Next Article in Journal
The Influence of Sulfurization and Carbonization on Mo-Based Catalysts for CH3SH Synthesis
Previous Article in Journal
Deciphering the Pivotal Reaction Conditions for Hydrogen Production from Tar Catalytic Cracking by Perovskite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Advanced Oxidation Processes for Degrading Pharmaceuticals in Wastewater—A Review

by
Nur Nabaahah Roslan
1,
Harry Lik Hock Lau
1,
Nurul Amanina A. Suhaimi
1,
Nurulizzatul Ningsheh M. Shahri
1,
Sera Budi Verinda
2,
Muhammad Nur
3,
Jun-Wei Lim
4,5 and
Anwar Usman
1,*
1
Department of Chemistry, Faculty of Science, University Brunei Darussalam, Jalan Tungku Link, Gadong BE1410, Brunei
2
Biomedical Graduate Program, Faculty of Medicine, Universitas Diponegoro, Tembalang Campus, Semarang 50275, Indonesia
3
Integrated Laboratory, Center for Plasma Research, Universitas Diponegoro, Tembalang Campus, Semarang 50275, Indonesia
4
HICoE-Centre for Biofuel and Biochemical Research, Department of Fundamental and Applied Sciences, Institute of Self-Sustainable Building, University Teknologi PETRONAS, Seri Iskandar 32610, Perak Darul Ridzuan, Malaysia
5
Centre for Herbal Pharmacology and Environmental Sustainability, Chettinad Hospital and Research Institute, Chettinad Academy of Research and Education, Kelambakkam 603103, Tamil Nadu, India
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(3), 189; https://doi.org/10.3390/catal14030189
Submission received: 6 February 2024 / Revised: 4 March 2024 / Accepted: 8 March 2024 / Published: 10 March 2024

Abstract

:
A large variety of pharmaceutical compounds have recently been detected in wastewater and natural water systems. This review highlighted the significance of removing pharmaceutical compounds, which are considered indispensable emerging contaminants, from wastewater and natural water systems. Various advanced oxidation processes (AOPs), including UV-H2O2, Fenton and photo-Fenton, ozone-based processes, photocatalysis, and physical processes, such as sonolysis, microwave, and electron beam irradiation, which are regarded as the most viable methods to eliminate different categories of pharmaceutical compounds, are discussed. All these AOPs exhibit great promising techniques, and the catalytic degradation process of the emerging contaminants, advantages, and disadvantages of each technique were deliberated. Heterogeneous photocatalysis employing metal oxides, particularly anatase TiO2 nanoparticles as catalysts activated by UV light irradiation, was reviewed in terms of the electron–hole separation, migration of the charge carriers to the catalyst surfaces, and redox potential of the charge carriers. This brief overview also emphasized that anatase TiO2 nanoparticles and TiO2-based nanomaterials are promising photocatalysts, and a combination of photocatalysis and other AOPs enhanced photocatalytic degradation efficiency. Finally, the challenges of applying anatase TiO2-based photocatalysis in environmental remediation and wastewater treatments to degrade pharmaceutical compounds, including mass spectroscopic analysis and a biological activity test of by-products of the emerging contaminants resulting from photocatalysis, are summarized.

Graphical Abstract

1. Introduction

In recent years, the consumption of pharmaceuticals used in medical treatments, public healthcare, and upholding a high standard of living, increased inevitably along with the global population at an annual rate of 5.8% [1]. This prompts the occurrence of pharmaceuticals in the environment, particularly in wastewater and sewage systems of hospitals, clinics, pharmaceutical industries, and residential areas, due to their uncontrolled discharge, non-metabolized in humans and animals, or expiry date [2,3,4]. Pharmaceuticals that are mostly organic compounds with complex aromatic structures and persistent in aquatic systems have specific biological activities, so their existence in the environment is considered indispensable emerging contaminants [5,6,7]. The majority of pharmaceuticals found in wastewater are antibiotics, analgesics, painkillers, β-blockers, blood–lipid regulators, cytostatic drugs, steroidal hormones, anti-inflammatory drugs, antidepressants, stimulants, and other medications used for the treatment of various diseases [3,4,5]. As these pharmaceuticals are resistant to light, physical, and chemical treatments, they are not effectively degraded using the existing conventional methods in wastewater treatments and could reach and pollute surface water, groundwater, and water systems [8,9].
The accumulation of active pharmaceutical compounds in water systems results in the emergence of antimicrobial resistance genes and antibiotic resistance genes, causing serious threats to aquatic organisms, the environment, and eventually human health through trophic transfer [10,11]. Moreover, the presence of multiple pharmaceuticals in water systems could induce synergistic effects, enhancing their ecotoxicity and other considerable impacts [12]. Ecotoxicological risks of the active pharmaceutical compounds have been linked to various cancers in humans, as well as physiological perturbation and reproduction dysfunction of fishes and microorganisms. Therefore, the increasing emerging contaminants in the environment have become a serious global concern. With this in mind, in addition to regulations for the proper handling and disposal of unused pharmaceuticals, cost-effective and highly efficient methods are being developed and implemented in wastewater treatments as part of the global efforts to remove pharmaceuticals from water systems [13].
In general, there are various methods of removing pharmaceuticals from water systems, and most of them are based on biological, chemical, and physical approaches. Among the promising methods, adsorption [14,15,16], nanofiltration and nanoadsorption [17,18], membrane bioreactors [19,20], biological degradation [21,22], and advanced oxidation processes (AOPs) [23,24] have been explored and demonstrated to be successful in the removal of pharmaceuticals from aqueous solutions. Considering the respective advantages and disadvantages of each method, adsorption, biological degradation, and AOPs have attracted great attention [25].
Using the adsorption method, for instance, a large variety of adsorbents with their functional groups being protonated or deprotonated, pharmaceutical compounds are easily immobilized through hydrogen-bonding or electrostatic interactions. Given these advantages, the adsorptive removal of pharmaceuticals has been intensively explored on different adsorbents, such as sludge-derived biochar [26], biochar [27], activated carbon [28], clay [29], alkali-activated clay [16], and polymeric materials [15]. This method has been demonstrated to be successful in removing and decreasing the bioavailability of a large number of pharmaceuticals from aquatic media. These extensive research studies also revealed that the adsorption behavior of pharmaceuticals depends strongly on their functional groups, molecular conformation, ionic nature, and solubility in water, while the functional groups and net surface charge on the adsorbent surfaces play an important role in adsorbent–pharmaceutical interactions. It is worth noting that several key issues related to the post-adsorption steps to treat pharmaceuticals remain crucial challenges [30]. Similar problems are also accounted for in other techniques, such as flocculation, coagulation, and membrane filtration. From this point of view, these techniques are less practical in comparison with AOPs, which could degrade the organic pollutants and pharmaceuticals instantaneously into smaller chemical compounds that could possibly be less toxic.
In order to eliminate the pharmaceuticals in water systems, a biological degradation method, employing aerobic or anaerobic biological reactors using microorganisms, such as bacteria, fungi, algae, yeast, nematodes, and protozoa, has also been developed [31,32]. This method relies on breaking down organic polymers using normal cellular processes of those microorganisms. Although this method is the most cost-effective and ecological, its application is limited to biodegradable pharmaceuticals. Furthermore, the biological process in employed microorganisms can be altered in the presence of pharmaceuticals [33]. In addition, this method requires excessive energy consumption and intensive labor to grow the employed microbes, and it has been pointed out that this method has low efficacy, rapid saturation, and unpleasant smell [34].
An interesting breakthrough in eliminating pharmaceuticals in aqueous solutions is the use of AOPs that can generate reactive oxygen species (ROS) in aqueous solutions or on the catalyst surfaces, which are capable of oxidizing and degrading pharmaceuticals into small compounds. AOPs have been proven to be a good alternative for the rapid degradation and elimination of refractory and non-biodegradable pharmaceutical compounds, such as antibiotics [35,36]. Therefore, AOPs are considered to be highly effective in the remediation of pharmaceuticals and other emerging pollutants. Although the detailed fragmentation pathways of pharmaceuticals upon oxidation still remain a research challenge, the by-products have been reported to be biologically inactive, suggesting that ROS eventually oxidizes the emerging contaminants into smaller sizes that are biodegradable and environmentally benign oxidation end products [37]. There are various established AOPs, which are dominantly classified into UV–hydrogen peroxide (UV/H2O2), Fenton and photo-Fenton, ozone-based processes, photocatalysis, and physical processes (including sonolysis, microwave, and electron beam irradiation) [24]. Among these AOPs, photocatalysis is the simplest method, and photocatalysts can be reused multiple times in wastewater treatment [38]. A large number of research efforts employing metal oxides, such as TiO2, ZnO, WO3, and ZnWO3 nanoparticles as photocatalysts activated by UV light irradiation, have been devoted to the removal of pharmaceutical compounds from aqueous solutions [39].
Considering that all AOPs are flexible and are based on the same principle, where they generate ROS with high oxidation capacities to initiate the oxidization of organic pollutants, they can be combined to improve the removal efficiency of pharmaceutical compounds [40]. The combination of AOPs could also enhance cost-effectivity in terms of speed, the use of chemicals, and energy consumption. Therefore, it is crucial to outline the knowledge of photocatalytic degradation of pharmaceutical compounds by AOPs, including UV/H2O2, Fenton and photo-Fenton, ozone-based processes, photocatalysis, and sonolysis. Several articles reviewing this topic have been devoted by different research groups [24,41,42]. Thus, this review article focuses on pharmaceuticals in wastewater and provides an overview of the recent research and its efforts on photocatalytic degradation, highlighting the AOPs and degradation process of pharmaceuticals and other organic pollutants in aqueous solutions. The degradation mechanism of pharmaceutical compounds and the advantages and disadvantages of each AOP are discussed. In particular, this review also summarizes the important role of the photocatalytic activity of different catalysts in the heterogeneous photocatalytic degradation of pharmaceutical compounds in aqueous solutions, highlighting that photocatalysis on anatase-TiO2 nanoparticles activated by UV light irradiation is the most promising method. This review article is divided into several sections, providing detailed pharmaceuticals found in wastewater and their degradation by various AOPs, delving into the mechanisms and recent studies on the degradation of the pharmaceutical compounds by various AOPs. Finally, future perspectives on the prospects and potential practical applications of anatase TiO2-based nanoparticles and nanocomposites in removing pharmaceuticals in wastewater treatment are highlighted.

2. Pharmaceuticals in the Environment

Global consumption and demand for pharmaceuticals have significantly increased along with population growth, economic development, and standard of living improvement. Increasing the awareness of chronic diseases and the establishment of healthcare practices also have elevated the market of pharmaceuticals [43]. For instance, it is reported that over thirty million tons of pharmaceuticals have been consumed worldwide in 2023 [44]. As some portions (10−90%) of pharmaceuticals consumed by patients who receive medical treatments are not metabolized, the pharmaceutical compounds and their derivatives are excreted into the sewage system. The presence of pharmaceuticals in the environment, particularly in urban wastewater, should be proportional to their consumption, and the pharmaceuticals in wastewater finally enter wastewater treatment plants and water systems [45].
Based on their therapeutic applications, the pharmaceuticals found in wastewater mainly include antibiotics, analgesics, antidepressants, antihistamines, anti-inflammatory drugs, painkillers, cardiovascular drugs, antidyskinetic medicines, diuretic drugs, statins, anti-epileptic drugs, and antihypertensive agents [46,47]. These pharmaceuticals are commonly used to treat fever, reduce inflammation and allergic diseases, and alleviate symptoms, such as pain, headaches, cold, flu, arthritis, musculoskeletal injuries, dyskinesia and Parkinson’s disease, high blood pressure, and high cholesterol, as summarized in Table 1. As mentioned above, these pharmaceutical compounds are not effectively degraded by conventional methods in water treatment plants, and these emerging contaminants could eventually enter the water systems. In this sense, the presence of these pharmaceutical compounds in the aquatic environment, such as water systems, freshwater ecosystems, marine environments, and groundwater [20], was detected in the order between ng/L and μg/L [48,49]. Although, with this concentration, the therapeutic effects of these pharmaceutical compounds are negligible, and bioaccumulation, chronic toxicity, and mutagenicity of their pharmacologically active components in water systems pose potential hazards to aquatic life, the environment, and human health. In fact, long-term exposure to low-concentrated antibiotics, nonsteroid anti-inflammatory drugs (NSAIDs), antihypertensive drugs, and antidepressants has been demonstrated to affect the cardiovascular and cardiac irregularities in zebrafish [50,51] and water fleas (Daphnia magna) [52], as well as hypoglycemia, growth inhibition, and sexual development alteration in aquatic organisms [2,53].
In addition, Rodríguez et al. evaluated the toxicity and risk of pharmaceuticals present in wastewater and reported highly hazardous pharmaceuticals detected in the hospital wastewater samples, such as azithromycin, clarithromycin, doxycycline, cephalexin, ciprofloxacin, and ofloxacin [54]. Other antibiotics, such as sulfamethoxazole, could cause mutations in genes and chronic health effects [55]. In general, the presence of antibiotics in the environment can lead to the enhancement of antibiotic-resistant genes and multi-drug-resistant microbes that have been part of the global public health crisis [56,57]. Therefore, the removal of these emerging contaminants from wastewater is indispensable.
As displayed in Figure 1, pharmaceutical compounds generally have homocyclic and/or heterocyclic aromatic rings attached to primary and secondary amine (−NH2 and −NH), carbonyl (−CO), ether (−O−), hydroxyl (−OH), methoxyl (−OCH3), or sulfoxide (−SO) functional groups. These functional groups could be protonated or deprotonated, depending on the acidity of the aqueous medium, as represented by their respective pKa values. With the pKa value in either an acidic or basic condition, as listed in Table 1, most pharmaceutical compounds are in their protonated, deprotonated, or zwitterionic forms at ambient pH. From the chemistry viewpoint, pharmaceutical compounds easily make intermolecular interactions through hydrogen-bonding and/or ionic interactions with any biological and non-biological substances [58]. Therefore, they can diffuse and immobilize efficiently on the surfaces of solid substances. By optimizing these facts, adsorption and AOPs have been demonstrated to degrade pharmaceutical compounds efficiently [59,60].

3. Degradation of Pharmaceuticals by Advanced Oxidation Processes in Aqueous Solution

As mentioned above, pharmaceuticals are highly resistant organic compounds that are toxic and difficult to eliminate using conventional methods in wastewater treatment plants, while AOPs have attracted great attention due to their notable performance in degrading pharmaceuticals in aqueous solutions. The COVID-19 virus outbreak globally has prompted more studies on AOPs to decontaminate and disinfect hospital wastewater. AOPs are believed to be ecofriendly and innovative methods that could be involved in wastewater treatments to enhance the existing anti-pollutant technology procedures. AOPs are also considered as clean methods, generating free radicals that have at least a single unpaired electron, such as superoxide anion radicals (O2 + eO2 ; E = +0.33 V), hydroxyl radicals (H2O → OH + H+ + e; E = +2.8 V), hydroperoxyl radicals (O2 + H+ + eHO2; E = +1.7 V), singlet oxygen (1O2 + eO2; E = +1.48 V), ozone (O3 + 2H+ + 2e → O2 + H2O; E = +2.07 V), hydroperoxide anions (HO2 + 2e + H2O → 3OH; E = −0.88 V), sulfate anion radicals (SO42−SO4 + e; E = +2.6 V), and carbonate anion radicals (CO32−CO3 + e; E = +1.81 V). These free radicals are the most common ROS that are responsible for oxidizing pharmaceuticals in aqueous solutions or on catalyst surfaces [61,62]. Among the AOPs, UV-H2O2, Fenton and photo-Fenton, ozone-based processes, sonolysis, and photocatalysis, as individuals or a combination of them, have been demonstrated to be feasible to degrade pharmaceutical compounds in aqueous solutions [24]. These methods will be discussed in detail in the subsequent sections, taking into account the specific characteristics and capabilities of each AOP as well as the oxidation process related to the chemical structure of pharmaceutical compounds.

3.1. UV–Hydrogen Peroxide

In the UV-H2O2 process, UV light irradiation accelerates the photolysis of H2O2, generating active OH radicals in aqueous solutions due to the O-O bond cleavage of H2O2 [24,63], according to the following general reaction:
H 2 O 2 + h v 2 O H
The Gibbs free energy and enthalpy of this light-induced dissociation reaction of H2O2 are G = 237 kJ mol−1 and H = 190 kJ mol−1 [64], suggesting that the reaction is spontaneous and thermodynamically feasible. The rate of photolysis is enhanced either with the pH of the medium or with the concentration of H2O2, suggesting that the light-induced dissociation of H2O2 is favorable in solutions with less acidic conditions and follows the first-order reaction kinetics. The OH radical is a strong oxidant with a standard reduction potential of +2.80 V, which is higher compared with that of most oxidizing agents, so this radical is able to oxidize a wide variety of organic compounds, including pharmaceuticals [65,66]. It is important to note that the oxidation reaction by the radical is ultrafast with a rate constant typically in the range of 108 − 1011 ms−1, and the OH radical has a short lifetime within a few tens of ns [67,68]. These suggest that the reaction is actually controlled by the diffusion of both the radical and organic compounds and, hence, the degradation rate of the organic compounds is extremely low. In this sense, the degradation rate can be enhanced by increasing the concentration of H2O2 in the contaminated water solution. Moreover, in solutions with high H2O2 concentrations, the OH radicals could react with H2O2 to form a hydroperoxyl (HO2) radical, which is also capable of oxidizing organic compounds [68]. Although the UV-H2O2 method is considered to consume a relatively cheap chemical agent, the use of high concentrations of H2O2 makes this method non-cost-effective [68].

3.2. Fenton, Photo-Fenton, and Electro-Fenton

In principle, Fenton utilizes homogenous catalytic ferric (Fe2+) ions in the photolysis of H2O2 in aqueous solutions under acidic conditions [69]. The spontaneous dissociation of H2O2 to the formation of OH radicals is catalyzed by dissolved Fe2+ ions, according to the Haber–Weiss mechanism [70] as follows:
F e 2 + + H 2 O 2 F e 3 + + O H +   O H
In the Fenton reaction, the Fe2+ ion is the most frequently employed due to its abundance, low toxicity, and simplicity in the removal of pollutants from water. The additional UV radiation in the photo-Fenton method accelerates the dissociation of H2O2 and degrades organic compounds more efficiently compared to the Fenton method. Both Fenton and photo-Fenton are preferable due to their less technological complexity and less specialized equipment, and they could be environmentally beneficial upon the cyclic reaction to from Fe2+ ions as follows:
F e 3 + + H 2 O 2 F e 2 + + H + +   O 2 H
Reactions (2)–(3) have to occur efficiently to realize the entire Fenton process. Therefore, the drawback of the Fenton and photo-Fenton methods is that they are non-regenerated catalysts. When the latter reaction is inefficient, the former reaction leads to the formation of Fe-containing sludge, and the buildup of Fe3+ ions inhibits the whole process and produces secondary iron pollution [68]. Another downside of the Fenton and photo-Fenton processes is the excessive usage of acid to maintain acidic conditions at pH 3 [24].
Electro-Fenton has also been demonstrated to be a promising water treatment technology to remove a variety of organic pollutants from water [71]. This method also utilizes Fe2+ ions to accelerate the dissociation of H2O2 to form OH radical as an effective oxidation agent for the decomposition of stubborn pollutants [71]. In an electro-Fenton cell, the OH radical is produced in situ in the acidic aqueous solution and is positively related to the applied electric field. In this method, there are few possibilities to generate OH, including (i) the use of a sacrificial Fe anode as a source of Fe2+ ions and the direct addition of H2O2, (ii) the generation of H2O2 by an oxygen-sparging cathode and the addition of Fe2+ ions into the solution, (iii) the use of a sacrificial Fe anode and the generation of H2O2 by an oxygen-sparging cathode, and (iv) the generation of an OH radical and Fe2+ through the dissociation of H2O2 in the solution and the reduction in Fe3+ ions on the cathode. The disadvantages of this method are a decrease in the catalytic activity associated with the scavenging of the OH radical by Fe2+ ion, the precipitation of Fe(OH)3 at high pHs, and the poisoning effect due to the formation of a polymer layer on the anode surface.

3.3. Ozone-Based Processes

The application of ozone has been successfully demonstrated to degrade pharmaceuticals, such as antibiotic, anticancer, antipsychotic, and painkiller drugs contained in artificial wastewater [72] and other organic pollutants, industrial effluents, and pesticides [24]. Ozone is a strong oxidant with a reduction potential of 2.07 V and can oxidize organic compounds in two ways: (i) a direct electrophilic attack by ozone molecules and (ii) an indirect attack through the generation of an H2O2 and OH radical when ozone breaks down in water, as given by
2 O 3 + 2 H 2 O H 2 O 2 + 2 O 2 + 2   OH
Ozone-based processes employ either ozonation or the formation of ozone microbubbles to produce ROS by dissolving ozone in an aqueous medium. The degradation efficiency of pharmaceuticals or organic pollutants in water by ozone-based processes can be increased by additional H2O2 or UV light irradiation in order to enhance the production of the OH radical [73].
The drawbacks of ozone-based processes include low solubility of ozone in water, high energy consumption to generate ozone that leads to high costs, and the formation of by-products with genotoxic aldehydes, like formaldehyde [74]. Another disadvantage is the low efficiency against compounds with amide linkages due to their ozone-resistant properties [75]. Therefore, ozonation could be used as a pre-treatment process rather than in the purification of water [24]. However, adding a catalyst can alleviate the formation of toxic by-products with the same principle. This method is commonly employed and known as catalytic ozonation. Exploiting several metals, metal oxides, minerals, and carbon materials as a catalyst promotes the complete degradation of organic compounds and reduces partial oxidation in forming harmful intermediates [76]. In this sense, activated carbon is widely utilized in achieving the complete mineralization of antibiotics due to their inertness, stability, large specific surface area, and porosity [77,78]. Their high adsorptive properties enhance an excellent mass transfer in developing an efficient catalytic ozonation process. These properties and ozone decomposition determine catalytic activity and can influence the complete removal of pollutants. Incorporating catalysts into ozone-based processes favors indirect oxidation compared to direct electrophilic ozone reactions with organic compounds. In other words, ozone saturation adsorbing onto the catalyst surface reduces the ozone availability for direct electrophilic attack.
A study of the removal of cephalexin via catalytic ozonation catalyzed using Mn-anchored zeolite reported excellent catalytic performance with a removal efficiency of 97% within only 2 min of retention time [78]. A comparison between non-catalytic ozonation was also conducted, and the difference between the two approaches showed that the degradation of cephalexin was 79.2-fold higher with the addition of the Mn@zeolite catalyst [78].

3.4. Ultrasonic Irradiation

Ultrasonic irradiation or sonolysis is an innovative and sustainable method for the removal of pharmaceutical waste from an aqueous medium. This technique employs ultrasound to alter and modify the target compounds, without any addition of chemicals, through water fragmentation upon collapsing the generated sonic microcavities. In principle, high-frequency sound waves are passed through a specific area within a liquid medium to generate sonic cavitation, which results in the formation of microcavities that rapidly expand and then violently collapse, producing shock waves within a short period of time [79,80,81]. The extremely high local temperature and pressure induce a pyrolysis process inside the cavity, where the dissociation of water generates an OH radical, which oxidizes organic pollutants if the radicals efficiently diffuse into the solution. As a result, volatile and non-volatile molecules within and in close proximity to the microcavities are oxidized and degraded [82]. Diffusion of the generated OH radical in the solution is governed by several parameters, including reactor geometry and frequency of the ultrasound wave. The application of sonolysis has been recently demonstrated in the degradation of azo dyes, but the sonolysis of pharmaceuticals is very new and is rarely investigated [83]. The drawback of sonolysis is mainly related to its high energy consumption and low electrical efficiency.
An interesting finding has been reported by Alishiri et al. on the ultrasonic removal of 15 mg/L ciprofloxacin and cephalexin using Fe3O4/GO nanocomposites, where the sonolysis of ciprofloxacin and cephalexin resulted in 96.39% and 97.69% removal efficiency, respectively [84]. In another study, the removal of 20 mg/L sulfamethoxazole, sulfadimethoxine, and sulfamerazine antibiotics using a magnetic Fe3O4–bentonite nanocomposite (Fe3O4-Bt) with a 17 min ultrasonication time showed the removal of 98.06%, 95.49%, and 96.52% [85].

3.5. Photocatalysis

Photocatalysis utilizes the transformation of the photonic energy into chemical energy, employing the intrinsic photoactivity of metal oxide, perovskite-type metal oxide, or metal sulfide nanoparticles activated by light irradiation [86,87]. Compared with homogeneous photocatalysis, which utilizes all the reactants and reagents in the same physical states, heterogeneous photocatalysis is preferable to degrade organic pollutants due to the easy separation of the catalysts after the photocatalytic process [88]. Moreover, heterogeneous photocatalysis has been demonstrated as a promising approach for destroying organic pollutants in wastewater due to its high efficiency, non-selectivity, and robustness with respect to the regeneration of photocatalysts. Photocatalytic degradation of organic pollutants is initiated by light irradiation, causing the excitation of electrons and creating holes in the lattice of the photocatalysts. When electrons and holes escape from recombination and migrate to the catalyst surfaces, depending on their reduction potential, the charge carriers could generate O2 and OH radicals, which initiate oxidation reactions to degrade organic pollutants [89].
Intensive studies have been devoted to finding photocatalysts with good photoactivity, biological and chemical inertness, relatively low bandgap, strong catalytic activity, non-toxicity, long charge carrier diffusion length, high charge mobility, abundant quantity, and inexpensive [90]. These criteria could be fulfilled by several metal oxides and perovskite-type metal oxides, including TiO2, ZnO, CeO2, Mn3O4, SrTiO3, BiVO4, and Bi2WO6 nanoparticles. Supplementing to this, metal oxides, such as TiO2 and ZnO nanoparticles, not only show good photoactivity in the removal of pharmaceuticals, pesticides, and organic dyes, but they can also exhibit antibacterial, antioxidant, and self-cleaning properties due to their morphology, structure, capacity to scavenge free radicals, inhibition of toxic by-products, and hydrophilicity [91,92]. These multifunctional properties give this photocatalytic material potential for the future of nanotechnology, biomedical applications, coatings, and window glass applications. Hence, these photoactive metal oxides can be synthesized with a variety of morphologies and nanostructures, such as nanospheres, nanofibers, nanowires, and nanorods [93]. With the bandgap energy being in the range of 3.0−3.5 eV, the photocatalytic activity of these catalysts depends on light irradiation in the UV region below 420 nm [94,95,96].
Incorporation of a small number of metallic dopants into the crystal lattice of the photocatalysts could reduce their bandgap energy to allow them to be activated by visible light irradiation. However, visible light also excites some organic pollutants, particularly dyes, leading to their spontaneous photolysis, so that the photocatalytic degradation of the dyes on such visible-light-responsive catalysts competes with their spontaneous photolysis [97,98]. In many studies, it has been pointed out that the incorporation of dopants creates trapping states, accelerating electron–hole recombination and decreasing the photocatalytic activity of the metal oxides and perovskite-type metal oxides [99,100].
In efforts to increase active sites, narrow the band gap for wider light absorption, and improve charge migration dynamics, current research highlighted the employment of structural modification, surface sensitization, and the formation of heterojunction [101]. The specific engineering schemes have been designed toward per surface and interfacial engineering, defects engineering, metals as cocatalysts and plasmonic materials, and heterostructured materials [101]. For instance, Zhou et al. have developed novel 0D/2D Bi4V2O11/g-C3N4 S-scheme heterojunctions to absorb light within the visible spectrum with a wavelength of up to 450 nm [102]. This study revealed high photocatalytic degradation efficiency of oxytetracycline and reactive red 2 to be 74.1% and 84.2%, respectively. The combination of two n-type photocatalysts utilizes a more positive valance band and a more negative conduction band, which facilitates high redox capacity and delays charge recombination. Nevertheless, although sunlight can be considered as an alternative cost-effective light source for practical purposes, photocatalyst activated by UV light irradiation has a higher photoactivity, resulting in higher degradation efficiency of organic contaminants and better reproducibility [103]. Another study on the development of a photocatalyst highlighted the fabrication of a Fenton-like catalyst using FeOCl/MoS2 [104]. The catalytic performance has shown 97.3% degradation of methylene blue (dye) and over 89% degradation of tetracycline (antibiotic), atrazine (pesticides), Bisphenol A (endocrine disruptor), carbamazepine (pharmaceutical), and rhodamine B (dye). This high catalytic performance was achieved due to the increased OH radicals from the reduction in Fe3+ to Fenton-active Fe2+ via the oxidation of Mo4+ to Mo6+ [104].
It is worth noting that the photocatalytic activity of metal oxides and perovskite-type metal oxides depends on the suppression of the electron–hole recombination and the efficient migration of the charge carriers to the catalyst surfaces, as well as the redox potential of the charge carriers, which is associated with the reduction potential of conduction and a valence band [105]. The lower reduction potential of the conduction band with respect to that of O2/O2 (E = +0.33 V) facilitates the reduction in solvated oxygen to form an O2 radical, whereas a higher reduction potential of the valence band with respect to that of OH/OH (E = +2.8 V) is required to oxidize the water molecules on the catalyst surfaces to generate the OH radical [106]. Based on the reduction potentials of the conduction band and valence band, TiO2 and SrTiO3 nanoparticles are the most promising photocatalysts, as the excited electrons and holes could generate O2 and OH radicals, respectively. Other metal oxides, such as ZnO and WO3 nanoparticles, could only generate either an O2 or OH radical. This fact emphasizes the high photocatalytic activity of TiO2 and SrTiO3 nanoparticles compared with ZnO and WO3 nanoparticles. Supporting this notion, the degradation rate constant of amoxicillin, ampicillin, and cloxacillin on TiO2 nanoparticles is higher than those on ZnO nanoparticles [107,108]. Figure 2 illustrates the band edge positions of the most commonly used photocatalysts [109]. Metal oxides, such as TiO2, ZnO, WO3, MnO, and CuO nanoparticles, are suitable for pollution degradation due to their strong oxidation properties.
It should be noted that O2 and OH radicals, which are generated on the catalyst surfaces, have short lifetimes within ns to μs, and photocatalysis is an ultrafast process [67,68]. Therefore, the photocatalytic degradation rate of organic pollutants depends strongly on the diffusion and immobilization of the target organic pollutants onto the catalyst surfaces, where the organic pollutants are readily oxidized by the radicals [87]. Using several established diffusion models, highly effective photocatalytic degradation of the organic contaminants has been revealed to take place in early irradiation times due to their rapid external film diffusion to the catalyst surfaces. It was then followed by a slower photocatalytic degradation before saturation at longer irradiation times. It is conceivable that the organic pollutants should be in contact with the radicals on the catalyst surface. With this in mind, one could consider that, if the external mass transfer and intraparticle diffusion of the organic pollutants onto the catalyst surfaces is much slower than the radical formation, the generated active radicals are dissolved in the solution and quickly decay into non-reactive species, so that no photocatalytic processes are observed.

4. Parameters Governing Advanced Oxidation Processes in the Degradation of Pharmaceuticals

The photocatalytic degradation of most organic pollutants triggered by reactive radicals has been attributed to the breaking down of their conjugated system. The oxidation reaction follows several fragmentation pathways and reductive hydrolysis, producing small organic molecules, carbon dioxide, and ammonium hydroxide as final products [110,111]. The photocatalytic degradation of organic compounds should be an ultrafast reaction [112]. The degradation kinetics should be governed by several conventional non-fundamental parameters, including light irradiation power, irradiation time, catalyst dosage, initial concentration, and temperature. The degradation kinetics of the heterogeneous photocatalysis has been well described by the Langmuir–Hinshelwood (L–H) model [113,114], emphasizing that the oxidation reaction of organic pollutants is the pseudo-first-order. The L–H model provides the average photocatalytic degradation process observable in the steady state measurement, although the photocatalytic degradation of organic pollutants should involve rapid multistep or complex reactions. In this sense, photocatalytic degradation should involve many reaction pathways and intermediates, and in most cases, the existence of backward reactions would result in an equilibrium at prolonged irradiation times, so that a complete photocatalytic degradation cannot be reached.
Nevertheless, based on the generalized Smoluchowski equation, the observed rate constant obtained from the L–H model could be positively related to the diffusion constant of organic pollutants in the solution, highlighting an important role of molecular size, molecular conformation, functional groups, and classes of the organic pollutants. Supporting this notion, Suhaimi et al. demonstrated that the photocatalytic degradation of the planar conformation and relatively smaller size of methylene blue on TiO2 nanoparticles surpasses rhodamine B in their binary solution [96]. A similar finding has also been reported by Wang et al. in the photocatalytic degradation of the ternary solution of methylene blue, rhodamine B, and methyl orange on BaTiO3/Bi2WO6 heterojunction photocatalysts [115]. Furthermore, the energetic diffusion and oxidation reaction between organic contaminants and reactive oxygen species can be enhanced by increasing the temperature of the photocatalytic degradation.
Taking into account that the photocatalytic degradation of pharmaceutical compounds depends on the generated O2 and OH radicals, adding a small amount of H2O2, which efficiently produces OH radicals on the catalyst surfaces, should enhance the degradation rate of organic pollutants. Although dissociation of H2O2 is also accelerated under UV light irradiation, leading to the formation of OH radicals in the solution [116], as given by reaction (1) in Section 3.1, the direct oxidation reaction of organic pollutants by the formation of OH radicals in the solution is relatively slow and inefficient and undergoes through different mechanisms from that of the photocatalytic process on the catalyst surfaces. Moreover, the excessive addition of H2O2 would negatively affect degradation efficiency due to the increased scavenging effect of OH radicals. Based on similar phenomena, one could anticipate that a similar effect of additional H2O2 would be observed in the photocatalytic degradation of pharmaceutical compounds.
It is worth noting that the stability and reusability of photocatalysts have also been of interest [117]. Several metal oxides, such as TiO2 and ZnO nanoparticles, are amphoteric and have poor chemical stability at a high pH in the medium [118]. BiVO4 and Fe2O3 nanoparticles have poor charge carrier mobility in the crystal lattice and slow oxygen kinetics on the particle surface, whereas WO3 nanoparticles undergo peroxidation and photocorrosion during the photocatalytic process [119]. However, as the crystal structure and the surface of TiO2, ZnO, SiO2, and Fe3O4 nanoparticles remain unchanged after photoexcitation, these photoactive catalysts can be regenerated and reused [120].
The effect of the pH of a medium, the pKa value of the organic contaminants, and the pHpzc (point of zero charge) of the photocatalyst play important roles in the degradation of pharmaceuticals. By taking into account the pHpzc value of the catalyst, one could predict the changeability of ionic charges of the catalyst surface. The effect of pH involves the protonation and deprotonation of functional groups for electrostatic and ionic interactions between organic compounds and catalyst surfaces. Hence, a variation of the results was reported by several studies on the effect of pH on pharmaceutical degradation. The investigation of photocatalytic degradation of 60.76 μmol/L rifampicin with the addition of 10 mg TiO2 NPs under UV light irradiation shows an increase in the efficiency values from pH 3 to pH 5.63. This indicates favorable degradation of rifampicin antibiotic in an acidic medium, which can be rationalized by the conservation of the rifampicin zwitterion structure between pH 3 and pH 7.9. A decreasing degradation efficiency was reported with increasing basicity of the medium from pH 6 to pH 10 due to the deprotonation of piperazine nitrogen of rifampicin above its pKb value of 7.9, which induces electrostatic repulsion among negative surface charges of TiO2 NPs [121]. In another finding by Yousefi et al., the photocatalytic degradation efficiency of rifampicin by BiOCl/BiOI nanocomposites was optimal in an acidic environment (pH 3 and pH 5) due to the attraction of rifampicin-free electron pairs to the positive surface of the catalyst [122]. The pHpzc of BiOCl/BiOI was 7.5, leading to an unsatisfactory rifampicin removal at pH 8 to pH 10. In addition to this discussion, the pH of the reaction medium also affects the degradation efficiency of pharmaceutical compounds using different AOP techniques, particularly homogenous Fenton and photo-Fenton. Several studies reveal that the optimal degradation efficiency by these techniques could only be achieved at pH 3 [123]. Above this pH, the formation of hydroxyl radicals is insufficient, as the concentration of H+ ions decreases. On the other hand, below pH 3, the degradation of organic compounds is inefficient due to slow H2O2 dissociation in the formation of HO2 radicals.

5. A Combination of Advanced Oxidation Processes

Although individual AOPs can be used to degrade pharmaceutical compounds and other organic pollutants in wastewater, a combination of AOPs has been employed in many studies and reached satisfactory results [24]. In particular, a combination of sonolysis with ozonation has been demonstrated to efficiently degrade various pharmaceuticals, such as diclofenac, sulfamethoxazole, and carbamazepine in an aqueous solution [124]. A coupled Fenton reaction with high-frequency sonolysis (862 kHz) has been demonstrated to degrade ibuprofen from both distilled water and wastewater [125]. Another important combination is photocatalysis and sonication [126,127], where applying ultrasonic waves (20–1000 kHz) in photocatalysis could induce cavitation bubbles, enhancing the generation of OH radicals and the degradation of organic pollutants. It is worth noting that, in these combinations of AOPs, the mechanisms of respective sonolysis, ozonation, and photocatalytic degradation of the organic pollutants are unchanged. Applying additional sonication provides more kinetic energy and enhances the diffusion and immobilization of the organic pollutants on the microbubble or catalyst surfaces, resulting in higher degradation efficiency.

6. Reported Studies of the Advanced Oxidation Processes of Pharmaceuticals

Various AOPs, such as UV-H2O2, Fenton-based, ozone-based, and semiconductor-based photocatalytic processes, have been employed for the degradation of organic pollutants, including pharmaceutical compounds [128,129,130,131,132]. Among these AOPs, a heterogeneous photocatalytic process employing metal oxides, metal sulfides, and perovskite-type materials as photocatalysts is more widely used for the removal of persistent pollutants, creating safe end products [133]. More specifically, photocatalysis using anatase TiO2 nanoparticles is the most promising method due to their high photocatalytic activity, cost-effectivity, environmental friendliness, and chemical and physical inertness [134,135,136,137]. Therefore, several research groups have published numerous papers on the usage of AOPs using TiO2-based nanoparticles as catalysts in the photocatalytic degradation of pharmaceuticals in water [138,139,140].
Light-driven AOPs have been applied to degrade carbamazepine, ciprofloxacin, ibuprofen, and a mixture of these pharmaceutical compounds on TiO2 nanoparticles as catalysts. It was found that ciprofloxacin showed the highest degradation rate, followed by ibuprofen and carbamazepine [141]. The removal of levofloxacin upon photocatalysis using TiO2 nanoparticles as catalysts has also been investigated. The degradation of levofloxacin was confirmed by decreasing the antibacterial activity of its by-products against Escherichia coli (E. coli), where the antibacterial activity of the treated levofloxacin solution decreases gradually with the irradiation time [142].
Effective photocatalytic degradation of three antiviral pharmaceuticals, namely 1-amantadine, 2-amantadine, and rimantadine, under UV irradiation in the presence of TiO2 nanoparticles has been established [143]. The removal of cefotaxime upon photocatalysis under UV light radiation of two types of nanocatalysts, i.e., TiO2 and TiO2/kaolin, has been investigated and compared with the degradation of this β-lactam antibiotic under UV-H2O2 treatment. It was reported that the UV-H2O2 treatment is more effective and promising in removing cefotaxime from an aqueous solution [144].
The photocatalytic degradation of ibuprofen from synthetic wastewater has been reported under UV-LED irradiation in the presence of TiO2, ZnO, H2O2, potassium peroxomonosulfate (KSO5), and potassium peroxodisulfate (K2S2O8). Compared to other systems investigated, ibuprofen was found to degrade most efficiently on TiO2 nanoparticles [145]. On the other hand, the photodegradation of sulfamethoxazole, metronidazole, and ciprofloxacin on TiO2 nanoparticles under UV light irradiation has also been investigated [146]. The effective degradation of metoprolol, a medication to treat blood pressure and heart rate, through photocatalytic activity under the visible light of a Xenon lamp in the presence of B-doped TiO2 nanoparticles has been performed. The photocatalytic degradation of this antibiotic was enhanced by a B dopant in the lattice structure of TiO2 nanoparticles [147]. The photocatalytic degradation of isoniazid and rifampicin using TiO2 and ZnO catalysts has been investigated and compared with the degradation of these antibiotics upon the UV-A photo-Fenton process [148]. A radical scavenging assay was also conducted to determine the effects of free radicals, electron vacancy, and singlet oxygen, and it was revealed that isoniazid was degraded by various ROS, while rifampicin was degraded by electron vacancy.
The degradation of norfloxacin has been investigated using different AOPs, including UV/H2O2, Fenton, photo-Fenton process, and UV/TiO2 [149], highlighting that this antibiotic is efficiently degraded by the photo-Fenton process. The degradation of ribavirin using the UV/TiO2/H2O2 system has also been reported. The results revealed that ribavirin was degraded, and the formation of its transformation products was found to contribute to ecotoxicity. Although the method was effective in degrading ribavirin, it also led to the formation of toxic transformation products, which require further processes to degrade [150]. Similarly, upon the photocatalytic degradation of memantine using UV-C/H2O2 and UV-A/TiO2, it has been reported that the presence of ROS, especially OH radicals, is more prominent than O2 radicals, while intermediate species in the degradation process of memantine are non-biodegradable and are considered relatively toxic [151].
It is interesting to note that the photocatalytic degradation of anti-epileptic carbamazepine in the presence of the BiOCl/AgCl composite catalyst activated by simulated sunlight irradiation is more efficient than that in the presence of BiOCl [152]. The high photocatalytic activity of the BiOCl/AgCl composite is attributed to efficient electron–hole separation and the migration of the charge carriers to the catalyst surfaces. The degradation pathways of carbamazepine involve oxidation by photo-induced radicals, aromatic ring rearrangement, and hydroxylation [152]. In contrast, the photodegradation of carbamazepine in estuarine water is enhanced by the presence of chloride ions [153], which form reactive chlorine species (RCS). This highlights that the UV/H2O2, Fenton, photo-Fenton processes, and UV/TiO2 in the presence of chlorine, have multiple reactive species (ROS and RCS) to oxidize the emerging contaminants [154].
A study of the degradation of metronidazole treated using various methods, including ultrasonic irradiation, UV light irradiation, H2O2, ozonation, and peroxonium, has been recently reported [155]. In addition, the degradation of this antibiotic was also investigated via UV and UV/H2O2/O3 associated with heterogeneous catalysis of TiO2 nanoparticles (UV/H2O2/O3/TiO2@PU) that have been functionalized in polyurethane macroporous structures. It was found that both ultrasonic and UV light irradiation had less efficacy in degrading metronidazole, but this antibiotic was efficiently degraded in the UV/H2O2/O3/TiO2@PU system [155].
Metal-doped TiO2 nanoparticles have been intensively utilized as photocatalysts in the degradation of a variety of pharmaceuticals in water under solar light. For instance, Ni-doped TiO2 nanoparticles have been utilized and found to be an excellent photocatalyst to degrade ofloxacin [156], cephalexin, and tetracycline in water and wastewater with high photocatalytic degradation efficiency [157]. On the other hand, Ag-doped TiO2 nanoparticles under UV light irradiation also could efficiently degrade chloramphenicol [158]. Fe-doped TiO2 nanoparticles under UV light have successfully mediated the photocatalytic degradation of metronidazole from aqueous solutions and other pharmaceutical compounds from industrial pharmaceutical wastewater [159]. Carbon-doped TiO2 nanoparticles activated by visible UV light were successfully utilized to degrade carbamazepine and diclofenac, where complete depletion was achieved at a concentration of the catalyst of 244 mg/L [160].
The effect of supporting materials in the photodegradation of pharmaceuticals is demonstrated in the degradation of carbamazepine on mesoporous dendritic silica-supported TiO2 nanoparticles (TiO2/MDS) in the presence of persulfate [161]. The high photocatalytic activity of the TiO2/MDS/persulfate system is further revealed in the photocatalytic degradation of amoxicillin, sulfamethoxazole, bisphenol A, and naproxen [161]. It was speculated that the high photocatalytic activity of TiO2/MDS is due to the suppression of electron–hole recombination. The degradation of ceftriaxone through photocatalytic degradation using light irradiation in the presence of heterogeneous activated carbon-supported TiO2 nanoparticles has also been investigated [162]. The findings suggest that the photocatalyst made of activated carbon–TiO2 composites can be utilized in the removal of pharmaceutical wastes by the photocatalytic process in an environmentally friendly, reliably efficient, effective, reusable, and sustainable manner [162]. Another important TiO2-based composite to degrade pharmaceutical wastes from aqueous solutions is multi-walled carbon nanotube–TiO2-SiO2 nanocomposites. It has been demonstrated that the introduction of MWCNT reduces the bang gap, allowing visible light excitation, enhancing the surface area, and improving the photocatalytic degradation efficiency of acetaminophen in an aqueous solution [163].
Ali et al. investigated the photocatalytic degradation of the endocrine-disrupting drug (estrone hormone (E1)) using silica-supported g-C3N4/WO3 under UV and visible light irradiation [164]. The result shows excellent estrone hormone removal of 100% and 96.2% using UV light and visible light irradiation, respectively, under the set parameters of 3000 µg of photocatalyst dosage, pH 7, and 300 µg/L hormone concentration.
Finally, a combination of nanofiltration with AOPs, including UV, UV/TiO2, and UV/H2O2, has been employed to degrade commonly consumed anticancer drugs, such as etoposide, paclitaxel, cyclophosphamide, and ifosfamide. The results revealed that direct photolysis was able to effectively degrade etoposide and paclitaxel, whereas the degradation of cyclophosphamide and ifosfamide was found to be ineffective [165]. It is also notable that a combination of ozonation and the photocatalytic degradation of metronidazole and cephalexin using the Urea/TiO2/ZnFe2O4/Clinoptilolite catalyst under visible-light irradiation and ozone injection showed higher potential for the degradation of these pharmaceuticals due to a large number of generated ROS [166].

7. Pilot Test of Advanced Oxidation Processes on Industrial Wastewater

Intending to promote the various AOPs on an industrial scale, some studies conducted pilot tests using advanced technological prototypes on the practical applications of AOPs on wastewater remediation. Actual wastewater samples were collected from the influent drains of industries, hospitals, and urban wastewater plants and experimented for their removal efficiencies and toxicities. The pre- and post-assessments for water quality in the potentiality of AOPs in eliminating toxicity were analyzed based on several parameters and types of contaminants. The physical assessment may include parameters such as temperature, pH, electrical conductivity, chemical oxygen demand (COD), total organic carbon (TOC), dissolved organic carbon (DOC), particulate matter (turbidity), hardness, color, and salinity. Supplemental chemical analysis may require the identification of organic substances, heavy metals, nitrogen, nitrates, nitrites, ammonia, chlorine, phosphates, and phosphorus found in water. In addition to this, the presence of E. coli, coliform bacteria pathogenic microorganisms, Klepsiella pneumonia (K. pneumonia), Staphylococcus aureus (S. aureus), abiotic particles, and viruses in water content are often investigated. Hence, most methods of identification use inductively coupled plasma for detecting heavy metals; a UV-Vis spectrophotometer for analyzing color, COD, DOC, TOC, organic substances, and turbidity; ion chromatography for measuring significant anions and cations in water content; and a real-time bacteria sensor, polymerase chain reaction, surface-enhanced Raman scattering, and laser-induced fluorescence spectroscopy for detecting E.coli, K. pneumonia, S. aureus, bacteria, and viruses [167].
Most of the reported pilot tests, as tabulated in Table 2, show successful pharmaceutical waste removal that follow a realistic wastewater treatment approach. Research focusing on developing small-scale AOP models still requires some improvements, with economic and operational costs being the main limitations. The cost can be reduced by improving the light absorptivity of a photocatalyst toward a wider solar spectrum. Solar irradiation can lower the expenses for electrical consumption and UV light generation. Hence, several strategies were implemented to optimize the photocatalytic activity to overcome its limitations and the occurrence of charge recombination, as mentioned in Section 3.5.
Wastewater contains several impurities and contaminants; for example, soluble ions, such as nitrate ions, chloride ions, and carbonate ions, could act as free radical scavengers and inhibit the efficient degradation of organic compounds and pharmaceuticals. The laboratory-scale removal efficiency of antibiotics in ultra-pure water has resulted in a high degradation rate. However, evidence of intermediate by-products and possible toxicity caused by secondary products is insufficient and requires additional examination [172].

8. Conclusions and Future Perspectives

In this review, the presence of pharmaceuticals in water systems that pose potential hazards to aquatic life, the environment, and humans due to their bioaccumulation, chronic toxicity, and mutagenicity was comprehensively summarized. Advanced oxidation processes (AOPs), including UV-H2O2, Fenton and photo-Fenton, ozone-based processes, photocatalysis, and physical processes, are the most viable and fascinating methods to eliminate pharmaceuticals and other organic pollutants in wastewater and water systems. AOPs are radical-based processes where the generated radicals act as reactive oxygen species (ROS) and are apparently considered the most feasible radicals to degrade the emerging contaminants. Among these AOPs, photocatalysis is initiated by light radiation in the presence of metal oxide or perovskite-type metal oxide nanoparticles as photocatalysts to degrade pharmaceuticals in water systems due to their simplicity and eco-friendliness. The photoactivity of the photocatalysts is determined by the electron–hole separation and migration of the charge carriers to the catalyst surfaces and the redox potential of the charge carriers. With these criteria, TiO2 and SrTiO3 nanoparticles were found to be the most promising photocatalysts, as their excited electrons and holes could generate O2 and OH radicals, respectively, which are responsible for initiating the degradation of pharmaceutical compounds. Until now, anatase TiO2 nanoparticles have been revealed to have the best photocatalytic activity and have been intensively explored in the degradation of various pharmaceuticals and other organic pollutants in aqueous solutions. For that reason, anatase TiO2-based nanomaterials activated by UV light irradiation have also been successfully applied in the photocatalysis of pharmaceutical compounds. In this brief overview, anatase TiO2 nanoparticles and their nanocomposites have been emphasized to be the most promising photocatalysts. A combination of photocatalysis and other AOPs has been highlighted to be a promising efficient method for enhancing the photocatalytic degradation of pharmaceutical compounds. Finally, the effects of diffusion, polarizability, steric hindrance, molecular structure, and the optical properties of pharmaceutical compounds on their degradation by AOPs are conceivable factors to be evaluated in future studies to provide a better understanding of the photocatalytic degradation of these indispensable emerging contaminants. This should include mass spectroscopic analysis and biological activity tests of by-products of pharmaceutical compounds resulting from photocatalysis, as well as the development of new metal oxide, perovskite-type metal oxide, or metal sulfide and their nanocomposite photocatalysts for environmental remediation and wastewater treatment.

Author Contributions

Conceptualization, A.U.; resources, N.N.R., H.L.H.L. and A.U.; data curation, N.A.A.S., N.N.M.S. and S.B.V.; writing—original draft preparation, N.N.R., H.L.H.L. and A.U.; validation, A.U., J.-W.L. and M.N.; writing—review and editing, N.A.A.S. and A.U.; supervision, A.U. and M.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Peña, O.I.G.; Zavala, M.Á.L.; Ruelas, H.C. Pharmaceuticals market, consumption trends and disease incidence are not driving the pharmaceutical research on water and wastewater. Int. J. Environ. Res. Public Health 2021, 18, 2532. [Google Scholar] [CrossRef] [PubMed]
  2. Hejna, M.; Kapuścińska, D.; Aksmann, A. Pharmaceuticals in the aquatic environment: A review on eco-toxicology and the remediation potential of algae. Int. J. Environ. Res. Public Health 2022, 19, 7717. [Google Scholar] [CrossRef]
  3. Aus der Beek, T.; Weber, F.-A.; Bergmann, A.; Hickmann, S.; Ebert, I.; Hein, A.; Kuster, A. Pharmaceuticals in the environment—Global occurrences and perspectives. Environ. Toxicol. Chem. 2016, 35, 823–835. [Google Scholar] [CrossRef]
  4. Polianciuc, S.I.; Gurzău, A.E.; Kiss, B.; Ștefan, M.G.; Loghin, F. Antibiotics in the environment: Causes and consequences. Med. Pharm. Rep. 2020, 93, 231–240. [Google Scholar] [CrossRef]
  5. Samal, K.; Mahapatra, S.; Ali, M.H. Pharmaceutical wastewater as emerging contaminants (EC): Treatment technologies, impact on environment and human health. Energy Nexus 2022, 6, 100076. [Google Scholar] [CrossRef]
  6. Rogowska, J.; Zimmermann, A. Household pharmaceutical waste disposal as a global problem—A review. Int. J. Environ. Res. Public Health 2022, 19, 15798. [Google Scholar] [CrossRef] [PubMed]
  7. Aliste, M.; Garrido, I.; Hernandez, V.; Flores, P.; Hellín, P.; Navarro, S.; Fenoll, J. Assessment of reclaimed agro-wastewater polluted with insecticide residues for irrigation of growing lettuce (Lactuca sativa L.) using solar photocatalytic technology. Environ. Pollut. 2022, 292, 118367. [Google Scholar] [CrossRef] [PubMed]
  8. Frascaroli, G.; Reid, D.; Hunter, C.; Roberts, J.; Helwig, K.; Spencer, J.; Escudero, A. Pharmaceuticals in wastewater treatment plants: A systematic review on the substances of greatest concern responsible for the development of antimicrobial resistance. Appl. Sci. 2021, 11, 6670. [Google Scholar] [CrossRef]
  9. Shigei, M.; Assayed, A.; Hazaymeh, A.; Dalahmeh, S.S. Pharmaceutical and antibiotic pollutant levels in wastewater and the waters of the Zarqa river, Jordan. Appl. Sci. 2021, 11, 8638. [Google Scholar] [CrossRef]
  10. Xie, Z.; Lu, G.; Yan, Z.; Liu, J.; Wang, P.; Wang, Y. Bioaccumulation and trophic transfer of pharmaceuticals in food webs from a large freshwater lake. Environ. Pollut. 2017, 222, 356–366. [Google Scholar] [CrossRef]
  11. Yang, H.; Lu, G.; Yan, Z.; Liu, J.; Dong, H.; Bao, X.; Zhang, X.; Sun, Y. Residues, bioaccumulation, and trophic transfer of pharmaceuticals and personal care products in highly urbanized rivers affected by water diversion. J. Hazard. Mater. 2020, 391, 122245. [Google Scholar] [CrossRef] [PubMed]
  12. 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]
  13. Letsoalo, M.R.; Sithole, T.; Mufamadi, S.; Mazhandu, Z.; Sillanpaa, M.; Kaushik, A.; Mashifana, T. Efficient detection and treatment of pharmaceutical contaminants to produce clean water for better health and environmental. J. Clean. Prod. 2023, 387, 135798. [Google Scholar] [CrossRef]
  14. Gil, A.; Taoufik, N.; Gacía, A.M.; Korili, A.S. Comparative removal of emerging contaminants from aqueous solution by adsorption on an activated carbon. Environ. Technol. 2019, 40, 3017–3030. [Google Scholar] [CrossRef] [PubMed]
  15. Shahrin, E.W.E.S.; Narudin, N.A.H.; Shahri, N.N.M.; Nur, M.; Lim, J.-W.; Bilad, M.R.; Mahadi, A.H.; Hobley, J.; Usman, A. A comparative study of adsorption behavior of rifampicin, streptomycin, and ibuprofen contaminants from aqueous solutions onto chitosan: Dynamic interactions, kinetics, diffusions, and mechanisms. Emerg. Contam. 2023, 9, 100199. [Google Scholar] [CrossRef]
  16. Majid, A.F.A.; Dewi, R.; Shahri, N.N.M.; Shahri, E.W.E.S.; Kusrini, E.; Shamsuddin, N.; Lim, J.-W.; Thongratkaew, S.; Kajornsak Faungnawakij, K.; Usman, A. Enhancing adsorption performance of alkali activated kaolinite in the removal of antibiotic rifampicin from aqueous solution. Colloids Surf. A Physicochem. Eng. Asp. 2023, 676, 132209. [Google Scholar] [CrossRef]
  17. Foureaux, A.F.S.; Reis, E.O.; Lebron, Y.; Moreira, V.; Santos, L.V.; Amaral, M.S.; Lange, L.C. Rejection of pharmaceutical compounds from surface water by nanofiltration and reverse osmosis. Sep. Purif. Technol. 2019, 212, 171–179. [Google Scholar] [CrossRef]
  18. Moreira, J.B.; Santos, T.D.; Zaparoli, M.; de Almeida, A.C.A.; Costa, J.A.V.; de Morais, M.G. An overview of nanofiltration and nanoadsorption technologies to emerging pollutants treatment. Appl. Sci. 2022, 12, 8352. [Google Scholar] [CrossRef]
  19. Oberoi, A.S.; Surendra, K.C.; Wu, D.; Lu, H.; Wong, J.W.C.; Khana, S.K. Anaerobic membrane bioreactors for pharmaceutical-laden wastewater treatment: A critical review. Bioresour. Technol. 2022, 361, 127667. [Google Scholar] [CrossRef]
  20. Hosseinpour, S.; Azimian-Kivi, M.; Jafarzadeh, Y.; Yegani, R. Pharmaceutical wastewater treatment using polypropylene membranes incorporated with carboxylated and PEG-grafted nanodiamond in membrane bioreactor (MBR). Water Environ. J. 2021, 35, 1249–1259. [Google Scholar] [CrossRef]
  21. Dhiman, N.; Chaudhary, S.; Singh, A.; Chauhan, A.; Kumar, R. Sustainable degradation of pharmaceutical waste using different fungal strains: Enzyme induction, kinetics and isotherm studies. Environ. Technol. Innov. 2022, 25, 102156. [Google Scholar] [CrossRef]
  22. Moghaddam, A.; Khayatan, D.; Barzegar, P.E.F.; Ranjbar, R.; Yazdanian, M.; Tahmasebi, E.; Alam, M.; Abbasi, K.; Ghaleh, H.E.G.; Tebyaniyan, H. Biodegradation of pharmaceutical compounds in industrial wastewater using biological treatment: A comprehensive overview. Int. J. Environ. Sci. Technol. 2023, 20, 5659–5696. [Google Scholar] [CrossRef]
  23. Thakur, A.K.; Kumar, R.; Kumar, A.; Shankar, R.; Khan, N.A.; Gupta, K.N.; Ram, M.; Arya, R.K. Pharmaceutical waste-water treatment via advanced oxidation based integrated processes: An engineering and economic perspective. J. Water Process Eng. 2023, 54, 103977. [Google Scholar] [CrossRef]
  24. Pandis, P.K.; Kalogirou, C.; Kanellou, E.; Vaitsis, C.; Savvidou, M.G.; Sourkouni, G.; Zorpas, A.A.; Argirusis, C. Key points of advanced oxidation processes (AOPs) for wastewater, organic pollutants and pharmaceutical waste treatment: A mini review. ChemEngineering 2022, 6, 8. [Google Scholar] [CrossRef]
  25. Ahmed, M.; Mavukkandy, M.O.; Giwa, A.; Elektorowicz, M.; Katsou, E.; Khelifi, O.; Naddeo, V.; Hasan, S.W. Recent developments in hazardous pollutants removal from wastewater and water reuse within a circular economy. NPJ Clean Water 2022, 5, 12. [Google Scholar] [CrossRef]
  26. Ihsanullah, I.; Khan, M.T.; Zubair, M.; Bilal, M.; Sajid, M. Removal of pharmaceuticals from water using sewage sludge-derived biochar: A review. Chemosphere 2022, 289, 133196. [Google Scholar] [CrossRef] [PubMed]
  27. Chauhan, S.; Shafi, T.; Dubey, B.K.; Chowdhury, S. Biochar-mediated removal of pharmaceutical compounds from aqueous matrices via adsorption. Waste Dispos. Sustain. Energy 2023, 5, 37–62. [Google Scholar] [CrossRef]
  28. Ilavský, J.; Barloková, D. The removal of selected pharmaceuticals from water by adsorption with granular activated carbons. Eng. Proc. 2023, 57, 33. [Google Scholar]
  29. Kryuchkova, M.; Batasheva, S.; Akhatova, F.; Babaev, V.; Buzyurova, D.; Vikulina, A.; Volodkin, D.; Fakhrullin, R.; Rozhina, E. Pharmaceuticals removal by adsorption with montmorillonite nanoclay. Int. J. Mol. Sci. 2021, 22, 9670. [Google Scholar] [CrossRef]
  30. Ganthavee, V.; Trzcinski, A.P. Removal of pharmaceutically active compounds from wastewater using adsorption coupled with electrochemical oxidation technology: A critical review. J. Ind. Eng. Chem. 2023, 126, 20–35. [Google Scholar] [CrossRef]
  31. Kummerer, K. Antibiotics in the aquatic environment—A review. Part I. Chemosphere 2009, 75, 417–434. [Google Scholar] [CrossRef] [PubMed]
  32. Silva, A.; Delerue-Matos, C.; Figueiredo, S.A.; Freitas, O.M. The use of algae and fungi for removal of pharmaceuticals by bioremediation and biosorption processes: A review. Water 2019, 11, 1555. [Google Scholar] [CrossRef]
  33. Fernandes, J.P.; Almeida, C.M.R.; Salgado, M.A.; Carvalho, M.F.; Mucha, A.P. Pharmaceutical Compounds in Aquatic Environments—Occurrence, fate and bioremediation prospective. Toxics 2021, 9, 257. [Google Scholar] [CrossRef]
  34. Lakhani, S.; Acharya, D.; Sakariya, R.; Sharma, D.; Patel, P.; Shah, M.; Prajapati, M. A comprehensive study of bioremediation for pharmaceutical wastewater treatment. Clean. Chem. Eng. 2022, 4, 100073. [Google Scholar] [CrossRef]
  35. Yu, S.Y.; Xie, Z.H.; Wu, X.; Zheng, Y.Z.; Shi, Y.; Xiong, Z.K.; Zhou, P.; Liu, Y.; He, C.S.; Pan, Z.C.; et al. Review of advanced oxidation processes for treating hospital sewage to achieve decontamination and disinfection. Chin. Chem. Lett. 2024, 35, 108714. [Google Scholar] [CrossRef]
  36. Lupu, G.-I.; Orbeci, C.; Bobirica, L.; Bobirica, C.; Pascu, L.F. Key principles of advanced oxidation processes: A systematic analysis of current and future perspectives of the removal of antibiotics from wastewater. Catalysts 2023, 13, 1280. [Google Scholar] [CrossRef]
  37. Verinda, S.B.; Muniroh, M.; Yulianto, E.; Maharani, N.; Gunawan, G.; Amalia, N.F.; Hobley, J.; Usman, A.; Nur, M. Degradation of ciprofloxacin in aqueous solution using ozone microbubbles: Spectroscopic, kinetics, and antibacterial analysis. Heliyon 2022, 8, e10137. [Google Scholar] [CrossRef]
  38. Chung, J.; Chung, S.; Lee, G.; Lee, Y.-W. Application of wastewater reuse with photocatalyst prepared by sol-gel method and its kinetics on the decomposition of low molecular weight pollutants. Int. J. Environ. Res. Public Health 2020, 17, 4203. [Google Scholar] [CrossRef]
  39. Friedmann, D. A general overview of heterogeneous photocatalysis as a remediation technology for wastewaters containing pharmaceutical compounds. Water 2022, 14, 3588. [Google Scholar] [CrossRef]
  40. Kumari, P.; Kumar, A. Advanced oxidation process: A remediation technique for organic and non-biodegradable pollutant. Results Surf. Interfaces 2023, 11, 100122. [Google Scholar] [CrossRef]
  41. Dong, C.; Fang, W.; Yi, Q.; Zhang, J. A comprehensive review on reactive oxygen species (ROS) in advanced oxidation processes (AOPs). Chemosphere 2022, 308, 136205. [Google Scholar] [CrossRef]
  42. Varma, K.S.; Tayade, R.J.; Shah, K.J.; Joshi, P.A.; Shukla, A.D.; Gandhi, V.G. Photocatalytic degradation of pharmaceutical and pesticide compounds (PPCs) using doped TiO2 nanomaterials: A review. Water-Energy Nexus 2020, 3, 46–61. [Google Scholar] [CrossRef]
  43. Tawfik, E.A.; Tawfik, A.F.; Alajmi, A.M.; Badr, M.Y.; Al-Jedai, A.; Almozain, N.H.; Bukhary, H.A.; Halwani, A.A.; Al Awadh, S.A.; Alshamsan, A.; et al. Localizing pharmaceuticals manufacturing and its impact on drug security in Saudi Arabia. Saudi Pharm. J. 2022, 30, 28–38. [Google Scholar] [CrossRef] [PubMed]
  44. Vora, L.K.; Gholap, A.D.; Jetha, K.; Thakur, R.R.S.; Solanki, H.K.; Chavda, V.P. Artificial intelligence in pharmaceutical technology and drug delivery design. Pharmaceutics 2023, 15, 1916. [Google Scholar] [CrossRef] [PubMed]
  45. Pouzola, T.; Levi, Y.; Bertrand-Krajewski, J.-L. Modelling daily and hourly loads of pharmaceuticals in urban wastewater. Int. J. Hyg. Environ. Health 2020, 229, 113552. [Google Scholar] [CrossRef] [PubMed]
  46. Patel, M.; Kumar, R.; Kishor, K.; Mlsna, T.; Pittman, C.U., Jr.; Mohan, D. Pharmaceuticals of emerging concern in aquatic systems: Chemistry, occurrence, effects, and removal methods. Chem. Rev. 2019, 119, 3510–3673. [Google Scholar] [CrossRef] [PubMed]
  47. Ślósarczyk, K.; Jakóbczyk-Karpierz, S.; Różkowski, J.; Witkowski, A.J. Occurrence of pharmaceuticals and personal care products in the water environment of Poland: A review. Water 2021, 13, 2283. [Google Scholar] [CrossRef]
  48. Hernández-Tenorio, R.; González-Juárez, E.; Guzmán-Mar, J.L.; Hinojosa-Reyes, L.; Hernández-Ramírez, A. Review of occurrence of pharmaceuticals worldwide for estimating concentration ranges in aquatic environments at the end of the last decade. J. Hazard. Mater. Adv. 2002, 8, 100172. [Google Scholar] [CrossRef]
  49. Bavumiragira, J.P.; Ge, J.; Yin, H. Fate and transport of pharmaceuticals in water systems: A processes review. Sci. Total Environ. 2022, 823, 153635. [Google Scholar] [CrossRef]
  50. Pindling, S.; Azulai, D.; Zheng, B.; Dahan, D.; Perron, G.G. Dysbiosis and early mortality in zebrafish larvae exposed to subclinical concentrations of streptomycin. FEMS Microbiol. Lett. 2018, 365, fny188. [Google Scholar] [CrossRef]
  51. Matijević, G.; Babić, S.; Maršavelski, A.; Stipaničev, D.; Repec, S.; Čož-Rakovac, R.; Klobučar, G. Estimating risk of cardiovascular pharmaceuticals in freshwaters using zebrafish embryotoxicity test—Statins threat revealed. Chemosphere 2023, 313, 137574. [Google Scholar] [CrossRef]
  52. Walter, R.B.; Milano, F.; Brasky, T.M.; White, E. Long-term use of acetaminophen, aspirin, and other nonsteroidal anti-inflammatory drugs and risk of hematologic malignancies: Results from the prospective Vitamins and Lifestyle (VITAL) study. J. Clin. Oncol. 2011, 29, 2424–2431. [Google Scholar] [CrossRef]
  53. Porretti, M.; Arrigo, F.; Di Bella, G.; Faggio, C. Impact of pharmaceutical products on zebrafish: An effective tool to assess aquatic pollution. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2022, 261, 109439. [Google Scholar] [CrossRef] [PubMed]
  54. Rodríguez-Rodríguez, C.E.; Ramírez-Morales, D.; Masis-Mora, M.; Montiel-Mora, J.R.; Soto-Garita, C.; Araya-Valverde, E.; Cambronero-Heinrichs, J.C.; Sànchez-Melsió, A.; Briceño-Guevara, S.; Mendez-Rivera, M.; et al. Occurrence and risk assessment of pharmaceuticals in hospital wastewater in Costa Rica. Chemosphere 2023, 339, 139746. [Google Scholar] [CrossRef] [PubMed]
  55. Manyahi, J.; Moyo, S.; Aboud, S.; Langeland, N.; Blomberg, B. High rate of antimicrobial resistance and multiple mutations in the dihydrofolate reductase gene among Streptococcus pneumoniae isolated from HIV-infected adults in a community setting in Tanzania. J. Glob. Antimicrob. Resist. 2020, 22, 749–753. [Google Scholar] [CrossRef]
  56. Chaturvedi, P.; Chaurasia, D.; Pandey, A.; Gupta, P. Co-occurrence of multidrug resistance, b-lactamase and plasmid mediated AmpC genes in bacteria isolated from river Ganga, northern India. Environ. Pollut. 2020, 267, 115502. [Google Scholar] [CrossRef] [PubMed]
  57. Stachurová, T.; Sýkorová, N.; Semerád, J.; Malachová, K. Resistant genes and multidrug-resistant bacteria in wastewater: A study of their transfer to the water reservoir in the Czech Republic. Life 2022, 12, 147. [Google Scholar] [CrossRef] [PubMed]
  58. Hutchins, K.M. Functional materials based on molecules with hydrogen-bonding ability: Applications to drug co-crystals and polymer complexes. R. Soc. Open Sci. 2018, 5, 180564. [Google Scholar] [CrossRef]
  59. Mansouri, F.; Chouchene, K.; Roche, N.; Ksibi, M. Removal of pharmaceuticals from water by adsorption and advanced oxidation processes: State of the art and trends. Appl. Sci. 2021, 11, 6659. [Google Scholar] [CrossRef]
  60. 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]
  61. Phaniendra, A.; Jestadi, D.B.; Periyasamy, L. Free radicals: Properties, sources, targets, and their implication in various diseases. Ind. J. Clin. Biochem. 2015, 30, 11–26. [Google Scholar] [CrossRef]
  62. Aziz, K.H.H. Heterogeneous catalytic activation of peroxydisulfate toward degradation of pharmaceuticals diclofenac and ibuprofen using scrap printed circuit board. RSC Adv. 2023, 13, 115. [Google Scholar] [CrossRef]
  63. Giri, A.S.; Golder, A.K.; Chakma, S. Kinetics, degradation mechanisms and antibiotic activity reduction of chloramphenicol in aqueous solution by UV/H2O2 process. Water Sci. Technol. 2021, 84, 524. [Google Scholar] [CrossRef]
  64. Cataldo, F. Hydrogen peroxide photolysis with different UV light sources including a new UV-LED light source. New Front. Chem. 2014, 23, 99–110. [Google Scholar]
  65. Scaria, J.; Nidheesh, P.V. Comparison of Hydroxyl-Radical-Based Advanced Oxidation Processes with Sulfate Radical-Based Advanced Oxidation Processes. Curr. Opin. Chem. Eng. 2022, 36, 1–2. [Google Scholar] [CrossRef]
  66. Sievers, M. Advanced oxidation processes. Treatise Water Sci. 2011, 4, 377–408. [Google Scholar]
  67. Barroso-Martinez, J.S.; Romo, A.I.B.; Pudar, S.; Putnam, S.T.; Bustos, E.; Rodriguez-Lopez, J. Real-time detection of hydroxyl radical generated at operating electrodes via redox-active adduct formation using scanning electrochemical microscopy. J. Am. Chem. Soc. 2022, 144, 18896–18907. [Google Scholar] [CrossRef] [PubMed]
  68. Kulišťáková, A. Removal of pharmaceutical micropollutants from real wastewater matrices by means of photochemical advanced oxidation processes—A review. J. Water Process Eng. 2023, 53, 103727. [Google Scholar] [CrossRef]
  69. Litter, M.I.; Mariel Slodowicz, M. An overview on heterogeneous Fenton and photoFenton reactions using zerovalent iron materials. J. Adv. Oxid. Technol. 2017, 20, 20160164. [Google Scholar] [CrossRef]
  70. Lin, P.J.; Yeh, C.H.; Jiang, J.C. Theoretical insight into hydroxyl production via H2O2 decomposition over the Fe3O4(311) surface. RSC Adv. 2021, 11, 36257. [Google Scholar] [CrossRef] [PubMed]
  71. Puga, A.; Moreira, M.M.; Figueiredo, S.A.; Delerue-Matos, C.; Pazos, M.; Rosales, E.; Sanromán, M.Á. Electro-Fenton degradation of a ternary pharmaceutical mixture and its application in the regeneration of spent biochar. J. Electroanal. Chem. 2021, 886, 2–9. [Google Scholar] [CrossRef]
  72. Patel, S.; Mondal, S.; Majumder, S.K.; Das, P.; Ghosh, P. Treatment of a pharmaceutical industrial effluent by a hybrid process of advanced oxidation and adsorption. ACS Omega 2020, 5, 32305–32317. [Google Scholar] [CrossRef]
  73. Du, J.; Wang, C.; Zhao, Z.; Cui, F.; Ou, Q.; Liu, J. Role of oxygen and superoxide radicals in promoting H2O2 production during VUV/UV radiation of water. Chem. Eng. Sci. 2021, 241, 116683. [Google Scholar] [CrossRef]
  74. Rekhate, C.V.; Srivastava, J.K. Recent advances in ozone-based advanced oxidation processes for treatment of wastewater—A review. Chem. Eng. J. Adv. 2020, 3, 100031. [Google Scholar] [CrossRef]
  75. Nakada, N.; Shinohara, H.; Murata, A.; Kiri, K.; Managaki, S.; Sato, N.; Takada, H. Removal of selected pharmaceuticals and personal care products (PPCPs) and endocrine-disrupting chemicals (EDCs) during sand filtration and ozonation at a municipal sewage treatment plant. Water Res. 2007, 41, 4373–4382. [Google Scholar] [CrossRef]
  76. Orge, C.A.; Graça, C.A.L.; Restivo, J.; Pereira, M.F.R.; Soares, O.S.G.P. Catalytic ozonation of pharmaceutical compounds using carbon-based catalysts. Catal. Commun. 2024, in press. [Google Scholar] [CrossRef]
  77. Francoeur, M.; Yacou, C.; Petit, E.; Granier, D.; Flaud, V.; Gaspard, S.; Brosillon, S.; Ayral, A. Removal of antibiotics by adsorption and catalytic ozonation using magnetic activated carbons prepared from Sargassum sp. J. Water Process Eng. 2023, 53, 103602. [Google Scholar] [CrossRef]
  78. Zeng, Y.; Zhuo, Q.; Dai, L.; Guan, B. Mn anchored zeolite molecular nest for enhanced catalytic ozonation of cephalexin. Chemosphere 2023, 335, 139058. [Google Scholar] [CrossRef] [PubMed]
  79. Pilli, S.; Bhunia, P.; Yan, S.; LeBlanc, R.J.; Tyagi, R.D.; Surampalli, R.Y. Ultrasonic pretreatment of sludge: A review. Ultrason. Sonochem. 2011, 18, 1–18. [Google Scholar] [CrossRef] [PubMed]
  80. Li, Z.; Dong, J.; Wang, L.; Zhang, Y.; Zhuang, T.; Wang, H.; Cui, X.; Wang, Z. A power-triggered preparation strategy of nanostructured inorganics: Sonosynthesis. Nanoscale Adv. 2021, 3, 2423. [Google Scholar] [CrossRef] [PubMed]
  81. Fetyan, N.A.H.; Attia, T.M.S. Water purification using ultrasound waves: Application and challenges. Arab. J. Basic Appl. Sci. 2020, 27, 194–207. [Google Scholar] [CrossRef]
  82. Sakkas, P.M.; Schneider, O.; Sourkouni, G.; Argirusis, C. Sonochemistry in the service of SOFC research. Ultrason. Sonochem. 2014, 21, 1939–1947. [Google Scholar] [CrossRef] [PubMed]
  83. Naddeo, V.; Landi, M.; Scannapieco, D.; Belgiorno, V. Sonochemical degradation of twenty-three emerging contaminants in urban wastewater. Desalin. Water Treat. 2013, 51, 6601–6608. [Google Scholar] [CrossRef]
  84. Alishiri, M.; Abdollahi, S.M.; Neysari, A.N.; Ranjbar, S.F.; Abdoli, N.; Afsharjahanshahi, M. Removal of ciprofloxacin and cephalexin antibiotics in water environment by magnetic graphene oxide nanocomposites; optimization using response surface methodology. Results Eng. 2023, 20, 101507. [Google Scholar] [CrossRef]
  85. Hjazi, A.; Almajidi, Y.Q.; Kadhum, W.R.; Aly, M.; Malviya, J.; Fenjan, M.N.; Alawadi, A.; Alsaalamy, A.; Chandramauli, A.; Baharinikoo, L. Optimization of removal of sulfonamide antibiotics by magnetic nanocomposite from water samples using central composite design. Water Resour. Ind. 2023, 30, 100229. [Google Scholar] [CrossRef]
  86. Kong, C.P.Y.; Suhaimi, N.A.A.; Shahri, N.N.M.; Lim, J.-W.; Nur, M.; Hobley, J.; Usman, A. Auramine O UV photocatalytic degradation on TiO2 nanoparticles in a heterogeneous aqueous solution. Catalysts 2022, 12, 975. [Google Scholar] [CrossRef]
  87. Suhaimi, N.A.A.; Kong, C.P.Y.; Shahri, N.N.M.; Nur, M.; Hobley, J.; Usman, A. Dynamics of diffusion- and immobilization-limited photocatalytic degradation of dyes by metal oxide nanoparticles in binary or ternary solutions. Catalysts 2022, 12, 1254. [Google Scholar] [CrossRef]
  88. Qahtan, T.F.; Owolabi, T.O.; Olubi, O.E.; Hezam, A. State-of-the-art, challenges and prospects of heterogeneous tandem photocatalysis. Coord. Chem. Rev. 2023, 492, 2–20. [Google Scholar] [CrossRef]
  89. Wang, J.L.; Xu, L.J. Advanced oxidation processes for wastewater treatment: Formation of hydroxyl radical and application. Crit. Rev. Environ. Sci. Technol. 2012, 42, 251–325. [Google Scholar] [CrossRef]
  90. Zhu, J.; Zhu, Z.; Zhang, H.; Lu, J.; Qiu, Y. Efficient degradation of organic pollutants by peroxymonosulfate activated with MgCuFe-layered double hydroxide. RSC Adv. 2019, 9, 2284. [Google Scholar] [CrossRef]
  91. Thakur, N.; Thakur, N.; Kumar, A.; Thakur, V.K.; Kalia, S.; Arya, V.; Kumar, A.; Kumar, S.; Kyzas, G.Z. A critical review on the recent trends of photocatalytic, antibacterial, antioxidant and nanohybrid applications of anatase and rutile TiO2 nanoparticles. Sci. Total Environ. 2024, 914, 169815. [Google Scholar] [CrossRef] [PubMed]
  92. Yang, L.; Yang, J.; Yang, D.-Q. A durable superhydrophilic self-cleaning coating based on TiO2–SiO2-PAA nanocomposite for photovoltaic applications: Long-term outdoor study. Sol. Energy Mater. Sol. Cells 2024, 268, 112731. [Google Scholar] [CrossRef]
  93. Danish, M.S.S.; Bhattacharya, A.; Stepanova, D.; Mikhaylov, A.; Grilli, M.L.; Khosravy, M.; Senjyu, T. A systematic review of metal oxide applications for energy and environmental sustainability. Metals 2020, 10, 1604. [Google Scholar] [CrossRef]
  94. Malathi, A.; Madhavan, J.; Ashokkumar, M.; Arunachalam, P. A review on BiVO4 photocatalyst: Activity enhancement methods for solar photocatalytic applications. Appl. Catal. A Gen. 2018, 555, 47–74. [Google Scholar]
  95. Zulmajdi, S.L.N.; Zamri, N.I.I.; Yasin, H.M.; Kusrini, E.; Hobley, J.; Usman, A. Comparative study on the adsorption, kinetics, and thermodynamics of the photocatalytic degradation of six different synthetic dyes on TiO2 nanoparticles. React. Kinet. Mech. Catal. 2020, 129, 519–534. [Google Scholar] [CrossRef]
  96. Suhaimi, N.A.A.; Shahri, N.N.M.; Samat, J.H.; Kusrini, E.; Lim, J.-W.; Hobley, J.; Usman, A. Domination of methylene blue over rhodamine B during simultaneous photocatalytic degradation by TiO2 nanoparticles in an aqueous binary solution under UV irradiation. React. Kinet. Mech. Catal. 2022, 135, 511–527. [Google Scholar] [CrossRef]
  97. Xu, C.; Rangaiah, G.P.; Zhao, X.S. Photocatalytic degradation of methylene blue by titanium dioxide: Experimental and modeling study. Ind. Eng. Chem. Res. 2014, 53, 14641–14649. [Google Scholar] [CrossRef]
  98. Bae, S.; Kim, S.; Lee, S.; Choi, W. Dye decolorization test for the activity assessment of visible light photocatalysts: Realities and limitations. Catal. Today 2014, 224, 21–28. [Google Scholar] [CrossRef]
  99. Gatti, T.; Lamberti, F.; Mazzaro, R.; Kriegel, I.; Schlettwein, D.; Enrichi, F.; Lago, N.; Di Maria, E.; Meneghesso, G.; Vomiero, A.; et al. Opportunities from doping of non-critical metal oxides in last generation light-conversion devices. Adv. Energy Mater. 2021, 11, 2101041. [Google Scholar] [CrossRef]
  100. Sakuna, P.; Ketwong, P.; Ohtani, B.; Trakulmututa, J.; Kobkeatthawin, T.; Luengnaruemitchai, A.; Smit, S.M. The influence of metal-doped graphitic carbon nitride on photocatalytic conversion of acetic acid to carbon dioxide. Front. Chem. 2022, 10, 825786. [Google Scholar] [CrossRef]
  101. Ali, H.; Masar, M.; Yasir, M.; Machovsky, M.; Monteiro, O.C.; Kuritka, I. Current trends in environmental and energy photocatalysis and ISO standardization. J. Environ. Chem. Eng. 2023, 11, 111541. [Google Scholar] [CrossRef]
  102. Zhou, L.; Li, Y.; Zhang, Y.; Qiu, L.; Xing, Y. A 0D/2D Bi4V2O11/g-C3N4 S-scheme heterojunction with rapid interfacial charges migration for photocatalytic antibiotic degradation. Acta Phys. Chim. Sin. 2022, 38, 2112027. [Google Scholar] [CrossRef]
  103. Wang, L.; Zhao, J.; Liu, H.; Huang, J. Design, Modification and application of semiconductor photocatalysts. J. Taiwan Inst. Chem. Eng. 2018, 93, 590–602. [Google Scholar] [CrossRef]
  104. Qu, S.; Wang, W.; Pan, X.; Li, C. Improving the Fenton catalytic performance of FeOCl using an electron mediator. J. Hazard. Mater. 2020, 384, 121494. [Google Scholar] [CrossRef]
  105. Pastor, E.; Sachs, M.; Selim, S.; Durrant, J.R.; Bakulin, A.A.; Walsh, A. Electronic defects in metal oxide photocatalysts. Nat. Rev. Mater. 2022, 7, 503–521. [Google Scholar] [CrossRef]
  106. Janczarek, M.; Kowalska, E. On the origin of enhanced photocatalytic activity of copper-modified titania in the oxidative reaction systems. Catalysts 2017, 7, 317. [Google Scholar] [CrossRef]
  107. 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]
  108. Elmolla, E.S.; Chaudhuri, M. Degradation of amoxicillin, ampicillin and cloxacillin antibiotics in aqueous solution by the UV/ZnO photocatalytic process. J. Hazard. Mater. 2010, 173, 445–449. [Google Scholar] [CrossRef]
  109. Suma, D.; Deka, T.; Ningthoukhongjam, P.; Chowdhury, A.; Nair, R.G. A critical review on prospects and challenges of metal-oxide embedded g-C3N4-based direct Z-scheme photocatalysts for water splitting and environmental remediation. Appl. Surf. Sci. Adv. 2022, 11, 100273. [Google Scholar]
  110. Javaid, R.; Qazi, U.Y. Catalytic oxidation process for the degradation of synthetic dyes: An overview. Int. J. Environ. Res. Public Health 2019, 16, 2066. [Google Scholar] [CrossRef] [PubMed]
  111. Ren, G.; Han, H.; Wang, Y.; Liu, S.; Zhao, J.; Meng, X.; Li, Z. Recent advances of photocatalytic application in water treatment: A review. Nanomaterials 2021, 11, 1804. [Google Scholar] [CrossRef]
  112. Gan, W.; Fu, X.; Guo, J.; Zhang, M.; Yu, H.; Ding, C.; Qi, S.; Cao, X.; Sun, Z. Facile synthesis of mesoporous hierarchical TiO2 micro-flowers serving as the scaffolding of 0D Ag3PO4 nanoparticles for the ultra-fast degradation of organic pollutants. J. Alloys Compd. 2022, 909, 164737. [Google Scholar] [CrossRef]
  113. Kumar, K.V.; Porkodi, K.; Rocha, F. Langmuir-Hinshelwood kinetics—A theoretical study. Catal. Commun. 2008, 9, 82–84. [Google Scholar] [CrossRef]
  114. Armenise, S.; García-Bordejé, E.; Valverde, J.L.; Romeo, E.; Monzón, A. A Langmuir-Hinshelwood approach to the kinetic modelling of catalytic ammonia decomposition in an integral reactor. Phys. Chem. Chem. Phys. 2013, 15, 12104–12117. [Google Scholar] [CrossRef]
  115. Wang, Y.; Sun, X.; Xian, T.; Liu, G.; Yang, H. Photocatalytic purification of simulated dye wastewater in different pH environments by using BaTiO3/Bi2WO6 heterojunction photocatalysts. Opt. Mater. 2021, 113, 110853. [Google Scholar] [CrossRef]
  116. Hirakawa, T.; Yawata, K.; Nosaka, Y. Photocatalytic reactivity for O2•− and OH radical formation in anatase and rutile TiO2 suspension as the effect of H2O2 addition. Appl. Catal. A General 2007, 325, 105–111. [Google Scholar] [CrossRef]
  117. Hassan, M.A.; El-Nemr, M.A.; Elkatory, M.R.; Ragab, S.; Niculescu, V.-C.; El-Nemr, A. Principles of photocatalysts and their different applications: A review. Top. Curr. Chem. 2023, 31, 381. [Google Scholar] [CrossRef]
  118. Wang, J.; van Ree, T.; Wu, Y.; Zhang, P.; Gao, L. Metal oxide semiconductors for solar water splitting. In Metal Oxides Energy Technologies; Wu, Y., Ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2018; pp. 205–249. [Google Scholar]
  119. Zheng, G.; Wang, J.; Liu, H.; Murugadoss, V.; Zu, G.; Che, H.; Lai, C.; Li, H.; Ding, T.; Gao, Q. Tungsten oxide nanostructures and nanocomposites for photoelectrochemical water splitting. Nanoscale 2019, 11, 18968–18994. [Google Scholar] [CrossRef] [PubMed]
  120. Bockenstedt, J.; Vidwans, N.A.; Gentry, T.; Vaddiraju, S. Catalyst recovery, regeneration and reuse during large-scale disinfection of water using photocatalysis. Water 2021, 13, 2623. [Google Scholar] [CrossRef]
  121. Suhaimi, N.A.A.; Umar, M.K.H.; Lau, H.L.H.; Roslan, N.N.; Lim, J.-W.; Hobley, J.; Nur, M.; Usman, A. An insight into the photocatalytic degradation of the antibiotic rifampicin by titanium dioxide nanoparticles in aqueous solution under UV light irradiation. Reac. Kinet. Mech. Cat. 2024, in press. [Google Scholar] [CrossRef]
  122. Yousefi, A.; Nezamzadeh-Ejhieh, A. Characterization of BiOCl/BiOI binary catalyst and its photocatalytic activity towards rifampin. J. Photochem. Photobiol. A Chem. 2022, 433, 114135. [Google Scholar] [CrossRef]
  123. Kadıoğlu, E.N.; Öztürk, H.; Eroğlu, H.A.; Akbal, F.; Kuleyin, A.; Özkaraova, E.B. Artificial neural network modeling of Fenton-based advanced oxidation processes for recycling of textile wastewater. J. Ind. Eng. Chem. 2024, in press. [Google Scholar] [CrossRef]
  124. Naddeo, V.; Uyguner-Demirel, C.S.; Prado, M.; Cesaro, A.; Belgiorno, V.; Ballesteros, F. Enhanced ozonation of selected pharmaceutical compounds by sonolysis. Environ. Technol. 2015, 36, 1876–1883. [Google Scholar] [CrossRef] [PubMed]
  125. Adityosulindro, S.; Barthe, L.; González-Labrada, K.; Haza, U.J.J.; Delmas, H.; Julcour, C. Sonolysis and sono-Fenton oxidation for removal of ibuprofen in (waste)water. Ultrason. Sonochem. 2017, 39, 889–896. [Google Scholar] [CrossRef] [PubMed]
  126. Patel, J.; Singh, A.K.; Jain, B.; Yadav, S.; Carabineiro, S.A.C.; Susan, M.A.B.H. Lochrome dark blue azo dye removal by sonophotocatalysis using Mn2+ doped ZnS quantum dots. Catalysts 2021, 11, 1025. [Google Scholar] [CrossRef]
  127. Panda, D.; Manickam, S. Recent advancements in the sonophotocatalysis (SPC) and doped-sonophotocatalysis (DSPC) for the treatment of recalcitrant hazardous organic water pollutants. Ultrason. Sonochem. 2017, 36, 481–496. [Google Scholar] [CrossRef] [PubMed]
  128. Porcar-Santos, O.; Cruz-Alcalde, A.; López-Vinent, N.; Zanganas, D.; Sans, C. Photocatalytic degradation of sulfamethoxazole using TiO2 in simulated seawater: Evidence for direct formation of reactive halogen species and halogenated by-products. Sci. Total Environ. 2020, 736, 2–9. [Google Scholar] [CrossRef] [PubMed]
  129. Ryu, B.; Wong, K.T.; Choong, C.E.; Kim, J.R.; Kim, H.; Kim, S.H.; Jeon, B.H.; Yoon, Y.; Snyder, S.A.; Jang, M. Degradation synergism between sonolysis and photocatalysis for organic pollutants with different hydrophobicity: A perspective of mechanism and application for high mineralization efficiency. J. Hazard. Mater. 2021, 416, 2–11. [Google Scholar] [CrossRef] [PubMed]
  130. Issaka, E.; AMU-Darko, J.N.O.; Yakubu, S.; Fapohunda, F.O.; Ali, N.; Bilal, M. Advanced catalytic ozonation for degradation of pharmaceutical pollutants―A review. Chemosphere 2022, 289, 2–15. [Google Scholar] [CrossRef]
  131. Huang, B.; Yang, C.; Zeng, H.; Zhou, L. Multivalent iron-based magnetic porous biochar from peach gum polysaccharide as a heterogeneous Fenton catalyst for degradation of dye pollutants. Int. J. Biol. Macromol. 2023, 253, 2–8. [Google Scholar] [CrossRef]
  132. Dhawle, R.; Frontistis, Z.; Mantzavinos, D.; Lianos, P. Production of hydrogen peroxide with a photocatalytic fuel cell and its application to UV/H2O2 degradation of dyes. Chem. Eng. J. Adv. 2021, 6, 100109. [Google Scholar] [CrossRef]
  133. Pavel, M.; Anastasescu, C.; State, R.N.; Vasile, A.; Papa, F.; Balint, I. Photocatalytic degradation of organic and inorganic pollutants to harmless end products: Assessment of practical application potential for water and air cleaning. Catalysts 2023, 13, 380. [Google Scholar] [CrossRef]
  134. Nickheslat, A.; Amin, M.M.; Izanloo, H.; Fatehizadeh, A.; Mousavi, S.M. Phenol photocatalytic degradation by advanced oxidation process under ultraviolet radiation using titanium dioxide. J. Environ. Public Health 2013, 2013, 815310. [Google Scholar] [CrossRef] [PubMed]
  135. Luttrell, T.; Halpegamage, S.; Tao, J.; Kramer, A.; Sutter, E.; Batzill, M. Why is anatase a better photocatalyst than rutile?—Model studies on epitaxial TiO2 films. Sci. Rep. 2015, 4, 1–8. [Google Scholar] [CrossRef]
  136. Zulmajdi, S.L.N.; Zamri, N.I.I.; Mahadi, A.H.; Rosli, M.Y.H.; Ja’afar, F.; Yasin, H.M.; Kusrini, E.; Hobley, J.; Usman, A. Sol-gel preparation of different crystalline phases of TiO2 nanoparticles for photocatalytic degradation of methylene blue in aqueous solution. Am. J. Nanomater. 2019, 7, 39–45. [Google Scholar] [CrossRef]
  137. Ahmed, S.N.; Haider, W. Heterogeneous photocatalysis and its potential applications in water and wastewater treatment: A review. Nanotechnology 2018, 29, 2–22. [Google Scholar] [CrossRef]
  138. Yilmaz, E.; Salem, S.; Sarp, G.; Aydin, S.; Sahin, K.; Korkmaz, I.; Yuvali, D. TiO2 nanoparticles and C-nanofibers modified magnetic Fe3O4 nanospheres (TiO2@Fe3O4@C–NF): A multifunctional hybrid material for magnetic solid-phase extraction of ibuprofen and photocatalytic degradation of drug molecules and azo dye. Talanta 2020, 213, 1–9. [Google Scholar] [CrossRef]
  139. Khani, Z.; Schieppati, D.; Bianchi, C.L.; Boffito, D.C. The sonophotocatalytic degradation of pharmaceuticals in water by MnOx-TiO2 systems with tuned band-gaps. Catalysts 2019, 9, 949. [Google Scholar] [CrossRef]
  140. Malesic-Eleftheriadou, N.; Evgenidou, E.N.; Kyzas, G.Z.; Bikiaris, D.N.; Lambropoulou, D.A. Removal of antibiotics in aqueous media by using new synthesized bio-based poly(ethylene terephthalate)-TiO2 photocatalysts. Chemosphere 2019, 234, 746–755. [Google Scholar] [CrossRef]
  141. 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]
  142. Kansal, S.K.; Kundu, P.; Sood, S.; Lamba, R.; Umar, A.; Mehta, S.K. Photocatalytic degradation of the antibiotic levofloxacin using highly crystalline TiO2 nanoparticles. New J. Chem. 2014, 38, 3220–3226. [Google Scholar] [CrossRef]
  143. An, J.; Li, G.; An, T.; Song, W.; Feng, H.; Lu, Y. Photocatalytic degradation of three amantadine antiviral drugs as well as their eco-toxicity evolution. Catal. Today 2015, 258, 602–609. [Google Scholar] [CrossRef]
  144. Abbood, N.S.; Ali, N.S.; Khader, E.H.; Majdi, H.S.; Albayati, T.M.; Saady, N.M.C. Photocatalytic degradation of cefotaxime pharmaceutical compounds onto a modified nanocatalyst. Res. Chem. Intermed. 2023, 49, 43–56. [Google Scholar] [CrossRef]
  145. Wang, Z.; Srivastava, V.; Ambat, I.; Safaei, Z.; Sillanpää, M. Degradation of Ibuprofen by UV-LED/catalytic advanced oxidation process. J. Water Process Eng. 2019, 31, 100808. [Google Scholar] [CrossRef]
  146. Akter, S.; Islam, M.S.; Kabir, M.H.; Shaikh, M.A.A.; Gafur, M.A. UV/TiO2 photodegradation of metronidazole, ciprofloxacin and sulfamethoxazole in aqueous solution: An optimization and kinetic study. Arab. J. Chem. 2022, 15, 103900. [Google Scholar] [CrossRef]
  147. Cavalcante, R.P.; Dantas, R.F.; Bayarri, B.; González, O.; Giménez, J.; Esplugas, S.; Machulek, A. Synthesis and characterization of B-doped TiO2 and their performance for the degradation of metoprolol. Catal. Today 2015, 252, 27–34. [Google Scholar] [CrossRef]
  148. Stets, S.; do Amaral, B.; Schneider, J.T.; de Barros, I.R.; de Liz, M.V.; Ribeiro, R.R.; Nagata, N.; Peralta-Zamora, P. Antituberculosis drugs degradation by UV-based advanced oxidation processes. J. Photochem. Photobiol. A Chem. 2018, 353, 26–33. [Google Scholar] [CrossRef]
  149. Shankaraiah, G.; Poodari, S.; Bhagawan, D.; Himabindu, V.; Vidyavathi, S. Degradation of antibiotic norfloxacin in aqueous solution using advanced oxidation processes (AOPs)—A comparative study. Desalin. Water Treat. 2016, 57, 27804–27815. [Google Scholar] [CrossRef]
  150. Wu, X.; Zhang, J.; Hu, S.; Zhang, G.; Lan, H.; Peng, J.; Liu, H. Evaluation of degradation performance toward antiviral drug ribavirin using advanced oxidation process and its relations to ecotoxicity evolution. Sci. Total Environ. 2022, 850, 157851. [Google Scholar] [CrossRef] [PubMed]
  151. Papac, J.; Ballesteros, S.C.; Tonkovic, S.; Kovacic, M.; Tomic, A.; Cvetnić, M.; Kusic, H.; Senta, I.; Terzić, S.; Ahel, M.; et al. Degradation of pharmaceutical memantine by photo-based advanced oxidation processes: Kinetics, pathways and environmental aspects. J. Environ. Chem. Eng. 2023, 11, 109334. [Google Scholar] [CrossRef]
  152. Meribout, R.; Zuo, Y.; Khodja, A.A.; Piram, A.; Lebarillier, S.; Cheng, J.; Wang, C.; Wong-Wah-Chung, P. Photocatalytic degradation of antiepileptic drug carbamazepine with bismuth oxychlorides (BiOCl and BiOCl/AgCl compsite) in water: Efficiency evaluation and elucidation degradation pathways. J. Photochem. Photobiol. A Chem. 2016, 328, 105–113. [Google Scholar] [CrossRef]
  153. Chiron, S.; Minero, C.; Vione, D. Photodegradation processes of the antiepileptic drug carbamazepine, relevant to estuarine waters. Environ. Sci. Technol. 2006, 40, 5977–5983. [Google Scholar] [CrossRef] [PubMed]
  154. Guo, K.; Wu, Z.; Chen, C.; Fang, J. UV/Chlorine Process: An efficient advanced oxidation process with multiple radicals and functions in water treatment. Acc. Chem. Res. 2022, 55, 286–297. [Google Scholar] [CrossRef]
  155. Sanches-Simões, E.; Testolin, R.C.; Müller, F.M.; Gerlach, O.M.S.; Pimentel-Almeida, W.; Conti-Lampert, A.D.; Corrêa, A.X.R.; Almerindo, G.I.; González, S.Y.G.; Radetski, C.M. Metronidazole degradation by physico-chemical and advanced oxidative processes: Influence of pH and nTiO2-functionalized macroporous support. Water Air Soil Pollut. 2022, 233, 466. [Google Scholar] [CrossRef]
  156. Kundu, P.; Kaur, A.; Mehta, S.K.; Kansal, S.K. Removal of ofloxacin from aqueous phase using Ni-doped TiO2 nanoparticles under solar irradiation. J. Nanosci. Nanotechnol. 2014, 14, 6991–6995. [Google Scholar] [CrossRef]
  157. Nguyen, T.L.; Pham, T.H.; Viet, N.M.; Thang, P.Q.; Rajagopal, R.; Sathya, R.; Jung, S.H.; Kim, T. Improved photodegradation of antibiotics pollutants in wastewaters by advanced oxidation process based on Ni-doped TiO2. Chemosphere 2022, 302, 134837. [Google Scholar] [CrossRef]
  158. Shokri, M.; Jodat, A.; Modirshahla, N.; Behnajady, M.A. Photocatalytic degradation of chloramphenicol in an aqueous suspension of silver-doped TiO2 nanoparticles. Environ. Technol. 2013, 34, 1161–1166. [Google Scholar] [CrossRef]
  159. Malakootian, M.; Olama, N.; Malakootian, M.; Nasiri, A. Photocatalytic degradation of metronidazole from aquatic solution by TiO2-doped Fe3+ nano-photocatalyst. Int. J. Environ. Sci. Technol. 2019, 16, 4275–4284. [Google Scholar] [CrossRef]
  160. Surenjan, A.; Sambandam, B.; Pradeep, T.; Philip, L. Synthesis, characterization and performance of visible light active C-TiO2 for pharmaceutical photodegradation. J. Environ. Chem. Eng. 2017, 5, 757–767. [Google Scholar] [CrossRef]
  161. Ntelane, T.S.; Feleni, U.; Mthombeni, N.H.; Kuvarega, A.T. Sulfate radical-based advanced oxidation process (SR-AOP) on titania supported mesoporous dendritic silica (TiO2/MDS) for the degradation of carbamazepine and other water pollutants. Colloids Surf. A Physicochem. Eng. Asp. 2022, 655, 130276. [Google Scholar] [CrossRef]
  162. Abdullah, M.; Iqbal, J.; Rehman, M.S.U.; Khalid, U.; Mateen, F.; Arshad, S.N.; Al-Sehemi, A.G.; Algarni, H.; Al-Hartomy, O.A.; Fazal, T. Removal of ceftriaxone sodium antibiotic from pharmaceutical wastewater using an activated carbon based TiO2 composite: Adsorption and photocatalytic degradation evaluation. Chemosphere 2023, 317, 137834. [Google Scholar] [CrossRef]
  163. Czech, B.; Tyszczuk-Rotko, K. Visible-light-driven photocatalytic removal of acetaminophen from water using a novel MWCNT-TiO2-SiO2 photocatalysts. Sep. Purif. Technol. 2018, 206, 343–355. [Google Scholar] [CrossRef]
  164. Ali, H.; Yasir, M.; Ngwabebhoh, F.A.; Sopik, T.; Zandraa, O.; Sevcik, J.; Masar, M.; Machovsky, M.; Kuritka, I. Boosting photocatalytic degradation of estrone hormone by silica-supported g-C3N4/WO3 using response surface methodology coupled with Box-Behnken design. J. Photochem. Photobiol. A Chem. 2023, 441, 114733. [Google Scholar] [CrossRef]
  165. Janssens, R.; Cristovao, M.B.; Bronze, M.R.; Crespo, J.G.; Pereira, V.J.; Luis, P. Coupling of nanofiltration and UV, UV/TiO2 and UV/H2O2 processes for the removal of anti-cancer drugs from real secondary wastewater effluent. J. Environ. Chem. Eng. 2019, 7, 103351. [Google Scholar] [CrossRef]
  166. Aram, M.; Farhadian, M.; Nazar, A.R.S.; Tangestaninejad, S.; Eskandari, P.; Byong-Hun Jeon, B.-H. Metronidazole and Cephalexin degradation by using of Urea/TiO2/ZnFe2O4/Clinoptiloite catalyst under visible-light irradiation and ozone injection. J. Mol. Liq. 2020, 304, 112764. [Google Scholar] [CrossRef]
  167. Zainurin, S.N.; Ismail, W.Z.W.; Mahamud, S.N.I.; Ismail, I.; Jamaludin, J.; Ariffin, K.N.Z.; Wan Ahmad Kamil, W.M.W.A. Advancements in monitoring water quality based on various sensing methods: A systematic review. Int. J. Environ. Res. Public Health 2022, 19, 14080. [Google Scholar] [CrossRef] [PubMed]
  168. Cibati, A.; Gonzalez-Olmos, R.; Rodriguez-Mozaz, S.; Buttiglieri, G. Unravelling the performance of UV/H2O2 on the removal of pharmaceuticals in real industrial, hospital, grey and urban wastewaters. Chemosphere 2022, 290, 133315. [Google Scholar] [CrossRef] [PubMed]
  169. Lumbaque, E.C.; Cardoso, R.M.; de Araújo Gomes, A.; Malato, S.; Sánchez Pérez, J.A.; Sirtori, C. Removal of pharmaceuticals in hospital wastewater by solar photo-Fenton with Fe3+-EDDS using a pilot raceway pond reactor: Transformation products and in silico toxicity assessment. Microchem. J. 2021, 164, 106014. [Google Scholar] [CrossRef]
  170. Pan, J.; Niu, X.-Z.; Yang, H.; Zheng, X.; Guan, B.; Wang, H. Pilot test of Mn-Fe/Al2O3 catalytic ozonation in tertiary wastewater treatment. J. Environ. Chem. Eng. 2024, 12, 111937. [Google Scholar] [CrossRef]
  171. Majumder, A.; Otter, P.; Röher, D.; Bhatnagar, A.; Khalil, N.; Gupta, A.K.; Bresciani, R.; Arias, C.A. Combination of advanced biological systems and photocatalysis for the treatment of real hospital wastewater spiked with carbamazepine: A pilot-scale study. J. Environ. Manag. 2024, 351, 119672. [Google Scholar] [CrossRef] [PubMed]
  172. Zheng, J.; Zhang, P.; Li, X.; Ge, L.; Niu, J. Insight into typical photo-assisted AOPs for the degradation of antibiotic micropollutants: Mechanisms and research gaps. Chemosphere 2023, 343, 140211. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of pharmaceuticals commonly found in wastewater.
Figure 1. Chemical structures of pharmaceuticals commonly found in wastewater.
Catalysts 14 00189 g001
Figure 2. The reduction potential and band edge position of some common semiconductors.
Figure 2. The reduction potential and band edge position of some common semiconductors.
Catalysts 14 00189 g002
Table 1. Pharmaceutical compounds commonly found in wastewater, their chemical formula, pKa, and therapeutic applications.
Table 1. Pharmaceutical compounds commonly found in wastewater, their chemical formula, pKa, and therapeutic applications.
PharmaceuticalChemical FormulapKaTherapeutic Applications
Acetaminophen
(Paracetamol)
C8H9NO29.4Analgesic and antipyretic pharmaceuticals widely used to treat pain and fever.
Aspirin
(Acetylsalicylic acid)
C9H8O43.5An NSAID to reduce pain, fever, and/or inflammation.
AmantadineC10H17N10.5An antidyskinetic medicine to treat dyskinesia and Parkinson’s disease.
AmpicillinC16H19N3O4S2.5An antibiotic to treat acute otitis media caused by susceptible organisms.
AmoxicillinC16H19N3O5S3.2, 11.7An antibiotic widely used to treat bacterial infections, including chest infections, such as pneumonia and odontogenic abscesses
AzithromycinC38H72N2O12 8.7An antibiotic prescribed to children for the treatment of acute otitis media caused by Haemophilus influenzae, Moraxella catarrhalis, or Streptococcus pneumoniae.
BupropionC13H18ClNO8.2An antidepressant medication used for treating conditions like depression and mental disorders and aiding in smoking cessation.
CarbamazepineC15H12N2O13.9An anticonvulsant or anti-epileptic drug for the treatment of nerve pain and seizures.
CefotaximeC16H17N5O7S23.75A β-lactam antibiotic used to treat gram-positive, gram-negative, and anaerobic bacteria.
CephalexinC16H17N3O4S4.5A β-lactam antibiotic used to treat bacterial infections caused by bacteria such as pneumonia.
ChloramphenicolC11H12Cl2N1.14An antibiotic useful for the treatment of severe bacterial infections.
CeftriaxoneC18H18N8O7S33.0A cephalosporin antibiotic used to treat various bacterial infections.
CetirizineC21H27Cl3N2O33.6, 7.6A non-drowsy antihistamine used to relieve allergy symptoms.
CiprofloxacinC17H18FN3O36.1, 8.7A fluoroquinolone antibiotic used to treat various bacterial infections.
ClarithromycineC38H69NO138.5An antibiotic used to treat skin problems and chest infections.
CloxacillinC19H18ClN3O5S2.8A penicillin-type antibiotic to treat a wide range of bacterial infections.
DiclofenacC14H11Cl2NO24.15An NSAID used to treat pain and inflammation, such as gout and arthritis.
ErythromycinC37H67NO138.88An antibiotic used to treat a variety of bacterial infections, such as skin infections and respiratory tract infections.
FenofibrateC20H21ClO44.0A fibrate class of medication used to treat abnormal blood lipid levels.
HydrochlorothiazideC7H8ClN3O4S27.9, 9.2A diuretic drug used to treat high blood pressure, edema, and swelling due to fluid build-up.
IbuprofenC13H18O24.85An NSAID used to relieve fever, pain, and inflammation.
IsoniazidC6H7N3O1.82An antibiotic used in the treatment of latent mycobacterium tuberculosis infection.
KetoprofenC16H14O33.88An NSAID used to treat inflammation, pain, swelling, stiffness, rheumatoid arthritis, and osteoarthritis.
LevofloxacinC18H20FN3O45.7, 7.9A fluoroquinolone antibiotic used to treat accute sinusitis and pneumonia bacterial infections.
MemantineC12H21N10.7An antiparkinson agent used to suppress memory loss, dementia, and Alzheimer’s disease.
MetoprololC12H21N 13.9A β-blocker and antihypertensive medication used to treat high blood pressure, fast heart rate, and chest pain.
MetamizoleC13H17N3O4S1.4An antipyretic, analgesic, painkiller drug used to relieve severe and persistent fever and pain.
MetronidazoleC6H9N3O32.38An antibiotic and antiprotozoal medication used to treat bacterial infections and inflammatory diseases.
NabumetoneC15H16O24.8An NSAID used to treat mild to moderate pain and help to relieve symptoms of arthritis and reduce pain.
NaproxenC14H14O34.15An NSAID used to treat pain, fever, rheumatoid arthritis, and inflammatory diseases.
NorfloxacinC16H18FN3O36.34, 8.75A fluoroquinolone antibiotic used in the treatment of a variety of bacterial infections.
OflaxacinC18H20FN3O45.45A quinolone antibiotic used for the treatment of various bacterial infections.
OxytetracyclineC22H24N2O93.22, 7.46,
8.94
A tetracycline class of antibiotic used to treat various infectious diseases.
PromethazineC17H20N2S9.05Antihistamine used to relieve allergy symptoms.
RanitidineC13H22N4O3S8.4Antihistamine used to decrease acid produced in the stomach.
RibavirinC8H12N4O512.25An antiviral drug for the treatment of hepatitis C and respiratory viral infections.
RifampicinC43H58N4O121.7, 7.9An antibiotic used to treat a variety of mycobacterial and gram-positive bacterial infections.
RimantadineC12H21N10.4An antiviral drug used to prevent and treat respiratory tract infections caused by influenza A virus.
SimvastatinC25H38O54.7Hydroxymethylglutaryl-CoA reductase inhibitors used to treat high cholesterol and reduce the risk of heart disease.
StreptomycinC21H39N7O127.4, 13.5An antibiotic isolated from Streptomyces griseus used to inhibit gram-positive and gram-negative bacteria and that is useful to treat cavitary lung disease.
SulfamethoxazoleC10H11N3O3S3.9An antibiotic used to treat bacterial infections and is useful against gram-positive and gram-negative bacteria.
TetracyclineC22H24N2O83.3, 7.8,
9.6
An antibiotic used to treat a variety of infections, acne, brucellosis, cholera, malaria, and plague.
TramadolC16H25NO29.2An opioid painkiller to treat moderately severe pain.
TrimethoprimC14H18N4O37.6, 8.3An antibiotic used to treat and prevent different types of infections.
ValsartanC24H29N5O33.9, 4.7An angiotensin receptor blocker used to treat high blood pressure, heart failure, and diabetic kidney disease.
Table 2. Reported studies using AOPs and combination methods for real wastewater treatment and evidence of their effectiveness.
Table 2. Reported studies using AOPs and combination methods for real wastewater treatment and evidence of their effectiveness.
AOP MethodWastewaterLight SourceCatalystResultsRef.
UV-H2O2Northeast of Spain: two from urban wastewater plants. One from graywater, a psychiatric hospital, and the pharmaceutical industry.
The wastewater was composed of 30 pharmaceuticals and 13 transformation products.
UV-C lamp
(30 min)
NoneThe pharmaceuticals were approximately 6–86% removed from urban wastewater plants, and 59% from graywater, with electrical energy per order (EEO) values of 0.9–1.5 kWh/(m3·order).
The removals were lower for hospitals and industries (36% and 17%), with electrical energy per order (EEO) values of 7.3–9.1 kWh/(m3·order).
[168]
Photo-FentonPorto Alegre City, local hospital: paracetamol (297 μg/L) and dipyrone (55 μg/L).Solar irradiation
(30 min)
Fe3+-EDDS
(1:2 ratio)
77% pharmaceutical removal.
Most of the transformation products show low toxicity, mutagenicity, and bioaccumulation.
[169]
Catalytic
ozonation
Zhejiang Province, China, industrial treatment plant: 8% pharmaceutical waste.NoneMn-Fe/Al2O399% removal of DOC and 14.5 mg/L average removal of COD.
The effluent has a low total transcriptional effect level index value of ≤1.50.
[170]
Biological-photocatalysisKharagpur Subdivisional Hospital, India: containing carbamazepine.UV-A irradiationAl–ZnO/Fe85% carbamazepine removal from aerated horizontal flow-constructed wetland.
The biological method removed 30% carbamazepine.
[171]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Roslan, N.N.; Lau, H.L.H.; Suhaimi, N.A.A.; Shahri, N.N.M.; Verinda, S.B.; Nur, M.; Lim, J.-W.; Usman, A. Recent Advances in Advanced Oxidation Processes for Degrading Pharmaceuticals in Wastewater—A Review. Catalysts 2024, 14, 189. https://doi.org/10.3390/catal14030189

AMA Style

Roslan NN, Lau HLH, Suhaimi NAA, Shahri NNM, Verinda SB, Nur M, Lim J-W, Usman A. Recent Advances in Advanced Oxidation Processes for Degrading Pharmaceuticals in Wastewater—A Review. Catalysts. 2024; 14(3):189. https://doi.org/10.3390/catal14030189

Chicago/Turabian Style

Roslan, Nur Nabaahah, Harry Lik Hock Lau, Nurul Amanina A. Suhaimi, Nurulizzatul Ningsheh M. Shahri, Sera Budi Verinda, Muhammad Nur, Jun-Wei Lim, and Anwar Usman. 2024. "Recent Advances in Advanced Oxidation Processes for Degrading Pharmaceuticals in Wastewater—A Review" Catalysts 14, no. 3: 189. https://doi.org/10.3390/catal14030189

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