Next Article in Journal
Carbonaceous Nanomaterials Employed in the Development of Electrochemical Sensors Based on Screen-Printing Technique—A Review
Next Article in Special Issue
Exploring the Photothermo-Catalytic Performance of Brookite TiO2-CeO2 Composites
Previous Article in Journal
Stable Continuous Production of γ-Valerolactone from Biomass-Derived Levulinic Acid over Zr–Al-Beta Zeolite Catalyst
Previous Article in Special Issue
Enhanced Photocatalytic CO2 Reduction over TiO2 Using Metalloporphyrin as the Cocatalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Developments of Advanced Ti3+-Self-Doped TiO2 for Efficient Visible-Light-Driven Photocatalysis

1
Department of Chemistry, Sungkyunkwan University, 2066 Seoburo, Jangan-gu, Suwon 16419, Korea
2
Center for Integrated Nanostructure Physics (CINAP), Institute for Basic Science (IBS), 2066 Seoburo, Jangan-gu, Suwon 16419, Korea
3
Department of Biophysics, Sungkyunkwan University, 2066 Seoburo, Jangan-gu, Suwon 16419, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(6), 679; https://doi.org/10.3390/catal10060679
Submission received: 30 April 2020 / Revised: 5 June 2020 / Accepted: 13 June 2020 / Published: 17 June 2020
(This article belongs to the Special Issue Recent Advances in TiO2 Photocatalysts)

Abstract

:
Research into the development of efficient semiconductor photocatalytic materials is a promising approach to solving environmental and energy problems worldwide. Among these materials, TiO2 photocatalysts are one of the most commonly used due to their efficient photoactivity, high stability, low cost and environmental friendliness. However, since the UV content of sunlight is less than 5%, the development of visible light-activated TiO2-based photocatalysts is essential to increase the solar energy efficiency. Here, we review recent works on advanced visible light-activated Ti3+-self-doped TiO2 (Ti3+–TiO2) photocatalysts with improved electronic band structures for efficient charge separation. We analyze the different methods used to produce Ti3+–TiO2 photocatalysts, where Ti3+ with a high oxygen defect density can be used for energy production from visible light. We categorize advanced modifications in electronic states of Ti3+–TiO2 by improving their photocatalytic activity. Ti3+–TiO2 photocatalysts with large charge separation and low recombination of photogenerated electrons and holes can be practically applied for energy conversion and advanced oxidation processes in natural environments and deserve significant attention.

Graphical Abstract

1. Introduction

Environmental pollution and sustainable energy development are controversial issues we all face today [1]. In keeping with the pace of technological development, the development of renewable energy technologies must be studied in depth. Solar energy is the most abundantly available energy source. An amount as large as almost 99.9% of the energy used on Earth is produced by the sun [2,3,4]. Harnessing this enormous amount of solar energy would provide sustainable and environmentally friendly energy. It is also a promising option for economically feasible technology development.
Semiconducting photocatalysts are an eco-friendly and promising technology that converts solar energy into chemical energy. It can be used in various fields such as photolysis of harmful chemicals [5,6,7,8], artificial photosynthesis [9,10,11,12], photocatalytic water splitting [1,13,14,15,16,17] and electrochemical energy conversion [18,19,20,21,22,23]. Semiconducting photocatalysts using solar energy can solve environmental problems and have been actively researched around the world; as such, they have the potential to become a sustainable energy source [24,25,26]. Recently, various materials have been studied, including conjugated polymers (CPs) [27,28,29,30], 2D-layered materials [25,26,31,32,33,34], semiconductor quantum dots (QDs) [35,36,37,38,39], graphene quantum dots (GQDs) [40,41], isolated single-atom site catalyst (ISAS-catalyst) [42,43,44,45] and MOF-based materials [46,47,48]. However, semiconducting photocatalysts containing TiO2 remain the most popular. There are several advantages of TiO2 photocatalysts, including their low cost, good chemical and physical stability, lack of toxicity to humans and high photoreactivity [49,50,51,52,53].
However, there are also disadvantages that limit the practical application of TiO2. First, the recombination of light-generated electrons and holes easily occurs, hindering the efficient use of absorbed energy [54,55,56]. Second, the TiO2 surface is generally hydrophilic, so its absorption capacity for aromatic organic pollutants is low, limiting its responsiveness [54,57,58,59,60]. Though there are many other issues to overcome, the wide bandgap (~3.2 eV) of TiO2 requires ultraviolet light for photoactivity [24,54]. Only 5% or less of sunlight is UV light [24,53,59]. This means that TiO2 cannot use the visible-near IR energy that comprises most sunlight. Therefore, great efforts have focused on research to improve charge separation, surface reactivity and light absorption to overcome the disadvantages of TiO2 [12,61,62,63,64]. Increasing solar energy efficiency and using more energy are major challenges for expanding the light-activated region into the visible region.
Adjusting the band structure to harness visible light can result in higher photon harvesting. To achieve this goal, attempts have been made to understand and modify the physical and chemical properties of TiO2, such as the slow-light effect in periodic TiO2 photonic crystals [65] and the multi-reflection effect in semi-hollow TiO2 spheres [66]. Attempts have also been made to improve TiO2 by classic chemical element doping [67,68,69,70,71], dye sensitizing [72,73], formation of semiconducting junctions [74,75,76,77], localized surface plasmon resonance (LSPR) photosensitization [78,79], stoichiometric adjustment or coordination of surface atoms or surface oxygen vacancies [61,80,81,82,83] and tuning inter-particle interactions of TiO2 nanocrystals [84,85]. In particular, Ti3+-self-doped TiO2 (Ti3+–TiO2) such as blue-black TiO2 has been found to induce high efficiency in visible-light-driven photocatalysis [86,87,88,89,90,91,92,93,94]. Nonetheless, improvements in photogenerated charge separation in Ti3+–TiO2 remain challenges for a high photon harvesting efficiency. Thus, Ti3+–TiO2 has been developed with modifications of the electronic structures to extend TiO2 photocatalyst to highly efficient visible light harvesting.
In this review, we focus on the development of advanced Ti3+–TiO2 for efficient solar energy harvesting of TiO2 photocatalysts. Ti3+–TiO2 is oxygen-deficient (TiO2−x) and was first reported in 2011 by Chem et al. They produced defects on the TiO2 surface and called the resulting material ‘disorder-engineered black TiO2.’ This method reduced the light absorption band gap of to 1.5 eV, suggesting the possibility of using it for visible light catalysis by increasing solar absorption efficiency. [92] Since then, TiO2 has been reduced to produce blue TiO2 [87], and great efforts have been put into developing blue-black TiO2 to use visible light, and much progress has been made. Figure 1 shows the schematic of the nanoparticle structure of Ti3+–TiO2 and the band location for the insertion of each dopant.
We provide the latest reports on synthesis methods of Ti3+–TiO2-based photocatalysts and their application efficiencies. First, we analyze the physical concept of Ti3+–TiO2 photocatalysts and their production and advancement. We discuss the shortcomings in the production of Ti3+–TiO2 and recent reports on the use of other materials to enhance the absorption of visible light. Moreover, many efforts have been discussed to overcome the disadvantages of photocatalytic efficiencies of Ti3+–TiO2 due to low charge separation and high recombination of photogenerated electrons and holes. Finally, we present some perspectives on the future of modified Ti3+–TiO2 photocatalysts.

2. Preparation of Ti3+–TiO2

Ti3+–TiO2 (TiO2−x) has been demonstrated to enhance visible light absorption [95]. Ti3+-self doping causes defect levels due to oxygen vacancies in the lattice structure of self-doped TiO2. The defect levels overlap the conduction band of TiO2 and can induce vacancies in the electronic state just below the conduction band [96]. As a result, the band gap is reduced, resulting in enhanced visible light absorption. Methods for producing Ti3+–TiO2 have been a focus of many studies. In general, high pressure and high temperature are used to reduce oxygen to form oxygen vacancies. Ti3+–TiO2 doped with high oxygen vacancies is super hydrophilic. Thus, Ti3+–TiO2 on hydrophilic substrates can result in strong bonding to the surface, thus increasing the coating ability.
According to recent trends, there are many reduction methods of TiO2, which we divide into five main branches: hydrogenation, hydrothermal reaction, alkaline metal reduction, sol-gelation and phase-selective reduction. As generations go through, the preparation method of Ti3+–TiO2 is not always superior, but it is considered to be a little more relaxed and more sophisticated.

2.1. Hydrogenation

Hydrogenation using hydrogen gas to react with oxygen atoms in the lattice of TiO2 is the most common method for Ti+3 doping. Ti3+ is self-doped in an abundance of TiO2 by high thermal or electromagnetic energy [97,98]. Hydrogenation requires harsh synthetic conditions and dangerous production processes (i.e., high heat and pressure, risk of explosion of hydrogen gas) [92,97,98,99]. To overcome this, other reducing agents that can provide hydrogen such as NaBH4 and TiH2 have been introduced as alternatives to hydrogen gas [98,100,101,102]. Additionally, increasing the rich Ti3+ doping involves higher heat, higher pressure and longer reaction times. Well-crystallized samples can effectively enhance visible and infrared absorption and exhibit hydrophilicity by oxygen doping with H in the lattice. However, in hydrogenated TiO2, a high concentration of hydrogen atoms was required for efficient photocatalytic activity [103].
Additionally, if the temperature increases under firing conditions, phase conversion from anatase to rutile may occur. This can result in a band gap shift from anatase (~3.2 eV) to rutile (~3.0 eV). Phase transformation requires 600 to 700 degrees of energy in the absence of impurities, but the presence and amount of oxygen vacancies can lower the firing temperature. The increased oxygen deficiency causes relaxation of the large oxygen sublattice and facilitates the rearrangement and modification of atoms from anatase to rutile. In a high-pressure reducing atmosphere, heat transfer is also advantageous because of the high material density. Fast heat transfer will further promote anatase-to-rutile phase transformation [104]. However, the abundant doping of Ti3+ results in an irregular crystalline phase, called a disordered phase. Disordered crystals may behave differently from the normal physicochemical properties of anatase and rutile crystals.

2.2. Hydrothermal Reaction

A method of manufacturing Ti3+–TiO2 with abundant defects has been reported using TiCl3 and TiF4 as precursors. Ti3+ can easily be oxidized under high heat and pressure conditions. Introduction of Ti4+ can inhibit oxidation of Ti3+ in this reaction, leading to abundant doping [105]. In addition, Fang et al. synthesized various reduced TiO2 samples using Zn powder as a reducing agent and HF as a solvent in a simple one-pot hydrothermal process. However, Ti3+ introduced by Zn reduction is not stable in air and can be easily oxidized [106].
This method, which has been slightly improved under hydrogenation conditions, still requires high heat and pressure. It has an advantage in that the reaction proceeds in a liquid state that is easier to handle than hydrogen gas. However, it is unstable in the air and hydrogen saturation is not effective. A new solvent is needed to overcome this stability issue.

2.3. Alkaline Metal Reduction

Researchers tired of using high heat and pressure gases and liquids. In general, attempts have been made to reduce TiO2 by using alkali metal, the most commonly used solid reducing agent. It is possible to obtain Ti3+–TiO2 by simply replacing the high heat and pressure and simply mixing the powder. In 2013, Yin et al. produced gray TiO2 rutile nanowires by reducing the TiO2 surface by directly introducing aluminum metal at a temperature of 700 °C [107]. There have also been attempts to generate Ti3+ under milder conditions. In 2017, Zhang et al. sufficiently reduced TiO2 using Na/NaCl powder under room temperature and an argon atmosphere and treated it with Ru to synthesize nanoparticles of TiO2 with Ti3+ and Ru. Crystalline TiO2 was milled with Na and NaCl powder at room temperature for 0.25 to 4 h under an argon atmosphere. It can be seen from Figure 2 that the color of P25 becomes darker as it undergoes harsh reaction conditions. Ti3+-rich TiO2 nanocrystals can be used as effective support for Ru particles and the Ru/Ti3+–TiO2 catalysts showed excellent performance in catalytic hydrogenation of N-methyl pyrrole [108].

2.4. Sol-Gelation

Since then, people have shown how to synthesize TiO2 doped with Ti3+ abundantly, starting with the precursor building block, rather than doping Ti3+ into the already prepared TiO2. Zhang et al. synthesized TiO2 single-crystal nanorods using a sol-gelation method and succeeded in obtaining Ti3+-self-doped blue TiO2 single-crystal nanorods by further annealing at 350 °C. Blue TiO2 showed a 97.01% higher decomposition rate of RhB than did white TiO2 under visible light, and the photocatalytic hydrogen generation rate increased. As a result of these investigations, the band gap was reduced to 2.61 eV due to the synergistic effect of the 1D-shaped single crystal structure and this increased the photocatalyst’s visible light absorption ability and photocatalytic reactivity [109]. In 2018, Yao et al. synthesized N/Ti3+-doped TiO2 in one step using TiN. Additionally, using BiOBr, N/Ti3+-doped TiO2 and heterojunctions were formed to enhance sonocatalytic activity. Doping of BiOBr and N/Ti3+ leads to effective charge separation of electron—electron pairs, increasing the methylene blue removal efficiency by 1.81 times [110].
In general, the grown TiO2 crystal phase is anatase. However, depending on the reaction conditions, rutile anatase may also be present. Among the synthetic conditions, there is a possibility that BiOBr promoted rutile growth [110]. Optimal performance is achieved by mixing anatase and rutile in appropriate proportions (85:15). However, it is generally found that the photocatalytic efficiency of anatase is higher. Due to the high density of localized states in anatase, it is used for surface adsorbed hydroxyl radicals and slow charge carrier recombination. Rutile has a high charge recombination rate and consequently has a reduced ability to adsorb functional species, which is detrimental to photocatalytic performance [104].

2.5. Phase-Selective Reduction

There are many ways to produce Ti3+–TiO2, but almost all of these require high temperature and high pressure. In addition, it is not yet possible to selectively reduce anatase and rutile phase on TiO2 crystals in harsh conditions of high temperature and high pressure. In 2016, Zhang et al. first developed a method for selectively reducing the crystal phase of Degussa P25 TiO2 nanoparticles at room temperature and atmospheric pressure using a simple Li-Ethylene diamine (EDA) solution. They melted Li metal in EDA solution at room temperature and pressure to produce a blue electride solution of a strong reducing agent to reduce TiO2 nanocrystals. The Li-EDA solution is unique in reducing only the rutile phase while the anatase phase remained the same as in the composition of Degussa P25. They confirmed that Ti3+ was selectively introduced into TiO2 to significantly improve visible and near-infrared absorption and increase the charge-hole charge separation efficiency through type II band gap alignment (Figure 3). These effects promoted a strong hydrogen-generating surface reaction. Therefore, when using phase-selective disorder engineering, 0.5 wt% Pt and 3.46-mmol were used to show high stability and a high hydrogen generation rate of 13.89 mmol·h−1·g−1 [87]. In addition, in 2019, Hwang et al. discovered that when P25 is treated with Na-EDA solution, only the anatase phase is reduced. Engineered Ad/Ro TiO2 selectively treated with Na-EDA to reduce the anatase phase contained only Ti3+ defect sites in the anatase phase and the internal energy band gap was narrowed. This indicates enhanced visible light activity. Hwang et al. subsequently reduced CO2 using Ti3+-self-doped Ad/Ro TiO2 that selectively reduced only the anatase phase to produce CH4 using light in the visible region [111]. The phase-selective method of reducing TiO2 using a metal-EDA electride solution has attracted attention as a promising method of self-doping rich Ti3+ on the surface of TiO2 at room temperature and atmospheric pressure to improve the visible light activity of TiO2. The developed Blue TiO2 (BTO) was used for organic arylation reactions, the yield of which was supported with untreated P25 (40%) and BTO (63.4%) [112]. In addition, the BTO had excellent reactive oxygen species (ROS) generation and showed strong efficacy in decomposing algae (Chlamydomonas) under sunlight [113].
The phase-selective reduction of P25 does not require heat or pressure. In addition, while selectively reducing rutile and anatase TiO2 crystal phases, it became possible to impart selectivity [114]. With the selective reduction of rutile and anatase, their physicochemical properties are clearly different. We expect to pioneer new applications in various fields by applying this method.

3. Enhancement of Photogenerated Charge Separation and Photocatalytic Activity in Advanced Ti3+–TiO2

Photocatalysts exhibit catalytic activity when introduced to light. The photogenerated electrons and holes move to the photocatalytic surface. The moved electrons and holes react with adsorbed electron acceptors and donors, respectively, to complete the catalytic reaction. Abundant Ti3+ doping can lead the excitation region from ultraviolet to visible light. Ti3+-doped TiO2 can narrow the wide band gap of TiO2 for harvesting visible light and can provide an increase in electronic conductivity due to doped Ti3+ defects. However, after photoexcitation, the generated electrons and holes may undergo significant electron recombination in the depletion region of Ti3+-doped TiO2 [115,116,117].
To overcome this and ultimately enhance photocatalytic activity in all applications, deformation studies with several different materials have been conducted. Performance can be improved by using Ti3+–TiO2 in combination with other materials or elements. Studies using various other materials have been conducted to overcome the previously mentioned limitations of Ti3+–TiO2 deformation.
We focus on the synthetic methods of advanced Ti3+–TiO2 with high photogenerated charge separation and efficiency and classify them into four categories according to recent developments. Ti3+–TiO2 modification has been extensively studied, including metal and nonmetal doping, semiconducting coupling, and stoichiometry modification to prepare and stabilize composites with other materials.

3.1. Metal-Doping

It has long been a major concern to increase the efficiency of TiO2 photocatalysts and expand their use in the visible region using easily accessible metal dopants. Metal dopants can be mixed with Ti3+ atoms or replaced by Ti atoms, and the band gap can be adjusted in such a state. Moreover, a new band was produced as a conduction band to help better separate photogenerated charges after photoexcitation. Metal dopants act as an electron acceptor as well as a hole acceptor in the valence band. The introduction of metal dopants is still a promising option for Ti3+–TiO2. We summarized these efforts in Table 1.

3.1.1. Surface Plasmon Effects on Ti3+–TiO2

LSPR-based photosensitization can promote the absorption of visible or near infrared rays [122,123]. If the frequency of the incident photon coincides with the natural frequency of the vibrating surface electrons, resonant group oscillations of the valence electrons can occur to promote absorption of Ti3+-self-doped TiO2 visible light [124]. Usually, LSPR excitation occurs in the visible region when Au [125], Ag [126], and Pt [127] are combined with TiO2 [128]. Attempts to combine Ti3+-self-doped TiO2 and LSPR metals have continued. However, in almost all cases, noble metals should be used for plasmon resonance [125,126,127]. The use of precious metals is still valid. In 2019, phase-selective room-temperature solution engineering was used to enhance HER performance by attaching Pt particles to TiO2 richly doped with Ti3+ [114]. This indicates that the simple phase-selective room-temperature solution engineering method does not lag behind TiO2 manufactured in other harsh conditions of different reducing conditions.
Recently, He et al. used Ti3+ and Ni for plasmon-mediated carrier transfer to enhance the degradation of methylene blue in visible light. They deposited Ni–TiO2 structures on SiO2 spheres based on the modified nanosphere lithography method. It was found that plasmon-generated hot electrons and Ni holes can be transferred to TiO2 in a heterogeneous structure (Figure 4). The transferred hot electrons and holes occupy oxygen vacancies or produce Ti3+. TiO2 self-doped with Ti3+ showed enhanced removal performance of methylene blue of 0.11 ± 0.04 μmol·L−1·min−1 in visible light [69].
Furthermore, it is valuable to be able to achieve the same effect as LSPR by using inexpensive non-precious metal transition metals other than Ni. We believe that developing a method of applying transition metals will become the main approach to metal-doped Tt3+ TiO2 in the future.

3.1.2. Single Atom Site Doping Effects on Ti3+–TiO2

Isolated single atom site (ISAS) catalysts are a new approach in the field of catalysis due to their increased catalytic activity and excellent selectivity. Hejazi et al. created a stable platinum single atom site by grafting Pt single atoms to the Ti3+-self-doped TiO2 anatase (001) nanosheet by hydrogenation [89]. After reducing the thin Ti3+layer in the surface Ar/H2 environment, it was used as a Pt single atom support and tightly adhered from the diluted aqueous Pt solution. The amount of Pt deposition can be adjusted by considering the heat treatment temperature in Ar/H2. This showed a 150-fold improvement in hydrogen generation rate over that observed using the same amount of platinum nanoparticles (Figure 5) [89]. Despite many issues, including the stability of ISAS catalysts, this is a very attractive approach. With a small amount of metal, a photocatalytic performance of tens to hundreds of times improvement is observed. In addition, we believe that a single atom leads to a widely applicable approach for many. The development of the ISAS catalyst using Ti3+ as support can provide a new approach for Ti3+-self-doped TiO2 for absorbing visible light.

3.1.3. Transition or Rare Metal Grafting on Ti3+–TiO2

In general, the bonding of transition metals and TiO2 is a classic method. Similarly, much effort has been put into grafting transition metals into Ti3+–TiO2. Recently, research on photocatalysts for removing microorganisms and viruses was actively conducted in addition to the removal of existing classical organic or inorganic contaminants. Zhang et al. reported cell lysis of E. coli and agropathogenic fungal spores by preparing classical Cu and Ti3+–TiO2 in a new way. It was reported that anhydrous ethanol can be used as a reducing agent in the hydrothermal process leading to the formation of metallic copper and Ti3+. They confirmed the photocatalytic disinfection effect on five agricultural pathogenic fungal spores, including F. graminearum and B. dothidea spores, by fluorescence staining using a FungaLight CFDA-AM/propidium iodide yeast vitality kit. Almost all bacteria were killed upon irradiation for 3 h. In Figure 6, killed bacteria are identified by fluorescent staining. Both copper and other metals are expected to be effective in bacterial and viral cell lysis when combined with Ti3+-self-doped TiO2 [120].
Recently, grafting of rare earth metals such as Bi (rather than classic commercially available metals such as Cu, Zn and Co) into Ti3+–TiO2 has been reported. Gul et al. increased the solar reactivity by grafting Bi to Ti3+–TiO2. They performed extensive doping with Ti3+ and Bi-by-Bi doping and sol-gel technology and showed better photocatalytic performance than undoped TiO2 for flumequine decomposition [118].
As shown in Figure 7, Zhou et al. introduced doping with six rare earth metals (i.e., La, Ce, Pr, Nd, Eu and Gd) and sepiolite with Ti3+–TiO2 to improve visible light activity and provide low band gap energy. In addition, this strategy resulted in a high adsorption capacity of contaminants through the sepiolite and improved the photocatalytic performance. Figure 7 shows the schematic structure for modification of the conduction band maximum and the photodegradation effect of Orange G for each metal. They found that, upon doping of rare earth metal ions, the structure and chemical properties of the nanocomposite of TiO2 depended heavily on the radius of the rare earth metal ion. Rare-earth ions with a valence of +3 (i.e., RE+3) can be abundantly doped with Ti3+ because the rare earth metal ions are much larger than that of Ti4+. However, this can cause a charge imbalance that can contribute to the adsorption of OH. It is also reported that the empty 4f level of rare earth metals can act as a scavenger for light-generating electrons. Of the six rare earth metals, the Eu complex showed the best photocatalytic performance [121]. The combination of common transition metals and rare earth metals with Ti3+-self-doping can be an excellent way to enhance photocatalytic performance, and promising materials for cell lysis of viruses and bacteria can be developed.

3.2. Nonmetal Doping

Much effort has been made to expand the applications of TiO2 photocatalysts with metal and nonmetal based dopants into the visible light regime and increase the photocatalytic efficiency. In particular, many common elements such as nitrogen, carbon, sulfur and phosphorus have been used for grafting. The grafting of nonmetallic elements such as these can lower the surface energy for adsorption of the photocatalytic surface [129]. Unlike metal dopants, nonmetal dopants can be directly involved in the band of TiO2. In addition, nonmetallic elements with a different electronic environment from oxygen can have positive and negative effects on the band gap energy. We believe that such modification through band engineering is a unique possibility for nonmetal dopants. We summarize the recent Ti3+ and nonmetallic dopant grafting in Table 2.

3.2.1. N-Doped Ti3+–TiO2

Nitrogen is one of the most frequently used nonmetal elements for adjusting the valence band of TiO2. Doping with nitrogen may cause the maximum upward movement of the valence band, and visible light absorption may also be improved by controlling the red-shift of the absorption edge. Nitrogen can improve both carrier density and electron conductivity in the neutral region [129]. Additionally, it can act as an absorption site for organic pollutants. Recently, efforts have been made to enhance visible light absorption and enhance photocatalytic performance using Ti3+ and N dopants together. By doping N and Ti3+ together, Ti3+ in the band decreases the conduction band, and N increases the valence band to reduce the band gap energy and enhance the photocatalytic performance. Jia et al. improved the charge separation by introducing carbon after N doping, introducing carbon components, and using carbon as a site to accept the separated charge after photoexcitation. This showed a significantly faster Rhodamine B decomposition rate than conventional nitrogen-doped Ti3+-self-doped TiO2 (Figure 8) [70].

3.2.2. Sulfur-Doped Ti3+–TiO2

Among nonmetal dopants, sulfur has the same valence as oxygen, has adequate energy, and may not cause charge imbalance in TiO2. The energy band occupied by the charge imbalance can act as a carrier recombination site and degrades the carrier transport process [129,135]. 4 Based on these advantages, sulfur has been noted as another promising nonmetallic dopant to drive the performance of TiO2 photocatalysts. Active research with Ti3+ is ongoing. Ji et al. introduced S2− into anatase TiO2 together with Ti3+, which significantly lowered the band gap energy and increased photocatalytic hydrogen production. They confirmed that, by introducing sulfur and Ti3+, the band gap decreased from 3.2 eV for the existing anatase TiO2 to 2.0 eV after co-doping [132]. As shown in Figure 9, Meng et al. added plasmonic Ag particles to TiO2 nanorods introduced with sulfur and Ti3+ to synergistically reduce the band gap through impurity dopants and expand LSPR into the visible light regime. These are photocatalytic reactions where phenol’s visible-light-driven photocatalytic decomposition rate and hydrogen production rate are as high as 98.67% and 209.2 μmol·h−1·g−1, approximately 2 and 5 times higher than TiO2 alone [133].

3.2.3. Multi-Doped Ti3+–TiO2

Several elemental dopants can be introduced at the same time. Yan et al. reported black TiO2 nanosheets with Ti3+ by doping nitrogen, carbon and sulfur. The doped C, N, S elements can result in additional impurity levels above the valence band. The mass concentrations of doped C, N and S are 14.20%, 4.80% and 3.30%, respectively, and these worked synergistically with Ti3+. They used visible light to photocatalytically decompose methyl orange up to 92.13%. The amount of hydrogen generated by the photocatalyst was 149 μmol·h−1·g−1 [134]. An attempt to use multiple materials without being limited to a single material may provide a new way to improve photocatalytic performance.

3.3. Semiconducting Coupling

Semiconducting heterogeneous catalytic junction enhances various properties by adding a new band to Ti3+–TiO2. The newly added band may additionally absorb visible light, promote optical junction carrier separation and transfer and enhance interactions with the target material. Semiconducting heterojunctions are advantageous for absorbing visible light of photocatalysts and increase photocatalytic activity by promoting separation and transfer of photojunction carriers of semiconductors [136,137]. Previous bands of semiconductor materials such as CdS [138], Ag3PO4 [139], Cu2O [140] and WO3 [141], have been combined with TiO2. Recently, g–C3N4 has shown overall water splitting band potential and high visible light response, leading to the current trend [142,143]. Likewise, when Ti3+–TiO2 is combined with semiconducting materials, better synergy can be expected than before. We looked at the recent trends of Ti3+–TiO2 and organized them in Table 3.

3.3.1. Transition Metal Dichalcogenide or Quantum Dot/Ti3+–TiO2

Transition metal dichalcogenide (TMD) materials based on quasi-dimensional (2D) structures have recently attracted attention because of their excellent electronic and optical properties. These materials have great potential for high-tech applications and have different properties compared to conventional bulk counterpart materials. The state density of the TMD material increases in a semi-continuous stepwise fashion. Therefore, TMD materials can exhibit new electronic and optical properties [25,26]. MoS2 is a representative TMD material, and it is a popular material in HER photocatalysts and electrocatalysts. This is because the sandwiched MoS2 atomic layer can provide various unsaturated bonds at the edge of the layer. There have been attempts to enhance the photocatalytic performance with recent Ti3+–TiO2 using these new electronic and optical properties. There have been attempts to dope the MoS2 layer and other semiconducting or metal materials over spear-shaped black Ti3+–TiO2. Attempts have been made to bond N/Ti3+–TiO2 with spear and MoS2 and to apply hollow Ti3+–TiO2 and MoS2/CdS together [76,147]. All of the materials were able to clearly identify the vertically grown layer and showed excellent performance in removing HER and Methylene orange. Figure 10 shows the schematic structure and mechanism of action of hollow black TiO2/MoS2/CdS. In addition to TMD metal, there have been attempts to combine Ti3+–TiO2 with a layered material represented by bismuth oxyhalides. Yao et al. confirmed the well-grown TiO2 and BiOBr nanostructures by synthesizing N/Ti3+–TiO2 and BiOBr in one step. These materials showed high efficiency in removing methylene blue due to increased visible light absorption [110].
In addition to 2D-based layered materials, zero-dimensional quantum dots using TMD materials are also attracting attention because of their small size, large specific surface area and shortened charge travel distance. Many recent TiO2 quantum dot complexes have attracted considerable attention. Coupling of TiO2 and semiconducting quantum dots can improve the charge separation of photo-generated electron-electron pairs and smaller quantum dot materials are uniformly dispersed on the TiO2 surface, preventing aggregation of quantum dots and providing more active sites [153,154]. However, the synergy of the existing TiO2/QD materials was not satisfactory [155,156]. Therefore, a catalyst using visible solar energy was achieved with the introduction of Ti3+–TiO2. Likewise, there have been recent attempts to modify the MoS2 and CdS QDs through conjugation of Ti3+–TiO2 as shown in Figure 11 [146,150]. Abundant Ti3+self-doping sufficiently improves the visible light response, and QD MoS2 and CdS can promote photocatalytic carrier separation through a built-in electric field to provide a rich specific surface area to improve photocatalytic performance.
In addition to TMD materials, there have also been attempts to integrate Ti3+–TiO2 with QD. Ag3PO4 attracted attention because of its high quantum yield and band gap energy (2.45 eV) suitable for absorbing visible light. However, Ag3PO4 was only used as a photocatalyst due to fast charge recombination and relatively large particle size. Overcoming these shortcomings with QD Ag3PO4 and bonding with Ti3+–TiO2 increased visible light responsiveness and improved photocatalytic performance (Figure 12) [145]. The introduction of quantum dots suggests a new paradigm to solve problems that could not be solved with the existing classical materials. Richly doped Ti3+ can serve as an effective support for quantum dot materials and as an electron carrier delivery site. Thus, richly doped Ti3+ has very promising and potential possibilities, requiring more in-depth research into this area.

3.3.2. Carbon-Based Nonmetallic Semiconductors/Ti3+–TiO2

Carbon-based nonmetallic semiconducting materials are attracting great attention in the energy industry because of their attractive properties such as high conductivity and easy accessibility. The two-dimensional electronic network of sp2 carbon atoms exhibits high electron mobility and conduction [157]. In the past few years, carbon-based materials have been extensively explored as photocatalysts. Recently, graphitic-carbon nitride (g–C3N4) has attracted much attention as a visible light active catalyst due to its appropriate band gap (2.69 eV). In particular, TiO2 and Z-structure photocatalyst band alignments are similar to the plant photosynthesis process and can provide high photocatalytic activity through more powerful charge separation than type I and type II band alignment [158,159]. Ti3+–TiO2 can strengthen this Z-structure with abundant Ti3+ doping in the conduction band. Figure 13 shows the Z structure diagram between TiO2 and g–C3N4. Recently, studies on enhancing the performance of visible light-activated photocatalysts have been actively conducted through the combination of Ti3+–TiO2 and g–C3N4. There have been attempts to improve charge separation by doping plasmon metal between Ti3+–TiO/g–C3N4. Cao et al. introduced Ag-doped black TiO2 nanosphere structures and g–C3N4. Figure 14 shows a schematic of the Ti3+–TiO2/Meso-g–C3N4 manufacturing method. They showed that Ag could act as an electron-conducting bridge between TiO2 and g–C3N4, contributing to electron donors and acceptors in the Z-structure [151].

3.3.3. Metal Oxide/Ti3+–TiO2

Heterojunctions of metal oxide and TiO2 are a classic (and the easiest) method to enhance TiO2 photocatalytic activity after metal doping. To date, new heterojunctions of various metal oxides and TiO2 have been studied. Heterojunctions of Ti3+–TiO2 and metal oxides are also of great interest. Although many metal oxides are candidates, heterojunctions of WO3 and Ti3+–TiO2 have recently been developed. WO3 can achieve Z-structure band alignment with Ti3+–TiO2. The resulting Z-structure has a band structure similar to natural photosynthesis [158,159]. This can strongly increase the charge separation efficiency. In addition, two different redox sites can suppress the reverse reaction of radical production. In 2020, Nguyen et al. developed a catalyst that introduced blue Ti3+–TiO2 and WO3 produced by lithium (Li–EDA) of ethylenediamine, a strong reducing agent of the previous superbase. They show 100% selectivity in reducing CO2 to CO, and that blue Ti3+–TiO2/WO3 formed a perfect Z-structure (Figure 15) [91]. Consequently, the introduction of a complete Z-structure can be a powerful modification of Ti3+–TiO2. This is a direct way to overcome the weakness of charge separation and recombination and dramatically enhance the performance of photocatalysts.
There have also been attempts to make heterojunctions with CeO2, which has a band gap energy similar to TiO2. CeO2 has recently attracted attention due to the special redox properties between Ce (III) and Ce (IV) oxidation states, which include high thermal stability, excellent oxygen storage capacity and easy conversion. Xiu et al. developed Ti3+/Ce3+ self-doped TiO2/CeO2 nanosheets by reducing TiO2 and CeO2 together. They showed that Ti3+ and Ce3+ complement each other and enhance electron separation (Figure 16) [148].

3.4. Stoichiometry Modification

There have been attempts to change the stoichiometry of materials beyond controlling Ti3+-self-doped intrinsic defects to enhance the visible light absorption of TiO2 photocatalysts. Introduction of defect states such as Ti3+ is sensitive to the chemical state and spatial distribution of electronic structure modifiers. Therefore, photocatalytic activity may be reduced through the formation of an undesirable band structure. It is important to create favorable defects at the atomic level to improve absorption in the visible light band. However, it is still difficult to correct structural defects at the atomic level and produce an optimal TiO2 material. A study on correcting atomic-level structural defects is ongoing beyond the recent Ti3+–TiO2 work. Efforts have been made to replace hydrogen vacancies in Ti3+–TiO2 with atomic levels of hydrogen [81], carbon [82] and nitrogen [82,83]. We summarized in Table 4.
In general, reduced Ti3+–TiO2 has a blue-black color, whereas substituted TiO2 has a different color. Red TiO2 appears when the space for a single oxygen atom is filled with two hydrogen atoms, called Ti3+H TiO2 [81]. As shown in Figure 17, the color and atomic/band structure of TiO2 was changed according to the vacancies in the crystal structure. In addition, the introduction of heterogeneous elements results in a new band state. The band gap is reduced due to changes in the conduction band and valence band, and both show excellent activity in visible light.
We believe that engineering the oxidation state of Ti atoms (beyond Ti3+, which is simply reduced) and the surrounding electronic environment through complete stoichiometric control is a new paradigm for developing effective TiO2 visible photocatalysts. In the future, we can expect to expand and improve the photocatalytic performance through visible light for TiO2 photocatalysts via uniformly controlled stoichiometric modifications.

4. Conclusion and Outlook

TiO2-based photocatalytic reactions were first reported in the early 1930 s. They were first used to bleach dye using reactive oxygen species generated under UV light. TiO2 photocatalyst (called a “photosensitizer”) has been studied extensively, and it is expected to play a role in solving the limited energy problem [24]. In particular, efforts have been made to improve the absorption band of TiO2 photocatalyst into the visible light regime to efficiently use solar energy. These efforts continued to Ti3+–TiO2, and since the report of black Ti3+–TiO2 reduced by Ward et al. in 2011, Ti3+–TiO2 research has made great progress [92]. We first studied the development of TiO2 reduction to produce abundant Ti3+ for overcoming the high heat and pressure of the hydrogenation reaction (as the first generation of Ti3+–TiO2-based photocatalysts). Then, due to their long reaction times and risks, increasing mild and easy to handle methods have been developed as the next generation. A lithium-ethylenediamine solution (Li-EDA) has been introduced for self-doping Ti3+ effectively at room temperature and atmospheric pressure, as well as controlling the crystalline phase of reduced TiO2. Selective reduced crystal phases with different physicochemical properties will pioneer photocatalytic utilization for several properties [87]. However, although Ti3+ is abundantly doped such that the band region for visible light is expanded, there are still limitations. The interaction and reactivity between target materials and photocatalysts, as well as effective suppression of charge separation and recombination, are still challenges that need to be overcome. The best way to improve the limited performance of these single materials is with several different composite materials.
In this review, we focused on the types and methods of Ti3+–TiO2 and the materials introduced and categorized them into four categories according to recent developments. Ti3+ doping can be used as sufficient support for an isolated single-atom site catalyst and can stabilize the catalyst atom. The isolated single-atom catalyst has excellent selectivity and reactivity with a small amount of catalyst [89]. Introducing metals other than plasmonic precious metals will make commercialization easier. The introduction of LSPR using isolated single atom site catalysts and non-precious metals in metal dopants is the most effective modification method. Going further, bimetallic sites, which have been extensively studied recently, are being proposed as a new way to solve the selectivity and stability issues that a single atom catalyst cannot. There are different advantages depending on the metal combination, and there are even advantages to not using precious metals. It has recently been shown that bimetallic site catalysts can also be applied with TiO2 [160]. However, this approach requires more research of bimetal sites.
The introduction of nonmetal dopants is a way to engineer the fundamental band structure of TiO2. Modification of the band structure can make it easier to expand this approach to visible photocatalysts, and most nonmetallic dopants can improve the decomposition of harmful substances through strong interactions with organic harmful substances. Another approach is doping two or three heterogeneous elements beyond single element nonmetallic doping. The introduction of a suitable semiconducting material can be used to overcome the shortcomings of the aforementioned TiO2 photocatalysts. Beyond the classic 3D dopant, the introduction of 2D TMD and 0D QD materials shortens both the large specific surface area and charge travel distance. Because of this, it is possible to improve the charge separation of electron—electron pairs, which could not be solved before. The improved charge separation has high expectations for HER and water splitting. The introduction of metal oxides can also be a great alternative. Due to their excellent accessibility and stability, carbon-based materials can be promisingly combined with Ti3+ TiO2. The Z-structure has a structure similar to that of photosynthesis. The formation of Z-structures through the introduction of metal oxides and carbon-based materials is a powerful approach to improving photocatalytic performance. Similar to a plant’s photosynthetic structure, the photocatalyst can replace various functions of plants in the future. This research field has potential in the energy industry.
Furthermore, advances in material engineering technology can dominate atomic defects. This presents a new base of TiO2 photocatalysts through heteroatom insertion into the TiO2 grid beyond Ti3+ single atom doping [81,82,83]. It is expected that the next generation of Ti3+–TiO2 is in stoichiometric modification. If the electronic environment is completely controlled by controlling the oxidation state of the Ti atom and the surrounding atoms rather than simply doping the reduced Ti3+, there is a possibility to produce a TiO2 photocatalyst suitable for each application method. TiO2 has been studied as a practical photocatalytic material. However, in addition to applications that can be directly applied, in-depth studies on the electron environment, oxidation state of various dopants in the crystal structure, enhanced charge separation, and defect-level band energy are feasible. By investigating the atomic state and bonding level in the final crystal structure, it is possible to identify the principles and mechanisms of theoretical photocatalytic applications. Such an understanding is essential to improving photocatalytic performance.
Advanced Ti3+–TiO2 approaches have recently been extended to visible light in solar energy harvesting and air purification. In particular, we believe that by applying air purification [152], diseases caused by viruses and bacteria [120] in the air—that have emerged as among the biggest problems faced by humanity—can be prevented by using Ti3+–TiO2-based photocatalysts, a visible photocatalyst that is harmless to humans. For this, it is necessary to study the interactions of cells with photocatalysts in a view point of biochemistry, as well as conduct theoretical studies about classical interactions between organic and inorganic and their mechanisms. Nonetheless, the application of visible light photocatalysts in this industry is still insufficient. It is essential to explore sufficient functions required for commercialization [91], further study of Ti3+–TiO2-based materials and the development of more effective heterogeneous materials. This makes Ti3+–TiO2 a viable option for visible photocatalysts in the future.

Author Contributions

S.N. wrote the original draft. S.S. and H.L. edited the draft and supervised it. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Institute for Basic Science grant number IBS-R011-D1, Korea Evaluation Institute of Industrial Technology grant number 20004627, INNOPOLIS Foundation grant number 2019-DD-SB-0602.

Acknowledgments

This work was supported by the Institute for Basic Science (IBS-R011-D1). This work was partially supported by the Korea Evaluation Institute of Industrial Technology (20004627) and the INNOPOLIS Foundation (2019-DD-SB-0602).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ran, J.; Zhang, J.; Yu, J.; Jaroniec, M.; Qiao, S.Z. Earth-abundant cocatalysts for semiconductor-based photocatalytic water splitting. Chem. Soc. Rev. 2014, 43, 7787–7812. [Google Scholar] [CrossRef]
  2. Dickinson, R.E. Land Surface Processes and Climate—Surface Albedos and Energy Balance. In Theory of Climate, Proceedings of a Symposium Commemorating the Two-Hundredth Anniversary of the Academy of Sciences of Lisbon; Saltzman, B., Ed.; Elsevier: Amsterdam, The Netherlands, 1983; Volume 25, pp. 305–353. [Google Scholar]
  3. Dines, W.H. The heat balance of the atmosphere. Q. J. R. Meteorol. Soc. 1917, 43, 151–158. [Google Scholar] [CrossRef]
  4. Stephens, G.L.; Li, J.; Wild, M.; Clayson, C.A.; Loeb, N.; Kato, S.; L’Ecuyer, T.; Stackhouse, P.W.; Lebsock, M.; Andrews, T. An update on Earth’s energy balance in light of the latest global observations. Nat. Geosci. 2012, 5, 691–696. [Google Scholar] [CrossRef]
  5. Kumar, A. A Review on the Factors Affecting the Photocatalytic Degradation of Hazardous Materials. Mater. Sci. Eng. C 2017, 1, 106–114. [Google Scholar] [CrossRef] [Green Version]
  6. Santhosh, C.; Malathi, A.; Daneshvar, E.; Kollu, P.; Bhatnagar, A. Photocatalytic degradation of toxic aquatic pollutants by novel magnetic 3D–TiO2@HPGA nanocomposite. Sci. Rep. 2018, 8, 15531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Panthi, G.; Park, M.; Kim, H.-Y.; Lee, S.-Y.; Park, S.-J. Electrospun ZnO hybrid nanofibers for photodegradation of wastewater containing organic dyes: A review. J. Ind. Eng. Chem. 2015, 21, 26–35. [Google Scholar] [CrossRef]
  8. Gupta, V.K.; Jain, R.; Nayak, A.; Agarwal, S.; Shrivastava, M. Removal of the hazardous dye—Tartrazine by photodegradation on titanium dioxide surface. Mater. Sci. Eng. C 2011, 31, 1062–1067. [Google Scholar] [CrossRef]
  9. Zhang, T.; Lin, W. Metal-organic frameworks for artificial photosynthesis and photocatalysis. Chem. Soc. Rev. 2014, 43, 5982–5993. [Google Scholar] [CrossRef] [PubMed]
  10. Hoffmann, M.R.; Moss, J.A.; Baum, M.M. Artificial photosynthesis: Semiconductor photocatalytic fixation of CO2 to afford higher organic compounds. Dalton Trans. 2011, 40, 5151–5158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Lee, S.H.; Kim, J.H.; Park, C.B. Coupling Photocatalysis and Redox Biocatalysis toward Biocatalyzed Artificial Photosynthesis. Chem. Eur. 2013, 19, 4392–4406. [Google Scholar] [CrossRef]
  12. Ong, W.-J.; Tan, L.-L.; Ng, Y.H.; Yong, S.-T.; Chai, S.-P. Graphitic Carbon Nitride (g–C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer to Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef] [PubMed]
  13. Osterloh, F.E. Inorganic nanostructures for photoelectrochemical and photocatalytic water splitting. Chem. Soc. Rev. 2013, 42, 2294–2320. [Google Scholar] [CrossRef] [PubMed]
  14. Takanabe, K. Photocatalytic Water Splitting: Quantitative Approaches toward Photocatalyst by Design. ACS Catal. 2017, 7, 8006–8022. [Google Scholar] [CrossRef]
  15. She, X.; Wu, J.; Xu, H.; Zhong, J.; Wang, Y.; Song, Y.; Nie, K.; Liu, Y.; Yang, Y.; Rodrigues, M.-T.F.; et al. High Efficiency Photocatalytic Water Splitting Using 2D α-Fe2O3/g–C3N4 Z-Scheme Catalysts. Adv. Energy Mater. 2017, 7, 1700025. [Google Scholar] [CrossRef]
  16. Lin, L.; Yu, Z.; Wang, X. Crystalline Carbon Nitride Semiconductors for Photocatalytic Water Splitting. Angew. Chem. Int. Ed. 2019, 58, 6164–6175. [Google Scholar] [CrossRef]
  17. Zhu, M.; Zhu, C.; Wu, D.; Wang, X.; Wang, H.; Gao, J.; Huang, H.; Shi, C.; Liu, Y.; Kang, Z. Efficient photocatalytic water splitting through titanium silicalite stabilized CoO nanodots. Nanoscale 2019, 11, 15984–15990. [Google Scholar] [CrossRef]
  18. Collins, G.; Armstrong, E.; McNulty, D.; O’Hanlon, S.; Geaney, H.; O’Dwyer, C. 2D and 3D photonic crystal materials for photocatalysis and electrochemical energy storage and conversion. Sci. Technol. Adv. Mater. 2016, 17, 563–582. [Google Scholar] [CrossRef] [Green Version]
  19. Rangan, K.; Arachchige, S.M.; Brown, J.R.; Brewer, K.J. Solar energy conversion using photochemical molecular devices: Photocatalytic hydrogen production from water using mixed-metal supramolecular complexes. Energy Environ. Sci. 2009, 2, 410–419. [Google Scholar] [CrossRef]
  20. Bard, A.J. Inner-Sphere Heterogeneous Electrode Reactions. Electrocatalysis and Photocatalysis: The Challenge. J. Am. Chem. Soc. 2010, 132, 7559–7567. [Google Scholar] [CrossRef]
  21. Zhang, P.; Lou, X.W. Design of Heterostructured Hollow Photocatalysts for Solar-to-Chemical Energy Conversion. Adv. Mater. 2019, 31, 1900281. [Google Scholar] [CrossRef]
  22. Shiraishi, Y.; Toi, S.; Ichikawa, S.; Hirai, T. Photocatalytic NH3 Splitting on TiO2 Particles Decorated with Pt–Au Bimetallic Alloy Nanoparticles. ACS Appl. Nano Mater. 2020, 3, 1612–1620. [Google Scholar] [CrossRef]
  23. Misra, M.; Chowdhury, S.R.; Singh, N. TiO2@Au@CoMn2O4 core–shell nanorods for photo‒electrochemical and photocatalytic activity for decomposition of toxic organic compounds and photo reduction of Cr6+ ion. J. Alloys Compd. 2020, 824, 153861. [Google Scholar] [CrossRef]
  24. Xu, H.; Ouyang, S.; Liu, L.; Reunchan, P.; Umezawa, N.; Ye, J. Recent advances in TiO2-based photocatalysis. J. Mater. Chem. A 2014, 2, 12642–12661. [Google Scholar] [CrossRef]
  25. Rao, C.N.R.; Ramakrishna Matte, H.S.S.; Maitra, U. Graphene Analogues of Inorganic Layered Materials. Angew. Chem. Int. Ed. 2013, 52, 13162–13185. [Google Scholar] [CrossRef] [PubMed]
  26. Sun, Z.; Talreja, N.; Tao, H.; Texter, J.; Muhler, M.; Strunk, J.; Chen, J. Catalysis of Carbon Dioxide Photoreduction on Nanosheets: Fundamentals and Challenges. Angew. Chem. Int. Ed. 2018, 57, 7610–7627. [Google Scholar] [CrossRef]
  27. Sprick, R.S.; Bonillo, B.; Clowes, R.; Guiglion, P.; Brownbill, N.J.; Slater, B.J.; Blanc, F.; Zwijnenburg, M.A.; Adams, D.J.; Cooper, A.I. Visible-Light-Driven Hydrogen Evolution Using Planarized Conjugated Polymer Photocatalysts. Angew. Chem. Int. Ed. 2016, 55, 1792–1796. [Google Scholar] [CrossRef] [Green Version]
  28. Wang, S.; Hai, X.; Ding, X.; Jin, S.; Xiang, Y.; Wang, P.; Jiang, B.; Ichihara, F.; Oshikiri, M.; Meng, X.; et al. Intermolecular cascaded π-conjugation channels for electron delivery powering CO2 photoreduction. Nat. Commun. 2020, 11, 1149. [Google Scholar] [CrossRef] [Green Version]
  29. Zhao, P.; Wang, L.; Wu, Y.; Yang, T.; Ding, Y.; Yang, H.G.; Hu, A. Hyperbranched Conjugated Polymer Dots: The Enhanced Photocatalytic Activity for Visible Light-Driven Hydrogen Production. Macromolecules 2019, 52, 4376–4384. [Google Scholar] [CrossRef]
  30. Lan, Z.-A.; Ren, W.; Chen, X.; Zhang, Y.; Wang, X. Conjugated donor-acceptor polymer photocatalysts with electron-output “tentacles” for efficient hydrogen evolution. Appl. Catal. B 2019, 245, 596–603. [Google Scholar] [CrossRef]
  31. Hasani, A.; Tekalgne, M.; Le, Q.V.; Jang, H.W.; Kim, S.Y. Two-dimensional materials as catalysts for solar fuels: Hydrogen evolution reaction and CO2 reduction. J. Mater. Chem. A 2019, 7, 430–454. [Google Scholar] [CrossRef]
  32. Di, J.; Yan, C.; Handoko, A.D.; Seh, Z.W.; Li, H.; Liu, Z. Ultrathin two-dimensional materials for photo- and electrocatalytic hydrogen evolution. Mater. Today 2018, 21, 749–770. [Google Scholar] [CrossRef]
  33. Li, J.; Zhan, G.; Yu, Y.; Zhang, L. Superior visible light hydrogen evolution of Janus bilayer junctions via atomic-level charge flow steering. Nat. Commun. 2016, 7, 11480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Bao, W.; Cai, X.; Kim, D.; Sridhara, K.; Fuhrer, M.S. High mobility ambipolar MoS2 field-effect transistors: Substrate and dielectric effects. Appl. Phys. Lett. 2013, 102, 042104. [Google Scholar] [CrossRef] [Green Version]
  35. Lian, S.; Kodaimati, M.S.; Dolzhnikov, D.S.; Calzada, R.; Weiss, E.A. Powering a CO2 Reduction Catalyst with Visible Light through Multiple Sub-picosecond Electron Transfers from a Quantum Dot. J. Am. Chem. Soc. 2017, 139, 8931–8938. [Google Scholar] [CrossRef]
  36. Li, X.-B.; Gao, Y.-J.; Wang, Y.; Zhan, F.; Zhang, X.-Y.; Kong, Q.-Y.; Zhao, N.-J.; Guo, Q.; Wu, H.-L.; Li, Z.-J.; et al. Self-Assembled Framework Enhances Electronic Communication of Ultrasmall-Sized Nanoparticles for Exceptional Solar Hydrogen Evolution. J. Am. Chem. Soc. 2017, 139, 4789–4796. [Google Scholar] [CrossRef]
  37. Wu, L.-Z.; Chen, B.; Li, Z.-J.; Tung, C.-H. Enhancement of the Efficiency of Photocatalytic Reduction of Protons to Hydrogen via Molecular Assembly. Acc. Chem. Res. 2014, 47, 2177–2185. [Google Scholar] [CrossRef]
  38. Zrazhevskiy, P.; Sena, M.; Gao, X. Designing multifunctional quantum dots for bioimaging, detection, and drug delivery. Chem. Soc. Rev. 2010, 39, 4326–4354. [Google Scholar] [CrossRef] [Green Version]
  39. Smith, A.M.; Nie, S. Semiconductor Nanocrystals: Structure, Properties, and Band Gap Engineering. Acc. Chem. Res. 2010, 43, 190–200. [Google Scholar] [CrossRef] [Green Version]
  40. Zheng, L.; Su, H.; Zhang, J.; Walekar, L.S.; Vafaei Molamahmood, H.; Zhou, B.; Long, M.; Hu, Y.H. Highly selective photocatalytic production of H2O2 on sulfur and nitrogen co-doped graphene quantum dots tuned TiO2. Appl. Catal. B 2018, 239, 475–484. [Google Scholar] [CrossRef]
  41. Diao, S.; Zhang, X.; Shao, Z.; Ding, K.; Jie, J.; Zhang, X. 12.35% efficient graphene quantum dots/silicon heterojunction solar cells using graphene transparent electrode. Nano Energy 2017, 31, 359–366. [Google Scholar] [CrossRef]
  42. Di, J.; Chen, C.; Yang, S.-Z.; Chen, S.; Duan, M.; Xiong, J.; Zhu, C.; Long, R.; Hao, W.; Chi, Z.; et al. Isolated single atom cobalt in Bi3O4Br atomic layers to trigger efficient CO2 photoreduction. Nat. Commun. 2019, 10, 2840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Wang, A.; Li, J.; Zhang, T. Heterogeneous single-atom catalysis. Nat. Rev. Chem. 2018, 2, 65–81. [Google Scholar] [CrossRef]
  44. Liu, J. Catalysis by Supported Single Metal Atoms. ACS Catal. 2017, 7, 34–59. [Google Scholar] [CrossRef]
  45. Li, X.; Bi, W.; Zhang, L.; Tao, S.; Chu, W.; Zhang, Q.; Luo, Y.; Wu, C.; Xie, Y. Single-Atom Pt as Co-Catalyst for Enhanced Photocatalytic H2 Evolution. Adv. Mater. 2016, 28, 2427–2431. [Google Scholar] [CrossRef]
  46. Dhakshinamoorthy, A.; Li, Z.; Garcia, H. Catalysis and photocatalysis by metal organic frameworks. Chem. Soc. Rev. 2018, 47, 8134–8172. [Google Scholar] [CrossRef] [PubMed]
  47. Nasalevich, M.A.; Van der Veen, M.; Kapteijn, F.; Gascon, J. Metal-organic frameworks as heterogeneous photocatalysts: Advantages and challenges. CrystEngComm 2014, 16, 4919–4926. [Google Scholar] [CrossRef] [Green Version]
  48. Corma, A.; García, H.; Llabrés i Xamena, F.X. Engineering Metal Organic Frameworks for Heterogeneous Catalysis. Chem. Rev. 2010, 110, 4606–4655. [Google Scholar] [CrossRef]
  49. Baloyi, J.; Seadira, T.; Raphulu, M.; Ochieng, A. Preparation, Characterization and Growth Mechanism of Dandelion-like TiO2 Nanostructures and their Application in Photocatalysis towards Reduction of Cr(VI). Mater. Today 2015, 2, 3973–3987. [Google Scholar] [CrossRef]
  50. Goodeve, C.F.; Kitchener, J.A. The mechanism of photosensitisation by solids. J. Chem. Soc. Faraday Trans. 1938, 34, 902–908. [Google Scholar] [CrossRef]
  51. Jiang, L.; Wang, Y.; Feng, C. Application of Photocatalytic Technology in Environmental Safety. Procedia Eng. 2012, 45, 993–997. [Google Scholar] [CrossRef] [Green Version]
  52. Hashimoto, K.; Irie, H.; Fujishima, A. TiO2 Photocatalysis: A Historical Overview and Future Prospects. Jpn. J. Appl. Phys. 2005, 44, 8269–8285. [Google Scholar] [CrossRef]
  53. Dong, H.; Zeng, G.; Tang, L.; Fan, C.; Zhang, C.; He, X.; He, Y. An overview on limitations of TiO2-based particles for photocatalytic degradation of organic pollutants and the corresponding countermeasures. Water Res. 2015, 79, 128–146. [Google Scholar] [CrossRef]
  54. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  55. Higashimoto, S. Titanium-dioxide-based visible-light-sensitive photocatalysis: Mechanistic insight and applications. Catalysts 2019, 9, 201. [Google Scholar] [CrossRef] [Green Version]
  56. Etacheri, V.; Di Valentin, C.; Schneider, J.; Bahnemann, D.; Pillai, S.C. Visible-light activation of TiO2 photocatalysts: Advances in theory and experiments. J. Photochem. Photobiol. C 2015, 25, 1–29. [Google Scholar] [CrossRef] [Green Version]
  57. Torimoto, T.; Okawa, Y.; Takeda, N.; Yoneyama, H. Effect of activated carbon content in TiO2-loaded activated carbon on photodegradation behaviors of dichloromethane. J. Photochem. Photobiol. C 1997, 103, 153–157. [Google Scholar] [CrossRef]
  58. Bhattacharyya, A.; Kawi, S.; Ray, M.B. Photocatalytic degradation of orange II by TiO2 catalysts supported on adsorbents. Catal. Today 2004, 98, 431–439. [Google Scholar] [CrossRef]
  59. Szczepanik, B. Photocatalytic degradation of organic contaminants over clay–TiO2 nanocomposites: A review. Appl. Clay Sci. 2017, 141, 227–239. [Google Scholar] [CrossRef]
  60. Yang, J.; Du, J.; Li, X.; Liu, Y.; Jiang, C.; Qi, W.; Zhang, K.; Gong, C.; Li, R.; Luo, M. Highly hydrophilic TiO2 nanotubes network by alkaline hydrothermal method for photocatalysis degradation of methyl orange. Nanomaterials 2019, 9, 526. [Google Scholar] [CrossRef] [Green Version]
  61. Pan, X.; Yang, M.-Q.; Fu, X.; Zhang, N.; Xu, Y.-J. Defective TiO2 with oxygen vacancies: Synthesis, properties and photocatalytic applications. Nanoscale 2013, 5, 3601–3614. [Google Scholar] [CrossRef]
  62. Wang, Y.; Wang, Q.; Zhan, X.; Wang, F.; Safdar, M.; He, J. Visible light driven type II heterostructures and their enhanced photocatalysis properties: A review. Nanoscale 2013, 5, 8326–8339. [Google Scholar] [CrossRef] [PubMed]
  63. Lazar, M.A.; Daoud, W.A. Achieving selectivity in TiO2-based photocatalysis. RSC Adv. 2013, 3, 4130–4140. [Google Scholar] [CrossRef]
  64. Liu, L.; Liu, Z.; Liu, A.; Gu, X.; Ge, C.; Gao, F.; Dong, L. Engineering the TiO2—Graphene Interface to Enhance Photocatalytic H2 Production. ChemSusChem 2014, 7, 618–626. [Google Scholar] [CrossRef]
  65. Chen, J.I.; Von Freymann, G.; Choi, S.Y.; Kitaev, V.; Ozin, G.A. Amplified photochemistry with slow photons. Adv. Mater. 2006, 18, 1915–1919. [Google Scholar] [CrossRef]
  66. Li, H.; Bian, Z.; Zhu, J.; Zhang, D.; Li, G.; Huo, Y.; Li, H.; Lu, Y. Mesoporous Titania Spheres with Tunable Chamber Stucture and Enhanced Photocatalytic Activity. J. Am. Chem. Soc. 2007, 129, 8406–8407. [Google Scholar] [CrossRef] [PubMed]
  67. Zhu, J.; Deng, Z.; Chen, F.; Zhang, J.; Chen, H.; Anpo, M.; Huang, J.; Zhang, L. Hydrothermal doping method for preparation of Cr3+–TiO2 photocatalysts with concentration gradient distribution of Cr3+. Appl. Catal. B 2006, 62, 329–335. [Google Scholar] [CrossRef]
  68. Mathew, S.; Ganguly, P.; Rhatigan, S.; Kumaravel, V.; Byrne, C.; Hinder, S.J.; Bartlett, J.; Nolan, M.; Pillai, S.C. Cu-doped TiO2: Visible light assisted photocatalytic antimicrobial activity. Appl. Sci. 2018, 8, 2067. [Google Scholar] [CrossRef] [Green Version]
  69. He, S.; Huang, J.; Goodsell, J.L.; Angerhofer, A.; Wei, W.D. Plasmonic Nickel–TiO2 Heterostructures for Visible-Light-Driven Photochemical Reactions. Angew. Chem. Int. Ed. 2019, 58, 6038–6041. [Google Scholar] [CrossRef]
  70. Jia, T.; Fu, F.; Yu, D.; Cao, J.; Sun, G. Facile synthesis and characterization of N-doped TiO2/C nanocomposites with enhanced visible-light photocatalytic performance. Appl. Surf. Sci. 2018, 430, 438–447. [Google Scholar] [CrossRef]
  71. Zhou, F.; Song, H.; Wang, H.; Komarneni, S.; Yan, C. N-doped TiO2/sepiolite nanocomposites with enhanced visible-light catalysis: Role of N precursors. Appl. Clay Sci. 2018, 166, 9–17. [Google Scholar] [CrossRef]
  72. Romero Saez, M.; Jaramillo, L.; Saravanan, R.; Benito, N.; Pabón, E.; Mosquera, E.; Gracia Caroca, F. Notable photocatalytic activity of TiO2-polyethylene nanocomposites for visible light degradation of organic pollutants. EXPRESS Polym. Lett. 2017, 11, 899–909. [Google Scholar] [CrossRef]
  73. Saravanan, R.; Aviles, J.; Gracia, F.; Mosquera, E.; Gupta, V.K. Crystallinity and lowering band gap induced visible light photocatalytic activity of TiO2/CS (Chitosan) nanocomposites. Int. J. Biol. Macromol. 2018, 109, 1239–1245. [Google Scholar] [CrossRef] [PubMed]
  74. Raziq, F.; Sun, L.; Wang, Y.; Zhang, X.; Humayun, M.; Ali, S.; Bai, L.; Qu, Y.; Yu, H.; Jing, L. Synthesis of Large Surface-Area g–C3N4 Comodified with MnOx and Au–TiO2 as Efficient Visible-Light Photocatalysts for Fuel Production. Adv. Energy Mater. 2018, 8, 1701580. [Google Scholar] [CrossRef]
  75. Delsouz Khaki, M.R.; Shafeeyan, M.S.; Raman, A.A.A.; Daud, W.M.A.W. Evaluating the efficiency of nano-sized Cu doped TiO2/ZnO photocatalyst under visible light irradiation. J. Mol. Liq. 2018, 258, 354–365. [Google Scholar] [CrossRef]
  76. Liu, X.; Xing, Z.; Zhang, Y.; Li, Z.; Wu, X.; Tan, S.; Yu, X.; Zhu, Q.; Zhou, W. Fabrication of 3D flower-like black N–TiO2-x@MoS2 for unprecedented-high visible-light-driven photocatalytic performance. Appl. Catal. B 2017, 201, 119–127. [Google Scholar] [CrossRef]
  77. Zhang, L.; Qin, M.; Yu, W.; Zhang, Q.; Xie, H.; Sun, Z.; Shao, Q.; Guo, X.; Hao, L.; Zheng, Y. Heterostructured TiO2/WO3 nanocomposites for photocatalytic degradation of toluene under visible light. J. Electrochem. Soc. 2017, 164, H1086. [Google Scholar] [CrossRef]
  78. Yao, G.-Y.; Liu, Q.-L.; Zhao, Z.-Y. Studied Localized Surface Plasmon Resonance Effects of Au Nanoparticles on TiO2 by FDTD Simulations. Catalysts 2018, 8, 236. [Google Scholar] [CrossRef]
  79. Lin, Z.; Wang, X.; Liu, J.; Tian, Z.; Dai, L.; He, B.; Han, C.; Wu, Y.; Zeng, Z.; Hu, Z. On the role of localized surface plasmon resonance in UV-Vis light irradiated Au/TiO2 photocatalysis systems: Pros and cons. Nanoscale 2015, 7, 4114–4123. [Google Scholar] [CrossRef]
  80. Schaub, R.; Wahlström, E.; Rønnau, A.; Lægsgaard, E.; Stensgaard, I.; Besenbacher, F. Oxygen-Mediated Diffusion of Oxygen Vacancies on the TiO2110) Surface. Science 2003, 299, 377. [Google Scholar] [CrossRef]
  81. 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]
  82. Wu, T.; Niu, P.; Yang, Y.; Yin, L.-C.; Tan, J.; Zhu, H.; Irvine, J.T.S.; Wang, L.; Liu, G.; Cheng, H.-M. Homogeneous Doping of Substitutional Nitrogen/Carbon in TiO2 Plates for Visible Light Photocatalytic Water Oxidation. Sci. Rep. 2019, 29, 1901943. [Google Scholar]
  83. Cao, Y.-Q.; Zhao, X.-R.; Chen, J.; Zhang, W.; Li, M.; Zhu, L.; Zhang, X.-J.; Wu, D.; Li, A.-D. TiOxNy Modified TiO2 Powders Prepared by Plasma Enhanced Atomic Layer Deposition for Highly Visible Light Photocatalysis. Sci. Rep. 2018, 8, 12131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ismail, A.A.; Bahnemann, D.W.; Bannat, I.; Wark, M. Gold nanoparticles on mesoporous interparticle networks of titanium dioxide nanocrystals for enhanced photonic efficiencies. J. Phys. Chem. 2009, 113, 7429–7435. [Google Scholar] [CrossRef]
  85. Forsyth, M.; MacFarlane, D.R.; Best, A.; Adebahr, J.; Jacobsson, P.; Hill, A.J. The effect of nano-particle TiO2 fillers on structure and transport in polymer electrolytes. Solid State Ion. 2002, 147, 203–211. [Google Scholar] [CrossRef]
  86. 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]
  87. Zhang, K.; Wang, L.; Kim, J.K.; Ma, M.; Veerappan, G.; Lee, C.-L.; Kong, K.-J.; Lee, H.; Park, J.H. An order/disorder/water junction system for highly efficient co-catalyst-free photocatalytic hydrogen generation. Energy Environ. Sci. 2016, 9, 499–503. [Google Scholar] [CrossRef]
  88. Naldoni, A.; Allieta, M.; Santangelo, S.; Marelli, M.; Fabbri, F.; Cappelli, S.; Bianchi, C.L.; Psaro, R.; Dal Santo, V. Effect of Nature and Location of Defects on Bandgap Narrowing in Black TiO2 Nanoparticles. J. Am. Chem. Soc. 2012, 134, 7600–7603. [Google Scholar] [CrossRef]
  89. Hejazi, S.; Mohajernia, S.; Osuagwu, B.; Zoppellaro, G.; Andryskova, P.; Tomanec, O.; Kment, S.; Zbořil, R.; Schmuki, P. On the Controlled Loading of Single Platinum Atoms as a Co-Catalyst on TiO2 Anatase for Optimized Photocatalytic H2 Generation. Adv. Mater. 2020, 32, 1908505. [Google Scholar] [CrossRef] [Green Version]
  90. Yu, X.; Fan, X.; An, L.; Liu, G.; Li, Z.; Liu, J.; Hu, P. Mesocrystalline Ti3+ TiO2 hybridized g–C3N4 for efficient visible-light photocatalysis. Carbon 2018, 128, 21–30. [Google Scholar] [CrossRef]
  91. Nguyen, C.T.K.; Quang Tran, N.; Seo, S.; Hwang, H.; Oh, S.; Yu, J.; Lee, J.; Anh Le, T.; Hwang, J.; Kim, M.; et al. Highly efficient nanostructured metal-decorated hybrid semiconductors for solar conversion of CO2 with almost complete CO selectivity. Mater. Today 2020. [Google Scholar] [CrossRef]
  92. Chen, X.; Liu, L.; Yu, P.Y.; Mao, S.S. Increasing Solar Absorption for Photocatalysis with Black Hydrogenated Titanium Dioxide Nanocrystals. Science 2011, 331, 746. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, S.; Cai, J.; Mao, J.; Li, S.; Shen, J.; Gao, S.; Huang, J.; Wang, X.; Parkin, I.P.; Lai, Y. Defective black Ti3+ self-doped TiO2 and reduced graphene oxide composite nanoparticles for boosting visible-light driven photocatalytic and photoelectrochemical activity. Appl. Surf. Sci. 2019, 467, 45–55. [Google Scholar] [CrossRef]
  94. Tan, S.; Xing, Z.; Zhang, J.; Li, Z.; Wu, X.; Cui, J.; Kuang, J.; Zhu, Q.; Zhou, W. Ti3+–TiO2/g–C3N4 mesostructured nanosheets heterojunctions as efficient visible-light-driven photocatalysts. J. Catal. 2018, 357, 90–99. [Google Scholar] [CrossRef]
  95. Sasikala, R.; Shirole, A.; Sudarsan, V.; Sakuntala, T.; Sudakar, C.; Naik, R.; Bharadwaj, S.R. Highly dispersed phase of SnO2 on TiO2 nanoparticles synthesized by polyol-mediated route: Photocatalytic activity for hydrogen generation. Int. J. Hydrog. Energy 2009, 34, 3621–3630. [Google Scholar] [CrossRef]
  96. Justicia, I.; Ordejón, P.; Canto, G.; Mozos, J.L.; Fraxedas, J.; Battiston, G.A.; Gerbasi, R.; Figueras, A. Designed Self-Doped Titanium Oxide Thin Films for Efficient Visible-Light Photocatalysis. Adv. Mater. 2002, 14, 1399–1402. [Google Scholar] [CrossRef]
  97. Buso, D.; Pacifico, J.; Martucci, A.; Mulvaney, P. Gold-nanoparticle-doped TiO2 semiconductor thin films: Optical characterization. Adv. Funct. Mater. 2007, 17, 347–354. [Google Scholar] [CrossRef]
  98. Zhu, G.; Shan, Y.; Lin, T.; Zhao, W.; Xu, J.; Tian, Z.; Zhang, H.; Zheng, C.; Huang, F. Hydrogenated blue titania with high solar absorption and greatly improved photocatalysis. Nanoscale 2016, 8, 4705–4712. [Google Scholar] [CrossRef]
  99. Qiu, J.; Li, S.; Gray, E.; Liu, H.; Gu, Q.-F.; Sun, C.; Lai, C.; Zhao, H.; Zhang, S. Hydrogenation Synthesis of Blue TiO2 for High-Performance Lithium-Ion Batteries. J. Phys. Chem. 2014, 118, 8824–8830. [Google Scholar]
  100. Sun, L.; Li, Z.; Li, Z.; Hu, Y.; Chen, C.; Yang, C.; Du, B.; Sun, Y.; Besenbacher, F.; Yu, M. Design and mechanism of core-shell TiO2 nanoparticles as a high-performance photothermal agent. Nanoscale 2017, 9, 16183–16192. [Google Scholar] [CrossRef]
  101. Wang, J.; Wang, Y.; Wang, W.; Peng, T.; Liang, J.; Li, P.; Pan, D.; Fan, Q.; Wu, W. Visible light driven Ti3+ self-doped TiO2 for adsorption-photocatalysis of aqueous U(VI). Environ. Pollut. 2020, 262, 114373. [Google Scholar] [CrossRef]
  102. Cheng, D.; Li, Y.; Yang, L.; Luo, S.; Yang, L.; Luo, X.; Luo, Y.; Li, T.; Gao, J.; Dionysiou, D.D. One-step reductive synthesis of Ti3+ self–doped elongated anatase TiO2 nanowires combined with reduced graphene oxide for adsorbing and degrading waste engine oil. J. Hazard. Mater. 2019, 378, 120752. [Google Scholar] [CrossRef] [PubMed]
  103. Li, J.-J.; Weng, B.; Cai, S.-C.; Chen, J.; Jia, H.-P.; Xu, Y.-J. Efficient promotion of charge transfer and separation in hydrogenated TiO2/WO3 with rich surface-oxygen-vacancies for photodecomposition of gaseous toluene. J. Hazard. Mater. 2018, 342, 661–669. [Google Scholar] [CrossRef] [PubMed]
  104. Hanaor, D.A.H.; Sorrell, C.C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2011, 46, 855–874. [Google Scholar] [CrossRef] [Green Version]
  105. Zhu, Q.; Peng, Y.; Lin, L.; Fan, C.-M.; Gao, G.-Q.; Wang, R.-X.; Xu, A.-W. Stable blue TiO2−x nanoparticles for efficient visible light photocatalysts. J. Mater. Chem. 2014, 2, 4429–4437. [Google Scholar] [CrossRef]
  106. Fang, W.; Dappozze, F.; Guillard, C.; Zhou, Y.; Xing, M.; Mishra, S.; Daniele, S.; Zhang, J. Zn-Assisted TiO2–x Photocatalyst with Efficient Charge Separation for Enhanced Photocatalytic Activities. J. Phys. Chem. 2017, 121, 17068–17076. [Google Scholar] [CrossRef]
  107. Yin, H.; Lin, T.; Yang, C.; Wang, Z.; Zhu, G.; Xu, T.; Xie, X.; Huang, F.; Jiang, M. Gray TiO2 Nanowires Synthesized by Aluminum-Mediated Reduction and Their Excellent Photocatalytic Activity for Water Cleaning. Chem. Eur. 2013, 19, 13313–13316. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, M.; Pei, Q.; Chen, W.; Liu, L.; He, T.; Chen, P. Room temperature synthesis of reduced TiO2 and its application as a support for catalytic hydrogenation. RSC Adv. 2017, 7, 4306–4311. [Google Scholar] [CrossRef] [Green Version]
  109. Zhang, Y.; Xing, Z.; Liu, X.; Li, Z.; Wu, X.; Jiang, J.; Li, M.; Zhu, Q.; Zhou, W. Ti3+ Self-Doped Blue TiO2(B) Single-Crystalline Nanorods for Efficient Solar-Driven Photocatalytic Performance. ACS Appl. Mater. Interfaces 2016, 8, 26851–26859. [Google Scholar] [CrossRef] [PubMed]
  110. Yao, Y.; Sun, M.; Yuan, X.; Zhu, Y.; Lin, X.; Anandan, S. One-step hydrothermal synthesis of N/Ti3+ co-doping multiphasic TiO2/BiOBr heterojunctions towards enhanced sonocatalytic performance. Ultrason Sonochem 2018, 49, 69–78. [Google Scholar] [CrossRef]
  111. Hwang, H.M.; Oh, S.; Shim, J.-H.; Kim, Y.-M.; Kim, A.; Kim, D.; Kim, J.; Bak, S.; Cho, Y.; Bui, V.Q.; et al. Phase-Selective Disordered Anatase/Ordered Rutile Interface System for Visible-Light-Driven, Metal-Free CO2 Reduction. ACS Appl. Mater. Interfaces 2019, 11, 35693–35701. [Google Scholar] [CrossRef]
  112. Bak, S.; Lee, S.M.; Hwang, H.M.; Lee, H. Phase-selective modulation of TiO2 for visible light-driven CH arylation: Tuning of absorption and adsorptivity. J. Mol. Catal. B Enzym. 2019, 471, 71–76. [Google Scholar] [CrossRef]
  113. Kim, Y.; Hwang, H.M.; Wang, L.; Kim, I.; Yoon, Y.; Lee, H. Solar-light photocatalytic disinfection using crystalline/amorphous low energy bandgap reduced TiO2. Sci. Rep. 2016, 6, 25212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Kim, N.Y.; Lee, H.K.; Moon, J.T.; Joo, J.B. Synthesis of Spherical TiO2 Particles with Disordered Rutile Surface for Photocatalytic Hydrogen Production. Catalysts 2019, 9, 491. [Google Scholar] [CrossRef] [Green Version]
  115. Zhang, C.; Chen, S.; Mo, L.e.; Huang, Y.; Tian, H.; Hu, L.; Huo, Z.; Dai, S.; Kong, F.; Pan, X. Charge Recombination and Band-Edge Shift in the Dye-Sensitized Mg2+-Doped TiO2 Solar Cells. J. Phys. Chem. 2011, 115, 16418–16424. [Google Scholar]
  116. Cao, J.; Zhang, Y.; Tong, H.; Li, P.; Kako, T.; Ye, J. Selective local nitrogen doping in a TiO2 electrode for enhancing photoelectrochemical water splitting. ChemComm 2012, 48, 8649–8651. [Google Scholar] [CrossRef] [PubMed]
  117. Cao, J.; Zhang, Y.; Liu, L.; Ye, J. A p-type Cr-doped TiO2 photo-electrode for photo-reduction. ChemComm 2013, 49, 3440–3442. [Google Scholar] [CrossRef]
  118. Gul, I.; Sayed, M.; Shah, N.S.; Ali Khan, J.; Polychronopoulou, K.; Iqbal, J.; Rehman, F. Solar light responsive bismuth doped titania with Ti3+ for efficient photocatalytic degradation of flumequine: Synergistic role of peroxymonosulfate. Chem. Eng. J. 2020, 384, 123255. [Google Scholar] [CrossRef]
  119. Liu, H.; Shen, B.; Xing, M.; Zhang, J.; Tian, B. Vacuum-activated Co2+ and Ti3+ co-modified TiO2 with stable and enhanced photocatalytic activity. Res. Chem. Intermed. 2016, 42, 3459–3471. [Google Scholar] [CrossRef]
  120. Zhang, J.; Yuan, M.; Liu, X.; Wang, X.; Liu, S.; Han, B.; Liu, B.; Shi, H. Copper modified Ti3+ self-doped TiO2 photocatalyst for highly efficient photodisinfection of five agricultural pathogenic fungus. Chem. Eng. J. 2020, 387, 124171. [Google Scholar] [CrossRef]
  121. Zhou, F.; Yan, C.; Sun, Q.; Komarneni, S. TiO2/Sepiolite nanocomposites doped with rare earth ions: Preparation, characterization and visible light photocatalytic activity. Microporous Mesoporous Mater. 2019, 274, 25–32. [Google Scholar] [CrossRef]
  122. Wang, P.; Huang, B.; Dai, Y.; Whangbo, M.-H. Plasmonic photocatalysts: Harvesting visible light with noble metal nanoparticles. Phys. Chem. Chem. Phys. 2012, 14, 9813–9825. [Google Scholar] [CrossRef]
  123. Wang, Z.; Liu, Y.; Huang, B.; Dai, Y.; Lou, Z.; Wang, G.; Zhang, X.; Qin, X. Progress on extending the light absorption spectra of photocatalysts. Phys. Chem. Chem. Phys. 2014, 16, 2758–2774. [Google Scholar] [CrossRef]
  124. Linic, S.; Christopher, P.; Ingram, D.B. Plasmonic-metal nanostructures for efficient conversion of solar to chemical energy. Nat. Mater. 2011, 10, 911–921. [Google Scholar] [CrossRef] [PubMed]
  125. Hu, M.; Chen, J.; Li, Z.-Y.; Au, L.; Hartland, G.V.; Li, X.; Marquez, M.; Xia, Y. Gold nanostructures: Engineering their plasmonic properties for biomedical applications. Chem. Soc. Rev. 2006, 35, 1084–1094. [Google Scholar] [CrossRef] [PubMed]
  126. Rycenga, M.; Cobley, C.M.; Zeng, J.; Li, W.; Moran, C.H.; Zhang, Q.; Qin, D.; Xia, Y. Controlling the Synthesis and Assembly of Silver Nanostructures for Plasmonic Applications. Chem. Rev. 2011, 111, 3669–3712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Bigall, N.C.; Härtling, T.; Klose, M.; Simon, P.; Eng, L.M.; Eychmüller, A. Monodisperse Platinum Nanospheres with Adjustable Diameters from 10 to 100 nm: Synthesis and Distinct Optical Properties. Nano Lett. 2008, 8, 4588–4592. [Google Scholar] [CrossRef]
  128. Hutter, E.; Fendler, J.H. Exploitation of Localized Surface Plasmon Resonance. Adv. Mater. 2004, 16, 1685–1706. [Google Scholar] [CrossRef]
  129. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides. Science 2001, 293, 269. [Google Scholar] [CrossRef]
  130. Zhao, Y.; Zhu, L.; Yu, Y.; Gao, F.; Wang, W.; Chen, D.; Zhao, X. Facile one-pot preparation of Ti3+, N co-doping TiO2 nanotube arrays and enhanced photodegradation activities by tuning tube lengths and diameters. Catal. Today 2019. [Google Scholar] [CrossRef]
  131. Cao, Y.; Xing, Z.; Shen, Y.; Li, Z.; Wu, X.; Yan, X.; Zou, J.; Yang, S.; Zhou, W. Mesoporous black Ti3+/N–TiO2 spheres for efficient visible-light-driven photocatalytic performance. Chem. Eng. J. 2017, 325, 199–207. [Google Scholar] [CrossRef]
  132. Ji, L.; Zhou, X.; Schmuki, P. Sulfur and Ti3+ co-Doping of TiO2 Nanotubes Enhance Photocatalytic H2 Evolution without the Use of Any co-catalyst. Chem.: Asian J. 2019, 14, 2724–2730. [Google Scholar]
  133. Li, M.; Xing, Z.; Jiang, J.; Li, Z.; Yin, J.; Kuang, J.; Tan, S.; Zhu, Q.; Zhou, W. Surface plasmon resonance-enhanced visible-light-driven photocatalysis by Ag nanoparticles decorated S–TiO2−x nanorods. J. Taiwan Inst. Chem. Eng. 2018, 82, 198–204. [Google Scholar] [CrossRef]
  134. Yan, X.; Xing, Z.; Cao, Y.; Hu, M.; Li, Z.; Wu, X.; Zhu, Q.; Yang, S.; Zhou, W. In-situ C-N-S-tridoped single crystal black TiO2 nanosheets with exposed {001} facets as efficient visible-light-driven photocatalysts. Appl. Catal. B 2017, 219, 572–579. [Google Scholar] [CrossRef]
  135. Irie, H.; Watanabe, Y.; Hashimoto, K. Nitrogen-Concentration Dependence on Photocatalytic Activity of TiO2-xNx Powders. J. Phys. Chem. B 2003, 107, 5483–5486. [Google Scholar] [CrossRef]
  136. Wang, X.; Cui, Y.; Li, T.; Lei, M.; Li, J.; Wei, Z. Recent Advances in the Functional 2D Photonic and Optoelectronic Devices. Adv. Opt. Mater. 2019, 7, 1801274. [Google Scholar] [CrossRef]
  137. Wei, H.; McMaster, W.A.; Tan, J.Z.Y.; Chen, D.; Caruso, R.A. Tricomponent brookite/anatase TiO2/g–C3N4 heterojunction in mesoporous hollow microspheres for enhanced visible-light photocatalysis. J. Mater. Chem. 2018, 6, 7236–7245. [Google Scholar] [CrossRef]
  138. Li, S.-Y.; Liu, Z.-L.; Xiang, G.-X.; Ma, B.-H.; Meng, X.-D.; He, Y.-L. Influence of calcination temperature on the photocatalytic performance of the hierarchical TiO2 pinecone-like structure decorated with CdS nanoparticles. Ceram. Int. 2019, 45, 767–776. [Google Scholar] [CrossRef]
  139. Zhao, F.-M.; Pan, L.; Wang, S.; Deng, Q.; Zou, J.-J.; Wang, L.; Zhang, X. Ag3PO4/TiO2 composite for efficient photodegradation of organic pollutants under visible light. Appl. Surf. Sci. 2014, 317, 833–838. [Google Scholar] [CrossRef]
  140. Zhang, Y.-H.; Li, Y.-L.; Jiu, B.-B.; Gong, F.-L.; Chen, J.; Fang, S.; Zhang, H. Highly enhanced photocatalytic H2 evolution of Cu2O microcube by coupling with TiO2 nanoparticles. Nat. Nanotechnol. 2019, 30. [Google Scholar] [CrossRef]
  141. Chen, X.; Ye, J.; Ouyang, S.; Kako, T.; Li, Z.; Zou, Z. Enhanced Incident Photon-to-Electron Conversion Efficiency of Tungsten Trioxide Photoanodes Based on 3D-Photonic Crystal Design. ACS Nano 2011, 5, 4310–4318. [Google Scholar] [CrossRef]
  142. Yang, Y.; Qiu, M.; Li, L.; Pi, Y.; Yan, G.; Yang, L. A Direct Z-Scheme Van Der Waals Heterojunction (WO3·H2O/g–C3N4) for High Efficient Overall Water Splitting under Visible-Light. Sol. RRL 2018, 2, 1800148. [Google Scholar] [CrossRef]
  143. Pan, J.; Jiang, Z.; Feng, S.; Zhao, C.; Dong, Z.; Wang, B.; Wang, J.; Song, C.; Zheng, Y.; Li, C. The enhanced photocatalytic hydrogen production of the fusiform g–C3N4 modification CaTiO3 nano-heterojunction. Int. J. Hydrog. Energy 2018, 43, 19019–19028. [Google Scholar] [CrossRef]
  144. Zhao, T.; Xing, Z.; Xiu, Z.; Li, Z.; Chen, P.; Zhu, Q.; Zhou, W. Synergistic effect of surface plasmon resonance, Ti3+ and oxygen vacancy defects on Ag/MoS2/TiO2-x ternary heterojunctions with enhancing photothermal catalysis for low-temperature wastewater degradation. J. Hazard. Mater. 2019, 364, 117–124. [Google Scholar] [CrossRef]
  145. Ma, L.; Han, H.; Pan, L.; Tahir, M.; Wang, L.; Zhang, X.; Zou, J.-J. Fabrication of TiO2 nanosheets via Ti3+ doping and Ag3PO4 QD sensitization for highly efficient visible-light photocatalysis. RSC Adv. 2016, 6, 63984–63990. [Google Scholar] [CrossRef]
  146. Zhao, T.; Xing, Z.; Xiu, Z.; Li, Z.; Shen, L.; Cao, Y.; Hu, M.; Yang, S.; Zhou, W. CdS quantum dots/Ti3+–TiO2 nanobelts heterojunctions as efficient visible-light-driven photocatalysts. Mater. Res. Bull. 2018, 103, 114–121. [Google Scholar] [CrossRef]
  147. Sun, B.; Zhou, W.; Li, H.; Ren, L.; Qiao, P.; Li, W.; Fu, H. Synthesis of Particulate Hierarchical Tandem Heterojunctions toward Optimized Photocatalytic Hydrogen Production. Adv. Mater. 2018, 30, 1804282. [Google Scholar] [CrossRef]
  148. Xiu, Z.; Xing, Z.; Li, Z.; Wu, X.; Yan, X.; Hu, M.; Cao, Y.; Yang, S.; Zhou, W. Ti3+–TiO2/Ce3+-CeO2 Nanosheet heterojunctions as efficient visible-light-driven photocatalysts. Mater. Res. Bull. 2018, 100, 191–197. [Google Scholar] [CrossRef]
  149. Wang, Y.; Cao, X.; Hu, Q.; Liang, X.; Tian, T.; Lin, J.; Yue, M.; Ding, Y. FeOx Derived from an Iron-Containing Polyoxometalate Boosting the Photocatalytic Water Oxidation Activity of Ti3+-Doped TiO2. ACS Appl. Mater. Interfaces 2019, 11, 23135–23143. [Google Scholar] [CrossRef]
  150. Pan, J.; Dong, Z.; Jiang, Z.; Zhao, C.; Wang, B.; Zhao, W.; Wang, J.; Song, C.; Zheng, Y.; Li, C. MoS2 Quantum Dots Modified Black Ti3+–TiO2/g–C3N4 Hollow Nanosphere Heterojunction toward Photocatalytic Hydrogen Production Enhancement. Solar RRL 2019, 3, 1900337. [Google Scholar] [CrossRef]
  151. Cao, Y.; Xing, Z.; Li, Z.; Wu, X.; Hu, M.; Yan, X.; Zhu, Q.; Yang, S.; Zhou, W. Mesoporous black TiO2-x/Ag nanospheres coupled with g–C3N4 nanosheets as 3D/2D ternary heterojunctions visible light photocatalysts. J. Hazard. Mater. 2018, 343, 181–190. [Google Scholar] [CrossRef]
  152. Huang, Y.; Wang, P.; Wang, Z.; Rao, Y.; Cao, J.-j.; Pu, S.; Ho, W.; Lee, S.C. Protonated g–C3N4/Ti3+ self-doped TiO2 nanocomposite films: Room-temperature preparation, hydrophilicity, and application for photocatalytic NOx removal. Appl. Catal. B 2019, 240, 122–131. [Google Scholar] [CrossRef]
  153. Ye, M.-Y.; Zhao, Z.-H.; Hu, Z.-F.; Liu, L.-Q.; Ji, H.-M.; Shen, Z.-R.; Ma, T.-Y. 0D/2D Heterojunctions of Vanadate Quantum Dots/Graphitic Carbon Nitride Nanosheets for Enhanced Visible-Light-Driven Photocatalysis. Angew. Chem. Int. Ed. 2017, 56, 8407–8411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Huo, N.; Gupta, S.; Konstantatos, G. MoS2–HgTe Quantum Dot Hybrid Photodetectors beyond 2 µm. Adv. Mater. 2017, 29, 1606576. [Google Scholar] [CrossRef] [PubMed]
  155. Lian, Z.; Xu, P.; Wang, W.; Zhang, D.; Xiao, S.; Li, X.; Li, G. C60-Decorated CdS/TiO2 Mesoporous Architectures with Enhanced Photostability and Photocatalytic Activity for H2 Evolution. ACS Appl. Mater. Interfaces 2015, 7, 4533–4540. [Google Scholar] [CrossRef] [PubMed]
  156. Huo, Y.; Yang, X.; Zhu, J.; Li, H. Highly active and stable CdS–TiO2 visible photocatalyst prepared by in situ sulfurization under supercritical conditions. Appl. Catal. B 2011, 106, 69–75. [Google Scholar] [CrossRef]
  157. Yu, X.; Li, Z.; Liu, J.; Hu, P. TaOC chemical bond enhancing charge separation between Ta4+ doped Ta2O5 quantum dots and cotton-like g–C3N4. Appl. Catal. B 2017, 205, 271–280. [Google Scholar] [CrossRef]
  158. Zhou, P.; Yu, J.; Jaroniec, M. All-Solid-State Z-Scheme Photocatalytic Systems. Adv. Mater. 2014, 26, 4920–4935. [Google Scholar] [CrossRef]
  159. Zhang, L.J.; Li, S.; Liu, B.K.; Wang, D.J.; Xie, T.F. Highly Efficient CdS/WO3 Photocatalysts: Z-Scheme Photocatalytic Mechanism for Their Enhanced Photocatalytic H2 Evolution under Visible Light. ACS Catal. 2014, 4, 3724–3729. [Google Scholar] [CrossRef]
  160. Liu, L.; Zhou, X.; Liu, L.; Jiang, S.; Li, Y.; Guo, L.; Yan, S.; Tai, X. Heterogeneous Bimetallic Cu–Ni Nanoparticle-Supported Catalysts in the Selective Oxidation of Benzyl Alcohol to Benzaldehyde. Catalysts 2019, 9, 538. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic structures of (a) TiO2 and (b) Ti3+–TiO2 according to band positions of (c) TiO2 and (d) Ti3+–TiO2 modified with heterogeneous materials.
Figure 1. Schematic structures of (a) TiO2 and (b) Ti3+–TiO2 according to band positions of (c) TiO2 and (d) Ti3+–TiO2 modified with heterogeneous materials.
Catalysts 10 00679 g001
Figure 2. Scheme of the route for the preparation of Ru/TiO2−xf. (ai) Photographs of P25 nanocrystals and TiO2−x prepared by Na/NaCl reduction corresponding to reaction times from 1 to 4 h. Reproduced from [108]. Copyright (2017), The Royal Society of Chemistry.
Figure 2. Scheme of the route for the preparation of Ru/TiO2−xf. (ai) Photographs of P25 nanocrystals and TiO2−x prepared by Na/NaCl reduction corresponding to reaction times from 1 to 4 h. Reproduced from [108]. Copyright (2017), The Royal Society of Chemistry.
Catalysts 10 00679 g002
Figure 3. (A) Photographs of P25 (left), disordered anatase—TiO2 (middle) and disordered rutile TiO2 (right) suspensions (0.05 g L−1) after Li-Ethylene diamine (EDA) treatment for 6 days; (B) x-ray diffraction (XRD) patterns of the Li-EDA-treated P-25 crystals with different treatment times for the (A) anatase phase and the (R) rutile phase. (C) HR-TEM images and selected area electron diffraction pattern of P-25; scale bar: 10 nm; (D) HR-TEM images and the selected-area electron diffraction pattern of blue P-25; scale bar: 10 nm. Enlarged TEM images of the junction area (red squares: P-25 and green squares: blue P-25); (E) calculated bandgap diagrams of P25 (left) and blue P-25 (right). Reproduced from [1]. Copyright (2016), The Royal Society of Chemistry.
Figure 3. (A) Photographs of P25 (left), disordered anatase—TiO2 (middle) and disordered rutile TiO2 (right) suspensions (0.05 g L−1) after Li-Ethylene diamine (EDA) treatment for 6 days; (B) x-ray diffraction (XRD) patterns of the Li-EDA-treated P-25 crystals with different treatment times for the (A) anatase phase and the (R) rutile phase. (C) HR-TEM images and selected area electron diffraction pattern of P-25; scale bar: 10 nm; (D) HR-TEM images and the selected-area electron diffraction pattern of blue P-25; scale bar: 10 nm. Enlarged TEM images of the junction area (red squares: P-25 and green squares: blue P-25); (E) calculated bandgap diagrams of P25 (left) and blue P-25 (right). Reproduced from [1]. Copyright (2016), The Royal Society of Chemistry.
Catalysts 10 00679 g003
Figure 4. Fabrication and morphology of plasmonic Ni–TiO2 substrates. (A) Spin-coating silica nanospheres on glass or a Si wafer; (B) deposition of TiO2 on (A); (C) deposition of Ni on (B); (DF) False-colored SEM images of the sample surfaces described in (AC); (G) monitoring PMCT in Ni–TiO2 heterostructures by X-band EPR spectroscopy; (H) illustration of PMCT in Ni–TiO2. Hot electrons (red) are transferred from Ni to TiO2 upon the LSPR excitation of Ni and spatially separated from hot holes (blue). Reproduced from [69]. Copyright (2019), Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 4. Fabrication and morphology of plasmonic Ni–TiO2 substrates. (A) Spin-coating silica nanospheres on glass or a Si wafer; (B) deposition of TiO2 on (A); (C) deposition of Ni on (B); (DF) False-colored SEM images of the sample surfaces described in (AC); (G) monitoring PMCT in Ni–TiO2 heterostructures by X-band EPR spectroscopy; (H) illustration of PMCT in Ni–TiO2. Hot electrons (red) are transferred from Ni to TiO2 upon the LSPR excitation of Ni and spatially separated from hot holes (blue). Reproduced from [69]. Copyright (2019), Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 10 00679 g004
Figure 5. (a) Schematic of the sputtered layer, (b) top-surface SEM image and (inset) optical image of the TiO2 layer deposited on SiO2–Si TEM grid; (c) HAADF-TEM image and (d) EDS map of Pt–SA-decorated TiO2 layer; (e) surface density of atomic Pt for different samples; (f) schematic of Pt–SA-decorated TiO2 for H2 evolution application. Reproduced from [89]. Copyright (2020). Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2020.
Figure 5. (a) Schematic of the sputtered layer, (b) top-surface SEM image and (inset) optical image of the TiO2 layer deposited on SiO2–Si TEM grid; (c) HAADF-TEM image and (d) EDS map of Pt–SA-decorated TiO2 layer; (e) surface density of atomic Pt for different samples; (f) schematic of Pt–SA-decorated TiO2 for H2 evolution application. Reproduced from [89]. Copyright (2020). Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2020.
Catalysts 10 00679 g005
Figure 6. Fluorescent images showing cell viability of the F. graminearum macroconidia after illumination for 100 min: without photocatalyst (a and b) or treated with the 10% Cu/Ti3+–TiO2 composite (c and d); (b,d) are the enlarged observations of macroconidia in (a,c). The cell was stained by CFDA-AM/propidium iodide and the live cells appear green, while the dead ones are red in the images. (e) is ESR spectrum of the pure TiO2 and Cu/Ti3+–TiO2. Reproduced from [120]. Copyright (2020), Elsevier.
Figure 6. Fluorescent images showing cell viability of the F. graminearum macroconidia after illumination for 100 min: without photocatalyst (a and b) or treated with the 10% Cu/Ti3+–TiO2 composite (c and d); (b,d) are the enlarged observations of macroconidia in (a,c). The cell was stained by CFDA-AM/propidium iodide and the live cells appear green, while the dead ones are red in the images. (e) is ESR spectrum of the pure TiO2 and Cu/Ti3+–TiO2. Reproduced from [120]. Copyright (2020), Elsevier.
Catalysts 10 00679 g006
Figure 7. (a) Proposed visible-light-induced photocatalytic mechanism over the RE–TiO2/Sep nanocomposites and (b) photocatalytic degradation rate of orange G under visible light irradiation for 10 h over various types of photocatalysts. Reproduced from [121]. Copyright (2018), Elsevier.
Figure 7. (a) Proposed visible-light-induced photocatalytic mechanism over the RE–TiO2/Sep nanocomposites and (b) photocatalytic degradation rate of orange G under visible light irradiation for 10 h over various types of photocatalysts. Reproduced from [121]. Copyright (2018), Elsevier.
Catalysts 10 00679 g007
Figure 8. Schematic diagram of the charge transfer and the proposed reaction mechanism over N-doped TiO2/C nanocomposites under visible light irradiation. Reproduced from [70]. Copyright (2017), Elsevier.
Figure 8. Schematic diagram of the charge transfer and the proposed reaction mechanism over N-doped TiO2/C nanocomposites under visible light irradiation. Reproduced from [70]. Copyright (2017), Elsevier.
Catalysts 10 00679 g008
Figure 9. Schematic diagram for the formation of Ag/S–TiO2−x nanorods. Reproduced from [133]. Copyright (2017), Elsevier.
Figure 9. Schematic diagram for the formation of Ag/S–TiO2−x nanorods. Reproduced from [133]. Copyright (2017), Elsevier.
Catalysts 10 00679 g009
Figure 10. The synthesis process of particulate b–TiO2/MoS2/CdS tandem heterojunctions and schematics of the b–TiO2/MoS2/CdS tandem heterojunctions used for solar-driven water splitting. Reproduced from [147]. Copyright (2018), published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 10. The synthesis process of particulate b–TiO2/MoS2/CdS tandem heterojunctions and schematics of the b–TiO2/MoS2/CdS tandem heterojunctions used for solar-driven water splitting. Reproduced from [147]. Copyright (2018), published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 10 00679 g010
Figure 11. Process diagram of the preparation and photocatalytic hydrogen evolution reaction (HER) mechanism of the MoS2 QDs modified black Ti3+–TiO2/g–C3N4 hollow nanosphere heterojunction. Reproduced from [150]. Copyright (2019), Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 11. Process diagram of the preparation and photocatalytic hydrogen evolution reaction (HER) mechanism of the MoS2 QDs modified black Ti3+–TiO2/g–C3N4 hollow nanosphere heterojunction. Reproduced from [150]. Copyright (2019), Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 10 00679 g011
Figure 12. Schematic of the Ag3PO4 QDs–TiO2 nano spheres (NS) composite and its energy diagram and photocatalytic mechanism under visible light irradiation. Reproduced from [145]. Copyright (2016), The Royal Society of Chemistry.
Figure 12. Schematic of the Ag3PO4 QDs–TiO2 nano spheres (NS) composite and its energy diagram and photocatalytic mechanism under visible light irradiation. Reproduced from [145]. Copyright (2016), The Royal Society of Chemistry.
Catalysts 10 00679 g012
Figure 13. Possible visible light photocatalytic mechanism of Ti3+-doped meso–TiO2/g–C3N4 composites for hydrogen production. Reproduced from [90]. Copyright (2017), Elsevier.
Figure 13. Possible visible light photocatalytic mechanism of Ti3+-doped meso–TiO2/g–C3N4 composites for hydrogen production. Reproduced from [90]. Copyright (2017), Elsevier.
Catalysts 10 00679 g013
Figure 14. Energy diagrams of Ti3+–TiO2/Meso-g–C3N4 nanosheet heterojunctions. Reproduced from [94]. Copyright (2017), Elsevier.
Figure 14. Energy diagrams of Ti3+–TiO2/Meso-g–C3N4 nanosheet heterojunctions. Reproduced from [94]. Copyright (2017), Elsevier.
Catalysts 10 00679 g014
Figure 15. Band alignments of P25 and blue TiO2 NPs and the proposed electron transfer mechanism of the 7BT/W1-A1 HNPs photocatalyst. The photoelectrons generated and excited electrons from the CB of WO3 to migrate to trap holes at the VB of blue TiO2 sites through a Z-scheme system. After separation, the excited electrons rapidly jump from the CB of blue TiO2 onto Ag NPs acting as active sites for conversion of CO2 into CO, while all of the photogenerated holes directly oxidize H2O to form O2 from the VB of the WO3 site. Reproduced from [91]. Copyright (2019) Elsevier.
Figure 15. Band alignments of P25 and blue TiO2 NPs and the proposed electron transfer mechanism of the 7BT/W1-A1 HNPs photocatalyst. The photoelectrons generated and excited electrons from the CB of WO3 to migrate to trap holes at the VB of blue TiO2 sites through a Z-scheme system. After separation, the excited electrons rapidly jump from the CB of blue TiO2 onto Ag NPs acting as active sites for conversion of CO2 into CO, while all of the photogenerated holes directly oxidize H2O to form O2 from the VB of the WO3 site. Reproduced from [91]. Copyright (2019) Elsevier.
Catalysts 10 00679 g015
Figure 16. Visible-light-driven photocatalytic mechanism for Ti3+–TiO2/Ce3+-CeO2 nanosheet heterojunctions. Reproduced from [148]. Copyright (2017), Elsevier.
Figure 16. Visible-light-driven photocatalytic mechanism for Ti3+–TiO2/Ce3+-CeO2 nanosheet heterojunctions. Reproduced from [148]. Copyright (2017), Elsevier.
Catalysts 10 00679 g016
Figure 17. The color change of anatase TiO2 powder from white to blue by introducing hydrogen-free oxygen vacancies and from white to red by introducing atomic hydrogen-mediated oxygen vacancies. The top panel shows digital images of the samples and the bottom panel shows the atomic structures of anatase TiO2 and oxygen-deficient TiO2 with and without atomic hydrogen. Blue and red spheres indicate titanium and oxygen atoms, respectively and green spheres represent hydrogen. The calculated band structures of TiO2 with (a) OV-2H, (b) 4 (OV-2H) and (c) a neutral OV. Reproduced from [81]. Copyright (2018), Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 17. The color change of anatase TiO2 powder from white to blue by introducing hydrogen-free oxygen vacancies and from white to red by introducing atomic hydrogen-mediated oxygen vacancies. The top panel shows digital images of the samples and the bottom panel shows the atomic structures of anatase TiO2 and oxygen-deficient TiO2 with and without atomic hydrogen. Blue and red spheres indicate titanium and oxygen atoms, respectively and green spheres represent hydrogen. The calculated band structures of TiO2 with (a) OV-2H, (b) 4 (OV-2H) and (c) a neutral OV. Reproduced from [81]. Copyright (2018), Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 10 00679 g017
Table 1. Summary of metal-doped Ti3+self-doped TiO2 photocatalysts.
Table 1. Summary of metal-doped Ti3+self-doped TiO2 photocatalysts.
CatalystsLightApplication TargetEfficiencyRef.
Plasmonic Ni/Ti3+ TiO2/SiO2 nanospheres300 W Xe lampMethylene blue0.11 ± 0.04 μmol·L−1·min−1[69]
Pt single atom-anatase(001) Ti3+ TiO2 nano sheets325 nm, 365 nmHydrogen evolution reaction400 µmol·h−1·g−1[89]
Zn-assisted Ti3+ TiO2500 W Tungsten halogen lampFormic acid100% 90 min−1[106]
Li-EDA TiO2–Pt150 W Xe lampHydrogen evolution reaction350 µmol·h−1·g−1[114]
Bi3+/Ti3+ TiO2Xenon lampFlumequine under HSO5−65%[118]
Co2+/Ti3+ TiO2300 W Xe lampAcid Orange 7100% 5 h−1[119]
Cu/Ti3+ TiO2300 W Xe lampF. graminearum and B. dothidea spores100%[120]
RE-doped/Ti3+ TiO2/Sep300 W Xe lampOrange G72%[121]
Table 2. Summary of nonmetal doped Ti3+self-doped TiO2 photocatalysts.
Table 2. Summary of nonmetal doped Ti3+self-doped TiO2 photocatalysts.
CatalystsLightApplication TargetEfficiencyRef.
N/Ti3+/C TiO2500 W Xe lampRhodamine B100% 1 h−1[70]
N/Ti3+ TiO2 nanotube400 W Halogen lampRhodamine B100% 4 h−1[130]
N/Ti3+ TiO2 spheres300 W Xe lampMethylene orange100% 3 h−1[131]
S/Ti3+ TiO2150 W Xenon lampHydrogen evolution reaction22.5 µL·h−1·cm2[132]
S/Ti3+ TiO2–Ag nanorodsXe lampHydrogen evolution reaction209.2 μmol·h−1·g−1[133]
C–N–S-tridoped Ti3+ TiO2300 W Xe lampHydrogen evolution reaction149.7 μmol·h−1·g−1[134]
Table 3. Summary of semiconducting coupling with Ti3+self-doped TiO2 photocatalysts.
Table 3. Summary of semiconducting coupling with Ti3+self-doped TiO2 photocatalysts.
CatalysisLightApplication TargetEfficiencyRef.
Ti3+ TiO2/gC3N4Xe lampHydrogen evolution reaction3748.46 μmol·h−1·g−1[90]
Ti3+ TiO2/WO3–AgSolarCO2 to CO1166.72 μmol·h−1·g−1[91]
Ti3+ TiO2/rGO500 W Xe lampMethylene blue100% 2 h−1[93]
Ti3+ TiO2/meso-gC3N4300 W Xenon lampPhenol100% 90 min−1[94]
Ti3+ TiO2/WO3420 nm cutoffToluene100% 60 min−1[103]
Ti3+ TiO2 nanowires/rGO300 W Xe lampWaste oil100% 5 h−1[102]
N/Ti3+ TiO2/BiOBr120 mW lampMethylene blue100% 50 min−1[110]
Ti3+ TiO2/MoS2/Ag500 W Xe lampBisphenol A100% 2 h−1[144]
Ti3+ TiO2/Ag3PO4 QD300 W Xe lampMethylene orange100% 100 min−1[145]
Ti3+ TiO2/CdS QD300 W Xe lampMethylene blue100% 150 min −1[146]
Ti3+ hollow TiO2/MoS2/CdS300 W Xe lampHydrogen evolution reaction8950 μmol·h−1·g−1[147]
N/Ti3+ TiO2/MoS2300 W Xe lampMethylene orange100% 2 h−1[76]
Ti3+ TiO2/Ce3+ CeO2300 W Xe lampMethylene blue, Methylene orange100% 3 h −1[148]
Ti3+ TiO2/FeOx420 nm LEDOxygen evolution reaction410 μmol·h−1·g−1[149]
Ti3+ TiO2/g–C3N4 hollow nanosphere/MoS2 QD300 W Xe lampHydrogen evolution reaction1524.37 μmol·h−1·g−1[150]
Ti3+ TiO2/g–C3N4 nanospheres/Ag300 W Xe lampMethylene orange100% 3 h−1[151]
Ti3+ TiO2/g–C3N4Xe lampNOx75%[152]
Table 4. Summary of stoichiometry modification of Ti3+self-doped TiO2 photocatalysts.
Table 4. Summary of stoichiometry modification of Ti3+self-doped TiO2 photocatalysts.
CatalystColorApplication TargetEfficiencyRef.
Ti3+2H TiO2Red[81]
Ti3+CN TiO2Oxygen evolution reaction4.125 μmol·h−1·g−1[82]
Ti3+N TiO2Yellow-greenMethylene orange100% 2 h−1[83]

Share and Cite

MDPI and ACS Style

Na, S.; Seo, S.; Lee, H. Recent Developments of Advanced Ti3+-Self-Doped TiO2 for Efficient Visible-Light-Driven Photocatalysis. Catalysts 2020, 10, 679. https://doi.org/10.3390/catal10060679

AMA Style

Na S, Seo S, Lee H. Recent Developments of Advanced Ti3+-Self-Doped TiO2 for Efficient Visible-Light-Driven Photocatalysis. Catalysts. 2020; 10(6):679. https://doi.org/10.3390/catal10060679

Chicago/Turabian Style

Na, Siyoung, Sohyeon Seo, and Hyoyoung Lee. 2020. "Recent Developments of Advanced Ti3+-Self-Doped TiO2 for Efficient Visible-Light-Driven Photocatalysis" Catalysts 10, no. 6: 679. https://doi.org/10.3390/catal10060679

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