Next Article in Journal
Defining Seasonal Functional Traits of a Freshwater Zooplankton Community Using δ13C and δ15N Stable Isotope Analysis
Next Article in Special Issue
Treatment of Dairy Wastewater by Oxygen Injection: Occurrence and Removal Efficiency of a Benzotriazole Based Anticorrosive
Previous Article in Journal
Cross-Analysis of Land and Runoff Variations in Response to Urbanization on Basin, Watershed, and City Scales with/without Green Infrastructures
Previous Article in Special Issue
High-Strength Domestic Wastewater Treatment and Reuse with Onsite Passive Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Carbamazepine as a Possible Anthropogenic Marker in Water: Occurrences, Toxicological Effects, Regulations and Removal by Wastewater Treatment Technologies

1
Strategic Water Infrastructure Lab, School of Civil, Mining and Environmental Engineering, University of Wollongong, Wollongong, NSW 2522, Australia
2
School of Mechanical, Materials and Mechatronics Engineering, University of Wollongong, Wollongong, NSW 2522, Australia
3
Illawarra Health and Medical Research Institute, University of Wollongong, Wollongong, NSW 2522, Australia
4
Centre for Medical and Molecular Bioscience, University of Wollongong, Wollongong, NSW 2522, Australia
5
Environmental Science Centre, Department of Urban Engineering, University of Tokyo, Tokyo 113-0033, Japan
*
Author to whom correspondence should be addressed.
Water 2018, 10(2), 107; https://doi.org/10.3390/w10020107
Submission received: 8 December 2017 / Revised: 15 January 2018 / Accepted: 21 January 2018 / Published: 26 January 2018
(This article belongs to the Special Issue Recent Advances in Water Management: Saving, Treatment and Reuse)

Abstract

:
Carbamazepine (CBZ), a pharmaceutical compound, has been proposed as an anthropogenic marker to assess water quality due to its persistence in conventional treatment plants and widespread presence in water bodies. This paper presents a comprehensive literature review on sources and occurrences of CBZ in water bodies, as well as toxicological effects and regulations of the drug. Given the documented side effects of CBZ on the human body when taken medicinally, its careful monitoring in water is recommended. CBZ residues in drinking water may provide a pathway to embryos and infants via intrauterine exposure or breast-feeding, which may cause congenital malformations and/or neurodevelopmental problems over long term exposure. An in-depth technical assessment of the conventional and advanced treatment technologies revealed the inadequacy of the standalone technologies. Compared to conventional activated sludge and membrane bioreactor processes, effective removal of CBZ can be achieved by nanofiltration and reverse osmosis membranes. However, recent studies have revealed that harsh chemical cleaning, as required to mitigate membrane fouling, can often reduce the long-term removal efficiency. Furthermore, despite the efficient performance of activated carbon adsorption and advanced oxidation processes, a few challenges such as cost of chemicals and regeneration of activated carbon need to be carefully considered. The limitations of the individual technologies point to the advantages of combined and hybrid systems, namely, membrane bioreactor coupled with nanofiltration, adsorption or advanced oxidation process.

1. Introduction

The occurrence of pharmaceutically active compounds (PhACs) in environmental systems such as freshwater bodies has become a topic of growing concern over the last decade due to their potential detrimental impacts on aquatic life and human health [1,2]. Because a large proportion of PhACs ends up in sewage via bodily excretion and indiscriminate disposal of unwanted/expired pharmaceuticals, disposal of untreated or ineffectively treated wastewater is considered a major source of their occurrence in environmental systems [3,4,5,6]. The need for effective removal of PhACs has resulted in the emergence of various advanced wastewater treatment technologies such as membrane technology and advanced oxidation processes [4].
The widespread occurrence of PhACs in wastewater and wastewater-impacted freshwater has triggered the establishment of water quality standards for their regular monitoring [5,7]. In this context, carbamazepine (CBZ) has been proposed as an anthropogenic marker of sewage contamination in freshwater bodies [8,9]. CBZ is an anticonvulsant and mood stabilizing drug used primarily in the treatment of epilepsy and bipolar disorder [10]. Table 1 summarizes the salient physicochemical properties of CBZ. CBZ is one of the most frequently detected pharmaceutical compounds in environmental systems [4,11]. It is ubiquitously present in raw wastewater in the high ng/L to low µg/L range and is only poorly removed by the conventional wastewater treatment plants (WWTP) [12,13]. Almost all the known advanced technologies have been tested for CBZ removal, however, none has appeared as a universal solution [14,15].
A number of interesting review papers have been published on the occurrences of micropollutants in environmental systems such as wastewater [18], surface water [4] and groundwater [5], as well as the performance of conventional treatment technologies for the removal of micropollutants [19,20,21]. The persistence of CBZ in conventional treatment processes leads to its widespread occurrence in water bodies. Thus, CBZ has been proposed as an anthropogenic marker to assess water treatment quality. CBZ removal by conventional WWTPs has been reviewed previously by Zhang et al. [16]. However, the efficacy of advanced treatment technologies has not been critically analyzed to date. In addition, toxicity of CBZ to aquatic species and human and relevant regulations have not been comprehensively reviewed.
In this paper, the occurrence of CBZ in wastewater and freshwater bodies (e.g., surface and groundwater) along with the associated influencing factors are systematically analyzed. In addition, the toxicological effects of CBZ on the aquatic ecosystem are critically discussed. Importantly, the factors governing the resistance of CBZ to the available wastewater treatment processes are elucidated, and the efficacy of advanced/emerging treatment processes is comprehensively discussed.

2. Occurrence in Aquatic Systems

Following consumption, up to 10% of CBZ is excreted from human body [4,22]. Recent studies have reported a few tens to several thousands of ng/L CBZ in municipal wastewater [22,23,24]. CBZ is poorly removed (typically less than 10%) by the conventional WWTPs [25,26,27]. Hence, treatment plant effluents are an important gateway for CBZ to enter surface and groundwater. Table 2 depicts the reported levels of occurrence of CBZ in WWTP effluent, surface water and groundwater. Generally, CBZ concentration has been reported to be higher in WWTP effluents as compared to surface water (Table 2) because dilution and natural attenuation can significantly reduce the concentration of pollutants [28]. CBZ is most likely to reach groundwater via bank infiltration of WWTP effluent [5,29]. In addition, seepage of landfill leachate and combined sewer overflows can contaminate groundwater [30]. In this section, the factors influencing the occurrence of PhACs including CBZ in raw wastewater, WWTP effluent and freshwater bodies are critically discussed.

2.1. Wastewater Treatment Plant Effluent

Since a number of factors can affect CBZ occurrence in wastewater, its concentration in WWTP effluent has been observed to be highly variable (30–6300 ng/L) (Table 2). The variation in WWTP effluent CBZ concentration (Table 2 and Table 3) can be attributed to a number of factors such as CBZ production/consumption rate, environmental regulations, effectiveness of the WWTPs and the seasonal factors affecting WWTP performance [4,68,69].
Some studies have reported good correlation between CBZ concentration in wastewater with its production and consumption rate at the corresponding locations. For instance, Kasprzyk-Hordern et al. [22] and Choi et al. [48] studied the occurrence of PhACs such as carbamazepine, acetaminophen, sulfamethoxazole and codeine in wastewater from selected locations of UK and Korea, respectively. They observed that the concentration of pharmaceuticals in wastewater correlated well with their consumption rates in the respective countries [22,48]. By contrast, CBZ concentration in the WWTP effluent of Canada, Finland and Switzerland (Table 3) did not correlate well with CBZ consumption rates, indicating that there are other influencing factors.
PhAC occurrence in wastewater can be governed by, among other factors, their excretion rate after metabolism within human body. However, it is important to note that low excretion rate from human body does not necessarily lead to their detection in low concentration or frequency in water. For instance, excretion rate of CBZ and a few other PhACs such as ibuprofen, clofibric acid and gemfibrozil is generally low (1–10%), while excretion rates of up to 70% were reported for some PhACs such as atenolol and paracetamol [76,77]. However, as noted above, CBZ is detected in significantly high concentration in raw and treated wastewater because of its low removal by the natural attenuation process and by the WWTPs [4,78].
Seasonal variations that affect the flow pattern of wastewater in combined sewerage system can lead to the change in wastewater composition. In a study by Kasprzyk-Hordern et al. [22], an increase of up to two-folds in the concentration of PhACs including CBZ was observed during dry weather. Moreover, WWTP effluent CBZ concentration in some countries such as Australia [70], Switzerland [26,74] and USA [79] were reported to be consistently higher in winter than in summer (Table 3), which can also be attributed to drier weather conditions in winter [22,26].

2.2. Surface Water

The major source of CBZ in surface water is the disposal of WWTP effluent [80,81]. After its release to freshwater bodies, different natural attenuation processes such as photolysis, aerobic biodegradation, sorption onto sediments and dilution in surface water play an important role in reducing CBZ concentration [81,82]. However, in-stream attenuation rate varies depending on the physicochemical properties of the PhACs and the local environmental conditions. For instance, Kunkel and Radke [83] observed different attenuation rates for 10 pharmaceuticals including CBZ in river water, and this variation was generally attributed to the physicochemical properties of the compounds. Similarly, in a study that investigated the relationship between attenuation rate and physicochemical properties of 225 micropollutants [82], high attenuation rate was obtained for compounds having medium to low volatility (−4 < log Kaw < −2) and significant hydrophilicity (0 < log Kow < 4.5). This is because these micropollutants are better exposed to in-stream biotic (e.g., biotransformation) and abiotic (e.g., photolysis) attenuation processes as compared to hydrophobic micropollutants that are adsorbed onto river sediments [84]. Since log Kow value for CBZ falls between 0-4.5 [17,81], its concentration is expected to be reduced via in-stream attenuation processes [85].
Water dilution can reduce CBZ concentration in surface water. Indeed, higher concentrations of PhACs including CBZ was reported in surface water bodies during dry weather as compared to that observed in wet weather [86]. In a study by Wang et al. [87], CBZ concentration in surface water was lower in samples that were collected during summer than those collected during winter. This is probably because of the enhanced biodegradation rate due to higher temperature in summer [4,81]. Heavy rainfall can often cause increased leaching of PhACs from openly dumped municipal and hospital solid waste. Storm water runoff can lead these compounds to surface water, consequently increasing the concentration of PhACs in surface water [23,88].
Sorption onto river sediments has been reported to reduce the aqueous phase concentration of hydrophobic PhACs along the river segment. However, in this case, concentration of these pollutants is not significantly reduced by the in-stream biotic and abiotic attenuation processes [81,84,89]. In a study by Riml et al. [90], concentration of two PhACs, namely bezafibrate and metoprolol, was observed to be mainly reduced by sorption onto sediments, while the role of biotransformation and photolysis was insignificant. Since it is a moderately hydrophobic compound (log Kow = 2.45), reduction in CBZ concentration has been attributed to photolysis and sorption [85,91].
Residues of CBZ are introduced into sea water via surface runoff and groundwater discharge [92]. Concentrations of CBZ in sea water are very low. Weigel et al. [93] detected CBZ in sea water at a concentration of 2 ng/L by using a method that comprised solid-phase extraction and GC/MS quantification with a detection limit of 0.1 and 0.7 ng/L. Although this value seems insignificant, the fact that CBZ is detectable in sea water indicates that it is an extremely persistent compound. Also interesting to note, is that residues of CBZ can accumulate in soil through seepage of irrigation water, and due to sewage sludge used as fertilizer [94,95]. A study conducted by the US Geological Survey found an average CBZ concentration of 41.6 ng/mg in the sediment of 44 rivers across the US [96].

2.3. Groundwater

Groundwater constitutes approximately 30% of the total freshwater resources in the world. Because 70% of the freshwater resources are frozen, groundwater represents 97% of freshwater available for human use [97]. Groundwater is the major source of freshwater for domestic and industrial use in many countries. PhACs can contaminate groundwater through different pathways such as percolation of landfill leachate, artificial recharge, percolation of storm water runoff and leakages from sewers and septic tanks [5]. Depending on the organic fraction of the soil, high attenuation of some PhACs such as hydrophobic compounds (log D > 3) can occur in soil strata en route to groundwater [98,99]. Nevertheless, CBZ concentration in groundwater has been reported to be in the range of 1–100 ng/L in available studies from Canada, Germany, Japan and USA (Table 2). However, a higher concentration of CBZ (425 to 3600 ng/L) in groundwater was observed at a site in UK [30,59,60]. This is probably because CBZ concentration was also higher in WWTP effluents and surface water in that location (Table 2 and Table 3). Indeed, Stuart et al. [30] reported that the occurrence of CBZ in groundwater is most likely to be derived from the bank infiltration of WWTP effluent or through surface water/groundwater interaction. Concentration of CBZ in groundwater may not be as high as in surface water but it is still an issue that should be addressed on a priority basis.

3. Toxicological Effects

CBZ is widely detected in water bodies, hence it is essential to evaluate its effects on the ecosystems. A number of studies have assessed the ecotoxicity of CBZ (Table 4). In an experiment conducted by Ying et al. [25], the respiratory quotient value for CBZ was 4.69, indicating potential risks to aquatic organisms. However, other experimental studies have shown that CBZ may not pose an immediate risk. For example, Ferrari et al. [100] studied the toxicological effects of CBZ on bacteria, algae, microcrustaceans and fish. It was observed to have a relatively limited acute ecotoxicity on the tested organisms.
Possible human health effects of long-term exposure to different pharmaceuticals may include endocrine disruption, induction of antibiotic resistance in human pathogens, genotoxicity, carcinogenicity, allergic reactions, and reproductive and/or developmental effects [109,110]. Limited research has been conducted on the potential human health risks of long-term exposure to CBZ residues in water. Risk assessments conducted to date have generally shown that the trace concentrations of CBZ detected in drinking water does not pose an unacceptable health risk to humans [111]. However, careful monitoring must continue, given the documented side effects of CBZ on the human body when taken medicinally.
CBZ is the main cause of the Stevens–Johnson syndrome and its associated disease toxic epidermal necrolysis in Southeast Asian countries due to its intake medicinally [112]. These are two forms of a life-threatening skin condition with an overall mortality rate of 30%, in which cell death causes the epidermis to separate from the dermis [113]. Recent studies have also revealed that intrauterine exposure of CBZ is associated with spina bifida [114] and neuro developmental problems [115] of human embryo when gravidas were exposed to CBZ monotherapy. Atkinson et al. [116] also reported higher fetal losses and congenital malformation rates among women who were prescribed carbamazepine during pregnancy. Because the residue of CBZ in drinking water may provide a pathway to embryo and infant via intrauterine exposure or breast-feeding, the presence of CBZ in groundwater, and drinking water remains a significant concern warranting further systematic risk assessment studies.

4. Regulations

Strict regulations were introduced in the 1960s in many countries for pharmaceutical production [117]. For example, in Australia, the Therapeutic Goods Administration (TGA) undertakes assessments, similar to those of the US Food and Drug Administration (FDA), to “ensure that prescription and ‘over-the-counter’ medicines, medical devices, and related products, supplied in or exported from Australia, meet appropriate standards” [118].
Despite the release of a significant amount of PhACs into the environment, there is limited literature available on regulations for the presence of pharmaceuticals in water [117]. The regulatory framework set out for pharmaceuticals governs the quality and safety of use, rather than the health and environmental risks of long-term exposure to drinking water containing trace concentrations of pharmaceuticals. In the U.S., the FDA requires that an environmental assessment report be carried out when the expected concentration of the active ingredient of the pharmaceutical in the aquatic environment is equal to or higher than 1 µg/L. However, some state departments, such as the Minnesota Department of Health regulates that the CBZ concentration in drinking water must not exceed 40 µg/L [119]. Australian regulations require CBZ concentration in drinking water to be less than 100 µg/L [120]. In Europe, authorization for pharmaceutical production requires an environmental risk assessment [118]. In many countries, however, health risk-based standards and limit values for the presence of pharmaceuticals in drinking water have either not been set or are insufficient [111].

5. Biological Treatment Technologies for CBZ Removal

5.1. Activated Sludge Based Processes

Removal of PhACs by different biological treatment processes has been studied extensively [3,4]. In conventional activated sludge (CAS) processes, microorganisms generate energy by utilizing bulk organics present in wastewater as a primary source of food (also known as substrate). A part of this energy is used by the microorganisms for their cell growth and remaining energy is used for cell maintenance [121,122,123]. Since some PhACs such as antibiotics can be toxic to microorganisms and can inhibit their growth, an additional growth substrate (i.e., co-metabolism) is required to maintain microbial growth and diversity for adequate biodegradation [124,125].
The CAS process involves the application of microorganisms for the degradation of pollutants [16]. A membrane bioreactor (MBR) is an integration of the CAS process with an ultrafiltration (UF) or a microfiltration (MF) membrane for effective solid-liquid separation [3,126,127]. Removal of CBZ by CAS and MBR at different operating conditions such as hydraulic retention time (HRT), solids retention time (SRT) and initial concentration is presented in Table 5.
In the CAS process, a settling tank is used to separate the treated water from the sludge. In MBR, this solid-liquid separation is performed by filtration via MF or UF membranes. Effective retention of the activated sludge by the membrane in MBR decouples SRT from HRT, thereby allowing the operation of the activated sludge based bioreactor at higher mixed liquor suspended solid concentration (MLSS) and longer SRT [135,136]. It has been reported in several studies that MBR provides better aqueous phase removal of moderately biodegradable PhACs as compared to the CAS process [126,137]. For example, removal of the nonsteroidal anti-inflammatory drug diclofenac by MBR was 56%, while its removal was 26% in CAS [138]. Similarly, MBR achieved up to 20% better removal of another nonsteroidal anti-inflammatory drug naproxen [126]. However, CBZ removal by both CAS and MBR has been reported to be poor and unstable (Table 5). Poor removal of CBZ can be attributed to its physicochemical properties such as molecular structure and hydrophilicity [71,134,139,140].
Sorption onto the activated sludge can increase the overall removal of PhACs. CAS and MBR are observed to achieve high removal (>80%) for hydrophobic PhACs (log D > 3) but lower removal (typically < 20%) for hydrophilic PhACs [141,142]. Since CBZ is moderately hydrophobic, its removal via sorption onto activated sludge has been reported to range between 5 to 20% only [16,27]. This suggests that CBZ removal depends on its intrinsic biodegradability, which is governed by its molecular properties. In general, simple structured PhACs, especially without branched/multi chain groups, are readily degradable [134,143]. Moreover, PhACs containing an electron withdrawing functional group (EWG), such as carboxyl, halogen and amide, are resistant to biological treatment [134]. Indeed, CBZ contains an EWG (i.e., amide) that makes it resistant to biodegradation.
It is important to note that operating conditions such as SRT, HRT and MLSS concentration can also influence the removal of some PhACs by activated sludge [3,139]. However, because CBZ is a hardly biodegradable compound, available reports indicate limited influence of these parameters on CBZ removal. Zhang et al. [16] did not observe any CBZ removal by CAS even at an SRT of 100 days. Similarly, Radjenovic et al. [128], observed no improvement in CBZ removal following the increase in the SRT of both CAS and MBR. By contrast, Wijekoon et al. [27] achieved 40% removal of CBZ in MBR at a SRT of 88 days. Notwithstanding the fact that the experimental conditions may have been different in these studies, the observations here suggest that the removal and fate of CBZ during biological treatment processes depend on multiple factors.
The effect of redox conditions or dissolved oxygen on the removal of PhACs in MBR has been reported in a few studies [144,145,146]. In a study by Suarez et al. [147], PhACs were classified based on their biodegradation potential under aerobic and anoxic conditions: they observed that readily degradable PhACs such as fluoxetine and ibuprofen were biodegradable under both anoxic and aerobic conditions, while a few such as roxithromycin, naproxen, diclofenac and erythromycin were persistent in anoxic conditions but highly biodegradable under aerobic conditions. However, hydrophilic PhACs including CBZ were resistant to biodegradation in both aerobic and anoxic conditions [147], and a negligible difference in CBZ removal by sequential anoxic—aerobic MBR versus conventional aerobic MBR was noted. Notably, however, Hai et al. [139] reported that near-anoxic conditions (DO = 0.5 mg/L) can be a favorable operating regime for CBZ removal. They explained that ‘sequential anoxic-aerobic’ and ‘continuous near-anoxic (DO = 0.5 mg/L)’ operation modes were different. In the former, oxygen transfer from the aerated compartments to the anoxic zone due to the sludge recirculation may influence the removal efficiency [139].

5.2. White-Rot Fungi and Their Extracellular Enzymes

White-rot fungi (WRF) can degrade a variety of recalcitrant pollutants (e.g., poly-aromatic hydrocarbons and PhACs) that are poorly degraded by bacteria-dominated activated sludge [148,149,150,151,152]. In presence of a readily degradable substrate, WRF produce one or more type of extracellular enzymes such as laccase and lignin peroxidases (LiP). These enzymes catalyze the degradation of recalcitrant pollutants over a wide range of pH [153,154]. In addition to the extracellular enzymes, Golan-Rozen et al. [155] observed that the intracellular enzyme viz cytochrome P450 plays a vital role in CBZ degradation by whole-cell WRF. They demonstrated that the degradation of CBZ reduced from 99% to approximately 15% when cytochrome P450 was inhibited [155]. Whole-cell WRF and their extracellular enzymes have been studied extensively for enhanced removal of PhACs as depicted in Table 6.
Depending on the fugal species, efficient removal of CBZ has been achieved by whole-cell WRF in sterile batch bioreactors (Table 6). Since WRF species produce different combinations of extracellular enzymes, their performance of CBZ degradation might be different. For instance, Rodarte-Morales et al. [156] observed that Bjerkandera sp. R1 and Bjerkandera adusta both achieved almost complete removal (99%) of CBZ at an initial concentration of 1 mg/L in a batch bioreactor over an incubation time of 14 days. On the other hand, Trametes versicolor was reported to achieve less than 5% removal when CBZ was incubated in a whole-cell batch bioreactor at an initial concentration of 0.1 mg/L and an incubation time of 24 h [158]. Difference in performance was not only observed in case of different WRF species but also in different strains of a WRF species. For instance, Golan-Rozen et al. [155] studied the removal of CBZ by three different strains of Pleurotus ostreatus under identical operating conditions. They observed that P. ostreatus (PC9) achieved 99% CBZ removal, while a moderate removal (50–60%) was achieved by other two strains, namely P. ostreatus (Florida N001) and P. ostreatus (Florida F6).
Extracellular enzymes produced by WRF species have been studied for the removal of PhACs including CBZ in both batch and continuous-flow enzymatic bioreactors (Table 6). Degradation of PhACs by extracellular enzymes such as laccase occurs due to the transfer of a single electron from the pollutant to the active sites of the enzyme. Similar to the activated sludge based treatment process, the extent of degradation by an enzyme also depends on the molecular properties of the PhACs. Since CBZ contains a recalcitrant EWG (i.e., amide), its degradation by extracellular enzymes has been reported to range only between 5–15% [157,162]. High removal of CBZ in whole-cell fungal bioreactor as compared to enzymatic membrane bioreactor was explained by Golan-Rozen et al. [155]. They observed that the intracellular enzyme viz cytochrome P450 plays a vital role in CBZ degradation by whole-cell WRF. They demonstrated that the degradation of CBZ reduced from 99% to approximately 15% when cytochrome P450 was inhibited [155].
Performance of WRF for PhAC removal has been predominantly assessed under sterile conditions to avoid bacterial contamination [158,165,166]. This is because bacterial contamination under non-sterile conditions can negatively affect the performance of whole-cell WRF [154,167]. Indeed, poor removal of CBZ has been reported in fungal bioreactors operated under non-sterile conditions as compared to sterile fungal bioreactors [168,169,170]. For instance, Nguyen et al. [161] reported only 5% CBZ removal in a whole-cell fungal membrane bioreactor. In another study, no CBZ removal was observed in a non-sterile fluidized bed fungal bioreactor during the treatment of hospital wastewater [171]. To avoid bacterial contamination, a number of strategies such as fungal biomass replacement/renovation and pre-treatment of wastewater have been proposed. These strategies have been reviewed by Asif et al. [154].

6. CBZ Removal by Advanced Physicochemical Treatment Technologies

6.1. Performance of Nanofiltration and Reverse Osmosis Membranes

Nanofiltration (NF) and reverse osmosis (RO) are pressure-driven membrane filtration technologies [172,173]. They utilize semi-permeable membranes to primarily target the removal of dissolved contaminants. Both NF and RO have been studied for the removal of PhACs from secondary treated wastewater and freshwater, producing excellent quality effluent [173,174,175]. Several studies have shown that both NF and RO membranes can effectively retain CBZ, with a typical removal efficiency of greater than 95% [176,177]. Table 7 illustrates representative examples of CBZ removal by NF/RO membranes under a wide range of operating conditions.
Conceptually, NF membranes can retain PhACs via following mechanisms: (i) sorption of a solute on the membrane surface; (ii) size exclusion i.e., the sieving property of the membrane; and (iii) charge repulsion. However, electrostatic interaction cannot contribute to CBZ removal by the charged NF membrane, since CBZ remains neutral over a wide range of pH [14]. Hence, the molecular weight cut-off (MWCO) of NF/RO membranes is an important parameter for CBZ removal. In a study by Bellona et al. [177], efficient retention of CBZ by the RO membrane was attributed to size exclusion mechanism because the molecular weight of CBZ (i.e., 236 g/mole) was greater than the MWCO of the RO membrane. In a study by Comerton et al. [178], a loose NF membrane (MWCO = 400 g/mole) and a tight NF (MWCO = 200 g/mole) achieved 7 and 67% CBZ removal, respectively, thus, exemplifying the role of membrane MWCO in CBZ removal. In another study by Nghiem & Hawkes [185], efficient rejection of CBZ was reported for the NF90 membrane (MWCO < 200 g/mole) as compared to the NF270 membrane (MWCO > 300 g/mole).
Membrane fouling can affect the rejection of PhACs by NF membranes due to change in membrane surface properties (Table 7). Notably, CBZ rejection by the NF membranes is governed by the type of foulants. For instance, CBZ rejection by the NF270 membrane was reduced by 5 and 10% due to fouling caused by humic acid and sodium alginate, respectively [186]. This reduction in CBZ removal can be attributed to its diffusion into permeate following its adsorption on the fouling layer formed on the membrane surface. A more dramatic reduction in CBZ rejection (by 50%) was reported when MBR permeate (comprising multiple foulant materials) was fed to the NF-filtration system [186,187]. In general, the combination of membrane fouling and scaling can affect pollutant removal more severely. Chemical cleaning is performed to clean fouled membranes. However, recurrent chemical cleaning can affect membrane properties and, in turn, CBZ rejection. For example, significant reduction in CBZ rejection was observed due to cleaning the NF membrane by using caustic soda [184].
Natural organic matter (NOM) are ubiquitously present in surface water bodies. Since NF/RO processes are widely used for surface water treatment [188], the effectiveness of NF/RO for CBZ rejection in presence of NOM is vital. Comerton et al. [189] observed a statistically significant improvement in CBZ rejection during nanofiltration of pure water spiked with NOM. This is because CBZ, which is a moderately hydrophobic compound, can adsorb on NOM. However, a significant decrease in the rejection of CBZ was observed with concentration of the cations (calcium, magnesium, sodium) doubled. It has been hypothesized that increases in ionic strength and divalent cation concentration can cause conformational changes to NOM macromolecules. This may alter the presentation of sites for compound association leading to a reduction in NOM–compound complexation [189]. Therefore, the decrease in CBZ rejection in the natural waters with increase in cation concentration may be due to reduced association of CBZ with NOM.

6.2. Adsorption of CBZ by Activated Carbon

Activated carbons including granular activated carbon (GAC) and powdered activated carbon (PAC) are widely used as tertiary treatment processes primarily for color and odor removal from drinking water. GAC and PAC have also shown great potential for the removal of PhACs from secondary (i.e., biologically treated) wastewater [4,190,191]. Adsorption/removal of a pollutant by GAC/PAC is governed by the following mechanisms: (i) the electron donor–acceptor complex; (ii) the π–π dispersion interactions; (iii) hydrophobic interactions; and (iv) solvent effects that controls the solubility, reactivity and reaction kinetics [192,193]. The key properties of an adsorbent that can affect the efficacy of adsorption process include but are not limited to surface area, dose, surface chemistry and morphology, while water partitioning coefficient (log Kow), acid dissociation coefficient (pKa), molecular structure and size of the pollutants can influence the extent of adsorption by GAC/PAC [194]. In previous studies, efficient removal of PhACs has been achieved by GAC having larger pore size, because it can effectively adsorb pollutants with different shapes and size. Moreover, it was noted that pore volume has more influence on the adsorption of PhACs than specific area, and larger pore volume can achieve higher removal efficiency [195,196]. Representative examples of CBZ removal by GAC/PAC from a variety of water matrix (e.g., surface water and MBR effluent) is provided in Table 8.
Since CBZ is neutral at pH ranging from 0–14, its removal by activated carbon is governed by hydrophobic interaction that depends on water partitioning coefficient [4,17]. In a study Yu et al. [17], better adsorption of CBZ (log Kow = 2.45) by activated carbon as compared to naproxen (log Kow = 3.18) and 4-n-nonylphenol (log Kow = 5.8) was reported. Better adsorption of CBZ can be attributed to influence of pH on hydrophobicity of ionizable micropollutants [3]. Indeed, Yu et al. [17] demonstrated that activated carbon achieved better CBZ (log Kow = 2.45) removal as compared to naproxen because actual log Kow value for naproxen was 0.89 at the operating pH (i.e., 6.4). On the other hand, 4-n-nonylphenol contains both hydrophobic and hydrophilic functional groups in its molecule. The hydrophilic groups of 4-n-nonylphenol can affect its adsorption by activated carbon [208,209]. Therefore, better adsorption of other micropollutants including CBZ as compared to 4-n-nonylphenol is possible [17].
Effectiveness of GAC/PAC was also investigated at pilot-scale plants treating secondary effluent. For instance, Ternes et al. [197] observed almost complete removal (>99%) of CBZ in a pilot-scale GAC plant. In another study, CBZ removal was 95% in pilot-scale GAC/PAC systems treating secondary effluent at an initial activated carbon dose of 30–100 mg/L [190]. GAC was observed to provide better removal of PhACs including CBZ as compared to two other water treatment processes, namely, ozonation and sand filtration [198].
Increasing activated carbon dose can improve CBZ removal. For instance, CBZ removal improved from 36 to 97% when the PAC dose was increased seven folds [205]. In a study by Nguyen et al. [210], instead of using PAC as a post-treatment, it was added directly to the mixed liquor of an MBR. In that study, CBZ removal improved from 50% to 90% following the increase in PAC dose to MBR from 0.1 to 0.5 g/L [210].
Compared to GAC, PAC has a larger surface area that can conceptually provide faster reaction kinetics and better removal efficiency. However, survey of the available literature suggests that both GAC and PAC are effective for CBZ removal (Table 8). Notably, to date the performance of PAC and GAC has been assessed in short term experiments [204,211]. Since saturation of binding sites reduces the removal of pollutants over time, research is required to investigate PAC/GAC regeneration aspects. Grover et al. [202] monitored the performance of a full-scale post-treatment GAC plant over a period of seven months. They observed that CBZ removal reduced over time, and GAC plant could only achieve 30% removal of CBZ during long term operation. In addition to the saturation of GAC binding sites, impurities such as humic substances can compete for GAC/PAC binding sites that may result in ineffective CBZ removal [4,202].

6.3. CBZ Degradation by Advanced Oxidation Processes

Due to the molecular properties of CBZ, conventional biological processes are not effective for its removal (see Section 5). On the other hand, despite the effective removal of CBZ by NF/RO membrane filtration and activated carbon adsorption (Table 7 and Table 8), an additional step is required for the treatment of the produced concentrate. In this context it is noteworthy that advanced oxidation processes (AOP) may achieve effective degradation of CBZ (Table 9). Post treatment of biologically treated wastewater by AOPs may simultaneously achieve disinfection and PhAC removal [212].
Formation of hydroxyl radicals (OH) are mainly responsible for the degradation of PhACs by AOPs, while formation of ozone radicals (O3) in ozonation process can also contribute to the degradation process [213,214]. Some PhACs such as naproxen are degraded by both OH and O3 radicals, while some are only susceptible to degradation by OH radicals. CBZ, which is resistant to biological treatment, is effectively degraded by both OH and O3 radicals. For instance, Ternes et al. [197] reported almost complete removal (>99%) of CBZ (35-1000 ng/L) during ozonation process at an initial dose of 0.5 mg/L.
Although UV photlysis alone has been observed to be ineffective for CBZ removal (0–20%) in a number of studies [220,231], removal can be significantly improved by adding H2O2 or a photocatalyst such as TiO2 [214,220]. For instance, adding a single dose of H2O2 (5–15%) to UV photolysis process resulted in an enhanced CBZ removal of 60–75% [225,226]. In another study, H2O2-concentration dependent increase in CBZ removal by UV/H2O2 process was reported, and 99% removal was achieved at an initial H2O2 concentration of 120 mg/L [220]. CBZ removal can improve with increasing H2O2 concentration but it will reach a plateau beyond a threshold H2O2 concentration [220].
Fenton’s reagent has been reported to efficiently oxidize (>99%) CBZ (Table 9). Notably, Fe+2 based Fenton process achieved better CBZ removal than UV alone, UV/TiO2 and UV/H2O2. However, the requirement of acidic conditions for Fenton’s process is a considerable drawback for its practical application [232].
AOP are undoubtedly very efficient for the removal of CBZ but their practical applications are constrained by associated high cost of chemicals. Moreover, transformation products formed following the oxidation of PhACs including CBZ can be more toxic than the parent compound [220]. To overcome this issue, the use of biological filters or ACs can be a suitable option [233,234].

6.4. Combined/Hybrid Treatment Systems

A single treatment option may not be universally applicable, so a combined (sequential) or integrated treatment may be more effective for CBZ removal. Combined or hybrid treatment options are also conceptually beneficial, usually leading to improved treatment efficiencies [235,236,237]. Examples of possible combined/hybrid water treatment processes include membrane filtration followed by activated carbon, MBR followed by activated carbon, activated carbon adsorption followed by UV, integrated MBR-TiO2 photocatalysis, and integrated MBR-PAC adsorption. A summary of CBZ removal by combined/hybrid treatment systems is presented in Table 10.
Kleywegt et al. [244] investigated the removal of CBZ by GAC adsorption followed by UV photolysis that achieved effective removal of CBZ (93%). Serrano et al. [203] investigated the removal of PhACs including CBZ in a CAS process followed by GAC adsorption. They reported that adsorption onto activated carbon improved the overall removal of CBZ by as much as 40% [203]. Nguyen et al. [210] also reported an overall CBZ removal of 98% by an MBR followed by GAC adsorption. When a sequencing batch reactor coupled to an external microfiltration membrane was investigated by Serrano et al. [241], up to 93% CBZ removal was achieved following the addition of a single dose (1 g/L) of PAC directly into the bioreactor. Likewise, the integrated MBR-PAC system was also reported to to achieve 92% CBZ removal [211]. CBZ removal by MBR treatment followed by GAC filtration was investigated by Nguyen et al. [210]. While MBR alone showed a removal of less than 20%, MBR-GAC achieved an extremely efficient removal of 98% [210]. This result demonstrates that GAC post-treatment could significantly improve the removal of PhACs, which are resistant to degradation by the activated sludge.
Other integrated/hybrid technologies have also shown effective CBZ removal. AOPs cannot mineralize PhACs, however, it is important to note that the metabolites formed following the oxidation of CBZ have been reported to be readily mineralized by the activated sludge [245,246,247]. For instance, Hübner et al. [243] studied CBZ removal by combining ozonation with a sand column mimicking a soil aquifer treatment (SAT) systems. They observed that the degradation products of CBZ formed after ozonation were significantly mineralized (>80%) in the sand column at an HRT of 5-6 days [243]. In another study [239], a combination of gamma radiation and activated sludge based biological treatment achieved up to 79% CBZ mineralization. They reported that significant CBZ removal (>99%) was mainly achieved by gamma radiation at an initial dose of 800 Gy, while activated sludge was responsible for mineralization of CBZ [239]. In a study by Laeraet al. [238], an integrated MBR-TiO2/UV system achieved up to 95% CBZ removal from pharmaceutical industrial effluent, showing that the integration of biological and chemical oxidation processes can be an effective strategy for enhanced CBZ removal. Despite the efficacy of combined/integrated oxidation and biological processes for mineralization, the cost associated with the application of oxidation processes needs to be considered.
Carbamazepine has been consistently shown to be poorly removed by coagulation despite being a neutral compound [21,197,248,249,250]. However, a properly designed coagulation/flocculation unit can efficiently remove suspended solids and can thereby enhance the performance of a subsequent activated carbon adsorption unit by reducing competitive adsorption [151]. Coagulation pre-treatment has been found to significantly enhance carbamazepine removal efficiency by adsorption [197]. Because high concentrations of suspended or colloidal solids in the wastewater may impede the advanced oxidation processes, sufficient prior removal of these materials by a physicochemical treatment such as coagulation is required [251,252].

7. Fate and Metabolites of CBZ

CBZ transformation products following its metabolism in human body or following its degradation by different treatment processes have been reported (Figure 1). Figure 1 sheds light on different pathways of CBZ metabolism/degradation by AOPs, human liver, and microorganisms. Huerta-Fontela et al. [199] reported that CBZ degradation by ozonation occurs via a ring opening mechanism due to the attack of ozone on the non-aromatic carbon-carbon double bond of CBZ, forming the metabolite epoxy-carbamazepine. Compared to the CBZ degradation products reported by McDowell et al. [253] in UV photolysis (Figure 1), Vogna et al. [187] observed that the addition of H2O2 to UV photolysis yields different degradation products (Figure 1), indicating a difference in the degradation pathway of CBZ in presence of H2O2. In biological systems (such as human liver, fungus and activated sludge), CBZ degradation products usually contain the azepine structure (Figure 1).
As depicted in Figure 1, several metabolites/degradation products have been reported for CBZ treatment by different processes. Toxicity of CBZ degradation products has been reported in a number of studies as summarized here. Based on the Yeast Estrogen Screen (YES) assay, Mohapatra et al. [257] reported that CBZ and its degradation products showed no estrogenic activity. In another study, Jelic et al. [160] investigated the toxicity of the media in a pulsed fluidized bed bioreactor containing whole-cell T. versicolor and CBZ at an initial concentration of 500 µg/L. The acute toxicity test (Microtox) showed the toxicity induced by CBZ was reduced from 95% to 24% after an incubation time of 10 days, suggesting that the degradation products were non-toxic [258]. Similarly, degradation products following the oxidation of CBZ by ozonation, UV and UV/H2O2 processes exhibited no genotoxic, cytotoxic or estrogenic effects [258]. CBZ metabolites including 10,11-dihydro-trans-10,11-dihydroxy-carbamazepine, acidone and acridine formed during CAS process [256] have also been reported to be non-toxic [258,259]. Although degradation products or metabolites of CBZ formed during CAS and AOPs are non-toxic, Bu et al. [255] reported that CBZ-2,3-arene (one oxide intermediate) is believed to cause idiosyncratic effect after CBZ consumption for medicinal purposes [255].

8. Conclusions and Outlook

Although its occurrence in freshwater may not pose an immediate threat to aquatic ecosystems or human health, effective removal of CBZ is still required for safe water reuse applications and drinking water treatment. Biological wastewater treatment processes such as conventional activated sludge and membrane bioreactor are not effective for CBZ removal due to its resistance to biodegradation. However, advanced wastewater treatment processes seem to be effective for efficient CBZ removal. For instance, CBZ removal by the reverse osmosis and nanofiltration membranes is above 90%. Similarly, post-treatment with granular and powdered activated carbon provides efficient CBZ removal ranging from 90–99%. However, membrane fouling in case of membrane technologies and regeneration of activated carbon are obstacles that warrant technical solutions. Depending on the type of fungal species and operating condition, white-rot fungi can achieve almost complete removal of CBZ but bacterial contamination may affect the efficacy of fungal bioreactors during long term operations. Although advanced oxidation processes such as ozonation, UV/H2O2, UV/TiO2 and Fenton processes are effective for CBZ removal, costs associated with the addition of chemicals, and separation of catalysts needs to be carefully considered. Finally, the literature to date suggests that degradation products formed following the degradation of CBZ by biological and chemical oxidation processes may not induce toxic effects on aquatic ecosystems.

Acknowledgments

This research has been conducted with the support of the Australian Government Research Training Program Scholarship. This study was partially funded by the GeoQuEST Research Centre and Faculty of EIS strategic partnership grant, University of Wollongong, Australia.

Author Contributions

F.I.H. conceived and led the project. F.I.H., M.B.A. and S.Y. planned and conducted the literature survey and prepared the manuscript in consultation with the coauthors. The co-authors contributed to specific sections of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hughes, S.R.; Kay, P.; Brown, L.E. Global synthesis and critical evaluation of pharmaceutical data sets collected from river systems. Environ. Sci. Technol. 2012, 47, 661–677. [Google Scholar] [CrossRef] [PubMed]
  2. Gavrilescu, M.; Demnerová, K.; Aamand, J.; Agathos, S.; Fava, F. Emerging pollutants in the environment: Present and future challenges in biomonitoring, ecological risks and bioremediation. New Biotechnol. 2015, 32, 147–156. [Google Scholar] [CrossRef] [PubMed]
  3. Hai, F.I.; Nghiem, L.D.; Khan, S.J.; Price, W.E.; Yamamoto, K. Wastewater reuse: Removal of emerging trace organic contaminants. In Membrane biological Reactors: Theory, Modeling, Design, Management and Applications to Wastewater Reuse; Hai, F.I., Yamamoto, K., Lee, C., Eds.; IWA publishing: London, UK, 2014; pp. 165–205. ISBN 9781780400655. [Google Scholar]
  4. 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, 619–641. [Google Scholar] [CrossRef] [PubMed]
  5. Lapworth, D.; Baran, N.; Stuart, M.; Ward, R. Emerging organic contaminants in groundwater: A review of sources, fate and occurrence. Environ. Pollut. 2012, 163, 287–303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Vieno, N.; Tuhkanen, T.; Kronberg, L. Elimination of pharmaceuticals in sewage treatment plants in finland. Water Res. 2007, 41, 1001–1012. [Google Scholar] [CrossRef] [PubMed]
  7. Brack, W.; Dulio, V.; Ågerstrand, M.; Allan, I.; Altenburger, R.; Brinkmann, M.; Bunke, D.; Burgess, R.M.; Cousins, I.; Escher, B.I. Towards the review of the european union water framework directive: Recommendations for more efficient assessment and management of chemical contamination in european surface water resources. Sci. Total Environ. 2017, 576, 720–737. [Google Scholar] [CrossRef] [PubMed]
  8. Kumar, A.; Batley, G.E.; Nidumolu, B.; Hutchinson, T.H. Derivation of water quality guidelines for priority pharmaceuticals. Environ. Toxicol. Chem. 2016, 35, 1815–1824. [Google Scholar] [CrossRef] [PubMed]
  9. Osorio, M.V.; Reis, S.; Lima, J.L.; Segundo, M.A. Analytical features of diclofenac evaluation in water as a potential marker of anthropogenic pollution. Curr. Pharm. Anal. 2017, 13, 39–47. [Google Scholar] [CrossRef]
  10. Arye, G.; Dror, I.; Berkowitz, B. Fate and transport of carbamazepine in soil aquifer treatment (sat) infiltration basin soils. Chemosphere 2011, 82, 244–252. [Google Scholar] [CrossRef] [PubMed]
  11. Ferrer, I.; Thurman, E.M. Analysis of 100 pharmaceuticals and their degradates in water samples by liquid chromatography/quadrupole time-of-flight mass spectrometry. J. Chromatogr. A 2012, 1259, 148–157. [Google Scholar] [CrossRef] [PubMed]
  12. Wick, A.; Fink, G.; Joss, A.; Siegrist, H.; Ternes, T.A. Fate of beta blockers and psycho-active drugs in conventional wastewater treatment. Water Res. 2009, 43, 1060–1074. [Google Scholar] [CrossRef] [PubMed]
  13. Alvarino, T.; Suarez, S.; Lema, J.; Omil, F. Understanding the removal mechanisms of ppcps and the influence of main technological parameters in anaerobic uasb and aerobic cas reactors. J. Hazard. Mater. 2014, 278, 506–513. [Google Scholar] [CrossRef] [PubMed]
  14. Siegrist, H.; Joss, A. Review on the fate of organic micropollutants in wastewater treatment and water reuse with membranes. Water Sci. Technol. 2012, 66, 1369–1376. [Google Scholar] [CrossRef] [PubMed]
  15. Zhou, S.; Xia, Y.; Li, T.; Yao, T.; Shi, Z.; Zhu, S.; Gao, N. Degradation of carbamazepine by uv/chlorine advanced oxidation process and formation of disinfection by-products. Environ. Sci. Pollut. Res. 2016, 23, 16448–16455. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, Y.; Geißen, S.-U.; Gal, C. Carbamazepine and diclofenac: Removal in wastewater treatment plants and occurrence in water bodies. Chemosphere 2008, 73, 1151–1161. [Google Scholar] [CrossRef] [PubMed]
  17. Yu, Z.; Peldszus, S.; Huck, P.M. Adsorption characteristics of selected pharmaceuticals and an endocrine disrupting compound-naproxen, carbamazepine and nonylphenol-on activated carbon. Water Res. 2008, 42, 2873–2882. [Google Scholar] [CrossRef] [PubMed]
  18. Deblonde, T.; Cossu-Leguille, C.; Hartemann, P. Emerging pollutants in wastewater: A review of the literature. Int. J. Hyg. Environ. Health 2011, 214, 442–448. [Google Scholar] [CrossRef] [PubMed]
  19. Bolong, N.; Ismail, A.; Salim, M.R.; Matsuura, T. A review of the effects of emerging contaminants in wastewater and options for their removal. Desalination 2009, 239, 229–246. [Google Scholar] [CrossRef]
  20. Verlicchi, P.; Al Aukidy, M.; Zambello, E. Occurrence of pharmaceutical compounds in urban wastewater: Removal, mass load and environmental risk after a secondary treatment—A review. Sci. Total Environ. 2012, 429, 123–155. [Google Scholar] [CrossRef] [PubMed]
  21. Alexander, J.T.; Hai, F.I.; Al-aboud, T.M. Chemical coagulation-based processes for trace organic contaminant removal: Current state and future potential. J. Environ. Manag. 2012, 111, 195–207. [Google Scholar] [CrossRef] [PubMed]
  22. Kasprzyk-Hordern, B.; Dinsdale, R.M.; Guwy, A.J. Illicit drugs and pharmaceuticals in the environment—Forensic applications of environmental data. Part 1: Estimation of the usage of drugs in local communities. Environ. Pollut. 2009, 157, 1773–1777. [Google Scholar] [CrossRef] [PubMed]
  23. Singer, H.; Jaus, S.; Hanke, I.; Lück, A.; Hollender, J.; Alder, A.C. Determination of biocides and pesticides by on-line solid phase extraction coupled with mass spectrometry and their behaviour in wastewater and surface water. Environ. Pollut. 2010, 158, 3054–3064. [Google Scholar] [CrossRef] [PubMed]
  24. Dvory, N.Z.; Kuznetsov, M.; Livshitz, Y.; Gasser, G.; Pankratov, I.; Lev, O.; Adar, E.; Yakirevich, A. Modeling sewage leakage and transport in carbonate aquifer using carbamazepine as an indicator. Water Res. 2018, 128, 157–170. [Google Scholar] [CrossRef] [PubMed]
  25. Ying, G.-G.; Kookana, R.S.; Kolpin, D.W. Occurrence and removal of pharmaceutically active compounds in sewage treatment plants with different technologies. J. Environ. Monit. 2009, 11, 1498–1505. [Google Scholar] [CrossRef] [PubMed]
  26. Tixier, C.; Singer, H.P.; Oellers, S.; Müller, S.R. Occurrence and fate of carbamazepine, clofibric acid, diclofenac, ibuprofen, ketoprofen, and naproxen in surface waters. Environ. Sci. Technol. 2003, 37, 1061–1068. [Google Scholar] [CrossRef] [PubMed]
  27. Wijekoon, K.C.; Hai, F.I.; Kang, J.; Price, W.E.; Guo, W.; Ngo, H.H.; Nghiem, L.D. The fate of pharmaceuticals, steroid hormones, phytoestrogens, uv-filters and pesticides during mbr treatment. Bioresour. Technol. 2013, 144, 247–254. [Google Scholar] [CrossRef] [PubMed]
  28. Guo, Y.C.; Krasner, S.W. Occurence of primidone, carbamazepine, caffeine and precursors for n-nitrosodimethylamine in drinding water sources impacted by wastewater. J. Am. Water Resour. Assoc. 2009, 45, 58–67. [Google Scholar] [CrossRef]
  29. Heberer, T. Occurrence, fate, and removal of pharmaceutical residues in the aquatic environment: A review of recent research data. Toxicol. Lett. 2002, 131, 5–17. [Google Scholar] [CrossRef]
  30. Stuart, M.; Lapworth, D.; Crane, E.; Hart, A. Review of risk from potential emerging contaminants in UK groundwater. Sci. Total Environ. 2012, 416, 1–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Metcalfe, C.D.; Miao, X.-S.; Koenig, B.G.; Struger, J. Distribution of acidic and neutral drugs in surface waters near sewage treatment plants in the lower great lakes, canada. Environ. Toxicol. Chem. 2003, 22, 2881–2889. [Google Scholar] [CrossRef] [PubMed]
  32. Miao, X.-S.; Metcalfe, C.D. Determination of carbamazepine and its metabolites in aqueous samples using liquid chromatography−electrospray tandem mass spectrometry. Anal. Chem. 2003, 75, 3731–3738. [Google Scholar] [CrossRef] [PubMed]
  33. Miao, X.-S.; Yang, J.-J.; Metcalfe, C.D. Carbamazepine and its metabolites in wastewater and in biosolids in a municipal wastewater treatment plant. Environ. Sci. Technol. 2005, 39, 7469–7475. [Google Scholar] [CrossRef] [PubMed]
  34. Gagné, F.; Blaise, C.; André, C. Occurrence of pharmaceutical products in a municipal effluent and toxicity to rainbow trout (Oncorhynchus mykiss) hepatocytes. Ecotoxicol. Environ. Saf. 2006, 64, 329–336. [Google Scholar] [CrossRef] [PubMed]
  35. Yu, Z.; Peldszus, S.; Huck, P.M. Optimizing gas chromatographic–mass spectrometric analysis of selected pharmaceuticals and endocrine-disrupting substances in water using factorial experimental design. J. Chromatogr. A 2007, 1148, 65–77. [Google Scholar] [CrossRef] [PubMed]
  36. Gottschall, N.; Topp, E.; Metcalfe, C.; Edwards, M.; Payne, M.; Kleywegt, S.; Russell, P.; Lapen, D.R. Pharmaceutical and personal care products in groundwater, subsurface drainage, soil, and wheat grain, following a high single application of municipal biosolids to a field. Chemosphere 2012, 87, 194–203. [Google Scholar] [CrossRef] [PubMed]
  37. Edwards, M.; Topp, E.; Metcalfe, C.D.; Li, H.; Gottschall, N.; Bolton, P.; Curnoe, W.; Payne, M.; Beck, A.; Kleywegt, S.; et al. Pharmaceutical and personal care products in tile drainage following surface spreading and injection of dewatered municipal biosolids to an agricultural field. Sci. Total Environ. 2009, 407, 4220–4230. [Google Scholar] [CrossRef] [PubMed]
  38. Hummel, D.; Löffler, D.; Fink, G.; Ternes, T.A. Simultaneous determination of psychoactive drugs and their metabolites in aqueous matrices by liquid chromatography mass spectrometry. Environ. Sci. Technol. 2006, 40, 7321–7328. [Google Scholar] [CrossRef] [PubMed]
  39. Ternes, T.A.; Stüber, J.; Herrmann, N.; McDowell, D.; Ried, A.; Kampmann, M.; Teiser, B. Ozonation: A tool for removal of pharmaceuticals, contrast media and musk fragrances from wastewater? Water Res. 2003, 37, 1976–1982. [Google Scholar] [CrossRef]
  40. Ternes, T.A. Occurrence of drugs in german sewage treatment plants and rivers. Water Res. 1998, 32, 3245–3260. [Google Scholar] [CrossRef]
  41. Wiegel, S.; Aulinger, A.; Brockmeyer, R.; Harms, H.; Löffler, J.; Reincke, H.; Schmidt, R.; Stachel, B.; von Tümpling, W.; Wanke, A. Pharmaceuticals in the river elbe and its tributaries. Chemosphere 2004, 57, 107–126. [Google Scholar] [CrossRef] [PubMed]
  42. Musolff, A.; Leschik, S.; Möder, M.; Strauch, G.; Reinstorf, F.; Schirmer, M. Temporal and spatial patterns of micropollutants in urban receiving waters. Environ. Pollut. 2009, 157, 3069–3077. [Google Scholar] [CrossRef] [PubMed]
  43. Osenbrück, K.; Gläser, H.-R.; Knöller, K.; Weise, S.M.; Möder, M.; Wennrich, R.; Schirmer, M.; Reinstorf, F.; Busch, W.; Strauch, G. Sources and transport of selected organic micropollutants in urban groundwater underlying the city of halle (saale), Germany. Water Res. 2007, 41, 3259–3270. [Google Scholar] [CrossRef] [PubMed]
  44. Okuda, T.; Kobayashi, Y.; Nagao, R.; Yamashita, N.; Tanaka, H.; Tanaka, S.; Fujii, S.; Konishi, C.; Houwa, I. Removal efficiency of 66 pharmaceuticals during wastewater treatment process in Japan. Water Sci. Technol. 2008, 57, 65–71. [Google Scholar] [CrossRef] [PubMed]
  45. Nakada, N.; Komori, K.; Suzuki, Y.; Konishi, C.; Houwa, I.; Tanaka, H. Occurrence of 70 pharmaceutical and personal care products in tone river basin in Japan. Water Sci. Technol. 2007, 56, 133–140. [Google Scholar] [CrossRef] [PubMed]
  46. Nakada, N.; Kiri, K.; Shinohara, H.; Harada, A.; Kuroda, K.; Takizawa, S.; Takada, H. Evaluation of pharmaceuticals and personal care products as water-soluble molecular markers of sewage. Environ. Sci. Technol. 2008, 42, 6347–6353. [Google Scholar] [CrossRef] [PubMed]
  47. Kuroda, K.; Murakami, M.; Oguma, K.; Muramatsu, Y.; Takada, H.; Takizawa, S. Assessment of groundwater pollution in tokyo using ppcps as sewage markers. Environ. Sci. Technol. 2011, 46, 1455–1464. [Google Scholar] [CrossRef] [PubMed]
  48. Choi, K.; Kim, Y.; Park, J.; Park, C.K.; Kim, M.; Kim, H.S.; Kim, P. Seasonal variations of several pharmaceutical residues in surface water and sewage treatment plants of Han River, Korea. Sci. Total Environ. 2008, 405, 120–128. [Google Scholar] [CrossRef] [PubMed]
  49. Behera, S.K.; Kim, H.W.; Oh, J.-E.; Park, H.-S. Occurrence and removal of antibiotics, hormones and several other pharmaceuticals in wastewater treatment plants of the largest industrial city of Korea. Sci. Total Environ. 2011, 409, 4351–4360. [Google Scholar] [CrossRef] [PubMed]
  50. Kim, J.-W.; Jang, H.-S.; Kim, J.-G.; Ishibashi, H.; Hirano, M.; Nasu, K.; Ichikawa, N.; Takao, Y.; Shinohara, R.; Arizono, K. Occurrence of pharmaceutical and personal care products (ppcps) in surface water from Mankyung River, South Korea. J. Health Sci. 2009, 55, 249–258. [Google Scholar] [CrossRef]
  51. Yoon, Y.; Ryu, J.; Oh, J.; Choi, B.-G.; Snyder, S.A. Occurrence of endocrine disrupting compounds, pharmaceuticals, and personal care products in the Han river (Seoul, South Korea). Sci. Total Environ. 2010, 408, 636–643. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, H.-C.; Wang, P.-L.; Ding, W.-H. Using liquid chromatography–ion trap mass spectrometry to determine pharmaceutical residues in taiwanese rivers and wastewaters. Chemosphere 2008, 72, 863–869. [Google Scholar] [CrossRef] [PubMed]
  53. Lin, W.-C.; Chen, H.-C.; Ding, W.-H. Determination of pharmaceutical residues in waters by solid-phase extraction and large-volume on-line derivatization with gas chromatography–mass spectrometry. J. Chromatogr. A 2005, 1065, 279–285. [Google Scholar] [CrossRef] [PubMed]
  54. 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]
  55. Zhang, Z.L.; Zhou, J.L. Simultaneous determination of various pharmaceutical compounds in water by solid-phase extraction–liquid chromatography–tandem mass spectrometry. J. Chromatogr. A 2007, 1154, 205–213. [Google Scholar] [CrossRef] [PubMed]
  56. Zhou, J.L.; Zhang, Z.L.; Banks, E.; Grover, D.; Jiang, J.Q. Pharmaceutical residues in wastewater treatment works effluents and their impact on receiving river water. J. Hazard. Mater. 2009, 166, 655–661. [Google Scholar] [CrossRef] [PubMed]
  57. Kasprzyk-Hordern, B.; Dinsdale, R.M.; Guwy, A.J. The occurrence of pharmaceuticals, personal care products, endocrine disruptors and illicit drugs in surface water in south wales, UK. Water Res. 2008, 42, 3498–3518. [Google Scholar] [CrossRef] [PubMed]
  58. Kasprzyk-Hordern, B.; Dinsdale, R.M.; Guwy, A.J. Multi-residue method for the determination of basic/neutral pharmaceuticals and illicit drugs in surface water by solid-phase extraction and ultra performance liquid chromatography–positive electrospray ionisation tandem mass spectrometry. J. Chromatogr. A 2007, 1161, 132–145. [Google Scholar] [CrossRef] [PubMed]
  59. Stuart, M.E.; Manamsa, K.; Talbot, J.C.; Crane, E.J. Emerging Contaminants in Groundwater, 2011. Available online: https://nora.nerc.ac.uk/14557/ (accessed on 15 October 2017).
  60. Lapworth, D.; Stuart, M.; Hart, A.; Crane, E.; Baran, N. Emerging Contaminants in Groundwater; Groundwater Forum: London, UK, 2011; Available online: http://nora.nerc.ac.uk/14093/ (accessed on 15 October 2017).
  61. Spongberg, A.L.; Witter, J.D. Pharmaceutical compounds in the wastewater process stream in Northwest Ohio. Sci. Total Environ. 2008, 397, 148–157. [Google Scholar] [CrossRef] [PubMed]
  62. Vanderford, B.J.; Snyder, S.A. Analysis of pharmaceuticals in water by isotope dilution liquid chromatography/tandem mass spectrometry. Environ. Sci. Technol. 2006, 40, 7312–7320. [Google Scholar] [CrossRef] [PubMed]
  63. Glassmeyer, S.T.; Furlong, E.T.; Kolpin, D.W.; Cahill, J.D.; Zaugg, S.D.; Werner, S.L.; Meyer, M.T.; Kryak, D.D. Transport of chemical and microbial compounds from known wastewater discharges:  Potential for use as indicators of human fecal contamination. Environ. Sci. Technol. 2005, 39, 5157–5169. [Google Scholar] [CrossRef] [PubMed]
  64. Kolpin, D.W.; Skopec, M.; Meyer, M.T.; Furlong, E.T.; Zaugg, S.D. Urban contribution of pharmaceuticals and other organic wastewater contaminants to streams during differing flow conditions. Sci. Total Environ. 2004, 328, 119–130. [Google Scholar] [CrossRef] [PubMed]
  65. Conley, J.M.; Symes, S.J.; Kindelberger, S.A.; Richards, S.M. Rapid liquid chromatography–tandem mass spectrometry method for the determination of a broad mixture of pharmaceuticals in surface water. J. Chromatogr. A 2008, 1185, 206–215. [Google Scholar] [CrossRef] [PubMed]
  66. Katz, B.G.; Griffin, D.W.; Davis, J.H. Groundwater quality impacts from the land application of treated municipal wastewater in a large karstic spring basin: Chemical and microbiological indicators. Sci. Total Environ. 2009, 407, 2872–2886. [Google Scholar] [CrossRef] [PubMed]
  67. Fram, M.S.; Belitz, K. Occurrence and concentrations of pharmaceutical compounds in groundwater used for public drinking-water supply in california. Sci. Total Environ. 2011, 409, 3409–3417. [Google Scholar] [CrossRef] [PubMed]
  68. Jelić, A.; Gros, M.; Petrović, M.; Ginebreda, A.; Barceló, D. Occurrence and elimination of pharmaceuticals during conventional wastewater treatment. In Emerging and Priority Pollutants in Rivers; Guasch, H., Ginebreda, A., Geiszinger, A., Eds.; Springer: Berlin, Germany, 2012; pp. 1–23. ISBN 9783642257223. [Google Scholar]
  69. Petrovic, M.; de Alda, M.J.L.; Diaz-Cruz, S.; Postigo, C.; Radjenovic, J.; Gros, M.; Barcelo, D. Fate and removal of pharmaceuticals and illicit drugs in conventional and membrane bioreactor wastewater treatment plants and by riverbank filtration. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2009, 367, 3979–4003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Al-Rifai, J.H.; Gabelish, C.L.; Schäfer, A.I. Occurrence of pharmaceutically active and non-steroidal estrogenic compounds in three different wastewater recycling schemes in Australia. Chemosphere 2007, 69, 803–815. [Google Scholar] [CrossRef] [PubMed]
  71. Clara, M.; Strenn, B.; Gans, O.; Martinez, E.; Kreuzinger, N.; Kroiss, H. Removal of selected pharmaceuticals, fragrances and endocrine disrupting compounds in a membrane bioreactor and conventional wastewater treatment plants. Water Res. 2005, 39, 4797–4807. [Google Scholar] [CrossRef] [PubMed]
  72. Clara, M.; Strenn, B.; Kreuzinger, N. Carbamazepine as a possible anthropogenic marker in the aquatic environment: Investigations on the behaviour of carbamazepine in wastewater treatment and during groundwater infiltration. Water Res. 2004, 38, 947–954. [Google Scholar] [CrossRef] [PubMed]
  73. Vieno, N.M.; Tuhkanen, T.; Kronberg, L. Analysis of neutral and basic pharmaceuticals in sewage treatment plants and in recipient rivers using solid phase extraction and liquid chromatography–tandem mass spectrometry detection. J. Chromatogr. A 2006, 1134, 101–111. [Google Scholar] [CrossRef] [PubMed]
  74. Joss, A.; Keller, E.; Alder, A.C.; Göbel, A.; McArdell, C.S.; Ternes, T.; Siegrist, H. Removal of pharmaceuticals and fragrances in biological wastewater treatment. Water Res. 2005, 39, 3139–3152. [Google Scholar] [CrossRef] [PubMed]
  75. Vochezer, K. Modelling of Carbamazepine and Diclofenac in a River Network—Photolytic Degradation in Swiss Rivers; Swedish University of Agricultural Sciences: Uppsala, Sweden, 2010. [Google Scholar]
  76. Modick, H.; Weiss, T.; Dierkes, G.; Brüning, T.; Koch, H.M. Ubiquitous presence of paracetamol in human urine: Sources and implications. Reproduction 2014, 147, R105–R117. [Google Scholar] [CrossRef] [PubMed]
  77. Camacho-Muñoz, D.; Martín, J.; Santos, J.L.; Aparicio, I.; Alonso, E. Concentration evolution of pharmaceutically active compounds in raw urban and industrial wastewater. Chemosphere 2014, 111, 70–79. [Google Scholar] [CrossRef] [PubMed]
  78. Rivera-Jaimes, J.A.; Postigo, C.; Melgoza-Alemán, R.M.; Aceña, J.; Barceló, D.; de Alda, M.L. Study of pharmaceuticals in surface and wastewater from cuernavaca, morelos, mexico: Occurrence and environmental risk assessment. Sci. Total Environ. 2018, 613, 1263–1274. [Google Scholar] [CrossRef] [PubMed]
  79. Yu, Y.; Wu, L.; Chang, A.C. Seasonal variation of endocrine disrupting compounds, pharmaceuticals and personal care products in wastewater treatment plants. Sci. Total Environ. 2013, 442, 310–316. [Google Scholar] [CrossRef] [PubMed]
  80. Al Aukidy, M.; Verlicchi, P.; Jelic, A.; Petrovic, M.; Barcelò, D. Monitoring release of pharmaceutical compounds: Occurrence and environmental risk assessment of two wwtp effluents and their receiving bodies in the Po Valley, Italy. Sci. Total Environ. 2012, 438, 15–25. [Google Scholar] [CrossRef] [PubMed]
  81. Pal, A.; Gin, K.Y.-H.; Lin, A.Y.-C.; Reinhard, M. Impacts of emerging organic contaminants on freshwater resources: Review of recent occurrences, sources, fate and effects. Sci. Total Environ. 2010, 408, 6062–6069. [Google Scholar] [CrossRef] [PubMed]
  82. Gioia, R.; Dachs, J. The riverine input–output paradox for organic pollutants. Front. Ecol. Environ. 2012, 10, 405–406. [Google Scholar] [CrossRef]
  83. Kunkel, U.; Radke, M. Fate of pharmaceuticals in rivers: Deriving a benchmark dataset at favorable attenuation conditions. Water Res. 2012, 46, 5551–5565. [Google Scholar] [CrossRef] [PubMed]
  84. Acuña, V.; von Schiller, D.; García-Galán, M.J.; Rodríguez-Mozaz, S.; Corominas, L.; Petrovic, M.; Poch, M.; Barceló, D.; Sabater, S. Occurrence and in-stream attenuation of wastewater-derived pharmaceuticals in iberian rivers. Sci. Total Environ. 2015, 503, 133–141. [Google Scholar] [CrossRef] [PubMed]
  85. Writer, J.H.; Antweiler, R.C.; Ferrer, I.; Ryan, J.N.; Thurman, E.M. In-stream attenuation of neuro-active pharmaceuticals and their metabolites. Environ. Sci. Technol. 2013, 47, 9781–9790. [Google Scholar] [CrossRef] [PubMed]
  86. Gómez, M.J.; Herrera, S.; Solé, D.; García-Calvo, E.; Fernández-Alba, A.R. Spatio-temporal evaluation of organic contaminants and their transformation products along a river basin affected by urban, agricultural and industrial pollution. Sci. Total Environ. 2012, 420, 134–145. [Google Scholar] [CrossRef] [PubMed]
  87. Wang, C.; Shi, H.; Adams, C.D.; Gamagedara, S.; Stayton, I.; Timmons, T.; Ma, Y. Investigation of pharmaceuticals in missouri natural and drinking water using high performance liquid chromatography-tandem mass spectrometry. Water Res. 2011, 45, 1818–1828. [Google Scholar] [CrossRef] [PubMed]
  88. Christoffels, E.; Brunsch, A.; Wunderlich-Pfeiffer, J.; Mertens, F.M. Monitoring micropollutants in the swist river basin. Water Sci. Technol. 2016, 74, 2280–2296. [Google Scholar] [CrossRef] [PubMed]
  89. Kickham, P.; Otton, S.; Moore, M.M.; Ikonomou, M.G.; Gobas, F.A. Relationship between biodegradation and sorption of phthalate esters and their metabolites in natural sediments. Environ. Toxicol. Chem. 2012, 31, 1730–1737. [Google Scholar] [CrossRef] [PubMed]
  90. Riml, J.; Wörman, A.; Kunkel, U.; Radke, M. Evaluating the fate of six common pharmaceuticals using a reactive transport model: Insights from a stream tracer test. Sci. Total Environ. 2013, 458, 344–354. [Google Scholar] [CrossRef] [PubMed]
  91. Yamamoto, H.; Nakamura, Y.; Moriguchi, S.; Nakamura, Y.; Honda, Y.; Tamura, I.; Hirata, Y.; Hayashi, A.; Sekizawa, J. Persistence and partitioning of eight selected pharmaceuticals in the aquatic environment: Laboratory photolysis, biodegradation, and sorption experiments. Water Res. 2009, 43, 351–362. [Google Scholar] [CrossRef] [PubMed]
  92. Weigel, S.; Berger, U.; Jensen, E.; Kallenborn, R.; Thoresen, H.; Hühnerfuss, H. Determination of selected pharmaceuticals and caffeine in sewage and seawater from tromsø/norway with emphasis on ibuprofen and its metabolites. Chemosphere 2004, 56, 583–592. [Google Scholar] [CrossRef] [PubMed]
  93. Weigel, S.; Bester, K.; Hühnerfuss, H. New method for rapid solid-phase extraction of large-volume water samples and its application to non-target screening of north sea water for organic contaminants by gas chromatography–mass spectrometry. J. Chromatogr. A 2001, 912, 151–161. [Google Scholar] [CrossRef]
  94. Kinney, C.A.; Furlong, E.T.; Werner, S.L.; Cahill, J.D. Presence and distribution of wastewater-derived pharmaceuticals in soil irrigated with reclaimed water. Environ. Toxicol. Chem. 2006, 25, 317–326. [Google Scholar] [CrossRef] [PubMed]
  95. Schlenker, G. Pharmaceuticals in the environment. In Encyclopedia of Quantitative Risk Analysis and Assessment; Melnick, E.L., Everitt, B.S., Eds.; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2008; Volume 3, ISBN 9780470035498. [Google Scholar]
  96. Thacker, P. Pharmaceutical data elude researchers. Environ. Sci. Technol. 2005, 39, 193A. [Google Scholar] [PubMed]
  97. Postigo, C.; Barceló, D. Synthetic organic compounds and their transformation products in groundwater: Occurrence, fate and mitigation. Sci. Total Environ. 2015, 503, 32–47. [Google Scholar] [CrossRef] [PubMed]
  98. Chefetz, B.; Mualem, T.; Ben-Ari, J. Sorption and mobility of pharmaceutical compounds in soil irrigated with reclaimed wastewater. Chemosphere 2008, 73, 1335–1343. [Google Scholar] [CrossRef] [PubMed]
  99. Christou, A.; Agüera, A.; Bayona, J.M.; Cytryn, E.; Fotopoulos, V.; Lambropoulou, D.; Manaia, C.M.; Michael, C.; Revitt, M.; Schröder, P. The potential implications of reclaimed wastewater reuse for irrigation on the agricultural environment: The knowns and unknowns of the fate of antibiotics and antibiotic resistant bacteria and resistance genes—A review. Water Res. 2017, 123, 448–467. [Google Scholar] [CrossRef] [PubMed]
  100. Ferrari, B.T.; Paxéus, N.; Giudice, R.L.; Pollio, A.; Garric, J. Ecotoxicological impact of pharmaceuticals found in treated wastewaters: Study of carbamazepine, clofibric acid, and diclofenac. Ecotoxicol. Environ. Saf. 2003, 55, 359–370. [Google Scholar] [CrossRef]
  101. Han, G.H.; Hur, H.G.; Kim, S.D. Ecotoxicological risk of pharmaceuticals from wastewater treatment plants in Korea: Occurrence and toxicity to daphnia magna. Environ. Toxicol. Chem. 2006, 25, 265–271. [Google Scholar] [CrossRef] [PubMed]
  102. Dussault, E.B.; Balakrishnan, V.K.; Sverko, E.; Solomon, K.R.; Sibley, P.K. Toxicity of human pharmaceuticals and personal care products to benthic invertebrates. Environ. Toxicol. Chem. 2008, 27, 425–432. [Google Scholar] [CrossRef] [PubMed]
  103. Jos, A.; Repetto, G.; Rios, J.C.; Hazen, M.J.; Molero, M.L.; del Peso, A.; Salguero, M.; Fernández-Freire, P.; Pérez-Martín, J.M.; Cameán, A. Ecotoxicological evaluation of carbamazepine using six different model systems with eighteen endpoints. Toxicol. In Vitro 2003, 17, 525–532. [Google Scholar] [CrossRef]
  104. Cleuvers, M. Aquatic ecotoxicity of pharmaceuticals including the assessment of combination effects. Toxicol. Lett. 2003, 142, 185–194. [Google Scholar] [CrossRef]
  105. Li, Z.-H.; Zlabek, V.; Velisek, J.; Grabic, R.; Machovaa, J.; Randaka, T. Physiological condition status and musclebased biomarkers in rainbow trout (Oncorhynchus mykiss), after long-term exposure to carbamazepine. J. Appl. Toxicol. 2009, 30, 197–203. [Google Scholar]
  106. Van den Brandhof, E.-J.; Montforts, M. Fish embryo toxicity of carbamazepine, diclofenac and metoprolol. Ecotoxicol. Environ. Saf. 2010, 73, 1862–1866. [Google Scholar] [CrossRef] [PubMed]
  107. Hillis, D.G.; Antunes, P.; Sibley, P.K.; Klironomos, J.N.; Solomon, K.R. Structural responses of daucus carota root-organ cultures and the arbuscular mycorrhizal fungus, glomus intraradices, to 12 pharmaceuticals. Chemosphere 2008, 73, 344–352. [Google Scholar] [CrossRef] [PubMed]
  108. Quinn, B.; Gagné, F.; Blaise, C. An investigation into the acute and chronic toxicity of eleven pharmaceuticals (and their solvents) found in wastewater effluent on the cnidarian, hydra attenuata. Sci. Total Environ. 2008, 389, 306–314. [Google Scholar] [CrossRef] [PubMed]
  109. Nash, J.P.; Kime, D.E.; Van der Ven, L.T.; Wester, P.W.; Brion, F.; Maack, G.; Stahlschmidt-Allner, P.; Tyler, C.R. Long-term exposure to environmental concentrations of the pharmaceutical ethynylestradiol causes reproductive failure in fish. Environ. Health Perspect. 2004, 112, 1725. [Google Scholar] [CrossRef] [PubMed]
  110. Corcoran, J.; Winter, M.J.; Tyler, C.R. Pharmaceuticals in the aquatic environment: A critical review of the evidence for health effects in fish. Crit. Rev. Toxicol. 2010, 40, 287–304. [Google Scholar] [CrossRef] [PubMed]
  111. Snyder, S.A.; Vanderford, B.J.; Drewes, J.; Dickenson, E.; Snyder, E.M.; Bruce, G.M.; Pleus, R.C. State of Knowledge of Endocrine Disruptors and Pharmaceuticals in Drinking Water; IWA Publishing: London, UK, 2008; p. 264. ISBN 9781843392415. [Google Scholar]
  112. Chen, P.; Lin, J.-J.; Lu, C.-S.; Ong, C.-T.; Hsieh, P.F.; Yang, C.-C.; Tai, C.-T.; Wu, S.-L.; Lu, C.-H.; Hsu, Y.-C.; et al. Carbamazepine-induced toxic effects and HLA-B* 1502 screening in taiwan. N. Engl. J. Med. 2011, 364, 1126–1133. [Google Scholar] [CrossRef] [PubMed]
  113. Pereira, F.A.; Mudgil, A.V.; Rosmarin, D.M. Toxic epidermal necrolysis. J. Am. Acad. Dermatol. 2007, 56, 181–200. [Google Scholar] [CrossRef] [PubMed]
  114. Jentink, J.; Dolk, H.; Loane, A.M.; Morris, K.J.; Wellesley, D.; Garne, E.; de Jong-van den Berg, L. Intrauterine exposure to carbamazepine and specific congenital malformations: Systematic review and case-control study. Br. Med. J. 2010, 341, 6581–6588. [Google Scholar] [CrossRef] [PubMed]
  115. Cummings, C.; Stewart, M.; Stevenson, M.; Morrow, J.; Nelson, J. Neurodevelopment of children exposed in utero to lamotrigine, sodium valproate and carbamazepine. Arch. Dis. Child. 2011, 96, 643–647. [Google Scholar] [CrossRef] [PubMed]
  116. Atkinson, E.D.; Brice-Bennett, S.; D’Souza, W.S. Antiepileptic medication during pregnancy: Does fetal genotype affect outcome? Pediatr. Res. 2007, 62, 120–127. [Google Scholar] [CrossRef] [PubMed]
  117. Virkutyte, J.; Varma, R.S.; Jegatheesan, V. Treatment of Micropollutants in Water and Wastewater; IWA Publishing: London, UK, 2010; p. 520. ISBN 9781843393160. [Google Scholar]
  118. Lofgren, H.; Boer, R.D. Pharmaceuticals in Australia: Developments in regulation and governance. Soc. Sci. Med. 2004, 58, 2397–2407. [Google Scholar] [CrossRef] [PubMed]
  119. MDH. Carbamazepine in Drinking Water; Health Risk Assessment Unit, Environmental Health Division: St. Paul, MN, USA, 2011.
  120. Cunliffe, D. Australian Guidelines for Water Recycling: Augmentation of Drinking Water Supplies; Environment Protection and Heritage Council: Canberra, Australia; National Health and Medical Research Council: Canberra, Australia; Natural Resource Management Ministerial Council: Canberra, Australia, 2008. [Google Scholar]
  121. 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]
  122. Metcalf, E.; EDDY, M. Wastewater Engineering: Treatment and Resource Recovery, 5th ed.; McGraw-Hill Professional: New York, NY, USA, 2013; p. 2048. ISBN 9780073401188. [Google Scholar]
  123. Habib, R.; Asif, M.B.; Iftekhar, S.; Khan, Z.; Gurung, K.; Srivastava, V.; Sillanpää, M. Influence of relaxation modes on membrane fouling in submerged membrane bioreactor for domestic wastewater treatment. Chemosphere 2017, 181, 19–25. [Google Scholar] [CrossRef] [PubMed]
  124. Tran, N.H.; Urase, T.; Ngo, H.H.; Hu, J.; Ong, S.L. Insight into metabolic and cometabolic activities of autotrophic and heterotrophic microorganisms in the biodegradation of emerging trace organic contaminants. Bioresour. Technol. 2013, 146, 721–731. [Google Scholar] [CrossRef] [PubMed]
  125. Fernandez-Fontaina, E.; Carballa, M.; Omil, F.; Lema, J. Modelling cometabolic biotransformation of organic micropollutants in nitrifying reactors. Water Res. 2014, 65, 371–383. [Google Scholar] [CrossRef] [PubMed]
  126. 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] [PubMed]
  127. Hai, F.I.; Yamamoto, K.; Fukushi, K. Different fouling modes of submerged hollow-fiber and flat-sheet membranes induced by high strength wastewater with concurrent biofouling. Desalination 2005, 180, 89–97. [Google Scholar] [CrossRef]
  128. Radjenovic, J.; Petrovic, M.; Barceló, D. Analysis of pharmaceuticals in wastewater and removal using a membrane bioreactor. Anal. Bioanal. Chem. 2007, 387, 1365–1377. [Google Scholar] [CrossRef] [PubMed]
  129. Nakada, N.; Tanishima, T.; Shinohara, H.; Kiri, K.; Takada, H. Pharmaceutical chemicals and endocrine disrupters in municipal wastewater in tokyo and their removal during activated sludge treatment. Water Res. 2006, 40, 3297–3303. [Google Scholar] [CrossRef] [PubMed]
  130. Clara, M.; Kreuzinger, N.; Strenn, B.; Gans, O.; Kroiss, H. The solids retention time—A suitable design parameter to evaluate the capacity of wastewater treatment plants to remove micropollutants. Water Res. 2005, 39, 97–106. [Google Scholar] [CrossRef] [PubMed]
  131. Yan, Q.; Gao, X.; Chen, Y.-P.; Peng, X.-Y.; Zhang, Y.-X.; Gan, X.-M.; Zi, C.-F.; Guo, J.-S. Occurrence, fate and ecotoxicological assessment of pharmaceutically active compounds in wastewater and sludge from wastewater treatment plants in chongqing, the three gorges reservoir area. Sci. Total Environ. 2014, 470, 618–630. [Google Scholar] [CrossRef] [PubMed]
  132. Subedi, B.; Kannan, K. Occurrence and fate of select psychoactive pharmaceuticals and antihypertensives in two wastewater treatment plants in new york state, USA. Sci. Total Environ. 2015, 514, 273–280. [Google Scholar] [CrossRef] [PubMed]
  133. Hai, F.I.; Li, X.; Price, W.E.; Nghiem, L.D. Removal of carbamazepine and sulfamethoxazole by MBR under anoxic and aerobic conditions. Bioresour. Technol. 2011, 102, 10386–10390. [Google Scholar] [CrossRef] [PubMed]
  134. Tadkaew, N.; Hai, F.I.; McDonald, J.A.; Khan, S.J.; Nghiem, L.D. Removal of trace organics by mbr treatment: The role of molecular properties. Water Res. 2011, 45, 2439–2451. [Google Scholar] [CrossRef] [PubMed]
  135. Asif, M.B.; Habib, R.; Iftekhar, S.; Khan, Z.; Majeed, N. Optimization of the operational parameters in a submerged membrane bioreactor using box behnken response surface methodology: Membrane fouling control and effluent quality. Desalination 2017, 82, 26–38. [Google Scholar] [CrossRef]
  136. Jumat, M.R.; Hasan, N.A.; Subramanian, P.; Heberling, C.; Colwell, R.R.; Hong, P.-Y. Membrane bioreactor-based wastewater treatment plant in saudi arabia: Reduction of viral diversity, load, and infectious capacity. Water 2017, 9, 534. [Google Scholar] [CrossRef]
  137. Kimura, K.; Hara, H.; Watanabe, Y. Elimination of selected acidic pharmaceuticals from municipal wastewater by an activated sludge system and membrane bioreactors. Environ. Sci. Technol. 2007, 41, 3708–3714. [Google Scholar] [CrossRef] [PubMed]
  138. Pérez, S.; Barceló, D. First evidence for occurrence of hydroxylated human metabolites of diclofenac and aceclofenac in wastewater using QqLIT-MS and QqTOF-MS. Anal. Chem. 2008, 80, 8135–8145. [Google Scholar] [CrossRef] [PubMed]
  139. Cirja, M.; Ivashechkin, P.; Schäffer, A.; Corvini, P. Factors affecting the removal of organic micropollutants from wastewater in conventional treatment plants (CTP) and membrane bioreactors (MBR). Rev. Environ. Sci. Biotechnol. 2008, 7, 61–78. [Google Scholar] [CrossRef]
  140. Luo, W.; Hai, F.I.; Kang, J.; Price, W.E.; Guo, W.; Ngo, H.H.; Yamamoto, K.; Nghiem, L.D. Effects of salinity build-up on biomass characteristics and trace organic chemical removal: Implications on the development of high retention membrane bioreactors. Bioresour. Technol. 2015, 177, 274–281. [Google Scholar] [CrossRef] [PubMed]
  141. Jiang, Q.; Ngo, H.H.; Nghiem, L.D.; Hai, F.I.; Price, W.E.; Zhang, J.; Liang, S.; Deng, L.; Guo, W. Effect of hydraulic retention time on the performance of a hybrid moving bed biofilm reactor-membrane bioreactor system for micropollutants removal from municipal wastewater. Bioresour. Technol. 2018, 247, 1228–1232. [Google Scholar] [CrossRef] [PubMed]
  142. Gurung, K.; Ncibi, M.C.; Sillanpää, M. Assessing membrane fouling and the performance of pilot-scale membrane bioreactor (mbr) to treat real municipal wastewater during winter season in nordic regions. Sci. Total Environ. 2017, 579, 1289–1297. [Google Scholar] [CrossRef] [PubMed]
  143. Kimura, K.; Hara, H.; Watanabe, Y. Removal of pharmaceutical compounds by submerged membrane bioreactors (mbrs). Desalination 2005, 178, 135–140. [Google Scholar] [CrossRef]
  144. Abegglen, C.; Joss, A.; McArdell, C.S.; Fink, G.; Schlüsener, M.P.; Ternes, T.A.; Siegrist, H. The fate of selected micropollutants in a single-house mbr. Water Res. 2009, 43, 2036–2046. [Google Scholar] [CrossRef] [PubMed]
  145. Joss, A.; Andersen, H.; Ternes, T.; Richle, P.R.; Siegrist, H. Removal of estrogens in municipal wastewater treatment under aerobic and anaerobic conditions: Consequences for plant optimization. Environ. Sci. Technol. 2004, 38, 3047–3055. [Google Scholar] [CrossRef] [PubMed]
  146. Phan, H.V.; Hai, F.I.; Kang, J.; Dam, H.K.; Zhang, R.; Price, W.E.; Broeckmann, A.; Nghiem, L.D. Simultaneous nitrification/denitrification and trace organic contaminant (troc) removal by an anoxic–aerobic membrane bioreactor (mbr). Bioresour. Technol. 2014, 165, 96–104. [Google Scholar] [CrossRef] [PubMed]
  147. Suarez, S.; Lema, J.M.; Omil, F. Removal of pharmaceutical and personal care products (ppcps) under nitrifying and denitrifying conditions. Water Res. 2010, 44, 3214–3224. [Google Scholar] [CrossRef] [PubMed]
  148. 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, 141, 97–108. [Google Scholar] [CrossRef] [PubMed]
  149. Asif, M.B.; Hai, F.I.; Hou, J.; Price, W.E.; Nghiem, L.D. Impact of wastewater derived dissolved interfering compounds on growth, enzymatic activity and trace organic contaminant removal of white rot fungi—A critical review. J. Environ. Manag. 2017, 201, 89–109. [Google Scholar] [CrossRef] [PubMed]
  150. Hai, F.I.; Yamamoto, K.; Fukushi, K. Development of a submerged membrane fungi reactor for textile wastewater treatment. Desalination 2006, 192, 315–322. [Google Scholar] [CrossRef]
  151. Hai, F.I.; Yamamoto, K.; Fukushi, K. Hybrid treatment systems for dye wastewater. Crit. Rev. Environ. Sci. Technol. 2007, 37, 315–377. [Google Scholar] [CrossRef]
  152. Hai, F.I.; Yamamoto, K.; Fukushi, K.; Nakajima, F. Fouling resistant compact hollow-fiber module with spacer for submerged membrane bioreactor treating high strength industrial wastewater. J. Membr. Sci. 2008, 317, 34–42. [Google Scholar] [CrossRef]
  153. Margot, J.; Maillard, J.; Rossi, L.; Barry, D.A.; Holliger, C. Influence of treatment conditions on the oxidation of micropollutants by trametes versicolor laccase. New Biotechnol. 2013, 30, 803–813. [Google Scholar] [CrossRef] [PubMed]
  154. 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]
  155. Golan-Rozen, N.; Chefetz, B.; Ben-Ari, J.; Geva, J.; Hadar, Y. Transformation of the recalcitrant pharmaceutical compound carbamazepine by pleurotus ostreatus: Role of cytochrome p450 monooxygenase and manganese peroxidase. Environ. Sci. Technol. 2011, 45, 6800–6805. [Google Scholar] [CrossRef] [PubMed]
  156. Rodarte-Morales, A.; Feijoo, G.; Moreira, M.; Lema, J. Degradation of selected pharmaceutical and personal care products (ppcps) by white-rot fungi. World J. Microbiol. Biotechnol. 2011, 27, 1839–1846. [Google Scholar] [CrossRef]
  157. Tran, N.H.; Urase, T.; Kusakabe, O. Biodegradation characteristics of pharmaceutical substances by whole fungal culture trametes versicolor and its laccase. J. Water Environ. Technol. 2010, 8, 125–140. [Google Scholar] [CrossRef]
  158. Nguyen, L.N.; Hai, F.I.; Yang, S.; Kang, J.; Leusch, F.D.L.; Roddick, F.; Price, W.E.; Nghiem, L.D. Removal of pharmaceuticals, steroid hormones, phytoestrogens, uv-filters, industrial chemicals and pesticides by trametes versicolor: Role of biosorption and biodegradation. Int. Biodeterior. Biodegrad. 2014, 88, 169–175. [Google Scholar] [CrossRef]
  159. Rodarte-Morales, A.; Feijoo, G.; Moreira, M.; Lema, J. Operation of stirred tank reactors (strs) and fixed-bed reactors (fbrs) with free and immobilized phanerochaete chrysosporium for the continuous removal of pharmaceutical compounds. Biochem. Eng. J. 2012, 66, 38–45. [Google Scholar] [CrossRef]
  160. Jelic, A.; Cruz-Morató, C.; Marco-Urrea, E.; Sarrà, M.; Perez, S.; Vicent, T.; Petrović, M.; Barcelo, D. Degradation of carbamazepine by trametes versicolor in an air pulsed fluidized bed bioreactor and identification of intermediates. Water Res. 2012, 46, 955–964. [Google Scholar] [CrossRef] [PubMed]
  161. Nguyen, L.N.; Hai, F.I.; Yang, S.; Kang, J.; Leusch, F.D.; 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] [PubMed]
  162. Nguyen, L.N.; Hai, F.I.; Price, W.E.; Kang, J.; Leusch, F.D.; Roddick, F.; van de Merwe, J.P.; Magram, S.F.; Nghiem, L.D. Degradation of a broad spectrum of trace organic contaminants by an enzymatic membrane reactor: Complementary role of membrane retention and enzymatic degradation. Int. Biodeterior. Biodegrad. 2015, 99, 115–122. [Google Scholar] [CrossRef]
  163. Zhang, Y.; Geißen, S.-U. In vitro degradation of carbamazepine and diclofenac by crude lignin peroxidase. J. Hazard. Mater. 2010, 176, 1089–1092. [Google Scholar] [CrossRef] [PubMed]
  164. Nguyen, L.N.; van de Merwe, J.P.; Hai, F.I.; Leusch, F.D.; 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]
  165. Rodarte-Morales, A.; Feijoo, G.; Moreira, M.; Lema, J. 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]
  166. Yang, S.; Hai, F.I.; Nghiem, L.D.; Roddick, F.; Price, W.E. Removal of trace organic contaminants by nitrifying activated sludge and whole-cell and crude enzyme extract of trametes versicolor. Water Sci. Technol. 2013, 67, 1216–1223. [Google Scholar] [CrossRef] [PubMed]
  167. Espinosa-Ortiz, E.J.; Rene, E.R.; Pakshirajan, K.; van Hullebusch, E.D.; Lens, P.N. Fungal pelleted reactors in wastewater treatment: Applications and perspectives. Chem. Eng. J. 2016, 283, 553–571. [Google Scholar] [CrossRef]
  168. Mir-Tutusaus, J.; Sarrà, M.; Caminal, G. Continuous treatment of non-sterile hospital wastewater by trametes versicolor: How to increase fungal viability by means of operational strategies and pretreatments. J. Hazard. Mater. 2016, 318, 561–570. [Google Scholar] [CrossRef] [PubMed]
  169. Mir-Tutusaus, J.A.; Parladé, E.; Llorca, M.; Villagrasa, M.; Barceló, D.; Rodriguez-Mozaz, S.; Martinez-Alonso, M.; Gaju, N.; Caminal, G.; Sarrà, M. Pharmaceuticals removal and microbial community assessment in a continuous fungal treatment of non-sterile real hospital wastewater after a coagulation-flocculation pretreatment. Water Res. 2017, 116, 65–75. [Google Scholar] [CrossRef] [PubMed]
  170. Badia-Fabregat, M.; Lucas, D.; Pereira, M.A.; Alves, M.; Pennanen, T.; Fritze, H.; Rodríguez-Mozaz, S.; Barceló, D.; Vicent, T.; Caminal, G. Continuous fungal treatment of non-sterile veterinary hospital effluent: Pharmaceuticals removal and microbial community assessment. Appl. Microbiol. Biotechnol. 2016, 100, 2401–2415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Cruz-Morató, C.; Lucas, D.; Llorca, M.; Rodriguez-Mozaz, S.; Gorga, M.; Petrovic, M.; Barceló, D.; Vicent, T.; Sarrà, M.; Marco-Urrea, E. Hospital wastewater treatment by fungal bioreactor: Removal efficiency for pharmaceuticals and endocrine disruptor compounds. Sci. Total Environ. 2014, 493, 365–376. [Google Scholar] [CrossRef] [PubMed]
  172. Srivastava, S.; Goyal, P. Novel Biomaterials: Decontamination of Toxic Metals from Wasterwater; Springer: Berlin/Heidelberg, Germany, 2010. [Google Scholar]
  173. Taheran, M.; Brar, S.K.; Verma, M.; Surampalli, R.; Zhang, T.; Valero, J. Membrane processes for removal of pharmaceutically active compounds (phacs) from water and wastewaters. Sci. Total Environ. 2016, 547, 60–77. [Google Scholar] [CrossRef] [PubMed]
  174. Racar, M.; Dolar, D.; Špehar, A.; Košutić, K. Application of uf/nf/ro membranes for treatment and reuse of rendering plant wastewater. Process Saf. Environ. Prot. 2017, 105, 386–392. [Google Scholar] [CrossRef]
  175. Schulte-Herbrüggen, H.M.; Semião, A.J.; Chaurand, P.; Graham, M.C. Effect of ph and pressure on uranium removal from drinking water using nf/ro membranes. Environ. Sci. Technol. 2016, 50, 5817–5824. [Google Scholar] [CrossRef] [PubMed]
  176. Radjenović, J.; Petrović, M.; Ventura, F.; Barceló, D. Rejection of pharmaceuticals in nanofiltration and reverse osmosis membrane drinking water treatment. Water Res. 2008, 42, 3601–3610. [Google Scholar] [CrossRef] [PubMed]
  177. Bellona, C.; Drewes, J.E.; Oelker, G.; Luna, J.; Filteau, G.; Amy, G. Comparing nanofiltration and reverse osmosis for drinking water augmentation. Am. Water Work. Assoc. J. 2008, 100, 102–116. [Google Scholar]
  178. Comerton, A.M.; Andrews, R.C.; Bagley, D.M.; Hao, C. The rejection of endocrine disrupting and pharmaceutically active compounds by nf and ro membranes as a function of compound and water matrix properties. J. Membr. Sci. 2008, 313, 323–335. [Google Scholar] [CrossRef]
  179. Röhricht, M.; Krisam, J.; Weise, U.; Kraus, U.R.; Düring, R.-A. Elimination of carbamazepine, diclofenac and naproxen from treated wastewater by nanofiltration. CLEAN Soil Air Water 2009, 37, 638–641. [Google Scholar] [CrossRef]
  180. Gur-Reznik, S.; Koren-Menashe, I.; Heller-Grossman, L.; Rufel, O.; Dosoretz, C.G. Influence of seasonal and operating conditions on the rejection of pharmaceutical active compounds by ro and nf membranes. Desalination 2011, 277, 250–256. [Google Scholar] [CrossRef]
  181. Beier, S.; Köster, S.; Veltmann, K.; Schröder, H.; Pinnekamp, J. Treatment of hospital wastewater effluent by nanofiltration and reverse osmosis. Water Sci. Technol. 2010, 61, 1691–1698. [Google Scholar] [CrossRef] [PubMed]
  182. Nghiem, L.D.; Coleman, P.J.; Espendiller, C. Mechanisms underlying the effects of membrane fouling on the nanofiltration of trace organic contaminants. Desalination 2010, 250, 682–687. [Google Scholar] [CrossRef]
  183. Vogel, D.; Simon, A.; Alturki, A.A.; Bilitewski, B.; Price, W.E.; Nghiem, L.D. Effects of fouling and scaling on the retention of trace organic contaminants by a nanofiltration membrane: The role of cake-enhanced concentration polarisation. Sep. Purif. Technol. 2010, 73, 256–263. [Google Scholar] [CrossRef]
  184. Simon, A.; Price, W.E.; Nghiem, L.D. Effects of chemical cleaning on the nanofiltration of pharmaceutically active compounds (phacs). Sep. Purif. Technol. 2012, 88, 208–215. [Google Scholar] [CrossRef]
  185. Nghiem, L.D.; Hawkes, S. Effects of membrane fouling on the nanofiltration of pharmaceutically active compounds (phacs): Mechanisms and role of membrane pore size. Sep. Purif. Technol. 2007, 57, 176–184. [Google Scholar] [CrossRef]
  186. Hajibabania, S.; Verliefde, A.; Drewes, J.E.; Nghiem, L.D.; McDonald, J.; Khan, S.; Le-Clech, P. Effect of fouling on removal of trace organic compounds by nanofiltration. Drink. Water Eng. Sci. 2011, 4, 117–149. [Google Scholar] [CrossRef]
  187. Vogna, D.; Marotta, R.; Andreozzi, R.; Napolitano, A.; D’Ischia, M. Kinetic and chemical assessment of the uv/h2o2 treatment of antiepileptic drug carbamazepine. Chemosphere 2004, 54, 497–505. [Google Scholar] [CrossRef]
  188. Mohammad, A.W.; Teow, Y.; Ang, W.; Chung, Y.; Oatley-Radcliffe, D.; Hilal, N. Nanofiltration membranes review: Recent advances and future prospects. Desalination 2015, 356, 226–254. [Google Scholar] [CrossRef]
  189. Comerton, A.M.; Andrews, R.C.; Bagley, D.M. The influence of natural organic matter and cations on the rejection of endocrine disrupting and pharmaceutically active compounds by nanofiltration. Water Res. 2009, 43, 613–622. [Google Scholar] [CrossRef] [PubMed]
  190. Kårelid, V.; Larsson, G.; Björlenius, B. Pilot-scale removal of pharmaceuticals in municipal wastewater: Comparison of granular and powdered activated carbon treatment at three wastewater treatment plants. J. Environ. Manag. 2017, 193, 491–502. [Google Scholar] [CrossRef] [PubMed]
  191. Skouteris, G.; Saroj, D.; Melidis, P.; Hai, F.I.; Ouki, S. The effect of activated carbon addition on membrane bioreactor processes for wastewater treatment and reclamation—A critical review. Bioresour. Technol. 2015, 185, 399–410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Moreno-Castilla, C. Adsorption of organic molecules from aqueous solutions on carbon materials. Carbon 2004, 42, 83–94. [Google Scholar] [CrossRef]
  193. Real, F.J.; Benitez, F.J.; Acero, J.L.; Casas, F. Adsorption of selected emerging contaminants onto pac and gac: Equilibrium isotherms, kinetics, and effect of the water matrix. J. Environ. Sci. Health Part A 2017, 1–8. [Google Scholar] [CrossRef] [PubMed]
  194. Kovalova, L.; Siegrist, H.; Von Gunten, U.; Eugster, J.; Hagenbuch, M.; Wittmer, A.; Moser, R.; McArdell, C.S. Elimination of micropollutants during post-treatment of hospital wastewater with powdered activated carbon, ozone, and UV. Environ. Sci. Technol. 2013, 47, 7899–7908. [Google Scholar] [CrossRef] [PubMed]
  195. Choi, K.-J.; Kim, S.-G.; Kim, S.-H. Removal of antibiotics by coagulation and granular activated carbon filtration. J. Hazard. Mater. 2008, 151, 38–43. [Google Scholar] [CrossRef] [PubMed]
  196. Rossner, A.; Snyder, S.A.; Knappe, D.R. Removal of emerging contaminants of concern by alternative adsorbents. Water Res. 2009, 43, 3787–3796. [Google Scholar] [CrossRef] [PubMed]
  197. Ternes, T.A.; Meisenheimer, M.; McDowell, D.; Sacher, F.; Brauch, H.-J.; Haist-Gulde, B.; Preuss, G.; Wilme, U.; Zulei-Seibert, N. Removal of pharmaceuticals during drinking water treatment. Environ. Sci. Technol. 2002, 36, 3855–3863. [Google Scholar] [CrossRef] [PubMed]
  198. Pojana, G.; Fantinati, A.; Marcomini, A. Occurrence of environmentally relevant pharmaceuticals in italian drinking water treatment plants. Int. J. Environ. Anal. Chem. 2011, 91, 537–552. [Google Scholar] [CrossRef]
  199. Huerta-Fontela, M.; Galceran, M.T.; Ventura, F. Occurrence and removal of pharmaceuticals and hormones through drinking water treatment. Water Res. 2011, 45, 1432–1442. [Google Scholar] [CrossRef] [PubMed]
  200. Kim, S.D.; Cho, J.; Kim, I.S.; Vanderford, B.J.; Snyder, S.A. Occurrence and removal of pharmaceuticals and endocrine disruptors in South Korean surface, drinking, and waste waters. Water Res. 2007, 41, 1013–1021. [Google Scholar] [CrossRef] [PubMed]
  201. Stackelberg, P.E.; Gibs, J.; Furlong, E.T.; Meyer, M.T.; Zaugg, S.D.; Lippincott, R.L. Efficiency of conventional drinking-water-treatment processes in removal of pharmaceuticals and other organic compounds. Sci. Total Environ. 2007, 377, 255–272. [Google Scholar] [CrossRef] [PubMed]
  202. Grover, D.P.; Zhou, J.L.; Frickers, P.E.; Readman, J.W. Improved removal of estrogenic and pharmaceutical compounds in sewage effluent by full scale granular activated carbon: Impact on receiving river water. J. Hazard. Mater. 2011, 185, 1005–1011. [Google Scholar] [CrossRef] [PubMed]
  203. Serrano, D.; Lema, J.M.; Omil, F. Influence of the employment of adsorption and coprecipitation agents for the removal of ppcps in conventional activated sludge (cas) systems. Water Sci. Technol. 2010, 62, 728–735. [Google Scholar] [CrossRef] [PubMed]
  204. Meinel, F.; Ruhl, A.; Sperlich, A.; Zietzschmann, F.; Jekel, M. Pilot-scale investigation of micropollutant removal with granular and powdered activated carbon. Water Air Soil Pollut. 2015, 226, 2260. [Google Scholar] [CrossRef]
  205. Snyder, S.A.; Adham, S.; Redding, A.M.; Cannon, F.S.; DeCarolis, J.; Oppenheimer, J.; Wert, E.C.; Yoon, Y. Role of membranes and activated carbon in the removal of endocrine disruptors and pharmaceuticals. Desalination 2007, 202, 156–181. [Google Scholar] [CrossRef]
  206. Lipp, P.; Groß, H.-J.; Tiehm, A. Improved elimination of organic micropollutants by a process combination of membrane bioreactor (mbr) and powdered activated carbon (pac). Desalin. Water Treat. 2012, 42, 65–72. [Google Scholar] [CrossRef]
  207. Westerhoff, P.; Yoon, Y.; Snyder, S.; Wert, E. Fate of endocrine-disruptor, pharmaceutical, and personal care product chemicals during simulated drinking water treatment processes. Environ. Sci. Technol. 2005, 39, 6649–6663. [Google Scholar] [CrossRef] [PubMed]
  208. Soares, A.; Guieysse, B.; Jefferson, B.; Cartmell, E.; Lester, J. Nonylphenol in the environment: A critical review on occurrence, fate, toxicity and treatment in wastewaters. Environ. Int. 2008, 34, 1033–1049. [Google Scholar] [CrossRef] [PubMed]
  209. Nevskaia, D.; Guerrero-Ruiz, A. Comparative study of the adsorption from aqueous solutions and the desorption of phenol and nonylphenol substrates on activated carbons. J. Colloid Interface Sci. 2001, 234, 316–321. [Google Scholar] [CrossRef] [PubMed]
  210. Nguyen, L.N.; Hai, F.I.; Kang, J.; Nghiem, L.D.; Price, W.E.; Guo, W.; Ngo, H.H.; Tung, K.-L. Comparison between sequential and simultaneous application of activated carbon with membrane bioreactor for trace organic contaminant removal. Bioresour. Technol. 2013, 130, 412–417. [Google Scholar] [CrossRef] [PubMed]
  211. Li, X.; Hai, F.I.; Nghiem, L.D. Simultaneous activated carbon adsorption within a membrane bioreactor for an enhanced micropollutant removal. Bioresour. Technol. 2011, 102, 5319–5324. [Google Scholar] [CrossRef] [PubMed]
  212. Hernández-Leal, L.; Temmink, H.; Zeeman, G.; Buisman, C. Removal of micropollutants from aerobically treated grey water via ozone and activated carbon. Water Res. 2011, 45, 2887–2896. [Google Scholar] [CrossRef] [PubMed]
  213. Klavarioti, M.; Mantzavinos, D.; Kassinos, D. Removal of residual pharmaceuticals from aqueous systems by advanced oxidation processes. Environ. Int. 2009, 35, 402–417. [Google Scholar] [CrossRef] [PubMed]
  214. Alharbi, S.K.; Price, W.E. Degradation and fate of pharmaceutically active contaminants by advanced oxidation processes. Curr. Pollut. Rep. 2017, 3, 268–280. [Google Scholar] [CrossRef]
  215. Andreozzi, R.; Marotta, R.; Pinto, G.; Pollio, A. Carbamazepine in water: Persistence in the environment, ozonation treatment and preliminary assessment on algal toxicity. Water Res. 2002, 36, 2869–2877. [Google Scholar] [CrossRef]
  216. Hua, W.; Bennett, E.R.; Letcher, R.J. Ozone treatment and the depletion of detectable pharmaceuticals and atrazine herbicide in drinking water sourced from the upper detroit river, ontario, canada. Water Res. 2006, 40, 2259–2266. [Google Scholar] [CrossRef] [PubMed]
  217. Huber, M.M.; Canonica, S.; Park, G.-Y.; von Gunten, U. Oxidation of pharmaceuticals during ozonation and advanced oxidation processes. Environ. Sci. Technol. 2003, 37, 1016–1024. [Google Scholar] [CrossRef] [PubMed]
  218. Wert, E.C.; Rosario-Ortiz, F.L.; Snyder, S.A. Effect of ozone exposure on the oxidation of trace organic contaminants in wastewater. Water Res. 2009, 43, 1005–1014. [Google Scholar] [CrossRef] [PubMed]
  219. Buffle, M.-O.; Schumacher, J.; Salhi, E.; Jekel, M.; von Gunten, U. Measurement of the initial phase of ozone decomposition in water and wastewater by means of a continuous quench-flow system: Application to disinfection and pharmaceutical oxidation. Water Res. 2006, 40, 1884–1894. [Google Scholar] [CrossRef] [PubMed]
  220. Alharbi, S.K.; Kang, J.; Nghiem, L.D.; van de Merwe, J.P.; Leusch, F.D.; Price, W.E. Photolysis and UV/H2O2 of diclofenac, sulfamethoxazole, carbamazepine, and trimethoprim: Identification of their major degradation products by ESI–LC–MS and assessment of the toxicity of reaction mixtures. Process Saf. Environ. Prot. 2017, 112, 222–234. [Google Scholar] [CrossRef]
  221. Im, J.-K.; Cho, I.-H.; Kim, S.-K.; Zoh, K.-D. Optimization of carbamazepine removal in O3/UV/H2O2 system using a response surface methodology with central composite design. Desalination 2012, 285, 306–314. [Google Scholar] [CrossRef]
  222. Monteagudo, J.; Durán, A.; González, R.; Expósito, A. In situ chemical oxidation of carbamazepine solutions using persulfate simultaneously activated by heat energy, UV light, Fe2+ ions, and H2O2. Appl. Catal. B Environ. 2015, 176, 120–129. [Google Scholar] [CrossRef]
  223. Shirazi, E.; Torabian, A.; Nabi-Bidhendi, G. Carbamazepine removal from groundwater: Effectiveness of the Tio2/UV, nanoparticulate zero-valent iron, and fenton (NZVI/H2O2) processes. CLEAN Soil Air Water 2013, 41, 1062–1072. [Google Scholar] [CrossRef]
  224. Yang, X.; Sun, J.; Fu, W.; Shang, C.; Li, Y.; Chen, Y.; Gan, W.; Fang, J. Ppcp degradation by uv/chlorine treatment and its impact on dbp formation potential in real waters. Water Res. 2016, 98, 309–318. [Google Scholar] [CrossRef] [PubMed]
  225. Rosario-Ortiz, F.L.; Wert, E.C.; Snyder, S.A. Evaluation of UV/H2O2 treatment for the oxidation of pharmaceuticals in wastewater. Water Res. 2010, 44, 1440–1448. [Google Scholar] [CrossRef] [PubMed]
  226. Sichel, C.; Garcia, C.; Andre, K. Feasibility studies: UV/chlorine advanced oxidation treatment for the removal of emerging contaminants. Water Res. 2011, 45, 6371–6380. [Google Scholar] [CrossRef] [PubMed]
  227. Pereira, V.J.; Linden, K.G.; Weinberg, H.S. Evaluation of uv irradiation for photolytic and oxidative degradation of pharmaceutical compounds in water. Water Res. 2007, 41, 4413–4423. [Google Scholar] [CrossRef] [PubMed]
  228. Doll, T.E.; Frimmel, F.H. Photocatalytic degradation of carbamazepine, clofibric acid and iomeprol with p25 and hombikat UV100 in the presence of natural organic matter (nom) and other organic water constituents. Water Res. 2005, 39, 403–411. [Google Scholar] [CrossRef] [PubMed]
  229. Bernabeu, A.; Palacios, S.; Vicente, R.; Vercher, R.F.; Malato, S.; Arques, A.; Amat, A.M. Solar photo-fenton at mild conditions to treat a mixture of six emerging pollutants. Chem. Eng. J. 2012, 198–199, 65–72. [Google Scholar] [CrossRef]
  230. Klamerth, N.; Rizzo, L.; Malato, S.; Maldonado, M.I.; Agüera, A.; Fernández-Alba, A.R. Degradation of fifteen emerging contaminants at μg/L initial concentrations by mild solar photo-fenton in mwtp effluents. Water Res. 2010, 44, 545–554. [Google Scholar] [CrossRef] [PubMed]
  231. Liu, N.; Zheng, M.; Sijak, S.; Tang, L.; Xu, G.; Wu, M. Aquatic photolysis of carbamazepine by UV/H2O2 and UV/Fe(ii) processes. Res. Chem. Interm. 2015, 41, 7015–7028. [Google Scholar] [CrossRef]
  232. Dai, C.-m.; Zhou, X.-F.; Zhang, Y.-L.; Duan, Y.-P.; Qiang, Z.-M.; Zhang, T.C. Comparative study of the degradation of carbamazepine in water by advanced oxidation processes. Environ. Technol. 2011, 33, 1101–1109. [Google Scholar] [CrossRef] [PubMed]
  233. Hollender, J.; Zimmermann, S.G.; Koepke, S.; Krauss, M.; McArdell, C.S.; Ort, C.; Singer, H.; von Gunten, U.; Siegrist, H. Elimination of organic micropollutants in a municipal wastewater treatment plant upgraded with a full-scale post-ozonation followed by sand filtration. Environ. Sci. Technol. 2009, 43, 7862–7869. [Google Scholar] [CrossRef] [PubMed]
  234. Reungoat, J.; Macova, M.; Escher, B.; Carswell, S.; Mueller, J.; Keller, J. Removal of micropollutants and reduction of biological activity in a full scale reclamation plant using ozonation and activated carbon filtration. Water Res. 2010, 44, 625–637. [Google Scholar] [CrossRef] [PubMed]
  235. Comninellis, C.; Kapalka, A.; Malato, S.; Parsons, S.A.; Poulios, I.; Mantzavinos, D. Advanced oxidation processes for water treatment: Advances and trends for R&D. J. Chem. Technol. Biotechnol. 2008, 83, 769–776. [Google Scholar]
  236. Hai, F.I.; Nguyen, L.N.; Nghiem, L.D.; Liao, B.-Q.; Koyuncu, I.; Price, W.E. Trace organic contaminants removal by combined processes for wastewater reuse. In Advanced Treatment Technologies for Urban Wastewater Reuse; Fatta-Kassinos, D., Dionysiou, D.D., Kümmerer, K., Eds.; Springer: Cham, Switzerland, 2014; pp. 39–77. ISBN 9783319238869. [Google Scholar]
  237. Nguyen, L.N.; Hai, F.I.; Kang, J.; Price, W.E.; Nghiem, L.D. Removal of emerging trace organic contaminants by mbr-based hybrid treatment processes. Int. Biodeterior. Biodegrad. 2013, 85, 474–482. [Google Scholar] [CrossRef]
  238. Laera, G.; Chong, M.N.; Jin, B.; Lopez, A. An integrated MBR–Tio2 photocatalysis process for the removal of carbamazepine from simulated pharmaceutical industrial effluent. Bioresour. Technol. 2011, 102, 7012–7015. [Google Scholar] [CrossRef] [PubMed]
  239. Wang, S.; Wang, J. Carbamazepine degradation by gamma irradiation coupled to biological treatment. J. Hazard. Mater. 2017, 321, 639–646. [Google Scholar] [CrossRef] [PubMed]
  240. Ivancev-Tumbas, I.; Hobby, R. Removal of organic xenobiotics by combined out/in ultrafiltration and powdered activated carbon adsorption. Desalination 2010, 255, 124–128. [Google Scholar] [CrossRef]
  241. Serrano, D.; Suárez, S.; Lema, J.M.; Omil, F. Removal of persistent pharmaceutical micropollutants from sewage by addition of pac in a sequential membrane bioreactor. Water Res. 2011, 45, 5323–5333. [Google Scholar] [CrossRef] [PubMed]
  242. Nguyen, L.N.; Hai, F.I.; Kang, J.; Price, W.E.; Nghiem, L.D. Removal of trace organic contaminants by a membrane bioreactor–granular activated carbon (mbr–gac) system. Bioresour. Technol. 2012, 113, 169–173. [Google Scholar] [CrossRef] [PubMed]
  243. Hübner, U.; Seiwert, B.; Reemtsma, T.; Jekel, M. Ozonation products of carbamazepine and their removal from secondary effluents by soil aquifer treatment–indications from column experiments. Water Res. 2014, 49, 34–43. [Google Scholar] [CrossRef] [PubMed]
  244. Kleywegt, S.; Pileggi, V.; Yang, P.; Hao, C.; Zhao, X.; Rocks, C.; Thach, S.; Cheung, P.; Whitehead, B. Pharmaceuticals, hormones and bisphenol a in untreated source and finished drinking water in Ontario, Canada—Occurrence and treatment efficiency. Sci. Total Environ. 2011, 409, 1481–1488. [Google Scholar] [CrossRef] [PubMed]
  245. Hu, L.; Martin, H.M.; Arce-Bulted, O.; Sugihara, M.N.; Keating, K.A.; Strathmann, T.J. Oxidation of carbamazepine by Mn(vii) and Fe(vi): Reaction kinetics and mechanism. Environ. Sci. Technol. 2008, 43, 509–515. [Google Scholar] [CrossRef]
  246. Kosjek, T.; Andersen, H.R.; Kompare, B.; Ledin, A.; Heath, E. Fate of carbamazepine during water treatment. Environ. Sci. Technol. 2009, 43, 6256–6261. [Google Scholar] [CrossRef] [PubMed]
  247. Prado, M.; Borea, L.; Cesaro, A.; Liu, H.; Naddeo, V.; Belgiorno, V.; Ballesteros, F. Removal of emerging contaminant and fouling control in membrane bioreactors by combined ozonation and sonolysis. Int. Biodeterior. Biodegrad. 2017, 119, 577–586. [Google Scholar] [CrossRef]
  248. Suarez, S.; Lema, J.M.; Omil, F. Pre-treatment of hospital wastewater by coagulation-flocculation and flotation. Bioresour. Technol. 2009, 100, 2138–2146. [Google Scholar] [CrossRef] [PubMed]
  249. Vieno, N.; Tuhkanen, T.; Kronberg, L. Removal of pharmaceuticals in drinking water treatment: Effect of chemical coagulation. Environ. Technol. 2006, 27, 183–192. [Google Scholar] [CrossRef] [PubMed]
  250. Vieno, N.M.; Härkki, H.; Tuhkanen, T.; Kronberg, L. Occurrence of pharmaceuticals in river water and their elimination in a pilot-scale drinking water treatment plant. Environ. Sci. Technol. 2007, 41, 5077–5084. [Google Scholar] [CrossRef] [PubMed]
  251. Wert, E.C.; Gonzales, S.; Dong, M.M.; Rosario-Ortiz, F.L. Evaluation of enhanced coagulation pretreatment to improve ozone oxidation efficiency in wastewater. Water Res. 2011, 45, 5191–5199. [Google Scholar] [CrossRef] [PubMed]
  252. Ciardelli, G.; Ranieri, N. The treatment and reuse of wastewater in the textile industry by means of ozonation and electroflocculation. Water Res. 2001, 35, 567–572. [Google Scholar] [CrossRef]
  253. McDowell, D.C.; Huber, M.M.; Wagner, M.; von Gunten, U.; Ternes, T.A. Ozonation of carbamazepine in drinking water: Identification and kinetic study of major oxidation products. Environ. Sci. Technol. 2005, 39, 8014–8022. [Google Scholar] [CrossRef] [PubMed]
  254. Marco-Urrea, E.; Radjenović, J.; Caminal, G.; Petrović, M.; Vicent, T.; Barceló, D. Oxidation of atenolol, propranolol, carbamazepine and clofibric acid by a biological fenton-like system mediated by the white-rot fungus trametes versicolor. Water Res. 2010, 44, 521–532. [Google Scholar] [CrossRef] [PubMed]
  255. Bu, H.-Z.; Zhao, P.; Dalvie, D.K.; Pool, W.F. Identification of primary and sequential bioactivation pathways of carbamazepine in human liver microsomes using liquid chromatography/tandem mass spectrometry. Rapid Commun. Mass Spectrom. 2007, 21, 3317–3322. [Google Scholar] [CrossRef] [PubMed]
  256. Leclercq, M.; Mathieu, O.; Gomez, E.; Casellas, C.; Fenet, H.; Hillaire-Buys, D. Presence and fate of carbamazepine, oxcarbazepine, and seven of their metabolites at wastewater treatment plants. Arch. Environ. Contam. Toxicol. 2009, 56, 408–415. [Google Scholar] [CrossRef] [PubMed]
  257. Mohapatra, D.P.; Brar, S.K.; Tyagi, R.D.; Picard, P.; Surampalli, R.Y. A comparative study of ultrasonication, fenton’s oxidation and ferro-sonication treatment for degradation of carbamazepine from wastewater and toxicity test by yeast estrogen screen (yes) assay. Sci. Total Environ. 2013, 447, 280–285. [Google Scholar] [CrossRef] [PubMed]
  258. Vom Eyser, C.; Börgers, A.; Richard, J.; Dopp, E.; Janzen, N.; Bester, K.; Tuerk, J. Chemical and toxicological evaluation of transformation products during advanced oxidation processes. Water Sci. Technol. 2013, 68, 1976–1983. [Google Scholar] [CrossRef] [PubMed]
  259. Li, Z.; Fenet, H.; Gomez, E.; Chiron, S. Transformation of the antiepileptic drug oxcarbazepine upon different water disinfection processes. Water Res. 2011, 45, 1587–1596. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Metabolites/degradation products formed following CBZ degradation by (a) ozonation [199]; (b) UV/H2O2 [187]; (c) fungal degradation [254]; (d) UV photolysis [253]; (e) human liver [255]; (f) activated sludge process [256].
Figure 1. Metabolites/degradation products formed following CBZ degradation by (a) ozonation [199]; (b) UV/H2O2 [187]; (c) fungal degradation [254]; (d) UV photolysis [253]; (e) human liver [255]; (f) activated sludge process [256].
Water 10 00107 g001
Table 1. Physicochemical and pharmacological properties of carbamazepine (modified after [16,17]).
Table 1. Physicochemical and pharmacological properties of carbamazepine (modified after [16,17]).
Structure Water 10 00107 i001
FormulaC15H12N2O
CAS No.298-46-4
Molecular Weight236.2686 g·mol−1
UsageAnticonvulsant/mood stabilizing drug
Water solubility17.7 mg/L (25 °C)
Log P (octanol-water partition coefficient)2.45
Log D at pH = 7 a1.32
Henry’s Law Constant1.09 × 10−5 Pa·m3·mol−1 (25 °C)
Half-life (t1/2)25–65 h
Excretion72% absorbed and metabolized in liver, 28% excreted in feces
Metabolites in urineCBZ, CBZ-epoxide, CBZ-diol, CBZ-acridan, 2-OH-CBZ, 3-OH-CBZ
Dosage800–1200 mg/day
Other informationAutoinduction i.e., induces its own metabolism during continued intake
Note: a Log D is the logarithm of the distribution coefficient, which is the ratio of the sum of concentrations of all forms of the compound (ionised and unionised) in octanol and water at a given pH [3].
Table 2. Occurrences of carbamazepine (CBZ) in municipal wastewater treatment plant (WWTP) effluent and freshwater.
Table 2. Occurrences of carbamazepine (CBZ) in municipal wastewater treatment plant (WWTP) effluent and freshwater.
CountryW WWTP EffluentSurface WaterGroundwater
No. of WWTPs *Concentration (ng/L)Concentration (ng/L)Concentration (ng/L)
Canada733–426 [a]0.7–126 [b]10–49 [c]
Germany51075–6300 [d]81–1100 [e]1–100 [f]
Japan2081–86 [g]0.1–34.7 [h]1.64–97 [i]
South Korea1173–729 [j]6–61 [k]NA
Taiwan4290–960 [l]0.5–120 [m]NA
UK3152–4596 [n]9–327 [o]425–3600 [p]
USA1633–270 [q]2–172 [r]1.5–42 [s]
Notes: Data sources: [a] [31,32,33,34]; [b] [31,32,35]; [c] [36,37]; [d] [29,38,39,40]; [e] [38,40,41]; [f] [42,43]; [g] [44,45,46]; [h] [45,47]; [i] [46,47]; [j] [48,49]; [k] [48,49,50,51]; [l] [52,53]; [m] [52]; [n] [54,55,56]; [o] [54,55,57,58]; [p] [30,59,60]; [q] [28,61,62,63]; [r] [61,63,64,65]; [s] [66,67]. * number of WWTPs surveyed for analyzing the concentration of CBZ; NA: not available.
Table 3. Seasonal variations in CBZ concentration along with its annual consumption rate.
Table 3. Seasonal variations in CBZ concentration along with its annual consumption rate.
CountryWWTP EffluentConsumption Rate (tons/year)
No. of WWTPs*Concentration (ng/L)
WinterSummer
Australia31480 [70]NA10 [16]
Austria11952 [71]1337–1594 [3,71]6 [16]
111000 [72]1500 [72]
Canada4426 [32]300 [33]28 [16]
Finland12500 [6]NA4.8 [16]
3380–470 [73]NA
Germany31900 [38]2100 [40]76 [16]
Korea4103–195 [48]5–6 [48]9.2 [48]
Switzerland21000 [74]950 [74]4.1 [75]
3400–800 [26]200–600 [26]
UK3 (Winter)
2 (Summer)
637–950 [56]2499 [54]40 [16]
Note: * Number of WWTPs surveyed for analyzing the concentration of CBZ in winter and summer; NA: not available.
Table 4. Effect of CBZ exposure on aquatic species under different exposure conditions.
Table 4. Effect of CBZ exposure on aquatic species under different exposure conditions.
SpeciesCritical EffectsExposure TimeLC50 (mg/L)EC50 (mg/L)NOEC (mg/L)LOEC (mg/L)References
Water flea
Daphnia magnaMortality2 days111---[101]
Chironomus dilutusSurvival10 days47.310.2--[102]
Hyalella aztecaGrowth10 days1.59.5--[102]
Brachionus calyciflorusReproduction inhibition2 days--0.40.8[100]
Ceriodaphnia dubiaReproduction inhibition7 days--0.0250.1[100]
Daphnia magnaMobility inhibition2 days-77.7--[100]
Daphnia magnaImmobilization2 days-97.8--[103]
Bacteria
Aliivibrio fischeriBioluminescence30 min-64.2--[100]
Vibrio fischeriBioluminescence5 min-87.4--[103]
Algae
Chlorella vulgarisGrowth inhibition24 h-110.9--[103]
Desmodesmus subspicatusInhibition of average growth rate3 days-74--[104]
Raphidocelis subcapitataGrowth inhibition4 days 89>100>100[100]
Rainbow trout
Oncorhynchus mykissCondition factor42 days---2[105]
Oncorhynchus mykissAntioxidant responses in muscle42 days---0.001[105]
Oncorhynchus mykissChanges in RNA-DNA ratio42 days---2[105]
Zebrafish
Danio rerioDevelopmental effects3 days>24585.630.6-[106]
Danio rerioEmbryos and larvae mortality10 days--2550[100]
Mycorrhizal fungus
Glomus intraradicesSpore production28 days-0.1--[107]
Duckweed
Lemna minorInhibition of average growth rate7 days-25.5--[104]
Cnidarian
Hydra attenuateMorphological changes4 days29.415.515[108]
Notes: LC50: lethal concentration to kill/inactivate 50% of population; EC50: effective concentration that gives half-maximal response; NOEC: no-observed effect concentration; and LOEC: lowest-observed effect concentration. “-”: not available.
Table 5. CBZ removal by conventional activated sludge (CAS) and membrane bioreactor (MBR) under different operating conditions.
Table 5. CBZ removal by conventional activated sludge (CAS) and membrane bioreactor (MBR) under different operating conditions.
Process TypeInfluent Concentration (ng/L)SRT (days)HRT (h)Aerobic/AnoxicRemoval (%)References
CAS240312aerobicnegligible[128]
1561011.5bothnegligible[126]
350–185052–11412.5–13.6bothnegligible[71]
3502–201.5–20bothnegligible[6]
15–2703.8–8.47.1–9.4aerobicnegligible–80[129]
1000107.3bothnegligible[74]
670–70419NAaerobic0[130]
10–2011–159–17bothnegligible–25[131]
200–60015–2516–24aerobicnegligible[132]
MBR240infinite14aerobicnegligible[128]
156>6015aerobicnegligible[126]
156>607.2aerobicnegligible[126]
704–185010–550.5–4bothnegligible[71]
10001613both25[74]
704–120022NAbothnegligible[130]
750,000infinite24near-anoxic68[133]
58826aerobic40[27]
14007024aerobic10[134]
Notes: CAS: conventional activated sludge process; MBR: membrane bioreactor; SRT: solid retention time; HRT: hydraulic retention time; NA: not available.
Table 6. Removal of CBZ by white-rot fungi and their extracellular enzymes.
Table 6. Removal of CBZ by white-rot fungi and their extracellular enzymes.
Bioreactor TypeWRF Species/Enzyme TypeHRT (h)/Incubation Time (days)Initial Concentration (ng/L)Removal Efficiency (%)References
Removal by whole-cell WRF
Stirred tank (batch)Bjerkandera sp. R1141,000,00099[156]
Stirred tank (batch)B. adusta (Laccase, LiP and MnP)141,000,00099[156]
Stirred tank (batch)T. versicolor (Laccase, LiP and MnP)210,00075[157]
Stirred tank (batch)T. versicolor (Laccase, LiP and MnP)1100,0002[158]
Stirred tank (batch)P. ostreatus (Florida N001) (Laccase, MnP)32100050[155]
Stirred tank (batch)P. ostreatus (Florida F6) (Laccase, MnP)32100060[155]
Stirred tank (batch)P. ostreatus (PC9) (Laccase, MnP)32100099[155]
Stirred tank a (continuous)P. chrysosporium (MnP, LiP)24500,00025–60[159]
Fluidized bed a (continuous)T. versicolor (Laccase, MnP, LiP)72200,00095.6[160]
Membrane bioreactor (continuous) aT. versicolor (Laccase, MnP, LiP)48500020[161]
Removal by extracellular enzymes
Stirred tank (batch)Laccase from A. oryzae1100,000<5[162]
Stirred tank (batch)Laccase from T. versicolor210,0005[157]
Stirred tank (batch)LiP from P. chrysosporium4NA10–15[163]
Membrane bioreactor (continuous)Laccase from A. oryzae85000<5[162]
Membrane bioreactor (continuous)Laccase from A. oryzae850007[164]
Notes: a Indicates fungal bioreactor operated under non-sterile conditions; NA: not available.
Table 7. CBZ removal from various water matrices by nanofiltration (NF) and reverse osmosis (RO) membranes under different operating conditions.
Table 7. CBZ removal from various water matrices by nanofiltration (NF) and reverse osmosis (RO) membranes under different operating conditions.
Membrane Type (Pore Size)ConfigurationWater MatrixInitial CBZ Concentration (ng/L)Applied Pressure (psi)Removal (%)References
NF (0.34 nm)Flat-sheetGroundwater84.5NA>98[176]
NF (0.27 nm)Flat-sheet (tight)MBR permeate15015097.3 ± 0.6[178]
NF (0.42 nm)Flat-sheet (loose)MBR permeate1507571.2 ± 3.1[178]
NFFlat-sheetWWTP effluent500–8504.356[179]
NFFlat-sheetWWTP effluent500–850108[179]
NF (0.84 nm)Flat-sheetPrimary effluent20007274[180]
NFSpiral-woundHospital wastewater10009888[181]
NF (0.84 nm)Flat-sheetSynthetic wastewater750,00026180 a, 60 b[182]
NF (0.68 nm)Flat-sheetSynthetic wastewater750,00026195 a, 90 b[182]
NF (0.84 nm)Flat-sheetSynthetic wastewater750,00098670 a, 20 b[183]
NF (0.84 nm)Flat-sheetSynthetic wastewater750,00026180 a, 90 b[184]
RO (0.34 nm)Flat-sheetGroundwater84.5NA>98[176]
ROFlat-sheetMBR permeate15025091 ± 8.4[178]
ROFlat-sheetMBR permeate15015097.9 ± 1.5[178]
ROFlat-sheetPrimary effluent100072100[180]
ROSpiral-wound moduleHospital wastewater100019699[181]
Notes: a Virgin membrane; b fouled membrane; NA: not available.
Table 8. Performance of granular activated carbon (GAC) and powdered activated carbon (PAC) for CBZ removal from various water matrices.
Table 8. Performance of granular activated carbon (GAC) and powdered activated carbon (PAC) for CBZ removal from various water matrices.
AC TypeWater MatrixInitial CBZ concentration (ng/L)Contact Time (min)Removal (%)References
GACSynthetic wastewater1000144099–80 a[197]
Ozonation effluent3688 b[198]
Groundwater915>75 c[199]
Surface water2599[200]
Disinfected surface water6001.5–379 d[201]
WWTP effluent6730 e[202]
GAC added to an activated sludge based bioreactor22,000144043 f[203]
WWTP effluent400010080 g[204]
WWTP effluent30–100130>99[190]
PACSurface water7830095 h[205]
Surface water7830036 i[205]
MBR permeate10003099.4 j[206]
Surface water5024080[207]
WWTP effluent30–10020–4095–100[190]
Notes: a The specific throughput of CBZ in a carbon layer of 80 cm was 50 m3/kg when removal decreased to 80%; b the daily production of this drinking water treatment plant was 28,000 m3, and this removal was observed in winter; c the system comprised 20 granular activated filters (volume = 150 m3 each), experimental duration was 4 months; d the daily production of this drinking water treatment plant was 235,000 m3 (experimental duration = 3 weeks); e total GAC volume was 1900 m3; f the GAC concentration in the bioreactor was 1000 ppm, and experimental duration was 33 days; g the initial GAC concentration was 20 mg/L; h the PAC dose was 35 mg/L; i the PAC dose was 5 mg/L; and j the PAC dose was 10 g/L, “–”: not available.
Table 9. Performance of various advanced oxidation processes (AOPs) for CBZ removal.
Table 9. Performance of various advanced oxidation processes (AOPs) for CBZ removal.
AOP TypeCBZ Initial Concentration (ng/L)Operating ConditionsRemoval (%)References
Ozonation 1000Dose = 0.5 mg/L
Contact time = 20 min
>99[197]
35Dose = 1-1.5 mg/L
Contact time = 10 min
>97[197]
9Dose = 0.2 mg/L
Contact time = 15 min
>99[199]
8 × 105Dose = 1 mg/L
Contact time = 10 min
>99[215]
3.8Dose = 1.5–2 mg/L
Contact time = 20 min
80–99[216]
1.18 × 105Dose = 0.1–2 mg/L
Contact time = 10 h
80–99[217]
170Dose = 0.8 mg/L
Contact time = 24 min
100[218]
4.72 × 105Dose = 1.5–4 mg/L
Contact time = 20 min
>99[219]
UV alone5000UV Wavelength = 254 nm
Energy output = 83 W
Irradiation time = 60 min
20[220]
NAUV Wavelength = 200–280 nm
Energy output = 120 W
Irradiation time = NA
7[221]
1.5 × 107UV Wavelength = 254 nm
Energy output = 220 W
Irradiation time = 2 h
16[222]
5 × 106UV Wavelength = 254 nm
Energy output = 400 W
Irradiation time = 30 min
<5[223]
19–59UV Wavelength = 254 nm
Energy output = 10 W
Irradiation time = 3 min
<10[224]
UV/H2O24.72 × 106UV Wavelength = 254 nm
Energy output = 83 W
H2O2 dose = 170 mg/L
Irradiation time = 60 min
90[187]
210UV Wavelength = 254 nm
Energy output = 30 W
H2O2 dose = 2–20 mg/L
Irradiation time = 20 min
14–74[225]
1000UV Wavelength = 254 nm
Energy output = 20 W
H2O2 dose = 5 mg/L
Irradiation time = 15 min
60[226]
50,000UV Wavelength = 254 nm
Energy output = 83 W
H2O2 dose: 10–200 mg/L
Irradiation time = 60 min
90–99[220]
19–59UV Wavelength = 254 nm
Energy output = 10 W
H2O2 dose = 5 mg/L
Irradiation time = 3 min
20[224]
2.36 × 105UV Wavelength = 254 nm
Energy output = 1000 W
H2O2 dose = 5 mg/L
Irradiation time = NA
90[227]
UV/Cl21000UV Wavelength = 254 nm
Energy output = 80 W
Cl2 dose = 1 mg/L
Irradiation time = 20 min
55[226]
19–59UV Wavelength = 254 nm
Energy output = 10 W
Cl2 dose = 5 mg/L
Irradiation time = 1.5 min
30–60[224]
19–59UV Wavelength = 254 nm
Energy output = 10 W
Cl2 dose = 3 mg/L
Irradiation time = 3 min
50[224]
UV/TiO24 × 106UV Wavelength = 200–296 nm
Energy output = 1000 W
TiO2 dose = 100 mg/L
Irradiation time = 9 min
99[228]
5 × 106UV Wavelength = 254 nm
Energy output = 250 W
TiO2 dose = 20–500 mg/L
Irradiation time = 30 min
80–98[223]
5 × 106UV Wavelength = 254 nm
Energy output = 400 W
TiO2 dose = 20–500 mg/L
Irradiation time = 30 min
90–99[223]
Photo-Fenton5 × 107UV Wavelength = 254 nm
Energy output = 400 W
Fe2+ dose = 5 mg/L
Irradiation time = 15 min
>99[229]
1 × 105UV Wavelength = 200–296 nm
Energy output = 30 W
Fe2+ dose = 5 mg/L
Irradiation time = 1.5 h
95[230]
Note: “NA”: not available.
Table 10. CBZ removal by combined and integrated treatment systems.
Table 10. CBZ removal by combined and integrated treatment systems.
Treatment SystemsCBZ Initial Concentration (ng/L)Operating ConditionsRemoval (%)References
Integrated MBR-UV/TiO21 × 107SRT = 60 d
HRT = 50 h
UV wavelength = 360 nm
TiO2 dose = NA
up to 95[238]
Integrated MBR-PAC5000SRT = infinite
HRT = 24 h
PAC dose = 0.1 g/L
50[210]
Integrated MBR-PAC5000SRT = infinite
HRT = 24 h
PAC dose = 0.5 g/L
90[210]
Integrated Gamma radiation-CAS (batch experiment)1.7 × 107Incubation time = 10 d
Radiation dose = 800 Gy
>99[239]
Integrated PAC-UF3 × 105PAC dose = 5–10 mg/L
Contact time = 1.5 h
40[240]
Integrated MBR-PAC7.5 × 105SRT = infinite
HRT = 24 h
PAC dose = 0.1 and 1 g/L
34 and 90[211]
Integrated MBR-PAC2 × 104SRT = infinite
HRT = 24 h
PAC dose = 1 g/L
>99[241]
Integrated MBR-PAC390–1800SRT = 20 d
HRT = 24 h
PAC dose = 1 g/L
99.4[206]
Integrated CAS-GAC2 × 104SRT = NA
HRT = NA
GAC dose = 100–1000 mg/L
10–50[203]
MBR followed by GAC5000SRT of MBR = infinite
HRT of MBR = 24 h
GAC contact time = 7 min
98[242]
Ozonation followed by biological sand column2.06HRT of sand column = 5–6 d
Ozone contact time = NA
80[243]
Note: NA: not available.

Share and Cite

MDPI and ACS Style

Hai, F.I.; Yang, S.; Asif, M.B.; Sencadas, V.; Shawkat, S.; Sanderson-Smith, M.; Gorman, J.; Xu, Z.-Q.; Yamamoto, K. Carbamazepine as a Possible Anthropogenic Marker in Water: Occurrences, Toxicological Effects, Regulations and Removal by Wastewater Treatment Technologies. Water 2018, 10, 107. https://doi.org/10.3390/w10020107

AMA Style

Hai FI, Yang S, Asif MB, Sencadas V, Shawkat S, Sanderson-Smith M, Gorman J, Xu Z-Q, Yamamoto K. Carbamazepine as a Possible Anthropogenic Marker in Water: Occurrences, Toxicological Effects, Regulations and Removal by Wastewater Treatment Technologies. Water. 2018; 10(2):107. https://doi.org/10.3390/w10020107

Chicago/Turabian Style

Hai, Faisal I., Shufan Yang, Muhammad B. Asif, Vitor Sencadas, Samia Shawkat, Martina Sanderson-Smith, Jody Gorman, Zhi-Qiang Xu, and Kazuo Yamamoto. 2018. "Carbamazepine as a Possible Anthropogenic Marker in Water: Occurrences, Toxicological Effects, Regulations and Removal by Wastewater Treatment Technologies" Water 10, no. 2: 107. https://doi.org/10.3390/w10020107

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