Open access peer-reviewed chapter

Modified Titanium Dioxide for Photocatalytic Applications

Written By

John Moma and Jeffrey Baloyi

Submitted: 12 March 2018 Reviewed: 07 June 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.79374

From the Edited Volume

Photocatalysts - Applications and Attributes

Edited by Sher Bahadar Khan and Kalsoom Akhtar

Chapter metrics overview

4,743 Chapter Downloads

View Full Metrics

Abstract

Titanium dioxide (TiO2) has been widely used as a photocatalyst in many environmental and energy applications due to its efficient photoactivity, high stability, low cost, and safety to the environment and humans. However, its large band gap energy, ca. 3.2 eV limits its absorption of solar radiation to the UV light range which accounts for only about 5% of the solar spectrum. Furthermore, the photocatalytic activity of TiO2 is also limited by the rapid recombination of the photogenerated electron-hole pairs. When used in water treatment applications, TiO2 has a poor affinity toward organic pollutants, especially hydrophobic organic pollutants. Several strategies have been employed to reduce its band gap energy, its electron-hole recombination rates as well as enhance its absorption of organic pollutants. In this chapter, we review some of the most recent works that have employed the doping, decoration, and structural modification of TiO2 particles for applications in photocatalysis. Additionally, we discuss the effectiveness of these dopants and/or modifiers in enhancing TiO2 photoactivity as well as some perspective on the future of TiO2 photocatalysis.

Keywords

  • titanium dioxide
  • photocatalysis
  • environmental pollution
  • modification

1. Introduction

The rapid growth of global population as well as industrialization has led to a concomitant increase in environmental pollution. This has very negative effects on natural elements that are vital for life on earth such as air and water. It becomes very crucial therefore to find sustainable ways to mitigate pollution in order to provide a clean and safe environment for humans. Photocatalysis has attracted worldwide interest due to its potential to use solar energy not only to solve environmental problems but also provide a renewable and sustainable energy source. An efficient photocatalyst converts solar energy into chemical energy which can be used for environmental and energy applications such as water treatment, air purification, self-cleaning surfaces, hydrogen production by water cleavage and CO2 conversion to hydrocarbon fuels.

Research in the development of efficient photocatalytic materials has seen significant progress in the last 2 decades with a large number of research papers published every year. Improvements in the performance of photocatalytic materials have been largely correlated with advances in nanotechnology. Of many materials that have been studied for photocatalysis, titanium dioxide (TiO2; titania) has been extensively researched because it possesses may merits such as high photocatalytic activity, excellent physical and chemical stability, low cost, non-corrosive, nontoxicity and high availability [1, 2, 3, 4]. The photocatalytic activity of titania depends on its phase. It exists in three crystalline phases; the anatase, rutile and brookite. The anatase phase is metastable and has a higher photocatalytic activity, while the rutile phase is more chemically stable but less active. Some titania with a mixture of both anatase and rutile phases exhibit higher activities compared to pure anatase and rutile phases [5, 6, 7]. When titania is irradiated with light of sufficient energy, electrons from the valence band are promoted to the conduction band, leaving an electron deficiency or hole, h+, in the valence band and an excess of negative charge in the conduction band. The free electrons in the conduction band are good reducing agents while the resultant holes in the valence band are strong oxidizing agents and can both participate in redox reactions.

Titania however suffers from a number of drawbacks that limit its practical applications in photocatalysis. Firstly, the photogenerated electrons and holes coexist in the titania particle and the probability of their recombination is high. This leads to low rates of the desired chemical transformations with respect to the absorbed light energy [8, 9]. The relatively large band gap energy (~ 3.2 eV) requires ultraviolet light for photoactivation, resulting in a very low efficiency in utilizing solar light. UV light accounts for only about 5% of the solar spectrum compared to visible light (45%) [1, 10]. In addition to these, because titania is non-porous and has a polar surface, it exhibits low absorption ability for non-polar organic pollutants [10, 11, 12, 13]. There is also the challenge to recover nano-sized titania particles from treated water in regards to both economic and safety concern [14]. The TiO2 nanoparticles also suffer from aggregation and agglomeration which affect the photoactivity as well as light absorption [15, 16, 17, 18]. Several strategies have been employed in the open literature to overcome these drawbacks. These strategies aim at extending the wavelength of photoactivation of TiO2 into the visible region of the spectrum thereby increasing the utilization of solar energy; preventing the electron/hole pair recombination and thus allowing more charge carriers to successfully diffuse to the surface; increasing the absorption affinity of TiO2 towards organic pollutants as well as preventing the aggregation and agglomeration of the nano-titania particles while easing their recovery from treated water. Several reviews have been published in recent years on the development of strategies to eliminate the limitations of titania photocatalysis [1, 19, 20, 21, 22, 23, 24, 25]. Most of these however focus on pollutant removal from wastewater, water splitting for hydrogen production, CO2 conversion and reaction mechanisms [1, 21, 25, 26, 27, 28, 29, 30, 31]. In this chapter, we review some of the latest publications mainly covering the last 5 years, on strategies that have been researched to overcome the limitations of TiO2 for general photocatalytic applications and the level of success that these strategies have been able to achieve. Based on the current level of research in this field, we also present some perspectives on the future of modified TiO2 photocatalysis.

Advertisement

2. Modification of TiO2 photocatalysts

A large number of research works have been published on TiO2 modification to enhance its photocatalytic properties. These modifications have been done in many different ways which include metal and non-metal doping, dye sensitization, surface modification, fabrication of composites with other materials and immobilization and stabilization on support structures. The properties of modified TiO2 are always intrinsically different from the pure TiO2 with regards to light absorption, charge separation, adsorption of organic pollutants, stabilization of the TiO2 particles and ease of separation of TiO2 particles.

2.1. Metal doping

Metal doping has been extensively used to advance efforts at developing modified TiO2 photocatalysts to operate efficiently under visible light. The photoactivity of metal-doped TiO2 photocatalysts depends to a large extent on the nature of the dopant ion and its nature, its level, the method used in the doping, the type of TiO2 used as well as the reaction for which the catalyst is used and the reaction conditions [32]. The mechanism of the lowering of the band gap energy of TiO2 with metal doping is shown in Figure 1. It is believed that doping TiO2 with metals results in an overlap of the Ti 3d orbitals with the d levels of the metals causing a shift in the absorption spectrum to longer wavelengths which in turn favours the use of visible light to photoactivate the TiO2.

Figure 1.

Band-gap lowering mechanism of metal-doped TiO2.

Doping of TiO2 nanoparticles with Li, Na, Mg, Fe and Co by high energy ball milling with the metal nitrates was found to widen the TiO2 visible light response range. In the Na-doped sample, Ti existed as both Ti4+ and Ti3+ and the conversion between Ti4+ and Ti3+ was found to prevent the recombination of electrons and (e) and holes (h+). The metal ion doping promoted crystal phase transformations that generated electrons (e) and holes (h+) [33]. Mesoporous TiO2 prepared by sol gel technique and doped with different levels of Pt (1–5 wt% nominal loading) resulted in a high surface area TiO2 with an enhanced catalytic performance in photocatalytic water splitting for the Pt-doped samples. The 2.5 wt%Pt-TiO2 had showed the optimum catalytic performance and a reduction in the TiO2 band gap energy from 3.00 to 2.34 eV with an enhanced electron storage capacity, leading to a minimization of the electron-hole recombination rate [34]. Noble metal nanoparticles such as Ag [35], Pt [34], Pd [36], Rh [37] and Au [38] have also been used to modify TiO2 for photocatalysis and have been reported to efficiently hinder electron-hole recombination due to the resulting Schottky barrier at the metal-TiO2 interface. The noble metal nanoparticles act as a mediator in storing and transporting photogenerated electrons from the surface of TiO2 to an acceptor. The photocatalytic activity increases as the charge carriers recombination rate is decreased.

In a recent review by Low et al. [21] the deposition of Au onto TiO2 surface is reported to result in electron transfer from photo-excited Au particles (> 420 nm) to the conduction band of TiO2, which showed a decrease in their absorption band (∼550 nm) and the band was recovered by the addition of electron donors such as Fe2+ and alcohols. Zhang et al. [39] reported that the visible light activity of coupled Au/TiO2 can be ascribed to the electric field enhancement near the metal nanoparticles. Moreover, numerous researchers coupled Au and Ag nanoparticles onto TiO2 surface to use their properties of localized surface plasmonic resonance (LSPR) in photocatalysis [40]. Wang et al. [41] and Hu et al. [42] reported an improved photocatalytic performance due to the Pt nanoparticle which increased the electron transfer rate to the oxidant. It was observed that photocatalytic sacrificial hydrogen generation was influenced by several parameters such as platinum loading (wt%) on TiO2, solution pH, and light (UV, visible and solar) intensities [43]. Moreover, complete discoloration and dye mineralization were achieved using Pt/TiO2 as catalyst; the results were attributed to the higher Pt content of the photocatalyst prepared with the highest deposition time. For Pt-TiO2 catalysts the best discoloration and dye mineralization were obtained over the catalyst prepared by photochemical deposition method and using 120 min of deposition time in the synthesis. These results may be due to the higher Pt content of the photocatalyst prepared with the highest deposition time.

Haung et al. [44] prepared Pt/TiO2 nanoparticles from TiO2 prepared at various hydrolysis pH values and found that the phase of TiO2 obtained depended largely on the hydrolysis pH. The anatase/rutile intersection of a Pt/TiO2 sample had a lower recombination rate compared to the anatase phase of Pt/TiO2 due to the longer recombination pathway. Though, the Pt/TiO2 anatase phase showed better degradation efficiency than the Pt/TiO2 anatase/rutile intersection. The decrease in the anatase composition of TiO2, and the decrease in the composition of TiO2 resulted in the degradation rate decrease, suggesting that anatase composition in the Pt/TiO2 system played a crucial role of increasing the photocatalytic degradation of Acid Red 1 dye.

Liu et al. [45] prepared the palladium doped TiO2 (Pd-TiO2) photocatalyst using chemical reduction method and tested it the photocatalytic degradation of organic pollutant. It was found that the TiO2 grain size was reduced while the specific surface area increased and the absorption of ultraviolet light also enhanced after using chemical reduction method, however, all these changes had no effect on degradation of organic pollutant. But the degradation was significantly improved due to the deposition of Pd nanoparticles; the Pd/TiO2 organic pollutant degradation was 7.3 times higher compared to TiO2 (P25).

Repouse et al. [46] prepared a series of noble metal promoted TiO2 (P25) by wet impregnation and found that the dispersion of the small metal crystallites on TiO2 did not affect the optical band gap of TiO2. The Pt-promoted catalyst exhibited the highest photocatalytic efficiency in the degradation of bisphenol A under solar irradiation. They also found the presence of humic acid to considerably improve the reaction rate of Rh/TiO2 but had a clearly adverse effect with P25 TiO2 photocatalyst. Fluorescence data revealed that humic acid is capable of photosensitizing the Rh/TiO2 catalyst.

Indium-doped TiO2 have recently been used for photocatalytic reduction of CO2 [47]. Indium doping resulted in an increase in surface area because of suppression of TiO2 particle growth during the TiO2 synthesis. The light absorption ability of the In-TiO2 was enhanced due to the introduction of the impurity level below the conduction band level of the TiO2. The photocatalytic CO2 reduction activity of the In-TiO2 was about 8 time that of pure TiO2 as a consequence of the high surface area and extended light absorption range.

The doping of TiO2 with transition metals such as Cr [48], Co [48], Fe [48, 49, 50], Ni [48, 51], Mn [48, 52], V [53], Cu [54], Ni [51] and Zn [55], has been studied by different research groups. Numerous studies reported that doping of TiO2 with transition metals improve the photocatalytic activity, attributable to a change in the electronic structure resulting in the absorption region being shifted from UV to visible light. The shift results from charge-transfer transition between the d electrons of the transition metals and the conduct or valence band of TiO2 nanoparticles. Inturi et al. [48] compared the doping of TiO2 nanoparticles with Cr, Fe, V, Mn, Mo, Ce, Co, Cu, Ni, Y and Zr and it was found that Cr, Fe and V showed improved conversions in the visible region while, the incorporation of the other transition metals (Mn, Mo, Ce, Co, Cu, Ni, Y and Zr) exhibited an inhibition effect on the photocatalytic activity. The Cr-doped TiO2 demonstrated a superior catalytic performance and the rate constant was found to be approximately 8–19 times higher than the rest of the metal doped catalysts. It was reported that the reduction peaks in Cr-doped TiO2 shifted to much lower temperatures, due to the increase in the reduction potential of titania and chromium. Therefore, the higher photocatalytic efficiency of Cr/TiO2 in the visible light can be attributed to strong interaction (formation of Cr-O-Ti bonds). Fe-doped TiO2 nanoparticles were used in the visible light degradation of para-nitrophenol and it was found that the Fe-dopant concentration was crucially important in determining the activity of the catalyst. The maximum degradation rate of para-nitrophenol observed was 92% in 5 h when the Fe(3+) molar concentration was 0.05 mol%, without addition of any oxidizing reagents. The excellent photocatalytic activity was as a result of an increase in the threshold wavelength response as well as maximum separation of photogenerated charge carriers [49]. On the other hand, Fe-doped TiO2 evaluated for solar photocatalytic activity for the degradation of humic acid showed a retardation effect for the doped catalysts compared to the bare TiO2 specimens, which could be attributed to surface complexation reactions rather than the reactions taking place in aqueous medium. The faster removal rates attained by using bare TiO2 could be regarded as substrate specific rather than being related to the inefficient visible light activated catalytic performance [50]. Ola et al. [56] reported that the properties of V doped TiO2 were tuned towards visible light because of the substitution of the Ti4+ by V4+ or V5+ ions since the V4+ is centred at 770 nm while the absorption band of V5+ is lower than 570 nm. Moradi et al. [57] obtained high photocatalytic activity of Fe doped TiO2 and studied the effects of Fe3+ doping content on the band gap and size of the nanoparticles. It was found that the increase in the doping content decreased the band gap energy and particle size from 3.3 eV and 13 nm for bare TiO2 to 2.9 eV and 5 nm for Fe10-TiO2, respectively.

The rare earth metals doped TiO2 catalyst also have good electron trapping properties which can result in a stronger absorption edge shift towards longer wavelength, obtaining narrow band gap. Bethanabotla et al. [58] carried out a comprehensive study on the rare earth doping into a TiO2 and found that the rare earth dopants improved the aqueous-phase photodegradation of phenol at low loadings under simulated solar irradiation, with improvements varying by catalyst composition. Differences in defect chemistry on key kinetic steps were given as the explanation for the enhanced performance of the rare earth doped samples compared to pure titania. Reszczyńska et al. [59] prepared a series of Y3+, Pr3+, Er3+ and Eu3+ modified TiO2 nanoparticles photocatalysts and results demonstrate that the incorporation of RE3+ ions into TiO2 nanoparticles resulted in blue shift of absorption edges of TiO2 nanoparticles and could be ascribed to movement of conduction band edge above the first excited state of RE3+. Moreover, incorporated RE3+ ions at the first excited state interact with the electrons of the conduction band of TiO2, resulting in a higher energy transfer from the TiO2 to RE3+ ions. But observed blue shift could be also attributed to decrease in crystallite size of RE3+–TiO2 in comparison to TiO2. The Y3+, Pr3+, Er3+ and Eu3+ modified TiO2 nanoparticles exhibited higher activity under visible light irradiation compared to pure P25 TiO2 and can be excited under visible light in the range from 420 to 450 nm. In a similar work on rare earths (Er, Yb, Ho, Tb, Gd and Pr) titania nanotubes (RE-NTs), [60] the RE3+ species were found to be located at the crystal boundaries rather than inside the TiO2 unit cell and an observed excitation into the TiO2 absorption band with resulting RE3+ emission confirmed energy migration between the TiO2 matrix and RE3+. The presence of the rare earth component was found to reduce recombination of the electrons and holes successfully by catching them and also by promoting their rapid development along the surface of TiO2 nanoparticles. Lanthanide ions doping did not impact the energy gap of TiO2 nanoparticles, however this enhanced the light absorption of catalyst. The surface range of TiO2 nanoparticles generally increases by La3+ particle doping by diminishing the crystallite size and accordingly, the doped TiO2 nanoparticle displayed higher adsorption capacity. Based on theoretical calculations, it was proposed that during the electrochemical process, new Ho-f states and surface vacancies were formed and may reduce the photon excitation energy from the valence to the conduction band under visible light irradiation. The photocatalytic activity under visible light irradiation was attributed not to ·OH but to other forms of reactive oxygen species (O2·, HO2, H2O2).

2.2. Non-metal doping

TiO2 nanoparticles have been comprehensively doped at the O sites with non-metals such as C [61], B [62], I [63], F [64], S [65], and N [66]. Non-metal dopants are reported to be more appropriate for the extension of the photocatalytic activity of TiO2 into visible region compared to metal dopant [67, 68]. This can be ascribed to the impurity states which are near the valence band edge, however, they do not act as charge carriers, and their role as recombination centres might be minimized [53]. As shown in Figure 2, the mixing of the p states of the doped non-metal with the O2p states shifts the valence band edge upward and narrows the band-gap energy of the doped TiO2 photocatalyst. The nitrogen and carbon doped TiO2 nanoparticles has been reported to exhibit greater photocatalytic activity under visible light irradiation compared to other non-metal dopants.

Figure 2.

Band-gap energy narrowing mechanism for non-metal-doped TiO2.

N-doped TiO2 (N-TiO2) appears to be the most efficient and extensively investigated photocatalyst for non-metal doping. Zeng et al. [69] reported the preparation of a highly active modified N-TiO2 nanoparticle via a novel modular calcination method. The excellent photocatalytic performance of the photocatalyst was ascribed to excellent crystallinity, strong light harvesting and fast separation of photogenerated carriers. Moreover, the enhancement of charge separation was attributed to the formation of paramagnetic [O-Ti4+-N2−-Ti4+-VO] cluster. The surface oxygen vacancy induced by vacuum treatment trapped electron and promoted to generate super oxygen anion radical which was a necessary active species in photocatalytic process. Phongamwong et al. [70] investigated the photocatalytic activity of CO2 reduction under visible light over modified N-TiO2 photocatalyst and they have found that the band gap of N-TiO2 photocatalyst slightly decreases with increasing N content. In addition, the sub-band energies related to the impurity energy level were observed in the N-TiO2 photocatalyst because of the interstitial N species and the sub-band gap energies were found to have decreased from 2.18 eV with 10 wt% N-TiO2 photocatalyst. In contrast, the replacement of O by N is difficult because of the radius of N (17.1 nm) being higher compared to O (14 nm) and the electroneutrality can be maintained by oxygen vacancies, that are provided by replacement of three oxygen vacancies by two nitrogen atom [71]. N-TiO2 photocatalyst reduces the oxygen energy vacancies from 4.2 to 0.6 eV, suggesting that N favors the formation of oxygen vacancies [72].

In contrast, O atoms (14 nm) could be substituted easily by F atoms (13.3 nm) because of their similar ionic radius [73]. Yu et al. [64] reported that the F-doped TiO2 (F-TiO2) is able to absorb visible light due to the high-density states that were evaluated to be below the maxima valence band, although there was no shift in the band edge of TiO2. Samsudin et al. found a synergistic effect between fluorine and hydrogen in hydrogenated F-doped TiO2 which enabled light absorption in UV, visible and infrared light illumination with enhanced electrons and holes separation. Surface vacancies and Ti3+ centres of the hydrogenated F-doped catalyst coupled with enhanced surface hydrophilicity facilitated the production of surface-bound and free hydroxyl radicals. Species present on the surface of the catalyst triggered the formation of new Ti3+ occupied states under the conduction band of the hydrogenated F-doped TiO2, thus narrowing the band gap energy [73]. Enhanced photocatalytic performance of N-doped TiO2 over pure TiO2 has also been ascribed to efficient separation of electron-hole pairs as well as an increased creation of surface radicals such as hydroxyl The band gap can also be narrowed by doping TiO2 with S, since replacement of S into TiO2 can be performed easily due to larger radius of S atoms (18 nm) compared to O atoms (14 nm). S incorporation in TiO2 has been reported to change the lattice spacing of the TiO2 with a reduction in the band gap width from 3.2 to 1.7 eV allowing for higher photocatalytic activity [74]. N, S and C co-doped TiO2 samples photocatalytic reduction of Cr(IV) showed that the co-doping and calcination played an important role in the microstructure and photocatalytic activity of the catalysts. The co-doped samples calcined at 500°C showed the highest activities ascribed to the synergistic effect in enhancing crystallization of anatase and (N, S and C) co-doping. The carbon doped TiO2 (C-TiO2) is reported to be more active than N-TiO2, therefore, C-TiO2 has received special attention [75]. Noorimotlagh et al. [76] investigated the photocatalytic removal of nonylphenol (NP) compound using visible light active C-TiO2 with anatase/rutile. It was found that the doping of C into TiO2 lattice may enhance the visible light utilization and affect the structural properties of the as-synthesized photocatalysts. Moreover, it was reported that after C doping and changing the calcination temperature, the band gap was narrowed from 3.17 to 2.72 eV and from 2.72 to 2.66 eV, respectively. Ji et al. [61] reported the preparation of C-TiO2 with a diameter of around 200 nm and the tube wall was composed of anatase TiO2, amorphous carbon, crystalline carbon and carbon element doping into the lattice of TiO2. The C-TiO2 nanotubes exhibited much better performance in photocatalytic activity than bare TiO2 under UV and visible light. The obtained results were ascribed to the C doping, which narrowed the band gap energy of TiO2, extended the visible light adsorption toward longer wavelength and hindered charge recombination.

2.3. Co-doping and tri-doping

Although single metal doped and non-metal doped TiO2 have exhibited excellent performance in decreasing the electrons and holes recombination, but they suffer from thermal stability and losing a number of dopants during catalyst preparation process [77]. Therefore, co-doping of two kinds of atoms into TiO2 has recently attracted much interest [78]. The electronic structure of TiO2 can be altered by co-doping on TiO2 by formation of new doping levels inside its band gap. Abdullah et al. [77] reported that the doping levels situated within the band gap of TiO2 can either accept photogenerated electrons from TiO2 valence band or absorb photons with longer wavelengths. Therefore, suggesting that the TiO2 absorption range can be expanded.

Zang et al. [79] evaluated the photocatalytic degradation of atrazine under UV and visible light irradiation by N,F-codoped TiO2 nanowires and nanoparticles in aqueous phase. It was found that photocatalytic degradation of atrazine was higher in the presence of N,F-codoped TiO2 nanowires than that of N,F-codoped TiO2 nanoparticles. The higher photocatalytic performance in the presence of N,F-codoped TiO2 nanowires was attributed to the higher charge carrier mobility and lower carrier recombination rate. Moreover, the speed of electron diffusion across nanoparticle intersections is several orders of magnitude smaller compared to that of nanowire because of frequent electron trapping at the intersections of nanoparticles and increasing the recombination of separated charges before they reach the TiO2 nanoparticles surface. Park et al. [80] showed the best performance for novel Cu/N-doped TiO2 photoelectrodes for dye-sensitized solar cells. It was found that the Cu/N-doped TiO2 nanoparticles provided higher surface area, active charge transfer and decreased charge recombination. Moreover, the addition of suitable content of Cu- to N-doped TiO2 electrode effectively inhibited the growth of TiO2 nanoparticles and improved the optical response of the photoelectrode under visible light irradiation. Chatzitakis et al. [81] studied the photoelectrochemical properties of C, N, F codoped TiO2 nanotubes. It was found that increasing surface area is not followed by increase in the photoconversion efficiency, but rather that an optimal balance between electroactive surface area and charge carrier concentration occurs.

Zhao et al. [82] investigated the photocatalytic H2 evolution performance of Ir-C-N tridoped TiO2 under UV-visible light irradiation. The photocatalytic activity of TiO2 nanoparticles was reported to be improved by Ir-C-N tridoped TiO2 under UV-visible light, due the synergistic effect between Ir, C and N on the electron structure of TiO2. It was found that Ir existed as Ir4+ by substituting Ti in the lattice of TiO2 nanoparticles, whereas the C and N were also incorporated into the surface of TiO2 nanoparticles in interstitial mode. The absorption of TiO2 nanoparticles was expanded into the visible light region and the band gap was narrowed to ~3.0 eV, resulting in improved photocatalytic H2 evolution under UV-visible light irradiation. Tan et al. [83] investigated the photocatalytic degradation of methylene blue by W–Bi–S-tridoped TiO2 nanoparticles. It was found that the absorption edge of TiO2 was expanded into visible-light region after doping with W, Bi and S and the catalytst showed the best photocatalytic activity, than that of TiO2, S-TiO2, W–S–TiO2 and Bi–S–TiO2. This might be attributed to the synergistic effect of W, Bi and S.

2.4. Nano-structured TiO2

Amongst the various strategies that have been used to enhance TiO2 photocatalytic activity, improvement of morphology, crystal structure and surface area have also been considered important and widely investigated approach to achieve better photocatalytic performance. The nanotitania crystallinity can simply be enhanced by optimizing the annealing temperature. However, the stability of the structure and geometries have to be considered when annealing [84]. For the nanotitania morphology and surface area, various ordered structures have been studied. TiO2 nanotubes [85, 86], nanowires [79], nanospheres [87], etc. Tang et al. fabricated monodisperse mesoporous anatase TiO2 nanospheres using a template material and found the resulting catalysts to show high photocatalytic degradation efficiency and selectivity towards different target dye molecules and could be readily separated from a slurry system after photocatalytic reaction [87]. Anodic TiO2 nanotubes have been reported to allow a high control over the separation of photogenerated charge carriers in photocatalytic reactions. The nanotube array has as key advantage the fact that nanotube modifications can be embedded site specifically into the tube wall or at defined locations along the tube wall. This allows for engineering of reaction sites giving rise to enhanced photocatalytic efficiencies and selectivities [88].

2.5. Nanocarbon modified TiO2

The design and preparation of graphene-based composites containing metal oxides and metal nanoparticles have attracted attention for photocatalytic performances. For example, Tan et al. [89] prepared a novel graphene oxide-doped-oxygen-rich TiO2 (GO–OTiO2) hybrid heterostructure and evaluated its activity for photoreduction of CO2 under the irradiation of low-power energy-saving daylight bulbs. It was found that the photostability of O2–TiO2 was significantly improved by the addition of GO, at which the resulting hybrid composite retained a high reactivity. The photoactivity attained was about 1.6 and 14.0 folds higher than that of bare O2–TiO2 and the commercial Degussa P25, respectively. This high photocatalytic performance of GO–OTiO2 was attributed to the synergistic effect of the visible-light-responsiveness of O2–TiO2 and an enhanced separation and transfer of photogenerated charge carriers at the intimate interface of GO–OTiO2 heterojunctions. This study is reported to have opened up new possibilities in the development of novel, next generation heterojunction photocatalysts for energy and environmental related applications. Lin et al. [90] also investigated photoreduction of CO2 with H2O vapor in the gas-phase under the irradiation of a Xe lamp using TiO2/nitrogen (N) doped reduced graphene oxide (TiO2/NrGO) nanocomposites. They found that the quantity and configuration of N dopant in the TiO2/NrGO nanocomposites strongly influenced the photocatalytic efficiency, and the highest catalytic activity was observed for TiO2/NrGO nanocomposites with the highest N doping content. Moreover, modified TiO2/rGO demonstrated a synergistic effect, enhancing CO2 adsorption on the catalyst surface and promoting photogenerated electron transfer that resulted in a higher CO2 photoreduction rate of TiO2/NrGO. Qu et al. [91] prepared the graphene quantum dots (GQDs) with high quantum yield (about 23.6% at an excitation wavelength of 320 nm) and GQDs/TiO2 nanotubes (GQDs/TiO2 nanoparticles) nanocomposites and the photocatalytic activity was tested towards the degradation of methyl orange. It was found that the GQDs deposited on TiO2 nanoparticles can expand the visible light absorption of TiO2 nanoparticles and enhance the activity on photocatalytic degradation of methyl orange under UV-vis light irradiation (ʎ = 380–780 nm). Furthermore, the photocatalytic activity of GQDs/TiO2 nanoparticles was approximately 2.7 times as higher than that of bare TiO2 nanoparticles. Tian et al. [92] reported the preparation of N, S co-doped graphene quantum dots (N, S-GQDs)-reduced graphene oxide- (rGO)-TiO2 nanotubes (TiO2NT) nanocomposites for photodegradation of methyl orange under visible light irradiation. It was found that the S-GQDs+rGO + TiO2 nanocomposites simultaneously showed an extended photoresponse range, improved charge separation and transportation properties. Moreover, the apparent rate constant of N, S-GQDs+rGO + TiO2NT is 1.8 and 16.3 times higher compared to rGO + TiO2NT and pure TiO2NT, respectively. Suggesting that GQDs can improve the utilization of solar light for energy conversion and environmental therapy.

2.6. Immobilized TiO2

Another drawback of TiO2 nanoparticles mentioned above is the formation of uniform suspension in water which makes its recovery difficult, therefore hindering the application of photocatalytic in an industrial scale. As a result, many studies have attempted the modification of TiO2 nanoparticles on support materials such as clays [93, 94] quartz [95], stainless steel [96], etc. Clays have been reported to be a significant support material for TiO2 nanoparticles because of their layered morphology, chemical as well as mechanical stability, cation exchange capacity, non-toxic nature, low cost and availability. Therefore, TiO2/clay nanocomposites have attracted much attention for application in both water and air purification and have been prepared by numerous researchers. Belver et al. [97] investigated the removal of atrazine under solar light using a novel W-TiO2/clay photocatalysts. It was found that the photocatalytic activity of W-TiO2/clay catalyst exhibited higher photocatalytic performance than that of an un-doped TiO2/clay, which was explained by the presence of W ions in the TiO2 nanostructure. The substitution of Ti ions with W resulted in the increase of its crystal size and the distortion of its lattice and moderately narrower band gap of photocatalysts. Mishra et al. [98] reported the preparation of TiO2/clay nanocomposites for photocatalytic degradation of VOC and dye. They found that the photocatalytic performance of TiO2/clay nanocomposites is highly dependent on the clay texture (as 2:1 clays show highest activity than 1:1) apart from their surface area and porosity. Moreover, the reactions involving TiO2/Clay photocatalyst were fast with rate constant of 0.02886 and 0.04600 min−1 for dye and VOC respectively than the other nanocomposites.

Advertisement

3. Conclusions

In this chapter, we have given an overview of the development of modified TiO2 catalysts and its future prospects from a scientific point of view. We note that the field has experienced major advances in the last 5 years especially in the area of modifying TiO2 with carbon nanomaterials. Based on the literature we have covered here, we believe that there is still quite a lot that can be achieved in improving the performance of TiO2 catalysts for photocatalytic applications.

Advertisement

Conflict of interest

There are no conflicts of interest to declare.

References

  1. 1. Dong H, Zeng G, Tang L, Fan C, Zhang C, He X. An overview on limitations of TiO2-based particles for photocatalytic degradation of organic pollutants and the corresponding countermeasures. Water Research. 2015;79:128-146
  2. 2. Jiang L, Wang Y, Feng C. Application of photocatalytic technology in environmental safety. Procedia Engineering. 2012;45:993-997
  3. 3. Tasbihi M, Călin I, Šuligoj A, Fanetti M, Lavrenčič Štangar U. Photocatalytic degradation of gaseous toluene by using TiO2 nanoparticles immobilized on fiberglass cloth. Journal of Photochemistry and Photobiology A: Chemistry. 2017;336:89-97
  4. 4. Baloyi J, Seadira T, Raphulu M, Ochieng A. Preparation, characterization and growth mechanism of dandelion-like TiO2 nanostructures and their application in photocatalysis towards reduction of Cr (VI). Materials Today: Proceedings. 2015;2(7):3973-3987
  5. 5. Valencia S, Marín JM, Restrepo G. Study of the bandgap of synthesized titanium dioxide nanoparticules using the sol-gel method and a hydrothermal treatment. The Open Materials Science Journal. 2010;4:9-14
  6. 6. Jing L, Li S, Song S, Xue L, Fu H. Investigation on the electron transfer between anatase and rutile in nano-sized TiO2 by means of surface photovoltage technique and its effects on the photocatalytic activity. Solar Energy Materials and Solar Cells. 2008;92(9):1030-1036
  7. 7. Ohno T, Sarukawa K, Tokieda K, Matsumura M. Morphology of a TiO2 photocatalyst (Degussa, P-25) consisting of anatase and rutile crystalline phases. Journal of Catalysis. 2001;203(1):82-86
  8. 8. Etacheri V, Di Valentin C, Schneider J, Bahnemann D, Pillai SC. Visible-light activation of TiO2 photocatalysts: Advances in theory and experiments. Journal of Photochemistry and Photobiology C: Photochemistry Reviews. 2015;25:1-29
  9. 9. Riboni F, Bettini LG, Bahnemann DW, Selli E. WO3–TiO2 vs. TiO2 photocatalysts: Effect of the W precursor and amount on the photocatalytic activity of mixed oxides. Catalysis Today. 2013;209:28-34
  10. 10. Szczepanik B. Photocatalytic degradation of organic contaminants over clay-TiO2 nanocomposites: A review. Applied Clay Science. 2017;141:227-239
  11. 11. Bhattacharyya A, Kawi S, Ray MB. Photocatalytic degradation of orange II by TiO2 catalysts supported on adsorbents. Catalysis Today. 2004;98(3):431-439
  12. 12. Torimoto T, Okawa Y, Takeda N, Yoneyama H. Effect of activated carbon content in TiO2-loaded activated carbon on photodegradation behaviors of dichloromethane. Journal of Photochemistry and Photobiology A: Chemistry. 1997;103(1):153-157
  13. 13. Lepore GP, Persaud L, Langford CH. Supporting titanium dioxide photocatalysts on silica gel and hydrophobically modified silica gel. Journal of Photochemistry and Photobiology A: Chemistry. 1996;98(1):103-111
  14. 14. Tang Y, Zhang G, Liu C, Luo S, Xu X, Chen L, et al. Magnetic TiO2-graphene composite as a high-performance and recyclable platform for efficient photocatalytic removal of herbicides from water. Journal of Hazardous Materials. 2013;252(253):115-122
  15. 15. Pellegrino F, Pellutiè L, Sordello F, Minero C, Ortel E, Hodoroaba V-D, et al. Influence of agglomeration and aggregation on the photocatalytic activity of TiO2 nanoparticles. Applied Catalysis B: Environmental. 2017;216:80-87
  16. 16. Melcher J, Barth N, Schilde C, Kwade A, Bahnemann D. Influence of TiO2 agglomerate and aggregate sizes on photocatalytic activity. Journal of Materials Science. 2017;52(2):1047-1056
  17. 17. De Temmerman P-J, Verleysen E, Lammertyn J, Mast J. Semi-automatic size measurement of primary particles in aggregated nanomaterials by transmission electron microscopy. Powder Technology. 2014;261:191-200
  18. 18. Ivanova I, Mendive CB, Bahnemann D. The role of nanoparticulate agglomerates in TiO2 photocatalysis: Degradation of oxalic acid. Journal of Nanoparticle Research. 2016;18(7):187
  19. 19. Ni M, Leung MKH, Leung DYC, Sumathy K. A review and recent developments in photocatalytic water-splitting using TiO2 for hydrogen production. Renewable and Sustainable Energy Reviews. 2007;11(3):401-425
  20. 20. Schneider J, Matsuoka M, Takeuchi M, Zhang J, Horiuchi Y, Ampo M, et al. Understanding TiO2 photocatalysis: Mechanisms and materials. Chemical Reviews. 2014;114(19):9919-9986
  21. 21. Low J, Cheng B, Yu J. Surface modification and enhanced photocatalytic CO2 reduction performance of TiO2: A review. Applied Surface Science. 2017;392:658-686
  22. 22. Verbruggen SW. TiO2 photocatalysis for the degradation of pollutants in gas phase: From morphological design to plasmonic enhancement. Journal of Photochemistry and Photobiology C: Photochemistry Reviews. 2015;24:64-82
  23. 23. Liu N, Chen X, Zhang J, Schwank JW. A review on TiO2-based nanotubes synthesized via hydrothermal method: Formation mechanism, structure modification, and photocatalytic applications. Catalysis Today. 2014;225:34-51
  24. 24. Wen J, Li X, Li W, Fang Y, Xie J, Xu Y. Photocatalysis fundamentals and surface modification of TiO2 nanomaterials. Chinese Journal of Catalysis. 2015;36(12):2049-2070
  25. 25. Daghrir R, Drogui P, Robert D. Modified TiO2 for environmental photocatalytic applications: A review. Industrial & Engineering Chemistry Research. 2013;52(10):3581-3599
  26. 26. Pelaez M, Nolan NT, Pillai SC, Seery MK, Falaras P, Kontos AG, et al. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Applied Catalysis B: Environmental. 2012;125:331-349
  27. 27. Ahmad H, Kamarudin SK, Minggu LJ, Kassim M. Hydrogen from photo-catalytic water splitting process: A review. Renewable and Sustainable Energy Reviews. 2015;43:599-610
  28. 28. Liu L, Li Y. Understanding the reaction mechanism of photocatalytic reduction of CO2 with H2O on TiO2-based photocatalysts: A review. Aerosol and air quality research. 2014;14(2):453-469
  29. 29. Morales-Torres S, Pastrana-Martínez LM, Figueiredo JL, Faria JL, Silva AMT. Design of graphene-based TiO2 photocatalysts—A review. Environmental Science and Pollution Research. 2012;19(9):3676-3687
  30. 30. Zangeneh H, Zinatizadeh AAL, Habibi M, Akia M, Hasnain Isa M. Photocatalytic oxidation of organic dyes and pollutants in wastewater using different modified titanium dioxides: A comparative review. Journal of Industrial and Engineering Chemistry. 2015;26:1-36
  31. 31. Kumar SG, Devi LG. Review on modified TiO2 photocatalysis under UV/visible light: Selected results and related mechanisms on interfacial charge carrier transfer dynamics. The Journal of Physical Chemistry A. 2011;115(46):13211-13241
  32. 32. Kuvarega AT, Mamba BB. TiO2-based photocatalysis: Toward visible light-responsive photocatalysts through doping and fabrication of carbon-based nanocomposites. Critical Reviews in Solid State and Materials Sciences. 2017;42(4):295-346
  33. 33. Zhao Q, Wang M, Yang H, Shi D, Wang Y. Preparation, characterization and the antimicrobial properties of metal ion-doped TiO2 nano-powders. Ceramics International. 2018;44(5):5145-5154
  34. 34. Guayaquil-Sosa JF, Serrano-Rosales B, Valadés-Pelayo PJ, de Lasa H. Photocatalytic hydrogen production using mesoporous TiO2 doped with Pt. Applied Catalysis B: Environmental. 2017;211:337-348
  35. 35. Hossain MA, Elias M, Sarker DR, Diba ZR, Mithun JM, Azad MAK, et al. Synthesis of Fe-or Ag-doped TiO2–MWCNT nanocomposite thin films and their visible-light-induced catalysis of dye degradation and antibacterial activity. Research on Chemical Intermediates. 2018;44:2667-2683
  36. 36. Lavorato C, Argurio P, Mastropietro TF, Pirri G, Poerio T, Molinari R. Pd/TiO2 doped faujasite photocatalysts for acetophenone transfer hydrogenation in a photocatalytic membrane reactor. Journal of Catalysis. 2017;353:152-161
  37. 37. Jin C, Dai Y, Wei W, Ma X, Li M, Huang B. Effects of single metal atom (Pt, Pd, Rh and Ru) adsorption on the photocatalytic properties of anatase TiO2. Applied Surface Science. 2017;426:639-646
  38. 38. Zou Z, Zhou Z, Wang H, Yang Z. Effect of Au clustering on ferromagnetism in au doped TiO2 films: Theory and experiments investigation. Journal of Physics and Chemistry of Solids. 2017;100:71-77
  39. 39. Zhang J, Jin X, Morales-Guzman PI, Yu X, Liu H, Zhang H, et al. Engineering the absorption and field enhancement properties of Au–TiO2 nanohybrids via whispering gallery mode resonances for photocatalytic water splitting. ACS Nano. 2016;10(4):4496-4503
  40. 40. Sousa-Castillo A, Comesaña-Hermo M, Rodriguez-Gonzalez B, Pérez-Lorenzo M, Wang Z, Kong X-T, et al. Boosting hot electron-driven photocatalysis through anisotropic plasmonic nanoparticles with hot spots in Au–TiO2 nanoarchitectures. The Journal of Physical Chemistry C. 2016;120(21):11690-11699
  41. 41. Wang F, Wong RJ, Ho JH, Jiang Y, Amal R. Sensitization of Pt/TiO2 using plasmonic Au nanoparticles for hydrogen evolution under visible-light irradiation. ACS applied materials & interfaces. 2017;9(36):30575-30582
  42. 42. Hu Y, Song X, Jiang S, Wei C. Enhanced photocatalytic activity of Pt-doped TiO2 for NOx oxidation both under UV and visible light irradiation: A synergistic effect of lattice Pt4+ and surface PtO. Chemical Engineering Journal. 2015;274:102-112
  43. 43. Chowdhury P, Gomaa H, Ray AK. Sacrificial hydrogen generation from aqueous triethanolamine with eosin Y-sensitized Pt/TiO2 photocatalyst in UV, visible and solar light irradiation. Chemosphere. 2015;121:54-61
  44. 44. Huang B-S, Su E-C, Huang Y-Y, Tseng H-H. Tailored Pt/TiO2 photocatalyst with controllable phase prepared via a modified sol–gel process for dye degradation. Journal of Nanoscience and Nanotechnology. 2018;18(3):2235-2240
  45. 45. Liu Q, Yu ZB, Zhang RH, Li MJ, Chen Y, Wang L, et al. Photocatalytic degradation of perfluorooctanoic acid by Pd-TiO2 photocatalyst. Huan jing ke xue= Huanjing kexue. 2015;36(6):2138-2146
  46. 46. Repousi V, Petala A, Frontistis Z, Antonopoulou M, Konstantinou I, Kondarides DI, et al. Photocatalytic degradation of bisphenol a over Rh/TiO2 suspensions in different water matrices. Catalysis Today. 2017;284:59-66
  47. 47. Tahir M, Amin NS. Indium-doped TiO2 nanoparticles for photocatalytic CO2 reduction with H2O vapors to CH4. Applied Catalysis B: Environmental. 2015;162:98-109
  48. 48. Inturi SNR, Boningari T, Suidan M, Smirniotis PG. Visible-light-induced photodegradation of gas phase acetonitrile using aerosol-made transition metal (V, Cr, Fe, co, Mn, Mo, Ni, cu, Y, Ce, and Zr) doped TiO2. Applied Catalysis B: Environmental. 2014;144:333-342
  49. 49. Sood S, Umar A, Mehta SK, Kansal SK. Highly effective Fe-doped TiO2 nanoparticles photocatalysts for visible-light driven photocatalytic degradation of toxic organic compounds. Journal of Colloid and Interface Science. 2015;450:213-223
  50. 50. Birben NC, Uyguner-Demirel CS, Kavurmaci SS, Gürkan YY, Turkten N, Cinar Z, et al. Application of Fe-doped TiO2 specimens for the solar photocatalytic degradation of humic acid. Catalysis Today. 2017;281:78-84
  51. 51. Yadav HM, Otari SV, Bohara RA, Mali SS, Pawar SH, Delekar SD. Synthesis and visible light photocatalytic antibacterial activity of nickel-doped TiO2 nanoparticles against gram-positive and gram-negative bacteria. Journal of Photochemistry and Photobiology A: Chemistry. 2014;294:130-136
  52. 52. Tripathi AK, Mathpal MC, Kumar P, Singh MK, Soler MAG, Agarwal A. Structural, optical and photoconductivity of Sn and Mn doped TiO2 nanoparticles. Journal of Alloys and Compounds. 2015;622:37-47
  53. 53. Khan H, Berk D. Synthesis, physicochemical properties and visible light photocatalytic studies of molybdenum, iron and vanadium doped titanium dioxide. Reaction Kinetics, Mechanisms and Catalysis. 2014;111(1):393-414
  54. 54. Choudhury B, Choudhury A, Borah D. Interplay of dopants and defects in making cu doped TiO2 nanoparticle a ferromagnetic semiconductor. Journal of Alloys and Compounds. 2015;646:692-698
  55. 55. Kaviyarasu K, Geetha N, Kanimozhi K, Magdalane CM, Sivaranjani S, Ayeshamariam A, et al. In vitro cytotoxicity effect and antibacterial performance of human lung epithelial cells A549 activity of zinc oxide doped TiO2 nanocrystals: Investigation of bio-medical application by chemical method. Materials Science and Engineering: C. 2017;74:325-333
  56. 56. Ola O, Maroto-Valer MM. Transition metal oxide based TiO2 nanoparticles for visible light induced CO2 photoreduction. Applied Catalysis A: General. 2015;502:114-121
  57. 57. Moradi V, Jun MBG, Blackburn A, Herring RA. Significant improvement in visible light photocatalytic activity of Fe doped TiO2 using an acid treatment process. Applied Surface Science. 2018;427:791-799
  58. 58. Bhethanabotla VC, Russell DR, Kuhn JN. Assessment of mechanisms for enhanced performance of Yb/Er/titania photocatalysts for organic degradation: Role of rare earth elements in the titania phase. Applied Catalysis B: Environmental. 2017;202:156-164
  59. 59. Reszczyńska J, Grzyb T, Wei Z, Klein M, Kowalska E, Ohtani B, et al. Photocatalytic activity and luminescence properties of RE3+–TiO2 nanocrystals prepared by sol–gel and hydrothermal methods. Applied Catalysis B: Environmental. 2016;181:825-837
  60. 60. Mazierski P, Lisowski W, Grzyb T, Winiarski MJ, Klimczuk T, Mikołajczyk A, et al. Enhanced photocatalytic properties of lanthanide-TiO2 nanotubes: An experimental and theoretical study. Applied Catalysis B: Environmental. 2017;205:376-385
  61. 61. Ji L, Zhang Y, Miao S, Gong M, Liu X. In situ synthesis of carbon doped TiO2 nanotubes with an enhanced photocatalytic performance under UV and visible light. Carbon. 2017;125:544-550
  62. 62. Simsek EB. Solvothermal synthesized boron doped TiO2 catalysts: Photocatalytic degradation of endocrine disrupting compounds and pharmaceuticals under visible light irradiation. Applied Catalysis B: Environmental. 2017;200:309-322
  63. 63. Zhang L, Qingrui SUN, Hongxiao YAN. Method for Preparing Iodine-Doped TiO2 Nano-Catalyst and Use Thereof in Heterogeneously Catalyzing Configuration Transformation of Trans-Carotenoids. Google Patents; 2017
  64. 64. Yu W, Liu X, Pan L, Li J, Liu J, Zhang J, et al. Enhanced visible light photocatalytic degradation of methylene blue by F-doped TiO2. Applied Surface Science. 2014;319:107-112
  65. 65. Helmy ET, El Nemr A, Mousa M, Arafa E, Eldafrawy S. Photocatalytic degradation of organic dyes pollutants in the industrial textile wastewater by using synthesized TiO2, C-doped TiO2, S-doped TiO2 and C, S co-doped TiO2 nanoparticles. Journal of Water and Environmental Nanotechnology. 2018;3:116-127
  66. 66. Jia T, Fu F, Yu D, Cao J, Sun G. Facile synthesis and characterization of N-doped TiO2/C nanocomposites with enhanced visible-light photocatalytic performance. Applied Surface Science. 2018;430:438-447
  67. 67. Lee HC, Park HS, Cho SK, Nam KM, Bard AJ. Direct photoelectrochemical characterization of photocatalytic H, N doped TiO2 powder suspensions. Journal of Electroanalytical Chemistry. 2017;819:38-45
  68. 68. Jaiswal R, Bharambe J, Patel N, Dashora A, Kothari DC, Miotello A. Copper and nitrogen co-doped TiO2 photocatalyst with enhanced optical absorption and catalytic activity. Applied Catalysis B: Environmental. 2015;168:333-341
  69. 69. Zeng L, Lu Z, Li M, Yang J, Song W, Zeng D, et al. A modular calcination method to prepare modified N-doped TiO2 nanoparticle with high photocatalytic activity. Applied Catalysis B: Environmental. 2016;183:308-316
  70. 70. Phongamwong T, Chareonpanich M, Limtrakul J. Role of chlorophyll in Spirulina on photocatalytic activity of CO2 reduction under visible light over modified N-doped TiO2 photocatalysts. Applied Catalysis B: Environmental. 2015;168:114-124
  71. 71. Mahy JG, Cerfontaine V, Poelman D, Devred F, Gaigneaux EM, Heinrichs B, et al. Highly efficient low-temperature N-doped TiO2 catalysts for visible light photocatalytic applications. Materials. 2018;11(4):584
  72. 72. Wang F, He X, Sun L, Chen J, Wang X, Xu J, et al. Engineering an N-doped TiO 2@ N-doped C butterfly-like nanostructure with long-lived photo-generated carriers for efficient photocatalytic selective amine oxidation. Journal of Materials Chemistry A. 2018;6(5):2091-2099
  73. 73. Samsudin EM, Hamid SBA, Juan JC, Basirun WJ, Centi G. Synergetic effects in novel hydrogenated F-doped TiO2 photocatalysts. Applied Surface Science. 2016;370:380-393
  74. 74. McManamon C, O'Connell J, Delaney P, Rasappa S, Holmes JD, Morris MA. A facile route to synthesis of S-doped TiO2 nanoparticles for photocatalytic activity. Journal of Molecular Catalysis A: Chemical. 2015;406:51-57
  75. 75. Lei XF, Xue XX, Yang H, Chen C, Li X, Niu MC, et al. Effect of calcination temperature on the structure and visible-light photocatalytic activities of (N, S and C) co-doped TiO2 nano-materials. Applied Surface Science. 2015;332:172-180
  76. 76. Noorimotlagh Z, Kazeminezhad I, Jaafarzadeh N, Ahmadi M, Ramezani Z, Silva Martinez S. The visible-light photodegradation of nonylphenol in the presence of carbon-doped TiO2 with rutile/anatase ratio coated on GAC: Effect of parameters and degradation mechanism. Journal of Hazardous Materials. 2018;350:108-120
  77. 77. Abdullah H, Khan MMR, Ong HR, Yaakob Z. Modified TiO 2 photocatalyst for CO 2 photocatalytic reduction: An overview. Journal of CO2 Utilization. 2017;22:15-32
  78. 78. Jin Q, Nie C, Shen Q, Xu Y, Nie Y. Cobalt and sulfur co-doped TiO2 nanostructures with enhanced photo-response properties for photocatalyst. Functional Materials Letters. 2017;10(05):1750061
  79. 79. Zhang Y, Han C, Zhang G, Dionysiou DD, Nadagouda MN. PEG-assisted synthesis of crystal TiO2 nanowires with high specific surface area for enhanced photocatalytic degradation of atrazine. Chemical Engineering Journal. 2015;268:170-179
  80. 80. Park J-Y, Kim C-S, Okuyama K, Lee H-M, Jang H-D, Lee S-E, et al. Copper and nitrogen doping on TiO2 photoelectrodes and their functions in dye-sensitized solar cells. Journal of Power Sources. 2016;306:764-771
  81. 81. Chatzitakis A, Grandcolas M, Xu K, Mei S, Yang J, Jensen IJT, et al. Assessing the photoelectrochemical properties of C, N, F codoped TiO2 nanotubes of different lengths. Catalysis Today. 2017;287:161-168
  82. 82. Zhao XG, Huang LQ. Iridium, carbon and nitrogen multiple-doped TiO2 nanoparticles with enhanced photocatalytic activity. Ceramics International. 2017;43(5):3975-3980
  83. 83. Tan D, Liu Y, Guo W, Wang J, Chang Z, Li D. Synthesis and characterisation of W–Bi–S-tridoped TiO2 nanoparticles with enhanced photocatalytic activity. Materials Science and Technology. 2016;32(16):1673-1677
  84. 84. Lin J, Liu X, Zhu S, Liu Y, Chen X. Anatase TiO(2) nanotube powder film with high crystallinity for enhanced photocatalytic performance. Nanoscale Research Letters. 2015;10:110
  85. 85. Podporska-Carroll J, Panaitescu E, Quilty B, Wang L, Menon L, Pillai SC. Antimicrobial properties of highly efficient photocatalytic TiO2 nanotubes. Applied Catalysis B: Environmental. 2015;176-177:70-75
  86. 86. Weon S, Choi W. TiO2 nanotubes with open channels as deactivation-resistant photocatalyst for the degradation of volatile organic compounds. Environmental Science & Technology. 2016;50(5):2556-2563
  87. 87. Tang H, Zhang D, Tang G, Ji X, Li C, Yan X, et al. Low temperature synthesis and photocatalytic properties of mesoporous TiO2 nanospheres. Journal of Alloys and Compounds. 2014;591:52-57
  88. 88. Zhou X, Liu N, Schmuki P. Photocatalysis with TiO2 nanotubes: “Colorful” reactivity and designing site-specific Photocatalytic centers into TiO2 nanotubes. ACS Catalysis. 2017;7(5):3210-3235
  89. 89. Tan L-L, Ong W-J, Chai S-P, Goh BT, Mohamed AR. Visible-light-active oxygen-rich TiO2 decorated 2D graphene oxide with enhanced photocatalytic activity toward carbon dioxide reduction. Applied Catalysis B: Environmental. 2015;179:160-170
  90. 90. Lin L-Y, Nie Y, Kavadiya S, Soundappan T, Biswas P. N-doped reduced graphene oxide promoted nano TiO2 as a bifunctional adsorbent/photocatalyst for CO2 photoreduction: Effect of N species. Chemical Engineering Journal. 2017;316:449-460
  91. 91. Qu A, Xie H, Xu X, Zhang Y, Wen S, Cui Y. High quantum yield graphene quantum dots decorated TiO2 nanotubes for enhancing photocatalytic activity. Applied Surface Science. 2016;375:230-241
  92. 92. Tian H, Shen K, Hu X, Qiao L, Zheng WN. S co-doped graphene quantum dots-graphene-TiO2 nanotubes composite with enhanced photocatalytic activity. Journal of Alloys and Compounds. 2017;691:369-377
  93. 93. Mishra A, Mehta A, Sharma M, Basu S. Impact of ag nanoparticles on photomineralization of chlorobenzene by TiO2/bentonite nanocomposite. Journal of Environmental Chemical Engineering. 2017;5(1):644-651
  94. 94. Seftel EM, Niarchos M, Mitropoulos C, Mertens M, Vansant EF, Cool P. Photocatalytic removal of phenol and methylene-blue in aqueous media using TiO2@ LDH clay nanocomposites. Catalysis Today. 2015;252:120-127
  95. 95. He Y, Sutton NB, Rijnaarts HHH, Langenhoff AAM. Degradation of pharmaceuticals in wastewater using immobilized TiO2 photocatalysis under simulated solar irradiation. Applied Catalysis B: Environmental. 2016;182:132-141
  96. 96. Feng D, Xu S, Liu G. Application of immobilized TiO2 photocatalysis to improve the inactivation of Heterosigma akashiwo in ballast water by intense pulsed light. Chemosphere. 2015;125:102-107
  97. 97. Belver C, Han C, Rodriguez JJ, Dionysiou DD. Innovative W-doped titanium dioxide anchored on clay for photocatalytic removal of atrazine. Catalysis Today. 2017;280:21-28
  98. 98. Mishra A, Mehta A, Sharma M, Basu S. Enhanced heterogeneous photodegradation of VOC and dye using microwave synthesized TiO2/clay nanocomposites: A comparison study of different type of clays. Journal of Alloys and Compounds. 2017;694:574-580

Written By

John Moma and Jeffrey Baloyi

Submitted: 12 March 2018 Reviewed: 07 June 2018 Published: 05 November 2018