Next Article in Journal
Microstructure and Mechanical Properties of Friction Stir Welded DP1180 Steel Plates
Previous Article in Journal
Terahertz Metamaterial Absorber Based on Ni–Mn–Sn Ferromagnetic Shape Memory Alloy Films
Previous Article in Special Issue
Analysis of the Use of Recycled Aluminum to Generate Green Hydrogen in an Electric Bicycle
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Oxygen-Deficient Titanium Oxide for Photocatalytic Hydrogen Production

1
School of Science, China University of Geosciences, Beijing 100083, China
2
School of Science, Beijing University of Posts and Telecommunications, Beijing 100876, China
*
Author to whom correspondence should be addressed.
Metals 2023, 13(7), 1163; https://doi.org/10.3390/met13071163
Submission received: 18 May 2023 / Revised: 15 June 2023 / Accepted: 20 June 2023 / Published: 22 June 2023

Abstract

:
Photocatalytic technology based on the specific band structure of semiconductors offers a promising way to solve the urgent energy and environmental issues in modern society. In particular, hydrogen production from water splitting over semiconductor photocatalysts attracts great attention owing to the clean source and application of energy, which highly depends on the performance of photocatalysts. Among the various photocatalysts, TiO2 has been intensively investigated and used extensively due to its outstanding photocatalytic activity, high chemical stability, non-toxicity, and low cost. However, pure TiO2 has a wide band gap of approximately 3.2 eV, which limits its photocatalytic activity for water splitting to generate hydrogen only under ultraviolet light, excluding most of the inexhaustible sunlight for human beings. Fortunately, the band gap of semiconductors can be manipulated, in which introducing oxygen defects is one of the most effective measures to narrow the band gap of titanium oxides. This review considers the fundamentals of photocatalytic water splitting for hydrogen production over TiO2, discusses the latest progress in this field, and summarizes the various methods and strategies to induce oxygen defects in TiO2 crystals. Then, the next section outlines the modification approaches of oxygen-deficient titanium oxide (TiO2−δ) to further improve its photocatalytic performance. Finally, a brief summary and outlook of the studies on TiO2−δ photocatalysts for water splitting to produce hydrogen are presented.

Graphical Abstract

1. Introduction

The exploitation and utilization of fossil fuels, such as coal, oil, and natural gas, facilitates the development of industrialization and urbanization. However, fossil fuels are non-renewable resources with limited reserves, which will certainly become scarce. In addition, the use of fossil fuels has dramatically induced negative effects on the ecological environment. For instance, carbon dioxide emitted during fossil fuel burning is one of the main greenhouse gases. The industrial by-products and wastes cause severe pollution to the environment. Some pollutants even harm human health by accumulating through the food chain [1,2]. Hence, it is necessary and urgent to develop sustainable energy to replace fossil fuels. Renewable energy sources, such as solar energy, wind power, and geothermal power, are being developed and widely used. However, the replacement of fossil fuels still remains elusive due to the restriction on techniques and the economy [3,4,5].
Among the many candidates, hydrogen energy is considered as one of the most promising energy carriers. Hydrogen is one of the most abundant elements on earth, and hydrogen energy can be obtained from a variety of natural resources. Moreover, hydrogen has superb combustibility, a high ignition point (585 °C), and a high heat of combustion (1.42 × 105 kJ·kg−1). Compared with most of the common fuels, it has unparalleled superiority (see Table 1). Additionally, the combustion product of hydrogen contains only water (see Equation (1)), while the burning of fossil fuels will produce a large quantity of carbon dioxide, sulfur oxide, nitrogen oxide, and so on, which are associated with a series of severe environmental issues, including greenhouse effects, photochemical smog, and acid rain [6,7]. Comparatively, hydrogen is certainly a clean, efficient, and sustainable energy source with tremendous prospects for development. Nowadays, more than 95% of hydrogen in industry is produced from fossil energy, such as natural gas, petroleum, and coal. However, these traditional processes for hydrogen production have low efficiency and emit a large amount of exhaust gases such as carbon dioxide [7,8,9]. Producing hydrogen by water electrolysis is also an important method to prepare hydrogen on a large scale, but it will consume a large amount of electric energy [10,11]. Fortunately, the emergence of photocatalytic technology provides a new option for hydrogen production: producing hydrogen by photocatalytic water splitting. The abundant water resources and inexhaustible solar energy on Earth provide significant advantages for hydrogen production methods [12].
H 2   ( g ) + 1 2 O 2   ( g ) = H 2 O   ( l ) Δ H = 285.8   kJ · mol 1
Photocatalysts are the key for producing hydrogen efficiently from the photolysis of water. In literature, TiO2 was the first reported photocatalyst, which has been studied extensively and already applied in some specific areas due to its high photocatalytic activity, non-toxicity, good stability, low cost, and so on. Since the 1990s, the TiO2 photocatalyst has made great progress in the fields of photodegradation of environmental pollutants and photocatalytic water splitting to produce hydrogen [14,15,16]. However, the utilization rate of solar energy by TiO2 photocatalyst is very low due to the fact that TiO2 can be excited only by short-wavelength ultraviolet light, which accounts for only approximately 5% of solar light. This drawback urges scientists to develop methods to modify TiO2 photocatalysts which can be driven by visible light. Among them, ion doping, constructing heterojunctions, noble or transition metals decoration, dye sensitization, structural designing, and construction of oxygen defects have proven effective strategies [17,18,19,20]. In particular, the construction of oxygen defects is one of the most efficient ways to manipulate the band gap of titanium oxides. Literature surveys indicate that oxygen-deficient titanium oxide (TiO2−δ) can absorb more visible light than stoichiometric TiO2, and the formation of oxygen defects in titanium oxide could also enhance its electrical conductivity, thus facilitating the transfer of photogenerated electrons [21,22]. As a result, many TiO2−δ-based photocatalysts with superb performance have been developed to generate hydrogen from water splitting [23,24]. In fact, oxygen defects are often consciously or unconsciously introduced into TiO2 in various modifying processes. Hence, some approaches, including but not limited to ion doping, deposition of noble metals, and loading on supports, are often adopted to enhance the photocatalytic activity of TiO2 jointly with introducing oxygen defects [25,26,27,28,29].
Therefore, in this review, the mechanism of photocatalytic hydrogen production by water splitting over TiO2 is firstly discussed in detail. Then, the effect of introducing oxygen defects on the photocatalytic activity of TiO2 is analyzed. The last part of this section provides a brief overview of the research progress in photocatalytic water splitting to generate hydrogen over TiO2−δ-based photocatalysts. In Section 2, a variety of methods to introduce oxygen defects into TiO2 are summarized, and their merits and shortcomings are analyzed. This will guide proper techniques to develop TiO2−δ based materials. In the following Section 3, we will discuss the modification methods of TiO2 photocatalysts in addition to the introduction of oxygen defects, such as ion doping, deposition of noble metals, dye sensitization, and so on, which are helpful for further enhancing the photocatalytic activity of TiO2−δ. Finally, the perspectives and existing challenges of photocatalytic water splitting into hydrogen over TiO2−δ based photocatalysts are presented in the short section of Conclusions and Outlooks.

2. Fundamentals of Producing H2 by Photocatalytic Water Splitting over TiO2

2.1. Mechanism of Photocatalytic Water Splitting to Generate H2

Photocatalysis technology is based on the special energy band structure of semiconductors. In ground state, the valence band (VB) of a semiconductor is fully occupied by electrons and the conduction band (CB) is empty. There is a quantized and discontinuous band gap between the low energy VB and high energy CB. The band gap energy (Eg) of semiconductors is narrower than that of insulators (>5 eV). Therefore, the electrons in VB of a semiconductor can be excited and leap into CB when it is stimulated by photons with certain energy (higher than Eg), leaving the same number of holes in VB (Figure 1). Photogenerated electrons (eCB) and holes (hVB+) possess strong reducing and oxidizing abilities, respectively, and will migrate quickly to the surface of photocatalysts to participate in redox reactions [30]. The photocatalysts can directly decompose water when they are suspended in water, which does not require a complex reaction system. Photocatalytic water splitting over semiconductors generally involves the following five steps: (i) water molecules are adsorbed on the surface of a photocatalyst; (ii) the electrons in VB leap into CB, producing eCB and hVB+ under the irradiation by light; (iii) the photogenerated eCB and hVB+ transfer to the surface of the photocatalyst; (iv) the eCB reduces H+ into hydrogen and hVB+ oxidizes H2O to oxygen, which are commonly referred to as the hydrogen evolution reaction and oxygen evolution reaction; and (v) the produced hydrogen and oxygen are desorbed from the surface of the photocatalyst. Among them, steps II–IV are the rate-determining steps on the photocatalytic water splitting (see Equations (2)–(5)). As the O2 dissolved in water will markedly compete eCB with the hydrogen evolution reaction, sacrificial agent is added into the system to improve the photocatalytic efficiency. The commonly used sacrificial agents include EDTA-2Na, methanol, and so on. In those cases, hVB+ will be quickly captured by sacrificial agents instead of reacting with H2O because a single-electron process usually is faster than an O2 evolution reaction.
Semiconductor + 2hv 2e + 2h+
2H+ + 2e  H2
2H2O + 2h+  O2 + 2H+
Overall reaction: H2O + 2hv H2 (g) + O2 (g)
During photocatalytic water splitting, only the photons carrying energy greater than the Eg value of a semiconductor can excite the valance electrons into CB. As the Eg value of TiO2 is approximately 3.2 eV, only ultraviolet light with a wavelength less than 380 nm can excite its valance electrons. Next, apart from moving to the surface of the semiconductor to participate in redox reactions, those excited electrons will also recombine with the holes, releasing light and/or heat energy. The recombination of eCB and hVB+ is the deactivation process of the photogenerated carriers, which does not contribute to the photocatalytic water splitting and should be avoided as much as possible (Figure 2a). Moreover, the reducing ability of eCB- depends on the bottom of CB (CB minimum) and the oxidizing ability of hVB+ relies on the top of VB (VB maximum). The necessary conditions for photocatalytic water splitting are that the CB minimum is more negative than the reduction potential of H+/H2 (0 V vs. NHE at pH = 0) and the VB maximum is more positive than the oxidation potential of H2O/O2 (1.23 V vs. NHE at pH = 0). This requires an Eg value of no less than 1.23 eV, covering the oxidation-reduction potential of H2O. In fact, the Eg value of photocatalysts for photocatalytic water splitting is generally required to be more than 1.9 eV due to the influence of mechanical and thermodynamic losses. Specifically, the CB minimum and VB maximum of TiO2 are approximately −0.2 and 3 eV, respectively [31]. Therefore, TiO2 can split water into hydrogen and oxygen efficiently through photocatalysis (Figure 2b).

2.2. Impact of Oxygen Defects on the Photocatalytic Activity of TiO2

As mentioned above, TiO2 can only absorb ultraviolet light because of its wide band gap. However, a large proportion (about 50%) of the solar spectrum is visible light. Thus, enhancing the capability of harvesting visible light is an effective way to improve the photocatalytic performance of TiO2. In the literature, it was reported that the original white TiO2 would be turned black after it was thermally treated with H2, indicating that the light absorption capability of the reduced TiO2 (actually TiO2−δ) was significantly enhanced. Moreover, it has been proven that the light absorption spectrum edge of TiO2−δ will shift to a long wavelength as the density of oxygen defects increases (Figure 3a) and the corresponding Eg decreases (Figure 3b) [21,32,33]. When there is an oxygen vacancy, one atom of oxygen in TiO2 is bonded with three Ti atoms, and two redundant electrons are shared by the surrounding three Ti atoms (see Figure 3c). A portion of Ti4+ will be converted into Ti3+ after trapping the redundant electrons. The appearance of Ti3+ species in the nonstoichiometric TiO2−δ is generally considered as the main reason that causes its absorption to visible light. Ti3+ species caused by oxygen defects can introduce new intermediate defect states (shallow donor) below the bottom of CB and modify the band gap structure of TiO2 (Figure 3d), which means that TiO2−δ has a narrower band gap and thus can absorb visible light [34,35,36,37].
On the other hand, the presence of oxygen vacancies enlarges the lattice spaces of TiO2. As a result, the resistance to electron transfer will decrease. A low resistance for electron transfer is beneficial for the quick transfer of photogenerated electrons, thus suppressing the recombination of photogenerated eCB and hVB+ [38]. For example, Hao et al. [39] prepared an oxygen-deficient blue titanium oxide, reporting that the prepared TiO2−δ electrode would present a much lower charge transfer resistance (87 Ω) compared with its TiO2 counterpart (356 Ω) [39]. Additionally, the bridging oxygen vacancies tend to cause the Ti 3d defect state in the band gap of TiO2. The Ti interstitials in the near-surface region can provide the electronic charges that the photocatalytic reactions need [40]. As a result, TiO2−δ will present a higher photocatalytic performance than TiO2.

2.3. Brief Overview on Photocatalytic Water Splitting to Generate H2 over TiO2−δ

Since the earliest report on light-driven water splitting by Fujishima and Honda in 1972 [41], semiconductor photocatalysis has attracted great attention in the field of catalysis. However, for quite a long period, semiconductor photocatalysis developed at a mild speed and many studies were focused on the photodegradation of pollutants [42,43]. After entering the 21st century, studies on semiconductor photocatalysis have grown explosively and quite a lot of photocatalysts with excellent performance have been developed [25,44]. In particular, although oxygen vacancy was reported to generate a defect state in the band gap of TiO2 leading to a narrower band gap of TiO2−δ in 1980s, TiO2−δ-based photocatalysts were promptly developed and applied to water splitting until recently [45,46,47].
In earlier times, oxygen defects were often introduced into TiO2 unconsciously during the modifying process, and it was then discovered that those titanium oxides with oxygen defects perform better on photocatalysis than those without oxygen defects. Therefore, researchers began to develop oxygen-deficient titania photocatalysts and explored the detailed mechanisms of how oxygen defects influence the photocatalytic performance of titanium oxides [37,48]. In 2008, Sasikala et al. [49] synthesized a series of Sn- and Eu-doped TiO2 (Ti1−(x+0.001)Eu0.001SnxO2−δ, where 0.05 < x < 0.3) nanoparticles, which showed an onset of light absorption at approximately 450 nm and high activity for hydrogen generation. Liu et al. [50] subsequently reported an oxygen-deficient anatase TiO2 nanosheet with a dominant (001) crystalline plane, indicating that a special electron transfer process on the reconstructed surface of TiO2 substantially enhanced the hydrogen evolution rate from photocatalytic water splitting. TiO2 treated by H2 at high temperatures also presented enhanced photocatalytic activity for water oxidation and high apparent quantum efficiency for O2 evolution (41% under light irradiation at 365 nm) [51]. An electron-beam irradiated titania film shows a wider range of absorbed light and higher efficiency of hydrogen production owing to the oxygen vacancies or defects enhancing mobility and separation of electrons and holes [52]. Other oxygen-deficient TiO2 samples can be obtained by using the ion layer gas reaction (Spray-ILGAR) technique, microwave induced reduction, and the solution plasma process. They show high photocatalytic hydrogen evolution activity [53,54]. In summary, many TiO2−δ-based photocatalysts have been developed, but most of them are used to degrade pollutants and only a limited number of them are used to split water [22,23,27,55,56,57]. Among these limited reports, thermal treatment in hydrogen is the most widely used method of introducing oxygen defects in TiO2 [22]. The introduced oxygen defects in TiO2 are generally combined with other strategies, such as ion doping and composition with other semiconductors, to achieve high hydrogen evolution activity, which has been the focus of recent studies [47,58,59,60,61].

3. Methods of Introducing Oxygen Defects in TiO2

3.1. Reductive Treatment

Reductive treatment is the most direct way to introduce oxygen defects in TiO2. TiO2 can be reduced into TiO2−δ by adding a proper reducing agent. Among the many reductants, H2 is the most widely used option because of its strong reducing ability without introducing impurities [22,62,63,64,65]. However, H2 treatment is usually carried out at high temperatures and the explosion limit of H2 falls in a very wide range of 4.0~75.6 vol.%. In other words, the operation of H2 treatment on TiO2 is quite dangerous and requires very accurate processes. Moreover, treating TiO2 with H2 is usually a time-consuming task. For example, Xu et al. [66] reported black TiO2 through H2 treatment in a 20.0 bar of H2 atmosphere at approxmately 200 °C for 5 days. Zhang et al. [67] prepared defective TiO2−δ hollow microspheres also by high-temperature H2 reduction for 3 h at 550 °C. Wierzbicka et al. [68] synthesized a reduced “grey” brookite TiO2 photocatalyst by hydrogenating it at 500 °C, showing a remarkable noble metal free photocatalytic H2 evolution performance, substantially higher than that of hydrogenated anatase or rutile TiO2. The density of defects can be adjusted by tuning the H2 treatment temperature, soaking time, and H2 concentration. For instance, Samsudin et al. [69] put TiO2 into a continuous flow of 1 atm of pure H2 at 500 °C for different times, finally obtaining TiO2−δ with different densities of oxygen defects. They indicated that with time of H2 treatment, the density of oxygen defects increased, the color of the products becomes deeper from white to dark gray and to bluish gray (Figure 4a), and the light absorption ability of the resultant TiO2−δ was significantly enhanced (Figure 4b). However, more defects do not always guarantee better photocatalytic performance. Here, the photocatalytic performance of TiO2 hydrogen treated for 24 h is inferior to that of the sample treated for 12 h. This might be due to more defects acting as recombination centers of photogenerated carriers. Thus, the control of oxygen defect density in TiO2 is also important. In addition, because treating TiO2 with H2 will not introduce other impurities, the shallow donor levels of Ti3+ are the major factor narrowing the band gap of titanium oxides. The increased electron density on the catalyst surface led by Ti3+ and oxygen vacancies is also beneficial for improving photocatalytic performance.
Apart from H2, some other gases have been also used as reductants. For example, NH3 is also often used to reduce TiO2. Chen et al. [56] synthesized a N-doped and oxygen-deficient TiO2 photocatalyst by heating the commercially available pure TiO2 in a NH3 atmosphere at 550 °C for 5 h. It is easy to introduce N into TiO2 (N doping) when using NH3 as the reducing agent. Similarly, Ihara et al. [70] prepared a N-doped oxygen-deficient titanium oxide by calcinating the hydrolytic product of Ti(SO4)2 with ammonia in dry air at 400 °C for 1 h. Additionally, some familiar reducing substances such as carbon, NaBH4, and Li can be also used to prepare oxygen-deficient TiO2 [71]. Guan et al. [72] prepared a product of oxygen-deficient TiO2 by a three-step process, which showed strong absorbance over the whole visible-light region. In their process, a Ti coating was first pretreated in carbon powder at 1073 K for 2 h, which was then oxidized at 1073 K for 15 h in air. Next, the obtained samples were treated in carbon powder again at 973 K for 30 min, finally obtaining the product of oxygen-deficient TiO2. Zhao et al. [73] first prepared TiO2 anatase nanorods by a two-step hydrothermal method. Then, the obtained sample was mixed with NaBH4 (1:1 in mole) in a mortar and thermally treated in Ar at 300 °C for 30 min, finally acquiring the reduced anatase nanorods. Interestingly, Martinze et al. [74] prepared a reduced blue TiO2 by using Li foil and TiO2 which were solved in ethylene diamine, stirring in anhydrous and dark conditions for 1440 h. Treating with these non-hydrogen reductants avoids the risk of explosion compared with hydrogen treatment.
In addition, providing an anoxic environment in the treatment process of TiO2 can also result in the same effect as adding reducing agents. For example, Pereira et al. [75] obtained oxygen-deficient TiO2 films with enhanced visible and near-infrared optical absorption by periodically interrupting the O2 gas supply in the process of magnetron sputtering. Dhumal et al. [76] synthesized oxygen-deficient titanium suboxide (TiOx with x < 2) nanoparticles by using a diffusion flame aerosol reactor under an oxygen lean environment in the formation zone of particles. Xiao et al. [77] reported the formation of oxygen vacancies in TiO2 during the process of calcining TiO2 in Ar or N2 atmospheres. Kushwaha et al. [78] prepared a black oxygen-deficient TiO2-graphite nanocomposite by calcining Ti-EDTA complex under hypoxic conditions. Singh et al. [36] investigated the effect of thermal treatment on TiO2 thin films under an oxygen anoxic environment, reporting a reduction in the band gap of 0.36 eV.

3.2. Pulsed Laser Irradiation

The excimer laser is a powerful tool and is often used to manipulate the composition and structure of material surfaces. Pulsed laser irradiation is a simple process for producing black, oxygen-deficient TiO2. A photochemical reduction reaction will take place during the pulsed laser absorption, thereby resulting in the evolution of oxygen deficiencies. The absorption of focused laser irradiation accompanied by fast heating/cooling processes will promote the formation of a porous surface [79,80,81]. As mentioned before, the dangers involved in hydrogenation operation greatly limit its application, while hydrogen plasma irradiation overcomes this shortcoming well [82,83]. For example, Wang et al. [82] synthesized a black titania with a core/shell structure (TiO2@TiO2−xHx) assisted by hydrogen plasma and its photocatalytic activity for water splitting and cleaning pollutants was much better than that of TiO2. In addition, Nd:YAG, ArF, KrF, and XeCl excimer lasers are also frequently used methods besides hydrogen plasma [84,85]. Nakajima et al. [85] indicated that the TiO2 crystal surface would be successfully reduced through ArF, KrF, and XeCl excimer laser irradiation, forming an oxygen-deficient TiO2−δ layer with a thickness of 160 nm. Moreover, as shown in Figure 5a, the resistance of TiO2 decreased after laser irradiation. Significant diffuse scattering around the (220) reflection for a wide range of Qx (0.04~0.04) over the irradiated sample (Figure 5b) indicated a strong local lattice distortion near the surface of the sample. Pulsed laser irradiation is very suitable for surface treatment. It is simple to get high reductive efficiency based on photochemical reactions due to high-power laser irradiation. Meanwhile, the resultant surface of photocatalysts generally has a large specific surface area which is beneficial for improving photocatalytic performance.

3.3. Pulsed Laser Deposition

Pulsed laser deposition (PLD) is a good technique to prepare functional thin films by depositing the ablated substances on a substrate. The oxygen deficiency in the film can be adjusted by controlling the partial pressure of O2 and laser power density. For instance, Leichtweiss et al. [86] prepared oxygen-deficient titanium oxide films with an average composition of TiO1.6 by PLD at room temperature, which presented high efficiency for the water-splitting reaction. Kunti et al. [87] deposited TiO2-SiO2 composite films on amorphous quartz substrates at different partial pressures of O2 by PLD technique, revealing the generation of oxygen defects and Ti3+ states in the films. Moreover, ion-doped, oxygen-deficient TiO2 films can be obtained by changing the humidity of the environment, atmosphere, and ion implantation [88,89,90]. For instance, Socol et al. [90] fabricated N-doped crystalline TiO2 thin films by PLD in N2 or N2-O2 mixtures. Nath et al. [91] synthesized TiO1.5 nanoparticles by varying the focusing conditions of pulsed laser ablation. Rahman et al. [92] prepared TiO2 nanostructures with different morphologies and incorporation of oxygen vacancy defects on a Si substrate by a single-step, catalyst-assisted PLD method (Figure 6). The morphology can be controlled by adjusting the deposition temperature and template. The film materials with special morphological structures can be prepared by PLD, thus adjusting the specific surface area of the catalysts. Ion doping can also be achieved by changing the reacting atmosphere. Thus, the band gap of titanium oxides can be reduced jointly by oxygen defects and ion doping.

3.4. Ion Doping

Due to the difference in electronegativity between various elements, the introduction of impurity atoms into TiO2 will change the partial concentration of electrons in TiO2, thus producing oxygen defects in it. For instance, Ti4+ will be converted into Ti3+ when the oxygen atoms in TiO2 are replaced by highly electronegative F atoms due to the increased electron density around Ti4+ caused by the doped F atoms [93]. As shown in Figure 7a, clear Ti3+ signals can be observed in the EPR spectrum of fluorine-treated anatase. The corresponding Raman spectra also display a slight shift to a higher frequency at the peak of 144 cm−1, which is attributed to the presence of oxygen vacancies and Ti3+. The oxygen vacancies are spontaneously introduced during N doping [94]. Pu et al. [95] successfully prepared N-doped, oxygen-deficient TiO2 microspheres through a two-step synthesis method. Firstly, TiO2 microspheres are synthesized by solvothermal synthesis. Then, the final oxygen-deficient titanium oxide products were obtained by electron beam irradiation using urea as the nitrogen source, and the concentration of Ti3+ increased with an increasing dose of the electron beam irradiation. Wang et al. [26] reported a N-doped TiO2 (TiO2−xNx) by a simple wet method: hydrolyzing acidic tetra-butyl titanate in ammonia solution followed by calcination at 350 °C for 1 h. Of course, the nitrogen source for doping generally directly or indirectly originates from reducing NH3, which can also promote the reduction of TiO2. Moreover, the doping of some metal ions, such as Eu3+, La3+, and Gd3+, can introduce oxygen defects in TiO2 as well. Those ions with a lower valence than Ti4+ can generate anion vacancies in TiO2 [96,97,98], thereby forming Ti3+. Zhang et al. [99] proved that the formation energy of a vacancy on the La-doped TiO2 surface was lower than that formed on the pure TiO2 surface treated in reducing conditions or oxidizing conditions by calculation (Figure 7b). Wang et al. [96] synthesized 0.4 mol% Gd and 2.0 mol% La co-doped TiO2 microspheres via a hydrothermal method, which exhibited enhanced visible-light absorption. The doped La3+ and Gd3+ create abundant oxygen deficiencies and surface defects in the sample, decreasing the excitation energy of TiO2. Doping TiO2 with highly electronegative elements will inevitably result in oxygen defects. Thus, introducing oxygen defects during the ion-doping process usually occurs unconsciously and controlling the density of oxygen defects is a great challenge. However, scientists can combine the advantages of ion doping and oxygen defects to improve the photocatalytic performance of TiO2 [100,101].

3.5. Plasma-Assisted Deposition

Plasma-enhanced chemical vapor deposition (PECVD) has the features of low deposition temperature, high purity, uniform thickness and composition of films, as well as easy control of reaction parameters. It can be used to prepare various metal films, inorganic films, and organic films. The structure and properties of films can be adjusted by controlling reaction conditions. Specifically, highly active species can be produced by plasma treatment under mild conditions. For example, Hatanaka et al. [102] prepared TiOx:OH films using a remote PECVD technique, which showed high photoconductivity. Sakai et al. [103] obtained oxygen-deficient TiO2 anatase films by using oxygen plasma-assisted reactive evaporation by increasing the supply of titanium atoms, and the resultant oxygen-deficient TiO2 films showed excellent hydrophilicity, which was conducive to thorough contact with water and facilitated its splitting reaction. Li et al. [104] introduced numerous oxygen deficiencies and Ti3+ defects on the surface of TiO2 nanoparticles via Ar plasma. Similarly, Hojo et al. [105] also successfully introduced oxygen defects in a TiO2:Nb film by annealing the sample with Ar plasma irradiation. Recently, Kawakami et al. [106] reported a kind of anatase/rutile mixed phase TiO2 nanoparticle with many oxygen deficiencies, which were obtained by annealing the sample with low-temperature O2 plasma. There are also excited species, such as ozone and OH generated during the plasma discharge in water. Thus, the plasma-liquid interaction has been widely applied to prepare nanomaterials. For instance, An et al. [107] prepared gray hydrogenated TiO2 spheres using a plasma-modified sol-gel system. Mizukoshi et al. [108] obtained a blue TiO2 containing oxygen defects by generating discharge plasma in an aqueous ammonia solution containing TiO2 powder. TiO2 was reduced by a reducing species, such as hydrogen radicals generated during the plasma discharge process in aqueous ammonia. The color of TiO2 was gradually deepened with treating time and the capacity of light absorption was enhanced simultaneously, mainly because of the increasing amount of oxygen defects in the samples. Apart from introducing oxygen defects, plasma-assisted treating also leads to more bridging/terminal oxygen groups adsorbed on the surface of the samples, thus facilitating the charge transfer and suppressing the recombination of electrons and holes.

3.6. Ultrasonic-Assisted Techniques

Ultrasonic spray pyrolysis is a simple, low-cost, and scalable technique [54,109,110,111,112]. In the literature, Nakaruk et al. [110] successfully prepared fully dense TiO2 films with oxygen deficiencies by using ultrasonic spray pyrolysis and proved that the concentration of oxygen deficiencies could be controlled by changing the annealing temperature. Oxygen vacancies can also be directly induced in TiO2 by low-frequency ultrasound (LFUS) treating because the high-speed particle collisions and shock waves from LFUS can impact the atomic arrangement in the TiO2 lattice. For instance, Osorio-Vargas et al. [113] prepared visible-light responsive TiO2-based photocatalysts by dispersing P25 powder into water and exposed this to a LFUS environment for 6 h. Bellardita et al. [114] reported that ultrasonic treating P25 powder dispersed in water induced oxygen deficiency in TiO2, thus narrowing the bandgap of TiO2 from 3.18 to 3.04 eV.

3.7. Calcination under Anoxic Conditions

Thermal treatment atmosphere exerts an important influence on the formation of oxygen deficiencies [115,116,117,118]. The ratio of O and Ti in the resultant titanium oxides will be less than 2 when there is not enough oxygen in the preparation process. For instance, Albetran et al. [119] revealed that the color of titania changed from white to gray and black as the ratio of Ar/air of the thermal treating atmosphere increased (Figure 8a), and the light absorption of the corresponding products was also improved (Figure 8b). Sang et al. [120] fabricated oxygen-deficient TiO2 nanotube arrays by calcining in nitrogen, or a mixture gas of 5% hydrogen in nitrogen, which exhibited higher photocurrent density and smaller charge transfer resistance than that of the samples calcined in air (Figure 8c,d). This is because the large lattice spaces caused by oxygen vacancies decreases the electrical resistance for electron transfer. Qi et al. [121] prepared a defective TiO2 sample with oxygen deficiencies by thermally treating TiO2 at 200 °C under vacuum conditions. The defect concentration in the sample is positively proportional to the thermal treatment time. Li et al. [122] reported an oxygen-deficient dumbbell-shaped anatase TiO2−x product. In detail, a TiCl3-HAc mixed solution was solvothermally treated at 180 °C for 5 h and the solvothermally synthesized product was calcined under vacuum at 400 °C for 1 h.

3.8. Molten Salt Calcination

Du et al. [118] reported a facile strategy based on molten salt calcination to construct oxygen deficiencies in TiO2. A flower-like TiO2 precursor was synthesized via a solvothermal method using tetrabutyl titanate and acetic acid (HAc)/N,N-dimethyl formamide (DMF) as the titanium source and solvent, respectively. The as-prepared precursor was mixed with eutectic salts of LiCl/KCl (45/55 by weight) and calcined in a muffle furnace at 400 °C for 2 h. The lattice oxygen of TiO2 was consumed during the calcination because of the low partial pressure of O2 in the molten salt, thereby introducing numerous oxygen deficiencies and Ti3+ in the final product.
In summary, up to now, hydrogen reduction is still the most extensively used method to prepare oxygen-deficient TiO2 owing to the strong deoxidizing ability and purity. However, it is time consuming and has high energy consumption and a high explosion risk. Thus, some other reductants such as carbon, NH3, and Li are also used to reduce TiO2 in the literature. Synthesizing titanium oxide in an anoxic environment is widely used because it is easily implemented. Pulsed laser irradiation is a simple process for producing oxygen-deficient TiO2; however, this is more suitable for treating films because the radiation response mainly happens in the surface layer. Similarly, oxygen-deficient TiO2 films can be easily obtained through adjusting the partial pressure of O2 and the laser power density of PLD. Introducing oxygen defects through ion doping is a natural process and the density of oxygen defects mainly depends on the doped species of ions and their concentration. Plasma discharge in water will provide reductively excited species, which can easily reduce TiO2. However, it is currently not widely applied. Oxygen-deficient TiO2 can be prepared by ultrasonic spray pyrolysis or by calcinating under anoxic conditions, and the density of oxygen deficiencies can be controlled by controlling the experimental temperature. Molten salt calcination is simple and easily operated. Introducing oxygen defects improves the photocatalytic performance of TiO2 in 2 major ways: one is narrowing band gap to absorb more light, and the other is changing the lattice structure to decrease resistance to electron transfer. Table 2 compares the different methods of introducing oxygen defects in TiO2.

4. Modification Methods of TiO2−δ Photocatalysts

TiO2−δ has been proven to perform better than stoichiometric TiO2 in the process of photocatalytic water splitting. Many strategies such as ion doping, constructing heterojunction and deposition of noble metals have been proved to effectively improve the photocatalytic activity of TiO2. Thus, the photocatalytic activity of TiO2−δ should be enhanced further by these strategies.

4.1. Ion Doping

Ion doping can introduce defects into TiO2 which could act as the capture traps of photogenerated carriers, thereby suppressing the recombination of photogenerated eCB and hVB+. The lattice distortion caused by the doped atoms with different ionic sizes would increase the asymmetry of the crystal structure, which could promote the separation of photogenerated eCB and hVB+. Additionally, the energy band structure of TiO2 can be effectively manipulated by ion doping. The narrowed band gap can extend the light absorption and enhance the utilization efficiency on solar energy of the resultant photocatalysts.

4.1.1. Metal Ion Doping

The doping of transition metals has been proven an effective method for regulating the band positions of TiO2. The main principle is to insert an additional energy level between the original conduction band and valence band. For example, Sheng et al. [123] reported a Pd-doped TiO2, revealing that the photogenerated eCB and hVB+ were efficiently separated after Pd doping. Sasirekha et al. [124] prepared a Ru-doped anatase TiO2 supported on silica by a solid-state dispersion method, which performed well in the photocatalytic reduction of carbon dioxide. Gao et al. [125] indicated that the doping of Mo, Pd, Ru, and Rh could narrow the band gap of TiO2, thus enhancing the probability of activation by visible light. Their theoretically calculated results through density functional theory revealed that the impurity states of 4d electrons would form new degenerate energy levels, thus narrowing the band gap of TiO2. Thalgaspitiya et al. [126] synthesized mesoporous composites of M-doped titanium dioxide (M = Mn, Co, Ni, Mo, and W) with reduced graphene oxide (rGO), indicating that the indirect band gap of the composites could be adjusted into the range of 2.20–2.48 eV.
Rare earth ions have rich energy levels and unique features of 4f electronic transitions, providing unique opportunities for manipulating the band gap of semiconductors by elemental doping. For instance, Wang et al. [127] fabricated samples of La3+- or Yb3+-doped TiO2 supported on r-GO, reporting that the anionic vacancies in the TiO2 lattice caused by La3+ and Yb3+ would generate Ti3+, thus enhancing the visible-light response of the samples. Stengl et al. [128] prepared several samples of rare earth (La, Ce, Pr, Nd, Sm, Eu, Dy, Gd)-doped TiO2, which were all visible-light sensitive. Fang et al. [129] synthesized rare earth ion (Er3+ and/or Yb3+)-doped TiO2 photocatalysts by a hydrothermal method, indicating that the doping of Er3+ and/or Yb3+ could decrease the recombination rate of photogenerated electron-hole pairs, finally leading to a higher photocatalytic efficiency of TiO2. In addition, the phase transition from anatase to rutile can be significantly delayed by the doping of rare earth ions [130,131].
Alkali metal and alkali earth metal ions were also used to improve the photocatalytic activity of TiO2. Liu et al. [132] prepared a mesoporous Na-doped titanium dioxide with a band gap of 3.08 eV. The doped Na ions could enter into the (004) crystalline plane of anatase TiO2, finally leading to the dislocation defects in TiO2. Lv et al. [133] successfully fabricated AM-TiO2−x samples (AM = Mg, Ca, Sr, and Ba), revealing that the CB position of TiO2 became more negative after AM doping, thus improving the hydrogen production ability of TiO2. The separation of carriers and transfer efficiency were also dramatically promoted (Figure 9a–c).

4.1.2. Nonmetallic Ion Doping

The doping of nonmetallic ions can expand the light-absorption region of TiO2 and suppress the recombination of photogenerated eCB and hVB+. Normally, the p orbital in the most outer electronic layer of the doped ions would hybridize with the 2p orbital of O in TiO2, forming new shallow levels near the top of the valence band. For example, N doping is widely studied because the ion radius of N is closest to that of O [134,136,137,138]. Li et al. [136] prepared a N-doped TiO2 which performed better in photocatalytic hydrogen evolution than the undoped TiO2 under the same conditions (Figure 9d). Yuan et al. [139] prepared a N-doped TiO2 with a high specific surface area by heating a mixture of urea and TiO2. The absorption spectrum of the N-doped TiO2 shifted to the wavelength of 600 nm and the sample showed high photocatalytic activity on hydrogen evolution. Momeni et al. [140] prepared S-doped TiO2 nanostructure photocatalyst films which performed well in the removal of RhB and hydrogen generation under visible-light radiation. Carmichael et al. [141] reported B-doped titanium dioxide films with a hydrogen evolution rate of 24 µL·cm−2·h−1, which far exceeded the undoped TiO2 at 2.6 µL·cm−2·h−1. Wu et al. [142] fabricated F-doped TiO2 particulate thin films, which could be applied in the photodegradation of organic pollutants and photoinduced splitting of water into hydrogen under the irradiation of either UV or visible light.

4.1.3. Multiple Ion Co-Doping

Different ions have different impacts on TiO2; thus, the co-doping of multiple ions is an effective method to obtain higher photocatalytic activity. In the literature, Zhu et al. [143] studied the electronic and optical properties of C-, Mo-, and (Mo,C)-co-doped anatase TiO2 using the first principle calculations. The results show that the optical absorption edges of the (Mo,C)-co-doped TiO2 will shift towards the visible-light region. Diao et al. [144] reported K, Na, and Cl co-doped rutile TiO2, exhibiting good photocatalytic degradation of gaseous formaldehyde under visible-light irradiation. Li et al. [134] reported the photocatalytic activity for hydrogen production over (B,N)-co-doped TiO2 under visible-light irradiation. N doping extends the absorption edge to the visible-light region and B doping acts as the shallow trap for photogenerated electrons to prolong the life of the electrons and holes. Consequently, stronger photocurrents were observed on (B,N)-co-doped TiO2 than those of N-doped TiO2, B-doped TiO2, and undoped TiO2 (Figure 9e). Barakat et al. [135] prepared FexCo1−x-co-doped titanium oxide nanotubes, achieving distinct enhancement of the visible-light absorption capacity (Figure 9f). Filippatos et al. [145] even reported a photocatalyst of H, F, and Cl co-doped titanium dioxide with a high hydrogen production rate.
In short, improving the photocatalytic performance of TiO2 by ion doping is mainly based on introducing defects, changing the lattice structure, and adjusting the band gap. Metal ion doping also affects the electron distribution and lattice size. Nonmetallic ions, such as N-, S-, and P-doping, generates new shallow levels by the hybridization of Ti 2p orbital with O 2p, thus narrowing the band gap of TiO2. Ion doping can be achieved through lots of ways, so it is easy to carry out in various experimental environments. However, the results may be quite different when using different doping methods.

4.2. Composite

The heterostructure formed by the recombination of two or more semiconductors with matched energy band structures can effectively improve the separation efficiency of photogenerated eCB and hVB+. As shown in Figure 10, there are usually four types of heterostructures based on different relative positions of the energy band, including type I, type II, type III, and the Z-scheme system [146,147]. The built-in electric field formed along the interface will promote the transfer of electrons. Additionally, the combination with narrow band semiconductors could allow TiO2 to respond to visible light.
For instance, Smith et al. [148] synthesized a nanotubular composite of TiO2-WO3. This composite demonstrated an increase of 46% in water-splitting efficiency compared to TiO2 nanotubes prepared under similar conditions. Choudhury et al. [149] prepared ultra-thin PdO-TiO2 composite films which could be used to photogenerate hydrogen efficiently from methanol/water for a long period of time. Navarrete et al. [150] synthesized β-Ga2O3/TiO2 composite photocatalysts for H2 production from a water/methanol mixture (Figure 11a). The high activity is attributed to the slow charge recombination of the photogenerated eCB and hVB+ (Figure 11b). Gholami et al. [151] confirmed that the activity of the ZnO-TiO2 composite for photodegradation of bentazon was better than that of ZnO and TiO2 separately. Chen et al. [152] constructed a NiO/TiO2 heterojunction on the surface of TiO2 film. The strong inner electrical field effectively separates the photogenerated electron-hole pairs, and thus the composite exhibited much better photocatalytic activity than the original TiO2 film (Figure 11c,d). The graphene-TiO2 composite has been widely studied because of its excellent mobility of charge carriers, large specific surface area, flexible structure, high transparency, and good electrical and thermal conduction [153,154,155,156,157,158,159]. Zhang et al. [157] prepared a TiO2/graphene sheet composite by a sol-gel method, exhibiting a hydrogen evolution rate of 8.6 µmol·h−1 which was nearly two times that over the commercially available Degussa P25 (4.5 µmol·h−1). Fu et al. [159] constructed a g-C3N4/graphene-CNTs/TiO2 Z-scheme photocatalytic system, in which the graphene CNTs effectively promoted the transfer of photogenerated carriers, thereby generating a stronger photocurrent (Figure 11e,f). The built-in electric field along the interface of the composite can promote the transfer of electrons, thus suppressing the recombination of photogenerated eCB and hVB+. Therefore, scientists could purposefully design the structure of composites according to the band structure of semiconductors, which can reduce the uncertainty of experiments.

4.3. Surface Noble Metal Deposition

The photogenerated carriers will be redistributed when the surface semiconductor comes into contact with metal. The electrons will transfer from the n-type semiconductor to metals because of the lower Fermi levels of metals. Moreover, the surface plasmon polaritons can enhance the light response of TiO2 [160,161,162,163]. In the literature, Zheng et al. [164] investigated the photocatalytic performance of TiO2 deposited with Au, Ag, and AuAg bimetallic nanoparticles. The results showed that the local surface plasmon resonance of noble metals improved the photocatalytic activity TiO2 under visible-light irradiation. Luo et al. [165] reported a visible-light-driven responsive Au/rGO/hydrogenated TiO2 nanotube array ternary composite with a high hydrogen evolution rate of 45 mmol·cm−2·h−1. The visible-light harvesting was significantly improved by the Au nanoparticles due to the localized surface plasmon resonance effect. Ag, Pd, and Rh have also been used to modify TiO2 by depositing them on its surface [61,166,167,168,169]. For example, Ge et al. [167] decorated Ag nanoparticles onto vertically aligned TiO2 nanotube arrays. The Ag-decorated TiO2 can efficiently drive photocatalytic water splitting under visible-light irradiation owing to the surface plasmon resonance of Ag. Due to the local surface plasmon resonance, the photocatalytic performance of noble-metal-modified TiO2 is significantly greater than that of the modified TiO2 by other methods such as ion doping and composites. However, the high cost incurred by expensive noble metals restricts the application of this strategy.

4.4. Dye Sensitization

The excitation potential of some dyes is more negative than the CB potential of TiO2. Thus, the light response range of TiO2 can be effectively expanded by dye sensitization. Dye molecules can deliver photogenerated electrons to the CB of TiO2 and then the electrons transfer further to participate in reactions [170,171,172]. For example, Shi et al. [171] prepared Eosin Y-sensitized nanosheet-stacked hollow-sphere TiO2 for efficient photocatalytic H2 production under visible-light irradiation. Vallejo et al. [170] reported the enhancement on light absorption and photocatalytic activity over rGO-TiO2 thin films after they were sensitized by natural dyes extracted from Bactris guineensis (Figure 12). In fact, lots of dyes have been used to sensitize TiO2, such as complexes of Fe (II) and polypyridyl, quinacridone, hydroxoaluminum-tricarboxymonoamide phthalocyanine, and so on [173,174,175]. Dye sensitization is easy to realize and has a low cost. Although many natural dyes can be used as raw materials for the sensitization of TiO2, sensitized TiO2 is not always stable which limits its wide application.

4.5. Loading on Supports

Loading on supports is an effective way to solve the problems of agglomeration and tough recycling of TiO2 nanoparticles. In addition, the supporting materials of high electrical conductivity could provide channels for quick transfer of electrons, thereby decreasing the recombination rate of photogenerated eCB and hVB+. For example, Li et al. [176] reported a catalyst of nitrogen-doped carbon nanofiber supporting MoS2/TiO2, in which the photogenerated electrons could quickly transfer to the carbon fiber along the basal plane of MoS2. Many natural materials are preeminent supports, so it is not essential for scientists to design and construct supports with special structures. In literature, zeolite, SiO2, and carbon materials are frequently used as the supporting materials for TiO2 [177,178,179,180,181,182]. Najafabadi et al. [180] reported four kinds of zeolites (Na-Y, Na-mordenite, H-Y, and H-beta) supporting TiO2, which all exhibited high hydrogen evolution rates. For the Na-Y zeolite supporting TiO2, the rate reached 250.8 µmol·g−1·h−1, which was almost three times that of Degussa P25 (84.2 µmol·g−1·h−1) under the same conditions. Kim et al. [182] prepared TiO2 supported by SiO2, showing much higher photocatalytic activity than pure TiO2 which could be attributed to the large specific surface area. Ti-O-Si bonds modified the narrow band gap and the local structure. Loading on supports is frequently associated with other reactions such as ion doping and forming heterojunctions. Thus, it can combine the advantages of varied strategies. Yin et al. [183] synthesized Bi plasmon-enhanced mesoporous Bi2MoO6/Ti3+ self-doped TiO2 microsphere heterojunctions. The formation of heterojunctions, Ti3+, and surface plasmon resonance (SPR) of Bi jointly achieved high catalytical activity of TiO2 under visible light. Xing et al. [184] combined ion doping with supports and synthesized a F-doped-TiO2−x/MCF composite, which exhibited high photocatalytic activity for hydrogen evolution.

4.6. Crystal Facet Engineering

The exposed facets of traditional TiO2 photocatalysts are the thermodynamically stable (101) facets. However, the specific surface energy of (001) facets is higher than that of (101) facets, implying that the (001) facets have higher reaction activity. In addition, the uncoordinated Ti5c atoms in the (001) facets can narrow the band gap of TiO2. Therefore, exposing more (001) facets will help to improve the photocatalytic performance of TiO2, which is generally realized by controlling the synthesis conditions [185,186]. For instance, Wang et al. [187] synthesized a series of (001) facet-dominated TiO2 nanosheets with high visible-light photoactivity by a simple hydrothermal method at different temperatures. Shang et al. [188] synthesized graphene-TiO2 nanocomposites with dominantly exposed (001) facets through various dosages of graphite oxide (GO) and hydrofluoric acid (HF) during a facile solvothermal process. The well-conductive and highly reactive (001) facets enhanced the photocatalytic properties and facilitated the separation of photogenerated carriers.
As a summary, Table 3 lists the hydrogen evolution efficiency from photocatalytic water splitting over typical titanium oxide-based photocatalysts. Obviously, noble-metal-modified TiO2 photocatalysts have incomparable advantages on hydrogen evolution over the other titanium oxide-based counterparts. However, the searches for alternative non-noble metals are still one of the focuses in this field because of the high cost and scarcity of noble metals. Additionally, combining multiple modification methods can achieve better results than using a single method. Table 3 lists also lists some typical non-TiO2-based photocatalysts for comparison. As can be seen, noble-metal-modified TiO2 photocatalysts obviously perform much better than metal sulfides and phosphides in hydrogen evolution reactions. The composites clearly perform better than single materials for TiO2, metal sulfides, and phosphides.

5. Conclusions and Outlooks

(1) Oxygen-deficient titanium oxide (TiO2−δ) shows higher photocatalytic activity than stoichiometric TiO2, which can be mainly attributed to the presence of Ti3+ species and oxygen deficiencies. The Ti3+ species would lead to new intermediate defect states (shallow donor) forming below the bottom of the conduction band of TiO2, which narrows the band gap of TiO2. The presence of oxygen deficiencies can decrease the transfer resistance of electrons. Resultantly, the photogenerated electrons can quickly transfer, thereby avoiding recombining with holes.
(2) Reductive treatment is the most direct and effective method to introduce oxygen defects in titanium oxides, for which H2 is the most common reductant, while other reductants such as carbon, NaBH4, and NH3 can also be selected. Moreover, ion doping, pulsed laser irradiation, calcination under anoxic conditions, plasma assistance, and so forth, have also been proven efficient strategies for introducing oxygen defects into titanium oxides. Other modification methods for TiO2, including ion doping, composite, surface noble metal deposition, dye sensitization, and loading on supports are also exploited to broaden the light-absorption region and suppress the recombination of photogenerated eCB and hVB+ for TiO2−δ. The photocatalytic activity of titanium oxides is hopefully improved further by the combination of introducing oxygen defects with these modification methods, which have reached some remarkable results.
(3) Hydrogen production by photocatalytic water splitting over TiO2−δ-based photocatalysts shows a strong development momentum. However, there exists at least three major challenges at present. The first is how to control the concentration of oxygen defects in TiO2−δ. Although the density of oxygen deficiencies can be controlled by adjusting the conditions of the reduction treatment, the spontaneously introduced oxygen defects during other modification processes, such as ion doping and surface treatment, are difficult to control and predict accurately. Secondly, current studies on regulating energy band structures mainly concentrate on enhancing light harvesting. Actually, the positions of CB and VB are also critical for photocatalytic water splitting, especially the position of CB. The CB position of TiO2 is very close to the reduction potential of H+/H2 (0 V vs. NHE at pH = 0). The decrease in the CB minimum can lead to a wider light-absorption region but the reducing ability of photogenerated electrons is also impaired at the same time. If the CB minimum is more positive than the reduction potential of H+/H2, the photocatalytic hydrogen evolution activity will take a mighty blow. Thus, regulating the band gap of TiO2 is a challenging task because there are numerous factors that can affect the band position of TiO2−δ during the modifying process. Combining theoretical calculation prediction with precise control of synthesis conditions may be a solution to solve this issue. In addition, the current studies pay little attention to the adsorption of reactants (H2O) and the desorption of products (H2 and O2). The dissolved O2 and H2 can react with each other at the cocatalyst surface. O2 dissolved in water will also compete photogenerated electrons with the hydrogen evolution reaction. These factors weaken the efficiency of photocatalytic hydrogen evolution. Therefore, this might be the next hot topic in studies of this nature. Although photocatalytic hydrogen evolution remains in the laboratory stage, further study may bring promising results.

Author Contributions

Y.C.: literature surveying, drawing of figures and tables, sorting data, writing of original manuscript; X.F.: literature surveying, revision and finalizing of the manuscript; Z.P.: revision and finalizing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Many thanks for the financial support for this work by the National Natural Science Foundation of China (grant nos. 12174035 and 61274015).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable. No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sial, M.H.; Arshed, N.; Amjad, M.A.; Khan, Y.A. Nexus between fossil fuel consumption and infant mortality rate: A non-linear analysis. Environ. Sci. Pollut. Res. 2022, 29, 58378–58387. [Google Scholar] [CrossRef] [PubMed]
  2. Vohra, K.; Vodonos, A.; Schwartz, J.; Marais, E.A.; Sulprizio, M.P.; Mickley, L.J. Global mortality from outdoor fine particle pollution generated by fossil fuel combustion: Results from GEOS-Chem. Environ. Res. 2021, 195, 110754. [Google Scholar] [CrossRef]
  3. Aslan, M.; Isik, H. Green energy for the battlefield. Int. J. Green Energy 2017, 14, 1020–1026. [Google Scholar] [CrossRef]
  4. Koroneos, C.; Spachos, T.; Moussiopoulos, N. Exergy analysis of renewable energy sources. Renew. Energy 2003, 28, 295–310. [Google Scholar] [CrossRef]
  5. Kumar, G.; Kim, S.-H.; Lay, C.-H.; Ponnusamy, V.K. Recent developments on alternative fuels, energy and environment for sustainability Preface. Bioresour. Technol. 2020, 317, 124010. [Google Scholar] [CrossRef] [PubMed]
  6. Veziroglu, T.N.; Sahin, S. 21st Century’s energy: Hydrogen energy system. Energy Convers. Manag. 2008, 49, 1820–1831. [Google Scholar] [CrossRef]
  7. Sharma, S.; Agarwal, S.; Jain, A. Significance of Hydrogen as Economic and Environmentally Friendly Fuel. Energies 2021, 14, 7389. [Google Scholar] [CrossRef]
  8. Muradov, N. Low to near-zero CO2 production of hydrogen from fossil fuels: Status and perspectives. Int. J. Hydrogen Energy 2017, 42, 14058–14088. [Google Scholar] [CrossRef]
  9. Qureshi, F.; Yusuf, M.; Ibrahim, H.; Kamyab, H.; Chelliapan, S.; Pham, C.Q.; Vo, D.V.N. Contemporary avenues of the Hydrogen industry: Opportunities and challenges in the eco-friendly approach. Environ. Res. 2023, 229, 115963. [Google Scholar] [CrossRef]
  10. Sapountzi, F.M.; Gracia, J.M.; Weststrate, C.J.; Fredriksson, H.O.A.; Niemantsverdriet, J.W. Electrocatalysts for the generation of hydrogen, oxygen and synthesis gas. Prog. Energy Combust. Sci. 2017, 58, 1–35. [Google Scholar] [CrossRef] [Green Version]
  11. Tee, S.Y.; Win, K.Y.; Teo, W.S.; Koh, L.-D.; Liu, S.; Teng, C.P.; Han, M.-Y. Recent Progress in Energy-Driven Water Splitting. Adv. Sci. 2017, 4, 1600337. [Google Scholar] [CrossRef] [Green Version]
  12. Ismail, A.A.; Bahnemann, D.W. Photochemical splitting of water for hydrogen production by photocatalysis: A review. Sol. Energy Mater. Sol. Cells 2014, 128, 85–101. [Google Scholar] [CrossRef]
  13. Dean, J.A.; Wei, J. Lange’s Handbook of Chemistry, 2nd ed.; Science Press: Beijing, China, 2003. [Google Scholar]
  14. Djurisic, A.B.; He, Y.; Ng, A.M.C. Visible-light photocatalysts: Prospects and challenges. Apl Mater. 2020, 8, 033001. [Google Scholar] [CrossRef] [Green Version]
  15. Long, X.; Wei, X.; Qiu, Y.; Song, Y.; Bi, L.; Tang, P.; Yan, X.; Wang, S.; Liao, J. TiO2 aerogel composite high-efficiency photocatalysts for environmental treatment and hydrogen energy production. Nanotechnol. Rev. 2023, 12, 20220490. [Google Scholar] [CrossRef]
  16. Nur, A.S.M.; Sultana, M.; Mondal, A.; Islam, S.; Robel, F.N.; Islam, A.; Sumi, M.S.A. A review on the development of elemental and codoped TiO2 photocatalysts for enhanced dye degradation under UV-vis irradiation. J. Water Process Eng. 2022, 47, 102728. [Google Scholar] [CrossRef]
  17. Zhang, T.; Han, X.; Nguyen, N.T.; Yang, L.; Zhou, X. TiO2-based photocatalysts for CO2 reduction and solar fuel generation. Chin. J. Catal. 2022, 43, 2500–2529. [Google Scholar] [CrossRef]
  18. Wang, J.; Wang, Z.; Wang, W.; Wang, Y.; Hu, X.; Liu, J.; Gong, X.; Miao, W.; Ding, L.; Li, X.; et al. Synthesis, modification and application of titanium dioxide nanoparticles: A review. Nanoscale 2022, 14, 6709–6734. [Google Scholar] [CrossRef]
  19. Feng, J.; Pan, L.; Liu, H.; Yuan, S.; Zhang, L.; Yin, H.; Song, H.; Li, L. Synergistic degradation of the aqueous antibiotic norfloxacin by nonthermal plasma combined with defective titanium dioxide exposed {001} facets. Sep. Purif. Technol. 2022, 300, 121761. [Google Scholar] [CrossRef]
  20. Zhu, C.; Shen, Y.; Yang, F.; Zhu, P.; An, C. Engineering hydrogenated TiO2 nanosheets by rational deposition of Ni clusters and Pt single atoms onto exposing facets for high-performance solar fuel production. Chem. Eng. J. 2023, 466, 143174. [Google Scholar] [CrossRef]
  21. Yang, Y.; Yin, L.-C.; Gong, Y.; Niu, P.; Wang, J.-Q.; Gu, L.; Chen, X.; Liu, G.; Wang, L.; Cheng, H.-M. An Unusual Strong Visible-Light Absorption Band in Red Anatase TiO2 photocatalyst induced by atomic hydrogen-occupied oxygen vacancies. Adv. Mater. 2018, 30, 1704479. [Google Scholar] [CrossRef]
  22. Cushing, S.K.; Meng, F.; Zhang, J.; Ding, B.; Chen, C.K.; Chen, C.-J.; Liu, R.-S.; Bristow, A.D.; Bright, J.; Zheng, P.; et al. Effects of defects on photocatalytic activity of hydrogen-treated titanium oxide nanobelts. ACS Catal. 2017, 7, 1742–1748. [Google Scholar] [CrossRef] [Green Version]
  23. Hong, R.; Deng, C.; Jing, M.; Lin, H.; Tao, C.; Zhang, D. Oxygen flows-dependent photocatalytic performance in Ti3+ doped TiO2 thin films. Opt. Mater. 2019, 95, 109224. [Google Scholar]
  24. Liu, L.; Liu, J.; Zong, S.; Huang, Z.; Feng, X.; Zheng, J.; Fang, Y. Facile synthesis of nitrogen-doped TiO2 microspheres containing oxygen vacancies with excellent photocatalytic H2 evolution activity. J. Phys. Chem. Solids 2022, 170, 110930. [Google Scholar] [CrossRef]
  25. Zhao, Z.; Tan, H.; Zhao, H.; Lv, Y.; Zhou, L.-J.; Song, Y.; Sun, Z. Reduced TiO2 rutile nanorods with well-defined facets and their visible-light photocatalytic activity. Chem. Commun. 2014, 50, 2755–2757. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, Y.Q.; Yu, X.J.; Sun, D.Z. Synthesis, characterization, and photocatalytic activity of TiO2-xNx nanocatalyst. J. Hazard. Mater. 2007, 144, 328–333. [Google Scholar] [CrossRef] [PubMed]
  27. Yuan, J.; Liu, Y.; Bo, T.; Zhou, W. Activated HER performance of defected single layered TiO2 nanosheet via transition metal doping. Int. J. Hydrogen Energy 2020, 45, 2681–2688. [Google Scholar] [CrossRef]
  28. Li, R.; Luan, Q.; Dong, C.; Dong, W.; Tang, W.; Wang, G.; Lu, Y. Light-facilitated structure reconstruction on self-optimized photocatalyst TiO2@BiOCl for selectively efficient conversion of CO2 to CH4. Appl. Catal. B-Environ. 2021, 286, 119832. [Google Scholar] [CrossRef]
  29. Zhang, Y.-P.; Han, W.; Yang, Y.; Zhang, H.-Y.; Wang, Y.; Wang, L.; Sun, X.-J.; Zhang, F.-M. S-scheme heterojunction of black TiO2 and covalent-organic framework for enhanced evolution. Chem. Eng. J. 2022, 446, 137213. [Google Scholar] [CrossRef]
  30. Rajeshwar, K. Hydrogen generation at irradiated oxide semiconductor-solution interfaces. J. Appl. Electrochem. 2007, 37, 765–787. [Google Scholar] [CrossRef]
  31. Lettieri, S.; Pavone, M.; Fioravanti, A.; Santamaria Amato, L.; Maddalena, P. Charge carrier processes and optical properties in TiO2 and TiO2-based heterojunction photocatalysts: A review. Materials 2021, 14, 1645. [Google Scholar] [CrossRef] [PubMed]
  32. Nakano, T.; Ito, R.; Kogoshi, S.; Katayama, N. Optimal levels of oxygen deficiency in the visible light photocatalyst TiO2-x and long-term stability of catalytic performance. J. Phys. Chem. Solids 2016, 98, 136–142. [Google Scholar] [CrossRef]
  33. Wu, H.; Wang, Z.; Jin, S.; Cao, X.; Ren, F.; Wu, L.; Xing, Z.; Wang, X.; Cai, G.; Jiang, C. Enhanced photoelectrochemical performance of TiO2 through controlled Ar+ ion irradiation: A combined experimental and theoretical study. Int. J. Hydrogen Energy 2018, 43, 6936–6944. [Google Scholar] [CrossRef]
  34. Tian, M.; Mahjouri-Samani, M.; Eres, G.; Sachan, R.; Yoon, M.; Chisholm, M.F.; Wang, K.; Puretzky, A.A.; Rouleau, C.M.; Geohegan, D.B.; et al. Structure and formation mechanism of black TiO2 nanoparticles. ACS Nano 2015, 9, 10482–10488. [Google Scholar] [CrossRef]
  35. Amano, F.; Nakata, M.; Yamamoto, A.; Tanaka, T. Effect of Ti3+ ions and conduction band electrons on photocatalytic and photoelectrochemical activity of rutile titania for water oxidation. J. Phys. Chem. C 2016, 120, 6467–6474. [Google Scholar] [CrossRef]
  36. Singh, A.P.; Kodan, N.; Mehta, B.R. Enhancing the photoelectrochemical properties of titanium dioxide by thermal treatment in oxygen deficient environment. Appl. Surf. Sci. 2016, 372, 63–69. [Google Scholar] [CrossRef]
  37. Xiu, Z.; Guo, M.; Zhao, T.; Pan, K.; Xing, Z.; Li, Z.; Zhou, W. Recent advances in Ti3+ self-doped nanostructured TiO2 visible light photocatalysts for environmental and energy applications. Chem. Eng. J. 2020, 382, 123011. [Google Scholar] [CrossRef]
  38. Li, Y.; Cooper, J.K.; Liu, W.; Sutter-Fella, C.M.; Amani, M.; Beeman, J.W.; Javey, A.; Ager, J.W.; Liu, Y.; Toma, F.M.; et al. Defective TiO2 with high photoconductive gain for efficient and stable planar heterojunction perovskite solar cells. Nat. Commun. 2016, 7, 12446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Hao, Z.; Chen, Q.; Dai, W.; Ren, Y.; Zhou, Y.; Yang, J.; Xie, S.; Shen, Y.; Wu, J.; Chen, W.; et al. Oxygen-deficient blue TiO2 for ultrastable and fast lithium storage. Adv. Energy Mater. 2020, 10, 1903107. [Google Scholar] [CrossRef]
  40. Wendt, S.; Sprunger, P.T.; Lira, E.; Madsen, G.K.H.; Li, Z.; Hansen, J.O.; Matthiesen, J.; Blekinge-Rasmussen, A.; Laegsgaard, E.; Hammer, B. The role of interstitial sites in the Ti 3d defect state in the band gap of Titania. Science 2008, 320, 1755–1759. [Google Scholar] [CrossRef] [Green Version]
  41. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  42. Long, Z.; Li, Q.; Wei, T.; Zhang, G.; Ren, Z. Historical development and prospects of photocatalysts for pollutant removal in water. J. Hazard. Mater. 2020, 395, 122599. [Google Scholar] [CrossRef]
  43. Matthews, R.W. Photooxidation of organic impurities in water using thin-films of titanium-dioxide. J. Phys. Chem. 1987, 91, 3328–3333. [Google Scholar] [CrossRef]
  44. Chan, S.H.S.; Wu, T.Y.; Juan, J.C.; Teh, C.Y. Recent developments of metal oxide semiconductors as photocatalysts in advanced oxidation processes (AOPs) for treatment of dye waste-water. J. Chem. Technol. Biotechnol. 2011, 86, 1130–1158. [Google Scholar] [CrossRef]
  45. Pan, L.; Ai, M.; Huang, C.; Yin, L.; Liu, X.; Zhang, R.; Wang, S.; Jiang, Z.; Zhang, X.; Zou, J.-J.; et al. Manipulating spin polarization of titanium dioxide for efficient photocatalysis. Nat. Commun. 2020, 11, 418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Pennington, A.M.; Yang, R.A.; Munoz, D.T.; Celik, F.E. Metal-free hydrogen evolution over defect-rich anatase titanium dioxide. Int. J. Hydrogen Energy 2018, 43, 15176–15190. [Google Scholar] [CrossRef]
  47. Yamazaki, Y.; Mori, K.; Kuwahara, Y.; Kobayashi, H.; Yamashita, H. Defect engineering of Pt/TiO2-x photocatalysts via reduction treatment assisted by hydrogen spillover. ACS Appl. Mater. Interfaces 2021, 13, 48669–48678. [Google Scholar] [CrossRef]
  48. Zhang, Y.-C.; Afzal, N.; Pan, L.; Zhang, X.; Zou, J.-J. Structure-activity relationship of defective metal-based photocatalysts for water splitting: Experimental and theoretical perspectives. Adv. Sci. 2019, 6, 1900053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Sasikala, R.; Sudarsan, V.; Sudakar, C.; Naik, R.; Sakuntala, T.; Bharadwaj, S.R. Enhanced photocatalytic hydrogen evolution over nanometer sized Sn and Eu doped titanium oxide. Int. J. Hydrogen Energy 2008, 33, 4966–4973. [Google Scholar] [CrossRef]
  50. Liu, G.; Yang, H.G.; Wang, X.; Cheng, L.; Lu, H.; Wang, L.; Lu, G.Q.; Cheng, H.-M. Enhanced photoactivity of oxygen-deficient anatase TiO2 sheets with dominant {001} facets. J. Phys. Chem. C 2009, 113, 21784–21788. [Google Scholar] [CrossRef]
  51. Amano, F.; Nakata, M. High-temperature calcination and hydrogen reduction of rutile TiO2: A method to improve the photocatalytic activity for water oxidation. Appl. Catal. B-Environ. 2014, 158, 202–208. [Google Scholar] [CrossRef]
  52. Mali, M.G.; Yoon, H.; An, S.; Choi, J.-Y.; Kim, H.-Y.; Lee, B.C.; Kim, B.N.; Park, J.H.; Al-Deyab, S.S.; Yoon, S.S. Enhanced solar water splitting of electron beam irradiated titania photoanode by electrostatic spray deposition. Appl. Surf. Sci. 2014, 319, 205–210. [Google Scholar] [CrossRef]
  53. Pitchaimuthu, S.; Honda, K.; Suzuki, S.; Naito, A.; Suzuki, N.; Katsumata, K.-I.; Nakata, K.; Ishida, N.; Kitamura, N.; Idemoto, Y.; et al. Solution plasma process-derived defect-induced heterophase anatase/brookite TiO2 nanocrystals for enhanced gaseous photocatalytic performance. ACS Omega 2018, 3, 898–905. [Google Scholar] [CrossRef] [Green Version]
  54. Ennaceri, H.; Boujnah, M.; Taleb, A.; Khaldoun, A.; Saez-Araoz, R.; Ennaoui, A.; El Kenz, A.; Benyoussef, A. Thickness effect on the optical properties of TiO2-anatase thin films prepared by ultrasonic spray pyrolysis: Experimental and ab initio study. Int. J. Hydrogen Energy 2017, 42, 19467–19480. [Google Scholar] [CrossRef]
  55. Nakajima, T.; Nakamura, T.; Shinoda, K.; Tsuchiya, T. Rapid formation of black titania photoanodes: Pulsed laser-induced oxygen release and enhanced solar water splitting efficiency. J. Mater. Chem. A 2014, 2, 6762–6771. [Google Scholar] [CrossRef]
  56. Chen, Y.; Cao, X.; Lin, B.; Gao, B. Origin of the visible-light photoactivity of NH3-treated TiO2: Effect of nitrogen doping and oxygen vacancies. Appl. Surf. Sci. 2013, 264, 845–852. [Google Scholar] [CrossRef]
  57. Zhang, Y.-X.; Wu, S.-M.; Tian, G.; Zhao, X.-F.; Wang, L.-Y.; Yin, Y.-X.; Wu, L.; Li, Q.-N.; Zhang, Y.-X.; Wu, J.-S.; et al. Titanium vacancies in TiO2 nanofibers enable highly efficient photodriven seawater splitting. Chem.-A Eur. J. 2021, 27, 14202–14208. [Google Scholar] [CrossRef] [PubMed]
  58. An, X.; Hu, C.; Liu, H.; Qu, J. Oxygen vacancy mediated construction of anatase/brookite heterophase junctions for high-efficiency photocatalytic hydrogen evolution. J. Mater. Chem. A 2017, 5, 24989–24994. [Google Scholar] [CrossRef]
  59. Yuan, D.; Jiao, Y.; Li, Z.; Chen, X.; Ding, J.; Dai, W.-L.; Wan, H.; Guan, G. TiN bridged all-solid Z-Scheme CNNS/TiN/TiO2-x heterojunction by a facile in situ reduction strategy for enhanced photocatalytic hydrogen evolution. Adv. Mater. Interfaces 2021, 8, 2100695. [Google Scholar] [CrossRef]
  60. Park, E.; Patil, S.S.; Lee, H.; Kumbhar, V.S.; Lee, K. Photoelectrochemical H2 evolution on WO3/BiVO4 enabled by single-crystalline TiO2 overlayer modulations. Nanoscale 2021, 13, 16932–16941. [Google Scholar] [CrossRef]
  61. Balayeva, N.O.; Mamiyev, Z.; Dillert, R.; Zheng, N.; Bahnemann, D.W. Rh/TiO2-photocatalyzed acceptorless dehydrogenation of N-heterocycles upon visible-light illumination. ACS Catal. 2020, 10, 5542–5553. [Google Scholar] [CrossRef]
  62. Mo, L.-B.; Bai, Y.; Xiang, Q.-Y.; Li, Q.; Wang, J.-O.; Ibrahim, K.; Cao, J.-L. Band gap engineering of TiO2 through hydrogenation. Appl. Phys. Lett. 2014, 105, 202114. [Google Scholar] [CrossRef]
  63. Wang, H.; Wang, G.; Ling, Y.; Lepert, M.; Wang, C.; Zhang, J.Z.; Li, Y. Photoelectrochemical study of oxygen deficient TiO2 nanowire arrays with CdS quantum dot sensitization. Nanoscale 2012, 4, 1463–1466. [Google Scholar] [CrossRef] [PubMed]
  64. Liu, N.; Zhou, X.; Nhat Truong, N.; Peters, K.; Zoller, F.; Hwang, I.; Schneider, C.; Miehlich, M.E.; Freitag, D.; Meyer, K.; et al. Black Magic in Gray Titania: Noble-Metal-Free Photocatalytic H2 evolution from hydrogenated anatase. ChemSusChem 2017, 10, 62–67. [Google Scholar] [CrossRef] [PubMed]
  65. Abdelmaksoud, M.K.; Sayed, A.; Sayed, S.; Abbas, M. A novel solar radiation absorption enhancement of TiO2 nanomaterial by a simple hydrogenation method. J. Mater. Res. 2021, 36, 2118–2131. [Google Scholar] [CrossRef]
  66. Xu, Y.F.; Zhang, C.; Zhang, L.X.; Zhang, X.H.; Yao, H.L.; Shi, J.L. Pd-catalyzed instant hydrogenation of TiO2 with enhanced photocatalytic performance. Energy Environ. Sci. 2016, 9, 2410–2417. [Google Scholar] [CrossRef]
  67. Zhang, J.W.; Wang, S.; Liu, F.S.; Fu, X.J.; Ma, G.Q.; Hou, M.S.; Tang, Z. Preparation of defective TiO2-x hollow microspheres for photocatalytic degradation of methylene blue. Acta Phys.-Chim. Sin. 2019, 35, 885–895. [Google Scholar] [CrossRef]
  68. Wierzbicka, E.; Altomare, M.; Wu, M.; Liu, N.; Yokosawa, T.; Fehn, D.; Qin, S.; Meyer, K.; Unruh, T.; Spiecker, E.; et al. Reduced grey brookite for noble metal free photocatalytic H2 evolution. J. Mater. Chem. A 2021, 9, 1168–1179. [Google Scholar] [CrossRef]
  69. Samsudin, E.M.; Hamid, S.B.A.; Juan, J.C.; Basirun, W.J.; Kandjani, A.E. Surface modification of mixed-phase hydrogenated TiO2 and corresponding photocatalytic response. Appl. Surf. Sci. 2015, 359, 883–896. [Google Scholar] [CrossRef]
  70. Ihara, T.; Miyoshi, M.; Iriyama, Y.; Matsumoto, O.; Sugihara, S. Visible-light-active titanium oxide photocatalyst realized by an oxygen-deficient structure and by nitrogen doping. Appl. Catal. B-Environ. 2003, 42, 403–409. [Google Scholar] [CrossRef]
  71. Qiu, J.-Y.; Feng, H.-Z.; Chen, Z.-H.; Ruan, S.-H.; Chen, Y.-P.; Xu, T.-T.; Su, J.-Y.; Ha, E.-N.; Wang, L.-Y. Selective introduction of surface defects in anatase TiO2 nanosheets for highly efficient photocatalytic hydrogen generation. Rare Met. 2022, 41, 2074–2083. [Google Scholar] [CrossRef]
  72. Guan, S.; Hao, L.; Lu, Y.; Yoshida, H.; Pan, F.; Asanuma, H. Fabrication of oxygen-deficient TiO2 coatings with nano-fiber morphology for visible-light photocatalysis. Mater. Sci. Semicond. Process. 2016, 41, 358–363. [Google Scholar] [CrossRef]
  73. Zhao, H.; Chen, J.; Rao, G.; Deng, W.; Li, Y. Enhancing photocatalytic CO2 reduction by coating an ultrathin Al2O3 layer on oxygen deficient TiO2 nanorods through atomic layer deposition. Appl. Surf. Sci. 2017, 404, 49–56. [Google Scholar] [CrossRef] [Green Version]
  74. Martinez-Oviedo, A.; Ray, S.K.; Hoang Phuc, N.; Lee, S.W. Efficient photo-oxidation of NOx by Sn doped blue TiO2 nanoparticles. J. Photochem. Photobiol. A-Chem. 2019, 370, 18–25. [Google Scholar] [CrossRef]
  75. Pereira, A.L.J.; Lisboa Filho, P.N.; Acuna, J.; Brandt, I.S.; Pasa, A.A.; Zanatta, A.R.; Vilcarromero, J.; Beltran, A.; Dias da Silva, J.H. Enhancement of optical absorption by modulation of the oxygen flow of TiO2 films deposited by reactive sputtering. J. Appl. Phys. 2012, 111, 113513. [Google Scholar] [CrossRef] [Green Version]
  76. Dhumal, S.Y.; Daulton, T.L.; Jiang, J.; Khomami, B.; Biswas, P. Synthesis of visible light-active nanostructured TiOx (x < 2) photocatalysts in a flame aerosol reactor. Appl. Catal. B-Environ. 2009, 86, 145–151. [Google Scholar]
  77. Xiao, P.; Liu, D.; Garcia, B.B.; Sepehri, S.; Zhang, Y.; Cao, G. Electrochemical and photoelectrical properties of titania nanotube arrays annealed in different gases. Sens. Actuators B-Chem. 2008, 134, 367–372. [Google Scholar] [CrossRef]
  78. Kushwaha, S.; Nagarajan, R. Black TiO2-graphitic carbon nanocomposite from a single source precursor and its interaction with colored and colorless contaminants under visible radiation. Mater. Res. Bull. 2020, 132, 110983. [Google Scholar] [CrossRef]
  79. Starbova, K.; Yordanova, V.; Nihtianova, D.; Hintz, W.; Tomas, J.; Starbov, N. Excimer laser processing as a tool for photocatalytic design of sol-gel TiO2 thin films. Appl. Surf. Sci. 2008, 254, 4044–4051. [Google Scholar] [CrossRef]
  80. Lau, M.; Reichenberger, S.; Haxhiaj, I.; Barcikowski, S.; Mueller, A.M. Mechanism of laser-induced bulk and surface defect generation in ZnO and TiO2 nanoparticles: Effect on photoelectrochemical performance. ACS Appl. Energy Mater. 2018, 1, 5366–5385. [Google Scholar] [CrossRef] [Green Version]
  81. Nair, P.R.; Ramirez, C.R.S.; Pinilla, M.A.G.; Krishnan, B.; Avellaneda, D.A.; Pelaes, R.F.C.; Shaji, S. Black titanium dioxide nanocolloids by laser irradiation in liquids for visible light photo-catalytic/electrochemical applications. Appl. Surf. Sci. 2023, 623, 157096. [Google Scholar] [CrossRef]
  82. Wang, Z.; Yang, C.; Lin, T.; Yin, H.; Chen, P.; Wan, D.; Xu, F.; Huang, F.; Lin, J.; Xie, X.; et al. H-doped black titania with very high solar absorption and excellent photocatalysis enhanced by localized surface plasmon resonance. Adv. Funct. Mater. 2013, 23, 5444–5450. [Google Scholar] [CrossRef]
  83. Filice, S.; Fiorenza, R.; Reitano, R.; Scalese, S.; Scire, S.; Fisicaro, G.; Deretzis, I.; La Magna, A.; Bongiorno, C.; Compagnini, G. TiO2 colloids laser-treated in ethanol for photocatalytic H2 production. ACS Appl. Nano Mater. 2020, 3, 9127–9140. [Google Scholar] [CrossRef]
  84. Fiorenza, R.; Scire, S.; D’Urso, L.; Compagnini, G.; Bellardita, M.; Palmisano, L. Efficient H2 production by photocatalytic water splitting under UV or solar light over variously modified TiO2-based catalysts. Int. J. Hydrogen Energy 2019, 44, 14796–14807. [Google Scholar] [CrossRef]
  85. Nakajima, T.; Tsuchiya, T.; Kumagai, T. Pulsed laser-induced oxygen deficiency at TiO2 surface: Anomalous structure and electrical transport properties. J. Solid State Chem. 2009, 182, 2560–2565. [Google Scholar] [CrossRef]
  86. Leichtweiss, T.; Henning, R.A.; Koettgen, J.; Schmidt, R.M.; Hollaender, B.; Martin, M.; Wuttig, M.; Janek, J. Amorphous and highly nonstoichiometric titania (TiOx) thin films close to metal-like conductivity. J. Mater. Chem. A 2014, 2, 6631–6640. [Google Scholar] [CrossRef] [Green Version]
  87. Kunti, A.K.; Chowdhury, M.; Sharma, S.K.; Gupta, M.; Chaudhary, R.J. Influence of O2 pressure on structural, morphological and optical properties of TiO2-SiO2 composite thin films prepared by pulsed laser deposition. Thin Solid Film. 2017, 629, 79–89. [Google Scholar] [CrossRef]
  88. Ali, N.; Bashir, S.; Umm-i-Kalsoom; Akram, M.; Mahmood, K. Effect of dry and wet ambient environment on the pulsed laser ablation of titanium. Appl. Surf. Sci. 2013, 270, 49–57. [Google Scholar] [CrossRef]
  89. Davila, Y.; Petitmangin, A.; Hebert, C.; Perriere, J.; Seiler, W. Oxygen deficiency in oxide films grown by PLD. Appl. Surf. Sci. 2011, 257, 5354–5357. [Google Scholar] [CrossRef]
  90. Socol, G.; Gnatyuk, Y.; Stefan, N.; Smirnova, N.; Djokic, V.; Sutan, C.; Malinovschi, V.; Stanculescu, A.; Korduban, O.; Mihailescu, I.N. Photocatalytic activity of pulsed laser deposited TiO2 thin films in N2, O2 and CH4. Thin Solid Film. 2010, 518, 4648–4653. [Google Scholar] [CrossRef]
  91. Nath, A.; Laha, S.S.; Khare, A. Effect of focusing conditions on synthesis of titanium oxide nanoparticles via laser ablation in titanium-water interface. Appl. Surf. Sci. 2011, 257, 3118–3122. [Google Scholar] [CrossRef]
  92. Rahman, M.A.; Bazargan, S.; Srivastava, S.; Wang, X.; Abd-Ellah, M.; Thomas, J.P.; Heinig, N.F.; Pradhan, D.; Leung, K.T. Defect-rich decorated TiO2 nanowires for super-efficient photoelectrochemical water splitting driven by visible light. Energy Environ. Sci. 2015, 8, 3363–3373. [Google Scholar] [CrossRef]
  93. Bellardita, M.; Garlisi, C.; Ozer, L.Y.; Venezia, A.M.; Sa, J.; Mamedov, F.; Palmisano, L.; Palmisano, G. Highly stable defective TiO2-x with tuned exposed facets induced by fluorine: Impact of surface and bulk properties on selective UV/visible alcohol photo-oxidation. Appl. Surf. Sci. 2020, 510, 145419. [Google Scholar] [CrossRef]
  94. Lo, H.-H.; Gopal, N.O.; Ke, S.-C. Origin of photoactivity of oxygen-deficient TiO2 under visible light. Appl. Phys. Lett. 2009, 95, 083126. [Google Scholar] [CrossRef]
  95. Pu, X.; Hu, Y.; Cui, S.; Cheng, L.; Jiao, Z. Preparation of N-doped and oxygen-deficient TiO2 microspheres via a novel electron beam-assisted method. Solid State Sci. 2017, 70, 66–73. [Google Scholar] [CrossRef]
  96. Wang, M.; Xu, X.Y.; Lin, L.; He, D.N. Gd-La codoped TiO2 nanoparticles as solar photocatalysts. Prog. Nat. Sci.-Mater. Int. 2015, 25, 6–11. [Google Scholar] [CrossRef] [Green Version]
  97. Kunti, A.K.; Sharma, S.K. Structural and spectral properties of red light emitting Eu3+ activated TiO2 nanophosphor for white LED application. Ceram. Int. 2017, 43, 9838–9845. [Google Scholar] [CrossRef]
  98. Zhang, J.; Zhang, J.; Ren, H.; Yu, L.; Wu, Z.; Zhang, Z. High rate capability and long cycle stability of TiO2-delta-La composite nanotubes as anode material for lithium ion batteries. J. Alloys Compd. 2014, 609, 178–184. [Google Scholar] [CrossRef]
  99. Zhang, J.; Zhao, Z.; Wang, X.; Yu, T.; Guan, J.; Yu, Z.; Li, Z.; Zou, Z. Increasing the oxygen vacancy density on the TiO2 surface by La-doping for dye-sensitized solar cells. J. Phys. Chem. C 2010, 114, 18396–18400. [Google Scholar] [CrossRef]
  100. Aguinaco, A.; Amaya, B.; Ramirez-del-Solar, M. Facile fabrication of Fe-TiO2 thin film and its photocatalytic activity. Environ. Sci. Pollut. Res. 2022, 29, 23292–23302. [Google Scholar] [CrossRef] [PubMed]
  101. Bhowmick, S.; Saini, C.P.; Santra, B.; Walczak, L.; Semisalova, A.; Gupta, M.; Kanjilal, A. Modulation of the work function of TiO2 nanotubes by nitrogen doping: Implications for the photocatalytic degradation of dyes. ACS Appl. Nano Mater. 2023, 6, 50–60. [Google Scholar] [CrossRef]
  102. Hatanaka, Y.; Naito, H.; Itou, S.; Kando, M. Photocatalytic characteristics of hydro-oxygenated amorphous titanium oxide films prepared using remote plasma enhanced chemical vapor deposition. Appl. Surf. Sci. 2005, 244, 554–557. [Google Scholar] [CrossRef]
  103. Sakai, T.; Kuniyoshi, Y.; Aoki, W.; Ezoe, S.; Endo, T.; Hoshi, Y. High-rate deposition of photocatalytic TiO2 films by oxygen plasma assist reactive evaporation method. Thin Solid Film. 2008, 516, 5860–5863. [Google Scholar] [CrossRef]
  104. Li, Y.; Wang, W.; Wang, F.; Di, L.; Yang, S.; Zhu, S.; Yao, Y.; Ma, C.; Dai, B.; Yu, F. Enhanced photocatalytic degradation of organic dyes via defect-rich TiO2 prepared by dielectric barrier discharge plasma. Nanomaterials 2019, 9, 720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Hojo, M.; Okimura, K. Effect of annealing with Ar plasma irradiation for transparent conductive Nb-doped TiO2 films on glass substrate. Jpn. J. Appl. Phys. 2009, 48, 08hk06. [Google Scholar] [CrossRef]
  106. Kawakami, R.; Mimoto, Y.; Yanagiya, S.-I.; Shirai, A.; Niibe, M.; Nakano, Y.; Mukai, T. Photocatalytic activity enhancement of anatase/rutile-mixed phase TiO2 nanoparticles annealed with low-temperature O2 plasma. Phys. Status Solidi A-Appl. Mater. Sci. 2021, 218, 2100536. [Google Scholar] [CrossRef]
  107. An, H.-R.; Hong, Y.C.; Kim, H.; Huh, J.Y.; Park, E.C.; Park, S.Y.; Jeong, Y.; Park, J.-I.; Kim, J.-P.; Lee, Y.-C.; et al. Studies on mass production and highly solar light photocatalytic properties of gray hydrogenated-TiO2 sphere photocatalysts. J. Hazard. Mater. 2018, 358, 222–233. [Google Scholar] [CrossRef]
  108. Mizukoshi, Y.; Ohwada, M.; Seino, S.; Horibed, H.; Nishimura, Y.; Terashima, C. Synthesis of oxygen-deficient blue titanium oxide by discharge plasma generated in aqueous ammonia solution. Appl. Surf. Sci. 2019, 489, 255–261. [Google Scholar] [CrossRef]
  109. Ji, M.; Choa, Y.-H.; Lee, Y.-I. One-step synthesis of black TiO2-x microspheres by ultrasonic spray pyrolysis process and their visible-light-driven photocatalytic activities. Ultrason. Sonochem. 2021, 74, 105557. [Google Scholar] [CrossRef]
  110. Nakaruk, A.; Reece, P.J.; Ragazzon, D.; Sorrell, C.C. TiO2 films prepared by ultrasonic spray pyrolysis. Mater. Sci. Technol. 2010, 26, 469–472. [Google Scholar] [CrossRef]
  111. da Silva, A.L.; Trindade, F.J.; Dalmasso, J.-L.; Ramos, B.; Teixeira, A.C.S.C.; Gouvea, D. Synthesis of TiO2 microspheres by ultrasonic spray pyrolysis and photocatalytic activity evaluation. Ceram. Int. 2022, 48, 9739–9745. [Google Scholar] [CrossRef]
  112. Nguyen Thi Khanh, V.; Nguyen Nang, D.; Nguyen Van, C.; Nguyen Nhat, H.; Nguyen Thanh, T.; Tran Quoc, T.; Dang Van, T. A simple and efficient ultrasonic-assisted electrochemical approach for scalable production of nitrogen-doped TiO2 nanocrystals. Nanotechnology 2021, 32, 465602. [Google Scholar] [CrossRef] [PubMed]
  113. Osorio-Vargas, P.A.; Pulgarin, C.; Sienkiewicz, A.; Pizzio, L.R.; Blanco, M.N.; Torres-Palma, R.A.; Petrier, C.; Rengifo-Herrera, J.A. Low-frequency ultrasound induces oxygen vacancies formation and visible light absorption in TiO2 P-25 nanoparticles. Ultrason. Sonochem. 2012, 19, 383–386. [Google Scholar] [CrossRef]
  114. Bellardita, M.; El Nazer, H.A.; Loddo, V.; Parrino, F.; Venezia, A.M.; Palmisano, L. Photoactivity under visible light of metal loaded TiO2 catalysts prepared by low frequency ultrasound treatment. Catal. Today 2017, 284, 92–99. [Google Scholar] [CrossRef]
  115. Langhammer, D.; Thyr, J.; Osterlund, L. Surface properties of reduced and stoichiometric TiO2 as probed by SO2 adsorption. J. Phys. Chem. C 2019, 123, 24549–24557. [Google Scholar] [CrossRef]
  116. Xu, M.; Chen, Y.; Qin, J.; Feng, Y.; Li, W.; Chen, W.; Zhu, J.; Li, H.; Bian, Z. Unveiling the role of defects on oxygen activation and photodegradation of organic pollutants. Environ. Sci. Technol. 2018, 52, 13879–13886. [Google Scholar] [CrossRef]
  117. Zhang, X.; Cai, M.; Cui, N.; Chen, G.; Zou, G.; Zhou, L. Defective black TiO2: Effects of annealing atmospheres and urea addition on the properties and photocatalytic activities. Nanomaterials 2021, 11, 2648. [Google Scholar] [CrossRef]
  118. Du, M.; Chen, Q.; Wang, Y.; Hu, J.; Meng, X. Synchronous construction of oxygen vacancies and phase junction in TiO2 hierarchical structure for enhancement of visible light photocatalytic activity. J. Alloys Compd. 2020, 830, 154649. [Google Scholar] [CrossRef]
  119. Albetran, H.; O’Connor, B.H.; Low, I.M. Effect of calcination on band gaps for electrospun titania nanofibers heated in air-argon mixtures. Mater. Des. 2016, 92, 480–485. [Google Scholar] [CrossRef] [Green Version]
  120. Sang, L.X.; Zhang, Z.Y.; Ma, C.F. Photoelectrical and charge transfer properties of hydrogen-evolving TiO2 nanotube arrays electrodes annealed in different gases. Int. J. Hydrogen Energy 2011, 36, 4732–4738. [Google Scholar] [CrossRef]
  121. Qi, W.; Zhang, F.; An, X.; Liu, H.; Qu, J. Oxygen vacancy modulation of {010}-dominated TiO2 for enhanced photodegradation of Sulfamethoxazole. Catal. Commun. 2019, 118, 35–38. [Google Scholar] [CrossRef]
  122. Li, Y.; Ye, X.; Cao, S.; Yang, C.; Wang, Y.; Ye, J. Oxygen-deficient dumbbell-shaped anatase TiO2-x mesocrystals with nearly 100% exposed {101} facets: Synthesis, growth mechanism, and Photocatalytic Performance. Chem.-A Eur. J. 2019, 25, 3032–3041. [Google Scholar] [CrossRef]
  123. Sheng, Z.; Song, S.; Wang, H.; Wu, Z.; Liu, Y. One-step hydrothermal synthesis of Pd-modified TiO2 with high photocatalytic activity for nitric oxide oxidation in gas phase. Environ. Eng. Sci. 2012, 29, 972–978. [Google Scholar] [CrossRef]
  124. Sasirekha, N.; Basha, S.J.S.; Shanthi, K. Photocatalytic performance of Ru doped anatase mounted on silica for reduction of carbon dioxide. Appl. Catal. B-Environ. 2006, 62, 169–180. [Google Scholar] [CrossRef]
  125. Gao, P.; Yang, L.; Xiao, S.; Wang, L.; Guo, W.; Lu, J. Effect of Ru, Rh, Mo, and Pd adsorption on the electronic and optical properties of anatase TiO2(101): A DFT investigation. Materials 2019, 12, 814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Thalgaspitiya, W.R.K.; Kapuge, T.K.; He, J.; Deljoo, B.; Meguerdichian, A.G.; Aindow, M.; Suib, S.L. Multifunctional transition metal doped titanium dioxide reduced graphene oxide composites as highly efficient adsorbents and photocatalysts. Microporous Mesoporous Mater. 2020, 307, 110521. [Google Scholar] [CrossRef]
  127. Wang, T.; Li, B.-R.; Wu, L.-G.; Yin, Y.-B.; Jiang, B.-Q.; Lou, J.-Q. Enhanced performance of TiO2/reduced graphene oxide doped by rare-earth ions for degrading phenol in seawater excited by weak visible light. Adv. Powder Technol. 2019, 30, 1920–1931. [Google Scholar] [CrossRef]
  128. Stengl, V.; Bakardjieva, S.; Murafa, N. Preparation and photocatalytic activity of rare earth doped TiO2 nanoparticles. Mater. Chem. Phys. 2009, 114, 217–226. [Google Scholar] [CrossRef]
  129. Fang, X.; Chen, X.; Zhu, Z. Optical and photocatalytic properties of Er3+ and/or Yb3+ doped TiO2 photocatalysts. J. Mater. Sci.-Mater. Electron. 2017, 28, 474–479. [Google Scholar] [CrossRef]
  130. Setiawati, E.; Kawano, K. Stabilization of anatase phase in the rare earth; Eu and Sm ion doped nanoparticle TiO2. J. Alloys Compd. 2008, 451, 293–296. [Google Scholar] [CrossRef]
  131. Yu, Y.; Chen, G.; Zhou, Y.; Han, Z. Recent advances in rare-earth elements modification of inorganic semiconductor-based photocatalysts for efficient solar energy conversion: A review. J. Rare Earths 2015, 33, 453–462. [Google Scholar] [CrossRef]
  132. Liu, S.-Y.; Zuo, C.-G.; Xia, J. Solid-state synthesis and photodegradation property of anatase TiO2 micro-nanopowder by sodium replacement. Solid State Sci. 2021, 115, 106589. [Google Scholar] [CrossRef]
  133. Lv, C.; Lan, X.; Wang, L.; Yu, Q.; Zhang, M.; Sun, H.; Shi, J. Alkaline-earth-metal-doped TiO2 for enhanced photodegradation and H2 evolution: Insights into the mechanisms. Catal. Sci. Technol. 2019, 9, 6124–6135. [Google Scholar] [CrossRef]
  134. Li, Y.; Ma, G.; Peng, S.; Lu, G.; Li, S. Boron and nitrogen co-doped titania with enhanced visible-light photocatalytic activity for hydrogen evolution. Appl. Surf. Sci. 2008, 254, 6831–6836. [Google Scholar] [CrossRef]
  135. Barakat, N.A.M.; Zaki, A.H.; Ahmed, E.; Farghali, A.A.; Al-Mubaddel, F.S. FexCo1-x-doped titanium oxide nanotubes as effective photocatalysts for hydrogen extraction from ammonium phosphate. Int. J. Hydrogen Energy 2018, 43, 7990–7997. [Google Scholar] [CrossRef]
  136. 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]
  137. Chaudhari, N.S.; Warule, S.S.; Dhanmane, S.A.; Kulkarni, M.V.; Valant, M.; Kale, B.B. Nanostructured N-doped TiO2 marigold flowers for an efficient solar hydrogen production from H2S. Nanoscale 2013, 5, 9383–9390. [Google Scholar] [CrossRef]
  138. Wang, Y.; Huang, Y.; Ho, W.; Zhang, L.; Zou, Z.; Lee, S. Biomolecule-controlled hydrothermal synthesis of C-N-S-tridoped TiO2 nanocrystalline photocatalysts for NO removal under simulated solar light irradiation. J. Hazard. Mater. 2009, 169, 77–87. [Google Scholar] [CrossRef]
  139. Yuan, J.; Chen, M.; Shi, J.; Shangguan, W. Preparations and photocatalytic hydrogen evolution of N-doped TiO2 from urea and titanium tetrachloride. Int. J. Hydrogen Energy 2006, 31, 1326–1331. [Google Scholar] [CrossRef]
  140. Momeni, M.M.; Ghayeb, Y.; Ghonchegi, Z. Visible light activity of sulfur-doped TiO2 nanostructure photoelectrodes prepared by single-step electrochemical anodizing process. J. Solid State Electrochem. 2015, 19, 1359–1366. [Google Scholar] [CrossRef]
  141. Carmichael, P.; Hazafy, D.; Bhachu, D.S.; Mills, A.; Darr, J.A.; Parkin, I.P. Atmospheric pressure chemical vapour deposition of boron doped titanium dioxide for photocatalytic water reduction and oxidation. Phys. Chem. Chem. Phys. 2013, 15, 16788–16794. [Google Scholar] [CrossRef] [Green Version]
  142. Wu, G.; Chen, A. Direct growth of F-doped TiO2 particulate thin films with high photocatalytic activity for environmental applications. J. Photochem. Photobiol. A-Chem. 2008, 195, 47–53. [Google Scholar] [CrossRef]
  143. Zhu, H.X.; Liu, J.M. First principles calculations of electronic and optical properties of Mo and C co-doped anatase TiO2. Appl. Phys. A-Mater. Sci. Process. 2014, 117, 831–839. [Google Scholar] [CrossRef]
  144. Diao, W.; He, J.; Wang, Q.; Rao, X.; Zhang, Y. K, Na and Cl co-doped TiO2 nanorod arrays on carbon cloth for efficient photocatalytic degradation of formaldehyde under UV/visible LED irradiation. Catal. Sci. Technol. 2021, 11, 230–238. [Google Scholar] [CrossRef]
  145. Filippatos, P.-P.; Soultati, A.; Kelaidis, N.; Petaroudis, C.; Alivisatou, A.-A.; Drivas, C.; Kennou, S.; Agapaki, E.; Charalampidis, G.; Yusoff, A.R.b.M.; et al. Preparation of hydrogen, fluorine and chlorine doped and co-doped titanium dioxide photocatalysts: A theoretical and experimental approach. Sci. Rep. 2021, 11, 5700. [Google Scholar] [CrossRef]
  146. Meng, S.; Zhang, J.; Chen, S.; Zhang, S.; Huang, W. Perspective on construction of heterojunction photocatalysts and the complete utilization of photogenerated charge carriers. Appl. Surf. Sci. 2019, 476, 982–992. [Google Scholar] [CrossRef]
  147. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef]
  148. Smith, Y.R.; Sarma, B.; Mohanty, S.K.; Misra, M. Formation of TiO2-WO3 nanotubular composite via single-step anodization and its application in photoelectrochemical hydrogen generation. Electrochem. Commun. 2012, 19, 131–134. [Google Scholar] [CrossRef]
  149. Choudhury, S.; Sasikala, R.; Saxena, V.; Aswal, D.K.; Bhattacharya, D. A new route for the fabrication of an ultrathin film of a PdO-TiO2 composite photocatalyst. Dalton Trans. 2012, 41, 12090–12095. [Google Scholar] [CrossRef]
  150. Navarrete, M.; Cipagauta-Diaz, S.; Gomez, R. Ga2O3/TiO2 semiconductors free of noble metals for the photocatalytic hydrogen production in a water/methanol mixture. J. Chem. Technol. Biotechnol. 2019, 94, 3457–3465. [Google Scholar] [CrossRef]
  151. Gholami, M.; Shirzad-Siboni, M.; Farzadkia, M.; Yang, J.-K. Synthesis, characterization, and application of ZnO/TiO2 nanocomposite for photocatalysis of a herbicide (Bentazon). Desalination Water Treat. 2016, 57, 13632–13644. [Google Scholar] [CrossRef]
  152. Chen, J.-Z.; Chen, T.-H.; Lai, L.-W.; Li, P.-Y.; Liu, H.-W.; Hong, Y.-Y.; Liu, D.-S. Preparation and characterization of surface photocatalytic activity with NiO/TiO2 nanocomposite structure. Materials 2015, 8, 4273–4286. [Google Scholar] [CrossRef] [Green Version]
  153. Wang, W.; Wu, Z.; Eftekhari, E.; Huo, Z.; Li, X.; Tade, M.O.; Yan, C.; Yan, Z.; Li, C.; Li, Q.; et al. High performance heterojunction photocatalytic membranes formed by embedding Cu2O and TiO2 nanowires in reduced graphene oxide. Catal. Sci. Technol. 2018, 8, 1704–1711. [Google Scholar] [CrossRef] [Green Version]
  154. Morales-Torres, S.; Pastrana-Martinez, L.M.; Figueiredo, J.L.; Faria, J.L.; Silva, A.M.T. Design of graphene-based TiO2 photocatalysts-a review. Environ. Sci. Pollut. Res. 2012, 19, 3676–3687. [Google Scholar] [CrossRef]
  155. Gao, P.; Sun, D.D. Ultrasonic preparation of hierarchical Graphene-Oxide/TiO2 composite microspheres for efficient photocatalytic hydrogen production. Chem.-Asian J. 2013, 8, 2779–2786. [Google Scholar] [CrossRef]
  156. Kong, D.; Zhao, M.; Li, S.; Huang, F.; Song, J.; Yuan, Y.; Shen, Y.; Xie, A. Synthesis of TiO2/rGO nanocomposites with enhanced photoelectrochemical performance and photocatalytic activity. Nano 2016, 11, 1650007. [Google Scholar] [CrossRef]
  157. Zhang, X.-Y.; Li, H.-P.; Cui, X.-L. Preparation and Photocatalytic Activity for Hydrogen evolution of TiO2/graphene sheets composite. Chin. J. Inorg. Chem. 2009, 25, 1903–1907. [Google Scholar]
  158. Shen, J.; Shi, M.; Yan, B.; Ma, H.; Li, N.; Ye, M. Ionic liquid-assisted one-step hydrothermal synthesis of TiO2-reduced graphene oxide composites. Nano Res. 2011, 4, 795–806. [Google Scholar] [CrossRef]
  159. Fu, Z.; Wang, H.; Wang, Y.; Wang, S.; Li, Z.; Sun, Q. Construction of three-dimensional g-C3N4/Gr-CNTs/TiO2 Z-scheme catalyst with enhanced photocatalytic activity. Appl. Surf. Sci. 2020, 510, 145494. [Google Scholar] [CrossRef]
  160. Chiang, H.-H.; Wang, S.-H.; Chou, H.-Y.; Huang, C.-C.; Tsai, T.-L.; Yang, Y.-C.; Lee, J.-W.; Lin, T.-Y.; Wu, Y.-J.; Chen, C.-C. Surface modification of ato photocatalyst on its bactericidal effect against Escherichia coli. J. Mar. Sci. Technol.-Taiwan 2014, 22, 269–276. [Google Scholar]
  161. Zhou, X.M. TiO2-supported single-atom catalysts for photocatalytic reactions. Acta Phys.-Chim. Sin. 2021, 37, 2008064. [Google Scholar]
  162. Bernareggi, M.; Chiarello, G.L.; West, G.; Ratova, M.; Ferretti, A.M.; Kelly, P.; Selli, E. Cu and Pt clusters deposition on TiO2 powders by DC magnetron sputtering for photocatalytic hydrogen production. Catal. Today 2019, 326, 15–21. [Google Scholar] [CrossRef] [Green Version]
  163. Dozzi, M.V.; Brocato, S.; Marra, G.; Tozzola, G.; Meda, L.; Selli, E. Aqueous ammonia abatement on Pt- and Ru-modified TiO2: Selectivity effects of the metal nanoparticles deposition method. Catal. Today 2017, 287, 148–154. [Google Scholar] [CrossRef]
  164. Zheng, Z.; Murakamia, N.; Liu, J.; Teng, Z.; Zhang, Q.; Cao, Y.; Cheng, H.; Ohno, T. Development of plasmonic photocatalyst by site-selective loading of bimetallic nanoparticles of Au and Ag on titanium(IV) oxide. ChemCatChem 2020, 12, 3783–3792. [Google Scholar] [CrossRef]
  165. Luo, J.; Li, D.; Yang, Y.; Liu, H.; Chen, J.; Wang, H. Preparation of Au/reduced graphene oxide/hydrogenated TiO2 nanotube arrays ternary composites for visible-light-driven photoelectrochemical water splitting. J. Alloys Compd. 2016, 661, 380–388. [Google Scholar] [CrossRef]
  166. Liu, E.; Kang, L.; Yang, Y.; Sun, T.; Hu, X.; Zhu, C.; Liu, H.; Wang, Q.; Li, X.; Fan, J. Plasmonic Ag deposited TiO2 nano-sheet film for enhanced photocatalytic hydrogen production by water splitting. Nanotechnology 2014, 25, 165401. [Google Scholar] [CrossRef] [PubMed]
  167. Ge, M.-Z.; Cao, C.-Y.; Li, S.-H.; Tang, Y.-X.; Wang, L.-N.; Qi, N.; Huang, J.-Y.; Zhang, K.-Q.; Al-Deyab, S.S.; Lai, Y.-K. In situ plasmonic Ag nanoparticle anchored TiO2 nanotube arrays as visible-light-driven photocatalysts for enhanced water splitting. Nanoscale 2016, 8, 5226–5234. [Google Scholar] [CrossRef]
  168. Barrios, C.E.; Albiter, E.; Jimenez, J.; Tiznado, H.; Romo-Herrera, J.; Zanella, R. Photocatalytic hydrogen production over titania modified by gold–Metal (palladium, nickel and cobalt) catalysts. Int. J. Hydrogen Energy 2016, 41, 23287–23300. [Google Scholar] [CrossRef]
  169. Monamary, A.; Vijayalakshmi, K. Substantial effect of palladium overlayer deposition on the H2 sensing performance of TiO2/ITO nanocomposite. Ceram. Int. 2018, 44, 22957–22962. [Google Scholar] [CrossRef]
  170. Vallejo, W.; Rueda, A.; Diaz-Uribe, C.; Grande, C.; Quintana, P. Photocatalytic activity of graphene oxide-TiO2 thin films sensitized by natural dyes extracted from Bactris guineensis. R. Soc. Open Sci. 2019, 6, 181824. [Google Scholar] [CrossRef] [Green Version]
  171. Shi, J.W.; Guan, X.J.; Zhou, Z.H.; Liu, H.P.; Guo, L.J. Eosin Y-sensitized nanosheet-stacked hollow-sphere TiO2 for efficient photocatalytic H2 production under visible-light irradiation. J. Nanopart. Res. 2015, 17, 252. [Google Scholar] [CrossRef]
  172. Murcia, J.J.; Avila-Martinez, E.G.; Rojas, H.; Cubillos, J.; Ivanova, S.; Penkova, A.; Laguna, O.H. Powder and nanotubes titania modified by dye sensitization as photocatalysts for the organic pollutants elimination. Nanomaterials 2019, 9, 517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Rodriguez, H.B.; Di Iorio, Y.; Meichtry, J.M.; Grela, M.A.; Litter, M.I.; San Roman, E. Evidence on dye clustering in the sensitization of TiO2 by aluminum phthalocyanine. Photochem. Photobiol. Sci. 2013, 12, 1984–1990. [Google Scholar] [CrossRef] [PubMed]
  174. Ding, H.M.; Sun, H.; Shan, Y.K. Preparation and characterization of mesoporous SBA-15 supported dye-sensitized TiO2 photocatalyst. J. Photochem. Photobiol. A-Chem. 2005, 169, 101–107. [Google Scholar] [CrossRef]
  175. Wahyuningsih, S.; Purnawan, C.; Kartikasari, P.A.; Praistia, N. Visible light photoelectrocatalytic degradation of rhodamine B using a dye-sensitised TiO2 electrode. Chem. Pap. 2014, 68, 1248–1256. [Google Scholar] [CrossRef]
  176. Li, Y.; Li, H.M.; Lu, X.L.; Yu, X.; Kong, M.H.; Duan, X.D.; Qin, G.; Zhao, Y.H.; Wang, Z.L.; Dionysiou, D.D. Molybdenum disulfide nanosheets vertically grown on self-supported titanium dioxide/nitrogen-doped carbon nanofiber film for effective hydrogen peroxide decomposition and “memory catalysis”. J. Colloid Interface Sci. 2021, 596, 384–395. [Google Scholar] [CrossRef] [PubMed]
  177. Ikeue, K.; Yamashita, H.; Anpo, M. Photocatalytic reduction of CO2 with H2O on titanium oxides prepared within zeolites and mesoporous molecular sieves. Electrochemistry 2002, 70, 402–408. [Google Scholar] [CrossRef] [Green Version]
  178. Rasalingam, S.; Kibombo, H.S.; Wu, C.-M.; Budhi, S.; Peng, R.; Baltrusaitis, J.; Koodali, R.T. Influence of Ti-O-Si hetero-linkages in the photocatalytic degradation of Rhodamine B. Catal. Commun. 2013, 31, 66–70. [Google Scholar] [CrossRef]
  179. Liu, B.; Zeng, H.C. Carbon nanotubes supported mesoporous mesocrystals of anatase TiO2. Chem. Mater. 2008, 20, 2711–2718. [Google Scholar] [CrossRef]
  180. Najafabadi, A.T.; Taghipour, F. Physicochemical impact of zeolites as the support for photocatalytic hydrogen production using solar-activated TiO2-based nanoparticles. Energy Convers. Manag. 2014, 82, 106–113. [Google Scholar] [CrossRef]
  181. Xu, Y.M.; Langford, C.H. Enhanced photoactivity of a titanium(iv) oxide-supported on ZSM5 and zeolite-a at low-coverage. J. Phys. Chem. 1995, 99, 11501–11507. [Google Scholar] [CrossRef]
  182. Kim, H.J.; Shul, Y.G.; Han, H.S. Photocatalytic properties of silica-supported TiO2. Top. Catal. 2005, 35, 287–293. [Google Scholar] [CrossRef]
  183. Yin, J.W.; Xing, Z.P.; Kuang, J.Y.; Li, Z.Z.; Li, M.; Jiang, J.J.; Tan, S.Y.; Zhu, Q.; Zhou, W. Bi plasmon-enhanced mesoporous Bi2MoO6/Ti3+ self-doped TiO2 microsphere heterojunctions as efficient visible-light-driven photocatalysts. J. Alloys Compd. 2018, 750, 659–668. [Google Scholar] [CrossRef]
  184. Xing, M.Y.; Zhang, J.L.; Qiu, B.C.; Tian, B.Z.; Anpo, M.; Che, M. A Brown mesoporous TiO2-x/MCF composite with an extremely high quantum yield of solar energy photocatalysis for H2 evolution. Small 2015, 11, 1920–1929. [Google Scholar] [CrossRef] [PubMed]
  185. Wen, C.Z.; Jiang, H.B.; Qiao, S.Z.; Yang, H.G.; Lu, G.Q. Synthesis of high-reactive facets dominated anatase TiO2. J. Mater. Chem. 2011, 21, 7052–7061. [Google Scholar] [CrossRef] [Green Version]
  186. Li, H.; Zeng, Y.; Huang, T.; Piao, L.; Liu, M. Controlled synthesis of anatase TiO2 single crystals with dominant {001} facets from TiO2 powders. ChemPlusChem 2012, 77, 1017–1021. [Google Scholar] [CrossRef]
  187. Wang, W.; Lu, C.-H.; Ni, Y.-R.; Song, J.-B.; Su, M.-X.; Xu, Z.-Z. Enhanced visible-light photoactivity of {001} facets dominated TiO2 nanosheets with even distributed bulk oxygen vacancy and Ti3+. Catal. Commun. 2012, 22, 19–23. [Google Scholar] [CrossRef]
  188. Shang, Q.; Tan, X.; Yu, T.; Zhang, Z.; Zou, Y.; Wang, S. Efficient gaseous toluene photoconversion on graphene-titanium dioxide nanocomposites with dominate exposed {001} facets. J. Colloid Interface Sci. 2015, 455, 134–144. [Google Scholar] [CrossRef] [PubMed]
  189. Lv, Z.; Cheng, X.; Liu, B.; Guo, Z.; Jin, M.; Zhang, C. Enhanced photoredox water splitting of Sb-N donor-acceptor pairs in TiO2. Inorg. Chem. Front. 2019, 6, 2404–2411. [Google Scholar] [CrossRef]
  190. Koci, K.; Troppova, I.; Edelmannova, M.; Starostka, J.; Matejova, L.; Lang, J.; Reli, M.; Drobna, H.; Rokicinska, A.; Kustrowski, P.; et al. Photocatalytic decomposition of methanol over La/TiO2 materials. Environ. Sci. Pollut. Res. 2018, 25, 34818–34825. [Google Scholar] [CrossRef]
  191. Bharatvaj, J.; Preethi, V.; Kanmani, S. Hydrogen production from sulphide wastewater using Ce3+-TiO2 photocatalysis. Int. J. Hydrogen Energy 2018, 43, 3935–3945. [Google Scholar] [CrossRef]
  192. Agegnehu, A.K.; Pan, C.J.; Tsai, M.C.; Rick, J.; Su, W.N.; Lee, J.F.; Hwang, B.J. Visible light responsive noble metal-free nanocomposite of V-doped TiO2 nanorod with highly reduced graphene oxide for enhanced solar H2 production. Int. J. Hydrogen Energy 2016, 41, 6752–6762. [Google Scholar] [CrossRef]
  193. Barakat, N.A.M.; Ahmed, E.; Amen, M.T.; Abdelkareem, M.A.; Farghali, A.A. N-doped Ni/C/TiO2 nanocomposite as effective photocatalyst for water splitting. Mater. Lett. 2018, 210, 317–320. [Google Scholar] [CrossRef]
  194. Gao, L.S.; Zhang, S.N.; Zou, X.X.; Wang, J.F.; Su, J.; Chen, J.S. Oxygen vacancy engineering of titania-induced by Sr2+ dopants for visible-light-driven hydrogen evolution. Inorg. Chem. 2021, 60, 32–36. [Google Scholar] [CrossRef]
  195. Xu, J.C.; Zhang, J.J.; Cai, Z.Y.; Huang, H.; Huang, T.H.; Wang, P.; Wang, X.Y. Facile and large-scale synthesis of defective black TiO2-x(B) nanosheets for efficient visible-light-driven photocatalytic hydrogen evolution. Catalysts 2019, 9, 1048. [Google Scholar] [CrossRef] [Green Version]
  196. Barakat, N.A.M.; Erfan, N.A.; Mohammed, A.A.; Mohamed, S.E.I. Ag-decorated TiO2 nanofibers as Arrhenius equation-incompatible and effective photocatalyst for water splitting under visible light irradiation. Colloids Surf. A-Physicochem. Eng. Asp. 2020, 604, 125307. [Google Scholar] [CrossRef]
  197. Dang, H.F.; Dong, X.F.; Dong, Y.C.; Zhang, Y.; Hampshire, S. TiO2 nanotubes coupled with nano-Cu(OH)2 for highly efficient photocatalytic hydrogen production. Int. J. Hydrogen Energy 2013, 38, 2126–2135. [Google Scholar] [CrossRef]
  198. Diaz, L.; Rodriguez, V.D.; Gonzalez-Rodriguez, M.; Rodriguez-Castellon, E.; Algarra, M.; Nunez, P.; Moretti, E. M/TiO2 (M = Fe, Co, Ni, Cu, Zn) catalysts for photocatalytic hydrogen production under UV and visible light irradiation. Inorg. Chem. Front. 2021, 8, 3491–3500. [Google Scholar] [CrossRef]
  199. El-Bery, H.M.; Abdelhamid, H.N. Photocatalytic hydrogen generation via water splitting using ZIF-67 derived Co3O4@C/TiO2. J. Environ. Chem. Eng. 2021, 9, 105702. [Google Scholar] [CrossRef]
  200. Fujita, S.; Kawamori, H.; Honda, D.; Yoshida, H.; Arai, M. Photocatalytic hydrogen production from aqueous glycerol solution using NiO/TiO2 catalysts: Effects of preparation and reaction conditions. Appl. Catal. B-Environ. 2016, 181, 818–824. [Google Scholar] [CrossRef]
  201. Han, C.; Wang, Y.; Lei, Y.; Wang, B.; Wu, N.; Shi, Q.; Li, Q. In situ synthesis of graphitic-C3N4 nanosheet hybridized N-doped TiO2 nanofibers for efficient photocatalytic H2 production and degradation. Nano Res. 2015, 8, 1199–1209. [Google Scholar] [CrossRef]
  202. Kokporka, L.; Onsuratoom, S.; Puangpetch, T.; Chavadej, S. Sol-gel-synthesized mesoporous-assembled TiO2-ZrO2 mixed oxide nanocrystals and their photocatalytic sensitized H2 production activity under visible light irradiation. Mater. Sci. Semicond. Process. 2013, 16, 667–678. [Google Scholar] [CrossRef]
  203. Mou, Z.G.; Wu, Y.J.; Sun, J.H.; Yang, P.; Du, Y.K.; Lu, C. TiO2 Nanoparticles-functionalized N-doped graphene with superior interfacial contact and enhanced charge separation for photocatalytic hydrogen generation. ACS Appl. Mater. Interfaces 2014, 6, 13798–13806. [Google Scholar] [CrossRef] [PubMed]
  204. Sharma, A.; Thuan, D.V.; Pham, T.D.; Tung, M.H.T.; Truc, N.T.T.; Vo, D.V.N. Advanced surface of fibrous activated carbon immobilized with FeO/TiO2 for photocatalytic evolution of hydrogen under visible light. Chem. Eng. Technol. 2020, 43, 752–761. [Google Scholar] [CrossRef]
  205. Wang, C.; Hu, Q.Q.; Huang, J.Q.; Zhu, C.; Deng, Z.H.; Shi, H.L.; Wu, L.; Liu, Z.G.; Cao, Y.G. Enhanced hydrogen production by water splitting using Cu-doped TiO2 film with preferred (001) orientation. Appl. Surf. Sci. 2014, 292, 161–164. [Google Scholar] [CrossRef]
  206. Ma, J.; Tan, X.; Yu, T.; Li, X. Fabrication of g-C3N4/TiO2 hierarchical spheres with reactive {001} TiO2 crystal facets and its visible-light photocatalytic activity. Int. J. Hydrogen Energy 2016, 41, 3877–3887. [Google Scholar] [CrossRef]
  207. Xiang, Q.; Yu, J.; Jaroniec, M. Enhanced photocatalytic H2 production activity of graphene-modified titania nanosheets. Nanoscale 2011, 3, 3670–3678. [Google Scholar] [CrossRef]
  208. Chu, L.; Lin, Y.; Liu, Y.; Wang, H.; Zhang, Q.; Li, Y.; Cao, Y.; Yu, H.; Peng, F. Preparation of CdS-CoSx photocatalysts and their photocatalytic and photoelectrochemical characteristics for hydrogen production. Int. J. Hydrogen Energy 2019, 44, 27795–27805. [Google Scholar] [CrossRef]
  209. Dong, J.; Duan, L.; Wu, Q.; Yao, W. Facile preparation of Pt/CdS photocatalyst by a modified photoreduction method with efficient hydrogen production. Int. J. Hydrogen Energy 2018, 43, 2139–2147. [Google Scholar] [CrossRef]
  210. Kondo, T.; Nagata, M. Cu-doped ZnS/zeolite composite photocatalysts for hydrogen production from aqueous S2−/SO32− Solutions. Chem. Lett. 2017, 46, 1797–1799. [Google Scholar] [CrossRef]
  211. Qiao, F.; Liu, W.; Yang, J.; Yuan, J.; Sun, K.; Liu, P.F. Hydrogen production performance and theoretical mechanism analysis of chain-like ZnO/ZnS heterojunction. Int. J. Hydrogen Energy 2023, 48, 953–963. [Google Scholar] [CrossRef]
  212. Ma, W.; Zheng, D.; Xiao, B.; Xian, Y.; Zhang, Q.; Wang, S.; Liu, J.; Wang, P.; Hu, X. An efficient photocatalytic system under visible light: In-situ growth cocatalyst Ni2P on the surface of CdS. J. Environ. Chem. Eng. 2022, 10, 107822. [Google Scholar] [CrossRef]
  213. Zou, Y.; Guo, C.; Cao, X.; Zhang, L.; Chen, T.; Guo, C.; Wang, J. Synthesis of CdS/CoP hollow nanocages with improved photocatalytic water splitting performance for hydrogen evolution. J. Environ. Chem. Eng. 2021, 9, 106270. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the energy band structure of semiconductors.
Figure 1. Schematic illustration of the energy band structure of semiconductors.
Metals 13 01163 g001
Figure 2. (a) Behavior of photogenerated carriers in a semiconductor. (b) Schematic illustration of the mechanism during photocatalytic water splitting over a semiconductor.
Figure 2. (a) Behavior of photogenerated carriers in a semiconductor. (b) Schematic illustration of the mechanism during photocatalytic water splitting over a semiconductor.
Metals 13 01163 g002
Figure 3. (a) Optical absorption of various TiO2-x samples and (b) their corresponding band gaps [32]. (c) Illustrations on oxygen vacancy and (d) donor states owing to Ti3+.
Figure 3. (a) Optical absorption of various TiO2-x samples and (b) their corresponding band gaps [32]. (c) Illustrations on oxygen vacancy and (d) donor states owing to Ti3+.
Metals 13 01163 g003aMetals 13 01163 g003b
Figure 4. (a) Digital images and (b) absorption spectra together with K-M functions showing the calculated band gap interpolation for TiO2 hydrogenated at different times [69].
Figure 4. (a) Digital images and (b) absorption spectra together with K-M functions showing the calculated band gap interpolation for TiO2 hydrogenated at different times [69].
Metals 13 01163 g004
Figure 5. (a) Surface resistance of TiO2 (100) substrates as a function of pulse number irradiated by ArF, KrF, and XeCl lasers. (b) Reciprocal space mappings around the (220) reflection for the unirradiated TiO2 (100) substrate and laser-irradiated TiO2−δ/TiO2 (100) substrate. The insets show the Qx profiles at the (220) reflection [85].
Figure 5. (a) Surface resistance of TiO2 (100) substrates as a function of pulse number irradiated by ArF, KrF, and XeCl lasers. (b) Reciprocal space mappings around the (220) reflection for the unirradiated TiO2 (100) substrate and laser-irradiated TiO2−δ/TiO2 (100) substrate. The insets show the Qx profiles at the (220) reflection [85].
Metals 13 01163 g005
Figure 6. (a) Schematic models of TiO2 nanostructures grown on gold nanoisland (GNI)-modified Si (100) templates at 675, 700, and 720 °C. (b) Photographs and XPS spectra of O 1 s, Ti 2p3/2, and Si 2p regions of TiO2 films consisting of nanobelts, corrugated nanowires (NWs), straight NWs, and decorated NWs. (c) SEM images of TiO2 nanostructures grown in 20 mTorr Ar at 675–750 °C on GNI-modified, H-terminated Si (GNI/H-Si), GNI-modified, RCA-cleaned Si (GNI/RCA-Si), and GNI-modified, thermally-oxidized (GNI/Ox-Si) templates. The corresponding lower left insets show schematic models of the as-grown nanostructures, and the upper right ones display the magnified SEM images of the selected nanostructures [92].
Figure 6. (a) Schematic models of TiO2 nanostructures grown on gold nanoisland (GNI)-modified Si (100) templates at 675, 700, and 720 °C. (b) Photographs and XPS spectra of O 1 s, Ti 2p3/2, and Si 2p regions of TiO2 films consisting of nanobelts, corrugated nanowires (NWs), straight NWs, and decorated NWs. (c) SEM images of TiO2 nanostructures grown in 20 mTorr Ar at 675–750 °C on GNI-modified, H-terminated Si (GNI/H-Si), GNI-modified, RCA-cleaned Si (GNI/RCA-Si), and GNI-modified, thermally-oxidized (GNI/Ox-Si) templates. The corresponding lower left insets show schematic models of the as-grown nanostructures, and the upper right ones display the magnified SEM images of the selected nanostructures [92].
Metals 13 01163 g006
Figure 7. (a) EPR spectra of the anatase samples synthesized by a hydrothermal treatment with different HF amounts [93]. (b) Formation energies of oxygen vacancies as a function of ∆µO (the difference in oxygen chemical potentials [99].
Figure 7. (a) EPR spectra of the anatase samples synthesized by a hydrothermal treatment with different HF amounts [93]. (b) Formation energies of oxygen vacancies as a function of ∆µO (the difference in oxygen chemical potentials [99].
Metals 13 01163 g007
Figure 8. (a) Digital photograph and (b) band gaps of the electrospun TiO2 nanofibers prepared by non-isothermally heating from 25 to 900 °C at 10 °C/min in argon–air mixtures [119]. (c) Photocurrent density vs. the applied potential of the TiO2 nanotube arrays annealed in air (TNT-A), N2 (TNT-N), and 5% H2/N2 mixture gas (TNT-H) under ultraviolet light (365 ± 15 nm) irradiation and the control tests in the dark [120]. (d) Electrochemical impedance spectroscopy plots of the anodized TiO2 nanotubes annealed in air (TNT-A), N2 (TNT-N) and 5% H2/N2 mixture gas (TNT-H) under ultraviolet light illumination [120].
Figure 8. (a) Digital photograph and (b) band gaps of the electrospun TiO2 nanofibers prepared by non-isothermally heating from 25 to 900 °C at 10 °C/min in argon–air mixtures [119]. (c) Photocurrent density vs. the applied potential of the TiO2 nanotube arrays annealed in air (TNT-A), N2 (TNT-N), and 5% H2/N2 mixture gas (TNT-H) under ultraviolet light (365 ± 15 nm) irradiation and the control tests in the dark [120]. (d) Electrochemical impedance spectroscopy plots of the anodized TiO2 nanotubes annealed in air (TNT-A), N2 (TNT-N) and 5% H2/N2 mixture gas (TNT-H) under ultraviolet light illumination [120].
Metals 13 01163 g008
Figure 9. (a) UV–Vis diffuse reflectance spectra of TiO2 and alkaline earth metal-doped TiO2 [133] and (b) gaps in TiO2 and alkaline earth metal-doped TiO2 [133]. (c) Photocatalytic H2 production from water splitting over TiO2 and alkaline earth metal-doped TiO2 under the condition of adding Pt as a co-catalyst [133]. (d) Photocatalytic H2 generation over TiO2 doped with different amounts of N [129]. (e) Photocurrent response curves of TiO2, B-doped TiO2, N-doped TiO2, and (B,N)-co-doped TiO2 to visible light [134]. (f) Photocatalytic H2 generation over (FexCo1−x)-co-doped TiO2 [135].
Figure 9. (a) UV–Vis diffuse reflectance spectra of TiO2 and alkaline earth metal-doped TiO2 [133] and (b) gaps in TiO2 and alkaline earth metal-doped TiO2 [133]. (c) Photocatalytic H2 production from water splitting over TiO2 and alkaline earth metal-doped TiO2 under the condition of adding Pt as a co-catalyst [133]. (d) Photocatalytic H2 generation over TiO2 doped with different amounts of N [129]. (e) Photocurrent response curves of TiO2, B-doped TiO2, N-doped TiO2, and (B,N)-co-doped TiO2 to visible light [134]. (f) Photocatalytic H2 generation over (FexCo1−x)-co-doped TiO2 [135].
Metals 13 01163 g009aMetals 13 01163 g009b
Figure 10. Schematic illustration on the separation ways of photogenerated electron-hole pairs over heterojunction photocatalysts: (a) type-I, (b) type-II, (c) type-III, and (d) Z-scheme [147].
Figure 10. Schematic illustration on the separation ways of photogenerated electron-hole pairs over heterojunction photocatalysts: (a) type-I, (b) type-II, (c) type-III, and (d) Z-scheme [147].
Metals 13 01163 g010
Figure 11. (a) Average hydrogen evolution rates of TiO2, Ga2O3, TG3 (3% Ga2O3/TiO2), TG5 (5% Ga2O3/TiO2), TG10 (10% Ga2O3/TiO2) photocatalysts, and TPt reference (Pt modified TiO2) [150]. (b) Mechanism for H2 production over the TG5 photocatalyst [150]. (c) Transient current response curves of TiO2 and NiO/TiO2 nanocomposite under ultraviolet light irradiation [152]. (d) Schematic diagram on the energy band of a p-NiO/n-TiO2 heterojunction structure [152]. (e) Transient current response curves of 3D g-C3N4/graphene-CNTs/TiO2 samples with different amounts of TiO2 under an Xe lamp [159]. (f) Schematic diagram of the photocatalytic processes over 3D g-C3N4/graphene- CNTs/TiO2 [159].
Figure 11. (a) Average hydrogen evolution rates of TiO2, Ga2O3, TG3 (3% Ga2O3/TiO2), TG5 (5% Ga2O3/TiO2), TG10 (10% Ga2O3/TiO2) photocatalysts, and TPt reference (Pt modified TiO2) [150]. (b) Mechanism for H2 production over the TG5 photocatalyst [150]. (c) Transient current response curves of TiO2 and NiO/TiO2 nanocomposite under ultraviolet light irradiation [152]. (d) Schematic diagram on the energy band of a p-NiO/n-TiO2 heterojunction structure [152]. (e) Transient current response curves of 3D g-C3N4/graphene-CNTs/TiO2 samples with different amounts of TiO2 under an Xe lamp [159]. (f) Schematic diagram of the photocatalytic processes over 3D g-C3N4/graphene- CNTs/TiO2 [159].
Metals 13 01163 g011aMetals 13 01163 g011b
Figure 12. Band gaps estimated on the basis of the Kubelka–Munk plots for (a) TiO2-GO thin films and (b) TiO2-GO thin films sensitized with anthocyanin that was extracted from the fruit of Bactris guineensis (TiO2-GO-CO). The samples A, B, C, and D were prepared by adding 0.15%, 0.26%, 0.51%, and 1.1% GO in mass into TiO2. (c) Schematic illustration of the energy levels for the TiO2-GO thin films sensitized with natural dye [170].
Figure 12. Band gaps estimated on the basis of the Kubelka–Munk plots for (a) TiO2-GO thin films and (b) TiO2-GO thin films sensitized with anthocyanin that was extracted from the fruit of Bactris guineensis (TiO2-GO-CO). The samples A, B, C, and D were prepared by adding 0.15%, 0.26%, 0.51%, and 1.1% GO in mass into TiO2. (c) Schematic illustration of the energy levels for the TiO2-GO thin films sensitized with natural dye [170].
Metals 13 01163 g012
Table 1. Heat of combustion and ignition points of some commonly used fuels [13].
Table 1. Heat of combustion and ignition points of some commonly used fuels [13].
FuelsHeat of Combustion
(kJ·mol−1)
Heat of Combustion
(kJ·kg−1)
Ignition Point
(°C)
hydrogen285.81.42 × 105 585
coal-8.36 × 103~3.06 × 104300~700
gasoline-4.31 × 104427
diesel-4.26 × 104220
kerosene-4.31 × 10480
natural gas-3.89 × 104 kJ·m−3650
wood-1.2 × 104200~290
ethanol1366.82.97 × 10412
methane890.35.55 × 104538
butane26534.56 × 104365
acetone1788.73.08 × 104465
graphite393.73.28 × 104~650
Table 2. Comparison of different methods of introducing oxygen defects in TiO2.
Table 2. Comparison of different methods of introducing oxygen defects in TiO2.
Methods AdvantagesDisadvantagesRef.
Reduced by H2Strong reducing ability, no impurities introduced, and easy control on the density of oxygen defects by adjusting reaction time.High temperature, high energy consumption, time-consuming, and high risk.[65,67,68,69]
Reduced by chemical reductants such as NH3, NaBH4, and carbonMild reaction conditions and low energy consumption.Difficult control on the density of oxygen defects and easy introduction of impurities.[56,70,71,72]
Prepared in anoxic environmentConvenient operation and can easily obtain products in large quantities.High temperature, high energy consumption, and time consuming.[78,79,82]
Pulsed laser irradiation High reactivity and reducing efficiency and convenient operation.Special equipment needed.[88,91,92]
Pulsed laser deposition Convenient operation, easy control on the density of oxygen defects by adjusting the partial pressure of O2 and laser power density, and easily obtains special morphological structures of products.Special equipment needed.[80,81]
Ion dopingMild reaction conditions and wide selection of approaches.Unconsciously introduces oxygen defects without controllling their density and easily introduces impurities.[95,97,98,100,101,107]
Plasma-assisted depositionMild and controllable reaction conditions which is suitable for preparing films.Special equipment needed and low productivity.[104,106]
Ultrasonic-assisted techniquesConvenient operation, low cost, and scalable.The density of oxygen defects is not easily controlled.[54,109,112,114]
Calcination under anoxic conditionsConvenient operation and easy to control the density of oxygen defects by adjusting the partial pressure of O2.High temperature, high energy consumption, and time consuming.[117,120,121]
Molten salt calcinationConvenient operation and easily obtains products in large quantities.Special TiO2 precursor needed.[118]
Table 3. Hydrogen evolution efficiency of photocatalytic water splitting over various TiO2-based photocatalysts in comparison with those of typical photocatalysts reported in the literature.
Table 3. Hydrogen evolution efficiency of photocatalytic water splitting over various TiO2-based photocatalysts in comparison with those of typical photocatalysts reported in the literature.
CatalystLight SourceReaction ConditionH2 Production
(mmol h−1)
Ref.
N-doped TiO2>400 nmWater0.315[139]
N-doped TiO2>420 nmEDTA-2Na solution2.21[134]
(B,N)-co-doped TiO2>420 nmEDTA-2Na solution10.45[134]
(Sb,N)-co-doped TiO2Xe lamp10% aqueous TEOA solution2.33[189]
B-doped TiO2365 nm0.2 M HCl and absolute ethanol aqueous solution (1:1)0.099[141]
N-doped TiO2visible lightH2S/0.25 M KOH solution8.8[137]
N-doped TiO2Xe lamp20% aqueous methanol solution2.98[136]
S-doped TiO2Xe lamp1 M NaOH aqueous solution0.17[140]
Fe-doped TiO2solar light radiationtriammonium phosphate aqueous solution4.01[135]
Co-doped TiO2solar light radiationtriammonium phosphate aqueous solution9.82[135]
(Fe,Co)-co-doped TiO2solar light radiationtriammonium phosphate aqueous solution17.41[135]
La-doped TiO2Hg UVA lamp12 M aqueous methanol solution80[190]
Ce-doped TiO2visible lightsulphide wastewater from refinery6.789[191]
H-doped TiO2365 nm25% aqueous methanol solution0.286[145]
F-doped TiO2365 nm25% aqueous methanol solution0.0928[145]
Cl-doped TiO2365 nm25% aqueous methanol solution0.336[145]
V-doped TiO2/rGOXe lamp20% aqueous methanol solution0.12[192]
N-doped Ni/C/TiO2Hg lamp30% aqueous methanol solution0.383[193]
Sr-doped TiO2−δ>400 nmwater1.092[194]
TiO2−δ>420 nm30% aqueous methanol solution0.00058[195]
Pt/TiO2−δvisible light50% aqueous methanol solution4.9[47]
Ag-decorated TiO2Hg lampwater120[196]
Au-decorated TiO2254 nmaqueous methanol solution106[168]
Au,Pd-decorated TiO2254 nmaqueous methanol solution266[168]
Au,Ni-decorated TiO2254 nmaqueous methanol solution256[168]
Au,Co-decorated TiO2254 nmaqueous methanol solution171[168]
Pd-decorated TiO2254 nmaqueous methanol solution59[168]
Ni-decorated TiO2254 nmaqueous methanol solution20[168]
Co-decorated TiO2254 nmaqueous methanol solution10[168]
Cu(OH)2/TiO2ultraviolet light10% aqueous methanol solution14.94[197]
Cu/TiO2UV lamp25% aqueous methanol solution5[198]
Cu/TiO2visible light25% aqueous methanol solution0.22[198]
Co3O4@C/TiO2365 nm25% aqueous methanol solution11.4[199]
NiO/TiO2Hg lampglycerol and distilled water1.2[200]
g-C3N4/N-TiO2Xe lamp20% aqueous methanol solution8.931[201]
EosinY-sensitized TiO2/ZrO2Xe arc lamp15% DEA aqueous solution1.87[202]
β-Ga2O3/TiO2254 nm50% aqueous methanol solution0.244[150]
N-doped TiO2/N-doped grapheneXe lamp10% aqueous TEOA solution0.039[203]
FeO-TiO2/ACFvisible light20% aqueous methanol solution6.178[204]
TiO2/ACFvisible light20% aqueous methanol solution1.672[204]
Cu-doped TiO2 with preferred (001) orientationXe lamp10% aqueous methanol solution0.81[205]
g-C3N4/TiO2 with preferred (001) orientation>420 nm10% aqueous TEOA solution0.033[206]
TiO2/graphene with exposed (001) facetsXe lamp25% aqueous methanol solution0.736[207]
CdS>420 nm0.5 M Na2S-0.5 M Na2SO3 aqueous solution0.063[208]
CdS-CoSx>420 nm0.5 M Na2S-0.5 M Na2SO3 aqueous solution0.1686[208]
Pt/CdS>420 nm1.0 M aqueous (NH4)2SO3 solution1.158[209]
ZnSXe lamp0.1 M Na2S-0.1 M Na2SO3 aqueous solution0.04[210]
Cu-ZnS/ZeoliteXe lamp0.1 M Na2S-0.1 M Na2SO3 aqueous solution0.48[210]
ZnO/ZnSXe lamp0.064 M Na2S aqueous solution0.228[211]
ZnOXe lamp0.064 M Na2S aqueous solution0.138[211]
Ni2P>420 nm0.35 M Na2S-0.25 M Na2SO3 aqueous solution0.28[212]
Ni2P/CdS>420 nm0.35 M Na2S-0.25 M Na2SO3 aqueous solution16.02[212]
CoPvisible lightNa2S-Na2SO3 aqueous solution1.75[213]
CdS/CoPvisible lightNa2S-Na2SO3 aqueous solution15.74[213]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, Y.; Fu, X.; Peng, Z. A Review on Oxygen-Deficient Titanium Oxide for Photocatalytic Hydrogen Production. Metals 2023, 13, 1163. https://doi.org/10.3390/met13071163

AMA Style

Chen Y, Fu X, Peng Z. A Review on Oxygen-Deficient Titanium Oxide for Photocatalytic Hydrogen Production. Metals. 2023; 13(7):1163. https://doi.org/10.3390/met13071163

Chicago/Turabian Style

Chen, Yan, Xiuli Fu, and Zhijian Peng. 2023. "A Review on Oxygen-Deficient Titanium Oxide for Photocatalytic Hydrogen Production" Metals 13, no. 7: 1163. https://doi.org/10.3390/met13071163

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