Next Article in Journal
Multi-Step Sequence Flood Forecasting Based on MSBP Model
Next Article in Special Issue
Removal of Pharmaceuticals from Wastewater: Analysis of the Past and Present Global Research Activities
Previous Article in Journal
Testing Landscape, Climate and Lithology Impact on Carbon, Major and Trace Elements of the Lena River and Its Tributaries during a Spring Flood Period
Previous Article in Special Issue
Removal of Carbamazepine onto Modified Zeolitic Tuff in Different Water Matrices: Batch and Continuous Flow Experiments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Effectiveness of Advanced Oxidation Processes in Wastewater Treatment: State of the Art

by
Laura Antiñolo Bermúdez
1,
Jaime Martín Pascual
1,
María del Mar Muñio Martínez
2 and
Jose Manuel Poyatos Capilla
1,*
1
Department of Civil Engineering and Institute of Water Research, University of Granada, 18071 Granada, Spain
2
Department of Chemical Engineering, University of Granada, 18071 Granada, Spain
*
Author to whom correspondence should be addressed.
Water 2021, 13(15), 2094; https://doi.org/10.3390/w13152094
Submission received: 31 May 2021 / Revised: 23 July 2021 / Accepted: 27 July 2021 / Published: 30 July 2021

Abstract

:
In recent years, many scientific studies have focused their efforts on quantifying the different types of pollutants that are not removed in wastewater treatment plants. Compounds of emerging concern (CECs) have been detected in different natural environments. The presence of these compounds in wastewater is not new, but they may have consequences in the future. These compounds reach the natural environment through various routes, such as wastewater. This review focuses on the study of tertiary treatment with advanced oxidation processes (AOPs) for the degradation of CECs. The main objective of the different existing AOPs applied to the treatment of wastewater is the degradation of pollutants that are not eliminated by means of traditional wastewater treatment.

Graphical Abstract

1. Introduction

Today, there is strong environmental concern and a focus on minimising the impact of human activities on the environment. For years, the overexploitation to which the planet has been subjected has been unsustainable, highlighting the pressure exerted on the world’s water reserves. The high demand of population centres, massive industrial production, and agricultural activities have caused an unsustainable situation for the available water resources [1]. Due to this, wastewater treatment becomes absolutely essential for the subsequent return to the natural environment in the most efficient way possible. However, current technologies used in wastewater treatment are not efficient against certain types of pollutants that are able to remain in the treated wastewater, which is returned to the natural environment, thus reaching the water reserves [2,3]. Once in the environment, they are characterized by environmental persistence and threats to human health [4,5]. Due to the low concentrations of contaminants and diversity in nature, the removal of these contaminants poses numerous challenges [6].
Due to this, the application of tertiary treatments that are effective and economically viable in wastewater treatment plants has been investigated in recent years.
These pollutants are a global problem and are not limited to just one region; they have been detected in waters around the world. In América, the Biobío River (Chile) has population settlements that, due to their economic activities, such as agriculture or forestry, incorporate compounds of emerging concern, such as pharmaceuticals, personal care products, and pesticides, into the aquatic environment [7]. The presence of 28 antibiotics was detected in six rivers and a drinking water storage catchment in watersheds of South-East Queensland, Australia [8]. In the Somes River (Transylvania, Romania), compounds including musk fragrances, pharmaceutics, metabolites and intermediates were detected [9]. An analysis of the Ebro River (Spain) revealed the presence of 31 pharmaceuticals; these compounds were detected in wastewater treatment plant (WWTP) effluents, with the highest concentrations found in effluents of the Zaragoza WWTP [10]. It was detected the occurrence of 20 illicit and therapeutic pharmaceuticals and metabolites in surface waters influenced by WWTP discharge and in wastewater effluents in Nebraska [11]. In the Pearl River (China), eight types of antibiotics were detected in the water [12]. Laboratory analysis revealed that antibiotics were widely distributed in Baiyangdian Lake, China [13]. In rivers and lakes in the agricultural zone of Northeast Denmark, 17 emerging contaminants (diclofenac, MCPA, caffeine and TCEPT were the most abundant) were identified and quantified in surface waters (including pesticides, personal care products [PCPs], pharmaceuticals, plasticisers, and fire retardants) [14]. In the affluents of the Delaware River (Pennsylvania, EE.UU), Triclosan (PCP), which is widely used as an antiseptic, was found in high concentrations, which proved to be a high-risk compound for aquatic life. In this river, Diphenhydramine (antibiotic) was also found whose effect is similar to Triclosan [15].
Pollution is not just limited to water; it is also ‘absorbed’ by sediments and living organisms that are found where waters contaminated by CECs are found. In Baiyangdian Lake (China), quinolones were detected in sediments and aquatic plants. Quinolones and macrolides are often found in aquatic animals and birds [13].
The present study focuses on the different types of advanced oxidations and their efficiency against the different polluting CECs. These pollutants are a global problem and are not limited to just one region but have been detected in waters around the world. This paper tries to give a global view of the problem by compiling the most relevant studies carried out during the last decade and centralizes other studies that review the problem.

2. Advanced Oxidation Processes for Wastewater Treatment

Advanced oxidation processes (AOPs) have been used in urban and industrial wastewater treatment as an effective option for the treatment of non-biodegradable compounds [16].
The complete mineralisation to water, inorganic compounds and CO2, water of the contaminants in the wastewater, or at least their transformation into more innocuous compounds, can be achieved by applying a chemical treatment using AOPs [17]. There is a wide range of advanced oxidation treatments where different reagents are used with the aim of producing OH radicals, which are very powerful oxidising agents. AOPs are classified as either homogeneous or heterogeneous. Homogeneous processes include processes that use energy and those that do not. This classification is shown in Figure 1.
Homogeneous processes that use energy and ultraviolet radiation use reagent systems that include O3, H2O2, and photo-Fenton reactions. They all produce OH radicals. For the degradation of compounds whose UV absorption range is within the corresponding range of the spectrum, these AOPs are generally used [17,18]. Homogeneous processes that use ultrasound energy form hydroxyl radicals because of the extreme conditions generated by ultrasound, which are capable of cracking water molecules. They can be combined with other AOPs. Finally, homogeneous processes that use energy have anodic oxidation, electrochemical oxidation, and electro-Fenton processes. This type of process breaks down compounds and molecules for degradation using electrical energy. The electron transfer occurs by the intervention of hydroxyl radicals [17].
Homogeneous processes that do not use energy are hydrogen peroxide and catalyst processes, ozonation with hydrogen peroxide, and ozonation in an alkaline medium. The difference between the hydrogen peroxide process and catalysation with H2O2/UV is found in the area of radical formation of OH, which is less in the process that does not use energy [17].
In heterogeneous AOPs, catalysts are often used to produce the degradation of the contaminant compounds. Compared to homogeneous processes, the catalysts used in heterogeneous processes have the advantage of being able to separate the product more easily [17].
The reactions produced in the different types of AOPs are exposed in Table 1.

3. Methodology

The search for relevant literature was carried out by searching the Scopus database using the following keywords: “advanced oxidation processes”, “concern emerging contaminants” and “wastewater treatment”. The first search resulted in 51 articles, and a review article for 2010 (67 additional references) was used to provide an overview since 2000. The less relevant articles were discarded following the criterion of looking for articles of removal of the same pollutants with more easily reproducible operating conditions. Once the search list was generated, it was manually checked to exclude studies that were not relevant and to give preference to articles that eliminated several contaminants. In addition, relevant studies found in the bibliography of the selected studies were included in the literature identification list. Articles were selected from the years 2000 to 2021 to check the progress of the technology studied and only articles written in English were selected. This extended range of years was selected because it was considered important to give an overview of the evolution of advanced oxidation processes.

4. Compounds of Emerging Concern (CECs)

Water is the essential resource for life; everybody depends on it, and because of this it is one of the most threatened resources on the planet. Aquifers and water reserves are constantly affected by overexploitation and pollution associated with human activities. For this reason, the European Union (EU), as well as most governments around the world, is focusing efforts on achieving good quality of their resources. The EU created an observation list of emerging pollutants. These substances do not currently follow an elimination legislation, but it has been studied and shown that they are present in rivers, lakes, aquifers, and natural environments; therefore, they can pose a somewhat long-term risk and the study of their elimination is essential to prevent their entry into ecosystems. These substances can be divided into four main groups: pharmaceuticals (PHs) (e.g., antibiotics and analgesics), PCPs (e.g., fragrances and antiseptics), pesticides (e.g., atrazine and dimethoate) and illicit drugs (e.g., opioids and amphetamines).
The introduction of this type of compound into the environment has its origin in human activity. Transformation products and metabolites may enter the aquatic environment and eventually reach drinking water if the CECs are not eliminated during wastewater treatment [43].
The EU, through Directive 2008/105/EC, establishes an observation list of substances for monitoring purposes at the European level. This list is updated every 24 months by the Water Framework Directive of the European Union. These modifications are made taking into account several criteria, among which are the research programmes in force and the characterisation and results obtained by the member states of the EU of their river basin districts. Each member state selects a representative monitoring station, and the monitoring frequency shall not be less than once a year. The last update of this list was made in December 2018, and a final programme of measures is foreseen for December 2021 to be fully operational by December 2024 (https://www.boe.es/buscar/doc.php?id=DOUE-L-2013-81677 accessed on 7 April 2020). In addition, in 2005, the European Commission funded the NORMAN project. (https://www.norman-network.net accessed on 7 April 2020). The NORMAN network enhances the exchange of information on emerging environmental substances and encourages the validation and harmonisation of common measurement methods and monitoring tools so that the requirements of risk assessors and risk managers can be better met. It specifically seeks both to promote and to benefit from the synergies between research teams from different countries in the field of emerging substances. This network includes the competent authorities/reference laboratories, research centres and academia, industry stakeholders, government institutions and standardisation bodies. The mission of the NORMAN network is to enhance the exchange of information and collection data environmental substances, encourage the validation and harmonisation of common measurement methods and monitoring tools so that the demands of risk assessors can be better met and ensure that knowledge of emerging pollutants is maintained and developed by stimulating coordinated, interdisciplinary projects on problem-oriented research and knowledge transfer to address identified needs.
In the case of the US, the Environmental Protection Agency (EPA; https://www.epa.gov/ccl accessed on 28 May 2020) publishes a Contaminant Candidate List (CCL) every five years, with the last being published in 2016. The EPA has a candidate list of 97 chemical contaminants/groups of contaminants and 12 microbial contaminants (including PHs and PCPs). Within this area, each member state, as with Europe, adapts this policy to its territory. In California, the State Water Resources Control Board has a Recycled Water Policy where scientific advisors provide guidance on how to monitor CECs for prioritisation in legislation [44].
In the case of Japan and its Ministry of Health, they focus their efforts on evaluating the environmental risk posed by pharmaceutical products, and with it their marketing is approved or denied. The same occurs in the case of Canada and Australia, which are based on standard physicochemical tests to determine the biodegradability of the compounds, without following a specific list that is monitored or regulated (World Health Organization, 2012).
In China, the Ministry of Environmental Protection establishes a plan for the control and prevention of environmental risks of chemical substances, including pharmaceuticals. This was launched in 2013 [45].
In general, the release of antibiotics into the environment is believed to be a major concern as it can increase the appearance of resistant bacteria in the environment. However, treatments done by man, as well as natural filtration, can be penetrated by emerging compounds and cause them to pose a potential risk to the supply of drinking water [46].

4.1. Pharmaceuticals Compounds (PH)

The appearance of different types of pharmaceuticals in waters has been extensively studied in recent years. A large number of compounds associated with the removal of urban and industrial wastewater have been detected in groundwater and surface. The current wastewater treatment technology is insufficient for these pollutants, and they are not retained in the treatment plants, which is why many of these pharmaceutical residues reach the ecosystem, thus joining the water cycle. In primary treatments, while some pharmaceuticals remain in the water, others can be removed by adsorption, e.g., naproxen, ibuprofen, iopromide, and sulfamethoxazole [47,48]. Antibiotics and anti-inflammatories are eliminated in subsequent biological treatments by 30–75%.
The concern about pharmaceuticals is their chronic toxicity and not their acute toxic effects. In general, pharmaceuticals are biologically active compounds that are not easily biodegradable. They may cause side effects in non-target organisms or cause a similar function as they do in their intended users [49]. The evolution and spread of antimicrobial resistance are recognised as one of the major global health challenges of the 21st century by major regulatory, economic, and political bodies, including the European Commission (EC), the United Nations (UN), and the World Health Organisation (WHO), and the surveillance of critical hotspots through intensive monitoring is recommended, including urban wastewater treatment plants, aimed at reducing its propagation [50].
The uncontrolled consumption of pharmaceuticals by the population, where many of them are consumed without medical control, such as ibuprofen or diclofenac, has led to their detection in water intended for human consumption [8]. Thus, hospital wastewaters, landfill leachates, and municipal wastewaters are always considered ‘hot spots’ for environmental water contamination by pharmaceuticals [51].
During therapeutic periods, humans excrete pharmaceuticals and their metabolites. In the case of pharmaceuticals used for veterinary medicine, when animal wastes are sprayed on agricultural fields to fertilise them, these substances are released into the environment [52].
The different physico-chemical properties of these compounds, as well as their metabolites or degradation by-products thereof, can affect the waters they arrive at and/or be retained in the soil into which the treated waters are discharged, which causes a negative effect on the ecosystem. The annual usage of antibiotics has been estimated to be between 100.000 and 200.000 tons globally, with more than 25.000 tons used each year in China [12,53].
The pharmaceuticals are divided into several types, analgesics and antibiotics are the most widely used of them.
Antibiotics: These are pharmaceuticals that are used to fight bacterial infections in organisms. Their use is not limited to humans, but they are also used in animals and in the food industry to preserve food. Therefore, they are widely produced and consumed, causing large amounts of them to be present in wastewater.
Analgesics and anti-inflammatories: These are the pharmaceuticals with the highest global consumption and those most used by the population as self-medication [54]. Compounds such as ibuprofen and diclofenac belong to this group.
Antidepressants and anticonvulsants: These medications help improve the way the brain uses certain natural chemicals and can help combat depression and epileptic attacks.
Lipid regulators: These are regulators of compounds present in blood, such as enzymes.
β-blockers: These compounds work primarily in the heart and blood vessels.
Many PHs are also thought to mimic natural hormones in the body, hence their classification as endocrine-disrupting chemicals (EDCs) [52]. These compounds are defined as natural and/or synthetic substances that can include naturally generated estrogenic hormones, e.g., estrone (E1) and 17b-oestradiol (E2), and, therefore, are ubiquitous in aquatic environments receiving wastewater effluents [49]. Oestrogens and xenoestrogens are excreted into wastewater and reach wastewater treatment plants as organic contaminants, where they are only partially eliminated [55]. Some of the most studied PHs and their classifications are listed in Table 2.
The application of AOPs based on studies of the removal of pharmaceutical products present in wastewater that are not eliminated by other technologies will be analysed. During the oxidation of urban wastewater, transformation products (TPs) of the antibiotics present can be formed, which may be more biologically potent, less biodegradable, or more toxic compared to the parent compounds [56].
One of the most studied types of AOPs for the treatment of wastewater by the authors is the UV/H2O2 process. Photolysis of hydrogen peroxide is produced under ultraviolet radiation and it does not depend on pH. A H2O2/UV system can totally mineralise any organic compound, reducing it to CO2 and water [17]. In addition, in this method, you can use sunlight instead of UV lamps. This is very interesting as in places where there are many hours with sunlight, this would be a low-cost application. In the case of antibiotics, this process can degrade compounds such as ciprofloxacin, which is a compound that is observed in wastewater. Ciprofloxacin is a compound included in the observation List of substances for the EU. Michael et al. [50] studied ciprofloxacin removal using H2O2/UV and H2O2/sunlight processes by comparing them to each other. H2O2/UV treatment was able to remove ciprofloxacin (90 min, 0.9 kJ L–1), while the H2O2/sunlight process was able to remove ciprofloxacin in 60 min (8 kJ L–1). Monteoliva-García et al. [57] reported the elimination of ciprofloxacin (concentrations from 22.30 to 98.53 μg L−1) by applying H2O2/UV treatment ([H2O2] = 25, 50, and 100 mg L−1), with total elimination after 20 min of treatment. Yuan et al. [58] completely eliminated ciprofloxacin by applying H2O2/UV from an initial CIP concentration of 5 μM and using an LP-Hg lamp. Rosal et al. [59] studied ciprofloxacin removal using O3 and O3/H2O2-based AOPs under similar conditions. While removal rates were very good at 98% and more than 93%, respectively, TOC removal was much more efficient in the O3/H2O2 process, obtaining more than 90% removal compared to 15% in the other process. De Witte et al. [60] obtained a very similar result of eliminating the TOC of 95% by applying O3-based AOPs. Most of the treatments range between pH 7.5 and 8, this is very interesting and useful as this is within the pH range detected in urban wastewater. Therefore, for ciprofloxacin, it can be accepted that it can be almost completely removed from wastewater by AOPs, with the most favourable carbon treatment being O3/H2O2 treatment.
Another commonly used antibiotic is sulfamethoxazole. Michael et al. [50] studied its elimination by means of H2O2/UV and H2O2/sunlight treatment to determine which was more efficient. The H2O2/UV treatment was able to eliminate sulfamethoxazole after 90 min (0.9 kJ L−1), while the process with H2O2/UV only eliminated 46% after 300 min and 42 kJ L−1. Lekkerkerker-Teunissen et al. [61] also studied the elimination of sulfamethoxazole by applying UV/H2O2 treatment, which was able to eliminate it by more than 90% (UV doses ranged from 300–700 mJ cm–2). Yang et al. [62] applied the UV/H2O2 treatment and succeeded in removing sulfamethoxazole (SMX) after 60 min. Other authors applied O3-based AOPs for SMX elimination, obtaining excellent elimination results (99.9%) from a high concentration of SMX (30 mg L–1, 22 °C) [63]. The pH range of the various treatments is very wide, ranging from 2–10. This is really useful; although, the water can be treated effectively at a natural pH but would achieve satisfactory eliminations in the event of temporary fluctuations in the plant.
For the treatment of amoxicillin, an antibiotic widely used because of its great efficacy, its elimination has been studied using various POP processes. By applying UV/H2O2 treatment, it was possible to irradiate the compound with a low-pressure lamp at an incident light intensity of 8 × 10–7 Einstein L–1 s–1 and one at an H2O2 concentration of 0.4–10 mM at 99% in only 20 min [64]. However, in the study by Elmolla et al. [65], which carried out UV/TiO2 treatment, only 20% elimination (UV 365 nm) and a high pH dependency were reported, achieving the highest degradation at pH 11. The authors also studied the addition of H2O2 at pH 5 (ambient) with a TiO2 concentration of 1 g/L, achieving complete degradation of amoxicillin in only 30 min of heterogeneous photocatalysis treatment (UV/H2O2/TiO2). The amoxicillin treatment with O3-based AOPs also provided very good results. At an initial amoxicillin concentration of 5.0 × 10–4 M, with pH 5.5 (buffer) and a flow of O3 1.6 × 10–4 M, 90% was eliminated [66]. Therefore, amoxicillin is a compound that reacts very well to advanced oxidation treatments, achieving very good removal performance. In conclusion, the highest degradation for this pharmaceutical compound occurs at pH 11, which can be problematic; although, high clearance is also achieved at pH 5, where mineralisation of the pharmaceutical (phenolic ring hydroxylation) occurs.
Another important antibiotic is metronidazole, which is used for bacterial infections affecting various parts of the body (vaginal infection primarily and others, such as the stomach, liver, skin, and brain). Several authors have studied their removal by applying different types of AOPs and comparing them with each other. Shemmet et al. [67] applied the UV/H2O2 method under four operating conditions, obtaining elimination rates of between 58% and 67% (6.0 μM metronidazole; 1.5 mW cm–2; 50 mg L–1 H2O2). Furthermore, ref. [67] applied the UV/H2O2/Fe2+ method and obtained 94% elimination as the best result, also starting from a 6.0 μM Metronidazole concentration, and they compared it with the H2O2/Fe2+ process under the same reaction conditions, where the maximum elimination result was 76%. These results show significant differences in behaviour under equal conditions, making the advanced photo-Fenton oxidation process the most effective against this compound. Rosal et al. [59] studied metronidazole elimination using ozone technology, comparing O3-based AOPs and O3/H2O2 under the same conditions, where they obtained elimination percentages of 91% (15% TOC) and 92% (above 90% TOC), respectively, highlighting the role H2O2 key to eliminating toxicity, the same as other antibiotics such as ciprofloxacin. The pH ranges studied are very broad, but the data seem to indicate that better removal rates are achieved at increasingly acidic pH values in UV/H2O2/Fe2+ processes, with this AOP showing a removal rate of over 90%. In the case of lincomycin, Andreozzi et al. [68] applying different types of AOPs (UV/H2O2 and O3-based AOPs). O3-based AOPs shown a total elimination in 2 min of treatment and no toxicity in one hour of treatment. UV/H2O2 obtained the elimination percentages of 80% in 3 min of treatment and no toxic product generation. De Witte et al. [60] applying O3-based AOPs in the elimination of levofloxacin and obtained a removal of 99.9 %.
The antiepileptic pharmaceutical carbamazepine, which is present in wastewater, has been studied for its removal by the advanced UV/H2O2 removal process. At low H2O2 concentration, carbamazepine removal is not appropriate despite UV treatment with medium- and low-pressure lamps [61]. However, at H2O2 concentrations of 5 mg/L and above, up to 99.7% removal is achieved [69]. Monteoliva-García et al. [57] achieved at laboratory scale eliminations of more than 80% until complete elimination under real plant conditions (pH natural and real wastewater). The results of the different authors seem to indicate that working with natural pH is sufficient, but nevertheless the initial H2O2 concentration in the treatment is very important, since at low concentrations the UV flux applied, even if it is high, is not sufficient to achieve optimum removal yields.
Their elimination has been extensively studied lipid-lowering pharmaceuticals (lipid regulators). An example of this compound is bezafibrate, treating the waters with O3-based AOPs, eliminations of the compound have been reported ranging from 80% [70], 94% (15% TOC) [59] until reaching above 95% [71]. Another compound of the same nature, gemfibrozil also obtained elimination percentages close to 100%, but with better TOC elimination performance (>90%) when the ozone process is combined with H2O2, being the case similar to that of antibiotics [59].
Another type of pharmaceutical widely used is those known as anti-inflammatories. One example is diclofenac, a widely used anti-inflammatory that has been found in many wastewater treatment plant effluents. The advanced oxidation process H2O2/UV is able after 90 min of treatment to degrade this compound with a mineralisation of 39% [72]. Lekkerkerker-Teunissen et al. [61] reported its elimination in more than 80% by applying different intensities with low and medium pressure UV lamps two ranged from 300–700 mJ cm−2 and variable concentrations of H2O2 (0–10 mg L−1). Andreozzi et al. [73] obtained a 100% elimination in 2 min starting from an initial diclofenac concentration of 2.8 mg L−1 and applying a UV intensity (LP-Hg lamp) of 2.51 × 10−6 E s−1 and variable concentrations of H2O2. It is treatment by means of an AOP based on O3 also eliminates it completely [74]. At a concentration of O3 of 5 mg L–1 and H2O2 of 1.8 mg L–1, it showed high efficacy in the elimination of ibuprofen and diclofenac, obtaining 98% mineralisation [75]. The Photo-Fenton system completely oxidised diclofenac after 60 min [76].
Another anti-inflammatory widely used by the population and detectable in practically all studies of wastewater effluents is ibuprofen. Several authors have applied O3/H2O2 treatment and O3-based AOPs obtaining very good results, with eliminations reaching up to 99.4% starting from an initial ibuprofen concentration of 2 μg L−1 and an ozone flow 1–5 mg L–1 with a molar ratio O3:H2O2 (2:1) [75]. Huber et al. [71] reported an elimination of between 40–70% treating the effluent with O3 (0.1–2 mg L−1) starting from an ibuprofen concentration of 0.5 μM. Eliminations greater than 62% were also achieved for water with an initial concentration of 0.13 μg L−1 and an O3 (5–15 mg L−1) [77]. Monteoliva-García et al. [57] reported an ibuprofen elimination (concentrations from 54.6 to 275.0 μg L−1) applying the H2O2/UV treatment ([H2O2] = 25, 50, and 100 mg L−1) from 89.8 to 100%. This seems to indicate that despite the high elimination obtained with both treatments, the most efficient is O3/H2O2, practically eliminating it completely.
In the case of naproxen, a very common anti-inflammatory, it seems to indicate that its behaviour is similar to that of ibuprofen. Rosal et al. [59] compared two AOPs to each other, obtaining with O3-based AOPs an elimination percentage of 93% and approximately 15% TOC, but nevertheless applying O3/H2O2 the removal was very similar obtaining 94% and more than 90% TOC. This same author also carried out the study of another anti-inflammatory, ketoprofen, and the behaviour was the same as in the case of ibuprofen and naproxen. By means of O3-based AOPs, it was able to degrade ketoprofen by 69% and eliminate approximately 15% TOC, applying O3/H2O2 the elimination performance was 70%, very similar but with a toxicity elimination of 90% TOC [59]. Although the removal result is effective, the O3/H2O2 treatment manages to remove a much higher percentage of the total organic carbon, making it a much more efficient treatment.
The degradation of metoprolol (MTP), a β-blocker commonly used for cardiovascular diseases, by UV/H2O2 (LP lamp, 254 nm) process was comparatively evaluated. The results indicate that MTP can rapidly degrade and was marginally affected by changing pH, with a removal rate of 56.7% within 15 min. The MTP degradation treatment satisfied pseudo-first-order kinetic [51]. The treatment of clofibric acid, a lipid regulator metabolite, was completely degraded with no toxic product generation in 2 min by applying two different AOPs (UV/H2O2, O3-based AOPs) [73].
Table S1 in Supplementary Materials summarises the results of different authors in the elimination of pharmaceutical compounds by applying AOPs [78,79,80,81,82].

4.2. Personal Care Products (PCPs)

Personal care products are designated for use on the human body. They comprise a high range of products, such as moisturisers, gels, fragrances, sun creams, and mosquito repellents. Hygienic products alone include at least 8000 preparations available within the European market, with some 140.000 tonnes of shampoo used annually in Germany [83]. Most PCPs are discharged into wastewater and make their way to wastewater treatment plants in their biologically altered or original form [49]. Personal care products may be completely degraded during the wastewater treatment process, or they may remain unchanged or partially converted to metabolites in some cases [84,85]. Therefore, to ensure that TPs are nontoxic and safe, detoxifying the parent compounds becomes a vital and meaningful research focus for promoting the large-scale application of AOPs [86]. Some of the most studied PCPs and their classifications are listed in Table 3.
Most of the studies show that treatment by O3-based AOP is highly efficient in the elimination of PCPs, reaching, in some cases, up to 100%; however, this method has the drawback that it uses a large amount of energy for some minimal O3 productions. Another one of the most used AOPs is UV/H2O2, achieving excellent removal results [52].
UV/H2O2 treatment was applied in the study of the elimination of triclosan, an antimicrobial agent widely used in many household and personal care products. The result was an elimination of greater than 80%, associated with this elimination at pH 8.5 the most favourable [87]. Triclosan also has a good response against the treatment of O3-based AOPs, being almost completely eliminated (94–99.9% removal, Tric 10 mg L–1, pH 7, 25 °C), but identifying several toxic by-products that are detrimental to the resulting water quality [82].
The degradation of methylparaben, ethylparaben, propylparaben, and butylparaben, was carried out by applying an O3-based AOP. An ozone dose of 0.67 g/h resulted in the removal of 99% of parabens in 12 min, as well as the removal of 32% TOC and 61% of chemical oxygen demand [COD] in 3 h of ozonation time for a 500 μM of solution of parabens [88]. Another paraben studied their elimination by UV/H2O2, butylparaben, was able to eliminate more than 95% of its presence after 20 min of treatment (29.6 W/cm2, 0.01 M H2O2, pH 7.0, 22 °C) [89].

4.3. Illicit Drugs

There are studies that analyse the presence of this type of compound. Zuccato et al. [90] tested the presence of drugs of abuse (opioids, amphetamines, cocaine, and cannabis derivatives), some related opioid pharmaceuticals (methadone and codeine), and/or their metabolites in Italian and British surface waters. In 2005, it was reported that kilograms of cocaine residues travel daily down the Po River (Italy). Opioids and non-opioids have for a long-time been used for pain medication because pain is a global health priority and the detection of such contaminants in wastewater serves to estimate the treatment burden of a population using wastewater-based epidemiology (WBE) [91]. In the past decade, wastewater-based epidemiology has been applied in many countries such as China [92] (where influent and effluent wastewater samples were collected from 49 wastewater treatment plants (WWTPs) in 24 major cities that cover all the geographic regions of the country), EE.UU [93], but mainly in Europe [94,95,96,97].
In the case of cocaine, it is largely excreted in the urine as metabolites (its main metabolite is benzoylecgonine) and in a small percentage without changes. If it is consumed with ethanol, the metabolite cocaethylene is also excreted. The group of opioids comprising morphine (metabolic residue like heroin), codeine, and morphine itself is also excreted, but are easily hydrolysed; they are also present in wastewater. Amphetamines, another group of illicit drugs, are excreted as is, not as metabolites. Cannabis is mainly excreted in the urine. Finally, opioid pharmaceuticals are also detected since some drugs of abuse may also have some medical applications, such as codeine for the treatment of heroin addiction, morphine as a strong pain reliever, and dronabinol THC for use in chemotherapy or AIDS patients, and all are present in the wastewater. All these compounds were detected in rivers in Italy by Zuccato et al. [90], demonstrating that they are common contaminants of the aquatic environment in populated areas.
In South Wales in the United Kingdom, a study was conducted for the detection of illicit drugs, including amphetamine, cocaine, and its main metabolite, benzoylecgonine. The illicit drugs studied were found in rivers at low levels of ng L−1. The average daily load of amphetamine was 8 g day−1. For cocaine, it was 1.2 g day−1, and for its main metabolite, benzoylecgonine, it was 39 g day−1. This frequent presence is associated with their high illegal usage and is strongly associated with the discharge of insufficiently treated wastewater effluent [98]. Methamphetamine (0.8 ng/L) and MDMA (0.5 ng/L) were detected in effluent samples from three WWTPs in the USA. The USA estimated an annual release of 0.02 kg of ecstasy in the receiving waters of one of these WWTPs [99]. A study in Northeast Spain in which influent and effluent waters from wastewater treatment plants were monitored, amphetamine was detected (maximum concentration of 210 ng/L), MDA was detected in 10 plants (maximum concentration of 200 ng/L), and methamphetamine was detected in 12 plants (3–90 ng/L). MDEA was also found in 18 effluents, and MDEA was detected in 1 plant (12 ng/L). The main metabolite of cocaine (benzoylecgonine [BE]) was detected in 40 of 42 plants monitored at maximum concentrations of 4.7–7.5 µg/L [100].
The few studies on the application of AOP technology in the elimination of illicit drugs have mainly focused on the elimination of cocaine and its metabolite, BE, as it is one of the most consumed. Spasiano et al. [101] applied the UV/H2O2 process (LP lamp mercury) for different water matrices as a reliable method. This AOP process is effective against this contaminant. BE removal was insensitive in a pH range of 4–8. Russo et al. [102] carried out a study comparing the efficacy of direct photolysis against UV/H2O2 (in a microcapillary film photoreactor) for the degradation of BE. The authors did not obtain differences in BE removal for pH values of 4–8, coinciding with the results of [101]. Indirect photolysis UV254 at different residence times up to 140 s obtained BE removal between 50–70% for different water matrices (pH 6–7.8), while for UV/H2O2 removal rates of up to 95% were achieved for residence times of 2 s, observing various transformation products resulting from the applied treatments, with the potential toxicity of these compounds being unknown.

4.4. Pesticides

These compounds are chemical substances widely used in agriculture to protect crops against insects, fungi, plants considered weeds, and other pests. They have been highly studied products due to their extensive use, and, therefore, their presence and destination in groundwater are well known. Studies have described the transport of pesticides from the soil to depths within different groundwater systems, as well as the occurrence of their metabolites [103,104,105,106].
The spatial occurrence of pesticides in an aquifer is closely related to present or historical land use. A major pathway for contaminant transport to groundwater is the leaching from soils following pesticide application. Fractures, drains, soak ways, and sumps can provide rapid transport routes to groundwater, which bypass some of the natural attenuation potential provided by shallow unsaturated zones and the soil [107]. Pesticides and insecticides have become pervasive environmental problems due to their widespread use [29].
Examples of pesticides are isoproturon, MCPP, atrazine, imazapyr, 2,4-D, glyphosate, simazine, diuron, and picloram. The presence of these compounds in the natural environment is not usual; there are data for at least three decades in which they have been found in the natural environment. Organic pesticides have been used since the discovery of their qualities as insecticides by the Swiss chemist Paul H. Muller in 1939. Due to their great utility, their world consumption has been increasing. Large amounts of pesticides applied in uncontrolled agriculture have caused environmental problems worldwide [51]. Studies have shown that the detection of herbicides is often seasonal. Detectable amounts of some s-triazine herbicides were found in 18 Swiss lakes that were analysed. Atrazine was the major herbicide present, accompanied by smaller amounts of simazine and terbuthylazine, and occasionally (methyl-thio)-s-triazines. Seasonal trends were also observed [108]. Atrazine, simazine, and lindane were found in the Guadiana River from March to November in 1993, observing a seasonal variation [4].
The new vacuum ultraviolet/ultraviolet/chlorine (VUV/UV/chlorine) vacuum process is considered a new AOP, but little is yet known about its kinetics for the degradation of contaminants in water treatment. Using this method, six typical pesticides (dimethoate [DMT], atrazine [ATZ], prometon [PMT], propoxur [PPX], bromacil [BRM], and propachlor [PPC]) were removal with 95% efficiency after 60 s [51]. This seems to indicate that this new VUV/UV/Chlorine process is a very promising AOP. This type of pollutant remains in the soil where it is directly applied. There are AOPs used in situ (ISCO) that are used for soil shadowing. An example of this is the activated persulfate process (SR-AOP), which is a promising technology [109].
Table 4 shows a representative list of PCPs, illicit drugs, and pesticides that can be removed by advanced oxidation processes.

5. Discussion

In this paper, the different types of existing AOPs and their effectiveness in the removal of CECs from wastewater have been discussed. Despite being an efficient technology, they have their own limitations in terms of practical application. Two factors that limit its wide application are hard reaction conditions and high treatment costs [111]. In addition, degradation of organic pollutants leads to formation of by-products, which may be more toxic than the original contaminants. The most studied AOPs are those based on photocatalytic processes, where complete mineralization of the pollutants has generally not been possible [112,113,114]. Therefore, the toxicity of wastewater treated by photocatalysis is a topical issue, it is of high importance to evaluate the toxicity of treated wastewater effluent [115]. It is essential to study their toxicity conventional methods are limited to detecting reaction by-products.
Another aspect to be considered in the application of this advanced technology is its economic cost, which is very high. The trend is to combine biological and physicochemical processes to optimise the overall process cost and for each process to be pollutant selective. To make these processes more energetically more efficient and promote their application, hybrid techniques of AOP and biological treatment processes have been developed [116]. AOPs combined with biological Fenton [Fenton/biological processes] in wastewater treatment improved biodegradability with a TOC reduction of approximately 35% [117]. It was also reported to be used in synthetic wastewater where it reduced the operating costs required for the complete mineralisation of pharmaceuticals [furosemide and ranitidine] over extended periods of time [118]. The electro-Fenton/biological treatment process applied to synthetic water improved its biodegradability for a short treatment of 3 h [119]. For the removal of the pharmaceutical sulfamethoxazole and erythromycin in synthetic water, the combined membrane bioreactor/solar Fenton technology achieved complete removal [120]. The MBR/ozonation process applied to synthetic water achieved complete removal of the pharmaceutical sulfadiazine [121]. The combined MBR and H2O2/UV process applied to urban wastewater allowed high removal efficiencies of over 95% for carbamazepine and up to 98% for ibuprofen, resulting in complete removal for ciprofloxacin [1]. Studies seem to determine that in terms of contaminants degradation, the efficiency of combined processes of AOPs and biological treatments is higher than that of single AOPs [111].
Therefore, there are several limitations to the application of advanced oxidation processes such as the economic cost of their implementation, the intermediate compounds are in some cases more toxic than the starting compounds, as well as the lack of complete mineralisation in some cases. This remains a challenge to be considered for future research.

6. Conclusions

This paper provides a summary of the studies that have been carried out by different authors for the treatment of emerging contamination compounds that have been detected in natural environments by means of AOPs. Although these products were present in the aquatic environment for decades, due to their levels or detection in recent studies are now recognised as potentially dangerous for natural water reserves. These contaminants are not completely removed in conventional treatment systems.
The combination of AOPs as tertiary treatment in wastewater treatment plants can provide biodegradation mechanisms for contaminating compounds that are not eliminated in previous purification processes, thus avoiding their integration into the environment.
Advanced oxidation processes are very promising alternatives for the removal of emerging contamination compounds. With them, very high removal percentages can be achieved. In the case of compounds such as amoxicillin and sulfamethoxazole, their removal percentage is around 90%. In the case of compounds such as diclofenac, ibuprofen, and ciprofloxacin, even their complete elimination from water can be achieved. In the case of the contaminant sulfamethoxazole, removal by AOPs produces removals of more than 90%, reaching complete removal for the UV/H2O2 process. Other pharmaceuticals that are removed with very high or complete removal rates are oxytetracycline and doxycycline. Very promising removal rates of 69% were achieved for ketoprofen, as well as removal rates of 53–94% for metronidazole. In the case of lipid regulators bezafibrate and clofibric acid, clearance rates of 50–90% were achieved.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/w13152094/s1, Table S1: Removal percentage of pharmaceutical compounds using AOPs.

Author Contributions

Conceptualization, L.A.B., J.M.P., M.d.M.M.M. and J.M.P.C.; methodology J.M.P., M.d.M.M.M. and J.M.P.C.; formal analysis L.A.B.; investigation L.A.B., J.M.P., M.d.M.M.M. and J.M.P.C.; resources L.A.B.; data curation, L.A.B., J.M.P., M.d.M.M.M. and J.M.P.C.; writing—original draft preparation, L.A.B.; writing—review and editing, L.A.B., J.M.P.C., M.d.M.M.M. and J.M.P.C.; supervision, J.M.P.C., M.d.M.M.M. and J.M.P.C.; project administration, J.M.P.C.; funding acquisition, J.M.P.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This study was financed by the Spanish Ministry of Science, Innovation and University Project RTI2018-101270-B-I00.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

AOPs, advanced oxidation processes; BE, benzoylecgonine; BOD5, biochemical oxygen demand; CAS, activated sludge system; CECs, compounds of emerging concern; CIP, ciprofloxacin; COD, chemical oxygen demand; DOC, dissolved organic carbon; Fe2+, ferrous ion; GAC, granular activated carbon; Hv, photolysis; HP, high pressure; HPLC-MS, high pressure liquid chromatography; H2O2, hydrogen peroxide; ISCO, in situ chemical oxidation; LP, low pressure; MCPP, mecoprop (methylchlorophenoxypropionic acid); MDA, tenamfetamine; MDEA, methylenedioxyethylamphetamine; MDMA, 3,4-metilendioxi-methamphetamine; MTP, metoprolol; O3, ozonation; PCP, personal care product; PH, pharmaceutical; SMX, sulfamethoxazole; SR-AOP, activated persulfate oxidation processes; TiO2, titanium dioxide; TOC, total organic carbon; Tric, triclosan; t-BuOH, tert-butyl alcohol, also called tert-butanol; US, ultrasound energy; UV, ultraviolet radiation; VUV, vacuum ultraviolet; WWTP, wastewater treatment plant; 2,4-D, 2,4-dichlorophenoxyacetic acid.

References

  1. Monteoliva-García, A.; Martín-Pascual, J.; Muñío, M.M.; Poyatos, J.M. Removal of a pharmaceutical mix from urban wastewater coupling membrane bioreactor with advanced oxidation processes. J. Environ. Eng. 2019, 145, 04019055. [Google Scholar] [CrossRef]
  2. Lee, S.-H.; Kim, K.-H.; Lee, M.; Lee, B.-D. Detection status and removal characteristics of pharmaceuticals in wastewater treatment effluent. J. Water Process. Eng. 2019, 31, 100828. [Google Scholar] [CrossRef]
  3. Brillas, E. A review on the photoelectro-Fenton process as efficient electrochemical advanced oxidation for wastewater re-mediation. Treatment with UV light, sunlight, and coupling with conventional and other photo-assisted advanced technologies. Chemosphere 2020, 250, 126198. [Google Scholar] [CrossRef] [PubMed]
  4. Miniero, R.; Abate, V.; Brambilla, G.; Davoli, E.; De Felip, E.; De Filippis, S.P.; Dellatte, E.; De Luca, S.; Fanelli, R.; Fattore, E.; et al. Persistent toxic substances in Mediterranean aquatic species. Sci. Total Environ. 2014, 494–495, 18–27. [Google Scholar] [CrossRef] [PubMed]
  5. Teodosiu, C.; Gilca, A.-F.; Barjoveanu, G.; Fiore, S. Emerging pollutants removal through advanced drinking water treatment: A review on processes and environmental performances assessment. J. Clean. Prod. 2018, 197, 1210–1221. [Google Scholar] [CrossRef]
  6. Bui, X.T.; Vo, T.P.T.; Ngo, H.H.; Guo, W.S.; Nguyen, T.T. Multicriteria assessment of advanced treatment technologies for mi-cropollutants removal at large-scale applications. Sci. Total Environ. 2016, 563–564, 1050–1067. [Google Scholar] [CrossRef] [PubMed]
  7. Rozas, O.; Vidal, C.; Baeza, C.; Jardim, W.F.; Rossner, A.; Mansilla, H.D. Organic micropollutants (OMPs) in natural waters: Oxidation by UV/H2O2 treatment and toxicity assessment. Water Res. 2016, 98, 109–118. [Google Scholar] [CrossRef]
  8. Watkinson, A.; Murby, E.; Kolpin, D.; Costanzo, S. The occurrence of antibiotics in an urban watershed: From wastewater to drinking water. Sci. Total Environ. 2009, 407, 2711–2723. [Google Scholar] [CrossRef]
  9. Moldovan, Z. Occurrences of pharmaceutical and personal care products as micropollutants in rivers from Romania. Chemosphere 2006, 64, 1808–1817. [Google Scholar] [CrossRef]
  10. Silva, B.F.D.; Jelic, A.; López-Serna, R.; Mozeto, A.A.; Petrovic, M.; Barceló, D. Occurrence and distribution of pharmaceuticals in surface water, suspended solids and sediments of the Ebro river basin, Spain. Chemosphere 2011, 85, 1331–1339. [Google Scholar] [CrossRef]
  11. Bartelt-Hunt, S.L.; Snow, D.D.; Damon, T.; Shockley, J.; Hoagland, K. The occurrence of illicit and therapeutic pharmaceuticals in wastewater effluent and surface waters in Nebraska. Environ. Pollut. 2009, 157, 786–791. [Google Scholar] [CrossRef] [Green Version]
  12. Xu, W.-H.; Zhang, G.; Zou, S.-C.; Li, X.; Liu, Y.-C. Determination of selected antibiotics in the Victoria Harbour and the Pearl River, South China using high-performance liquid chromatography-electrospray ionization tandem mass spectrometry. Environ. Pollut. 2007, 145, 672–679. [Google Scholar] [CrossRef] [Green Version]
  13. Li, W.; Shi, Y.; Gao, L.; Liu, J.; Cai, Y. Occurrence of antibiotics in water, sediments, aquatic plants, and animals from Baiyangdian Lake in North China. Chemosphere 2012, 89, 1307–1315. [Google Scholar] [CrossRef]
  14. Matamoros, V.; Arias, C.A.; Nguyen, L.X.; Salvado, V.; Brix, H. Occurrence and behavior of emerging contaminants in surface water and a restored wetland. Chemosphere 2012, 88, 1083–1089. [Google Scholar] [CrossRef] [PubMed]
  15. Vilimanovic, D.; Andaluri, G.; Hannah, R.; Suri, R.; MacGillivray, A.R. Occurrence and aquatic toxicity of contaminants of emerging concern (CECs) in tributaries of an urbanized section of the Delaware River Watershed. AIMS Environ. Sci. 2020, 7, 302–319. [Google Scholar] [CrossRef]
  16. Boczkaj, G.; Fernandes, A. Wastewater treatment by means of advanced oxidation processes at basic pH conditions: A review. Chem. Eng. J. 2017, 320, 608–633. [Google Scholar] [CrossRef]
  17. Poyatos, J.M.; Muñio, M.M.; Almecija, M.C.; Torres, J.C.; Hontoria, E.; Osorio, F. Advanced oxidation processes for wastewater treatment: State of the art. Water Air Soil Pollut. 2010, 205, 187–204. [Google Scholar] [CrossRef]
  18. Rosenfeldt, E.J.; Chen, P.-J.; Kullman, S.; Linden, K.G. Destruction of estrogenic activity in water using UV advanced oxidation. Sci. Total Environ. 2007, 377, 105–113. [Google Scholar] [CrossRef] [PubMed]
  19. Baghirzade, B.S.; Yetis, U.; Dilek, F.B. Imidacloprid elimination by O3 and O3/UV: Kinetics study, matrix effect, and mechanism insight. Environ. Sci. Pollut. Res. 2021, 28, 24535–24551. [Google Scholar] [CrossRef] [PubMed]
  20. Maniakova, G.; Salmerón, I.; Polo-López, M.I.; Oller, I.; Rizzo, L.; Malato, S. Simultaneous removal of contaminants of emerging concern and pathogens from urban wastewater by homogeneous solar driven advanced oxidation processes. Sci. Total Environ. 2021, 766, 144320. [Google Scholar] [CrossRef]
  21. Ferro, G.; Guarino, F.; Castiglione, S.; Rizzo, L. Antibiotic resistance spread potential in urban wastewater effluents disinfected by UV/H2O2 process. Sci. Total Environ. 2016, 560-561, 29–35. [Google Scholar] [CrossRef] [PubMed]
  22. Climent, L.F.; Olmos, R.G.; Anfruns, A.; Aymerich, I.; Corominas, L.; Barceló, D.; Rodriguez-Mozaz, S. Elimination study of the chemotherapy drug tamoxifen by different advanced oxidation processes: Transformation products and toxicity assessment. Chemosphere 2017, 168, 284–292. [Google Scholar] [CrossRef]
  23. Bianco, A.; Polo−López, M.; Fernandez-Ibañez, P.; Brigante, M.; Mailhot, G. Disinfection of water inoculated with Enterococcus faecalis using solar/Fe(III)EDDS-H2O2 or S2O82—process. Water Res. 2017, 118, 249–260. [Google Scholar] [CrossRef]
  24. He, Z.; Song, S.; Ying, H.; Xu, L.; Chen, J. p-Aminophenol degradation by ozonation combined with sonolysis: Operating conditions influence and mechanism. Ultrason. Sonochem. 2007, 14, 568–574. [Google Scholar] [CrossRef]
  25. Suri, R.P.S.; Andaluri, G. Oxidative sonication of estrogen hormones in water and municipal wastewater. Res. J. Environ. Sci. 2017, 11, 71–81. [Google Scholar] [CrossRef]
  26. Shemer, H.; Narkis, N. Trihalomethanes aqueous solutions sono-oxidation. Water Res. 2005, 39, 2704–2710. [Google Scholar] [CrossRef] [PubMed]
  27. Amadelli, R.; De Battisti, A.; Girenko, D.; Kovalyov, S.; Velichenko, A. Electrochemical oxidation of trans-3,4-dihydroxycinnamic acid at PbO2 electrodes: Direct electrolysis and ozone mediated reactions compared. Electrochim. Acta 2000, 46, 341–347. [Google Scholar] [CrossRef]
  28. Soares, I.C.D.C.; Da Silva, Á.R.L.; Santos, E.C.M.D.M.; dos Santos, E.V.; Da Silva, D.R.; Martínez-Huitle, C.A. Understanding the electrochemical oxidation of dyes on platinum and boron-doped diamond electrode surfaces: Experimental and computational study. J. Solid State Electrochem. 2020, 24, 3245–3256. [Google Scholar] [CrossRef]
  29. Damodhar, G.; Ghosh, P. Removal of organic compounds found in the wastewater through electrochemical advanced oxidation processes: A review. Russ. J. Electrochem. 2019, 55, 591–620. [Google Scholar] [CrossRef]
  30. Florenza, X.; Solano AM, S.; Centellas, F.; Martínez-Huitle, C.A.; Brillas, E.; Garcia-Segura, S. Degradation of the azo dye Acid Red 1 by anodic oxidation and indirect electrochemical processes based on Fenton’s reaction chemistry. Relationship between decolorization, mineralization and products. Electrochim. Acta. 2014, 142, 276–288. [Google Scholar] [CrossRef]
  31. Peralta-Hernández, J.; Meas-Vong, Y.; Rodríguez, F.J.; Chapman, T.W.; Maldonado, M.I.; Godínez, L.A. Comparison of hydrogen peroxide-based processes for treating dye-containing wastewater: Decolorization and destruction of Orange II azo dye in dilute solution. Dye Pigment. 2008, 76, 656–662. [Google Scholar] [CrossRef]
  32. Raschitor, A.; Llanos, J.; Rodrigo, M.A.; Cañizares, P. Combined electrochemical processes for the efficient degradation of non-polar organochlorine pesticides. J. Environ. Manag. 2019, 248, 109289. [Google Scholar] [CrossRef] [PubMed]
  33. Kourdali, S.; Badis, A.; Boucherit, A. Degradation of direct yellow 9 by electro-Fenton: Process study and optimization and, monitoring of treated water toxicity using catalase. Ecotoxicol. Environ. Saf. 2014, 110, 110–120. [Google Scholar] [CrossRef] [PubMed]
  34. Alaton, I.A.; Balcioglu, I.A.; Bahnemann, D.W. Advanced oxidation of a reactive dyebath effluent: Comparison of O3, H2O2/UV-C and TiO2/UV-A processes. Water Res. 2002, 36, 1143–1154. [Google Scholar] [CrossRef]
  35. Dogruel, S.; Atesci, Z.C.; Aydin, E.; Pehlivanoglu-Mantas, E. Ozonation in advanced treatment of secondary municipal wastewater effluents for the removal of micropollutants. Environ. Sci. Pollut. Res. 2020, 27, 45460–45475. [Google Scholar] [CrossRef]
  36. Al Momani, F. Degradation of cyanobacteria anatoxin-a by advanced oxidation processes. Sep. Purif. Technol. 2007, 57, 85–93. [Google Scholar] [CrossRef]
  37. Lou, W.; Kane, A.; Wolbert, D.; Rtimi, S.; Assadi, A.A. Study of a photocatalytic process for removal of antibiotics from wastewater in a falling film photoreactor: Scavenger study and process intensification feasibility. Chem. Eng. Process. Process. Intensif. 2017, 122, 213–221. [Google Scholar] [CrossRef]
  38. Hong, P.; Li, Y.; He, J.; Saeed, A.; Zhang, K.; Wang, C.; Kong, L.; Liu, J. Rapid degradation of aqueous doxycycline by surface CoFe2O4/H2O2 system: Behaviors, mechanisms, pathways and DFT calculation. Appl. Surf. Sci. 2020, 526, 146557. [Google Scholar] [CrossRef]
  39. Aguilar-Melo, C.; Rodríguez, J.L.; Chairez, I.; Tiznado, H.; Poznyak, T. Naphthalene degradation by catalytic ozonation based on nickel oxide: Study of the ethanol as cosolvent. Environ. Sci. Pollut. Res. 2016, 24, 25550–25560. [Google Scholar] [CrossRef]
  40. Tanaka, K.; Abe, K.; Hisanaga, T. Photocatalytic water treatment on immobilized TiO2 combined with ozonation. J. Photochem. Photobiol. A Chem. 1996, 101, 85–87. [Google Scholar] [CrossRef]
  41. Aguinaco, A.; Beltran, F.J.; García-Araya, J.F.; Oropesa, A.L. Photocatalytic ozonation to remove the pharmaceutical diclofenac from water: Influence of variables. Chem. Eng. J. 2012, 189-190, 275–282. [Google Scholar] [CrossRef]
  42. García, J.; Oliveira, J.; Silva, A.; Oliveira, C.; Nozaki, J.; De Souza, N. Comparative study of the degradation of real textile effluents by photocatalytic reactions involving UV/TiO2/H2O2 and UV/Fe2+/H2O2 systems. J. Hazard. Mater. 2007, 147, 105–110. [Google Scholar] [CrossRef]
  43. Kümmerer, K. The presence of pharmaceuticals in the environment due to human use—Present knowledge and future challenges. J. Environ. Manag. 2009, 90, 2354–2366. [Google Scholar] [CrossRef] [PubMed]
  44. Naidu, R.; Jit, J.; Kennedy, B.; Arias, V. Emerging contaminant uncertainties and policy: The chicken or the egg conundrum. Chemosphere 2016, 154, 385–390. [Google Scholar] [CrossRef] [PubMed]
  45. Bu, Q.; Wang, B.; Huang, J.; Deng, S.; Yu, G. Pharmaceuticals and personal care products in the aquatic environment in China: A review. J. Hazard. Mater. 2013, 262, 189–211. [Google Scholar] [CrossRef]
  46. Petrović, M.; Gonzalez, S.; Barceló, D. Analysis and removal of emerging contaminants in wastewater and drinking water. Trends Anal. Chemistry 2003, 22, 685–696. [Google Scholar] [CrossRef] [Green Version]
  47. Carballa, M.; Omil, F.; Lema, J.M.; Llompart, M.; García-Jares, C.; Rodríguez, I.; Gomez, M.; Ternes, T. Behavior of pharmaceuticals, cosmetics and hormones in a sewage treatment plant. Water Res. 2004, 38, 2918–2926. [Google Scholar] [CrossRef]
  48. Rivera-Utrilla, J.; Sánchez-Polo, M.; Ferro-García, M.Á.; Prados-Joya, G.; Ocampo-Pérez, R. Pharmaceuticals as emerging contaminants and their removal from water. A review. Chemosphere 2013, 93, 1268–1287. [Google Scholar] [CrossRef]
  49. Jiang, J.-Q.; Zhou, Z.; Sharma, V. Occurrence, transportation, monitoring and treatment of emerging micro-pollutants in waste water—A review from global views. Microchem. J. 2013, 110, 292–300. [Google Scholar] [CrossRef]
  50. Michael, S.G.; Michael-Kordatou, I.; Nahim-Granados, S.; Polo-López, M.I.; Rocha, J.; Martínez-Piernas, A.B.; Fernández-Ibáñez, P.; Agüera, A.; Manaia, C.M.; Fatta-Kassinos, D. Investigating the impact of UV-C/H2O2 and sunlight/H2O2 on the removal of antibiotics, antibiotic resistance determinants and toxicity present in urban wastewater. Chem. Eng. J. 2020, 388, 124383. [Google Scholar] [CrossRef]
  51. Gao, Y.-Q.; Zhang, J.; Li, C.; Tian, F.-X.; Gao, N.-Y. Comparative evaluation of metoprolol degradation by UV/chlorine and UV/H2O2 processes. Chemosphere 2020, 243, 125325. [Google Scholar] [CrossRef]
  52. Tayo, L.L.; Caparanga, A.R.; Doma, B.T.; Liao, C.-H. A Review on the removal of Pharmaceutical and Personal Care Products (PPCPs) using advanced oxidation processes. J. Adv. Oxid. Technol. 2018, 21, 196–214. [Google Scholar] [CrossRef]
  53. Yan, C.; Yang, Y.; Zhou, J.; Liu, M.; Nie, M.; Shi, H.; Gu, L. Antibiotics in the surface water of the Yangtze estuary: Occurrence, distribution and risk assessment. Environ. Pollut. 2013, 175, 22–29. [Google Scholar] [CrossRef]
  54. Oviedo Cordoba, H.; Cortina Navarro, C.E.; Osorio Coronel, J.A.; Romero Torres, S.M. Realidades de la práctica de la auto-medicación en estudiantes de la Universidad del Magdalena. Enfermería Global. 2021, 20, 531–556. [Google Scholar] [CrossRef]
  55. Jardim, W.F.; Montagner, C.C.; Pescara, I.C.; Umbuzeiro, G.; Bergamasco, A.M.D.D.; Eldridge, M.L.; Sodré, F.F. An integrated approach to evaluate emerging contaminants in drinking water. Sep. Purif. Technol. 2012, 84, 3–8. [Google Scholar] [CrossRef]
  56. Bletsou, A.A.; Jeon, J.; Hollender, J.; Archontaki, E.; Thomaidis, N.S. Targeted and non-targeted liquid chromatography-mass spectrometric workflows for identification of transformation products of emerging pollutants in the aquatic environment. TrAC Trends Anal. Chem. 2015, 66, 32–44. [Google Scholar] [CrossRef] [Green Version]
  57. Monteoliva-García, A.; Martín-Pascual, J.; Muñío, M.M.; Poyatos, J.M. Removal of carbamazepine, ciprofloxacin and ibuprofen in real urban wastewater by using light-driven advanced oxidation processes. Int. J. Environ. Sci. Technol. 2019, 16, 6005–6018. [Google Scholar] [CrossRef]
  58. Yuan, F.; Hu, C.; Hu, X.; Wei, D.; Chen, Y.; Qu, J. Photodegradation and toxicity changes of antibiotics in UV and UV/H2O2 process. J. Hazard. Mater. 2011, 185, 1256–1263. [Google Scholar] [CrossRef]
  59. Rosal, R.; Rodríguez, A.; Melon, J.A.P.; Mezcua, M.; Hernando, D.; Leton, P.; Calvo, E.G.; Agüera, A.; Fernández-Alba, A. Removal of pharmaceuticals and kinetics of mineralization by O3/H2O2 in a biotreated municipal wastewater. Water Res. 2008, 42, 3719–3728. [Google Scholar] [CrossRef]
  60. De Witte, B.; Dewulf, J.; Demeestere, K.; Van Langenhove, H. Ozonation and advanced oxidation by the peroxone process of ciprofloxacin in water. J. Hazard. Mater. 2009, 161, 701–708. [Google Scholar] [CrossRef] [PubMed]
  61. Lekkerkerker-Teunissen, K.; Benotti, M.J.; Snyder, S.; van Dijk, H.C. Transformation of atrazine, carbamazepine, diclofenac and sulfamethoxazole by low and medium pressure UV and UV/H2O2 treatment. Sep. Purif. Technol. 2012, 96, 33–43. [Google Scholar] [CrossRef]
  62. Yang, Y.; Lu, X.; Jiang, J.; Ma, J.; Liu, G.; Cao, Y.; Liu, W.; Li, J.; Pang, S.; Kong, X.; et al. Degradation of sulfamethoxazole by UV, UV/H2O2 and UV/persulfate (PDS): Formation of oxidation products and effect of bicarbonate. Water Res. 2017, 118, 196–207. [Google Scholar] [CrossRef] [PubMed]
  63. Kim, T.-H.; Kim, S.D.; Kim, H.Y.; Lim, S.J.; Lee, M.; Yu, S. Degradation and toxicity assessment of sulfamethoxazole and chlortetracycline using electron beam, ozone and UV. J. Hazard. Mater. 2012, 227-228, 237–242. [Google Scholar] [CrossRef] [PubMed]
  64. Jung, Y.J.; Kim, W.G.; Yoon, Y.; Kang, J.-W.; Hong, Y.M.; Kim, H.W. Removal of amoxicillin by UV and UV/H2O2 processes. Sci. Total Environ. 2012, 420, 160–167. [Google Scholar] [CrossRef]
  65. Elmolla, E.S.; Chaudhuri, M. Photocatalytic degradation of amoxicillin, ampicillin and cloxacillin antibiotics in aqueous solution using UV/TiO2 and UV/H2O2/TiO2 photocatalysis. Desalination 2010, 252, 46–52. [Google Scholar] [CrossRef]
  66. Torre, J.A.C.; Lehmann Ah González, C.; Carmona Em Muñoz, A.H. Presencia de Fármacos en Aguas Residuales y Eficacia de los Procesos Convencionales en su Eliminación; Cátedra de Ingeriería Sanidad y Ambiente de la ETS Ingenieros de Caminos, Canales y Puertos, Universidad Politécnica de Madrid: Madrid, Spain, 2006. [Google Scholar]
  67. Shemer, H.; Kunukcu, Y.K.; Linden, K.G. Degradation of the pharmaceutical Metronidazole via UV, Fenton and photo-Fenton processes. Chemosphere 2006, 63, 269–276. [Google Scholar] [CrossRef]
  68. Andreozzi, R.; Canterino, M.; Giudice, R.L.; Marotta, R.; Pinto, G.; Pollio, A. Lincomycin solar photodegradation, algal toxicity and removal from wastewaters by means of ozonation. Water Res. 2006, 40, 630–638. [Google Scholar] [CrossRef] [PubMed]
  69. Alharbi, S.K.; Kang, J.; Nghiem, L.; van de Merwe, J.; Leusch, F.; 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] [Green Version]
  70. 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]
  71. 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]
  72. Vogna, D.; Marotta, R.; Napolitano, A.; Andreozzi, R.; D’Ischia, M. Advanced oxidation of the pharmaceutical drug diclofenac with UV/H2O2 and ozone. Water Res. 2004, 38, 414–422. [Google Scholar] [CrossRef] [PubMed]
  73. Andreozzi, R.; Campanella, L.; Fraysse, B.; Garric, J.; Gonnella, A.; Giudice, R.L.; Marotta, R.; Pinto, G.; Pollio, A. Effects of advanced oxidation processes (AOPs) on the toxicity of a mixture of pharmaceuticals. Water Sci. Technol. 2004, 50, 23–28. [Google Scholar] [CrossRef] [PubMed]
  74. Sein, M.M.; Zedda, M.; Tuerk, J.; Schmidt, T.C.; Golloch, A.; Von Sonntag, C. Oxidation of diclofenac with ozone in aqueous solution. Environ. Sci. Technol. 2008, 42, 6656–6662. [Google Scholar] [CrossRef] [PubMed]
  75. Zwiener, C.; Frimmel, F.H. Oxidative treatment of pharmaceuticals in water. Water Res. 2000, 34, 1881–1885. [Google Scholar] [CrossRef]
  76. Ravina, M.; Campanella, L.; Kiwi, J. Accelerated mineralization of the drug Diclofenac via Fenton reactions in a concentric photo-reactor. Water Res. 2002, 36, 3553–3560. [Google Scholar] [CrossRef]
  77. 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]
  78. Witte, B.D.; Langenhove, H.V.; Hemelsoet, K.; Demeestere, K.; Wispelaere, P.D.; Van Speybroeck, V.; Dewulf, J. Levofloxacin ozonation in water: Rate determining process parameters and reaction pathway elucidation. Chemosphere 2009, 76, 683–689. [Google Scholar] [CrossRef]
  79. Lima, V.B.; Goulart, L.A.; Rocha, R.S.; Steter, J.R.; Lanza, M.R. Degradation of antibiotic ciprofloxacin by different AOP systems using electrochemically generated hydrogen peroxide. Chemosphere 2020, 247, 125807. [Google Scholar] [CrossRef] [PubMed]
  80. Lange, F.; Cornelissen, S.; Kubac, D.; Sein, M.M.; von Sonntag, J.; Hannich, C.B.; Golloch, A.; Heipieper, H.J.; Möder, M.; von Sonntag, C. Degradation of macrolide antibiotics by ozone: A mechanistic case study with clarithromycin. Chemosphere 2006, 65, 17–23. [Google Scholar] [CrossRef]
  81. Garoma, T.; Umamaheshwar, S.K.; Mumper, A. Removal of sulfadiazine, sulfamethizole, sulfamethoxazole, and sulfathiazole from aqueous solution by ozonation. Chemosphere 2010, 79, 814–820. [Google Scholar] [CrossRef]
  82. Chen, X.; Richard, J.; Liu, Y.; Dopp, E.; Tuerk, J.; Bester, K. Ozonation products of triclosan in advanced wastewater treatment. Water Res. 2012, 46, 2247–2256. [Google Scholar] [CrossRef]
  83. Ellis, J. Pharmaceutical and personal care products (PPCPs) in urban receiving waters. Environ. Pollut. 2006, 144, 184–189. [Google Scholar] [CrossRef]
  84. Ebele, A.J.; Abdallah, M.A.-E.; Harrad, S. Pharmaceuticals and personal care products (PPCPs) in the freshwater aquatic environment. Emerg. Contam. 2017, 3, 1–16. [Google Scholar] [CrossRef]
  85. Xia, K.; Bhandari, A.; Das, K.; Pillar, G. Occurrence and fate of pharmaceuticals and personal care products (PPCPs) in biosolids. J. Environ. Qual. 2005, 34, 91–104. [Google Scholar] [CrossRef] [PubMed]
  86. Lu, G.; Hu, J. Effect of alpha-hydroxy acids on transformation products formation and degradation mechanisms of carbamaz-epine by UV/H2O2 process. Sci. Total. Environ. 2019, 689, 70–78. [Google Scholar] [CrossRef]
  87. Huang, Y.; Kong, M.; Coffin, S.; Cochran, K.H.; Westerman, D.C.; Schlenk, D.; Richardson, S.D.; Lei, L.; Dionysiou, D.D. Degradation of contaminants of emerging concern by UV/H2O2 for water reuse: Kinetics, mechanisms, and cytotoxicity analysis. Water Res. 2020, 174, 115587. [Google Scholar] [CrossRef]
  88. Tay, K.S.; Rahman, N.A.; Bin Abas, M.R. Kinetic studies of the degradation of parabens in aqueous solution by ozone oxidation. Environ. Chem. Lett. 2010, 8, 331–337. [Google Scholar] [CrossRef]
  89. De la Cruz, N.; Giménez, J.; Esplugas, S.; Grandjean, D.; de Alencastro, L.F.; Pulgarín, C. Degradation of 32 emergent contaminants by UV and neutral photo-Fenton in domestic wastewater effluent previously treated by activated sludge. Water Res. 2012, 46, 1947–1957. [Google Scholar] [CrossRef] [PubMed]
  90. Zuccato, E.; Castiglioni, S.; Bagnati, R.; Chiabrando, C.; Grassi, P.; Fanelli, R. Illicit drugs, a novel group of environmental con-taminants. Water Res. 2008, 42, 961–968. [Google Scholar] [CrossRef]
  91. Ahmed, F.; Tscharke, B.; O’Brien, J.W.; Cabot, P.J.; Hall, W.D.; Mueller, J.F.; Thomas, K.V. Can wastewater analysis be used as a tool to assess the burden of pain treatment within a population? Environ. Res. 2020, 188, 109769. [Google Scholar] [CrossRef] [PubMed]
  92. Du, P.; Zhou, Z.; Bai, Y.; Xu, Z.; Gao, T.; Fu, X.; Li, X. Estimating heroin abuse in major Chinese cities through wastewater-based epidemiology. Sci. Total. Environ. 2017, 605–606, 158–165. [Google Scholar] [CrossRef]
  93. Subedi, B.; Kannan, K. Mass loading and removal of select illicit drugs in two wastewater treatment plants in New York State and estimation of illicit drug usage in communities through wastewater analysis. Environ. Sci. Technol. 2014, 48, 6661–6670. [Google Scholar] [CrossRef]
  94. Been, F.; Benaglia, L.; Lucia, S.; Gervasoni, J.-P.; Esseiva, P.; Delémont, O. Data triangulation in the context of opioids monitoring via wastewater analyses. Drug Alcohol Depend. 2015, 151, 203–210. [Google Scholar] [CrossRef] [PubMed]
  95. McCall, A.-K.; Bade, R.; Kinyua, J.; Lai, F.Y.; Thai, P.; Covaci, A.; Bijlsma, L.; van Nuijs, A.; Ort, C. Critical review on the stability of illicit drugs in sewers and wastewater samples. Water Res. 2016, 88, 933–947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Mercan, S.; Kuloglu, M.; Tekin, T.; Turkmen, Z.; Dogru, A.O.; Safran, A.N.; Acikkol, M.; Asicioglu, F. Wastewater-based monitoring of illicit drug consumption in Istanbul: Preliminary results from two districts. Sci Total Environ. 2019, 656, 231–238. [Google Scholar] [CrossRef] [PubMed]
  97. Damien, D.A.; Thomas, N.; Hélène, P.; Sara, K.; Yves, L. First evaluation of illicit and licit drug consumption based on wastewater analysis in Fort de France urban area (Martinique, Caribbean), a transit area for drug smuggling. Sci. Total Environ. 2014, 490, 970–978. [Google Scholar] [CrossRef] [PubMed]
  98. Kasprzyk-Hordern, B.; Dinsdale, R.; 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]
  99. Jones-Lepp, T.L.; Alvarez, D.A.; Petty, J.D.; Huckins, J.N. Polar organic chemical integrative sampling and liquid chromatography? Electrospray/ion-trap mass spectrometry for assessing selected prescription and illicit drugs in treated sewage effluents. Arch. Environ. Contam. Toxicol. 2004, 47, 427–439. [Google Scholar] [CrossRef]
  100. Huerta-Fontela, M.; Galceran, M.T.; Martin-Alonso, J.; Ventura, F. Occurrence of psychoactive stimulatory drugs in wastewaters in north-eastern Spain. Sci. Total Environ. 2008, 397, 31–40. [Google Scholar] [CrossRef]
  101. Spasiano, D.; Russo, D.; Vaccaro, M.; Siciliano, A.; Marotta, R.; Guida, M.; Reis, N.; Puma, G.L.; Andreozzi, R. Removal of benzoylecgonine from water matrices through UV254/H2O2 process: Reaction kinetic modeling, ecotoxicity and genotoxicity assessment. J. Hazard. Mater. 2016, 318, 515–525. [Google Scholar] [CrossRef] [Green Version]
  102. Russo, D.; Spasiano, D.; Vaccaro, M.; Cochran, K.H.; Richardson, S.D.; Andreozzi, R.; Puma, G.L.; Reis, N.; Marotta, R. Investigation on the removal of the major cocaine metabolite (benzoylecgonine) in water matrices by UV254 /H2O2 process by using a flow microcapillary film array photoreactor as an efficient experimental tool. Water Res. 2016, 89, 375–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Gooddy, D.C.; Bloomfield, J.P.; Chilton, P.J.; Johnson, A.C.; Williams, R.J. Assessing herbicide concentrations in the saturated and unsaturated zone of a Chalk aquifer in southern England. Ground Water 2001, 39, 262–271. [Google Scholar] [CrossRef] [PubMed]
  104. Gooddy, D.C.; Chilton, P.J.; Harrison, I. A field study to assess the degradation and transport of diuron and its metabolites in a calcareous soil. Sci. Total Environ. 2002, 297, 67–83. [Google Scholar] [CrossRef] [Green Version]
  105. Kolpin, D.W.; Thurman, E.M.; Linhart, S.M. Occurrence of cyanazine compounds in groundwater: Degradates More Prevalent than the parent compound. Environ. Sci. Technol. 2001, 35, 1217–1222. [Google Scholar] [CrossRef] [Green Version]
  106. Panno, S.; Kelly, W. Nitrate and herbicide loading in two groundwater basins of Illinois’ sinkhole plain. J. Hydrol. 2004, 290, 229–242. [Google Scholar] [CrossRef]
  107. Lapworth, D.; Gooddy, D.; Stuart, M.; Chilton, P.; Cachandt, G.; Knapp, M.; Bishop, S. Pesticides in groundwater: Some observations on temporal and spatial trends. Water Environ. J. 2006, 20, 55–64. [Google Scholar] [CrossRef] [Green Version]
  108. Hans-Rudolf, B. Atrazine and others-triazine herbicides in lakes and in rain in Switzerland. Environ. Sci. Technol. 1990, 24, 1049–1058. [Google Scholar] [CrossRef]
  109. Chen, L.; Hu, X.; Yang, Y.; Jiang, C.; Bian, C.; Liu, C.; Zhang, M.; Cai, T. Degradation of atrazine and structurally related s-triazine herbicides in soils by ferrous-activated persulfate: Kinetics, mechanisms and soil-types effects. Chem. Eng. J. 2018, 351, 523–531. [Google Scholar] [CrossRef]
  110. Serna-Galvis, E.A.; Botero-Coy, A.M.; Martínez-Pachón, D.; Moncayo-Lasso, A.; Ibáñez, M.; Hernández, F.; Torres-Palma, R.A. Degradation of seventeen contaminants of emerging concern in municipal wastewater effluents by sonochemical advanced oxidation pro-cesses. Water Res. 2019, 154, 349–360. [Google Scholar] [CrossRef]
  111. Ma, D.; Yi, H.; Lai, C.; Liu, X.; Huo, X.; An, Z.; Li, L.; Fu, Y.; Li, B.; Zhang, M.; et al. Critical review of advanced oxidation processes in organic wastewater treatment. Chemosphere 2021, 275, 130104. [Google Scholar] [CrossRef]
  112. Berberidou, C.; Kitsiou, V.; Lambropoulou, D.; Antoniadis, A.; Ntonou, E.; Zalidis, G.C.; Poulios, I. Evaluation of an alternative method for wastewater treatment containing pesticides using solar photocatalytic oxidation and constructed wetlands. J. Environ. Manag. 2017, 195, 133–139. [Google Scholar] [CrossRef]
  113. Karaolia, P.; Michael-Kordatou, I.; Hapeshi, E.; Drosou, C.; Bertakis, Y.; Christofilos, D.; Armatas, G.S.; Sygellou, L.; Schwartz, T.; Xekoukoulotakis, N.P.; et al. Removal of antibiotics, antibi-otic-resistant bacteria and their associated genes by graphene-based TiO2 composite photocatalysts under solar radiation in urban wastewaters. Appl. Catal. B Environ. 2018, 224, 810–824. [Google Scholar] [CrossRef]
  114. Talwar, S.; Sangal, V.K.; Verma, A. Feasibility of using combined TiO2 photocatalysis and RBC process for the treatment of real pharmaceutical wastewater. J. Photochem. Photobiol. A Chem. 2018, 353, 263–270. [Google Scholar] [CrossRef]
  115. Rueda-Marquez, J.J.; Levchuk, I.; Fernández Ibañez, P.; Sillanpää, M. A critical review on application of photocatalysis for toxicity reduction of real wastewaters. J. Clean. Prod. 2020, 258, 120694. [Google Scholar] [CrossRef]
  116. Ibáñez, M.; Lor, E.G.; Bijlsma, L.; Morales, E.; Pastor, L.; Hernandez, F. Removal of emerging contaminants in sewage water subjected to advanced oxidation with ozone. J. Hazard. Mater. 2013, 260, 389–398. [Google Scholar] [CrossRef]
  117. Martínez, F.; Molina, R.; Rodríguez, I.; Pariente, M.I.; Segura, Y.; Melero, J.A. Techno-economical assessment of coupling Fen-ton/biological processes for the treatment of a pharmaceutical wastewater. J. Environ. Chem. Eng. 2018, 6, 485–494. [Google Scholar] [CrossRef]
  118. Olvera-Vargas, H.; Oturan, N.; Buisson, D.; Oturan, M.A. A coupled Bio-EF process for mineralization of the pharmaceuticals furosemide and ranitidine: Feasibility assessment. Chemosphere 2016, 155, 606–613. [Google Scholar] [CrossRef]
  119. Ganzenko, O.; Trellu, C.; Oturan, N.; Huguenot, D.; Péchaud, Y.; Van Hullebusch, E.D.; Oturan, M.A. Electro-Fenton treatment of a complex pharmaceutical mixture: Mineralization efficiency and biodegradability enhancement. Chemosphere 2020, 253, 126659. [Google Scholar] [CrossRef] [PubMed]
  120. Karaolia, P.; Michael-Kordatou, I.; Hapeshi, E.; Alexander, J.; Schwartz, T.; Fatta-Kassinos, D. Investigation of the potential of a membrane BioReactor followed by solar Fenton oxidation to remove antibiotic-related microcontaminants. Chem. Eng. J. 2017, 310, 491–502. [Google Scholar] [CrossRef]
  121. Lastre-Acosta, A.M.; Palharim, P.H.; Barbosa, I.M.; Mierzwa, J.C.; Teixeira, A.C.S.C. Removal of sulfadiazine from simulated industrial wastewater by a membrane bioreactor and ozonation. J. Environ. Manag. 2020, 271, 111040. [Google Scholar] [CrossRef]
Figure 1. Classification of advanced oxidation processes. Abbreviations used: O3, ozonation; H2O2, hydrogen peroxide; UV, ultraviolet radiation; US, ultrasound energy; Fe2+, ferrous ion, modified of [17].
Figure 1. Classification of advanced oxidation processes. Abbreviations used: O3, ozonation; H2O2, hydrogen peroxide; UV, ultraviolet radiation; US, ultrasound energy; Fe2+, ferrous ion, modified of [17].
Water 13 02094 g001
Table 1. Reactions of the different types of advanced oxidation processes.
Table 1. Reactions of the different types of advanced oxidation processes.
Advanced Oxidation ProcessesReactionsDescriptionExamples of Pollutants That Can Be Removed
O3/UV H 2 O + O 3   h v   2 OH + O 2
2 OH H 2 O 2
The photolysis of the ozone produces the formation of hydroxyl radicals [17].Imidacloprid [19]
H2O2/UV 2 OH h v H 2 O 2 The formation of hydroxyl radicals is generated by photolysis of H2O2 [17].Bacterias (E. Coli, S. Enteritidis and E. Faecalis) [20]
Inactivation of a multidrug resistant E. Coli [21]
O3/H2O2/UV 2 O 3 + H 2 O 2   h v   2 OH + 3 O 2 Combination of systems O3/UV and H2O2/UV. The use of H2O2 it accelerates the decomposition of ozone and increases the generation of OH [17].Tamoxifen [22]
Fe2+/H2O2/UV Fe 2 + + H 2 O 2 Fe 3 + + OH + OH The Fenton reaction is the process most often applied when it is necessary to remove recalcitrant compounds [17].E. Faecalis [23]
Inactivation of a multidrug resistant E. Coli [21]
O3/US H 2 O ) ) ) H + OH
O 3 ) ) )   O 2 g + O ( 3 P )
O ( 3 P ) g + H 2 O 2 OH
O 3 + OH O 2 + HO 2
O 2 + H HO 2
O 3 + HO 2 2 O 2 + OH
OH + OH H 2 O
HO 2 + OH H 2 O + O 2
Ultrasounds are represented as ‘(((‘ [24].7α -estradiol, 17β-estradiol, estrone, 17α-dihydroequilin, 17α-ethinyl estradiol, estriol and equilin [25]
H2O2/US H 2 O 2 ) ) ) OH + OH
H 2 O 2 + O 2 ) ) ) HO 2 + HO 2
H 2 O 2 + OH HO 2 + H 2 O
The combination of ultrasound and H2O2 achieves the formation of hydroxyl radicals [26].7α -estradiol, 17β-estradiol, estrone, 17α-dihydroequilin, 17α-ethinyl estradiol, estriol and equilin [25]
Electrochemical oxidation[Anode]
H 2 O H + + OH ads + e
OH ads O ads orO 2 + H + + e
O ads + O 2 O 3
[Cathode]
O 2 + 2 H + + 2 e H 2 O 2
In this type of oxidation, the compound to be degraded reacts to the oxidants, which have been electrochemically generated in situ [27].Reactive Orange 16, Reactive Violet 4, Reactive Red 228, and Reactive Black 5 [28]
Anodic oxidation Anode + H 2 O H + + Anode OH + e In this process, organic compounds directly react with heterogeneous hydroxyl radicals (OH), formed by oxidation of water at the surface of anode with high oxygen overpotential [29].Acid Red 1 azo dye [30]
Electro-Fenton O 2 g + 2 H + + 2 e H 2 O 2 Hydrogen peroxide is produced electrochemically through the cathodic reduction of dissolved oxygen on a carbon electrode [31].Non-polar organochlorine pesticides [32]
Direct yellow 9 azo dye [33]
Ozonation in an alkaline medium O 3 + OH O 3 + OH
O 3 O 2 + O
O + H + OH
The degradation of the compound occurs through the action of the ozone itself as well as through the radicals generated in the alkaline medium [34].Bisphenol A, paraxantina [35]
Ozonation with hydrogen peroxide H 2 O 2 + 2 O 3 2 OH + 3 O 2
HO 2 + O 3 HO 2 + O 3
Hydrogen peroxide in an aqueous solution is partially dissociated to hydroperoxide anion which reacts with ozone [36].Fluoroquinolone antibiotics and Clarithromycin [37]
Hydrogen peroxide and catalyst Fe 2 + + H 2 O 2 Fe 3 + + OH + OH This type of reaction is very similar to photo-Fenton processes but the formation rate of OH radicals is lower [17].Doxycycline (catalyst CoFe2O4) [38]
Catalytic ozonation Fe 2 + + O 3 FeO 2 + + O 2
FeO 2 + + H 2 O Fe 3 + + OH + OH
In catalytic ozonation the most widely used catalyst is Fe2+ [36].Naphthalene [39]
Photocatalytic ozonation:
O3/UV/TiO2
TiO 2 h v h + + e
e + O 3 O 3
O 3 O 2 + O
O + H 2 O OH + OH
h + + OH OH
Titanium dioxide (TiO2) is the most effective catalyst of those used in AOPs.
The basic mechanism is described by [40].
Diclofenac [41]
Heterogeneous photocatalysis:
UV/TiO2/H2O2
TiO 2 h v TiO 2 ( e + h + )
TiO 2 h + + O H a d TiO 2 + OH
H 2 O 2 + e OH + OH
In these processes, titanium dioxide is combined with hydrogen peroxide and UV radiation [42].Inactivation of A multidrug resistant E. Coli [21]
Table 2. Classification and functions of pharmaceutical compounds.
Table 2. Classification and functions of pharmaceutical compounds.
Pharmaceutical CompoundsUseExamples [52]
AntibioticsAntibacterial activity. Human and veterinary useAmoxicillin, ampicillin, cefaclor, cefalexin, ciprofloxacin, chlortetracycline, clarithromycin, difloxacine, doxycycline, enoxacin, erythromycin, lincomycin, levofloxacin, metromidazole, mecillinam, ofloxacin, oxytetracycline, penicillin, sulfamethoxazole, sulfadiazine, sulfamethizole, sulfathiazole, sulphapyridine, tetracyclines, trimethoprim, tylosin…
Analgesics and anti-Inflammatory PharmaceuticalsPain relief and the reduction of inflammationIbuprofen, diclofenac, paracetamol, acetaminophen, acetylsalicylic acid, fenoprofen, indomethacin, naproxen, nimesulide, mefenamic acid, fluoxetine, ketoprofen, phenazone…
Antidepressants and AnticonvulsantsRelief from mental symptoms and the treatment of epileptic seizuresDiazepam, carbamazepine, doxepin, imipramine, amitriptyline, primidone, salbutamol, meprobamate, fluxetine, oxazepam, gabapentin, phenobarbital, thioridazine, dilantin…
Lipid RegulatorsRegulation of cholesterol and blood triglyceridesClofibric acid, clofibrate, benzafibrate, fenofibric acid, etofibrate, gemfibrozil, simvastatin, furosemide, Bendroflumethiazide…
β-blockersReduction of blood pressureAtenolol, metoprolol, propranolol, sotalol, timolol…
X-ray contrastsDiagnostic contrast (organ visibility)Iopromide, iopamidol, diatrizoate…
Eestrogens, Progestogens, Androgens, Glucocorticoids, Phytoestrogens& HormonesRegulation of female/male sexual development, maintenance of pregnancy, growth promotion in meat-producing animals, control of immune function, treatment of breast cancer, lymphomas and leukaemias17b-Oestradiol (E2), estrone (E1), estriol, diethylstilbestrol (DES), 17-α ethynylestradiol, mestranol, zerranol, trenbolone acetate, melengestrol acetate, tamoxifen, testosterone, phytosterols, sesquiterpenes, androstenedione, beclomethasone, progesterone, norethindrone…
Table 3. Classification and functions of personal care products.
Table 3. Classification and functions of personal care products.
Personal Care ProductsUseExamples [52]
FragrancesPleasant odoursPolycyclic and macrocyclic musks (musk xylol and musk ketone), phthalates, tonalide, celestolide, galaxolide….
Sun–screen agentsSkin protection from solar ultraviolet radiationBenzophenone, 4-methylbenzylidene camphor, homosalate, octocrylene, oxybenzone, octylmethoxycinnamate, octyldimethyl-PABA…
Insect repellentsRepellentN,N-diethyltoluamide,…
AntisepticsDisinfectantsClorophene, triclosan, bromoprene, 2-phenylphenol, 4-chlorocresol, 4-chloroxylenol…
PreservativesPrevention of undesirable chemical changes or microbial decompositionMethylparaben, 2-phenoxyethanol, ethyl 4-hydroxybenzoate, butyl 4-hydroxybenzoate…
Table 4. Examples of representative PCPs, illicit drugs, and pesticides that can be removed by AOPs.
Table 4. Examples of representative PCPs, illicit drugs, and pesticides that can be removed by AOPs.
Advanced Oxidation ProcessesExamples of Representative PCPs, Illicit Drugs and Pesticides That Can Be Removed
O3-based AOPsPCPs: Triclosan [84]; Ethylparaben, propylparaben, and butylparaben [90]
O3/UVPCPs: Triclosan [88]; Butylparaben [91];
Illicit drugs: Cocaine and its metabolite [103,104,110]
VUV/UV/chlorinePesticides: Dimethoate [DMT], atrazine [ATZ], prometon [PMT], propoxur [PPX], bromacil [BRM], and propachlor [PPC] [109]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bermúdez, L.A.; Pascual, J.M.; Martínez, M.d.M.M.; Poyatos Capilla, J.M. Effectiveness of Advanced Oxidation Processes in Wastewater Treatment: State of the Art. Water 2021, 13, 2094. https://doi.org/10.3390/w13152094

AMA Style

Bermúdez LA, Pascual JM, Martínez MdMM, Poyatos Capilla JM. Effectiveness of Advanced Oxidation Processes in Wastewater Treatment: State of the Art. Water. 2021; 13(15):2094. https://doi.org/10.3390/w13152094

Chicago/Turabian Style

Bermúdez, Laura Antiñolo, Jaime Martín Pascual, María del Mar Muñio Martínez, and Jose Manuel Poyatos Capilla. 2021. "Effectiveness of Advanced Oxidation Processes in Wastewater Treatment: State of the Art" Water 13, no. 15: 2094. https://doi.org/10.3390/w13152094

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