Next Article in Journal
Fixed-Bed Studies of Landfill Leachate Treatment Using Chitosan-Coated Carbon Composite
Next Article in Special Issue
Background Content of Polycyclic Aromatic Hydrocarbons during Monitoring of Natural and Anthropogenically Transformed Landscapes in the Coastal Area Soils
Previous Article in Journal
Influence of Terrain Factors on Urban Pluvial Flooding Characteristics: A Case Study of a Small Watershed in Guangzhou, China
Previous Article in Special Issue
Assessing the Combined Pollution and Risks of Potentially Toxic Elements and PAHs in the Urban Soils of the Oldest City in Western Siberia: A Case Study of Tyumen, Russia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fungal-Based Remediation in the Treatment of Anthropogenic Activities and Pharmaceutical-Pollutant-Contaminated Wastewater

1
Amity Institute of Biotechnology, Amity University Jharkhand, Ranchi 834001, Jharkhand, India
2
School of Applied and Life Sciences, Uttaranchal University, Dehradun 248007, Uttarakhand, India
3
Guru Nanak College of Pharmaceutical Sciences, Jhajra, Cjakrata Road, Dehradun 248007, Uttarakhand, India
4
Department of Biotechnology, School of Biosciences & Technology, Vellore Institute of Technology (VIT), Tiruvalam Road, Vellore 632014, Tamil Nadu, India
5
Amity School of Engineering & Technology, Amity University Jharkhand, Ranchi 834001, Jharkhand, India
6
Department of Biotechnology, Heritage Institute of Technology, Kolkata 700107, West Bengal, India
7
Department of Geology, St. Columba’s College, Hazaribag 825302, Jharkhand, India
8
Academy of Biology and Biotechnology, Southern Federal University, 344090 Rostov-on-Don, Russia
9
Department of Biomedical Sciences, College of Health Sciences, Abu Dhabi University, Abu Dhabi P.O. Box 59911, United Arab Emirates
10
Department of Medical Biotechnology, College of Applied Medical Sciences, Qassim University, Buraydah 52571, Saudi Arabia
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Water 2023, 15(12), 2262; https://doi.org/10.3390/w15122262
Submission received: 6 April 2023 / Revised: 21 May 2023 / Accepted: 8 June 2023 / Published: 16 June 2023

Abstract

:
Pharmaceutical personal care products (PPCPs) have increased in consumption due to the worldwide post-pandemic situation, marking them as chemical and pathogenic pollutants in significantly higher concentrations than ever in the ecosystem. Considering the inexplicable levels of these chemical residues discharged into the environment, concerns have been raised regarding their probable ecotoxicity to marine and terrestrial life. A further concern is the potential for developing and spreading antibiotic-resistant microorganisms and genes in aquatic ecosystems due to antibiotic exposure. Hence, knowing how these compounds impact aquatic ecosystem functioning is imperative, and thus is a critical area of research. The ecological risk analysis of PPCPs in aquatic ecosystems has been carried out using various strategies. Previous studies have reported numerous approaches for eliminating these PPCPs, including conventional treatment methods, activated sludge processes, generated wetlands, biological remediation, sequencing batch reactors, phytoremediation, and membrane bioreactors. In terms of green biotechnology approaches, the current research aims to discover effective procedures for removing PPCPs and their emerging resources as pollutants. Therefore, this review focuses on the over-extensive utilization of PPCPs and their emergent sources responsible for the contamination and environmental threat for future wastewater purposes. Further, as fungi and their enzymes and derivatives can remove pharmaceuticals and personal care products from wastewater through oxidation and several processes, they have attracted the attention of the scientific community due to their ability to remove PPCPs as pollutants and their status as emerging resources in wastewater. This review examines the fundamental approach and progress of the bioremediation of pharmaceutical- and personal-care-contaminated wastewater using fungal-based systems. It also discusses mechanistic approaches through hybridizing cultures and other biological systems with fungal strains, current technologies, and prospects for future research on PPCPs in wastewater treatment.

Graphical Abstract

1. Introduction

Water resources are becoming scarcer, and the quality of water bodies is seriously challenged by several contaminants that influence both human health and aquatic environments due to their long-lasting nature, resistance to degradation, and toxicity [1,2]. Emerging pollutants are frequently detected in water environments, including fire retardants, pharmaceuticals, and skin care products [3]. Pharmaceuticals and personal care products (PPCPs) are one major category of the recently emerging organic micropollutants causing much concern [4].
The US Environmental Protection Agency defines pharmaceuticals and personal care products (PPCPs) as “any product used by persons for personal health or cosmetic purposes or used by agribusiness to increase growth or health of livestock” (USEPA). This description covers the wide range of chemicals used in cosmetics, fragrances, over-the-counter drugs, and veterinary treatments (Figure 1).
Since they are designed to have the most substantial effect at low concentrations, PPCPs significantly impact the environment and people, even in trace quantities [5]. As a result, concerns have been raised concerning the potential damage PPCPs could bring to people and ecosystems, notably by promoting the spread of antibiotic-resistant genes. In addition to improving daily life, PPCPs are used to prevent and treat illnesses in animals and humans. Under typical circumstances, PPCPs might not evaporate quickly, even though they might easily dissolve in water. Because of these characteristics, PPCPs can access water sources in various ways [6]. PPCPs are frequently discovered in concentrations between ng/L and g/L in water and wastewater samples. In aquatic environments, PPCPs have detrimental ecological effects that affect the ecosystem and human health [7].
Different methods, including photo-Fenton, electrochemical treatment and activated carbon methods and emerging technologies such as graphene-based nano-systems, ZnO nanostructures, and zirconia nanostructures, have been utilized as effective adsorbents for the elimination of pharmaceutical compounds [8,9,10]. However, despite using these physicochemical technologies for pharmaceutical waste treatment, they have certain drawbacks, such as producing by-products, high operational costs, and labor-intensive processes [11]. Further assistance is required to address these limitations. Moreover, most wastewater treatment facilities need help to eradicate the new micropollutants. Various physicochemical and biological approaches to micropollutant (PPCP) treatment have been investigated, including the advanced oxidation process and fungi-induced activated sludge methods. PPCPs are pollutants that are generated in the form of industrial waste.
Microbes such as fungi, which promote the breakdown of organic substances at this stage, are crucial [12]. Indeed, high biodegradation rates of PPCPs have been observed in activated sludge chambers.
Fungi are microorganisms that break down the bulk of organic chemicals in the environment. Fungi have been utilized in water and soil bioremediation techniques since the 1980s [13]. In recent years, using white-rot fungi for the biological treatment of wastewater containing pharmaceuticals and personal care products (PPCPs) has emerged as a promising approach [14]. Various studies conducted have demonstrated through batch-scale experiments that white-rot fungi are capable of secreting abundant enzymes and degrading a wide range of environmental pollutants, including pharmaceuticals, dyes, heavy metals, and organic contaminants which disrupt the endocrine system [15,16,17]. Moreover, research has indicated that fungi can also employ biosorption to remove pharmaceutical and personal care substances [18]. Several reports have shown the efficient removal of pharmaceutical compounds via filamentous fungi, achieving removal rates ranging from 78% to 100% in synthetic wastewater media [19,20,21,22]. However, previous studies have often exhibited slow degradation rates, with treatment durations exceeding 26 days [14,23,24]. Nonetheless, recent investigations have involved the white-rot fungus Trametes versicolor [25] and species of Aspergillus, Mucor, Penicillium, Rhizopus, Trametes, and Trichoderma [26,27]. They have demonstrated the significant degradation of pharmaceutical compounds with high removal rates achieved within a shorter timeframe, ranging from 4 h to 10 days [28,29,30,31]. Fungi provide several advantages over traditional physicochemical treatment procedures, including being highly effective, having a cheap cost, and being eco-friendly [32]. Furthermore, they can act in various pH circumstances and have a variety of enzymatic abilities for degradation (Figure 2). Additionally, the hyphal forms of fungus can aid in the biosorption process of water purification by allowing PPCPs and other chemicals to cling to their surface or be ingested inside the cell, where they are stored rather than transferred in the water [33]. The previous cases have reported the presence of PPCPs from the industrial effluents and the coastal water followed by their treatment through fungal strains [34].
Fungi have demonstrated remarkable capability in degrading diverse pharmaceutical and personal care compounds, making them highly valuable for the bioremediation of emerging contaminants in wastewater. This review will provide a comprehensive overview of the recent research, focusing specifically on using fungi in the bioremediation of PPCPs. The review examines cellular mechanisms, treatment processes, an overview of the fungi employed, their specific applications for different pollutants, and the beneficial byproducts generated through fungal activity. Furthermore, the review will address the challenges encountered in this field, highlight prospects, and explore the potential of innovative technologies to harness and enhance fungi’s remediation abilities in treating PPCPs.
The details which contributed to the studies of PPCPs as significant contributors to contamination and their removal through physical, chemical, biological, and integrated green approaches, explicitly focusing on fungal strains, hybrid fungal and other algal or bacterial strains, are mentioned in Table 1. Further, the new techniques through AOR approaches, including monitoring and devices, are also categorized in Table 1.

2. PPCP as Pollutants

Pollutants from pharmaceuticals and personal care products (PPCPs) have emerged due to their nature of causing diverse physiological impacts in humans, even at nominal doses. Numerous studies conducted in recent years have confirmed the presence of different PPCPs in various environmental niches, primarily due to the indiscriminate use of antimicrobials and other organic compounds to treat bacterial and fungal infections, cardiac and other major surgeries in the treatment of cancer, and patients undergoing organ transplant. In addition, the use of similar compounds in personal care products raises concerns about their potential adverse effects on humans and wildlife [98,99,100]. These products constitute a large and diverse group of organic compounds, including prescription and non-prescription medications, nutritional supplements, diagnostic agents, and daily personal care products (PCPs) such as soaps, lotions, toothpaste, fragrances, and cosmetics, together with their metabolites and transformation products [101]. In the USA, approximately two out of three prescription medications were reported to be either redundant or unused. It was also found that the total estimated cost of these wasted prescriptions ranged anywhere from USD 59,264.20 to USD 152,014.89. All healthcare centers are responsible for 3% of the world’s pharmaceutical waste contribution, followed by tablets and injectables [100]. Microbeads in a diversity of PPCP cosmetics, including shower gels, toothpaste, nail polishes, and other daily care products, account for the environmental plastic waste grouped as microplastics or nano plastics. Most of these microplastics cannot be treated in wastewater treatment plants (WWTPs). Therefore, they are discarded directly into the environment, causing health hazards; it is estimated that over 3 trillion microbeads are released via WWTPs. These plastic microbeads attract organic pollutants and chemicals linked to chemical hazards, including cancer [89]. At this hour, reliable, relevant, and valid scientific data are required to guide regulatory and policy responses to address possible human, environmental, and ecology hazards from PPCP pollutants. The government, academics, cosmetics manufacturers, pharmaceutical corporations, and consulting firms engaged in scientific research or policy and management activities are the main stakeholders in the management of these pollutants [1,102].

2.1. Sources and Routes

Sewage treatment facilities have mostly been found as reservoirs of trace antibiotics and PPCPs (upto µg/L). Most often, concentrations in treatment plants in China have ranged between not-detected (ND) and 7910 ng/L (influent) and between not-detected and 9460 ng/L (effluent) (54). Doxycycline, fluoroquinolones, and penicillin comprise most antimicrobial agents used among chemotherapeutics needed to prevent deadly illnesses in both people and animals [103]. Pharmaceutical industries are compelled to scale up the production of these antimicrobials to keep up with the rising demand for these medications. Although some drugs are used internally, they remain unchanged in urine and feces and are then deposited in sewage treatment facilities. Several antibiotics and their metabolic products [101], hormones [38], and other pharmaceuticals were frequently detected in sewage samples [2]. Such pharmaceutically active molecules include fluoroquinolones, macrolides, azoles, cephalexin, erythromycin, and its metabolic products (e.g., Roxithromycin). These sewage treatment plants also contain acetaminophen; caffeine; hormones such as estrogen and progesterone; glucocorticoids; and others. The two primary mechanisms influencing the PPCPs’ fates in the environment are adsorption and biodegradation. Penicillin, amoxicillin, tetracycline, quinolones, novobiocin, and gentamicin are the antibiotics most utilized when treating animals. The most common illnesses treated in animals used for meat production include diarrhea, shipping fever, and bovine pneumonia [39]. These compounds have been discovered in manure and soil that is heavily contaminated with manure. Antimicrobial agents and PCPs have been utilized in animal husbandry, aquaculture, and agriculture as growth promoters and to boost feed conversion efficiency. These practices include animal farming [36,49]. PPCPs are ubiquitously found in several environmental niches, such as in excretion products (urine and feces), and hence are most typically detected in WWTPs. They subsequently enter aquatic system components in medicines and household items which include estrogenic endocrine-disrupting chemicals (e-EDCs), toxic metals, and nanoparticles, and their subsequent discharge in rivers represents a significant risk to human and livestock health [91]. The distribution and hazardous outcomes of PPCPs are discussed in Table 2.

2.2. Mechanism of Action and Hazards Caused to Human Health

Pharmaceutical substances employ toxic or static action toward the microbial cell. This cytotoxic or static mechanism enables the immune system to eradicate pathogens. The inhibition of cell wall production, the synthesis of DNA/RNA, the synthesis of protein, ribonucleic acid (RNA) synthesis, protein synthesis, and membrane disorganization and the disruption of membranes are all mechanisms used by antibiotics to kill or inhibit the growth of bacteria or fungi [98]. Since antimicrobials are efficient against a wide variety of microbial strains, they are used worldwide to treat various infections. This widespread use has led to ubiquitous distribution in the different environmental milieus.
It is now established that PPCPs pose a significant threat to human health and wildlife and should be immediately addressed. Chronic exposure to these chemicals increases the risk of developing testicular germ line and breast cancer, even at low concentrations. Prolonged exposure to these compounds also causes hypospadias, cryptorchidism, and infertility due to decreased sperm count [98,115]. Experts believe that investments should be made in technologies aiding advanced wastewater treatment and educating medical professionals to reduce the overprescription of antibiotics. These steps can reduce the environmental burden of PPCPs. This requires support from pharmaceutical companies and governments to incorporate pharmaceutical-return programs in corroboration with education and asking all municipalities to adhere to minimum secondary wastewater treatment [115]. The administration should implement these strategies to reduce the burden of human-derived PPCPs on the environment.

3. Fungal Genera and Enzymes for the Remediation of PPCP

Bioremediation minimizes the pollutant level by converting recalcitrant and xenobiotic components to a less toxic or acceptable form [58]. Due to their great metabolic activity, microorganisms are omnipresent and can survive various environmental conditions. Their universality, diversity, colossal population and biomass, anaerobic activity, survival capacity in extreme conditions, and stupefied catalytic mechanisms make them suitable for bioremediation [55]. They have compelled scientists and researchers all over the globe to dive deep into understanding their biochemistry and genetics for their utilization in various applications. Bioremediation, accomplished by biodegradation, involves biotic and abiotic transformations which microorganisms achieve. It is a complex activity involving the disruption of the different organic compounds into their inorganic components. The physical and chemical interaction of microorganisms with various polluting factors leads to the complete or partial degradation of the molecule of interest. Pharmaceuticals and personal care products (PPCPs) mostly contain endocrine-disrupting chemicals and trace organic compounds that cause bioaccumulation, are highly toxic to aquatic life, and adversely affect human health. Although traditional wastewater treatment processes such as activated sludge and membrane bioreactors are adequate for removing most organic compounds, nutrients, and pathogens, certain hydrophilic compounds and compounds containing functional groups with strong electron-withdrawal capacities are barely eliminated. Thus, PPCPs demand specialized treatment processes other than conventional techniques. The details of the fungal enzymatic bioremediation of PPCPs and their resources through different fungal stains in different bioreactors are mentioned in Table 3 and Table 4.
Fungi have been utilized in bioremediation for more than three decades. The rich morphology; diverse metabolic capacity; and ability to thrive, survive, colonize, and propagate in diverse habitats make fungi one of the most suitable candidates for detoxifying PPCPs. Additionally, fungi are comparatively more effective for bioremediation than bacteria because of their various processes and enzymatic activities, as well as their efficiency in functioning under a wide range of pH [32,51,72,76]. Several factors, such as the composition of media, pH, spore suspension concentration, and incubation factors, including duration, mixing speed, etc., are the key factors influencing mycelium growth and enzyme production. Other crucial factors include various operational parameters such as cultivation time, the concentration of organic and inorganic constituents, stationary or submerged cultures, aeration, the concentration of inducers, and protease activation [76].
White-rot fungi (WRF), belonging to the phylum Basidiomycota, refer to a collection of an eco-physiological group of fungal species such as Pleurotus ostreatus, Phanerochaete chrysosporium, Trametes versicolor, Ganoderma lucidum, and Irpex lacteous that possess a lignin modification enzyme (LME) and are capable of lignin degradation. The fungi are named so because of their lignin degradation capacity. Woods attacked by them acquire a bleached appearance with a fibrous texture since dark-colored lignin gets eliminated in its presence. The non-specific WRF enzymes, which are non-stereoselective and extensively diverse, facilitate their utilization for the bioremediation of PPCPs [53]. WRF degrades lignin along with its aromatic compounds. The lignin-modifying system (LMS) possessed by WRF refers to the assembly of oxidoreductive metalloenzymes, heme peroxidases, laccases, secondary metabolites, and surfactants. Lignin peroxidases (LiPs), manganese-dependent peroxidases (MnPs), and laccases are the three major groups of enzymes secreted by WRF [81]. However, all three classes of enzymes may not be secreted by the same fungal strain. After the degradation or fragmentation of lignin, the surfeit of the fragments thus formed is taken up by the hyphae and passed on to the intercellular network for further metabolism [59].
Fungi, more specifically WRF, may utilize four distinct pathways or mechanisms for the bioremediation of PPCPs that can be efficient both independently and synergistically. The four pathways are (1) utilizing their hyphae for the bioabsorption and immobilization of various PPCPs, (2) involving the extracellular enzymes, peroxidases, and phenolases, (3) involving the intracellular enzyme complexes such as cytochrome P450, and (4) producing reactive oxygen species (ROS). The enzymes secreted by WRF, being extensively oxidative, do not require the compound’s internalization and subsequent toxicity, but instead can potentially act intracellularly or extracellularly [57].
The vegetative fungal body comprises masses of long or thread-like structures known as hyphae, a defining property of filamentous fungi such as WRF that aids in rapid colonization. This structure is involved in various functions, a crucial one among them being absorption. Various chemicals, including PPCPs, adhere to the surfaces of hyphae of WRF or might be internalized within the cell, thereby getting absorbed by the fungi. This absorption mechanism stops the PPCPs from flowing along with the effluent, proving the efficiency of WRF in bioabsorption and immobilization [51].
Lignin peroxidase is an enzyme belonging to the oxidoreductase family. It can be broadly categorized as a ligase enzyme capable of degrading lignin, a highly resistant biopolymer. Manganese-dependent peroxidase is another ligninolytic enzyme that aids in lysing lignin. Laccases constitute a family of copper-containing oxidase enzymes; the copper ions present in the enzyme’s active site undergo a cycle of oxidation and reduction reactions during the catalytic process in which two water molecules are formed on the concomitant electron loss of a single oxygen molecule. This abstracted electron further leads to the oxidation of various benzene-ring-containing compounds capable of degrading various aromatic compounds, including phenols or amines [60,65]. Trametes versicolor, Phanerochaete chrysosporium, and Phlebia tremellosa are some of the most prominent WRFs in bioremediation. T. versicolor is a versatile species that secretes all three extracellular ligninolytic enzymes; however, laccase is the predominant enzyme secreted by the strain ATCC 7731. It also secretes the cytochrome P450 complex. This species has relatively more significant redox potential and is stable in crude and purified forms. However, its bioremediation efficiencies in terms of PPCPs are majorly dependent on factors such as cultivation conditions, substrates, and the chemical components present in the PPCP effluent [52]. PPCPs, including steroids, nonylphenol, and octocrylene, which can lead to the dysfunctioning of the endocrine system, have been reported to be removed at a rate of 70–99% by T. versicolor. However, pharmaceutical compounds containing strong electron-withdrawing functional groups such as ciprofloxacin, salicylic acid, azithromycin, tetracycline, and carbamazepine have been reported to be removed at the rate of 35%, 0–5%, 26%, 0–5%, and 0–90%, respectively. Additionally, this fungal strain has been reported to be effective for removing compounds such as triclosan and oxybenzone, apart from octocrylene and nonylphenol found in personal care products. Some of the common drugs removed by T. versicolour are naxopren, carbamazepine, and ibuprofen. Among other WRF species, Phanerochaete chrysosporium, Bjerkandera adusta, Dichomitus squalene, and Pleurotus ostreatus have been investigated to be effective in the bioremediation of PPCPs such as ibuprofen, diclofenac, and triclosan owing to their LMEs [92]. Genetic engineering techniques targeting microbes can also be used in the treatment of wastewater [48]. P. chrysosporium is capable of degrading chlorinated xenobiotics, whereas B. adusta spp. TBB-03 effectively removes acetaminophen and bisphenol under varied conditions of temperature and pH [54]. Moreover, both fungal strains successfully degrade antidepressants such as citalopram and fluoxetine. Besides LMEs, WRF secrete natural mediators that enable crude enzymes to achieve better results in removing PPCPs than purified enzymes. Crude extracellular enzymes potentially degrade PPCPs, including oxytetracycline and tetracycline, at 84% and 72%, respectively, in batch and continuous cultures. Additionally, investigations on cell cultures of WRF have shown that submerged whole-cell cultures are more promising than solid media cultures.
In addition to the extracellular enzymatic machinery of WRF, several intracellular enzymes have also been investigated to be efficient in detoxifying xenobiotic compounds. The cytochrome P450 monooxygenase superfamily is one such complex with heme as a cofactor, capable of degrading PPCPs, including chlorinated hydrocarbons and polycyclic aromatic hydrocarbons. This enzyme complex aids in catalyzing deamination, heteroatom oxygenation, dehalogenation, dealkalinalization, and hydroxylation reactions, ultimately leading to the mineralization and degradation of organic pollutants [15]. Organic pollutants can be degraded through metal–carbon hybrid material-mediated PS-AOPs for wastewater decontamination [8]. In order to overcome these challenges, alternative approaches with excellent performance have been looked for [6,8].
The LME machinery of WRF makes these fungal species potential producers of superoxide anion radicals (O2•−), hydrogen peroxide (H2O2), and hydroxyl radicals (HO), which are reactive oxygen species (ROSs). ROSs are chemically reactive molecules with oxygen. Superoxide radicals are formed when an oxygen molecule gets reduced by gaining an electron. The volatile oxygen molecule in this form tries to get rid of the electron, which makes that molecule reactive when taken up by another molecule [51]. The dismutation of superoxide, thus formed, produces hydrogen peroxide, which in turn may get partially reduced to a hydroxyl radical or wholly reduced to water following the Fenton reaction. These ROSs oxidize various compounds to a high degree. The laccase oxidation and dismutation activity by Pleurotus eryngii on hydroquinone and other aromatic aldehydes enhances the ROS production process.
Hence, WRF can be considered promising and efficient weapons in combating PPCPs, which are extensively harmful contaminants. The efficiency of various fungal strains in detoxifying chemical and biochemical compounds makes this ‘fungal rescue’ path effective.

4. Mechanistic Approaches Using Fungal Strains to Remove PPCPs

Mycodegradation refers to the decomposition or degradation of various chemical compounds and other materials via fungi. It is a fact that fungi are capable of efficient degradation, since it has been witnessed in many circumstances that fungi destroy several kinds of materials such as leather, paper, wood, textiles, and plastic. Pleurotus ostreatus and Irpex lacteus successfully break down PAH present in polluted industrial soil. The cultures of white-rot fungus are very capable of not only decomposing PAH but also eliminating different PAHs. Several different kinds of enzymes such as laccases, manganese peroxidase (MnP), and lignin peroxidase (LiP) are produced by fungi, and these can very effectively be employed in the decomposition of hydrocarbons, pesticides, polychlorinated biphenyls, dyes, chlorinated and phenolic compounds, etc. [95].
An array of enzymes can be synthesized by fungi, such as laccase, esterase, manganese peroxidase, lipase, and unspecific peroxygenases (UPOs). These enzymes can remediate organic pollutants such as pharmaceuticals, pesticides, microplastics, recalcitrant PAHs, and personal care products. In addition to enzyme-mediated methods, a range of other approaches can be used with fungi, such as biotransformation, bioabsorption, and biomineralization (bio-precipitation), which helps them counteract recalcitrant toxic compounds [80]. As an analytical study exhibited that there was a decline in the activity of catalase as an increment in the concentration of oil occurred when the bioremediation of contaminated soil was underway, this indicated that catalase activity had the potential to be employed as a tool to monitor the efficacy of bioremediation. Hence, if the essentiality of catalase in making fungi tolerant to heavy metals is considered, fungi are a viable option for the remediation of areas contaminated with metals [70].
For the purification of contaminated water effluents polluted with phenolic elements and their derivatives, the decolorization of dyes used in the textile industry, the breakdown of polyaromatic hydrocarbons (PAHs) and pesticides, and for the effective removal of H2O2 residues from bleaching effluents, the enzyme catalase is an excellent option. Catalase also keeps track of the physiological response of organisms, especially when environmental pollutants are present [73].
Fungal laccases provide many benefits of immense value for biotechnological uses of commercial effluent treatment. Since these enzymes show broad specificity concerning the substrate, they are well suited for bleaching Kraft pulp or rendering agricultural by-products, including wastes from olive mills or coffee pulp. Effective decolorization was observed when the effluent from a Kraft paper mill bleach plant was treated with laccase from a fungal isolate of Flavodon flavus. Bisphenol A, an endocrine-disrupting chemical, was successfully degraded by laccase purified from a white-rot basidiomycete named Trametes villosa. It has also been documented that laccases isolated from aquatic hyphomycetes called Clavariopsis aquatica break down xenoestrogen nonylphenol. Apart from its possible use in organic degeneration in freshwater ecosystems, the enzyme laccase also presents novel ideas for biotechnological utilities, such as treating contaminated water [62]. Initially, the laccases derived from fungal species were classified as enzymes that degrade lignin (lignin-degrading). Later, it was discovered that they aid the coupling of the reduction of O2 with the oxidation of an array of substrates of organic nature, which include phenols, aromatic amines, aliphatic amines, polyphenols, and phenols that have -OCH3 substitution. Earlier works have indicated the role of fungal laccases in the decomposition of various upcoming pollutants, such as bisphenol A (BPA), sulfamethoxazole, atrazine, estrogens (e.g., 17α-ethinylestradiol (EE2), estrogen (E1), and 17β-estradiol (E2)), and other recalcitrant organic pollutants [74].
Depending on the origin and functional nature, peroxidases are classified as manganese peroxidase (MnP), lignin peroxidase (LiP), and versatile peroxidase (VP). Of these, lignin and manganese peroxidases are heme peroxidases and need hydrogen peroxidases and manganese’s presence for functionality. These peroxidases are mainly known for the breakdown of toxins via white-rot fungi and fungi belonging to Basidiomycetes. When it comes to versatile peroxidases, they have the potential to oxidize both phenolic as well as non-phenolic compounds and are essential for bioremediation purposes. Notably, versatile peroxidase (VP) has broad substrate-specificity [75].
Specific peroxidases cannot be adjusted into the abovementioned classification; they are dye-decolorizing peroxidases (DyPs) and unspecific peroxygenases (UPOs). They are heme peroxidases and utilize hydrogen peroxide to catalyze the oxidation of several non-phenolic lignin model compounds and other organic substances [63].
Lately, to improve the stability under acidic conditions (even at a pH as low as 2), a manganese peroxidase enzyme derived from C.subvermispora was engineered. The improved Mn2+-oxidizing nature and enhanced acid stability were incorporated by analyzing its crystalline structure as a scaffold. A stable enzyme was synthesized, which could oxidize veratryl alcohol and Reactive Black 5 [63].
In order to render the toxic elements in the environment non-toxic, fungi have sophisticated oxidative and hydrolytic enzymatic machinery [66,71]. Apart from this machinery, some fungi also have intracellular networks which comprise the genome, constituted by cytochrome (CYP) P450 monooxygenases and glutathione transferases for combating an array of contaminants.
The genome comprises the members of detoxification pathways, usually from multigenic families such as cytochrome P450 monooxygenases and glutathione transferases [56].
The fungal cytochrome P450 system can act as a versatile catalyst for the region- and stereospecific oxidation of non-activated hydrocarbons. It also can serve as an ideal substitute for chemical catalysts [78]. Mutagenic/carcinogenic fused-ring high-molecular-weight PAHs (HMW-PAHs), crude oil aliphatic hydrocarbon n-alkanes, and endocrine-disrupting long-chain alkylphenols (APs) were oxidized via CYP63A2 P450monooxygenase isolated from the white-rot fungus, i.e., P. chrysosporium [79]; the mechanism is summarized in Figure 3. The different mechanisms for treating pharmaceutics through fungal degradation are shown in Figure 4.
Fungal biodegradation can be carried out against diverse types of micropollutants due to the ability of fungi to produce a wide range of enzymes that can break down different types of pollutants. Different types of fungi are used depending on the specific micropollutant being targeted, and factors such as pH, temperature, and nutrient availability can also affect the efficiency of fungal biodegradation [64]. The mechanisms of fungal biodegradation are diverse and can be applied to various micropollutants, including pharmaceuticals, pesticides, and industrial chemicals. For example, studies have shown that certain species of fungi, such as Phanerochaete chrysosporium and Trametes versicolor, can degrade a range of micropollutants, including pharmaceuticals, pesticides, and personal care products. The key to successful biodegradation lies in optimizing fungal growth and activity conditions, such as pH, temperature, and nutrient availability. One example of the successful fungal biodegradation of micropollutants is using white-rot fungi to degrade pharmaceuticals in wastewater. In this case, researchers isolated various white-rot fungi and conducted culture experiments to determine which strains were most effective at breaking down specific pharmaceuticals. This method allowed for the targeted and efficient removal of micropollutants in the wastewater [64,67].

5. Current Methods of Fungal Biodegradation: Of PPCPs

Nowadays, several methods of fungal modification have been opted for and invented by researchers [69,97]. The genetic modification of fungi for enhanced biodegradation involves identifying and manipulating genes responsible for the degradation pathways of specific micropollutants. This can be achieved through gene knockout, gene overexpression, and gene editing using CRISPR/Cas9. However, there are still challenges in achieving the efficient and stable expression of modified genes in fungi and ensuring the safety and environmental impact of genetically modified organisms. Therefore, research in this area is ongoing, with the intention to develop more effective and sustainable methods of fungal biodegradation. Using fungal consortia for biodegradation is another promising approach to micropollutant removal. This involves combining different types of fungi with complementary degradation pathways, allowing for a broader range of pollutants to be targeted. Studies have shown that using fungal consortia can enhance the efficiency and stability of biodegradation compared to using single strains alone. Fungus-based biofilters are becoming increasingly popular in wastewater treatment plants as a low-cost and sustainable micropollutant method [69,97].
Genetic modification, nanobiotechnological approaches, and fungal consortia are promising approaches to improving the efficiency and sustainability of fungal biodegradation [41,45]. However, challenges remain in achieving the stable expression of modified genes and ensuring the safety and environmental impact of genetically modified organisms. Additionally, fungus-based biofilters are becoming increasingly popular in wastewater treatment plants as a low-cost and sustainable method for micropollutant removal. Nonetheless, the issue of how genetically modified fungi can become dominant strains in environments remains a challenge. For example, a study showed that a fungal consortium of five strains could efficiently degrade a mixture of pesticides and herbicides in contaminated soil [47]. Each strain had a specific degradation pathway, allowing for the breakdown of different pollutants. Using the fungal consortium resulted in higher degradation rates and increased stability than using each strain alone.
Overall, various novel methods are available for treating pharmaceutical waste, including chemical, physical, and biological approaches [96]. Recent studies have shown the effectiveness of using fungal consortia for degrading pesticide and herbicide contaminants in soil, which could have potential applications in pharmaceutical waste treatment [47]. Chemical methods involve oxidizing or reducing agents to break down the waste. Physical methods (Sinha, 2018) include filtration, evaporation, and distillation. Biological methods use microorganisms to degrade the waste [42,43,44,47]. However, these methods have limitations such as high cost and low efficiency. Recently, new devices have been developed for pharmaceutical waste treatment using advanced oxidation processes (AOPs) and nanotechnology. AOPs use reactive species, such as hydroxyl radicals, to break down the waste [12]. Nanotechnology involves the use of nanoparticles to enhance the efficiency of AOPs. Commercial devices such as cavitation and ultrasound devices were utilized and developed in various cases of pharmaceutical wastewater treatment [90]. These new devices have shown promising results with high degradation rates and low costs. However, further research is needed to optimize these devices for large-scale applications and ensure their safety for human health and the environment.

6. Environmental Regulations and Initiatives in the Removal of PPCP

The unintended accumulation of PPCPs in different aquatic and environmental niches (such as freshwater, sediments, and lifeforms) at amounts capable of deteriorating humanity and the ecosystem has come to our attention in recent years. PPCPs are often found in freshwater environments at relatively low concentrations. However, many of them and their intermediates are physiologically active and can potentially affect aquatic animals that are not their intended biological targets [116]. Due to the actual usage of PPCPs, rising applicability for human and veterinary medicine, and the ongoing leakage of these substances into the environment, this has become a severe issue. The European Union (EU) and the United States Environmental Protection Agency (USEPA) have created priority pollutant categories that highlight a wide range of compounds found in wastewater and the environment that may be detrimental. The Environment Agency (EA) of England and Wales developed a classification scheme for these compounds’ estimated hazard ratios to identify compounds with a significantly high potential to orchestrate concern for the aquatic environment and other biological lifeforms. This classification scheme combines conventional risk assessment techniques with propensity, biomagnification, and toxicological effects (PBT) standards. According to this process, the top 10 compounds were thioridazine, clotrimazole, lofepramine, procyclidine, dextropropoxyphene, paracetamol, tramadol, mebeverine, tamoxifen, and aminophylline [117]. Pharmaceuticals and personal care products were also detected wastewater from municipal and the marine water bodies [35].
In a related expedition, an alternative selection of priority intoxicants was recognized utilizing the Oil Spill Prevention, Administration, and Response (OSPAR) sampling and categorization method for hazardous material, which included lofepramine. Perfluoroalkyl substances (PFASs), polybrominated diphenyl ethers (PBDEs), and PCBs are examples of environmental contaminants with toxicities. These contaminants were primarily designed to optimize their physiological activity at low concentrations and to specifically target biochemical, enzymatic, or cell-signaling mechanisms [94].
Nevertheless, the cytotoxicity of PPCPs circulating in the environment has extended beyond the immediate effects fostered by therapeutic agents. Recent scientific expeditions and studies around the globe have shown that the orchestration of PPCP toxicity may vary depending upon the organism being exposed, the time duration of exposure, the concentration metric of the contaminant, and the route through which exposure occurs [77]. Moreover, the detrimental effects of acute exposure traces, especially at specific concentrations, are explanative enough for the physiological abnormalities observed in non-specific organisms than those recorded for an acutely higher concentration of dose exposure. The efficacious management and overall permissible circulation of PPCPs seem implausible without a global consensus furbishing innovative ideas to create and install a well-structured, sustainable Environmental Risk Assessment (ERA) protocol for these lethal contaminants. Pre-existing procedures and predicaments must be upgraded and evolved to acclimatize to the potential consequences of these toxic compounds in the environment [93].

7. Challenges and Future Directions

Fungi could now be considered a valuable tool for the remediation of contaminants from pharmaceutical and personal care origin. They make ideal biological agents to incorporate in wastewater treatment procedures due to their versatility in degrading a wide range of resistant compounds and ease of handling. However, fungus performance has usually been studied maximally in sterile batches. Since bacterial contamination can significantly impair fungal function, few experiments have been carried out in continuous-flow reactors under non-sterile circumstances. More work is therefore needed to address this limitation. The added limitation of the fungal strain treatment process is washing out the mediators and the enzymes with treated effluent. This presents a particular challenge for enzymatic reactors that use enzymes rather than fungi.
Moreover, the remediation pathways of a few PPCPs, such as ibuprofen, carbamazepine, naproxen, and diclofenac, have been described in the literature. The metabolites/intermediate products are more hazardous than the parent chemical. Therefore, efforts should be made to characterize and illuminate the intermediate product formation during the treatment of PPCPs via fungus strains, which would further help to ascertain the degradation pathways/mechanisms of PPCPs. Many enzymatic conversion/full-conversion gaps in pathways are also unresolved, obstructing fungal use in the treatment process. Genetic manipulation/engineering could help offset the mismatch, representing a vital step towards improving bioremediation via fungal systems. In this regard, it is essential to develop efficient strategies for genetic manipulation that would be supported for all fungi.
Moreover, to develop a potent tool usable in wastewater treatment plants to prevent PPCPs from entering ecosystems, many investigations on the physiology and functional genomics in representatives of various groups still require a complete understanding of the mechanisms of the metabolism of PPCPs to ensure the highest efficiency for pharmaceutical and personal care products’ removal. New genetic engineering expertise should be introduced to choose and amplify the best fungal strains to eliminate PPCPs. Essential metabolic genes are present in several bacteria and may be transferred to other organisms. Since these microorganisms can remove pharmaceuticals from wastewater and allow for their continuous use, considering legal restrictions and a practical market and cost rationale, genetically modified fungi will probably have a significant future in this field.

Author Contributions

Conceptualization, S.M. (Sumira Malik), J.B., A.G.A., and V.D.R.; methodology, S.M. (Sumira Malik), J.B., S.M. (Sagar Mondal), H.K., and S.R.; software, R.H.; validation, S.M. (Sumira Malik), A.G.A., V.D.R., and T.M.; formal analysis, S.M. (Sumira Malik), T.M. and R.H.; investigation, S.M. (Sumira Malik) and V.D.R.; resources, T.M., V.D.R., and A.G.A.; data curation, S.M. (Sumira Malik); writing—original draft preparation, S.M. (Sumira Malik), J.B., S.S., S.M. (Sagar Mondal), S.N., and H.K.; writing—review and editing, S.M. (Sumira Malik); visualization, S.M. (Sumira Malik); supervision, V.D.R., S.M. (Sumira Malik), N.S.S., and A.G.A.; project administration, S.M. (Sumira Malik), T.M., V.D.R., N.S.S., and A.G.A.; funding acquisition, N.S.S. and A.G.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

V.D.R. and T.M. acknowledge support by the laboratory «Soil Health» of the Southern Federal University with the financial support of the Ministry of Science and Higher Education of the Russian Federation, agreement No. 07-2022-1122.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Balakrishna, K.; Rath, A.; Praveenkumarreddy, Y.; Guruge, K.S.; Subedi, B. A review of the occurrence of pharmaceuticals and personal care products in Indian water bodies. Ecotoxicol. Environ. Saf. 2017, 137, 113–120. [Google Scholar] [CrossRef] [Green Version]
  2. Wu, C.; Huang, X.; Witter, J.D.; Spongberg, A.L.; Wang, K.; Wang, D.; Liu, J. Occurrence of pharmaceuticals and personal care products and associated environmental risks in the central and lower Yangtze River, China. Ecotoxicol. Environ. Saf. 2014, 106, 19–26. [Google Scholar] [CrossRef]
  3. Yan, H.; Lai, C.; Liu, S.; Wang, D.; Zhou, X.; Zhang, M.; Li, L.; Li, X.; Xu, F.; Nie, J. Metal-carbon hybrid materials induced persulfate activation: Application, mechanism, and tunable reaction pathways. Water Res. 2023, 234, 119808. [Google Scholar] [CrossRef]
  4. Mojiri, A.; Zhou, J.L.; Ohashi, A.; Ozaki, N.; Kindaichi, T. Comprehensive review of polycyclic aromatic hydrocarbons in water sources, their effects and treatments. Sci. Total Environ. 2019, 696, 133971. [Google Scholar] [CrossRef]
  5. Patel, M.; Kumar, R.; Kishor, K.; Mlsna, T.; Pittman, C.U., Jr.; Mohan, D. Pharmaceuticals of emerging concern in aquatic systems: Chemistry, occurrence, effects, and removal methods. Chem. Rev. 2019, 119, 3510–3673. [Google Scholar] [CrossRef] [Green Version]
  6. Fu, W.; Fu, J.; Li, X.; Li, B.; Wang, X. Occurrence and fate of PPCPs in typical drinking water treatment plants in China. Environ. Geochem. Health 2019, 41, 5–15. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, J.; Chu, L.; Wojnárovits, L.; Takács, E. Occurrence and fate of antibiotics, antibiotic resistant genes (ARGs) and antibiotic resistant bacteria (ARB) in municipal wastewater treatment plant: An overview. Sci. Total Environ. 2020, 744, 140997. [Google Scholar] [CrossRef]
  8. Lai, C.; Ma, D.; Yi, H.; Zhang, M.; Xu, F.; Huo, X.; Ye, H.; Li, L.; Yang, L.; Tang, L.; et al. Functional partition of Fe and Ti co-doped g-C3N4 for photo-Fenton degradation of oxytetracycline: Performance, mechanism, and DFT study. Sep. Purif. Technol. 2023, 306, 122546. [Google Scholar] [CrossRef]
  9. Bhardwaj, S.K.; Mujawar, M.; Mishra, Y.K.; Hickman, N.; Chavali, M.; Kaushik, A. Bio-inspired graphene-based nano-systems for biomedical applications. Nanotechnology 2021, 32, 502001. [Google Scholar] [CrossRef] [PubMed]
  10. Kaur, N.; Khunger, A.; Wallen, S.L.; Kaushik, A.; Chaudhary, G.R.; Varma, R.S. Advanced green analytical chemistry for environmental pesticide detection. Curr. Opin. Green Sustain. Chem. 2021, 30, 100488. [Google Scholar] [CrossRef]
  11. Caracciolo, A.B.; Topp, E.; Grenni, P. Pharmaceuticals in the environment: Biodegradation and effects on natural microbial communities. A review. J. Pharm. Biomed. Anal. 2015, 106, 25–36. [Google Scholar] [CrossRef] [PubMed]
  12. Raouf, M.E.A.; Maysour, N.E.; Farag, R.K.; Abdul-Raheim, A.M. Wastewater treatment methodologies, review article. Int. J. Environ. Agri. Sci. 2019, 3, 018. [Google Scholar]
  13. Rodríguez-Rodríguez, C.E.; Lucas, D.; Barón, E.; Gago-Ferrero, P.; Molins-Delgado, D.; Rodríguez-Mozaz, S.; Eljarrat, E.; Díaz-Cruz, M.S.; Barceló, D.; Caminal, G.; et al. Re-inoculation strategies enhance the degradation of emerging pollutants in fungal bioaugmentation of sewage sludge. Bioresour. Technol. 2014, 168, 180–189. [Google Scholar] [CrossRef]
  14. Mir-Tutusaus, J.A.; Baccar, R.; Caminal, G.; Sarrà, M. Can white-rot fungi be a real wastewater treatment alternative for organic micropollutants removal? A review. Water Res. 2018, 138, 137–151. [Google Scholar] [CrossRef]
  15. Asif, M.B.; Hai, F.I.; Singh, L.; Price, W.E.; Nghiem, L.D. Degradation of pharmaceuticals and personal care products by white-rot fungi—A critical review. Curr. Pollut. Rep. 2017, 3, 88–103. [Google Scholar] [CrossRef] [Green Version]
  16. Lucas, D.; Badia-Fabregat, M.; Vicent, T.; Caminal, G.; Rodríguez-Mozaz, S.; Balcázar, J.L.; Barceló, D. Fungal treatment for the removal of antibiotics and antibiotic resistance genes in veterinary hospital wastewater. Chemosphere 2016, 152, 301–308. [Google Scholar] [CrossRef]
  17. Stenholm, Å.; Hedeland, M.; Arvidsson, T.; Pettersson, C.E. Removal of diclofenac from a non-sterile aqueous system using Trametes versicolor with an emphasis on adsorption and biodegradation mechanisms. Environ. Technol. 2019, 40, 2460–2472. [Google Scholar] [CrossRef]
  18. Legorreta-Castañeda, A.J.; Lucho-Constantino, C.A.; Beltrán-Hernández, R.I.; Coronel-Olivares, C.; Vázquez-Rodríguez, G.A. Biosorption of water pollutants by fungal pellets. Water 2020, 12, 1155. [Google Scholar] [CrossRef] [Green Version]
  19. Kang, B.R.; Kim, M.S.; Lee, T.K. Unveiling of Concealed Processes for the Degradation of Pharmaceutical Compounds by Neopestalotiopsis sp. Microorganisms 2019, 7, 264. [Google Scholar] [CrossRef] [Green Version]
  20. Dalecka, B.; Juhna, T.; Rajarao, G.K. Constructive use of filamentous fungi to remove pharmaceutical substances from wastewater. J. Water Process Eng. 2020, 33, 100992. [Google Scholar] [CrossRef]
  21. Dhiman, N.; Chaudhary, S.; Singh, A.; Chauhan, A.; Kumar, R. Sustainable degradation of pharmaceutical waste using different fungal strains: Enzyme induction, kinetics and isotherm studies. Environ. Technol. Innov. 2022, 25, 102156. [Google Scholar] [CrossRef]
  22. Buratti, S.; Rinaldi, F.; Calleri, E.; Bernardi, M.; Oliva, D.; Malgaretti, M.; De Girolamo, G.; Barucco, B.; Girometta, C.E.; Savino, E. Ganoderma resinaceum and Perenniporia fraxinea: Two Promising Wood Decay Fungi for Pharmaceutical Degradation. J. Fungi 2023, 9, 555. [Google Scholar] [CrossRef]
  23. Li, X.; Xu, J.; de Toledo, R.A.; Shim, H. Enhanced removal of naproxen and carbamazepine from wastewater using a novel countercurrent seepage bioreactor immobilized with Phanerochaete chrysosporium under non-sterile conditions. Bioresour. Technol. 2015, 197, 465–474. [Google Scholar] [CrossRef]
  24. Badia-Fabregat, M.; Lucas, D.; Pereira, M.A.; Alves, M.; Pennanen, T.; Fritze, H.; Rodríguez-Mozaz, S.; Barceló, D.; Vicent, T.; Caminal, G. Continuous fungal treatment of non-sterile veterinary hospital effluent: Pharmaceuticals removal and microbial community assessment. Appl. Microbiol. Biotechnol. 2016, 100, 2401–2415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Shreve, M.J.; Brockman, A.; Hartleb, M.; Prebihalo, S.; Dorman, F.L.; Brennan, R.A. The white-rot fungus Trametes versicolor reduces the estrogenic activity of a mixture of emerging contaminants in wastewater treatment plant effluent. Int. Biodeterior. Biodegrad. 2016, 109, 132–140. [Google Scholar] [CrossRef] [Green Version]
  26. Mohapatra, D.; Rath, S.K.; Mohapatra, P.K. Soil fungi for bioremediation of pesticide toxicants: A perspective. Geomicrobiol. J. 2022, 39, 352–372. [Google Scholar] [CrossRef]
  27. Shourie, A.; Vijayalakshmi, U. Fungal diversity and its role in mycoremediation. Geomicrobiol. J. 2022, 39, 426–444. [Google Scholar] [CrossRef]
  28. Olicón-Hernández, D.R.; Camacho-Morales, R.L.; Pozo, C.; González-López, J.; Aranda, E. Evaluation of diclofenac biodegradation by the ascomycete fungus Penicillium oxalicum at flask and bench bioreactor scales. Sci. Total Environ. 2019, 662, 607–614. [Google Scholar] [CrossRef]
  29. Dalecka, B.; Oskarsson, C.; Juhna, T.; Kuttava Rajarao, G. Isolation of fungal strains from municipal wastewater for the removal of pharmaceutical substances. Water 2020, 12, 524. [Google Scholar] [CrossRef] [Green Version]
  30. Cruz del Álamo, A.; Pariente, M.I.; Sanchez-Bayo, A.; Puyol, D.; Rodríguez, R.; Morales, V.; Bautista, L.F.; Vicente, G.; Melero, J.A.; Molina, R.; et al. Assessment of Trametes versicolor, Isochrysis galbana, and Purple Phototrophic Bacteria for the Removal of Pharmaceutical Compounds in Hospital Wastewater. Adv. Environ. Eng. Res. 2021, 2, 027. [Google Scholar]
  31. Kasonga, T.K.; Kamika, I.; Ngole-Jeme, V.M. Ligninolytic enzyme activity and removal efficiency of pharmaceuticals in a water matrix by fungus Rhizopus sp. Isolated from cassava. Environ. Technol. 2022, 44, 2157–2170. [Google Scholar] [CrossRef]
  32. Tomasini, A.; León-Santiesteban, H.H. The role of the filamentous fungi in bioremediation. In Fungal Bioremediation; CRC Press: Boca Raton, FL, USA, 2019; pp. 3–21. [Google Scholar]
  33. Lu, S.; Yu, Y.; Ren, L.; Zhang, X.; Liu, G.; Yu, Y. Estimation of intake and uptake of bisphenols and triclosan from personal care products by dermal contact. Sci. Total Environ. 2018, 621, 1389–1396. [Google Scholar] [CrossRef]
  34. Ali, A.M.; Rønning, H.T.; Alarif, W.; Kallenborn, R.; Al-Lihaibi, S.S. Occurrence of pharmaceuticals and personal care products in effluent-dominated Saudi Arabian coastal waters of the Red Sea. Chemosphere 2017, 175, 505–513. [Google Scholar] [CrossRef]
  35. Krogh, J.; Lyons, S.; Lowe, C.J. Pharmaceuticals and personal care products in municipal wastewater and the marine receiving environment near Victoria Canada. Front. Mar. Sci. 2017, 4, 415. [Google Scholar] [CrossRef] [Green Version]
  36. Dey, S.; Bano, F.; Malik, A. Pharmaceuticals and personal care product (PPCP) contamination—A global discharge inventory. In Pharmaceuticals and Personal Care Products: Waste Management and Treatment Technology; Elsevier: Amsterdam, The Netherlands, 2019; pp. 1–26. [Google Scholar] [CrossRef]
  37. Hopkins, Z.R.; Blaney, L. An aggregate analysis of personal care products in the environment: Identifying the distribution of environmentally-relevant concentrations. Environ. Int. 2016, 92–93, 301–316. [Google Scholar] [CrossRef]
  38. Mo, J.; Yang, Q.; Zhang, N.; Zhang, W.; Zheng, Y.; Zhang, Z. A review on agro-industrial waste (AIW) derived adsorbents for water and wastewater treatment. J. Environ. Manag. 2018, 227, 395–405. [Google Scholar] [CrossRef]
  39. Kaczala, F.; Blum, S.E. The Occurrence of Veterinary Pharmaceuticals in the Environment: A Review. Curr. Anal. Chem. 2016, 12, 169–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Jin, Y.; Yuan, T.; Li, J.; Shen, Z.; Tian, Y. Occurrence, health risk assessment and water quality criteria derivation of six personal care products (PCPs) in Huangpu River, China. Environ. Monit. Assess. 2022, 194, 577. [Google Scholar] [CrossRef] [PubMed]
  41. Kishore, S.; Malik, S.; Shah, M.P.; Bora, J.; Chaudhary, V.; Kumar, L.; Sayyed, R.Z.; Ranjan, A. A comprehensive review on removal of pollutants from wastewater through microbial nanobiotechnology-based solutions. Biotechnol. Genet. Eng. Rev. 2022, 1–26. [Google Scholar] [CrossRef]
  42. Malik, S.; Dhasmana, A.; Preetam, S.; Mishra, Y.K.; Chaudhary, V.; Bera, S.P.; Ranjan, A.; Bora, J.; Kaushik, A.; Minkina, T.; et al. Exploring Microbial-Based Green Nanobiotechnology for Wastewater Remediation: A Sustainable Strategy. Nanomaterials 2022, 12, 4187. [Google Scholar] [CrossRef] [PubMed]
  43. Malik, S.; Kishore, S.; Dhasmana, A.; Kumari, P.; Mitra, T.; Chaudhary, V.; Kumari, R.; Bora, J.; Ranjan, A.; Minkina, T.; et al. A Perspective Review on Microbial Fuel Cells in Treatment and Product Recovery from Wastewater. Water 2023, 15, 316. [Google Scholar] [CrossRef]
  44. Malik, S.; Kishore, S.; Bora, J.; Chaudhary, V.; Kumari, A.; Kumari, P.; Kumar, L.; Bhardwaj, A. A Comprehensive Review on Microalgae-Based Biorefinery as Two-Way Source of Wastewater Treatment and Bioresource Recovery. CLEAN–Soil Air Water 2023, 51, 2200044. [Google Scholar] [CrossRef]
  45. Satpati, G.G.; Dikshit, P.K.; Mal, N.; Pal, R.; Sherpa, K.C.; Rajak, R.C.; Rather, S.-U.; Raghunathan, S.; Davoodbasha, M. A state-of-the-art review on the co-cultivation of microalgae-fungi in wastewater for biofuel production. Sci. Total Environ. 2023, 870, 161828. [Google Scholar] [CrossRef] [PubMed]
  46. Sinha, S. Derivatization of milk protein using poly (Acrylamide): Its characterization and application. In Macromolecular Symposia; Wiley: Hoboken, NJ, USA, 2018; Volume 381, No. 1; p. 1800120. [Google Scholar]
  47. Villaverde, J.; Rubio-Bellido, M.; Lara-Moreno, A.; Merchan, F.; Morillo, E. Combined use of microbial consortia isolated from different agricultural soils and cyclodextrin as a bioremediation technique for herbicide-contaminated soils. Chemosphere 2018, 193, 118–125. [Google Scholar] [CrossRef] [Green Version]
  48. Bora, J.; Imam, S.; Vaibhav, V.; Malik, S. Use of Genetic Engineering Approach in Bioremediation of Wastewater. In Modern Approaches in Waste Bioremediation; Shah, M.P., Ed.; Springer: Cham, Switzerland, 2023. [Google Scholar] [CrossRef]
  49. Cheng, W.; Li, J.; Wu, Y.; Xu, L.; Su, C.; Qian, Y.; Zhu, Y.-G.; Chen, H. The behavior of antibiotics and antibiotic resistance genes in the eco-agricultural system: A case study. J. Hazard. Mater. 2016, 304, 18–25. [Google Scholar] [CrossRef]
  50. Cajthaml, T.; Křesinová, Z.; Svobodová, K.; Möder, M. Biodegradation of endocrine-disrupting compounds and suppression of estrogenic activity by ligninolytic fungi. Chemosphere 2009, 75, 745–750. [Google Scholar] [CrossRef]
  51. Akerman-Sanchez, G.; Rojas-Jimenez, K. Fungi for the bioremediation of pharmaceutical-derived pollutants: A bioengineering approach to water treatment. Environ. Adv. 2021, 4, 100071. [Google Scholar] [CrossRef]
  52. Grandclément, C.; Seyssiecq, I.; Piram, A.; Wong-Wah-Chung, P.; Vanot, G.; Tiliacos, N.; Roche, N.; Doumenq, P. From the conventional biological wastewater treatment to hybrid processes, the evaluation of organic micropollutant removal: A review. Water Res. 2017, 111, 297–317. [Google Scholar] [CrossRef] [Green Version]
  53. Haroune, L.; Saibi, S.; Cabana, H.; Bellenger, J.P. Intracellular enzymes contribution to the biocatalytic removal of pharmaceuticals by Trametes hirsuta. Environ. Sci. Technol. 2017, 51, 897–904. [Google Scholar] [CrossRef]
  54. Kang, B.R.; Kim, S.Y.; Kang, M.; Lee, T.K. Removal of pharmaceuticals and personal care products using native fungal enzymes extracted during the ligninolytic process. Environ. Res. 2021, 195, 110878. [Google Scholar] [CrossRef]
  55. Lucas, D.; Castellet-Rovira, F.; Villagrasa, M.; Badia-Fabregat, M.; Barceló, D.; Vicent, T.; Caminal, G.; Sarrà, M.; Rodríguez-Mozaz, S. The role of sorption processes in the removal of pharmaceuticals by fungal treatment of wastewater. Sci. Total Environ. 2018, 610, 1147–1153. [Google Scholar] [CrossRef] [PubMed]
  56. Morel, M.; Meux, E.; Mathieu, Y.; Thuillier, A.; Chibani, K.; Harvengt, L.; Jacquot, J.-P.; Gelhaye, E. Xenomic networks variability and adaptation traits in wood-decaying fungi. Microb. Biotechnol. 2013, 6, 248–263. [Google Scholar] [CrossRef]
  57. Naghdi, M.; Taheran, M.; Brar, S.K.; Kermanshahi-Pour, A.; Verma, M.; Surampalli, R.Y. Removal of pharmaceutical compounds in water and wastewater using fungal oxidoreductase enzymes. Environ. Pollut. 2018, 234, 190–213. [Google Scholar] [CrossRef] [PubMed]
  58. Raychoudhury, T.; Prajapati, S.K. Bioremediation of pharmaceuticals in water and wastewater. In Microbial Bioremediation & Biodegradation; Springer: Singapore, 2020; pp. 425–446. [Google Scholar]
  59. Zhuo, R.; Fan, F. A comprehensive insight into the application of white rot fungi and their lignocellulolytic enzymes in the removal of organic pollutants. Sci. Total Environ. 2021, 778, 14613. [Google Scholar] [CrossRef]
  60. Yang, J.; Li, W.; Ng, T.B.; Deng, X.; Lin, J.; Ye, X. Laccases: Production, expression regulation, and applications in pharmaceutical biodegradation. Front. Microbiol. 2017, 8, 832. [Google Scholar] [CrossRef] [Green Version]
  61. Cruz-Morató, C.; Lucas, D.; Llorca, M.; Rodriguez-Mozaz, S.; Gorga, M.; Petrovic, M.; Barceló, D.; Vicent, T.; Sarrà, M.; Marco-Urrea, E. Hospital wastewater treatment by fungal bioreactor: Removal efficiency for pharmaceuticals and endocrine disruptor compounds. Sci. Total Environ. 2014, 493, 365–376. [Google Scholar] [CrossRef]
  62. Viswanath, B.; Rajesh, B.; Janardhan, A.; Kumar, A.P.; Narasimha, G. Fungal laccases and their applications in bioremediation. Enzym. Res. 2014, 2014, 163242. [Google Scholar] [CrossRef] [Green Version]
  63. Deshmukh, R.; Khardenavis, A.A.; Purohit, H.J. Diverse metabolic capacities of fungi for bioremediation. Indian J. Microbiol. 2016, 56, 247–264. [Google Scholar] [CrossRef] [Green Version]
  64. Blanco-Orta, M.F.; García-de la Cruz, R.F.; Paz-Maldonado, L.M.T.; Pedraza-González, D.A.; Morales-Avila, M.M.; Balderas-Hernández, V.E.; González-Ortega, O.; Pérez-Martínez, A.S. Assessing three industrially produced fungi for the bioremediation of diclofenac. J. Environ. Sci. Health Part A 2023, 1–10. [Google Scholar] [CrossRef]
  65. Chauhan, P.S.; Goradia, B.; Saxena, A. Bacterial laccase: Recent update on production, properties, and industrial applications. 3 Biotech 2017, 7, 323. [Google Scholar] [CrossRef]
  66. Holanda, F.H.; Birolli, W.G.; Morais, E.D.S.; Sena, I.S.; Ferreira, A.M.; Faustino, S.M.M.; Solon, L.G.d.S.; Porto, A.L.M.; Ferreira, I.M. Study of biodegradation of chloramphenicol by endophytic fungi isolated from Bertholletia excelsa (Brazil nuts). Biocatal. Agric. Biotechnol. 2019, 20, 101200. [Google Scholar] [CrossRef]
  67. Moghaddam, A.; Khayatan, D.; Esmaeili Fard Barzegar, P.; Ranjbar, R.; Yazdanian, M.; Tahmasebi, E.; Alam, M.; Abbasi, K.; Esmaeili Gouvarchin Ghaleh, H.; Tebyaniyan, H. Biodegradation of pharmaceutical compounds in industrial wastewater using biological treatment: A comprehensive overview. Int. J. Environ. Sci. Technol. 2023, 20, 5659–5696. [Google Scholar] [CrossRef]
  68. Negi, B.B.; Das, C. Mycoremediation of wastewater, challenges, and current status: A review. Bioresour. Technol. Rep. 2023, 11, 101409. [Google Scholar] [CrossRef]
  69. Narayanan, M.; Ali, S.S.; El-Sheekh, M. A comprehensive review on the potential of microbial enzymes in multipollutant bioremediation: Mechanisms, challenges, and future prospects. J. Environ. Manag. 2023, 334, 117532. [Google Scholar] [CrossRef] [PubMed]
  70. Soares, P.R.S.; Birolli, W.G.; Ferreira, I.M.; Porto, A.L.M. Biodegradation pathway of the organophosphate pesticides chlorpyrifos, methyl parathion, and profenofos by the marine-derived fungus Aspergillus sydowii CBMAI 935 and its potential for methylation reactions of phenolic compounds. Mar. Pollut. Bull. 2021, 166, 112185. [Google Scholar] [CrossRef]
  71. Vacondio, B.; Birolli, W.G.; Ferreira, I.M.; Seleghim, M.H.; Gonçalves, S.; Vasconcellos, S.P.; Porto, A.L. Biodegradation of pentachlorophenol by marine-derived fungus Trichoderma harzianum CBMAI 1677 isolated from ascidian Didemnun ligulum. Biocatal. Agric. Biotechnol. 2015, 4, 266–275. [Google Scholar] [CrossRef]
  72. Vaksmaa, A.; Guerrero-Cruz, S.; Ghosh, P.; Zeghal, E.; Hernando-Morales, V.; Niemann, H. Role of fungi in bioremediation of emerging pollutants. Front. Mar. Sci. 2023, 10, 1070905. [Google Scholar] [CrossRef]
  73. Takio, N.; Yadav, M.; Yadav, H.S. Catalase-mediated remediation of environmental pollutants and potential application–A review. Biocatal. Biotransformation 2021, 39, 389–407. [Google Scholar] [CrossRef]
  74. Chen, Y.; Stemple, B.; Kumar, M.; Wei, N. Cell surface displays fungal laccase as a renewable biocatalyst for degradation of persistent micropollutants bisphenol A and sulfamethoxazole. Environ. Sci. Technol. 2016, 50, 8799–8808. [Google Scholar] [CrossRef]
  75. Karigar, C.S.; Rao, S.S. Role of microbial enzymes in the bioremediation of pollutants: A review. Enzym. Res. 2011, 2011, 805187. [Google Scholar] [CrossRef] [Green Version]
  76. Silva, A.; Delerue-Matos, C.; Figueiredo, S.A.; Freitas, O.M. The use of algae and fungi for removal of pharmaceuticals by bioremediation and biosorption processes: A review. Water 2019, 11, 1555. [Google Scholar] [CrossRef] [Green Version]
  77. Hena, S.; Gutierrez, L.; Croué, J.P. Removal of pharmaceutical and personal care products (PPCPs) from wastewater using microalgae: A review. J. Hazard. Mater. 2021, 403, 124041. [Google Scholar] [CrossRef]
  78. Urlacher, V.B.; Girhard, M. Cytochrome P450 monooxygenases: An update on perspectives for synthetic application. Trends Biotechnol. 2012, 30, 26–36. [Google Scholar] [CrossRef] [PubMed]
  79. Syed, K.; Porollo, A.; Lam, Y.W.; Grimmett, P.E.; Yadav, J.S. CYP63A2 is a catalytically versatile fungal P450 monooxygenase capable of oxidizing higher-molecular-weight polycyclic aromatic hydrocarbons, alkylphenols, and alkanes. Appl. Environ. Microbiol. 2013, 79, 2692–2702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Madadi, R.; Bester, K. Fungi and biochar applications in bioremediation of organic micropollutants from aquatic media. Mar. Pollut. Bull. 2021, 166, 112247. [Google Scholar] [CrossRef]
  81. Li, F.; Ma, F.; Zhao, H.; Zhang, S.; Wang, L.; Zhang, X.; Yu, H. A lytic polysaccharide monooxygenase from a white-rot fungus drives the degradation of lignin by a versatile peroxidase. Appl. Environ. Microbiol. 2019, 85, e02803-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Cruz-Morató, C.; Ferrando-Climent, L.; Rodriguez-Mozaz, S.; Barceló, D.; Marco-Urrea, E.; Vicent, T.; Sarrà, M. Degradation of pharmaceuticals in non-sterile urban wastewater by Trametes versicolor in a fluidized bed bioreactor. Water Res. 2013, 47, 5200–5521. [Google Scholar] [CrossRef]
  83. Jelic, A.; Cruz-Morató, C.; Marco-Urrea, E.; Sarrà, M.; Perez, S.; Vicent, T.; Petrović, M.; Barcelo, D. Degradation of carbamazepine by Trametes versicolor in an air pulsed fluidized bed bioreactor and identification of intermediates. Water Res. 2012, 46, 955–964. [Google Scholar] [CrossRef] [PubMed]
  84. Arun, K.B.; Madhavan, A.; Tarafdar, A.; Sirohi, R.; Anoopkumar, A.N.; Kuriakose, L.L.; Awasthi, M.K.; Binod, P.; Varjani, S.; Sindhu, R. Filamentous fungi for pharmaceutical compounds degradation in the environment: A sustainable approach. Environ. Technol. Innov. 2023, 31, 103182. [Google Scholar]
  85. Aranda, E.; Godoy, P.; Reina, R.; Badia-Fabregat, M.; Rosell, M.; Marco-Urrea, E.; García-Romera, I. Isolation of Ascomycota fungi with capability to transform PAHs: Insights into the biodegradation mechanisms of Penicillium oxalicum. Int. Biodeterior. Biodegrad. 2017, 122, 141–150. [Google Scholar] [CrossRef]
  86. Blánquez, P.; Sarrà, M.; Vicent, T. Development of a continuous process to adapt the textile wastewater treatment by fungi to industrial conditions. Process Biochem. 2008, 43, 1–7. [Google Scholar] [CrossRef]
  87. Rodarte-Morales, A.I.; Feijoo, G.; Moreira, M.T.; Lema, J.M. Operation of stirred tank reactors (STRs) and fixed-bed reactors (FBRs) with free and immobilized Phanerochaete chrysosporium for the continuous removal of pharmaceutical compounds. Biochem. Eng. J. 2012, 66, 38–45. [Google Scholar] [CrossRef]
  88. Rodarte-Morales, A.I.; Feijoo, G.; Moreira, M.T.; Lema, J.M. Biotransformation of three active pharmaceutical compounds by the fungus Phanerochaete chrysosporium in a fed-batch stirred reactor under air and oxygen supply. Biodegradation 2012, 23, 145–156. [Google Scholar] [CrossRef]
  89. Bashir, S.M.; Kimiko, S.; Mak, C.-W.; Fang, J.K.-H.; Gonçalves, D. Personal Care and Cosmetic Products as a Potential Source of Environmental Contamination by Microplastics in a Densely Populated Asian City. Front. Mar. Sci. 2021, 8, 683482. [Google Scholar] [CrossRef]
  90. Saharan, V.K.; Rizwani, M.A.; Malani, A.A.; Pandit, A.B. Effect of the geometry of hydrodynamically cavitating device on the degradation of orange-G. Ultrason. Sonochemistry 2013, 20, 345–353. [Google Scholar] [CrossRef]
  91. Lim, F.Y.; Ong, S.L.; Hu, J. Recent Advances in the Use of Chemical Markers for Tracing Wastewater Contamination in Aquatic Environment: A Review. Water 2017, 9, 143. [Google Scholar] [CrossRef] [Green Version]
  92. Biel-Maeso, M.; Corada-Fernández, C.; Lara-Martín, P.A. Monitoring the occurrence of pharmaceuticals in soils irrigated with reclaimed wastewater. Environ. Pollut. 2018, 235, 312–321. [Google Scholar] [CrossRef]
  93. Rashid, S.S.; Liu, Y.Q. Comparison of life cycle toxicity assessment methods for municipal wastewater treatment with the inclusion of direct emissions of metals, PPCPs, and EDCs. Sci. Total Environ. 2021, 756, 143849. [Google Scholar] [CrossRef] [PubMed]
  94. Tarpani, R.R.Z.; Azapagic, A. Life cycle environmental impacts of advanced wastewater treatment techniques for removal of pharmaceuticals and personal care products (PPCPs). J. Environ. Manag. 2018, 215, 258–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Gaur, N.; Narasimhulu, K.; PydiSetty, Y. Recent advances in the bio-remediation of persistent organic pollutants and their effect on the environment. J. Clean. Prod. 2018, 198, 1602–1631. [Google Scholar] [CrossRef]
  96. Mancuso, G.; Langone, M.; Andreottola, G. A critical review of the current technologies in wastewater treatment plants by using hydrodynamic cavitation process: Principles and applications. J. Environ. Health Sci. Eng. 2020, 18, 311–333. [Google Scholar] [CrossRef] [PubMed]
  97. Pandey, A.; Govindwar, S.P.; Kurade, M.B.; Jeon, B.H. (Eds.) Current Developments in Bioengineering and Biotechnology: Advances in Eco-Friendly and Sustainable Technologies for the Treatment of Textile Wastewater; Elsevier: Amsterdam, The Netherlands, 2023. [Google Scholar]
  98. Chaturvedi, P.; Shukla, P.; Giri, B.S.; Chowdhary, P.; Chandra, R.; Gupta, P.; Pandey, A. Prevalence and hazardous impact of pharmaceutical and personal care products and antibiotics in the environment: A review on emerging contaminants. Environ. Res. 2021, 194, 110664. [Google Scholar] [CrossRef]
  99. Hutchings, M.I.; Truman, A.W.; Wilkinson, B. Antibiotics: Past, present, and future. Curr. Opin. Microbiol. 2019, 51, 72–80. [Google Scholar] [CrossRef] [PubMed]
  100. Mohammed, S.A.; Kahissay, M.H.; Hailu, A.D. Pharmaceuticals wastage and pharmaceuticals waste management in public health facilities of Dessie town, North East Ethiopia. PLoS ONE 2021, 16, e0259160. [Google Scholar] [CrossRef] [PubMed]
  101. Liu, J.-L.; Wong, M.-H. Pharmaceuticals and personal care products (PPCPs): A review on environmental contamination in China. Environ. Int. 2013, 59, 208–224. [Google Scholar] [CrossRef] [PubMed]
  102. Khalid, M.; Abdollahi, M. Environmental Distribution of Personal Care Products and Their Effects on Human Health. Iran. J. Pharm. Res. 2021, 20, 216–253. [Google Scholar] [CrossRef]
  103. Checcucci, A.; Trevisi, P.; Luise, D.; Modesto, M.; Blasioli, S.; Braschi, I.; Mattarelli, P. Exploring the Animal Waste Resistome: The Spread of Antimicrobial Resistance Genes through the Use of Livestock Manure. Front. Microbiol. 2020, 11, 1416. [Google Scholar] [CrossRef]
  104. Ribeiro, A.R.; Sures, B.; Schmidt, T.C. Cephalosporin antibiotics in the aquatic environment: A critical review of occurrence, fate, ecotoxicity and removal technologies. Environ. Pollut. 2018, 241, 1153–1166. [Google Scholar] [CrossRef]
  105. Elizalde-Velázquez, A.; Gómez-Oliván, L.M.; Galar-Martínez, M.; Islas-Flores, H.; Dublán-García, O.; SanJuan-Reyes, N. Amoxicillin in the Aquatic Environment, Its Fate and Environmental Risk. In Environmental Health Risk—Hazardous Factors to Living Species; Larramendy, M., Soloneski, S., Eds.; InTech: London, UK, 2016. [Google Scholar] [CrossRef] [Green Version]
  106. Grossman, T.H. Tetracycline Antibiotics and Resistance. Cold Spring Harb. Perspect. Med. 2016, 6, a025387. [Google Scholar] [CrossRef] [Green Version]
  107. Qandil, A.M.; Al-Zoubi, L.O.; Al-Bakri, A.G.; Amawi, H.A.; Al-Balas, Q.A.; Alkatheri, A.M.; Albekairy, A.M. Synthesis, Antibacterial Evaluation, and QSAR of α-Substituted-N4-Acetamides of Ciprofloxacin and Norfloxacin. Antibiotics 2014, 3, 244–269. [Google Scholar] [CrossRef] [Green Version]
  108. Gothwal, R.; Shashidhar, T. Antibiotic Pollution in the Environment: A Review. CLEAN—Soil Air Water 2015, 43, 479–489. [Google Scholar] [CrossRef]
  109. Peng, X.; Ou, W.; Wang, C.; Wang, Z.; Huang, Q.; Jin, J.; Tan, J. Occurrence and ecological potential of pharmaceuticals and personal care products in groundwater and reservoirs in the vicinity of municipal landfills in China. Sci. Total Environ. 2014, 490, 889–898. [Google Scholar] [CrossRef]
  110. Kung, T.A.; Lee, S.H.; Yang, T.C.; Wang, W.H. Survey of selected personal care products in surface water of coral reefs in Kenting National Park, Taiwan. Sci. Total Environ. 2018, 635, 1302–1307. [Google Scholar] [CrossRef] [PubMed]
  111. Berger, K.P.; Kogut, K.R.; Bradman, A.; She, J.; Gavin, Q.; Zahedi, R.; Parra, K.L.; Harley, K.G. Personal care product use as a predictor of urinary concentrations of certain phthalates, parabens, and phenols in the HERMOSA study. J. Expo. Sci. Environ. Epidemiol. 2019, 29, 21–32. [Google Scholar] [CrossRef] [PubMed]
  112. Dodson, R.E.; Nishioka, M.; Standley, L.J.; Perovich, L.J.; Brody, J.G.; Rudel, R.A. Endocrine Disruptors and Asthma-Associated Chemicals in Consumer Products. Environ. Health Perspect. 2012, 120, 935–943. [Google Scholar] [CrossRef] [Green Version]
  113. Rodríguez-Carmona, Y.; Ashrap, P.; Calafat, A.M.; Ye, X.; Rosario, Z.; Bedrosian, L.D.; Huerta-Montanez, G.; Vélez-Vega, C.M.; Alshawabkeh, A.; Cordero, J.F.; et al. Determinants and characterization of exposure to phthalates, DEHTP, and DINCH among pregnant women in the PROTECT birth cohort in Puerto Rico. J. Expo. Sci. Environ. Epidemiol. 2020, 30, 56–69. [Google Scholar] [CrossRef]
  114. Herzog, B.; Amorós-Galicia, L.; Sohn, M.; Hofer, M.; Quass, K.; Giesinger, J. Analysis of photokinetics of 2′-ethylhexyl-4-methoxycinnamate in sunscreens. Photochem. Photobiol. Sci. Off. J. Eur. Photochem. Assoc. Eur. Soc. Photobiol. 2019, 18, 1773–1781. [Google Scholar] [CrossRef]
  115. Le Thi Minh, T.; Nguyen Phuoc, D.; Dinh Quoc, T.; Ngo, H.H.; Do Hong Lan, C. Presence of e-EDCs in surface water and effluents of pollution sources in Sai Gon and Dong Nai river basin. Sustain. Environ. Res. 2016, 26, 20–27. [Google Scholar] [CrossRef] [Green Version]
  116. Jjemba, P.K. Pharma-Ecology: The Occurrence and Fate of Pharmaceuticals and Personal Care Products in the Environment; John Wiley & Sons: Hoboken, NJ, USA, 2018. [Google Scholar]
  117. Boxall, A.B.; Rudd, M.A.; Brooks, B.W.; Caldwell, D.J.; Choi, K.; Hickmann, S.; Innes, E.; Ostapyk, K.; Staveley, J.P.; Verslycke, T.; et al. Pharmaceuticals and personal care products in the environment: What are the big questions? Environ. Health Perspect. 2012, 120, 1221–1229. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Some common examples of pharmaceuticals and personal care products (PPCPs).
Figure 1. Some common examples of pharmaceuticals and personal care products (PPCPs).
Water 15 02262 g001
Figure 2. Illustration of the different possible mechanisms of water treatment through bioremediation of PPCPs.
Figure 2. Illustration of the different possible mechanisms of water treatment through bioremediation of PPCPs.
Water 15 02262 g002
Figure 3. Mechanism of action of fungal strains in PPCP removal from wastewater.
Figure 3. Mechanism of action of fungal strains in PPCP removal from wastewater.
Water 15 02262 g003
Figure 4. (A,B) shows the removal efficiency (%) of drugs diclofenac and carbamazepine (CAR) and (DIC) from a synthetic wastewater media of selected fungal isolates compared to T. versicolor. The figures are adapted from [29]. (CH) shows CAR’s removal efficiency (%) at pH 5.5 and 6.3 from synthetic wastewater media with the degradation percentages following 24 h (G) and 7 days (H) for each strain tested. The figures are adapted from [22]. (IK) The pharmaceutical removal efficiency of laccase enzymes (I) and removal of acetaminophen related to laccase activities (J) and sugar concentrations (K); decreases in acetaminophen, sulfamethoxazole, and carbamazepine concentrations due to the activities of crude laccase. The colors indicate nonwoody and woody lignocellulosic biomasses in (J,K). The figures are adapted from [19]. * This indicates significantly different concentrations compared to the control are marked by * based on a t-test (* p-value < 0.05).
Figure 4. (A,B) shows the removal efficiency (%) of drugs diclofenac and carbamazepine (CAR) and (DIC) from a synthetic wastewater media of selected fungal isolates compared to T. versicolor. The figures are adapted from [29]. (CH) shows CAR’s removal efficiency (%) at pH 5.5 and 6.3 from synthetic wastewater media with the degradation percentages following 24 h (G) and 7 days (H) for each strain tested. The figures are adapted from [22]. (IK) The pharmaceutical removal efficiency of laccase enzymes (I) and removal of acetaminophen related to laccase activities (J) and sugar concentrations (K); decreases in acetaminophen, sulfamethoxazole, and carbamazepine concentrations due to the activities of crude laccase. The colors indicate nonwoody and woody lignocellulosic biomasses in (J,K). The figures are adapted from [19]. * This indicates significantly different concentrations compared to the control are marked by * based on a t-test (* p-value < 0.05).
Water 15 02262 g004
Table 1. Model research table: significant and broader contribution of studies related to PPCPs.
Table 1. Model research table: significant and broader contribution of studies related to PPCPs.
Authors and DateTimeline of Proposed Hypothesis and ResearchProposed Hypothesis and Significant FindingsThe Significant Contribution of Studies in Current Articles from the Review
[1,2,4,5,6,7,11,33,34,35,36,37,38,39,40]2012–2021Presence of PPCPs and merging resources as sources of contamination in effluent and water bodies at the global level in waterbodies.In total, 28.5% of studies show PPCPs and other sources as contamination of wastewater.
[3,41,42,43,44,45,46,47,48]2018–2023Use of physical, chemical, and biological methods (bioremediation) through general microbes and nanotechnology.In total, 2.5% of data show the use of chemical treatment, bioremediation, and green nanobiotechnology in wastewater treatment.
[15,16,18,19,20,21,22,23,24,25,26,29,31,32,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74]2008–2023PPCP degradation through bioremediation using white-rot fungal strains through fungal-based enzymes (oxidoreductase, lignocellulolytic, and laccases), endophytic fungi, Neopestalotiopsis spp., Rhizopus, Phanerochaete chrysosporium, Trichoderma harzianum, and hybrid process with both the chemical and biological process.In total, 32.7% of derived data explain the function of various fungal enzymes in bioremediation and hybrid bioremediation.
[30,55,75,76,77,78,79,80,81]2011–2021PPCP degradation through combinatorial bioremediation using enzymatic algal and fungal strains (white rot), only algal strains and fungal + biochar, and fungus + bacteria.In total, 9.2% of data explain the role of only pure fungal strains and mixed cultures in wastewater treatment.
[14,17,82,83,84,85,86]2012–2014Biodegradation through Trametes versicolor and Penicillium oxalicum fungal strains using bioreactors.In total, 10% of data reveal involvement and culturing of specific fungal enzymes with the latest techniques using fluidized bed reactors (FBR).
[23,87,88]2012–2015Biodegradation through Phanerochaete chrysosporium fungal strains using bioreactors (fed-batch stirred reactor).
[6,8,9,10,33,51,89,90,91,92,93,94,95,96,97]2018–2023Advanced techniques, devices, and life cycle toxicity assessment methods for treating industrial wastewater contaminated by PPCP pollutants. Use of chemical methods, chemical markers, and dermal contact in tracing and monitoring aquatic contamination, bisphenols, and triclosan in soils irrigated wastewater.In total, 11.76% of studies focus on life cycle assessment techniques for wastewater treatment. The data show methods of monitoring and tracing contaminated wastewater.
Table 2. Distribution of PPCP pollutants and their hazardous responses.
Table 2. Distribution of PPCP pollutants and their hazardous responses.
PollutantUseSourceHazard ResponseReferences
Pharmaceutical Drugs
Penicillin GSTDs such as syphilis/gonorrhea. Bacterial pneumonia, wounds, and other infections,
meningitis, and anthrax
Manure (primarily liquid)Allergic reactions, diarrhea,
nausea/vomiting, neurotoxicity/seizure, fever, and angioedema
[99]
CephalosporinUTI, septic shock, infection of bones and joints, and OBGYN infectionsWater (surface and sewage)Anaphylaxis during surgery,
skin rashes/ urticaria, and positive Coomb’s test
[104]
AmoxicillinUTI and infection of genitals, tonsillitis, LRTI, and otolaryngological infectionsSTP and hospital
effluents, surface water, and WWTPs
Hypersensitivity of all types, including anaphylaxis, and GI disturbances[1,105]
TetracyclineParapharyngeal infections, UTI, and GI infectionsManure and groundwaterBloating, black hairy tongue, throat infection, and migraine[106]
CiprofloxacinUTI, neonatal sepsis, typhoid, and cystic fibrosisRiver water and domestic water supplyNausea, fatigue, malaise, pale
skin, aberrant liver function, and headache
[1,98]
NorfloxacinGenitourinary tract infectionTap water and river waterPain in joints and muscles, rectal pain, stomach upset, dizziness, nausea, and headache[1,107]
SulfamethoxazoleShigellosis, URTI, LRTI, and UTIPlant leaves,
sewage, and river and ground
water
Neuropsychological symptoms, pruritis, and migraine[1,6,98]
ErythromycinRTI, bronchitis, lung infection, diphtheria, and pertussisTap, river, and
sewage
water
Severe GI symptoms, including abdominal cramping[1,6,108]
TrimethoprimCystitisGroundwaterSevere GI manifestations including change in taste and apathy for food[109]
SulphadiazineUTI, Toxoplasma gondii
encephalitis, malaria, infection of the ear, and chancroid
Water (ground)Loss of appetite, emesis, diarrhea,
and headache
[98]
Ibuprofen/
Diclofenac
NSAIDGroundwater and
seawater
Bioaccumulation and toxicity[1]
MetforminAntidiabetic
(antihyperglycemic
agent)
SeawaterBioaccumulation and toxicity[1]
Acetaminophen/
Paracetamol
Antipyretic/analgesicSea and surface water, hospital, municipal effluents and sediments,
and WWTPs
Bioaccumulation and toxicity[1]
Caffeine/NicotineNeuro-stimulantSea, river, treated, and untreated WWTPsBioaccumulation and toxicity[1,6,98]
Naproxen/
Indomethacin
NSAIDGroundwater and river waterBioaccumulation and toxicity[1,109]
Triclocarban/
Triclosan
AntimicrobialWWTP (influent)Bioaccumulation and toxicity[1]
DiltiazemCalcium channel
blockers
Surface water and sedimentBioaccumulation and toxicity[1]
CarbamazepineAnticonvulsantSurface water and sedimentBioaccumulation and toxicity
Morphine/
Dihydrocodeine
Analgesics (opioid)River water and treated and untreated WWTPsBioaccumulation and toxicity[98]
AtenololAntihypertensiveSurface water, seawater, and sedimentBioaccumulation and toxicity[1]
17α-
ethinylestradiol/17
β-estradiol
Estrogen (endocrine-disrupting
compounds)
Sediment and
sludge/WWTPs
Bioaccumulation and toxicity[39]
Methylparaben/
Ethylparaben
Preservative
(endocrine disrupter)
River water and
reservoirs
Bioaccumulation and toxicity[109]
Clofibric acid/
Benzafibrate/
Gemfibrozil
Lipid regulatorGroundwaterBioaccumulation and toxicity[98]
Clotrimazole/
Econazole/
Miconazole/
Tebuconazole/
Ketoconazole
Antifungal and
pesticides
Reservoirs and groundwaterBioaccumulation and toxicity[109]
Personal care products
Alkylphenol polyethoxylated (APEOs):
NP, NPEOs, OP, 4OP, 4tOP, and OPEs
Detergents, disinfectants, and surface cleanersWastewater treatment plants (WWTPs), sludge and sediments, landfill leaks, and surface waterEndocrine disruption interferes with human reproduction,
inhibits progesterone/androstenedione, induces testosterone/17β-estradiol production, decreases the human sperm count and motility, alters hormone metabolism and aberrant hypothalamus–pituitary–adrenal axis activity, ulcerative colitis, hepatic infection, and various carcinomas
[102,110]
Antimicrobials: triclocarban,
Triclosan,
1,4-dichlorobenzene,
and ortho-phenylphenol
Detergents, toothpaste, soaps, and perfumesSolid, sediment, effluent, raw wastewater, drinking water, and surface and groundwaterROS generation, reduced GSH/GSSG ratio, and altered mRNA expressions, aberrant energy production and cell cycle regulation, apoptosis, lipid accumulation, decreased sperm count and motility, PCOS, aberrant thyroid function, autoimmunity, and congenital disabilities[111,112]
Bisphenols:
Bisphenol A, B, F, AF, and S (BPA/B/F/AF/S)
Shampoos and conditioners, sun protection lotion, washing detergents, nail polishes, and shaving creamsFreshwater, tap water, river and marine surface water, and sedimentsAdverse reproductive health, hampered pregnancy, fetal growth, preterm birth, uterine leiomyoma, risk of oxidative damage to nucleic acids, and risk of diabetes[40,112]
D3, 4, 5, and 6Antiperspirants, lotions, oils, shampoos, creams for baby care, deodorants, fragrances, hair care products, lotions, nail colors, and skin cleansersAir and sewage systemsPulmonary toxicity, estrogenic activity, and breast carcinogenesis[102,112]
Ethanolamines:
MEA, DEA, and TEA
Cleaners, detergents, shampoos, and dyesSoil and groundwaterAllergic irritation, contact dermatitis, bronchoconstriction and asthma, scalp irritation, hair loss, cuticle damage, and protein loss
Fragrances:
ATII, DPMI, ABDI, AHDI, HHCB, and AHTN
Room fresheners, body lotions, cleansing lotions, deodorants, fabric softeners, detergents, and other cosmeticsAir and sewage systems and river waterAtopic asthma, phototoxicity, photoallergy, hand eczema and allergy, contact dermatitis, antiprogestogenic effects, and human ovarian cancer
Glycol ethers: ME, BE, IPE, and EEFace wash, lotions, polish/wax, and shaving creamIndoor airLonger pregnancy time, reduced motile sperm count, lethal for developing children, asthma, eczema, allergic rhinosinusitis, IgE sensitization, skin erythema, and contact dermatitis
Insect repellents:
DEET, BR, IR3535, ID, and PBO
Insect repellents and cosmeticsRiver water and groundwaterManageable minor toxicities, erythematous–edematous dermatitis, urticaria, and carcinogenic potential
Parabens: BePB, EP, MP,
PP, and BP
Blush, cosmetics, foundation, mascara, and sunscreenRiver water and soilBreast tumors, endocrine disruptors to birth outcomes, decreased neonates’ body length, and impaired child cognitive abilities
Phthalate:
BBP, DEHP,
DEHTP,
DMP, DiBP, DBP, and DiNP
Several cosmetics, antiperspirants, baby care products, perfumes, hair gels/sprays, and moussesRiver water, fresh water, air, and soilSkin irritation, rhinitis, and eczema, asthma, altered liver and kidney function, metabolic disorder, cardiological problems, obesity, lipid accumulation, and behavioral problems among children[113]
UV filters:
BP1, BP2, BP3, BP4, 3BC, BS, 4DHB, EHMC, PEG25-PABA,
Et-PABA.
HMS, 4HB,
IMC, 4MBC
OC, OD-PABA, OMC, OS, PS, TiO2, and ZnO
CosmeticsGroundwater, surface water, drinking water, wastewater effluents, and seawaterSkin sensitization, impaired reproductive health, and Hirschsprung’s disease[37,102,111,112,114]
Note(s): Table abbreviations: BPA, bisphenol A; BPB, bisphenol B; BPF, bisphenol F; BPAF, bisphenol AF; BPS, bisphenol S; D3, hexamethylcyclotrisiloxane; D4, octamethylcyclotetrasiloxan; D5, decamethylcyclopentasiloxane; D6, dodecamethylcyclohexasiloxane; EMA, ethanolamines; MEA, monoethanolamine; DEA, diethanolamine; TEA, triethanolamine; DPMI, cashmeran; ABDI, celestolide; HHCB, galaxolide; AHDI, phantolide; AHTN, toxalide; ATII, traseolide; ME, 2-methoxyethanol; EE, 2-ethoxyethanol; IPE, 2-isopropoxyethanol; BE, 2-butoxyethanol; BR, bayrepel; IR3535, ethyl butyl-acetylamino propionate; ID, indole; MP, methylparaben; EP, ethylparaben; PP, propylparaben; BP, butylparaben; BePB, benzylparaben; BBP, benzyl butyl phthalate; DEHP, di (2-ethylhexyl) phthalate; DEHTP, di (2-ethylhexyl) terephthalate; DMP, dimethyl phthalate; DiBP, di-isobutyl phthalate; DBP, di-n-butyl phthalate; DiNP, di-isononyl phthalate; BP, benzophenone; 3BC, 3-benzylidene camphor; BS, benzyl salicylate; 4DHB, 4,4′-dihydroxybenzophenone; EHMC, 2-ethylhexyl 4-methoxycinnamate; PEG25-PABA, ethoxylated ethyl 4-amino benzoate; Et-PABA, ethyl-4-aminobenzoate; HMS, homosalate; 4HB, 4-hydroxy benzophenone; IMC, isopentyl-4-methoxycinnamate; 4MBC, 4-methylbenzylidene camphor; OC, octocrylene; OD-PABA, octyl dimethyl para amino benzoate; OMC, octyl-methoxycinnamate; OS, octyl salicylate; PS, phenyl salicylate; NPEOs, nonylphenol polyethoxylates; OP, octylphenol; 4OP, 4-n-octylphenol; 4tOP, 4-tert-octylphenol, OPEs, octylphenol ethoxylates; BR, bayrepel; IR3535, ethyl butyl-acetylamino propionate; ID, indole; PBO, piperonyl butoxide; DEET, N,N-diethyl-m-picaridin.
Table 3. Fungal stains and enzymatic bioremediation of PPCPs in different bioreactors.
Table 3. Fungal stains and enzymatic bioremediation of PPCPs in different bioreactors.
Fungal StrainEnzyme(s) InvolvedPPCPReactor TypeCulture TypeInitial Concentration
(mg/L)
Removal Efficiency
(%)
Reference
Trametes versicolorLaccase, LiP, and MnP
  • Ibuprofen
  • Acetaminophen
  • Ketoprofen
  • Propranolol
  • Azithromycin
  • Ciprofloxacin
Fluidized bedFed-batch
  • 2.34
  • 1.56
  • 0.08
  • 0.06
  • 4.31
  • 84.71
  • 100
  • 100
  • 100
  • 100
  • 100
  • 35
[82]
Trametes versicolorLaccase, LiP, and MnP
  • Carbamazepine
Fluidized bedFed batch
  • 0.05–9
  • 61–94
[83]
Trametes versicolorLaccase, LiP, and MnP
  • 17 β-estradiol (E1)
  • 17 α ethynylestradio
Fixed bedContinuous
  • 3–18.8
  • 7.3
  • >95
  • >95
[86]
Trametes versicolorLaccase, LiP, and MnP
  • Naproxen
  • Ibuprofen
  • Ketoprofen
  • Diclofenac
  • Phenazone
  • Ofloxacin
  • Ciprofloxacine
  • Metronidazole
  • Erithromycin
  • Caffeine
  • Atenolol
  • Carbamazepine
Fluidized bedContinuous
  • 1.62
  • 35.5
  • 2.17
  • 0.477
  • 0.497
  • 3.34
  • 13.0
  • 0.912
  • 0.008
  • 149
  • 2.99
  • 0.056
  • 100
  • 100
  • 95
  • 100
  • 96
  • 98
  • 99
  • 85
  • 100
  • 39
  • 75
  • 0
[61]
Trametes versicolorLaccase, LiP, and MnP
  • Diclofenac
Memnbrane reactorContinuous
  • 0.3–1.5
  • 55
[60]
Trametes versicolorLaccase, LiP, and MnP
  • Triclosan
  • 17 α ethinylestradiol
  • Nonylphenol
Erlenmeyer flaskFed-batch
  • 10
  • 10
  • 10
  • 98
  • 94
  • 90
[50]
Phanerochaet chrysosporiumLiP and MnP
  • Diclofenac
  • Ibuprofen
  • Naproxen
Stirred tankFed-batch
  • 0.8
  • 0.8
  • 0.8
  • >99
  • 75–99
  • >99
[12]
LiP and MnP
  • Diclofenac
  • Ibuprofen
  • Naproxen
  • Carbamazepine
  • Diazepam
Stirred tankContinuous
  • 1
  • 1
  • 1
  • 0.5
  • 0.25–0.5
  • 92
  • 95
  • 95
  • 20–60
  • 0
Pleurotus ostreatusLaccase and MnP
  • Triclosan
  • 17 α ethinylestradiol
  • Nonylphenol
Erlenmeyer flaskFed-batch
  • 10
  • 10
  • 10
  • 98
  • 62
  • 93
[50]
Dichomitus squalene
  • Triclosan
  • 17 α ethinylestradiol
  • Nonylphenol
Erlenmeyer flaskFed-batch
  • 10
  • 10
  • 10
  • 98
  • 78
  • 85
Bjerkandera adusta
  • Triclosan
  • 17 α ethinylestradiol
  • Nonylphenol
Erlenmeyer flaskFed-batch
  • 10
  • 10
  • 10
  • 98
  • 78
  • 85
Table 4. Fungal strains and enzymes in bioremediation of PPCP-contaminated wastewater.
Table 4. Fungal strains and enzymes in bioremediation of PPCP-contaminated wastewater.
Name of the Physiological ParameterType of Physiological ParameterAbbreviationSourceUsed for the Bioremediation ofReferences
LaccaseEnzymeLacCerrena unicolor,
Trametes hispida,
Daedalea quercina,
Coriolus versicolor,
Trametes versicolor
2,4-
Dichlorophenol,
pentachlorophenol
TNT
PAH
Anthracene and
benzo pyrene
Delor 106 (PCB)
PCP
Atrazine
(herbicide)
Di(2-ethylhexyl)
phthalate, heavy
metals
Pesticides
Phenols
PCB
Phenylurea
herbicide diuron
Gasoline
Anthracene
Naphthalene
Organic pollutants
[60]
TyrosinaseEnzymeTyrosAgaricus bisporusPhenolic compounds[95]
Lignin PeroxidaseEnzymeLiPPhanerochaete
chrysosporium
Bentazon
(herbicide)
Trichlorophenol
Poly aromatic
hydrocarbon
(PAH)
Delor 106
(Polychlorinated
biphenyl (PCB))
Phencyclidine
(PCP)
Remazol Brilliant
Blue R
PAH
[81]
Versatile PeroxidaseEnzymeVPPleurotus eryngii, Bjerkandera adustaPhenolic as well as non-phenolic compounds[75]
Manganese peroxidaseEnzymeMnPPhlebia radiata,
Lentinula edodes,
Pleurotus ostreatus,
Phanerochaete
chrysosporium
Trichlorophenol
PAH P.
Delor 106 (PCB)
Remazol Brilliant
Blue R
PCP
PAH
Di(2-ethylhexyl)
phthalate, heavy
metals,
PAH
Reactive black 5,
Veratryl alcohol
[95]
Dye-decolorizing peroxidasesEnzymeDyPsIrpex lacteusNon-phenolic lignin model compounds
Organic compounds
[63]
Unspecific peroxygenasesEnzymeUPOAgrocybe aegerita, Marasmius rotulaNon-phenolic lignin model compounds
Organic compounds
CatalaseEnzymeCatNeurospora crassaHeavy metals[73]
Cytochrome P450CytochromeCytPhanerochaete chrysosporium,
Saccharomyces cerevisiae
Mutagenic/carcinogenic fused-ring high molecular weight PAHs (HMW-PAHs)
Crude oil aliphatic hydrocarbon n-alkanes
Endocrine-disrupting long-chain alkylphenols (APs)
[57]
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

Malik, S.; Bora, J.; Nag, S.; Sinha, S.; Mondal, S.; Rustagi, S.; Hazra, R.; Kumar, H.; Rajput, V.D.; Minkina, T.; et al. Fungal-Based Remediation in the Treatment of Anthropogenic Activities and Pharmaceutical-Pollutant-Contaminated Wastewater. Water 2023, 15, 2262. https://doi.org/10.3390/w15122262

AMA Style

Malik S, Bora J, Nag S, Sinha S, Mondal S, Rustagi S, Hazra R, Kumar H, Rajput VD, Minkina T, et al. Fungal-Based Remediation in the Treatment of Anthropogenic Activities and Pharmaceutical-Pollutant-Contaminated Wastewater. Water. 2023; 15(12):2262. https://doi.org/10.3390/w15122262

Chicago/Turabian Style

Malik, Sumira, Jutishna Bora, Sagnik Nag, Sweta Sinha, Sagar Mondal, Sarvesh Rustagi, Richismita Hazra, Harshavardhan Kumar, Vishnu D. Rajput, Tatiana Minkina, and et al. 2023. "Fungal-Based Remediation in the Treatment of Anthropogenic Activities and Pharmaceutical-Pollutant-Contaminated Wastewater" Water 15, no. 12: 2262. https://doi.org/10.3390/w15122262

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