Open access peer-reviewed chapter

Recent Overview on the Abatement of Pesticide Residues in Water by Photocatalytic Treatment Using TiO2

Written By

Nuria Vela, Gabriel Pérez-Lucas, José Fenoll and Simón Navarro

Submitted: 30 November 2016 Reviewed: 27 March 2017 Published: 26 July 2017

DOI: 10.5772/intechopen.68802

From the Edited Volume

Application of Titanium Dioxide

Edited by Magdalena Janus

Chapter metrics overview

2,274 Chapter Downloads

View Full Metrics

Abstract

The water bodies’ pollution with phytosanitary products can pose a serious threat to aquatic ecosystems and drinking water resources. The usual appearance of pesticides in surface water, waste water and groundwater has driven the search for proper methods to remove persistent pesticides. Although typical biological treatments of water offer some advantages such as low cost and operability, many investigations referring to the removal of pesticides have suggested that in many cases they have low effectiveness due to the limited biodegradability of many agrochemicals. In recent years, research for new techniques for water detoxification to avoid these disadvantages has led to processes that involve light, which are called advanced oxidation processes (AOPs). Among the different semiconductor (SC) materials tested as potential photocatalysts, titanium dioxide (TiO2) is the most popular because of its photochemical stability, commercial availability, non-toxic nature and low cost, high photoactivity, ease of preparation in the laboratory, possibility of doping with metals and non-metals and coating on solid support. Thus, in the present review, we provide an overview of the recent research being developed to photodegrade pesticide residues in water using TiO2 as photocatalyst.

Keywords

  • titania
  • pesticides
  • water
  • photocatalytic degradation

1. Introduction

Pesticide may be defined as ‘any substance or mixture of substances intended for preventing, destroying, or controlling any pest including vectors of human or animal diseases, unwanted species of plants or animals causing harm during, or otherwise interfering with, the production, processing, storage, or marketing of food, agricultural commodities, wood and wood products, or animal feedstuffs, or which may be administered to animals for the control of insects, arachnids or other pests in or on their bodies. The term includes chemicals used as growth regulators, defoliants, desiccants, fruit thinning agents, or agents for preventing the premature fall of fruits, and substances applied to crops either before or after harvest to prevent deterioration during storage or transport’ [1]. There are many hundreds of them including herbicides, insecticides, fungicides, rodenticides, nematicides, plant growth regulators and others. Pesticide products contain both ‘active’ and ‘inert’ ingredients: active ingredients are the chemicals in pesticide products that kill, control or repel pests. All other ingredients are called ‘inert ingredients,’ which are important for product performance and usability.

Pesticides have been widely applied to protect agricultural crops since the 1940s, and their use increased steadily during the subsequent decades. The Green Revolution was the notable increase in cereal-grain production in many developing countries in the 1960s and 1970s. This tendency resulted from the introduction of hybrid strains of wheat, rice and corn and the adoption of modern agricultural technologies, including irrigation and large doses of agrochemicals, fertilisers and pesticides [2]. However, Rachel Carson (1907–1964) with the publishing of her sensational book Silent Spring in 1962 [3] warned of the dangers to all natural systems from the misuse of some pesticides such as DDT (1,1,1-trichloro-2,2-di(4-chlorophenyl) ethane). As a result, a wider audience was warned of the environmental effects of the widespread use of pesticides, and DDT was banned for agricultural use 10 years later in the USA, and the regulation of chemical pesticide use was strengthened. Currently, all pesticides are subject to strict registration as meaning ‘the process whereby the responsible national government authority approves the sale and use of a pesticide following the evaluation of comprehensive scientific data demonstrating that the product is effective for the purposes intended and not unduly hazardous to human or animal health or the environment’ [4]. In evaluating a pesticide registration application, a wide variety of potential human health and environmental effects associated with their use must be tested. Registrants must generate the necessary scientific information to address concerns corresponding to the identity, composition, potential adverse effects and environmental fate of each pesticide. These data allow evaluating whether a pesticide could harm certain nontarget organisms and endangered species.

Protection of crop losses/yield reduction and increase in food quality are key benefits associated with the use of pesticides in agriculture. However, most organic pesticides characterised as persistent in the environment can bioaccumulate through the food web and can be transported in long distances [5], as evidenced by the accumulation in regions where persistent pesticides have never been used [6]. Persistent organic pollutants (POPs) are chemical compounds that persist in the environment and adversely affect human health and the environment around the world [7]. Because they can be transported by wind and water, most POPs can affect human health and wildlife far from where they are applied. They have high persistence in the environment and can accumulate passing from one species to the next through the food chain. To treat this environmental concern, the USA joined forces with 90 other countries and the European Community to sign the agreement United Nations treaty in Stockholm (Sweden, May 2001). The Stockholm Convention on Persistent Organic Pollutants, approved by Council Decision 2006/507/EC [8], entered into force on 17 May 2004. The aim of the Convention was to protect human health and the environment from POPs. Under the treaty, known as the Stockholm Convention, countries agreed to reduce or eliminate the production, use and/or release of 12 key POPs (‘the dirty dozen’, mainly organochlorine (OC) insecticides) and specified under the Convention a scientific review process that has led to the addition of other POP chemicals of global concern. Currently, there are 30 substances catalogued as POPs including mainly pesticides, industrial chemicals and by-products.

In addition, many pesticides are endocrine-disrupting chemicals (EDCs), compounds that alter the normal functioning of the endocrine system of both wildlife and humans increasing incidence of breast cancer, abnormal growth patterns and neurodevelopmental delays in children, as well as changes in immune function [911]. Most of them are organochlorine (OC) pesticides that affect the reproductive function.

Worldwide consumption of pesticides for agricultural use is constantly increasing as increased human population and crop production, and it has undergone significant changes since the 1960s. Nowadays, the worldwide consumption of pesticides is about 2 million tonnes per year, which 45% is used by Europe alone, 25% is consumed in the USA and 30% in the rest of the world. The proportion of herbicides in pesticide consumption increased rapidly, from 20% in 1960 to 47.5% in 2015. However, the proportion of consumption of insecticides (29.5%) and fungicides/bactericides (17.5%) declined despite their sales increased with other accounts for 5.5% only [12]. The rapid increase of herbicide consumption enhanced agricultural intensification and productivity.

The application of chemical pesticides, in particular the organic-synthesised pesticides, has been a significant mark of human civilisation, which greatly protects and facilitates agricultural productivity. Worlwide, insect pests cause an estimated 14% loss, plant pathogens cause a 13% loss, and weeds causes a 13% loss [13]. Pesticides are so indispensable in agricultural production. About one-third of the agricultural products are produced by using pesticides [14]. Without pesticide application, the loss of fruits, vegetables and cereals from pest injury would reach 78, 54 and 32%, respectively [15]. Ideally, a pesticide must be lethal to the targeted pests, but not to nontarget species, including humans. Unfortunately, this is not the case, so the controversy of use and abuse of pesticides is obvious. Consequently, the risks of using pesticides are serious as well [13]. Most pesticides are not spontaneously generated, and they are toxic to humans and the environment in greater or lesser degree [16]. Pesticides and their degraded products can pass into the atmosphere, soils and water, resulting in the accumulation of toxic substances and thus threatening human health and the environment. In addition, accumulated application leads to loss of biodiversity. Because many pesticides are barely degradable, they persist in soil and pollute surface water and groundwater. Depending on their chemical properties, they can bioaccumulate in food chains and consequently affect human health.

For many pesticides, there is evidence of long-term ubiquity in the aquatic environment at the European Union (EU) level, and therefore they need special consideration as regards their impact on the presentation of chemical status under the European Water Framework Directive (EWFD) [17]. The EWFD establishing a framework for community action in the field of water policy lays down a strategy against the pollution of water. Directive 2008/105/EC [18], amended by Directive 2013/39/EU [19] on environmental quality standards in the field of water policy, lays down environmental quality standards in accordance with the EWFD, for the 33 priority substances. The EWFD also sets out general provisions for the protection and conservation of groundwater. The establishment of detailed quality criteria for the assessment of groundwater chemical status in Europe was laid down in the European Groundwater Directive [20]. For this reason, the EU established the following groundwater quality standards: 0.1 μg L−1 for individual pesticide and 0.5 μg L−1 for the sum of all individual pesticides to safeguard people from harmful effects.

During the 1990s of the last century, atrazine (herbicide) and endosulfan (insecticide) were found most often in surface waters in the USA and Australia due to their widespread use. In addition, although in lower proportions, other pesticides such as pronofos, dimethoate, chlordane, diuron, prometryn and/or fluometuron were detected [21]. Studies that are more recent also reported the presence of several pesticides in environmental waters (surface water, groundwater and seawater) close to agricultural lands over the world [2234]. In addition, different studies have corroborated the presence of some pesticides in drinking water [3538].

Pesticides are being continuously released into the aquatic environment through anthropogenic activities. Their detection in storm and wastewater effluent has been reported to be a major obstacle as regards wide-ranging acceptance of water recycling [39]. In addition, their variety, toxicity and persistence present a threat to humans through pollution of drinking water resources (e.g. surface water and groundwater). The frequent occurrence of pesticides in surface water and groundwater has prompted the search for suitable methods to destroy them. Although conventional biological treatments of water offer some advantages such as their low cost and easy operation, most studies concerning the treatment of pesticides have concluded that they are not very effective due to their low biodegradability [40, 41]. Other technologies such as adsorption or coagulation merely concentrate pesticides by transferring them in other phases but still remain and not being completely eliminated. To solve this problem, apart from reducing emissions, two main water strategies are followed: (i) chemical treatment of drinking water, surface water and groundwater and (ii) chemical treatment of wastewaters containing biocides and bio-recalcitrant pollutants as pesticides. Chemical treatments of polluted surface water, waste water and groundwater are part of a long-term strategy to improve the quality of water by removing toxic compounds of anthropogenic origin before returning the water to its natural cycle. The Directive 2013/39/EU [19] promotes the preventive action and the polluter pays principle, the identification of pollution causes, dealing with emissions of pollutants at the source, and finally the development of innovative wastewater treatment technologies, avoiding expensive solutions. Therefore, effective, low-cost and robust methods to decontaminate waters are needed, as long as they do not further stress the environment or endanger human health, particularly prior to direct or indirect reuse of reclaimed water. In this context, the development of Solar Chemistry Applications is of special relevance, especially photochemical processes where solar photons are absorbed by reactants and/or a catalyst causing a chemical reaction. Consequently, in recent years there has been growing interest in the use of advanced oxidation processes (AOPs) to remove pesticide residues as alternative to methods that are more conventional because they allow the abatement of them by mineralisation.

Advertisement

2. Advanced oxidation processes to remove pesticides from water

AOPs have been commonly defined as near-ambient temperature treatment processes based on highly reactive radicals, especially the hydroxyl radical (OH). Other radicals and active oxygen species involved are superoxide radical anions (O2•−), hydroperoxyl radicals (HO2•−), triplet oxygen (3O2) and organic peroxyl radicals (ROO). In all probability, the OH (E0 = 2.8 V) is among the strongest oxidising species used in water treatment and confers the potential to greatly accelerate the rates of pesticide oxidation. Hydroxyl radicals can degrade indiscriminately micropollutants with reaction rate constants usually around 109 L mol−1 s−1 [42], yielding CO2, H2O and, eventually, inorganic ions as final products. After fluorine (E0 = 3.1 V), OH is the strongest oxidant [43], and its production can be achieved by many pathways, which allows one to choose the appropriate AOP according to the specific characteristics of the target water/wastewater and treatment requirements. Regarding the methodology to generate hydroxyl radicals, AOPs can be divided into chemical, electrochemical, sono-chemical and photochemical processes. Typical AOPs can be also classified as homogeneous that occur in a single phase and heterogeneous processes because they make use of a heterogeneous catalyst like metal-supported catalysts, carbon materials or semiconductors such as TiO2, WO3, ZnO, CdS, SnO2, ZnS and others [44]. Figure 1 shows some homogeneous and heterogeneous processes [41, 45].

Figure 1.

Scheme for conventional AOPs.

When the chemical process (mineralisation) destroys the contaminants and their reaction intermediate products (metabolites), critical secondary wastes are not generated and, thus, post-treatment or final disposal is not required [46]. However, if complete mineralisation is not achieved or the reaction period is too long, a final post-treatment may be necessary. A higher biodegradability and lower toxicity of the reaction by-products, in comparison with the parent compounds, are desirable benefits of applying AOPs to treat wastewaters. However, in some cases, these by-products are less biodegradable and/or more toxic than the parent compounds. For this reason, AOPs can be applied as post- or pretreatment of biological processes. The integration of different AOPs in a sequence of treatment processes is a common approach to achieve a biodegradable effluent, which can be further treated by a conventional biological process, reducing the residence time and reagent consumption in comparison with AOPs alone. Figure 2 shows possible integration of AOPs in wastewater and drinking water treatment plants.

Figure 2.

Possible implementation of AOPs in waste water and drinking water treatment plants (adapted from Petrovic et al. [47].

One drawback of these processes is the presence of scavengers in wastewaters because these species consume OH, competing with pesticides. They can be organic matters (e.g. humic and/or fulvic acids, amino acids, proteins and carbohydrates) or inorganic ions (CO3=, HCO3, S=, Br, NO3 and others). Because most natural waters contain these scavengers, optimisation of AOPs must be performed bearing them in mind. The knowledge of their effect on the process efficiency is difficult because they have different reactivities, as well as due to the constant variations in the aqueous phase when the parent pollutants are continuously transformed into many different intermediates. Among these techniques, photocatalytic methods in the presence of artificial or solar light, like heterogeneous photocatalysis (HP), have been proven very effective for the degradation of a wide range of pesticides [48].

2.1. Basis of heterogeneous photocatalysis

Photocatalysis may be defined as the acceleration of a photoreaction by the presence of a catalyst. HP, the use under irradiation of a stable solid semiconductor for stimulating a reaction at the solid/solution interface, is a technique of environmental interest for the treatment of pesticide-polluted water combining the low cost, the mild conditions and the possibility of using natural sunlight as the source of irradiation [39, 49]. Progress and challenges of HP can be reviewed in recent published papers [50, 51]. In brief, HP is based on the irradiation of semiconductor (SC) particles, usually suspended in aqueous solutions, with wavelength energy hvEg (band-gap energy). Thus, an electron (e) is driven to the conduction band (cb), remaining a positive hole (h+) in the valence band (vb). Both the e and h+ migrate to the particle surface (Figure 2). The ecb and the h+vb can recombine on the surface or in the bulk of the particle in short time, and the energy dissipated as heat. In addition, they can be trapped on the surface reacting with donor (D) or acceptor (A) species adsorbed or close to the surface of the particle [52] as can be seen in Figure 3. The wavelength (λ) of radiation required to activate the catalyst must be equal or lower than the calculation by Planck’s equation, λ = hc/Eg, where h is Planck’s constant (6.626176 × 10−34 J s−1), c is the speed of light and Eg is the semiconductor band-gap energy.

Figure 3.

Scheme for the heterogeneous photocatalysis.

As a rule, TiO2 is considered the best photocatalyst due to different qualities such as high photochemical stability, high efficiency, non-toxic nature and low cost, whose behaviour is very well documented in the recent literature [53, 54]. Excellent reviews have been published during the last years on the photoactivity of TiO2 to purify pesticide-polluted waters [39, 48, 5560].

2.2. Properties and characteristics of titanium dioxide photocatalysts

Titanium dioxide is by far, the most investigated photocatalyst to remove organic pollutants from water. The photocatalytic activity of TiO2 slurries depends on physical properties of the catalyst (crystal and pore structure, surface area, porosity, band gap, particle size and surface hydroxyl density) [61, 62]. On the other hand, operating conditions such as light intensity and wavelength, initial concentration and type of pollutants, catalyst loading, oxygen content, interfering substances, presence of oxidants/electron acceptor, pH value and configuration of photoreactor have a key role [39, 58, 6366]. Finally, the mode of TiO2 application (suspended, immobilised or doped) is fundamental to rate the photocatalytic activity.

2.2.1. Composition of TiO2 and its types

Titanium dioxide is known to occur in nature as anatase (At), brookite (Bk) and rutile (Rl) (Figure 4). Rl is usually considered to be the high-temperature and high-pressure phase relative to At, whereas Bk is often considered to be of secondary origin. Rl is the most common, most stable and chemically inert and can be excited by both visible and ultraviolet (UV) light (wavelengths smaller than 390 nm) [67]. At is only excited by UV light and can be transformed into Rl at high temperatures. Both Rl and At have a tetragonal ditetragonal dipyramidal crystal system but have different space group lattices. UV light does not excite Bk, but its orthorhombic crystal system can be transformed into Rl with the application of heat.

Figure 4.

(1) The schematic conventional cells for anatase (a), rutile (b) and brookite (c) phases [73]; (2) SEM image of TiO2 P25 and (3) XRD pattern of TiO2 P25.

Generally, At exhibits higher photocatalytic activities than Rl. However, the reasons for the differences in photocatalytic activity between At and Rl are still being debated. Although At has lower absorbance ability to solar light than Rl because its band gap is larger (3.2 eV) than that of Rl (3.0 eV), the photocatalytic activity of At is higher than that of Rl. This can be explained because At has a higher surface adsorption capacity to hydroxyl groups and a lower charge carrier recombination rate than Rl [68]. Also, the lower photocatalytic activity of Rl is due to its larger grain size [69], lower specific surface areas and less capacity for surface adsorption. In addition, the lifetime of photo-generated e and h+ in Rl is about an order of magnitude smaller than that of At. Consequently, the chance of participation of photoexcited e- and h+ of At in surface chemical reactions is greatly enhanced. According to Zhang et al. [70], Rl and Bk belong to the direct band-gap SC category, while At appears to be an indirect band-gap SC. As a consequence, At exhibits a longer lifetime of photoexcited e and h+ than Rl and Bk because the direct transition of photo-generated electrons from the conduction band to valence band of At is not possible. Moreover, At has the lightest average effective mass of photo-generated e and h+ than Rl and Bk suggesting the fastest migration of photo-generated e and h+ from the interior to the surface of At. This results in the lowest recombination rate of photo-generated charge carriers within At. Therefore, it is not surprising that Rt and Bk show a smaller photocatalytic activity than At.

TiO2 can use natural sunlight because it has an appropriate energetic separation between its valence and conduction bands which can be surpassed by the energy content of a solar photon (380 nm > λ >300 nm). The sunlight puts 0.2–0.3 mol photons of m−2 h−1 in the 300–400 nm range with a typical UV flux of 20–30 W m−2. In addition, photons can be generated by artificial irradiation although it is the most important source of costs during the treatment of wastewater [65]. Different available TiO2 catalysts (with different surface areas, crystal sizes and compositions) such as Degussa P25, Hombikat UV100, PC500, PC10, PC50, Rhodia and others have been tested for the photolytic degradation of pesticides in aqueous environments [39]. From them, Degussa P25 has been the most used because it has good properties (i.e. typically a 70:30 At:Rl composition, non-porous, Brunauer, Emmett and Teller (BET) around 55 m2 g−1 and average particle size 30 nm) and a substantially higher photocatalytic activity than other commercial TiO2 [65]. The higher photocatalytic activity of P25 has been attributed to its crystalline composition of Rl and At. It is known that the smaller band gap of rutile absorbs the photons and generates e/h+ pairs. Then, the electron transfer takes place from the rutile to electron traps in the At phase. Thus, the recombination is inhibited and allows the hole to move to the surface of the particle to react [71].

2.2.2. Operating conditions

The e/h+ formation in the photochemical reaction is strongly dependent on the light intensity at a given wavelength [72]. Therefore, the dependency of pollutant degradation rate on the light intensity has been studied in numerous investigations of various organic pollutants. According to Herrmann [73], the reaction rate is proportional to the radiant flux (φ) < 25 mW cm−2, while above this value, the rate varies as φ½, which indicates a too high value of the flux increasing the e/h+ recombination rate. When the intensity is high, the reaction rate does not depend on light intensity because at low intensity, reactions involving e/h+ formation are predominant, while e/h+ recombination is not significant [74].

Several authors have indicated that when the level of the target pesticide increases, a large number of molecules of the compound are adsorbed on the photocatalyst surface and, consequently, the reactive species (OH and O2•−) required for pesticide degradation also increase. However, the formation of OH and O2•− on the catalyst surface remains constant for a given catalyst amount, light intensity and irradiation time. Hence, at higher concentrations the available OH is inadequate for pollutant degradation. Therefore, as the concentration increases, the pollutant degradation rate decreases [75]. In addition, an increase in pesticide concentration leads to the generation of intermediates (metabolites), which may be adsorbed on the surface of the catalyst.

Because TiO2 is often used as suspension, the photocatalytic degradation rate initially increases with catalyst loading and then decreases at high concentrations due to light scattering and screening effects. Although the number of active sites in solution will increase with catalyst loading, light penetration is compromised because of excessive particle concentration. The interaction between particles (agglomeration) increases at high concentration, and, consequently, the surface area available for light adsorption is reduced and photocatalytic activity decreases. The optimum catalyst loading has to be found in order to avoid excess catalyst and ensure the maximum absorption of photons. Although the results in the literature are very different, it may be deduced that the incident radiation and path length inside the photoreactor are of special interest in determining the optimum catalyst mass.

Waste water, surface water, groundwater and drinking water pHs vary significantly and play an important role in the photodegradation of pesticides since it determines the size of aggregates it forms and the surface charge of the photocatalyst. The surface charge of the photocatalyst and the ionisation or speciation (pKa) of a pesticide can be seriously affected by the solution pH. Electrostatic interaction between the surface of the semiconductor, substrate, solvent molecules and radicals formed during photocatalytic oxidation strongly depends on the solution pH. At pH below its pKa value, an organic compound exists as neutral state. Above this pKa value, organic compounds attain a negative charge, which can significantly influence their photocatalytic degradation. The point of zero charge (PZC, the pH at which the surface has a neutral net electrical charge) of TiO2 is not very sensitive to the crystallographic structure (At vs Rl) and the experimental method used. The common value is pH 5.9 for both phases. Although the PZC of TiO2 depends on the production method, the most frequent value for TiO2 P25 is 6.3 [76]. Below or above this value, the charge of the catalyst surface is positive or negative, respectively, according to the following reactions:

pH<pzcTiOH+H+TiOH2+,pH>pzcTiOH+OHTiO+H2OE1

At low pH, the positive holes are considered as the major oxidation step, whereas at neutral or high pHs, OH is the predominant species [77]. It is expected that the generation of OH will be higher due to the presence of more available OH on the TiO2 surface. Thus, the degradation efficiency of the process will be enhanced at high pH. A very important feature of the photocatalytic process is that in many cases, a great number of different metabolites are produced, which may behave in a different way depending on the pH of the solution. As regards temperature, photocatalytic systems do not require heating and operate at room temperature because of photonic activation [65].

The e/h+ recombination is one of the main drawbacks in the application of semiconductor photocatalysis as it causes waste of energy. In the absence of suitable electron acceptor, recombination step is predominant, and thus, it limits the quantum yield [39]. In HP reactions, O2 is generally used as electron acceptor. Addition of exogenous oxidant/electron acceptors into a semiconductor suspension has been shown to improve the photocatalytic degradation of many pesticides because they can eliminate the e/h+ recombination by accepting the conduction band electron, increase the OH concentration and oxidation rate of intermediate compound and produce more radicals and oxidising species to accelerate the degradation efficiency of intermediate compounds. Because OH plays an important role in photodegradation, several researchers have investigated the effect of addition of different electron acceptors (i.e. H2O2, KBrO3 or Na2S2O8) on the photocatalytic degradation of many pesticides [75, 78].

The ability of peroxydisulfate is not only attributed to the promotion of charge separation but also to the production of sulphate radicals (SO4•−), which are very strong oxidising agents (E0 =2.6 V), and the appearance of more hydroxyl radicals (OH) according to the following reactions:

S2O8=+eSO4=+SO4E2
SO4+eSO4=E3
SO4+H2OSO4=+OH+H+E4
SO4+RHIntermediatesSO4=+CO2E5

Besides, the addition of H2O2 enhances the degradation due to the increase in the OH concentration as follows:

H2O2+eOH+OHE6
H2O2+O2OH+OH+O2E7
H2O2+hv2OHE8

The quantum yield of S2O82− (1.8 mol Einstein−1) is much larger than that of H2O2 (1 mol Einstein−1), which can be related to the rate of recombination of OH (5.3×109 M−1 s−1) and SO4•− (8.1×108 M−1 s−1) [79].

As reviewed by Ahmed et al. [39], many studies have demonstrated that water components like Ca2+, Mg2+, Fe2+, Zn2+, Cu2+, HCO3, PO43−, NO3, SO42− and Cl and dissolved organic matter (DOM) can affect the photodegradation rate of organic pollutants since they can be adsorbed onto the surface of TiO2 [80, 81]. These dissolved components can compete with the pesticide for the active sites depending on the solution pH, reducing the formation of OH. Anions result in corresponding anion radicals scavenged by OH. However, they have lower oxidation potential. In addition, DOM, ubiquitously present in storm and wastewater effluent, also plays an important role regarding pesticide degradation. The observed slowdowns are related to the inhibition (surface deactivation), competition and light attenuation effects. Moreover, the presence of humic acids in the reaction solution has been reported to significantly reduce light transmittance and consequently the photooxidation rate.

As usual, two types of photoreactors are used for photocatalytic wastewater treatment processes: (i) slurry photoreactors and (ii) fixed-bed photoreactors [82]. The slurry photoreactors utilise suspended photocatalyst particles, while the other type utilises immobilised photocatalyst particles on a surface. Currently, the catalysts are applied in the form of a slurry in most cases. However, the separation of catalyst after the reaction in the slurry systems is an expensive and tedious stage, which adds to the overall running costs of the plant. Therefore, immobilised catalyst systems are preferred in order to avoid an increase in cost and time. However, the immobilised catalyst systems have low interfacial surface areas and consequently a very low activity, being also difficult to scale [83]. It is noted that configuration of photoreactor has an important role in the efficiency of the photocatalytic wastewater treatment processes.

2.2.3. Bare, doped and immobilised application of TiO2

Generally, the use of TiO2 slurries has been demonstrated to have higher photocatalytic activity as compared to the same immobilised catalyst. This is due to changes on the surface of the catalyst by blocking pores and the appearance of by-products causing the loss of active sites on its surface. Usually, TiO2 is prepared in the form of nanopowders, crystals, thin films, nanotubes and nanorods. As a rule, the immobilisation of TiO2 onto supporting material has been carried out via one of two major routes: (i) physical (thermal treatment) and (ii) chemical (sol-gel, electrodeposition, etc.). The evolution of different supports and the benefits and drawbacks of various immobilisation techniques to obtain a high-surface-area TiO2 can be seen in the reviews by Shan et al. [84] and Dahl et al. [85]. Nonetheless, the interest for the development of TiO2 supported on different materials is growing because the use of the bare TiO2 phases presents some drawbacks as (i) necessity of irradiation with UV light due to the small amount of photons absorbed in the Vis region, (ii) high recombination rate for the photoproduced e/h+ pairs, (iii) difficulty to improve the performance by doping with some materials that often act as recombination centres, (iv) deactivation in the absence of H2O vapour when aromatic molecules must be abated and (v) difficulty to support powdered TiO2 on some materials [45]. Consequently, the research line in HP has been driven to modify some electronic and morphological properties of TiO2 to enhance its photoefficacy. In this context, nanosized particles and films on glasses or other supports and powdered samples with high specific surface areas have been obtained to increase the possibility for the involved species to avoid the separation step [86]. Doping, loading and sensitisation of TiO2 are methods aimed to shift the light absorption towards visible light and/or to increase the lifetime of the photoproduced e/h+ pairs.

A number of approaches have been suggested in recent years to enhance photocatalytic activity of TiO2 in the visible light region for its use in water detoxification [87]. Metal ion doping and co-doping with non-metals can improve trapping of the photoexcited conduction band electrons at the surface, thereby minimising charge carrier recombination. Several dopants used (e.g. Sn, Ag, Pd, Re, Bi, V, Mo, Th or Pt among others), have been shown to enhance photocatalytic activity for the systems examined. However, the photoactivity of the metal-doped TiO2 photocatalyst significantly depends on the dopant ion nature and concentration, preparation method and operating conditions [88]. The deposition of metal ions on TiO2 can modify the photoconductive properties by increasing the charge separation efficiency between electrons and holes, which will enhance the formation of both free hydroxyl radicals and active oxygen species [89]. Recent research indicates that the desired narrowing on the band gap of TiO2 can be achieved using non-metal elements such as N, F, S and C. Thus, modified TiO2 showed a significant improvement on the absorption in the Vis light region due to band-gap narrowing and enhancing the degradation of pesticides under Vis light irradiation, mainly under sunlight [90, 91]. Also, it is possible to produce coupled colloidal structures using TiO2 such as TiO2-SnO2, TiO2-CdS, TiO2-Bi2S3, TiO2-WO3 or TiO2-Fe2O3, in which illumination of one semiconductor produces a response in the other at the interface between them by increasing the charge separation and extending the energy range of photoexcitation. On the other hand, the coating of one semiconductor or metal nanomaterial on the surface of another semiconductor or metal nanoparticle core is called capping. Semiconductor nanoparticles of TiO2 can be coated with another semiconductor (i.e. SnO2) with a different band gap to enhance its emissive properties [92].

2.3. Heterogeneous photocatalytic degradation of pesticide residues in water over titanium dioxide

A desirable feature in the photodegradation of pesticides in water is the transformation of the parent compounds in order to avoid their toxicity. However, the main objective is the mineralisation of the pesticides. As previously commented, OH is the main species involved for organic substrate oxidation, but the free radical HO2 and its conjugate O2•− also play an important role although those radicals are much less reactive than OH. All these free radicals react with pesticides by hydrogen abstraction or electrophilic addition to double bonds. Further, the radicals react with O2 to give organic peroxyl radicals (ROO) initiating different oxidative reactions that may lead to the complete mineralisation of the pesticides. Since OH is non-selective, numerous and different transformation products (intermediates) can be formed at low concentrations being in certain cases more persistent and toxic than the parent compounds.

Mineralization:OrganicpesticideIntermediatesCO2+H2O+Cl+NO3+SO42+PO43etc.E9

The presence of pesticide residues in water is usually monitored using chromatographic techniques such as gas chromatography (GC) and liquid chromatography (LC) coupled to mass spectrometry (MSn) and time-of-flight (TOF) detection systems. However, since identification of all the transformation products generated during the photooxidation is not possible, the measure of total organic carbon (TOC) and more specifically dissolved organic carbon (DOC) is crucial in the process because determination of CO2 must be stoichiometric with the organic carbon in the parent pesticide. This determination can be carried out in a simple and rapid way to know the mass balance and the remaining amount of metabolites. As example, Figure 5 shows the photodegradation pathways proposed for chlorantraniliprole [93] and tebuconazole [94] dissolved in water when illuminated in the presence of TiO2.

Figure 5.

Photometabolic pathways proposed for chlorantraniliprole (insecticide) and tebuconazole (fungicide) in water slurries when illuminated in the presence of TiO2.

In the case of chlorantraniliprole, a new class of anthranilic diamide insecticide, several transformation products are generated during irradiation as a result of different reactions such as hydroxylation, deamination, hydrolysis of the amide bridge and rearrangement followed by cyclisation, methyl amine transfer and fragmentation. For tebuconazole, a common triazole fungicide with numerous agricultural and urban uses, the transformation pathway was found to proceed through tert-butyl chain cleavage, hydroxylation, oxidation and dechlorination and showed that its degradation mechanism was mainly driven by OH and h+.

The use of HP has showed a powerful growth in recent years. Today, pesticides constitute an important group concerning pollutant treatment. The large number of papers published in the last years proves this interest. A review to the literature extracted from the Web of Science™ (formerly ISI Web of Knowledge, www.isiknowledge.com) managed by Thomson Reuters (Philadelphia, USA) using the following keywords, TiO2, pesticides and water, shows 463 papers only in the period 2005−2016.

Table 1 shows some of the most popular journal publishing on the topic “photocatalysis and pesticides” according to the following criteria: impact factor > 2 and Eigenfactor score > 0.01 (Journal Citation Reports (JCR) Science Edition 2015).

Journal title (ISO)Editorial/countryISSN aIF b5Y-IF cES d
Chemical ReviewsAmerican Chemical Society (USA)0009-266537.36951.5600.24503
Journal of Photochemistry and Photobiology C: Photochemistry ReviewsElsevier Science BV (Japan)1389-556712.16215.2680.00442
Applied Catalysis B: EnvironmentalElsevier Science BV (The Netherlands)0926-33738.3288.1420.05714
Journal of CatalysisAcademic Press INC Elsevier Science (USA)0021-95177.3547.4820.03438
Water ResearchPergamon-Elsevier Science Ltd. (England)0043-13545.9916.7690.08394
Environmental Science and TechnologyAmerican Chemical Society (USA)0013-936X5.3936.3960.20232
Chemical Engineering JournalElsevier Science SA (Switzerland)1385-89475.3605.4390.09715
Catalysis Science and TechnologyRoyal Society of Chemistry (England)2044-47535.2875.5470.02367
Journal of Hazardous MaterialsElsevier Science BV (The Netherlands)0304-38944.8365.6410.09715
Solar Energy Materials and Solar CellsElsevier Science BV (The Netherlands)0927-02484.7325.0160.04275
Catalysis TodayElsevier Science BV (The Netherlands)0920-58614.3124.1050.03376
Applied Catalysis A: GeneralElsevier Science BV (The Netherlands)0926-860X4.0124.4030.03476
Science of the Total EnvironmentElsevier Science BV (The Netherlands)0048-96973.9764.3170.08394
Journal of Molecular Catalysis A: ChemicalElsevier Science BV (The Netherlands)1381-11693.9584.0450.01727
ChemospherePergamon-Elsevier Science Ltd. (England)0045-65353.6984.0680.06483
Solar EnergyPergamon-Elsevier Science Ltd. (USA)0038-092X3.6854.4140.02277
Journal of Environmental ManagementAcademic Press-Elsevier Science Ltd. (England)0301-47973.1314.0490.03326
CatalystsMDPI AG (Switzerland)2073-43442.9643.1940.16915
Journal of Agricultural and Food ChemistryAmerican Chemical Society (USA)0021-85612.8573.3080.08776
Environmental Science and Pollution ResearchSpringer Heidelberg (Germany)0944-13442.7602.8760.02617
Journal of Chemical Technology and BiotechnologyWiley-Blackwell (England)0268-25752.7382.7440.01051
Journal of Photochemistry and Photobiology A: ChemistryElsevier Science SA (Switzerland)1010-60302.4772.5730.01025
Photochemical and Photobiological ScienceRoyal Society of Chemistry (England)1474-905X2.2352.6730.01040
Journal of Environmental ScienceScience Press (Mainland China)1001-07422.2082.6990.01282

Table 1.

Some of the main and influential journals of interest to readers where the authors usually publish their works about ‘photocatalytic degradation of pesticides in water’ (source: JCR, 2015).

aInternational Standard Serial Number.


bImpact factor.


c5-Year impact factor.


dEigenfactor score.


The Journal Impact Factor is defined as all citations to the journal in the current JCR year to items published in the previous 2 years, divided by the total number of scholarly items (these comprise articles, reviews and proceeding papers) published in the journal in the previous 2 years. The Eigenfactor score calculation is based on the number of times articles from the journal published in the past 5 years have been cited in the JCR year, but it also considers which journals have contributed these citations so that highly cited journals will influence the network more than lesser cited journals. References from one article in a journal to another article from the same journal are removed, so that Eigenfactor scores are not influenced by journal self-citation.

Following the criteria above, some of the most representative publications are presented in this summary (Table 2). Results show that sunlight photoalteration (photolysis) processes are well now to play an important role in the degradation of pesticides and other contaminants in the aquatic environment. These technologies allow the removal of pesticides by mineralisation. When the exciting energy used comes from the Sun, the process is called solar photocatalysis [95]. Photocatalytic oxidation by semiconductor oxides is an area of environmental interest for the treatment of polluted water, particularly relevant for Mediterranean agricultural areas, where solar irradiation is highly available making this process quite attractive. An ideal photocatalyst is characterised by photostability, biologically and chemically inert nature, low cost and availability and capability to adsorb reactants under efficient photonic activation. Due to these characteristics, titanium dioxide (TiO2) has been demonstrated to be an excellent catalyst, and its behaviour is very well documented for the photodegradation of pesticide residues in water.

PesticidesPhotocatalystsLight sourceMain findingsReferences
PyrimethanilTiO2 P25 and home madeSolarSimilar photoefficiences in mineralisation[81]
Acephate, omethoate, methyl parathionFe3O4@SiO2@mTiO2 nanomicrospheresUVDisappearance after 45–80 min[96]
ImidaclopridTiO2 and TiO2-based hybridSolarTiO2 + fly ash is 2–3 times less active than sol-gel TiO2[97]
Spirotetramat, spirodiclofen, spiromesifenZnO, TiO2 P25, TiO2 Kronos vlp 7000, Zn2TiO4 and ZnTiO3UVZnO > TiO2 P25 > TiO2 Kronos vlp 7000 > Zn2TiO4 > ZnTiO3[98]
FlubendiamideZnO/Na2S2O8 and TiO2 P25/Na2S2O8UVZnO and TiO2 oxides strongly enhance the degradation rate[99]
Thiamethoxam, imidacloprid, acetamipridZnO/Na2S2O8 and TiO2 P25/Na2S2O8UV/solarSolar irradiation was more efficient compared to artificial light for the removal of these insecticides (t½ = 0.3–2 min)[100]
ChlorantraniliproleZnO, TiO2 P25, TiO2 Kronos vlp 7000, Zn2TiO4UVHalf-lives of 53 and 71 min for ZnO/Na2S2O8 and TiO2/Na2S2O8 systems, respectively[93]
Metamitron, metribuzinDifferent TiO2 nanopowdersUVt½ lower than 8 min using TiO2 P25[101]
DichlorvosTiO2 supported on zeolitesUVComplete degradation in 360–540 min[102]
AtrazineN-Doped TiO2 supported by phosphorsUVEnhanced performance compared to either pure N-TiO2 nanoparticles or bare phosphors microparticles[103]
Spinosad, indoxacarbZnO/Na2S2O8 and TiO2 P25/Na2S2O8UV95% degradation rate was observed for both spinosyns after 2 min when using Na2S2O8[104]
MethabenzthiazuronZnO, TiO2 P25UVHalf-lives of 2 and 7 min for ZnO and TiO2, respectively[105]
Monocrotophos, endosulfan, chlorpyriphosTiO2 coated on polymeric beadsSolarRapid photodegradation using immobilised bead photoreactor[106]
IsoproturonTiO2 coated on inert cement beadsUV85% degradation rate after 6 h[107]
DiazinonTiO2 doped with N, NS and FeFNSUV LEDThe FeFNS-doped TiO2 was found to be an efficient catalyst (96% degradation after 100 min)[108]
Methyl oxidemeton, methidathion, carbaryl, dimethoatheTiO2 P25, Hombikat UV 100, Millennium PC (50, 100, 105), Kronos 7101UVTiO2 P25 was the most effective for pesticide degradation. Removal of pesticides in less than 300 min[109]
30 sulfonylurea herbicidesZnO and TiO2 P25 in tandem with Na2S2O8UVAverage time required for 90% degradation about 3 and 30 min for ZnO/Na2S2O8 and TiO2/Na2S2O8 systems, respectively[110]
Chlorotoluron, diuron, fluometuron, isoproturon, linuronZnO, TiO2, WO3, SnO2 and ZnSSolarThe time required for 90% degradation ranged from 23 to 47 min for isoproturon and linuron, respectively, when using the tandem ZnO/Na2S2O8[111]
CarbofuranZnO, TiO2 P25 DegussaSolarHalf-lives ranging from 6 to 385 min[112]
MalathionWO3/TiO22% WO3/TiO2 exhibited the best photocatalytic activity achieving abatement of 76% TOC after 300 min[113]
Ethoprophos, isoxaben, metalaxyl, metribuzin, pencycuron, pendimethalin, propanil, tolclofos-methylZnO, TiO2, WO3, SnO2, ZnSSolarHalf-lives for ZnO ranged from 9 to 38 min (t30W 0.3–1.7 min), while in the presence of TiO2, ranged from 41 to 260 min (t30W 1.9–16.3 min)[114]
Simazine, prometryn, terbutryn, atrazine, terbuthylazine, propachlor, S-metolachlor, alachlorZnO, TiO2SolarDegradation rate >70% after 240 min in the ZnO/Na2S2O8 system[115]
FenamiphosZnO, TiO2, WO3, SnO2SolarHalf-life <3 min for ZnO, TiO2 in tandem with Na2S2O8[116]
Cyprodinil, fludioxonilZnO, TiO2SolarDT75, referred to the normalised illumination time (t30W), was lower than 40 and 550 min (t30W = 2 and 40 min) for both fungicides using ZnO and TiO2, respectively[117]
Methyl parathion, dichlorvosN-doped and P25 TiO2UV/solarN-doped TiO2 showed higher photocatalytic activity under solar radiation compared to UV and visible light[118]
MCPA, clopyralid, mecopropFe- and N-doped TiO2UVLowering of the band gap of titanium dioxide by doping is not always favourable for increasing photocatalytic efficiency of degradation[119]
2,4-D, Diuron, ametryneTiO2 slurrySolarResults showed that the overall model fitted the experimental data of herbicides mineralisation in the solar CPC reactor satisfactorily for both cloudy and sunny days[120]
CarbendazimTiO2 slurryUVMore than 90% of fungicide was degraded after 75 min[121]
GlyphosateTiO2 slurryUVAddition of Fe3+, Cu2+, H2O2, K2S2O8 or KBrO3 enhances the photodegradation[122]
Alachlor, atrazine, chlorfenvinphos, diuron, isoproturon, pentachlorophenolTiO2 slurry and Fe2+SolarPhoto-Fenton treatment was found to be shorter than TiO2 and more appropriate for these compounds[123]
Triclopyr, dantinozidDifferent types of TiO2 slurries in tandem with electron acceptorsSolarThe photocatalyst Degussa P25 was found to be more efficient as compared with other photocatalysts[124]
Cymoxanil, methomyl, oxamyl, dimethoate, pyrimethanil, teloneTiO2 slurry and Fe2+SolarTotal disappearance of the parent compounds and nearly complete mineralisation were attained with all pesticides tested[125]
DimethoateZnO, TiO2UVBoth catalysts were unable to mineralise dimethoate, but the addition of oxidants improved the efficiency of the processes[126]
Dichlorvos, monocrotophos, parathion, phorateTiO2·SiO2 beadsSolarAfter 420 min illumination by sunlight, 0.65 × 10−4 mol dm−3 of four organophosphorus pesticides can be completely photocatalytically degraded into PO43−.[127]

Table 2.

A brief summary of recent research studies in which TiO2 was used for treating pesticide-polluted waters in the period 2005–2016.

2.4. Environmental impact and treatment cost

As previously stated, there is a very extensive literature (at laboratory and pilot plant scale) on the photocatalytic degradation of organic pollutants in water. However, there are not many works devoted to the study of the impact of the process from an environmental and economic point of view. In this context, the well-known life cycle impact assessment (LCIA) tool has been successfully used to assess the environmental impact of chemical processes. This tool finds the potential impacts associated with the entire life cycle of a product or a process. This methodology has been devised to study and compare processes at the industrial level but can be perfectly used at the laboratory level. On the other hand, there are several methods of estimating the costs of implementing each of the different AOPs. In general, the following items are proposed: (i) facility cost, (ii) project contingency, (iii) engineering project and (iv) replacement costs. The sum of these four concepts is the total installed cost, based on which the yearly economic impact can be evaluated. Then, operating costs have to be calculated. These costs are normally yearly and consist of the following items: (i) personnel, (ii) maintenance, (iii) electricity and (iv) materials and services. These costs added to the annual facility costs are the total annual costs [128].

Advertisement

3. Conclusion

Currently, population growth, technology development, inadequate agricultural practices and land use have created unprecedented water pollution problems. Agricultural wastewater is characterised by high organic matter content and traces of organic pollutants, mainly pesticides. In addition, effluents from the agro-food and other industries have a potentially environmental risk that requires appropriate and comprehensive treatment. The most often used methods for the treatment of the industrial wastewaters, including membrane filtration, chemical coagulation/flocculation, ion exchange, precipitation, adsorption, biological degradation and ozonation, are not efficient enough and have important limitations to remove bio-recalcitrant compounds as many pesticides. In recent years, advanced oxidation processes (AOPs) have been proposed as a most promising way for degradation of various pollutants. Among them, heterogeneous photocatalysis technology is an interesting route among AOPs, which can be conveniently used for the complete degradation of different hazardous compounds including pesticides. Results show that sunlight photoalteration processes are well now to play an important role in the degradation of pesticides and other contaminants in water. These technologies allow pesticides to be removed by mineralisation. Photocatalytic oxidation by semiconductor oxides is an area of environmental interest for the treatment of polluted water, particularly relevant for Mediterranean agricultural areas, where solar irradiation is highly available (more than 2800 h of sunshine per year on average) making this process quite attractive. Due to its specific characteristics, titanium dioxide (TiO2) has been demonstrated to be an excellent catalyst, and its behaviour is very well documented for the photodegradation of pesticide residues in water. Recently, many authors have also developed combined AOP and biological systems to implement in wastewater treatment plants. In addition to the experimental and modelling work, the aspect most lacking of this combination systems for the treatment of bio-recalcitrant specific industrial wastewater is the performance of complete economic studies, which could present this innovative technology as a cost-competitive one.

Advertisement

Acknowledgments

The authors acknowledge the financial support received from the European Commission through the LIFE+ program (LIFE 13 ENV/ES/000488) and the EU’s funding instrument for the environment and climate action.

Advertisement

Abbreviations

POPsPersistent organic pollutants
EDCsEndocrine-disrupting chemicals
OCOrganochlorine
EWFDEuropean Water Framework Directive
AOPsAdvanced oxidation processes
HPHeterogeneous photocatalysis
EgBand-gap energy
DDonor
AAcceptor
cbConduction band
e−Electron
h+Hole
vbValence band
SCSemiconductor
AtAnatase
BkBrookite
RlRutile
SEMScanning electron microscopy
XRDX-ray diffraction
UVUltraviolet
BETBrunauer, Emmett and Teller
PZCPoint of zero charge
TOCTotal organic matter
DOMDissolved organic matter
GCGas chromatography
LCLiquid chromatography
MSMass spectrometry
TOFTime of flight
JCRJournal Citation Reports
LCIALife cycle impact assessment
EUEuropean Union
ISSNInternational Standard Serial Number
IFImpact factor
5Y-IF5-year impact factor
ESEigenfactor score

References

  1. 1. Food and Agriculture Organization (FAO) of the United Nations. International Code of Conduct on the Distribution and Use of Pesticides. [Internet]. 2003. Available from: ftp://ftp.fao.org/docrep/fao/005/y4544e/y4544e00.pdf [Accessed: 24-01-2017]
  2. 2. Briggs, J. Green revolution. In: Kitchin R, Thrift N, editors. International Encyclopedia of Human Geography. Amsterdam: Elsevier; 2009. pp. 634-638
  3. 3. Carson, R. Silent Spring. Anniversary edition (2002 and 2012). Boston: Houghton Mifflin Harcourt; 2002. 400 p
  4. 4. Food and Agriculture Organization (FAO) of the United Nations. Pesticides Registration Legislation. [Internet]. 1995. Available from: http://www.fao.org/docrep/012/T0553E/T0553E.pdf [Accessed: 24 January 2017]
  5. 5. Shen L, Wania F, Lei YD, Teixeira C, Muir DCG, Bidleman TF. Atmospheric distribution and long-range transport behavior of organochlorine pesticides in North America. Environmental Science & Technology. 2005;39:409-420
  6. 6. Dietz R, Riget FF, Sonne C, Letcher R, Born EW, Muir, DCG. Seasonal and temporal trends in polychlorinated biphenyls and organochlorine pesticides in East Greenland polar bears (Ursus maritimus), 1990-2001. Environmental Science & Technology. 2004;331:107-124
  7. 7. Mehemetli E, Koumanova B. The Fate of Persistent Organic Pollutants in the Environment. Dordrecht: Springer; 2008. 471 p
  8. 8. European Community. Council Decision (2006/507/EC) of the 14 October 2004 concerning the conclusion of the Stockholm convention on persistent organic pollutants. Official Journal of the European Union. 2004;L 209:1-2
  9. 9. Mnif W, Hassine AIH, Bouaziz A, Bartegi A, Thomas O, Roig B. Effect of endocrine disruptor pesticides: A review. International Journal of Environmental Research and Public Health. 2011;8:2265-2303
  10. 10. Roig B, Mnif W, Hassine AI, Zidi I, Bayle S, Bartegi A, Thomas O. Endocrine disrupting chemicals and human risk assessment: A critical review. Critical Reviews in Environmental Science and Technology. 2013;21:2297-2351
  11. 11. Giulivo M, López de Alda M, Capri E, Barceló D. Human exposure to endocrine disrupting compounds: Their role in reproductive systems, metabolic syndrome and breast cancer. A review. Environmental Research. 2016;151:251-264
  12. 12. Food and Agriculture Organization (FAO) of the United Nations. FAOSTAT. Statistic Division. [Internet]. 2016. Available from: http://www.fao.org/faostat [Accessed: 24 January 2017]
  13. 13. Pimentel D. Pesticides and pest control. In: Rajinder P, Dhawan A, editors. Integrated Pest Management: Innovation-Development Process. Vol. 1. Dordrecht: Springer; 2009. pp. 83-87
  14. 14. Liu CJ, Men WJ, Liu YJ, Hao Z. The pollution of pesticides in soils and its bioremediation. System Sciences and Comprehensive Studies in Agriculture. 2002;18:291-297
  15. 15. Cai DW. Understand the role of chemical pesticides and prevent misuses of pesticides. Bulletin of Agricultural Science and Technology. 2008;1:36-38
  16. 16. Aktar W, Sengupta D, Chowdhury A. Impact of pesticides use in agriculture: Their benefits and hazards. Interdisciplinary Toxicology. 2009;2:1-12
  17. 17. The European Parliament and the Council of the European Union. Directive 2000/60/EC of the European Parliament and of the Council of 23 October 2000 establishing a framework for community action in the field of water policy, Official Journal of the European Community. 2000;L 327:1-69
  18. 18. The European Parliament and the Council of the European Union. Directive 2008/105/EC of the European Parliament and of the Council of 16 December 2008 on environmental quality standards in the field of water policy, amending and subsequently repealing Council Directives 82/176/EEC, 83/513/EEC, 84/156/EEC, 84/491/EEC, 86/280/EEC and amending Directive 2000/60/EC of the European Parliament and of the Council. Official Journal of the European Community. 2008;L 348:84-97
  19. 19. The European Parliament and the Council of the European Union. Directive 2013/39/EU of the European Parliament and of the Council of 12 August 2013 amending Directives 2000/60/EC and 2008/105/EC as regards priority substances in the field of water policy. Official Journal of the European Community. 2013;L 226;1-17
  20. 20. The European Parliament and the Council of the European Union. Directive 2006/118/EC of the European Parliament and of the Council of 12 December 2006 on the protection of groundwater against pollution and deterioration. Official Journal of the European Community. 2006;L 372:19-31
  21. 21. Cooper B. Central and North West Regions Water Quality Program 1995/1996: Report on Pesticide Monitoring. TS96.048. Sydney: NSW Department of Land & Water Conservation; 1996
  22. 22. Cerejeira MJ, Viana P, Batista S, Pereira T, Silva E, Valério MJ, Silva A, Ferreira M, Silva-Fernandes AM. Pesticides in Portuguese surface and ground waters. Water Research. 2003;37:1055-1063
  23. 23. Konstantinou IK, Hela DG, Albanis TA. The status of pesticide pollution in surface waters (rivers and lakes) of Greece. Part I. Review on occurrence and levels. Environmental Pollution. 2006;141:555-570
  24. 24. Gilliom RJ. Pesticides in U.S. streams and groundwater. Environmental Science & Technology. 2007;41:3408-3414
  25. 25. Arias-Estévez M, López-Periago E, Martínez-Carballo E, Simal-Gándara J, Mejuto JC, García-Río L. The mobility and degradation of pesticides in soils and the pollution of groundwater resources. Agriculture, Ecosystems & Environment. 2008;123:247-260
  26. 26. Woudneh MB, Ou Z, Sekela M, Tuominen T, Gledhil, M. Pesticide multiresidues in waters of the Lower Fraser Valley, British Columbia, Canada. Part I. Surface water. Journal of Environmental Quality. 2009;38:940-947
  27. 27. Añasco N, Un, S, Koyama J, Matsuoka T, Kuwahara N. Assessment of pesticide residues in freshwater areas affected by rice paddy effluents in Southern Japan. Environmental Monitoring and Assessment. 2010;160:371-383
  28. 28. Jurado A, Vázquez-Suñe E, Carrera J, López de Alda M, Pujades E, Barceló D. Emerging organic contaminants in groundwater in Spain: A review of sources, recent occurrence and fate in the European context. Science of the Total Environment. 2012;440:82-94
  29. 29. Lapworth DJ, Baran N, Stuart ME, Ward RS. Emerging organic contaminants in groundwater. A review of sources, fate and occurrence. Environmental Pollution. 2012;163:287-303
  30. 30. Bottoni P, Grenni P, Lucentini L, Caracciolo A. Terbuthylazine and other triazines in Italian water resources. Microchemical Journal. 2013;107:136-142
  31. 31. Moreno-Gónzález R, Campillo JA, León VM. Influence of an intensive agricultural drainage basin on the seasonal distribution of organic pollutants in seawater from a Mediterranean coastal lagoon (Mar Menor, SE Spain). Marine Pollution Bulletin. 2013;77:400-411
  32. 32. Grung M, Lin Y, Zhang H, Orderdalen Steen A, Huang J, Zhang G, Larssen T. Pesticide levels and environmental risk in aquatic environments in China—A review. Environment International. 2015;81:87-97
  33. 33. Yadav IC, Devi NL, Syed JH, Cheng Z, Li J, Zhang G, Jones KC. Current status of persistent organic pesticides residues in air, water, and soil, and their possible effect on neighboring countries: A comprehensive review of India. Science of the Total Environment. 2015;511:123-137
  34. 34. Ccanccapa A, Masia A, Navarro-Ortega A, Pico Y, Barceló D. Pesticides in the Ebro river basin: Occurrence and risk assessment. Environmental Pollution. 2016;211:414-424
  35. 35. Dolan T, Howsam P, Parsons DJ. Diffuse pesticide pollution of drinking water sources: Impact of legislation and UK responses. Water Policy. 2012;14:680-693
  36. 36. Kaushik CP, Sharma HR, Kaushik A. Organochlorine pesticide residues in drinking water in the rural areas of Haryana, India. Environmental Monitoring and Assessment. 2012;184:103-112
  37. 37. Barbosa AMC, Solano M, Umbuzeiro G. Pesticides in drinking water—The Brazilian monitoring program. Frontiers in Public Health. 2015;3:Article 246
  38. 38. Mekonen S, Argaw R, Simanesew A, Houbraken M, Senaeve D, Ambelu A, Spanoghe P. Pesticide residues in drinking water and associated risk to consumers in Ethiopia. Chemosphere. 2016;162:252-260
  39. 39. 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. Journal of Environmental Management. 2011;92:311-330
  40. 40. Stanisakis AS. Use of selected advanced oxidation processes (AOPs) for wastewater treatment—A mini review. Global NEST Journal. 2008;10:376-378
  41. 41. Ribeiro AR, Nunes OC, Pereira MFR, Silva AMT. An overview on the advanced oxidation processes applied for the treatment of water pollutants defined in the recently launched Directive 2013/39/EU. Environmental International. 2015;75:33-51
  42. 42. Hoigné J. Inter-calibration of OH radical sources and water quality parameters. Water Science and Technology. 1997;35:1-8
  43. 43. Pera-Titus M, Garcı́a-Molina V, Baños MA, Giménez J, Espluga, S. Degradation of chlorophenols by means of advanced oxidation processes: A general review. Applied Catalysis B. 2004;47:219-256
  44. 44. Di Paola A, García-López E, Marcì G, Palmisano, L. A survey of photocatalytic materials for environmental remediation. Journal of Hazardous Materials. 2012;211:3-29
  45. 45. Babuponnusami A, Muthukumar K. A review on Fenton and improvements to the Fenton process for wastewater treatment. Journal of Environmental Chemical Engineering. 2014;2:557-572
  46. 46. Andreozzi R, Caprio V, Insola A, Marotta R. Advanced oxidation processes (AOP) for water purification and recovery. Catalysis Today. 1999;53:51-59
  47. 47. Petrovic M, Radjenovic J, Barceló D. Advanced oxidation processes (AOPs) applied for wastewater and drinking water treatment. Elimination of pharmaceuticals. The Holistic Approach to Environment. 2011;1:63-74
  48. 48. Reddy PV, Kim K. A review of photochemical approaches for the treatment of a wide range of pesticides. Journal of Hazardous Materials. 2015;285:325-335
  49. 49. Bhatkande DS, Pangarkar VG, Beenackers A. Photocatalytic degradation for environmental applications: A review. Journal of Chemical Technology and Biotechnology. 2002;77:102-116
  50. 50. Fechete I. Wang, Y. Védrine C. The past, present and future of heterogeneous catalysis. Catalysis Today. 2012;189:2-27
  51. 51. Qu Y, Duan X. Progress, challenge and perspective of heterogeneous photocatalysis. Chemical Society Reviews. 2013;42:2568-2580.
  52. 52. Litter MI. Heterogeneous photocatalysis—Transition metal ions in photocatalytic systems. Applied Catalysis B: Environmental. 1999;23:89-114
  53. 53. Serpone N, Sauve G, Koch R, Tahiri H, Pichat P, Piccinini P, Pelizzetti E, Hidaka H. Standardization protocol of process efficiencies and activation parameters in heterogeneous photocatalysis: Relative photonic efficiencies. Journal of Photochemistry and Photobiology A: Chemistry. 1996;94:191-203
  54. 54. Lazar MA, Varghese S, Nair SS. Photocatalytic water treatment by titanium dioxide: Recent updates. Catalysts. 2012;2:572-601
  55. 55. Konstantinou IK, Albanis TA. Photocatalytic transformation of pesticides in aqueous titanium dioxide suspensions using artificial and solar light: Intermediates and degradation pathways. Applied Catalysis B: Environmental. 2003;42:319-335
  56. 56. Herrmann JM. Heterogeneous photocatalysis: State of the art and present applications. Topics in Catalysis. 2005;34:49-65
  57. 57. Devipriya S, Yesodharan S. Photocatalytic degradation of pesticide contaminants in water. Solar Energy Materials Solar Cells. 2005;86:309-348
  58. 58. Gaya UI, Abdullah AH. Heterogeneous photocatalytic degradation of organic contaminants over titanium dioxide: A review of fundamentals, progress and problems. Journal of Photochemistry and Photobiology C: Photochemistry Reviews. 2008;9:1-12
  59. 59. Plantard G, Janin, T, Goetz V, Brosillon S. Solar photocatalysis treatment of phytosanitary refuses: Efficiency of industrial photocatalysts. Applied Catalysis B: Environmental. 2012;115:38-44
  60. 60. Reddy PAK, Reddy PVL, Kwon E, Kim KH, Akter T, Kalagara S. Recent advances in photocatalytic treatment of pollutants in aqueous media. Environment International. 2016;91:94-103
  61. 61. Fujishima A, Zhang X, Tryk DA. TiO2 photocatalysis and related surface phenomena. Surface Science Reports. 2008;63:515-582
  62. 62. Henderson MA. A surface science perspective on TiO2 photocatalysis. Surface Science Reports. 2011;66:185-297
  63. 63. Yu JG, Jimmy CY, Leung MKP, Zhao XJ, Ho WK, Zhao JC. Effects of acidic and basic hydrolysis catalysts on the photocatalytic activity and microstructures of bimodal mesoporous titania. Journal of Catalysis. 2003;217:69-78
  64. 64. Kabra, K, Chaudhary R, Sawhney RL. Treatment of hazardous organic and inorganic compounds through aqueous-phase photocatalysis: A review. Industrial & Engineering Chemistry Research. 2004;43:7683-7696
  65. 65. Malato S, Fernández-Ibáñez P, Maldonado MI, Blanco J, Gernjak W. Decontamination and disinfection of water by solar photocatalysis: Recent overview and trends. Catalysis Today. 2009;147:1-59
  66. 66. Chong MN, Jin B, Chow CWK, Saint C. Recent developments in photocatalytic water treatment. Water Research. 2010;44:2997-3027
  67. 67. Hashimoto K, Irie H, Fujishima A. TiO2 photocatalysis: A historical overview and future prospects. Japanese Journal of Applied Physics. 2005;44:8269-8285
  68. 68. Hanaor DAH, Sorrell CC. Review of the anatase to rutile phase transformation. Journal of Materials Science. 2011;46:855-874
  69. 69. Zhang Z, Wang C, Zakaria R, Ying JY. Role of particle size in nanocrystalline TiO2-based photocatalysts. Journal of Physical Chemistry B. 1998;102:10871-10878
  70. 70. Zhang J, Zhou P, Liu J, Yu J. New understanding of the difference of photocatalytic activity among anatase, rutile and brookite TiO2. Physical Chemistry Chemical Physics. 2014;16:20382-20386.
  71. 71. Hurun DC, Agrios AG, Gray KA, Rajh T, Thurnaur MC. Explaining the enhanced photocatalytic activity of Degussa P 25 mixed-phase TiO2 using EPR. Journal of Physical Chemistry B. 2003;107:4545-4549
  72. 72. Cassano AE, Alfano OM. Reaction engineering of suspended solid heterogeneous photocatalytic reactors. Catalysis Today. 2000;58:167-197
  73. 73. Herrmann JM. Heterogeneous photocatalysis: Fundamentals and applications to the removal of various types of aqueous pollutants. Catalysis Today. 1999;53:115-129
  74. 74. Ollis DF, Pelizzetti E, Serpone N. Photocatalyzed destruction of wáter contaminants. Environmental Science and Technology. 1991;25:1522-1529
  75. 75. Bahnemann W, Muneer M, Haque MM. Titanium dioxide-mediated photocatalysed degradation of few selected organic pollutants in aqueous suspensions. Catalysis Today. 2007;124:133-148
  76. 76. Kosmulski M. Isoelectric points and points of zero charge of metal (hydr)oxides: 50 years after parks’ review. Advances in Colloid and Interface Science. 2016;238:1-61
  77. 77. Shifu C, Ghengyu C. Photocatalytic degradation of organophosphorus pesticides using floating photocatalyst TiO2/SiO2/beads by sunlight. Solar Energy. 2005;79:1-9
  78. 78. Wei L, Shifu C, Wei Z, Sujuan Z. Titanium dioxide mediated photocatalytic degradation of methamidophos in aqueous phase. Journal of Hazardous Materials. 2009;164:154-160
  79. 79. Kusic H, Peternel I, Ukic S, Koprivanac N, Bolanca T, Papic S, Loncaric Bozic A. Modeling of iron activated persulfate oxidation treating reactive azo dye in water matrix. Chemical Engineering Journal. 2011;172:109-121.
  80. 80. Carbajo J, García-Muñoz P, Tolosana-Moranchel A, Faraldos M, Bahamonde A. Effect of water composition on the photocatalytic removal of pesticides with different TiO2 catalysts. Environmental Science and Pollution Research. 2014;21:12233-12240.
  81. 81. Carbajo J, Jiménez M, Miralles S, Malato S, Faraldos M, Bahamonde A. Study of application of titania catalysts on solar photocatalysis: Influence of type of pollutants and water matrices. Chemical Engineering Journal. 2016;291:64-73
  82. 82. Kowalska E, Rau S. Photoreactors for wastewater treatment: A review. Recent Patents on Engineering. 2010;4:242-266
  83. 83. Charles G, Roques–Carmes T, Becheikh N, Falk L, Commenge JM, Corbel S. Determination of kinetic constants of a photocatalytic reaction in micro–channel reactors in the presence of mass–transfer limitation and axial dispersion. Journal of Photochemistry and Photobiology A: Chemistry. 2011;223:202-211
  84. 84. Shan AY, Ghazi TI, Rashid SA. Immobilisation of titanium dioxide onto supporting materials in heterogeneous photocatalysis: A review. Applied Catalysis A: General. 2010;389:1-8
  85. 85. Dahl M, Liu Y, Yin Y. Composite titanium dioxide nanomaterials. Chemical Review. 2014;114:9853-9889
  86. 86. Bourikas K, Kordulis C, Lycourghiotis A. Titanium dioxide (anatase and rutile): Surface chemistry, liquid−solid interface chemistry, and scientific synthesis of supported catalysts. Chemical Review. 2014;114:9754-9823
  87. 87. Kumar SG, Devi LG. Review on modified TiO2 photocatalysis under UV/visible light: Selected results and related mechanisms on interfacial charge carrier transfer dynamics. The Journal of Physical Chemistry A. 2011;115:13211-13241
  88. 88. Dvoranova D, Brezova V, Mazur M, Malati MA. Investigations of metaldoped titanium dioxide titanium dioxide photocatalysts. Applied Catalysis B: Environmental. 2002;37:91-105
  89. 89. Kato S, Hirano Y, Iwata M, Sano T, Takeuchi K, Matsuzawa S. Photocatalytic degradation of gaseous sulphur compounds by silver-deposited titanium dioxide. Applied Catalysis B: Environmental. 2005;57:109-115
  90. 90. Peláez M, Nolan N, Pillai S, Seery M, Falaras P, Kontos AG, Dunlop PSM, Hamiltone JWJ, Byrne JA, O’Shea K, Entezari MH, Dionysiou DD. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Applied Catalysis B: Environmental. 2012;125:331-349
  91. 91. Chatterjee D, Dasgupta S. Visible light induced photocatalytic degradation of organic pollutants. Journal of Photochemistry and Photobiology C: Photochemistry Reviews. 2005;6:186-205
  92. 92. Gupta SM, Tripathi M. A review of TiO2 nanoparticles. Chinese Science Bulletin. 2011;56:1639-1657
  93. 93. Fenoll J, Garrido I, Cava J, Hellín P, Flores P, Navarro S. Photometabolic pathways of chlorantraniliprole in aqueous slurries containing binary and ternary oxides of Zn and Ti. Chemical Engineering Journal. 2015;264:720-727
  94. 94. Stamatis N, Antonopoulou M, Konstantinou I. Photocatalytic degradation kinetics and mechanisms of fungicide tebuconazole in aqueous TiO2 suspensions. Catalysis Today. 2015;252:93-99
  95. 95. Spasiano D, Marotta R, Malato S, Fernández-Ibáñez P, Di Somma I. Solar photocatalysis: Materials, reactors, some commercial and pre-industrialized applications. A comprehensive approach. Applied Catalysis B: Environmental. 2015;170‐171:90-123
  96. 96. Zheng LL, Pi FW, Wang YF, Xu H, Zhang YZ, Sun XL. Photocatalytic degradation of acephate, omethoate, and methyl parathion by Fe3O4@SiO2@mTiO2 nanomicrospheres. Journal of Hazardous Materials. 2016;315:11-22
  97. 97. Andronic L, Isac L, Miralles-Cuevas S, Visa M, Oller I, Duta A, Malato S. Pilot-plant evaluation of TiO2 and TiO2-based hybrid photocatalysts for solar treatment of polluted water. Journal of Hazardous Materials. 2016;320:469-478
  98. 98. Fenoll J, Garrido I, Hellín P, Vela N, Flores P, Navarro S. Photooxidation of three spirocyclic acid derivative insecticides in aqueous suspensions as catalyzed by titanium and zinc oxides. Journal of Photochemistry and Photobiology A: Chemistry. 2016;328:189-197
  99. 99. Fenoll J, Vela N, Garrido I, Navarro G. Pérez-Lucas G, Navarro S. Reclamation of water polluted with flubendiamide residues by photocatalytic treatment with semiconductor oxides. Photochemistry & Photobiology. 2015;91:1088-1094
  100. 100. Fenoll J, Garrido I, Hellín P, Flores P, Navarro S. Photodegradation of neonicotinoid insecticides in water by semiconductor oxides. Environmental Science and Pollution Research. 2015;22:15055-15066
  101. 101. Vela N, Fenoll J, Garrido I, Navarro G, Gambín M, Navarro S. Photocatalytic mitigation of triazinone herbicide residues using titanium dioxide in slurry photoreactor. Catalysis Today. 2015;252:70-77
  102. 102. Gómez S, Marchena CL, Renzini MS, Pizzio L, Pierella L. In situ generated TiO2 over zeolitic supports as reusable photocatalysts for the degradation of dichlorvos. Applied Catalysis B: Environmental. 2015;162:167-173
  103. 103. Sacco O, Vaiano V, Han C, Sannino D, Dionysiou DD. Photocatalytic removal of atrazine using N-doped TiO2 supported on phosphors. Applied Catalysis B: Environmental. 2015;164:462-474
  104. 104. Fenoll J, Vela N, Garrido I, Pérez-Lucas G, Navarro S. Abatement of spinosad and indoxacarb residues in pure water by photocatalytic treatment using binary and ternary oxides of Zn and Ti. Environmental Science & Pollution Research. 2014;21:12143-12153
  105. 105. Fenoll J, Flores P, Hellín P, Hernández J, Navarro S. Minimization of methabenzthiazuron residues in leaching water using amended soils and photocatalytic treatment with TiO2 and ZnO. Journal of Environmental Sciences. 2014;26:757-764
  106. 106. Sivagami K, Krishna RR, Swaminathan T. Photo catalytic degradation of pesticides in immobilized bead photo reactor under solar irradiation. Solar Energy. 2014;103:488-493
  107. 107. Verma A, Prakash NT, Toor AP. Photocatalytic degradation of herbicide isoproturon in TiO2 aqueous suspensions: Study of reaction intermediates and degradation pathways. Environmental Progress & Sustainable Energy. 2014;33:402-409
  108. 108. Hossaini H, Moussavi G, Farrokhi M. The investigation of the LED-activated FeFNS-TiO2 nanocatalyst for photocatalytic degradation and mineralization of organophosphate pesticides in water. Water Research. 2014;59:130-144
  109. 109. Vicente R, Soler J, Arques A, Amat AM, Frontistis Z, Xekoukoulotakis N, Mantzavinos D. Comparison of different TiO2 samples as photocatalyst for the degradation of a mixture of four commercial pesticides. Journal of Chemical Technology and Biotechnology. 2014;89:1259-1264
  110. 110. Fenoll J, Sabater P, Navarro G, Vela N, Pérez-Lucas G, Navarro S. Abatement kinetics of 30 sulfonylurea herbicide residues in water by photocatalytic treatment with semiconductor materials. Journal of Environmental Management. 2013;130:361-368
  111. 111. Fenoll J, Martínez-Menchón M, Navarro G, Vela N, Navarro S. Photocatalytic degradation of substituted phenylurea herbicides in aqueous semiconductor suspensions exposed to solar energy. Chemosphere. 2013;91:571-578
  112. 112. Fenoll J, Hellín P, Flores P, Martínez CM, Navarro S. Degradation intermediates and reaction pathway of carbofuran in leaching water using TiO2 and ZnO as photocatalyst under natural sunlight. Journal of Photochemistry and Photobiology A: Chemistry. 2013;251:33-40
  113. 113. Ramos-Delgado NA, Gracia-Pinilla MA, Maya-Trevino L, Hinojosa-Reyes L, Guzman-Mar JL, Hernandez-Ramirez A. Solar photocatalytic activity of TiO2 modified with WO3 on the degradation of an organophosphorus pesticide. Journal of Hazardous Materials. 2013;263:36-44
  114. 114. Fenoll J, Flores P, Hellín P, Martínez CM, Navarro S. Photodegradation of eight miscellaneous pesticides in drinking water after treatment with semiconductor materials under sunlight at pilot plant scale. Chemical Engineering Journal. 2012;204‐206:54-64
  115. 115. Fenoll J, Hellín P, Martínez CM, Flores P, Navarro S. Semiconductor-sensitized photodegradation of s-triazine and chloroacetanilide herbicides in leaching water using TiO2 and ZnO as catalyst under natural sunlight. Journal of Photochemistry and Photobiology A: Chemistry. 2012;238:81-87
  116. 116. Fenoll J, Hellín P, Martínez CM, Flores P, Navarro S. Semiconductor oxides-sensitized photodegradation of fenamiphos in leaching water under natural sunlight. Applied Catalysis B: Environmental. 2012;115-116:31-37
  117. 117. Fenoll J, Ruiz E, Hellín P, Flores P, Navarro S. Heterogeneous photocatalytic oxidation of cyprodinil and fludioxonil on leaching water under solar irradiation. Chemosphere. 2011;85:1262-1268
  118. 118. Senthilnathan J, Philip L. Photodegradation of methyl parathion and dichlorvos from drinking water with N-doped TiO2 under solar radiation. Chemical Engineering Journal 2011;172:678-688
  119. 119. Sojic DV, Despotovic VN, Abazovic ND, Comor MI, Abramovic BF. Photocatalytic degradation of selected herbicides in aqueous suspensions of doped titania under visible light irradiation. Journal of Hazardous Materials. 2010;179:49-56
  120. 120. Colina-Márquez J, Machuca-Martínez F, Puma GL. Photocatalytic mineralization of commercial herbicides in a pilot-scale solar CPC reactor: Photoreactor modeling and reaction kinetics constants independent of radiation field. Environmental Science & Technology. 2009;43:8953-8960
  121. 121. Saien J, Khezrianjoo S. Degradation of the fungicide carbendazim in aqueous solutions with UV/TiO2 process: Optimization, kinetics and toxicity studies. Journal of Hazardous Materials. 2008;57:269-276
  122. 122. Chen SF, Liu YZ. Study on the photocatalytic degradation of glyphosate by TiO2 photocatalyst. Chemosphere. 2007;67:1010-1017
  123. 123. Maldonado MI, Passarinho PC, Oller I, Gernjak W, Fernández P, Blanco J, Malato S. Photocatalytic degradation of EU priority substances: A comparison between TiO2 and Fenton plus photo-Fenton in a solar pilot plant. Journal of Photochemistry and Photobiology A: Chemistry. 2007;185:354-363
  124. 124. Qamar M, Muneer M, Bahnemann D. Heterogeneous photocatalysed degradation of two selected pesticide derivatives, triclopyr and daminozid in aqueous suspensions of titanium dioxide. Journal of Environmental Management. 2007;80:99-106
  125. 125. Oller I, Gernjak W, Maldonado MI, Pérez-Estrada LA, Sánchez-Pérez JA, Malato S. Solar photocatalytic degradation of some hazardous water-soluble pesticides at pilot-plant scale. Journal of Hazardous Materials 2006;138:507-517
  126. 126. Evgenidou E, Fytianos K, Poulios I. Photocatalytic oxidation of dimethoate in aqueous solutions. Journal of Photochemistry and Photobiology A: Chemistry. 2006;175:29-38
  127. 127. Chen SF, Cao GY. Photocatalytic degradation of organophosphorus pesticides using floating photocatalyst TiO2 center dot SiO2/beads by sunlight. Solar Energy. 2005;79:1-9
  128. 128. Giménez J, Bayarri B, González O, Malato S, Peral J, Esplugas S. Advanced oxidation processes at laboratory scale: Environmental and economic impacts. ACS Sustainable Chemistry & Engineering. 2015;3:3188-3196

Written By

Nuria Vela, Gabriel Pérez-Lucas, José Fenoll and Simón Navarro

Submitted: 30 November 2016 Reviewed: 27 March 2017 Published: 26 July 2017