Next Article in Journal
UV and Zero-Valent Iron (ZVI) Activated Continuous Flow Persulfate Oxidation of Municipal Wastewater
Next Article in Special Issue
Electrochemical Characterisation of the Photoanode Containing TiO2 and SnS2 in the Presence of Various Pharmaceuticals
Previous Article in Journal
Samarium-Mediated Asymmetric Synthesis
Previous Article in Special Issue
Photocatalytic Degradation of Ciprofloxacin by UV Light Using N-Doped TiO2 in Suspension and Coated Forms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Titanium Dioxide as the Most Used Photocatalyst for Water Purification: An Overview

by
Sanja J. Armaković
1,2,*,
Maria M. Savanović
1,2 and
Stevan Armaković
2,3
1
University of Novi Sad, Faculty of Sciences, Department of Chemistry, Biochemistry and Environmental Protection, 21000 Novi Sad, Serbia
2
Association for the International Development of Academic and Scientific Collaboration (AIDASCO), 21000 Novi Sad, Serbia
3
University of Novi Sad, Faculty of Sciences, Department of Physics, 21000 Novi Sad, Serbia
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(1), 26; https://doi.org/10.3390/catal13010026
Submission received: 30 November 2022 / Revised: 14 December 2022 / Accepted: 21 December 2022 / Published: 24 December 2022
(This article belongs to the Special Issue Advances in Photocatalytic Wastewater Purification, 2nd Edition)

Abstract

:
Titanium dioxide (TiO2), one of the most frequently used materials in general, has emerged as an excellent photocatalytic material for environmental applications. In this review, principles and mechanisms of the photocatalytic activity of TiO2 have been analyzed. Structural and physical specificities of TiO2 nanoparticles, such as morphology, crystal structure, and electronic and optical properties, have been considered in the context of photocatalytic applications. A review of the influence of several factors, such as the type and dimensions of photocatalyst particles, pH of the solution, the influence of oxidants/electron acceptors, and light intensity on photocatalytic properties of TiO2, has been provided. Superhydrophilicity as an intrinsic property of the TiO2 surface was discussed through surface reconstruction on TiO2 during the reversible hydrophilic changes. Additionally, attention was paid to improving the photocatalytic properties of TiO2 particles through aggregation and agglomeration.

1. Introduction

As soon as it was determined that semiconductors could be efficiently used for obtaining products such as H2 and O2, they were also recognized as efficient photocatalysts for degrading toxic materials. Titanium dioxide (TiO2) is the most prominent photocatalyst [1,2,3], widely employed due to its great photocatalytic activity, chemical and biological stability, insolubility in water, acid and base environment, resistivity towards corrosion, nontoxicity, low price, and availability in comparison to oxide, sulfide, and other materials [4,5,6]. Heterogeneous photocatalytic oxidation of organic compounds in an aqueous solution in the presence of TiO2 as photocatalyst provides the opportunity for efficient treatment of waste, drinking, surface, and ground waters, as well as for obtaining the ultraclean water suitable for the pharmaceutical industry and microelectronics. Many organic substances and their intermediates are completely mineralized [7,8,9]. Depending on the structure of the parent compound, degradation proceeds to carbon dioxide, water, nitrates, chalcogenides, phosphates, and other inorganic ions [10]. However, besides all the mentioned advantages, TiO2 has certain drawbacks regarding its practical application. Due to the large band gap, a minimal amount of radiation from a close UV area of sunlight, only ca. 3–4%, is employed during the photocatalytic process. Therefore, TiO2 is practically inactive when exposed to sunlight [11]. This reduces the applicability of the sun as the most economical light source. Near the Earth’s surface, there are 0.2–0.3 mol photons m–2 h–1 in the range 300–400 nm with a typical UV flux of 20–30 W m–2 [12].
Taking into account the serious influence of organic pollutants on the environment, there is a constant need for the development of materials for the efficient elimination of toxic substances from drinking water [13,14], ground and surface water [15,16,17,18], ocean water, sediment and soil [19,20]. In this regard, oxide nanomaterials such as TiO2 have exhibited great potential for practical applications in many areas. Many advantages of TiO2 over the broad range of other nanomaterials have led to the situation where TiO2 is one of the most frequently employed photocatalytic materials.
In this review, we comprehensively introduced the most important features of TiO2 regarding its photocatalytic properties. It is essential to understand its photocatalytic properties, suitable physical and chemical properties, possible modifications using inexpensive methods, applicability to a wide range of compounds, and commercial availability. Finally, a better understanding of TiO2 photocatalytic properties can contribute to its more frequent practical application for water purification.

2. Structural Features and Physical Properties of TiO2 Nanoparticles

Understanding the fundamental properties of semiconductor materials governing their photoelectric performance (such as their crystal structure, lattice parameters, and optical and electronic properties) is essential to optimize their performance for photocatalytic applications. This section introduces the fundamentals underlying the photocatalytic performance of nanostructured TiO2.

2.1. Crystal Structures of TiO2

Morphology and the crystal structure of TiO2 principally determine its photocatalytic activity. Therefore, essential factors for TiO2 photocatalytic activity are crystallite size and the specific surface area [21,22,23]. TiO2 nanostructures with different shapes and titania-based nanocomposites have attracted much attention in research due to their diverse physicochemical characteristics. The 1D TiO2 nanostructures have gained more attention compared to their 0D and 2D counterparts due to the higher aspect ratio, increased surface area, and efficient electronic charge properties [24]. TiO2 exists in two tetragonal forms (rutile and anatase) and one rhombic form (brookite), Figure 1 [25,26,27,28]. Brookite is difficult to obtain in laboratory conditions, while rutile and anatase are easily prepared. As a bulk material, rutile is the stable phase; however, solution-phase preparation methods for TiO2 generally favor the anatase structure. These observations are attributed to two main effects: surface energy and precursor chemistry. It has been found that the surface energy of anatase is lower than those of rutile and brookite [28]. The concept of surface energy accurately explains the observed crossover size of about 30 nm where anatase nanoparticles transform to rutile [29]. Secondly, the crystal structure stability has been explained based on a molecular picture. The precursor chemistry determines the nucleation and growth of the different polymorphs of TiO2, which depends on the reactants used [30,31]. A complicating factor in understanding nanoparticle formation is the multitude of experimental conditions used to synthesize the different TiO2 phases, making it challenging to compare mechanisms [28,32].
Figure 1. 3D visualization crystal structures of TiO2 using visualization for electronic and structural analysis (VESTA) [33].
Figure 1. 3D visualization crystal structures of TiO2 using visualization for electronic and structural analysis (VESTA) [33].
Catalysts 13 00026 g001
These polymorphs, anatase, and rutile, exhibit different properties and, consequently, different photocatalytic performances. Of the two tetragonal forms mentioned, the anatase form shows significantly higher photocatalytic activity than rutile due to the higher presence and nature of surface hydroxyl groups. Transformation of anatase form to rutile happens at elevated temperatures of 700–1000 °C. Anatase form is stable at lower temperatures (it occurs in the form of a pyramidal crystal structure). At the same time, rutile (needle-shaped) is dominant during the synthesis at high temperatures [34,35,36]. Anatase has a higher energy gap, which additionally contributes to the photocatalytic activity of this form. The energy gap of anatase is 3.20 eV, while the rutile energy gap is 3.02 eV [37,38].
TiO2 Degussa P25 is the most frequently used commercial product containing 75% of anatase form and 25% of rutile form [39,40]. The mentioned mixture exhibits outstanding photocatalytic performance and superiority compared with other TiO2 [41]. The predominant form of titania used is anatase. It was found that anatase is the most photocatalytic active form within TiO2 polymorphs. TiO2 Hombikat is a pure anatase form of photocatalyst [39]. TiO2 also appears in another form named TiO2 Wackherr “oxide de titan standard”, which contains 100% of anatase form, exhibits interesting characteristics related to photocatalytic application, and is more efficient than TiO2 Degussa P25 [42,43,44]. The main characteristic of this photocatalyst, when compared with TiO2 Degussa P25, is a lower scattering of radiation in the UV area, which is most likely the consequence of the greater size of particles. Particle size is a crucial factor affecting the performance of nano-photocatalytic materials. The greater size of TiO2 Wackherr particles than TiO2 Degussa P25 results in a lower specific surface area. Particles with a lower specific surface area usually exhibit low photocatalytic activity. However, they can be more efficient in photocatalysis, where optical properties and low scattering are significant [44,45].

2.2. Electronic Structure of TiO2

Understanding the electronic properties of semiconductor materials, including their band structure, nonequilibrium carrier concentration, carrier mobility, and lifetime is essential to achieve inflection of charge carrier behavior and enhancing photocatalytic performance. Against this background, electronic properties of anatase, rutile, and brookite TiO2 phases were introduced [46]. The current understanding of the electronic structures of TiO2 is mainly based on the results of independent and combined theoretical calculations, usually in the framework of the density functional theory (DFT) [26] and experimental techniques (e.g., synchrotron radiation photoelectron spectroscopy, UV-vis spectroscopy, ultraviolet photoelectron spectroscopy, and photoluminescence) [47,48].
The application of various computational methods in elucidating the photocatalytic properties of novel TiO2-based materials is essential. Namely, applying the DFT approach enables researchers to predict the band gap of newly designed materials and focus their attention on synthesizing materials with target photocatalytic properties. The computational aspect of band-gap engineering is critical in developing new photocatalytic materials, as this parameter is essential for practical applications [49]. Additionally, the DFT approach, in combination with carefully selected density functionals and basis sets, is an excellent tool for understanding the light absorption properties of molecules that are degraded by photocatalysts, helping scientists to easier identify the degradation mechanism [50].
Regarding the computational design of novel materials, the DFT approach offers the best cost–efficiency ratio [51]. This flexible theoretical approach allows researchers to elucidate the structural and electronic properties of photocatalytic materials by analyzing how various structural alterations influence the electronic subsystem of nanomaterials.
However, the DFT approach’s severe drawback is that it severely underestimates the band gap values due to self-interaction error [52,53,54,55]. Considering the fundamental importance of band gaps in photocatalysis, this is a significant challenge in designing novel compounds to be applied in this area. In general, this issue can be tackled in two main ways. One approach is to create specially designed density functionals that reproduce electron density adequately. The second approach is to use the existing density functionals and implement specific corrections within them. Both directions have positive and negative sides. For example, specially designed hybrid density functionals yield outstanding results in accuracy, but they are computationally much more demanding and are usually applicable only for single-point band structure calculations. Conversely, corrected density functionals offer improved results over conventional density functionals, with computational costs slightly more demanding than traditional density functionals.
Considering the application of hybrid functionals for band structure calculations, it is worth mentioning the importance of the HSE06 hybrid functional [56]. Recent studies have confirmed this function’s importance in predicting band gaps of different materials with small, intermediate, and large values of this critical parameter [57,58]. However, in terms of computational costs, hybrid density functionals may still be unavailable to significant number of research groups, which may apply the computationally affordable DFT + U method [59]. This method adds a Hubbard-like term to the Hamiltonian to account for on-site interactions. The simplest version of the DFT + U approach relies on only one parameter—on-site Coulomb term (U). However, this method can be improved by incorporating the site exchange term J. U and J parameters can be obtained through ab initio applications. Still, they are frequently obtained empirically by testing a range of values. So far, many research studies have reported the values of these parameters for certain types of materials, so a literature survey is warmly recommended to find good starting values.
In Figure 2, we present the band structures and density of states as obtained by applying the DFT + U method, with U taking the value of 7.8 eV and J taking the value of 1 eV, as reported in studies by [60,61]. Inspection of band structures shows that the application of the DFT + U method provides results for band gaps that are in agreement with experimental results, which will be mentioned somewhat later. The density of states figures also show that the conduction band consists of Ti states, while valence band (VB) consists of O states. As reported in reference by Li et al. [62], CB consists of Ti3d states, while VB consists of O2p states.
Ordinarily, TiO2 phases, especially anatase, show good insulation in their ideal stoichiometric states due to their wide bandgap [63]. However, certain types of point defects are unavoidably introduced during the preparation process. These defects might be interstitial titanium ions (T3+), oxygen point defects, and substituted ions and can notably affect the charge carrier behavior, band structure, and, eventually, photocatalytic performance [64]. The characteristics of defects (e.g., concentration, type, distribution, and dimension) and their influence on the photoinduced charge carriers in TiO2 can be various [65]. Point defects can induce the generation of defect states, the position generally influenced by the surface and phase. For instance, the defect states of rutile-TiO2 (110) and anatase-TiO2 (101) are found at ≈0.8–1.0 eV and ≈0.4–1.1 eV, of which both are below the CB edge [66]. In the brookite phase, both rutile-like (Y-shaped) and anatase-like (T-shaped) OTi3 building blocks exist, inducing the O2p in VB to present characteristics from both tetragonal phases [26]. Considering the application of hybrid and recently developed density functionals, the study by Dharmale et al. is worth mentioning. In their work, electronic properties such as effective mass, the partial density of states, the total density of states, and the band structure of brookite TiO2 have been studied by applying using seven exchange-correlation functionals, including the already-mentioned HSE06 [67]. Band structures and projected density of states for the rutile), brookite, and anatase polymorphs of TiO2 are shown in Figure 2.

2.3. Optical Properties of TiO2

A semiconductor’s optical properties (e.g., photoconductivity, dielectric constant, refractive index, extinction coefficient, reflectivity, absorption coefficient, and loss function) are related to its bandgap [68]. TiO2 is commonly known as a wide-bandgap semiconductor with high susceptibility to UV light [69,70]. The optical absorption in the visible and near-infrared regions is insignificant because photons in the visible region do not have intrinsic excitation of carriers. When the electrons (e) in the VB are exposed to UV radiation, they are excited to the CB, leaving holes (h+) in the VB [71,72,73]. The CB e is now in a purely 3D state, and the possibility of transition of e to the VB is reduced due to the difference in parity. Therefore, the probability of e–h+ recombination is reduced [74]. Consequently, separating energy between these two states defines the sensitivity of TiO2 towards the light in the UV range. However, the optical properties at the surface differ considerably from those of the bulk material, providing extensive opportunities to optimize further the photoelectric and optical properties of TiO2 [75].
The bandgap energies (i.e., optical absorption edges) of rutile, anatase, and brookite TiO2 are estimated to be ≈3.00, 3.21, and 3.13 eV at room temperature, respectively. The photon energy values of optical absorption edges of rutile and anatase increase with the crystal growth temperature decrease [28]. Overall, rutile and anatase bandgaps in bulk are considered to be indirect [76]. Detailed examinations of rutile TiO2 at temperature 1.6 K discover an anisotropic optical response characterized by a direct forbidden transition at ≈3.06 eV. At the same time, the bandgap near the edge is prevailed by an indirect transition. The direct bandgap transition of anatase occurs at ≈3.8–4.0 eV [77]. Additionally, the orthorhombic TiO2 brookite bandgap energy has been experimentally determined to be 3.1–3.4 eV. This is a biaxial material with three independent components other than the dielectric tensor of the uniaxial rutile and anatase materials [5,28]. The significant value of the energy gap limits the exceptional photocatalytic characteristics of TiO2. Namely, the photon energy should be high enough to excite the TiO2 particles.
In recent experimental and theoretical investigations, efforts have been made to improve the optical properties of TiO2 by increasing its photosensitivity and identifying the correlation between its surface, interfacial, and microstructural characteristics and the corresponding mechanisms crucial to its photoelectric properties [78,79]. Generally used strategies to enhance the optical properties of TiO2 include element doping [80,81] coupled with other semiconductors to form heterojunctions [82,83], synthesis of nanostructures with particular morphologies [84], surface sensitization to improve optical characteristics using organic dyes, metal nanostructures or metal complexes [85], etc.
The doping technique can be explained as the deliberate addition of impurities into a semiconductor material (Figure 3). To enhance TiO2 catalytic activity under visible light, metal/non-metal-doped TiO2 structures have been extensively studied [86]. Adding metal/non-metal increases oxygen vacancies and reduces the band gap energy, resulting in higher photocatalytic activity [87]. This could be useful in wastewater treatment for the photocatalytic degradation of organic compounds under visible and UV light radiation. Many different metals, such as Pd [88], Pt [89], Au, Ag [90], Ce [91], Sr [92], V [93], etc., have been employed for the preparation of metal-doped TiO2 catalysts.
The sol–gel method is often used in the synthesis, where the photocatalyst is doped with metals such as Fe, Ni, Cr, V, La, Nd, and Sm. The sol–gel process represents one of the versatile methods for preparing nano-dimensional materials. Incorporating an active dopant allows the doped element to interact directly with the support, which is why the material has catalytic or photocatalytic properties. When metal nanoparticles are doped into the TiO2, a new energy level or an interband state is produced in the band gap close to the CB arising from the partially filled d orbitals of the doping metal ions modifying the electronic structure and hence lowering the band gap (Figure 3b).
Although metal-ion doping decreases the energy gap of TiO2, the aforementioned metal ions can also act as recombination centers for electrons and holes, thus reducing the overall activity of the TiO2 [94]. In recent years, non-metals such as C, N, P, S, and B have been the best candidates for obtaining the desired band-gap narrowing of TiO2 [95]. It was shown that non-metal doped TiO2 shows significant catalytic activity under visible light radiation [96]. Several researchers have depicted that TiO2 doping with non-metal ions improve photocatalytic efficiency compared to metals. Doping with cations changes the morphology of the photocatalysts and the photocatalytic activity. It also affects the photocatalyst’s electronic structure. Non-metal dopants influence the VB of TiO2 through interactions with 2p e of O and form an impurity level above the VB. Distinct non-metals, including N, C, S, F, B, etc., are used as dopants and have shown promising outcomes in recent research studies. Nitrogen doping in TiO2 has gained considerable attention due to its ability to narrow the band gap and promote the electron–hole pair transfer mechanism (Figure 3c). It has been observed that in the case of doping with lighter elements such as N, C, S, or B at substitutional sites of TiO2, lower atomic number elements with smaller effective nuclear charge appear at a higher localized energy level in the band gap. Doping TiO2 by carbon stabilizes anatase TiO2, enhances its conductivity, and extends the pollutant adsorption on the surface of TiO2 [97].
Žener et al. [87] showed that doping, co-doping, and modifying TiO2 samples with nitrogen, sulfur, and platinum increased their photocatalytic activity by up to 6 times. XRD measurements revealed that the replacement of HCl with H2SO4 during sol–gel synthesis reduced the size of the crystallites from ~30 nm to ~20 nm, increasing the surface area. This is consistent with the samples’ photocatalytic activity and the photocatalysts’ measured photocurrent behavior [87].
Coupling TiO2 with the metal oxides increases the charge carrier separation and thereby increases the lifetime of the charge carriers [98]. When the coupled catalyst, consisting of TiO2 and a semiconductor, is irradiated, both the TiO2 and the semiconductor will excite electrons from VB to the CB using UV and Visible irradiation. TiO2 and the semiconductor configuration for coupling are crucial for the enhancement activity. The CB of TiO2 should be more favorable than the corresponding band of the semiconductor, and the VB should be more cathodic than that of TiO2. The e from the CB of the semiconductor migrates to the CB of TiO2 and increases the concentration of electrons at the TiO2 conduction band. At the same time, the h+ generated at the VB of TiO2 will be transferred to the VB of the semiconductor, increasing the concentration of h+ in the coupled semiconductor/TiO2 (Figure 3d) [99,100,101]. Couplings of TiO2 with a metal oxide such as ZnO [102], SiO2 [103], Cu2O [104], Bi2O3 or ZnMn2O4 [105], graphene [106], etc. are reported for many photocatalytic applications including organic pollutant degradation, water splitting, pharmaceutical degradation, etc. TiO2–ZnO binary oxide systems containing various molar ratios of TiO2–ZnO were prepared using a sol–gel method. It was reported that the crystalline structure, thermal stability, and porous structure parameters were determined by the molar ratio of TiO2 to ZnO and the calcination process for the most part. TiO2–ZnO showed high photocatalytic activity towards the degradation of C.I. Basic Blue 9, C.I. Basic Red 1, and C.I. Basic Violet 10 dyes [102]. TiO2 nanoparticles synthesised via the acid-catalysed sol–gel method were used to prepare coupled TiO2/SiO2 mesoporous materials were prepared by deposition of TiO2 nanoparticles. TiO2/SiO2 showed photocatalytic activity towards the photodegradation of rhodamine 6G in aqueous solution using UV radiation [103]. One of the noticed studies focuses on TiO2 nanocomposite with graphene as photocatalyst, one of the most prominent representatives of carbon nanostructures. Coupling TiO2 with graphene proved to be beneficial, as the higher efficiency in photocatalysis has been observed compared to that of TiO2 alone. Graphene sheets are thought to act as an electron acceptor that enables the separation and transfer of photogenerated electrons during TiO2 excitation, simultaneously reducing recombination of electron–hole pairs [106]. Wang et al. reported one of the first studies where TiO2 (P25)-graphene nanocomposites were used for photocatalytic degradation of methylene blue. Further research has led to the photodegradation of many organic pollutants by these materials [107].

3. Photocatalytic Activity of TiO2: Basic Principles and Mechanism

Exciting species are formed by irradiating TiO2 semiconducting particles with a light of energy higher than the energy gap [108,109,110,111,112,113,114,115]. Namely, electron–hole (e–h+) pairs are formed (reaction (1)), together with holes in VB and electrons in CB, representing the initial phase of the photocatalytic process (Figure 4) [115].
TiO 2 h v   ( λ   <   387   nm ) TiO 2 e + h +
Although the lifetime of e–h+ pairs is only several nanoseconds, it is enough for a redox reaction with the semiconductor to be initiated [108,109,110]. Further, formed pairs can be separated or transferred to the surface. Thus, they can take part in the oxide reduction reactions of organic and inorganic compounds, or they can be subjected to the process of recombination (Figure 4), decreasing the quantum yield of the response [108,111]. In an aqueous solution, the surface of TiO2 reacts with OH and water molecules which are adsorbed at the surface of the catalyst, by which OH are formed (reactions (2) and (3)):
h +   +   H 2 O ads     OH ads   +   H +
h +   +   OH ads     OH ads
However, one condition has to be achieved for the oxidation of OH and water molecules, which are adsorbed on the surface. The redox potential of these reactions has to be lower than the redox potential of the VB of the semiconductor. Reactions (2) and (3) are thermodynamically possible since the aforementioned condition is fulfilled for the TiO2. Since newly formed OH are powerful oxidizing agents whose standard redox potential is +2.8 V [112,113,114], they can practically oxidize all organic compounds to mineral products (Figure 4) [115].
Since the redox potentials of many organic compounds are lower than the redox potentials of VB of anatase TiO2, it should be expected for direct oxidation through h+ to occur. However, Turchi and Ollis [116] reported that this does not happen during the processes carried in organic solvents, during which complete oxidation of organic compounds does not occur. Other authors believe that direct oxidation of adsorbed reactants, such as carboxyl acids or their anions, employing h+ is possible (reaction (4)) [117,118].
h +   +   organic   compound     CO 2   +   H 2 O   +   inorganic   ions
In the case of applying TiO2 as a photocatalyst during processes occurring in an aqueous solution, the share of the reactions through OH is probably higher (due to the presence of water). However, reactions through h+ could not be neglected [119,120,121]. In addition to this, applying a fluorescent technique for the quantitative determination of OH in TiO2 with coumaric and terephthalic acid, it was found that oxidation reactions occur through h+ rather than radicals [122]. Additionally, there are examples where certain authors report that captured h+ and bonded OH equally participate in the oxidation process [123,124]. h+ transfer to OHads after water oxidation on the surface, while e from the CB and water molecules are mutual competitors for the takeover of h+. OH quickly react with the molecules on the catalyst’s surface and within the solution and are usually the prominent participants in the photocatalytic reaction [120,125,126].
Phases of the photocatalytic processes are as follows [116,127,128,129]:
  • Excitation of the photocatalyst by photons of energy higher than the energy gap followed by the formation of e and h+ (reaction (1));
  • Adsorption on the surface of the photocatalyst is followed by reactions (5)–(8):
h +   +   P ads     TiO 2   +     P ads +
h + +   H 2 O ads   TiO 2 + OH +   H +
h + + OH ads   TiO 2 + OH ads
e +   P ads   TiO 2 +   P ad
where P ads represent the adsorbed organic molecule;
  • “Capturing” of the e with O2 as the most commonly used electron acceptor (reaction (9)):
e +   O 2   TiO 2 +   O 2
  • Photodegradation by OH, either adsorbed or free, under a variety of free or adsorbed organic species;
  • Recombination of the e−h+ pairs followed by the heat release (reaction (10)):
e + h +   Heat
According to the presented reactions, it can be concluded that the process is rather complex. The extent of the complexity of the process is increased when it is considered that formed radicals can be subjected to the so-called reverse reactions. In such a way, parent compounds can be regenerated [111,130]. The enactment of the reverse reactions influences the photocatalytic process [131]. Therefore, proposing a reaction scheme that considers all formed intermediates is complex. However, the number of researchers finding solutions to these complex mechanisms is increasing, indicating the importance of thie field [120,132,133].

4. Factors Affecting Photocatalytic Properties of TiO2

4.1. Type and Dimensions of TiO2

The surface area plays a vital role in the photocatalytic performance of the nanosized porous TiO2. Namely, the higher the surface area, the better the photocatalytic performance of TiO2 will be [134,135]. Different methods can be applied to synthesize TiO2, influencing the pore structure and crystallite size distribution. Crucial steps of the synthesis procedure can be adjusted to obtain a larger surface area which can offer more active adsorption sites and centers of photocatalytic reaction [136]. Improved photocatalytic activity of mesoporous TiO2 catalysts, in comparison with the microporous materials with restricted pore size, has been reported by several researchers [137,138,139]. Diffusion of the organic molecules to the active sites on the material surface and optimal exposition to the solar light is achieved through mesoporous TiO2 [138,140]. Since the anatase phase has a much higher photocatalytic activity than amorphous and rutile TiO2, it is still challenging to synthesize mesoporous TiO2 containing the anatase phase of high crystallinity and a large surface area [141,142]. The anatase phase is formed during the calcination process at a temperature of 350 °C and higher [143,144], while the rutile phase is formed as the calcination temperature reaches >600 °C [35,120]. However, high calcination temperature results in nanocrystalline particles’ growth and the rapid decrease in the specific surface area [145]. Sol–gel synthesis of nanostructured materials was also intensively explored in the last decade [146,147,148]. There are many advantages of this approach; some are concerned with the fact that ambient temperature and atmospheric pressure are employed, together with the fact that this process occurs in a solution, offering the control of purity, homogeneity, doping, composition, and stoichiometry. Due to its simplicity and the possibility of controlling structural and morphological properties, sol–gel synthesis is now widely used [149,150]. For nano-dimension particles, recombining photogenerated e–h+ is much slower than the diffusion of free charge carriers from the particle’s interior to its surface, which contributes to the higher photocatalytic activity [151,152]. For the degradation of large organic molecules such as metoprolol [153,154], methylene blue [137,155], paracetamol [156], 3-amino-2-chloropyridine [157], pindolol [158], clopyralid [159], picloram [160], ethylene [161], 4-chlorophenol and methylene orange [162], etc., it was determined that mesoporous TiO2 exhibits high photocatalytic activity. Additionally, photodegradation of stearic acid was achieved using mesoporous and microporous TiO2 films [163], TiO2 self-cleaning films [164], and TiO2 films prepared by template-assisted sol–gel [165]. Mills et al. [166] reported the findings of a detailed study of the photocatalytic properties of Pilkington ActivTM glass in the degradation of stearic acid. Pilkington Glass ActivTM represents a possible successor to Degussa P25 TiO2. Especially as a photocatalyst film for comparing other photocatalysts or superhydrophilitic self-cleaning films. Pilkington Glass ActivTM has superior mechanical stability, reproducible activity, and widespread commercial availability, making it highly attractive as a reference photocatalytic film [166].
Photodegradation of some dyes was very effective in the presence of TiO2. For example, methylene blue was photodegraded using various TiO2 films [167], TiO2 impregnated diatomite [168], TiO2 pretreated with varying concentrations of NaOH [169] brookite–rutile bi-crystalline phase of TiO2 [170], etc. Another dye that has attracted attention due to its frequent decomposition with TiO2 is methyl orange. It was reported that methyl orange was photodegraded using Ag-doped TiO2 photocatalysts [171], carbon nanotubes-TiO2 [172], TiO2-zeolite photocatalyst [173], etc., under visible light and UV irradiation.
These results indicate that it is possible to design photocatalysts possessing molecular size sensitivity. Comparing the properties of TiO2 Wackherr and Degussa P25, TiO2 Wackherr has a much lower specific surface area than Degussa P25 (almost six times). This is the consequence of the much larger particles of Wackherr compared to Degussa P25 [44]. On the other side, Degussa P25 has higher radiation scattering than Wackherr, due to which Degussa P25 is less efficient regarding radiation usage [44,111]. Back reactions often decrease the efficiency of the photocatalytic degradation rates; these reduce partially oxidized transients to give back the initial substrate [44,111,174].
It is well known that photocatalytic activity occurs at the surface of photocatalysts. Therefore, smaller TiO2 nanoparticles, which possess a much higher surface-to-volume ratio, have a greater specific surface area available for catalysis. TiO2 catalysts have been prepared in various nano forms, such as nanoparticle powders, nanotubes, nanorods, nanowires, and immobilized states (as thin films) [86]. Between extensive applications using surface chemical and physical properties of TiO2, the defects, and the TiO2 surface states, depending enormously on material preparation technologies, play an essential role in its optical, chemical, and electrical properties. Having that in mind, choosing a reliable and well-controllable technology to design the defects in TiO2 is crucial for their specific application [175]. For the preparation of TiO2 thin films, all viable physical and chemical deposition technologies have been used. Over others, atomic layer deposition (ALD) has different advantages due to its precise thickness control, low growth temperature, large area uniformity, and extremely conformal surface coating for nanostructures [176]. Although many precursors have been used successfully for deposition of TiO2 by ALD processes, the common precursor TiCl4, stands out as a liquid with a moderate vapor pressure [177,178]. Active research needs to be continued in the field of TiO2 thin films. This would lead to understanding the synthesis routes, photocatalytic characterization, film morphology, and developing techniques that produce TiO2 thin films with desirable photocatalytic properties while minimizing costs [86].
TiO2 immobilization can be performed on powder and pellet substrates, thick/rigid substrates, or thin/soft materials [179,180]. Titanium oxide-based photocatalytic films can be produced by fused deposition modelling (FDM) by applying biopolymers obtained from renewable biomass resources. This thermoplastic route allows shaping composites through the immobilization of photoactive TiO2 nanoparticles using environmentally friendly bioplastics such as polylactic acid (PLA). Composites synthesized by FDM with an inorganic charge of 30 wt% of TiO2 exhibited a 100% methyl orange (MO) photodegradation after 24 h of light exposition. This could be explained by the extremely uniform dispersion of the nanophase within the polymer matrix in the FDM feedstock [180,181].

4.2. UV Light Intensity

Light intensity is another important factor that influences the process of photocatalytic degradation. Among many factors, light intensity determines the extent of light absorption by semiconducting catalysts at a particular wavelength [12,182,183,184,185]. Related to the experimental setup, overall pollutant conversion and degradation efficiency are invariably determined by light intensity distribution within the reactor [186]. According to the reported results by Herrmann [187], the degradation rate can be correlated with the radiant flux, Φ, in the following manner: below the value of 25 mW cm–2 for Φ, the degradation rate is proportional to Φ, while for the values of above 25 mW cm–2 degradation rate is proportional to Φ1/2. The latter case indicates a high flux value, and an increase in degradation rate is proportional to Φ1/2. Recombination is insignificant for low light intensity. Additionally, the time interval in which the pollutant is exposed to a certain intensity of radiation plays an important role in its degradation, because the longer the pollutant is exposed to radiation, the greater efficiency of degradation could be achieved. Effects of light intensity on the solar photocatalytic degradation of Bisphenol A in water and the presence of TiO2 on sunny and cloudy days were investigated by Kaneco et al. [188]. It was determined that degradation efficiency increases rapidly with the light intensity as Φ reaches 0.35 mW cm–2, after which efficiency increases gradually. In the study of Venkatachalam et al. [147], the mineralization efficiency of 4-chlorophenol was compared when lamps of different wavelengths over TiO2 were used, namely 365 and 254 nm. The results indicated that at 365 nm mineralization rate was somewhat higher than at 254 nm.
It is reported in the literature that the curing behavior of unpigmented coatings is dependent on light intensity [189,190,191,192]. A faster cure and even more complete polymerization are achieved with increased light intensity. The influence of different light intensities (in the range from 10–70 mW cm–2) on a system which consisted of epoxy acrylate and tri-propylene glycol diacrylate was investigated by Doğruyol et al. [193]. With the increase in light intensity up to 50 mW cm–2, the conversion rate and overall conversion increased, while it decreased with the further increase in light intensity. The influence of the light intensity on the photo-polymerization of poly(methacrylic acid) hydrogels was studied by He et al. [191], and the results of the study indicated that as the UV intensity increased from 2 to 24 mW cm–2, the final conversion decreased from 99% to 61%.
However, Mills et al. [166] state that for Degussa P25 TiO2 films, it is difficult to estimate how much light is absorbed and lost through reflection and scattering. Accordingly, they recommend that the photocatalyst system’s formal quantum efficiency (FQE) be calculated instead. FQE is the ratio of photoreaction rates and incident light intensity. They further state that the FQE is ≤quantum yield for any photochemical process.
The dependence of the rate of conversion and final conversion of pigmented and unpigmented UV curable systems on UV light intensity, ranging from 2 to 80 mW cm–2, was investigated by Kardar et al. [189]. Related to unpigmented formulation, it was found that for the intensity ranging from 2 to 20 mW cm–2, maximum rates were achieved, and final conversion increased, while the increase in light intensity to 40 and then to 80 mW cm–2 decreased the rate of conversion. Related to pigmented formulation, the polymerization and ultimate modification rate increased with the increased light intensity up to 40 mW cm–2.
Additionally, the influence of different light intensities (in the range from 5–50 mW cm–2) on a system that consisted of UV and monomer benzyl acrylate was investigated by Thiher et al. [194]. Various UV, electron-beam, and hybrid processing conditions were presented for the studies. The processing conditions for electron-beam and UV-only systems were chosen such that the impact of oxygen and additives could be observed. Benzyl acrylate was photopolymerized for 30 s at two different effective irradiances, 20 mW cm–2 and 50 mW cm–2.

4.3. Effect of pH

The effect of the pH value on the photocatalytic properties of TiO2 is rather complex due to several factors: electrostatic interaction between semiconductor surface, solvent molecules, reactants, and radicals formed during the reaction, etc. [195]. Surface-charge properties of the photocatalyst are influenced by pH value, as shown by Zhu et al. [196]. This fact could be explained based on the point of zero charge (PZC). The adsorbent surface is positively charged, attracting anions and repealing cations when the pH is lower than the PZC. Above the PZC the surface is negatively charged, attracting cations and repealing anions. According to [21,197], this can be represented through the following (reactions (11) and (12)):
TiOH   +   H +     TiOH 2 +
TiOH   + OH ads     TiO +   H 2 O
The solution’s pH value is essential for heterogeneous photocatalytic reactions since it influences the photocatalyst’s surface charge and particle aggregates’ size. Based on this data and data for the pKa of the substrate, degradation can be influenced via electrostatic interactions between the substrate and photocatalyst surface. However, electrostatic interactions are not always crucial for the process. When the pH level is equal to 7 and higher, OH are the main reactive species, while for lower pH levels, the main reactive species are positively charged h+ [198,199,200,201,202]. This also represents essential information since it can be determined through which reactive species the process is being performed.

4.4. Effect of Temperature

The characteristics of TiO2 material could potentially be modified by the temperature [203]. This induces modifications in nanoparticle size, crystalline phase composition, state of aggregation, and optical properties. All these aspects are of underlying importance because they affect the photocatalytic properties of the TiO2 surface [204]. TiO2 potential is highly based on its crystalline structure, average particle size, and morphology [205]. The specific surface area, which depends on particle size, is an essential parameter for the improved catalytical activity of TiO2 [206]. Over time, multiple methods have been employed to prepare TiO2 at the nanoscale level, such as hydrothermal, microemulsion, precipitation, and sol–gel [205,207]. The sol–gel method proved a suitable technique for synthesizing nanosized TiO2 at low temperatures with high photocatalytic activity, high purity, and homogeneity [208]. The sol–gel process modifies a liquid into a solid phase, sol to gel. The final step of the sol–gel process includes the calcination of colloids at different temperatures and time intervals, which causes the particles’ agglomeration and results in causing polydispersity [205].
Chen et al. prepared anatase powder at a calcination temperature of 400 °C, possessing a diameter of 10 nm and a specific surface area of 106.9 m2/g. Additionally, they found a rutile phase at calcination temperatures above 600 °C [209]. Tripathi et al. prepared mixed-phase and pure TiO2 by the sol–gel method calcined at a temperature from 400 °C to 700 °C with a particle size between 19 nm to 68 nm [210]. Velardi et al. [211] synthesized TiO2 powders at different calcination temperatures, which influenced the degree of crystallinity, the phase composition, the optical properties, the nanoparticle dimension, and the photocatalytic activity of TiO2 powders produced by sol–gel method. The obtained results have shown that the temperature increase induces a phase transformation from anatase to rutile. Additionally, a temperature increase leads to the rise in the grain size of TiO2 nanoparticles [211].

4.5. Influence of Oxidants/Electron Acceptor

The addition of electron acceptors significantly increases the efficiency of the photocatalytic process [212,213]. Electron acceptors prevent the recombination of e–h+ by separation of free charge carriers and an influence of e [214,215,216]. VB and CB for TiO2 originate from the 2p orbital of O2 and 3d orbital of Ti. The role of electron acceptor is most frequently reserved for molecular O2, which is quickly oxidized to O 2 (reaction (13)) and which is usually adsorbed on the TiO2 surface [217,218]. In this way, the recombination of e–h+ is prevented, while oxygen radicals are collected and can further take part in the degradation of contaminants [219].
Newly formed O 2 in acid solution gives H2O2, which can further oxidize organic compounds (reactions (14) and (15)):
e +   O 2   ads   O 2   ads  
O 2 +   H +     HO 2
2 HO 2   H 2 O 2 +   O 2
In some cases, photocatalytic activity is completely suppressed without O2. Besides O2, other substances could be used as electron acceptors, such as H2O2, KBrO3, and (NH4)2S2O8 [199,220,221,222,223]. H2O2 has a significant role in this group of electron acceptors since it efficiently suppresses the recombination of e−h+ pairs through reduction with e from the CB, which separates the charges (reaction (16)) and it represents the source of additional OH (reactions (16) and (17)) [224]:
H 2 O 2 +   e     OH   + OH
H 2 O 2 + h v 2 OH
Or H2O2 reacts with O 2 btained from O2 in reaction (13) (reaction (18)):
H 2 O 2 + O 2 OH + OH +   O 2
where O2 forms, which can further provide and in such a way initiates the process again [199].
KBrO3 and (NH4)2S2O8 through reduction with e give radicals which are strong oxidants such as SO 4 and BrO 2 as in the case of H2O2. This increases the quantum yield of processes catalyzed with TiO2. Reactions that include these two radicals are the following (reactions (19)–(22)) [220,225,226,227]:
S 2 O 8 2 +   e   SO 4 +   SO 4 2
SO 4 2 +   H 2 O   SO 4 2 + OH + H +
BrO 3 + 2 H + +   e   BrO 2 +   H 2 O
BrO 3 + 6 H + + 6 e BrO 2 ,   HOBr   Br + 3 H 2 O
While e scavenging and potential involvement of BrO 2 (formed in reaction (21)) could enhance degradation at elevated BrO 3 levels, the system reactivity could be limited by the formation of Br (reaction (22)). The latter could be adsorbed on the photocatalyst surface and involved in h+ and OH scavenging [66,228,229]. However, although electron acceptors increase the photocatalytic degradation rate, it was found that H2O2 has an inhibiting effect above the optimal concentrations, i.e., it decreases the degradation rate [199]. Namely, at higher concentrations, a peroxo complex is formed, which interferes with the degradation process, i.e., H2O2 acts as an absorber of h+ and OH, generating the much less reactive hydroperoxyl/superoxide radicals ( HO 2 / O 2 , reaction (23)). Moreover, HO 2 can further react with OH to form O2 and H2O, which are not directly involved in degradation of organic compound (reaction (24)) [199,228,230,231].
H 2 O 2 +   OH     HO 2 +   H 2 O   or   O 2 + H + +   H 2 O
HO 2 +   OH     O 2 +   H 2 O
For this reason, it is necessary to experimentally determine the optimal concentration of the investigated electron acceptor to bypass the negative effect on the process. Regarding electron acceptors, it can be concluded that their influence depends on the nature of the examined substance and the amount of the applied electron acceptor.

4.6. Effect of Inorganic Salts

The ability of some cations to decrease the photodegradation efficiency is related to the fact that they compete with organic compounds for the active sites on the surface of TiO2 [183]. This results in decreased degradation of targeted compounds since the photocatalyst is deactivated. Some examples of cations producing this effect are Cu2+ and Fe3+. It is determined that Ca2+, Mg2+, and Zn2+ have a negligible effect on the photodegradation of organic compounds. This occurrence is explained by the fact that mentioned cations are in their highest oxidation state [232].
Inorganic anions such as HCO 3 , CO 3 2 , Cl, NO 3 and SO 4 2 can be found in wastewater, and their presence causes colloidal instability, an increase in mass transfer, and a decrease in the surface contact between the target molecule and photocatalyst [233]. Consequently, the presence of the mentioned ions results in decreased degradation of organic compounds. Scavenge of OH through CO 3 2 and HCO 3 can be represented by the following reactions (25)–(27) [183,234]:
CO 3 2 +   OH     CO 3 + OH
HCO 3 + OH     CO 3 + H 2 O
CO 3 + OH     HCO 3
Further, Cl and its h+ and OH scavenging effect (reactions (28)–(30)) decrease the degradation efficiency [235,236,237].
Cl +   h +   Cl
Cl +   OH     HOCl
HOCl +   H +   Cl +   H 2 O
Further, active sites of the catalyst surface can also be blocked by the formed chloride radical anions. The chloride radicals generated in reactions (28) and (29) have a lower oxidation potential, namely, +2.47 V, than hydroxyl radicals, which have a redox potential of +2.80 V [238].
Wang et al. [239] reported the effect of NO 3 and SO 4 2 on the photodegradation of reactive Red 2 dye under UV irradiation, indicating increased dye removal. Zhu et al. [240] reported that the presence of NO 3 accelerated the photocatalytic degradation of an azo dye under visible light radiation. The enhancement of the degradation rate by NO 3 may be principally attributed to the direct or indirect forms of OH, reactions (31)–(33) [240,241]:
NO 3 + h v     NO 2 + O
NO 3 + H 2 O + h v     NO 2 + OH + OH
O + H 2 O   2 OH
SO 4 2 in photocatalytic degradation has two crucial roles. It increases aqueous ionic strength by guiding the molecule to the bulk interface, resulting in improved photocatalytic efficiency [242]. Additionally, the adsorbed SO 4 2 reacts with generated VB h+ forming SO 4 2 . This reaction between SO 4 2 and photogenerated h+ on the photocatalyst surface can prevent the e–h+ recombination, enhancing the photodegradation rate (reactions (34) and (35)):
SO 4 2 +   h + SO 4
SO 4 + H 2 O   SO 4 2 + OH   + H
SO 4 is a powerful oxidant with a redox potential of +2.6 V, which can enhance the degradation of an organic compound. The degradation rates also depend on the type of salt used. This effect has been studied by various authors [228,243,244,245].

4.7. Superhydrophilicity of TiO2

The laboratories of TOTO Inc. were conducting experiments, and by accident, they discovered the superhydrophilicity of TiO2. If a TiO2 film was prepared with a certain percentage of SiO2, it was found to acquire superhydrophilic properties, i.e., a water contact angle of ~0° after UV irradiation [246,247]. Additionally, it was found that photo-induced superhydrophilicity was an intrinsic property of the TiO2 surface [248]. Based on the reconstruction of the surface hydroxyl groups under UV irradiation, the mechanism of this process was proposed [249]. Figure 5 illustrates surface reconstruction on TiO2 during the reversible hydrophilic changes. The holes diffuse to the TiO2 surface, trapped at lattice oxygen atoms, while molecular oxygen captures photoexcited electrons. After that, hole capture weakens the binding energy between oxygen in the lattice and Ti atoms. Additionally, another adsorbed water molecule breaks this bond, forming new hydroxyl groups [250]. In the dark, the hydroxyl groups progressively desorb from the surface in the form of H2O2 or H2O and O2. Although it seems similar at first glance, the photo-induced hydrophilic conversion is a different process from the photocatalytic degradation of organic pollutants. Strontium titanate, which has almost the same photocatalytic oxidation strength as TiO2, does not become hydrophilic using UV radiation [251]. However, although WO3 exhibits photo-induced hydrophilic conversion, it does not exhibit photocatalytic activity [252]. One unique feature of TiO2 is that there are two different photo-induced phenomena, i.e., photo-induced superhydrophilic phenomenon and photocatalytic phenomenon [253]. Notwithstanding the fact they are different processes, they can, and in fact must take place simultaneously on the same TiO2 surface. The surface can have more superhydrophilic character and less photocatalytic character, or vice versa, which depends on the composition and the processing [11,254].
Applications of superhydrophilic technology are widespread nowadays [255]. One example is an antifogging surface which the superhydrophilic effect can prepare. Fogging of glass and mirrors surfaces appears when humid air condenses, with the formation of plenty of tiny water droplets that scatter light. Water droplets can not be formed on a superhydrophilic surface. Instead, a uniform film of water that does not scatter light can be formed on the surface. Additionally, superhydrophilic properties of TiO2 can help the self-cleaning process of TiO2 [255,256,257,258].

4.8. Aggregation and Agglomeration of TiO2 Particles

Photocatalyst loading depends strongly on the aggregate particle size when incident radiation is of a wavelength lower than 320 nm. The effect is primarily due to a low penetration depth of the radiation inside the aggregates, limiting the volume fraction of the irradiated aggregates [259].
The fact that three-dimensional TiO2 networks exist in aqueous suspensions served as a solid basis for Wang et al. [260] to propose a novel mechanism for improved photocatalytic processes. From that aspect, it is crucial to understand how the catalyst particles exist in these solutions and, thus, to understand their influence on the catalytic process and its efficiency. If the photocatalytic particles are aggregated/agglomerated and possess the same crystallographic orientation, then the photon energy can be transferred from particle to particle (Figure 6) [259]. This means that even if the targeted molecule is adsorbed on the photocatalytic particle, which is not activated by the photon energy, the photocatalytic process still can take part since the photon energy will be transferred to the particle to which it is adsorbed [71,260,261,262].
Once the energy has been transferred to the particle “containing” the targeted molecule, the latter will act as a hole trap, and the separation of the original exciton will be induced. So, the basic idea related to the mentioned mechanism is that the agglomeration can be thought of as a long chain of properly aligned TiO2 particles, which allow the transfer of the photon energy from the absorption location to the location of the particle which adsorbed the targeted molecule [263,264].
The aggregation/agglomeration could significantly affect all photocatalytic systems [265,266]. Photocatalytic powders usually consist of nanocrystalline primary particles aggregated to form secondary structures having dimensions in the micrometer range. Improved photocatalytic activity should be achieved when the strong electronic coupling between primary nanoparticles exists [265,267]. As could be concluded, the aggregation and agglomeration could increase the photocatalytic activities of TiO2. Namely, the target molecule does not necessarily need to be located where photon absorption occurs [260,268]. Additionally, the described effect imposes the challenge of designing and synthesizing photocatalytic material with the improved electronic coupling between the semiconducting nanoparticles, which could lead to a breakthrough in the field of photocatalytic degradation [269,270].

5. Practical Application

The possibility to activate catalysts with sunlight and recent advances in synthesis methods of the catalyst with desirable band gaps opened the opportunity to design prominent solar collectors where photochemical processes are promoted with the absorption of sunlight. While for solar thermal processes, it is essential to collect as many as possible photons of all wavelengths, for the solar photochemical process, it is crucial to collect only high-energy short-wavelength photons, which are responsible for the initiation of photochemical processes. Usually, for the initiation of solar photochemical processes, UV or near UV sunlight is necessary. However, there are some cases where the sunlight of up to 580 nm can be employed, while 600 nm and higher wavelengths are not usable for these purposes [6]. Initially, photoreactors for photochemical applications were based on line-focus parabolic-trough concentrators. This hardware already existed for solar thermal applications and could be easily modified for photochemical processes. The first European facility for water detoxification based on solar photochemical processes was established in Spain by The Centre for Energy, Environmental and Technological Research. Twelve PTCs were used in this case, while non-concentrating collectors became popular since the influence of concentration and solar tracking does not reduce their efficiency.
Researchers from the Institute of Science and Technology for Ceramics, Italy, developed a TiO2-coated fabric to be used as a photocatalyst agent to degrade Rhodamine B (RhB) in water. They implemented the obtained photocatalytic materials in a 6 L capacity semi-pilot plant. They evaluated the degradation of RhB dye, simulating the water pollution. The good results encouraged the scale-up of the 6 L semi-pilot plant up to the 100 L pilot plant built [271].
Biologically pretreated industrial wastewater from the factories of the Volkswagen AG in Wolfsburg (Germany) and Taubate (Brazil) has been treated in laboratory conditions with great success, after which a pilot plant was installed in the Wolfsburg factory in 1998 [272]. Another project, called “SOLARDETOX”, is worth mentioning. The name is an abbreviation for Solar Detoxification Technology for the Treatment of Industrial Non-Biodegradable Persistent Chlorinated Water Contaminants. This project aimed to design and develop a commercial non-concentrating solar detoxification system using neither collecting nor non-collecting collectors but compound parabolic collector technology (CPC), having a concentration ratio equal to 1. CPCs present a particular class of solar collectors produced in the shape of two meeting parabolas. The CPC collector belongs to a non-imaging class of collectors and is considered one of the types with the highest possible concentration ratio. The SOLARDETOX treatment plant is installed at the Hidrocen factory (Madrid, Spain).
Fendrich et al. [273] presented solar concentration technologies for wastewater remediation. They concluded that though mostly on model systems, recent results open promising perspectives for using concentrated sunlight as the energy source powering advanced oxidation processes, such as photocatalytic degradation by TiO2. Additionally, they identified the photocatalyst materials capable of efficiently working with sunlight and the transition to real wastewater investigation as the most critical issues to be addressed by research in the field.

6. Conclusions and Outlook

Using clean and self-sustainable technologies is imperative in front of the scientific community. TiO2 belongs to a group of materials with immense practical potential and, by all means, contributes to solving some of the emerging challenges in front of humanity. The energy coming from the Sun is the focus of present and future technologies. In contrast, photocatalysis, a mechanism closely related to this limitless energy source, has been recognized as a rapid and essential strategy to remove environmental pollutants. TiO2 stands out as the best and most commonly used photocatalyst, with a significant advantage over a broad range of nanomaterials.
In this work, we comprehensively introduced the most important features of TiO2 regarding its photocatalytic properties. For the ecological and economical application use of TiO2, it is essential to present the structural characteristics and physical properties of TiO2 nanoparticles. Therefore, this review introduces the fundamentals underlying the photocatalytic performance of nanostructured TiO2 related to its crystal structures and electronic and optical properties. Further, the photocatalytic activity of TiO2 and the mechanism were discussed in detail. Factors that affect photocatalytic properties of TiO2 to the greatest extent, such as type and dimensions of TiO2 particles, UV light intensity, the effect of pH, the influence of oxidants/electron acceptors, and the influence of inorganic salts, were explained. Computational aspects of designing the novel TiO2-based photocatalysis have also been presented. Attention was paid to the superhydrophilicity of TiO2, which is described by two different photo-induced phenomena: photo-induced superhydrophilic and photocatalytic phenomena taking place simultaneously on the same TiO2 surface. Although TiO2 is by far the best nanomaterial, there are still possibilities for its improvement.

Author Contributions

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

Funding

This work is supported by the Ministry of Education, Science and Technological Development of the Republic of Serbia (Grant No. 451-03-68/2022-14/200125).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Krakowiak, R.; Musial, J.; Bakun, P.; Spychała, M.; Czarczynska-Goslinska, B.; Mlynarczyk, D.T.; Koczorowski, T.; Sobotta, L.; Stanisz, B.; Goslinski, T. Titanium Dioxide-Based Photocatalysts for Degradation of Emerging Contaminants Including Pharmaceutical Pollutants. Appl. Sci. 2021, 11, 8674. [Google Scholar] [CrossRef]
  2. Elgohary, E.A.; Mohamed, Y.M.A.; El Nazer, H.A.; Baaloudj, O.; Alyami, M.S.S.; El Jery, A.; Assadi, A.A.; Amrane, A. A Review of the Use of Semiconductors as Catalysts in the Photocatalytic Inactivation of Microorganisms. Catalysts 2021, 11, 1498. [Google Scholar] [CrossRef]
  3. Guerrero, M.; Altube, A.; García-Lecina, E.; Rossinyol, E.; Baró, M.D.; Pellicer, E.; Sort, J. Facile in Situ Synthesis of BiOCl Nanoplates Stacked to Highly Porous TiO2: A Synergistic Combination for Environmental Remediation. ACS Appl. Mater. Interfaces 2014, 6, 13994–14000. [Google Scholar] [CrossRef] [PubMed]
  4. Lee, S.Y.; Park, S.J. TiO2 Photocatalyst for Water Treatment Applications. J. Ind. Eng. Chem. 2013, 19, 1761–1769. [Google Scholar] [CrossRef]
  5. Bahadoran, A.; De Lile, J.R.; Masudy-Panah, S.; Sadeghi, B.; Li, J.; Sabzalian, M.H.; Ramakrishna, S.; Liu, Q.; Cavaliere, P.; Gopinathan, A. Photocatalytic Materials Obtained from E-Waste Recycling: Review, Techniques, Critique, and Update. J. Manuf. Mater. Process. 2022, 6, 69. [Google Scholar] [CrossRef]
  6. Robert, D.; Malato, S. Solar Photocatalysis: A Clean Process for Water Detoxification. Sci. Total Environ. 2002, 291, 85–97. [Google Scholar] [CrossRef]
  7. Gowland Dan, C.A.; Neil, R.; Efthalia, C. Photocatalytic Oxidation of Natural Organic Matter in Water. Water 2020, 13, 288. [Google Scholar] [CrossRef]
  8. Uzelac, M.M.; Srđenović Čonić, B.; Kladar, N.; Armaković, S.; Armaković, S.J. Removal of Hydrochlorothiazide from Drinking and Environmental Water: Hydrolysis, Direct and Indirect Photolysis. Energy Environ. 2022, Article in press. [Google Scholar] [CrossRef]
  9. Burrows, H.D.; Santaballa, J.A.; Steenken, S. Reaction Pathways and Mechanisms of Photodegradation of Pesticides. J. Photochem. Photobiol. B Biol. 2002, 67, 71–108. [Google Scholar] [CrossRef] [Green Version]
  10. Menacherry, S.P.M.; Aravind, U.K.; Aravindakumar, C.T. Oxidative Degradation of Pharmaceutical Waste, Theophylline, from Natural Environment. Atmosphere 2022, 13, 835. [Google Scholar] [CrossRef]
  11. Fujishima, A.; Zhang, X. Titanium Dioxide Photocatalysis: Present Situation and Future Approaches. Comptes Rendus Chim. 2006, 9, 750–760. [Google Scholar] [CrossRef]
  12. Ahmed, S.; Rasul, M.G.; Martens, W.N.; Brown, R.; Hashib, M.A. Advances in Heterogeneous Photocatalytic Degradation of Phenols and Dyes in Wastewater: A Review. Water Air Soil Pollut. 2011, 215, 3–29. [Google Scholar] [CrossRef] [Green Version]
  13. Krithiga, T.; Sathish, S.; Renita, A.A.; Prabu, D.; Lokesh, S.; Geetha, R.; Namasivayam, S.K.R.; Sillanpaa, M. Persistent Organic Pollutants in Water Resources: Fate, Occurrence, Characterization and Risk Analysis. Sci. Total Environ. 2022, 831, 154808. [Google Scholar]
  14. Marathe, D.; Balbudhe, S.; Kumari, K. Persistent Organic Pollutants: A Global Issue, a Global Response. In Persistent Organic Pollutants; CRC Press: Boca Raton, FL, USA, 2021; pp. 1–32. ISBN 1003046800. [Google Scholar]
  15. Singh, G.; Singh, A.; Singh, P.; Mishra, V.K. Organic Pollutants in Groundwater Resource. Groundw. Geochem. Pollut. Remediat. Methods 2021, 1, 139–163. [Google Scholar]
  16. 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. [Google Scholar] [CrossRef]
  17. Andreozzi, R.; Caprio, V.; Marotta, R.; Radovnikovic, A. Ozonation and H2O2/UV Treatment of Clofibric Acid in Water: A Kinetic Investigation. J. Hazard. Mater. 2003, 103, 233–246. [Google Scholar] [CrossRef]
  18. Sasakova, N.; Gregova, G.; Takacova, D.; Mojzisova, J.; Papajova, I.; Venglovsky, J.; Szaboova, T.; Kovacova, S. Pollution of Surface and Ground Water by Sources Related to Agricultural Activities. Front. Sustain. Food Syst. 2018, 2, 42. [Google Scholar] [CrossRef] [Green Version]
  19. Ma, Y.; Halsall, C.J.; Crosse, J.D.; Graf, C.; Cai, M.; He, J.; Gao, G.; Jones, K. Persistent Organic Pollutants in Ocean Sediments from the N Orth P Acific to the A Rctic O Cean. J. Geophys. Res. Ocean. 2015, 120, 2723–2735. [Google Scholar] [CrossRef] [Green Version]
  20. Dachs, J.; Méjanelle, L. Organic Pollutants in Coastal Waters, Sediments, and Biota: A Relevant Driver for Ecosystems during the Anthropocene? Estuaries Coasts 2010, 33, 1–14. [Google Scholar] [CrossRef]
  21. Behnajady, M.A.; Modirshahla, N. Kinetic Modeling on Photooxidative Degradation of C.I. Acid Orange 7 in a Tubular Continuous-Flow Photoreactor. Chemosphere 2006, 62, 1543–1548. [Google Scholar] [CrossRef]
  22. Xia, X.H.; Liang, Y.; Wang, Z.; Fan, J.; Luo, Y.S.; Jia, Z.J. Synthesis and Photocatalytic Properties of TiO2 Nanostructures. Mater. Res. Bull. 2008, 43, 2187–2195. [Google Scholar] [CrossRef]
  23. Ceballos-Chuc, M.C.; Ramos-Castillo, C.M.; Rodríguez-Pérez, M.; Ruiz-Gómez, M.Á.; Rodríguez-Gattorno, G.; Villanueva-Cab, J. Synergistic Correlation in the Colloidal Properties of TiO2 Nanoparticles and Its Impact on the Photocatalytic Activity. Inorganics 2022, 10, 125. [Google Scholar] [CrossRef]
  24. Reghunath, S.; Pinheiro, D.; KR, S.D. A Review of Hierarchical Nanostructures of TiO2: Advances and Applications. Appl. Surf. Sci. Adv. 2021, 3, 100063. [Google Scholar] [CrossRef]
  25. Verma, R.; Gangwar, J.; Srivastava, A.K. Multiphase TiO2 Nanostructures: A Review of Efficient Synthesis, Growth Mechanism, Probing Capabilities, and Applications in Bio-Safety and Health. RSC Adv. 2017, 7, 44199–44224. [Google Scholar] [CrossRef] [Green Version]
  26. Landmann, M.; Rauls, E.; Schmidt, W.G. The Electronic Structure and Optical Response of Rutile, Anatase and Brookite TiO 2. J. Phys. Condens. Matter 2012, 24, 195503. [Google Scholar] [CrossRef] [Green Version]
  27. Siddiqui, H. Modification of Physical and Chemical Properties of Titanium Dioxide (TiO 2) by Ion Implantation for Dye Sensitized Solar Cells. In Ion Beam Techniques and Applications; IntechOpen: London, UK, 2019; ISBN 1789845718. [Google Scholar]
  28. Reyes-Coronado, D.; Rodríguez-Gattorno, G.; Espinosa-Pesqueira, M.E.; Cab, C.; De Coss, R.; Oskam, G. Phase-Pure TiO2 Nanoparticles: Anatase, Brookite and Rutile. Nanotechnology 2008, 19, 145605. [Google Scholar] [CrossRef]
  29. Navrotsky, A. Energetics of Nanoparticle Oxides: Interplay between Surface Energy and Polymorphism. Geochem. Trans. 2003, 4, 34. [Google Scholar] [CrossRef]
  30. Li, J.-G.; Ishigaki, T.; Sun, X. Anatase, Brookite, and Rutile Nanocrystals via Redox Reactions under Mild Hydrothermal Conditions: Phase-Selective Synthesis and Physicochemical Properties. J. Phys. Chem. C 2007, 111, 4969–4976. [Google Scholar] [CrossRef]
  31. Pottier, A.; Chanéac, C.; Tronc, E.; Mazerolles, L.; Jolivet, J.-P. Synthesis of Brookite TiO2 Nanoparticlesby Thermolysis of TiCl4 in Strongly Acidic Aqueous Media. J. Mater. Chem. 2001, 11, 1116–1121. [Google Scholar] [CrossRef]
  32. Kondamareddy, K.K.; Neena, D.; Lu, D.; Peng, T.; Lopez, M.A.M.; Wang, C.; Yu, Z.; Cheng, N.; Fu, D.J.; Zhao, X.-Z. Ultra-Trace (Parts per Million-Ppm) W6+ Dopant Ions Induced Anatase to Rutile Transition (ART) of Phase Pure Anatase TiO2 Nanoparticles for Highly Efficient Visible Light-Active Photocatalytic Degradation of Organic Pollutants. Appl. Surf. Sci. 2018, 456, 676–693. [Google Scholar] [CrossRef]
  33. Momma, K.; Izumi, F. VESTA 3 for Three-Dimensional Visualization of Crystal, Volumetric and Morphology Data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  34. Hanaor, D.A.H.; Sorrell, C.C. Review of the Anatase to Rutile Phase Transformation. J. Mater. Sci. 2011, 46, 855–874. [Google Scholar] [CrossRef]
  35. Siwińska-Stefańska, K.; Krysztafkiewicz, A.; Ciesielczyk, F.; Paukszta, D.; Sójka-Ledakowicz, J.; Jesionowski, T. Physicochemical and Structural Properties of TiO2 Precipitated in an Emulsion System. Physicochem. Probl. Miner. Process. 2010, 44, 231–244. [Google Scholar]
  36. Patra, S.; Davoisne, C.; Bouyanfif, H.; Foix, D.; Sauvage, F. Phase Stability Frustration on Ultra-Nanosized Anatase TiO2. Sci. Rep. 2015, 5, 10928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Holm, A.; Hamandi, M.; Simonet, F.; Jouguet, B.; Dappozze, F.; Guillard, C. Impact of Rutile and Anatase Phase on the Photocatalytic Decomposition of Lactic Acid. Appl. Catal. B Environ. 2019, 253, 96–104. [Google Scholar] [CrossRef]
  38. Bhatkhande, D.S.; Pangarkar, V.G.; Beenackers, A.A.C.M. Photocatalytic Degradation for Environmental Applications–a Review. J. Chem. Technol. Biotechnol. Int. Res. Process. Environ. Clean Technol. 2002, 77, 102–116. [Google Scholar] [CrossRef]
  39. Lan, Y.; Lu, Y.; Ren, Z. Mini Review on Photocatalysis of Titanium Dioxide Nanoparticles and Their Solar Applications. Nano Energy 2013, 2, 1031–1045. [Google Scholar] [CrossRef]
  40. Lydakis-Simantiris, N.; Riga, D.; Katsivela, E.; Mantzavinos, D.; Xekoukoulotakis, N.P. Disinfection of Spring Water and Secondary Treated Municipal Wastewater by TiO2 Photocatalysis. Desalination 2010, 250, 351–355. [Google Scholar] [CrossRef]
  41. Chatzitakis, A.; Berberidou, C.; Paspaltsis, I.; Kyriakou, G.; Sklaviadis, T.; Poulios, I. Photocatalytic Degradation and Drug Activity Reduction of Chloramphenicol. Water Res. 2008, 42, 386–394. [Google Scholar] [CrossRef]
  42. Vijayabalan, A.; Selvam, K.; Krishnakumar, B.; Swaminathan, M. Photocatalytic Degradation of Reactive Orange 4 by Surface Fluorinated TiO2 Wackherr under UV-A Light. Sep. Purif. Technol. 2013, 108, 51–56. [Google Scholar] [CrossRef]
  43. Selvam, K.; Swaminathan, M. Photocatalytic Synthesis of 2-Methylquinolines with TiO2 Wackherr and Home Prepared TiO2–A Comparative Study. Arab. J. Chem. 2017, 10, S28–S34. [Google Scholar] [CrossRef] [Green Version]
  44. Vione, D.; Minero, C.; Maurino, V.; Carlotti, M.E.; Picatonotto, T.; Pelizzetti, E. Degradation of Phenol and Benzoic Acid in the Presence of a TiO2-Based Heterogeneous Photocatalyst. Appl. Catal. B Environ. 2005, 58, 79–88. [Google Scholar] [CrossRef]
  45. Li, D.; Song, H.; Meng, X.; Shen, T.; Sun, J.; Han, W.; Wang, X. Effects of Particle Size on the Structure and Photocatalytic Performance by Alkali-Treated TiO2. Nanomaterials 2020, 10, 546. [Google Scholar] [CrossRef]
  46. Könenkamp, R. Carrier Transport in Nanoporous TiO 2 Films. Phys. Rev. B 2000, 61, 11057. [Google Scholar] [CrossRef]
  47. Zhang, K.; Lin, Y.; Muhammad, Z.; Wu, C.; Yang, S.; He, Q.; Zheng, X.; Chen, S.; Ge, B.; Song, L. Active {010} Facet-Exposed Cu2MoS4 Nanotube as High-Efficiency Photocatalyst. Nano Res. 2017, 10, 3817–3825. [Google Scholar] [CrossRef]
  48. Che, M.; Védrine, J.C. Characterization of Solid Materials and Heterogeneous Catalysts: From Structure to Surface Reactivity; John Wiley & Sons: Hoboken, NJ, USA, 2012; ISBN 3527645330. [Google Scholar]
  49. Kowalska, E.; Wei, Z.; Janczarek, M. Band-Gap Engineering of Photocatalysts: Surface Modification versus Doping. In Visible-Light-Active Photocatalysis: Nanostructured Catalyst Design, Mechanisms, and Applications; Wiley: Hoboken, NJ, USA, 2018; pp. 449–484. [Google Scholar]
  50. Armaković, S.; Armaković, S.J. Atomistica. Online–Web Application for Generating Input Files for ORCA Molecular Modelling Package Made with the Anvil Platform. Mol. Simul. 2022, 1–7. [Google Scholar] [CrossRef]
  51. Tsyshevsky, R.V.; Pagoria, P.; Kuklja, M.M. Computational Design of Novel Energetic Materials: Dinitro-Bis-Triazolo-Tetrazine. J. Phys. Chem. C 2015, 119, 8512–8521. [Google Scholar] [CrossRef]
  52. Perdew, J.P. Density Functional Theory and the Band Gap Problem. Int. J. Quantum Chem. 1985, 28, 497–523. [Google Scholar] [CrossRef]
  53. Perdew, J.P.; Yang, W.; Burke, K.; Yang, Z.; Gross, E.K.U.; Scheffler, M.; Scuseria, G.E.; Henderson, T.M.; Zhang, I.Y.; Ruzsinszky, A. Understanding Band Gaps of Solids in Generalized Kohn–Sham Theory. Proc. Natl. Acad. Sci. USA 2017, 114, 2801–2806. [Google Scholar] [CrossRef] [Green Version]
  54. Perdew, J.P.; Levy, M. Physical Content of the Exact Kohn-Sham Orbital Energies: Band Gaps and Derivative Discontinuities. Phys. Rev. Lett. 1983, 51, 1884. [Google Scholar] [CrossRef]
  55. Mori-Sánchez, P.; Cohen, A.J.; Yang, W. Localization and Delocalization Errors in Density Functional Theory and Implications for Band-Gap Prediction. Phys. Rev. Lett. 2008, 100, 146401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid Functionals Based on a Screened Coulomb Potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef] [Green Version]
  57. Borlido, P.; Aull, T.; Huran, A.W.; Tran, F.; Marques, M.A.L.; Botti, S. Large-Scale Benchmark of Exchange–Correlation Functionals for the Determination of Electronic Band Gaps of Solids. J. Chem. Theory Comput. 2019, 15, 5069–5079. [Google Scholar] [CrossRef]
  58. Borlido, P.; Schmidt, J.; Huran, A.W.; Tran, F.; Marques, M.A.L.; Botti, S. Exchange-Correlation Functionals for Band Gaps of Solids: Benchmark, Reparametrization and Machine Learning. npj Comput. Mater. 2020, 6, 96. [Google Scholar] [CrossRef]
  59. Cococcioni, M. The LDA+ U Approach: A Simple Hubbard Correction for Correlated Ground States. Correl. Electrons From Model. Mater. Model. Simul. 2012, 2, 1–33. [Google Scholar]
  60. Zhang, R.; Zhao, J.; Yang, Y.; Lu, Z.; Shi, W. Understanding Electronic and Optical Properties of La and Mn Co-Doped Anatase TiO2. Comput. Condens. Matter 2016, 6, 5–17. [Google Scholar] [CrossRef] [Green Version]
  61. Chen, W.; Yuan, P.; Zhang, S.; Sun, Q.; Liang, E.; Jia, Y. Electronic Properties of Anatase TiO2 Doped by Lanthanides: A DFT+ U Study. Phys. B Condens. Matter 2012, 407, 1038–1043. [Google Scholar] [CrossRef]
  62. Li, Z.; Li, Z.; Zuo, C.; Fang, X. Application of Nanostructured TiO2 in UV Photodetectors: A Review. Adv. Mater. 2022, 34, 2109083. [Google Scholar] [CrossRef]
  63. Yang, H.G.; Sun, C.H.; Qiao, S.Z.; Zou, J.; Liu, G.; Smith, S.C.; Cheng, H.M.; Lu, G.Q. Anatase TiO2 Single Crystals with a Large Percentage of Reactive Facets. Nature 2008, 453, 638–641. [Google Scholar] [CrossRef] [Green Version]
  64. Zhao, H.; Pan, F.; Li, Y. A Review on the Effects of TiO2 Surface Point Defects on CO2 Photoreduction with H2O. J. Mater. 2017, 3, 17–32. [Google Scholar] [CrossRef]
  65. Wrana, D.; Gensch, T.; Jany, B.R.; Cieślik, K.; Rodenbücher, C.; Cempura, G.; Kruk, A.; Krok, F. Photoluminescence Imaging of Defects in TiO2: The Influence of Grain Boundaries and Doping on Charge Carrier Dynamics. Appl. Surf. Sci. 2021, 569, 150909. [Google Scholar] [CrossRef]
  66. Wen, B.; Hao, Q.; Yin, W.-J.; Zhang, L.; Wang, Z.; Wang, T.; Zhou, C.; Selloni, A.; Yang, X.; Liu, L.-M. Electronic Structure and Photoabsorption of Ti 3+ Ions in Reduced Anatase and Rutile TiO 2. Phys. Chem. Chem. Phys. 2018, 20, 17658–17665. [Google Scholar] [CrossRef]
  67. Dharmale, N.; Chaudhury, S.; Kar, J.K. Various Exchange-Correlation Effects on Structural, Electronic, and Optical Properties of Brookite TiO2. ECS J. Solid State Sci. Technol. 2021, 10, 83010. [Google Scholar] [CrossRef]
  68. Armaković, S.J.; Mary, Y.S.; Mary, Y.S.; Pelemiš, S.; Armaković, S. Optoelectronic Properties of the Newly Designed 1, 3, 5-Triazine Derivatives with Isatin, Chalcone and Acridone Moieties. Comput. Theor. Chem. 2021, 1197, 113160. [Google Scholar] [CrossRef]
  69. Del Angel, R.; Durán-Álvarez, J.C.; Zanella, R. TiO2-Low Band Gap Semiconductor Heterostructures for Water Treatment Using Sunlight-Driven Photocatalysis. In Titanium Dioxide: Material for a Sustainable Environment; IntechOpen: London, UK, 2018; Volume 305. [Google Scholar]
  70. Etacheri, V.; Di Valentin, C.; Schneider, J.; Bahnemann, D.; Pillai, S.C. Visible-Light Activation of TiO2 Photocatalysts: Advances in Theory and Experiments. J. Photochem. Photobiol. C Photochem. Rev. 2015, 25, 1–29. [Google Scholar] [CrossRef] [Green Version]
  71. Yang, H.; Yang, B.; Chen, W.; Yang, J. Preparation and Photocatalytic Activities of TiO2 -Based Composite Catalysts. Catalysts 2022, 12, 1263. [Google Scholar] [CrossRef]
  72. Pawar, M.; Topcu Sendoğdular, S.; Gouma, P. A Brief Overview of TiO2 Photocatalyst for Organic Dye Remediation: Case Study of Reaction Mechanisms Involved in Ce-TiO2 Photocatalysts System. J. Nanomater. 2018, 2018, 5953609. [Google Scholar] [CrossRef] [Green Version]
  73. Zhang, Z.; Yates, J.T., Jr. Direct Observation of Surface-Mediated Electron− Hole Pair Recombination in TiO2 (110). J. Phys. Chem. C 2010, 114, 3098–3101. [Google Scholar] [CrossRef]
  74. Kubovics, M.; Silva, G.; Ana, M.L.; Faria, J.L. Photocatalytic Hydrogen Production Using Porous 3D Graphene-Based Aerogels Supporting Pt/TiO2 Nanoparticles. Gels 2022, 8, 719. [Google Scholar] [CrossRef]
  75. Mosquera-Vargas, E.; Herrera-Molina, D.; Diosa, J.E. Structural and Optical Properties of TiO2 Nanoparticles and Their Photocatalytic Behavior under Visible Light. Ing. Compet. 2021, 23, 2. [Google Scholar]
  76. Luttrell, T.; Halpegamage, S.; Tao, J.; Kramer, A.; Sutter, E.; Batzill, M. Why Is Anatase a Better Photocatalyst than Rutile?-Model Studies on Epitaxial TiO2 Films. Sci. Rep. 2014, 4, 4043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Glassford, K.M.; Chelikowsky, J.R. Optical Properties of Titanium Dioxide in the Rutile Structure. Phys. Rev. B 1992, 45, 3874. [Google Scholar] [CrossRef] [PubMed]
  78. Govindasamy, G.; Murugasen, P.; Sagadevan, S. Investigations on the Synthesis, Optical and Electrical Properties of TiO2 Thin Films by Chemical Bath Deposition (CBD) Method. Mater. Res. 2016, 19, 413–419. [Google Scholar] [CrossRef] [Green Version]
  79. Soussi, A.; Ait Hssi, A.; Boujnah, M.; Boulkadat, L.; Abouabassi, K.; Asbayou, A.; Elfanaoui, A.; Markazi, R.; Ihlal, A.; Bouabid, K. Electronic and Optical Properties of TiO2 Thin Films: Combined Experimental and Theoretical Study. J. Electron. Mater. 2021, 50, 4497–4510. [Google Scholar] [CrossRef]
  80. Huang, F.; Yan, A.; Zhao, H. Influences of Doping on Photocatalytic Properties of TiO2 Photocatalyst. In Semiconductor Photocatalysis-Materials, Mechanisms and Applications; InTech: Rang-Du-Fliers, France, 2016; pp. 31–80. [Google Scholar]
  81. Madima, N.; Kefeni, K.K.; Mishra, S.B.; Mishra, A.K.; Kuvarega, A.T. Fabrication of Magnetic Recoverable Fe3O4/TiO2 Heterostructure for Photocatalytic Degradation of Rhodamine B Dye. Inorg. Chem. Commun. 2022, 145, 109966. [Google Scholar] [CrossRef]
  82. Amin, S.; Sher, M.; Ali, A.; Rehman, M.F.; Hayat, A.; Ikram, M.; Abbas, A.; Amin, H.M.A. Sulfonamide-Functionalized Silver Nanoparticles as an Analytical Nanoprobe for Selective Ni (II) Sensing with Synergistic Antimicrobial Activity. Environ. Nanotechnol. Monit. Manag. 2022, 18, 100735. [Google Scholar] [CrossRef]
  83. Kumar, A.; Khan, M.; He, J.; Lo, I.M.C. Recent Developments and Challenges in Practical Application of Visible–Light–Driven TiO2–Based Heterojunctions for PPCP Degradation: A Critical Review. Water Res. 2020, 170, 115356. [Google Scholar] [CrossRef]
  84. Hua, X.; Liu, Z.; Bruce, P.G.; Grey, C.P. The Morphology of TiO2 (B) Nanoparticles. J. Am. Chem. Soc. 2015, 137, 13612–13623. [Google Scholar] [CrossRef]
  85. Goulart, S.; Nieves, L.J.J.; Dal Bó, A.G.; Bernardin, A.M. Sensitization of TiO2 Nanoparticles with Natural Dyes Extracts for Photocatalytic Activity under Visible Light. Dye. Pigment. 2020, 182, 108654. [Google Scholar] [CrossRef]
  86. Varshney, G.; Kanel, S.R.; Kempisty, D.M.; Varshney, V.; Agrawal, A.; Sahle-Demessie, E.; Varma, R.S.; Nadagouda, M.N. Nanoscale TiO2 Films and Their Application in Remediation of Organic Pollutants. Coord. Chem. Rev. 2016, 306, 43–64. [Google Scholar] [CrossRef]
  87. Žener, B.; Matoh, L.; Reli, M.; Škapin, A.S.; Korošec, R.C. Metal and Non-Metal Modified Titania: The Effect of Phase Composition and Surface Area on Photocatalytic Activity. Acta Chim. Slov. 2022, 69, 217–226. [Google Scholar]
  88. Wu, J.; Lu, S.; Ge, D.; Zhang, L.; Chen, W.; Gu, H. Photocatalytic Properties of Pd/TiO2 Nanosheets for Hydrogen Evolution from Water Splitting. RSC Adv. 2016, 6, 67502–67508. [Google Scholar] [CrossRef]
  89. Yu, F.; Wang, C.; Ma, H.; Song, M.; Li, D.; Li, Y.; Li, S.; Zhang, X.; Liu, Y. Revisiting Pt/TiO2 Photocatalysts for Thermally Assisted Photocatalytic Reduction of CO2. Nanoscale 2020, 12, 7000–7010. [Google Scholar] [CrossRef]
  90. Sadrieyeh, S.; Malekfar, R. Photocatalytic Performance of Plasmonic Au/Ag-TiO2 Aerogel Nanocomposites. J. Non. Cryst. Solids 2018, 489, 33–39. [Google Scholar] [CrossRef]
  91. Cao, X.; Yang, X.; Li, H.; Huang, W.; Liu, X. Investigation of Ce-TiO2 Photocatalyst and Its Application in Asphalt-Based Specimens for NO Degradation. Constr. Build. Mater. 2017, 148, 824–832. [Google Scholar] [CrossRef]
  92. Sood, S.; Umar, A.; Mehta, S.K.; Sinha, A.S.K.; Kansal, S.K. Efficient Photocatalytic Degradation of Brilliant Green Using Sr-Doped TiO2 Nanoparticles. Ceram. Int. 2015, 41, 3533–3540. [Google Scholar] [CrossRef]
  93. Rossi, L.; Palacio, M.; Villabrille, P.I.; Rosso, J.A. V-Doped TiO2 Photocatalysts and Their Application to Pollutant Degradation. Environ. Sci. Pollut. Res. 2021, 28, 24112–24123. [Google Scholar] [CrossRef]
  94. Zheng, S.K.; Wang, T.M.; Hao, W.C.; Shen, R. Improvement of Photocatalytic Activity of TiO2 Thin Film by Sn Ion Implantation. Vacuum 2002, 65, 155–159. [Google Scholar] [CrossRef]
  95. Choudhury, B.; Bayan, S.; Choudhury, A.; Chakraborty, P. Narrowing of Band Gap and Effective Charge Carrier Separation in Oxygen Deficient TiO2 Nanotubes with Improved Visible Light Photocatalytic Activity. J. Colloid Interface Sci. 2016, 465, 1–10. [Google Scholar] [CrossRef]
  96. Yalçın, Y.; Kılıç, M.; Cinar, Z. The Role of Non-Metal Doping in TiO2 Photocatalysis. J. Adv. Oxid. Technol. 2010, 13, 281–296. [Google Scholar] [CrossRef]
  97. Arora, I.; Chawla, H.; Chandra, A.; Sagadevan, S.; Garg, S. Advances in the Strategies for Enhancing the Photocatalytic Activity of TiO2: Conversion from UV-Light Active to Visible-Light Active Photocatalyst. Inorg. Chem. Commun. 2022, 143, 109700. [Google Scholar] [CrossRef]
  98. Angela, S.; Lunardi, V.B.; Kusuma, K.; Soetaredjo, F.E.; Putro, J.N.; Santoso, S.P.; Angkawijaya, A.E.; Lie, J.; Gunarto, C.; Kurniawan, A. Facile synthesis of hierarchical porous ZIF-8@TiO2 for simultaneous adsorption and photocatalytic decomposition of crystal violet. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100598. [Google Scholar] [CrossRef]
  99. Janczarek, M.; Kowalska, E. On the Origin of Enhanced Photocatalytic Activity of Copper-Modified Titania in the Oxidative Reaction Systems. Catalysts 2017, 7, 317. [Google Scholar] [CrossRef] [Green Version]
  100. Tian, Y.; Yang, H.; Wu, S.; Gong, B.; Xu, C.; Yan, J.; Cen, K.; Bo, Z.; Ostrikov, K. High-performance Water Purification and Desalination by Solar-driven Interfacial Evaporation and Photocatalytic VOC Decomposition Enabled by Hierarchical TiO2-CuO Nanoarchitecture. Int. J. Energy Res. 2022, 46, 1313–1326. [Google Scholar] [CrossRef]
  101. Humayun, M.; Raziq, F.; Khan, A.; Luo, W. Modification Strategies of TiO2 for Potential Applications in Photocatalysis: A Critical Review. Green Chem. Lett. Rev. 2018, 11, 86–102. [Google Scholar] [CrossRef] [Green Version]
  102. Siwińska-Stefańska, K.; Kubiak, A.; Piasecki, A.; Goscianska, J.; Nowaczyk, G.; Jurga, S.; Jesionowski, T. TiO2-ZnO Binary Oxide Systems: Comprehensive Characterization and Tests of Photocatalytic Activity. Materials 2018, 11, 841. [Google Scholar] [CrossRef]
  103. Beyers, E.; Biermans, E.; Ribbens, S.; De Witte, K.; Mertens, M.; Meynen, V.; Bals, S.; Van Tendeloo, G.; Vansant, E.F.; Cool, P. Combined TiO2/SiO2 Mesoporous Photocatalysts with Location and Phase Controllable TiO2 Nanoparticles. Appl. Catal. B Environ. 2009, 88, 515–524. [Google Scholar] [CrossRef]
  104. Xiong, L.; Yang, F.; Yan, L.; Yan, N.; Yang, X.; Qiu, M.; Yu, Y. Bifunctional Photocatalysis of TiO2/Cu2O Composite under Visible Light: Ti3+ in Organic Pollutant Degradation and Water Splitting. J. Phys. Chem. Solids 2011, 72, 1104–1109. [Google Scholar] [CrossRef]
  105. Bessekhouad, Y.; Robert, D.; Weber, J.-V. Photocatalytic Activity of Cu2O/TiO2, Bi2O3/TiO2 and ZnMn2O4/TiO2 Heterojunctions. Catal. Today 2005, 101, 315–321. [Google Scholar] [CrossRef]
  106. Padmanabhan, N.T.; Thomas, N.; Louis, J.; Mathew, D.T.; Ganguly, P.; John, H.; Pillai, S.C. Graphene Coupled TiO2 Photocatalysts for Environmental Applications: A Review. Chemosphere 2021, 271, 129506. [Google Scholar] [CrossRef]
  107. Wang, W.-S.; Wang, D.-H.; Qu, W.-G.; Lu, L.-Q.; Xu, A.-W. Large Ultrathin Anatase TiO2 Nanosheets with Exposed {001} Facets on Graphene for Enhanced Visible Light Photocatalytic Activity. J. Phys. Chem. C 2012, 116, 19893–19901. [Google Scholar] [CrossRef]
  108. Chakhtouna, H.; Benzeid, H.; Zari, N.; Bouhfid, R. Recent Progress on Ag/TiO2 Photocatalysts: Photocatalytic and Bactericidal Behaviors. Environ. Sci. Pollut. Res. 2021, 28, 44638–44666. [Google Scholar] [CrossRef]
  109. Kerketta, U.; Tesler, A.B. Single-Atom Co-Catalysts Employed in Titanium Dioxide Photocatalysis. Catalysts 2022, 12, 1223. [Google Scholar] [CrossRef]
  110. Bussi, J.; Ohanian, M.; Vázquez, M.; Dalchiele, E.A. Photocatalytic Removal of Hg from Solid Wastes of Chlor-Alkali Plant. J. Environ. Eng. 2002, 128, 733–739. [Google Scholar] [CrossRef]
  111. Minero, C.; Vione, D. A Quantitative Evalution of the Photocatalytic Performance of TiO2 Slurries. Appl. Catal. B Environ. 2006, 67, 257–269. [Google Scholar] [CrossRef]
  112. Konstantinou, I.K.; Albanis, T.A. TiO2-Assisted Photocatalytic Degradation of Azo Dyes in Aqueous Solution: Kinetic and Mechanistic Investigations: A Review. Appl. Catal. B Environ. 2004, 49, 1–14. [Google Scholar] [CrossRef]
  113. Krystynik, P. Advanced Oxidation Processes (AOPs)–Utilization of Hydroxyl Radical and Singlet Oxygen; IntechOpen: London, UK, 2021; ISBN 1839682825. [Google Scholar]
  114. Wang, B.; Zhang, H.; Wang, F.; Xiong, X.; Tian, K.; Sun, Y.; Yu, T. Application of Heterogeneous Catalytic Ozonation for Refractory Organics in Wastewater. Catalysts 2019, 9, 241. [Google Scholar] [CrossRef] [Green Version]
  115. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  116. Turchi, C.S.; Ollis, D.F. Photocatalytic Degradation of Organic Water Contaminants: Mechanisms Involving Hydroxyl Radical Attack. J. Catal. 1990, 122, 178–192. [Google Scholar] [CrossRef]
  117. Jiang, D.; Zhao, H.; Zhang, S.; John, R. Kinetic Study of Photocatalytic Oxidation of Adsorbed Carboxylic Acids at TiO2 Porous Films by Photoelectrolysis. J. Catal. 2004, 223, 212–220. [Google Scholar] [CrossRef]
  118. Herrmann, J.-M.; Guillard, C. Photocatalytic Degradation of Pesticides in Agricultural Used Waters. Comptes Rendus L’académie Des Sci. IIC Chem. 2000, 3, 417–422. [Google Scholar] [CrossRef]
  119. Santiago-Morales, J.; Agüera, A.; del Gómez, M.M.; Fernández-Alba, A.R.; Giménez, J.; Esplugas, S.; Rosal, R. Transformation Products and Reaction Kinetics in Simulated Solar Light Photocatalytic Degradation of Propranolol Using Ce-Doped TiO2. Appl. Catal. B Environ. 2013, 129, 13–29. [Google Scholar] [CrossRef]
  120. Yang, H.; Li, G.; An, T.; Gao, Y.; Fu, J. Photocatalytic Degradation Kinetics and Mechanism of Environmental Pharmaceuticals in Aqueous Suspension of TiO2: A Case of Sulfa Drugs. Catal. Today 2010, 153, 200–207. [Google Scholar] [CrossRef]
  121. Anucha, C.B.; Altin, I.; Bacaksiz, E.; Stathopoulos, V.N. Titanium Dioxide (TiO₂)-Based Photocatalyst Materials Activity Enhancement for Contaminants of Emerging Concern (CECs) Degradation: In the Light of Modification Strategies. Chem. Eng. J. Adv. 2022, 10, 100262. [Google Scholar] [CrossRef]
  122. Ishibashi, K.; Fujishima, A.; Watanabe, T.; Hashimoto, K. Detection of Active Oxidative Species in TiO2 Photocatalysis Using the Fluorescence Technique. Electrochem. Commun. 2000, 2, 207–210. [Google Scholar] [CrossRef]
  123. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  124. Du, Y.; Rabani, J. The Measure of TiO2 Photocatalytic Efficiency and the Comparison of Different Photocatalytic Titania. J. Phys. Chem. B 2003, 107, 11970–11978. [Google Scholar] [CrossRef]
  125. Linsebigler, A.L.; Lu, G.; Yates, J.T., Jr. Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  126. Chiu, Y.-H.; Chang, T.-F.M.; Chen, C.-Y.; Sone, M.; Hsu, Y.-J. Mechanistic Insights into Photodegradation of Organic Dyes Using Heterostructure Photocatalysts. Catalysts 2019, 9, 430. [Google Scholar] [CrossRef] [Green Version]
  127. Alikarami, M.; Soltani, R.D.C.; Khataee, A. An Innovative Combination of Electrochemical and Photocatalytic Processes for Decontamination of Bisphenol A Endocrine Disruptor Form Aquatic Phase: Insight into Mechanism, Enhancers and Bio-Toxicity Assay. Sep. Purif. Technol. 2019, 220, 42–51. [Google Scholar] [CrossRef]
  128. Kabra, K.; Chaudhary, R.; Sawhney, R.L. Treatment of Hazardous Organic and Inorganic Compounds through Aqueous-Phase Photocatalysis: A Review. Ind. Eng. Chem. Res. 2004, 43, 7683–7696. [Google Scholar] [CrossRef]
  129. Chong, M.N.; Jin, B.; Chow, C.W.K.; Saint, C. Recent Developments in Photocatalytic Water Treatment Technology: A Review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef] [PubMed]
  130. Liochev, S.I. Reactive Oxygen Species and the Free Radical Theory of Aging. Free Radic. Biol. Med. 2013, 60, 1–4. [Google Scholar] [CrossRef] [PubMed]
  131. Sharma, S.; Kumar, R.; Raizada, P.; Ahamad, T.; Alshehri, S.M.; Nguyen, V.-H.; Thakur, S.; Nguyen, C.C.; Kim, S.Y.; Van Le, Q. An Overview on Recent Progress in Photocatalytic Air Purification: Metal-Based and Metal-Free Photocatalysis. Environ. Res. 2022, 214, 113995. [Google Scholar] [CrossRef] [PubMed]
  132. Mehara, J.; Roithová, J. Identifying Reactive Intermediates by Mass Spectrometry. Chem. Sci. 2020, 11, 11960–11972. [Google Scholar] [CrossRef]
  133. Halasz, A.; Hawari, J.; Perreault, N.N. LC-MS Analysis Data of Intermediate Products Used to Construct Photodegradation Pathways for Bis (1H-Tetrazol-5-Yl) Amine (H2BTA). Data Br. 2020, 28, 104936. [Google Scholar] [CrossRef]
  134. Gahlot, S.; Dappozze, F.; Mishra, S.; Guillard, C. High Surface Area G-C3N4 and g-C3N4-TiO2 Photocatalytic Activity under UV and Visible Light: Impact of Individual Component. J. Environ. Chem. Eng. 2021, 9, 105587. [Google Scholar] [CrossRef]
  135. Demirors, A.F.; van Blaaderen, A.; Imhof, A. Synthesis of Eccentric Titania− Silica Core− Shell and Composite Particles. Chem. Mater. 2009, 21, 979–984. [Google Scholar] [CrossRef]
  136. Kim, D.S.; Han, S.J.; Kwak, S.-Y. Synthesis and Photocatalytic Activity of Mesoporous TiO2 with the Surface Area, Crystallite Size, and Pore Size. J. Colloid Interface Sci. 2007, 316, 85–91. [Google Scholar] [CrossRef]
  137. Kim, E.-Y.; Kim, D.-S.; Ahn, B.-T. Synthesis of Mesoporous TiO2 and Its Application to Photocatalytic Activation of Methylene Blue and E. Coli. Bull. Korean Chem. Soc. 2009, 30, 193–196. [Google Scholar]
  138. Tan, Y.N.; Wong, C.L.; Mohamed, A.R. An Overview on the Photocatalytic Activity of Nano-Doped-TiO2 In the Degradation of Organic Pollutants. Int. Sch. Res. Not. 2011, 2011, 261219. [Google Scholar] [CrossRef] [Green Version]
  139. He, K.; Zhao, C.; Zhao, G.; Han, G. Effects of Pore Size on the Photocatalytic Activity of Mesoporous TiO2 Prepared by a Sol–Gel Process. J. Sol Gel Sci. Technol. 2015, 75, 557–563. [Google Scholar] [CrossRef]
  140. Alagarasi, A.; Rajalakshmi, P.U.; Shanthi, K.; Selvam, P. Solar-Light Driven Photocatalytic Activity of Mesoporous Nanocrystalline TiO2, SnO2, and TiO2-SnO2 Composites. Mater. Today Sustain. 2019, 5, 100016. [Google Scholar] [CrossRef]
  141. Luo, Z.; Poyraz, A.S.; Kuo, C.-H.; Miao, R.; Meng, Y.; Chen, S.-Y.; Jiang, T.; Wenos, C.; Suib, S.L. Crystalline Mixed Phase (Anatase/Rutile) Mesoporous Titanium Dioxides for Visible Light Photocatalytic Activity. Chem. Mater. 2015, 27, 6–17. [Google Scholar] [CrossRef]
  142. Yu, J.C.; Zhang, L.; Zheng, Z.; Zhao, J. Synthesis and Characterization of Phosphated Mesoporous Titanium Dioxide with High Photocatalytic Activity. Chem. Mater. 2003, 15, 2280–2286. [Google Scholar] [CrossRef]
  143. Ferreira, V.R.A.; Azenha, M. Crystallization of Hollow TiO2 into Anatase at Mild Conditions, for Improved Surface Recognition in Selective Photocatalysis. Available online: https://papers.ssrn.com/sol3/papers.cfm?abstract_id=4133307 (accessed on 23 December 2022.).
  144. Chang, H.-S.; Choo, K.-H.; Lee, B.; Choi, S.-J. The Methods of Identification, Analysis, and Removal of Endocrine Disrupting Compounds (EDCs) in Water. J. Hazard. Mater. 2009, 172, 1–12. [Google Scholar] [CrossRef]
  145. Kusuma, M.; Chandrappa, G.T. Effect of Calcination Temperature on Characteristic Properties of CaMoO4 Nanoparticles. J. Sci. Adv. Mater. Devices 2019, 4, 150–157. [Google Scholar] [CrossRef]
  146. Bokov, D.; Turki Jalil, A.; Chupradit, S.; Suksatan, W.; Javed Ansari, M.; Shewael, I.H.; Valiev, G.H.; Kianfar, E. Nanomaterial by Sol-Gel Method: Synthesis and Application. Adv. Mater. Sci. Eng. 2021, 2021, 5102014. [Google Scholar] [CrossRef]
  147. Venkatachalam, N.; Palanichamy, M.; Murugesan, V. Sol–Gel Preparation and Characterization of Alkaline Earth Metal Doped Nano TiO2: Efficient Photocatalytic Degradation of 4-Chlorophenol. J. Mol. Catal. A Chem. 2007, 273, 177–185. [Google Scholar] [CrossRef]
  148. Tajer-Kajinebaf, V.; Sarpoolaky, H.; Mohammadi, T. Sol–Gel Synthesis of Nanostructured Titania–Silica Mesoporous Membranes with Photo-Degradation and Physical Separation Capacities for Water Purification. Ceram. Int. 2014, 40, 1747–1757. [Google Scholar] [CrossRef]
  149. Navas, D.; Fuentes, S.; Castro-Alvarez, A.; Chavez-Angel, E. Review on Sol-Gel Synthesis of Perovskite and Oxide Nanomaterials. Gels 2021, 7, 275. [Google Scholar] [CrossRef] [PubMed]
  150. Danks, A.E.; Hall, S.R.; Schnepp, Z. The Evolution of ‘Sol–Gel’Chemistry as a Technique for Materials Synthesis. Mater. Horizons 2016, 3, 91–112. [Google Scholar] [CrossRef] [Green Version]
  151. Colmenares, J.C.; Luque, R.; Campelo, J.M.; Colmenares, F.; Karpiński, Z.; Romero, A.A. Nanostructured Photocatalysts and Their Applications in the Photocatalytic Transformation of Lignocellulosic Biomass: An Overview. Materials 2009, 2, 2228–2258. [Google Scholar] [CrossRef]
  152. Suhaimi, N.A.A.; Kong, C.P.Y.; Shahri, N.N.M.; Nur, M.; Hobley, J.; Usman, A. Dynamics of Diffusion-and Immobilization-Limited Photocatalytic Degradation of Dyes by Metal Oxide Nanoparticles in Binary or Ternary Solutions. Catalysts 2022, 12, 1254. [Google Scholar] [CrossRef]
  153. Czech, B.; Rubinowska, K. TiO2-Assisted Photocatalytic Degradation of Diclofenac, Metoprolol, Estrone and Chloramphenicol as Endocrine Disruptors in Water. Adsorption 2013, 19, 619–630. [Google Scholar] [CrossRef] [Green Version]
  154. De Ceglie, C.; Pal, S.; Murgolo, S.; Licciulli, A.; Mascolo, G. Investigation of Photocatalysis by Mesoporous Titanium Dioxide Supported on Glass Fibers as an Integrated Technology for Water Remediation. Catalysts 2022, 12, 41. [Google Scholar] [CrossRef]
  155. Ahmed, M.A.; Abou-Gamra, Z.M.; Salem, A.M. Photocatalytic Degradation of Methylene Blue Dye over Novel Spherical Mesoporous Cr2O3/TiO2 Nanoparticles Prepared by Sol-Gel Using Octadecylamine Template. J. Environ. Chem. Eng. 2017, 5, 4251–4261. [Google Scholar] [CrossRef]
  156. Khasawneh, O.F.S.; Palaniandy, P.; Teng, L.P. Large-Scale Study for the Photocatalytic Degradation of Paracetamol Using Fe2O3/TiO2 Nanocomposite Catalyst and CPC Reactor under Natural Sunlight Radiations. MethodsX 2019, 6, 2735–2743. [Google Scholar] [CrossRef]
  157. Ortega-Liébana, M.C.; Sánchez-López, E.; Hidalgo-Carrillo, J.; Marinas, A.; Marinas, J.M.; Urbano, F.J. A Comparative Study of Photocatalytic Degradation of 3-Chloropyridine under UV and Solar Light by Homogeneous (Photo-Fenton) and Heterogeneous (TiO2) Photocatalysis. Appl. Catal. B Environ. 2012, 127, 316–322. [Google Scholar] [CrossRef]
  158. Stojilovic, N.; Scepanovic, M.; Grujic-Brojcin, M.; Golubovic, A.; Armakovic, S.; Abramovic, B.; Sreckovic, T.; Vasilic, R.; Dordevic, S.V.; Popovic, Z.V. Synthesis and Characterization of Titania-Based Nanopowders for Photocatalytic Degradation of Pindolol. Bull. Am. Phys. Soc. 2017, 62, 18. [Google Scholar]
  159. Hazaraimi, M.H.; Goh, P.S.; Lau, W.J.; Ismail, A.F.; Wu, Z.; Subramaniam, M.N.; Lim, J.W.; Kanakaraju, D. The State-of-the-Art Development of Photocatalysts for Degrading of Persistent Herbicides in Aqueous Environment. Sci. Total Environ. 2022, 843, 156975. [Google Scholar] [CrossRef]
  160. Islam, M.R.; Islam, J.B.; Furukawa, M.; Tateishi, I.; Katsumata, H.; Kaneco, S. Photocatalytic Degradation of a Systemic Herbicide: Picloram from Aqueous Solution Using Titanium Oxide (TiO2) under Sunlight. ChemEngineering 2020, 4, 58. [Google Scholar] [CrossRef]
  161. De Chiara, M.L.V.; Pal, S.; Licciulli, A.; Amodio, M.L.; Colelli, G. Photocatalytic Degradation of Ethylene on Mesoporous TiO2/SiO2 Nanocomposites: Effects on the Ripening of Mature Green Tomatoes. Biosyst. Eng. 2015, 132, 61–70. [Google Scholar] [CrossRef]
  162. Paek, S.-M.; Jung, H.; Lee, Y.-J.; Park, M.; Hwang, S.-J.; Choy, J.-H. Exfoliation and Reassembling Route to Mesoporous Titania Nanohybrids. Chem. Mater. 2006, 18, 1134–1140. [Google Scholar] [CrossRef]
  163. Alofi, S.; O’Rourke, C.; Mills, A. Photocatalytic Destruction of Stearic Acid by TiO2 Films: Evidence of Highly Efficient Transport of Photogenerated Electrons and Holes. J. Photochem. Photobiol. A Chem. 2023, 435, 114273. [Google Scholar] [CrossRef]
  164. Alofi, S.; O’Rourke, C.; Mills, A. Kinetics of Stearic Acid Destruction on TiO2 ‘Self-Cleaning’Films Revisited. Photochem. Photobiol. Sci. 2022, 21, 2061–2069. [Google Scholar] [CrossRef]
  165. Smirnova, N.; Fesenko, T.; Zhukovsky, M.; Goworek, J.; Eremenko, A. Photodegradation of Stearic Acid Adsorbed on Superhydrophilic TiO2 Surface: In Situ FT-IR and LDI Study. Nanoscale Res. Lett. 2015, 10, 500. [Google Scholar] [CrossRef] [Green Version]
  166. Mills, A.; Lepre, A.; Elliott, N.; Bhopal, S.; Parkin, I.P.; O’neill, S.A. Characterisation of the Photocatalyst Pilkington ActivTM: A Reference Film Photocatalyst? J. Photochem. Photobiol. A Chem. 2003, 160, 213–224. [Google Scholar] [CrossRef]
  167. Matsunami, D.; Yamanaka, K.; Mizoguchi, T.; Kojima, K. Comparison of Photodegradation of Methylene Blue Using Various TiO2 Films and WO3 Powders under Ultraviolet and Visible-Light Irradiation. J. Photochem. Photobiol. A Chem. 2019, 369, 106–114. [Google Scholar] [CrossRef]
  168. Zuo, R.; Du, G.; Zhang, W.; Liu, L.; Liu, Y.; Mei, L.; Li, Z. Photocatalytic Degradation of Methylene Blue Using TiO2 Impregnated Diatomite. Adv. Mater. Sci. Eng. 2014, 2014, 170148. [Google Scholar] [CrossRef] [Green Version]
  169. Hou, C.; Hu, B.; Zhu, J. Photocatalytic Degradation of Methylene Blue over TiO2 Pretreated with Varying Concentrations of NaOH. Catalysts 2018, 8, 575. [Google Scholar] [CrossRef] [Green Version]
  170. Kim, M.G.; Lee, J.E.; Kim, K.S.; Kang, J.M.; Lee, J.H.; Kim, K.H.; Cho, M.; Lee, S.G. Photocatalytic Degradation of Methylene Blue under UV and Visible Light by Brookite–Rutile Bi-Crystalline Phase of TiO2. New J. Chem. 2021, 45, 3485–3497. [Google Scholar] [CrossRef]
  171. Kader, S.; Al-Mamun, M.R.; Suhan, M.B.K.; Shuchi, S.B.; Islam, M.S. Enhanced Photodegradation of Methyl Orange Dye under UV Irradiation Using MoO3 and Ag Doped TiO2 Photocatalysts. Environ. Technol. Innov. 2022, 27, 102476. [Google Scholar] [CrossRef]
  172. Wang, S.; Zhou, S. Photodegradation of Methyl Orange by Photocatalyst of CNTs/P-TiO2 under UV and Visible-Light Irradiation. J. Hazard. Mater. 2011, 185, 77–85. [Google Scholar] [CrossRef] [PubMed]
  173. Aziztyana, A.P.; Wardhani, S.; Prananto, Y.P.; Purwonugroho, D. Optimisation of Methyl Orange Photodegradation Using TiO2-Zeolite Photocatalyst and H2O2 in Acid Condition. In Proceedings of the IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2019; Volume 546, p. 42047. [Google Scholar]
  174. Dong, H.; Zeng, G.; Tang, L.; Fan, C.; Zhang, C.; He, X.; He, Y. An Overview on Limitations of TiO2-Based Particles for Photocatalytic Degradation of Organic Pollutants and the Corresponding Countermeasures. Water Res. 2015, 79, 128–146. [Google Scholar] [CrossRef]
  175. Jin, C.; Liu, B.; Lei, Z.; Sun, J. Structure and Photoluminescence of the TiO2 Films Grown by Atomic Layer Deposition Using Tetrakis-Dimethylamino Titanium and Ozone. Nanoscale Res. Lett. 2015, 10, 95. [Google Scholar] [CrossRef] [Green Version]
  176. Leskelä, M.; Ritala, M. Atomic Layer Deposition (ALD): From Precursors to Thin Film Structures. Thin Solid Films 2002, 409, 138–146. [Google Scholar] [CrossRef]
  177. Aarik, J.; Aidla, A.; Kiisler, A.-A.; Uustare, T.; Sammelselg, V. Effect of Crystal Structure on Optical Properties of TiO2 Films Grown by Atomic Layer Deposition. Thin Solid Films 1997, 305, 270–273. [Google Scholar] [CrossRef]
  178. Muñoz-Rojas, D.; Maindron, T.; Esteve, A.; Piallat, F.; Kools, J.C.S.; Decams, J.-M. Speeding up the Unique Assets of Atomic Layer Deposition. Mater. Today Chem. 2019, 12, 96–120. [Google Scholar] [CrossRef]
  179. Jin, L.; Dai, B. TiO2 Activation Using Acid-Treated Vermiculite as a Support: Characteristics and Photoreactivity. Appl. Surf. Sci. 2012, 258, 3386–3392. [Google Scholar] [CrossRef]
  180. Sangiorgi, A.; Gonzalez, Z.; Ferrandez-Montero, A.; Yus, J.; Sanchez-Herencia, A.J.; Galassi, C.; Sanson, A.; Ferrari, B. 3D Printing of Photocatalytic Filters Using a Biopolymer to Immobilize TiO2 Nanoparticles. J. Electrochem. Soc. 2019, 166, H3239–H3248. [Google Scholar] [CrossRef]
  181. Luo, Y.; Cao, Y.; Guo, G. Effects of TiO2 Nanoparticles on the Photodegradation of Poly (Lactic Acid). J. Appl. Polym. Sci. 2018, 135, 46509. [Google Scholar] [CrossRef]
  182. Yang, L.; Liu, Z. Study on Light Intensity in the Process of Photocatalytic Degradation of Indoor Gaseous Formaldehyde for Saving Energy. Energy Convers. Manag. 2007, 48, 882–889. [Google Scholar] [CrossRef]
  183. Ajmal, A.; Majeed, I.; Malik, R.N.; Idriss, H.; Nadeem, M.A. Principles and Mechanisms of Photocatalytic Dye Degradation on TiO2 Based Photocatalysts: A Comparative Overview. Rsc Adv. 2014, 4, 37003–37026. [Google Scholar] [CrossRef]
  184. Reza, K.M.; Kurny, A.S.W.; Gulshan, F. Parameters Affecting the Photocatalytic Degradation of Dyes Using TiO2: A Review. Appl. Water Sci. 2017, 7, 1569–1578. [Google Scholar] [CrossRef] [Green Version]
  185. Al-Nuaim, M.A.; Alwasiti, A.A.; Shnain, Z.Y. The Photocatalytic Process in the Treatment of Polluted Water. Chem. Pap. 2022, 1–25. [Google Scholar] [CrossRef]
  186. Pareek, V.; Chong, S.; Tadé, M.; Adesina, A.A. Light Intensity Distribution in Heterogenous Photocatalytic Reactors. Asia Pacific J. Chem. Eng. 2008, 3, 171–201. [Google Scholar] [CrossRef]
  187. Herrmann, J.-M. Heterogeneous Photocatalysis: Fundamentals and Applications to the Removal of Various Types of Aqueous Pollutants. Catal. Today 1999, 53, 115–129. [Google Scholar] [CrossRef]
  188. Kaneco, S.; Rahman, M.A.; Suzuki, T.; Katsumata, H.; Ohta, K. Optimization of Solar Photocatalytic Degradation Conditions of Bisphenol A in Water Using Titanium Dioxide. J. Photochem. Photobiol. A Chem. 2004, 163, 419–424. [Google Scholar] [CrossRef]
  189. Kardar, P.; Ebrahimi, M.; Bastani, S. Influence of Temperature and Light Intensity on the Photocuring Process and Kinetics Parameters of a Pigmented UV Curable System. J. Therm. Anal. Calorim. 2014, 118, 541–549. [Google Scholar] [CrossRef]
  190. Rusu, M.C.; Block, C.; Van Assche, G.; Van Mele, B. Influence of Temperature and UV Intensity on Photo-Polymerization Reaction Studied by Photo-DSC. J. Therm. Anal. Calorim. 2012, 110, 287–294. [Google Scholar] [CrossRef]
  191. He, H.; Li, L.; Lee, L.J. Photopolymerization and Structure Formation of Methacrylic Acid Based Hydrogels: The Effect of Light Intensity. React. Funct. Polym. 2008, 68, 103–113. [Google Scholar] [CrossRef]
  192. Terrones, G.; Pearlstein, A.J. Effects of Kinetics and Optical Attenuation on the Completeness, Uniformity, and Dynamics of Monomer Conversion in Free-Radical Photopolymerizations. Macromolecules 2001, 34, 8894–8906. [Google Scholar] [CrossRef]
  193. Doğruyol, Z.; Arsu, N.; Pekcan, Ö. Critical Exponents of Photoinitiated Gelation at Different Light Intensities. J. Macromol. Sci. Part B 2009, 48, 745–754. [Google Scholar] [CrossRef]
  194. Thiher, N.L.K.; Schissel, S.M.; Jessop, J.L.P. Quantifying UV/EB Dual Cure for Successful Mitigation of Oxygen Inhibition and Light Attenuation. Prog. Org. Coat. 2020, 138, 105378. [Google Scholar] [CrossRef]
  195. Alsheheri, S.Z. Nanocomposites Containing Titanium Dioxide for Environmental Remediation. Des. Monomers Polym. 2021, 24, 22–45. [Google Scholar] [CrossRef]
  196. Zhu, H.; Jiang, R.; Fu, Y.; Guan, Y.; Yao, J.; Xiao, L.; Zeng, G. Effective Photocatalytic Decolorization of Methyl Orange Utilizing TiO2/ZnO/Chitosan Nanocomposite Films under Simulated Solar Irradiation. Desalination 2012, 286, 41–48. [Google Scholar] [CrossRef]
  197. Takahatake, Y.; Shibata, A.; Nomura, K.; Sato, T. Effect of Flowing Water on Sr Sorption Changes of Hydrous Sodium Titanate. Minerals 2017, 7, 247. [Google Scholar] [CrossRef]
  198. Bahrudin, N.N. Evaluation of Degradation Kinetic and Photostability of Immobilized TiO2/Activated Carbon Bilayer Photocatalyst for Phenol Removal. Appl. Surf. Sci. Adv. 2022, 7, 100208. [Google Scholar] [CrossRef]
  199. Hapeshi, E.; Achilleos, A.; Vasquez, M.I.; Michael, C.; Xekoukoulotakis, N.P.; Mantzavinos, D.; Kassinos, D. Drugs Degrading Photocatalytically: Kinetics and Mechanisms of Ofloxacin and Atenolol Removal on Titania Suspensions. Water Res. 2010, 44, 1737–1746. [Google Scholar] [CrossRef]
  200. Maeng, S.K.; Cho, K.; Jeong, B.; Lee, J.; Lee, Y.; Lee, C.; Choi, K.J.; Hong, S.W. Substrate-Immobilized Electrospun TiO2 Nanofibers for Photocatalytic Degradation of Pharmaceuticals: The Effects of PH and Dissolved Organic Matter Characteristics. Water Res. 2015, 86, 25–34. [Google Scholar] [CrossRef]
  201. Fawzy, A.; Mahanna, H.; Mossad, M. Effective Photocatalytic Degradation of Amoxicillin Using MIL-53 (Al)/ZnO Composite. Environ. Sci. Pollut. Res. 2022, 29, 68532–68546. [Google Scholar] [CrossRef]
  202. Alam, U.; Khan, A.; Raza, W.; Khan, A.; Bahnemann, D.; Muneer, M. Highly Efficient Y and V Co-Doped ZnO Photocatalyst with Enhanced Dye Sensitized Visible Light Photocatalytic Activity. Catal. Today 2017, 284, 169–178. [Google Scholar] [CrossRef]
  203. Kalaivani, T.; Anilkumar, P. Role of Temperature on the Phase Modification of TiO2 Nanoparticles Synthesized by the Precipitation Method. Silicon 2018, 10, 1679–1686. [Google Scholar] [CrossRef]
  204. Henderson, M.A. A Surface Science Perspective on TiO2 Photocatalysis. Surf. Sci. Rep. 2011, 66, 185–297. [Google Scholar] [CrossRef]
  205. Jamil, S.; Fasehullah, M. Effect of Temperature on Structure, Morphology, and Optical Properties of TiO2 Nanoparticles. Mater. Innovafions 2021, 1, 22–28. [Google Scholar] [CrossRef]
  206. Panagiotopoulou, P.; Kondarides, D.I. Effect of Morphological Characteristics of TiO2-Supported Noble Metal Catalysts on Their Activity for the Water–Gas Shift Reaction. J. Catal. 2004, 225, 327–336. [Google Scholar] [CrossRef]
  207. Yu, K.; Zhao, J.; Guo, Y.; Ding, X.; Liu, Y.; Wang, Z. Sol–Gel Synthesis and Hydrothermal Processing of Anatase Nanocrystals from Titanium n-Butoxide. Mater. Lett. 2005, 59, 2515–2518. [Google Scholar] [CrossRef]
  208. Elsellami, L.; Dappozze, F.; Fessi, N.; Houas, A.; Guillard, C. Highly Photocatalytic Activity of Nanocrystalline TiO2 (Anatase, Rutile) Powders Prepared from TiCl4 by Sol–Gel Method in Aqueous Solutions. Process Saf. Environ. Prot. 2018, 113, 109–121. [Google Scholar] [CrossRef]
  209. Chen, Y.-F.; Lee, C.-Y.; Yeng, M.-Y.; Chiu, H.-T. The Effect of Calcination Temperature on the Crystallinity of TiO2 Nanopowders. J. Cryst. Growth 2003, 247, 363–370. [Google Scholar] [CrossRef]
  210. Bessekhouad, Y.; Robert, D.; Weber, J. V Preparation of TiO2 Nanoparticles by Sol-Gel Route. Int. J. Photoenergy 2003, 5, 153–158. [Google Scholar] [CrossRef] [Green Version]
  211. Velardi, L.; Scrimieri, L.; Serra, A.; Manno, D.; Calcagnile, L. Effect of Temperature on the Physical, Optical and Photocatalytic Properties of TiO2 Nanoparticles. SN Appl. Sci. 2020, 2, 707. [Google Scholar] [CrossRef] [Green Version]
  212. Nezar, S.; Laoufi, N.A. Electron Acceptors Effect on Photocatalytic Degradation of Metformin under Sunlight Irradiation. Sol. Energy 2018, 164, 267–275. [Google Scholar] [CrossRef]
  213. Mao, S.; Bao, R.; Fang, D.; Yi, J. Influence of Electron Acceptor and Carrier in Amorphous TiO2 upon the Photocatalytic Degradation of Methylene Orange. Mater. Res. Express 2018, 5, 45051. [Google Scholar] [CrossRef]
  214. Ahmed, S.; Rasul, M.G.; Brown, R.; Hashib, M.A. Influence of Parameters on the Heterogeneous Photocatalytic Degradation of Pesticides and Phenolic Contaminants in Wastewater: A Short Review. J. Environ. Manag. 2011, 92, 311–330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Grushevskaya, S.; Belyanskaya, I.; Kozaderov, O. Approaches for Modifying Oxide-Semiconductor Materials to Increase the Efficiency of Photocatalytic Water Splitting. Materials 2022, 15, 4915. [Google Scholar] [CrossRef]
  216. Pesci, F.M.; Wang, G.; Klug, D.R.; Li, Y.; Cowan, A.J. Efficient Suppression of Electron–Hole Recombination in Oxygen-Deficient Hydrogen-Treated TiO2 Nanowires for Photoelectrochemical Water Splitting. J. Phys. Chem. C 2013, 117, 25837–25844. [Google Scholar] [CrossRef]
  217. Ulusoy, T.G. Development of Photoanodes for Performance Enhanced Dye Sensitized Solar Cells. Doctoral Dissertation, Bilkent Universitesi, Çankaya, Turkey, 2015. [Google Scholar]
  218. Kashif, N.; Ouyang, F. Parameters Effect on Heterogeneous Photocatalysed Degradation of Phenol in Aqueous Dispersion of TiO2. J. Environ. Sci. 2009, 21, 527–533. [Google Scholar] [CrossRef]
  219. Pizzino, G.; Irrera, N.; Cucinotta, M.; Pallio, G.; Mannino, F.; Arcoraci, V.; Squadrito, F.; Altavilla, D.; Bitto, A. Oxidative Stress: Harms and Benefits for Human Health. Oxid. Med. Cell. Longev. 2017, 2017, 8416763. [Google Scholar] [CrossRef] [Green Version]
  220. Vučić, M.D.R.; Mitrović, J.Z.; Kostić, M.M.; Velinov, N.D.; Najdanović, S.M.; Bojić, D.V. Heterogeneous Photocatalytic Degradation of Anthraquinone Dye Reactive Blue 19: Optimization, Comparison between Processes and Identification of Intermediate Products. Water Sa 2020, 46, 291–299. [Google Scholar]
  221. Saquib, M.; Muneer, M. Photocatalytic Degradation of Two Selected Textile Dye Derivatives, Eosine Yellowish and p-Rosaniline, in Aqueous Suspensions of Titanium Dioxide. J. Environ. Sci. Heal. Part A 2003, 38, 2581–2598. [Google Scholar] [CrossRef]
  222. Liu, Q.T.; Williams, T.D.; Cumming, R.I.; Holm, G.; Hetheridge, M.J.; Murray-Smith, R. Comparative Aquatic Toxicity of Propranolol and Its Photodegraded Mixtures: Algae and Rotifer Screening. Environ. Toxicol. Chem. 2009, 28, 2622–2631. [Google Scholar] [CrossRef]
  223. Zebardast, M.; Fallah Shojaei, A.; Tabatabaeian, K. Enhanced Removal of Methylene Blue Dye by Bimetallic Nano-Sized MOF-5s. Iran. J. Catal. 2018, 8, 297–309. [Google Scholar]
  224. Sharma, P.; Slater, T.J.A.; Sharma, M.; Bowker, M.; Catlow, C.R.A. Enhanced H2O2 Production via Photocatalytic O2 Reduction over Structurally-Modified Poly (Heptazine Imide). Chem. Mater. 2022, 34, 5511–5521. [Google Scholar] [CrossRef]
  225. Mir, N.A.; Haque, M.M.; Khan, A.; Muneer, M.; Boxall, C. Photoassisted Degradation of a Herbicide Derivative, Dinoseb, in Aqueous Suspension of Titania. Sci. World J. 2012, 2012, 251527. [Google Scholar] [CrossRef] [Green Version]
  226. Mir, N.A.; Khan, A.; Dar, A.A.; Muneer, M. Photocatalytic Study of Two Azo Dye Derivatives, Ponceau Bs and Reactive Blue 160 in Aqueous Suspension of TiO2: Adsorption Isotherm and Decolorization Kinetics. IJIRSET 2014, 3, 933–9348. [Google Scholar]
  227. Yang, Y.; Cheng, B.; Yu, J.; Wang, L.; Ho, W. TiO2/In2S3 S-Scheme Photocatalyst with Enhanced H2O2-Production Activity. Nano Res. 2021, 1–9. [Google Scholar] [CrossRef]
  228. Muruganandham, M.; Swaminathan, M. Photocatalytic Decolourisation and Degradation of Reactive Orange 4 by TiO2-UV Process. Dye. Pigment. 2006, 68, 133–142. [Google Scholar] [CrossRef]
  229. Liu, X.; Zhang, T.; Shao, Y. Aqueous Bromate Reduction by UV Activation of Sulfite. CLEAN Soil Air Water 2014, 42, 1370–1375. [Google Scholar] [CrossRef]
  230. Poulios, I.; Kositzi, M.; Kouras, A. Photocatalytic Decomposition of Triclopyr over Aqueous Semiconductor Suspensions. J. Photochem. Photobiol. A Chem. 1998, 115, 175–183. [Google Scholar] [CrossRef]
  231. Pędziwiatr, P. Decomposition of Hydrogen Peroxide-Kinetics and Review of Chosen Catalysts. Acta Innov. 2018, 26, 45–52. [Google Scholar] [CrossRef]
  232. Zhang, W.; Sun, W.; Zhang, Y.; Yu, D.; Piao, W.; Wei, H.; Liu, X.; Sun, C. Effect of Inorganic Salt on the Removal of Typical Pollutants in Wastewater by RuO2/TiO2 via Catalytic Wet Air Oxidation. Chemosphere 2022, 312, 137194. [Google Scholar] [CrossRef] [PubMed]
  233. Ma, J.; Yang, Y.; Jiang, X.; Xie, Z.; Li, X.; Chen, C.; Chen, H. Impacts of Inorganic Anions and Natural Organic Matter on Thermally Activated Persulfate Oxidation of BTEX in Water. Chemosphere 2018, 190, 296–306. [Google Scholar] [CrossRef]
  234. Sánchez-Polo, M.; Ocampo-Pérez, R.; Rivera-Utrilla, J.; Mota, A.J. Comparative Study of the Photodegradation of Bisphenol A by HO, SO4− and CO3−/HCO3 Radicals in Aqueous Phase. Sci. Total Environ. 2013, 463, 423–431. [Google Scholar] [CrossRef] [PubMed]
  235. Liao, C.-H.; Kang, S.-F.; Wu, F.-A. Hydroxyl Radical Scavenging Role of Chloride and Bicarbonate Ions in the H2O2/UV Process. Chemosphere 2001, 44, 1193–1200. [Google Scholar] [CrossRef] [PubMed]
  236. Lutze, H.V.; Kerlin, N.; Schmidt, T.C. Sulfate Radical-Based Water Treatment in Presence of Chloride: Formation of Chlorate, Inter-Conversion of Sulfate Radicals into Hydroxyl Radicals and Influence of Bicarbonate. Water Res. 2015, 72, 349–360. [Google Scholar] [CrossRef]
  237. Al-Mamary, M.A.; Moussa, Z. Antioxidant Activity: The Presence and Impact of Hydroxyl Groups in Small Molecules of Natural and Synthetic Origin. In Antioxidants—Benefits, Sources, Mechanisms of Action; IntechOpen: London, UK, 2021; pp. 318–377. [Google Scholar]
  238. Lu, C.-S.; Chen, C.-C.; Mai, F.-D.; Li, H.-K. Identification of the Degradation Pathways of Alkanolamines with TiO2 Photocatalysis. J. Hazard. Mater. 2009, 165, 306–316. [Google Scholar] [CrossRef]
  239. Wang, Z.; Yuan, R.; Guo, Y.; Xu, L.; Liu, J. Effects of Chloride Ions on Bleaching of Azo Dyes by Co2+/Oxone Regent: Kinetic Analysis. J. Hazard. Mater. 2011, 190, 1083–1087. [Google Scholar] [CrossRef]
  240. Zhu, H.; Jiang, R.; Xiao, L.; Chang, Y.; Guan, Y.; Li, X.; Zeng, G. Photocatalytic Decolorization and Degradation of Congo Red on Innovative Crosslinked Chitosan/Nano-CdS Composite Catalyst under Visible Light Irradiation. J. Hazard. Mater. 2009, 169, 933–940. [Google Scholar] [CrossRef]
  241. Pandi, K.; Preeyanghaa, M.; Vinesh, V.; Madhavan, J.; Neppolian, B. Complete Photocatalytic Degradation of Tetracycline by Carbon Doped TiO2 Supported with Stable Metal Nitrate Hydroxide. Environ. Res. 2022, 207, 112188. [Google Scholar] [CrossRef]
  242. Mertah, O.; Gómez-Avilés, A.; Kherbeche, A.; Belver, C.; Bedia, J. Peroxymonosulfate Enhanced Photodegradation of Sulfamethoxazole with TiO2@ CuCo2O4 Catalysts under Simulated Solar Light. J. Environ. Chem. Eng. 2022, 10, 108438. [Google Scholar] [CrossRef]
  243. Riga, A.; Soutsas, K.; Ntampegliotis, K.; Karayannis, V.; Papapolymerou, G. Effect of System Parameters and of Inorganic Salts on the Decolorization and Degradation of Procion H-Exl Dyes. Comparison of H2O2/UV, Fenton, UV/Fenton, TiO2/UV and TiO2/UV/H2O2 Processes. Desalination 2007, 211, 72–86. [Google Scholar] [CrossRef]
  244. Mahvi, A.H.; Ghanbarian, M.; Nasseri, S.; Khairi, A. Mineralization and Discoloration of Textile Wastewater by TiO2 Nanoparticles. Desalination 2009, 239, 309–316. [Google Scholar] [CrossRef]
  245. Ganiyu, S.O.; El-Din, M.G. Insight into In-Situ Radical and Non-Radical Oxidative Degradation of Organic Compounds in Complex Real Matrix during Electrooxidation with Boron Doped Diamond Electrode: A Case Study of Oil Sands Process Water Treatment. Appl. Catal. B Environ. 2020, 279, 119366. [Google Scholar] [CrossRef]
  246. Wang, R.; Hashimoto, K.; Fujishima, A.; Chikuni, M.; Kojima, E.; Kitamura, A.; Shimohigoshi, M.; Watanabe, T. Light-Induced Amphiphilic Surfaces. Nature 1997, 388, 431–432. [Google Scholar] [CrossRef]
  247. Zhang, M.; Lei, E.; Zhang, R.; Liu, Z. The Effect of SiO2 on TiO2-SiO2 Composite Film for Self-Cleaning Application. Surf. Interfaces 2019, 16, 194–198. [Google Scholar] [CrossRef]
  248. Kawasaki, S.; Holmström, E.; Takahashi, R.; Spijker, P.; Foster, A.S.; Onishi, H.; Lippmaa, M. Intrinsic Superhydrophilicity of Titania-Terminated Surfaces. J. Phys. Chem. C 2017, 121, 2268–2275. [Google Scholar] [CrossRef] [Green Version]
  249. Sakai, N.; Fujishima, A.; Watanabe, T.; Hashimoto, K. Quantitative Evaluation of the Photoinduced Hydrophilic Conversion Properties of TiO2 Thin Film Surfaces by the Reciprocal of Contact Angle. J. Phys. Chem. B 2003, 107, 1028–1035. [Google Scholar] [CrossRef]
  250. Mao, C.; Wang, X.; Zhang, W.; Hu, B.; Deng, H. Super-Hydrophilic TiO2-Based Coating of Anion Exchange Membranes with Improved Antifouling Performance. Colloids Surf. A Physicochem. Eng. Asp. 2021, 614, 126136. [Google Scholar] [CrossRef]
  251. Siebenhofer, M.; Viernstein, A.; Morgenbesser, M.; Fleig, J.; Kubicek, M. Photoinduced Electronic and Ionic Effects in Strontium Titanate. Mater. Adv. 2021, 2, 7583–7619. [Google Scholar] [CrossRef]
  252. Kim, H.-M.; Kim, D.; Kim, B. Photoinduced Hydrophilicity of TiO2/WO3 Double Layer Films. Surf. Coatings Technol. 2015, 271, 18–21. [Google Scholar] [CrossRef]
  253. Banerjee, S.; Dionysiou, D.D.; Pillai, S.C. Self-Cleaning Applications of TiO2 by Photo-Induced Hydrophilicity and Photocatalysis. Appl. Catal. B Environ. 2015, 176, 396–428. [Google Scholar] [CrossRef] [Green Version]
  254. De Jesus, M.A.M.L.; da Neto, J.T.S.; Timò, G.; Paiva, P.R.P.; Dantas, M.S.S.; de Ferreira, A.M. Superhydrophilic Self-Cleaning Surfaces Based on TiO2 and TiO2/SiO2 Composite Films for Photovoltaic Module Cover Glass. Appl. Adhes. Sci. 2015, 3, 5. [Google Scholar] [CrossRef] [Green Version]
  255. Chen, H.; Li, X.; Li, D. Superhydrophilic–Superhydrophobic Patterned Surfaces: From Simplified Fabrication to Emerging Applications. Nanotechnol. Precis. Eng. 2022, 5, 35002. [Google Scholar] [CrossRef]
  256. Okkay, H.; Satı, S.; Cengiz, U. Mechanically Stable Superhydrophilic Antifog Surface by Microwave Assisted Sol-Gel Method. J. Taiwan Inst. Chem. Eng. 2021, 120, 360–367. [Google Scholar] [CrossRef]
  257. Adachi, T.; Latthe, S.S.; Gosavi, S.W.; Roy, N.; Suzuki, N.; Ikari, H.; Kato, K.; Katsumata, K.; Nakata, K.; Furudate, M. Photocatalytic, Superhydrophilic, Self-Cleaning TiO2 Coating on Cheap, Light-Weight, Flexible Polycarbonate Substrates. Appl. Surf. Sci. 2018, 458, 917–923. [Google Scholar] [CrossRef]
  258. Kameya, Y.; Yabe, H. Optical and Superhydrophilic Characteristics of TiO2 Coating with Subwavelength Surface Structure Consisting of Spherical Nanoparticle Aggregates. Coatings 2019, 9, 547. [Google Scholar] [CrossRef] [Green Version]
  259. Pellegrino, F.; Pellutiè, L.; Sordello, F.; Minero, C.; Ortel, E.; Hodoroaba, V.-D.; Maurino, V. Influence of Agglomeration and Aggregation on the Photocatalytic Activity of TiO2 Nanoparticles. Appl. Catal. B Environ. 2017, 216, 80–87. [Google Scholar] [CrossRef]
  260. Wang, C.; Böttcher, C.; Bahnemann, D.W.; Dohrmann, J.K. A Comparative Study of Nanometer Sized Fe (III)-Doped TiO2 Photocatalysts: Synthesis, Characterization and Activity. J. Mater. Chem. 2003, 13, 2322–2329. [Google Scholar] [CrossRef]
  261. Zhang, H.; Wang, D.; Han, Y.; Tang, Q.; Wu, H.; Mao, N. High Photoactivity Rutile-Type TiO2 Particles Co-Doped with Multiple Elements under Visible Light Irradiation. Mater. Res. Express 2018, 5, 105015. [Google Scholar] [CrossRef] [Green Version]
  262. Tolosana-Moranchel, A.; Pecharromán, C.; Faraldos, M.; Bahamonde, A. Strong Effect of Light Scattering by Distribution of TiO2 Particle Aggregates on Photocatalytic Efficiency in Aqueous Suspensions. Chem. Eng. J. 2021, 403, 126186. [Google Scholar] [CrossRef]
  263. Liu, Y.; Tang, A.; Zhang, Q.; Yin, Y. Seed-Mediated Growth of Anatase TiO2 Nanocrystals with Core–Antenna Structures for Enhanced Photocatalytic Activity. J. Am. Chem. Soc. 2015, 137, 11327–11339. [Google Scholar] [CrossRef]
  264. Yang, Y.; Li, J.; Zhang, H.; Wang, G.; Rao, Y.; Gan, G. TiO2 Tailored Low Loss NiCuZn Ferrite Ceramics Having Equivalent Permeability and Permittivity for Miniaturized Antenna. J. Magn. Magn. Mater. 2019, 487, 165318. [Google Scholar] [CrossRef]
  265. Han, C.; Gómez, D.E.; Xiao, Q.; Xu, J. Near-Field Enhancement by Plasmonic Antennas for Photocatalytic Suzuki-Miyaura Cross-Coupling Reactions. J. Catal. 2021, 397, 205–211. [Google Scholar] [CrossRef]
  266. Gao, S.; Ueno, K.; Misawa, H. Plasmonic Antenna Effects on Photochemical Reactions. Acc. Chem. Res. 2011, 44, 251–260. [Google Scholar] [CrossRef]
  267. Han, P.; Tana, T.; Xiao, Q.; Sarina, S.; Waclawik, E.R.; Gómez, D.E.; Zhu, H. Promoting Ni (II) Catalysis with Plasmonic Antennas. Chem 2019, 5, 2879–2899. [Google Scholar] [CrossRef]
  268. Gellé, A.; Price, G.D.; Voisard, F.; Brodusch, N.; Gauvin, R.; Amara, Z.; Moores, A. Enhancing Singlet Oxygen Photocatalysis with Plasmonic Nanoparticles. ACS Appl. Mater. Interfaces 2021, 13, 35606–35616. [Google Scholar] [CrossRef]
  269. Lei, S.H.I.; Duan, W. Highly Active Mixed-Phase TiO2 Photocatalysts Fabricated at Low Temperatureand the Correlation between Phase Compositionand Photocatalytic Activity. J. Environ. Sci. 2008, 20, 1263–1267. [Google Scholar]
  270. Porcu, S.; Secci, F.; Ricci, P.C. Advances in Hybrid Composites for Photocatalytic Applications: A Review. Molecules 2022, 27, 6828. [Google Scholar] [CrossRef]
  271. Faccani, L.; Ortelli, S.; Blosi, M.; Costa, A.L. Ceramized Fabrics and Their Integration in a Semi-Pilot Plant for the Photodegradation of Water Pollutants. Catalysts 2021, 11, 1418. [Google Scholar] [CrossRef]
  272. Malato, S.; Maldonado, M.I.; Fernandez-Ibanez, P.; Oller, I.; Polo, I.; Sánchez-Moreno, R. Decontamination and Disinfection of Water by Solar Photocatalysis: The Pilot Plants of the Plataforma Solar de Almeria. Mater. Sci. Semicond. Process. 2016, 42, 15–23. [Google Scholar] [CrossRef]
  273. Fendrich, M.A.; Quaranta, A.; Orlandi, M.; Bettonte, M.; Miotello, A. Solar Concentration for Wastewaters Remediation: A Review of Materials and Technologies. Appl. Sci. 2018, 9, 118. [Google Scholar] [CrossRef]
Figure 2. Band structures (left panel) and projected density of states (right panel) for the (a) rutile, (b) brookite, and (c) anatase polymorphs of TiO2, as obtained by DFT + U calculations (PBE functional, GBRV pseudopotentials, U = 7.8   eV , J = 1.0   eV ) in Quantum Espresso program as implemented in Schrödinger Materials Science Suite 2022-4.
Figure 2. Band structures (left panel) and projected density of states (right panel) for the (a) rutile, (b) brookite, and (c) anatase polymorphs of TiO2, as obtained by DFT + U calculations (PBE functional, GBRV pseudopotentials, U = 7.8   eV , J = 1.0   eV ) in Quantum Espresso program as implemented in Schrödinger Materials Science Suite 2022-4.
Catalysts 13 00026 g002
Figure 3. Illustrating the generation of photoinduced e and h+ in (a) pure TiO2, (b) metal-doped TiO2, (c) non-metal-doped TiO2, and (d) coupling TiO2 with the metal oxides.
Figure 3. Illustrating the generation of photoinduced e and h+ in (a) pure TiO2, (b) metal-doped TiO2, (c) non-metal-doped TiO2, and (d) coupling TiO2 with the metal oxides.
Catalysts 13 00026 g003
Figure 4. Photocatalytic degradation process utilizing TiO2. Adapted with permission from ref. [115] Copyright© 2022 American Chemical Society.
Figure 4. Photocatalytic degradation process utilizing TiO2. Adapted with permission from ref. [115] Copyright© 2022 American Chemical Society.
Catalysts 13 00026 g004
Figure 5. Surface reconstruction on TiO2 during the reversible hydrophilic changes: (a) before UV irradiation OH group is bound to oxygen vacancy; (b) at the transition state, where the photogenerated hole is trapped at the lattice oxygen; and (c) after UV irradiation, when new OH groups were formed.
Figure 5. Surface reconstruction on TiO2 during the reversible hydrophilic changes: (a) before UV irradiation OH group is bound to oxygen vacancy; (b) at the transition state, where the photogenerated hole is trapped at the lattice oxygen; and (c) after UV irradiation, when new OH groups were formed.
Catalysts 13 00026 g005
Figure 6. Illustration of the aggregation/agglomeration of TiO2 and increased photocatalytic activity through energy transfer.
Figure 6. Illustration of the aggregation/agglomeration of TiO2 and increased photocatalytic activity through energy transfer.
Catalysts 13 00026 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Armaković, S.J.; Savanović, M.M.; Armaković, S. Titanium Dioxide as the Most Used Photocatalyst for Water Purification: An Overview. Catalysts 2023, 13, 26. https://doi.org/10.3390/catal13010026

AMA Style

Armaković SJ, Savanović MM, Armaković S. Titanium Dioxide as the Most Used Photocatalyst for Water Purification: An Overview. Catalysts. 2023; 13(1):26. https://doi.org/10.3390/catal13010026

Chicago/Turabian Style

Armaković, Sanja J., Maria M. Savanović, and Stevan Armaković. 2023. "Titanium Dioxide as the Most Used Photocatalyst for Water Purification: An Overview" Catalysts 13, no. 1: 26. https://doi.org/10.3390/catal13010026

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop