Next Article in Journal
Polynuclear Cobalt Complexes as Catalysts for Light-Driven Water Oxidation: A Review of Recent Advances
Previous Article in Journal
BF3·Et2O-Promoted Decomposition of Cyclic α-Diazo-β-Hydroxy Ketones: Novel Insights into Mechanistic Aspects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Switchable Intrinsic Defect Chemistry of Titania for Catalytic Applications

by
Swaminathan Jayashree
1,2 and
Meiyazhagan Ashokkumar
1,*
1
Department of Materials Science & NanoEngineering, Rice University, Houston, TX 77005, USA
2
CSIR-Central Electrochemical Research Institute, Karaikudi 630003, India
*
Author to whom correspondence should be addressed.
Catalysts 2018, 8(12), 601; https://doi.org/10.3390/catal8120601
Submission received: 23 October 2018 / Revised: 16 November 2018 / Accepted: 18 November 2018 / Published: 2 December 2018

Abstract

:
The energy crisis is one of the most serious issue that we confront today. Among different strategies to gain access to reliable fuel, the production of hydrogen fuel through the water-splitting reaction has emerged as the most viable alternative. Specifically, the studies on defect-rich TiO2 materials have been proved that it can perform as an efficient catalyst for electrocatalytic and photocatalytic water-splitting reactions. In this invited review, we have included a general and critical discussion on the background of titanium sub-oxides structure, defect chemistries and the consequent disorder arising in defect-rich Titania and their applications towards water-splitting reactions. We have particularly emphasized the origin of the catalytic activity in Titania-based material and its effects on the structural, optical and electronic behavior. This review article also summarizes studies on challenging issues on defect-rich Titania and new possible directions for the development of an efficient catalyst with improved catalytic performance.

1. Introduction

Some of the world’s technologies are currently due to the manipulation of defects. For example, defects can transform a hard white solid α-Al2O3 (corundum) to colored gems like ruby and sapphire [1]. Thus, the intentional creation of defects in materials with control over their concentration, type, spatial distribution, and location can drastically modify their functionalities. In particular, defect engineering serves as a platform to tailor the intrinsic electrons of materials and, hence, to tune their properties, reactivity, and catalytic activity [2,3,4,5,6,7,8,9,10,11,12,13]. The past centuries have witnessed significant progress towards commercialization of water-splitting technologies predominantly due to the intentional creation of defects in various catalytically in-active materials [14,15,16,17,18,19,20,21,22], but still, we are distant from the basic understanding of structural defects and the critical parameters which play an essential role in water-splitting applications. The grand challenge lies in the development of an improved catalyst for water-splitting applications. A stream of research focuses only on the effect of intrinsic defects on the solid-state properties of the material [23,24,25,26,27,28,29,30,31,32,33,34,35]. Few other groups concentrate on improving the electronic, optical properties of materials and discussing issues related to water-splitting reaction [36,37]. However, the progress of water electrolyzer technology requires combining these two research areas in order to address the challenges in developing an enhanced catalyst for water-splitting. Hence, it is a desperate need to fully understand the role of defect disorder on water-splitting in order to enhance the performance of electro/photocatalyst for water-splitting reaction.
To understand the scientific origin of catalytic activity on the disordered structures herein, we focus our interest on: (i) defects and defect chemistry; (ii) standard experimental approaches to generate point defects; (iii) the effect of having defects on their structural, electrical, optical behavior and, hence, catalytic activity; (iv) role of defects associated with the electrocatalytic production of hydrogen and photo-oxidation of water; and (v) challenges and applications of defect-rich Titania. Along with this, we have included perspectives and outlook on the future of the defect engineering. Several reports and reviews exist summarizing various statements on defect-engineered catalyst for water-splitting reaction [38,39,40,41,42]. Hence, in this review, we attempted to articulate the impact of defects on the structural, optical and electronic properties, which are critical factors for defining the catalytic activity of a material. We believe that a fundamental understanding of the defect engineering and combination of optimization strategies would be the bottleneck for finding the appropriate candidates for water-splitting reaction. Notably, we collectively discussed intrinsic defect engineering (self-doping) of futuristic catalyst ‘Titania’ for water-splitting reaction. The striking characteristics of Titania such as strong oxidizing power, earth abundance, non-toxic, chemical, physical and thermal stability, easy availability and excellent photoactivity supports this material as the best option for water-splitting reaction. It is also a ninth earth abundant material, obtained from the mineral ilmenite [43,44], and its phase diagram (Figure 1) shows various stable phases at different Ti/O stoichiometry.

2. Defects in Titania

In general, four common native-point defects (Figure 2) exist in Titania: (i) Oxygen vacancies (Vo)—Missing of oxygen ion from the lattice site; (ii) Titanium vacancies (VTi)—Ti ion is missing from its lattice site; (iii) Titanium interstitials (Tii)—Ti ion is located in an interstitial site; and (iv) Oxygen interstitials (Oi)—Oxygen ion is located in an interstitial site.
General characteristics of point defects in Titania/Defect chemistry of Titania:
(i)
Point defects preferably gather as clusters rather than distributed randomly on the crystal structure and, hence, lead to line defects, planar and volume defects [45].
(ii)
Depending on the location of defects in the crystal structure, the activity varies as tabulated in Table 1.
(iii)
Point defects are generally represented using Kroger-Vink notation (Table 2).
(iv)
As charge-neutrality demands, point defects usually form as a charge-neutral complexes [Schottky defects (e.g., V Ti + 2 V O , 2 Ti Ti + V O ) and Frenkel defects (e.g., O Ti + V O , V Ti +   Ti i )]. Their formation defect equilibria are given below [48,49]:
(1)
Defect equilibria of oxygen vacancy (VO) formation: Oo Catalysts 08 00601 i001 Vo●● + 2e′ + ½ O2
(2)
Defect equilibria of Titanium vacancies (VTi): O2 Catalysts 08 00601 i001 2Oo + V Ti + 4h
(3)
Defect equilibria of Titanium interstitials (Tii): 2Oo + T i T i Catalysts 08 00601 i001 Tii●●● + 3e′ + O2
(4)
Defect equilibria of oxygen interstitials (Oi): V i x + ½ O2 Catalysts 08 00601 i001 Oi′′ + 2h
(V)
Intuitively, ionization of these ionic defects leads to the formation of electronic defects in the crystal system, which are called color centers [50,51]. There are two types of color center-based on their charge (i) electron center (F) and (ii) electron-hole center (V) and classified further as F, F+, F++ and V, V+, V++, respectively.

2.1. Electron (F, F+, F++) Center

This center occurs due to the absence of a negatively charged ions (i.e., oxygen vacancy) in Titania. As a result of charge imbalance, it possess positive charge and, hence, attracts, traps electron and comprises F centers [52,53].

2.1.1. F Center

The vacancy of oxygen (Vo) is compensated by a pair of unpaired electrons from Ti 3d orbitals, resulting in a (double) positive charged vacancy (F center). It has four Ti4+ ions in the neighborhood with no unpaired electrons.

2.1.2. F+ Center

One of the electrons from F center tends to occupy the neighborhood Ti4+ ion, so it is positively charged with respect to the lattice, yields T i l a t t i c e 3 + center (F+ center). It possesses one unpaired electron.

2.1.3. F++ Center

Both electrons from F center tend to occupy the neighboring Ti4+ ion, so it is positively charged with respect to the lattice and yield T i l a t t i c e 2 + center (F++ center).
The formation mechanism and defect equilibria are shown below (Figure 2):

2.2. Electron Holes

It is an antimorph of F center [54], consisting of Titanium vacancy and holes, which gives p-type conductivity to Titania as shown in Figure 3.
These point defects possess distinct energy and, therefore, have different formation energies [55]. Recently, formation energy of various defects in Titania has been calculated as a function of fermi energy since fermi level of a material is based on their charged defects, electron and hole carriers (Figure 4a,b) [56]. According to them, the formation of the surpassing defect type in Titania depends upon their non-stoichiometry level. At Ti-rich conditions, titanium interstitials and oxygen vacancies are formed, and they behave as weak donors. While the formation of titanium vacancy is favored at O-rich growth conditions, they act as deep acceptors. M. K. Nowotny et al. attempted methods to derive the defect disorder diagram of Titania as a function of oxygen partial pressure (Figure 4c) including illustrating the concentration of specific defects at different oxygen partial pressure [57]. Under low oxygen pressure/activity, a considerable reduction of Titania leads to the formation of titanium interstitials (Ti excess: T i 1 + x O2) and electrons. While at prolonged oxidizing conditions, titanium vacancies with electron holes (acceptors) may become as most dominant defects. Hence, the predominant defect in Titania over a wide range of stoichiometry (under strongly reduced, reduced and oxidized conditions) is oxygen vacancies (donors) and it is compensated by ionic charge (electrons). It is also supported by Kofstad, 1972 (Figure 5d). Thus, oxygen vacancies and titanium interstitials mostly dominate the defect chemistry of Titania due to their lower formation enthalpies and it leads to n-type conductivity in Titania. Formally, titanium vacancies are thermodynamically reversible defects (theoretically); hence, they are usually obtained by quenching to make it as an irreversible defect [58,59,60,61,62,63].
Based on the above discussion, it is understood that an appropriate growth condition is required to incorporate the required defects in Titania. It is also a challenge to integrate particular defects, since various complexities involve in the formation of defects.

3. Experimental Approaches to Generate Defects in Titania

The defects in Titania are induced either by an in-situ control (defects creation during synthesis or ex-situ controlled (the incorporation of defects after preparation) mechanism. The reported methods to produce Ti interstitial in Titania are prolonged oxidation (1000 °C for 24 h) [60,64,65,66,67]. The state-of-the-art method to prepare reduce Titania (oxygen deficient) include energetic ion or electron beam implantation, UV irradiation, heating TiO2 under vacuum, thermal annealing to high temperatures (above 500 K), reducing conditions (C, H2), plasma-treating, laser, and high-energy particle (neutron, Ar+, electron, or γ-ray) bombardment, chemical vapor deposition, vacuum activation, metal reduction, electrochemical reduction, partial oxidation starting from Ti, Ti(II) and Ti(III) precursors, etc. [68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83].
In general, the sequence of reduction of TiO2 is:
TiO2→TinO2n−1 (n > 10)→TinO2n−1(4 < n < 10)→Ti3O5→Ti2O3→TiO→Ti2O
Hence, defect formations and their concentration, as well as the properties depend on the preparative conditions. The design of synthesis with defects also has a significant effect on the intrinsic properties such as structural, electrical, optical behavior and, in turn, their catalytic activity.

3.1. Effect of Defects on Crystal Structure of Titania

Point defects in Titania will induce lattice relaxation and ionic polarization effect, and, hence, affect the surface geometries (i.e., atomic arrangement and coordination) [84,85,86]. Upon removal of the O atom, the nearest Ti–Ti bonds tend to relax away from the vacancy in order to reinforce their bonding with rest of the lattice. Simultaneously, the neighboring Ti–O bonds displace slightly inwards due to the electrostatic attraction induced by oxygen vacancies. This internal modulation induced by oxygen vacancy defects result in the rearrangement of atomic positions and decreases the overlap between Ti dangling bonds closer to the vacancy, and consequently leads to a reduction in Ti–O bond length [87,88,89]. This defect creates uneven internal stress/strain and leads to distortion in TiO6 octahedra, which diverges with the concentration of defects and constitutes the material with different octahedral packing and crystal symmetry [90] as depicted by E. Stoyanov et al. (Figure 5).
As the concentration of oxygen vacancy defects increases, ordering of vacancy occurs, and it leads to the formation of new phases of titanium oxide, (i.e., Structural condensation happens to avoid having entire layers of vacant sites and form extended planar defects [crystallographic shear (CS) planes] with random or regular spacing) [91,92]. The number of CS planes increases thickness of the blocks as well between adjacent CS planes decreases with an increase in vacancy [90].

3.2. Crystal Structures of Various Titanium Oxides

Stoichiometric Titania (TiO2) exists as a well-known phase such as rutile (tetragonal-P42/mnm), anatase (tetragonal-I41/amd) and brookite (orthorhombic-Pbca). Whereas the structure of non-stoichiometric Titanium oxide varies with O/Ti stoichiometry as summarized in Figure 6 and Table 3.
Crystal structures of some well-known Ti oxides are discussed below:
Titanium dioxide (TiO2): The crystal structure of TiO2 is considered as a two-dimensional chain of edge-sharing TiO6 octahedra with individual chains linked at a corner oxygen atom to form the 3D lattice. Each octahedron is made up of titanium atom at the center and oxygen at each corner.
Magneli phase − TinO2n−1 (n = 4–10) (e.g., Ti4O7, Ti5O9, Ti6O11…)
The crystal structure of Magneli phases is made of TiO2 octahedra blocks with an oxygen vacancy at every nth layer, which leads to shear planes in the structure as illustrated in Figure 7. The TiO2 octahedra blocks are coupled via edges and corners while shear planes have face-sharing octahedral. This atomic arrangement leads to titanium-atom positions in one block which matches with the interstitial or unoccupied positions in the next block. This organization ultimately decreases the symmetry of the crystal system from tetragonal to triclinic as the size of the unit cell increases.
The structure of magneli phase can be regarded as made of stacked alternating shear slabs and the family members of magneli phase differ in many shear planes. Thus, they also diverge in length of the pseudo-rutile chain segments.
Magneli phases with higher n value possess greater shear plane interval. The shear planes provides a pathway for electrons transport, and hence, magneli oxides are an excellent electrical conductor [93].
Magneli phase is a mixed valence compound of (n−2) Ti4+ (3d0) and two Ti3+ (3d1) configurations. The presence of both Ti3+ and Ti4+ ions offer several possible configurations of Ti in the crystal, leading to various charge-ordered states and viewed as electron-doped (oxygen deficient) TiO2 [94].
The value of n assures the crystal structure as well the electronic structure of the Magneli phase.
Ti3O5: The unit cell includes two Ti3+ ions and one Ti4+ ion with a different significant Ti–O distance leading to a distorted octahedron [95,96].
Ti2O3: The stoichiometry of Ti2O3 extends from TiO1.49 to TiO1.51 and comprises pairs of Ti3+ ions in octahedral sites as a fundamental component in the unit cell [97,98].
TiO:
TiO has a defective NaCl structure (space group Fm 3 ¯ m) and possesses 10–15 at % vacancies at both Ti and O sites [99]. Owing to higher structural vacancies in titanium monoxide, they are considered as strongly non-stoichiometric and has a broader homogeneity region from TiO0.70 to TiO1.25 [100,101].
TiO has two polymorphs based on the vacancy distribution. If the vacancies are randomly distributed it leads to disordered cubic phase, and the other polymorph has ordered vacancies with monoclinic symmetry [83,102,103,104,105,106].
It is metallic in nature due to the high content of vacancy channels [107].
Ti2O:
Ti2O has anti-CdI2 structure and consists of zigzag oxygen rows with alternate O layers vacant. It possesses more antiphase boundaries and stacking faults [108,109,110,111].
Based on the above discussion, it infers that striking changes in the crystal structure of titanium oxides have occurred due to the relatively small structural alterations through defects.

4. Effect of Defects on Optical Behavior of Titania

Stoichiometric TiO2 is a wide bandgap semiconductor (3.3 eV), and, hence, it absorbs UV light, while reduced Titania absorbs visible light and behaves as a semiconductor or metal depending on their oxygen vacancy (Figure 8) [74,112,113,114,115,116,117,118,119].
According to crystal field splitting (Figure 9), Ti (3d, 4s, 4p) orbitals are linearly coalesced with O (2s, 2p) orbitals to form molecular orbitals as depicted in Figure 9 [90]. The upper t2g* and, eg* are primarily of Ti 3d character and configure conduction band, while the lower t2g and, eg are mainly of O 2p character and form valence band, respectively [90].
The defects and non-stoichiometry in reduced Titania will induce alterations in Ti 3d occupancies and directly control the local structure of t2g and, eg orbitals due to Jahn-Teller effect. The following difference in the splitting of t2g and eg levels in different distorted environments of Titania are illustrated in Figure 10 [90].
As an outcome, the incorporated modifications in the crystal field as a function of increasing oxygen vacancy content are:
  • the systematic decrease in crystal field splitting between t2g and, eg levels.
  • the covalent bonding between Ti and O becomes ionic.
  • The degree of electron occupancy in the t2g level increases. TiO2 has none, Ti2O3 has one and TiO has two t2g electrons.
Besides the above-described changes in crystal symmetry and coordination number, the extent of mixing of Ti 3d and O 2p orbitals is also affected, and, hence, their electronic band structure and band gap vary on titanium oxide polymorphs. It is also known that the oxygen vacancy and titanium interstitial behave as donor states since their energy level is closer to the conduction band; while titanium vacancy forces the Fermi level towards the valence band, to accept electrons from their surrounding O atoms and act as an acceptor state [49,120,121,122,123,124,125]. The distinct energy levels of various defect states in titanium oxide are shown in Figure 11 [86].
With these continuous formations of defect states, localized midgap states in between the bandgap will enrich and ultimately narrow the band gap of TiO2. Thus, the formation of vacancies results in shifts of average band energies, change of band widths and position of the Fermi level with the creation of new “vacancy states” in the bandgap. It also formulates an accumulation layer in the near-surface region that leads to a downward band bending [126].
The structural model of certain reduced Titania with their characteristic optical behavior and preparation methods are illustrated in Figure 12 [127]. As seen in Figure 12, Ti and O reside in their lattice and give rise to a colorless compound in case of stoichiometric Titania (TiO2). While on heat treatments under the H2 atmosphere, the surface oxygen vacancies are created and spread into bulk, which resulted in a pale blue color of a crystal. On continuing the heat treatment further, some oxygen vacancies co-exist with the conduction electrons in dark blue and dark green crystals, whereas, the conduction electrons completely migrate to the oxygen vacancy site on extensive heat treatment under the N2 atmosphere, so that the crystal appears yellow in color. Thus, different defect concentration in reduced titanium oxide forms colorful titanium oxide.

5. Effect of Defects on Electrical Behavior of Titania

Defects turn Titania from semiconductor to metallic depending on the intensity or level of structural deformation. Usually, lower oxides of titanium are metallic conductors while higher oxide content is semiconductor or insulator. This transition of conductivity from electronic conduction in titanium to a hopping mechanism in Titania is due to the presence of oxygen. At lower oxygen content, the corresponding Ti atoms are under great proximity and interact electronically, which leads to good conductivity [128,129]. Besides, the shear planes in magneli’s phase offer easy access for the charge transport and hence improvement in the conductivity as shown in Figure 13a [130]. The relative variation in conductivity of magneli’s phases due to the variation in oxygen vacancy content (shear planes) is shown in Figure 13b [92].
J. Nowotny et al. described a transition from n-type to p-type conductivity in Titania under different oxygen activity conditions (Figure 14) [49]. These distinctly different electronic behaviors of reduced Titania under different oxygen activity conditions are due to the prevalence of predominant defects and due to self-structural modification.
J. Nostalt et al. reported predominant defects and their corresponding conductivity at different oxygen abundance by the slope (m) (Table 4) [131]. The change in electrical properties at different oxygen proportions is due to the dissimilarity in electron occupancy, bond type, strength of interplay between Ti 3d orbitals and interaction between Ti 3d and O 2p orbitals (i.e., the complex interaction of charge order, orbital order). Subsequently, the electrons delocalization on different Ti ions or localized onsimilar Ti ions, which eventually leads to enhanced conductivity.

6. Effect of Defects on the Catalytic Activity of Titania

Point defects in Titania act as catalytically active sites due to the vacancy-electron, vacancy-vacancy interactions, and vacancy-strain coupling. Furthermore, clusters of vacancy and associative vacancy voids behave as a transfer station for atom exchange or migration because of their associated ‘conducting vacancy channels’ and plays a governing role in the catalytic activity. Also, the vacancy-induced electrons on defective Titania are localized on the Ti atom and accelerate local lattice distortion, constituting a polaron [132]. These mobile polarons form movement with low activation energy and favor conductivity and catalytic activity [133]. Besides, a large number of donor/acceptor defect states and a large density of dangling bonds in reduced Titania cause high surface activities and provide cooperative electrical effects (e.g., improved electron transfer and transport) for catalytic activity [134]. Further, defects induce an internal stress/strain, which is crucial for faster kinetics and favorable reactivity [135,136]. Plus, point defects have a potential to stabilize the binding of surface adsorbed species and escort to stiffer molecule-surface interactions which enhance the activity. Thus, these defect-related interface effects and defect-mediated charge migration effects augment the catalytic activity towards water-splitting [70,137,138,139,140,141,142,143].

7. Role of Defects towards Electro-Catalytic Water Splitting

Point defects in reduced Titania build a lattice strain in the material, which induces hydrophilicity on their surface than that of (defect-free) Titania without strain [140]. Hence, water molecules are dissociatively adsorbed on the defect sites, and consequently, the number of bridging/terminal hydroxyl groups increases with increasing concentration of structural defects in reduced Titania. This hydroxyl group depresses the O+ yield (Knotek−Feibelman mechanism) and favors lateral long-range redistribution of defect-electrons on the surface [144]. Thus, surface chemistry and electrical behavior of Titania are modified by electrons associated with oxygen vacancy defect sites and hence catalytic activity. The mentioned K-F process inversely decays with the oxygen vacancy content (Figure 15).
Similarly, titanium vacancy offers favorable active sites for water-splitting as it possesses strong electron affinity, which allows effective charge transfer for oxygen evolution reaction [57]. Hence, overall water-splitting is possible with n- and p-type defective Titania.

8. Role of Defects towards Photoelectrochemical Water Splitting

In principle, efficient photoelectrochemical watersplitting requires proficient separation of photo-generated charge carriers (photon activation) and subsequent faster transportation to the semiconductor interface for water ionization [145].
Photon activation: TiO2 + hν → e + h+
Water ionization: H2O + h+ → OH + H+
Since many reactive oxygen species are generated as an intermediate during the photoelectrochemical watersplitting reaction (Figure 16), both bulk and surface characteristics of the semiconductor will significantly influence the mechanism and kinetics of charge separation and interfacial charge transfer. Sun et al. proposed the photoelectrochemical watersplitting reaction pathway in Titania (Figure 17). According to them, water-splitting occurs by the formation of hydroperoxide intermediate. Later, Ti–OH and Ti–OOH groups are formed due to nucleophilic attack of water following the coupling process and O–O bond formation. Then, it collapses to produce oxygen with regeneration of TiO2 [146].
Thus, based on the overall reaction mechanism, the presence of O2˚ and OH on the semiconductor’s surface are likely to be an active site for the photoelectrochemical water oxidation. The defective Titania is enriched with surface hydroxyl groups and as well, highly active species such as O2˚ and O22− will be produced on Ti3+ site when it reacts with dissolved oxygen (Figure 18) [147,148]. Whereas pristine Titania will barely afford chemisorptions sites. Hence, point defects in reduced Titania act as a reactive center for photoelectrochemical water-splitting reaction.
Apart from the surface phenomena, bulk characteristics of defective Titania like superficial and deep defect states offer a dominant pathway for charge transport [148,149,150,151,152,153,154,155,156,157,158]. Also, defective Titania absorbs visible light, and the defect concentration tunes their optical behavior [159,160,161,162]. Importantly, defective Titania possesses numerous dangling bonds and serve as a bridge to facilitate charge transfer once it is photo-excited. Thus, benefiting from the surface adsorption and bulk properties through point defects, reduced Titania could serve as an efficient photo-electrocatalyst and as a generic platform for other defective metal oxides.

9. Challenges in Reduced Titania

Point defects in reduced Titania are usually not stable in air, as Ti3+/Ti2+ is easily oxidized and even susceptible to oxidation by dissolved oxygen in the water. Furthermore, the strategies to develop reduced Titania involve multiple steps, harsh synthesis conditions, or expensive facilities [69,162,163,164,165,166,167,168]. Therefore, developing a simple and economic strategy to synthesize stable reduced TiO2 is still a great challenge. Besides, the control of the quality and quantity of defects during preparation is complicated. However, the optimal concentration of point defects in reduced Titania is essential for the best catalytic activity. At higher concentrations, it functions as an electron trap and in case of photo-electrochemical water-splitting it acts as a scatterer in electro-catalytic water-splitting [169]. Moreover, reduced Titania has mostly triclinic crystal symmetry with very similar lattice parameters. As a result, the formed material usually contains various sub-stoichiometries.

10. Applications

The diverse characteristics of reduced Titania such as robustness, electron richness, chemical inertness (stability), electrical conductivity, visible absorption makes them a suitable candidate for cathodic protection of steel in concrete, photodegradation of organic compounds, CO2 reduction, Oxygen Reduction Reaction, solar cells, fuel cells, coating for tribological improvement, remediation of aqueous waste and contaminated water, electronic and optoelectronic devices, field emission, supercapacitors, microwave absorbers, photoelectrochemical sensor etc. [12,68,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201]. It is also used as a chemical catalyst for ethylene polymerization, production of chemicals such as pyrite, glyoxylic acid and oxidation of flue gases, trichloroethylene. The enhanced catalytic activity has been frequently attributed to the presence of oxygen vacancies, Ti3+/2+ species, hydroxyl groups, mid-gap energy levels. It also acts as a promising catalytic support for platinum and shows performance comparable to carbon black in PEM fuel cell technologies.

11. Towards Commercial Development

Currently, certain companies like Atranova, Atraverda, are in the process of making batteries, fuel cells with an electrically conductive titanium suboxide (Ti4O7) with the trade name Ebonex®. Lithium-ion rechargeable battery shows excellent rate performance with reduced Titania due to the oxygen vacancies, which creates well-balanced Li+/e diffusion and provides short lithium-ion diffusion path and reduced charge diffusion resistance. Besides, the disordered surface of reduced Titania improves pseudocapacitive lithium storage. Vector Corrosion Technologies use titanium suboxides for cathodic protection of concrete [39,40,41].

12. Conclusions

Without a doubt, titanium sub-oxides with different crystal structures and colors have attracted tremendous interest due to their outstanding, electrical, optical and catalytic properties. The controlled oxygen vacancies in Titania impose significant structural modifications and substantially vary the conductivity from n-type to p-type semiconductors, and even quasi-metallic conductors. It also turns a colorless, inert Titania to a colorful, reactive defect-rich Titania. Despite the excellent electrical conductivity and visible absorption, their catalytic activity is still far from the ideal expectation. Thus, managing and optimizing the defects and the spatial distribution for catalytic activity is a critical issue. Different synthesis methods usually lead to various properties of defective titanium oxides. Importantly, much efforts and optimization are required to control the defects (n/p-type), concentration, and their spatial distribution. It is practically difficult to introduce defects of specific types into the desired location and at a controlled proportion, which requires more in-depth investigations. Also, new methods and tools must be developed to precisely characterize the defects in Titania lattice. For example, the atomic scale understanding, dynamics of defect, migration, and reconstruction of defects in Titania are still unexplored. The authors feel that the theoretical investigations and simulation methods will help to understand the structure–property relationship of defect-rich Titania. Besides, the proposed mechanism of catalytic activity in reduced TiO2 should be evidenced through a more specific way. Also, it is of dire importance to consider the overall changes of the chemical environment in Titania rather than to focus locally on the defects, such as Ti3+ or VO. However, comprehensively the defect-rich Titania will be a futuristic catalyst for water-splitting applications with a combinatorial attempt from materials scientists, physicists, and chemists.

Acknowledgments

S.J. thanks the Department of Science and Technology for the award of DST-INSPIRE Fellowship and Bhaskara Advanced Solar Energy (BASE) Internship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Askeland, D.R.; Pradeep, P.F.; Wright, W.J. Imperfections in the atomic and Ionic Arrangements. In The Science and Engineering of Materials, 6th ed.; Global Engineering: Boston, MA, USA, 2014; pp. 113–145. ISBN 978-0-495-29602-7. [Google Scholar]
  2. Heather, S. “DNA and Evolution” ThoughtCo. Available online: Thoughtco.com/dna-and-evolution-1224567 (accessed on 20 October 2018).
  3. Hu, W.; Liu, Y.; Withers, R.L.; Frankcombe, T.J.; Norén, L.; Snashall, A.; Kitchin, M.; Smith, P.; Gong, B.; Chen, H.; et al. Electron-pinned defect-dipoles for high-performance colossal permittivity materials. Nat. Mater. 2013, 12, 821–826. [Google Scholar] [CrossRef] [PubMed]
  4. Repp, S.; Weber, S.; Erdem, E. Defect evolution of nonstoichiometric ZnO Quantum Dots. J. Phys. Chem. C 2016, 120, 25124–25130. [Google Scholar] [CrossRef]
  5. Hersam, M.C. Defects at the Two-Dimensional Limit. J. Phys. Chem. Lett. 2015, 6, 2738–2739. [Google Scholar] [CrossRef] [PubMed]
  6. Martinez, U.; Hansen, J.; Lira, E.; Kristoffersen, H.H.; Huo, P.; Bechstein, R.; Lægsgaard, E.; Besenbacher, F.; Hammer, B.; Wendt, S. Reduced step edges on rutile TiO2(110) as competing defects to oxygen vacancies on the terraces and reactive sites for ethanol dissociation. Phys. Rev. Lett. 2012, 109, 1–5. [Google Scholar] [CrossRef] [PubMed]
  7. Cremer, T.; Jensen, S.C.; Friend, C.M. Enhanced photo-oxidation of formaldehyde on highly reduced o-TiO2(110). J. Phys. Chem. C 2014, 118, 29242–29251. [Google Scholar] [CrossRef]
  8. Sivula, K. Defects Give New Life to an Old Material: Electronically Leaky Titania as a Photoanode Protection Layer. ChemCatChem 2014, 6, 2796–2797. [Google Scholar] [CrossRef]
  9. Gai-Boyes, P.L. Defects in Oxide Catalysts: Fundamental Studies of Catalysis in Action. Catal. Rev. 1992, 34, 1–54. [Google Scholar] [CrossRef]
  10. Attariani, H.; Momeni, K.; Adkins, K. Defect Engineering: A Path toward Exceeding Perfection. ACS Omega 2017, 2, 663–669. [Google Scholar] [CrossRef]
  11. Barzan, C.; Groppo, E.; Bordiga, S.; Zecchina, A. Defect Sites in H2—Reduced TiO2 Convert Ethylene to High Density Polyethylene without Activator. ACS Catal. 2014, 4, 986–989. [Google Scholar] [CrossRef]
  12. Seo, H.; Park, C.J.; Cho, Y.J.; Kim, Y.B.; Choi, D.K. Correlation of band edge native defect state evolution to bulk mobility changes in ZnO thin films. Appl. Phys. Lett. 2010, 96, 65–68. [Google Scholar] [CrossRef]
  13. Myung, S.-T.; Kikuchi, M.; Yoon, C.S.; Yashiro, H.; Kim, S.-J.; Sun, Y.-K.; Scrosati, B. Black anatase titania enabling ultra high cycling rates for rechargeable lithium batteries. Energy Environ. Sci. 2013, 6, 2609. [Google Scholar] [CrossRef]
  14. Wang, X.; Huang, K.; Ma, W.; Cong, Y.; Ge, C.; Feng, S. Defect Engineering, Electronic Structure, and Catalytic Properties of Perovskite Oxide La0.5Sr0.5CoO3−δ. Chem. A Eur. J. 2017, 23, 1093–1100. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, N.; Li, X.; Ye, H.; Chen, S.; Ju, H.; Liu, D.; Lin, Y.; Ye, W.; Wang, C.; Xu, Q.; et al. Oxide Defect Engineering Enables to Couple Solar Energy into Oxygen Activation. J. Am. Chem. Soc. 2016, 138, 8928–8935. [Google Scholar] [CrossRef] [PubMed]
  16. Lin, Z.; Carvalho, B.R.; Kahn, E.; Lv, R.; Rao, R.; Terrones, H.; Pimenta, M.A.; Terrones, M. Defect engineering of two-dimensional transition metal dichalcogenides. 2D Mater. 2016, 3, 1–21. [Google Scholar] [CrossRef]
  17. Feng, H.; Xu, Z.; Ren, L.; Liu, C.; Zhuang, J.; Hu, Z.; Xu, X.; Chen, J.; Wang, J.; Hao, W.; et al. Activating Titania for Efficient Electrocatalysis by Vacancy Engineering. ACS Catal. 2018, 8, 4288–4293. [Google Scholar] [CrossRef]
  18. Zhang, H.; Lv, R. Defect engineering of two-dimensional materials for efficient electrocatalysis. J. Mater. 2018, 4, 95–107. [Google Scholar] [CrossRef]
  19. Chen, J.; Iyemperumal, S.K.; Fenton, T.; Carl, A.D.; Grimm, R.L.; Li, G.; Deskins, N.A. Synergy between Defects, Photoexcited Electrons, and Supported Single Atom Catalysts for CO2 Reduction. ACS Catal. 2018. [Google Scholar] [CrossRef]
  20. Swaminathan, J.; Subbiah, R.; Singaram, V. Defect-Rich Metallic Titania (TiO1.23)—An Efficient Hydrogen Evolution Catalyst for Electrochemical Water Splitting. ACS Catal. 2016, 6. [Google Scholar] [CrossRef]
  21. Liu, C.; Zhang, A.Y.; Pei, D.N.; Yu, H.Q. Efficient Electrochemical Reduction of Nitrobenzene by Defect-Engineered TiO2−x Single Crystals. Environ. Sci. Technol. 2016, 50, 5234–5242. [Google Scholar] [CrossRef]
  22. Li, Y.H.; Liu, P.F.; Pan, L.F.; Wang, H.F.; Yang, Z.Z.; Zheng, L.R.; Hu, P.; Zhao, H.J.; Gu, L.; Yang, H.G. Local atomic structure modulations activate metal oxide as electrocatalyst for hydrogen evolution in acidic water. Nat. Commun. 2015, 6. [Google Scholar] [CrossRef]
  23. Bak, T.; Nowotny, J.; Nowotny, M.K.; Sheppard, L.R. Defect engineering of titanium dioxide. J. Aust. Ceram. Soc. 2008, 44, 63–67. [Google Scholar] [CrossRef]
  24. Lee, H.; Clark, S.J.; Robertson, J. Calculation of point defects in rutile TiO2 by the Screened Exchange Hybrid Functional. Phys. Rev. B 2012, 86, 075209. [Google Scholar] [CrossRef]
  25. Lu, T.C.; Wu, S.Y.; Lin, L.B.; Zheng, W.C. Defects in the reduced rutile single crystal. Phys. B Condens. Matter 2001, 304, 147–151. [Google Scholar] [CrossRef]
  26. Eyert, V.; Schwingenschlogl, U.; Eckern, U. Charge order, orbital order, and electron localization in the Magneli phase Ti4O7. Chem. Phys. Lett. 2004, 390, 151–156. [Google Scholar] [CrossRef]
  27. Wang, Z.W.; Shu, D.J.; Wang, M.; Ming, N. Ben Strain effect on diffusion properties of oxygen vacancies in bulk and subsurface of rutile TiO2. Surf. Sci. 2012, 606, 186–191. [Google Scholar] [CrossRef]
  28. Boonchun, A.; Reunchan, P.; Umezawa, N. Energetics of native defects in anatase TiO2: A hybrid density functional study. Phys. Chem. Chem. Phys. 2016, 18, 30040–30046. [Google Scholar] [CrossRef]
  29. Yagi, E.; Hasiguti, R.R.; Aono, M. Electronic conduction above 4 K of slightly reduced oxygen-deficient rutile TiO2−x. Phys. Rev. B 1996, 54, 7945–7956. [Google Scholar] [CrossRef]
  30. Finazzi, E.; Di Valentin, C.; Pacchioni, G.; Selloni, A. Excess electron states in reduced bulk anatase TiO2: Comparison of standard GGA, GGA+U, and hybrid DFT calculations. J. Chem. Phys. 2008, 129. [Google Scholar] [CrossRef]
  31. Diebold, U. The surface science of titanium dioxide. Surf. Sci. Rep. 2003, 48, 53–229. [Google Scholar] [CrossRef]
  32. Bak, T.; Nowotny, J.; Rekas, M.; Sorrell, C.C. Defect chemistry and semiconducting properties of titanium dioxide: II. Defect diagrams. J. Phys. Chem. Solids 2003, 64, 1057–1067. [Google Scholar] [CrossRef]
  33. Xiao, P.; Fang, H.; Cao, G.; Zhang, Y.; Zhang, X. Effect of Tin+ defects on electrochemical properties of highly-ordered titania nanotube arrays. Thin Solid Films 2010, 518, 7152–7155. [Google Scholar] [CrossRef]
  34. Agrawal, A.; Lin, J.; Barth, M.; White, R.; Zheng, B.; Chopra, S.; Gupta, S.; Wang, K.; Gelatos, J.; Mohney, S.E.; et al. Fermi level depinning and contact resistivity reduction using a reduced titania interlayer in n-silicon metal-insulator-semiconductor ohmic contacts. Appl. Phys. Lett. 2014, 104, 8–12. [Google Scholar] [CrossRef]
  35. Nowotny, M.K.; Bak, T.; Nowotny, J. Electrical properties and defect chemistry of TiO2 single crystal. II. Thermoelectric power. J. Phys. Chem. B 2006, 110, 16283–16291. [Google Scholar] [CrossRef] [PubMed]
  36. Nowotny, J. Titanium dioxide-based semiconductors for solar-driven environmentally friendly applications: Impact of point defects on performance. Energy Environ. Sci. 2008, 1, 565–572. [Google Scholar] [CrossRef]
  37. Santo, V.D.; Naldoni, A. The Effect of Point Defects on the Ordered/Disorderd Morphology on the Electronic and Structural Features in the Black TiO2 Nanomaterials; World Scientific: Kansas city, KS, USA, 2017; Chapter 4; pp. 49–75. ISBN 978-1-78634-165-5. [Google Scholar]
  38. Nowotny, J.; Alim, M.A.; Bak, T.; Idris, M.A.; Ionescu, M.; Prince, K.; Sahdan, M.Z.; Sopian, K.; Mat Teridi, M.A.; Sigmund, W. Defect chemistry and defect engineering of TiO2-based semiconductors for solar energy conversion. Chem. Soc. Rev. 2015, 44, 8424–8442. [Google Scholar] [CrossRef] [PubMed]
  39. Su, J.; Zou, X.; Chen, J.-S. Self-modification of titanium dioxide materials by Ti3+ and/or oxygen vacancies: New insights into defect chemistry of metal oxides. RSC Adv. 2014, 4, 13979–13988. [Google Scholar] [CrossRef]
  40. 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]
  41. Chen, X.; Liu, L.; Huang, F. Black titanium dioxide (TiO2) nanomaterials. Chem. Soc. Rev. 2015, 44, 1861–1885. [Google Scholar] [CrossRef]
  42. Fang, W.; Xing, M.; Zhang, J. Modifications on reduced titanium dioxide photocatalysts: A review. J. Photochem. Photobiol. C Photochem. Rev. 2017, 32, 21–39. [Google Scholar] [CrossRef]
  43. Easteal, A.J.; Udy, D.J. Extraction of Titanium Dioxide (TiO2) from Ilmenite and Titaniferous Slag. J. Appl. Chem. Biotechnol. 1973, 23, 865–870. [Google Scholar] [CrossRef]
  44. Buettner, K.M.; Valentine, A.M. Bioinorganic chemistry of Titanium. Chem. Rev. 2012, 112, 1863–1881. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, L.; Chen, X. Titanium dioxide nanomaterials: Self-structural modifications. Chem. Rev. 2014, 114, 9890–9918. [Google Scholar] [CrossRef] [PubMed]
  46. Seebauer, E.G.; Kratzer, M.C. Charged point defects in semiconductors. Mater. Sci. Eng. R Rep. 2006, 55, 57–149. [Google Scholar] [CrossRef]
  47. Nowotny, J.; Bak, T.; Ali, M.A. Semiconducting Properties and Defect Disorder of Titanium Dioxide. ECS Trans. 2015, 64, 11–28. [Google Scholar] [CrossRef]
  48. Bak, T.; Nowotny, J.; Nowotny, M.K.; Sheppard, L.R. Defect chemistry of titanium dioxide effect of interfaces. J. Aust. Ceram. Soc. 2007, 43, 49–55. [Google Scholar]
  49. Nowotny, J.; Bak, T.; Burg, T. Electrical properties of polycrystalline TiO2 at elevated temperatures. Electrical conductivity. Phys. Status Solidi Basic Res. 2007, 244, 2037–2054. [Google Scholar] [CrossRef]
  50. Serpone, N. Is the band gap of pristine TiO2 narrowed by anion- and cation-doping of titanium dioxide in second-generation photocatalysts? J. Phys. Chem. B 2006, 110, 24287–24293. [Google Scholar] [CrossRef] [PubMed]
  51. Kuznetsov, V.N.; Serpone, N. On the Origin of the Spectral Bands in the Visible Absorption Spectra of Visible-Light-Active TiO2 Specimens Analysis and Assignments. J. Phys. Chem. C 2009, 113, 15110–15123. [Google Scholar] [CrossRef]
  52. Swaminathan, J.; Ravichandran, S. Insights into the Electrocatalytic Behavior of Defect-Centered Reduced Titania (TiO1.23). J. Phys. Chem. C 2018, 122. [Google Scholar] [CrossRef]
  53. Chen, J.; Lin, L.B.; Jing, F.Q. Theoretical study of F-type color center in rutile TiO2. J. Phys. Chem. Solids 2001, 62, 1257–1262. [Google Scholar] [CrossRef]
  54. Kittel, C. Introduction to Solid State Physics, 7th ed.; John Wiley & Sons: New York, NY, USA, 1996; pp. 585–595. ISBN 13: 9780471111818. [Google Scholar]
  55. Morgan, B.J.; Watson, G.W. Intrinsic n-type Defect Formation in TiO2: A Comparison of Rutile and Anatase from GGA+U Calculations. J. Phys. Chem. C 2010, 114, 2321–2328. [Google Scholar] [CrossRef]
  56. Na-Phattalung, S.; Smith, M.F.; Kim, K.; Du, M.H.; Wei, S.H.; Zhang, S.B.; Limpijumnong, S. First-principles study of native defects in anatase TiO2. Phys. Rev. B Condens. Matter Mater. Phys. 2006, 73, 1–6. [Google Scholar] [CrossRef]
  57. Nowotny, M.K.; Sheppard, L.R.; Bak, T.; Nowotny, J. Defect Chemistry of Titanium Dioxide. Application of Defect Engineering in Processing of TiO2 based photocatalysts. J. Phys. Chem. C 2008, 112, 5275–5300. [Google Scholar] [CrossRef]
  58. Bak, T.; Burg, T.; Kang, S.J.L.; Nowotny, J.; Rekas, M.; Sheppard, L.; Sorrell, C.C.; Vance, E.R.; Yoshida, Y.; Yamawaki, M. Charge transport in polycrystalline titanium dioxide. J. Phys. Chem. Solids 2003, 64, 1089–1095. [Google Scholar] [CrossRef]
  59. Kofstad, P. High-temperature oxidation of titanium. J. Less Common Met. 1967, 12, 449–464. [Google Scholar] [CrossRef]
  60. Nowotny, J.; Bak, T.; Burg, T. Electrical properties of polycrystalline TiO2. Ionics 2007, 13, 79–82. [Google Scholar] [CrossRef]
  61. Tao, J.; Luttrell, T.; Batzill, M. A two-dimensional phase of TiO2 with a reduced bandgap. Nat. Chem. 2011, 3, 296–300. [Google Scholar] [CrossRef]
  62. Etacheri, V.; Seery, M.K.; Hinder, S.J.; Pillai, S.C. Oxygen rich titania: A dopant free, high temperature stable, and visible-light active anatase photocatalyst. Adv. Funct. Mater. 2011, 21, 3744–3752. [Google Scholar] [CrossRef]
  63. Bowker, M.; Bennett, R.A. The role of Ti3+ interstitials in TiO2(110) reduction and oxidation. J. Phys. Condens. Matter 2010, 22, 059801. [Google Scholar] [CrossRef]
  64. Tanaka, T.; Sumiya, A.; Sawada, H.; Kondo, Y.; Takayanagi, K. Direct observation of interstitial titanium ions in TiO2 substrate with gold nanoparticle. Surf. Sci. 2014, 619, 39–43. [Google Scholar] [CrossRef]
  65. Tanaka, T.; Sano, K.; Ando, M.; Sumiya, A.; Sawada, H.; Hosokawa, F.; Okunishi, E.; Kondo, Y.; Takayanagi, K. Oxygen-rich Ti1−x O2 pillar growth at a gold nanoparticle-TiO2 contact by O2 exposure. Surf. Sci. 2010, 604, L75–L78. [Google Scholar] [CrossRef]
  66. Ohwada, M.; Kimoto, K.; Mizoguchi, T.; Ebina, Y.; Sasaki, T. Atomic structure of titania nanosheet with vacancies. Sci. Rep. 2013, 3, 2801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Nowotny, M.K.; Bak, T.; Nowotny, J.; Sorrell, C.C. Titanium vacancies in nonstoichiometric TiO2 single crystal. Phys. Status Solidi Basic Res. 2005, 242, 88–90. [Google Scholar] [CrossRef]
  68. Zheng, Q.; Lee, H.-J.; Lee, J.; Choi, W.; Park, N.-B.; Lee, C. Electrochromic titania nanotube arrays for the enhanced photocatalytic degradation of phenol and pharmaceutical compounds. Chem. Eng. J. 2014, 249, 285–292. [Google Scholar] [CrossRef]
  69. Wang, G.; Wang, H.; Ling, Y.; Tang, Y.; Yang, X.; Fitzmorris, R.C.; Wang, C.; Zhang, J.Z.; Li, Y. Hydrogen-Treated TiO2 Nanowire Arrays for Photoelectrochemical Water Splitting. Nano Lett. 2011, 11, 3026–3033. [Google Scholar] [CrossRef] [PubMed]
  70. Chen, X.; Liu, L.; Liu, Z.; Marcus, M.A.; Wang, W.-C.; Oyler, N.A.; Grass, M.E.; Mao, B.; Glans, P.-A.; Yu, P.Y.; et al. Properties of Disorder-Engineered Black Titanium Dioxide Nanoparticles through Hydrogenation. Sci. Rep. 2013, 3, 1510. [Google Scholar] [CrossRef] [Green Version]
  71. Chen, X.; Liu, L.; Yu, P.Y.; Mao, S.S. Increasing solar absorption for photocatalysis with black hydrogenated titanium dioxide nanocrystals. Science 2011, 331, 746–750. [Google Scholar] [CrossRef]
  72. Nakajima, T.; Nakamura, T.; Shinoda, K.; Tsuchiya, T. Rapid formation of black titania photoanodes: Pulsed laser-induced oxygen release and enhanced solar water splitting efficiency. J. Mater. Chem. A 2014, 2, 6762. [Google Scholar] [CrossRef]
  73. Rahman, M.A.; Bazargan, S.; Srivastava, S.; Wang, X.; Abd-Ellah, M.; Thomas, J.P.; Heinig, N.F.; Pradhan, D.; Leung, K.T. Defect-rich decorated TiO2 nanowires for super-efficient photoelectrochemical water splitting driven by visible light. Energy Environ. Sci. 2015, 8, 3363–3373. [Google Scholar] [CrossRef]
  74. Wang, Z.; Yang, C.; Lin, T.; Yin, H.; Chen, P.; Wan, D.; Xu, F.; Huang, F.; Lin, J.; Xie, X.; et al. Visible-light photocatalytic, solar thermal and photoelectrochemical properties of aluminium-reduced black titania. Energy Environ. Sci. 2013, 6, 3007. [Google Scholar] [CrossRef]
  75. Tang, C.; Zhou, D.; Zhang, Q. Synthesis and characterization of Magneli phases: Reduction of TiO2 in a decomposed NH3 atmosphere. Mater. Lett. 2012, 79, 42–44. [Google Scholar] [CrossRef]
  76. Simon, P.; Pignon, B.; Miao, B.; Coste-Leconte, S.; Leconte, Y.; Marguet, S.; Jegou, P.; Bouchet-Fabre, B.; Reynaud, C.; Herlin-Boime, N. N-doped titanium monoxide nanoparticles with TiO rock-salt structure, low energy band gap, and visible light activity. Chem. Mater. 2010, 22, 3704–3711. [Google Scholar] [CrossRef]
  77. Takeuchi, T.; Fukushima, J.; Hayashi, Y.; Takizawa, H. Synthesis of Ti4O7 Nanoparticles by Carbothermal Reduction Using Microwave Rapid Heating. Catalysts 2017, 7, 65. [Google Scholar] [CrossRef]
  78. Tominaka, S.; Tsujimoto, Y.; Matsushita, Y.; Yamaura, K. Synthesis of nanostructured reduced titanium Oxide: Crystal structure transformation maintaining nanomorphology. Angew. Chem. Int. Ed. 2011, 50, 7418–7421. [Google Scholar] [CrossRef]
  79. Guan, S.; Hao, L.; Lu, Y.; Yoshida, H.; Pan, F.; Asanuma, H. Fabrication of oxygen-deficient TiO2 coatings with nano-fiber morphology for visible-light photocatalysis. Mater. Sci. Semicond. Process. 2016, 41, 358–363. [Google Scholar] [CrossRef]
  80. Zou, X.; Liu, J.; Su, J.; Zuo, F.; Chen, J.; Feng, P. Facile synthesis of thermal- and photostable titania with paramagnetic oxygen vacancies for visible-light photocatalysis. Chem. A Eur. J. 2013, 19, 2866–2873. [Google Scholar] [CrossRef] [PubMed]
  81. Huang, S.S.; Lin, Y.H.; Chuang, W.; Shao, P.S.; Chuang, C.H.; Lee, J.F.; Lu, M.L.; Weng, Y.T.; Wu, N.L. Synthesis of High-Performance Titanium Sub-Oxides for Electrochemical Applications Using Combination of Sol-Gel and Vacuum-Carbothermic Processes. ACS Sustain. Chem. Eng. 2018, 6, 3162–3168. [Google Scholar] [CrossRef]
  82. Sinhamahapatra, A.; Jeon, J.P.; Yu, J.S. A new approach to prepare highly active and stable black titania for visible light-assisted hydrogen production. Energy Environ. Sci. 2015, 8, 3539–3544. [Google Scholar] [CrossRef] [Green Version]
  83. Grigorov, K.G.; Grigorov, G.I.; Drajeva, L.; Bouchier, D.; Sporken, R.; Caudano, R. Synthesis and characterization of conductive titanium monoxide films. Diffusion of silicon in titanium monoxide films. Vacuum 1998, 51, 153–155. [Google Scholar] [CrossRef]
  84. Xiong, L.B.; Li, J.L.; Yang, B.; Yu, Y. Ti3+ in the surface of titanium dioxide: Generation, properties and photocatalytic application. J. Nanomater. 2012, 2012. [Google Scholar] [CrossRef]
  85. Haubrich, J.; Kaxiras, E.; Friend, C.M. The role of surface and subsurface point defects for chemical model studies on TiO2: A first-principles theoretical study of formaldehyde bonding on rutile TiO2(110). Chem. A Eur. J. 2011, 17, 4496–4506. [Google Scholar] [CrossRef] [PubMed]
  86. Li, G.; Li, L.; Zheng, J. Understanding the defect chemistry of oxide nanoparticles for creating new functionalities: A critical review. Sci. China Chem. 2011, 54, 876–886. [Google Scholar] [CrossRef]
  87. Janotti, A.; Varley, J.B.; Rinke, P.; Umezawa, N.; Kresse, G.; Van de Walle, C.G. Hybrid functional studies of the oxygen vacancy in TiO2. Phys. Rev. B 2010, 81, 085212. [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] [PubMed]
  89. Parker, J.C.; Siegel, R.W. Raman Microprobe Study of Nanophase TiO2 and Oxidation-Induced Spectral Changes. J. Mater. Res. 1990, 5, 1246–1252. [Google Scholar] [CrossRef]
  90. Stoyanov, E.; Langenhorst, F.; Steinle-Neumann, G. The effect of valence state and site geometry on Ti L3,2 and O K electron energy-loss spectra of TixOy phases. Am. Mineral. 2007, 92, 577–586. [Google Scholar] [CrossRef]
  91. Parras, M.; Varela, Á.; Cortés-Gil, R.; Boulahya, K.; Hernando, A.; González-Calbet, J.M. Room-temperature ferromagnetism in reduced rutile TiO2−δ nanoparticles. J. Phys. Chem. Lett. 2013, 4, 2171–2176. [Google Scholar] [CrossRef]
  92. Walsh, F.C.; Wills, R.G.A. The continuing development of Magnéli phase titanium sub-oxides and Ebonex® electrodes. Electrochim. Acta 2010, 55, 6342–6351. [Google Scholar] [CrossRef]
  93. Arif, A.F.; Balgis, R.; Ogi, T.; Iskandar, F.; Kinoshita, A.; Nakamura, K.; Okuyama, K. Highly conductive nano-sized Magnéli phases titanium oxide (TiOx). Sci. Rep. 2017, 7, 3646. [Google Scholar] [CrossRef]
  94. Watanabe, M. Raman spectroscopy of charge-ordered states in Magneli titanium oxides. Phys. Status Solidi 2009, 6, 260–263. [Google Scholar] [CrossRef]
  95. Åsbrink, S.; Magnéli, A. Crystal structure studies on trititanium pentoxide, Ti3O5. Acta Crystallogr. 1959, 12, 575–581. [Google Scholar] [CrossRef]
  96. Onoda, M. Phase Transitions of Ti3O5. J. Solid State Chem. 1998, 136, 67–73. [Google Scholar] [CrossRef]
  97. Ovsyannikov, S.V.; Wu, X.; Shchennikov, V.V.; Karkin, A.E.; Dubrovinskaia, N.; Garbarino, G.; Dubrovinsky, L. Structural stability of a golden semiconducting orthorhombic polymorph of Ti2O3 under high pressures and high temperatures. J. Phys. Condens. Matter 2010, 22. [Google Scholar] [CrossRef]
  98. Wu, Y.; Zhang, Q.; Wu, X.; Qin, S.; Liu, J. High pressure structural study of β-Ti3O5: X-ray diffraction and Raman spectroscopy. J. Solid State Chem. 2012, 192, 356–359. [Google Scholar] [CrossRef]
  99. Kao, C.H.; Yeh, S.W.; Huang, H.L.; Gan, D.; Shen, P. Study of the TiO to anatase transformation by thermal oxidation of Ti film in air. J. Phys. Chem. C 2011, 115, 5648–5656. [Google Scholar] [CrossRef]
  100. Kostenko, M.G.; Lukoyanov, A.V.; Rempel, A.A. Electronic structure and stability of nonstoichiometric titanium monoxide TiOy with structural vacancies in one of the sublattices. Phys. Solid State 2013, 55, 2108–2115. [Google Scholar] [CrossRef]
  101. Bartkowski, S.; Neumann, M. Electronic structure of titanium monoxide. Phys. Rev. B 1997, 56, 10656–10667. [Google Scholar] [CrossRef]
  102. Kostenko, M.G.; Lukoyanov, A.V.; Zhukov, V.P.; Rempel, A.A. Vacancies in ordered and disordered titanium monoxide: Mechanism of B1 structure stabilization. J. Solid State Chem. 2013, 204, 146–152. [Google Scholar] [CrossRef] [Green Version]
  103. Barudzija, T.; Gusev, A.; Jugovic, D.; Marinovic-Cincovic, M.; Dramicanin, M.; Zdujic, M.; Jovalekic, C.; Mitric, M. Structural and magnetic properties of mechanochemically synthesized nanocrystalline titanium monoxide. Hem. Ind. 2012, 66, 309–315. [Google Scholar] [CrossRef]
  104. Guo, C.; Jia, S.; Meng, W.; Zheng, H.; Jin, L.; Liu, Y.; Shi, J.; Wang, J. Orientation domains in vacancy-ordered titanium monoxide. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 2013, 69, 589–594. [Google Scholar] [CrossRef]
  105. Valeeva, A.A.; Rempel, A.A.; Sprengel, W.; Schaefer, H. Identification and study of vacancies in titanium monoxide by means of positron annihilation techniques. Phys. Chem. Chem. Phys. 2003, 5, 2304–2307. [Google Scholar] [CrossRef]
  106. Fujimura, T.; Iwasaki, H.; Kkegawa, T.; Tsuchida, Y.; Shimomura, O.; Factory, P.; Corporation, K.S.; Iwasaki, H.; Kkegawa, T.; Tsuchida, Y.; et al. Structure changes in vacancy-rich titanium monoxide at high pressures and high temperatures. High Press. Res. 2006, 37–41. [Google Scholar] [CrossRef]
  107. Denker, S.P. Electronic properties of titanium monoxide. J. Appl. Phys. 1966, 37, 142–149. [Google Scholar] [CrossRef]
  108. Gunda, N.S.H.; Puchala, B.; Van der Ven, A. Resolving phase stability in the Ti-O binary with first-principles statistical mechanics methods. Phys. Rev. Mater. 2018, 2, 033604. [Google Scholar] [CrossRef]
  109. Jostsons, A.; McDougall, P.G. Fault Structures in Ti2O. Phys. Status Solidi 1968, 29, 873–889. [Google Scholar] [CrossRef]
  110. Wang, Q.; Fan, J.; Zhang, S.; Yun, Y.; Zhang, J.; Zhang, P.; Hu, J.; Wang, L.; Shao, G. In situ coupling of Ti2O with rutile TiO2 as a core-shell structure and its photocatalysis performance. RSC Adv. 2017, 7, 54662–54667. [Google Scholar] [CrossRef]
  111. Burton, B.P.; Van De Walle, A. First principles phase diagram calculations for the octahedral-interstitial system αTiOX, 0 ≤ X ≤ 1/2. Calphad 2012, 39, 97–103. [Google Scholar] [CrossRef]
  112. Hamdy, M.S.; Amrollahi, R.; Mul, G. Surface Ti3+-containing (blue) titania: A unique photocatalyst with high activity and selectivity in visible light-stimulated selective oxidation. ACS Catal. 2012, 2, 2641–2647. [Google Scholar] [CrossRef]
  113. Liu, G.; Yin, L.C.; Wang, J.; Niu, P.; Zhen, C.; Xie, Y.; Cheng, H.M. A red anatase TiO2 photocatalyst for solar energy conversion. Energy Environ. Sci. 2012, 5, 9603–9610. [Google Scholar] [CrossRef]
  114. Liu, G.; Yang, H.G.; Wang, X.; Cheng, L.; Lu, H.; Wang, L.; Lu, G.Q.; Cheng, H.M. Enhanced photoactivity of oxygen-deficient anatase TiO2 sheets with dominant {001} facets. J. Phys. Chem. C 2009, 113, 21784–21788. [Google Scholar] [CrossRef]
  115. 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. A 2014, 2, 4429–4437. [Google Scholar] [CrossRef]
  116. Mu, J.; Spiecker, E.; Schmuki, P. Black TiO2 Nanotubes: Cocatalyst-Free Open-Circuit Hydrogen Generation. Nano Lett. 2014, 14, 3309–3313. [Google Scholar] [CrossRef]
  117. Kuznetsov, V.N.; Emeline, A.V.; Rudakova, A.V.; Aleksandrov, M.S.; Glazkova, N.I.; Lovtcius, V.A.; Kataeva, G.V.; Mikhaylov, R.V.; Ryabchuk, V.K.; Serpone, N. Visible-NIR light absorption of titania thermochemically fabricated from titanium and its alloys; UV-and visible-light-induced photochromism of yellow titania. J. Phys. Chem. C 2013, 117, 25852–25864. [Google Scholar] [CrossRef]
  118. Wang, H.; Lin, T.; Zhu, G.; Yin, H.; Lü, X.; Li, Y.; Huang, F. Colored titania nanocrystals and excellent photocatalysis for water cleaning. Catal. Commun. 2015, 60, 55–59. [Google Scholar] [CrossRef]
  119. Kako, T.; Umezawa, N.; Xie, K.; Ye, J. Undoped visible-light-sensitive titania photocatalyst. J. Mater. Sci. 2013, 48, 108–114. [Google Scholar] [CrossRef]
  120. Regonini, D.; Adamaki, V.; Bowen, C.R.; Pennock, S.R.; Taylor, J.; Dent, A.C.E. AC electrical properties of TiO2 and Magnéli phases, TinO2n−1. Solid State Ion. 2012, 229, 38–44. [Google Scholar] [CrossRef]
  121. Zhang, H.; Zhou, M.; Fu, Q.; Lei, B.; Lin, W.; Guo, H.; Wu, M.; Lei, Y. Observation of defect state in highly ordered titanium dioxide nanotube arrays. Nanotechnology 2014, 25, 275603. [Google Scholar] [CrossRef]
  122. Nowotny, M.K.; Bak, T.; Nowotny, J. Electrical Properties and Defect Chemistry of TiO2 Single Crystal. IV. Prolonged Oxidation Kinetics and Chemical Diffusion. J. Phys. Chem. B 2006, 110, 16270–16282. [Google Scholar] [CrossRef]
  123. Santara, B.; Giri, P.K.; Imakita, K.; Fujii, M. Evidence for Ti Interstitial Induced Extended Visible Absorption and Near Infrared Photoluminescence from Undoped TiO2 Nanoribbons: An In Situ Photoluminescence Study. J. Phys. Chem. C 2013, 117, 23402–23411. [Google Scholar] [CrossRef]
  124. Kevane, C.J. Oxygen Vaccancies and Electrical conduction in Metal Oxides. Phys. Rev. 1964, 133, A1431. [Google Scholar] [CrossRef]
  125. Hossain, F.M.; Murch, G.E.; Sheppard, L.; Nowotny, J. The Effect of Defect Disorder on the Electronic Structure of Rutile TiO2−x. Defect Diffus. Forum 2006, 251–252, 1–12. [Google Scholar] [CrossRef]
  126. Guo, Q.; Zhou, C.; Ma, Z.; Ren, Z.; Fan, H. Fundamental Processes in Surface Photocatalysis on TiO2. In Heterogeneous Photocatalysis—From Fundamentals to Green Applications; Springer: London, UK; pp. 361–370. ISBN 978-3-662-48717-4.361-370.
  127. Sekiya, T.; Yagisawa, T.; Kamiya, N.; Mulmi, D.D.; Kurita, S.; Murakami, Y.; Kodaira, T. Defects in anatase TiO2 single crystal controlled by heat treatments. J. Phys. Soc. Jpn. 2004, 73, 703–710. [Google Scholar] [CrossRef]
  128. Kitada, A.; Hasegawa, G.; Kobayashi, Y.; Kanamori, K.; Nakanishi, K.; Kageyama, H. Selective preparation of macroporous monoliths of conductive titanium oxides TinO2n−1 (n = 2, 3, 4, 6). J. Am. Chem. Soc. 2012, 134, 10894–10898. [Google Scholar] [CrossRef] [PubMed]
  129. Eder, D.; Kramer, R. Stoichiometry of titanium suboxide Part 2 Electric properties. Phys. Chem. Chem. Phys. 2003, 5, 1314–1319. [Google Scholar] [CrossRef]
  130. Kieslich, G.; Burkhardt, U.; Birkel, C.S.; Veremchuk, I.; Douglas, J.E.; Gaultois, M.W.; Lieberwirth, I.; Seshadri, R.; Stucky, G.D.; Grin, Y.; et al. Enhanced thermoelectric properties of the n-type Magneli phase WO2.90: Reduced thermal conductivity through microstructure engineering. J. Mater. Chem. A 2014, 2, 13492–13497. [Google Scholar] [CrossRef]
  131. Nowotny, J.; Radecka, M.; Rekas, M. Semiconducting properties of undoped TiO2. J. Phys. Chem Solids 1997, 58, 927–937. [Google Scholar] [CrossRef]
  132. Shibuya, T.; Yasuoka, K.; Mirbt, S.; Sanyal, B. A systematic study of polarons due to oxygen vacancy formation at the rutile TiO(110) surface by GGA+U and HSE06 methods. J. Phys. Condens. Matter 2012, 24. [Google Scholar] [CrossRef] [PubMed]
  133. Gono, P.; Wiktor, J.; Ambrosio, F.; Pasquarello, A. Surface Polarons Reducing Overpotentials in the Oxygen Evolution Reaction. ACS Catal. 2018, 8, 5847–5851. [Google Scholar] [CrossRef]
  134. Gravelle, P.C.; Juillet, F.; Meriaudeau, P.; Teichner, S.J. Surface reactivity of reduced titanium dioxide. Discuss. Faraday Soc. 1971, 52, 140–148. [Google Scholar] [CrossRef]
  135. Amakawa, K.; Sun, L.; Guo, C.; Hävecker, M.; Kube, P.; Wachs, I.E.; Lwin, S.; Frenkel, A.I.; Patlolla, A.; Hermann, K.; et al. How Strain Affects the Reactivity of Surface Metal Oxide Catalysts. Angew. Chem. Int. Ed. 2013, 52, 13553–13557. [Google Scholar] [CrossRef]
  136. Horikoshi, S.; Minatodani, Y.; Tsutsumi, H.; Uchida, H.; Abe, M.; Serpone, N. Influence of lattice distortion and oxygen vacancies on the uv-driven/microwave-assisted TiO2 photocatalysis. J. Photochem. Photobiol. A Chem. 2013, 265, 20–28. [Google Scholar] [CrossRef]
  137. Aschauer, U.; He, Y.; Cheng, H.; Li, S.C.; Diebold, U.; Selloni, A. Influence of subsurface defects on the surface reactivity of TiO2: Water on anatase (101). J. Phys. Chem. C 2010, 114, 1278–1284. [Google Scholar] [CrossRef]
  138. Schaub, R.; Thostrup, P.; Lopez, N.; Lægsgaard, E.; Stensgaard, I.; Nørskov, J.K.; Besenbacher, F. Oxygen vacancies as active sites for water dissociation on rutile TiO2(110). Phys. Rev. Lett. 2001, 87. [Google Scholar] [CrossRef] [PubMed]
  139. Jensen, S.C.; Friend, C.M. The dynamic roles of interstitial and surface defects on oxidation and reduction reactions on titania. Top. Catal. 2013, 56, 1377–1388. [Google Scholar] [CrossRef]
  140. Tilocca, A.; Selloni, A. Reaction pathway and free energy barrier for defect-induced water dissociation on the (101) surface of TiO2-anatase. J. Chem. Phys. 2003, 119, 7445–7450. [Google Scholar] [CrossRef]
  141. Petrik, N.G.; Kimmel, G.A. Reaction Kinetics of Water Molecules with Oxygen Vacancies on Rutile TiO2(110). J. Phys. Chem. C 2015, 119, 23059–23067. [Google Scholar] [CrossRef]
  142. Kolbrecka, K.; Przyluski, J. Sub-stoichiometric titanium oxides as ceramic electrodes for oxygen evolution-structural aspects of the voltammetric behaviour of TinO2n−1. Electrochim. Acta 1994, 39, 1591–1595. [Google Scholar] [CrossRef]
  143. Shan, Z.; Archana, P.S.; Shen, G.; Gupta, A.; Bakker, M.G.; Pan, S. NanoCOT: Low-Cost Nanostructured Electrode Containing Carbon, Oxygen, and Titanium for Efficient Oxygen Evolution Reaction. J. Am. Chem. Soc. 2015, 137, 11996–12005. [Google Scholar] [CrossRef]
  144. Zhang, Z.; Cao, K.; Yates, J.T. Defect-electron spreading on the TiO2(110) semiconductor surface by water adsorption. J. Phys. Chem. Lett. 2013, 4, 674–679. [Google Scholar] [CrossRef]
  145. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  146. Imanishi, A.; Okamura, T.; Ohashi, N.; Nakamura, R.; Nakato, Y. Mechanism of water photooxidation, at n-TiO2 (rutile) (110) and (100) surfaces: Dependence on solution pH. J. Am. Chem. Soc. 2007, 2, 11569–11578. [Google Scholar] [CrossRef] [PubMed]
  147. Komaguchi, K.; Maruoka, T.; Nakano, H.; Imae, I.; Ooyama, Y.; Harima, Y. Electron-transfer reaction of oxygen species on TiO2 nanoparticles induced by Sub-band-gap illumination. J. Phys. Chem. C 2010, 114, 1240–1245. [Google Scholar] [CrossRef]
  148. Pan, L.; Wang, S.; Zou, J.J.; Huang, Z.F.; Wang, L.; Zhang, X. Ti3+-defected and V-doped TiO2 quantum dots loaded on MCM-41. Chem. Commun. 2014, 50, 988–990. [Google Scholar] [CrossRef]
  149. Fujishima, A.; Zhang, X.; Tryk, D.A. TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008, 63, 515–582. [Google Scholar] [CrossRef]
  150. Di Paola, A.; Bellardita, M.; Palmisano, L.; Barbieriková, Z.; Brezová, V. Influence of crystallinity and OH surface density on the photocatalytic activity of TiO2 powders. J. Photochem. Photobiol. A Chem. 2014, 273, 59–67. [Google Scholar] [CrossRef]
  151. Radecka, M.; Trenczek-Zajac, A.; Zakrzewska, K.; Rekas, M. Effect of oxygen nonstoichiometry on photo-electrochemical properties of TiO2−x. J. Power Sources 2007, 173, 816–821. [Google Scholar] [CrossRef]
  152. Wheeler, D.A.; Ling, Y.; Dillon, R.J.; Fitzmorris, R.C.; Dudzik, C.G.; Zavodivker, L.; Rajh, T.; Dimitrijevic, N.M.; Millhauser, G.; Bardeen, C.; et al. Probing the nature of bandgap states in hydrogen-treated TiO2 nanowires. J. Phys. Chem. C 2013, 117, 26821–26830. [Google Scholar] [CrossRef]
  153. Lo, H.H.; Gopal, N.O.; Ke, S.C. Origin of photoactivity of oxygen-deficient TiO2 under visible light. Appl. Phys. Lett. 2009, 95, 2007–2010. [Google Scholar] [CrossRef]
  154. Zhuang, J.; Weng, S.; Dai, W.; Liu, P.; Liu, Q. Effects of interface defects on charge transfer and photoinduced properties of TiO2 bilayer films. J. Phys. Chem. C 2012, 116, 25354–25361. [Google Scholar] [CrossRef]
  155. Ma, S.; Huang, S.D.; Fang, Y.H.; Liu, Z.P. TiH Hydride Formed on Amorphous Black Titania: Unprecedented Active Species for Photocatalytic Hydrogen Evolution. ACS Catal. 2018, 8, 9711–9721. [Google Scholar] [CrossRef]
  156. Wang, Y.; Sun, H.; Tan, S.; Feng, H.; Cheng, Z.; Zhao, J.; Zhao, A.; Wang, B.; Luo, Y.; Yang, J.; et al. Role of point defects on the reactivity of reconstructed anatase titanium dioxide (001) surface. Nat. Commun. 2013, 4, 1–8. [Google Scholar] [CrossRef]
  157. Wan, Z.; Huang, G.F.; Huang, W.Q.; Jiao, C.; Yan, X.G.; Yang, Z.M.; Zhang, Q. The enhanced photocatalytic activity of Ti3+ self-doped TiO2 by a reduction method. Mater. Lett. 2014, 122, 33–36. [Google Scholar] [CrossRef]
  158. Zhou, X.; Liu, N.; Schmuki, P. Photocatalysis with TiO2 Nanotubes: “Colorful” Reactivity and Designing Site-Specific Photocatalytic Centers into TiO2 Nanotubes. ACS Catal. 2017, 7, 3210–3235. [Google Scholar] [CrossRef]
  159. Choudhury, B.; Choudhury, A. Oxygen defect dependent variation of band gap, Urbach energy and luminescence property of anatase, anatase-rutile mixed phase and of rutile phases of TiO2 nanoparticles. Phys. E Low Dimens. Syst. Nanostruct. 2014, 56, 364–371. [Google Scholar] [CrossRef]
  160. Das, T.K.; Ilaiyaraja, P.; Mocherla, P.S.V.; Bhalerao, G.M.; Sudakar, C. Influence of surface disorder, oxygen defects and bandgap in TiO2 nanostructures on the photovoltaic properties of dye sensitized solar cells. Sol. Energy Mater. Sol. Cells 2016, 144, 194–209. [Google Scholar] [CrossRef]
  161. Pan, S.; Liu, X.; Guo, M.; Yu, S.F.; Huang, H.; Fan, H.; Li, G. Engineering the intermediate band states in amorphous Ti3+-doped TiO2 for hybrid dye-sensitized solar cell applications. J. Mater. Chem. A 2015, 3, 11437–11443. [Google Scholar] [CrossRef]
  162. Bharti, B.; Kumar, S.; Lee, H.-N.; Kumar, R. Formation of oxygen vacancies and Ti3+ state in TiO2 thin film and enhanced optical properties by air plasma treatment. Sci. Rep. 2016, 6, 32355. [Google Scholar] [CrossRef] [PubMed]
  163. Saputera, W.H.; Mul, G.; Hamdy, M.S. Ti3+-containing titania: Synthesis tactics and photocatalytic performance. Catal. Today 2015, 246, 60–66. [Google Scholar] [CrossRef]
  164. 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]
  165. Wei, W.; Yaru, N.; Chunhua, L.; Zhongzi, X. Hydrogenation of TiO2 nanosheets with exposed {001} facets for enhanced photocatalytc activity. RSC Adv. 2012, 2, 8286–8288. [Google Scholar] [CrossRef]
  166. Ali, A.; Ruzybayev, I.; Yassitepe, E.; Karim, A.; Shah, S.I.; Bhatti, A.S. Phase transformations in the pulsed laser deposition grown TiO2 thin films as a consequence of O2 partial pressure and Nd doping. J. Phys. Chem. C 2015, 119, 11578–11587. [Google Scholar] [CrossRef]
  167. Lepcha, A.; Maccato, C.; Mettenbörger, A.; Andreu, T.; Mayrhofer, L.; Walter, M.; Olthof, S.; Ruoko, T.P.; Klein, A.; Moseler, M.; et al. Electrospun Black Titania Nanofibers: Influence of Hydrogen Plasma-Induced Disorder on the Electronic Structure and Photoelectrochemical Performance. J. Phys. Chem. C 2015, 119, 18835–18842. [Google Scholar] [CrossRef]
  168. Huang, C.N.; Bow, J.S.; Zheng, Y.; Chen, S.Y.; Ho, N.J.; Shen, P. Nonstoichiometric titanium oxides via pulsed laser ablation in water. Nanoscale Res. Lett. 2010, 5, 972–985. [Google Scholar] [CrossRef] [PubMed]
  169. Forster, M.; Potter, R.J.; Ling, Y.; Yang, Y.; Klug, D.R.; Cowan, A.J. Oxygen deficient α-Fe2O3 photoelectrodes: A balance between enhanced electrical properties and trap-mediated losses. Chem. Sci. 2015, 6, 4009–4016. [Google Scholar] [CrossRef] [PubMed]
  170. Zhou, W.; Li, W.; Wang, J.; Qu, Y.; Yang, Y.; Xie, Y.; Zhang, K.; Wang, L.; Fu, H.; Zhao, D. Ordered Mesoporous Black TiO2 as Highly Efficient Hydrogen Evolution Photocatalyst. J. Am. Chem. Soc. 2014, 136, 9280–9283. [Google Scholar] [CrossRef]
  171. Zhou, H.; Zhang, Y. Enhancing the capacitance of TiO2 nanotube arrays by a facile cathodic reduction process. J. Power Sources 2013, 239, 128–131. [Google Scholar] [CrossRef]
  172. Lin, T. Effective nonmetal incorporation in black titania with enhanced solar energy utilization. Energy Environ. Sci. 2014, 6. [Google Scholar] [CrossRef]
  173. Cui, H.; Zhao, W.; Yang, C.; Yin, H.; Lin, T.; Shan, Y.; Xie, Y.; Gu, H.; Huang, F. Black TiO2 nanotube arrays for high-efficiency photoelectrochemical water-splitting. J. Mater. Chem. A 2014, 8612–8616. [Google Scholar] [CrossRef]
  174. Wang, Z.; Yang, C.; Lin, T.; Yin, H.; Chen, P.; Wan, D.; Xu, F.; Huang, F.; Lin, J.; Xie, X.; et al. H-doped black titania with very high solar absorption and excellent photocatalysis enhanced by localized surface plasmon resonance. Adv. Funct. Mater. 2013, 23, 5444–5450. [Google Scholar] [CrossRef]
  175. Pei, D.-N.; Gong, L.; Zhang, A.-Y.; Zhang, X.; Chen, J.-J.; Mu, Y.; Yu, H.-Q. Defective titanium dioxide single crystals exposed by high-energy {001} facets for efficient oxygen reduction. Nat. Commun. 2015, 6, 8696. [Google Scholar] [CrossRef] [Green Version]
  176. Bonanni, S.; Aït-Mansour, K.; Harbich, W.; Brune, H. Effect of the TiO2 reduction state on the catalytic CO oxidation on deposited size-selected Pt clusters. J. Am. Chem. Soc. 2012, 134, 3445–3450. [Google Scholar] [CrossRef] [PubMed]
  177. Pan, X.; Yang, M.; Fu, X.; Zhang, N.; Xu, Y. Defective TiO2 with oxygen vacancies: Synthesis, properties and photocatalytic applications. Nanoscale 2013, 5, 3601. [Google Scholar] [CrossRef]
  178. Tan, L.L.; Ong, W.J.; Chai, S.P.; Goh, B.T.; Mohamed, A.R. Visible-light-active oxygen-rich TiO2 decorated 2D graphene oxide with enhanced photocatalytic activity toward carbon dioxide reduction. Appl. Catal. B Environ. 2015, 179, 160–170. [Google Scholar] [CrossRef]
  179. Ramchiary, A.; Samdarshi, S.K. Hydrogenation based disorder-engineered visible active N-doped mixed phase titania. Sol. Energy Mater. Sol. Cells 2015, 134, 381–388. [Google Scholar] [CrossRef]
  180. Kumar, V.; Ntwaeaborwa, O.M.; Holsa, J.; Motaung, D.E.; Swart, H.C. The role of oxygen and titanium related defects on the emission of TiO2: Tb3+ nano-phosphor for blue lighting applications. Opt. Mater. 2015, 46, 510–516. [Google Scholar] [CrossRef]
  181. Kang, Q.; Cao, J.; Zhang, Y.; Liu, L.; Xu, H.; Ye, J. Reduced TiO2 nanotube arrays for photoelectrochemical water splitting. J. Mater. Chem. A 2013, 1, 5766–5774. [Google Scholar] [CrossRef]
  182. Su, J.; Zou, X.; Zou, Y.; Li, G.; Wang, P.; Chen, J. Porous Titania with Heavily Self-Doped Ti3+ for Specific Sensing of CO at Room Temperature. Inorg. Chem. 2013, 52, 5924–5930. [Google Scholar] [CrossRef] [PubMed]
  183. Zhao, A.; Masa, J.; Xia, W. Oxygen-deficient titania as alternative support for Pt catalysts for the oxygen reduction reaction. J. Energy Chem. 2014, 23, 701–707. [Google Scholar] [CrossRef]
  184. Zhou, H.; Zhang, Y. Electrochemically Self-Doped TiO2 Nanotube Arrays for Supercapacitors. J. Phys. Chem. C 2014, 118, 5626–5636. [Google Scholar] [CrossRef]
  185. Li, Y.; Weng, Y.; Zhang, J.; Ding, J.; Zhu, Y.; Wang, Q.; Yang, Y.; Cheng, Y.; Zhang, Q.; Li, P.; et al. Observation of superconductivity in structure-selected Ti2O3 thin films. NPG Asia Mater. 2018, 10, 522–532. [Google Scholar] [CrossRef]
  186. Wei, X.; Skomski, R.; Balamurugan, B.; Sun, Z.G.; Ducharme, S.; Sellmyer, D.J. Magnetism of TiO and TiO2 nanoclusters. J. Appl. Phys. 2009, 105. [Google Scholar] [CrossRef]
  187. Han, X.; Huang, J.; Jing, X.; Yang, D.; Lin, H.; Wang, Z.; Li, P.; Chen, Y. Oxygen-Deficient Black Titania for Synergistic/Enhanced Sonodynamic and Photoinduced Cancer Therapy at Near Infrared-II Biowindow. ACS Nano 2018, 12, 4545–4555. [Google Scholar] [CrossRef] [PubMed]
  188. Tan, L.L.; Ong, W.J.; Chai, S.P.; Mohamed, A.R. Band gap engineered, oxygen-rich TiO2 for visible light induced photocatalytic reduction of CO2. Chem. Commun. 2014, 50, 6923–6926. [Google Scholar] [CrossRef] [PubMed]
  189. Chong, S.V.; Kadowaki, K.; Xia, J.; Idriss, H. Interesting magnetic behavior from reduced titanium dioxide nanobelts. Appl. Phys. Lett. 2008, 92, 1–4. [Google Scholar] [CrossRef]
  190. Guillemot, F.; Porté, M.C.; Labrugère, C.; Baquey, C. Ti4+ to Ti3+ conversion of TiO2 uppermost layer by low-temperature vacuum annealing: Interest for titanium biomedical applications. J. Colloid Interface Sci. 2002, 255, 75–78. [Google Scholar] [CrossRef] [PubMed]
  191. Wu, Q.M.; Ruan, J.M.; Zhou, Z.C.; Sang, S.B. Magneli phase titanium sub-oxide conductive ceramic TinO2n−1 as support for electrocatalyst toward oxygen reduction reaction with high activity and stability. J. Cent. South Univ. 2015, 22, 1212–1219. [Google Scholar] [CrossRef]
  192. Li, X.; Zhu, A.L.; Qu, W.; Wang, H.; Hui, R.; Zhang, L.; Zhang, J. Magneli phase Ti4O7 electrode for oxygen reduction reaction and its implication for zinc-air rechargeable batteries. Electrochim. Acta 2010, 55, 5891–5898. [Google Scholar] [CrossRef]
  193. Yang, Y.; Zhang, T.; Le, L.; Ruan, X.; Fang, P.; Pan, C.; Xiong, R.; Shi, J.; Wei, J. Quick and facile preparation of visible light-driven TiO2 photocatalyst with high absorption and photocatalytic activity. Sci. Rep. 2014, 4, 7045. [Google Scholar] [CrossRef]
  194. Liu, Y.; Xing, M.; Zhang, J. Ti3+ and carbon co-doped TiO2 with improved visible light photocatalytic activity. Chin. J. Catal. 2014, 35, 1511–1519. [Google Scholar] [CrossRef]
  195. Shahid, M.; Choi, S.Y.; Liu, J.; Kwon, Y.U. Reduced titania films with ordered nanopores and their application to visible light water splitting. Bull. Korean Chem. Soc. 2013, 34, 2271–2275. [Google Scholar] [CrossRef]
  196. Ren, Y.; Li, J.; Yu, J. Enhanced electrochemical performance of TiO2 by Ti3+ doping using a facile solvothermal method as anode materials for lithium-ion batteries. Electrochim. Acta 2014, 138, 41–47. [Google Scholar] [CrossRef]
  197. Xia, T.; Zhang, W.; Wang, Z.; Zhang, Y.; Song, X.; Murowchick, J.; Battaglia, V.; Liu, G.; Chen., X. Amorphous carbon-coated TiO2 nanocrysals for improved lithium-ion battery and photocatalytic performance. Nano Energy 2014, 6, 109–118. [Google Scholar] [CrossRef]
  198. Liu, H.; Yang, W.; Ma, Y.; Yao, J. Extended visible light response of binary TiO2-Ti2O3 photocatalyst prepared by a photo-assisted sol-gel method. Appl. Catal. A Gen. 2006, 299, 218–223. [Google Scholar] [CrossRef]
  199. Yin, G.; Huang, X.; Chen, T.; Zhao, W.; Bi, Q.; Xu, J.; Han, Y.; Huang, F. Hydrogenated Blue Titania for Efficient Solar to Chemical Conversions: Preparation, Characterization, and Reaction Mechanism of CO2 Reduction. ACS Catal. 2018, 8, 1009–1017. [Google Scholar] [CrossRef]
  200. Yeh, Y.C.; Li, S.S.; Wu, C.C.; Shao, T.W.; Kuo, P.C.; Chen, C.W. Stoichiometric dependence of TiOx as a cathode modifier on band alignment of polymer solar cells. Sol. Energy Mater. Sol. Cells 2014, 125, 233–238. [Google Scholar] [CrossRef]
  201. Pan, S.S.; Lu, W.; Zhao, Y.H.; Tong, W.; Li, M.; Jin, L.M.; Choi, J.Y.; Qi, F.; Chen, S.G.; Fei, L.F.; et al. Self-doped rutile titania with high performance for direct and ultrafast assay of H2O2. ACS Appl. Mater. Interfaces 2013, 5, 12784–12788. [Google Scholar] [CrossRef]
Figure 1. Phase diagram of the Ti-O system. The region Ti2O3-TiO2 contains Ti3O5, Ti2O3 and seven discrete phases of the homologous series TinO2n−1 (Magneli phase) and TiO2. Reproduced from [35]; Copyright 2003, Elsevier.
Figure 1. Phase diagram of the Ti-O system. The region Ti2O3-TiO2 contains Ti3O5, Ti2O3 and seven discrete phases of the homologous series TinO2n−1 (Magneli phase) and TiO2. Reproduced from [35]; Copyright 2003, Elsevier.
Catalysts 08 00601 g001
Figure 2. Schematic illustration and description of F center formation in Titania. Reproduced from [52], Copyright 2009, American Chemical Society.
Figure 2. Schematic illustration and description of F center formation in Titania. Reproduced from [52], Copyright 2009, American Chemical Society.
Catalysts 08 00601 g002
Figure 3. Electron-hole formation mechanism.
Figure 3. Electron-hole formation mechanism.
Catalysts 08 00601 g003
Figure 4. (a) DFT—calculated defect formation energies in Titania as a function of fermi energy (EF) under Ti-rich and (b) O-rich growth conditions. Reproduced from [56], Copyright 2006, The American Physical Society. (c,d) Variation in defect concentration under different oxygen pressure conditions [57,58]. Reproduced from [57], Copyright 2008, American Chemical Society.
Figure 4. (a) DFT—calculated defect formation energies in Titania as a function of fermi energy (EF) under Ti-rich and (b) O-rich growth conditions. Reproduced from [56], Copyright 2006, The American Physical Society. (c,d) Variation in defect concentration under different oxygen pressure conditions [57,58]. Reproduced from [57], Copyright 2008, American Chemical Society.
Catalysts 08 00601 g004
Figure 5. Different types of octahedral packing (a) no distortion (cubic TiO), (b) tetragonal distortion(rutile) (c) trigonal distortion (Ti2O3) (d) orthorhombic- like distortion (Ti4O7 and Ti5O) Reproduced from [90], Copyright 2007, American Mineralogist.
Figure 5. Different types of octahedral packing (a) no distortion (cubic TiO), (b) tetragonal distortion(rutile) (c) trigonal distortion (Ti2O3) (d) orthorhombic- like distortion (Ti4O7 and Ti5O) Reproduced from [90], Copyright 2007, American Mineralogist.
Catalysts 08 00601 g005
Figure 6. Crystal structures of various titanium oxides.
Figure 6. Crystal structures of various titanium oxides.
Catalysts 08 00601 g006
Figure 7. (a) Difference between TiO2 and Ti4O7 (Magneli phase) structure (b) Different orientations of TiO2 octahedra in Magneli phase materials. Reproduced from [92], Copyright 2010, Elsevier.
Figure 7. (a) Difference between TiO2 and Ti4O7 (Magneli phase) structure (b) Different orientations of TiO2 octahedra in Magneli phase materials. Reproduced from [92], Copyright 2010, Elsevier.
Catalysts 08 00601 g007
Figure 8. Enhancement of visible absorption with increasing oxygen vacancy.
Figure 8. Enhancement of visible absorption with increasing oxygen vacancy.
Catalysts 08 00601 g008
Figure 9. Molecular orbital diagram of TiO2, Ti2O3, TiO. Reproduced from [90], Copyright 2007, American Mineralogist.
Figure 9. Molecular orbital diagram of TiO2, Ti2O3, TiO. Reproduced from [90], Copyright 2007, American Mineralogist.
Catalysts 08 00601 g009
Figure 10. Splitting of t2g and, eg levels in tetragonal (a), trigonal (b), and orthorhombic (c) distorted environments. Reproduced from [90], Copyright 2007, American Mineralogist.
Figure 10. Splitting of t2g and, eg levels in tetragonal (a), trigonal (b), and orthorhombic (c) distorted environments. Reproduced from [90], Copyright 2007, American Mineralogist.
Catalysts 08 00601 g010
Figure 11. Various defect levels of Titania. The numbered letters a1, a2, a3, a4, b1, b2, c1, c2, c3, c4, d1 and d2 represent the defect levels VTix/ V Ti , V Ti / V Ti , V Ti   /   V Ti   , V Ti   /   V Ti     , O o /   O i , O o x /   O i , Ti i   /   Ti i , Ti i   /   Ti i   , Ti i /   Ti i   , Ti i x /   Ti i , V o / V 0   and Vx/V, respectively [86]. Reproduced from [86], Copyright 2011, Science China Press and Springer-Verlag Berlin.
Figure 11. Various defect levels of Titania. The numbered letters a1, a2, a3, a4, b1, b2, c1, c2, c3, c4, d1 and d2 represent the defect levels VTix/ V Ti , V Ti / V Ti , V Ti   /   V Ti   , V Ti   /   V Ti     , O o /   O i , O o x /   O i , Ti i   /   Ti i , Ti i   /   Ti i   , Ti i /   Ti i   , Ti i x /   Ti i , V o / V 0   and Vx/V, respectively [86]. Reproduced from [86], Copyright 2011, Science China Press and Springer-Verlag Berlin.
Catalysts 08 00601 g011
Figure 12. Schematic representation of the structural model for defect states in reduced Titania with their characteristic optical behavior and preparation methods. White, grayish and black arrows represent heat-treatments under hydrogen, inert and oxygen atmosphere, respectively. Reproduced from [127], Copyright 2004, The Physical Society of Japan.
Figure 12. Schematic representation of the structural model for defect states in reduced Titania with their characteristic optical behavior and preparation methods. White, grayish and black arrows represent heat-treatments under hydrogen, inert and oxygen atmosphere, respectively. Reproduced from [127], Copyright 2004, The Physical Society of Japan.
Catalysts 08 00601 g012
Figure 13. (a) Magneli shear planes providing access for electrons transport. Reproduced from [130], Copyright 2014, Royal Society of Chemistry. (b) Relative electrical conductivities of Magneli phase titanium sub-oxides. Reproduced from [92], Copyright 2010, Elsevier.
Figure 13. (a) Magneli shear planes providing access for electrons transport. Reproduced from [130], Copyright 2014, Royal Society of Chemistry. (b) Relative electrical conductivities of Magneli phase titanium sub-oxides. Reproduced from [92], Copyright 2010, Elsevier.
Catalysts 08 00601 g013
Figure 14. Effect of oxygen activity on the electrical conductivity of Titania. Reproduced from [49], Copyright 2007, John Wiley and Sons.
Figure 14. Effect of oxygen activity on the electrical conductivity of Titania. Reproduced from [49], Copyright 2007, John Wiley and Sons.
Catalysts 08 00601 g014
Figure 15. Influence of point defects in reduced Titania on their electronic structure. Reproduced from [144], Copyright 2013, American Chemical Society.
Figure 15. Influence of point defects in reduced Titania on their electronic structure. Reproduced from [144], Copyright 2013, American Chemical Society.
Catalysts 08 00601 g015
Figure 16. Reactive oxygen species generated in the photocatalytic oxidation steps of water. Reproduced from [146], Copyright 2017, American Chemical Society.
Figure 16. Reactive oxygen species generated in the photocatalytic oxidation steps of water. Reproduced from [146], Copyright 2017, American Chemical Society.
Catalysts 08 00601 g016
Figure 17. Photoelectrochemical watersplitting reaction mechanism in Titania. Reproduced from [146], Copyright 2007, American Chemical Society.
Figure 17. Photoelectrochemical watersplitting reaction mechanism in Titania. Reproduced from [146], Copyright 2007, American Chemical Society.
Catalysts 08 00601 g017
Figure 18. Surface reactivity of reduced Titania under visible light illumination. Reproduced from [147], Copyright 2010, American Chemical Society.
Figure 18. Surface reactivity of reduced Titania under visible light illumination. Reproduced from [147], Copyright 2010, American Chemical Society.
Catalysts 08 00601 g018
Table 1. Behavioral difference of defects in the bulk and the surface [46].
Table 1. Behavioral difference of defects in the bulk and the surface [46].
BulkSurface
Interstitial atomAdatom
VacancyVacancy
Interstitial ClusterAdatom island
Vacancy ClusterVacancy island
Kick-in/Kick-outExhange diffusion
Vacancy-interstitial formationVacancy-adatom formation
Table 2. Kroger-Vink notation and their meaning [47].
Table 2. Kroger-Vink notation and their meaning [47].
Kroger-Vink NotationMeaning
Ti i   Ti3+ ion in the interstitial site
Ti i i   Ti4+ ion in the interstitial site
V o Oxygen vacancy
O i O2− ion in the interstitial site
V Ti Titanium vacancy
Table 3. Variation of the crystal structure with O/Ti stoichiometry.
Table 3. Variation of the crystal structure with O/Ti stoichiometry.
CompoundX in TiOxStructure
TiO22Rutile
Ti10O191.9Anatase
Ti9O171.89Triclinic
Ti8O151.875Triclinic
Ti7O131.857Triclinic
Ti6O111.833Triclinic
Ti5O91.8Triclinic
Ti4O71.75Triclinic
γTi3O51.67Monoclinic
Ti2O31.5Tetragonal
TiO1.0Hexagonal
Cubic Monoclinic
Ti2O0.5Hexagonal
Ti0Hexagonal
Table 4. Predominant defects and conductivity of Titania depending on the oxygen activity. Reproduced from [131], Copyright 1996, Elsevier.
Table 4. Predominant defects and conductivity of Titania depending on the oxygen activity. Reproduced from [131], Copyright 1996, Elsevier.
mPredominant DefectsConductivity Type
6 V 0 , en
4 V 0 ,   e n
5 Ti i 4 + , Ti i 3 + , e’n
hp
5 V Ti 4 , hp

Share and Cite

MDPI and ACS Style

Jayashree, S.; Ashokkumar, M. Switchable Intrinsic Defect Chemistry of Titania for Catalytic Applications. Catalysts 2018, 8, 601. https://doi.org/10.3390/catal8120601

AMA Style

Jayashree S, Ashokkumar M. Switchable Intrinsic Defect Chemistry of Titania for Catalytic Applications. Catalysts. 2018; 8(12):601. https://doi.org/10.3390/catal8120601

Chicago/Turabian Style

Jayashree, Swaminathan, and Meiyazhagan Ashokkumar. 2018. "Switchable Intrinsic Defect Chemistry of Titania for Catalytic Applications" Catalysts 8, no. 12: 601. https://doi.org/10.3390/catal8120601

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