Skip to main content

Photodegradation of six selected antipsychiatric drugs; carbamazepine, sertraline, amisulpride, amitriptyline, diazepam, and alprazolam in environment: efficiency, pathway, and mechanism—a review

Abstract

Psychiatric drugs do not vanish after being carried to wastewater treatment plants by the urine or feces of patients and, a variable portion of their dose and also unused or expired drugs are lost to the environment. This is because the technology of plants is not intended to eradicate pharmaceuticals and their metabolites. Above all, psychotropics can change population dynamics and behavior at lower doses. We believe that antipsychotics have not gotten enough attention when it comes to drug pollution and that their importance as environmental pollutants has been underestimated. An innovative approach to eliminating pharmaceutical pollutants from water is the application of advanced oxidation methods. Among these oxidation methods are photocatalysis, ozonation, UV/hydrogen peroxide oxidation, and photo-Fenton oxidation. Photocatalytic degradation of pharmaceuticals is now the most widely used method since it is affordable and ecologically beneficial due to the reusable nature of the photocatalyst. When light is absorbed during photocatalytic degradation, electrons in the valence band (VB) get excited and migrate into the conduction band (CB). Consequently, hydroxyl radicals (OH) are produced by VB’s holes carrying out oxidation processes on photocatalyst surfaces. The charge difference between the two bands encourages reduction reactions by CB electrons at the surface. To perform successfully, a photocatalyst has to have enough surface-active sites, a favorable band edge location, modest bandgap energy, increased charge separation, and charge transfer. Due to the above-mentioned concerns, the investigation and analysis of the photocatalytic degradation of six psychiatric drugs—carbamazepine, sertraline, amisulpride, amitriptyline, diazepam, and alprazolam—are the main objectives of this review.

1 Introduction

The presence of contaminants in environmental samples is widespread. Persistent organic pollutants (POPs) are one type of pollutant found in wastewater and have long-term negative consequences on ecosystems, human health, and water quality [1, 2]. Many POPs have currently been found in surface water, groundwater, and urban wastewater treatment plants (WWTPs) [3,4,5]. Antibiotics, analgesics, steroids, antidepressants, antipyretics, perfumes, and cosmetics make up the majority of POPs [1, 6]. The World Health Organization (WHO) has reported that they are the primary causes of the current global health crisis [7]. Additionally, the growth in POPs could cost society and the government [7, 8].

The manufacturing of synthetic chemicals has increased, particularly in emerging nations, as a result of the population’s rapid development and the expansion of agricultural and industrial activities. This has seriously contaminated the world’s natural streams. Personal care items and medications are some of the synthetic organic compounds. Even at low concentrations of ng L− 1, the numerous organic contaminants in surface waters remain in aquatic ecosystems and endanger aquatic animals when exposed for an extended period [9,10,11].

Pharmaceuticals are necessary for better public health, but their effects on the environment are extremely harmful and require rapid action. Additionally, some medications’ metabolites are more hazardous than the original drugs. Unfortunately, these medications are now commonplace in the aquatic environment as a result of their extensive usage and constant discharge. This is due to the ineffective removal of pharmaceuticals from various sources, including home sources, hospital effluents, agricultural, and pharmaceutical sectors, by traditional WWTPs [12]. The majority of industrial facilities discharge their effluents directly into the sewage system without sufficient processing or treatment [13]. This puts a strain on water ecosystems all across the world [14]. Different chemical stability and solubility in water are essential characteristics to measure the extent of contamination in pharmaceutical substances. Even though they are present in very small amounts, they are likely to bioaccumulate and reach the food chain. Some of them are tenacious and gradually build up in water bodies.

Different environmental conditions, such as heat intensity, temperature, humidity, etc., can have a significant impact on a drug’s stability [15]. Additionally, there is not much research on how different drugs interact with one another. These organic and inorganic compounds contaminate the water and render it unsafe for drinking and other uses, which could have long-term negative impacts on one’s health [16]. Toxic compounds disrupt aquatic life, creating structural changes in fish internal organs (such as kidneys and intestines) that influence fish reproductive and growth patterns [17]. Diverse conventional techniques, including adsorption, filtration, screening, sedimentation, coagulation, and flocculation, are used to reduce the water pollution; however, these techniques are unsuitable for the removal of pharmaceutical chemicals due to the nature of these POPs in water [18, 19]. The biological wastewater treatment methods that are being presented are as follows: (1) biological processes in WWTPs, which comprises anaerobic treatment (anaerobic bioreactors, anaerobic lagoons), aerobic treatment (oxidation ponds, aeration lagoons, aerobic bioreactors, activated sludge, percolating or trickling filters, biological filters, rotating biological contactors, biological removal of nutrients); (2) phytoremediation of wastewater, which comprises constructed wetlands, rhizofiltration, rhizodegradation, phytoaccumulation, phytotransformation, and hyperaccumulators; and (3) mycoremediation of wastewater [20].

Additionally, due to the formation of novel degradation products, straightforward biological treatments such as activated sludge treatment and microbe-mediated biodegradation are not considerably successful at breaking down these prescription medications [21]. If plants and animals consume contaminated water and it enters their metabolic processes, the water will become dangerous to them [22]. A few main causes of water contamination include the agricultural sector, illegal waste disposal, leaks from landfills, and the discharge of sewage and industrial effluents [23]. Long-term increases in the concentration of pollutants are inevitable if sewage is released daily, even though the majority of industrial enterprises have their WWTPs. River water is the primary supply for drinking water treatment plants (DWTPs), hence this phenomenon will provide a significant challenge in the future. The traditional treatment methods including screening, coagulation/flocculation, and chlorination/fluoridation are used by the majority of DWTPs worldwide; only some processes such as ultrafiltration, nanofiltration, and adsorbents (e.g., such as charcoal) to remove impurities [24,25,26]. Pollutants including large concentrations of organic carbon, nitrogen compounds, and heavy metals cannot be adequately treated by these treatment methods [22]. Typically, combined overflow causes problems in WWTPs.

The concentration and effects of pharmaceutical chemicals, including psychiatric medications, in the aquatic environment have been evaluated throughout the past few decades [27]. Advanced oxidation processes (AOPs) used in pilot-size WWTPs, including ozone or UV radiation, are not thought to be feasible at the industrial scale because of their high running costs and tendency to produce hazardous byproducts. In contrast, the most persistent active pharmaceutical ingredients (APIs) can be eliminated using photocatalytic AOPs by simply activating a reusable photocatalyst with light. Photocatalysis, is characterized by a straightforward instrumental process, an easy-to-control process, non-selective oxidation, being economical, and causing a complete breakdown of pharmaceutical pollutants [28]. According to the photocatalysts’ band energy gap, the breakdown of pollutants by photocatalysts happens by the electron-hole pairs produced by either visible or UV irradiation [29,30,31].

Regarding the treatment of POPs, various methods are cumbersome to use. Therefore, removing POPs in low quantities is extremely challenging and urgent [32]. It has to do with the production of reactive oxygen species (ROS), which are used as photocatalysts in semiconductors. If energy is larger than or equal to the energy gap of semiconductors, photo-excited holes (h+) and electrons (e) are created under irradiation in the VB and the CB, respectively [33]. When they move to the semiconductor surface, they may produce radicals (OH, O•2−, etc.), which could trigger the redox process [8, 34].

So far, many review articles have studied the environmental pollution of antipsychiatric drugs, but these articles have usually either studied a specific category of these drugs or studied psychiatric drugs in general without referring to a specific class of them. Also, the effect of photocatalytic degradation on this category of drugs has not been specifically investigated. There was a need to investigate the environmental pollution of antipsychiatric drugs in a categorized way and to review the studies that have been conducted on them using photodegradation, as well as to study the effect of photodegradation on the pollution of these drugs. In 2021, Escudero et al. investigated antipsychotic drugs as environmental pollutants [35]. In this study, they have evaluated the environmental risks of antipsychotic drugs. Argaluza et al. have also studied the effects of psychiatric drugs on the environment in another review article [36]. They reviewed the data of previous articles about the occurrence, environmental effects, and toxicity of these drugs on non-target organisms. In a review article, Cunha et al. discussed the environmental risk assessment of 7 psychoactive drugs in the aquatic environment [37]. In another review article, studies related to the occurrence and fate of neuropsychiatric drugs and illegal drugs in sewage treatment plants (STPs) have been reviewed by Asimakopoulos and Kannan [38]. We will concentrate on the possible impact of antipsychotic medications as environmental pollutants in this review. We think that this type of therapy is unjustly undervalued. The goal of this paper is to give a thorough understanding of photocatalytic degradation and photocatalytic applications for the breakdown of contaminants from psychiatric medicines. The utilization of photocatalytic methods in the photodegradation of antipsychiatric drugs will be reported with some examples from literature. In the current study, photodegradation of six different psychiatric medications—carbamazepine (CBZ), sertraline (SER), amisulpride, amitriptyline (AMI), diazepam (DZP), and alprazolam—is the main emphasis.

2 Environmental pollution of drugs

Chemicals that create toxic issues, such as pesticides, heavy metals, polychlorinated biphenyls, aromatic compounds, etc., are well-known. But worries about the so-called “pollutants of emerging interest” are growing, with pharmaceuticals being the biggest source of worry. Two crucial qualities of pharmaceutical items are what motivates this preoccupation: First of all, at low quantities, like those present in the environment, they have pharmacological effects. Second, they are more likely to interact with their target molecules since they are created with stability in mind. Since the end of the 1990s, there has been an exponential increase in scientific journal articles discussing the presence of pharmaceutical products in the environment. This is because advances in analytical techniques have made it possible to identify pharmaceuticals at lower quantities in various matrices [39]. The amount of knowledge currently accessible on drugs’ effects on the environment is just overwhelming. According to a recent study [36], over 4000 distinct pharmacologically active compounds, including prescription medicines for people and animals as well as over-the-counter medications, are currently being provided on a global scale. According to estimates, 4.5 trillion doses of drugs will be consumed worldwide in 2020 [36]. The following factors [36] indicate that the trend will probably continue: Populations are getting older and living longer, economies are expanding, especially in emerging economies, which raises the capacity and expectations for treating aging and chronic diseases, livestock and aquaculture practices are getting more intensive to meet demand, new pharmaceutical products are being developed, and climate change will worsen already-existing diseases (both communicable and non-communicable) [36].

Studies of pharmaceutical residues in the environment have been possible thanks to the rapid development of mass spectrometric analytical technologies during the past 25 years. Early reports [40,41,42,43] of APIs in the environment coming from factory discharges were sporadic and garnered little attention at the time.

When estrogens were identified in sewage effluents as the cause of fish feminization in the late 1990s, interest in pharmaceuticals in the environment, particularly the role of excreted medications, skyrocketed. Following that, ecotoxicologists focused mostly on sewage effluents and receiving rivers, which may have diverted attention from potential alternate sources of pharmaceutical residues [44].

The pharmaceutical industry also stated that it is unlikely that considerable amounts of APIs will leak out during manufacture using a variety of justifications, including the fact that pharmaceuticals have such a high value that their release would be prevented for simply financial considerations [45]. This presumption was eventually shown to be false. We learned that exposure routes are not always predictable from the identification of diclofenac residues in cow corpses as the reason for the vulture population collapse in India and Pakistan [46].

The first of several articles demonstrating extremely high pharmaceutical emissions from medication manufacturers in Patancheru, close to Hyderabad, India, was published in 2007 [47,48,49]. This region, which has a huge concentration of industries in a small space, is a key hub for the global manufacture of bulk pharmaceuticals. For certain medications, the quantities in the effluent from a treatment facility that received wastewater from around 90 production units were higher than those discovered in the blood of patients taking medication. The development of resistance is worrisome due to the elevated concentrations of multiple broad-spectrum antibiotics. The most common medication, ciprofloxacin, has concentrations up to 31,000 µg L− 1, which are more than 1000 times more than what some bacteria can withstand [47]. The anticipated total discharge of ciprofloxacin was 44 kg d− 1, which is equal to Sweden’s intake over five days or, put another way, enough to treat every person in a city of 44, 000 people. These discharges have caused previously unheard-of levels of contamination in river sediment [48], surface, ground, and drinking water [49], and a recent report also showed contamination of irrigated soils [50]. The research received a lot of media attention, which boosted interest in this exposure route among scientists and the public [44, 51].

3 Environmental pollution of antipsychiatric drugs

About 700 different drug residues have been found in several environmental compartments, including wastewater, surface and groundwater, soil, air, biota, and even the tap water [45]. There are comprehensive books that examine the use of psychiatric medications in the environment [52]. Due to its ubiquitous nature, CBZ has even been proposed as a marker for wastewater-influenced aquatic bodies [53].

Additionally, parent drugs (or metabolites) continue to change and go through intricate metabolic processes once they are in the environment due to various organisms and physicochemical mechanisms (e.g., photo-degradation, adsorption to solids, etc.). This results in “transformation compounds.” Trawinski and Skibinski [54], for instance, studied the photodegradation of psychiatric medications. Some compounds are particularly persistent because they can linger unaltered in the environment for decades. In this instance, oxazepam, which has been present unaltered in the bottom of Swedish lakes for more than 30 years [55], is a typical case.

3.1 Antipsychiatric drugs pollution’s effects on human health and ecotoxicology

The impact of drug contamination on human health has not been extensively researched. Drug concentrations in tap water should not be harmful, according to a WHO assessment in 2012 [36]. A subsequent study conducted in China [56] verified these results. However, the most susceptible patient populations, such as allergic [36] patients, may experience issues due to the presence of medications in the environment. Even while there is no proof that short-term consequences on human health exist, there are still some unknowns, especially regarding long-term exposure (chronic exposure) to a combination of pollutants [36]. The main exposure pathways are drinking water, and eating fruits, vegetables, meat, fish, shellfish, and dairy products [36]. The rise in bacteria resistant to antibiotics, which is currently acknowledged as the major public health issue in the world, is likely the best-known illustration of the harmful impact of drug pollution on human health. In this regard, we believe that the “One Health” approach or concept, which holds that human health and environmental health are inextricably linked, is crucial [57]. However, we think it is important to apply this strategy to other treatment classes, not just antibiotics, such as psychiatric medications [58, 59]. Since psychoactive drugs can bypass maternal barriers that affect the development of the embryonic brain, some authors have hypothesized that they may be linked to human neuropsychiatric disorders like autism, Alzheimer’s disease, and schizophrenia [60].

Drugs can have toxic effects on the environment that go much beyond those on growth, death, or reproduction. For instance, psychoactive substances can change population dynamics by influencing the behavior and fitness of organisms [61,62,63,64]. The physiological systems that medications affect as well as their therapeutic goals are not unique to humans. Numerous living creatures share many of these phylogenetically highly conserved features and signaling pathways [65]. For instance, numerous behavioral test for investigational medications intended for human use are carried out on fish, including tests for anxiety, fear, and stress [66]. Fish and humans have many the same neurotransmitter and signaling pathways. In reality, both vertebrates and invertebrates, such as amphibians, fish, insects, and echinoderms, contain biogenic monoamines (such as serotonin, dopamine, norepinephrine, etc.) [67, 68]. From an evolutionary perspective, some compounds are so old that they can be found in animals that do not belong to the animal world. Acetylcholine, for instance, can be found in bacteria and fungi [69] and serotonin in plants [70]. Crickets’ behavior has been demonstrated to vary in response to fluoxetine [71]. The antipsychotic clozapine also causes constipation in fish [72], plants collect benzodiazepines that may influence their GABAergic system [73], and SER alters the nitrification processes by changing the microbial trophic chain [36, 74].

3.2 Antipsychiatric drug toxicity for creatures that are not the intended targets

Like other pharmacological categories, psychiatric drugs can be found in the environment. It is important to keep in mind that these substances do not occur isolated but rather as complex combinations, even though these concentrations are below the range projected to harm humans and cause acute or even chronic toxicity to non-target animals [75]. Psychiatric medications are one of the most relevant groups when it comes to the evaluation of ecotoxicological effects in terrestrial and aquatic non-target organisms because of their intrinsic biological activity, which could disrupt nervous and endocrine systems [76]. Additionally, the increasing number of research on chronic toxicity in aquatic creatures that are not the intended targets highlighted the need for a unique strategy to be developed to more clearly address this issue [77]. Considering how frequently medications are exposed to aquatic creatures, it would be more crucial to comprehend life cycle toxicity than to carry out acute toxicity experiments [78,79,80].

4 Photodegradation of drugs

Pharmaceuticals are compounds or combinations of substances used to cure or prevent diseases in humans. They do, however, represent a significant group of point or diffuse pollution-related micropollutants. Conventional wastewater treatment methods have only been shown to partially remove pharmaceutical contaminants, and as a result, municipal and/or hospital WWTPs have become important sources of their release into the environment. Pharmaceutical substances are present in aquatic media in quantities between ng L− 1 and µg L− 1. Pharmaceuticals have been demonstrated to have harmful impacts on both human health and ecology (loss of biodiversity, extinction of aquatic creatures, etc.) even at low concentrations [81,82,83,84,85]. Innovative technologies, such as AOPs, have been developed for the effective removal and degradation of pharmaceutical compounds from wastewater and the prevention of their release to ground and surface waters. AOPs have been proposed as a tertiary treatment in WWTP effluent. AOPs are thought to be promising techniques for the treatment of contaminated wastewater including non-biodegradable organic contaminants since they comprise various mechanisms of producing ROS, primarily OH. Complete elimination of numerous pharmaceutical contaminants is possible using heterogeneous photocatalysis, which is based on the effective generation of ROS in aqueous media [86,87,88,89].

Novel approaches to treating pharmaceutically rich sewage influent have been developed to increase removal efficiency. AOPs [90], UV irradiation alone [91, 92], UV irradiation and hydrogen peroxide combined to produce OH as an efficient oxidizing agent [93,94,95,96], and combining hydrogen peroxide with Fe2+ ions (photo-Fenton) to increase the efficiency of OH radical generation [97,98,99,100,101] are some examples of these treatment methods. More recently, OH radicals have been produced using heterogeneous photocatalysts like suspended titanium dioxide (TiO2) [102]. TiO2’s physicochemical characteristics as a semiconductor determine how effective it is as a photocatalyst [103, 104].

4.1 Drug photodegradation by photocatalysts

Another new method that can be used to remove pharmaceutical contaminants from water is the use of sophisticated oxidation techniques for drug breakdown. Photocatalysis, ozonation, UV/hydrogen peroxide oxidation, and photo-Fenton oxidation are some of these oxidation mechanisms. Non-thermal plasma and sonolysis are two further cutting-edge methods for drug breakdown. The first method takes advantage of oxidizing agents produced by non-thermal plasma, including H2O2, OH, and ozone [105,106,107,108]. The oxidizing agents in sonolysis are created by ultrasonic radiation. Due to the reuse of the photocatalyst, which makes it economical and environmentally friendly, photocatalytic degradation of medicines is currently the most common approach [109]. Nano-based photocatalysts are easy to make and can effectively break down pollutants. ZnO, TiO2, and CeO2 nanoparticles (NPs) are among the most popular and efficient nanophotocatalysts used in this context [110]. In photocatalytic degradation, light is absorbed, which causes electrons in the VB to be excited and move into the CB. As a result, VB develops holes that carry out oxidation processes on photocatalyst surfaces, generating OH. CB electrons are responsible for carrying out reduction processes at the surface by the charge separation between the two bands. A photocatalyst must have adequate surface-active sites, a good band edge position, a small bandgap energy, a low rate of charge recombination, enhanced charge separation, and charge transfer to function well. Additionally, at the nanoscale level, NPs’ optical and electrical properties are impacted by a high surface-to-volume ratio and high atom concentration at the surface. As a result, when a particle gets smaller and smaller, the surface atoms become less strongly bonded to the bulk atoms and more chemically active. This enables them to participate in chemical reactions and act as catalysts because of the high surface energy that results from this. A quantum size effect results from these modifications, which are related to the material’s electrical structure. Their optical characteristics are likewise impacted by this [111]. Bandgap is a crucial characteristic that represents the conductance and photoactivity of the nanomaterial. With a bandgap energy of 3.4 eV, TiO2 NPs are the most studied photocatalysts, whereas ZnO NPs, which have a smaller bandgap of 3.2 eV, exhibit photocorrosion and a high rate of photoexcitation recombination. The bandgap energy of CeO2 NPs is 3.1 eV. Therefore, the bandgap must be reduced to absorb visible light. By mixing NPs with noble metals and doping, this problem can be resolved. To increase photoactivity, TiO2/ZnO and CuO/ZnO nanocomposites both possess synergistic qualities [26, 112].

A wide range of photocatalysts have been created over the past ten years. Among them, graphitic carbon nitride (g-C3N4) combines several benefits that make it suitable for a variety of photocatalytic applications [113, 114]. These include a mild band gap (2.7 eV), a good response to visible light (up to 460 nm), a simple preparation method, high chemical stability and non-toxicity, and low cost. However, there are not many published research studies that deal with hospital wastewater. In genuine wastewater effluent spiked with g-C3N4 and taken from a full-scale WWTP in Brisbane, Queensland, Australia, Zhong et al. [115] examined the photocatalytic degradation of CBZ under simulated solar irradiation. They found that 100% destruction of CBZ had occurred after 15 min. Furthermore, Kane et al. [116] investigated the photodegradation of CBZ using g-C3N4 when it was exposed to UV light. Using a composite g-C3N4-based photocatalytic material called Ag/TiO2/M-g-C3N4, Gao et al. [117] investigated the removal efficiency of amoxicillin in real hospital wastewater and in a WWTP. In the two sample types, a clearance of amoxicillin of 71 and 49%, respectively, was seen after 60 min of the photocatalytic process [89].

4.2 Psychiatric drugs’ photodegradation

After being transported to WWTPs by patients’ urine or feces, psychoactive drugs do not disappear. However, a variable portion of the dose is lost to the environment, where pharmaceuticals and their metabolites have been found in virtually every matrix, including hospital effluent, municipal wastewater, rivers, lakes, and even drinking water. This is because the technology of these plants is not intended to eradicate pharmaceuticals and their metabolites. We believe that the environmental quantities that have been recorded may be high enough to noticeably affect aquatic species. Antipsychotics and other medications may also be bioaccumulating through web food. Most importantly, psychotropics have the potential to alter behavior and, at lower concentrations, population dynamics. In terms of drug pollution, antipsychotics, in our opinion, have not yet received adequate attention, and their relevance as environmental pollutants has been undervalued [35].

Due to this, the investigation and analysis of the photocatalytic degradation of six different psychiatric drugs— CBZ, SER, amisulpride, AMI, DZP, and alprazolam—are the main objectives of this review.

5 CBZ

Psychiatric medication CBZ (Table S1 in the Supplementary materials file), which was first sold in 1963, has amassed a sizable market share to date. According to estimates, there were around 1.01 kt consumed annually worldwide and 34.9 t in the US in 2008 [118]. Due to its high consumption and stable chemical composition, CBZ is continuously released into the environment and persists there. CBZ cannot, however, be efficiently removed (10%) by conventional WWTPs [119], since it is a persistent pharmaceutically active compounds (PhACs). However, some investigations even found a higher concentration of CBZ in effluent than in influent [120, 121]. As a result, research has shown that CBZ is one of the most commonly found PhACs in WWTPs and surface water around the world [122,123,124,125,126,127].

5.1 Photocatalytic degradation of CBZ

Due to their high persistence during wastewater treatment operations, the presence of pharmaceutical and personal care products (PPCPs) in the environment is a growing source of concern [128,129,130].

One of the most often found pharmaceuticals among the numerous PPCPs is CBZ, which is used in human medicine to treat seizures. A daily dosage of 100–2000 mg of CBZ is given chronically to patients, leading to significant global production [131]. The removal efficiency of CBZ in typical WWTPs is extremely low because of its biorefractory nature. Approximately 50% of the CBZ and its metabolites discharged in urine and feces are directly released to the municipal WWTPs [132]. Numerous investigations have reported the presence of CBZ in ambient samples at quite high amounts [128, 133, 134].

AOPs were designed to further dispose of wastewater containing CBZ to cover the insufficient treatment from WWTPs to remove CBZ residues and reduce its harmful effects on ecological systems and human bodies. AOPs have been shown to have stronger degrading effects and efficiency when compared to traditional biological processes, according to methods like Fenton oxidation [135], UV irradiation combined with chemical oxidizers or catalysts [136], or ultrasonic irradiation [137]. AOPs have shown to be effective treatments and disposal methods for PhACs like CBZ. Although AOPs offer a variety of benefits, some current shortages should not be overlooked either. For instance, technologies that express a well-disposal impact always call for the employment of harsh chemical additives or high energy input, which can result in high operational costs as well as secondary pollution production [138]. Therefore, it is imperative to find methods for dealing with refractory PhACs that are both ecologically benign and energy-sustainable [139]. The treatment approach, CBZ efficiency, occurrence in the environmental samples, and concentration are presented in Table 1 [140,141,142].

Table 1 The treatment, CBZ’s uses, occurrence in the environmental samples, and concentration

5.1.1 Photocatalytic degradation of CBZ using heterostructures GQD/BiVO4

Solar-driven Graphene quantum dots (GQDs) loaded with BiVO4 heterostructure catalysts were created and used by Wu and colleagues to degrade CBZ under artificial sunlight. GQDs, which are sensitive to visible light, were made and loaded onto BiVO4 heterostructures in this study. The catalytic degradation of the drug CBZ was then carried out using the novel ternary heterojunction catalysts (GQD/m-BiVO4/t-BiVO4) (Fig. 1). Under simulated sun irradiation, the catalyst had outstanding performance and catalytic processes for CBZ [139].

Fig. 1
figure 1

Diagrammatic representation of suggested CBZ degradation routes using GQD/BiVO4 hybrids to replicate solar-induced photocatalysis [139]; TP: transformation products

5.1.2 Photocatalytic degradation of CBZ using BiOCl-10

In a different study, Gao et al. [143] directly manufactured visible-light-driven BiOCl microspheres using a simple one-pot ethylene glycol-mediated solvothermal process. The degradation of CBZ under visible light irradiation (> 420 nm) allows for the examination of the photocatalytic activity of the obtained photocatalysts. Additionally, an active species trapping experiment was used to systematically explore the photocatalytic process. Finally, the use of BiOCl-10 in real wastewater was investigated. The objective of this study was to improve BiOCl’s responsiveness to visible light so that CBZ could be degraded effectively (Fig. 2). Analysis was also done on the relationship between catalyst activity and characterization. Most notably, BiOCl-10 proved effective without any additives in natural water. This research expands the potential uses of visible-light-driven photocatalysts in the eradication of resistant pharmaceutical contaminants [143].

Fig. 2
figure 2

Potential CBZ degradation products and the BiOCl-10 pathway [143]

5.1.3 Photocatalytic degradation of CBZ using magnetic BiOCl/Fe3O4

Chen et al. [144] successfully created a magnetic BiOCl/Fe3O4 catalyst through precipitation with a simulated solar light response. The effectiveness of photodegrading CBZ was used to further study the photocatalytic activity of BiOCl/Fe3O4 nanocomposites. In this investigation, after 60 min of irradiation with the BiOCl/Fe3O4 catalyst, 90.3% of the CBZ was eliminated. Furthermore, by using an external magnetic field, the BiOCl/Fe3O4 photocatalyst could be easily removed from the solution. This kind of magnetic photocatalyst, which is distinguished by ease of separation, might find use in the treatment of wastewater. Liquid chromatography-mass spectroscopy was used to establish how quickly CBZ vanished and how its intermediates formed, and it also suggested potential photocatalytic breakdown pathways. In comparison to prior CBZ degradation investigations, it was noteworthy that some of the intermediates were reported for the first time (Fig. 3). By using an external magnetic field, the BiOCl/Fe3O4 catalyst may be easily recycled from the reaction system. It showed that BiOCl/Fe3O4 might be a promising catalyst for use in actual applications to degrade organic pollutants [144].

Fig. 3
figure 3

Fe3O4/BiOCl-mediated degradation mechanisms of CBZ under simulated solar light radiation [144]

5.1.4 Photocatalytic degradation of CBZ using TiO2

CBZ was photodegraded in 2015 by Carabin et al. [119] using suspended photocatalysts made of TiO2. This study looked at how the photocatalytic activity at 365 nm could reduce the CBZ in a water solution. Using P90 photocatalyst concentration of 1.5 g L− 1 for 90 min of treatment with HOF-5 and light intensity of 18.9 W m− 2, 4% less than the standard condition to be obtained, the best operating parameters for CBZ degradation in solution (99% CBZ elimination) were reached. Low concentrations of CBZ (37 g L− 1) or relatively high concentrations of CBZ (5 to 20 mg L− 1) in water can both be treated using this improved photocatalytic method with the reaction rate of 3.2 µg L− 1 min− 1 at 20 oC [119].

5.1.5 Photocatalytic degradation of CBZ using BiPO4

Xu et al. [134] investigated how specifically formulated BiPO4 catalyzed the breakdown of CBZ. Hydrothermal synthesis was used to create BiPO4. The obtained samples’ physicochemical characteristics were examined, and the findings showed that the hydrothermal temperature and reaction time had an impact on the phase, morphology, and optical characteristics of the BiPO4 catalysts, which may further dictate their unique photocatalytic performances. The catalysts’ crystal properties were somewhat reflected in the intrinsic microstructure and optical characteristics. The BiPO4 made at 180 °C for 72 h (BPO-180-72) demonstrated the best photocatalytic activity when exposed to UV light; CBZ was almost entirely removed from ultrapure water after 60 min of exposure. The monoclinic phase and reasonably organized shape of the resultant BiPO4 were said to have a synergistic effect that led to positive photocatalytic activity. For BiPO4, it was initially demonstrated that the monoclinic phase was more active than the hexagonal phase. After 60 min of exposure, BPO-180-72 eliminated roughly 72.4% of CBZ from laboratory-prepared simulated wastewater, indicating a possible use for this substance in wastewater treatment (Fig. 4). The system’s photodegradation of CBZ was shown to be mostly mediated by photogenerated holes and OH [134].

Fig. 4
figure 4

Suggested mechanism for BPO-180-72 to break down CBZ when exposed to UV rays. The starting dose of the catalyst was 1 g L− 1, and the concentration of CBZ was 5 mg L− 1 [134]

6 SER

In the past ten years, numerous reports of SER (Table S2 in the Supplementary materials file), an antidepressant of the selective serotonin reuptake inhibitors (SSRIs) family, in environmental matrices have been made [145].

One of the pharmaceuticals, SER hydrochloride, is primarily used to treat adult outpatient clinical depression as well as adult and pediatric obsessive-compulsive, panic, and social anxiety disorders [31, 146].

6.1 Photocatalytic degradation of SER

One of the primary contaminants in the aquatic environment is antidepressants. Due to their widespread usage, potential health impacts, and partial removal from WWTPs using traditional techniques, they are the subject of extensive research. As a result, SER is now among the medications that are most frequently found in surface water and WWTP effluents [147]. Traditional treatment processes are ineffective for removing contaminants from wastewater, thus it is important to look for more efficient treatment options [31].

While this SER molecule’s occurrence and toxicology have received the majority of research attention, little is known about how this chemical changes in response to environmental conditions like solar irradiation [31]. The treatment approach, SER efficiency, occurrence in the environmental samples, and concentration are presented in Table 2 [148,149,150].

Table 2 The treatment, SER’s uses, occurrence in the environmental samples, and concentration

6.1.1 Photocatalytic degradation of SER using photolysis in an aqueous environment

Gornik et al. [145] looked into the photodegradation of SER in an aqueous environment. By analyzing SER photodegradation in laboratory-size trials supported by additional tests on actual surface water, the current work attempted to close these gaps. It has been determined that direct photolysis dominates the first-order kinetics of SER degradation, however, the presence of some reactive species, such as OH, CO3, and 3CDOM*, may speed up the process. The anticipated SER half-life of 1.4 d was used to validate the projected results in surface water that had been injected with the drug and exposed to real sunshine. Overall, their research sheds light on how SER is transformed in surface waters and shows that photodegradation is a crucial—and likely the primary—transformation pathway for this pollutant [145].

6.1.2 Photocatalytic degradation of SER using xenon lamp irradiation

In the other study [151], the photodegradation of two of the most commonly used psychiatric medications, DZP and SER, were caused by exposure to xenon lamp radiation, which overlaps the spectrum of sunlight, and to natural. In this study, the photodegradation of the psychiatric medications DZP, and SER was examined using the hyphenated UHPLC-QTOF-MS method, which produced extremely precise and reliable results. To model a process that might occur in the environment, the kinetics of deterioration were examined under two different forms of lighting: (1) xenon lamp light, and (2) natural sunlight (Fig. 5). In all cases, the degradation kinetics were pseudo-first order; however, in the second experiment, autocatalytic processes were seen for a variety of natural water samples. In total, nine photoproducts of DZP and seven altered SER compounds were identified, and their presence in various water samples was examined. However, none of the compounds were found, most likely as a result of the medicines’ adsorption on sediment, sewage sludge, or other environmental elements [151].

Fig. 5
figure 5

SER’s suggested photodegradation pathway [151]

6.1.3 Photocatalytic degradation of SER using P25/Chitosan composites

In addition, Degussa P25/Chitosan composites for the photocatalytic removal of SER and Acid Red 18 from water were investigated in 2021. In this study, P25/chitosan composites were created utilizing three different preparation techniques and two different chitosan sources. On the chitosan matrix, titanium dioxide was equally distributed without creating any unnecessary agglomerates. The photocatalytic activity was shown to be unaffected by the preparation method, although the rate of SER breakdown increased after 240 min when medium molecular weight chitosan was used instead of high molecular weight chitosan. Due to the shielding effect occurring above a specific level of photocatalysts, the amount of applied photocatalysts has a significant impact on how successful the photocatalytic reaction is. SER and Acid Red 18 both required optimal concentrations of 0.3 and 0.6 g L− 1, respectively. Additionally, it has been demonstrated that under identical reaction circumstances, the photocatalytic degradation of SER operates more effectively than the decomposition of Acid Red 18. These chemicals decompose significantly more quickly through photocatalytic breakdown than through photolysis. In comparison to pure Degussa P25, the presence of chitosan in the photocatalytic materials greatly improves the separation of the catalyst from the reaction mixture. These nanocomposites can be used to successfully remove organic contaminants from water by photocatalytic degradation [152].

6.1.4 Photocatalytic degradation of SER using ZnO-NPs

The photocatalytic degradation of SER in aqueous solutions by zinc oxide (ZnO-NPs). The photocatalytic degradation of SER hydrochloride in water was examined in the work employing nanoscale ZnO-NPs [31]. Using the sol-gel process, the ZnO-NPs were created using zinc gluconate as a precursor. H2O2/UV, ZnO-NPs/H2O2/UV, and ZnO-NPs/UV were three different procedures that were investigated for the elimination of SER hydrochloride in water. The outcomes show that for the effective degradation of the medication, ZnO-NP as a catalyst and UV irradiation were both required. With the ZnO-NPs/UV method, nearly 100% of SER hydrochloride was removed (98.7%) in 30 min at room temperature. SER hydrochloride photodegrades according to first-order kinetics at a rate of 0.068 min− 1. The highest drug clearance was attained at pH 11, indicating that SER hydrochloride degradation with ZnO-NPs/UV is pH-dependent. Important factors in the elimination of SER hydrochloride included the initial drug concentration, catalyst dosage, and hydrogen peroxide concentration. The results show that ZnO-NP has an enhanced photocatalytic activity toward the photodegradation of SER hydrochloride due to its high specific surface area and porous structure.

7 Amisulpride

Amisulpride is an atypical antipsychotic medication whose pharmacological effect is based on the selective binding to D2 and D3 dopaminergic receptors (Table S3 in the Supplementary materials file). Compared to traditional antipsychotic medications, this medication has a lower risk of extrapyramidal side effects and generally greater tolerability.

Nowadays, amisulpride is mostly used to treat various forms of schizophrenia [153], while it can also be used to treat depression in low doses [154, 155]. According to the most recent studies, amisulpride can also be used to treat chronic fatigue syndrome [156, 157].

7.1 Photocatalytic degradation of amisulpride

The use of psychiatric medications, a class of pharmaceuticals that are often prescribed and used to treat a variety of mental health issues, is rising globally. Numerous investigations have so far been conducted to look into the existence of psychiatric substances in the environment. Amisulpride is a standard antipsychotic medication [158], and new monitoring investigations have drawn attention to its presence in the environment. For instance, amisulpride has been found in hospital WWTP effluent in Ioannina (northwestern Greece) at concentrations ranging from 102 to 929 ng L− 1 [159]. amisulpride was found in concentrations between 0.2 and 5.5 ng L− 1 in the seawater of the Eastern Mediterranean Sea [160]. Additionally, amisulpride was found in the effluent of the Athens WWTP at a concentration of 0.07 ng L− 1 [161]. Amesulpride was measured at 16.8 ng L− 1 during the first wave of COVID-19 in wastewater samples from the WWTPs in Milan and Monza (Italy) [162]. A persistent pollutant known as amisulpride has also been found to be able to enter the aquatic environment and get into groundwater. Furthermore, the requirement for enhanced treatment has been demonstrated by the verification of the production of its distinctive non-biodegradable N-oxide product during its standard treatment [163, 164]. New methods must be used for removing various contaminants, including pharmaceuticals, from conventional WWTPs [85, 165]. As a tertiary wastewater treatment method, AOPs have been suggested [166, 167]. Heterogeneous photocatalysis is one of the AOPs that holds promise for the elimination of organic chemicals [164, 168].

Today, the development and registration of pharmaceutical products must take into account photostability testing because it is a crucial component of the stability study of medications [169]. Drug photodegradation is a significant area of research since it can cause the drug to lose its effectiveness and have negative side effects because it produces hazardous degradation products. The manufacturing, storage, and administration of pharmaceuticals can benefit greatly by knowing exactly what is created from the drug during this process, which may also greatly increase the safety of therapy [157, 170, 171]. The treatment approach, amisulpride’s efficiency, occurrence in the environmental samples, and concentration are presented in Table 3 [159,160,161, 163].

Table 3 The treatment, amisulpride’s uses, occurrence in the environmental samples, and concentration

7.1.1 Photocatalytic degradation of amisulpride using g-C3N4

In 2023, Antonopoulou et al. [168] did research on the photocatalytic degradation of amisulpride using g-C3N4 as a catalyst while exposed to UV-A rays. The ability of the photocatalytic method to extract amisulpride from both ultrapure and actual municipal wastewater was assessed. Both aqueous matrices have high removal efficiencies. However, a slower rate of degradation was seen when wastewater was used as the matrix, which could be explained by its intricate chemical makeup. Liquid chromatography-mass spectrometry (LC-MS) was used to pinpoint the transformation products (TPs) in both ultrapure and actual municipal wastewater. The photocatalytic degradation routes of amisulpride, which mostly involve oxidation, dealkylation, and cleavage of the methoxy group, are hypothesized based on the detected TPs. Additionally, the importance of h+ and O2•− in the reaction mechanism was demonstrated while studying the contribution of reactive species to the degradation mechanism utilizing well-researched scavengers (Fig. 6). Additionally, Dunaliella tertiolecta and Chlorococcum sp. microalgae were used to determine the progression of ecotoxicity. When ultrapure water was employed as the matrix, little toxicity was seen throughout the entire process without the development of hazardous TPs. In the case of actual municipal wastewater, an early increase in toxicity was noted and is attributable to the matrix’s composition. Heterogeneous photocatalysis was used to minimize the toxicity, and at the end of the procedure, practically full detoxification had been accomplished [168].

Fig. 6
figure 6

Amisulpride’s photocatalytic degradation route [168]

7.1.2 Photocatalytic degradation of amisulpride using UVA irradiation in methanol solution

In 2011, Skibinski [157] examined the photostability of amisulpride under UVA irradiation in methanol solution and carried out structural elucidation of its photodegradation products. The reversed phase UHPLC-DAD technology coupled with an accurate mass hybrid ESI-Q-TOF mass spectrometer was utilized for the quantitative and qualitative analysis of amisulpride and the elucidation of the stress degradation products. The information for determining the photodegradation kinetics and elucidating the structural formulae of the products was gathered during a single run (10 min) using the auto MS/MS mode. Four degradation products were discovered and their masses with good accuracy (0.53–3.05 ppm) and formulas were determined – 258.0666 (C10H14N2O4S), 367.1564 (C17H25N3O4S), 341.1412 (C15H23N3O4S) and 385.1665 (C17H27N3O5S). For all the investigated substances MS/MS fragmentation spectra were obtained (collision energy 19.8–26.1 V) allowing structural elucidation of unknown degradation products and suggesting photodegradation pathways of amisulpride (Fig. 7). It was discovered that the UHPLC-DAD/ESI-Q-TOF system is an effective analytical instrument for the quick and precise stability study of pharmaceutical compounds [157].

Fig. 7
figure 7

Amisulpride’s photodegradation process in solution [157]

7.1.3 Photocatalytic degradation of amisulpride using UVA exposure

The other study [172] examined the UVA photostability of amisulpride, doxepin, haloperidol, risperidone, venlafaxine, and zopiclone. An ultrahigh-performance liquid chromatography approach combined with tandem mass spectrometry (UHPLC-MS/MS) was used to assess the photodegradation process. It was discovered that the type of medicine and any co-existing excipients had an impact on these examined compounds. From MS/MS data, fragmentation routes for the photodegradation products were discovered (Fig. 8). To verify the quality, efficacy, and safety of the formulated products during the production process, storage, as well as usual usage, photostability testing is a crucial component of the medication stability assessment [172].

Fig. 8
figure 8

Using UVA exposure, the photodegradation of amisulpride, doxepin, haloperidol, zopiclone, venlafaxine, and risperidone in bulk drug and pharmaceutical preparation was assessed

8 AMI

Antidepressants and psychoactive medications can be used to treat depression and lessen its effects. AMI (Table S4 in the Supplementary materials file), a tricyclic antidepressant, prevents the neurotransmitters noradrenaline and serotonin from being reabsorbed from the synapses in the central nervous system [11].

A common tricyclic antidepressant used to treat psychiatric illnesses is AMI hydrochloride [173]. Due to its inexpensive cost, AMI is nevertheless used often even though it is relatively more hazardous than other SSRIs [174, 175].

8.1 Photocatalytic degradation of AMI

Since standard wastewater treatments are ineffective for removing these toxins, the drug is primarily excreted through urine after ingestion, making its way to the WWTP and ultimately the surface waterways. AMI was shown to be removed between 65 and 85% in WWTPs, according to several studies [176, 177], and at quantities of 71.6 ng L− 1 [178] and 1.4 ng L− 1 [179], respectively, in surface water and tap water [180].

Previous research claimed that in the presence of naturally occurring HS, amine medicines experience photosensitized degradation via electron transfer interaction. However, even though a significant portion of these medications are metabolized by humans and animals and ultimately released into the environment with their metabolites [181] these investigations mostly focused on the photodegradation of the parent medicines in aqueous solutions [182,183,184,185]. Limited research has been done on the photochemical breakdown of their metabolites in the environment [180,181,182,183,184,185]. The treatment approach, AMI’s efficiency, occurrence in the environmental samples, and concentration are presented in Table 4 [176, 177, 179, 186].

Table 4 The treatment, AMI’s uses, occurrence in the environmental samples, and concentration

8.1.1 Photocatalytic degradation of AMI using ZnO/SnO2

In 2023, researchers looked into sunlight-driven degradation of alprazolam and AMI by application of binary ZnO and tin oxide powders [11]. This study suggests using solid-state synthesized ZnO and tin oxide (SnO2) photocatalysts for the photocatalytic degradation of two specific psychoactive drugs, AMI and alprazolam, in an aqueous system while simulating solar and UV radiation. The gathered information validated the novel materials’ effective synthesis and potential for photocatalytic use. Regarding the photocatalytic evaluation, the main findings show that the presence of ZnO/SnO2 synthesized in a molar ratio of 2:1 and calcined at 700 °C, under 1 mg mL− 1 catalyst loading, led to the highest removal efficiency of AMI and ALP. Based on the reutilization data, it can be inferred that the specified photocatalyst had not lost its efficiency after three sequential runs for the photodegradation of alprazolam. Additionally, pure ZnO powders demonstrated the maximum activity following calcination at 500 °C, in the case of both investigated contaminants. The tert-butanol, sodium fluoride, and ethylenediaminetetraacetic acid tests provided evidence that the proportional contributions of different reactive species altered in the following order: positively charged holes > OHfree > OHads. This is essential because solar radiation is a free source of energy that is renewable and should be used instead of more expensive radiation sources [11].

8.1.2 Photocatalytic degradation of AMI using fulvic acid (FA)

In 2017, Chen et al. [187] carried out a study. The objective of this work was to look at how the common tricyclic antidepressant AMI and its active metabolite nortriptyline are indirectly photodegraded in the presence of FA. AMI and its active metabolite nortriptyline hardly underwent photodegradation in deionized water, but under the simulated sunlight in the FA solutions, they underwent efficient photodegradation. Nortriptyline photodegraded more slowly than AMI did. The substrates’ shapes were related to the photodegradation. The photosensitized reaction in the FA solutions was made easier by the deprotonated form. AMI and nortriptyline both degraded more quickly after the dissolved oxygen was removed, by 10- and 16-fold, respectively. The electron transfer from the nitrogen atom to the excited triplet state FA and subsequent a-hydrogen transfer caused the photodegradation to take place. The amine compounds underwent photosensitized degradation, which reduced the amines to lower ranks, such as tertiary to secondary amine (AMI) and secondary to primary amine (nortriptyline), with the former being the simpler process (Fig. 9). The findings imply that a major method of AMI and its active metabolite removal from natural waters is indirect photodegradation [187].

Fig. 9
figure 9

The suggested photosensitized AMI breakdown process in FA solution [187]

8.1.3 Photocatalytic degradation of AMI using TiO2/WO3

The photocatalytic degradations of the tricyclic antidepressant AMI in aqueous solutions under UV irradiation were examined in the other study [175], which used TiO2 and TiO2/WO3 coatings as photocatalysts. Plasma electrolytic oxidation (PEO) was used to create coatings on a pure titanium substrate. Utilizing non-contact atomic force microscopy and micro-Raman spectroscopy, researchers have examined the influence of certain synthesis parameters on the structural characteristics of produced coatings. Different quantities of crystalline anatase TiO2 and monoclinic WO3 − x phases, as well as amorphous and/or disordered phases of both oxides, were found in coatings, according to Raman scattering tests. Additionally, the data showed that as the PEO’s time increased, the coatings’ surface became more disorganized and uneven. The coating TiO2/WO3 (T/W) (45 s) created during PEO has demonstrated the greatest effectiveness in the elimination of AMI induced by UV light. After four consecutive trials, T/W (45 s) showed a small decline in photodegradation efficiency. The findings demonstrated that the ionic by-products of AMI’s breakdown included nitrite, nitrate, ammonium ion, acetate, and formate, while the highest degree of mineralization (38.8%) was achieved employing coating T/W (45 s) after 120 min of irradiation. Additionally, no discernible harm was seen after four mammalian cell lines—rat hepatoma, mouse neuroblastoma, human colon adenocarcinoma, and human fetal lung—were exposed to radiation from AMI solutions using various coatings. The results of this work show that the catalyst-recovering (removing) phase following the degradation process might be eliminated if coatings employed in the current investigation are applied as photocatalysts. This could facilitate the use of photocatalytic technologies for bigger volumes of wastewater [175].

8.1.4 Photocatalytic degradation of AMI using fe(III)-citrate-oxalate

Wan et al. [188] investigated photodegradation of AMI in a binary Fe(III)-citrate-oxalate system. This study aims to examine the Fe(III)-citrate-oxalate binary system’s synergistic effect in a sun-simulation environment. With a Fe(III)/citrate/oxalate molar ratio of 10:150:500, the Fe(III)-carboxylate binary system displayed good photoreactivity and up to 90% AMT was removed after 30 min at pH 6. In the pH range of 5–8, a synergistic effect was seen in the binary system of Fe(III)-citrate-oxalate. The presence of oxalate citrate depletion. The synergistic impact in the Fe(III)-citrate-oxalate system was made possible by the increased concentration ratios of oxalate to citrate. A putative AMT breakdown pathway based on OH mechanism was suggested by LC-MS studies. The synergistic mechanism of Fe(III)-citrate-oxalate binary complexes may be better understood as a result of this discovery, which has significant potential for use in environmental photocatalysis at pH values close to neutral (Fig. 10) [188].

Fig. 10
figure 10

Potential AMI photodegradation route [188]

9 DZP

The most widely used benzodiazepine drug is DZP (Table S5 in the Supplementary materials file), which is used as a sedative, tranquilizer, anticonvulsant, and muscle relaxant. Residues of this drug were discovered in effluent at 1 g L− 1 and in river and drinkable water at 10 ng L− 1 [78]. DZP, also referred to as Valium, is the most commonly prescribed benzodiazepine used to treat anxiety and alcohol withdrawal [151, 189].

9.1 Photocatalytic degradation of DZP

The medicinal treatment of people and animals is the main source of medicines found in the environment. These medications are broken down by the body and then eliminated as metabolites or the original medication after partial or complete absorption. These pharmaceutical byproducts are either released into the environment or sent to the STP [140]. The presence of pharmaceutical residues in STP effluents demonstrates that these compounds cannot be effectively removed by treatment plants [190]. Because they can generate highly oxidizing species like OH, AOPs are creative and effective ways to eliminate organic contaminants. H2O2 and iron are combined in the Fenton reaction, which promotes the degradation of several pollutant classes in a straightforward procedure under benign operating circumstances [191].

The breakdown of PhACs may be influenced by a variety of mechanisms, including sorption, biodegradation, volatilization, and photolysis [192]. Biochemical responses only remove a small portion of these contaminants because of the pseudo-persistence of many medications [193, 194]. The key process thought to be involved in the alteration of PhACs in the surface layer of water bodies is photolysis [195, 196]. The pollutant can absorb a photon in direct photolysis, which causes it to degrade because it contains aromatic rings, heteroatoms, and other functional groups in its structure. In indirect photolysis, degradation happens as a result of a reaction with active molecules, such as OH, 1O2, ROO, 3DOM*, and eaq (produced by photosensitizers like dissolved organic matter) [151, 192, 197, 198]. The treatment approach, DZP’s efficiency, occurrence in the environmental samples, and concentration are presented in Table 5 [199,200,201,202,203].

Table 5 The treatment, DZP’s uses, occurrence in the environmental samples, and concentration

9.1.1 Photocatalytic degradation of DZP using Fe(NO3)3

Lincomycin (LCM) and DZP were two drugs whose photo-Fenton degradation was examined by Bautitz and Nogueira [191]. It was determined whether ferrioxalate or iron nitrate had any impact on the photo-Fenton degradation efficiency of the drugs LCM and DZP. Under black light or sun irradiation, the degradation of both medicines was sped up in the presence of ferrioxalate in comparison to Fe(NO3)3. The drugs’ deterioration was next assessed in an effluent from a STP while being exposed to black light. Comparing ferrioxalate to Fe(NO3)3, the photo-Fenton degradation of LCM and DZP was improved, leading to the total oxidation of both medicines in a matter of minutes. The initial oxidation of the medicines was not appreciably hampered by the STP effluent [191].

9.1.2 Photocatalytic degradation of DZP using xenon lamp irradiation

In the other study, Jakimska et al. [15] conducted experimental investigations to look into the photodegradation of DZP and SER, two of the most commonly used psychiatric medications, caused by exposure to xenon lamp radiation, which overlaps the spectrum of sunlight, and natural sunshine. The current study focuses on the photodegradation of two common psychiatric medications, SER, and DPZ, caused by exposure to xenon lamp radiation that overlaps the spectrum of sunshine and natural sunlight. In this study, photodegradation was examined using the hyphenated UHPLC-quadrupole time-of-ligh (QToF)-MS technology, which produced extremely precise and reliable results. The development of degradation kinetics revealed the presence of autocatalytic processes (Fig. 11). When liquid chromatography was used in conjunction with QToF-MS, reliable mass measurements and appropriate fragmentation of the target compounds were made possible, allowing for the structure elucidation of the numerous photoproducts of DZP and SER generated during the forced photolysis [151].

Fig. 11
figure 11

The suggested DZP photodegradation route [151]

10 Alprazolam

The environment has been found to contain a sizable number of pharmaceutically active chemicals [80, 204, 205]. The usage of benzodiazepines is primarily for the treatment of anxiety, forgetfulness, and drowsiness. They are a class of psychiatric medications that act on the central nervous system. These medications are often eliminated in the urine after being heavily metabolized in the liver to create glucuronides conjugates, which are easily broken down by bacteria and transformed back into the original active component. Due to the presence of the fecal bacteria Escherichia coli, it appears possible that the glucuronides conjugates indicated above are deconjugated in home wastewater in the same manner. In sewage circumstances, these reactions lead to a rise in the parent compound’s quantity [80, 206]. Alprazolam (Table S6 in the Supplementary materials file), a member of the new generation of 1,4-benzodiazepines, is an anxiolytic that is primarily used to treat anxiety disorders [207, 208]. In addition to this, alprazolam is commonly used for its antidepressant impact in the treatment of pathologies that implicate chronically severe anxiety disorders, such as social phobia, and it is also used to treat panic disorders with or without agoraphobia [209, 210].

10.1 Photocatalytic degradation of alprazolam

Pollutants are removed using a variety of chemical and physical processes, such as coagulation, chemical precipitation, etc., but one drawback of these procedures is that they are not destructive; rather, they merely move the contamination from one phase to another. So-called advanced oxidation techniques, which include homogeneous and heterogeneous photocatalytic oxidation, are an alternative to the methods already stated. These procedures rely on the production of OH, which oxidizes a variety of organic contaminants [211].

Because the aromatic triazole ring appears to prevent the hydrolytic ring opening, alprazolam is insensitive to acid or base hydrolysis [207]. Alprazolam showed resistance to photodegradation by sun radiation, and it is anticipated that the photodegradation process would require many months to completely remove alprazolam from surface waters [206]. Vystavna et al. [212] looked at the presence of 21 medicines in river waters in two distinct regions: Kharkiv, Ukraine, and Bordeaux, France. Less than 100 ng L− 1 of alprazolam revealed almost the same degree of contamination in the rivers under study. The treatment approach, alprazolam’s efficiency, occurrence in the environmental samples, and concentration are presented in Table 6 [148, 213, 214].

Table 6 The treatment, alprazolam’s uses, occurrence in the environmental samples, and concentration

10.1.1 Photocatalytic degradation of alprazolam using Mg-doped ZnO

In a work by Ivetic et al. [215], Mg-doped zinc powders are described structurally and optically using a straightforward solid-state procedure to account for variations in their photocatalytic efficacy in the breakdown of alprazolam. The structural and optical alterations in the ZnO NPs brought on by the manufacturing process and Mg-doping and their effects on the photocatalytic effectiveness in the breakdown of alprazolam were documented. It appeared that the Mg-doped ZnO photocatalyst developed at 700 °C was the most effective. Grain growth, altered shape, and surface reduction in the samples are causes of the Mg-doped ZnO catalyst’s lower photocatalytic activity at higher temperatures (900 °C) [215].

10.1.2 Photocatalytic degradation of alprazolam using UVA/ZnO system

Fincur et al. [210] looked into the effectiveness of heterogeneous photocatalytic degradation of alprazolam in water suspension in 2017. In this study, the effectiveness of alprazolam’s photocatalytic degradation in water suspension was examined. The study tested the effects of many factors on the photodegradation of alprazolam, including the kind of radiation (UVA, Vis, solar) and photocatalyst (ZnO, P25), loading of the photocatalyst, pH, and the presence of OH and positive hole scavengers. The findings of this investigation unequivocally show that the UVA/ZnO system was more effective at removing alprazolam. Here, reaction intermediates produced by both catalysts were carefully examined; hydroxyl derivatives constituted the majority of the intermediates. Following three successive cycles, ZnO demonstrated remarkable photodegradation stability [210].

10.1.3 Photocatalytic degradation of alprazolam using water suspension of brookite type TiO2 nanopowders

Another study [216] from 2015 looked at how the composition and morphology of brookite-type powders affected how well they removed alprazolam through photocatalysis when exposed to UV light. By combining the solegele hydrothermal method with TiCl4 as a precursor, the hydrothermal temperature and reaction time were varied to create two series of nanocrystalline brookite-type powders. To produce a pure brookite phase, the ideal hydrothermal conditions have been identified. The sodium-titanate-dominated powders produced at lower temperatures (120 and 160 °C) were almost inactive in the breakdown of alprazolam, but the brookite-rich powders synthesized at 200 °C demonstrated remarkable photocatalytic efficiency. The most effective sample was pure brookite, which removed more than 98% of alprazolam within the first 30 min of the photocatalytic reaction. Instead of brookite crystallite size and specific surface area of nanopowders, increased amounts of brookite phase in hydrothermally synthesized TiO2 samples have been attributed to improved photocatalytic efficiency in the photodegradation of alprazolam. This improvement may be related to the higher energy of direct and indirect transitions in pure brookite samples, as well as a slight increase in the most frequent pore diameter with increased brookite content in brookite [216].

11 Conclusions

Pharmaceuticals are compounds or combinations of substances used to cure or prevent diseases in humans. They do, however, represent a significant group of point or diffuse pollution-related micropollutants. Conventional wastewater treatment methods have only been shown to partially remove pharmaceutical contaminants, and as a result, municipal and/or hospital WWTPs have become important sources of their release into the environment. Pharmaceutical substances are present in aquatic media in quantities between ng L− 1 and µg L− 1. Like other pharmacological classes, psychiatric drugs can be found in the environment. It is important to remember that these substances do not occur alone but rather as complex combinations, even when these concentrations are below the range projected to damage humans and cause acute or even chronic toxicity to non-target animals. Psychiatric medications are one of the most important classes when it comes to evaluating ecotoxicological consequences in terrestrial and aquatic non-target organisms because of their inherent biological activity, which could disrupt nervous and endocrine systems. Additionally, the growing body of research on chronic toxicity in aquatic organisms that are not the intended targets has shown that extrapolating between acute and chronic toxicity is not advised, underscoring the necessity of coming up with a unique strategy to more clearly define this problem. Considering how frequently medications are exposed to aquatic creatures, it would be more crucial to comprehend life cycle toxicity than to carry out acute toxicity testing.

The environmental pollution brought on by organic pollutants is escalating into a global crisis as a result of industrialization and population increase. Adsorption, ultrafiltration, coagulation, and other traditional biological and physical treatment methods (such as those mentioned above) have traditionally been the go-to methods for removing organic contaminants from various fluids and wastewater. However, new methods must be developed to chemically change many emergent anthropogenic organic pollutants into non-hazardous molecules, especially those with significant toxicity but at extremely low concentrations. These methods ought to be effective enough to quickly eliminate the organic contaminants. In a perfect world, they would also be affordable and “green,” meaning that no secondary pollution is created and the energy source, catalysts, and chemical reagents employed are all readily available, affordable, and ecologically benign. Innovative technologies, such as AOPs, have been developed for the effective removal and degradation of pharmaceutical compounds from wastewater and the prevention of their release to ground and surface waters. AOPs have been proposed as a tertiary treatment in effluent wastewater. AOPs are thought to be promising techniques for the treatment of contaminated wastewater including non-biodegradable organic contaminants since they comprise various mechanisms of producing ROS, primarily OH. The efficient generation of ROS in aqueous environments, the foundation of heterogeneous photocatalysis, enables the total elimination of a variety of pharmaceutical contaminants. Due to this, a wide range of photocatalysts for the photodegradation of psychiatric medications have been created over the past ten years. Results showed that the primary reactive species for the photodegradation of psychiatric medications in the system were photogenerated holes and OH.

Data availability

All data generated or analyzed during this study are included in this published article.

References

  1. Daughton CG, Ternes TA. Pharmaceuticals and personal care products in the environment: agents of subtle change? Environ Health Perspect. 1999;107:907–38.

    Article  Google Scholar 

  2. Liu JL, Wong MH. Pharmaceuticals and personal care products (PPCPs): A review on environmental contamination in China. Environ Int. 2013;59:208–24.

    Article  Google Scholar 

  3. Yang B, Zuo J, Li P, Wang K, Yu X, Zhang M. Effective ultrasound electrochemical degradation of biological toxicity and refractory cephalosporin pharmaceutical wastewater. Chem Eng J. 2016;287:30–7.

    Article  Google Scholar 

  4. Verlicchi P, Galletti A, Petrovic M, Barcelo D. Hospital effluents as a source of emerging pollutants: An overview of micropollutants and sustainable treatment options. J Hydrol. 2010;389:416–28.

    Article  Google Scholar 

  5. Sui Q, Wang B, Zhao W, Huang J, Yu G, Deng S, et al. Identification of priority pharmaceuticals in the water environment of China. Chemosphere. 2012;89:280–6.

    Article  Google Scholar 

  6. Ohoro CR, Adeniji AO, Okoh AI, Okoh AOO. Distribution and chemical analysis of pharmaceuticals and personal care products (PPCPs) in the environmental systems: a review. Int J Environ Res Public Health. 2019;16:3026.

    Article  Google Scholar 

  7. Salimi M, Behbahani M, Sobhi HR, Gholami M, Jonidi Jafari A, Rezaei Kalantary R, et al. A new nano-photocatalyst based on Pt and Bi co-doped TiO2 for efficient visible-light photo degradation of amoxicillin. New J Chem. 2019;43:1562–8.

    Article  Google Scholar 

  8. Hatefi R, Mashinchian-Moradi A, Younesi H, Nojavan S. Graphene quantum dots based on maltose as a high yield photocatalyst for efficient photodegradation of imipramine in wastewater samples. J Environ Health Sci. 2020;18:1531–40.

    Google Scholar 

  9. Schoeman C, Dlamini M, Okonkwo OJ. The impact of a Wastewater Treatment Works in Southern Gauteng, South Africa on efavirenz and nevirapine discharges into the aquatic environment. Emerg Contam. 2017;3:95–106.

    Article  Google Scholar 

  10. Peng Y, Fang W, Krauss M, Brack W, Wang Z, Li F, et al. Screening hundreds of emerging organic pollutants (EOPs) in surface water from the Yangtze River Delta (YRD): Occurrence, distribution, ecological risk. Environ Pollut. 2018;241:484–93.

    Article  Google Scholar 

  11. Fincur N, Sojic Merkulov D, Putnik P, Despotovic V, Banic N, Bognar S, et al. Sunlight-driven degradation of alprazolam and amitriptyline by application of binary zinc oxide and tin oxide powders. Separations. 2023;10:316.

    Article  Google Scholar 

  12. Zuccato E, Calamari D, Natangelo M, Fanelli R. Presence of therapeutic drugs in the environment. Lancet. 2000;355:1789–90.

    Article  Google Scholar 

  13. Rehman MSU, Rashid N, Ashfaq M, Saif A, Ahmad N, Han JI. Global risk of pharmaceutical contamination from highly populated developing countries. Chemosphere. 2015;138:1045–55.

    Article  Google Scholar 

  14. Hejna M, Kapuscinska D, Aksmann A. Pharmaceuticals in the aquatic environment: A review on eco-toxicology and the remediation potential of algae. Int J Env Res Pub He. 2022;19:7717.

    Article  Google Scholar 

  15. Bavumiragira JP, Ge Jn, Yin H. Fate and transport of pharmaceuticals in water systems: A processes review. Sci Total Environ. 2022;823:153635.

    Article  Google Scholar 

  16. Majumder S, Chatterjee S, Basnet P, Mukherjee J. ZnO based nanomaterials for photocatalytic degradation of aqueous pharmaceutical waste solutions – A contemporary review. Environ Nanotechnol Monit Manag. 2020;14:100386.

    Google Scholar 

  17. Quesada HB, Baptista ATA, Cusioli LF, Seibert D, de Oliveira Bezerra C, Bergamasco R. Surface water pollution by pharmaceuticals and an alternative of removal by low-cost adsorbents: A review. Chemosphere. 2019;222:766–80.

    Article  Google Scholar 

  18. Majumder A, Gupta B, Gupta AK. Pharmaceutically active compounds in aqueous environment: A status, toxicity and insights of remediation. Environ Res. 2019;176:108542.

    Article  Google Scholar 

  19. Akerdi AG, Bahrami SH. Application of heterogeneous nano-semiconductors for photocatalytic advanced oxidation of organic compounds: A review. J Environ Chem Eng. 2019;7:103283.

    Article  Google Scholar 

  20. Samer M. Biological and chemical wastewater treatment processes. In: Wastewater Treatment Engineering. London: IntechOpen; 2015.

    Chapter  Google Scholar 

  21. Patel M, Kumar R, Kishor K, Mlsna T, Pittman Jr. CU, Mohan D. Pharmaceuticals of emerging concern in aquatic systems: Chemistry, occurrence, effects, and removal methods. Chem Rev. 2019;119:3510–673.

    Article  Google Scholar 

  22. Abu Hasan H, Muhammad MH, Ismail NI. A review of biological drinking water treatment technologies for contaminants removal from polluted water resources. J Water Process Eng. 2020;33:101035.

    Article  Google Scholar 

  23. Abu Hasan H, Sheikh Abdullah SR, Tan Kofli N, Yeoh SJ. Interaction of environmental factors on simultaneous biosorption of lead and manganese ions by locally isolated Bacillus cereus. J Ind Eng Chem. 2016;37:295–305.

    Article  Google Scholar 

  24. de Souza DI, Dottein EM, Giacobbo A, Siqueira Rodrigues MA, de Pinho MN, Bernardes AM. Nanofiltration for the removal of norfloxacin from pharmaceutical effluent. J Environ Chem Eng. 2018;6:6147–53.

    Article  Google Scholar 

  25. Jaria G, Lourenco MAO, Silva CP, Ferreira P, Otero M, Calisto V, et al. Effect of the surface functionalization of a waste-derived activated carbon on pharmaceuticals’ adsorption from water. J Mol Liq. 2020;299:112098.

    Article  Google Scholar 

  26. Quddus F, Shah A, Iftikhar FJ, Shah NS, Haleem A. Environmentally benign nanoparticles for the photocatalytic degradation of pharmaceutical drugs. Catalysts. 2023;13:511.

    Article  Google Scholar 

  27. Valdez-Carrillo M, Abrell L, Ramirez-Hernandez J, Reyes-Lopez JA, Carreon-Diazconti C. Pharmaceuticals as emerging contaminants in the aquatic environment of Latin America: a review. Environ Sci Pollut R. 2020;27:44863–91.

    Article  Google Scholar 

  28. Valdivia MT, Taggart MA, Pap S, Kean A, Pfleger S, Megson IL. Photocatalytic metallic nanomaterials immobilised onto porous structures: Future perspectives for at-source pharmaceutical removal from hospital wastewater and potential benefits over existing technologies. J Water Process Eng. 2023;52:103553.

    Article  Google Scholar 

  29. Isai KA, Shrivastava VS. Photocatalytic degradation of methylene blue using ZnO and 2%Fe–ZnO semiconductor nanomaterials synthesized by sol–gel method: a comparative study. Sn Appl Sci. 2019;1:1247.

    Article  Google Scholar 

  30. Taourati R, Khaddor M, El Kasmi A. Stable ZnO nanocatalysts with high photocatalytic activity for textile dye treatment. Nano-Struct Nano-Objects. 2019;18:100303.

    Article  Google Scholar 

  31. Mohamed ZH, Riyad YM, Hendawy HA, Abdelbary HMH. Enhanced photocatalytic degradation of the antidepressant sertraline in aqueous solutions by zinc oxide nanoparticles. Water. 2023;15:2074.

    Article  Google Scholar 

  32. Daghrir R, Drogui P, Robert D. Modified TiO2 for environmental photocatalytic applications: a review. Ind Eng Chem Res. 2013;52:3581–99.

    Article  Google Scholar 

  33. Gao B, Chen GZ, Li Puma G. Carbon nanotubes/titanium dioxide (CNTs/TiO2) nanocomposites prepared by conventional and novel surfactant wrapping sol–gel methods exhibiting enhanced photocatalytic activity. Appl Catal B Environ. 2009;89:503–9.

    Article  Google Scholar 

  34. Cozzoli PD, Comparelli R, Fanizza E, Curri ML, Agostiano A. Photocatalytic activity of organic-capped anatase TiO2 nanocrystals in homogeneous organic solutions. Mater Sci Eng C. 2003;23:707–13.

    Article  Google Scholar 

  35. Escudero J, Munoz JL, Morera-Herreras T, Hernandez R, Medrano J, Domingo-Echaburu S, et al. Antipsychotics as environmental pollutants: An underrated threat? Sci Total Environ. 2021;769:144634.

    Article  Google Scholar 

  36. Argaluza J, Domingo-Echaburu S, Orive G, Medrano J, Hernandez R, Lertxundi U. Environmental pollution with psychiatric drugs. World J. Psychia. 2021;11:791–804.

    Article  Google Scholar 

  37. Cunha DL, Mendes MP, Marques M. Environmental risk assessment of psychoactive drugs in the aquatic environment. Environ Sci Pollut R. 2019;26:78–90.

    Article  Google Scholar 

  38. Asimakopoulos AG, Kannan K. Neuropsychiatric pharmaceuticals and illicit drugs in wastewater treatment plants: a review. Environ Chem. 2016;13:541–76.

    Article  Google Scholar 

  39. Daughton CG. Pharmaceuticals and the Environment (PiE): Evolution and impact of the published literature revealed by bibliometric analysis. Sci Total Environ. 2016;562:391–426.

    Article  Google Scholar 

  40. Holm JV, Ruegge K, Bjerg PL, Christensen TH. Occurrence and distribution of pharmaceutical organic compounds in the groundwater downgradient of a landfill (Grindsted, Denmark). Environ Sci Technol. 1995;29:1415–20.

    Article  Google Scholar 

  41. Reddersen K, Heberer T, Dunnbier U. Identification and significance of phenazone drugs and their metabolites in ground- and drinking water. Chemosphere. 2002;49:539–44.

    Article  Google Scholar 

  42. Zuhlke S, Dunnbier U, Heberer T. Detection and identification of phenazone-type drugs and their microbial metabolites in ground and drinking water applying solid-phase extraction and gas chromatography with mass spectrometric detection. J Chromatogr A. 2004;1050:201–9.

    Article  Google Scholar 

  43. Lee WY, Arnold CR. Chronic toxicity of ocean-dumped pharmaceutical wastes to the marine amphipod Amphithoe valida. Mar Pollut Bull. 1983;14:150–3.

    Article  Google Scholar 

  44. Larsson DGJ. Pollution from drug manufacturing: review and perspectives. Philos Trans R Soc Lond B Biol Sci. 2014;369: 20130571.

    Article  Google Scholar 

  45. Williams RT. Human Pharmaceuticals: Assessing the Impacts on Aquatic Ecosystems. Pensacola: SETAC Press; 2005.

    Google Scholar 

  46. Oaks JL, Gilbert M, Virani MZ, Watson RT, Meteyer CU, Rideout BA, et al. Diclofenac residues as the cause of vulture population decline in Pakistan. Nature. 2004;427:630–3.

    Article  Google Scholar 

  47. Larsson DGJ, de Pedro C, Paxeus N. Effluent from drug manufactures contains extremely high levels of pharmaceuticals. J Hazard Mater. 2007;148:751–5.

    Article  Google Scholar 

  48. Kristiansson E, Fick J, Janzon A, Grabic R, Rutgersson C, Weijdegard B, et al. Pyrosequencing of antibiotic-contaminated river sediments reveals high levels of resistance and gene transfer elements. PloS One. 2011;6:e17038.

    Article  Google Scholar 

  49. Fick J, Soderstrom H, Lindberg RH, Phan C, Tysklind M, Larsson DGJ. Contamination of surface, ground, and drinking water from pharmaceutical production. Environ Toxicol Chem. 2009;28:2522–7.

    Article  Google Scholar 

  50. Rutgersson C, Fick J, Marathe N, Kristiansson E, Janzon A, Angelin M, et al. Fluoroquinolones and qnr genes in sediment, water, soil, and human fecal flora in an environment polluted by manufacturing discharges. Environ Sci Technol. 2014;48:7825–32.

    Article  Google Scholar 

  51. Lubick N. India’s drug problem: Chemists show how waste-water contamination affects ecosystem. Nature. 2009;457:640.

    Article  Google Scholar 

  52. Silva B, Costa F, Neves IC, Tavares T. Psychiatric Pharmaceuticals as Emerging Contaminants in Wastewater. Berlin: Springer; 2015.

    Book  Google Scholar 

  53. Hai FI, Yang S, Asif MB, Sencadas V, Shawkat S, Sanderson-Smith M, et al. Carbamazepine as a possible anthropogenic marker in water: Occurrences, toxicological effects, regulations, and removal by wastewater treatment technologies. Water. 2018;10:107.

    Article  Google Scholar 

  54. Trawinski J, Skibinski R. Studies on photodegradation process of psychotropic drugs: a review. Environ Sci Pollut R. 2017;24:1152–99.

    Article  Google Scholar 

  55. Klaminder J, Brodin T, Sundelin A, Anderson NJ, Fahlman J, Jonsson M, et al. Long-term persistence of an anxiolytic drug (oxazepam) in a large freshwater lake. Environ Sci Technol. 2015;49:10406–12.

    Article  Google Scholar 

  56. Wang Z, Gao S, Dai Q, Zhao M, Yang F. Occurrence and risk assessment of psychoactive substances in tap water from China. Environ Pollut. 2020;261:114163.

    Article  Google Scholar 

  57. Amuasi JH, Lucas T, Horton R, Winkler AS. Reconnecting for our future: The Lancet One Health commission. Lancet. 2020;395:1469–71.

    Article  Google Scholar 

  58. Lertxundi U, Hernandez R, Medrano J, Orive G. Drug pollution and pharmacotherapy in psychiatry: A “platypus” in the room. Eur Psychiatry. 2020;63:e33.

    Article  Google Scholar 

  59. Domingo-Echaburu S, Orive G, Lertxundi U. Ivermectin & COVID-19: Let’s keep a One Health perspective. Sustain Chem Pharm. 2021;21:100438.

    Article  Google Scholar 

  60. Kaushik G, Thomas MA. The potential association of psychoactive pharmaceuticals in the environment with human neurological disorders. Sustain Chem Pharm. 2019;13:100148.

    Article  Google Scholar 

  61. Ford AT. From gender benders to brain benders (and beyond!). Aquat Toxicol. 2014;151:1–3.

    Article  Google Scholar 

  62. Brodin T, Piovano S, Fick J, Klaminder J, Heynen M, Jonsson M. Ecological effects of pharmaceuticals in aquatic systems–impacts through behavioural alterations. Philos T Roy Soc B. 2014;369:20130580.

    Article  Google Scholar 

  63. Gilbert N. Drug waste harms fish. Nature. 2011;476:265.

    Article  Google Scholar 

  64. Agerstrand M, Arnold K, Balshine S, Brodin T, Brooks BW, Maack G, et al. Emerging investigator series: use of behavioural endpoints in the regulation of chemicals. Environ Sci Process Impacts. 2020;22:49–65.

    Article  Google Scholar 

  65. Gunnarsson L, Jauhiainen A, Kristiansson E, Nerman O, Larsson DGJ. Evolutionary conservation of human drug targets in organisms used for environmental risk assessments. Environ Sci Technol. 2008;42:5807–13.

    Article  Google Scholar 

  66. Stewart A, Wu N, Cachat J, Hart P, Gaikwad S, Wong K, et al. Pharmacological modulation of anxiety-like phenotypes in adult zebrafish behavioral models. Prog Neuro-Psychoph. 2011;35:1421–31.

    Article  Google Scholar 

  67. Bauknecht P, Jekely G. Ancient coexistence of norepinephrine, tyramine, and octopamine signaling in bilaterians. BMC Biol. 2017;15:6.

    Article  Google Scholar 

  68. Turlejski K. Evolutionary ancient roles of serotonin: long-lasting regulation of activity and development. Acta Neurobiol Exp. 1996;56:619–36.

    Article  Google Scholar 

  69. Horiuchi Y, Kimura R, Kato N, Fujii T, Seki M, Endo T, et al. Evolutional study on acetylcholine expression. Life Sci. 2003;72:1745–56.

    Article  Google Scholar 

  70. Mukherjee S. Novel perspectives on the molecular crosstalk mechanisms of serotonin and melatonin in plants. Plant Physiol Bioch. 2018;132:33–45.

    Article  Google Scholar 

  71. Abbey-Lee RN, Uhrig EJ, Garnham L, Lundgren K, Child S, Lovlie H. Experimental manipulation of monoamine levels alters personality in crickets. Sci Rep-Uk. 2018;8:16211.

    Article  Google Scholar 

  72. de Alvarenga KAF, Sacramento EK, Rosa DV, Souza BR, de Rezende VB, Romano-Silva MA. Effects of antipsychotics on intestinal motility in zebrafish larvae. Neurogastroent Motil. 2017;29:e13006.

    Article  Google Scholar 

  73. Carter LJ, Williams M, Martin S, Kamaludeen SPB, Kookana RS. Sorption, plant uptake and metabolism of benzodiazepines. Sci Total Environ. 2018;628–629:18–25.

    Article  Google Scholar 

  74. Li Y, Miao Y, Zhang W, Yang N, Niu L, Zhang H, et al. Sertraline inhibits top-down forces (predation) in microbial food web and promotes nitrification in sediment. Environ Pollut. 2020;267:115580.

    Article  Google Scholar 

  75. Bound JP, Kitsou K, Voulvoulis N. Household disposal of pharmaceuticals and perception of risk to the environment. Environ Toxicol Phar. 2006;21:301–7.

    Article  Google Scholar 

  76. van der Ven K, Keil D, Moens LN, Hummelen PV, van Remortel P, Maras M, et al. Effects of the antidepressant mianserin in zebrafish: Molecular markers of endocrine disruption. Chemosphere. 2006;65:1836–45.

    Article  Google Scholar 

  77. Cunningham VL, Buzby M, Hutchinson T, Mastrocco F, Parke N, Roden N. Effects of human pharmaceuticals on aquatic life: next steps. Environ Sci Technol. 2006;40:3456–62.

    Article  Google Scholar 

  78. Halling-Sorensen B, Nors Nielsen S, Lanzky PF, Ingerslev F, Holten Lutzhoft HC, Jorgensen SE. Occurrence, fate and effects of pharmaceutical substances in the environment- A review. Chemosphere. 1998;36:357–93.

    Article  Google Scholar 

  79. Petrovic M, Perez S, Barcelo D. Analysis, Removal, Effects and Risk of Pharmaceuticals in the Water Cycle: Occurrence and Transformation in the Environment. 2nd Ed. Amsterdam: Elsevier; 2013.

    Google Scholar 

  80. Calisto V, Esteves VI. Psychiatric pharmaceuticals in the environment. Chemosphere. 2009;77:1257–74.

    Article  Google Scholar 

  81. Kosma CI, Lambropoulou DA, Albanis TA. Investigation of PPCPs in wastewater treatment plants in Greece: Occurrence, removal and environmental risk assessment. Sci Total Environ. 2014;466–467:421–38.

    Article  Google Scholar 

  82. Prieto-Rodriguez L, Oller I, Klamerth N, Aguera A, Rodriguez EM, Malato S. Application of solar AOPs and ozonation for elimination of micropollutants in municipal wastewater treatment plant effluents. Water Res. 2013;47:1521–8.

    Article  Google Scholar 

  83. Sousa MA, Gonçalves C, Vilar VJP, Boaventura RAR, Alpendurada MF. Suspended TiO2-assisted photocatalytic degradation of emerging contaminants in a municipal WWTP effluent using a solar pilot plant with CPCs. Chem Eng J. 2012;198–199:301–9.

    Article  Google Scholar 

  84. Van Doorslaer X, Dewulf J, De Maerschalk J, Van Langenhove H, Demeestere K. Heterogeneous photocatalysis of moxifloxacin in hospital effluent: Effect of selected matrix constituents. Chem Eng J. 2015;261:9–16.

    Article  Google Scholar 

  85. Kanakaraju D, Glass BD, Oelgemoller M. Titanium dioxide photocatalysis for pharmaceutical wastewater treatment. Environ Chem Lett. 2014;12:27–47.

    Article  Google Scholar 

  86. Mahmoud WMM, Rastogi T, Kummerer K. Application of titanium dioxide nanoparticles as a photocatalyst for the removal of micropollutants such as pharmaceuticals from water. Curr Opin Green Sust. 2017;6:1–10.

    Article  Google Scholar 

  87. Salimi M, Esrafili A, Gholami M, Jonidi Jafari A, Rezaei Kalantary R, Farzadkia M, et al. Contaminants of emerging concern: a review of new approach in AOP technologies. Environ Monit Assess. 2017;189:414.

    Article  Google Scholar 

  88. Michael I, Hapeshi E, Michael C, Varela AR, Kyriakou S, Manaia CM, et al. Solar photo-Fenton process on the abatement of antibiotics at a pilot scale: Degradation kinetics, ecotoxicity and phytotoxicity assessment and removal of antibiotic resistant enterococci. Water Res. 2012;46:5621–34.

    Article  Google Scholar 

  89. Rapti I, Kourkouta T, Malisova EM, Albanis T, Konstantinou I. Photocatalytic degradation of inherent pharmaceutical concentration levels in real hospital WWTP effluents using g-C3N4 catalyst on CPC pilot-scale reactor. Molecules. 2023;28:1170.

    Article  Google Scholar 

  90. Klavarioti M, Mantzavinos D, Kassinos D. Removal of residual pharmaceuticals from aqueous systems by advanced oxidation processes. Environ Int. 2009;35:402–17.

    Article  Google Scholar 

  91. Liu QT, Williams HE. Kinetics and degradation products for direct photolysis of β-blockers in water. Environ Sci Technol. 2007;41:803–10.

    Article  Google Scholar 

  92. Li YNB, Moore DE, Tattam BN. Photodegradation of amiloride in aqueous solution. Int J Pharmaceut. 1999;183:109–16.

    Article  Google Scholar 

  93. Li W, Lu S, Qiu Z, Lin K. Clofibric acid degradation in UV254/H2O2 process: Effect of temperature. J Hazard Mater. 2010;176:1051–7.

    Article  Google Scholar 

  94. Andreozzi R, Caprio V, Marotta R, Vogna D. Paracetamol oxidation from aqueous solutions by means of ozonation and H2O2/UV system. Water Res. 2003;37:993–1004.

    Article  Google Scholar 

  95. Rosario-Ortiz FL, Wert EC, Snyder SA. Evaluation of UV/H2O2 treatment for the oxidation of pharmaceuticals in wastewater. Water Res. 2010;44:1440–8.

    Article  Google Scholar 

  96. Vogna D, Marotta R, Napolitano A, Andreozzi R, d’Ischia M. Advanced oxidation of the pharmaceutical drug diclofenac with UV/H2O2 and ozone. Water Res. 2004;38:414–22.

    Article  Google Scholar 

  97. Ravina M, Campanella L, Kiwi J. Accelerated mineralization of the drug Diclofenac via Fenton reactions in a concentric photo-reactor. Water Res. 2002;36:3553–60.

    Article  Google Scholar 

  98. Perez-Estrada LA, Maldonado MI, Gernjak W, Aguera A, Fernandez-Alba AR, Ballesteros MM, et al. Decomposition of diclofenac by solar driven photocatalysis at pilot plant scale. Catal Today. 2005;101:219–26.

    Article  Google Scholar 

  99. Goi A, Veressinina Y, Trapido M. Degradation of salicylic acid by Fenton and modified Fenton treatment. Chem Eng J. 2008;143:1–9.

    Article  Google Scholar 

  100. Trovo AG, Melo SAS, Nogueira RFP. Photodegradation of the pharmaceuticals amoxicillin, bezafibrate and paracetamol by the photo-Fenton process—Application to sewage treatment plant effluent. J Photoch Photobio A. 2008;198:215–20.

    Article  Google Scholar 

  101. Melo SAS, Trovo AG, Bautitz IR, Nogueira RFP. Degradation of residual drugs by advanced oxidative processes. Quim Nova. 2009;32:188–97 [in Portuguese]. https://doi.org/10.1590/S0100-40422009000100034 (Accessed 17 Apr 2024)

  102. Panayotov DA, Yates JT. Depletion of conduction band electrons in TiO2 by water chemisorption – IR spectroscopic studies of the independence of Ti–OH frequencies on electron concentration. Chem Phys Lett. 2005;410:11–7.

    Article  Google Scholar 

  103. Thompson TL, Yates JT. Monitoring hole trapping in photoexcited TiO2(110) using a surface photoreaction. J Photoch Photobio B. 2005;109:18230–6.

    Google Scholar 

  104. Tong A, Braund R, Warren D, Peake B. TiO2-assisted photodegradation of pharmaceuticals — a review. Open Chem. 2012;10:989–1027.

    Article  Google Scholar 

  105. Magureanu M, Mandache NB, Parvulescu VI. Degradation of pharmaceutical compounds in water by non-thermal plasma treatment. Water Res. 2015;81:124–36.

    Article  Google Scholar 

  106. Bruggeman P, Schram D, Gonzalez MA, Rego R, Kong MG, Leys C. Characterization of a direct dc-excited discharge in water by optical emission spectroscopy. Plasma Sources Sci T. 2009;18:025017.

    Article  Google Scholar 

  107. Kanazawa S, Kawano H, Watanabe S, Furuki T, Akamine S, Ichiki R, et al. Observation of OH radicals produced by pulsed discharges on the surface of a liquid. Plasma Sources Sci T. 2011;20:034010.

    Article  Google Scholar 

  108. Bruggeman P, Schram DC. On OH production in water containing atmospheric pressure plasmas. Plasma Sources Sci T. 2010;19:045025.

    Article  Google Scholar 

  109. Papagiannis I, Koutsikou G, Frontistis Z, Konstantinou I, Avgouropoulos G, Mantzavinos D, et al. Photoelectrocatalytic vs. photocatalytic degradation of organic waterborne pollutants. Catalysts. 2018;8:455.

    Article  Google Scholar 

  110. Basavarajappa PS, Patil SB, Ganganagappa N, Reddy KR, Raghu AV, Reddy CV. Recent progress in metal-doped TiO2, non-metal doped/codoped TiO2 and TiO2 nanostructured hybrids for enhanced photocatalysis. Int J Hydrogen Energ. 2020;45:7764–78.

    Article  Google Scholar 

  111. Ying S, Guan Z, Ofoegbu PC, Clubb P, Rico C, He F, et al. Green synthesis of nanoparticles: Current developments and limitations. Environ Technol Inno. 2022;26:102336.

    Article  Google Scholar 

  112. Khan I, Saeed K, Khan I. Nanoparticles: Properties, applications and toxicities. Arab J Chem. 2019;12:908–31.

    Article  Google Scholar 

  113. Rapti I, Bairamis F, Konstantinou I. g-C3N4/MoS2 heterojunction for photocatalytic removal of phenol and Cr(VI). Photochem. 2021;1:358–70.

    Article  Google Scholar 

  114. Bairamis F, Konstantinou I. WO3 fibers/g-C3N4 Z-scheme heterostructure photocatalysts for simultaneous oxidation/reduction of phenol/Cr(VI) in aquatic media. Catalysts. 2021;11:792.

    Article  Google Scholar 

  115. Zhong J, Jiang H, Wang Z, Yu Z, Wang L, Mueller JF, et al. Efficient photocatalytic destruction of recalcitrant micropollutants using graphitic carbon nitride under simulated sunlight irradiation. Environ Sci Ecotechnol. 2021;5:100079.

    Article  Google Scholar 

  116. Kane A, Chafiq L, Dalhatou S, Bonnet P, Nasr M, Gaillard N, et al. g-C3N4/TiO2 S-scheme heterojunction photocatalyst with enhanced photocatalytic Carbamazepine degradation and mineralization. J Photoch Photobio A. 2022;430:113971.

    Article  Google Scholar 

  117. Gao B, Wang J, Dou M, Xu C, Huang X. Enhanced photocatalytic removal of amoxicillin with Ag/TiO2/mesoporous g-C3N4 under visible light: property and mechanistic studies. Environ Sci Pollut R. 2020;27:7025–39.

    Article  Google Scholar 

  118. Zhang Y, Geißen SU, Gal C. Carbamazepine and diclofenac: Removal in wastewater treatment plants and occurrence in water bodies. Chemosphere. 2008;73:1151–61.

    Article  Google Scholar 

  119. Carabin A, Drogui P, Robert D. Photo-degradation of carbamazepine using TiO2 suspended photocatalysts. J Taiwan Inst Chem E. 2015;54:109–17.

    Article  Google Scholar 

  120. Joss A, Keller E, Alder AC, Göbel A, McArdell CS, Ternes T, et al. Removal of pharmaceuticals and fragrances in biological wastewater treatment. Water Res. 2005;39:3139–52.

    Article  Google Scholar 

  121. Vieno N, Tuhkanen T, Kronberg L. Elimination of pharmaceuticals in sewage treatment plants in Finland. Water Res. 2007;41:1001–12.

    Article  Google Scholar 

  122. Xie Z, Lu G, Liu J, Yan Z, Ma B, Zhang Z, et al. Occurrence, bioaccumulation, and trophic magnification of pharmaceutically active compounds in Taihu Lake, China. Chemosphere. 2015;138:140–7.

    Article  Google Scholar 

  123. Gurke R, Roßler M, Marx C, Diamond S, Schubert S, Oertel R, et al. Occurrence and removal of frequently prescribed pharmaceuticals and corresponding metabolites in wastewater of a sewage treatment plant. Sci Total Environ. 2015;532:762–70.

    Article  Google Scholar 

  124. Matongo S, Birungi G, Moodley B, Ndungu P. Occurrence of selected pharmaceuticals in water and sediment of Umgeni River, KwaZulu-Natal, South Africa. Environ Sci Pollut R. 2015;22:10298–308.

    Article  Google Scholar 

  125. Lopez B, Ollivier P, Togola A, Baran N, Ghestem JP. Screening of French groundwater for regulated and emerging contaminants. Sci Total Environ. 2015;518–519:562–73.

    Article  Google Scholar 

  126. Mohapatra DP, Brar SK, Tyagi RD, Picard P, Surampalli RY. Carbamazepine in municipal wastewater and wastewater sludge: Ultrafast quantification by laser diode thermal desorption-atmospheric pressure chemical ionization coupled with tandem mass spectrometry. Talanta. 2012;99:247–55.

    Article  Google Scholar 

  127. De Laurentiis E, Chiron S, Kouras-Hadef S, Richard C, Minella M, Maurino V, et al. Photochemical fate of carbamazepine in surface freshwaters: laboratory measures and modeling. Environ Sci Technol. 2012;46:8164–73.

    Article  Google Scholar 

  128. Arye G, Dror I, Berkowitz B. Fate and transport of carbamazepine in soil aquifer treatment (SAT) infiltration basin soils. Chemosphere. 2011;82:244–52.

    Article  Google Scholar 

  129. Ghauch A, Baydoun H, Dermesropian P. Degradation of aqueous carbamazepine in ultrasonic/Fe0/H2O2 systems. Chem Eng J. 2011;172:18–27.

    Article  Google Scholar 

  130. Abdelmelek SB, Greaves J, Ishida KP, Cooper WJ, Song W. Removal of pharmaceutical and personal care products from reverse osmosis retentate using advanced oxidation processes. Environ Sci Technol. 2011;45:3665–71.

    Article  Google Scholar 

  131. Zhang W, Ding Y, Boyd SA, Teppen BJ, Li H. Sorption and desorption of carbamazepine from water by smectite clays. Chemosphere. 2010;81:954–60.

    Article  Google Scholar 

  132. Kosjek T, Andersen HR, Kompare B, Ledin A, Heath E. Fate of carbamazepine during water treatment. Environ Sci Technol. 2009;43:6256–61.

    Article  Google Scholar 

  133. Williams CF, Williams CF, Adamsen FJ. Sorption-desorption of carbamazepine from irrigated soils. J Environ Qual. 2006;35:1779–83.

    Article  Google Scholar 

  134. Xu J, Li L, Guo C, Zhang Y, Meng W. Photocatalytic degradation of carbamazepine by tailored BiPO4: efficiency, intermediates and pathway. Appl Catal B-Environ. 2013;130–131:285–92.

    Article  Google Scholar 

  135. Li W, Nanaboina V, Zhou Q, Korshin GV. Effects of Fenton treatment on the properties of effluent organic matter and their relationships with the degradation of pharmaceuticals and personal care products. Water Res. 2012;46:403–12.

    Article  Google Scholar 

  136. Keen OS, Baik S, Linden KG, Aga DS, Love NG. Enhanced biodegradation of carbamazepine after UV/H2O2 advanced oxidation. Environ Sci Technol. 2012;46:6222–7.

    Article  Google Scholar 

  137. Naddeo V, Meric S, Kassinos D, Belgiorno V, Guida M. Fate of pharmaceuticals in contaminated urban wastewater effluent under ultrasonic irradiation. Water Res. 2009;43:4019–27.

    Article  Google Scholar 

  138. Mohapatra DP, Brar SK, Tyagi RD, Picard P, Surampalli RY. Analysis and advanced oxidation treatment of a persistent pharmaceutical compound in wastewater and wastewater sludge-carbamazepine. Sci Total Environ. 2014;470–471:58–75.

    Article  Google Scholar 

  139. Tang L, Wang JJ, Jia CT, Lv GX, Xu G, Li WT, et al. Simulated solar driven catalytic degradation of psychiatric drug carbamazepine with binary BiVO4 heterostructures sensitized by graphene quantum dots. Appl Catal B-Environ. 2017;205:587–96.

    Article  Google Scholar 

  140. Ternes TA. Occurrence of drugs in German sewage treatment plants and rivers. Water Res. 1998;32:3245–60.

    Article  Google Scholar 

  141. Drewes JE, Heberer T, Reddersen K. Fate of pharmaceuticals during indirect potable reuse. Water Sci Technol. 2002;46:73–80.

    Article  Google Scholar 

  142. Benotti MJ, Trenholm RA, Vanderford BJ, Holady JC, Stanford BD, Snyder SA. Pharmaceuticals and endocrine disrupting compounds in U.S. drinking water. Environ Sci Technol. 2009;43:597–603.

    Article  Google Scholar 

  143. Gao X, Peng W, Tang G, Guo Q, Luo Y. Highly efficient and visible-light-driven BiOCl for photocatalytic degradation of carbamazepine. J Alloy Compd. 2018;757:455–65.

    Article  Google Scholar 

  144. Chen H, Wang X, Bi W, Wu Y, Dong W. Photodegradation of carbamazepine with BiOCl/Fe3O4 catalyst under simulated solar light irradiation. J Colloid Interf Sci. 2017;502:89–99.

    Article  Google Scholar 

  145. Gornik T, Vozic A, Heath E, Trontelj J, Roskar R, Zigon D, et al. Determination and photodegradation of sertraline residues in aqueous environment. Environ Pollut. 2020;256:113431.

    Article  Google Scholar 

  146. Minagh E, Hernan R, O’Rourke K, Lyng FM, Davoren M. Aquatic ecotoxicity of the selective serotonin reuptake inhibitor sertraline hydrochloride in a battery of freshwater test species. Ecotox Environ Safe. 2009;72:434–40.

    Article  Google Scholar 

  147. Mole RA, Brooks BW. Global scanning of selective serotonin reuptake inhibitors: occurrence, wastewater treatment and hazards in aquatic systems. Environ Pollut. 2019;250:1019–31.

    Article  Google Scholar 

  148. Golbaz S, Zamanzadeh M, Yaghmaeian K, Nabizadeh R, Rastkari N, Esfahani H. Occurrence and removal of psychiatric pharmaceuticals in the Tehran South Municipal Wastewater Treatment Plant. Environ Sci Pollut R. 2023;30:27041–55.

    Article  Google Scholar 

  149. Vasskog T, Berger U, Samuelsen PJ, Kallenborn R, Jensen E. Selective serotonin reuptake inhibitors in sewage influents and effluents from Tromso, Norway. J Chromatogr A. 2006;1115:187–95.

    Article  Google Scholar 

  150. Nagarnaik P, Batt A, Boulanger B. Source characterization of nervous system active pharmaceutical ingredients in healthcare facility wastewaters. J Environ Manage. 2011;92:872–7.

    Article  Google Scholar 

  151. Jakimska A, Sliwka Kaszynska M, Nagorski P, Kot Wasik A, Namieanik J. Environmental fate of two psychiatric drugs, diazepam and sertraline: phototransformation and investigation of their photoproducts in natural waters. J Chromatogr Sep Tech. 2014;5:253.

    Google Scholar 

  152. Rejek M, Grzechulska-Damszel J, Schmidt B. Synthesis, characterization, and evaluation of Degussa P25/chitosan composites for the photocatalytic removal of sertraline and Acid Red 18 from water. J Polym Environ. 2021;29:3660–7.

    Article  Google Scholar 

  153. Mortimer AM. How do we choose between atypical antipsychotics? The advantages of amisulpride. Int J Neuropsychop. 2004;7:S21–5.

    Article  Google Scholar 

  154. Amore M, Jori MC. Faster response on amisulpride 50 mg versus sertraline 50–100 mg in patients with dysthymia or double depression: a randomized, double-blind, parallel group study. Int Clin Psychopharm. 2001;16:317–24.

    Article  Google Scholar 

  155. Cassano GB, Jori MC. Efficacy and safety of amisulpride 50 mg versus paroxetine 20 mg in major depression: a randomized, double-blind, parallel group study. Int Clin Psychopharm. 2002;17:27–32.

    Article  Google Scholar 

  156. Pardini M, Guida S, Primavera A, Krueger F, Cocito L, Gialloreti LE. Amisulpride vs. fluoxetine treatment of Chronic Fatigue Syndrome: A pilot study. Eur Neuropsychopharm. 2011;21:282–6.

    Article  Google Scholar 

  157. Skibinski R. Identification of photodegradation product of amisulpride by ultra-high-pressure liquid chromatography–DAD/ESI-quadrupole time-of-flight-mass spectrometry. J Pharmaceut Biomed. 2011;56:904–10.

    Article  Google Scholar 

  158. Gros M, Williams M, Llorca M, Rodriguez-Mozaz S, Barcelo D, Kookana RS. Photolysis of the antidepressants amisulpride and desipramine in wastewaters: Identification of transformation products formed and their fate. Sci Total Environ. 2015;530–531:434–44.

    Article  Google Scholar 

  159. Konstas PS, Kosma C, Konstantinou I, Albanis T. Photocatalytic treatment of pharmaceuticals in real hospital wastewaters for effluent quality amelioration. Water. 2019;11:2165.

    Article  Google Scholar 

  160. Alygizakis NA, Gago-Ferrero P, Borova VL, Pavlidou A, Hatzianestis I, Thomaidis NS. Occurrence and spatial distribution of 158 pharmaceuticals, drugs of abuse and related metabolites in offshore seawater. Sci Total Environ. 2016;541:1097–105.

    Article  Google Scholar 

  161. Gago-Ferrero P, Bletsou AA, Damalas DE, Aalizadeh R, Alygizakis NA, Singer HP, et al. Wide-scope target screening of > 2000 emerging contaminants in wastewater samples with UPLC-Q-ToF-HRMS/MS and smart evaluation of its performance through the validation of 195 selected representative analytes. J Hazard Mater. 2020;387:121712.

    Article  Google Scholar 

  162. Cappelli F, Longoni O, Rigato J, Rusconi M, Sala A, Fochi I, et al. Suspect screening of wastewaters to trace anti-COVID-19 drugs: Potential adverse effects on aquatic environment. Sci Total Environ. 2022;824:153756.

    Article  Google Scholar 

  163. Bollmann AF, Seitz W, Prasse C, Lucke T, Schulz W, Ternes T. Occurrence and fate of amisulpride, sulpiride, and lamotrigine in municipal wastewater treatment plants with biological treatment and ozonation. J Hazard Mater. 2016;320:204–15.

    Article  Google Scholar 

  164. Spyrou A, Tzamaria A, Dormousoglou M, Skourti A, Vlastos D, Papadaki M, et al. The overall assessment of simultaneous photocatalytic degradation of Cimetidine and Amisulpride by using chemical and genotoxicological approaches. Sci Total Environ. 2022;838:156140.

    Article  Google Scholar 

  165. Rivera-Utrilla J, Sanchez-Polo M, Ferro-Garcia MA, Prados-Joya G, Ocampo-Perez R. Pharmaceuticals as emerging contaminants and their removal from water. A review. Chemosphere. 2013;93:1268–87.

    Article  Google Scholar 

  166. Kosma CI, Kapsi MG, Konstas PSG, Trantopoulos EP, Boti VI, Konstantinou IK, et al. Assessment of multiclass pharmaceutical active compounds (PhACs) in hospital WWTP influent and effluent samples by UHPLC-Orbitrap MS: Temporal variation, removals and environmental risk assessment. Environ Res. 2020;191:110152.

    Article  Google Scholar 

  167. Parry E, Young TM. Comparing targeted and non-targeted high-resolution mass spectrometric approaches for assessing advanced oxidation reactor performance. Water Res. 2016;104:72–81.

    Article  Google Scholar 

  168. Antonopoulou M, Papadaki M, Rapti I, Konstantinou I. Photocatalytic degradation of pharmaceutical amisulpride using g-C3N4 Catalyst and UV-A irradiation. Catalysts. 2023;13:226.

    Article  Google Scholar 

  169. Aman W, Thoma K. The influence of formulation and manufacturing process on the photostability of tablets. Int J Pharmaceut. 2002;243:33–41.

    Article  Google Scholar 

  170. Tonnesen HH. Formulation and stability testing of photolabile drugs. Int J Pharmaceut. 2001;225:1–14.

    Article  Google Scholar 

  171. Clothier RH. Phototoxicity and acute toxicity studies conducted by the FRAME Alternatives Laboratory: a brief review. ATLA-Altern Lab Anim. 2007;35:515–9.

    Article  Google Scholar 

  172. Maslanka A, Zmudzki P, Szlosarczyk M, Talik P, Hubicka U. Photodegradation assessment of amisulpride, doxepin, haloperidol, risperidone, venlafaxine, and zopiclone in bulk drug and in the presence of excipients. Monatsh Chem. 2020;151:483–93.

    Article  Google Scholar 

  173. Fernandez-Navarro JJ, Ruiz-Angel MJ, Garcia-Alvarez-Coque MC. Reversed-phase liquid chromatography without organic solvent for determination of tricyclic antidepressants. J Sep Sci. 2012;35:1303–9.

    Article  Google Scholar 

  174. Nabais JMV, Ledesma B, Laginhas C. Removal of amitriptyline from aqueous media using activated carbons. Adsorpt Sci Technol. 2012;30:255–63.

    Article  Google Scholar 

  175. Fincur NL, Grujic-Brojcin M, Scepanovic MJ, Cetojevic-Simin DD, Maletic SP, Stojadinovic S, et al. UV-driven removal of tricyclic antidepressive drug amitriptyline using TiO2 and TiO2/WO3 coatings. React Kinet Mech Cat. 2021;132:1193–209.

    Article  Google Scholar 

  176. Baker DR, Kasprzyk-Hordern B. Spatial and temporal occurrence of pharmaceuticals and illicit drugs in the aqueous environment and during wastewater treatment: New developments. Sci Total Environ. 2013;454–455:442–56.

    Article  Google Scholar 

  177. Kasprzyk-Hordern B, Dinsdale RM, Guwy AJ. The removal of pharmaceuticals, personal care products, endocrine disruptors and illicit drugs during wastewater treatment and its impact on the quality of receiving waters. Water Res. 2009;43:363–80.

    Article  Google Scholar 

  178. Baker DR, Kasprzyk-Hordern B. Multi-residue analysis of drugs of abuse in wastewater and surface water by solid-phase extraction and liquid chromatography–positive electrospray ionisation tandem mass spectrometry. J Chromatogr A. 2011;1218:1620–31.

    Article  Google Scholar 

  179. Togola A, Budzinski H. Multi-residue analysis of pharmaceutical compounds in aqueous samples. J Chromatogr A. 2008;1177:150–8.

    Article  Google Scholar 

  180. A. Osawa R, T. Barrocas B, C. Monteiro O, Oliveira MC, Florencio MH. Visible light photocatalytic degradation of amitriptyline using cobalt doped titanate nanowires: Kinetics and characterization of transformation products. J Environ Chem Eng. 2020;8:103585.

    Article  Google Scholar 

  181. Chen Y, Hu C, Hu X, Qu J. Indirect photodegradation of amine drugs in aqueous solution under simulated sunlight. Environ Sci Technol. 2009;43:2760–5.

    Article  Google Scholar 

  182. Chen Y, Li H, Wang Z, Li H, Tao T, Zuo Y. Photodegradation of selected β-blockers in aqueous fulvic acid solutions: Kinetics, mechanism, and product analysis. Water Res. 2012;46:2965–72.

    Article  Google Scholar 

  183. Chen Y, Liang Q, Zhou D, Wang Z, Tao T, Zuo Y. Photodegradation kinetics, products and mechanism of timolol under simulated sunlight. J Hazard Mater. 2013;252–253:220–6.

    Google Scholar 

  184. Makunina MP, Pozdnyakov IP, Chen Y, Grivin VP, Bazhin NM, Plyusnin VF. Mechanistic study of fulvic acid assisted propranolol photodegradation in aqueous solution. Chemosphere. 2015;119:1406–10.

    Article  Google Scholar 

  185. Aymerich I, Acuna V, Barcelo D, Garcia MJ, Petrovic M, Poch M, et al. Attenuation of pharmaceuticals and their transformation products in a wastewater treatment plant and its receiving river ecosystem. Water Res. 2016;100:126–36.

    Article  Google Scholar 

  186. Pivetta RC, Rodrigues-Silva C, Ribeiro AR, Rath S. Tracking the occurrence of psychotropic pharmaceuticals in Brazilian wastewater treatment plants and surface water, with assessment of environmental risks. Sci Total Environ. 2020;727:138661.

    Article  Google Scholar 

  187. Chen Y, Liang J, Liu L, Lu X, Deng J, Pozdnyakov IP, et al. Photosensitized degradation of amitriptyline and its active metabolite nortriptyline in aqueous fulvic acid solution. J Environ Qual. 2017;46:1081–7.

    Article  Google Scholar 

  188. Wan D, Zuo J, Chen Y, Chen Q, Zuo Y. Photodegradation of amitriptyline in Fe(III)-citrate-oxalate binary system: Synergistic effect and mechanism. Chemosphere. 2018;210:224–31.

    Article  Google Scholar 

  189. West CE, Rowland SJ. Aqueous phototransformation of diazepam and related human metabolites under simulated sunlight. Environ Sci Technol. 2012;46:4749–56.

    Article  Google Scholar 

  190. Gomez MJ, Martinez Bueno MJ, Lacorte S, Fernandez-Alba AR, Aguera A. Pilot survey monitoring pharmaceuticals and related compounds in a sewage treatment plant located on the Mediterranean coast. Chemosphere. 2007;66:993–1002.

    Article  Google Scholar 

  191. Bautitz IR, Nogueira RFP. Photodegradation of lincomycin and diazepam in sewage treatment plant effluent by photo-Fenton process. Catal Today. 2010;151:94–9.

    Article  Google Scholar 

  192. Dong MM, Trenholm R, Rosario-Ortiz FL. Photochemical degradation of atenolol, carbamazepine, meprobamate, phenytoin and primidone in wastewater effluents. J Hazard Mater. 2015;282:216–23.

    Article  Google Scholar 

  193. Fatta-Kassinos D, Vasquez MI, Kummerer K. Transformation products of pharmaceuticals in surface waters and wastewater formed during photolysis and advanced oxidation processes – Degradation, elucidation of byproducts and assessment of their biological potency. Chemosphere. 2011;85:693–709.

    Article  Google Scholar 

  194. Michael I, Vasquez MI, Hapeshi E, Haddad T, Baginska E, Kummerer K, et al. Contaminants of emerging concern: metabolites and transformation products of pharmaceuticals in the aquatic environment. In: Lambropoulou DA, Nollet LML, editors. Transformation Products of Emerging Contaminants in the Environment. Hoboken: John Wiley & Sons; 2014, p. 413–58.

    Chapter  Google Scholar 

  195. Boreen AL, Arnold WA, McNeill K. Photodegradation of pharmaceuticals in the aquatic environment: A review. Aquat Sci. 2003;65:320–41.

    Article  Google Scholar 

  196. Lin AY, Reinhard M. Photodegradation of common environmental pharmaceuticals and estrogens in river water. Environ Toxicol Chem. 2005;24:1303–9.

    Article  Google Scholar 

  197. Alapi T, Dombi A. Comparative study of the UV and UV/VUV-induced photolysis of phenol in aqueous solution. J Photoch Photobio A. 2007;188:409–18.

    Article  Google Scholar 

  198. Andreozzi R, Raffaele M, Nicklas P. Pharmaceuticals in STP effluents and their solar photodegradation in aquatic environment. Chemosphere. 2003;50:1319–30.

    Article  Google Scholar 

  199. Huerta-Fontela M, Galceran MT, Ventura F. Occurrence and removal of pharmaceuticals and hormones through drinking water treatment. Water Res. 2011;45:1432–42.

    Article  Google Scholar 

  200. Martinez Bueno MJ, Hernando MD, Herrera S, Gomez MJ, Fernandez-Alba AR, Bustamante I, et al. Pilot survey of chemical contaminants from industrial and human activities in river waters of Spain. Int J Environ An Ch. 2010;90:321–43.

    Article  Google Scholar 

  201. Ternes T, Bonerz M, Schmidt T. Determination of neutral pharmaceuticals in wastewater and rivers by liquid chromatography–electrospray tandem mass spectrometry. J Chromatogr A. 2001;938:175–85.

    Article  Google Scholar 

  202. Shao B, Chen D, Zhang J, Wu Y, Sun C. Determination of 76 pharmaceutical drugs by liquid chromatography–tandem mass spectrometry in slaughterhouse wastewater. J Chromatogr A. 2009;1216:8312–8.

    Article  Google Scholar 

  203. Sun Q, Lv M, Hu A, Yang X, Yu CP. Seasonal variation in the occurrence and removal of pharmaceuticals and personal care products in a wastewater treatment plant in Xiamen, China. J Hazard Mater. 2014;277:69–75.

    Article  Google Scholar 

  204. Bottoni P, Caroli S, Caracciolo AB. Pharmaceuticals as priority water contaminants. Toxicol Environ Chem. 2010;92:549–65.

    Article  Google Scholar 

  205. Radovic T, Grujic S, Petkovic A, Dimkic M, Lausevic M. Determination of pharmaceuticals and pesticides in river sediments and corresponding surface and ground water in the Danube River and tributaries in Serbia. Environ Monit Assess. 2014;187:4092.

    Article  Google Scholar 

  206. Calisto V, Domingues MRM, Esteves VI. Photodegradation of psychiatric pharmaceuticals in aquatic environments – Kinetics and photodegradation products. Water Res. 2011;45:6097–106.

    Article  Google Scholar 

  207. Cabrera CG, de Waisbaum RG, Nudelman NS. Kinetic and mechanistic studies on the hydrolysis and photodegradation of diazepam and alprazolam. J Phys Org Chem. 2005;18:156–61.

    Article  Google Scholar 

  208. Ganjali MR, Haji-Hashemi H, Faridbod F, Norouzi P, Qomi M. Potentiometric Determination of alprazolam based on carbon paste and PVC membrane electrodes. Int J Electrochem Sc. 2012;7:1470–81.

    Article  Google Scholar 

  209. Castaneda B, Ortiz-Cala W, Gallardo-Cabrera C, Nudelman NS. Stability studies of alprazolam tablets: effects of chemical interactions with some excipients in pharmaceutical solid preparations. J Phys Org Chem. 2009;22:807–14.

    Article  Google Scholar 

  210. Fincur NL, Krstic JB, Sibul FS, Sojic DV, Despotovic VN, Banic ND, et al. Removal of alprazolam from aqueous solutions by heterogeneous photocatalysis: Influencing factors, intermediates, and products. Chem Eng J. 2017;307:1105–15.

    Article  Google Scholar 

  211. Ahmed S, Rasul MG, Brown R, Hashib MA. Influence of parameters on the heterogeneous photocatalytic degradation of pesticides and phenolic contaminants in wastewater: A short review. J Environ Manage. 2011;92:311–30.

    Article  Google Scholar 

  212. Vystavna Y, Huneau F, Grynenko V, Vergeles Y, Celle-Jeanton H, Tapie N, et al. Pharmaceuticals in rivers of two regions with contrasted socio-economic conditions: occurrence, accumulation, and comparison for Ukraine and France. Water Air Soil Poll. 2012;223:2111–24.

    Article  Google Scholar 

  213. Wu M, Xiang J, Que C, Chen F, Xu G. Occurrence and fate of psychiatric pharmaceuticals in the urban water system of Shanghai, China. Chemosphere. 2015;138:486–93.

    Article  Google Scholar 

  214. Nunes CN, Pauluk LE, dos Anjos VE, Lopes MC, Quinaia SP. New approach to the determination of contaminants of emerging concern in natural water: study of alprazolam employing adsorptive cathodic stripping voltammetry. Anal Bioanal Chem. 2015;407:6171–9.

    Article  Google Scholar 

  215. Ivetic TB, Dimitrievska MR, Fincur NL, Dacanin LR, Guth IO, Abramovic BF, et al. Effect of annealing temperature on structural and optical properties of Mg-doped ZnO nanoparticles and their photocatalytic efficiency in alprazolam degradation. Ceram Int. 2014;40:1545–52.

    Article  Google Scholar 

  216. Tomic N, Grujic-Brojcin M, Fincur N, Abramovic B, Simovic B, Krstic J, et al. Photocatalytic degradation of alprazolam in water suspension of brookite type TiO2 nanopowders prepared using hydrothermal route. Mater Chem Phys. 2015;163:518–28.

    Article  Google Scholar 

Download references

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

F.M1 and F.M2 designed the study, F.M1 and F.M2 analyzed and visualized the data, F.M1 and F.M2 wrote the first draft of the manuscript, F.M1, and F.M2 proofread the manuscript, and F.M1 and F.M2 contributed substantially to revisions.

Corresponding author

Correspondence to Farzaneh Mohamadpour.

Ethics declarations

Ethical approval

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mohamadpour, F., Mohamadpour, F. Photodegradation of six selected antipsychiatric drugs; carbamazepine, sertraline, amisulpride, amitriptyline, diazepam, and alprazolam in environment: efficiency, pathway, and mechanism—a review. Sustain Environ Res 34, 8 (2024). https://doi.org/10.1186/s42834-024-00214-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s42834-024-00214-0

Keywords