Next Article in Journal
Comparative Study of Movie Shot Classification Based on Semantic Segmentation
Next Article in Special Issue
In Situ Bioremediation Techniques to Reduce Total Organic Matter Oversaturation of Fluvial Sediments: An Experimental Study
Previous Article in Journal
Effects of Manure Waste Biochars in Mining Soils
Previous Article in Special Issue
Vermiremediation of Biomixtures from Biobed Systems Contaminated with Pesticides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

From Laboratory Tests to the Ecoremedial System: The Importance of Microorganisms in the Recovery of PPCPs-Disturbed Ecosystems

by
María Carmen Molina
1,
Luis Fernando Bautista
2,
Myriam Catalá
1,
María Rosa de las Heras
1,
Pilar Martínez-Hidalgo
1,
Jon San-Sebastián
3 and
Natalia González-Benítez
1,*
1
Department of Biology, Geology, Physics and Inorganic Chemistry (Area of Biodiversity and Conservation), ESCET, Rey Juan Carlos University, 28933 Madrid, Spain
2
Department of Chemical and Environmental Technology (Group of Chemical and Environmental Engineering), ESCET, Rey Juan Carlos University, 28933 Madrid, Spain
3
Department of Studies and Consulting (Area of Engineering and Building), Tragsatec, 28037 Madrid, Spain
*
Author to whom correspondence should be addressed.
Appl. Sci. 2020, 10(10), 3391; https://doi.org/10.3390/app10103391
Submission received: 21 April 2020 / Revised: 6 May 2020 / Accepted: 7 May 2020 / Published: 14 May 2020

Abstract

:
The presence of a wide variety of emerging pollutants in natural water resources is an important global water quality challenge. Pharmaceuticals and personal care products (PPCPs) are known as emerging contaminants, widely used by modern society. This objective ensures availability and sustainable management of water and sanitation for all, according to the 2030 Agenda. Wastewater treatment plants (WWTP) do not always mitigate the presence of these emerging contaminants in effluents discharged into the environment, although the removal efficiency of WWTP varies based on the techniques used. This main subject is framed within a broader environmental paradigm, such as the transition to a circular economy. The research and innovation within the WWTP will play a key role in improving the water resource management and its surrounding industrial and natural ecosystems. Even though bioremediation is a green technology, its integration into the bio-economy strategy, which improves the quality of the environment, is surprisingly rare if we compare to other corrective techniques (physical and chemical). This work carries out a bibliographic review, since the beginning of the 21st century, on the biological remediation of some PPCPs, focusing on organisms (or their by-products) used at the scale of laboratory or scale-up. PPCPs have been selected on the basics of their occurrence in water resources. The data reveal that, despite the advantages that are associated with bioremediation, it is not the first option in the case of the recovery of systems contaminated with PPCPs. The results also show that fungi and bacteria are the most frequently studied microorganisms, with the latter being more easily implanted in complex biotechnological systems (78% of bacterial manuscripts vs. 40% fungi). A total of 52 works has been published while using microalgae and only in 7% of them, these organisms were used on a large scale. Special emphasis is made on the advantages that are provided by biotechnological systems in series, as well as on the need for eco-toxicological control that is associated with any process of recovery of contaminated systems.

1. Introduction

Ecosystem degradation is mainly due to anthropogenic actions that directly and indirectly affect aquatic and terrestrial ecosystems. Among the different phenomena linked to global change, three stand out due to their ubiquity and high impact: habitat loss [1], global warming [2,3] and environmental pollution [4]. In the latter, the effect of emerging pollutants is overriding. Most of these substances are micro- or nanocontaminants dispersed almost ubiquitously with unknown impacts. While the effects of eutrophication and heavy metal pollution have been relatively well characterized, the outcome of emerging contamination has been barely elucidated, and less known regarding their interaction with abiotic factors as well as responses in non-target organisms [5]. Among the substances of emerging concern, pharmaceutical and personal hygiene products (PPCPs) such as medicines for veterinary and human use, fragrances, personal hygiene products, or food supplements and additives, are occupying the interest of researchers because many of them are being continuously and massively discharged into the environment and possess high biological activity [6]. In fact, more than 3000 active organic components have been approved in Europe just for human medical care [7] and a high consumption of pharmacological substances has been reported in North America, Brazil, Europe, Australia and some areas of Asia [8].

2. Pharmaceuticals and Personal Care Products (PPCPs) in the Ecosystems

PPCPs can enter the surface water via direct discharge by industries, hospitals, households etc. and there is emerging evidence showing that wastewater treatment plants (WWTP) can act as hotspots that release large amounts of PPCPs into the environment, which is of particular concern in the regions or WWTP without adequate monitoring management. The low concentrations at which some these pollutants are found in the WWTP represent a greater risk, since conventional analytical techniques do not often detect them. These compounds have been reported ubiquitously in the aquatic environment including surface water and groundwater and even seawater, as well as agricultural soils and terrestrial ecosystems even in pristine areas such as Antarctica [9,10]. They are adsorbed and immobilized on soil particles or absorbed, metabolized and/or bioaccumulated by microorganisms and plants for animal and/or human consumption [11,12]. In many cases, their toxicity or possible mode of action on non-target organisms is unknown [7,13,14] but many have chemical structures resembling natural signaling compounds and allelochemicals, resulting in gene expression, blockage of membrane receptors or disruption of signal transduction [15].
PPCPs affect a wide variety of living beings, from the simplest prokaryotes to the most complex eukaryotes (Table 1 and Table 2). In the case of bacteria, they can promote changes in the structure and function of the ecosystem, including the primary production of biofilms [16,17,18,19]. Nutrients and micropollutants exert combined effects on the configuration of α and β diversity of fluvial microbial communities. Surprisingly, PPCPs have a hitherto unrecognized disruptive effect on abundance variations of central species and on module communities, suggesting the key role that micropollutants play in relationships of microbial coexistence in lotic ecosystems [20].
In arid and semi-arid regions, recovered residual water is used to irrigate agricultural crops or gardens. Among the effects generated by these substances in soils is the appearance of bacteria resistant to multiple drugs, which could lead to the failure of antibiotic treatments [21,22]. PPCPs also affect the fungi from the rhizosphere [23]. The knowledge of the toxicological effects of residual pharmaceutical compounds on crops is very limited. Wheat shoot and root elongation decreased and the antioxidant defensive system in roots was damaged after exposing Triticum aestivum to paracetamol [24]. Little is known regarding the risks to wildlife species, but ecotoxicological relevant effects on fern spores germination and development, a vascular plant model, has been demonstrated in vitro [15,25,26,27,28].
Regarding vertebrates such as birds, fish and small mammals, for example, the inappropriate use of diclofenac and the associated cultural practices related to the disposal of animal carcasses, combined with the high sensitivity of the vultures to diclofenac, were responsible for the decline of the populations of three species of vultures in Asia and northern Africa [29,30,31,32]. Human populations are directly exposed to varying concentrations of PPCPs, such as cosmetics application or through diet and the health risks involved are not completely known [30,31,32,33,34,35,36]. Humans exposure, especially children, to some of the PPCPs may cause serious allergies, antibiotic resistance or even toxicity [33]. Using OMICs (genomic, proteomic, transcriptomic) approaches, other authors described the effect of these micropollutants in mice and confirmed injuries in liver and kidney because of induced perturbations of metabolism and dysregulation of signal transduction processes [34]. Their continuous discharge, bioaccumulation and synergistic combination can cause serious adverse effects to the environment and humans [35]. In addition, PPCPS can have combined synergistic effects (Table 1 and Table 2). Thus, [36,37] using mixtures of common PPCPs their individual and combined effect on the marine bacterium Vibrio fischeri was evaluated All of the compounds that presented narcosis as a toxic mode of action at high doses also showed stimulation at low concentrations, an effect known as hormesis. The maximum stimulation effect of a mixture was greater than the highest stimulation effect of each individually tested compound. In addition, when the exposure time increased, the hormetic effect decreased [37]. These synergistic or antagonistic effects can also be found with other pollutants such as microplastics [38].
PPCPs could be removed by physical adsorption processes (e.g., activated carbon, graphene or carbon nanotubes) and chemical advanced oxidation processes (e.g., ozonation, UV treatment or ionizing irradiation) [39]. However, the mineralization to CO2 is not always complete moreover, intermediate substances generated in physical-chemical processes may be even more toxic than the preexisting ones [40,41]. Techniques that use living organisms (or their by-products) for the elimination of contaminants (biological remediation) have a number of advantages, such as low cost, mild operational conditions and the transformation of PPCPs to non-toxic substances (or complete mineralization) because, in some cases, different microorganisms can cooperate among them for the complete mineralization of the pollutant [39,42]. Thus, some authors [43,44,45,46,47] suggest expanding research on the use of biodegradationtechniques of PPCPs, due to their benefits to the environment. Therefore, when considering that biological techniques are fast, clean and inexpensive, their study should be prioritized, to be able to corroborate or not their degradation efficiency and apply them promptly. Nonetheless, currently physic-chemical techniques are more common than biological techniques (Figure 1) probably due to their ease of standardization.
Circular supply chain management (CSCM) has been recently described as: the integration of circular thinking into the management of the supply chain and its surrounding industrial and natural ecosystems. It systematically restores technical materials and regenerates biological materials toward a zero-waste vision through system-wide innovation in business models and supply chain functions from product/service design to end-of-life and waste management, involving all stakeholders in a product/service lifecycle including parts/product manufacturers, service providers, consumers, and users [48]. Within this framework, ecoremediation is a tool of undoubted value, which refers to the restoration or prevention of disturbances in ecosystems, maintaining or recovering their original balance state with multi-purpose approaches for removing pollutants by a vast array of biological components [49,50]. The basic characteristics of ecorremedial measures are their high buffer and self-protective capacities and the preservation of natural habitats and biological diversity [51]. Hence, these measures can be considered as ecosystemic services, which include:
  • the use of organic compounds and natural substances used as adsorbents (activated carbon, clay, bamboo canes…) or biosurfactants [52,53];
  • living organisms and their active principles as remedial components or biological remediation [54,55,56,57];
  • improvements in ecological engineering [58];
  • development of ecological numerical models [59];
  • environmental and economic assessment [60]; and
  • software solutions development for multi-parameter decision making and management tools [49].
An integrative approach of all these measures will render high efficiency removal-systems [61,62,63]. Regarding PPCPs present in waters and soils ecosystems, biological remediation (or bioremediation) could be an eco-friendly solution, since some organisms can immobilize PPCPs in their biological structures (bioadsorption and bioaccumulation) or even use them as nutrients (carbon, phosphorous, nitrogen source) or electron source. Therefore, bioremediation could be an alternative towards zero-waste vision, with an adequate waste management, main objective for CSCM. Essentially, the biological techniques used to eliminate these contaminants have been defined after the organisms that are directly involved in the process: prokaryotes (bioremediation or bacterial remediation), fungi and yeast (mycoremediation) and plants (phytoremediation). The objective of this mini-review is to screen from basic recovery mediated by microorganisms in pure culture, going by the new biotechnological systems applied to WWTPs, up to the latest advances in in/ex situ construction of ecoremedial systems, emphasizing the importance of microorganisms in each of these processes. The search is framed within the period 2002–2018 and it includes some of the more frequent PPCPs in wastewater: diclofenac (DFC), estradiol (EST) and its derivatives, carbamazepine (CBZ), ibuprofen (IBU), naproxen (NPX), codeine (CDN), sulfamethoxazole (SMX), caffeine (CFN), gemfibrozil (GFZ), triclosan (TCS), nonylphenol (NP), artemisine (ATN) and flumequine (FLU). For the collection of bibliographic data, a systematic review of academic articles related to the descriptors was carried out using SCOPUS electronic databases. Review, original articles, chapter books and books were included in this study. Special emphasis will be made on the advantages provided by biotechnological systems arranged in series and we will end with the ecotoxicological tests, since often, after the depurative process, the toxicity persists since the intermediate metabolisms remain toxic.

2.1. Microbial Remediation in Culture

Although microorganisms can be an important ecosystem resource and offer eco-friendly solutions against PPCPs contamination, few researches have been done regarding the degradation of PPCPs in water and soil by microorganisms. Overall, the research works are focused on the description of the levels of these emerging contaminants in water and more sparingly, in which soil or plant matrices they are easily absorbed or adsorbed (Figure 1). When it comes to eliminating these contaminants, the most commonly used solution is the use of physicochemical techniques (Figure 1). However, bioremediation has the hallmark of effective and inexpensive means for the removal of PPCPs, provided that a well-directed and systematic approach is found to study and implement these complex mechanisms in existing or new WWTPs and in situ aquatic and terrestrial systems. The convenience of using combined physic-chemical and biological techniques should be evaluated [95].

2.1.1. Bioremediation or Bacterial Remediation

A total of 271 papers were found in this section. In 78% of them, the prokaryotes were used in advanced remediation systems, such as active sludge, bioreactors or constructed wetlands. Most pharmaceuticals are toxic to bacterial strains (especially antibiotics), but some naturally occurring bacteria have the ability to biodegrade these contaminants. Thereby, bacteria isolates from water, soil and sediment (usually contaminated) have been used to remove the frequently detected PPCPs [39]. They are able to remove the pollutants by aerobic and anaerobic oxidative degradation, intra and extracellular degradation, and biosorption (absorption and adsorption) (Figure 2).
The acclimatization of prokaryotes is an essential factor to improve the effectiveness of the process. Many of the enzymes involved in the PPCPs removal are inducible so that a previous contact of the microorganisms with the PPCPs is important. For example, the presence of TCS induces the production of ammonia monooxygenase, which can degrade this compound [112]. However, constitutive systems such as oxidation through cytochrome P450 via [113,114] or bacterial aerobic lignocellulolytic enzymes [115,116], are also involved in bacterial degradative processes. Other mechanisms for the removal of PPCPs have been described (Figure 2), for example, desorption and mobilization by biosurfactants [105]; photodegradation coupled to biodegradation [106,108,117] and a degradative pathway by biogenic manganese oxides [109,110].
Most of the PPCPs-degrader bacteria belong to the Proteobacteria or Actinobacteria group, although Planococcus (little used) and Bacillus genera (Figure 3A), both Firmicutes, are used less frequently. All in all, whenever the degradation capacity of a microorganism is detected in the laboratory, it is evaluated as beging potentially useful in biotechnology. Thus, bacterial taxa have been first studied in the laboratory and subsequently applied in situ, except for Actinobacteria that, frequently, have been directly identified in active sludge or bioreactors (Figure 3A). Among the most used, Pseudomonas spp. can be highlighted since they are able to degrade, among other compounds, CBZ, TCS, cephalexin, CFN, SMX and diagnostic contrast media [118,119,120,121,122,123]. A 47% degradation of CBZ in 20 days was reported by Pseudomonas species [124]. P. putida activates the successful degradation of 17α-ethinylestradiol (E2α) through redox reactions mediated biogenic manganese oxide [109,110] (Figure 2). The synergistic action of these processes with engineered Escherichia coli cells transformed with an anchoring motif from P. syringae, allows for the complete mineralization of bisphenol A and NP [125]. Recently, 17β-hydroxysteroid dehydrogenase and its regulators were characterized [126]. Bacillus thuringiensis is capable of degrading NPX, IBU, sulfonamides, trimethoprim, NP, TCS and GFZ, among others [45,108,127,128,129,130,131]. Recently, a new pathway for NPX utilization by B. thuringiensis has been described [132]. Several Sphingomonas are linked with successful PPCPs biodegradation [133,134,135,136,137]. They have been used as bioinocula in recovery processes with successful results [135,138]. The use of nanoparticles in Sphingomonas spp. could be a potential strategy for improving degradation [139]. Recently, it has been immobilized on polydopamine-coated Fe3O4 iron nanoparticles, showing high efficacy in the elimination of NP polyethoxylates and during a greater number of cycles. Furthermore, separation and recycling were more readily achieved for immobilized cells as compared to free cells [137].
Less frequently used (Figure 3A), but with a higher potential, we find other bacteria such as Stenotrophomonas maltophila, which can degrade CBZ, triclocarban, TCS [119] and NP [140]. S. maltophilia can degrade NPX by co-metabolism and enzymatic induction [141]. Acinetobacter sp. can degrade sulfadiazine, sulfamethazine (SMT) and SMX but with different mineralization efficiency [131] and degrade E2α in cometabolization with other hormones [142]. The biotransformation of sulfonamides by Arthrobacter denitrificans gave high degradation values [143,144]. The draft genome from two sulfadiazine Arthrobacter bacteria has been compared with other bacteria genomes in a metagenomic approach that might help to identify the functional genes that are involved in degradation of these PPCPs [145]. Achromobacter denitrificans bioaugmented into laboratory-scale membrane bioreactors increases the SMX degradation rates [146]. Nitrosomonas europaea is an ammonia-oxidizing bacterium that is frequently associated with the degradation of hormones (e.g., [42]). Studies on the evolution of microbial populations in the presence of 17β-estradiol established a relationship between the presence of some bacteria (Nitrosomonas, Bacillus, Pseudomonas, Sphingomonas, Novosphingobium, Alcaligenes, Rhodanobacter and Mycobacterium) and the biomineralization of the contaminant [147,148]. The complete sequencing of different species of Sphingobium are being carried out to characterize and compare the expression activities of NP-degrader genes [149]. Recently, the Cupriavidus sp. complete genome has been sequenced, a caffeine-utilizing bacterium [150] and other PPCPs such as CBZ [123]. Cupriavidus sp. with other bacteria such as Sphingomonas sp., Delftia sp., Acinetobacter sp. And Methylobacterium sp. obtained by a wood-rotting microbial consortium (BOS08) have shown an optimized CBZ biodegradation of 40 % in 8 days [123]. All these bacteria including Planococcus, Zoogloea, and Patulibacter can potentially be PPCPs degraders but they have been poorly studied (Figure 3A). Rhodococcus rhodochrous promote degradation of a variety of drugs such as CBZ [151], TCS [152] and endocrine disruptors [153] among others. Streptomyces sp. has been linked to the degradation of some medications such as CBZ [115].
Purple phototrophic bacteria are anoxygenic commonly found in rivers, lake sediments, and wastewater treatment systems. They are metabolically the most versatile among all prokaryotes; anaerobically photoautotrophic and photoheterotrophic in the light, or microaerobic light, condition. They are distinguished by their ability to perform photosynthesis under anaerobic conditions and without producing oxygen [154]. Some of them have been described as emergency pollutants degraders (Figure 2) and have a high value added because they produce, from pollutants, substances of commercial interest [155,156]. Rhodobacter spheroides has proven to be effective in ameliorating hazardous pollutants found in pharmaceutical wastewater [157]. Rhodopseudomonas palustris and Rhodobacter capsulatus are also capable of metabolizing emerging organic compounds present in domestic wastewater, establishing synergy or competition relationships depending on the oxidized organics [154,158]. Recently, a study showed that Rhodopseudomonas contained genes associated with xenobiotic degrading pathways [119].

2.1.2. Mycoremediation

272 papers have been published while using fungi, or their components, as a tool for bioremediation and, in 40% of them, they have been used in complex systems. White rot fungi (WRF), are the most frequently used (Figure 3B). They belong to Basidiomycota and are recognized for their tolerance to high concentrations of pollutants [47] and for their ability to metabolize and degrade a wide variety of persistent organic compounds. This remarkable capacity has been attributed to extracellular enzymes (ligninolytic enzymes) with low substrate specificity [47], and they can be amplified an 80% using redox mediators [159,160] Three enzymes make up fungal extracellular matrix: laccase, manganese peroxidase and lignin peroxidase, whose relative importance in the degradation processes is species-specific [161,162]. Bioabsorption and intracellular degradation, as well as bioadsorption have also been described [163,164,165] (Figure 2). Some of them have been successfully used in the PPCPs degradation [166], given the constitutive nature of their enzymatic battery, they do not require acclimatization processes [167], although inducible laccase from fungi have been also described [168]. Other key that are reactions involved in these WRF pharmaceutical transformations are formylation, hydroxylation, dehalogenation, deamination, conjugation and polymerization [164,169]. The degradative capacity has been traditionally tested in laboratory and subsequently incorporated into WWTPs through cellular or enzymatic products immobilized in bioreactors, although in some cases such as Irpex and Dichomitus, only the laboratory scale has been approached (Figure 3B). Many applications regarding WRF have been reported in the literature [47].
Trametes versicolor is the most widely studied basidiomycete, showing a high oxidative capability to completely degrade PPCPs [162] (Figure 3B). More than 20 PPCPs susceptible to degradation by Trametes versicolor have been described [121,128,129]. Because Trametes laccases are more active than those of other fungi and bacteria (e.g., [170]), they have been used in numerous biotechnological applications and commercial products (e.g., [46]). Phanerochaete chrysosporium can also degrade several PPCPs under different aeration conditions. Isolated laccases from Trametes versicolor and Phanerochaete chrysosporium have been used in the formation of hybrid-nanoparticles that are capable of eliminating both inorganic and drug compounds from contaminated water [171,172]. Other widely reported WRF are Phanerochaete sordina, Pleurotus ostreatus and Bjerkandra species (for more details see [162,163], Figure 3B). In recent years, WRF such as Panus tigrinus and Dichomitus squales are being evaluated as potentially degrading fungi (Figure 3B). Ganoderma lucidum has been tested as a degrader of different drugs, with good results for diclofenac and ifosfamide [173]. Manganese peroxidase aggregates from this fungus transform NP and TCS [174,175]. Irpex lactatus is another suitable WRF in the biotransformation of PPCPs such as FLU [176], hormones [177,178] NP, TCS [179,180], CBZ and diclofenac [173]. Immobilized laccases from Pycnoporus sanguineus have been used successfully in the biotransformation of different antibiotics [181], hormones [182,183], TCS and NP [184] and from Coriolopsis gallica to remove TCS, diclofenac and hormones, NP [185,186,187]. Recently, other WRFs such as Stropharia rugosoannulata, Gymnopilus luteofolius, Agrocybe erebia [173] and Moniliophra roreri [188] have been tested for their potential as PPCPs-degraders (Figure 3B).
Much less explored is the degradative capacity of other non-lignolitic fungi groups such as Ascomycota and Mucoromycota (Figure 3B). Cultures of Rhodococcus rhodochrous and Aspergillus niger have reported the removal of CBZ up to 10% [189]. Aspergillus species biotransform diclofenac, TCS, estradiol, CFN, NP [159,163,190,191,192,193,194]. Genetically modified Aspergillus increases laccase production, which can be used for biotechnological purposes [163,195,196]. The other study shows the optimization and scale-up of the degradation process regulated by Penicillium oxalicum in a batch reactor and compared the degradation between free and immobilized biomass [197]. Some species of Penicillium can be used effectively in TCS, NP and hormones degradation [198,199,200,201,202]. The yeast Yarrowia lipolytica has been studied for the expression of cytochrome P450 and it seems capable of degrading diclofenac and NPX [191,203]. Trichoderma harzianum is able to degrade 17 β-estradiol [204] and CBZ with a performance similar to that of Pleurotus ostreatus [205]. Free laccase from Myceliophthora thermophile and Lentinula edodes can be used to remove estrogens (e.g., [101,206]), diclofenac, NPX [207] and antibiotics [208]. A recent study has underlined the potential of ascomycetes in purification mycotreatments both for their degradative capacity and for their possibilities in bioadsorption [209].
Regarding Mucoromycota, Cunninghamella elegans can transform ATN [210], GFZ [211], NPX [212] and FLU, in two compounds with less antimicrobial activity [213]. Native lipases from Mucor javanicus were tested for the degradation of IBU [214] without very good results. Mucor hiematis associated with commercial fungi can establish synergistic relationships for water purification [215]. Fusarium species degrade estrogens and NP [216,217]. Umbellopsis isabellina is able to degrade and reduce the toxicity of hormones and NP [218,219].
Some Ascomycota and Mucoromycota could be better PPCPs degraders than basidiomycetes, not only for their high tolerance to extreme conditions, but also by the high presence of cytochrome P450 family (CYP) epoxidases and transferases involved in the metabolism of xenobiotics in both groups [164]. However, the improvement of these filamentous fungi through the development of commercial products, biotechnological advances or applications on an industrial scale are in their infancy (Figure 3B).

2.1.3. Phycoremediation

Traditionally, microalgae cultures have not been used for the elimination of PPCPs since these compounds are often toxic to photosynthetic organisms (e.g., [220,221,222]). In fact, some of them are used as ecotoxicity testing organisms due to their sensitiveness to pollutants (Table 1). This is probably the reason why, so far this century, only 52 manuscripts have been published using these microorganisms as PPCPs-removers and only 7% of them propose an important role in advanced biorecovery systems for microalgae (Figure 3C). Scenedesmus obliquus and Chlamydomonas mexicana have some capacity to eliminate CBZ at low concentrations, but higher concentrations inhibited algal growth [223]. IBU stimulated the algal growth of Navicula sp. at lower concentrations, but growth decreased substantially at high concentrations (Figure 4). In addition, the inoculation of Navicula sp. to water systems, can inhibit the degradation of IBU, implying that it could prolong the stress time by this anti-inflammatory drug [224]. However, using wastewater as a source of nitrogen and phosphorus for mass production of rich added-value microalgae with the simultaneous contaminant removal represents an extraordinarily attractive option. Overall, the energy (biogas and biodiesel), pharmaceutical, food and feed industries are an attractive market of microalgae applications (Figure 4, [225,226,227,228]). In fact, estimations for demand in microalgae products project an increase up to US$ 44.6 billion by 2023 [227,229]. Recently, a geographic information system model was built to identify the areas in Mexico with high potential for microalgae production [227].
Some studies have shown that algal treatments of wastewater considerably reduce a high number of PPCPs (Figure 3C), but the toxicity of the water persists, due to the presence of ammonium, a metabolism toxic derivative, [99,234,235]. Although in some analyzes, on a laboratory scale, this toxic compound has been eliminated [103], the number of laboratory tests should be increased in order to confirm these results. Most of the focuses mainly on studies in continuous open ponds (e.g., [236,237]). As far as we know, the PPCPs algae-removal mechanisms are like those described for bacteria and fungi: intracellular and extracellular biodegradation, bioadsorption or/and bioaccumulation, (Figure 2, [238]). Extracellular processes appear to be mediated by laccases, like fungi and bacteria. Thus, Tetracystis laccase converts bisphenol A, 17α-ethinylestradiol, NP and TCS in the presence of a redox mediator [102]. The intracellular degradative metabolic pathway would be mediated by cytochrome P450 and conjugation with the implication of several other enzymes [223], Figure 2). External digestion could be carried out by laccases and other extracelluar enzymes [102,103]. Metabolic pathways for phycodegradation of SMT and SMX have been proposed elsewhere [239]. The degradation of these antibiotics was higher at higher concentrations, indicating that biodegradation might be an efficient mechanism of microalgae acclimatization to antibiotics [240]. Desmodesmus subspicatus uptake and biotransform E2α but yields a very toxic brominated product [241]. Green algae Desmodesmus sp. and Scenedesmus obliquus were used to carry out hydrolysis and reductive dechlorination of TCS, suggesting the possibility of mineralization by these algae [38]. Similarly, [242] not only report the bioremediation of DFC using Chlorella sorokiniana, Chlorella vulgaris and Scenedesmus obliquus, but also show an important reduction in the toxicity of the effluents, especially with S. obliquus. Even non-living S. obliquus can be an effective option in the elimination of DFC and other PPCPs by bioadsorption [243,244] since pollutants with cationic groups are actively attracted to the cell wall through electrostatic interactions [238]. Microalgae such as Desmodesmus subspicatus, Chlamydomonas mexicana and Scenedesmus obliquus can bioaccumulate emerging contaminants along with growth nutrients [223]. The accumulation of PPCPs in microalgae cells can induce the generation of reactive oxygen species (ROS) that are related to defense and adaptation mechanisms [231]. Moreover, Nordic microalgae isolated from nature can remove lipophilic active pharmaceutical ingredients, equally or more efficiently than the investigated culture collection strains. While Coelastrella sp. and Coelastrum astroideum were most efficient in accumulating certain compounds in their biomass, two algae species, Chlorella vulgaris and Chlorella saccharophila, were not only highly efficient in removing all 19 pharmaceuticals but also only small amounts of these compounds accumulated in their biomass allowing for its further use [222]. Nannochloris sp. mediated photo- and bio-degradation were the primary pathways for the removal of 17α-, 17β-estradiol and salicylic acid but the incomplete removal of all steroidal hormones demonstrates that there is still an endocrine disrupting potential in the ecosystem [108]. Recently, the use of an unidentified microalgae consortium, isolated from ponds enriched with emerging contaminants, has improved the results of phycoremediation [236].

2.1.4. Mixed Culture

The complete mineralization of emerging contaminants is achieved while using mixed cultures because the synergistic enzymatic activities of the members of a consortium are often more effective than the individual action of each (Figure 2). However, the response of mixed cultures depend on multiple factors, so if synergistic interactions are not generated, or even if competitive interactions occurred, the results may not be better [151].
Controlled co-cultures have expanded our knowledge. Thus, Trametes versicolor and Ganoderma lucidum were used to remove 13 different pharmaceuticals and generate biodiesel from the generated sludge. Combining both strains increase the removal efficiency as a result of the interactions developed between them [161]. The antibacterial activity of antibiotics could be eliminated after treatments with pure and co-culture of P. chrysosporium and Pycnoporus sanguineus [245]. Co-culture of Alcaligenes faecalis and P. sanguineus degrade well SMX [246]. Cunninghamella sp. could be involved in the degradation of endocrine disrupting chemicals by a mixed culture with WRF [247]. The complete mineralization of E2α by metabolization was achieved by ammonia oxidizing bacteria, together with heterotrophic bacteria [42]. Microbial interactions between Arthrobacter sp. and Pimelobacter sp. also allow for the complete mineralization of sulfadiazine. Synergistic effects were observed in mixed cultures of bacteria and fungi that could effectively and completely remove some PPCPs, while neither fungi nor bacteria alone could remove the compound, suggesting that bacteria and fungi can adopt different but complementary metabolic pathways to remove organic compounds [104,248].
The consortia of cyanobacteria/microalgae and bacteria can be efficient in the detoxification of organic pollutants from wastewaters, as compared to the individual microorganisms. These associations have been frequently used in photoreactors (see Section 2.3.1). Chlorella-Aspergillus pellets were investigated for the degradation of seven pharmacological compounds. In this lab-test, only one of them was degraded and the presence of algae did not suppose a benefit in the final degradation [249]. One of the troubles associated with the use of microalgae in WWTP is the subsequent collection of them (clarification). Co-pelletization using T. versicolor provided a solution to problems that are associated with current energy-intensive and costly algae harvesting processes [250]. Microalga-duckweed association can completely eliminate estrogens from wastewater mediating processes of sorption and biodegradation [99].
Figure 2 shows the documented biodegradation of E2α where different types of microorganisms and natural processes can intervene. Hydrophobic bioabsorption [107] and bioaccumulation can derive in PPCPs stabilization within the biomass, white bioabsorption, and biodegradation up to mineralization (CO2 + H2O) can remove completely pollutants from the ecosystem. Fundamentally bacteria [97,98] and also microalgae [99] usually perform intracellular degradation, whereas fungi metabolize this compound by extracellular degradation that is mediated by laccase. So far, the intervention of other lignolithic enzymes has not been described [101]. Intermediate metabolites could be secreted into the medium and biodegraded through a chain of synergistic reactions that are mediated by other microorganisms [39,42] or bioaccumulated by conjugation and/or accumulation in cellular organelles. Natural adsorbents, such as active carbon, favor the elimination of this hormone [111] and the biosurfactants help with its mobilization and subsequent elimination [105]. Recently, it has been described how photooxidic humic acids (mainly semiquinone radical) not only favor the oxidation of E2α but can also be used as the last acceptors of electrons in respiration, activating the development of bio-degraders [106,107]. A degradative pathway by biogenic manganese oxides has been described [109,110].

2.2. Phytoremediation

Plants can retain PPCPs in their roots by adsorption or translocate soluble pollutants in the water by hydrophilicity-regulated transport via xylems [251]. The uptake of more than 100 PPCPs by plants has been studied both in soil and water systems [33]. Vegetables are known to successfully remove gaseous pollutants, such as benzene [252] although systems must be improved [253]. Several plants can uptake, stabilize and accumulate (Figure 5A), metabolize, mineralize, volatilize, secrete or detoxify contaminants (Figure 5B) through in situ treatments (e.g., [254]). Because of these processes, some plants can use these pollutants as nutrients (source of carbon, nitrogen and phosphorus) generating an increase in their biomass and canceling the toxic effect that they have on other organisms of the ecosystem (Figure 5B), while other plants accumulate them in their vegetable structures (Figure 5A). In both cases, biomass can be harvested and then, the by-products can have a second life or added value (Figure 5A,B).
Although phytoremediation is an excellent biotechnology strategy in soil and water systems, it does not recover 100% of the contaminated area and requires that prior to planting, complementary methods should be used to reduce the amount of pollutants in the area. Additionally, when selecting a plant for phytoremediation, it is necessary to consider its physiological requirements for growth, its condition as a native or introduced plant [261], as well as its effectiveness against a specific pollutant, as well as the secondary added value of the plants (Figure 5), for example, ornamental uses or primary biomass to produce bioethanol [255,258,262]. Hitherto, studies have focused on the effectiveness of recovery processes, the detection of ultra-trace substances along with the identification of metabolites that are produced in the plant upon treatment with active pharmaceutical ingredients [263,264] and in the understanding of metabolic pathways associated with their degradation [265]. However, it would be important to have a more integrated vision proposing a new paradigm.
A plant is a complex multi-genomic organism that is formed by the plant itself and its microbiome [266]. Microbiome (endosphere, rizhosphere and phylosphere) is mostly composed of prokaryotes and fungi. They produce plant growth beneficial effects by the production of nutrients, phytohormones, siderophores and enzymes. They also control plant pathogens and favor acclimation to stress situations. In addition, endophytes play an essential role in detoxification processes. Most of the studies have focused on organic pollutants, heavy metals and metalloids [267,268,269,270,271,272,273]. In relation to organic compounds, Burkholderia kururiensis and Agrobacterium rhizogenes are phenolic tolerant and degrading bacteria. When they are inoculated into root hairs of plant systems, the plant-degradative capacity of some of these pollutants significantly increases [274]. A shallow aquifer planted with Juncus acutus was able to remediate bisphenol-A and another organic compounds from contaminated groundwater, observing that several endophytic strains were tolerant and able to use them [275,276]. However, specific research of rizhophytic and endophytic microorganisms associated with PPCP detoxification is scarce. Some of them claim that associated rhizosphere bacteria are involved in accelerating the degradation of some PPCPs [277,278]. Another study described an efficient biotransformation of anti-inflammatory drugs by endophytic and epiphytic fungi isolated by Plantago lanceolate [279]. An interesting work [280] proposed that enantiomeric-ibuprofen phytodegradation by endophytic bacteria can be involved in its metabolism by lettuce. Rhizobium radiobacter and Chriseobacterium nitroreducens, endophytes in horseradish, improve CBZ removal plant. Of the four metabolic pathways described to degrade CBZ in plants [265], at least one (acridine pathway) is activated by the presence of endophytes [281,282]. It is necessary to highlight the need for further advanced studies on the synergistic effect between plants and PPCPs-degrader endophytic bacteria [265]; however, as far as we know, only the data referred to in this review are available.

2.3. Combined Chemical Engineering-Biological Methods

2.3.1. Bioreactors

Conventional municipal WWTPs, in general, have a very low efficiency for the elimination of PPCPs from polluted waters, as mentioned above These WWTPs could be improved with physical-chemical and biological process. Among the biological processes capable of transforming the original organic compounds into other simpler and less toxic molecules, ideally carbon dioxide, those using controlled bioreactors can be performed at large scale with high PPCPs removal efficiency (Figure 3A–C).
  • Conventional activated sludge
Although the secondary biological treatment in WWTPs is not specifically designed to remove organic compounds at very low concentrations or trace levels, conventional activated sludge (CAS) processes can be optimized to improve the PPCPs removal efficiency. CAS effectiveness depends on the synergistic effect of the consortium, or microorganism mixture, to remove the PPCPs [283], which is constituted by complex communities that are formed by fungi, bacteria and algae [284]. However, both to lab and full-scale, little evidence of this type of synergistic relationship has been described so far and what is the ecology that steers the degradation processes is [187,285].
Conventional municipal WWTPs achieved removal efficiencies of >80% for fragrances, such as galaxolide and tonalide, after the secondary stage, while the uptakes for IBU, NPX, 17 β-estradiol and SMX were within the range 50–65% and the X-ray contrast iopromide was not significantly degraded [286]. However, other studies reached 93% depletion of IBU and 71% of SMX in different sewage treatment plants in Italy, showing important seasonal variations in the removal efficiency of 19 different PPCPs [287]. These variations comprised two main types of factors that, in turn, variate along the year: on one hand, the physical factors that are associated to the environmental conditions and, on the other, the chemical composition of the wastewater influents. Among the physical factors, temperature has the largest impact on the microbial activity, increasing degrading efficiency as temperature increased from winter to summer. With respect to the chemical composition of wastewaters, the load of some PPCPs, especially antibiotics, was lower in summer than in winter because these pharmaceuticals are less used in summer [287]. The removal efficiency of beta-blockers in CAS systems is usually low, ranging from 10% to 81% [287,288] for different waters, seasons and geographical locations, although it has been hypothesized that the removal rate of these drugs also depend on the solid retention time (SRT) of the activated sludge process too [289]. IBU and NPX are compounds showing the desired behavior in CAS treatments since they usually reach very high removal rates (typically 90–99%). On the other hand, CBZ usually shows poor removal values (<25%) in most CAS systems developed [289]. Bioaugmentation might be one strategy for improving the effectiveness of active sludge. For example, adding Spingomonas sp. or Nitrosomonas europea to activate sludge increases the degradation of TCS [112,135]. In relation to the elimination of PPCPs, the composition of the bacterial community (i.e., the identity of the taxa) could be more relevant than the bacterial diversity (i.e., number of different taxa) [290]. The increase in hydraulic retention time and the use of CAS augmented with acclimatized sludge substantially improves the elimination of CBZ and DCF, among others [291].
  • Membrane bioreactors
Membrane bioreactors (MBR) combine the conventional activated sludge technology with the use of a filtration (micro- or ultra-) membrane system. The activated sludge is suspended inside the membrane and drives the biodegradation of organic matter. Meanwhile, membrane retains microorganisms and particles and allows for the permeation of the treated water, avoiding the need to use a settler [292]. MBR have been increasingly adopted by industrial and municipal wastewater treatment plants because of their high efficiency to remove emerging organic pollutants, yielding a high quality water effluent, which is especially suitable for reclamation and reuse [293]. MBR systems have demonstrated good performance for the elimination of some emerging organic pollutants that were not effectively removed by activated sludge treatments. The cause could be the concentration of the biological sludge on the membrane surface thus facilitating the degradation of persistent organic compounds and the retention of larger molecules than the corresponding molecular weight cut-off of the membrane [294,295].
Although MBR processes have been applied for the removal of a wide range of compounds, the effectiveness of elimination can vary greatly from one case to another depending on the nature of the microbial population, the chemical properties of the pollutant and the operation conditions of the process, as in CAS processes. In a pilot scale MBR system, which was equipped with a flat submerged membrane, 23 diverse emerging organic pollutants were evaluated [296]. Five steroid hormones (hydrocortisone, estriol, testosterone, estrone and progesterone) showed high removal degrees (>90%). Similar values were reached by other pharmaceutical products including diuretics, beta-blockers, antibiotics, analgesics and anti-inflammatory drugs such as atenolol, bisoprolol, enalapril, tetracycline, ciprofloxacin, trimethoprim, ketoprofen and NPX, with almost complete removal (>99%) of furosemide, paracetamol (acetaminophen), IBU and the stimulant caffeine. However, the removal of other compounds ranged from relatively high (propanolol 84.2%) to very low (DFC 38.8%, hydroxychlorothiazide 10.2%), or even negligible values (CBZ ~0%).
The solid retention time (SRT) is the most important operating parameter affecting the removal efficiency of PPCPs in MBR systems. In the case of labile and easily degradable compounds, such as IBU, high removal efficiencies were achieved when SRT was above a critical value of 10 d. For recalcitrant pharmaceutical compounds, for example, CBZ, the removal was low whatever the SRT value [293]. However, for PPCPs with intermediate stability, the removal efficiency can be significantly affected by SRT values. For example, when SRT increase from 8 d to 80 d the removal of ketoprofen and clofibric acid in an MBR increased from ~55% to ~91% and from ~4% to ~35%, respectively [297]. The higher degree of PPCPs removal is associated, not only to a higher microbial activity for larger SRT values, but also for the higher microbial diversity reached as SRT increases [293]. Membrane fouling and clogging are two major operational drawbacks with high economic impact on MBR systems [33]. Both aerobic and anaerobic granular sludge have been used to inoculate membrane bioreactors to overcome these limitations. These granular sludge membrane bioreactors (GMBR) allowed for a high biomass loading while nitrogen and organic compounds were effectively removed from wastewaters. The removal of five PPCPs (sulphamethoxazole, norfloxacin, prednisolone, NPX and IBU) from a synthetic wastewater loaded with 50 µg/L of each drug in a GMBR has been studied under aerobic conditions [39]. The continuous bioreactor was formed by a polyvinylidene hollow fiber membrane module (0.22 µm pore size) working under a hydraulic retention time (HRT) of 4 h. Prednisolone concentration decreased >98%, while norfloxacin, NPX and sulphamethoxazole reached 87.8, 84 and 79.8%, respectively. In fact, a dynamics in the microbial populations has been observed in aerobic granular sludge membrane bioreactor, where an increase in the presence of bacteria capable of degrading PPCPs is observed throughout the exposure time [39]. These studies show results that exemplify the complex relationships that are established in communities for the complete removal of emerging pollutants.
  • Photobioreactors
Photosynthetic microorganisms, such as microalgae and cyanobacteria, can, simultaneously, remove nutrients and produce biomass from atmospheric CO2 and solar radiation (photoautotrophic growth). However, some of these microorganisms are able to grow under mixotrophic condition by using organic compounds as a partial carbon source [298]. Therefore, microalgae photobioreactors have been described as an efficient tool for PPCPs and other emerging pollutants in wastewaters [299]. For large scale applications, continuous raceways (Figure 6) are the most viable types of photobioreactors for microalgae-based micropollutant removal due to their low cost and simple operation and maintenance [300,301]. Raceways consist in shallow channels (approximately 0.30 m), equipped with paddle wheels to keep water flowing and avoid settling. Because of it their open design, the microbial community is usually a dynamic consortium, changing its microbial population along time. [302] measured the removal of 52 different pharmaceuticals by a mixed microalgae culture grown in a pond, initially inoculated with Tetradesmus dimorphus. Their results showed a large variability in removal efficiency, ranging from >90% for nine compounds to <10% for 18 compounds. Aeration time improves microalgae activity [303].
A raceway system where microalgae and bacterial species are grown symbiotically is usually referred to as high rate algal pond (HRAP). In these systems, the photosynthesis process developed by microalgae supply the oxygen that is required by aerobic bacteria that, in turn, can degrade the organic pollutants. Therefore, the costly aeration that is necessary to support aerobic growth can be avoided. The metabolites from bacterial degradation of the contaminants and growth promoters synthesized by bacteria, such as vitamin B12 and indoleacetic acid, can act as promoters for the growth of algae. Algal exudates are the main carbon sources (carbohydrate, protein, and lipid) for bacteria. In addition, the cell surfaces of microalgae can provide a stable habitat for the bacteria [238]. Thereby, cyanobacterial/algal photosynthesis provides oxygen, a key electron acceptor to the pollutant-degrading heterotrophic bacteria. In turn, bacteria support photoautotrophic growth of the partners by providing carbon dioxide and other stimulatory means [304]. The synergistic relationship in Chlorella-bacterial microcosms, substantially improves the removal of aromatic pollutants [305]. Competitive interaction between algae and bacteria also exists, so it is substantial to increase the number of consortium studies, in order to select the most effective systems. An HRAP containing Stigeoclonium sp., Chlorella sp., Monoraphidium sp. and diatoms as predominant species, showed a very high removal efficiency (>95%) for CFN, acetaminophen, IBU, methyl dihydrojasmonate, hydrocinnamic acid and oxybenzone for a HRT of four days. When HRT increased to eight days, other pollutants, such as ketoprofen, 5-methyl benzotriazole, galaxolide and TCS were removed to an extent higher than 95% [103]. In the above study, other important pharmaceuticals reached lower degree of degradation: DFC (92%), NPX (89%) and CBZ (62%). Other results also showed that HRAP are efficient alternatives for the treatment of wastewaters containing emerging and PPCPs micropollutants. For example, high removal (98%) of anti-inflammatory drugs (IBU, paracetamol, salicylic acid, and CDN) and diuretics, such as hydrochlorothiazide (84%) and furosemide (100%). However, the removal degree for several antibiotics (azithromycin, ciprofloxacin, ofloxacin and erythromycin) and anxiolytics (lorazepam) were <48% and <57%, respectively [250]. Tetracycline can also be removed (>99%) from real wastewaters by HRAP operating at seven days of HRT under summer conditions [106].
Wastewater treatment processes to eliminate micropollutants have been developed in closed bioreactors in order to avoid microbial contamination and population changes that occurs in open systems and to develop axenic cultures, with better control of operating conditions. Although these systems have not been developed on an industrial scale, pilot and laboratory studies proved good performance in the removal of not readily biodegradable pharmaceuticals in conventional raceways or in HRAP, such as salicylic acid, which was removed (>93%) in a bubbling tubular photobioreactor containing a culture of Tetradesmus obliquus supplemented with CO2 [235]. Though more scientific investigations on scale-up studies and technical aspects of microalgae bioremedial system are needed, the future of the PPCPs phycoremediation look promising [106].

2.3.2. Constructed Wetlands

Constructed wetlands (CWs) are ecoremedial engineered systems that are designed to utilize the natural functions of wetland vegetation, sediments, and their associated microbial assemblages for wastewater, greywater or storm water treatment within a more controlled environment [58]. The use of CWs arises from the desire to mimic the ability of natural wetlands to survive and re-establish the functional relationships after out/in disturbances. Nevertheless, on the scale of the impact on planet Earth, anthropic perturbations are the most important detrimental elements [307], so that successful design, the integration of components and the improvement in efficiency are fundamental.
These systems have revealed as excellent solutions for the elimination of emerging contaminants and antibiotic resistance genes [308]. Overall, the effectiveness of these full-scale systems is equal to and even higher than the conventional wastewater treatment plant (WWTP) but at low cost [309,310]. Some of the removal processes that take place in these systems are: phytoremediation, sediment sorption (adsorption, absorption), microbial degradation (anoxic, anaerobic, and aerobic) located on sediment or associated to plants [311,312], and photodegradation [299]. The integration of microalgae-based biotechnologies (such as open algal ponds) and constructed wetlands could increase the efficiency of PPCPs removal systems (Figure 6), since microalgae can accumulate and degrade pollutants, produce oxygen through photosynthesis (a limiting factor in the degradation processes of CW), and the algal debris can be used as the organic substrate for microorganisms ([238], Figure 6). In algal ponds, photodegradation processes are dominant and removal by biosorption is usually low, except for very hydrophobic pollutants [106]. The association in series of both systems, together with anaerobic bioreactors, favors the elimination of pathogenic microorganisms [313] in addition to the degradation of PPCPs (Figure 6). Technical details on the types of wetlands developed, as well as the physical-chemical and biological parameters limiting factors, can be consulted in different reviews [58,299,306,314,315].
The CW success on PPCPs removal depends on multiple factors, such as the type of CW (subsurface flow, horizontal or vertical subsurface flow, etc.), the plant model used and its remediation functions (absorption, degradation, mineralization, volatilization, etc.) and radicular penetration [316], the interaction with the microorganisms present in the CW and the biome that is associated with the plant [316], the chemical nature of each compound [310], the background nutrients and bacterial community evolution in the sediments [125], as well as, abiotic factors such as temperatures, pH, season of the year, aeration systems, etc. [316,317,318,319]. These factors can be balanced to stabilize the system. For example, cold-resistant bacteria can compensate for the negative effects of low temperatures [320]. Among them, the selection of the plant is a determining factor. The roots provide surface area for attached microorganisms, and root growth maintains the hydraulic properties of the substrate. Vegetal species present in the CW determine the composition of the microbial community on the roots and the rhizosphere [290]. The vegetation cover protects the surface from erosion and shading prevents unwanted algae growth [321]. Plants are the main factor that regulates the amount of nutrients in aquatic systems, buffering saprophytic processes and reducing eutrophication [322]. Vegetables also generate a microhabitat that regulates light and temperature, favors sedimentation and prevention of the medium clogging in vertical flow systems [323]. Useful plants to constructed wetlands are: aquatic species, including free-floating species such as Nymphaea sp., Eichhornia sp., Lemna sp., Azolla sp., Salvinia sp., Pistia stratiotes, Eichhornia crassipes, etc., and helophytes as Sagittaria sp., Juncus sp., Phargmites sp., Thypha sp., Scirpus sp and Schoenoplectus sp. [259,324,325]. In many countries, local plant, especially ornamental plants are used, with an aesthetic and commercial added value (Figure 5B). Some examples are: Cannae lilies, Canna indica, Heliconias sp., Zantedeschia aethiopica, etc. [259,260,323].
Microorganisms seem to be less inhibited by PPCPs in constructed wetlands, because plants stabilize and activate the microbial communities by supplying nutrients and oxygen and an adequate microhabitat or biofilms [326,327] increasing aerobic degradation and nitrification [323]. Overall, and as it happens in other full scape systems (active sludge, bioreactors, etc.), the presence of a specific organic pollutant activates and favors certain microbial genomic sequences that may be related to the degradation of pollutants [200,327]. Thus, [200] showed that IBU reduces the diversity of the fungal community, especially within unplanted wetlands. The species enhanced in abundance after IBU exposure were Aspergillus sp. and Trichoderma sp., suggesting a relationship of these species with the degradation of this anti-inflammatory drug. The presence of plants in CW had a remarkable effect on the structure of microbial community, helping buffer against the stress associated with CFN exposure [200]. Other study indicated that the presence of ciprofloxacin had an adverse effect on the bacterial communities in CW, initially reducing their ability to assimilate anthropogenic carbon-based compounds, but the bacterial communities returned to normal functioning after a 25 weeks acclimatization period [328]. Ongoing monitoring and hydrological and vegetation management are essential to maintain site habitat diversity and its functionality due to the natural dynamics of populations that are established over time in all the living components of a CW.
As previously mentioned, the microbiome associated to plants seems to have a fundamental role in their response to stress situations under CW conditions (Figure 6). In fact, [282] suggests the application of the potential endophytes to improve the performance of these systems. Thereby, plant-endophyte synergism in CW enhances the remediation of nutrients [329,330], heavy metal [30,245,331], metalloids [332], textile industrial effluents [333,334] and rural domestic water [318]. Likewise, bioaugmentation with denitrifying strains can activate the non-endophytic microorganisms that are associated with the rhizosphere, favoring the denitrification of nutrients [335] altering the soil microbial community structure by changing the species of nitrogen-related bacteria, and establishing a new bacterial community balance in situ [300,336,337]. In some studies, it has been proven that the remediation effect of bacteria is more important than that of plants [338]. There are few studies analyzing the degradation of PPCPs in CW emphasizing the plant-microorganism synergy effect. Some antibiotics, such as DFC and SMX, have been partially degraded in CW implying degrader endophytes [339,340]. Inoculation of an estradiol-degrading bacterium in a CW favors the degradation of this hormone and modifies the bacterial community of the biofilm where it is stabilized [341].
As far as we know, WRF has not been used in PPCPs-degrader constructed wetlands probably because an aquatic system is not the most suitable habitat for the development of these organisms. However, whole-cell has been used successfully in P. ostreatus bioreactors, for water treatment, even under bacterial stress [342]. Additionally, P. ostreatus harvested biomass from constructed wetlands has been used as a substrate for production of high-quality oyster mushrooms [258]. Thus, commercial P. ostreatus strains used, usually in food, have important degradative potential [342]. The incorporation of the WRF in the constructed wetlands could be considered as a second phase of biomass exploitation (Figure 5B), as long as the plants do not retain the contaminants in their structures for which pollutants analysis protocols on an industrial scale have to be generated. Scarce are also the works in which, ascomycetes potentially beneficial for the performance of these mesocosms, are inoculated into the system. In this respect, constructed wetlands bioaugmented with Trichodema harzianum have been developed [343]. This helped in the establishment of Iris versicolor, but it did not aid on the performance of the treated effluent.
Thirty years of the research into the use of CW as the most common ecoremedial method, for various types of wastewater, has proven that the great number of early worries and negative arguments have been successfully negated [51]. However, despite the advances described in its use, are still developing systems. Abiotic parameters must be better controlled, and engineering improvements must be undertaken, such as effluent and aeration flow systems. In addition, these systems can be a perfect habitat for the development of a wide variety of annoying mosquitoes and potentially transmitters of diseases [344]. Metamorphosis and the appearance of aquatic insects occurred earlier and more frequently when insect larvae (Diptera) were exposed to fluoxetine and citalopram [77], and the cumulative emergency of insects increased up to 89% of exposure to amphetamines [17]. The presence of antibiotic resistance genes in the effluent might entail a risk of antibiotic resistance being spread in the external environments to the system [339], although it has been proven that these systems also eliminate resistance genes [308]. Finally, the benefits that can be potentially obtained from the waste coming from the CWs, such as production of bioethanol, compost, ornamentation, etc. (Figure 5B) are so promising, that they deserve to be studied in detail for their large-scale application.

3. Ecotoxicology

Often, studies do not evaluate the ecotoxicity of PPCPs metabolites or conjugate transformation products, which are sometimes suggested to be biologically damaging [345,346] and often the resulting products can be more toxic [347,348]. In addition, the effects of excipients of the drugs and the additives added to different PPCPs have been poorly studied [349]. NPX photo-transformation products have been suggested to be more toxic than the parent compounds for algae, rotifers and microcrustaceans [346]. When the toxicity of the abiotic degradation products of IBU and 4-acetylbenzoic acid using test organism (Bacillus megaterium, Pseudoaltermones atlantica, and algae of the Chlorella genus) was evaluated, the toxicity was reduced in the case of IBU but the same did not occur with 4-acetylbenzoic acid [350]. Therefore, ecotoxicological evaluation should be a prerequisite for risk assessment of intermediate or non-degraded chemicals in the water and soil recovery processes [351]. Thus, ecotoxicological tests using biomarkers have brought to light important changes in the behavior of populations of aquatic invertebrates [352], stress oxidative in Zebra mussels [68], and delay in fish development and in frog’s metamorphosis [353]. Table 1 contains a detailed review of the most commonly used biomarkers to determine the ecotoxicity of recovered water. For example, the global order of sensitivity of the species to 26 PPCPs considered in aquatic environments was: Vibrio fischeri (15 min) > > algae> crustaceans > fish > WWTP biomass [36]. Fewer biomarkers have been used in agricultural soils or terrestrial ecosystems, where some nematodes, anelids and only one species of arthropod (Table 2) stand out.
The selection of the appropriate biomarkers to environmental pollution detection is important, both in aquatic and terrestrial environments. This selection will come determined by multiple factors, among others, the PPCPs mixture composition (Table 1 and Table 2). In addition, the degree of sensitivity to pharmaceutical products does not depend on the complexity of the organizations [8]. The monogononta rotifers could be one of these organisms. They have high rates of population growth, what allows the study of multigenerational effects in short times. In addition, they are bioindicators sensitive to water changes are susceptible to a wide range of contaminants and reproduce in a parthenogenic way that allows for comparing genetically identical individuals [354]. A chronic exposure to PPCPs of the rotifer Plationus patulus decreased egg production and increased egg detachment [65]. Algae are at the base of aquatic food webs and they play important roles in the transfer of energy and nutrients to species at higher trophic levels. The algae have a rapid reproduction rate and high sensitivity to contamination [67]. Microalgae have ecological relevance, they are easily cultivated in the laboratory and are sensitive [355,356]. In many studies developed with microalgae, after the exposure to PPCPS, an inhibition of the growth has been observed [67,69]. Daphnia magna, a planktonic crustacean of the order Cladocera and Dreissena polymorpha, a mollusc bivalve, are used in ecotoxicological bioassays for their low phenotypic variation and high sensitivity to toxic compounds. In one study signs of oxidative stress with high levels of lipid peroxidation in zebra mussels (Dreissena polymorpha) exposed to GFZ and DFC were proven [68]. Ethinylestradiol and fluoxetine generated changes at hormonal levels, decreasing fertility and fecundity and increased mortality, affecting growth rates in Daphnia magna [76]. Meanwhile, antibiotic contaminants that were commonly released in fisheries reduced the growth of green algae (Scenedesmus vacuolatus) and duckweed (Lemna minor), as well as the viability of the crustacean Daphnia magna [66]. Cimetidine, a common antihistamine, reduced the biomass of Gammarus when exposed to low concentrations [357] and exposure to low concentrations of a mixture of antibiotics resulted in changes in leaf microbial communities, resulting in an increase in body mass of amphipods (Gammarus) [78]. Fishes of the Cyprinidae family, such as Pimephales promelas and Danio rerio are used for their molecular and genetic similarities with other vertebrates as well as their rapid embryogenesis [353,358]. Soil nematodes are useful indicators of soil health, because they are abundant, have high diversity and several trophic groups. They play an important role in soil ecological processes and they are well-adapted to a wide range of environmental conditions [359]. The diversity and structure of the soil nematode community resulted in being vastly altered after exposition to 26 PPCPs. Mammalian steroid hormones caused a decrease in the number of nematodes and a change in the sex ratio [91,92]. Earthworms are in contact with soil pollutants by their feeding and their skin, so that they are very relevant for the assessment of toxicity of organic contaminants and bioavailability [360]. TCS affected the growth and reproductive performance of the earthworm (Eisenia andrei) [93]. The inclusion of multi-omic analyses in these studies, provides improved characterization of biological response, thereby enhancing prediction of toxicological outcomes in whole animals in the absence of morphological effects [81].
Determining the ecotoxicity of PPCP metabolites and transformation products is logistically very demanding, because of the number of the substances produced and released to the environment, their metabolites and environmentally generated byproducts. Therefore, stakeholders demand some prioritization criteria for ecotoxicological risk assessment studies. Two general approaches are proposed to identify the most problematic substances. The first one is the identification of the environmental hazard when considering not only lethality as endpoint, but relevant biological parameters related with species fitness to characterize contaminants adverse effects [9,15]. This is a paradigmatical change slowly taking place in the last few years, which is transforming the traditional one substance—one species assessment towards a real multidisciplinary Ecotoxicology centered on complex ecological networks of interactions in the biosphere. The measurement of environmental levels of substances of concern allows for modeling increasingly sophisticated environmental risk assessment (ERA) methods allowing to prioritize substances [36,361]. Other environmental assessment tools have been proposed in combination with ERA tools, such as Life Cycle Assessment (LCA) and Material Flow Analysis (MFA) [362]. However, this effort might not be enough and limitation in the number of substances should also be considered. In this sense, the European Registration, Evaluation, Authorization and Restriction of Chemicals (REACH) [363] regulation transfers the burden of ecotoxicological testing to producers and traders of chemicals (EC 1907/2006). Application of this kind of strategies, will both help reduce the number of new chemicals and give insight into the amount being released into the environment. This knowledge helps performing a chemical—directed environmental monitoring and ecotoxicology testing. In particular REACH identifies and prioritizes three categories of substances of very high concern (SVHC): (i) carcinogens, mutagens and toxic for reproduction substances (CMR), (ii) persistent, bioaccumulative and toxic substances (PBT) and very persistent, very bioaccumulative (vPvB) substances. Finally, the increasing use of green and bioremediation technologies to decontaminate and buffer urban and industrial effluents such as the ones reviewed in this paper will certainly help reduce the number and load of substances, metabolites and byproducts of environmental concern.

4. Conclusions and Future Perspectives

Ecoremediation allows for the recovery of systems that are altered by organic contaminants and specifically PPCPs. Microorganisms (aerobic and anaerobic bacteria, basidiomycete and ascomycete fungi, and some resistant algae) can function as eco-factories, capable of restoring ecosystems through clean, low-cost technologies, thus being an important link in the CSMC. Since the mineralization pathways of PPCPs can be complex, further research is necessary to systematically establish collaborative relationships between microorganisms. These researches, to the meso and macro-scale, could allow for achieving complete metabolization, while avoiding intermediate metabolites which are also often ecotoxic compounds and with multiple target cells [364]. Furthermore, the association of these microorganisms with macroorganisms (plants and fungi) in series systems, complements the degradation processes and amplifies the possibilities of success [365]. The engineering of the systems as well as the optimization of the processes in the field are lines of research that can be improved in the future. Moreover, rhizo- and endophytes in this type of system not only degrade PPCPs directly but also accelerate plant growth by producing growth-promoting enzymes and, hence, the remediation potential of CWs [366], line of research that is in the early stages of discovery. On the other hand, certain microalgae are proposed as good eco-factories, not only for their capacity as degraders of pollutant compounds, but also as a source of bioenergy and natural pigments, especially in situations of saline stress, that is, for the treatment of polluted sea wastewater. The efficiency and profitability of these must be improved [367].
In terms of circular economy, microorganism and macro-organism workers can offer clean raw materials that can be used in the production chain (food and pharmacological supplements), clean energy (bioethanol, biodiesel or biogas), biofertilizers, etc. More targeted research could provide a focused scope that allows for a better link between renewable raw material sources, composite intermediates, and final products [368,369]. Finally, it is urgent to highlight the need for the integration of chemistry and engineering in order to design truly green and sustainable pharmaceutical or chemical processes [370]. This could allow (1) achieving completely metabolized PPCP in target organisms (animals or men) or (2) obtaining PCPPs there are specific to the target cells (without cross reactions) or (3) producing zero ecotoxic PPCPs. In this manner, we will avoid, from its origin, the presence of toxic organic compounds in ecosystems.
WWTPs should be equipped with advanced technology in order to minimize pharmaceutical release [8]. This requires capital investments and environmentally sustainable entrepreneurship, including the environmental costs as part of the investment assumed benefit to medium and long term. Ethical-environmental values must prevail over economic interest in the short term.

Author Contributions

Conceptualization, M.C.M., L.F.B. and N.G.-B.; Section 1, M.C.M.; Section 2, L.F.B., N.G.-B., M.C.M., J.S.-S. and P.M.-H.; Section 3, M.C. and M.R.d.l.H.; Section 4, M.C.M.; Figure 1, Figure 2 and Figure 3, P.M.-H.; Figure 4, Figure 5 and Figure 6, J.S.-S.; Table 1 and Table 2, M.C. and M.R.d.l.H.; Original draft preparation, M.C.M.; Revision of the draft manuscript, N.G.-B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministerio de Economía y Competitividad (Spain) and FEDER, grant CGL2016-79158-P; Generalitat Valenciana (Spain), grant PROMETEO/2017/039 and Rey Juan Carlos University, grant DRÍADES-2020/00006/009.

Acknowledgments

We are very grateful to Sturgis and Castro for the revision of the English text.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviation

PPCPspharmaceuticals and personal care products
SDGsustainable development goal
WWTPWastewater treatment plants
ROSreactive oxygen species
UVultraviolet
CSCMCircular supply chain management
DFCdiclofenac
ESTestradiol
CBZcarbamazepine
IBUibuprofen
NPXnaproxen
CDNcodeine
SMXsulfamethoxazole
CFNcaffeine
GFZgemfibrozil
TCStriclosan
NPnonylphenol
ATNartemisine
FLUflumequine
E2α17α-ethinylestradiol
MBRmembrane bioreactors
HRThydraulic retention time
HRAPhigh rate algal pond
CWsconstructed wetlands
WRFWhite rot fungi
AChEAcetylcholinesterase
ERODEthoxyresorufin-O-deethylase

References

  1. Hanski, I. Habitat connectivity, habitat continuity, and metapopulations in dynamic landscapes. Oikos 1999, 87, 209. [Google Scholar] [CrossRef]
  2. Quereda Sala, J.; Gil Olcina, A.; Perez Cuevas, A.; Olcina Cantos, J.; Rico Amoros, A.; Montón Chiva, E. Climatic warming in the Spanish Mediterranean: Natural trend or urban effect. Clim. Chang. 2000, 46, 473–483. [Google Scholar] [CrossRef]
  3. Parmesan, C. Ecological and evolutionary responses to recent climate change. Annu. Rev. Ecol. Evol. Syst. 2006, 37, 637–669. [Google Scholar] [CrossRef] [Green Version]
  4. Schmeller, D.S.; Loyau, A.; Bao, K.; Brack, W.; Chatzinotas, A.; De Vleeschouwer, F.; Friesen, J.; Gandois, L.; Hansson, S.V.; Haver, M.; et al. People, pollution and pathogens – Global change impacts in mountain freshwater ecosystems. Sci. Total Environ. 2018, 622, 756–763. [Google Scholar] [CrossRef] [Green Version]
  5. Leston, S.; Nunes, M.; Viegas, I.; Nebot, C.; Cepeda, A.; Pardal, M.Â.; Ramos, F. The influence of sulfathiazole on the macroalgae ulva lactuca. Chemosphere 2014, 100, 105–110. [Google Scholar] [CrossRef] [Green Version]
  6. Kar, S.; Roy, K. Risk assessment for ecotoxicity of pharmaceuticals an emerging issue. Expert Opin. Drug Saf. 2012, 11, 235–274. [Google Scholar] [CrossRef]
  7. Zwiener, C. Occurrence and analysis of pharmaceuticals and their transformation products in drinking water treatment. Anal. Bioanal. Chem. 2007, 387, 1159–1162. [Google Scholar] [CrossRef]
  8. Patel, M.; Kumar, R.; Kishor, K.; Mlsna, T.; Pittman, C.U.; 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]
  9. Esteban, S.; Moreno-Merino, L.; Matellanes, R.; Catalá, M.; Gorga, M.; Petrovic, M.; López de Alda, M.; Barceló, D.; Silva, A.; Durán, J.J.; et al. Presence of endocrine disruptors in freshwater in the northern Antarctic Peninsula region. Environ. Res. 2016, 147, 179–192. [Google Scholar] [CrossRef]
  10. González-Alonso, S.; Merino, L.M.; Esteban, S.; López de Alda, M.; Barceló, D.; Durán, J.J.; López-Martínez, J.; Aceña, J.; Pérez, S.; Mastroianni, N.; et al. Occurrence of pharmaceutical, recreational and psychotropic drug residues in surface water on the northern Antarctic Peninsula region. Environ. Pollut. 2017, 229, 241–254. [Google Scholar] [CrossRef]
  11. Mompelat, S.; Le Bot, B.; Thomas, O. Occurrence and fate of pharmaceutical products and by-products, from resource to drinking water. Environ. Int. 2009, 35, 803–814. [Google Scholar] [CrossRef] [PubMed]
  12. Xie, H.; Hao, H.; Xu, N.; Liang, X.; Gao, D.; Xu, Y.; Gao, Y.; Tao, H.; Wong, M. Pharmaceuticals and personal care products in water, sediments, aquatic organisms, and fish feeds in the Pearl River Delta: Occurrence, distribution, potential sources, and health risk assessment. Sci. Total Environ. 2019, 659, 230–239. [Google Scholar] [CrossRef] [PubMed]
  13. Bernard, G.R.; Wheeler, A.P.; James, A.R.; Roland, S.; Summer, W.R.; Kenneth, P.S.; Fulkerson, W.J.; Patrick, E.W.; Christman, B.; William, D.D.; et al. The Effects of Ibuprofen on the Physiology and Survival of Patients with Sepsis. N. Engl. J. Med. 1997, 152, 529–535. [Google Scholar] [CrossRef] [PubMed]
  14. Musmarra, D.; Prisciandaro, M.; Capocelli, M.; Karatza, D.; Iovino, P.; Canzano, S.; Lancia, A. Degradation of ibuprofen by hydrodynamic cavitation: Reaction pathways and effect of operational parameters. Ultrason. Sonochem. 2016, 29, 76–83. [Google Scholar] [CrossRef]
  15. García-Cambero, J.P.; García-Cortés, H.; Valcárcel, Y.; Catalá, M. Environmental concentrations of the cocaine metabolite benzoylecgonine induced sublethal toxicity in the development of plants but not in a zebrafish embryo-larval model. J. Hazard. Mater. 2015, 300, 866–872. [Google Scholar] [CrossRef]
  16. Rosi-Marshall, E.J.; Kincaid, D.W.; Bechtold, H.A.; Royer, T.V.; Rojas, M.; Kelly, J.J. Pharmaceuticals suppress algal growth and microbial respiration and alter bacterial communities in stream biofilms. Ecol. Appl. 2013, 23, 583–593. [Google Scholar] [CrossRef] [Green Version]
  17. Lee, S.S.; Paspalof, A.M.; Snow, D.D.; Richmond, E.K.; Rosi-Marshall, E.J.; Kelly, J.J. Occurrence and potential biological effects of amphetamine on stream communities. Environ. Sci. Technol. 2016, 50, 9727–9735. [Google Scholar] [CrossRef]
  18. Catalá, M.; Gasulla, F.; Pradas Del Real, A.E.; García-Breijo, F.; Reig-Armiñana, J.; Barreno, E. The organic air pollutant cumene hydroperoxide interferes with NO antioxidant role in rehydrating lichen. Environ. Pollut. 2013, 179, 277–284. [Google Scholar] [CrossRef]
  19. Domínguez-Morueco, N.; Moreno, H.; Barreno, E.; Catalá, M. Preliminary assessment of terrestrial microalgae isolated from lichens as testing species for environmental monitoring: lichen phycobionts present high sensitivity to environmental micropollutants. Ecotoxicol. Environ. Saf. 2014, 99, 35–44. [Google Scholar] [CrossRef] [Green Version]
  20. Feng, J.; Li, B.; Jiang, X.; Yang, Y.; Wells, G.F.; Zhang, T.; Li, X. Antibiotic resistome in a large-scale healthy human gut microbiota deciphered. Environ. Microbiol. 2013, 15, 3119–3120. [Google Scholar]
  21. Franklin, A.M.; Williams, C.F.; Andrews, D.M.; Woodward, E.E.; Watson, J.E. Uptake of Three Antibiotics and an Antiepileptic Drug by Wheat Crops Spray Irrigated with Wastewater Treatment Plant Effluent. J. Environ. Qual. 2015, 45, 546. [Google Scholar] [CrossRef] [PubMed]
  22. Camargo, M.C.; Garcia, A.; Riquelme, A.; Otero, W.; Camargo, C.A.; Hernandez-Garcia, T.; Candia, R.; Bruce, M.G.; Rabkin, C.S. The problem of Helicobacter pylori resistance to antibiotics: a systematic review in Latin America. Am. J. Gastroenterol. 2014, 109, 485–495. [Google Scholar] [CrossRef] [Green Version]
  23. Liu, Y.; Wang, Y.; Xia, Z.; Wang, Y.; Wu, Y.; Gong, Z. Rapid determination of phytosterols by NIRS and chemometric methods. Spectrochim. Acta - Part A Mol. Biomol. Spectrosc. 2019, 211, 336–341. [Google Scholar] [CrossRef] [PubMed]
  24. An, J.; Zhou, Q.; Sun, F.; Zhang, L. Ecotoxicological effects of paracetamol on seed germination and seedling development of wheat (Triticum aestivum L.). J. Hazard. Mater. 2009, 169, 751–757. [Google Scholar] [CrossRef] [PubMed]
  25. Rodriguez-Gil, J.L.; San Sebastián Sauto, J.; González-Alonso, S.; Sánchez Sanchez, P.; Valcarcel, Y.; Catalá, M. Development of cost-effective strategies for environmental monitoring of irrigated areas in Mediterranean regions: Traditional and new approaches in a changing world. Agric. Ecosyst. Environ. 2013, 181, 41–49. [Google Scholar] [CrossRef]
  26. Feito, R.; Valcárcel, Y.; Catalá, M. Biomarker assessment of toxicity with miniaturised bioassays: Diclofenac as a case study. Ecotoxicology 2012, 21, 289–296. [Google Scholar] [CrossRef] [PubMed]
  27. Feito, R.; Valcárcel, Y.; Catalá, M. Preliminary data suggest that venlafaxine environmental concentrations could be toxic to plants. Chemosphere 2013, 90, 2065–2069. [Google Scholar] [CrossRef]
  28. Esteban, S.; Llamas, P.M.M.; García-Cortés, H.; Catalá, M. The endocrine disruptor nonylphenol induces sublethal toxicity in vascular plant development at environmental concentrations: A risk for riparian plants and irrigated crops? Environ. Pollut. 2016, 216, 480–486. [Google Scholar] [CrossRef]
  29. Naidoo, V.; Wolter, K.; Cuthbert, R.; Duncan, N. Veterinary diclofenac threatens Africa’s endangered vulture species. Regul. Toxicol. Pharmacol. 2009, 53, 205–208. [Google Scholar] [CrossRef]
  30. Ashraf, S.; Naveed, M.; Zahir, Z.A.; Afzal, M.; Rehman, K. Plant-endophyte synergism in constructed wetlands enhances the remediation of tannery effluent. Water Sci. Technol. 2018, 77, 1262–1270. [Google Scholar] [CrossRef]
  31. Oaks, J.L.; Gilbert, M.; Virani, M.Z.; Watson, R.T.; Meteyer, C.U.; Rideout, B.A.; Shivaprasad, H.L.; Ahmed, S.; Jamshed, M.; Chaudhry, I.; et al. Diclofenac residues as the cause of vulture population decline in Pakistan. Nature 2004, 427, 630–633. [Google Scholar] [CrossRef] [PubMed]
  32. Velevski, M.; Nikolov, S.C.; Hallmann, B.; Dobrev, V.; Sidiropoulos, L.; Saravia, V.; Tsiakiris, R.; Arkumarev, V.; Galanaki, A.; Kominos, T.; et al. Population decline and range contraction of the Egyptian Vulture Neophron percnopterus in the Balkan Peninsula. Bird Conserv. Int. 2015, 25, 440–450. [Google Scholar] [CrossRef] [Green Version]
  33. Al-Farsi, R.S.; Ahmed, M.; Al-Busaidi, A.; Choudri, B.S. Translocation of pharmaceuticals and personal care products (PPCPs) into plant tissues: A review. Emerg. Contam. 2017, 3, 132–137. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Huang, K.; Deng, Y.; Zhao, Y.; Wu, B.; Xu, K.; Ren, H. Evaluation of the toxic effects of municipal wastewater effluent on mice using omic approaches. Environ. Sci. Technol. 2013, 47, 9470–9477. [Google Scholar] [CrossRef]
  35. Luo, Y.; Guo, W.; Ngo, H.H.; Nghiem, L.D.; Hai, F.I.; Zhang, J.; Liang, S.; Wang, X.C. A review on the occurrence of micropollutants in the aquatic environment and their fate and removal during wastewater treatment. Sci. Total Environ. 2014, 473–474, 619–641. [Google Scholar] [CrossRef]
  36. Ortiz de García, S.A.; Pinto Pinto, G.; García-Encina, P.A.; Irusta-Mata, R. Ecotoxicity and environmental risk assessment of pharmaceuticals and personal care products in aquatic environments and wastewater treatment plants. Ecotoxicology 2014, 23, 1517–1533. [Google Scholar] [CrossRef]
  37. Ortiz de García, S.; García-Encina, P.A.; Irusta-Mata, R. Dose–response behavior of the bacterium Vibrio fischeri exposed to pharmaceuticals and personal care products. Ecotoxicology 2016, 25, 141–162. [Google Scholar] [CrossRef]
  38. Wang, S.; Liu, G.; Wang, S.; Zhu, Z.; Zhao, F.; Liu, F. Joint toxicity of microplastics with triclosan to marine microalgae Skeletonema costatum. Environ. Pollut. 2018, 246, 509–517. [Google Scholar]
  39. Wang, J.; Wang, S. Removal of pharmaceuticals and personal care products (PPCPs) from wastewater: A review. J. Environ. Manag. 2016, 182, 620–640. [Google Scholar] [CrossRef]
  40. Molina, M.C.; González, N.; Bautista, L.F.; Sanz, R.; Simarro, R.; Sánchez, I.; Sanz, J.L. Isolation and genetic identification of PAH degrading bacteria from a microbial consortium. Biodegradation 2009, 20, 789–800. [Google Scholar] [CrossRef]
  41. Marugán, J.; Bru, D.; Pablos, C.; Catalá, M. Comparative evaluation of acute toxicity by Vibrio fischeri and fern spore based bioassays in the follow-up of toxic chemicals degradation by photocatalysis. J. Hazard. Mater. 2012, 213–214, 117–122. [Google Scholar]
  42. Khunjar, W.O.; Mackintosh, S.A.; Baik, S.; Aga, D.S.; Love, N.G. Elucidating the Relative Roles of Ammonia Oxidizing and Heterotrophic Bacteria during the Biotransformation of 17alpha-Ethinylestradiol and Trimethoprim. Environ. Sci. Technol. 2011, 45, 3605–3612. [Google Scholar] [CrossRef] [PubMed]
  43. Domaradzka, D.; Guzik, U.; Hupert-Kocurek, K.; Wojcieszyńska, D. Cometabolic Degradation of Naproxen by Planococcus sp. Strain S5. Water Air Soil Pollut. 2015, 226, 297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Álvarez, P.M.; Jaramillo, J.; López-Piñero, F.; Plucinski, P.K. Preparation and characterization of magnetic TiO2 nanoparticles and their utilization for the degradation of emerging pollutants in water. Appl. Catal. B Environ. 2010, 100, 338–345. [Google Scholar] [CrossRef]
  45. Grenni, P.; Patrolecco, L.; Ademollo, N.; Di Lenola, M.; Barra Caracciolo, A. Capability of the natural microbial community in a river water ecosystem to degrade the drug naproxen. Environ. Sci. Pollut. Res. 2014, 21, 13470–13479. [Google Scholar] [CrossRef]
  46. Rodarte-Morales, A.I.; Feijoo, G.; Moreira, M.T.; Lema, J.M. Biotransformation of three pharmaceutical active 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] [PubMed]
  47. Tortella, G.; Durán, N.; Rubilar, O.; Parada, M.; Diez, M.C. Are white-rot fungi a real biotechnological option for the improvement of environmental health? Crit. Rev. Biotechnol. 2015, 35, 165–172. [Google Scholar] [CrossRef]
  48. Farooque, M.; Zhang, A.; Thürer, M.; Qu, T.; Huisingh, D. Circular supply chain management: A definition and structured literature review. J. Clean. Prod. 2019, 228, 882–900. [Google Scholar] [CrossRef]
  49. Glavan, M.; Ojstršek Zorčič, P.; Pintar, M. A tool for the selection and implementation of eco-remediation mitigation measures. Ecol. Eng. 2019, 130, 53–66. [Google Scholar] [CrossRef]
  50. Lahiri, S.; Ghosh, D.; Bhakta, J.N. Role of Microbes in Eco-Remediation of Perturbed Aquatic Ecosystem. Oceanogr. Coast. Informatics 2018, 25–61. [Google Scholar] [CrossRef]
  51. Svirčev, Z.; Krstić, S.; Vazić, T. The phylosophy and applicability of ecoremediations for the protection of water ecosystems. Acta Geogr. Slov. 2014, 54, 179–188. [Google Scholar] [CrossRef] [Green Version]
  52. El Bakouri, H.; Morillo, J.; Usero, J.; Ouassini, A. Potential use of organic waste substances as an ecological technique to reduce pesticide ground water contamination. J. Hydrol. 2008, 353, 335–342. [Google Scholar] [CrossRef]
  53. Rai, P.K.; Lee, S.S.; Zhang, M.; Tsang, Y.F.; Kim, K.H. Heavy metals in food crops: Health risks, fate, mechanisms, and management. Environ. Int. 2019, 125, 365–385. [Google Scholar] [CrossRef] [PubMed]
  54. Xiang, W.; Xiao, E.Y.; Rengel, Z. Phytoremediation facilitates removal of nitrogen and phosphorus from eutrophicated water and release from sediment. Environ. Monit. Assess. 2009, 157, 277–285. [Google Scholar] [CrossRef] [PubMed]
  55. Li, H.B.; Du, L.N.; Zou, Y.; Li, Y.H. Eco-remediation of branch river in plain river-net at estuary area. Procedia Environ. Sci. 2011, 10, 1085–1091. [Google Scholar] [CrossRef] [Green Version]
  56. Huang, L.; Du, S.; Fan, L.; Lin, X.; Wang, H.; Zhang, Y. Microbial activity facilitates phosphorus adsorption to shallow lake sediment. J. Soils Sediments 2011, 11, 185–193. [Google Scholar] [CrossRef]
  57. Sun, J.H.; Li, Y.; Yang, Y.J.; Cui, P.; Cheng, Z.Y.; Qiao, X.T.; Xu, Y.J. Developing a way to select plants for eutrophication eco-remediation by their nutrient uptake and growth kinetics characteristics. IOP Conf. Ser. Earth Environ. Sci. 2018, 199, 022070. [Google Scholar] [CrossRef]
  58. Kadlec, R.H.; Wallace, S. Treatment Wetlands; CRC Press: Boca Raton, FL, USA, 2008. [Google Scholar]
  59. Yanagi, T.; Yamada, M.; Suzuki, M. A challenge of water purification in Dokai Bay, Japan. Mar. Pollut. Bull. 1999, 38, 1063–1069. [Google Scholar] [CrossRef]
  60. Hong, J.; Yu, Z.; Fu, X.; Hong, J. Life cycle environmental and economic assessment of coal seam gas-based electricity generation. Int. J. Life Cycle Assess. 2019, 24, 1828–1839. [Google Scholar] [CrossRef]
  61. Cottin, N.; Merlin, G. Fate of chlorinated benzenes in laboratory peat and pozzolana filters. Water Air Soil Pollut. 2010, 213, 425–435. [Google Scholar] [CrossRef]
  62. Bozic, M.; Nikolic, G.; Rudic, Z.; Raicevic, V.; Lalevic, B. Constructed wetlands as an alternative restoration measure for shallow lakes. Water Sci. Technol. 2013, 68, 1672–1678. [Google Scholar] [CrossRef] [PubMed]
  63. Ning, D.; Huang, Y.; Pan, R.; Wang, F.; Wang, H. Effect of eco-remediation using planted floating bed system on nutrients and heavy metals in urban river water and sediment: A field study in China. Sci. Total Environ. 2014, 485–486, 596–603. [Google Scholar] [CrossRef] [PubMed]
  64. Maiti, S.; Sasmal, K.; Sinha, S.S.; Singh, M. Analysis of cytotoxicity and genotoxicity on E. coli, human blood cells and Allium cepa suggests a greater toxic potential of hair dye. Ecotoxicol. Environ. Saf. 2016, 124, 248–254. [Google Scholar] [CrossRef] [PubMed]
  65. Martinez Gomez, D.A.; Baca, S.; Walsh, E.J. Lethal and sublethal effects of selected PPCPs on the freshwater rotifer, Plationus patulus. Environ. Toxicol. Chem. 2015, 34, 913–922. [Google Scholar] [CrossRef] [PubMed]
  66. Kołodziejska, M.; Maszkowska, J.; Białk-Bielińska, A.; Steudte, S.; Kumirska, J.; Stepnowski, P.; Stolte, S. Aquatic toxicity of four veterinary drugs commonly applied in fish farming and animal husbandry. Chemosphere 2013, 92, 1253–1259. [Google Scholar] [CrossRef]
  67. Mu, L.; Zeng, X.; Liu, W.; Chen, H.; Bi, R.; Xie, L.; Bouchez, A.; Li, P.; Li, D.; Tang, J.; et al. Sensitivities of seven algal species to triclosan, fluoxetine and their mixtures. Sci. Rep. 2018, 8, 1–10. [Google Scholar]
  68. Quinn, B.; Schmidt, W.; O’Rourke, K.; Hernan, R. Effects of the pharmaceuticals gemfibrozil and diclofenac on biomarker expression in the zebra mussel (Dreissena polymorpha) and their comparison with standardised toxicity tests. Chemosphere 2011, 84, 657–663. [Google Scholar] [CrossRef]
  69. Seoane, M.; Rioboo, C.; Herrero, C.; Cid, Á. Toxicity induced by three antibiotics commonly used in aquaculture on the marine microalga Tetraselmis suecica (Kylin) Butch. Mar. Environ. Res. 2014, 101, 1–7. [Google Scholar] [CrossRef] [Green Version]
  70. Gagné, F.; Blaise, C.; Fournier, M.; Hansen, P.D. Effects of selected pharmaceutical products on phagocytic activity in Elliptio complanata mussels. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2006, 143, 179–186. [Google Scholar] [CrossRef]
  71. de Solla, S.R.; Campbell, S.D.; McInnis, R.; Gillis, P.L.; Gilroy, È.A.M.; Klinck, J.S. Toxicity and bioconcentration of the pharmaceuticals moxifloxacin, rosuvastatin, and drospirenone to the unionid mussel Lampsilis siliquoidea. Sci. Total Environ. 2014, 487, 537–544. [Google Scholar]
  72. Falfushynska, H.I.; Gnatyshyna, L.L.; Osadchuk, O.Y.; Farkas, A.; Vehovszky, A.; Carpenter, D.O.; Gyori, J.; Stoliar, O.B. Diversity of the molecular responses to separate wastewater effluents in freshwater mussels. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2014, 164, 51–58. [Google Scholar] [CrossRef] [PubMed]
  73. Jindra, M.; Palli, S.R.; Riddiford, L.M. The juvenile hormone signaling pathway in insect development. Annu. Rev. Entomol. 2012, 58, 181–204. [Google Scholar] [CrossRef] [PubMed]
  74. Watts, M.M.; Pascoe, D.; Carroll, K. Chronic exposure to 17α-ethinylestradiol and bisphenol A-effects on development and reproduction in the freshwater invertebrate Chironomus riparius (Diptera: Chironomidae). Aquat. Toxicol. 2001, 55, 113–124. [Google Scholar] [CrossRef]
  75. Watts, M.M.; Pascoe, D.; Carroll, K. Exposure to 17α-ethinylestradiol and bisphenol A - Effects on larval moulting and mouthpart structure of Chironomus riparius. Ecotoxicol. Environ. Saf. 2003, 54, 207–215. [Google Scholar] [CrossRef]
  76. Luna, T.O.; Plautz, S.C.; Salice, C.J. Chronic effects of 17α-ethinylestradiol, fluoxetine, and the mixture on individual and population-level end points in Daphnia magna. Arch. Environ. Contam. Toxicol. 2015, 68, 603–611. [Google Scholar] [CrossRef]
  77. Richmond, E.K.; Rosi-Marshall, E.J.; Lee, S.S.; Thompson, R.M.; Grace, M.R. Antidepressants in stream ecosystems: Influence of selective serotonin reuptake inhibitors (SSRIs) on algal production and insect emergence. Freshw. Sci. 2016, 35, 845–855. [Google Scholar] [CrossRef] [Green Version]
  78. Bundschuh, M.; Hahn, T.; Gessner, M.O.; Schulz, R. Antibiotic mixture effects on growth of the leaf-shredding stream detritivore Gammarus fossarum. Ecotoxicology 2017, 26, 547–554. [Google Scholar] [CrossRef] [Green Version]
  79. Gómez-Canela, C.; Miller, T.H.; Bury, N.R.; Tauler, R.; Barron, L.P. Targeted metabolomics of Gammarus pulex following controlled exposures to selected pharmaceuticals in water. Sci. Total Environ. 2016, 562, 777–788. [Google Scholar] [CrossRef]
  80. Garcia, R.N.; Chung, K.W.; Delorenzo, M.E.; Curran, M.C. Individual and mixture effects of caffeine and sulfamethoxazole on the daggerblade grass shrimp Palaemonetes pugio following maternal exposure. Environ. Toxicol. Chem. 2014, 33, 2120–2125. [Google Scholar] [CrossRef]
  81. Huang, S.S.Y.; Benskin, J.P.; Veldhoen, N.; Chandramouli, B.; Butler, H.; Helbing, C.C.; Cosgrove, J.R. A multi-omic approach to elucidate low-dose effects of xenobiotics in zebrafish (Danio rerio) larvae. Aquat. Toxicol. 2017, 182, 102–112. [Google Scholar] [CrossRef]
  82. Galus, M.; Jeyaranjaan, J.; Smith, E.; Li, H.; Metcalfe, C.; Wilson, J.Y. Chronic effects of exposure to a pharmaceutical mixture and municipal wastewater in zebrafish. Aquat. Toxicol. 2013, 132–133, 212–222. [Google Scholar] [CrossRef] [PubMed]
  83. Simmons, D.B.D.; Trudeau, V.L.; Marlatt, V.L.; Moon, T.W.; Sherry, J.P.; Metcalfe, C.D. Interaction of stilbene compounds with human and rainbow trout estrogen receptors. Environ. Toxicol. Chem. 2008, 27, 442–451. [Google Scholar] [CrossRef] [PubMed]
  84. Grabicova, K.; Fedorova, G.; Burkina, V.; Steinbach, C.; Schmidt-Posthaus, H.; Zlabek, V.; Kocour Kroupova, H.; Grabic, R.; Randak, T. Presence of UV filters in surface water and the effects of phenylbenzimidazole sulfonic acid on rainbow trout (Oncorhynchus mykiss) following a chronic toxicity test. Ecotoxicol. Environ. Saf. 2013, 96, 41–47. [Google Scholar] [CrossRef] [PubMed]
  85. Fernández, C.; Carbonell, G.; Babín, M. Effects of individual and a mixture of pharmaceuticals and personal-care products on cytotoxicity, EROD activity and ROS production in a rainbow trout gonadal cell line (RTG-2). J. Appl. Toxicol. 2013, 33, 1203–1212. [Google Scholar] [CrossRef]
  86. Yokota, H.; Eguchi, S.; Hasegawa, S.; Okada, K.; Yamamoto, F.; Sunagawa, A.; Tanaka, M.; Yamamoto, R.; Nakano, E. Assessment of in vitro antiovulatory activities of nonsteroidal anti-inflammatory drugs and comparison with in vivo reproductive toxicities of medaka (Oryzias latipes). Environ. Toxicol. 2015, 24, 296–303. [Google Scholar] [CrossRef]
  87. Nassef, M.; Matsumoto, S.; Seki, M.; Khalil, F.; Kang, I.J.; Shimasaki, Y.; Oshima, Y.; Honjo, T. Acute effects of triclosan, diclofenac and carbamazepine on feeding performance of Japanese medaka fish (Oryzias latipes). Chemosphere 2010, 80, 1095–1100. [Google Scholar] [CrossRef]
  88. Zenobio, J.E.; Sanchez, B.C.; Archuleta, L.C.; Sepulveda, M.S. Effects of triclocarban, N,N-diethyl-meta-toluamide, and a mixture of pharmaceuticals and personal care products on fathead minnows (Pimephales promelas). Environ. Toxicol. Chem. 2014, 33, 910–919. [Google Scholar] [CrossRef]
  89. Yan, L.; Mu, L.; Chen, H.X.; Guo, Z.B.; Luo, Y.J.; Xie, L.T. Combined effects of fluoxetine and triclosan on Pseudorasbora parva. Ying Yong Sheng Tai Xue Bao J. Appl. Ecol. 2018, 29, 3058–3066. [Google Scholar]
  90. Northcott, G.; Graham, L.; Emnet, P.; Storey, B.; Gaw, S. Personal care products and steroid hormones in the Antarctic coastal environment associated with two Antarctic research stations, McMurdo Station and Scott Base. Environ. Res. 2014, 136, 331–342. [Google Scholar]
  91. Hu, C.; Hermann, G.; Pen-Mouratov, S.; Shore, L.; Steinberger, Y. Mammalian steroid hormones can reduce abundance and affect the sex ratio in a soil nematode community. Agric. Ecosyst. Environ. 2011, 142, 275–279. [Google Scholar] [CrossRef]
  92. Beltrán Rodríguez, M.E.; Carbonell, G.; Escuer, M.; Fernández, C.; Gutiérrez, C.; Rodríguez Martín, J.A.; Campos-Herrera, R. Effect of soil properties, heavy metals and emerging contaminants in the soil nematodes diversity. Environ. Pollut. 2016, 213, 184–194. [Google Scholar]
  93. Chevillot, F.; Guyot, M.; Desrosiers, M.; Cadoret, N.; Veilleux, É.; Cabana, H.; Bellenger, J.P. Accumulation and sublethal effects of triclosan and its transformation product methyl-triclosan in the earthworm Eisenia andrei exposed to environmental concentrations in an artificial soil. Environ. Toxicol. Chem. 2018, 37, 1940–1948. [Google Scholar] [CrossRef] [PubMed]
  94. Pennington, M.J.; Rivas, N.G.; Prager, S.M.; Walton, W.E.; Trumble, J.T. Pharmaceuticals and personal care products alter the holobiome and development of a medically important mosquito. Environ. Pollut. 2015, 203, 199–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Larcher, S.; Yargeau, V. The effect of ozone on the biodegradation of 17α-ethinylestradiol and sulfamethoxazole by mixed bacterial cultures. Appl. Microbiol. Biotechnol. 2013, 97, 2201–2210. [Google Scholar] [CrossRef]
  96. Huang, Y.; Guo, J.; Yan, P.; Gong, H.; Fang, F. Sorption-desorption behavior of sulfamethoxazole, carbamazepine, bisphenol A and 17A-ethinylestradiol in sewage sludge. J. Hazard. Mater. 2019, 739–745. [Google Scholar] [CrossRef]
  97. Sarmah, A.K.; Halling-Sørensen, B. Biodegradation of selected emerging organic contaminants in the environment-an overview. In Leading-Edge Environmental Biogradation Research; Pawley, L.E., Ed.; Nova Science Publishers: Hauppauge, NY, USA, 2007; ISBN 1600219039. [Google Scholar]
  98. Combalbert, S.; Hernandez-Raquet, G. Occurrence, fate, and biodegradation of estrogens in sewage and manure. Appl. Microbiol. Biotechnol. 2010, 86, 1671–1692. [Google Scholar] [CrossRef]
  99. Shi, W.; Wang, L.; Rousseau, D.P.L.; Lens, P.N.L. Removal of estrone, 17α-ethinylestradiol, and 17ß-estradiol in algae and duckweed-based wastewater treatment systems. Environ. Sci. Pollut. Res. 2010, 17, 824–833. [Google Scholar] [CrossRef]
  100. Al-Ansari, A.M.; Saleem, A.; Kimpe, L.E.; Sherry, J.P.; McMaster, M.E.; Trudeau, V.L.; Blais, J.M. Bioaccumulation of the pharmaceutical 17α-ethinylestradiol in shorthead redhorse suckers (Moxostoma macrolepidotum) from the St. Clair River, Canada. Environ. Pollut. 2010, 158, 2566–2571. [Google Scholar] [CrossRef]
  101. Eldridge, H.C.; Milliken, A.; Farmer, C.; Hampton, A.S.; Wendland, N.; Coward, L.; Gregory, D.J.; Johnson, C.M. Efficient remediation of 17 α-ethinylestradiol by Lentinula edodes (shiitake) laccase. Biocatal. Agric. Biotechnol. 2017, 10, 64–68. [Google Scholar] [CrossRef] [Green Version]
  102. Otto, B.; Beuchel, C.; Liers, C.; Reisser, W.; Harms, H.; Schlosser, D. Laccase-like enzyme activities from chlorophycean green algae with potential for bioconversion of phenolic pollutants. FEMS Microbiol. Lett. 2015, 362. [Google Scholar] [CrossRef] [Green Version]
  103. Matamoros, V.; Uggetti, E.; García, J.; Bayona, J.M. Assessment of the mechanisms involved in the removal of emerging contaminants by microalgae from wastewater: A laboratory scale study. J. Hazard. Mater. 2016, 301, 197–205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Mikesková, H.; Novotný, C.; Svobodová, K. Interspecific interactions in mixed microbial cultures in a biodegradation perspective. Appl. Microbiol. Biotechnol. 2012. [CrossRef] [PubMed]
  105. Guo, Y.P.; Hu, Y.Y.; Lin, H.; Ou, X.L. Sorption and desorption of 17α-ethinylestradiol onto sediments affected by rhamnolipidic biosurfactants. J. Hazard. Mater. 2018, 344, 707–715. [Google Scholar] [CrossRef] [PubMed]
  106. Norvill, Z.N.; Shilton, A.; Guieysse, B. Emerging contaminant degradation and removal in algal wastewater treatment ponds: Identifying the research gaps. J. Hazard. Mater. 2016, 313, 291–309. [Google Scholar] [CrossRef] [PubMed]
  107. Huang, B.; Lai, C.; Dai, H.; Mu, K.; Xu, Z.; Gu, L.; Pan, X. Microbially reduced humic acid promotes the anaerobic photodegradation of 17α-ethinylestradiol. Ecotoxicol. Environ. Saf. 2019, 171, 313–320. [Google Scholar] [CrossRef] [PubMed]
  108. Bai, X.; Acharya, K. Removal of seven endocrine disrupting chemicals (EDCs) from municipal wastewater effluents by a freshwater green alga. Environ. Pollut. 2019, 247, 534–540. [Google Scholar] [CrossRef]
  109. Furgal, K.M.; Meyer, R.L.; Bester, K. Removing selected steroid hormones, biocides and pharmaceuticals from water by means of biogenic manganese oxide nanoparticles in situ at ppb levels. Chemosphere 2014, 136, 321–326. [Google Scholar] [CrossRef]
  110. Tran, T.N.; Kim, D.G.; Ko, S.O. Synergistic effects of biogenic manganese oxide and Mn(II)-oxidizing bacterium Pseudomonas putida strain MnB1 on the degradation of 17 A-ethinylestradiol. J. Hazard. Mater. 2018, 344, 350–359. [Google Scholar] [CrossRef]
  111. Rovani, S.; Censi, M.T.; Pedrotti, S.L.; Lima, É.C.; Cataluña, R.; Fernandes, A.N. Development of a new adsorbent from agro-industrial waste and its potential use in endocrine disruptor compound removal. J. Hazard. Mater. 2014, 344, 350–359. [Google Scholar] [CrossRef]
  112. Roh, H.; Subramanya, N.; Zhao, F.; Yu, C.P.; Sandt, J.; Chu, K.H. Biodegradation potential of wastewater micropollutants by ammonia-oxidizing bacteria. Chemosphere 2009, 77, 1084–1089. [Google Scholar] [CrossRef]
  113. Barth, S.; Fischer, M.; Schmid, R.D.; Pleiss, J. Sequence and structure of epoxide hydrolases: A systematic analysis. Proteins Struct. Funct. Genet. 2004, 55, 846–855. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Shrestha, P.; Oh, T.J.; Liou, K.; Sohng, J.K. Cytochrome P450 (CYP105F2) from Streptomyces peucetius and its activity with oleandomycin. Appl. Microbiol. Biotechnol. 2008, 79, 555–562. [Google Scholar] [CrossRef] [PubMed]
  115. Popa, C.; Favier, L.; Dinica, R.; Semrany, S.; Djelal, H.; Amrane, A.; Bahrim, G. Potential of newly isolated wild Streptomyces strains as agents for the biodegradation of a recalcitrant pharmaceutical, carbamazepine. Environ. Technol. (U. K.) 2014, 35, 3082–3091. [Google Scholar]
  116. Woo, H.L.; Hazen, T.C.; Simmons, B.A.; DeAngelis, K.M. Enzyme activities of aerobic lignocellulolytic bacteria isolated from wet tropical forest soils. Syst. Appl. Microbiol. 2014, 37, 60–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Huang, W.; Zhang, Y.; Zhang, Y.; Fang, D.; Schauer, J.J. Optimization of the measurement of particle-bound reactive oxygen species with 2′,7′-dichlorofluorescin (DCFH). Water Air Soil Pollut. 2016, 227, 1–10. [Google Scholar] [CrossRef]
  118. Quandt, E.M.; Summers, R.M.; Subramanian, M.V.; Barrick, J.E. Draft genome sequence of the bacterium Pseudomonas putida CBB5, which can utilize caffeine as a sole carbon and nitrogen source. Genome Announc. 2015, 3, 5–6. [Google Scholar] [CrossRef] [Green Version]
  119. Thelusmond, J.R.; Strathmann, T.J.; Cupples, A.M. Carbamazepine, triclocarban and triclosan biodegradation and the phylotypes and functional genes associated with xenobiotic degradation in four agricultural soils. Sci. Total Environ. 2019, 657, 1138–1149. [Google Scholar] [CrossRef]
  120. Devatha, C.P.; Pavithra, N. Isolation and identification of Pseudomonas from wastewater, its immobilization in cellulose biopolymer and performance in degrading Triclosan. J. Environ. Manag. 2019, 232, 584–591. [Google Scholar] [CrossRef]
  121. Kumari, R.; Ghosh Sachan, S. Bioconversion of toxic micropollutant triclosan to 2,4-dichlorophenol using a wastewater isolate Pseudomonas aeruginosa KS2002. Int. J. Environ. Sci. Technol. 2019, 16, 7663–7672. [Google Scholar] [CrossRef]
  122. Mahmoud, I.S.; Altaif, K.I.; Sini, M.K.A.; Daoud, S.; Aqel, N.N. Determination of antimicrobial drug resistance among bacterial isolates in two hospitals of Baghdad. Jordan J. Pharm. Sci. 2020, 13, 1–9. [Google Scholar]
  123. González-Benítez, N.; Molina, M.C.; Arrayás, M. Empirical evidence and mathematical modelling of carbamazepine degradative kinetics by a wood-rotting microbial consortium. Waste Biomass Valori. 2020. [Google Scholar] [CrossRef]
  124. Li, H.; Sumarah, M.W.; Topp, E. Persistence of the tricyclic antidepressant drugs amitriptyline and nortriptyline in agriculture soils. Environ. Toxicol. Chem. 2013, 32, 509–516. [Google Scholar] [CrossRef] [PubMed]
  125. Zhang, Z.; Ruan, Z.; Liu, J.; Liu, C.; Zhang, F.; Linhardt, R.J.; Li, L. Complete degradation of bisphenol A and nonylphenol by a composite of biogenic manganese oxides and Escherichia coli cells with surface-displayed multicopper oxidase CotA. Chem. Eng. J. 2019, 897–908. [Google Scholar] [CrossRef]
  126. Wang, P.; Zheng, D.; Peng, W.; Wang, Y.; Wang, X.; Xiong, W.; Liang, R. Characterization of 17β-hydroxysteroid dehydrogenase and regulators involved in estrogen degradation in Pseudomonas putida SJTE-1. Appl. Microbiol. Biotechnol. 2019, 103, 2413–2425. [Google Scholar] [CrossRef] [PubMed]
  127. Kjeldal, H.; Zhou, N.A.; Wissenbach, D.K.; Von Bergen, M.; Gough, H.L.; Nielsen, J.L. Genomic, proteomic, and metabolite characterization of gemfibrozil-degrading organism Bacillus sp. GeD10. Environ. Sci. Technol. 2016, 50, 744–755. [Google Scholar] [CrossRef]
  128. Marchlewicz, A.; Domaradzka, D.; Guzik, U.; Wojcieszyńska, D. Bacillus thuringiensis B1(2015b) is a gram-positive bacteria able to degrade naproxen and ibuprofen. Water Air Soil Pollut. 2016, 227, 1–8. [Google Scholar] [CrossRef] [Green Version]
  129. Liu, Q.; Li, M.; Liu, X.; Zhang, Q.; Liu, R.; Wang, Z.; Shi, X.; Quan, J.; Shen, X.; Zhang, F. Removal of sulfamethoxazole and trimethoprim from reclaimed water and the biodegradation mechanism. Front. Environ. Sci. Eng. 2018, 12, 6. [Google Scholar] [CrossRef]
  130. Zheng, G.; Wang, T.; Niu, M.; Chen, X.; Liu, C.; Wang, Y.; Chen, T. Biodegradation of nonylphenol during aerobic composting of sewage sludge under two intermittent aeration treatments in a full-scale plant. Environ. Pollut. 2018, 238, 783–791. [Google Scholar] [CrossRef]
  131. Wang, S.; Hu, Y.; Wang, J. Biodegradation of typical pharmaceutical compounds by a novel strain Acinetobacter sp. J. Environ. Manag. 2018, 217, 240–246. [Google Scholar] [CrossRef]
  132. Górny, D.; Guzik, U.; Hupert-Kocurek, K.; Wojcieszyńska, D. A new pathway for naproxen utilisation by Bacillus thuringiensis B1(2015b) and its decomposition in the presence of organic and inorganic contaminants. J. Environ. Manag. 2019, 239, 1–7. [Google Scholar] [CrossRef]
  133. Murdoch, R.W.; Hay, A.G. Genetic and chemical characterization of ibuprofen degradation by Sphingomonas Ibu-2. Microbiology 2013, 159, 621–632. [Google Scholar] [CrossRef] [Green Version]
  134. Thelusmond, J.R.; Strathmann, T.J.; Cupples, A.M. The identification of carbamazepine biodegrading phylotypes and phylotypes sensitive to carbamazepine exposure in two soil microbial communities. Sci. Total Environ. 2016, 571, 1241–1252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Zhou, N.A.; Lutovsky, A.C.; Andaker, G.L.; Ferguson, J.F.; Gough, H.L. Kinetics modeling predicts bioaugmentation with Sphingomonad cultures as a viable technology for enhanced pharmaceutical and personal care products removal during wastewater treatment. Bioresour. Technol. 2014, 166, 158–167. [Google Scholar] [CrossRef] [PubMed]
  136. Kim, Y.M.; Murugesan, K.; Schmidt, S.; Bokare, V.; Jeon, J.R.; Kim, E.J.; Chang, Y.S. Triclosan susceptibility and co-metabolism - A comparison for three aerobic pollutant-degrading bacteria. Bioresour. Technol. 2011, 102, 2206–2212. [Google Scholar] [CrossRef] [PubMed]
  137. Bai, N.; Wang, S.; Sun, P.; Abuduaini, R.; Zhu, X.; Zhao, Y. Degradation of nonylphenol polyethoxylates by functionalized Fe3O4 nanoparticle-immobilized Sphingomonas sp. Y2. Sci. Total Environ. 2018, 615, 462–468. [Google Scholar] [CrossRef] [PubMed]
  138. Cirja, M.; Hommes, G.; Ivashechkin, P.; Prell, J.; Schäffer, A.; Corvini, P.F.X.; Lenz, M. Impact of bio-augmentation with Sphingomonas sp. strain TTNP3 in membrane bioreactors degrading nonylphenol. Appl. Microbiol. Biotechnol. 2009, 84, 183–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Murugesan, K.; Bokare, V.; Jeon, J.R.; Kim, E.J.; Kim, J.H.; Chang, Y.S. Effect of Fe-Pd bimetallic nanoparticles on Sphingomonas sp. PH-07 and a nano-bio hybrid process for triclosan degradation. Bioresour. Technol. 2011, 102, 6019–6025. [Google Scholar] [CrossRef]
  140. Wang, Z.; Dai, Y.; Zhao, Q.; Li, N.; Zhou, Q.; Xie, S. Nonylphenol biodegradation, functional gene abundance and bacterial community in bioaugmented sediment: effect of external carbon source. Environ. Sci. Pollut. Res. 2015, 22, 12083–12091. [Google Scholar] [CrossRef]
  141. Wojcieszyńska, D.; Domaradzka, D.; Hupert-Kocurek, K.; Guzik, U. Bacterial degradation of naproxen - Undisclosed pollutant in the environment. J. Environ. Manag. 2014, 145, 157–161. [Google Scholar] [CrossRef]
  142. Pauwels, B.; Wille, K.; Noppe, H.; De Brabander, H.; Van De Wiele, T.; Verstraete, W.; Boon, N. 17 α-ethinylestradiol cometabolism by bacteria degrading estrone, 17β-estradiol and estriol. Biodegradation 2008, 19, 683–693. [Google Scholar] [CrossRef]
  143. Reis, P.J.M.; Reis, A.C.; Ricken, B.; Kolvenbach, B.A.; Manaia, C.M.; Corvini, P.F.X.; Nunes, O.C. Biodegradation of sulfamethoxazole and other sulfonamides by Achromobacter denitrificans PR1. J. Hazard. Mater. 2014, 280, 741–749. [Google Scholar] [CrossRef]
  144. Reis, A.C.; Čvančarová, M.; Liu, Y.; Lenz, M.; Hettich, T.; Kolvenbach, B.A.; Corvini, P.F.X.; Nunes, O.C. Biodegradation of sulfamethoxazole by a bacterial consortium of Achromobacter denitrificans PR1 and Leucobacter sp. GP. Appl. Microbiol. Biotechnol. 2018, 102, 10299–10314. [Google Scholar] [CrossRef] [PubMed]
  145. Deng, Y.; Mao, Y.; Li, B.; Yang, C.; Zhang, T. Aerobic degradation of sulfadiazine by arthrobacter spp.: Kinetics, pathways, and genomic characterization. Environ. Sci. Technol. 2016, 50, 9566–9575. [Google Scholar] [CrossRef] [PubMed]
  146. Nguyen, P.Y.; Silva, A.F.; Reis, A.C.; Nunes, O.C.; Rodrigues, A.M.; Rodrigues, J.E.; Cardoso, V.V.; Benoliel, M.J.; Reis, M.A.M.; Oehmen, A.; et al. Bioaugmentation of membrane bioreactor with Achromobacter denitrificans strain PR1 for enhanced sulfamethoxazole removal in wastewater. Sci. Total Environ. 2019, 648, 44–55. [Google Scholar] [CrossRef]
  147. Navrozidou, E.; Melidis, P.; Ntougias, S. Biodegradation aspects of ibuprofen and identification of ibuprofen-degrading microbiota in an immobilized cell bioreactor. Environ. Sci. Pollut. Res. 2019, 26, 14238–14249. [Google Scholar] [CrossRef] [PubMed]
  148. Zhang, H.; Wang, L.; Li, Y.; Wang, P.; Wang, C. Background nutrients and bacterial community evolution determine 13 C-17β-estradiol mineralization in lake sediment microcosms. Sci. Total Environ. 2019, 651, 2304–2311. [Google Scholar] [CrossRef]
  149. Ootsuka, M.; Nishizawa, T.; Hasegawa, M.; Kurusu, Y.; Ohta, H. Comparative analysis of the genetic basis of branched nonylphenol degradation by Sphingobium amiense DSM 16289 T and Sphingobium cloacae JCM 10874 T. Microbes Environ. 2018, 33, 450–454. [Google Scholar] [CrossRef]
  150. Watahiki, S.; Kimura, N. Draft genome sequence of a caffeine-utilizing bacterium, Cupriavidus sp. strain D384. Genome Announc. 2017, 5. [Google Scholar] [CrossRef] [Green Version]
  151. Ha, H.; Mahanty, B.; Yoon, S.; Kim, C.G. Degradation of the long-resistant pharmaceutical compounds carbamazepine and diatrizoate using mixed microbial culture. J. Environ. Sci. Heal. Part A Toxic/Hazardous Subst. Environ. Eng. 2016, 51, 467–471. [Google Scholar] [CrossRef]
  152. Lee, D.G.; Chu, K.H. Effects of growth substrate on triclosan biodegradation potential of oxygenase-expressing bacteria. Chemosphere 2013, 93, 1904–1911. [Google Scholar] [CrossRef]
  153. Villemur, R.; Cunha dos Santos, S.C.; Ouellette, J.; Juteau, P.; Lépine, F.; Déziel, E. Biodegradation of endocrine disruptors in solid-liquid two-phase partitioning systems by enrichment cultures. Appl. Environ. Microbiol. 2013, 79, 4701–4711. [Google Scholar] [CrossRef] [Green Version]
  154. Wang, D.; Sui, Q.; Lu, S.G.; Zhao, W.T.; Qiu, Z.F.; Miao, Z.W.; Yu, G. Occurrence and removal of six pharmaceuticals and personal care products in a wastewater treatment plant employing anaerobic/anoxic/aerobic and UV processes in Shanghai, China. Environ. Sci. Pollut. Res. 2014, 21, 4276–4285. [Google Scholar] [CrossRef] [PubMed]
  155. de las Heras, I.; Padrino, B.; Molina, R.; Segura, Y.; Melero, J.A.; Mohedano, A.F.; Martinez, F.; Puyol, D. Efficient treatment of synthetic wastewater contaminated with emerging pollutants by anaerobic purple phototrophic bacteria. Front. Wastewater Treat. Model. 2017, 1, 324–330. [Google Scholar]
  156. Puyol, D.; Batstone, D.J.; Hülsen, T.; Astals, S.; Peces, M.; Krömer, J.O. Resource recovery from wastewater by biological technologies: Opportunities, challenges, and prospects. Front. Microbiol. 2017, 7, 1–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Madukasi, E.I.; Dai, X.; He, C.; Zhou, J. Potentials of phototrophic bacteria in treating pharmaceutical wastewater. Int. J. Environ. Sci. Technol. 2010, 7, 165–174. [Google Scholar] [CrossRef] [Green Version]
  158. Merugu, R.; Prashanthi, Y.; Sarojini, T.; Badgu, N. Bioremediation of waste waters by the anoxygenic photosynthetic bacterium Rhodobacter sphaeroides SMR 009. Int. J. Res. Environ. Sci. Technol. 2014, 4, 16–19. [Google Scholar]
  159. Yang, Z.; Shi, Y.; Zhang, Y.; Cheng, Q.; Li, X.; Zhao, C.; Zhang, D. Different pathways for 4-n-nonylphenol biodegradation by two Aspergillus strains derived from estuary sediment: Evidence from metabolites determination and key-gene identification. J. Hazard. Mater. 2018, 359, 203–212. [Google Scholar] [CrossRef]
  160. Vasiliadou, I.A.; Molina, R.; Pariente, M.I.; Christoforidis, K.C.; Martinez, F.; Melero, J.A. Understanding the role of mediators in the efficiency of advanced oxidation processes using white-rot fungi. Chem. Eng. J. 2019, 359, 1427–1435. [Google Scholar] [CrossRef]
  161. Vasiliadou, I.A.; Sánchez-Vázquez, R.; Molina, R.; Martínez, F.; Melero, J.A.; Bautista, L.F.; Iglesias, J.; Morales, G. Biological removal of pharmaceutical compounds using white-rot fungi with concomitant FAME production of the residual biomass. J. Environ. Manag. 2016, 180, 228–237. [Google Scholar] [CrossRef]
  162. Yang, S.; Hai, F.I.; Nghiem, L.D.; Price, W.E.; Roddick, F.; Moreira, M.T.; Magram, S.F. Understanding the factors controlling the removal of trace organic contaminants by white-rot fungi and their lignin modifying enzymes: A critical review. Bioresour. Technol. 2013, 359, 203–212. [Google Scholar] [CrossRef] [Green Version]
  163. 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]
  164. Olicón-Hernández, D.R.; González-López, J.; Aranda, E. Overview on the biochemical potential of filamentous fungi to degrade pharmaceutical compounds. Front. Microbiol. 2017, 8, 1–7. [Google Scholar] [CrossRef] [PubMed]
  165. 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, 610–611, 1147–1153. [Google Scholar] [CrossRef] [PubMed]
  166. Camacho-Morales, R.L.; Gerardo-Gerardo, J.L.; Guillén Navarro, K.; Sánchez, J.E. Ligninolytic enzyme production by white rot fungi during paraquat (herbicide) degradation. Rev. Argent. Microbiol. 2017, 49, 189–196. [Google Scholar] [PubMed]
  167. Singh, H. Mycoremediation: Fungal Bioremediation; John Wiley & Sons: Hoboken, NJ, USA, 2006. [Google Scholar]
  168. González, T.; Terrón, M.C.; Yagüe, S.; Junca, H.; Carbajo, J.M.; Zapico, E.J.; Silva, R.; Arana-Cuenca, A.; Téllez, A.; González, A.E. Melanoidin-containing wastewaters induce selective laccase gene expression in the white-rot fungus Trametes sp. I-62. Res. Microbiol. 2008, 59, 103–109. [Google Scholar] [CrossRef]
  169. 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, 5200–5210. [Google Scholar] [CrossRef]
  170. Margot, J.; Bennati-Granier, C.; Maillard, J.; Blánquez, P.; Barry, D.A.; Holliger, C. Bacterial versus fungal laccase: Potential for micropollutant degradation. AMB Express 2013, 3, 63. [Google Scholar] [CrossRef] [Green Version]
  171. Maryšková, M.; Ardao, I.; García-González, C.A.; Martinová, L.; Rotková, J.; Ševců, A. Polyamide 6/chitosan nanofibers as support for the immobilization of Trametes versicolor laccase for the elimination of endocrine disrupting chemicals. Enzyme Microb. Technol. 2016, 89, 31–38. [Google Scholar] [CrossRef]
  172. Ardao, I.; Magnin, D.; Agathos, S.N. Bioinspired production of magnetic laccase-biotitania particles for the removal of endocrine disrupting chemicals. Biotechnol. Bioeng. 2015, 112, 1986–1996. [Google Scholar] [CrossRef]
  173. Castellet-Rovira, F.; Lucas, D.; Villagrasa, M.; Rodríguez-Mozaz, S.; Barceló, D.; Sarrà, M. Stropharia rugosoannulata and Gymnopilus luteofolius: Promising fungal species for pharmaceutical biodegradation in contaminated water. J. Environ. Manag. 2018, 207, 396–404. [Google Scholar] [CrossRef]
  174. Bilal, M.; Asgher, M.; Iqbal, H.M.N.; Hu, H.; Zhang, X. Bio-based degradation of emerging endocrine-disrupting and dye-based pollutants using cross-linked enzyme aggregates. Environ. Sci. Pollut. Res. 2017, 24, 7035–7041. [Google Scholar] [CrossRef]
  175. Liao, C.-S.; Yuan, S.-Y.; Hung, B.-H.; Chang, B. Removal of organic toxic chemicals using the spent mushroom compost of Ganoderma lucidum. J. Environ. Monit. 2012, 14, 1983–1988. [Google Scholar] [CrossRef] [PubMed]
  176. Čvančarová, M.; Moeder, M.; Filipová, A.; Reemtsma, T.; Cajthaml, T. Biotransformation of the antibiotic agent flumequine by ligninolytic fungi and residual antibacterial activity of the transformation mixtures. Environ. Sci. Technol. 2013, 47, 14128–14136. [Google Scholar] [CrossRef] [PubMed]
  177. Přenosilová, L.; Křesinová, Z.; Amemori, A.S.; Cajthaml, T.; Svobodová, K. Transcriptional response of lignin-degrading enzymes to 17 α-ethinyloestradiol in two white rots. Microb. Biotechnol. 2013, 6, 300–306. [Google Scholar] [CrossRef] [PubMed]
  178. Loffredo, E.; Castellana, G.; Taskin, E. A Two-step approach to eliminate pesticides and estrogens from a wastewater and reduce its phytotoxicity: Adsorption onto plant-derived materials and fungal degradation. Water Air Soil Pollut. 2016, 227. [Google Scholar] [CrossRef]
  179. 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]
  180. Moon, D.S.; Song, H.G. Degradation of alkylphenols by white rot fungus Irpex lacteus and its manganese peroxidase. Appl. Biochem. Biotechnol. 2012, 168, 542–549. [Google Scholar] [CrossRef]
  181. Parra Guardado, A.L.; Belleville, M.P.; de Jesús Rostro Alanis, M.; Parra Saldivar, R.; Sanchez-Marcano, J. Effect of redox mediators in pharmaceuticals degradation by laccase: A comparative study. Process. Biochem. 2019, 78, 123–131. [Google Scholar] [CrossRef]
  182. Golveia, J.C.S.; Santiago, M.F.; Sales, P.T.F.; Sartoratto, A.; Ponezi, A.N.; Thomaz, D.V.; de Gil, E.S.; Maria, M.T. Cupuaçu (Theobroma grandiflorum) residue and its potential application in the bioremediation of 17-A-ethinylestradiol as a Pycnoporus sanguineus laccase inducer. Prep. Biochem. Biotechnol. 2018, 48, 541–548. [Google Scholar] [CrossRef]
  183. Liu, J.; Luo, Q.; Huang, Q. Removal of 17 β-estradiol from poultry litter via solid state cultivation of lignolytic fungi. J. Clean. Prod. 2016, 139, 1400–1407. [Google Scholar] [CrossRef]
  184. Garcia-Morales, R.; Rodríguez-Delgado, M.; Gomez-Mariscal, K.; Orona-Navar, C.; Hernandez-Luna, C.; Torres, E.; Parra, R.; Cárdenas-Chávez, D.; Mahlknecht, J.; Ornelas-Soto, N. Biotransformation of endocrine-disrupting compounds in groundwater: Bisphenol A, nonylphenol, ethynylestradiol and triclosan by a laccase cocktail from Pycnoporus sanguineus CS43. Water Air Soil Pollut. 2015, 226, 251. [Google Scholar] [CrossRef] [Green Version]
  185. Nair, R.R.; Demarche, P.; Agathos, S.N. Formulation and characterization of an immobilized laccase biocatalyst and its application to eliminate organic micropollutants in wastewater. (Special Issue: Biotechnology for the bio and green economy.). N. Biotechnol. 2013, 30, 814–823. [Google Scholar] [CrossRef] [PubMed]
  186. Torres-Duarte, C.; Viana, M.T.; Vazquez-Duhalt, R. Laccase-mediated transformations of endocrine disrupting chemicals abolish binding affinities to estrogen receptors and their estrogenic activity in zebrafish. Appl. Biochem. Biotechnol. 2012, 168, 864–876. [Google Scholar] [CrossRef] [PubMed]
  187. Chen, X.; Zhuang, J.; Bester, K. Degradation of triclosan by environmental microbial consortia and by axenic cultures of microorganisms with concerns to wastewater treatment. Appl. Microbiol. Biotechnol. 2018, 102, 5403–5417. [Google Scholar] [CrossRef] [PubMed]
  188. Bronikowski, A.; Hagedoorn, P.L.; Koschorreck, K.; Urlacher, V.B. Expression of a new laccase from Moniliophthora roreri at high levels in Pichia pastoris and its potential application in micropollutant degradation. AMB Express 2017, 7, 73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Gauthier, H.; Yargeau, V.; Cooper, D.G. Biodegradation of pharmaceuticals by Rhodococcus rhodochrous and Aspergillus niger by co-metabolism. Sci. Total Environ. 2010, 408, 1701–1706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Zhou, B.; Ma, C.; Wang, H.; Xia, T. Biodegradation of caffeine by whole cells of tea-derived fungi Aspergillus sydowii, Aspergillus niger and optimization for caffeine degradation. BMC Microbiol. 2018, 18, 1–10. [Google Scholar] [CrossRef] [PubMed]
  191. Aracagök, Y.D.; Göker, H.; Cihangir, N. Biodegradation of diclofenac with fungal strains. Arch. Environ. Prot. 2018, 44, 55–62. [Google Scholar]
  192. Ertit Taştan, B.; Dönmez, G. Biodegradation of pesticide triclosan by A. versicolor in simulated wastewater and semi-synthetic media. Pestic. Biochem. Physiol. 2015, 118, 33–37. [Google Scholar] [CrossRef]
  193. Pai, P.V.; Pai, A.; Pai, S.; Devadiga, S.Y.; Nayak, V.; Rao, C. V Effect of glucose and nitrogen source on caffeine degradation by four filamentous fungi. Indian J. Biotechnol. 2013, 12, 432–434. [Google Scholar]
  194. Hussain, J.; Muhammad, Z.; Ullah, R.; Jamila, N.; Ahmad, S.; Khan, N.; Ayaz, S.; Haider, S. Biotransformation of ß-estradiol isolated from Sonchus eruca. African J. Biotechnol. 2011, 10, 5529–5533. [Google Scholar]
  195. Rodríguez, E.; Ruiz-Dueñas, F.J.; Kooistra, R.; Ram, A.; Martínez, Á.T.; Martínez, M.J. Isolation of two laccase genes from the white-rot fungus Pleurotus eryngii and heterologous expression of the pel3 encoded protein. J. Biotechnol. 2008, 134, 9–19. [Google Scholar] [CrossRef] [PubMed]
  196. Nguyen, L.N.; van de Merwe, J.P.; Hai, F.I.; Leusch, F.D.L.; Kang, J.; Price, W.E.; Roddick, F.; Magram, S.F.; Nghiem, L.D. Laccase-syringaldehyde-mediated degradation of trace organic contaminants in an enzymatic membrane reactor: Removal efficiency and effluent toxicity. Bioresour. Technol. 2016, 200, 477–484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. 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] [PubMed]
  198. Tian, H.; Ma, Y.J.; Li, W.Y.; Wang, J.W. Efficient degradation of triclosan by an endophytic fungus Penicillium oxalicum B4. Environ. Sci. Pollut. Res. 2018, 25, 8963–8975. [Google Scholar] [CrossRef]
  199. Kuzikova, I.; Safronova, V.; Zaytseva, T.; Medvedeva, N. Fate and effects of nonylphenol in the filamentous fungus Penicillium expansum isolated from the bottom sediments of the Gulf of Finland. J. Mar. Syst. 2017, 171, 111–119. [Google Scholar] [CrossRef]
  200. Zhang, D.; Luo, J.; Lee, Z.M.P.; Gersberg, R.M.; Liu, Y.; Tan, S.K.; Ng, W.J. Characterization of microbial communities in wetland mesocosms receiving caffeine-enriched wastewater. Environ. Sci. Pollut. Res. 2016, 23, 14526–14539. [Google Scholar] [CrossRef] [Green Version]
  201. Tastan, B.E.; Özdemir, C.; Tekinay, T. Effects of different culture media on biodegradation of triclosan by Rhodotorula mucilaginosa and Penicillium sp. Water Sci. Technol. 2016, 74, 473–481. [Google Scholar] [CrossRef] [Green Version]
  202. Shan, L.; Li, Y.; Chen, Y.; Yin, M.; Huang, J.; Zhang, Z.; Shi, X.; Liu, H. Microbial hydroxylation of 17β-estradiol by Penicillium brevicompactum. Biocatal. Biotransform. 2016, 34, 137–143. [Google Scholar] [CrossRef]
  203. Aracagök, Y.D.; Göker, H.; Cihangir, N. Biodegradation of micropollutant naproxen with a selected fungal strain and identification of metabolites. Zeitschrift fur Naturforsch. 2017, 72, 173–179. [Google Scholar] [CrossRef]
  204. Chatterjee, A.; Abraham, J. Mycoremediation of 17β-estradiol using Trichoderma citrinoviride strain AJAC3 along with enzyme studies. Environ. Prog. Sustain. Energy 2019, 8, 13142. [Google Scholar] [CrossRef]
  205. Buchicchio, A.; Bianco, G.; Sofo, A.; Masi, S.; Caniani, D. Biodegradation of carbamazepine and clarithromycin by Trichoderma harzianum and Pleurotus ostreatus investigated by liquid chromatography - high-resolution tandem mass spectrometry (FTICR MS-IRMPD). Sci. Total Environ. 2016, 557, 733–739. [Google Scholar] [CrossRef]
  206. Lloret, L.; Eibes, G.; Feijoo, G.; Moreira, M.T.; Lema, J.M. Degradation of estrogens by laccase from Myceliophthora thermophila in fed-batch and enzymatic membrane reactors. J. Hazard. Mater. 2012, 213, 175–183. [Google Scholar] [CrossRef]
  207. Lloret, L.; Eibes, G.; Lú-Chau, T.A.; Moreira, M.T.; Feijoo, G.; Lema, J.M. Laccase-catalyzed degradation of anti-inflammatories and estrogens. Biochem. Eng. J. 2010, 51, 124–131. [Google Scholar] [CrossRef]
  208. García-Delgado, C.; Eymar, E.; Camacho-Arévalo, R.; Petruccioli, M.; Crognale, S.; D’Annibale, A. Degradation of tetracyclines and sulfonamides by stevensite- and biochar-immobilized laccase systems and impact on residual antibiotic activity. J. Chem. Technol. Biotechnol. 2018, 93, 3394–3409. [Google Scholar] [CrossRef]
  209. González-Abradelo, D.; Pérez-Llano, Y.; Peidro-Guzmán, H.; Sánchez-Carbente, M.D.R.; Folch-Mallol, J.L.; Aranda, E.; Vaidyanathan, V.K.; Cabana, H.; Gunde-Cimerman, N.; Batista-García, R.A. First demonstration that ascomycetous halophilic fungi (Aspergillus sydowii and Aspergillus destruens) are useful in xenobiotic mycoremediation under high salinity conditions. Bioresour. Technol. 2019, 279, 287–296. [Google Scholar]
  210. Parshikov, I.A.; Muraleedharan, K.M.; Avery, M.A.; Williamson, J.S. Transformation of artemisinin by Cunninghamella elegans. Appl. Microbiol. Biotechnol. 2004, 64, 782–786. [Google Scholar] [CrossRef]
  211. Kang, S.I.; Kang, S.Y.; Kanaly, R.A.; Lee, E.; Lim, Y.; Hur, H.G. Rapid oxidation of ring methyl groups is the primary mechanism of biotransformation of gemfibrozil by the fungus Cunninghamella elegans. Arch. Microbiol. 2009, 191, 509–517. [Google Scholar] [CrossRef]
  212. Zhong, D.F.; Sun, L.; Liu, L.; Huang, H.H. Microbial transformation of naproxen by Cunninghamella species. Acta Pharmacol. Sin. 2003, 24, 442–447. [Google Scholar]
  213. Williams, A.J.; Deck, J.; Freeman, J.P.; Paul Chiarelli, M.; Adjei, M.D.; Heinze, T.M.; Sutherland, J.B. Biotransformation of flumequine by the fungus Cunninghamella elegans. Chemosphere 2007, 67, 240–243. [Google Scholar] [CrossRef]
  214. Goto, M.; Noda, S.; Kamiya, N.; Nakashio, F. Enzymatic resolution of racemic ibuprofen by surfactant-coated lipases in organic media. Biotechnol. Lett. 1996, 18, 839–844. [Google Scholar] [CrossRef]
  215. Djelal, H.; Amrane, A. Biodegradation by bioaugmentation of dairy wastewater by fungal consortium on a bioreactor lab-scale and on a pilot-scale. J. Environ. Sci. 2013, 25, 1906–1912. [Google Scholar] [CrossRef] [Green Version]
  216. Shi, J.H.; Suzuki, Y.; Lee, B.D.; Nakai, S.; Hosomi, M. Isolation and characterization of the ethynylestradiol-biodegrading microorganism Fusarium proliferatum strain HNS-1. Water Sci. Technol. 2002, 45, 175–179. [Google Scholar] [CrossRef] [PubMed]
  217. Dubroca, J.; Brault, A.; Kollmann, A.; Touton, I.; Jolivalt, C.; Kerhoas, L.; Mougin, C. Biotransformation of nonylphenol surfactants in soils amended with contaminated sewage sludges. In Environmental Chemistry: Green Chemistry and Pollutants in Ecosystems; Springer-Verlag: Berlin, Germany, 2005; ISBN 3540228608. [Google Scholar]
  218. Janicki, T.; Długoński, J.; Krupiński, M. Detoxification and simultaneous removal of phenolic xenobiotics and heavy metals with endocrine-disrupting activity by the non-ligninolytic fungus Umbelopsis isabellina. J. Hazard. Mater. 2018, 360, 661–669. [Google Scholar] [CrossRef]
  219. Janicki, T.; Krupiński, M.; Długoński, J. Degradation and toxicity reduction of the endocrine disruptors nonylphenol, 4-tert-octylphenol and 4-cumylphenol by the non-ligninolytic fungus Umbelopsis isabellina. Bioresour. Technol. 2016, 200, 223–229. [Google Scholar] [CrossRef] [PubMed]
  220. Czarny, K.; Szczukocki, D.; Krawczyk, B.; Skrzypek, S.; Zieliński, M.; Gadzała-Kopciuch, R. Toxic effects of single animal hormones and their mixtures on the growth of Chlorella vulgaris and Scenedesmus armatus. Chemosphere 2019, 224, 93–102. [Google Scholar] [CrossRef]
  221. Gosset, A.; Durrieu, C.; Barbe, P.; Bazin, C.; Bayard, R. Microalgal whole-cell biomarkers as sensitive tools for fast toxicity and pollution monitoring of urban wet weather discharges. Chemosphere 2019, 217, 522–533. [Google Scholar] [CrossRef]
  222. Gojkovic, Z.; Lindberg, R.H.; Tysklind, M.; Funk, C. Northern green algae have the capacity to remove active pharmaceutical ingredients. Ecotoxicol. Environ. Saf. 2019, 170, 644–656. [Google Scholar] [CrossRef]
  223. Xiong, J.Q.; Kurade, M.B.; Abou-Shanab, R.A.I.; Ji, M.K.; Choi, J.; Kim, J.O.; Jeon, B.H. Biodegradation of carbamazepine using freshwater microalgae Chlamydomonas mexicana and Scenedesmus obliquus and the determination of its metabolic fate. Bioresour. Technol. 2016, 205, 183–190. [Google Scholar] [CrossRef]
  224. Ding, T.; Yang, M.; Zhang, J.; Yang, B.; Lin, K.; Li, J.; Gan, J. Toxicity, degradation and metabolic fate of ibuprofen on freshwater diatom Navicula sp. J. Hazard. Mater. 2017, 330, 127–134. [Google Scholar] [CrossRef]
  225. Wu, N.; Moreira, C.M.; Zhang, Y.; Doan, N.; Yang, S.; Phlips, E.J.; Svoronos, S.A.; Pullammanappallil, P.C. Techno-Economic Analysis of Biogas Production from Microalgae through Anaerobic Digestion. In Anaerobic Digestion; IntechOpen Ltd.: London, UK, 2019; ISBN 9781838818500. [Google Scholar]
  226. Shomal, R.; Hisham, H.; Mlhem, A.; Hassan, R.; Al-Zuhair, S. Simultaneous extraction–reaction process for biodiesel production from microalgae. Energy Rep. 2019, 5, 37–40. [Google Scholar] [CrossRef]
  227. Lozano-Garcia, D.F.; Cuellar-Bermudez, S.P.; del Rio-Hinojosa, E.; Betancourt, F.; Aleman-Nava, G.S.; Parra-Saldivar, R. Potential land microalgae cultivation in Mexico: From food production to biofuels. Algal Res. 2019, 39, 101459. [Google Scholar] [CrossRef]
  228. de Souza, M.H.B.; Calijuri, M.L.; Assemany, P.P.; de Siqueira Castro, J.; de Oliveira, A.C.M. Soil application of microalgae for nitrogen recovery: A life-cycle approach. J. Clean. Prod. 2019, 211, 342–349. [Google Scholar] [CrossRef]
  229. Cuellar-Bermudez, S.P.; Aleman-Nava, G.S.; Chandra, R.; Garcia-Perez, J.S.; Contreras-Angulo, J.R.; Markou, G.; Muylaert, K.; Rittmann, B.E.; Parra-Saldivar, R. Nutrients utilization and contaminants removal. A review of two approaches of algae and cyanobacteria in wastewater. Algal Res. 2017, 24, 438–449. [Google Scholar] [CrossRef]
  230. Wang, S.; Poon, K.; Cai, Z. Removal and metabolism of triclosan by three different microalgal species in aquatic environment. J. Hazard. Mater. 2018, 342, 643–650. [Google Scholar] [CrossRef]
  231. Gao, Q.T.; Wong, Y.S.; Tam, N.F.Y. Antioxidant responses of different microalgal species to nonylphenol-induced oxidative stress. J. Appl. Phycol. 2017, 29, 1317–1329. [Google Scholar] [CrossRef]
  232. de Siqueira Castro, J.; Calijuri, M.L.; Mattiello, E.M.; Ribeiro, V.J.; Assemany, P.P. Algal biomass from wastewater: Soil phosphorus bioavailability and plants productivity. Sci. Total Environ. 2020, 711, 135088. [Google Scholar] [CrossRef]
  233. Abd El-Hakim, Y.M.; Mohamed, W.A.; El-Metwally, A.E. Spirulina platensis attenuates furan reprotoxicity by regulating oxidative stress, inflammation, and apoptosis in testis of rats. Ecotoxicol. Environ. Saf. 2018, 161, 25–33. [Google Scholar] [CrossRef]
  234. Hamjinda, N.S.; Chiemchaisri, W.; Watanabe, T.; Honda, R.; Chiemchaisri, C. Toxicological assessment of hospital wastewater in different treatment processes. Environ. Sci. Pollut. Res. 2018, 25, 7271–7279. [Google Scholar] [CrossRef]
  235. Escapa, C.; Coimbra, R.N.; Paniagua, S.; García, A.I.; Otero, M. Paracetamol and salicylic acid removal from contaminated water by microalgae. J. Environ. Manag. 2017, 203, 799–806. [Google Scholar] [CrossRef]
  236. López-Serna, R.; Posadas, E.; García-Encina, P.A.; Muñoz, R. Removal of contaminants of emerging concern from urban wastewater in novel algal-bacterial photobioreactors. Sci. Total Environ. 2019, 662, 32–40. [Google Scholar] [CrossRef] [Green Version]
  237. de Godos, I.; Muñoz, R.; Guieysse, B. Tetracycline removal during wastewater treatment in high-rate algal ponds. J. Hazard. Mater. 2012, 229, 446–449. [Google Scholar] [CrossRef] [PubMed]
  238. Xiong, J.Q.; Kurade, M.B.; Jeon, B.H. Can microalgae remove pharmaceutical contaminants from water? Trends Biotechnol. 2018, 36, 30–44. [Google Scholar] [CrossRef] [PubMed]
  239. Xiong, J.Q.; Kim, S.J.; Kurade, M.B.; Govindwar, S.; Abou-Shanab, R.A.I.; Kim, J.R.; Roh, H.S.; Khan, M.A.; Jeon, B.H. Combined effects of sulfamethazine and sulfamethoxazole on a freshwater microalga, Scenedesmus obliquus: Toxicity, biodegradation, and metabolic fate. J. Hazard. Mater. 2019, 370, 138–146. [Google Scholar] [CrossRef] [PubMed]
  240. Xiong, J.Q.; Govindwar, S.; Kurade, M.B.; Paeng, K.J.; Roh, H.S.; Khan, M.A.; Jeon, B.H. Toxicity of sulfamethazine and sulfamethoxazole and their removal by a green microalga, Scenedesmus obliquus. Chemosphere 2019, 218, 551–558. [Google Scholar] [CrossRef]
  241. Maes, H.M.; Maletz, S.X.; Ratte, H.T.; Hollender, J.; Schaeffer, A. Uptake, elimination, and biotransformation of 17 α-ethinylestradiol by the freshwater alga Desmodesmus subspicatus. Environ. Sci. Technol. 2014, 48, 12354–12361. [Google Scholar] [CrossRef]
  242. Escapa, C.; Torres, T.; Neuparth, T.; Coimbra, R.N.; García, A.I.; Santos, M.M.; Otero, M. Zebrafish embryo bioassays for a comprehensive evaluation of microalgae efficiency in the removal of diclofenac from water. Sci. Total Environ. 2018, 1024–1033. [Google Scholar] [CrossRef]
  243. Coimbra, R.N.; Escapa, C.; Vázquez, N.C.; Noriega-Hevia, G.; Otero, M. Utilization of non-living microalgae biomass from two different strains for the adsorptive removal of diclofenac from water. Water 2018, 10, 1401. [Google Scholar] [CrossRef] [Green Version]
  244. Ali, M.E.M.; Abd El-Aty, A.M.; Badawy, M.I.; Ali, R.K. Removal of pharmaceutical pollutants from synthetic wastewater using chemically modified biomass of green alga Scenedesmus obliquus. Ecotoxicol. Environ. Saf. 2018, 151, 144–152. [Google Scholar] [CrossRef]
  245. Gao, T.; Shi, X. Preparation of a synthetic seed for the common reed harboring an endophytic bacterium promoting seedling growth under cadmium stress. Environ. Sci. Pollut. Res. 2018, 25, 8871–8879. [Google Scholar] [CrossRef]
  246. Li, X.; Xu, Q.M.; Cheng, J.S.; Yuan, Y.J. Improving the bioremoval of sulfamethoxazole and alleviating cytotoxicity of its biotransformation by laccase producing system under coculture of Pycnoporus sanguineus and Alcaligenes faecalis. Bioresour. Technol. 2016, 220, 333–340. [Google Scholar] [CrossRef]
  247. Cabana, H.; Jones, J.P.; Agathos, S.N. Elimination of endocrine disrupting chemicals using white rot fungi and their lignin modifying enzymes: A review. Eng. Life Sci. 2007, 7, 429–456. [Google Scholar] [CrossRef]
  248. Hai, F.I.; Modin, O.; Yamamoto, K.; Fukushi, K.; Nakajima, F.; Nghiem, L.D. Pesticide removal by a mixed culture of bacteria and white-rot fungi. J. Taiwan Inst. Chem. Eng. 2012, 43, 459–462. [Google Scholar] [CrossRef] [Green Version]
  249. Bodin, H.; Daneshvar, A.; Gros, M.; Hultberg, M. Effects of biopellets composed of microalgae and fungi on pharmaceuticals present at environmentally relevant levels in water. Ecol. Eng. 2016, 91, 169–172. [Google Scholar] [CrossRef]
  250. Hom-Diaz, A.; Jaén-Gil, A.; Bello-Laserna, I.; Rodríguez-Mozaz, S.; Vicent, T.; Barceló, D.; Blánquez, P. Performance of a microalgal photobioreactor treating toilet wastewater: Pharmaceutically active compound removal and biomass harvesting. Sci. Total Environ. 2017, 91, 169–172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Wu, X.; Ernst, F.; Conkle, J.L.; Gan, J. Comparative uptake and translocation of pharmaceutical and personal care products (PPCPs) by common vegetables. Environ. Int. 2013, 60, 15–22. [Google Scholar] [CrossRef] [PubMed]
  252. Treesubsuntorn, C.; Thiravetyan, P. Botanical biofilter for indoor toluene removal and reduction of carbon dioxide emission under low light intensity by using mixed C3 and CAM plants. J. Clean. Prod. 2018, 194, 94–100. [Google Scholar] [CrossRef]
  253. Treesubsuntorn, C.; Dolphen, R.; Dhurakit, P.; Siswanto, D.; Thiravetyan, P. Green Technology Innovation in a Developing Country. 2017. Available online: https://aip.scitation.org/doi/abs/10.1063/1.5012704 (accessed on 29 November 2017).
  254. Bhatnagar, N.; Bhatnagar, N. Phytoremediation: A Promising Technology of Environmental Pollution Control. Glob. J. Microbiol. Biotechnol. 2016, 4, 39–42. [Google Scholar]
  255. Kouki, S.; Saidi, N.; M’hiri, F.; Hafiane, A.; Hassen, A. Co-Composting of Macrophyte Biomass and Sludge as an Alternative for Sustainable Management of Constructed Wetland By-Products. Clean Soil Air Water 2016, 44, 694–702. [Google Scholar] [CrossRef]
  256. Sánchez-Galván, G.; Bolaños-Santiago, Y. Phytofiltration of anaerobically digested sugarcane ethanol stillage using a macrophyte with high potential for biofuel production. Int. J. Phytoremediation 2018, 20, 805–812. [Google Scholar] [CrossRef]
  257. da Luz, J.M.R.; Paes, S.A.; Torres, D.P.; Nunes, M.D.; da Silva, J.S.; Mantovani, H.C.; Kasuya, M.C.M. Production of edible mushroom and degradation of antinutritional factors in jatropha biodiesel residues. LWT Food Sci. Technol. 2013, 50, 575–580. [Google Scholar] [CrossRef] [Green Version]
  258. Hultberg, M.; Prade, T.; Bodin, H.; Vidakovic, A.; Asp, H. Adding benefit to wetlands – Valorization of harvested common reed through mushroom production. Sci. Total Environ. 2018, 637, 1395–1399. [Google Scholar] [CrossRef] [PubMed]
  259. Calheiros, C.S.C.; Bessa, V.S.; Mesquita, R.B.R.; Brix, H.; Rangel, A.O.S.S.; Castro, P.M.L. Constructed wetland with a polyculture of ornamental plants for wastewater treatment at a rural tourism facility. Ecol. Eng. 2015, 79, 1–7. [Google Scholar] [CrossRef]
  260. Macci, C.; Peruzzi, E.; Doni, S.; Iannelli, R.; Masciandaro, G. Ornamental plants for micropollutant removal in wetland systems. Environ. Sci. Pollut. Res. 2015, 22, 2406–2415. [Google Scholar] [CrossRef]
  261. Salamanca, E.J.P.; Madera-Parra, C.A.; Avila-Williams, C.A.; Rengifo-Gallego, A.L.; Ríos, D.A. Phytoremediation using terrestrial plants. Phytoremediation 2015, 305–319. [Google Scholar]
  262. Sopajarn, A.; Sangwichien, C. Optimization of enzymatic saccharification of alkali pretreated Typha angustifolia for Glucose Production. Int. J. Chem. Eng. Appl. 2015, 6, 232–236. [Google Scholar] [CrossRef] [Green Version]
  263. Emhofer, L.; Himmelsbach, M.; Buchberger, W.; Klampfl, C.W. High-performance liquid chromatography – mass spectrometry analysis of the parent drugs and their metabolites in extracts from cress (Lepidium sativum) grown hydroponically in water containing four non-steroidal anti-inflammatory drugs. J. Chromatogr. A 2017, 1491, 137–144. [Google Scholar] [CrossRef]
  264. Emhofer, L.; Himmelsbach, M.; Buchberger, W.; Klampfl, C.W. Insights into the uptake, metabolization, and translocation of four non-steroidal anti-inflammatory drugs in cress (Lepidium sativum) by HPLC-MS2. Electrophoresis 2018, 39, 1294–1300. [Google Scholar] [CrossRef]
  265. Klampfl, C.W. Metabolization of pharmaceuticals by plants after uptake from water and soil: A review. TrAC Trends Anal. Chem. 2019, 111, 13–26. [Google Scholar] [CrossRef]
  266. Thijs, S.; Sillen, W.; Rineau, F.; Weyens, N.; Vangronsveld, J. Towards an enhanced understanding of plant-microbiome interactions to improve phytoremediation: Engineering the metaorganism. Front. Microbiol. 2016, 7, 416. [Google Scholar] [CrossRef]
  267. Mallick, I.; Bhattacharyya, C.; Mukherji, S.; Dey, D.; Sarkar, S.C.; Mukhopadhyay, U.K.; Ghosh, A. Effective rhizoinoculation and biofilm formation by arsenic immobilizing halophilic plant growth promoting bacteria (PGPB) isolated from mangrove rhizosphere: A step towards arsenic rhizoremediation. Sci. Total Environ. 2018, 610, 1239–1250. [Google Scholar] [CrossRef] [PubMed]
  268. Barac, T.; Taghavi, S.; Borremans, B.; Provoost, A.; Oeyen, L.; Colpaert, J.V.; Vangronsveld, J.; Van Der Lelie, D. Engineered endophytic bacteria improve phytoremediation of water-soluble, volatile, organic pollutants. Nat. Biotechnol. 2004, 22, 583–588. [Google Scholar] [CrossRef] [PubMed]
  269. Afzal, M.; Khan, Q.M.; Sessitsch, A. Endophytic bacteria: Prospects and applications for the phytoremediation of organic pollutants. Chemosphere 2014, 17, 232–242. [Google Scholar] [CrossRef] [PubMed]
  270. Feng, N.X.; Yu, J.; Zhao, H.M.; Cheng, Y.T.; Mo, C.H.; Cai, Q.Y.; Li, Y.W.; Li, H.; Wong, M.H. Efficient phytoremediation of organic contaminants in soils using plant–endophyte partnerships. Sci. Total Environ. 2017, 583, 352–368. [Google Scholar] [CrossRef] [PubMed]
  271. Dolphen, R.; Thiravetyan, P. Reducing arsenic in rice grains by leonardite and arsenic–resistant endophytic bacteria. Chemosphere 2019, 223, 448–454. [Google Scholar] [CrossRef]
  272. Ma, Y.; Oliveira, R.S.; Freitas, H.; Zhang, C. Biochemical and molecular mechanisms of plant-microbe-metal interactions: Relevance for phytoremediation. Front. Plant Sci. 2016, 7, 753. [Google Scholar] [CrossRef]
  273. Molina, M.C.; Bautista, L.F.; Belda, I.; Carmona, M.; Díaz, E.; Durante-Rodríguez, G.; García-Salgado, S.; López, J.; Martínez-Hidalgo, P.; Quijano, M.A.; et al. Bioremediation of Soil Contaminated with Arsenic; Ashok, K., Swati, S., Eds.; Springer: Singapore, 2019; pp. 321–351. [Google Scholar]
  274. González, P.S.; Ontañon, O.M.; Armendariz, A.L.; Talano, M.A.; Paisio, C.E.; Agostini, E. Brassica napus hairy roots and rhizobacteria for phenolic compounds removal. Environ. Sci. Pollut. Res. 2013, 20, 1310–1317. [Google Scholar] [CrossRef]
  275. Syranidou, E.; Christofilopoulos, S.; Politi, M.; Weyens, N.; Venieri, D.; Vangronsveld, J.; Kalogerakis, N. Bisphenol-A removal by the halophyte Juncus acutus in a phytoremediation pilot: Characterization and potential role of the endophytic community. J. Hazard. Mater. 2017, 323, 350–358. [Google Scholar] [CrossRef] [Green Version]
  276. Syranidou, E.; Christofilopoulos, S.; Kalogerakis, N. Juncus spp.—The helophyte for all (phyto)remediation purposes? N. Biotechnol. 2017, 38, 43–55. [Google Scholar] [CrossRef]
  277. Toyama, T.; Murashita, M.; Kobayashi, K.; Kikuchi, S.; Sei, K.; Tanaka, Y.; Ike, M.; Mori, K. Acceleration of nonylphenol and 4-tert-octylphenol degradation in sediment by phragmites australis and associated rhizosphere bacteria. Environ. Sci. Technol. 2011, 45, 6524–6530. [Google Scholar] [CrossRef]
  278. Toyama, T.; Ojima, T.; Tanaka, Y.; Mori, K.; Morikawa, M. Sustainable biodegradation of phenolic endocrine-disrupting chemicals by Phragmites australis-rhizosphere bacteria association. Water Sci. Technol. 2013, 68, 522–529. [Google Scholar] [CrossRef]
  279. Gonda, S.; Kiss-Szikszai, A.; Szucs, Z.; Balla, B.; Vasas, G. Efficient biotransformation of non-steroid anti-inflammatory drugs by endophytic and epiphytic fungi from dried leaves of a medicinal plant, Plantago lanceolata L. Int. Biodeterior. Biodegrad. 2016, 108, 115–121. [Google Scholar] [CrossRef]
  280. Hurtado, C.; Domínguez, C.; Pérez-Babace, L.; Cañameras, N.; Comas, J.; Bayona, J.M. Estimate of uptake and translocation of emerging organic contaminants from irrigation water concentration in lettuce grown under controlled conditions. J. Hazard. Mater. 2016, 305, 139–148. [Google Scholar] [CrossRef] [PubMed]
  281. Sauvêtre, A.; May, R.; Harpaintner, R.; Poschenrieder, C.; Schröder, P. Metabolism of carbamazepine in plant roots and endophytic rhizobacteria isolated from Phragmites australis. J. Hazard. Mater. 2018, 342, 85–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Sauvêtre, A.; Schröder, P. Uptake of carbamazepine by rhizomes and endophytic bacteria of Phragmites australis. Front. Plant Sci. 2015, 6, 1–11. [Google Scholar] [CrossRef]
  283. Nguyen, L.N.; Hai, F.I.; Yang, S.; Kang, J.; Leusch, F.D.L.; Roddick, F.; Price, W.E.; Nghiem, L.D. Removal of trace organic contaminants by an MBR comprising a mixed culture of bacteria and white-rot fungi. Bioresour. Technol. 2013, 148, 234–241. [Google Scholar] [CrossRef] [Green Version]
  284. Gallardo-Altamirano, M.J.; Maza-Márquez, P.; Peña-Herrera, J.M.; Rodelas, B.; Osorio, F.; Pozo, C. Removal of anti-inflammatory/analgesic pharmaceuticals from urban wastewater in a pilot-scale A 2 O system: Linking performance and microbial population dynamics to operating variables. Sci. Total Environ. 2018, 643, 1481–1492. [Google Scholar] [CrossRef]
  285. Deng, Y.; Wang, Y.; Mao, Y.; Zhang, T. Partnership of Arthrobacter and Pimelobacter in aerobic degradation of sulfadiazine revealed by metagenomics analysis and isolation. Environ. Sci. Technol. 2018, 52, 2963–2972. [Google Scholar] [CrossRef]
  286. Carballa, M.; Omil, F.; Lema, J.M.; Llompart, M.; García-Jares, C.; Rodríguez, I.; Gómez, M.; Ternes, T. Behavior of pharmaceuticals, cosmetics and hormones in a sewage treatment plant. Water Res. 2004, 38, 2918–2926. [Google Scholar] [CrossRef]
  287. Castiglioni, S.; Bagnati, R.; Fanelli, R.; Pomati, F.; Calamari, D.; Zuccato, E. Removal of pharmaceuticals in sewage treatment plants in Italy. Environ. Sci. Technol. 2006, 40, 357–363. [Google Scholar] [CrossRef]
  288. Kasprzyk-Hordern, B.; Dinsdale, R.M.; Guwy, A.J. The removal of pharmaceuticals, personal care products, endocrine disruptors and illicit drugs during wastewater treatment and its impact on the quality of receiving waters. Water Res. 2009, 43, 363–380. [Google Scholar] [CrossRef] [PubMed]
  289. Tiwari, B.; Sellamuthu, B.; Ouarda, Y.; Drogui, P.; Tyagi, R.D.; Buelna, G. Review on fate and mechanism of removal of pharmaceutical pollutants from wastewater using biological approach. Bioresour. Technol. 2017, 224, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  290. Hijosa-Valsero, M.; Sidrach-Cardona, R.; Pedescoll, A.; Sánchez, O.; Bécares, E. Role of Bacterial Diversity on PPCPs Removal in Constructed Wetlands. In Constructed Wetlands for Industrial Wastewater Treatment; Wiley-Blackwell: Hoboken, NJ, USA, 2018; pp. 405–426. ISBN 9781119268345. [Google Scholar]
  291. Boonnorat, J.; Kanyatrakul, A.; Prakhongsak, A.; Honda, R.; Panichnumsin, P.; Boonapatcharoen, N. Effect of hydraulic retention time on micropollutant biodegradation in activated sludge system augmented with acclimatized sludge treating low-micropollutants wastewater. Chemosphere 2019, 230, 606–615. [Google Scholar] [CrossRef] [PubMed]
  292. Hoinkis, J.; Deowan, S.A.; Panten, V.; Figoli, A.; Huang, R.R.; Drioli, E. Membrane bioreactor (MBR) technology - A promising approach for industrial water reuse. Procedia Eng. 2012, 33, 234–241. [Google Scholar] [CrossRef] [Green Version]
  293. Li, C.; Cabassud, C.; Guigui, C. Evaluation of membrane bioreactor on removal of pharmaceutical micropollutants: A review. Desalin. Water Treat. 2015, 55, 845–858. [Google Scholar] [CrossRef]
  294. Radjenović, J.; Petrović, M.; Barceló, D. Fate and distribution of pharmaceuticals in wastewater and sewage sludge of the conventional activated sludge (CAS) and advanced membrane bioreactor (MBR) treatment. Water Res. 2009, 43, 831–841. [Google Scholar] [CrossRef]
  295. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Thomaidis, N.S.; Xu, J. Progress in the biological and chemical treatment technologies for emerging contaminant removal from wastewater: A critical review. J. Hazard. Mater. 2017, 323, 274–298. [Google Scholar] [CrossRef]
  296. Gurung, K.; Ncibi, M.C.; Sillanpää, M. Removal and fate of emerging organic micropollutants (EOMs) in municipal wastewater by a pilot-scale membrane bioreactor (MBR) treatment under varying solid retention times. Sci. Total Environ. 2019, 667, 671–680. [Google Scholar] [CrossRef]
  297. Maeng, S.K.; Choi, B.G.; Lee, K.T.; Song, K.G. Influences of solid retention time, nitrification and microbial activity on the attenuation of pharmaceuticals and estrogens in membrane bioreactors. Water Res. 2013, 47, 3151–3162. [Google Scholar] [CrossRef]
  298. de Melo, R.G.; de Andrade, A.F.; Bezerra, R.P.; Correia, D.S.; de Souza, V.C.; Brasileiro-Vidal, A.C.; de Araújo Viana Marques, D.; Porto, A.L.F. Chlorella vulgaris mixotrophic growth enhanced biomass productivity and reduced toxicity from agro-industrial by-products. Chemosphere 2018, 204, 344–350. [Google Scholar] [CrossRef]
  299. Garcia-Rodríguez, A.; Matamoros, V.; Fontàs, C.; Salvadó, V. The ability of biologically based wastewater treatment systems to remove emerging organic contaminants—a review. Environ. Sci. Pollut. Res. 2014, 21, 11708–11728. [Google Scholar] [CrossRef] [PubMed]
  300. Costa, M.C.M.S.; Carvalho, L.; Leal, C.D.; Dias, M.F.; Martins, K.L.; Garcia, G.B.; Mancuelo, I.D.; Hipólito, T.; Macconell, E.F.A.; Okada, D.; et al. Impact of inocula and operating conditions on the microbial community structure of two anammox reactors. Environ. Technol. 2014, 35, 1811–1822. [Google Scholar] [CrossRef] [PubMed]
  301. Sawant, S.S.; Khadamkar, H.P.; Mathpati, C.S.; Pandit, R.; Lali, A.M. Computational and experimental studies of high depth algal raceway pond photo-bioreactor. Renew. Energy 2018, 118, 152–159. [Google Scholar] [CrossRef]
  302. Gentili, F.G.; Fick, J. Algal cultivation in urban wastewater: An efficient way to reduce pharmaceutical pollutants. J. Appl. Phycol. 2017, 29, 255–262. [Google Scholar] [CrossRef] [Green Version]
  303. Tolboom, S.N.; Carrillo-Nieves, D.; de Jesús Rostro-Alanis, M.; de la Cruz Quiroz, R.; Barceló, D.; Iqbal, H.M.N.; Parra-Saldivar, R. Algal-based removal strategies for hazardous contaminants from the environment – A review. Sci. Total Environ. 2019, 665, 358–366. [Google Scholar] [CrossRef]
  304. Subashchandrabose, S.R.; Ramakrishnan, B.; Megharaj, M.; Venkateswarlu, K.; Naidu, R. Consortia of cyanobacteria/microalgae and bacteria: Biotechnological potential. Biotechnol. Adv. 2011, 29, 896–907. [Google Scholar] [CrossRef]
  305. Borde, X.; Guieysse, B.; Delgado, O.; Muoz, R.; Hatti-Kaul, R.; Nugier-Chauvin, C.; Patin, H.; Mattiasson, B. Synergistic relationships in algal-bacterial microcosms for the treatment of aromatic pollutants. Bioresour. Technol. 2003, 86, 293–300. [Google Scholar] [CrossRef]
  306. Kumar, S.; Dutta, V. Constructed wetland microcosms as sustainable technology for domestic kwastewater treatment: An overview. Environ. Sci. Pollut. Res. Int. 2019, 26, 11662–11673. [Google Scholar] [CrossRef]
  307. Rockström, J.; Steffen, W.; Noone, K.; Persson, Å.; Chapin, F.S.; Lambin, E.F.; Lenton, T.M.; Scheffer, M.; Folke, C.; Schellnhuber, H.J.; et al. A safe operating space for humanity. Nature 2009, 461, 472–475. [Google Scholar] [CrossRef]
  308. Yi, X.; Tran, N.H.; Yin, T.; He, Y.; Gin, K.Y.H. Removal of selected PPCPs, EDCs, and antibiotic resistance genes in landfill leachate by a full-scale constructed wetlands system. Water Res. 2017, 121, 46–60. [Google Scholar] [CrossRef]
  309. Hijosa-Valsero, M.; Matamoros, V.; Martín-Villacorta, J.; Bécares, E.; Bayona, J.M. Assessment of full-scale natural systems for the removal of PPCPs from wastewater in small communities. Water Res. 2010, 44, 1429–1439. [Google Scholar] [CrossRef] [PubMed]
  310. Zhang, D.; Gersberg, R.M.; Ng, W.J.; Tan, S.K. Removal of pharmaceuticals and personal care products in aquatic plant-based systems: A review. Environ. Pollut. 2014, 184, 620–639. [Google Scholar] [CrossRef] [PubMed]
  311. Alexandrino, D.A.M.; Mucha, A.P.; Almeida, C.M.; Gao, W.; Jia, Z.; Carvalho, M.F. Biodegradation of the veterinary antibiotics enrofloxacin and ceftiofur and associated microbial community dynamics. Sci. Total Environ. 2017, 581–582, 359–368. [Google Scholar] [CrossRef]
  312. Harrabi, M.; Alexandrino, D.A.M.; Aloulou, F.; Elleuch, B.; Liu, B.; Jia, Z.; Almeida, C.M.R.; Mucha, A.P.; Carvalho, M.F. Biodegradation of oxytetracycline and enrofloxacin by autochthonous microbial communities from estuarine sediments. Sci. Total Environ. 2019, 648, 962–972. [Google Scholar] [CrossRef] [PubMed]
  313. VanKempen-Fryling, R.J.; Stein, O.R.; Camper, A.K. Presence and persistence of wastewater pathogen Escherichia coli O157:H7 in hydroponic reactors of treatment wetland species. Water Sci. Technol. 2015, 72, 135–140. [Google Scholar] [CrossRef] [PubMed]
  314. Wang, M.; Zhang, D.Q.; Dong, J.W.; Tan, S.K. Constructed wetlands for wastewater treatment in cold climate—A review. J. Environ. Sci. 2017, 57, 293–311. [Google Scholar] [CrossRef]
  315. Maiga, Y.; von Sperling, M.R.; Mihelcic, J. Constructed Wetlands. 2017. Available online: http://www.waterpathogens.org/book/constructed-wetlands (accessed on 22 March 2017).
  316. Salomo, S.; Münch, C.; Röske, I. Evaluation of the metabolic diversity of microbial communities in four different filter layers of a constructed wetland with vertical flow by BiologTM analysis. Water Res. 2009, 43, 4569–4578. [Google Scholar] [CrossRef]
  317. Olson, M.R.; Axler, R.P.; Hicks, R.E. Effects of freezing and storage temperature on MS2 viability. J. Virol. Methods 2004, 122, 147–152. [Google Scholar] [CrossRef]
  318. Shao, Y.; Yan Pei, H.; Rong Hu, W. Nitrogen removal by bioaugmentation in constructed wetlands for rural domestic wastewater in autumn. Desalin. Water Treat. 2013, 51, 147–152. [Google Scholar] [CrossRef]
  319. Pei, H.; Shao, Y.; Chanway, C.P.; Hu, W.; Meng, P.; Li, Z.; Chen, Y.; Ma, G. Bioaugmentation in a pilot-scale constructed wetland to treat domestic wastewater in summer and autumn. Environ. Sci. Pollut. Res. 2016, 23, 7776–7785. [Google Scholar] [CrossRef]
  320. Tang, M.; Zhang, F.; Yao, S.; Liu, Y.; Chen, J. Application of Pseudomonas flava WD-3 for sewage treatment in constructed wetland in winter. Environ. Technol. 2015, 36, 1205–1211. [Google Scholar] [CrossRef] [PubMed]
  321. Haberl, R.; Grego, S.; Langergraber, G.; Kadlec, R.H.; Cicalini, A.R.; Dias, S.M.; Novais, J.M.; Aubert, S.; Gerth, A.; Thomas, H.; et al. Constructed wetlands for the treatment of organic pollutants. J. Soils Sediments 2003, 3, 109–124. [Google Scholar] [CrossRef]
  322. Silvan, N.; Vasander, H.; Laine, J. Vegetation is the main factor in nutrient retention in a constructed wetland buffer. Plant Soil 2004, 258, 179–187. [Google Scholar] [CrossRef]
  323. Vymazal, J. Plants used in constructed wetlands with horizontal subsurface flow: A review. Hydrobiologia 2011, 674, 133–156. [Google Scholar] [CrossRef]
  324. Christofilopoulos, S.; Syranidou, E.; Gkavrou, G.; Manousaki, E.; Kalogerakis, N. The role of halophyte Juncus acutus L. in the remediation of mixed contamination in a hydroponic greenhouse experiment. J. Chem. Technol. Biotechnol. 2016, 91, 1665–1674. [Google Scholar] [CrossRef]
  325. Madera-Parra, C.A.; Peña-Salamanca, E.J.; Peña, M.R.; Rousseau, D.P.L.; Lens, P.N.L. Phytoremediation of Landfill Leachate with Colocasia esculenta, Gynerum sagittatum and Heliconia psittacorum in Constructed Wetlands. Int. J. Phytoremediation 2015, 7, 16–24. [Google Scholar] [CrossRef]
  326. Yan, Q.; Xu, Y.; Yu, Y.; Zhu, Z.W.; Feng, G. Effects of pharmaceuticals on microbial communities and activity of soil enzymes in mesocosm-scale constructed wetlands. Chemosphere 2018, 212, 245–253. [Google Scholar] [CrossRef]
  327. Tong, X.; Wang, X.; He, X.; Xu, K.; Mao, F. Effects of ofloxacin on nitrogen removal and microbial community structure in constructed wetland. Sci. Total Environ. 2019, 656, 503–511. [Google Scholar] [CrossRef]
  328. Weber, K.P.; Mitzel, M.R.; Slawson, R.M.; Legge, R.L. Effect of ciprofloxacin on microbiological development in wetland mesocosms. Water Res. 2011, 45, 3185–3196. [Google Scholar] [CrossRef]
  329. Shao, Y.; Pei, H.; Hu, W.; Chanway, C.P.; Meng, P.; Ji, Y.; Li, Z. Bioaugmentation in lab scale constructed wetland microcosms for treating polluted river water and domestic wastewater in northern China. Int. Biodeterior. Biodegrad. 2014, 95, 151–159. [Google Scholar] [CrossRef]
  330. Lingua, G.; Copetta, A.; Musso, D.; Aimo, S.; Ranzenigo, A.; Buico, A.; Gianotti, V.; Osella, D.; Berta, G. Effect of arbuscular mycorrhizal and bacterial inocula on nitrate concentration in mesocosms simulating a wastewater treatment system relying on phytodepuration. Environ. Sci. Pollut. Res. 2015, 22, 18616–18625. [Google Scholar] [CrossRef] [PubMed]
  331. Pat-Espadas, A.M.; Portales, R.L.; Amabilis-Sosa, L.E.; Gómez, G.; Vidal, G. Review of constructed wetlands for acid mine drainage treatment. Water 2018, 10, 1685. [Google Scholar] [CrossRef] [Green Version]
  332. Prum, C.; Dolphen, R.; Thiravetyan, P. Enhancing arsenic removal from arsenic-contaminated water by Echinodorus cordifolius−endophytic Arthrobacter creatinolyticus interactions. J. Environ. Manag. 2018, 213, 11–19. [Google Scholar] [CrossRef] [PubMed]
  333. Shehzadi, M.; Afzal, M.; Khan, M.U.; Islam, E.; Mobin, A.; Anwar, S.; Khan, Q.M. Enhanced degradation of textile effluent in constructed wetland system using Typha domingensis and textile effluent-degrading endophytic bacteria. Water Res. 2014, 58, 152–159. [Google Scholar] [CrossRef] [PubMed]
  334. Hussain, Z.; Arslan, M.; Malik, M.H.; Mohsin, M.; Iqbal, S.; Afzal, M. Treatment of the textile industry effluent in a pilot-scale vertical flow constructed wetland system augmented with bacterial endophytes. Sci. Total Environ. 2018, 645, 966–973. [Google Scholar] [CrossRef] [PubMed]
  335. Ben Saad, M.; Ben Said, M.; Bousselmi, L.; Ghrabi, A. Application of bioinoculation to enhance rhizocompetence of horizontal subsurface flow constructed wetland system. Desalin. Water Treat. 2016, 57, 22133–22139. [Google Scholar] [CrossRef]
  336. Zhao, X.; Yang, J.; Bai, S.; Ma, F.; Wang, L. Microbial population dynamics in response to bioaugmentation in a constructed wetland system under 10°C. Bioresour. Technol. 2016, 205, 166–173. [Google Scholar] [CrossRef]
  337. Zhang, S.; Yang, X.L.; Li, H.; Song, H.L.; Wang, R.C.; Dai, Z.Q. Degradation of sulfamethoxazole in bioelectrochemical system with power supplied by constructed wetland-coupled microbial fuel cells. Bioresour. Technol. 2017, 244, 345–352. [Google Scholar] [CrossRef]
  338. Paredes, D.; Kuschk, P.; Köser, H. Influence of plants and organic matter on the nitrogen removal in laboratory-scale model subsurface flow constructed wetlands inoculated with anaerobic ammonium oxidizing bacteria. Eng. Life Sci. 2007, 7, 565–576. [Google Scholar] [CrossRef]
  339. Nowrotek, M.; Kotlarska, E.; Łuczkiewicz, A.; Felis, E.; Sochacki, A.; Miksch, K. The treatment of wastewater containing pharmaceuticals in microcosm constructed wetlands: The occurrence of integrons (int1–2) and associated resistance genes (sul1–3, qacEΔ1). Environ. Sci. Pollut. Res. 2017, 24, 15055–15066. [Google Scholar] [CrossRef] [Green Version]
  340. Węgrzyn, A.; Felis, E. Isolation of bacterial endophytes from Phalaris arundinacea and their potential in diclofenac and sulfamethoxazole degradation. Polish J. Microbiol. 2018, 67, 321–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  341. Iasur-Kruh, L.; Hadar, Y.; Minz, D. Isolation and bioaugmentation of an estradiol-degrading bacterium and its integration into a mature biofilm. Appl. Environ. Microbiol. 2011, 77, 3734–3740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  342. Křesinová, Z.; Linhartová, L.; Filipová, A.; Ezechiáš, M.; Mašín, P.; Cajthaml, T. Biodegradation of endocrine disruptors in urban wastewater using Pleurotus ostreatus bioreactor. N. Biotechnol. 2018, 43, 53–61. [Google Scholar] [CrossRef]
  343. Gravel, V.; Dorais, M.; Gruyer, N.; Ménard, C. Constructed wetlands implanted with iris versicolor, Juncus sp. and Phragmites australis as a potential treatment for greenhouse effluents. Acta Hortic. 2011, 1139–1146. [Google Scholar] [CrossRef]
  344. Schäfer, M.L.; Lundström, J.O.; Pfeffer, M.; Lundkvist, E.; Landin, J. Biological diversity versus risk for mosquito nuisance and disease transmission in constructed wetlands in southern Sweden. Med. Vet. Entomol. 2004, 18, 256–267. [Google Scholar] [CrossRef]
  345. Boucher, J.G.; Boudreau, A.; Ahmed, S.; Atlas, E. Response to “Comment on ‘In Vitro Effects of Bisphenol A β-D-Glucuronide (BPA-G) on Adipogenesis in Human and Murine Preadipocytes. Env. Heal. Perspect 2015, 123, 1287–1293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  346. Parrella, A.; Rubino, M.; Isidori, M.; Nardelli, A.; Previtera, L.; Lavorgna, M. Ecotoxicity of naproxen and its phototransformation products. Sci. Total Environ. 2005, 348, 93–101. [Google Scholar]
  347. González, N.; Simarro, R.; Molina, M.C.; Bautista, L.F.; Delgado, L.; Villa, J.A. Effect of surfactants on PAH biodegradation by a bacterial consortium and on the dynamics of the bacterial community during the process. Bioresour. Technol. 2011, 102, 9438–9446. [Google Scholar] [CrossRef]
  348. Molina, M.D.C.; González Benítez, N.; Simarro, R.; Bautista, L.F.; Vargas, C.; García Cambero, J.P.; Díaz, E.M.; Arrayás, M.; Quijano, M.A. Bioremediation techniques for naproxen and carbamazepine elimination. Toxicity evaluation test. Chim. Oggi/Chemistry Today 2016, 34, 52–55. [Google Scholar]
  349. Anderson, J.C.; Beyger, L.; Guchardi, J.; Holdway, D. Chronic effects of hydroxypropyl-β-cyclodextrin on reproduction in the American flagfish (Jordanella floridae) over one complete life cycle. Environ. Toxicol. Chem. 2016, 35, 1358–1363. [Google Scholar] [CrossRef]
  350. Rubasinghege, G.; Gurung, R.; Rijal, H.; Maldonado-Torres, S.; Chan, A.; Acharya, S.; Rogelj, S.; Piyasena, M. Abiotic degradation and environmental toxicity of ibuprofen: Roles of mineral particles and solar radiation. Water Res. 2018, 131, 22–32. [Google Scholar] [CrossRef] [PubMed]
  351. Kase, R.; Korkaric, M.; Werner, I.; Ågerstrand, M. Criteria for Reporting and Evaluating ecotoxicity Data (CRED): Comparison and perception of the Klimisch and CRED methods for evaluating reliability and relevance of ecotoxicity studies. Environ. Sci. Eur. 2016, 28, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  352. Jarvis, A.L.; Bernot, M.J.; Bernot, R.J. Relationships between the psychiatric drug carbamazepine and freshwater macroinvertebrate community structure. Sci. Total Environ. 2014, 496, 499–509. [Google Scholar] [CrossRef] [PubMed]
  353. Mottaleb, M.A. Use of LC-MS and GC-MS Methods to measure emerging contaminants pharmaceutical and personal care products (PPCPs) in fish. J. Chromatogr. Sep. Tech. 2015, 06. [Google Scholar] [CrossRef]
  354. Dahms, H.U.; Hagiwara, A.; Lee, J.S. Ecotoxicology, ecophysiology, and mechanistic studies with rotifers. Aquat. Toxicol. 2011, 101, 1–12. [Google Scholar] [CrossRef]
  355. Prado, R.; Rioboo, C.; Herrero, C.; Cid, A. The herbicide paraquat induces alterations in the elemental and biochemical composition of non-target microalgal species. Chemosphere 2009, 76, 1440–1444. [Google Scholar] [CrossRef] [Green Version]
  356. McCormick, P.V.; Cairns, J. Algae as indicators of environmental change. J. Appl. Phycol. 1994, 6, 509–526. [Google Scholar] [CrossRef]
  357. Hoppe, P.D.; Rosi-Marshall, E.J.; Bechtold, H.A. The antihistamine cimetidine alters invertebrate growth and population dynamics in artificial streams. Freshw. Sci. 2012, 31, 379–388. [Google Scholar] [CrossRef]
  358. Stancová, V.; Ziková, A.; Svobodová, Z.; Kloas, W. Effects of the non-steroidal anti-inflammatory drug (NSAID) naproxen on gene expression of antioxidant enzymes in zebrafish (Danio rerio). Environ. Toxicol. Pharmacol. 2015, 40, 343–348. [Google Scholar] [CrossRef]
  359. Neher, D.A. Role of nematodes in soil health and their use as indicators. J. Nematol. 2001, 33, 161–168. [Google Scholar]
  360. Karlsson, M.V.; Marshall, S.; Gouin, T.; Boxall, A.B.A. Routes of uptake of diclofenac, fluoxetine, and triclosan into sediment-dwelling worms. Environ. Toxicol. Chem. 2016, 35, 836–842. [Google Scholar] [CrossRef] [PubMed]
  361. García-Santiago, X.; Franco-Uría, A.; Omil, F.; Lema, J.M. Risk assessment of persistent pharmaceuticals in biosolids: Dealing with uncertainty. J. Hazard. Mater. 2016, 302, 72–81. [Google Scholar] [CrossRef] [PubMed]
  362. Chen, Z.; Ngo, H.H.; Guo, W. A critical review on sustainability assessment of recycled water schemes. Sci. Total Environ. 2012, 426, 13–31. [Google Scholar] [CrossRef] [PubMed]
  363. Regulation (EC) No 1907/2006 of the European Parliament and of the Council of 18 December 2006 concerning the Registration, Evaluation, Authorisation and Restriction of Chemicals, (REACH) establishing a European Chemicals Agency, amending Directive 1999/45/EC, repealing Council Regulation (EEC) No 793/93, Commission Regulation (EC) No 1488/94, Council Directive 76/769/EEC, Commission Directives 91/155/EEC, 93/67/EEC, 93/105/EC, 2000/21/EC. Available online: https://eur-lex.europa.eu/LexUriServ/LexUriServ.do?uri=OJ:L:2007:136:0003:0280:en:PDF (accessed on 29 May 2007).
  364. Leng, L.; Wei, L.; Xiong, Q.; Xu, S.; Li, W.; Lv, S.; Lu, Q.; Wan, L.; Wen, Z.; Zhou, W. Use of microalgae based technology for the removal of antibiotics from wastewater: A review. Chemosphere 2020, 238, 124680. [Google Scholar] [CrossRef] [PubMed]
  365. Chen, R.Z.; Wong, M.H. Integrated wetlands for food production. Environ. Res. 2016, 426, 13–31. [Google Scholar] [CrossRef] [Green Version]
  366. Nguyen, P.M.; Afzal, M.; Ullah, I.; Shahid, N.; Baqar, M.; Arslan, M. Removal of pharmaceuticals and personal care products using constructed wetlands: Effective plant-bacteria synergism may enhance degradation efficiency. Environ. Sci. Pollut. Res. 2019, 26, 21109–21126. [Google Scholar] [CrossRef]
  367. Vo, H.N.P.; Ngo, H.H.; Guo, W.; Chang, S.W.; Nguyen, D.D.; Chen, Z.; Wang, X.C.; Chen, R.; Zhang, X. Microalgae for saline wastewater treatment: A critical review. Crit. Rev. Environ. Sci. Technol. 2020, 50, 1224–1265. [Google Scholar] [CrossRef]
  368. Homrich, A.S.; Galvão, G.; Abadia, L.G.; Carvalho, M.M. The circular economy umbrella: Trends and gaps on integrating pathways. J. Clean. Prod. 2018, 175, 525–543. [Google Scholar] [CrossRef]
  369. Tsolakis, N.; Bam, W.; Srai, J.S.; Kumar, M. Renewable chemical feedstock supply network design: The case of terpenes. J. Clean. Prod. 2019, 222, 802–822. [Google Scholar] [CrossRef]
  370. Jiménez-González, C.; Poechlauer, P.; Broxterman, Q.B.; Yang, B.S.; Am Ende, D.; Baird, J.; Bertsch, C.; Hannah, R.E.; Dell’Orco, P.; Noorman, H.; et al. Key green engineering research areas for sustainable manufacturing: A perspective from pharmaceutical and fine chemicals manufacturers. Org. Process Res. Dev. 2011, 51, 900–911. [Google Scholar] [CrossRef]
Figure 1. Number of papers published each year on the topic of pharmaceuticals and personal hygiene products (PPCP) and its removal according to Scopus database. Dark blue represents all papers that contain the search term “PPCP”, the red line encloses the terms PPCP and the following terms associated with their biological degradation (bacteria, archaea, fungi, microbial consortium, biological or degradation). The green line represents the terms PPCP and other terms associated with physicochemical removal of PPCP, namely: ozonation, Fenton, oxidation, UV, ionizing, irradiation, chemical, activated, carbon, graphene or nanotubes. Purple line counts the articles that reference PPCP and their biological degradation stated before, but excluding all terms related to physicochemical degradation to avoid overlap. Light blue considers the search terms PPCP and the physicochemical degradation terms stated before while excluding any term related to biological degradation of PPCP.
Figure 1. Number of papers published each year on the topic of pharmaceuticals and personal hygiene products (PPCP) and its removal according to Scopus database. Dark blue represents all papers that contain the search term “PPCP”, the red line encloses the terms PPCP and the following terms associated with their biological degradation (bacteria, archaea, fungi, microbial consortium, biological or degradation). The green line represents the terms PPCP and other terms associated with physicochemical removal of PPCP, namely: ozonation, Fenton, oxidation, UV, ionizing, irradiation, chemical, activated, carbon, graphene or nanotubes. Purple line counts the articles that reference PPCP and their biological degradation stated before, but excluding all terms related to physicochemical degradation to avoid overlap. Light blue considers the search terms PPCP and the physicochemical degradation terms stated before while excluding any term related to biological degradation of PPCP.
Applsci 10 03391 g001
Figure 2. Different mechanisms to PPCPs bioremoval, case of 17α-ethinylestradiol from wastewater. A, hydrophobic bioadsorption [96]; B, Absorption and bacterial intracellular degradation [97,98]; C, Absorption and microalgae intracellular degradation [99]; D, bioaccumulation [100]; E, extracellular degradation by laccase, other lignolytic enzymes do not seem to be involved [101,102,103]; F, intermediate metabolites secretion; G, synergistic degradation mediated by other microorganisms [39,42,104]; H, Desorption and mobilization by biosurfactants [105]; I, photodegradation coupled to biodegradation [106,107,108]; J, Degradative way by biogenic manganese oxides [109,110]; and, K, adsorption to natural activated carbon [111]. Discontinuous lines probable degradative pathways, but not proven.
Figure 2. Different mechanisms to PPCPs bioremoval, case of 17α-ethinylestradiol from wastewater. A, hydrophobic bioadsorption [96]; B, Absorption and bacterial intracellular degradation [97,98]; C, Absorption and microalgae intracellular degradation [99]; D, bioaccumulation [100]; E, extracellular degradation by laccase, other lignolytic enzymes do not seem to be involved [101,102,103]; F, intermediate metabolites secretion; G, synergistic degradation mediated by other microorganisms [39,42,104]; H, Desorption and mobilization by biosurfactants [105]; I, photodegradation coupled to biodegradation [106,107,108]; J, Degradative way by biogenic manganese oxides [109,110]; and, K, adsorption to natural activated carbon [111]. Discontinuous lines probable degradative pathways, but not proven.
Applsci 10 03391 g002
Figure 3. Scientific production on bacteria (A), fungi (B) and algae (C) associated with the elimination of PPCPs from contaminated systems. Left column. Laboratory-scale publications related to identification of useful strains, determination of degradative and/or accumulative capacity, metabolic pathways, toxicology and OMIC (proteomic, genomic, transcriptomic) approach. Right column. Cellular or enzymatic immobilization, activated sludge with or without bioaugmentation, membrane bioreactors, photobioreactors, commercial products, nanoparticles).
Figure 3. Scientific production on bacteria (A), fungi (B) and algae (C) associated with the elimination of PPCPs from contaminated systems. Left column. Laboratory-scale publications related to identification of useful strains, determination of degradative and/or accumulative capacity, metabolic pathways, toxicology and OMIC (proteomic, genomic, transcriptomic) approach. Right column. Cellular or enzymatic immobilization, activated sludge with or without bioaugmentation, membrane bioreactors, photobioreactors, commercial products, nanoparticles).
Applsci 10 03391 g003aApplsci 10 03391 g003b
Figure 4. Towards the circular supply chain management (CSCM) reactivation. a. High PPCPs concentration inhibit algae growth [196,197] but low concentrations can allow bioaccumulation (b) or metabolism (c) and therefore, the increase in biomass [180]. b. PPCP bioadsorption and bioabsorpiton associated to glucosil-conjugation bioaccumulation of PPCP or PPCP toxic derivatives [230]. Increase of ROS defense reactions [231]. d. Increase in the production of algal biomass useful by acclimatization in the production of biodiesel and biogas [182]. c. Bioabsorption and intracellular degradation by cytochrome P450 and associated enzymes [223] and extracellular degradation mediated by laccases [102,103]. e. Biomass usable in food and feed industry [227], biofertilizants [232], and pharmaceutical industry [233].
Figure 4. Towards the circular supply chain management (CSCM) reactivation. a. High PPCPs concentration inhibit algae growth [196,197] but low concentrations can allow bioaccumulation (b) or metabolism (c) and therefore, the increase in biomass [180]. b. PPCP bioadsorption and bioabsorpiton associated to glucosil-conjugation bioaccumulation of PPCP or PPCP toxic derivatives [230]. Increase of ROS defense reactions [231]. d. Increase in the production of algal biomass useful by acclimatization in the production of biodiesel and biogas [182]. c. Bioabsorption and intracellular degradation by cytochrome P450 and associated enzymes [223] and extracellular degradation mediated by laccases [102,103]. e. Biomass usable in food and feed industry [227], biofertilizants [232], and pharmaceutical industry [233].
Applsci 10 03391 g004
Figure 5. Towards the CSCM reactivation. (A). Pathway of PPCPs uptake and bioaccumulation in plants (roots, leaves and endophytes) and chemical adsorption on substrate. Harvest of macrophytes biomass, use in secondary applications. a. [255]; b. [256]. (B). Pathway of PPCPs uptake and complete mineralization (metabolic interactions between plants and associated microorganisms), air and plant-mediated volatilization in constructed wetlands. Harvest of macrophytes biomass, use in secondary applications. a. [103]; b. [255]; c. [256]; d. [257,258]; e. [259,260].
Figure 5. Towards the CSCM reactivation. (A). Pathway of PPCPs uptake and bioaccumulation in plants (roots, leaves and endophytes) and chemical adsorption on substrate. Harvest of macrophytes biomass, use in secondary applications. a. [255]; b. [256]. (B). Pathway of PPCPs uptake and complete mineralization (metabolic interactions between plants and associated microorganisms), air and plant-mediated volatilization in constructed wetlands. Harvest of macrophytes biomass, use in secondary applications. a. [103]; b. [255]; c. [256]; d. [257,258]; e. [259,260].
Applsci 10 03391 g005aApplsci 10 03391 g005b
Figure 6. Towards the CSCM reactivation, serial system. A. Anaerobic bioreactor-pond. Removal of Aerobic microorganisms and PPCPs [229,236]. B. Algal canal-contracted wetland. Removal of anaerobic bacteria, PPCPS and PPCPs residue removal [223]. C. Mixed treatment (constructed wetland). Remove of nutrients and contaminants. Reuse of zero waste [306].
Figure 6. Towards the CSCM reactivation, serial system. A. Anaerobic bioreactor-pond. Removal of Aerobic microorganisms and PPCPs [229,236]. B. Algal canal-contracted wetland. Removal of anaerobic bacteria, PPCPS and PPCPs residue removal [223]. C. Mixed treatment (constructed wetland). Remove of nutrients and contaminants. Reuse of zero waste [306].
Applsci 10 03391 g006
Table 1. Effects of different types of mixture of PPCPs on several species of aquatic organisms.
Table 1. Effects of different types of mixture of PPCPs on several species of aquatic organisms.
PhylumSpeciesMixture TypeEndpoint/Biomarker EffectReference
BacteroidetesFlavobacterium sp.Psychostimulant
+ Antihistamine
+ Antibiotic
+ Antidiabetic, cholesterol reducer
Relative decreases on bacterial community composition[16]
Pseudomonas sp.Psychostimulant
+ Antihistamine
+ Antibiotic
+ Antidiabetic, cholesterol reducer
Relative increases on bacterial community composition[16]
FirmicutesParacoccus yeeiAnalgesic
and antipyretic drug
+ Nonsteroidal anti-inflammatory drug
Susceptibility[13]
Staphylococcus aureusAnalgesic
and antipyretic drug
+ Nonsteroidal anti-inflammatory drug
Susceptibility[13]
ProteobacteriaEnterobacter aerogenesAnalgesic
and antipyretic drug
+ Nonsteroidal anti-inflammatory drug
Resistance[13]
Enterobacter cloacaeAnalgesic
and antipyretic drug
+ Nonsteroidal anti-inflammatory drug
Resistance[13]
Escherichia coliHair dyesCytotoxicity and mutagenic effect[64]
Vibrio fischeriAntibiotics
+ Disinfectant
+ H2 blocker
+ Lipid regulators
+ Nonsteroidal anti-inflammatory drug
+ Preservatives
Narcosis[37]
RotiferaPlationus patulusAntidepressant
+ Psychostimulant
+ Disinfectants
+ Nonsteroidal anti-inflammatory drugs
Egg reduction and detachment[65]
Phanerogams Magnoliophyte Lemna minorAntibioticsGrowth reduction[66]
ChlorophytaChlamydomonas microsphaeraAntidepressant
+ Disinfectants
Inhibitory effect on algal growth[67]
Chlorella pyrenoidosaAntidepressant
+ Disinfectants
Inhibitory effect on algal growth[67]
Chlorella ellipsoideaAntidepressant
+ Disinfectants
Inhibitory effect on algal growth[67]
Dunaliella salinaAntidepressant
+ Disinfectants
Inhibitory effect on algal growth[67]
Dunaliella parvaAntidepressant
+ Disinfectants
Inhibitory effect on algal growth[67]
Pseudokirchneriella subcapitataLipid regulators
+ Nonsteroidal anti-inflammatory drugs
Increased the level of lipid peroxidation, glutathione transferase and metallothioneins. DNA damage. Destabilization of the lysosomal membrane[68]
Scendesmus obliquusAntidepressant
+ Disinfectants
Inhibitory effect on algal growth[67]
Scendesmus quadricaudaAntidepressant
+ Disinfectants
Inhibitory effect on algal growth[67]
Scenedesmus vacuolatusAntibioticsGrowth redution[66]
Tetraselmis suecicaAntibioticsInhibited growth. Decrease in esterase activity and alteration of chlorophyll a, cellular content and autofluorescence[69]
MolluscaElliptio complanataAnalgesic
+ Antibiotics
+ Antidepressant
+ Anticonvulsants
+ Lipid regulators
+ Nonsteroidal anti-inflammatory drugs
Adverse effects on the immune system[70]
Dreissena polymorphaLipid regulators
+ Nonsteroidal anti-inflammatory drugs
High levels of lipid peroxidation[68]
Lampsilis siliquoideaAntibiotics
+ Hormones
+ Lipid regulators
Reduction in filter-feeding[71]
Unio tumidusHormones
+ Disinfectants
+ Nonsteroidal anti-inflammatory drugs
Elevated levels of lactate/pyruvate ratio, lipofuscin, DNA fragmentation and caspase-3 activity[72]
ArthropodaChironomus ripariusHormonesDeformities in the mouth, decreased fertility[73,74,75]
Daphnia magnaAntidepressant
+ Hormones
Population growth rate[76]
AntibioticsViability reduction[66]
AntochaAntidepressantsMetamorphosis occurred earlier and more frequently[77]
CorydalusAntidepressantsMetamorphosis occurred earlier and more frequently[77]
EctopriaAntidepressantsMetamorphosis occurred earlier and more frequently[77]
PsephenusAntidepressantsMetamorphosis occurred earlier and more frequently[77]
Gammarus fossarumAntibioticsIncrease in body mass[78]
Gammarus pulexBeta blockers
+ Disinfectants
+ Nonsteroidal anti-inflammatory drugs
Alterations in metabolite concentrations[79]
Palaemonetes pugioAntibiotics
+ Psychostimulant
Negative effects on offspring survival and development[80]
ChordataDanio rerioAntidepressant
+ Anticonvulsant
+ Antihistamine
+ Nonsteroidal anti-inflammatory drug
Perturbations in both the metabolome and transcriptome[81]
Antidepressant
+ Anticonvulsant
+ Lipid regulators
+ Nonsteroidal anti-inflammatory drugs
Decreased embryo production[82]
Oncorhynchus mykissHormonesInhibitor of estrogenicity[83]
Sunscreen agentsIncreased activity of certain P450 cytochromes[84]
Antidepressant
+ Anticonvulsants
+ Beta blockers
+ Psychostimulant
+ Hormones
+ Nonsteroidal anti-inflammatory drugs
+ Thiazide diuretic
+ Fragrances
ROS production and the ß-gal inhibition (RTG-2 cell line)[85]
Oryzias latipesNonsteroidal anti-inflammatory drugsAntiovulatory activity[86]
Anticonvulsant
+ Disinfectant
+ Nonsteroidal anti-inflammatory drug
Effects on feeding behavior and swimming speed[87]
Pimephales promelasAntihistamine
+ Beta blocker
+ Psychostimulant
+ Disinfectants
+ Hormone
+ Lipid regulators
+ Nonsteroidal anti-inflammatory drugs
Molecular estrogenic effects[88]
Pseudorasbora parvaAntidepressant
+ Disinfectants
AChE and EROD activities inhibitionLipid peroxidation[89]
Trematomus bernachiiHormones
+ Component in plastic and epoxy resins
+ Disinfectants
+ Precursors to the non-ionic surfactants
+ Preservatives
Preferential accumulation in fish correlating negatively with fillet size [90]
Table 2. Effects of different types of mixture of PPCPs on several species of terrestrial organisms.
Table 2. Effects of different types of mixture of PPCPs on several species of terrestrial organisms.
PhylumSpeciesMixture TypeEndpoint/Biomarker EffectReference
NematodaSoil nematode communityHormonesNumber of nematodes was decreased and change the sex ratio in a free-living nematode community[91]
Soil nematode communityAntibiotics
+ Anticonvulsants
+ Antipsychotic, antiemetic
+ Beta blockers
+ Psychostimulant
+ Component in plastic and epoxy resins
+ Diuretic
+ Fragrances
+ Hormones
+ Lipid regulators
+ Nonsteroidal anti-inflammatory drugs
+ Preservatives
The diversity and structure of the soil nematode community vastly altered[92]
AnnelidaEisenia andreiDisinfectantsAffect the growth and reproductive performance[93]
ArthropodaCulex quinquefasciatusAntibiotics
+ Psychostimulant
+ Hormones
+ Nonsteroidal anti-inflammatory drugs
Increase developmental time of larvae. Altered the mosquito bacterial microbiome[94]
MagnoliophyteAllium cepaHair dyesCytotoxic and mutagenic [64]
Triticum aestivum L.Analgesic
and antipyretic drug
Wheat shoot and root elongation decreased. The antioxidative defensive system in roots was damaged[24]

Share and Cite

MDPI and ACS Style

Molina, M.C.; Bautista, L.F.; Catalá, M.; de las Heras, M.R.; Martínez-Hidalgo, P.; San-Sebastián, J.; González-Benítez, N. From Laboratory Tests to the Ecoremedial System: The Importance of Microorganisms in the Recovery of PPCPs-Disturbed Ecosystems. Appl. Sci. 2020, 10, 3391. https://doi.org/10.3390/app10103391

AMA Style

Molina MC, Bautista LF, Catalá M, de las Heras MR, Martínez-Hidalgo P, San-Sebastián J, González-Benítez N. From Laboratory Tests to the Ecoremedial System: The Importance of Microorganisms in the Recovery of PPCPs-Disturbed Ecosystems. Applied Sciences. 2020; 10(10):3391. https://doi.org/10.3390/app10103391

Chicago/Turabian Style

Molina, María Carmen, Luis Fernando Bautista, Myriam Catalá, María Rosa de las Heras, Pilar Martínez-Hidalgo, Jon San-Sebastián, and Natalia González-Benítez. 2020. "From Laboratory Tests to the Ecoremedial System: The Importance of Microorganisms in the Recovery of PPCPs-Disturbed Ecosystems" Applied Sciences 10, no. 10: 3391. https://doi.org/10.3390/app10103391

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