)
P. chrysosporium maxium with Nigrosin (90.15%) > BF (89.8%), MG (83.25%) [96]
Aspergillus niger, Asp. terreus and Rhizopus oligosporus.Acid Blue 161 e Procion Red MX-5B58–68%[97]
Penicillium simplicissimum INCQS 40211(RR198),
R B214 (RB214), RB21, and a mixture of all three
[98]
P. simplicissimum (n-WRF)Triphenylmethane Dyes: CV, MV, MG, CB96.1–98.7%[99]
Diutina rugosaIndigo100% [100]
Asp. lentulusVAT Novatic Grey (100%), RR-Red (98.47%), RY (91.55%), and Indanthrane Blue (99.28%)
Reactive S
91–100%
55.33%
[101]
Aspergillus foetidusAzo reactive dyes, Drimarene dyes>95%[102]
Aspergillus oryzaeReactive dyes PR-HE7B and PV-H3R [103]
Aspergillus sp.,Fast Red A99.5%
Aspergillus flavusAzo textile dye, RR19884.96%[104]
Aspergillusoryzaedead cellsRB, MB, and MO100%[105]
Trichoderma harzianumAR and anthraquinone (Basic blue)70% and
51%
[106]
Saccharomyces cerevisiaeRamazole Blue100%[107]
Mucor circinelloidesRamazole Blue95% [108]
Table 4. Algae-mediated decolourization of textile dyes.
Table 4. Algae-mediated decolourization of textile dyes.
AlgaeDyes UsedEfficiencyReferences
Chlorella vulgarisMB and AO783.04  ±  2.94%[111]
C. vulgarisAzo dye: Metanil yellow, FO, FR, DB, AF Red, Direct Fast Scarlet, CR, and Acid-Fast N BlueUp to (75.68%)[112]
C. vulgaris UMACC 001Supranol Red 3BW (SR 3BW)41.8% to 50.0%.[113]
C. vulgaris
(Immobilised)
SB 3BW, Lanaset Red 2GA and Levafix Navy Blue EBNA44.0–49%[114]
Chlorella vulgarisBlue dye
Green dye
Reactive black
Blue: 63.89% (1 mg/L)
Green: 45.71% (1 mg/L)
80%
[115,116]
Sphaerocystis schroeteriBlue dye
Green dye
Blue: 63.87% (1 mg/L)
Green: 60.00% (1 mg/L)
[115]
Anabaena flos-aquae UTCC64,
Phormidium autumnale UTEX1580
and Synechococcus sp. PCC7942
Indigo, RBBR and Sulphur Black)80–90%[117]
Haematococcus sp., Chlorella sp.,
Chlorella vulgaris, Scenedesmus obliquuss,
S. officinalis, and S. quadricauda
Arthospira maxima (Blue green algae)
Up to 98%[118]
Phormidium valderianum, a marine cyanobacteriumAR, AR-119 and DB-155, (at Ph > 11)>90%[119]
Chlorella pyrenoidosaMB
Rhodamine B
84%[120]
Gloeocapsa pleurocapsoides and Phormidium ceylanicumAR 97 and FF Sky Blue dyes 80% after 26 days[121]
Achaetomium strumariumAcid red99.5% [122]
Anabaena flos aquae, Nostoc elepsosporum, Nostoc linkia,
Anabaena variabilis and Chlorella vulgaris
Azo dyesN. elepsosporum (100%),
C. vulgaris (96.16%),
A. variabilis (88.71%),
N. linkia (88.71) and
A. flos aquae (50.81).
[123]
Arthrospira platensis NIOF17/003Methylene Blue93% [23]
A. platensis complete drybiomss and Lipid-Free BiomassIsmate violet 2R, IV2R75.7% and 61.11%[25]
A. platensis complete drybiomss and Lipid-Free BiomassAmmonium ions (NH4+)In synthetic aquaculture: ACDW (64.24%)
LFB (89.68%)
In real aquaculture:
ACDW (25.7%)
LFB (37.8%)
[24]
Oedogonium subplagiostomumMO97%[124]
Cyanobacterium phormidiumIndigo 91%[125]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Patel, H.; Yadav, V.K.; Yadav, K.K.; Choudhary, N.; Kalasariya, H.; Alam, M.M.; Gacem, A.; Amanullah, M.; Ibrahium, H.A.; Park, J.-W.; et al. A Recent and Systemic Approach Towards Microbial Biodegradation of Dyes from Textile Industries. Water 2022, 14, 3163. https://doi.org/10.3390/w14193163

AMA Style

Patel H, Yadav VK, Yadav KK, Choudhary N, Kalasariya H, Alam MM, Gacem A, Amanullah M, Ibrahium HA, Park J-W, et al. A Recent and Systemic Approach Towards Microbial Biodegradation of Dyes from Textile Industries. Water. 2022; 14(19):3163. https://doi.org/10.3390/w14193163

Chicago/Turabian Style

Patel, Heli, Virendra Kumar Yadav, Krishna Kumar Yadav, Nisha Choudhary, Haresh Kalasariya, M. Mujahid Alam, Amel Gacem, Mohammed Amanullah, Hala A. Ibrahium, Jae-Woo Park, and et al. 2022. "A Recent and Systemic Approach Towards Microbial Biodegradation of Dyes from Textile Industries" Water 14, no. 19: 3163. https://doi.org/10.3390/w14193163

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

Article Metrics

Back to TopTop