Next Article in Journal
Photocatalytic Reduction of Hexavalent Chromium Using Cu3.21Bi4.79S9/g-C3N4 Nanocomposite
Next Article in Special Issue
Waste Biomass Selective and Sustainable Photooxidation to High-Added-Value Products: A Review
Previous Article in Journal
Magnetically Recoverable Biomass-Derived Carbon-Aerogel Supported ZnO (ZnO/MNC) Composites for the Photodegradation of Methylene Blue
Previous Article in Special Issue
Recent Advancements in Photocatalysis Coupling by External Physical Fields
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthesis and Application of Innovative and Environmentally Friendly Photocatalysts: A Review

by
Antonietta Mancuso
and
Giuseppina Iervolino
*
Department of Industrial Engineering, University of Salerno, Via Giovanni Paolo II 132, 84084 Fisciano, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1074; https://doi.org/10.3390/catal12101074
Submission received: 15 July 2022 / Revised: 9 September 2022 / Accepted: 9 September 2022 / Published: 20 September 2022
(This article belongs to the Special Issue 10th Anniversary of Catalysts—Feature Papers in Photocatalysis)

Abstract

:
Modern society faces two major challenges: removing pollutants from water and producing energy from renewable sources. To do this, science proposes innovative, low-cost, and environmentally friendly methods. The heterogeneous photocatalysis process fits perfectly in this scenario. In fact, with photocatalysis, it is possible both to mineralize contaminants that are not easily biodegradable and to produce hydrogen from the water splitting reaction or from the conversion of organic substances present in water. However, the main challenge in the field of heterogeneous photocatalysis is to produce low-cost and efficient photocatalysts active under visible light or sunlight. The objective of this review is to compare the new proposals for the synthesis of innovative photocatalysts that reflect the requirements of green chemistry, applied both in the removal of organic contaminants and in hydrogen production. From this comparison, we want to bring out the strengths and weaknesses of the proposals in the literature, but above all, new ideas to improve the efficiency of heterogeneous photocatalysis guaranteeing the principles of environmental and economic sustainability.

1. Introduction

In recent years, the development of industrialization and the constant growth of the population have led to an increase in the need for effective processes for the removal of contaminants, in particular from water, which at the same time, can facilitate the enhancement of waste substances present in wastewater. However, it is becoming increasingly difficult to have clean water because many organic compounds are widespread in industrial effluents, such as dyes, aromatic hydrocarbons, pharmaceutical compounds, pesticides, and heavy metal ions [1]. The latest advances in sensing and monitoring technologies have made it possible to classify some pollutants as emerging contaminants (EC) that can be divided into persistent organic pollutants (POPs), pharmaceuticals and personal care products (PPCPs), and endocrine disrupting chemicals (EDCs), according to their use and origin, in order to facilitate their removal [2,3]. Conventional physical, chemical, and biological methods have been used for wastewater treatment. However, many organic compounds are resistant to such treatments, which are, therefore, limited by low removal efficiency, high operating costs, process complexity, and slow operation [4,5,6]. Nowadays, advanced oxidation processes (AOPs) appear as promising technologies for the degradation of nonbiodegradable pollutants [7]. AOPs are oxidation processes carried out under ambient conditions. AOPs are involved in the generation of highly reactive species, such as hydroxyl radicals (•OH), in order to remove organic or even some inorganic compounds that cannot be oxidized by conventional oxidants (gaseous oxygen, ozone, and chlorine) until the assurance of their complete mineralization, in which CO2 and H2O are obtained [8,9]. Among different AOPs, heterogenous photocatalysis is a low-cost and environmentally benign technology widely used in water and in the air purification field. Heterogenous photocatalysis is based on the use of metal oxide semiconductors (TiO2, ZnO, WO3, Fe2O3, ZnS, CdS, etc.) as catalysts and an irradiation source [10,11,12]. The interest in heterogeneous photocatalysis started many decades ago, when in 1972 Fujishima and Honda developed the photochemical splitting of water into hydrogen and oxygen in the presence of TiO2 [13]. In recent years, research has focused attention on the application of semiconductors as photocatalysts for the removal of organic and inorganic substances from aqueous or gaseous systems in environmental remediation, in the treatment of drinking water, and in plants’ industrial and sanitary applications. The semiconductor metal oxide, especially TiO2, is able to convert light energy into chemical energy in order to decompose pollutants in air and water through oxidation and reduction processes [10].
Over the last few years, research has focused on the development of innovative heterogeneous photocatalysts through the design of sustainable and green synthetic approaches [14,15]. This need arose in order to try to satisfy the principles of green chemistry, including eliminating the use of toxic reagents and solvents dangerous to the environment [16]. These strategies range from the use of plant extracts to the valorization and recycling of metals in industrial waste, or from the use of solventless techniques to the elaboration of mild-reaction condition synthetic tools [17].
Currently, scientific literature is highlighting advances in the development of visible-light-active heterogeneous photocatalysts based on two different approaches: the design of sustainable synthetic methodologies and the use of biomass and waste as sources of chemicals incorporated into the final photoactive materials. The innovative spirit of the scientists guided the research on active photoactive materials with visible light for numerous applications. Particular attention should be paid to photocatalysts developed from environmentally friendly syntheses. This approach, in fact, follows the principles of sustainable development and fully fulfills the idea of green transition: there is no real benefit in the development of a material with high photocatalytic activity, but with a negative impact on the environment due to its unsustainable production. Many research groups have, therefore, developed several ecological and environmentally friendly photocatalysts.
The purpose of this review is to provide information about the latest developments in innovative photocatalysts, focusing on materials active in the presence of visible light, prepared through innovative doping methods (such as codoping) or prepared through the valorization of biomass and waste. In detail, this work describes, in separate sections, innovative photocatalysts (specifically TiO2-based photocatalysts) produced through sustainable processes (i.e., focusing on synthetic steps) used for water depollution, and innovative and noble metal-free photocatalysts, used for hydrogen production. In particular, the first part of this review concerns the application of active photocatalysts in visible light, then doped through different techniques and used for the removal of contaminants from water. In the second part, some alternative synthesis methods are reported, which involve the use of waste materials or avoid the use of harmful solvents used for the production of hydrogen. Finally, the conclusions report some considerations regarding the so-called “green” methods proposed for the synthesis of photocatalysts, which are also reported in this review.

2. Innovative Methods for Doping Photocatalysts: Codoped Photocatalysts

The photocatalytic process, in particular based on TiO2 materials, is applied in different fields, including medical research, drug delivery, antibacterial products, energy storage, pollutant degradation, water splitting, dye-sensitive solar cells, and self-cleaning [18]. TiO2 material is widely used as a photocatalyst for its low cost, good optical activity, high chemical stability, and nontoxic nature. TiO2 is an n-type semiconductor, and it exists in three crystalline phases: anatase, rutile, and brookite. The anatase phase evidences the best photocatalytic activity of TiO2 semiconductors because of its separation capacity of photogenerated holes and electrons between VB and CB under UV light [11]. Several factors affect the TiO2 photocatalytic activity, such as pore volume, specific surface area, particle size, morphology, and bandgap energy.
The main drawbacks of bare TiO2 are due to its large bandgap equal to 3.0 eV for the rutile crystalline phase and 3.2 eV for the anatase phase and the high-rate recombination of photogenerated electron-hole (e/h+) pairs, which limits the photocatalytic efficiency. In order to activate the degradation process of water pollutants, TiO2 is normally excited by UV light at a wavelength <387 nm. However, UV light represents only 5–8% of the solar spectrum at sea level. Therefore, the aim of researchers is to activate the catalysts by natural sunlight and prevent the photogenerated electron-hole pair recombination. A possible strategy to reduce the bandgap by extending the photo response of the catalyst to the visible light region is chemical modification through the doping of the crystal structure of TiO2 with metals or with nonmetals. The addition of the dopant elements can inhibit the recombination of photo-induced holes and electrons by increasing the charge separation in order to achieve complete degradation in a very short time, also in the presence of a small amount of photocatalyst. The TiO2 lattice can be modified by the introduction of transition metals, noble metals, and rare earth (RE) elements. Doping TiO2 with transition metal ions (for example, Fe, Zn, V, Cr, Co, Mn, and Ni) and with Ag, Au, and Ru, generates a red shift in the absorption band of semiconductors from the UV to the visible region, resulting in an improvement in photocatalytic efficiency [19]. The inserted energy levels provide sub-bandgap irradiation from which electrons can be excited by lower energy photons than are required by undoped photocatalysts, and thus transferred from dopant d-band to CB or from VB to dopant d-band. In some cases, the presence of metal dopants enhances charge separation as well as interfacial charge transfer, while in many other cases, the metal dopants actually generate the rapid charge recombination, thereby reducing the electron diffusion length and lifetime [10,20,21]. In the literature, TiO2 doping with Fe ions was widely investigated because the ionic radius of Fe3+ (0.64 Å) is smaller than that of Ti4+ (0.68 Å). Therefore, Fe ions can be easily incorporated into the TiO2 lattice in order to increase charge transfer, and thus cause a reduction in the size of particles. This can lead to a high surface area, which could favor the stronger adsorption of contaminants on the surface photocatalyst [22,23]. Sood et al. described the application of Fe-doped TiO2 photocatalyst in the field of environmental remediation for visible light-driven degradation of highly stable and toxic molecules, such as para-nitrophenol and methylene blue dye [24]. Likewise, the photocatalytic activity of Fe-doped TiO2 and Ni-doped TiO2 was tested for wastewater treatment to remove methylene blue from water under visible light [25,26]. Furthermore, researchers studied visible light-induced degradation of the acid orange 7 azo dye present in textile wastewater, the removal of the 2,4-Dichlorophenoxyacetic toxic herbicide and nitrobenzene from water and soil using Fe-doped TiO2 photocatalysts [27,28], while Bi-doped TiO2 and Ni-doped TiO2 photocatalysts were used for the degradation of ibuprofen anti-inflammatory drugs under solar irradiation [29]. In addition to the application in the field of water and soil decontamination, TiO2 doped with metals, for example, Co, evidenced high ferromagnetic properties, resulting in the adequate use of spintronic devices, such as spin transistors, nonvolatile storage, and spin logic devices [30]. Cu2+ (0.73 Å) is also a good dopant element for decreasing the bandgap due to its similar ionic radius [31] with respect to that of Ti4+ (0.68 Å). Many researchers confirmed the presence of Cu2+ ions in TiO2 matrices in an interstitial way, which acts as electron acceptors or substitutes for one of the Ti4+. Krishnakumar et al. focused their study on photocatalytic mineralization of reactive red 198 using Cu-doped TiO2 powders. The photocatalyst doping with Cu results in the generation of new energy states and the optical bandgap energy decreases, improving the photocatalytic degradation process [32]. The photocatalytic performances of TiO2-based photocatalysts doped with noble metals have been considered in the field of air purification for the removal of toluene, which is one of the volatile organic compounds present in the urban atmosphere polluted by vehicle exhausts and by the use of solvents [33]. On the other hand, various literature studies have demonstrated that the TiO2 doping process with rare-earth (RE) elements prevents the photogenerated electron-hole pair recombination, confirming a better visible light degradation of organic compounds present in water and in air [34,35,36]. The visible light activity of Pr-doped TiO2 powders in the degradation of rhodamine B and crystal violet [37] and also in the mineralization of phenol [38] was investigated. Among different EDCs, bisphenol A was treated in the presence of visible light using a Pr-doped TiO2 photocatalyst [39]. Sm-, Er-, Gd-, and Ce-doped TiO2 photocatalysts reported an interesting visible light-driven activity in the degradation of dyes and water-soluble phenolic compounds [40,41,42]. The photocatalytic activity of nonmetal-doped photocatalysts under visible light irradiation is commonly investigated for the removal of several dyes [43,44]. However, the dyes tend to absorb visible light, and consequently, the photocatalytic reaction could be induced not only by the absorption of visible light from the photocatalyst, but also from the photo absorption of radiation by the dye (i.e., sensitization of the dye) [45,46,47,48,49].
Among all the nonmetal dopants, nitrogen is the most commonly used due to its small ionization energy and its atomic size compared to that of oxygen [45]. The introduction of nitrogen as an impurity in TiO2 is an efficient way of improving the solar light harvesting ability and photocatalytic activity of N-doped TiO2 compared to bare TiO2. N-doped TiO2 photocatalysts can be used for a wide range of applications, such as the degradation of organic pollutants (i.e., dyes and drugs) [50,51] and air purification, sensors, solar energy conversion, and photovoltaics owing to their nontoxicity, ease of synthesis, low cost, and high chemical stability [52,53,54]. In addition to the degradation of industrial textile effluents, N-doped TiO2 is utilized as a self-cleaning material for air purification and teeth whitening [55,56]. The effect of N and its amount in the TiO2 structure on photocatalytic performance is widely described in the literature [57,58]. It is indeed reported that the incorporation of N in the TiO2 structure induces the generation of an N 2p band above the original O 2p valance band, leading to the bandgap reduction with an increase in the photocatalytic activity under visible light [59]. In particular, the oxygen sites of TiO2 can be partially replaced with nitrogen ions with the formation of oxygen vacancies and an increase in Ti amount as the content of N increases. Both oxygen vacancies and Ti species act as hole traps, inhibiting the recombination of electron-hole pairs and, therefore, improving the photocatalytic activity under visible light [60]. However, an optimal amount of dopant elements is required to ensure an improvement in photocatalytic performance because the presence of an excessive number of dopant ions in TiO2 could favor the recombination phenomena of the electron-hole pairs, causing a worsening of the photocatalytic activity [61]. Bakre et al. demonstrated the effectiveness of N-doped TiO2 catalysts in the degradation of cationic dye, methylene blue, and rhodamine B, under sunlight in comparison to commercial TiO2 [62].
Other nonmetals, such as C, S, and P that use less than N, have been considered by several researchers to improve the optical response of TiO2 in the visible light spectrum.
C-doped TiO2 photocatalysts were also employed for the removal of organic compounds, such as dyes, pharmaceuticals, personal care products, and even bacteria from water [63], as well as for hydrogen production [64] and photoreduction of CO2 [65]. Nonetheless, a majority of the literature is still focused on dye discoloration, while degradation and mineralization of other contaminants are less extensively investigated. It was observed that TiO2 doping with C can occur by substitution of O ions with C ions, leading to a replacement of Ti-O by Ti-C bonds, by the replacement of Ti with C due to the rupture of Ti-O and creation of C-O bonds, or by the stabilization of C at the interstitial sites.
The application of S on TiO2-based materials to provide positive effects for the photocatalytic activity is also discussed in the literature [66,67,68,69]. In general, the doping of TiO2 with S element can be via cationic or anionic S routes. It was reported that cationic doping takes place by the substitution of titanium ions with S4+ or S6+, while anionic doping occurs with oxygen ion substitution with S2−.
Additionally, being that the ionic radius of S2− (1.7 Å) is higher than that of O2− (1.22 Å), the incorporation of sulfur by substitution of Ti4+ with S6+ is energetically more favorable than the replacement of O2− with S2− [70]. Both cationic and anionic S-doped photocatalysts revealed enhanced visible light photocatalytic activity. However, S6+ was found to be responsible for reducing the TiO2 crystallite size and the cationic S-doped TiO2; the photoinduced holes and hydroxyls played a major role during photocatalysis. On the other hand, the S2− doping resulted in an increased TiO2 crystallite size, and in anionic S-doped TiO2, the photoinduced holes and electrons played approximately the same role in the photocatalytic activity. In the case of the cationic and anionic S-doped TiO2, the enhancement of the photoactivity was justified by the formation of Ti-O-S and O-Ti-S bands in the crystal lattice under the cationic and anionic S-doping, respectively, which introduced new impurity levels between the VB and CB. The photocatalytic activity of S-doped TiO2 under visible light is usually investigated for the mineralization of different organic pollutants. In fact, Sharotri et al. examined the photocatalytic response of S-doped TiO2 for various model compounds, such as rhodamine 6G (synthetic dye), quinalphos (an organophosphate pesticide), and diclofenac (a pharmaceutical waste), confirming the improved performance of doped nanoparticles with respect to bare TiO2 [71]. Boningari et al. investigated the photodegradation of acetaldehyde with a cationic S-doped TiO2 in the presence of visible light and revealed a higher efficient total organic carbon (TOC) removal of approximately 60% [72]. Bakar et al. examined the decomposition of methylene orange and phenol using both anionic and cationic S-doped TiO2 in the presence of visible light and found that the efficiency of phenol photodegradation was similar to that observed for methylene orange [73]. The photodegradation of atrazine was also investigated. The studies revealed higher effectiveness of atrazine oxidation by the solar/S-doped TiO2 system in comparison with the solar/TiO2 system. The authors also emphasized that lower water pH favored the photodegradation of herbicide, while the presence of humic acids (HA) decreased the decomposition rate due to competition with the atrazine molecules for hydroxyl radicals and partial absorption of sunlight before it reached TiO2 [74]. Lin et al. examined the synthesized S-doped TiO2 in terms of the degradation of gaseous 1,2-dichloroethane [75]. S-doped TiO2 photocatalysts exhibited superior photocatalytic activity under visible light compared to that of pure TiO2. The conversion rate of 1,2-dichloroethane was 55.3 nmoL min−1 g −1, whereas in the case of undoped TiO2 it was 2.16 nmoL /min∙g after 60 min of irradiation. The S-doped TiO2 synthesized by Baeissa et al. was applied for photo-oxidation of cyanide in water, reaching an efficiency of photocatalytic oxidation of cyanide equal to 100% after 30 min of irradiation with a blue fluorescent lamp [76]. Furthermore, the S-doped TiO2 showed potential for application in the antimicrobial field. Dunnill et al. investigated the inhibition of E. coli growth in the presence of the S-doped TiO2 films, recording the complete inhibition efficiency of bacterial growth after 24 h of irradiation. Recently, Zhang et al. studied the activity of P-doped TiO2 photocatalysts for oxidative coupling of benzylamines and photodegradation of phenol under visible light irradiation. The enhanced performance of photocatalyst was attributed to the TiO2 doping with P element, which can exist as a pentavalent oxidation state and can replace Ti4+ ions in the crystal lattice of anatase, by trapping photogenerated electrons in order to greatly reduce the recombination rate of electron/hole pairs [43].
In the last decade, for the treatment of wastewater containing biorecalcitrant organic pollutants, research interest has focused on TiO2 codoping with two different ions, allowing better separation of photogenerated electrons and holes, and thus ensuring an effective enhancement of visible light adsorption together with the increased photocatalytic activity of TiO2 [77].
The codoping process of a TiO2 semiconductor can take place with metal/metal, metal/nonmetal and nonmetal/nonmetal. Thambiliyagodage et al. studied the effect of TiO2 codoping with two transition metals, Fe and Cu, in the visible light-driven degradation of methylene blue. Fe3+ is introduced into the TiO2 lattice via substitution because the cationic radii of Fe3+ and Ti4+ are quite similar. While, as Cu2+ (0.870 Å) ions are much larger than Ti4+ the introduction of Cu2+ into TiO2 ions may occur in interstitial sites of the lattice.
Cu2+ belongs to the same group in the periodic table as noble metals, gold and silver, and therefore, it has similar properties due to its electronic configuration and the face-centered cubic structure. Furthermore, the authors observed that Cu can exist as copper oxides, Cu2O and CuO, generating heterojunctions with TiO2 that lead to an improvement of the charge carrier separation with an increase in their lifetime, and enhancement of the interfacial charge transfer efficiency, resulting in higher activity of the photocatalyst under visible light [78,79,80]. Mancuso et al. investigated the photocatalytic performance of Fe-Pr-codoped TiO2 photocatalysts in terms of color removal and mineralization efficiency for the treatment of aqueous solutions containing acid orange 7 dye under visible light irradiation by noting that the main reactive oxygen species which induce the contaminant photodegradation mechanism are superoxides and positive holes. It revealed the improvement in the photocatalytic activity, which should be related to the simultaneous doping of TiO2 with Pr3+ and Fe3+ ions that produce a narrow bandgap together with an enhanced generation of oxygen vacancies, which can trap electrons, and therefore, reduce their recombination with photogenerated positive holes. The Fe-Pr-codoped TiO2 material was also used for visible light degradation of phenol, confirming the absence of possible sensitization phenomena of the photocatalyst surface by acid orange 7 dye [81]. In order to develop efficient photocatalysts working with sunlight, some researchers tested the incorporation of transition metals Fe and Cr into the TiO2 lattice by the codoping process. The photocatalytic systems exhibited two times higher for visible light decomposition of gaseous isopropyl alcohol (IPA) than single metal-doped TiO2 nanoparticles due to the better optical properties in the range of 400–600 nm [82]. Liu et al. synthesized Zn-Mg codoped TiO2 nanoparticles to evaluate the performance of the dye-sensitized solar cell, resulting in an improved conversion efficiency with respect to the solar cell based on undoped TiO2 nanoparticles due to the faster carrier transport [83]. Additionally, in this case, the codoping of TiO2 with metal and nonmetal substantially enhances the photodegradation activity of organic compounds with respect to undoped TiO2 and single-doped TiO2, thanks to the further improvement of optical absorption towards the visible region. Qin et al. investigated the removal of pyrene, which is a polycyclic aromatic hydrocarbon used as a model pollutant, through a coupled photocatalytic-bacterial degradation process in presence of the Cu-N codoped TiO2, showing a better degradation efficiency under visible irradiation than UV light [84]. Visible light-induced degradation of organophosphate pesticides quinalphos and 2-chlorophenol using Mn-N-codoped TiO2 photocatalyst and color removal of acid orange 7 with Fe-N-codoped TiO2 powders with visible light were evaluated [85]. The use of C-N-codoped TiO2 photocatalysts described allows for the shifting of light absorption from the UV region to the visible light region, assuring the increase in photocatalytic activity in terms of photodegradation of rhodamine B [86]. The potential application of the N-F-codoped TiO2 powders to air purification was also investigated by simulating a commercial LED air cleaner system. The photocatalytic activity of N-F-codoped TiO2 nanoparticles on the decomposition of acetaldehyde was superior to that of commercial TiO2. This outcome is due to the substitution of oxygen atoms both with N and P, which generates an increase in oxygen vacancies in the lattice, acting as a site for the formation of the active radicals involved in the degradation reaction [87]. P–N–TiO2 photocatalysts were prepared for water splitting, which is a solar conversion process of energy into chemical energy (hydrogen and oxygen) for energy saving. In particular, as is largely discussed in the literature, the TiO2 doping with N improves the optical absorption towards the visible region, ensuring a good photocatalytic performance. However, doping with N can induce the lattice long-range distortion in the TiO2 crystal structure due to the substitution of O ions with N ions with the formation of oxygen vacancies, which, if present in excessive quantities, could act as recombination centers for holes and electrons, causing a worsening of the catalytic activity in the water splitting process. On the other hand, the presence of P together with N in the lattice reduces the long-range structural defects induced by N-doped TiO2 and recovers the long-range ordered structure of TiO2 crystals with a significant improvement in the separation of photoexcited electrons and holes by realizing the overall water splitting under visible light irradiation [88]. Below is a summary table with the main doped and codoped photocatalysts mentioned in this paragraph. Below, in the Table 1, a summary of the described photocatalyst doped with different element, with the light source, is reported.

3. Main Synthesis Methods of Doped and Codoped Catalysts

The growing interest in the TiO2-based photocatalysts led to the development of different preparation methods. Precipitation and coprecipitation, hydrothermal synthesis and sol-gel processes for the preparation of nanostructured materials are “wet” chemical methods, as they allow us to obtain solids with a high specific surface area and high porosity. Sol-gel is one of the most used methods for the preparation of visible active photocatalysts [89]. Solid particles homogeneously dispersed in a liquid medium in colloidal form and generate a sol, whereas an organized three-dimensional continuous solid arrangement having submicrometer size pores and containing the liquid phase produces a gel. Hydrolysis and polymerization reactions of different precursors (usually metal alkoxides) produce sol during a sol-gel preparation. The complete polymerization and the loss of solvent can induce the transition from the sol to the gel phase. Therefore, in a general way, the sol-gel method involves four basic steps: hydrolysis, polycondensation, drying, and calcination for precursor decomposition. In the case of UV activated photocatalysts doping, to induce activity by visible light irradiation, a solution containing the precursor salt of the doping element (metal or nonmetal) is inserted in the sol formation step. In fact, in this way, strong covalent bonds between the dopant element and the very reactive monomeric species of the semiconductor precursor (e.g., TiO2 or ZnO) are created. Several photocatalysts active under visible light were prepared by sol-gel process, including TiO2 doped with nitrogen, [90] boron [91], phosphorous [92], iron [22], cobalt [93], magnesium [94], nickel [95], cerium [96], zinc [97], barium [98] molybdenum [99], neodymium [100], and praseodymium [38]. The steps of the sol-gel method can be varied in order to enhance the doping efficiency. For example, visible active N-doped TiO2 has been synthesized through a modified sol-gel process [101], in which the reactants were titanium isopropoxide (as TiO2 precursor) and ammonia aqueous solution (as nitrogen precursor). The N-doped TiO2 catalyst was obtained by a hydrolysis reaction between the two precursors. The preparation was carried out at 0 °C, under stirring, until the formation of a gel. This gel was centrifuged and washed with distilled water and then thermally treated in a furnace at 450 °C for 30 min. The characterization of the N-doped TiO2 catalyst, yellow in color, evidenced the ability to absorb visible light, as indicated by the bandgap value (2.5 eV) [102] Similarly, the synthesis of Cr-doped TiO2 nanoparticles was performed by a modified sol-gel method using a water-alcohol mixture as a solvent. A solution containing ethanol and isopropyl alcohol were mixed slowly with titanium (IV) butoxide under sonication at 40 °C in the ultrasonic bath for 30 min for good dispersion. To this solution, alcoholic gum powder was added drop by drop with constant magnetic stirring. The resultant precursor solution was mixed with distilled water containing chromium nitrate nonahydrate under constant stirring at 80 °C for 4 h. The excess solvent was removed by evaporation with continuous stirring; the resultant precursor was dried at 110 °C for 12 h and then finally calcined at 500 °C for 4 h in a high temperature muffle furnace under static air atmosphere [103]. The sol-gel synthesis was also adopted for the preparation of codoped photocatalysts. For example, Fe-N-codoped TiO2 was prepared using distilled water containing urea and mixed with a solution of titanium tetraisopropoxide and iron(II) acetylacetonate [77]. In the same way, to obtain Fe-Pr-codoped TiO2 distilled water containing praseodymium nitrate was mixed with a solution of titanium tetraisopropoxide and iron acetylacetonate [81]. The systems were maintained at room temperature under continuous stirring for 10 min. Both suspensions were centrifuged for the separation of precipitates, which were washed with distilled water three times and finally placed in a furnace at 450 C for 30 min in static air. The hydrothermal method requires both high temperature and water pressure values as operating conditions. When a solvent, different from water is employed, the method is called “solvothermal” [104]. The preparation of photocatalysts by the hydrothermal method is typically carried out in a steel vessel operating at high pressure (autoclave) under controlled temperatures; the formation of nanocatalyst occurs in the liquid medium.
When the reaction mixture is heated, two zones at different temperatures are formed inside the autoclave. In the higher temperature, upper zone, the reactants of the mixture form a solution, while the saturated solution present in the lower part of the autoclave moves to the upper section of the reactor because of the convective motion. When the solution in the upper part of the autoclave becomes cooler and denser, it moves down. Simultaneously, due to the temperature decrease, the solubility limit is exceeded and consequently, crystallization can start. This technique allows us to directly obtain the catalyst in powder form, while the crystalline degree can be tuned, depending on the operating conditions. In addition, particle size, shape, and chemical composition can be controlled by changing only two parameters, namely, the temperature of the reaction mixture and the solvent used in the synthesis, in order to reach high pressure values and consequently, supersaturation at lower temperatures. A literature study reported a comparison between the sol-gel and the hydrothermal synthesis to prepare TiO2 photocatalyst for propene oxidation [105]. In both cases, a mixture prepared with titanium tetraisopropoxide and ethanol was stirred at room temperature, first magnetically for 1 h and then in ultrasonic bath (30 min). Then, a solution containing HCl and ethanol was added dropwise, and the mixture was left under continuous stirring for 1 h. The synthesis was also performed without HCl, and H2O was added instead. In the case of the sol-gel method, the product of the above-described procedure was dried at 100 °C for 12 h. In the case of the hydrothermal synthesis, the product was transferred to an autoclave and maintained at 180 °C for 12 h and, afterwards, it was dried at 100 °C for 12 h. Finally, both the sol-gel and the hydrothermal samples were heated in air in a muffle at 5 °C/min moved up to 350 °C and kept for 2 h at this temperature. It has been found that TiO2 materials with high crystallinity, between 67% and 81%, and surface area (up to 134 m2/g) were obtained with both sol-gel and hydrothermal methods. In general, the surface and pore volume of TiO2 samples obtained by the hydrothermal method are larger than those of TiO2 by sol-gel. TiO2 catalysts for hydrothermal synthesis are, in general, more active than sol-gel materials or commercial TiO2 (P25) in the photo-oxidation of propene. The effects, due to the presence of HCl during the synthesis of TiO2 and of the post synthesis heat treatment are much more marked in the case of materials prepared via the sol-gel method. In particular, the hydrothermal method leads, in general, to higher crystallinity and allows us to obtain a larger amount of anatase using a low HCl concentration. In the sol-gel method, the presence of acid is necessary to enhance the crystallization and to obtain samples with acceptable surface area. Both sol-gel and hydrothermal methods are appropriate for obtaining photocatalysts with large surface areas (larger than that of P25) and, in general, TiO2 samples obtained by the hydrothermal method have larger surface areas and pore volumes than TiO2 by sol-gel technique. The post synthesis heat treatment of the sol-gel samples is necessary to develop crystallinity, particularly when no HCl is used in the preparation, whereas, the hydrothermal method leads to high crystallinity even without the use of HCl in the synthesis and without the post synthesis heat treatment [105]. Furthermore, the hydrothermal method is a very effective way of introducing dopant elements into the crystalline structure of TiO2 or ZnO in the photocatalysts synthesis [106]. In particular, many papers have been devoted to the controlled synthesis of TiO2 hollow particles, due to their high photocatalytic activity. For example, Zhou et al. prepared flower-like F-doped TiO2 hollow microspheres with the hydrothermal synthesis method, controlling the hydrolysis of TiF4 in a Teflon autoclave at a reaction temperature of 180 °C [107]. A visible active N-doped TiO2 photocatalyst was prepared using triethylamine as a nitrogen source by a low-temperature hydrothermal method without any thermal treatment [108]. Fe-doped TiO2 nanoparticles were directly prepared with TiO2 nanoparticles and iron(III) nitrate ninehydrate was added into a Teflon-lined autoclave to hydrothermal react at 200 °C for 24 h [109]. It was found that the hydrothermal synthesis allowed the incorporation of Fe3+ into the inner matrix of TiO2 and not on its surface in order to obtain the photocatalytic powders active under visible light with an improvement of degradation activity of the pollutant. Fe-doped TiO2 anatase powders produced by a hydrothermal process with titanium (IV) tetra-tert-butoxide and FeCl3 or FeCl2 dissolved n-octanol [110] evidenced small crystal sizes (11.4 nm) and high specific surface (101.4 m2/g) with respect to undoped TiO2 (78 nm and 29 m2/g, respectively), while Fe-doped TiO2 prepared by the sol-gel method using titanium tetraisopropoxide, ferric nitrate, glacial acetic acid, and absolute ethanol have high specific surface areas and exhibited smaller average crystallite sizes (4–8 nm) than undoped TiO2 (11 nm), demonstrating that Fe3+ doping contains the grain growth of the TiO2 photocatalyst. The hydrothermal method was also used to prepare photocatalysts different from TiO2 and ZnO, which formed very high crystallinity structures. For example, Amano et al. [111] obtained a bismuth tungstate (Bi2WO6) catalyst possessing high photocatalytic activity under visible light irradiation. The hydrothermal method was also applied to prepare codoped TiO2 photocatalysts. Ce-N-codoped TiO2 was obtained with two solutions: solution A contained urea used as a nitrogen precursor and cerium nitrate hexahydrate as a cerium precursor with the addition of titanium(IV) butoxide and, solution B, which consisted of nitric acid and titanium(IV) butoxide. Solution B was added slowly into solution A at the rate of 1 mL/min under moderate stirring for approximately 2 h. The mixture was then transferred into a 100 mL of Teflon-lined autoclave, which was kept in an oven at 120 C for 16 h. After the autoclave was cooled to room temperature, the precipitates at the bottom of the autoclave were washed with deionized water several times to remove the impurities. The obtained samples were kept in the oven at 100 °C overnight for drying. In order to have better crystallization and to remove the solvent and other impurities, the samples were calcined at 450 °C for 3 h [112]. It was observed that the codoped samples showed a remarkable increase in the visible light-induced degradation of acid orange 7, which can be associated with the smaller crystal size (8 nm for codoped sample), larger surface area (150 m2/g for codoped sample) than N–TiO2 (16 nm and 67 m2/g), and pure TiO2 (18 nm and 57 m2/g) and also related to the slow recombination rate of photoexcited electrons and holes. The same Ce-NcodopedTiO2 sample was also synthetized by the sol-gel method [113]. Ceric ammonium nitrate and urea were dissolved in an aqueous solution with the addition of cetyltrimethyl ammonium bromide adjusting the acidity (pH 1) using phosphoric acid. Afterward, titanium(IV) butoxide was added dropwise to the acidified solution. When the titanium(IV) butoxide was completely hydrolyzed, the mixture was refluxed at 348 K for 24 h and, subsequently, dried under vacuum at 333 K for 4 h. The photocatalytic activity of the Ce-Ncodoped TiO2 nanoparticles was determined by measuring the removal of methylene blue under solar light irradiation, and its photodegradation efficiency (85%) was higher than the activity of N-doped TiO2 (71%) and Ce-doped TiO2 (58%). All the photocatalysts showed the anatase phase, but the codoped sample reported a smaller crystallite size (9.8 nm) even than the Ce-N-codoped TiO2 obtained by hydrothermal treatment. Moreover, the codoped sample had a structure with wormlike channels and a mesoporous structure. The photocatalysts preparation through the precipitation method [114] involves the chemical transformation of a highly soluble metal precursor into another substance of lower solubility, which precipitates in solution. The conversion to the low solubility compound (and then to the precipitate) is usually obtained by increasing the pH of the solution. In practice, the metal precursor is dissolved in a solvent (generally water). The obtained solution is thoroughly stirred and, after a given time, a precipitating agent is gradually, in a controlled amount, added. The formed solid precipitate is collected, washed, dried, and finally activated through thermal treatment. As declared above, raising the pH is the most common way to perform deposition-precipitation. To avoid a rapid precipitation in a solution that could lead to a remarkable increase in particle size, the mixing and precipitation steps should be carried out separately. At the laboratory scale, this is possible using a base, and this kind of precipitation method has been applied for the preparation of various catalysts. Upon raising the solution’s pH, the precipitation of a hydroxide is expected. Fe-doped TiO2 photocatalyst was synthetized by simple a deposition–precipitation route for the photodegradation of thiacloprid in the presence of H2O2 as an oxidant and UV light irradiation [115]. Fe-doped TiO2 particles showed good sedimentation properties and stability toward photocorrosion, which indicates the possibility of the industrial applications of this photocatalyst for wastewater treatment. TiO2 and Fe-doped TiO2 particles for the photodegradation of gaseous toluene as a typical pollutant in air were prepared with the direct precipitation method with an impinging stream-rotating packed bed (IS-RPB) [116] as a reactor. TiO2 and Fe-doped TiO2 catalysts show narrow pore size distributions and a similar total pore volume (0.45–0.48 cm3/g) and average pore diameter (9.6 nm) with high specific surface area (123.5–148.5 m2/g), in comparison with commercial TiO2 (P25) that presents a wide pore size distributions and a pore diameter of 30.6 nm with specific surface area of approximately 106 m2/g. Similarly, also for the catalysts formulated with the precipitation technique, the beneficial effect of the doping of iron in TiO2 was observed, confirming the improvement of the photocatalytic activity under UV light with degradation efficiency of toluene equal to approximately 96% compared to 55% and 56% values obtained with pure TiO2 and P25, respectively [117]. Moreover, the authors reported the preparation of Fe-Co codoped TiO2 photocatalyst using the impinging stream-rotating packed bed and precipitation method mentioned above, in order to evaluate the decomposition of phenol both under UV and visible light [118]. The codoping of TiO2 with Fe and Co ions increased the specific surface area (194 m2/g). The doping of TiO2 with Fe ions can inhibit crystal growth, and thus increase the specific surface area. However, the specific surface area of the catalyst first increases (194 m2/g) and then decreases (88 m2/g); the total pore volume rises from 0.56 to 0.77 cm3 g−1 as the Co2+ content increases in comparison with pure TiO2 that exhibits a specific surface area of 148.5 m2/g and total pore volume of 0.45 cm3 g−1. This result can be related to the introduction of an appropriate amount of Co2+ that can cause defects on the crystal surface, and Co2+ and Fe3+ act synergistically to increase the specific surface area of the catalyst. However, excess Co2+ can form new phases that may be deposited on the TiO2 surface and block the pores, resulting in a decrease in specific surface area. The enhanced crystalline and optical properties of codoped TiO2 demonstrated that the presence of Fe and Co ions into crystal lattice inhibits the recombination of photogenerated hole-electron pairs resulting in a better degradation activity. In fact, the decomposition efficiency of phenol after 2.5 h of UV and visible light irradiation was 88.71% and 24.58% for the Fe-Co codopedTiO2 catalyst, which represents 1.59 and 2.07 times that of pure TiO2 [118].

4. Recent Application of Innovative Photocatalysts

4.1. Photocatalytic Removal of Orgainic Dyes from Aqueous Solution

The dyes degradation cannot occur through biological techniques because the complex structure of dye molecules is resistant to microbial decomposition. Industrial activities involve a large consumption of dyes and their discharge into water bodies causes serious damage to the environment. A dye is a colored organic substance that absorbs visible light and can attach strongly to a fiber by means of chemical or physical bonding between the groups of fiber and dyes. Generally, dyes can be classified on the basis of their chemical structure as well as the existence of specific chromophores [119]. Dye is a highly toxic organic substance, such as benzidine and aromatic amine, and many of them are very difficult to treat [120]. It causes eye and skin irritation, asthma, allergic contact dermatitis, and respiratory tract infections [121]. Azo dyes are readily converted into aromatic amines under potentially dangerous and environmentally toxic anaerobic treatment conditions. Dye hinders the photosynthetic activity of aquatic plants by delaying the penetration of light into the water, which makes oxygen unavailable for the biodegradation of micro-organisms in the water [122].
On the contrary, the photocatalytic degradation process of dye is based on in situ generations of highly reactive hydroxyl radicals (OH•) that can rapidly and nonselectively oxidize any dye contaminant present in the wastewater.
Table 2 shows the visible light-driven performance of doped and codoped TiO2 powders prepared by the above-mentioned synthesis techniques on the degradation of dyes (methylene blue, rhodamine B, and acid orange 7) used as model contaminants reported in the literature.

4.2. Photocatalytic Removal of Emerging Contaminants from Aqueous Solution

Emerging contaminants (ECs) are unregulated pollutants, for which legislative measures are required. They are not new chemicals, but they have persisted in the environment for some time [138]. They consist of the main components of different products used in daily life.
In particular, ECs include endocrine-disrupting chemicals (EDCs), pharmaceuticals and personal care products (PPCPs), persistent organic pollutants (POPs), and so on [139]. The development of techniques and methods of environmental analysis allowed the detection of these harmful substances in municipal wastewater [140], surface water [141], and drinking water [142].
Although the detected concentrations of these ECs in municipal waste, surface, and drinking water are very low, their chemical properties are relatively stable and susceptible to bioaccumulation. Therefore, they can cause severe damage to the ecological environment and human health [143,144]. For example, EDCs are highly dangerous for humans as they attack human organs, and some psychoactive substances could damage brain tissue in humans [145].
ECs are usually noxious residues caused by the frequent and improper consumption of products during human production and living activities [146]. These toxic residues are excreted or discarded by humans and eventually enter wastewater treatment systems [147]. There are hardly any biodegradable drug residues discharged into surface waters together with wastewater from purification plants. Furthermore, these drug residues enter groundwater by runoff, diffusion, and infiltration of surface waters and, thus, can even pollute drinking water. There are also many ECs in the sludge from wastewater treatment systems. These sewage sludges are used for agriculture and cultural fertilization, and therefore, also surface waters are polluted.
In recent years, many studies have focused on photocatalytic technologies for EC removal [148].
For instance, Prieto-Rodriguez et al. reported that a low concentration of TiO2 is suitable for the removal of EC in the effluents of wastewater treatment plants. Visible light-induced degradation of ECs using TiO2-based photocatalysts doped and codoped with metals and/or nonmetals is largely described in the literature. Table 3 displays the activity of doped and codoped TiO2 powders prepared by the above-mentioned synthesis techniques on the degradation of ECs under visible light [149].

4.3. Photocatalytic Degradation of Phenolic Compounds

Phenolic compounds consist of aromatic compounds whose hydrogen is substituted with hydroxyl groups (−OH) on aromatic rings. They can be substituted with chlorine atoms or nitrated or alkylated to produce a series of corresponding derivative compounds. The hydroxyl groups in phenolic compounds can form hydrogen bonds with water molecules. Phenolic compounds can be divided into endogenous and exogenous phenolic compounds [160]. Endogenous phenols exist in nature and are generally derived from roots, stems, leaves, fruits, and other parts of plants or crops. While exogenous phenols are usually produced by smoke from biomass combustion and vehicle exhaust. In the atmosphere, phenolic compounds with substituent groups, such as methoxy (−OCH3), methyl (−CH3), carbonyl (C=O), and carboxylic acid (−COOH) are ubiquitous aromatic compounds. In fact, the important emissions from biomass burning of catechol and guaiacol were observed [161].
Moreover, some phenolic compounds are also produced during the manufacturing and industrial processing of products. For example, they are obtained in the coking and refining process as main products extracted from oil and coal mines [162]. Another important source of phenolic compounds is organic chemical synthesis. Domestic and industrial activities, such as phenolic resin manufacturing, oil industry, pesticide synthesis, plastics, coking operations, wood products, pulp, textiles, detergents, and pharmaceuticals, have discharged harmful and toxic phenolic compounds to water bodies, including rivers, lakes, and seas [163]. In particular, chlorophenols can be produced by the paper industry [164]. Methylphenols, chlorophenols and 2,4-dinitrophenol are widely used as synthetic intermediates for pesticides, preservatives, herbicides, and dyes [165]. Cyanophenols can be used as a precursor to synthetic drugs and antifungal agents [166]. Phenols are a class of organic weak acidic compounds with higher solubility in water, which are hard to degrade since they require a longer time degrade and tend to accumulate in the environment. Therefore, the excessive use of phenols and their derivatives further increases the risk of environmental pollution. Some phenolic compounds, such as nitrophenols can influence the photochemistry of the atmosphere, the quality of the air and the earth’s climate due to their strong absorption of UV light and visible light. In soil or water, phenolic pollutants can change their and pH, having negative effects on animals, plants, and micro-organisms. Phenolic compounds are highly toxic organic pollutants that invade the human body through the mucosa, digestive tract, respiratory tract, and skin, chemically reacting with proteins the body. In this way, the protein is denatured and coagulated, resulting in cellular inactivation. Chlorophenols alkylphenols are typical environmental endocrine disruptors (EDCs), present in water and soil which can damage human organs due to their embryotoxic and teratogenic nature [167]. For this reason, to avoid threatening the water environment and air quality, phenolic pollutants need to be treated effectively. AOPs have been successfully applied to industrial wastewater treatment, leading to the generation of powerful reactive species, principally the hydroxyl radicals and sulfate radicals.
AOPs can destroy the structures of aromatic rings, reduce biological toxicity, and break long-chain molecules that result in high-efficiency treatment of organic pollutants. For example, H2O2/UV-C and photo Fenton treatment can completely degrade 2,4-dichlorophenol with high TOC removal [168]. In the field of heterogeneous photocatalysis, high activity in phenol degradation under visible light irradiation was observed using TiO2 materials made active by codoping with metals and nonmetals [169,170]. In this regard, many studies in the literature discuss the visible light driven-degradation of different phenolic compounds in aqueous solutions using doped and codoped TiO2 materials. Table 4 shows the activity of doped and codoped TiO2 powders prepared with the above-mentioned synthesis techniques on degradation of phenolic compounds under visible light.
At this point, considering the results summarized in the previous tables, it is possible to make some reflections regarding the pros and cons of the synthesis methods typically proposed in the literature, and some consideration about the role of doping and codoping elements and their effect on the photocatalytic performance.
Regarding the synthesis methods, it is clear that it is mainly the “sol gel” method or the “hydrothermal method”. In particular, the recurrent choice of the sol-gel synthesis method is certainly explained by its intrinsic ease of execution. In fact, it is a method characterized by easily manageable operating conditions (P and T) [179]. Therefore, it is possible to define “sol gel” an ecological and simple method as it does not require complicated synthesis conditions or tools and it is known to be an inexpensive technique. Furthermore, it allows us to control the particle size and is quite flexible with the addition of other elements, which act as dopants for the semiconductor [180,181]. In fact the sol phase increases the covalent bond between the dopant species and the monomer species of the semiconductor [179] On the other hand, there is the risk of the phenomenon of “shrinkage” and aging of the gel, which involves a failure of the synthesis process. In addition, long processing times may be required and sometimes, the presence of impurities may be found. Considering the hydrothermal synthesis, this method also presents pros and cons in the preparation of heterogeneous catalysts. In particular, in regard to the advantages, it can be highlighted that with this method most of the precursor material can be made soluble in a suitable solvent by heating and pressing the system near its critical point. Moreover, easy and precise control of the size, shape distribution, and crystallinity of the final product by adjusting parameters, such as reaction temperature, reaction time, type of solvent, type of surfactant, and type of precursor, can be obtained. Specifically, this method is effective in the realization of high quality 1D nanostructures, especially nanorods. Moreover, easy addition of additives for doping is possible with hydrothermal synthesis [182]. On the other hand, this method is configured as a “black box”. In fact, using the autoclave, it is not possible to observe the reaction and correct the operating parameters in real time. Furthermore, the required temperature and pressure values are typically high, and therefore, require suitable equipment (the autoclave).
As for the considerations of the elements for doping and codoping, it can be said that the modification of the TiO2 semiconductor is essential to improve its performance in several respects. Surely an important fact concerns the possibility of using light sources other than UV (both for economic, ecological and safety reasons), therefore, for example, visible or solar light. This can only be possible through the doping of TiO2. The literature proposes various solutions to achieve this goal. In fact, it is possible to exploit the potential of noble metals (such as Au, Pt, Ag, and Pd), which greatly improve the photocatalytic performance of TiO2 both under UV irradiation and visible light. It should be emphasized that the use of noble metals is proposed above all to induce a modification of the surface of the semiconductor rather than its crystal lattice [183]. However, this solution does not produce a modification of the bandgap, therefore, of the absorption capacity of light sources other than UV, as it avoids the recombination of the electron-hole pair that typically affects semiconductors such as TiO2, and which must be kept stable to ensure the performance of the photocatalyst [184]. When nonmetallic elements are used, they typically enter the crystal lattice of the semiconductor, replacing an atom whose atomic radius is more similar in size, and acting with the oxygen atoms, and thus causing a change in the electron densities of the oxygen [185,186,187]. This causes a change in the atomic energy level of the oxygen atom with a consequent modification of the catalyst bandgap [188]. This allows the semiconductor (TiO2) to absorb light at a different wavelength.
The codoping approach allows for further improving the properties of the final material. In fact, with a codopage it is possible to act on the electronic, but also the chemical and magnetic properties of the starting semiconductor [189]. Literature studies show that titania codoping improved the ability of absorbing visible light compared to single element doped titania [181,190]. The role of codopants is not only in narrowing the bandgap, but also contributes to the formation of different heterostructures, such as the formation of different electronic structures with TiO2 [191]. Studies have shown that codoping of TiO2 produces synergistic effects that can further enhance the photocatalytic properties of TiO2 [188,192]. Furthermore, the choice of compatible codopants and the most suitable synthesis methods for the introduction of the dopant are crucial as they affect doping levels [181,191].

5. Innovative Photocatalysts for Hydrogen Production

Hydrogen is an attractive carrier of green energy due to its energy content and the lack of greenhouse gas emissions after its combustion. Currently, the main hydrogen production technologies involve fossil fuels: hydrogen is produced mainly through steam reforming and water gas shift processes [193]. Additional collateral technologies for its production include chemical, electrochemical, biological, and thermal processes. The possibility of producing energy using solar energy has attracted increasing interest in recent years. In particular, the thermochemical [194], photoelectrochemical [195] and photochemical [196] processes can be used for the generation of solar hydrogen. The photocatalytic generation of hydrogen can be obtained through the photoscission of water or the photoreformation of organic species [197]. The photocatalytic splitting of water consists of the decomposition of water into hydrogen and oxygen by reaction with photogenerated charge carriers. Photocatalytic reforming is based on the ability of organic species (i.e., sacrificial agents) to be oxidized by photogenerated positive holes and to release protons, which in turn generate hydrogen by reacting with photoelectrons. The scientific literature proposes several studies on this topic, and several catalytic formulations have been studied. At this point, when can we define a photocatalyst as “innovative?”
We certainly know when a catalyst can be considered “efficient.” As also reported by Clarizia et al., an efficient photocatalyst should meet several requirements. An efficient catalyst should be (i) photohydrostable in water, therefore, it must not undergo photocorrosion [19] and (ii) be easily separable from the mixture at the end of the treatment and, therefore, reusable [198]. Moreover, an efficient photocatalyst should be a semiconductor in which the probability of photogenerated electron/hole recombination is very limited. To achieve these objectives it is possible to act as follows: adjust the size and shape of the semiconductor particles; study the positions of the electronic bands and think about the possibility of introducing cocatalysts to limit the recombination of the electron-hole pair; and use sacrificial substances as electron donors [197]. However, when it comes to innovation, it is necessary to refer not only to the performance of the catalyst, and therefore, its efficiency, but above all to the method of synthesis and recyclability [199]. To date, among various semiconductor photocatalysts that can produce hydrogen via the photocatalytic splitting of water, one of the most used semiconductors is TiO2 as it has the features of superior photocatalytic activity, chemical stability, low-cost and nontoxicity, so it is regarded as a suitable material for use in photocatalytic water splitting for hydrogen production [199].
However, the photocatalytic performance of TiO2 is influenced by many factors, such as the bandgap, the active facet, the degree of crystallinity, the size, and the morphology. In particular, for anatase TiO2, the high activity of the (001) facet has been verified to influence the photocatalytic activity, through both theory and experiment [199,200].
If we consider only the TiO2, the preparation of the anatase TiO2 materials characterized by an exposed (001) facet is complicated and elaborate. In the literature there is confirmation of this limitation. For example, it is reported that HF in a hydrothermal synthetic process is used to restrain the growth of the (001) facet, and obtain TiO2 with an exposed (001) facet [200], or hydrothermal or hydrothermal synthesis is envisaged at high pressure. In particular, the use of extremely corrosive and toxic HF or NH4F is provided as supporting materials for the growing of crystal facets. These are just some examples concerning the synthesis of the semiconductor most used (TiO2) in the photocatalytic production of hydrogen. However, other obstacles are also faced in the synthesis of other types of photocatalysts. This is the case of MoS2 [201]. MoS2 is considered a very good cocatalyst, and it has been experimentally confirmed that hydrogen production on CdS with MoS2 loading is even more efficient than that of CdS with noble metal loading. However, for the production of the MoS2/CdS photocatalyst, the poisonous gas H2S must be used as a source of sulfur and the calcination temperature is relatively high (above 400 °C [202]).
Another typical difficulty encounter with the heterogeneous powder photocatalyst is the recovery and reuse of the catalyst. The photocatalyst in powder dispersed in the solution. It certainly has interesting advantages: a large specific surface available in contact with the reagents and excellent exposure to light radiation.
Unfortunately, the recovery and reuse of the powder photocatalyst requires an effective separation step which can be quite expensive (e.g., membrane or coagulation/filtration). To overcome this drawback, the solution immobilizes the photocatalysts on macroscopic supports, such as glass beads [203], fiberglass mesh [204], silica [205], alumina, and zeolite [206,207].
Although immobilization makes the photocatalyst recoverable, it reduces the photocatalytic performance due to mass transfer limiting phenomena [208,209]. Alternatively, in order to preserve the advantages of suspended photocatalysts, the ideal would be to synthesize a photocatalyst that is easy to remove after treatment, but which at the same time can work in suspension in aqueous solutions.
The exposed disadvantages limit the development of promising, low-cost and ecofriendly photocatalyst [201]. As for the synthesis method, there are several, but the innovation lies in proposing a short method, with few steps, which does not require the use of harmful and expensive reagents, but which at the same time leads to a photocatalyst with the required characteristics. Consequently, an easy, green, and efficient method is needed for the preparation photocatalyst for hydrogen production. Some of the most significant and (in our opinion) most innovative and environmentally friendly synthesis methods are reported below.

5.1. Noble Metal Free Photocatalyst for H2 Production

As previously reported, the photocatalytic production of hydrogen is left to all the uses of photocatalytics (based on TiO2) decorated with noble metals. The presence of noble metal dispersed on the surface of the photocatalyst is important to avoid recombination of the copy with the electron. The most effective catalyst for the reduction in the proton used in the photocatalytic production of hydrogen is Pt, which however, leads to an increase in the cost of hydrogen production and limits its practical application. Therefore, numerous efforts have been made to replace Pt with low-cost alternatives, in which transition metal complexes have also been investigated. Recent studies have shown that interesting results in terms of hydrogen production (under UV or visible light) are also possible for the use of noble metals. The literature proposes the application of metals, such as Ru, Ni, Cu, and Co that present similar efficiencies to that of noble metals, but with a very low cost [210]. Ni is cheaper than noble metals and could represent an alternative giving the interesting characteristics that have led to its use as a cocatalyst combined with semiconductors in photocatalytic processes [184]. Ni at optimal loading of TiO2 has demonstrated to perform better than Au in photocatalytic hydrogen production [211]. Moreover, relevant amount of molecular hydrogen was found on a Ni-modified perovskite (Ni/La-NaTaO3), at the optimal nickel load on the surface of the photocatalyst. Indeed the performances of the photocatalyst are strongly influenced by the Ni-load tuning [184]. In the literature low-cost Ni-modified CdS prepared via a hydrothermal reduction method has been proposed for the photocatalytic production of hydrogen starting from lactic acid solution. NiCl2·6H2O was used as a precursor, and it was dissolved in 60 mL ethylene glycol to form a light green solution. After that, 200 mg CdS was added. The solution was transferred in autoclave and heated at 200 ° C for 10 h. If we consider this method of synthesis, as for all methods that involve “hydrothermal synthesis”, the observation we can make is that, although low-cost materials can be used, the times required for this process are always very long. However, hydrothermal synthesis is a method that allows us to obtain catalytic particles of nanometric dimensions, with very uniform distribution, ideal for heterogeneous photocatalysis applications.
Another interesting proposal reported in the literature is the deposition of Ni on the surface of LaFeO3, a catalyst confirmed in the photocatalytic production of hydrogen starting from aqueous matrices containing glucose [184]. In this case the LaFeO3 was prepared with the solution combustion synthesis, a simple and fast method (that requires short times). Ni was photo-deposited in a reducing environment starting from NiCl2·6H2O as precursor. The results of the photocatalytic tests showed that Ni improves the photocatalytic hydrogen evolution activity of LaFeO3 under UV light because it improves the charge separation and transfer. In particular, the H2 production was approximately 2700 μmol/L after 4 h of irradiation time. This result was also confirmed by the reciclability tests which demonstrated the stability of this catalyst even after 5 cycles. Another recent study on the application of noble metal free photocatalyst for H2 production, proposes the copper phthalocyanine (CuPc) supported TiO2 photocatalyst [210]. This photocatalyst is active in the presence of visible light and it has been observed that the amorphous layer of CuPc has the fundamental role of efficiently separating the photogenerated charges and promoting the absorption of light in the visible region. This photocatalyst was prepared with a precipitation method starting directly from CuPc and TiO2 (P25) in suitable quantities.
In practice, the amorphous CuPc coating layer created on the TiO2 surface acted as a sensitizer to visible light and also an efficient charge transfer mediator. The idea of using TiO2 sensitized with dyes for the production of photocatalytic H2 has been widely reported in the literature. However, in order to free itself from the use of noble metals, research on the use of low-cost transition metal complexes as catalysts for the reduction in protons on TiO2 has yielded interesting results. This is the case with CoL2/TiO2 prepared using cobalt complexes [212]. Co(phen)2Cl2]Cl and degussa P25 (ass TiO2) were used for the synthesis. In particular the procedure was a precipitation that provide a step for the filtration and dry in vacuo for 24 h [212].

5.2. Green Synthesis of Photocatalysts for Hydrogen Production

The fundamental point in the production of photocatalytic hydrogen is the synthesis of photocatalysts which have high activities. The catalysts most studied for the production of hydrogen are TiO2 and all the photocatalysts based on TiO2, due to their photochemical stability and low cost [213]. Very often, to promote the photocatalytic hydrogen production and to improve the photocatalytic performance of TiO2, noble metal nanoparticles (NPs), especially Pt NPs, are used as cocatalysts [214]. Indeed, the Schottky barrier formed between Pt NPs and TiO2 can act as an efficient electron trap to prevent photogenerated electron/hole recombination, and thus promote the photocatalytic reaction [215]. The main need is to produce Pt nanoparticles (or other noble metals) that are uniformly dispersed since the production of hydrogen occurs precisely on the Pt particles. Therefore, it is essential to control the size of the noble metal particles and their distribution. The deposition method for the platinum nanoparticles, and the regulation of their distribution on the TiO2 surface is the focus of photocatalytic and photoelectrochemical hydrogen productio. For most TiO2/PtNP photocatalytic systems, photodeposition is an efficient and commonly used method [213].
However, among the traditional strategies, most of the reducing and stabilizing reagents used for the synthesis of Pt NP are not compatible with the environment [216,217]. The main drawbacks of Pt NP photodeposition are based on the poor dispersion and morphology of the formed Pt NPs, which is difficult to control in the photodeposition process. Pt growth sites are always randomly distributed, which would lead to nonuniform Pt dispersion and morphology. It is possible to improve this morphological aspect. However, it is necessary to do it with green methods. In this case, green methods for the deposition of metal nanoparticles involve the green synthesis of Pt NP through the reduction in sugars and other biomasses. In this case too, the challenge remains to obtain uniformly dispersed metal nanoparticles. The direct use of biomass to design TiO2/Pt NP catalysts with precise dispersion and morphology remains a challenge, as well as an opportunity, to open a promising and “green” path for solar to fuel conversion [213]. For example cubic Pt NPs and spherical Pt nanoclusters were synthesized using wood nanomaterials [217]. In fact natural wood is the most abundant renewable biomass in nature which has attracted growing attention for the preparation of functional materials. The efficient use of this natural resource would be an ideal path for a sustainable approach for future technologies. However, the extraction of cellulose from wood (18) and chemical modifications of cellulose (19) still pose obstacles to the use of wood. It has been shown that nanomaterials derived from natural wood can be used to synthesize Pt NP in aqueous systems with the ability to control particle shapes and sizes with reaction conditions. In this way it is possible to easily prepare spherical and cubic NP Pt supported by wood nanomaterials and spherical Pt nanoclusters [217]. These nanoparticles have been used in the catalytic reactions for the reduction in p-nitrophenol as a model reaction. However, it is possible to transfer the idea of the “green synthesis” of platinum nanoparticles to the TiO2 used in the photocatalytic reactions of hydrogen production. Therefore, hydrolysis of soluble hemicelluloses can influence the morphology of Pt NPs. Among the most abundant soluble biomasses there is sucrose. In the literature, the use of this biomass is proposed precisely for the chemical deposition of Pt NP on TiO2. Sucrose was found to be very efficient for forming well dispersed spherical Pt NP cocatalysts. The photocatalysts prepared in this way show higher photocatalytic efficiency than TiO2/PtNP prepared through the conventional photodeposition process. Furthermore, sucrose could act as an electron donor to generate hydrogen [213]. The synthesis procedure of these spherical Pt NP nanoparticles involved a chemical deposition with the use of chloroplatinic acid (H2PtCl6) added to 60 mL of sucrose solution (at different sucrose concentrations) This mixture was then subjected to stirring and heating (100 degrees for 20 min). After heating, the authors observed that the solution had turned brown, which demonstrated the formation of Pt NP. Furthermore, the study carried out showed that the morphology of the Pt NPs formed was mainly related to the sucrose concentrations [213]. Another interesting data reported by the same authors concerned the anchoring contact between TiO2 and Pt NPs. It plays a key role in the photocatalytic performance of TiO2 with Pt. In order to obtain TiO2 catalysts with uniform Pt NP loading, it has been found that the addition sequence of the precursors is also important. Both the TiO2 suspension and the H2PtCl6 were added to the sucrose solutions and heated at the same time. The prepared TiO2/Pt NP catalysts showed that the Pt NP cocatalysts are well dispersed. Moreover, the surface area of the Pt cocatalysts is a main parameter for explaining the photoactivity behavior of Pt NPs because smaller particles with better dispersity imply higher number of surface-active centers for the photoreactions. In terms of hydrogen production, after 4 h of simulated sunlight, approximately 40 μmol of H2 has been produced [213].
Another simple, efficient and environmentally friendly way to prepare TiO2 nanoplates and nanoparticle, which can be applied to H2-production via photocatalytic water splitting, is the low-heating solid-state chemical method (LSCM) [199]. This method was used to prepare uniform NH4TiOF3 nanoplates with a thickness of approximately 100 nm in the absence of surfactants. Solid-solid transformation was used to convert NH4TiOF3 nanoplates to TiO2 nanoplates with an exposed (001) facet via a simple sintering process at different temperatures. The TiO2 anatase nanoplates obtained in this work have excellent thermal stability. The low heating solid state chemical (LSCM) method, which requires only room temperature or a temperature below 100 C, is an easy, environmentally friendly, and simple method of synthesizing nanomaterials. LSCM has been used to prepare many inorganic nanomaterials, including oxides, sulfides, metals, and composite. NH4TiOF3 nanoplates were converted to TiO2 nanoplates via solid-solid transformation. In particular, an anatase titania was obtained with an exposed 001 facet, without the use of toxic substances, such as HF and NH4F. The resulting photocatalyst was then used for the production of hydrogen [199]. Typically, this method of synthesis involved 8 mmol of ammonium hexafluorotitanate weighed and ground for 10 min in an agate mortar. Subsequently, 6 mmol of boric acid was added. After grinding at room temperature for 0.5 h, the resulting solid samples were heated in a water bath at 60 °C to perform the entire reaction. The as-prepared samples were washed with distilled water and dried under natural conditions for 12 h. The final products were fabricated through the calcination of precursor at different temperature. The TiO2 nanoplates obtained with this method presented a single layer thickness of approximately 50 nm. Moreover, this photocatalyst was characterized by a very good heat-stability, and it was able to withstand high temperatures and maintain a stable anatase phase. The anatase TiO2 nanoplates exhibit a high H2 production under different light source (UV and visible), due to the exposed (001) facet and the high crystallinity [199]. Not only TiO2 but also other types of semiconductors are proposed in the literature for the photocatalytic production of hydrogen. Additionally, for these there is an alternative synthesis, which can be defined green. For example, among these photocatalysts, CdS has received the most attentions, due to its superior light absorption and appropriate conduction-band level [218,219]. However, the CdS, as such suffers from the phenomenon of photocorrosion [220], which can be limited by using a cocatalyst, such as noble metals or tungsten, dispersed on the surface of the CdS [221] However, the use of noble metals and tungsten is a problem due to their limited availability and their high cost. It is necessary to think of alternative solutions for cocatalysts. An interesting proposal in the literature is MoS2, for which excellent results are reported in terms of hydrogen production if coupled to CdS, and therefore, used as a cocatalyst. Unfortunately, the problem with MoS2, as well as for CdS, is the use of the poisonous gas H2S which must be used as a source of sulfur during the preparation of these two semiconductors, at relatively high calcination temperatures (above 400 °C) [202]. This is a major drawback that has limited the production of two very effective semiconductors. For this reason, and for the principles related to the concept of green chemistry [16], the need arises to find an alternative for the synthesis of MoS2/CdS photocatalyst, which excludes the use of poisonous substances such as H2S, but which allows the necessary source of sulfur to be obtained in an ecological way and at a low temperature. In the literature, a hydrothermal synthesis method is proposed for the production of MoS2/CdS photocatalyst, at a temperature of 200 °C and which uses thiourea as a source of sulfur [201].
For the synthesis CdCl2·2H2O (precursor of Cd) and thiourea with the molar ratio of 1:3 were dissolved in 80 mL of deionized water; then in the above solution various quantities of Na2MoO4·2H2O (Mo precursor) have been added. The solution was homogeneously mixed in an autoclave and brought to a temperature of 200 °C for 24 h. Finally, the precipitate obtained was washed with ethanol and water and dried in a vacuum chamber overnight at room temperature. A similar method, always based on the use of thiourea instead of H2S, was applied for the synthesis of TiO2/MoS2/graphene composite used for the photocatalytic production of hydrogen [222]. The loading of MoS2 enhanced the light absorption of the MoS2/CdS composite, which would result in higher light energy utilization, and it was observed that the H2 evolution capability of MoS2/CdS was 17 times greater than that of CdS. This is certainly a method that allows to obtain a very effective catalyst, a competitive sulfide photocatalysts for efficient solar hydrogen production, but synthesized through an environmentally friendly and inexpensive method.
An effective method for synthesizing photocatalysts for the production of hydrogen from biomass was first reported. Another interesting method to produce photocatalysts in a green way is starting from waste materials. This is the case reported in the literature on the synthesis of Ta₂O₅ photocatalyst prepared started from Ta present in Waste tantalum capacitors.
Waste tantalum capacitors contain approximately 30–40 wt% of rare metal (Ta) tantalum. They are, therefore, one of the main secondary resources of Ta. From these it is possible to recover the tantalum and foresee its reuse for a low-cost preparation as a Ta2O5 photocatalyst. It is possible to obtain Ta2O5 starting from TaCl5 used as the precursor for synthesizing Ta2O5 by chlorination-sintering process as reported in the literature [223]. To improve the photocatalytic activity, the prepared Ta2O5 was further decorated with polyaniline PANI through a simple chemisorption approach. Photocatalytic performance of the samples was evaluated in terms of H2 evolution and photodegradation of RhB. Polyaniline (PANI), used to decorate Ta2O5, is essential to increase the absorption of visible light by the catalyst. With increasing the PANI content, the visible light absorption of Ta2O5/PANI gradually enhanced. It is well known that PANI can absorb visible light. The visible absorption can be related to the electron transitions in PANI from polar on band to p* band [223]. In particular, with this prepared photocatalyst approximately 200.8 μmoLg−1h−1 of H2 have been obtained have been obtained under solar simulated light. Although this is a solution that sees the giving of new life to a waste, the production of a photocatalyst starting from tantalum capacitors to be disposed of consists of various steps. First approximately 10 g of sample must be pretreated by pyrolysis, crushing and screening to obtain particles rich in Ta. Ta-rich particles and FeCl2 powder with the adding amount of 50% were blended well in a quartz boat and put in a tube furnace, in presence of Argon, AT 500 °C for 2 h. With the chlorination process, Ta was selectively reacted with the chlorinating agent (FeCl2) and only TaCl5 was evaporated and collected in the ethanol solution. The TaCl5 ethanol solution was transferred to a water bath and kept at 70 ° C overnight. After being cooled to room temperature, the precipitates were collected by filtration and washed with deionized water. Finally, the precipitates were heated to 700 °C for 6 h under air and approximately 4.12 g of Ta2O5 nanoparticles were obtained. The advantage of this method proposed in the literature lies in the fact that it is a method that allows not only the recovery of Ta, but also its transformation into a compound with added value. However, as the authors of the paper rightly state, this recycling technology is valid but is still at the laboratory stage. It is necessary to determine more detailed technical parameters and the true environmental impact of this method (waste and CO2 emissions). The application of the life cycle analysis to evaluate the potential impact of the whole process on the environment and its sustainable definition compared to other methods is certainly necessary, not only for this method, but for all the proposed synthesis methods defined green.
Another peculiarity required for low-cost and environmentally friendly photocatalyst is their recyclability. The reuse of a photocatalyst for different process cycles is a fundamental requirement to define the sustainable and low environmental impact of photocatalysts. A catalyst is reusable if it can be easily separated from the stream (gas or liquid), in which it is used and above all if it does not lose its performance during the various reuses. The heterogeneous photocatalysts in powder form have the limitation of having to be separated by a filtration method, which will certainly be expensive and limiting in the reuse of the photocatalyst itself. Therefore, the solution supports the photocatalyst on a material that makes it easy to recover and reuse. Specifically, for the H2 production, there is little information in the literature in terms of supported photocatalyst. As specified above, the immobilization of the active species on a macroscopic support could generate limitations of mass transport phenomena with a consequent reduction in the performance of the process. The solution could be to preserve the advantages of a “suspended photocatalyst” but at the same time it is easy to remove and reuse. An interesting solution proposed in the literature concerns the use of particles with magnetic properties that, coupled with a photocatalyst, form the so called magnetic photocatalyst [224]. The use of this type of support is reported above all for environmental purposes (degradation of contaminants in water) and little has been developed specifically for the production of hydrogen. In any case, the synthesis of this type of supported photocatalyst is certainly very interesting. Magnetic photocatalysts include two components: a magnetic core to make the composite easy to remove by an external magnetic force, and a semiconductor shell with photocatalytic activity under irradiation. Magnetite (Fe3O4), maghemite (γ-Fe2O3), and M-Fe2O4 (where M stands for Ba2+, Ni 2+, Co2+, Zn2+ divalent metallic cations) are the three most widely used magnetic materials coupled with semiconductors and used as heterogeneous photocatalysts nostro ferromagnetiche [225].
Ru-doped LaFeO3 photocatalyst coupled with magnetic Fe2O3 particles was tested in the photocatalytic hydrogen production from glucose solution under visible light irradiation. The catalyst was successfully synthesized by solution combustion synthesis and immobilized on Fe2O3 magnetic particles by a physical mixture between Ru-LaFeO3 catalyst and Fe2O3 particles. The physical mixture was dispersed in distilled water and Tetraethyl orthosilicate was added to the solution. the recovered sample was calcined at 450 °C for 30 min to obtain the Ru-LaFeO3/Fe2O3 composite. The obtained catalyst is easy to recover after treatment, simply by using an external magnetic force. Furthermore, it has shown an excellent performance in terms of hydrogen production starting from aqueous glucose solution (5460 of μmoL/L after 4 h of visible irradiation time), but above all an efficient stability having maintained its performances unchanged even after several cycles of use [208]. Therefore, this is a definable “sustainable solution” as it allows to avoid the separation phase by filtering the photocatalyst at the end of the process. Furthermore, the photocatalyst is stable and active in the presence of visible light, which makes the process certainly advantageous from an economic point of view. Other solutions are proposed to immobilize the photocatalytic active phase on macroscopic supports, but such solutions have not yet been applied for the production of hydrogen. However, these are interesting solutions that could also be studied in the photocatalytic production of hydrogen. It is for example immobilization on supports such as glass spheres, quartz flakes, polystyrene pellets [226,227,228]. The deposition methods proposed are almost always simple and short, which is important for a low-cost and ecofriendly synthesis process. It must be specified that whatever the support is, it must be a material that does not absorb light in order not to compete with the ability of the photocatalytic phase to absorb light. This condition can be overcome only if the support is itself a semiconductor. In this case, a synergistic effect between the photocatalytic active phase and the relative support will be desirable, as reported in the Ru-LaFeO3/Fe2O3 composite case [208].

6. Conclusions

The observation that is certainly needed concerns the nature of the so-called “green” synthesis processes for the photocatalysts. These are certainly interesting processes that allow the production of semiconductors to be adopted in different photocatalytic contexts prepared from waste substances or biomasses, capable of absorbing visible or solar light very well, thanks also to the presence of codopages. The photocatalysts mentioned in this paper are effective for various purposes, from the abatement of pollutants to the production of hydrogen. However, it is believed that in order to define the synthesis of a photocatalyst “green” (whatever its final use), an environmental impact assessment and a life cycle assessment are necessary. From the evaluation of the literature reported in this work, it can be said that the study of ecofriendly and low-cost photocatalyst synthesis focuses more on the efficiency of the synthesis method and the prepared catalyst. No consideration is given to the environmental impact that a preparation method can have with respect to another, especially in the case of low-cost synthesis proposals. However, it must be said that the approaches reported in the literature are absolutely correct, because the proposal of any synthesis method must be justified by the functionality of the photocatalyst and its characteristics, which must respect those required. The photocatalysts prepared with green methods, starting from waste or biomass, must be competitive (in terms of functionality and stability) with the photocatalysts prepared with the classic synthesis methods. The same goes for the choice of codoping, which must be able to significantly enhance the performance of the photocatalysts, otherwise, the choice of using two elements for doping a semiconductor would not make sense. What emerged from this study is the possibility of obtaining promising photocatalysts using alternative synthesis techniques, which can be applied in different contexts.

7. Future Development

Based on the information reported in this review, future research regarding the synthesis methods of heterogeneous photocatalysts will certainly have to include evaluations and insights in terms of LCA between traditional and innovative methods in order to make the choice of innovative methods even more convincing. Moreover, we believe that, in order to consider an innovative photocatalyst, it must be prepared not only with simple procedures, but above all, it must be an environmentally and economically friendly catalyst, and therefore, usable for different process cycles. Experiments are rarely aimed at evaluating the recyclability and stability of a photocatalyst. Another interesting evaluation that could be proposed in future research could be the study of the photocatalytic performance in the degradation of a mix of pollutants (not of a single pollutant, as typically reported in the literature). This could be of strong interest to the scientific community. Formulating a photocatalyst capable of mineralizing a mix of contaminants at the same time means being very close to the real wastewater conditions and therefore, proposing a real and fast application of the photocatalytic process. Therefore, future studies should be carried out on these aspects.

Author Contributions

Conceptualization, G.I.; investigation, A.M. and G.I.; data curation, G.I.; writing—original draft preparation, A.M.; writing—review and editing, G.I.; visualization, G.I. and A.M.; supervision, G.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external founding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lee, K.M.; Lai, C.W.; Ngai, K.S.; Juan, J.C. Recent Developments of Zinc Oxide Based Photocatalyst in Water Treatment Technology: A Review. Water Res. 2016, 88, 428–448. [Google Scholar] [CrossRef] [PubMed]
  2. Gopinath, K.P.; Madhav, N.V.; Krishnan, A.; Malolan, R.; Rangarajan, G. Present Applications of Titanium Dioxide for the Photocatalytic Removal of Pollutants from Water: A Review. J. Environ. Manag. 2020, 270, 110906. [Google Scholar] [CrossRef] [PubMed]
  3. Zhu, D.; Zhou, Q. Action and Mechanism of Semiconductor Photocatalysis on Degradation of Organic Pollutants in Water Treatment: A Review. Environ. Nanotechnol. Monit. Manag. 2019, 12, 100255. [Google Scholar] [CrossRef]
  4. Zinicovscaia, I. Conventional methods of wastewater treatment. Cyanobacteria Bioremediation Wastewaters 2016, 17–25. [Google Scholar] [CrossRef]
  5. Rajasulochana, P.; Preethy, V. Comparison on Efficiency of Various Techniques in Treatment of Waste and Sewage Water–A Comprehensive Review. Resour. Technol. 2016, 2, 175–184. [Google Scholar] [CrossRef]
  6. Gutierrez-Urbano, I.; Villen-Guzman, M.; Perez-Recuerda, R.; Rodriguez-Maroto, J.M. Removal of Polycyclic Aromatic Hydrocarbons (Pahs) in Conventional Drinking Water Treatment Processes. J. Contam. Hydrol. 2021, 243, 103888. [Google Scholar] [CrossRef] [PubMed]
  7. Dong, G.; Chen, B.; Liu, B.; Hounjet, L.J.; Cao, Y.; Stoyanov, S.R.; Yang, M.; Zhang, B. Advanced Oxidation Processes in Microreactors for Water and Wastewater Treatment: Development, Challenges, and Opportunities. Water Res. 2022, 211, 118047. [Google Scholar] [CrossRef]
  8. Andreozzi, R.; Caprio, V.; Insola, A.; Marotta, R. Advanced Oxidation Processes (AOP) for Water Purification and Recovery. Catal. Today 1999, 53, 51–59. [Google Scholar] [CrossRef]
  9. Dewil, R.; Mantzavinos, D.; Poulios, I.; Rodrigo, M.A. New Perspectives for Advanced Oxidation Processes. J. Environ. Manag. 2017, 195, 93–99. [Google Scholar] [CrossRef]
  10. Ibhadon, A.O.; Fitzpatrick, P. Heterogeneous Photocatalysis: Recent Advances and Applications. Catalysts 2013, 3, 189–218. [Google Scholar] [CrossRef] [Green Version]
  11. Al-Mamun, M.R.; Kader, S.; Islam, M.S.; Khan, M.Z.H. Photocatalytic Activity Improvement and Application of UV-TiO2 Photocatalysis in Textile Wastewater Treatment: A Review. J. Environ. Chem. Eng. 2019, 7, 103248. [Google Scholar] [CrossRef]
  12. Vaiano, V.; Iervolino, G.; Rizzo, L.; Sannino, D. Advanced Oxidation Processes for the Removal of Food Dyes in Wastewater. Curr. Org. Chem. 2017, 21, 1068–1073. [Google Scholar] [CrossRef]
  13. Fujishima, A.; Rao, T.N.; Tryk, D.A. Titanium Dioxide Photocatalysis. J. Photochem. Photobiol. C Photochem. Rev. 2000, 1, 1–21. [Google Scholar] [CrossRef]
  14. Xu, L.; Zhang, S.; Li, W.; Zhang, Z. Visible-Light-Mediated Oxidative Amidation of Aldehydes by Using Magnetic CdS Quantum Dots as a Photocatalyst. Chem. A Eur. J. 2021, 27, 5483–5491. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, M.; Li, H.; Zhang, J.; Lv, H.; Yang, G.-Y. Research Advances of Light-Driven Hydrogen Evolution Using Polyoxometalate-Based Catalysts. Chin. J. Catal. 2021, 42, 855–871. [Google Scholar] [CrossRef]
  16. Li, C.J.; Trost, B.M. Green Chemistry for Chemical Synthesis. Proc. Natl. Acad. Sci. USA 2008, 105, 13197–13202. [Google Scholar] [CrossRef]
  17. Mian, M.; Liu, G. Recent Progress in Biochar-Supported Photocatalysts: Synthesis, Role of Biochar, and Applications. RSC Adv. 2018, 8, 14237–14248. [Google Scholar] [CrossRef]
  18. Basavarajappa, P.S.; Patil, S.B.; Ganganagappa, N.; Reddy, K.R.; Raghu, A.V.; Reddy, C.V. Recent Progress in Metal-Doped TiO2, Non-Metal Doped/Codoped TiO2 and TiO2 Nanostructured Hybrids for Enhanced Photocatalysis. Int. J. Hydrogen Energy 2019, 45, 7764–7778. [Google Scholar] [CrossRef]
  19. Cheng, H.-H.; Chen, S.-S.; Yang, S.-Y.; Liu, H.-M.; Lin, K.-S. Sol-Gel Hydrothermal Synthesis and Visible Light Photocatalytic Degradation Performance of Fe/N Codoped TiO2 Catalysts. Materials 2018, 11, 939. [Google Scholar] [CrossRef]
  20. Lam, S.W.; Chiang, K.; Lim, T.M.; Amal, R.; Low, G.K.-C. The Effect of Platinum and Silver Deposits in the Photocatalytic Oxidation of Resorcinol. Appl. Catal. B Environ. 2007, 72, 363–372. [Google Scholar] [CrossRef]
  21. Kamat, P.V. Manipulation of Charge Transfer Across Semiconductor Interface. A Criterion That Cannot Be Ignored in Photocatalyst Design. J. Phys. Chem. Lett. 2012, 3, 663–672. [Google Scholar] [CrossRef] [PubMed]
  22. Marami, M.B.; Farahmandjou, M.; Khoshnevisan, B. Sol–Gel Synthesis of Fe-Doped TiO2 Nanocrystals. J. Electron. Mater. 2018, 47, 3741–3748. [Google Scholar] [CrossRef]
  23. Jafari, A.J.; Moslemzadeh, M. Synthesis of Fe-Doped TiO2 for Photocatalytic Processes under UV-Visible Light: Effect of Preparation Methods on Crystal Size—A Systematic Review Study. Comments Inorg. Chem. 2020, 40, 327–346. [Google Scholar] [CrossRef]
  24. Sood, S.; Umar, A.; Mehta, S.K.; Kansal, S.K. Highly Effective Fe-Doped TiO2 Nanoparticles Photocatalysts for Visible-Light Driven Photocatalytic Degradation of Toxic Organic Compounds. J. Colloid Interface Sci. 2015, 450, 213–223. [Google Scholar] [CrossRef]
  25. Ali, T.; Tripathi, P.; Azam, A.; Raza, W.; Ahmed, A.S.; Ahmed, A.; Muneer, M. Photocatalytic Performance of Fe-Doped TiO2 Nanoparticles Under Visible-Light Irradiation. Mater. Res. Express 2017, 4, 015022. [Google Scholar] [CrossRef]
  26. Guan, B.; Yu, J.; Guo, S.; Yu, S.; Han, S. Porous Nickel Doped Titanium Dioxide Nanoparticles with Improved Visible Light Photocatalytic Activity. Nanoscale Adv. 2020, 2, 1352–1357. [Google Scholar] [CrossRef]
  27. Shen, J.-H.; Chuang, H.-Y.; Jiang, Z.-W.; Liu, X.-Z.; Horng, J.-J. Novel Quantification of Formation Trend and Reaction Efficiency of Hydroxyl Radicals for Investigating Photocatalytic Mechanism of Fe-Doped TiO2 During UV And Visible Light-Induced Degradation of Acid Orange 7. Chemosphere 2020, 251, 126380. [Google Scholar] [CrossRef]
  28. Ebrahimi, R.; Maleki, A.; Rezaee, R.; Daraei, H.; Safari, M.; McKay, G.; Lee, S.-M.; Jafari, A. Synthesis and Application of Fe-Doped TiO2 Nanoparticles for Photodegradation of 2,4-D from Aqueous Solution. Arab. J. Sci. Eng. 2020, 46, 6409–6422. [Google Scholar] [CrossRef]
  29. Bhatia, V.; Dhir, A. Transition Metal Doped TiO2 Mediated Photocatalytic Degradation of Anti-Inflammatory Drug Under Solar Irradiations. J. Environ. Chem. Eng. 2016, 4, 1267–1273. [Google Scholar] [CrossRef]
  30. Santara, B.; Pal, B.; Giri, P.K. Signature of Strong Ferromagnetism and Optical Properties of Co Doped TiO2 Nanoparticles. J. Appl. Phys. 2011, 110, 114322. [Google Scholar] [CrossRef]
  31. Kayestha, R.; Sumati; Hajela, K. ESR Studies on the Effect of Ionic Radii on Displacement of Mn2+ Bound to a Soluble Β-Galactoside Binding Hepatic Lectin. FEBS Lett. 1995, 368, 285–288. [Google Scholar] [CrossRef]
  32. Krishnakumar, V.; Boobas, S.; Jayaprakash, J.; Rajaboopathi, M.; Han, B.; Louhi-Kultanen, M. Effect of Cu Doping on TiO2 Nanoparticles and Its Photocatalytic Activity Under Visible Light. J. Mater. Sci. Mater. Electron. 2016, 27, 7438–7447. [Google Scholar] [CrossRef]
  33. Klein, M.; Grabowska, E.; Zaleska-Medynska, A. Noble Metal Modified TiO2 for Photocatalytic Air Purification. Physicochem. Probl. Miner. Processing 2015, 51, 49–57. [Google Scholar] [CrossRef]
  34. Xu, A.-W.; Gao, Y.; Liu, H.Q. The Preparation, Characterization, and their Photocatalytic Activities of Rare-Earth-Doped TiO2 Nanoparticles. J. Catal. 2002, 207, 151–157. [Google Scholar] [CrossRef]
  35. Li, F.; Li, X.; Ao, C.; Lee, S.; Hou, M. Enhanced Photocatalytic Degradation of Vocs Using Ln3+–TiO2 Catalysts for Indoor Air Purification. Chemosphere 2005, 59, 787–800. [Google Scholar] [CrossRef]
  36. Hewer, T.; Souza, E.; Martins, T.S.; Muccillo, E.; Freire, R. Influence of Neodymium Ions on Photocatalytic Activity of TiO2 Synthesized by Sol–Gel and Precipitation Methods. J. Mol. Catal. A Chem. 2010, 336, 58–63. [Google Scholar] [CrossRef]
  37. Ramya, R.; Krishnan, P.S.; Neelaveni, M.; Gurulakshmi, M.; Sivakumar, T.; Shanthi, K. Enhanced Visible Light Activity of Pr–TiO2 Nanocatalyst in the Degradation of Dyes: Effect of Pr Doping and TiO2 Morphology. J. Nanosci. Nanotechnol. 2019, 19, 3971–3981. [Google Scholar] [CrossRef]
  38. Reszczynska, J.; Esteban, D.A.; Gazda, M.; Zaleska-Medynska, A. Pr-doped TiO2. The Effect of Metal Content on Photocatalytic Activity. Physicochem. Probl. Miner. Processing 2014, 50, 515–524. [Google Scholar] [CrossRef]
  39. Cordeiro, D.S.; Cassio, F.L.; Ciccotti, L.; Hewer, T.L.R.; Corio, P.; Freire, R.S. Photocatalytic Activity of Pr-Modified TiO2 for the Degradation of Bisphenol A. SN Appl. Sci. 2021, 3, 258. [Google Scholar] [CrossRef]
  40. Xiao, Q.; Si, Z.; Yu, Z.; Qiu, G. Sol–Gel Auto-Combustion Synthesis of Samarium-Doped TiO2 Nanoparticles and Their Photocatalytic Activity Under Visible Light Irradiation. Mater. Sci. Eng. B 2007, 137, 189–194. [Google Scholar] [CrossRef]
  41. Singh, K.; Harish, S.; Kristy, A.P.; Shivani, V.; Archana, J.; Navaneethan, M.; Shimomura, M.; Hayakawa, Y. Erbium Doped TiO2 Interconnected Mesoporous Spheres as an Efficient Visible Light Catalyst for Photocatalytic Applications. Appl. Surf. Sci. 2018, 449, 755–763. [Google Scholar] [CrossRef]
  42. Silva, A.M.; Silva, C.G.; Dražić, G.; Faria, J.L. Ce-Doped TiO2 for Photocatalytic Degradation of Chlorophenol. Catal. Today 2009, 144, 13–18. [Google Scholar] [CrossRef]
  43. Zhang, Z.; Zhao, C.; Duan, Y.; Wang, C.; Zhao, Z.; Wang, H.; Gao, Y. Phosphorus-Doped TiO2 for Visible Light-Driven Oxidative Coupling of Benzyl Amines and Photodegradation of Phenol. Appl. Surf. Sci. 2020, 527, 146693. [Google Scholar] [CrossRef]
  44. Piątkowska, A.; Janus, M.; Szymański, K.; Mozia, S. C-, N-and S-Doped TiO2 Photocatalysts: A Review. Catalysts 2021, 11, 144. [Google Scholar] [CrossRef]
  45. Vaiano, V.; Sacco, O.; Sannino, D.; Ciambelli, P. Nanostructured N-doped TiO2 Coated on Glass Spheres for the Photocatalytic Removal of Organic Dyes Under UV or Visible Light Irradiation. Appl. Catal. B: Environ. 2015, 170–171, 153–161. [Google Scholar] [CrossRef]
  46. Sun, J.; Qiao, L.; Sun, S.; Wang, G. Photocatalytic Degradation of Orange G on Nitrogen-Doped TiO2 Catalysts Under Visible Light and Sunlight Irradiation. J. Hazard. Mater. 2008, 155, 312–319. [Google Scholar] [CrossRef]
  47. Liu, Y.; Chen, X.; Li, J.; Burda, C. Photocatalytic Degradation of Azo Dyes by Nitrogen-Doped TiO2 Nano catalysts. Chemosphere 2005, 61, 11–18. [Google Scholar] [CrossRef]
  48. Ragesh Nath, R.; Nethravathi, C.; Rajamathi, M. Hierarchically Porous, Biphasic, and C-Doped TiO2 for Solar Photocatalytic Degradation of Dyes and Selective Oxidation of Benzyl Alcohol. ACS Omega 2021, 6, 12124–12132. [Google Scholar] [CrossRef]
  49. Umebayashi, T.; Yamaki, T.; Tanaka, S.; Asai, K. Visible Light-Induced Degradation of Methylene Blue on S-doped TiO2. Chem. Lett. 2003, 32, 330–331. [Google Scholar] [CrossRef]
  50. Kuźmiński, K.; Morawski, A.W.; Janus, M. Adsorption and Photocatalytic Degradation of Anionic and Cationic Surfactants on Nitrogen-Modified TiO2. J. Surfactants Deterg. 2018, 21, 909–921. [Google Scholar] [CrossRef]
  51. Sacco, O.; Vaiano, V.; Rizzo, L.; Sannino, D. Photocatalytic Activity of a Visible Light Active Structured Photocatalyst Developed for Municipal Wastewater Treatment. J. Clean. Prod. 2018, 175, 38–49. [Google Scholar] [CrossRef]
  52. Sun, P.; Zhang, J.; Liu, W.; Wang, Q.; Cao, W. Modification to L-H Kinetics Model and Its Application in the Investigation on Photodegradation of Gaseous Benzene by Nitrogen-Doped TiO2. Catalysts 2018, 8, 326. [Google Scholar] [CrossRef]
  53. Kamaei, M.; Rashedi, H.; Dastgheib, S.M.M.; Tasharrofi, S. Comparing Photocatalytic Degradation of Gaseous Ethylbenzene Using N-doped and Pure TiO2 Nano-Catalysts Coated on Glass Beads under Both UV and Visible Light Irradiation. Catalysts 2018, 8, 466. [Google Scholar] [CrossRef]
  54. Preethi, L.; Antony, R.P.; Mathews, T.; Loo, J.; Wong, L.H.; Dash, S.; Tyagi, A. Nitrogen Doped Anatase-Rutile Heterostructured Nanotubes for Enhanced Photocatalytic Hydrogen Production: Promising Structure for Sustainable Fuel Production. Int. J. Hydrogen Energy 2016, 41, 5865–5877. [Google Scholar] [CrossRef]
  55. Ansari, S.A.; Khan, M.M.; Ansari, M.O.; Cho, M.H. Nitrogen-Doped Titanium Dioxide (N-Doped TiO2) for Visible Light Photocatalysis. New J. Chem. 2016, 40, 3000–3009. [Google Scholar] [CrossRef]
  56. Bersezio, C.; Pardo, C.; Miranda, S.; Maran, B.M.; Jorquera, G.; da Silva, A., Jr.; Rodrigues, M.T.; Fernández, E. Evaluation of the Effectiveness in Teeth Whitening of a Single Session with 6% Hydrogen Peroxide Laser/LED System. Photodiagnosis Photodyn. Ther. 2021, 36, 102532. [Google Scholar] [CrossRef]
  57. Huang, J.; Song, H.; Chen, C.; Yang, Y.; Xu, N.; Ji, X.; Li, C.; You, J.-A. Facile Synthesis of N-Doped TiO2 Nanoparticles Caged In MIL-100(Fe) For Photocatalytic Degradation of Organic Dyes Under Visible Light Irradiation. J. Environ. Chem. Eng. 2017, 5, 2579–2585. [Google Scholar] [CrossRef]
  58. Kawashita, M.; Endo, N.; Watanabe, T.; Miyazaki, T.; Furuya, M.; Yokota, K.; Abiko, Y.; Kanetaka, H.; Takahashi, N. Formation of Bioactive N-Doped TiO2 on Ti with Visible Light-Induced Antibacterial Activity Using Naoh, Hot Water, and Subsequent Ammonia Atmospheric Heat Treatment. Colloids Surf. B Biointerfaces 2016, 145, 285–290. [Google Scholar] [CrossRef]
  59. Kim, T.H.; Go, G.-M.; Cho, H.-B.; Song, Y.; Lee, C.-G.; Choa, Y.-H. A Novel Synthetic Method for N Doped TiO2 Nanoparticles Through Plasma-Assisted Electrolysis and Photocatalytic Activity in the Visible Region. Front. Chem. 2018, 6, 458. [Google Scholar] [CrossRef]
  60. 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]
  61. Qin, H.-L.; Gu, G.-B.; Liu, S. Preparation of nitrogen-Doped Titania with Visible-Light Activity and Its Application. Comptes Rendus. Chim. 2008, 11, 95–100. [Google Scholar] [CrossRef]
  62. Bakre, P.V.; Tilve, S.; Shirsat, R. Influence of N Sources on the Photocatalytic Activity of N-Doped TiO2. Arab. J. Chem. 2020, 13, 7637–7651. [Google Scholar] [CrossRef]
  63. Matos, J.; Miralles-Cuevas, S.; Ruíz-Delgado, A.; Oller, I.; Malato, S. Development of TiO2-C Photocatalysts for Solar Treatment of Polluted Water. Carbon 2017, 122, 361–373. [Google Scholar] [CrossRef]
  64. Jia, G.; Wang, Y.; Cui, X.; Zheng, W. Highly Carbon-Doped TiO2 Derived from MXene Boosting the Photocatalytic Hydrogen Evolution. ACS Sustain. Chem. Eng. 2018, 6, 13480–13486. [Google Scholar] [CrossRef]
  65. Wang, P.; Yin, G.; Bi, Q.; Huang, X.; Du, X.; Zhao, W.; Huang, F.-Q. Efficient Photocatalytic Reduction of CO2 Using Carbon-Doped Amorphous Titanium Oxide. ChemCatChem 2018, 10, 3854–3861. [Google Scholar] [CrossRef]
  66. Ohno, T.; Akiyoshi, M.; Umebayashi, T.; Asai, K.; Mitsui, T.; Matsumura, M. Preparation of S-Doped TiO2 Photocatalysts and Their Photocatalytic Activities Under Visible Light. Appl. Catal. A Gen. 2004, 265, 115–121. [Google Scholar] [CrossRef]
  67. McManamon, C.; O’Connell, J.; Delaney, P.; Rasappa, S.; Holmes, J.; Morris, M. A Facile Route to Synthesis of S-Doped TiO2 Nanoparticles for Photocatalytic Activity. J. Mol. Catal. A Chem. 2015, 406, 51–57. [Google Scholar] [CrossRef]
  68. Helmy, E.T.; Abouellef, E.M.; Soliman, U.A.; Pan, J.H. Novel Green Synthesis of S-Doped TiO2 Nanoparticles Using Malva Parviflora Plant Extract and Their Photocatalytic, Antimicrobial and Antioxidant Activities Under Sunlight Illumination. Chemosphere 2020, 271, 129524. [Google Scholar] [CrossRef]
  69. Zhu, M.; Zhai, C.; Qiu, L.; Lu, C.; Paton, A.S.; Du, Y.; Goh, M.C. New Method to Synthesize S-Doped TiO2 with Stable and Highly Efficient Photocatalytic Performance under Indoor Sunlight Irradiation. ACS Sustain. Chem. Eng. 2015, 3, 3123–3129. [Google Scholar] [CrossRef]
  70. Bu, X.; Wang, Y.; Li, J.; Zhang, C. Improving the Visible Light Photocatalytic Activity of TiO2 by Combining Sulfur Doping and Rectorite Carrier. J. Alloys Compd. 2015, 628, 20–26. [Google Scholar] [CrossRef]
  71. Sharotri, N.; Gupta, S.; Sud, D. Visible Light Responsive S-Doped TiO2 Nanoparticles: Synthesis, Characterization and Photocatalytic Degradation of Pollutants. Nanotechnol. Environ. Eng. 2022, 7, 503–515. [Google Scholar] [CrossRef]
  72. Boningari, T.; Inturi, S.N.R.; Suidan, M.; Smirniotis, P.G. Novel One-Step Synthesis of Sulfur Doped-TiO2 by Flame Spray Pyrolysis for Visible Light Photocatalytic Degradation of Acetaldehyde. Chem. Eng. J. 2018, 339, 249–258. [Google Scholar] [CrossRef]
  73. Bakar, S.A.; Ribeiro, C. A comparative Run for Visible-Light-Driven Photocatalytic Activity of Anionic and Cationic S-Doped TiO2 Photocatalysts: A Case Study of Possible Sulfur Doping Through Chemical Protocol. J. Mol. Catal. A Chem. 2016, 421, 1–15. [Google Scholar] [CrossRef]
  74. Liu, C.; Chen, W.; Sheng, Y.; Li, L. Atrazine Degradation in Solar Irradiation/S-Doped Titanium Dioxide Treatment. In Proceedings of the 2009 3rd International Conference on Bioinformatics and Biomedical Engineering, Beijing, China, 11–13 June 2009; pp. 1–4. [Google Scholar]
  75. Lin, Y.-H.; Chou, S.-H.; Chu, H. A Kinetic Study for the Degradation of 1,2-Dichloroethane By S-Doped TiO2 Under Visible Light. J. Nanoparticle Res. 2014, 16, 1–12. [Google Scholar] [CrossRef]
  76. Baeissa, E. Synthesis and Characterization of Sulfur-Titanium Dioxide Nanocomposites for Photocatalytic Oxidation of Cyanide Using Visible Light Irradiation. Chin. J. Catal. 2015, 36, 698–704. [Google Scholar] [CrossRef]
  77. Mancuso, A.; Sacco, O.; Sannino, D.; Pragliola, S.; Vaiano, V. Enhanced Visible-Light-Driven Photodegradation of Acid Orange 7 Azo Dye in Aqueous Solution Using Fe-N Co-Doped TiO2. Arab. J. Chem. 2020, 13, 8347–8360. [Google Scholar] [CrossRef]
  78. Thambiliyagodage, C.; Usgodaarachchi, L. Photocatalytic activity of N, Fe and Cu co-doped TiO2 Nanoparticles Under Sunlight. Curr. Res. Green Sustain. Chem. 2021, 4, 100186. [Google Scholar] [CrossRef]
  79. Iervolino, G.; Vaiano, V.; Rizzo, L. Visible light active Fe-doped TiO2 for the Oxidation of Arsenite to Arsenate in Drinking Water. Chem. Eng. Trans. 2018, 70, 1573–1578. [Google Scholar]
  80. Iervolino, G.; Vaiano, V.; Matarangolo, M.; Rizzo, L. Photocatalytic Oxidation of Arsenite to Arsenate Using a Continuous Packed Bed Photoreactor. Chem. Eng. Trans. 2019, 73, 253–258. [Google Scholar] [CrossRef]
  81. Mancuso, A.; Sacco, O.; Vaiano, V.; Sannino, D.; Pragliola, S.; Venditto, V.; Morante, N. Visible Light Active Fe-Pr Co-Doped TiO2 for Water Pollutants Degradation. Catal. Today 2021, 380, 93–104. [Google Scholar] [CrossRef]
  82. Jeong, E.D.; Borse, P.H.; Jang, J.S.; Lee, J.S.; Jung, O.S.; Chang, H.; Jin, J.S.; Won, M.S.; Kim, H.G. Hydrothermal Synthesis of Cr and Fe Co-Doped TiO2 Nanoparticle Photocatalyst. J. Ceram. Processing Res. 2008, 9, 250–253. [Google Scholar]
  83. Liu, Q.; Zhou, Y.; Duan, Y.; Wang, M.; Lin, Y. Improved Photovoltaic Performance of Dye-Sensitized Solar Cells (Dsscs) By Zn+Mg Co-Doped TiO2 Electrode. Electrochim. Acta 2013, 95, 48–53. [Google Scholar] [CrossRef]
  84. Qin, Z.; Zhao, Z.; Jiao, W.; Han, Z.; Xia, L.; Fang, Y.; Wang, S.; Ji, L.; Jiang, Y. Coupled Photocatalytic-Bacterial Degradation of Pyrene: Removal Enhancement and Bacterial Community Responses. Environ. Res. 2020, 183, 109135. [Google Scholar] [CrossRef]
  85. Sharotri, N.; Sharma, D.; Sud, D. Experimental and Theoretical Investigations of Mn-N-Co-Doped TiO2 Photocatalyst for Visible Light Induced Degradation of Organic Pollutants. J. Mater. Res. Technol. 2019, 8, 3995–4009. [Google Scholar] [CrossRef]
  86. Le, T.T.T.; Tran, T.D. Photocatalytic Degradation of Rhodamine B by C and N Codoped TiO2 Nanoparticles under Visible-Light Irradiation. J. Chem. 2020, 2020, 1–8. [Google Scholar] [CrossRef]
  87. Li, D.; Haneda, H.; Hishita, S.; Ohashi, N. Visible-Light-Driven N− F− Codoped TiO2 Photocatalysts. 2. Optical Characterization, Photocatalysis, And Potential Application to Air Purification. Chem. Mater. 2005, 17, 2596–2602. [Google Scholar] [CrossRef]
  88. Zheng, P.; Wu, H.; Guo, J.; Dong, J.; Jia, S.; Zhu, Z. P–N Co-Doping Induced Structural Recovery of TiO2 For Overall Water Splitting Under Visible Light Irradiation. J. Alloys Compd. 2014, 615, 79–83. [Google Scholar] [CrossRef]
  89. Qaderi, J.; Mamat, C.R.; Jalil, A.A. Preparation and Characterization of Copper, Iron, and Nickel Doped Titanium Dioxide Photocatalysts for Decolorization of Methylene Blue. Sains Malays. 2021, 50, 135–149. [Google Scholar] [CrossRef]
  90. Li, H.; Hao, Y.; Lu, H.; Liang, L.; Wang, Y.; Qiu, J.; Shi, X.; Wang, Y.; Yao, J. A Systematic Study on Visible-Light N-Doped TiO2 Photocatalyst Obtained from Ethylenediamine by Sol–Gel Method. Appl. Surf. Sci. 2015, 344, 112–118. [Google Scholar] [CrossRef]
  91. Grabowska-Musiał, E.; Zaleska-Medynska, A.; Sobczak, J.; Gazda, M.; Hupka, J. Boron-Doped TiO2: Characteristics and Photoactivity Under Visible Light. Procedia Chem. 2009, 1, 1553–1559. [Google Scholar] [CrossRef]
  92. Gopal, N.O.; Lo, H.-H.; Ke, T.-F.; Lee, C.-H.; Chou, C.-C.; Wu, J.-D.; Sheu, S.-C.; Ke, S.-C. Visible Light Active Phosphorus-Doped TiO2 Nanoparticles: An EPR Evidence for the Enhanced Charge Separation. J. Phys. Chem. C 2012, 116, 16191–16197. [Google Scholar] [CrossRef]
  93. Mugundan, S.; Rajamannan, B.; Viruthagiri, G.; Shanmugam, N.; Gobi, R.; Praveen, P. Synthesis and Characterization of Undoped and Cobalt-Doped TiO2 Nanoparticles Via Sol–Gel Technique. Appl. Nanosci. 2015, 5, 449–456. [Google Scholar] [CrossRef] [Green Version]
  94. Giahi, M.; Pathania, D.; Agarwal, S.; Gomaa, A.M.A.; Chong, K.F.; Gupta, V.K. Preparation of Mg-Doped TiO2 Nanoparticles for Photocatalytic Degradation of Some Organic Pollutants. Stud. Univ. Babes-Bolyai Chem. 2019, 64, 7–18. [Google Scholar] [CrossRef]
  95. Manzoor, M.; Rafiq, A.; Ikram, M.; Nafees, M.; Ali, S. Structural, Optical, And Magnetic Study of Ni-Doped TiO2 Nanoparticles Synthesized by Sol–Gel Method. Int. Nano Lett. 2018, 8, 1–8. [Google Scholar] [CrossRef]
  96. Zhang, X.; Liu, Q. Preparation and Characterization of Titania Photocatalyst Co-Doped with Boron, Nickel, and Cerium. Mater. Lett. 2008, 62, 2589–2592. [Google Scholar] [CrossRef]
  97. Srinivasan, S.S.; Wade, J.; Stefanakos, E.K.; Goswami, Y. Synergistic Effects of Sulfation and Co-Doping on the Visible Light Photocatalysis of TiO2. J. Alloys Compd. 2006, 424, 322–326. [Google Scholar] [CrossRef]
  98. Mugundan, S.; Praveen, P.; Sridhar, S.; Prabu, S.; Mary, K.L.; Ubaidullah, M.; Shaikh, S.F.; Kanagesan, S. Sol-Gel Synthesized Barium Doped TiO2 Nanoparticles for Solar Photocatalytic Application. Inorg. Chem. Commun. 2022, 139, 109340. [Google Scholar] [CrossRef]
  99. Huang, J.-G.; Guo, X.-T.; Wang, B.; Li, L.-Y.; Zhao, M.-X.; Dong, L.-L.; Liu, X.-J.; Huang, Y.-T. Synthesis and Photocatalytic Activity of Mo-Doped TiO2Nanoparticles. J. Spectrosc. 2015, 2015, 1–8. [Google Scholar] [CrossRef]
  100. Nithya, N.; Bhoopathi, G.; Magesh, G.; Kumar, C.D.N. Neodymium Doped TiO2 Nanoparticles by Sol-Gel Method for Antibacterial and Photocatalytic Activity. Mater. Sci. Semicond. Process. 2018, 83, 70–82. [Google Scholar] [CrossRef]
  101. Sacco, O.; Stoller, M.; Vaiano, V.; Ciambelli, P.; Chianese, A.; Sannino, D. Photocatalytic Degradation of Organic Dyes Under Visible Light on N-Doped TiO2 Photocatalysts. Int. J. Photoenergy 2012, 2012, 1–9. [Google Scholar] [CrossRef]
  102. Vaiano, V.; Sacco, O.; Iervolino, G.; Sannino, D.; Ciambelli, P.; Liguori, R.; Bezzeccheri, E.; Rubino, A. Enhanced Visible Light Photocatalytic Activity by Up-Conversion Phosphors Modified N-Doped TiO2. Appl. Catal. B Environ. 2015, 176, 594–600. [Google Scholar] [CrossRef]
  103. Tian, B.Z.; Li, C.Z.; Gu, F.; Jiang, H.B.; Hu, Y.J. Visible-Light Photocatalytic Activity of Cr-Doped TiO2 Nanoparticles Synthesized by Flame Spray Pyrolysis. J. Inorg. Mater. 2009, 24, 661–665. [Google Scholar] [CrossRef]
  104. Liu, J.; Zhao, Y.; Shi, L.; Yuan, S.; Fang, J.; Wang, Z.; Zhang, M. Solvothermal Synthesis of Crystalline Phase and Shape Controlled Sn4+-Doped TiO2 Nanocrystals: Effects of Reaction Solvent. ACS Appl. Mater. Interfaces 2011, 3, 1261–1268. [Google Scholar] [CrossRef]
  105. Cano-Casanova, L.; Amorós-Pérez, A.; Lillo-Ródenas, M.Á.; Román-Martínez, M.D.C. Effect of the Preparation Method (Sol-Gel or Hydrothermal) and Conditions on the TiO2 Properties and Activity for Propene Oxidation. Materials 2018, 11, 2227. [Google Scholar] [CrossRef] [PubMed]
  106. Hidalgo, M.C.; Aguilar-Moncayo, M.; Maicu, M.; Navío, J.A.; Colón, G. Hydrothermal Preparation of Highly Photoactive TiO2 Nanoparticles. Catal. Today 2007, 129, 50–58. [Google Scholar] [CrossRef]
  107. Zhou, J.K.; Lv, L.; Yu, J.; Li, H.L.; Guo, P.-Z.; Sun, A.H.; Zhao, X.S. Synthesis of Self-Organized Polycrystalline F-doped TiO2 Hollow Microspheres and Their Photocatalytic Activity under Visible Light. J. Phys. Chem. C 2008, 112, 5316–5321. [Google Scholar] [CrossRef]
  108. Wu, D.; Long, M.; Cai, W.; Chen, C.; Wu, Y. Low Temperature Hydrothermal Synthesis of N-Doped TiO2 Photocatalyst with High Visible-Light Activity. J. Alloys Compd. 2010, 502, 289–294. [Google Scholar] [CrossRef]
  109. Li, Z.; Shen, W.; He, W.; Zu, X. Effect of Fe-Doped TiO2 Nanoparticle Derived from Modified Hydrothermal Process on the Photocatalytic Degradation Performance on Methylene Blue. J. Hazard. Mater. 2008, 155, 590–594. [Google Scholar] [CrossRef]
  110. Zhu, J.; Zheng, W.; He, B.; Zhang, J.; Anpo, M. Characterization of Fe–TiO2 Photocatalysts Synthesized by Hydrothermal Method and Their Photocatalytic Reactivity for Photodegradation of XRG Dye Diluted in Water. J. Mol. Catal. A Chem. 2004, 216, 35–43. [Google Scholar] [CrossRef]
  111. Amano, F.; Yamakata, A.; Nogami, K.; Osawa, M.; Ohtani, B. Visible Light Responsive Pristine Metal Oxide Photocatalyst: Enhancement of Activity by Crystallization under Hydrothermal Treatment. J. Am. Chem. Soc. 2008, 130, 17650–17651. [Google Scholar] [CrossRef]
  112. Nasir, M.; Bagwasi, S.; Jiao, Y.; Chen, F.; Tian, B.; Zhang, J. Characterization and Activity of the Ce and N co-doped TiO2 Prepared Through Hydrothermal Method. Chem. Eng. J. 2014, 236, 388–397. [Google Scholar] [CrossRef]
  113. Xu, H.; Wang, K.; Liu, S.; Li, Y.; Chen, W. Preparation of Ce, N-Codoped Mesoporous Titania Photocatalyst with High Photocatalytic Activity under Visible Light. In Materials in Environmental Engineering; Hadi, H., Ed.; De Gruyter: Berlin, Germany, 2017; pp. 777–786. [Google Scholar] [CrossRef]
  114. Wang, Z.; Liu, S.; Cao, X.; Wu, S.; Liu, C.; Li, G.; Jiang, W.; Wang, H.; Wang, N.; Ding, W. Preparation and Characterization of TiO2 Nanoparticles by Two Different Precipitation Methods. Ceram. Int. 2020, 46, 15333–15341. [Google Scholar] [CrossRef]
  115. Banić, N.; Abramović, B.; Krstić, J.; Šojić, D.; Lončarević, D.; Cherkezova-Zheleva, Z.; Guzsvány, V. Photodegradation of Thiacloprid Using Fe/TiO2 as a Heterogeneous Photo-Fenton Catalyst. Appl. Catal. B Environ. 2011, 107, 363–371. [Google Scholar] [CrossRef]
  116. Jiao, W.; Liu, Y.; Qi, G. A New Impinging Stream–Rotating Packed Bed Reactor for Improvement of Micromixing Iodide and Iodate. Chem. Eng. J. 2010, 157, 168–173. [Google Scholar] [CrossRef]
  117. Zeng, G.; Zhang, Q.; Liu, Y.; Zhang, S.; Guo, J. Preparation of TiO2 and Fe-TiO2 with an Impinging Stream-Rotating Packed Bed by the Precipitation Method for the Photodegradation of Gaseous Toluene. Nanomaterials 2019, 9, 1173. [Google Scholar] [CrossRef]
  118. He, R.; Zhang, Q.; Liu, Y.; Guo, J.; Shen, H. Preparation of Fe and Co co-doped TiO2 by Precipitation Method in an Impinging Stream-Rotating Packed Bed for Photodegradation of Phenol Wastewater. Adv. Appl. Ceram. 2021, 120, 134–143. [Google Scholar] [CrossRef]
  119. Javaid, R.; Qazi, U.Y. Catalytic Oxidation Process for the Degradation of Synthetic Dyes: An Overview. Int. J. Environ. Res. Public Health 2019, 16, 2066. [Google Scholar] [CrossRef]
  120. Susan, E.F.; Elizabeth, A.D. Sensitization to Dyes in Textiles and Other Consumer Products. J. Toxicol. Cutan. Ocul. Toxicol. 1988, 7, 195–222. [Google Scholar] [CrossRef]
  121. Svedman, C.; Engfeldt, M.; Malinauskiene, L. Textile Contact Dermatitis: How Fabrics Can Induce Dermatitis. Curr. Treat. Options Allergy 2019, 6, 103–111. [Google Scholar] [CrossRef]
  122. Jamee, R.; Siddique, R. Biodegradation of Synthetic Dyes of Textile Effluent by Microorganisms: An Environmentally and Economically Sustainable Approach. Eur. J. Microbiol. Immunol. 2019, 9, 114–118. [Google Scholar] [CrossRef]
  123. Bhosale, R.R.; Pujari, S.R.; Muley, G.G.; Patil, S.H.; Patil, K.R.; Shaikh, M.F.; Gambhire, A.B. Solar Photocatalytic Degradation of Methylene Blue Using Doped TiO2 Nanoparticles. Sol. Energy 2014, 103, 473–479. [Google Scholar] [CrossRef]
  124. Shahmoradi, B.; Negahdary, M.; Maleki, A. Hydrothermal Synthesis of Surface-Modified, Manganese-Doped TiO2 Nanoparticles for Photodegradation of Methylene Blue. Environ. Eng. Sci. 2012, 29, 1032–1037. [Google Scholar] [CrossRef]
  125. Umar, K.; Aris, A.; Ahmad, H.; Parveen, T.; Jaafar, J.; Majid, Z.A.; Reddy, A.V.B.; Talib, J. Synthesis of Visible Light Active Doped TiO2 for the Degradation of Organic Pollutants—Methylene Blue and Glyphosate. J. Anal. Sci. Technol. 2016, 7, 725. [Google Scholar] [CrossRef]
  126. Nakhate, G.; Nikam, V.; Kanade, K.; Arbuj, S.; Kale, B.; Baeg, J. Hydrothermally Derived Nanosized Ni-Doped TiO2: A Visible Light Driven Photocatalyst for Methylene Blue Degradation. Mater. Chem. Phys. 2010, 124, 976–981. [Google Scholar] [CrossRef]
  127. Bayan, E.M.; Lupeiko, T.G.; Pustovaya, L.E.; Volkova, M.G. Synthesis and Photocatalytic Properties of Sn–TiO2 Nanomaterials. J. Adv. Dielectr. 2020, 10, 2060018. [Google Scholar] [CrossRef]
  128. Afshar, S.; Pordel, S.; Tahmouresilerd, B.; Azad, A. Improving the Photocatalytic Activity of Modified Anatase TiO2 with Different Concentrations of Aluminum under Visible Light: Mechanistic Survey. Photochem. Photobiol. 2016, 92, 783–789. [Google Scholar] [CrossRef]
  129. Luo, J.-L.; Wang, S.-F.; Liu, W.; Tian, C.-X.; Wu, J.-W.; Zu, X.-T.; Zhou, W.-L.; Yuan, X.-D.; Xiang, X. Influence of Different Aluminum Salts on the Photocatalytic Properties of Al Doped TiO2 Nanoparticles Towards the Degradation of AO7 Dye. Sci. Rep. 2017, 7, 1–16. [Google Scholar] [CrossRef]
  130. Arul, A.R.; Manjulavalli, T.E.; Venckatesh, R. Visible Light Proven Si Doped TiO2 Nanocatalyst for the Photodegradation of Organic Dye. Mater. Today Proc. 2019, 18, 1760–1769. [Google Scholar] [CrossRef]
  131. Alamelu, K.; Ali, J. TiO2-Pt Composite Photocatalyst for Photodegradation and Chemical Reduction of Recalcitrant Organic Pollutants. J. Environ. Chem. Eng. 2018, 6, 5720–5731. [Google Scholar] [CrossRef]
  132. Nguyen, C.H.; Fu, C.-C.; Juang, R.-S. Degradation of Methylene Blue and Methyl Orange by Palladium-Doped TiO2 Photocatalysis for Water Reuse: Efficiency and Degradation Pathways. J. Clean. Prod. 2018, 202, 413–427. [Google Scholar] [CrossRef]
  133. Ahadi, S.; Moalej, N.S.; Sheibani, S. Characteristics and Photocatalytic Behavior of Fe and Cu Doped TiO2 Prepared by Combined Sol-Gel and Mechanical Alloying. Solid State Sci. 2019, 96, 105975. [Google Scholar] [CrossRef]
  134. Xiao, Q.; Zhang, J.; Xiao, C.; Si, Z.; Tan, X. Solar Photocatalytic Degradation of Methylene Blue in Carbon-Doped TiO2 Nanoparticles Suspension. Sol. Energy 2008, 82, 706–713. [Google Scholar] [CrossRef]
  135. Barkul, R.P.; Patil, M.K.; Patil, S.M.; Shevale, V.B.; Delekar, S.D. Sunlight-Assisted Photocatalytic Degradation of Textile Effluent and Rhodamine B By Using Iodine Doped TiO2 Nanoparticles. J. Photochem. Photobiol. A Chem. 2017, 349, 138–147. [Google Scholar] [CrossRef]
  136. Jaimy, K.B.; Safeena, V.P.; Ghosh, S.; Hebalkar, N.Y.; Warrier, K.G.K. Photocatalytic Activity Enhancement in Doped Titanium Dioxide by Crystal Defects. Dalton Trans. 2012, 41, 4824–4832. [Google Scholar] [CrossRef] [PubMed]
  137. Cheng, J.; Chen, J.; Lin, W.; Liu, Y.; Kong, Y. Improved Visible Light Photocatalytic Activity of Fluorine and Nitrogen Co-Doped TiO2 with Tunable Nanoparticle Size. Appl. Surf. Sci. 2015, 332, 573–580. [Google Scholar] [CrossRef]
  138. Ahmed, S.; Mofijur, M.; Nuzhat, S.; Chowdhury, A.T.; Rafa, N.; Uddin, A.; Inayat, A.; Mahlia, T.; Ong, H.C.; Chia, W.Y.; et al. Recent Developments in Physical, Biological, Chemical, And Hybrid Treatment Techniques for Removing Emerging Contaminants from Wastewater. J. Hazard. Mater. 2021, 416, 125912. [Google Scholar] [CrossRef]
  139. Kasonga, T.K.; Coetzee, M.A.A.; Kamika, I.; Ngole-Jeme, V.M.; Momba, M.N.B. Endocrine-Disruptive Chemicals as Contaminants of Emerging Concern in Wastewater and Surface Water: A Review. J. Environ. Manag. 2021, 277, 111485. [Google Scholar] [CrossRef]
  140. Purba, L.D.A.; Othman, F.S.; Yuzir, A.; Mohamad, S.E.; Iwamoto, K.; Abdullah, N.; Shimizu, K.; Hermana, J. Enhanced Cultivation and Lipid Production of Isolated Microalgae Strains Using Municipal Wastewater. Environ. Technol. Innov. 2022, 27, 102444. [Google Scholar] [CrossRef]
  141. Zhou, Q.; Li, X.; Wu, S.; Zhong, Y.; Yang, C. Enhanced Strategies for Antibiotic Removal from Swine Wastewater in Anaerobic Digestion. Trends Biotechnol. 2020, 39, 8–11. [Google Scholar] [CrossRef]
  142. Valbonesi, P.; Profita, M.; Vasumini, I.; Fabbri, E. Contaminants of Emerging Concern in Drinking Water: Quality Assessment by Combining Chemical and Biological Analysis. Sci. Total Environ. 2020, 758, 143624. [Google Scholar] [CrossRef]
  143. Akhbarizadeh, R.; Dobaradaran, S.; Schmidt, T.C.; Nabipour, I.; Spitz, J. Worldwide Bottled Water Occurrence of Emerging Contaminants: A Review of the Recent Scientific Literature. J. Hazard. Mater. 2020, 392, 122271. [Google Scholar] [CrossRef] [PubMed]
  144. Pal, A.; Gin, K.Y.-H.; Lin, A.y.-C.; Reinhard, M. Impacts of Emerging Organic Contaminants on Freshwater Resources: Review of Recent Occurrences, Sources, Fate and Effects. Sci. Total Environ. 2010, 408, 6062–6069. [Google Scholar] [CrossRef] [PubMed]
  145. Jin, S.; Guerrero-Juarez, C.F.; Zhang, L.; Chang, I.; Ramos, R.; Kuan, C.-H.; Myung, P.; Plikus, M.V.; Nie, Q. Inference and Analysis of Cell-Cell Communication Using Cellchat. Nat. Commun. 2021, 12, 1–20. [Google Scholar] [CrossRef] [PubMed]
  146. Jiang, J.-J.; Lee, C.-L.; Fang, M.-D.; Tu, B.-W.; Liang, Y.-J. Impacts of Emerging Contaminants on Surrounding Aquatic Environment from a Youth Festival. Environ. Sci. Technol. 2014, 49, 792–799. [Google Scholar] [CrossRef]
  147. Saidulu, D.; Gupta, B.; Gupta, A.K.; Ghosal, P.S. A Review on Occurrences, Eco-Toxic Effects, And Remediation of Emerging Contaminants from Wastewater: Special Emphasis on Biological Treatment Based Hybrid Systems. J. Environ. Chem. Eng. 2021, 9, 105282. [Google Scholar] [CrossRef]
  148. Zhang, B.; He, X.; Yu, C.; Liu, G.; Ma, D.; Cui, C.; Yan, Q.; Zhang, Y.; Zhang, G.; Ma, J.; et al. Degradation of Tetracycline Hydrochloride by Ultrafine TiO2 Nanoparticles Modified G-C3N4 Heterojunction Photocatalyst: Influencing Factors, Products and Mechanism Insight. Chin. Chem. Lett. 2021, 33, 1337–1342. [Google Scholar] [CrossRef]
  149. Prieto-Rodriguez, L.; Miralles-Cuevas, S.; Oller, I.; Agüera, A.; Puma, G.L.; Malato, S. Treatment of Emerging Contaminants in Wastewater Treatment Plants (WWTP) Effluents by Solar Photocatalysis Using Low TiO2 Concentrations. J. Hazard. Mater. 2012, 211–212, 131–137. [Google Scholar] [CrossRef]
  150. Bhosale, M.G.; Sutar, R.S.; Londhe, S.S.; Patil, M.K. Sol–Gel Method Synthesized Ce-Doped TiO2 Visible Light Photocatalyst for Degradation of Organic Pollutants. Appl. Organomet. Chem. 2022, 36, e6586. [Google Scholar] [CrossRef]
  151. Šojić, D.; Despotović, V.; Abramović, B.; Todorova, N.; Giannakopoulou, T.; Trapalis, C. Photocatalytic Degradation of Mecoprop and Clopyralid in Aqueous Suspensions of Nanostructured N-doped TiO2. Molecules 2010, 15, 2994–3009. [Google Scholar] [CrossRef]
  152. Kusumawardani, L.J.; Syahputri, Y.; Iryani, A. Photocatalytic Degradation of Paraquat Dichloride using TiO2-Fe Nano Powder under Visible and Sunlight Irradiation. J. Kim. Val. 2020, 6, 55–61. [Google Scholar] [CrossRef]
  153. Quiroz, N.A.; Gutierrez, D.J.R.; Martínez, S.S.; Bahena, C.L. Degradation of Gesaprim Herbicide by Heterogeneous Photocatalysis Using Fe-Doped TiO2. Int. J. Geosci. 2011, 2, 669–675. [Google Scholar] [CrossRef]
  154. Senthilnathan, J.; Philip, L. Photocatalytic Degradation of Lindane Under UV and Visible Light Using N-Doped TiO2. Chem. Eng. J. 2010, 161, 83–92. [Google Scholar] [CrossRef]
  155. Manoharan, R.K.; Sankaran, S. Photocatalytic Degradation of Organic Pollutant Aldicarb by Non-Metal-Doped Nanotitania: Synthesis and Characterization. Environ. Sci. Pollut. Res. 2017, 25, 20510–20517. [Google Scholar] [CrossRef]
  156. Aba-Guevara, C.G.; Medina-Ramírez, I.E.; Hernández-Ramírez, A.; Jáuregui-Rincón, J.; Lozano-Álvarez, J.A.; Rodríguez-López, J.L. Comparison of Two Synthesis Methods on the Preparation of Fe, N-Co-Doped TiO2 Materials For Degradation of Pharmaceutical Compounds Under Visible Light. Ceram. Int. 2017, 43, 5068–5079. [Google Scholar] [CrossRef]
  157. Ortiz-Bustos, J.; Gómez-Ruiz, S.; Mazarío, J.; Domine, M.E.; del Hierro, I.; Pérez, Y. Copper and Sulphur Co-Doped Titanium Oxide Nanoparticles with Enhanced Catalytic and Photocatalytic Properties. Catal. Sci. Technol. 2020, 10, 6511–6524. [Google Scholar] [CrossRef]
  158. Mancuso, A.; Navarra, W.; Sacco, O.; Pragliola, S.; Vaiano, V.; Venditto, V. Photocatalytic Degradation of Thiacloprid Using Tri-Doped TiO2 Photocatalysts: A Preliminary Comparative Study. Catalysts 2021, 11, 927. [Google Scholar] [CrossRef]
  159. Joseph, A.I.J.; Thiripuranthagan, S. Non-Metal Doped Titania Photocatalysts for the Degradation of Neonicotinoid Insecticides Under Visible Light Irradiation. J. Nanosci. Nanotechnol. 2018, 18, 3158–3164. [Google Scholar] [CrossRef]
  160. López-Gámez, G.; Elez-Martínez, P.; Martín-Belloso, O.; Soliva-Fortuny, R. Pulsed Electric Fields Affect Endogenous Enzyme Activities, Respiration and Biosynthesis of Phenolic Compounds in Carrots. Postharvest Biol. Technol. 2020, 168, 111284. [Google Scholar] [CrossRef]
  161. Hays, M.D.; Fine, P.M.; Geron, C.D.; Kleeman, M.J.; Gullett, B.K. Open Burning of Agricultural Biomass: Physical and Chemical Properties of Particle-Phase Emissions. Atmos. Environ. 2005, 39, 6747–6764. [Google Scholar] [CrossRef]
  162. Batura, I.I.; Gogotov, A.F.; Cherepanov, V.I.; Baranov, O.I.; Levchuk, A.A.; Parilova, M.V. Utilizing the Phenol Byproducts of Coke Production: 3. Phenols as Coinhibitors of Thermopolymerization During Styrene Production. Coke Chem. 2009, 52, 23–27. [Google Scholar] [CrossRef]
  163. Villegas, L.G.C.; Mashhadi, N.; Chen, M.; Mukherjee, D.; Taylor, K.E.; Biswas, N. A Short Review of Techniques for Phenol Removal from Wastewater. Curr. Pollut. Rep. 2016, 2, 157–167. [Google Scholar] [CrossRef] [Green Version]
  164. Pedroza, A.M.; Mosqueda, R.; Alonso-Vante, N.; Rodriguez-Vazquez, R. Sequential Treatment Via Trametes Versicolor And UV/TiO2/Ruxsey to Reduce Contaminants in Wastewater Resulting from the Bleaching Process During Paper Production. Chemosphere 2007, 67, 793–801. [Google Scholar] [CrossRef] [PubMed]
  165. She, Z.; Gao, M.; Jin, C.; Chen, Y.; Yu, J. Toxicity and Biodegradation of 2,4-Dinitrophenol and 3-Nitrophenol in Anaerobic Systems. Process Biochem. 2005, 40, 3017–3024. [Google Scholar] [CrossRef]
  166. Gopi, P.K.; Srinithi, S.; Chen, S.-M.; Ravikumar, C.H. Designing of Cerium-Doped Bismuth Vanadate Nanorods/Functionalized-MWCNT Nanocomposite for the High Toxicity of 4-Cyanophenol Herbicide Detection in Human Urine Sample. Colloids Surf. A Physicochem. Eng. Asp. 2022, 639, 128371. [Google Scholar] [CrossRef]
  167. Tyler, C.R.; Jobling, S.; Sumpter, J.P. Endocrine Disruption in Wildlife: A Critical Review of the Evidence. Crit. Rev. Toxicol. 1998, 28, 319–361. [Google Scholar] [CrossRef]
  168. Karci, A.; Arslan-Alaton, I.; Olmez-Hanci, T.; Bekbölet, M. Transformation of 2, 4-Dichlorophenol by H2O2/UV-C, Fenton and Photo-Fenton Processes: Oxidation Products and Toxicity Evolution. J. Photochem. Photobiol. A Chem. 2012, 230, 65–73. [Google Scholar] [CrossRef]
  169. Niu, Y.; Xing, M.; Zhang, J.; Tian, B. Visible Light Activated Sulfur and Iron Co-Doped TiO2 Photocatalyst for the Photocatalytic Degradation of Phenol. Catal. Today 2013, 201, 159–166. [Google Scholar] [CrossRef]
  170. Xu, J.; Wang, F.; Liu, W.; Cao, W. Nanocrystalline N-Doped Powders: Mild Hydrothermal Synthesis and Photocatalytic Degradation of Phenol Under Visible Light Irradiation. Int. J. Photoenergy 2013, 2013, 1–8. [Google Scholar] [CrossRef]
  171. Chen, X.; Kuo, D.-H.; Lu, D. N-Doped Mesoporous TiO2 Nanoparticles Synthesized by Using Biological Renewable Nanocrystalline Cellulose as Template for The Degradation of Pollutants Under Visible and Sun Light. Chem. Eng. J. 2016, 295, 192–200. [Google Scholar] [CrossRef]
  172. Wang, Z.; Cai, W.; Hong, X.; Zhao, X.; Xu, F.; Cai, C. Photocatalytic Degradation of Phenol in Aqueous Nitrogen-Doped TiO2 Suspensions with Various Light Sources. Appl. Catal. B Environ. 2005, 57, 223–231. [Google Scholar] [CrossRef]
  173. Deng, B.; Fu, S.; Zhang, Y.; Wang, Y.; Ma, D.; Dong, S. Simultaneous Pollutant Degradation and Power Generation in Visible-Light Responsive Photocatalytic Fuel Cell with An Ag-TiO2 Loaded Photoanode. Nano-Struct. Nano-Objects 2018, 15, 167–172. [Google Scholar] [CrossRef]
  174. Borji, S.H.; Nasseri, S.; Mahvi, A.H.; Nabizadeh, R.; Javadi, A.H. Investigation of Photocatalytic Degradation of Phenol by Fe(III)-Doped TiO2 And TiO2 Nanoparticles. J. Environ. Health Sci. Eng. 2014, 12, 101. [Google Scholar] [CrossRef] [PubMed]
  175. Huang, D.; Liao, S.; Zhou, W.; Quan, S.; Liu, L.; He, Z.; Wan, J. Synthesis of Samarium- and Nitrogen-Co-Doped TiO2 by Modified Hydrothermal Method and Its Photocatalytic Performance for the Degradation of 4-Chlorophenol. J. Phys. Chem. Solids 2009, 70, 853–859. [Google Scholar] [CrossRef]
  176. Te Lin, J.C.; Tolosa, N.C.; Sopajaree, K.; Gongglom, A.; Jitjanesuwan, T.; Lu, M.C. Visible-Light Degradation of 2-Chlorophenol by TiO2 doped with Neighboring Transition Metal Cerium via Various Approaches and Operation Parameters. Research Squar 2021. Available online: https://www.researchsquare.com/article/rs-508006/latest.pdf (accessed on 14 July 2022).
  177. Benkhennouche-Bouchene, H.; Mahy, J.; Wolfs, C.; Vertruyen, B.; Poelman, D.; Eloy, P.; Hermans, S.; Bouhali, M.; Souici, A.; Bourouina-Bacha, S.; et al. Green Synthesis of N/Zr Co-Doped TiO2 for Photocatalytic Degradation of p-Nitrophenol in Wastewater. Catalysts 2021, 11, 235. [Google Scholar] [CrossRef]
  178. Choi, J.; Park, H.; Hoffmann, M.R. Combinatorial Doping of TiO2 with Platinum (Pt), Chromium (Cr), Vanadium (V), and Nickel (Ni) To Achieve Enhanced Photocatalytic Activity with Visible Light Irradiation. J. Mater. Res. 2010, 25, 149–158. [Google Scholar] [CrossRef]
  179. Tseng, T.K.; Lin, Y.S.; Chen, Y.J.; Chu, H. A Review of Photocatalysts Prepared by Sol-Gel Method for VOCs Removal. Int. J. Mol. Sci. 2010, 11, 2336–2361. [Google Scholar] [CrossRef] [PubMed]
  180. 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]
  181. Kanakaraju, D.; Kutiang, F.D.A.; Lim, Y.C.; Goh, P.S. Recent Progress of Ag/TiO2 Photocatalyst for Wastewater Treatment: Doping, Co-Doping, and Green Materials Functionalization. Appl. Mater. Today 2022, 27, 101500. [Google Scholar] [CrossRef]
  182. Kaur, N.; Singh, M.; Moumen, A.; Duina, G.; Comini, E. 1D Titanium Dioxide: Achievements in Chemical Sensing. Materials 2020, 13, 2974. [Google Scholar] [CrossRef]
  183. Vaiano, V.; Lara, M.A.; Iervolino, G.; Matarangolo, M.; Navío, J.A.; Hidalgo, M.C. Photocatalytic H2 Production from Glycerol Aqueous Solutions Over Fluorinated Pt-TiO2 with High {001} Facet Exposure. J. Photochem. Photobiol. A Chem. 2018, 365, 52–59. [Google Scholar] [CrossRef]
  184. Iervolino, G.; Vaiano, V.; Sannino, D.; Puga, F.; Navío, J.A.; Hidalgo, M.C. LaFeO3 Modified with Ni for Hydrogen Evolution via Photocatalytic Glucose Reforming in Liquid Phase. Catalysts 2021, 11, 1558. [Google Scholar] [CrossRef]
  185. Qian, W.; Greaney, P.A.; Fowler, S.; Chiu, S.-K.; Goforth, A.M.; Jiao, J. Low-Temperature Nitrogen Doping in Ammonia Solution for Production of N-Doped TiO2-Hybridized Graphene as a Highly Efficient Photocatalyst for Water Treatment. ACS Sustain. Chem. Eng. 2014, 2, 1802–1810. [Google Scholar] [CrossRef]
  186. Nadeem, M.A.; Idriss, H. Photo-Thermal Reactions of Ethanol Over Ag/TiO2 Catalysts. The Role of Silver Plasmon Resonance in the Reaction Kinetics. Chem. Commun. 2018, 54, 5197–5200. [Google Scholar] [CrossRef] [PubMed]
  187. Liu, X.; Xing, Z.; Zhang, Y.; Li, Z.; Wu, X.; Tan, S.; Yu, X.; Zhu, Q.; Zhou, W. Fabrication of 3D Flower-Like Black N-TiO2-X@ Mos2 for Unprecedented-High Visible-Light-Driven Photocatalytic Performance. Appl. Catal. B Environ. 2017, 201, 119–127. [Google Scholar] [CrossRef]
  188. Wang, W.; Zhao, H.; Pan, P.; Xue, K.; Zhang, Z.; Duan, J. Enhanced the Photocatalytic Activity of B–C–N–TiO2 Under Visible Light:Synergistic Effect of Element Doping and Z-Scheme Interface Heterojunction Constructed with Ag Nanoparticles. Ceram. Int. 2020, 47, 6094–6104. [Google Scholar] [CrossRef]
  189. Soni, V.; Raizada, P.; Kumar, A.; Hasija, V.; Singal, S.; Singh, P.; Hosseini-Bandegharaei, A.; Thakur, V.K.; Nguyen, V.-H. Indium Sulfide-Based Photocatalysts for Hydrogen Production and Water Cleaning: A Review. Environ. Chem. Lett. 2021, 19, 1065–1095. [Google Scholar] [CrossRef]
  190. He, J.; Kumar, A.; Khan, M.; Lo, I.M. Critical Review of Photocatalytic Disinfection of Bacteria: From Noble Metals-And Carbon Nanomaterials-TiO2 Composites to Challenges of Water Characteristics and Strategic Solutions. Sci. Total Environ. 2021, 758, 143953. [Google Scholar] [CrossRef]
  191. Jeyalakshmi, V.; Mahalakshmy, R.; Krishnamurthy, K.; Viswanathan, B. Photocatalytic Reduction of Carbon Dioxide by Water: A Step towards Sustainable Fuels and Chemicals. Mater. Sci. Forum 2012, 734, 1–62. [Google Scholar]
  192. Kunnamareddy, M.; Diravidamani, B.; Rajendran, R.; Singaram, B.; Varadharajan, K. Synthesis of Silver and Sulphur Codoped TiO2 Nanoparticles for Photocatalytic Degradation of Methylene Blue. J. Mater. Sci. Mater. Electron. 2018, 29, 18111–18119. [Google Scholar] [CrossRef]
  193. Chadwick, S.S. Ullmann’s Encyclopedia of Industrial Chemistry. Ref. Serv. Rev. 1988, 16, 31–34. [Google Scholar] [CrossRef]
  194. Steinfeld, A. Solar Thermochemical Production of Hydrogen—A Review. Sol. Energy 2005, 78, 603–615. [Google Scholar] [CrossRef]
  195. Crespo-Quesada, M.; Reisner, E. Emerging Approaches to Stabilise Photocorrodible Electrodes and Catalysts for Solar Fuel Applications. Energy Environ. Sci. 2017, 10, 1116–1127. [Google Scholar] [CrossRef] [Green Version]
  196. Wang, Z.; Roberts, R.; Naterer, G.; Gabriel, K. Comparison of Thermochemical, Electrolytic, Photoelectrolytic and Photochemical Solar-To-Hydrogen Production Technologies. Int. J. Hydrogen Energy 2012, 37, 16287–16301. [Google Scholar] [CrossRef]
  197. Clarizia, L.; Russo, D.; Di Somma, I.; Andreozzi, R.; Marotta, R. Hydrogen Generation through Solar Photocatalytic Processes: A Review of the Configuration and the Properties of Effective Metal-Based Semiconductor Nanomaterials. Energies 2017, 10, 1624. [Google Scholar] [CrossRef]
  198. Kitano, M.; Hara, M. Heterogeneous Photocatalytic Cleavage of Water. J. Mater. Chem. 2009, 20, 627–641. [Google Scholar] [CrossRef]
  199. Hu, J.; Cao, Y.; Wang, K.; Jia, D. Green Solid-State Synthesis and Photocatalytic Hydrogen Production Activity of Anatase TiO2 Nanoplates with Super Heat-Stability. RSC Adv. 2017, 7, 11827–11833. [Google Scholar] [CrossRef]
  200. 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]
  201. Liu, Y.; Yu, H.; Quan, X.; Chen, S. Green synthesis of Feather-Shaped Mos2/Cds Photocatalyst for Effective Hydrogen Production. Int. J. Photoenergy 2013, 2013, 247516. [Google Scholar] [CrossRef]
  202. Chen, G.; Li, D.; Li, F.; Fan, Y.; Zhao, H.; Luo, Y.; Yu, R.; Meng, Q. Ball-Milling Combined Calcination Synthesis of Mos2/Cds Photocatalysts for High Photocatalytic H2 Evolution Activity Under Visible Light Irradiation. Appl. Catal. A Gen. 2012, 443, 138–144. [Google Scholar] [CrossRef]
  203. Serpone, N.; Pelizzetti, E. Photocatalysis: Fundamentals and Applications; Wiley-Interscience: Hoboken, NJ, USA, 1989. [Google Scholar]
  204. Al–Ekabi, H.; Safarzadeh–Amiri, A.; Sifton, W.; Story, J. Advanced Technology for Water Purification by Heterogeneous Photocatalysis. Int. J. Environ. Pollut. 1991, 1, 125–136. [Google Scholar]
  205. Ilisz, I.; Dombi, A.; Mogyorósi, K.; Dékány, I. Photocatalytic Water Treatment with Different TiO2 Nanoparticles and Hydrophilic/Hydrophobic Layer Silicate Adsorbents. Colloids Surf. A Physicochem. Eng. Asp. 2003, 230, 89–97. [Google Scholar] [CrossRef]
  206. Reddy, E.P.; Davydov, L.; Smirniotis, P. TiO2-Loaded Zeolites and Mesoporous Materials in the Sonophotocatalytic Decomposition of Aqueous Organic Pollutants: The Role of the Support. Appl. Catal. B Environ. 2003, 42, 1–11. [Google Scholar] [CrossRef]
  207. Iervolino, G.; Sacco, O.; Vaiano, V.; Palma, V. Non-Thermal Plasma Technology for the Effective Regeneration of Macroscopic Adsorbent Materials Used in the Removal of Patent Blue V Dye from Aqueous Solutions. Chem. Eng. Trans. 2019, 73, 151–156. [Google Scholar] [CrossRef]
  208. Iervolino, G.; Vaiano, V.; Sannino, D.; Rizzo, L.; Galluzzi, A.; Polichetti, M.; Pepe, G.; Campiglia, P. Hydrogen Production from Glucose Degradation in Water and Wastewater Treated by Ru-Lafeo3/Fe2O3 Magnetic Particles Photocatalysis and Heterogeneous Photo-Fenton. Int. J. Hydrogen Energy 2018, 43, 2184–2196. [Google Scholar] [CrossRef]
  209. Liu, S.-Q. Magnetic Nano-Photocatalysts: Preparation, Structure, and Application. In Environmental Chemistry for a Sustainable World; Springer: Berlin/Heidelberg, Germany, 2012; pp. 99–117. [Google Scholar]
  210. Moon, H.S.; Yong, K. Noble-Metal Free Photocatalytic Hydrogen Generation of Cupc/TiO2 Nanoparticles Under Visible-Light Irradiation. Appl. Surf. Sci. 2020, 530, 147215. [Google Scholar] [CrossRef]
  211. Chen, W.-T.; Chan, A.; Sun-Waterhouse, D.; Moriga, T.; Idriss, H.; Waterhouse, G.I. Ni/TiO2: A Promising Low-Cost Photocatalytic System for Solar H2 Production from Ethanol–Water Mixtures. J. Catal. 2015, 326, 43–53. [Google Scholar] [CrossRef]
  212. Yin, M.; Ma, S.; Wu, C.; Fan, Y. A Noble-Metal-Free Photocatalytic Hydrogen Production System Based on Cobalt(iii) Complex And Eosin Y-Sensitized TiO2. RSC Adv. 2014, 5, 1852–1858. [Google Scholar] [CrossRef]
  213. Lin, X.; Wang, J. Green Synthesis of Well Dispersed TiO2/Pt Nanoparticles Photocatalysts and Enhanced Photocatalytic Activity Towards Hydrogen Production. Int. J. Hydrogen Energy 2019, 44, 31853–31859. [Google Scholar] [CrossRef]
  214. Wang, X.; Cao, L.; Chen, D.; Caruso, R.A. Engineering of Monodisperse Mesoporous Titania Beads for Photocatalytic Applications. ACS Appl. Mater. Interfaces 2013, 5, 9421–9428. [Google Scholar] [CrossRef]
  215. Leung, D.Y.C.; Fu, X.; Wang, C.; Ni, M.; Leung, M.K.H.; Wang, X.; Fu, X. Hydrogen Production over Titania-Based Photocatalysts. ChemSusChem 2010, 3, 681–694. [Google Scholar] [CrossRef]
  216. Cai, J.; Kimura, S.; Wada, M.; Kuga, S. Nanoporous Cellulose as Metal Nanoparticles Support. Biomacromolecules 2008, 10, 87–94. [Google Scholar] [CrossRef]
  217. Lin, X.; Wu, M.; Wu, D.; Kuga, S.; Endo, T.; Huang, Y. Platinum Nanoparticles Using Wood Nanomaterials: Eco-Friendly Synthesis, Shape Control and Catalytic Activity for P-Nitrophenol Reduction. Green Chem. 2010, 13, 283–287. [Google Scholar] [CrossRef]
  218. Jing, D.; Guo, L. A Novel Method for the Preparation of a Highly Stable and Active CdS Photocatalyst with a Special Surface Nanostructure. J. Phys. Chem. B 2006, 110, 11139–11145. [Google Scholar] [CrossRef] [PubMed]
  219. Meissner, D.; Memming, R.; Kastening, B. Photoelectrochemistry of Cadmium Sulfide. 1. Reanalysis of Photocorrosion and Flat-Band Potential. J. Phys. Chem. 1988, 92, 3476–3483. [Google Scholar] [CrossRef]
  220. Chen, Y.; Wang, L.; Lu, G.; Yao, X.; Guo, L. Nanoparticles Enwrapped with Nanotubes: A Unique Architecture of Cds/Titanate Nanotubes for Efficient Photocatalytic Hydrogen Production from Water. J. Mater. Chem. 2011, 21, 5134–5141. [Google Scholar] [CrossRef]
  221. Garcia-Esparza, A.T.; Cha, D.; Ou, Y.; Kubota, J.; Domen, K.; Takanabe, K. Tungsten Carbide Nanoparticles as Efficient Cocatalysts for Photocatalytic Overall Water Splitting. ChemSusChem 2012, 6, 168–181. [Google Scholar] [CrossRef] [PubMed]
  222. Xiang, Q.; Yu, J.; Jaroniec, M. Synergetic Effect of Mos2 and Graphene as Cocatalysts for Enhanced Photocatalytic H2 Production Activity of TiO2 Nanoparticles. J. Am. Chem. Soc. 2012, 134, 6575–6578. [Google Scholar] [CrossRef]
  223. Niu, B.; Chen, Z.; Xu, Z. Recycling Waste Tantalum Capacitors to Synthesize High Value-Added Ta2O5 and Polyaniline-Decorated Ta2O5 Photocatalyst by an Integrated Chlorination-Sintering-Chemisorption Process. J. Clean. Prod. 2019, 252, 117206. [Google Scholar] [CrossRef]
  224. Gao, Y.; Chen, B.; Li, H.; Ma, Y. Preparation and Characterization of a Magnetically Separated Photocatalyst and Its Catalytic Properties. Mater. Chem. Phys. 2003, 80, 348–355. [Google Scholar] [CrossRef]
  225. Vaiano, V.; Sacco, O.; Sannino, D.; Stoller, M.; Ciambelli, P.; Chianese, A. Photocatalytic Removal of Phenol by Ferromagnetic N-TiO2/SiO2/Fe3O4 Nanoparticles in Presence of Visible Light Irradiation. Chem. Eng. Trans. 2016, 47, 235–240. [Google Scholar] [CrossRef]
  226. Vaiano, V.; Iervolino, G.; Sannino, D.; Rizzo, L.; Sarno, G. MoOx/TiO2 Immobilized on Quartz Support as Structured Catalyst for the Photocatalytic Oxidation of as (III) to as (V) in Aqueous Solutions. Chem. Eng. Res. Des. 2016, 109, 190–199. [Google Scholar] [CrossRef]
  227. Vaiano, V.; Chianese, L.; Rizzo, L.; Iervolino, G. Visible Light Driven Oxidation of Arsenite to Arsenate in Aqueous Solution Using Cu-Doped Zno Supported on Polystyrene Pellets. Catal. Today 2020, 361, 69–76. [Google Scholar] [CrossRef]
  228. Vaiano, V.; Iervolino, G. Facile Method to Immobilize Zno Particles on Glass Spheres for the Photocatalytic Treatment of Tannery Wastewater. J. Colloid Interface Sci. 2018, 518, 192–199. [Google Scholar] [CrossRef] [PubMed]
Table 1. Doped and codoped TiO2 based catalyst.
Table 1. Doped and codoped TiO2 based catalyst.
CatalystDoping or Codoping ElementLight SourceApplication
Fe-TiO2FevisiblePara-niro phenol and methylene blue degradation
Ni-TiO2NisolarIbuprofen degradation
Bi-TiO2BisolarIbuprofen degradation
Cu-TiO2CuvisibleRed 198 dye removal
Pr-TiO2(RE Element): PrvisibleRhB and Crystal violet dyes removal
N-TiO2Non metal: NsolarDyes removal
S-TiO2Non metal: SvisibleMethyl orange and phenol removal
Fe-Cu-TiO2Fe and CuvisibleMethylene blue removal
Fe-Pr-TiO2Fe and PrvisibleAO7 dye removal
Cu-N-TiO2Fe and CuvisiblePyrene removal
Table 2. List of doped and codoped catalysts used on visible light degradation of dye contaminant.
Table 2. List of doped and codoped catalysts used on visible light degradation of dye contaminant.
CatalystPreparation MethodTreatment Time
(min)
Target DyeDye
Concentration (ppm)
Degradation
Efficiency
(%)
Refs.
Fe-TiO2sol gel60methylene blue3093[123]
Mn-TiO2hydrothermal180methylene blue1097[124]
Mn-TiO2hydrothermal300methylene blue16075[125]
Ni-TiO2hydrothermal180methylene blue598[126]
Ni-TiO2hydrothermal60methylene blue1093[26]
Sn-TiO2sol gel120methylene blue2077[127]
Al-TiO2sol gel180rhodamine B1089[128]
Al-TiO2sol gel60Acid orange 7599[129]
Si-TiO2sol gel120methylene blue1087[130]
Pt-TiO2Hydrothermal90rhodamine B1099.5[131]
Pd-TiO2sol gel120methylene blue2099.4[132]
Fe-TiO2sol gel60acid orange 71073[77].
Fe-TiO2sol gel and mechanical alloying240methylene blue278[133]
Pr-TiO2sol gel60acid orange 710-[81]
C-TiO2sol gel100methylene blue10100[134]
N-TiO2sol gel60acid orange 71055[77]
I-TiO2sol gel240rhodamine B2095[135]
S-TiO2hydrothermal70rhodamine B1095[57]
C-N-TiO2solvothermal90Rhodamine B2094[86]
Fe-N-TiO2sol gel60Acid orange 71083[77]
Fe-Pr-TiO2sol gel60Acid orange 71080[81]
Fe-Ce-TiO2sol gel120methylene blue486[136]
Fe-La-TiO2sol gel180methylene blue0.144[65]
F-N-TiO2hydrothermal300methylene blue597[137]
Table 3. List of doped and codoped catalysts used on visible light degradation of ECs.
Table 3. List of doped and codoped catalysts used on visible light degradation of ECs.
CatalystPreparation MethodTreatment Time
(min)
Target ECEC Concentration (ppm)Degradation Efficiency
(%)
Refs.
Ce-TiO2sol gel60tetracycline2093[150]
Bi-TiO2sol gel360Ibuprofen2589[29]
Ni-TiO2sol gel360Ibuprofen2550[29]
N-TiO2sol gel1200mecoprop58050[151]
Fe-TiO2sol gel75Paraquat dichloride3098[152]
Fe-TiO2sol gel200Gesaprim Herbicide4090[153]
Mn-TiO2hydrothermal300glyphosate675[154]
N-TiO2sol gel330lindane100100[154]
Cr-TiO2sol gel330lindane10030[154]
Ag-TiO2sol gel330lindane10023[154]
Fe-TiO2sol gel330lindane10010[154]
Fe-Cr-TiO2sol gel330lindane10037[154]
N-S-TiO2sol gel90aldicarb pesticide19>80%[155]
Fe-N-TiO2sol gel240amoxicillin3059[156]
Fe-N-TiO2sol gel240streptomycin3050[156]
Cu-S-TiO2sol gel90ciprofloxacin15100[157]
Fe-N-P-TiO2sol gel180thiachloprid0.529[158]
C-N-B-F-TiO2sol gel480thiamethoxam87.565[159]
C-N-B-F-TiO2sol gel450imidacloprid2471[159]
Table 4. List of doped and codoped catalysts used on visible light degradation of phenolic compounds.
Table 4. List of doped and codoped catalysts used on visible light degradation of phenolic compounds.
CatalystPreparation MethodTreatment Time
(min)
Target Phenolic CompoundPhenolic Compound
(ppm)
Degradation Efficiency
(%)
Refs.
N-TiO2hydrolysis precipitation120phenol50100[171]
N-TiO2hydrothermal120phenol2036[172]
N-TiO2hydrothermal600phenol2581[170]
Ag-TiO2sol gel impregnation3604-chlorophenol1073.5[173]
Fe-TiO2sol gel210phenol1030[174]
N-TiO2hydrothermal6004-chlorophenol13017[99]
Sm-TiO2hydrothermal6004-chlorophenol1301.5[175]
Ce-TiO2sol gel2402-chlorophenol10100[176]
Fe -TiO2sol-gel and mechanical alloying240phenol540[133]
Sm–N-TiO2hydrothermal6004-chlorophenol13040[175]
N-Zr-TiO2sol–gel14004-nitrophenol1468[177]
Cr-V-TiO2sol gel180phenol550[178]
Pt(II)-Cr-TiO2sol gel180phenol570[178]
Fe-Pr-TiO2sol gel240phenol1066[81]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mancuso, A.; Iervolino, G. Synthesis and Application of Innovative and Environmentally Friendly Photocatalysts: A Review. Catalysts 2022, 12, 1074. https://doi.org/10.3390/catal12101074

AMA Style

Mancuso A, Iervolino G. Synthesis and Application of Innovative and Environmentally Friendly Photocatalysts: A Review. Catalysts. 2022; 12(10):1074. https://doi.org/10.3390/catal12101074

Chicago/Turabian Style

Mancuso, Antonietta, and Giuseppina Iervolino. 2022. "Synthesis and Application of Innovative and Environmentally Friendly Photocatalysts: A Review" Catalysts 12, no. 10: 1074. https://doi.org/10.3390/catal12101074

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