Next Article in Journal
Methane Conversion over C2N-Supported Fe2 Dimers
Next Article in Special Issue
Enhanced Visible and Ultraviolet Light-Induced Gas-Phase Photocatalytic Activity of TiO2 Thin Films Modified by Increased Amount of Acetylacetone in Precursor Solution for Spray Pyrolysis
Previous Article in Journal
Asymmetric Friedel–Crafts Alkylation of Indoles Catalyzed by Chiral Aziridine-Phosphines
Previous Article in Special Issue
Application of Novel C-TiO2-CFA/PAN Photocatalytic Membranes in the Removal of Textile Dyes in Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of First-Principles Studies of TiO2: Nanocluster, Bulk, and Material Interface

1
Department of NanoEngineering, University of California San Diego, 9500 Gilman Drive, Mail Code 0448, La Jolla, CA 92093-0448, USA
2
School of Physics, State Key Laboratory of Crystal Materials, Shandong University, Jinan 250100, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(9), 972; https://doi.org/10.3390/catal10090972
Submission received: 31 July 2020 / Revised: 26 August 2020 / Accepted: 26 August 2020 / Published: 28 August 2020
(This article belongs to the Special Issue TiO2 for Photocatalytic Applications)

Abstract

:
TiO2 has extensive applications in the fields of renewable energy and environmental protections such as being used as photocatalysts or electron transport layers in solar cells. To achieve highly efficient photocatalytic and photovoltaic applications, ongoing efforts are being devoted to developing novel TiO2-based material structures or compositions, in which a first-principles computational approach is playing an increasing role. In this review article, we discuss recent computational and theoretical studies of structural, energetic, electronic, and optical properties of TiO2-based nanocluster, bulk, and material interface for photocatalytic and photovoltaic applications. We conclude the review with a discussion of future research directions in the field.

1. Introduction

Titanium dioxide (TiO2) is an ideal semiconductor photocatalyst because of its excellent properties (e.g., high activity, good stability, nontoxicity, and low cost), which has many promising applications in the fields of renewable energy and environmental protections [1,2,3,4,5]. However, the larger bandgap of TiO2 (~3.2 eV for anatase and ~3.0 eV for rutile) makes it inefficient for visible light to excite the electron–hole pairs, which are necessary to initiate a photocatalytic process [2,3,4]. Therefore, the photocatalytic applications of TiO2 in the visible light range are heavily limited. To improve the spectra response and photocatalytic activity of TiO2 in the whole solar spectra, numerous efforts have been carried out [2,3,4,6,7,8]. For instance, in recent years, a number of attempts have been made to improve the visible light absorption of TiO2 by nonmetal doping, which either introduces some impurity states in the bandgap or modifies the fundamental bandgap of TiO2, and promote the photocatalytic activity of TiO2 to some degree [4,6,7,8,9,10]. To achieve high photocatalytic efficiency, in addition to increasing the number of photoinduced electron–hole pairs, another approach is to increase the separation rate of photoinduced electron–hole pairs in TiO2. To do so, increasing efforts are being made to synthesize TiO2-based materials heterostructures and/or composites that utilize the interfacial charge transfer mechanism to promote the separation rate of electron–hole pairs [11,12,13,14,15]. A high charge separation rate is beneficial not only for photocatalysts, but also for photovoltaic cells in which TiO2 is often used as electron transport layers as they both use the separated electrons and holes [16,17,18,19,20,21,22,23,24].
In this review article, we review and discuss recent computational and theoretical progress on the TiO2 nanoclusters, bulk, and material interfaces from the viewpoints of first-principles calculations. This review is divided into three sections. First, we review the energetic stability and electronic properties of TiO2 nanoclusters. Second, we focus on the nonmetal doping and co-doping effects on the electronic structure and optical absorption of TiO2 along with the relationship between the doping preference of nonmetal atoms and their electronegativities as well as the growth conditions of doped TiO2. Next, we discuss TiO2 material interfaces for photocatalytic and photovoltaic applications from first-principles computational perspective. Finally, we conclude this article with several exciting future research directions.

2. Nanocluster

TiO2 nanoclusters are one class of extensively studied TiO2 nanostructures for photocatalytic applications because of their high surface to volume ratio [25,26]. Therefore, TiO2 clusters serve as ideal models to study catalytic properties of nanostructured TiO2 for pollutants degradation in the first-principles calculations. Qu and co-workers have studied the structural and electronic properties of (TiO2)n (n = 1–16) nanoclusters in size dependence using first-principles calculations [27,28,29]. In spite of these earlier studies, to develop a full understanding of the catalytic mechanism of TiO2 nanoclusters, the first step is generally to choose appropriate structural models with relatively high stability and sizes [25,26,30,31]. For example, in 2016, Arab et al. evaluated the structural stability of (TiO2)n (n = 1–10) nanoclusters using first-principles approach by analyzing their energetic properties including binding energy, second-order energy difference, and fragmentation energy [31]. The definition of these energetic terms and discussion on their results are below.
(i) The binding energy per atom is defined as E b = [ n E (Ti) + m E (O) −E(Ti n O m )]/(n + m), where E(Ti), E(O), and E(Ti n O m ) are the energies of Ti, O, and T i n O m species, respectively. The calculated E b increases continuously as cluster size increases, a common phenomenon for nanoclusters [32].
(ii) The second-order energy difference is defined as Δ 2 E n = E(TiO2) n 1 + E(TiO2) n + 1 − 2E(TiO2) n , where E(TiO2) n , E(TiO2) n 1 , and E(TiO2) n + 1 are the total energies of (TiO2) n , (TiO2) n 1 , and (TiO2) n + 1 clusters, respectively. The calculated second-order energy difference again cluster sizes show maximum values of Δ 2 E n at n = 3, 5, and 7, indicating that these clusters are energetically more favorable than their neighboring clusters. Interestingly, the calculated variation of the second-order energy difference shows an odd-even effect as a function of cluster size [32].
(iii) The fragmentation energy is defined as E f = E(TiO2) + E(TiO2) n 1 E(TiO2) n . The calculated fragmentation energy as a function of cluster size shows again that the (TiO2)3, (TiO2)5, and (TiO2)7 nanoclusters are more stable than their neighboring clusters, as the calculated E f of these clusters is higher than neighboring clusters.
This work presents an effective approach to evaluate the relative stability of nanomaterials beyond TiO2 nanoclusters.
Hernandez et al. studied structural and electronic properties of larger (TiO2)n (n = 15–20) clusters using particle swarm optimization and first-principles calculations [30]. At odd n = 15, 17, and 19, these clusters tend to form less stable structures with ionic bond character and relatively low binding energy; at even n = 16, 18, and 20, they tend to form more stable structures with large covalent nature and larger binding energy. Interestingly, the (TiO2)18 cluster is the most stable structure compared to its neighbors and shows the highest value of the electronic energy gap, which is thus likely to be formed via some experimental approach. Additionally, TiO6 embryo is found to be the common precursor of growth for TiO2 clusters with high symmetry.
In 2020, Bai et al. carried out first-principle studies on the adsorption mechanism of dimethylbenzoquinones (2,5-DMBQ) onto TiO2 clusters (n = 1–6) [26]. The adsorption ability of 2,5-DMBQ onto the TiO2 clusters were evaluated from the calculated adsorption energy and changes of Gibbs free energy. The authors found two absorption patterns by searching for the most stable structural configurations: (i) adsorption of the quinone ring on to TiO2 cluster with the bonding interaction between the C atoms of quinone and Ti atoms and between H atoms of methyl and O atoms, see Figure 1a, and (ii) ketonic O of the 2,5-DMBQ absorbed on the Ti atoms of TiO2 clusters with the bonding interactions between the O atoms of TiO2 and H atoms of methyl, see Figure 1b.
By comparing the two absorption patterns (i, ii), the authors found that all these adsorptions are spontaneous and the pattern (ii) is energetically more favorable than the pattern (i) because of more negative adsorption free energy. More interestingly, in both patterns, the (TiO2)3 cluster has the most negative adsorption energy among all the adsorption configurations, i.e., approximately −28.44 kcal/mol for pattern (i) and −48.59 kcal/mol for pattern (ii). It is coincidental that, in another independent work, a (TiO2)3 cluster with the same size was also proposed as one ideal model to study the fundamental mechanism of TiO2 photocatalyst, as discussed below. Note that the calculated adsorption energy is much more negative than the typical values of physorption energies (−1.20 to 9.56 kcal/mol) [33], indicating the chemisorption nature of the two absorption patterns. In addition, the water environment is also found unfavorable for adsorption because gaseous phase has a more negative adsorption energy. This work indicates that the fundamental photocatalytic mechanism of TiO2 nanoclusters can be well understood from first-principles density functional theory (DFT) calculations. This approach can be applied to evaluate the photocatalytic performance of TiO2 nanoclusters on other types of pollutant molecules.
In 2017, Gan et al. reported a first-principles computational study on the mechanism of photo-selective catalytic reduction of 4-bromobenzaldehyde (4-BBA) in different solvents using OH-defected TiO2 cluster model [34]. In this work, a neutral cluster model of Ti3O9H6 was used to represent the reaction sites with photocatalytic activity on TiO2 surfaces. The authors selected the Ti3O9H6 cluster based on three considerations: (i) the dangling bonds of TiO2 clusters can be saturated by replacing the Ti-O bonds with the Ti-OH bonds without altering Ti4+ oxidation state; (ii) the chemical environment of each Ti atom is equivalent and the small cluster size also saves computational cost; and (iii) the surface species such as OH defects on the TiO2 surface were already involved in the Ti3O9H6 cluster. This work shows that the Ti3O9H6 cluster is an ideal molecular model to study photocatalytic reactions on the TiO2 surface, which servers as an important reference in future work.
TiO2 photocatalysts are usually nanosized materials with highly hydrated surfaces. Hydrogen is thus one of the most common impurities and hydroxyl groups of different nature often appear on TiO2 surfaces. The size-dependent evolution of the chemical reactivity and site selectivity of (TiO2)n nanoclusters (n = 5–10) toward hydrogen peroxide were explored from first-principles computational studies by Mohajeri et al. [25]. The adsorption energy of H2O2 onto the TiO2 nanoclusters is found to have a converging trend at n ≥ 8. After adsorption, the H2O2 can be easily decomposed into two OH radicals that can be stabilized through forming bonding interactions with Ti and O atoms on the cluster surface.
The O2 molecule adsorption on the TiO2 nanoparticles was also modeled from first-principles DFT calculations [35]. The authors studied two possible adsorption modes: The first one is that O2 was adsorbed on TiO2 nanoclusters via the interaction between H and O atoms. This reduction process of TiO2 nanoparticles have three following features: (i) forming stable OH group, (ii) low activation energy, and (iii) introducing Ti3+ ion related states within the energy gap. Figure 2 shows the structural model, electronic structure diagram, and spin density surface of Ti8O16 cluster. The second one is that O2 was directly adsorbed on each Ti3+ ion of the reduced Ti8O16H cluster. This process shows three features: (i) formation of stable O2- species, (ii) no any energetic barrier, and (iii) no Ti3+ defect states within the energy gap. These calculated results are generally consistent with the experimental phenomena of H plasma reduction and interaction between reduced TiO2 and O2.

3. Bulk

In 2001, Asahi et al. reported the visible light photocatalytic activity of nitrogen-doped TiO2 [9,36]. Soon after, various nonmetal-doped (including B, C, Si, P, S, and halogen elements) TiO2 were studied to explore their photocatalytic performance under visible light. Although most of these nonmetal-doped TiO2 show the visible light optical absorption and photocatalytic activity in some degree, there are some “tricky” fundamental issues in their experimental optical absorption spectra and visible light photocatalytic mechanism. For example, what leads to the controversy on the origin of the visible light absorption in N-doped TiO2? Why can B-doped TiO2 show both redshift and blueshift of the optical absorption edge? What is responsible for the different optical absorption thresholds in C-doped TiO2? Which form (anion or cation) of the Si (P and S) ion in TiO2 is more effective to promote the visible light absorption? Can F-doping lead to the visible light absorption of TiO2? Why can Cl- and Br-doped TiO2 show the stronger photocatalytic ability than that of undoped TiO2? What is the difference on the mechanism of the visible light absorption between the I-anion and I-cation doped TiO2? To answer these questions, first-principle theoretical calculation is an effective approach and it has successfully explained many experimental phenomena, including the issues mentioned above. In this section, we review the recent theoretical progress in understanding the electronic structure and optical absorption of nonmetal-doped TiO2 from first-principles calculations. A brief summary of commonly used nonmetal dopants and the resulting properties of doped TiO2 is shown in Table 1.

3.1. Formation Energy

First-principles DFT calculations for the relative stability of nonmetal-doped TiO2 can help us understand the formation of the doped structures and provide useful guidance to prepare samples [37,38,39,40,41,42,43,44,45]. First, let us focus on the structural stability of substitutional doping models. In principle, there are two possible substitutional doping ways for X-doped TiO2, i.e., an X anion at an O site (X@O) or an X cation at a Ti site (X@Ti). The defect formation energy required for X substituting for either O or Ti in TiO2 could be calculated from the following formulas, respectively [43,44,45].
E f X = E X d o p e d E u n d o p e d μ X + μ O
E f X = E X d o p e d E u n d o p e d μ X + μ T i
E X d o p e d is the total energy of X-doped TiO2 and E u n d o p e d is the total energy of undoped TiO2. μ X is the chemical potential of dopant X, and μ O ( μ T i ) is the chemical potential of the O (Ti). The chemical potentials of Ti and O depend on whether TiO2 is grown under an O-rich or Ti-rich growth condition. Under the Ti-rich condition, the Ti chemical potential can be assumed as the energy of bulk Ti, while the O chemical potential can be obtained by the growth condition:
μ T i + 2 μ O = μ T i O 2
Under the O-rich condition, the chemical potential of O can be calculated from the ground-state energy of O2 molecule, while the chemical potential of Ti is then fixed by condition (3). Therefore, a link between the defect formation energy and the external growth condition of doped TiO2 can be created.
The relationship between the formation energies of substitutional nonmetal-doped TiO2 and its growth condition has been studied systematically [38,42,43,44,45]. For Si-doped TiO2, the formation energy of the substitutional Si-cation-doped model is much less than that of the substitutional Si-anion-doped model under both Ti-rich and O-rich growth conditions [44,46,47]. This indicates that Si is energetically more favorable to substitute Ti than O under both Ti-rich and O-rich growth conditions. For S- and P-doped TiO2, the doping sites of S and P strongly depend on the preparing method and growth condition of the doped TiO2 [43,48,49]. Under O-rich growth condition, S (P) prefers to replace Ti and form substitutional S (P)-cation doped structure. On the contrary, under the Ti-rich growth condition, S (P) prefers to replace O and form S (P)-anion doped structure. This is consistent with the experiments reported by several independent groups [50,51,52,53]. Yu et al. and Ohno et al. used the titanium isopropoxide and thiourea as the titanium and sulfur original materials, respectively, which corresponds to the O-rich growth condition, and prepared the S-cation -doped TiO2 [50,51,52]. Moreover, it is further confirmed that replacing Ti by S is energetically more favorable than replacing O under the O-rich growth condition [50]. In contrast, Umebayashi et al. used TiS2 as the starting material, which corresponds to the Ti-rich growth condition, and prepared S-anion -doped TiO2 [53,54]. Therefore, the first-principles theoretical calculations demonstrated a basic experimental fact that the ionic form and site of S (P) dopant in TiO2 can be controlled by the growth condition and preparation method of doped TiO2 sample [43,48,49]. For C-doped TiO2, Di Valentin et al.’s theoretical calculations also gave a similar conclusion [38]. In addition, the first-principles theoretical calculations for halogen-doped TiO2 further revealed the relationship between the doping sites of halogen atoms and their electronegativities as well as the growth condition of doped TiO2 [45,55]. The following conclusions were drawn [45].
(i) Substitutional X-anion-doped TiO2 (X = F, Cl, Br, and I) is energetically preferred to form under the Ti-rich rather than under the O-rich growth condition, and the formation energy increases in the order F < Cl < Br < I. This indicates that it is more difficult to replace an O atom using a larger and less electronegative X atom.
(ii) Substitutional X-cation-doped TiO2 (X = F, Cl, Br, and I) is energetically preferred to form under the O-rich rather than under the Ti-rich growth condition, and the formation energy increases in the order I < Br < Cl < F. This indicates that it is more difficult to replace a Ti atom using a smaller and more electronegative X atom.
(iii) Under O-rich growth conditions, it is energetically more favorable to substitute Ti than O using Br and I, while it is energetically more favorable to substitute O than Ti using F and Cl.
(iv) Under Ti-rich growth conditions, it is energetically more favorable to substitute O than Ti using all the X atoms (X = F, Cl, Br, and I).
To qualitatively show the relationship between the formation energy and the growth condition of doped TiO2 as well as the electronegativities of dopants, we plot the function of the formation energies of nonmetal-doped TiO2 as the oxygen chemical potential (corresponding to the growth condition of TiO2), see Figure 3. The following conclusions can be obtained.
(i) For high electronegative F (Cl and N), it is energetically more favorable to substitute O than Ti under both O-rich and Ti-rich conditions.
(ii) For low electronegative Si (Ge), it is energetically more favorable to substitute Ti than O under both O-rich and Ti-rich conditions.
(iii) For other nonmetal main group elements X- (X = B, C, S, Se, Te, P, As, Sb, Br, and I) doped TiO2, substitutional X-anion-doped TiO2 is energetically preferred to form under Ti-rich condition while substitutional X-cation-doped TiO2 is energetically preferred to form under O-rich conditions. This discrepancy can be partially understood from the electronegativity difference between nonmetal dopants and O (Ti) as well as the bond strength of X-O and X-Ti bonds. For high electronegative F (Cl and N), it is preferred to form a F-Ti (Cl-Ti and N-Ti) bond by picking up electrons from Ti rather than form a F-O (Cl-O and N-O) bond by losing electrons. In addition to the bulk doping, a recent computational study shows that F surface adsorption could tune the relative stability of (001) and (101) surfaces of anatase TiO2, thus providing one effective approach to synthesize TiO2 surfaces with desired properties [56,57,58,59]. For low electronegative Si (Ge), it is preferred to form a Si-O (Ge-O) bond by losing electrons to O rather than form a Si-Ti (Ge-Ti) bond by picking up electrons. For other nonmetal main group elements X (X = B, C, S, Se, Te, P, As, Sb, Br, and I), their electronegativities are between Ge and N, and thus both X-O and X-Ti bonds are possible to form, depending on the growth conditions of doped TiO2. These results are expected to provide some useful guidance to prepare nonmetal-doped TiO2 and other semiconductor oxides.
Next, let us discuss the possibility of interstitial nonmetal-doped TiO2. Generally speaking, only the dopant with a small atomic size can easily form interstitially doped structure. Therefore, B, C, and N are most possible to be located at the interstitial site. First-principles theoretical calculations show that substitutional N-anion and interstitial N-doped structures have nearly the same formation energy under O-rich growth condition [40]. Di Valentin et al. found that interstitial C-doped TiO2 even has lower formation energy than substitutional C-anion-doped TiO2 under O-rich growth condition [60]. Geng et al. reported that interstitial B-doped TiO2 has slightly lower formation energy than substitutional B-anion-doped TiO2 [61]. Finazzi et al. also suggested that the substitutional-anion doped B dopants can be converted into interstitial B after annealing at high temperature on the basis of the first-principles analysis [62]. In summary, for the nonmetal dopants with small atomic size, both the substitutional (at O site) and interstitial structures are possible to form.
Table 1. Summary of nonmetal dopants and the resulting properties of doped TiO2. X@Int (X = B and H) denotes that X dopant at an interstitial site.
Table 1. Summary of nonmetal dopants and the resulting properties of doped TiO2. X@Int (X = B and H) denotes that X dopant at an interstitial site.
DopantDoping SitePropertiesRef.
NN@OVisible light absorption due to either narrowed bandgap at a high doping
level (≥~4.2 at.%) or gap states at a low doping level (≤~2.1 at.%)
[9,40,63]
SS@ORedshift of optical absorption edge due to either narrowed bandgap
or gap states, depending on doping levels, as in the case of N@O doping.
[43]
PP@OVisible light absorption due to reduced optical band gap.[43]
P@TiUnchanged band gap.
BB@ORedshift due to gap states.[64]
B@IntBlueshift due to Moss–Burstein shift.
CC@ODifferent optical absorption thresholds due to discrete gap states.[65]
C@TiVisible light absorption due to narrowed band gap;
forming O=C double bond.
SiSi@TiVisible light optical absorption due to narrowed bandgap by 0.25 eV.[44]
FF@ONearly unchanged band gap.[45]
ClCl@OBandgap reduces by 0.2 eV; reduced oxidation and reduction ability
due to band-edge shift.
[45]
BrBr@OBandgap reduces about 0.3 eV; CBM shifts downwards by 0.3 eV
and VBM keeps unchanged.
[45]
II@Tin-type conductivity, visible light photocatalytic activity due to gap states.[45]
HH@Intn-type conductivity due to interstitial H doping.[66,67,68,69]

3.2. Nonmetal Doping

N Doping

Since the discovery that N doping can promote the visible light photocatalytic activity of TiO2 [9,36], numerous experimental efforts have been made to study the N doping influence on the optical absorption and photocatalytic properties of TiO2 [39,70,71,72,73,74,75,76,77,78,79,80,81,82,83]. With the progress of the theoretical and experimental research, the controversy on the origin of the redshift of the optical absorption edge in N-doped TiO2 appears [7,42,84]. Three different opinions have been proposed to explain the redshift of the optical absorption edge in N-doped TiO2:
(i) The mixing of N 2p states and valence band (VB) leads to the bandgap narrowing. A representative example is Asahi et al.’s theoretical calculation and experiment [9,36], and it is further confirmed by the later experiments [76,78,85].
(ii) N doping introduces isolated impurity states in the bandgap. This is widely confirmed by the experiments [70,71,77,78] and the first-principles theoretical calculations [40,63,86,87,88,89].
(iii) N doping creates oxygen vacancies in TiO2, which introduces defect states in the bandgap, and this leads to the visible light absorption [83,90,91,92].
In addition, several independent research groups reported that there exists an optimal nitrogen concentration to achieve the highest photocatalytic activity [78,81] or maximize visible light absorption [85]. It is also found that a bandgap narrowing can be realized in high-concentration N-doped TiO2 [78,85]. As a consequence, it is speculated that the N-doped TiO2 may show different electronic structure characteristics under different doping concentration, which may lead to the controversy on the origin of the visible light absorption. This speculation was confirmed through the first-principles electronic structure calculations [40].
Yang et al. analyzed electronic structures of N-doped anatase and rutile TiO2 at different doping levels and found following conclusions [42].
(i) At lower doping concentration (≤~2.1 at.%), N doping introduces some isolated impurity states above the VB in the bandgap. They act as transition levels, and thus the electronic transition among the VB, conduction band (CB), and impurity states may be responsible for the redshift of the optical absorption edge [63,86,87,88].
(ii) At higher doping levels (≥~4.2 at.%), more N 2p states are introduced, and they mix with the O 2p states, leading to the bandgap narrowing [9].
Zhao et al. also found similar conclusion from first-principles calculations [84]. Therefore, these results can clarify the controversy on the origin of the redshift of the optical absorption edge in N-doped TiO2, i.e., bandgap narrowing and N 2p states in the bandgap can be both responsible for the visible light absorption of N-doped TiO2, which depends on the N doping levels. It is noted that the same conclusion can be drawn using spin-polarized calculations despite some differences of the positions of N 2p states, which is caused by spin split. It also shows that when nitrogen concentration exceeds ~2.1 at.%, the optical energy gap has little further narrowing compared with that at lower doping levels. In contrast, increasing the nitrogen concentration leads to larger formation energy [42]. This is well consistent with the experiment, in which the higher nitrogen concentration does not lead to further optical energy-gap narrowing but makes the growth of N-doped TiO2 more difficult [85]. In addition, nonmetal doping (C, S, and P) can lower the formation energy of oxygen vacancy [38,93,94,95], and thus the combined effects of the oxygen vacancy and nonmetal doping cannot be excluded to be responsible for the visible light absorption in N- (C-, S-, and P-) doped TiO2.
TiO2 has three kinds of crystal phases: anatase, rutile, and brookite. Anatase and rutile are two common phases, and thus they are also generally considered to share similar electronic properties despite the different bandgap (3.2 eV for anatase phase [96] and 3.0 eV for rutile phase [97]). It is known that N doping in TiO2 can induce a redshift of the optical absorption edge [9,36,39,70,71,72,73,74,75,76,77,78,79,80,81,82,83]. Surprisingly, Diwald et al. observed a blueshift of the optical absorption edge ~0.2 eV in N-implanted rutile TiO2 [98]. Soon after, Di Valentin et al. proposed that the bandgap increasing of approximately 0.08 eV in N-doped rutile TiO2 can be responsible for the experimentally observed blueshift on the basis of the first-principles calculations [86]. However, this explanation is not in harmony with a basic experimental fact that the redshift is also observed in N-doped rutile TiO2 [36,72,74,77,82]. In principle, there are three possible N-doping sites, including substitutional N-anion (N at O site, i.e., N@O), N-cation (N at Ti site, i.e., N@Ti), and interstitial N-doped structures. A prior first-principles computational study for all these three possible models shows that N doping either introduces some impurity states in the bandgap for substitutional N-anion-doped and interstitial N-doped TiO2 or narrows the bandgap for substitutional N-cation doped TiO2, which both can lead to a redshift of the optical absorption edge [40] Therefore, the possibility that N doping leads to the blueshift of rutile TiO2 can be excluded. A phase transition from rutile to anatase by nitrogen doping can easily explain the blueshift of the optical absorption edge ~0.2 eV, though Diwald et al. did not observe the anatase phase in N-doped rutile TiO2 [98] In contrast, Henderson proposed that the blueshift may be explained according to hole trapping effects [8].

3.3. S Doping

Experiments show that there are two possible substitutional doping sites for S dopants in TiO2, i.e., S anion at O2− site (S@O) and S cation at Ti4+ site (S@Ti) [50,51,52,53,54,99,100,101,102,103], and both the S-anion- and S-cation-doped TiO2 show the high photocatalytic activity under visible light. For example, Umebayashi et al. prepared S-anion-doped anatase and rutile TiO2 by oxidation annealing of titanium disulfide (TiS2) and ion implantation in the rutile single crystal, respectively, and observed the visible absorption spectrum in these samples [53,54,99]. Similar visible light absorption property of S-anion-doped TiO2 was also reported by several other groups [101,102], and the S doping concentration was found to have a great influence on the visible light photocatalytic activity of S-anion TiO2 [100]. For substitutional S-cation-doped TiO2, interestingly, Yu et al. observed a bactericidal effect under visible light irradiation [50]. Ohno et al. even found that it exhibits stronger visible light absorption than N-, C-, and the S-anion-doped TiO2 [51,52,103]. As in the case of N-doped TiO2, the controversy on the mechanism of the visible light absorption in S-anion-doped TiO2 also exists. Earlier theoretical calculations indicated that S-anion doped TiO2 has an obvious bandgap narrowing [53,99,104,105], while the later first-principles calculation shows that S dopants introduce S 3p impurity states above the VB, which might be responsible for the redshift of the absorption edge of S-anion doped TiO2 [106].
First-principles studies revealed that S-anion- and S-cation-doped TiO2 show different mechanisms for the visible light absorption [43]. For S-anion-doped TiO2, different doping levels lead to the different electronic structure characteristics. At lower S doping concentration (≤2.08 at.%), the bandgap narrows slightly but some S 3p localized states are introduced in the bandgap. Therefore, electron excitations from these occupied S 3p states to CB might lead to a more significant redshift of the optical absorption edge than the slight bandgap narrowing [100]. At higher S doping concentration (≥4.17 at.%), the mixing of the S 3p states with the VB causes an obvious bandgap narrowing (~0.7 eV), thus leading to a substantial redshift of absorption spectra.
For S-cation-doped TiO2, S dopants introduce some occupied impurity states consisting of S 3s and O 2p states in the bandgap despite the unchanged bandgap. This indicates that the S dopant has an electron configuration resembling a S4+ (s2p0) ion in TiO2. Therefore, these impurity states can act as transition levels, and the electron excitations from these transition levels to CB may be responsible for the experimental redshift of the optical absorption edge. Similar S doping influence also appear in rutile TiO2 [43].

3.4. P Doping

Visible light photocatalytic activity was also reported in P-anion- and P-cation-doped TiO2, respectively. Shi et al. [107] found that P-cation-doped (P@Ti) TiO2 nanoparticles exhibit a stronger visible light absorption than undoped sample, which is thought to be induced by the impurity states in the bandgap. Furthermore, their X-ray photoelectron spectroscopy (XPS) measurements also indicated that the doped P ions are in the pentavalent-oxidation states (P5+) [107]. Yu et al. also found that P-doped TiO2 shows a better photocatalytic ability than that the pure TiO2 [108]. Li et al. observed a visible light absorption in substitutional P-anion-doped (P@O) anatase TiO2, and attributed it to a narrowed bandgap [109]. On the contrary, Yu et al. found a larger bandgap in phosphor-modified TiO2 than that of pure TiO2 [110].
To understand the mechanism underlying these inconsistent experimental observations, first-principles calculations for P-anion (P@O) and P-cation (P@Ti) doped TiO2 were performed [43]. It is found that P-anion doping does not cause a large bandgap narrowing but introduce P 3p states in the bandgap. In contrast, P-cation doping neither narrows the bandgap nor introduces impurity states in the bandgap. The first-principles results indicated that the high photocatalytic activity in P-cation-doped TiO2 may be caused by the large surface area and the crystallinity of TiO2 instead of the formation of an impurity energy level in the bandgap [108,110].

3.5. B Doping

There are three possible doping sites for boron in TiO2, i.e., substitutional B for O (B@O), substitutional B for Ti (B@Ti), and the interstitial B site. The atomic radius of B is larger than that of O (0.85 Å vs. 0.6 Å) but smaller than that of Ti (0.85 Å vs. 1.4 Å). Therefore, substitutional B-anion (B@O) doping and interstitial B doping are expected to cause a lattice expansion of TiO2, while substitutional B-cation (B@Ti) doping causes a lattice shrinking. Variable-cell structure optimizations for the three B-doped anatase models were carried out to qualitatively study the B doping influence on the lattice structure of TiO2 [64]. Compared with the undoped anatase model, the volumes of the substitutional B-anion and interstitial B-doped models expand approximately 7.7% and 7.0%, respectively. In contrast, the volume of substitutional B-cation-doped model shrinks by ~5%. These results are consistent with Jung et al.’s experiment in which the grain size of anatase phase was enlarged when the incorporation of boron oxides is more than 10% [111]. However, Chen et al. found that the doping of boron ions inhibited the crystal size [112]. Although substitutional B-cation doping can lead to a decrease of the volume of TiO2, this kind of structure is not consistent with Chen et al.’s experiment in which the B ions are sited at an interstitial position. Therefore, substitutional B-cation doping cannot account for Chen et al.’s experimental phenomenon.
The calculated electronic properties of three B-doped TiO2 models show that, for substitutional B-anion-doped TiO2, some impurity states mostly consisting of B 2p states are introduced in the bandgap [64] The electron excitation energy from VB to the unoccupied gap states above the Fermi level decreases about 0.3 eV with respect to the optical bandgap of undoped TiO2. This is consistent with the experimental redshift of the optical absorption edge in B-doped TiO2 [113]. It is noted that the substitutional B-anion doping leads to a spin-polarized electron state [62,64,114,115], and thus forms a paramagnetic defect consisting of B and adjacent Ti ions [62]. A detailed discussion can be found in Section 4. For the substitutional B-cation-doped TiO2, its optical bandgap decreases by ~0.3 eV due to the downward shift of the CB. On the contrary, for the interstitial B-doped TiO2, its optical bandgap increases by ~0.2 eV for anatase TiO2 and 0.3 eV for rutile TiO2 because of the well-known “band-filling mechanism” [116] or “Moss–Burstein shift” [117,118], which is often associated with the optical absorption shift in n-type semiconductors. This is in good agreement with the experimental blueshift of the optical absorption edge in interstitial B-doped TiO2, in which the optical absorption energy increases by ~0.12 eV. Therefore, the blueshift of optical absorption edge in B-doped TiO2 can be attributed to the intrinsic property of interstitial B-doped structure instead of the quantum size effects [111,112]. Similar doping effects also occur in B-doped rutile TiO2 [64,119]. It is also noted that the several different interstitial positions of B dopants should exist [62], and both the standard and hybrid DFT calculations have been carried out to study the trigonal-planar-coordinated [BO3] and pseudo-tetrahedral-coordinated [BO4] species [62,120]. It is found that the [BO3] species leads to an increase in the bandgap, while the [BO4] species leads to a decrease in the bandgap [120]. However, the [BO3] is shown to be more stable than [BO4] [62].

3.6. C Doping

C doping can extend the optical absorption of TiO2 effectively. Interestingly, different degrees of the reduction of the optical bandgap have been observed in C-doped TiO2 [10,121,122,123,124,125,126,127,128,129,130,131,132,133,134]. Khan et al. firstly found that C-anion-doped rutile TiO2 exhibits two optical absorption thresholds at 535 and 440 nm, which corresponds to the optical bandgap decrease of approximately 0.18 and 0.68 eV, respectively [10]. Soon after, more experimental studies confirmed that C-anion doping could induce different degrees of the redshift of the optical absorption edge of TiO2. For example, the optical bandgap reduction of approximately 0.3 eV and 0.45 eV was observed [121,122,123,124,125], and more pronounced optical bandgap reduction of approximately 0.72 eV [126], 0.86 eV [127,128], 0.95 eV [129,130], and 1.0 eV [131,132,133] was also observed. Xu et al. found two regions of photo-response from ultraviolet (UV) to 450 nm and 575 nm in C-doped anatase TiO2, which equals to the optical bandgap decrease of 0.45 eV and 1.05 eV, respectively [134]. These experiments both indicate that substitutional C-anion doping can lead to several different optical absorption thresholds in TiO2. In addition, carbon can also be incorporated into TiO2 as a cation at a Ti site because of its low electronegativity, and the C-cation doping influence on the optical absorption spectra of TiO2 is open. Kamisaka et al.’s first-principle calculations indicate that C-cation doping neither introduces gap states nor induces visible light absorption [135]. On the contrary, Ren et al. observed a visible absorption in the 400–450 nm range in their C-cation-doped TiO2 sample using UV–Vis diffuse reflectance spectroscopy [136].
Spin-polarized first-principles generalized gradient approximation (GGA) plus U calculations were carried to study the electronic and optical absorption properties of C-anion and C-cation-doped TiO2 [65]. With respect to the undoped TiO2, the calculated bandgap of C-anion-doped TiO2 changes slightly but some spin-polarized impurity states are introduced in the bandgap. Correspondingly, associated electron excitations among the VB, the CB, and the impurity states can be responsible for the various visible light absorption thresholds in C anion-doped TiO2. The electron excitation energy from the occupied gap states just above the valence band maximum (VBM) to the conduction band minimum (CBM) decreases approximately 0.33–0.53 eV, which is consistent with the reduction of the optical absorption energy approximately 0.30–0.45 eV [121,122,123,124,125,134]. The electron excitation energies from the occupied gap states to the CB and from the VB to the empty gap states reduce by approximately 0.63, 0.83, and 1.18 eV. This may be responsible for the large redshift (approximately 0.70–1.05 eV) of the absorption edge in C-doped anatase TiO2 and TiO2 nanotubes [126,127,128,129,130,131,132,133,134]. C-doped rutile TiO2 also shows same electronic characteristics [65].
Figure 4 shows density of states (DOS) plots of undoped and C-cation-doped anatase TiO2. For C-cation-doped TiO2, the calculated bandgap in the frame of GGA+U is approximately 2.85 eV, less than that of the undoped anatase TiO2 by approximately 0.18 eV [65]. This can explain the experimental redshift of the optical absorption and enhanced visible light absorption in the range of 400 to 450 nm [136]. Similar electronic structure modifications induced by C-cation doping also occur in C-doped rutile TiO2. It is worth mentioning that the local structure around the C dopant in anatase TiO2 significantly influences its electronic properties. Kamisaka et al. performed first-principles calculations for C-cation-doped TiO2, in which the cell size and shape were fixed, and found that the C dopant forms a planar CO3 species. Neither in-gap impurity states nor visible light absorbance is found in their calculations [135]. In contrast, Di Valentin et al. reported a local structure of CO4 unit and found the bandgap of C-cation-doped TiO2 shrank slightly approximately 0.1 eV [38]. In Yang et al.’s first-principles calculations, the lattice parameters and all the atomic positions are allowed to relax, in which the doped C atom forms a linear O-C-O unit with a short C-O distance (~1.2 Å) like carbon dioxide O=C=O. Interestingly, this structural characteristic of linear O-C-O unit was confirmed by experimental X-ray Photoelectron Spectroscopy (XPS) study, in which the XPS measurement of the C1s binding energy (288.6 eV) shows the presence of C=O bonds [136].
Interstitial C-doped TiO2 has also been studied using first-principles calculations [38,137]. It is found that the interstitial C dopants introduce impurity states in the bandgap, which can lead to the visible light absorption in C-doped TiO2.

3.7. Si Doping

Visible light photocatalytic activity of Si-doped TiO2 has also been reported [138,139,140,141]. For example, Oh et al. found that Si doping at low level could improve the photocatalytic activity of TiO2, while a high doping over 2% could decrease its photocatalytic activity [138]. Yan et al. observed that substitutional Si-cation doping (Si at Ti site) could cause a redshift of the absorption spectra of TiO2 and favor its photocatalytic activity [139]. Ozaki et al. prepared nitrogen-doped silica-modified TiO2 and found a high visible light photocatalytic activity [140,141]. First-principles calculations were carried out to understand the mechanism of extended visible light optical absorption [44]. In this work, substitutional Si-anion- (Si@O) and Si-cation- (Si@Ti) doped TiO2 are modeled, respectively. Compared with the undoped anatase TiO2, it is found that substitutional Si-cation doping leads to a bandgap narrowing of approximately 0.25 eV, which is consistent with the experimental visible light optical absorption [138,139,140,141]. The calculations also show that high silicon concentration doping cannot lead to further bandgap narrowing, but requires larger formation energy [47]. For the substitutional Si-anion-doped anatase TiO2, visible light absorption is expected because it shows the lower photon absorption energy than that of undoped anatase TiO2 [142].

3.8. Halogen Doping

In recent years, many attempts have been made to explore whether halogen doping can improve the visible light photocatalytic activity of TiO2 [143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159]. Among all the halogen elements, F is one of the most extensively studied dopants [143,144,145,146,147,148,149,150,151,152,153,154,155,156]. In particular, whether F doping could lead to visible light absorption is controversial. F-doped TiO2 has shown an improved photocatalytic activity [143,144] and a stronger absorption in the Ultraviolet/Visible (UV/Vis) region, which corresponds to a slight bandgap narrowing of approximately 0.05 eV [145]. Similar optical bandgap reduction also occurs in F-doped TiO2 nanotubes [146] and F-ion-implanted TiO2 [147,148]. On the contrary, more experimental studies show that F-doping neither induces any redshifts of the absorption edge of TiO2 [149,151,152,153] nor changes its optical absorption strength [154,155,156]. The photocatalytic properties of TiO2 doped with other halogen elements are also reported. Compared with the undoped and F-doped TiO2, Cl and Br co-doped TiO2 exhibits a smaller optical bandgap [157], and I-doped TiO2 shows a much better photocatalytic activity than undoped TiO2 under both visible light and UV–visible light [158,159].
To clarify the halogen doping influences on the photocatalytic properties of TiO2 under the UV/Vis light, it is essential to understand whether the doping introduces impurity states in the bandgap and how the doping affects the CB and VB edges, i.e., the CBM and VBM of TiO2. First-principle calculations of halogen-doped TiO2 have been done to understand its electronic properties and origin of the associated visible light photocatalytic activity [45]. The calculated bandgap of the undoped anatase TiO2 is approximately 2.10 eV. For F-, Cl-, Br-, and I-doped TiO2, the calculated bandgaps are approximately 2.06, 1.90, 1.80, and 1.40 eV, respectively. The calculated bandgap of the F-doped TiO2 is slightly less than that of the undoped TiO2 (by approximately 0.04 eV). As a result, it is understandable that either a very slight decrease (~0.05 eV [145]) or no optical bandgap change was found in experiments [149,151,152,153,154,155,156]. It is also noted each F atom ( s 2 p 5 ) requires one less electron from TiO2 than does each O atom ( s 2 p 4 ), and thus the substitution of one F ion for one O2− introduces one additional electron in the TiO2 lattice. As a result, the treatment of standard DFT calculations for the additional electron leads to the n-type conductive property of F-doped TiO2 [160]. In contrast, in the hybrid DFT or GGA + U calculations, the additional electron is shown to reduce one Ti4+ (d0) to Ti3+ (d1) ion and introduces localized gap states in the bandgap [161,162,163]. However, the lack of the visible light absorption in F-doped TiO2 indicates that the electron transition from these occupied gap states to the CB is not effective to lead to visible light absorption. Therefore, F is not a good dopant to extend the optical absorption edge of TiO2 into the visible light region. In contrast, an obvious bandgap narrowing (approximately 0.2 and 0.3 eV) is observed in Cl- and Br-doped TiO2, respectively. This is consistent with the experimentally observed reduction of the bandgap by approximately 0.2–0.3 eV in Cl/Br co-doped TiO2 [157]. For I-doped TiO2, the bandgap narrows by approximately 0.7 eV, and thus a significant visible light optical absorption is expected.
The photocatalytic ability of a semiconductor, i.e., the ability to transfer photon-excited electron–hole pairs to the absorbed species on the surface of the semiconductor, can be partially determined by the relative positions of its CBM and VBM with respect to the redox potentials of the adsorbate [2]. Therefore, the photocatalytic ability of TiO2 can be qualitatively evaluated by the positions of its CBM and VBM. Thermodynamically, the VBM of a photocatalytic semiconductor should lie below the redox level of the adsorbed species, while the CBM should lie above the redox level, so that the photoinduced hole in VB can capture an electron from the adsorbed species and the photoexcited electron in CB can be transferred to the adsorbed species. Figure 5 shows the comparison of the calculated VBM and CBM positions of X-anion doped anatase TiO2 (X = F, Cl, Br, and I) with the experimental values of undoped TiO2. A scissor operation of 1.10 eV was applied to make the calculated bandgap comparable to the experimental value. The corrected bandgaps for F-, Cl-, Br-, and I-doped TiO2 are 3.16, 3.00, 2.90, and 2.50 eV, respectively. For F-doped TiO2, its CBM and VBM shift downwards by 0.20 and 0.16 eV with respect to the undoped TiO2. This indicates that F-doped TiO2 should have a stronger oxidation ability, which can explain the experimentally observed higher photocatalytic activity in F-doped TiO2 than that of the undoped TiO2 [143,144,145,149,155,156]. For Cl-doped TiO2, its CBM shifts downwards by approximately 0.15 eV while the VBM shifts upwards by 0.05 eV relative to the corresponding values of undoped TiO2. Therefore, Cl doping may reduce the ability of oxidation and reduction of TiO2, which is in good agreement with the experimental fact that Cl-doped TiO2 shows lower water-splitting power than that of undoped TiO2 under UV irradiation [157]. For Br-doped TiO2, its VBM is nearly same with that of undoped TiO2, and thus Br-doped TiO2 should have the same oxidation ability with that of undoped TiO2. In contrast, the CMB of Br-doped TiO2 shifts downwards by approximately 0.3eV compared with that of undoped TiO2, so that Br-doped TiO2 might show lower ability to reduce H+ to H2 than does undoped TiO2. For I-doped TiO2, its CBM shifts downward slightly but VBM is raised strongly with respect to that of undoped TiO2, and thus visible light photocatalytic activity might appear in I-doped TiO2. In summary, the changes of the VBM and CBM can qualitatively explain some experimental facts that the photocatalytic ability was improved upon F, Cl, and Br doping [143,144,145,149,155,156,157]. In addition, the local internal field induced by the dipole moment is considered to be effective to separate the electron–hole pairs and inhibit their recombination [164,165], and thus the photocatalytic performance of doped TiO2 can be evaluated from the variation of the dipole moment of TiO6 octahedron adjacent to the dopants [165].
For I-doped anatase TiO2, its VBM and CBM are nearly the same as those of undoped TiO2. However, a double-occupied bandgap state above the VBM approximately 0.6 eV is introduced in the bandgap. Its PDOS plot shows that this gap state mostly consists of I 5s states and the 2p states of the neighboring O atoms around the I dopant, and that I 5p states contribute to its CB. This indicates that I dopant exists as an I5+ (s2) cation, as observed experimentally [158,159]. The Fermi level of I-doped TiO2 is pinned above the CBM approximately 0.2 eV, indicating that iodine is a good n-type dopant, as in the case of I-doped ZnTe [166]. Therefore, the optical bandgap of I-doped anatase TiO2 decreases approximately 0.4 eV with respect to that of undoped TiO2. This is mainly responsible for the optical bandgap decrease of I-doped TiO2 and the improved photocatalytic efficiency in the UV/Vis region [158,159]. For substitutional X-cation (X = F, Cl, and Br) doped TiO2, unlike the case of I-doped TiO2, Cl 3s and Br 4s states do not appear above the VBM. In contrast, singly-filled Cl 3p and Br 4p states appear in the bandgap of Cl- and Br-doped TiO2, respectively. This indicates that Cl and Br at Ti sites exist as Cl4+ (s2p1) and Br4+ (s2p1) ions, respectively. For F-doped TiO2, the F 2p states lie in the VB, and the empty F 2p states lie just above the Fermi level. Therefore, with respect to the Cl- and Br-doped TiO2, the doped F atom at a Ti site should exist as a F3+ (s2p2) ion to a first approximation [45].

3.9. Hydrogen Impurities in TiO2

As a ubiquitous impurity, hydrogen (H) widely exists in metal oxides, which forms either deep gap states or shallow donor levels [66,67,68,69]. Owing to its small ionic radius, H at an interstitial site is more stable than that at a substitutional site [167], though in principle, it can form substitutional structure by replacing the O atom. First-principles calculations show that the interstitial H introduces a shallow donor level in TiO2 [66,67,68,69]. This is consistent with the recent experiments in which the H is identified as a either a shallow donor or metastable donor in rutile TiO2 single crystal [168,169]. Interstitial H introduces an additional electron in TiO2, and thus the n-type conductivity can be understandable. However, recent GGA+U and hybrid functional calculations show that the additional electrons can be trapped by the Ti ions, reducing Ti4+ to Ti3+, and introduce the gap states [161,170], although the position of the gap states strongly depends on the choice of U parameters [162]. First-principles calculations also show that substitution of H for O can lead to the n-type conductivity, though its structure is not stable [167].
It is also interesting to explore the effects of H impurity on the structural stability and electronic property of nonmetal-doped TiO2 [167,171,172]. With respect to the N-doped TiO2, the combinational doping of (N, H) can enhance the structural stability and lead to a significant bandgap narrowing for anatase and brookite phases of TiO2 [167,171]. For N-anion-doped TiO2, the doped N dopant exists as N2 (s2p5) ion, and introduces an unoccupied N 2p impurity state in the bandgap [173]. For (N, H)-doped TiO2, the interstitial H atom adjacent to the N dopant introduces an additional electron into the lattice, and forms the N-H bond by transferring the additional electron to one empty N 2p orbital. As a result, all the N 2p impurity states become occupied, leading to either an obvious bandgap narrowing or double-filled N 2p impurity states above the VB. This type of charge compensation mechanism also occurs in (N, Ta)- and (N, F)-doped TiO2 [60,174], in which the substitutional Ta-cation (Ta5+) and F-anion (F) doping introduce one additional electron into the TiO2 lattice separately, which fills the empty N 2p orbital. For (C, H)-doped TiO2, it is thermodynamically more stable than C-doped TiO2, and a bandgap narrowing also occurs [172]. However, in this case, one electron donated by the interstitial H cannot compensate the two holes of C2− (s2p4) dopant [175], and thus this incomplete charge compensation results in an impurity level just below the CB [172]. It is expected that two interstitial H and one substitutional C at an O site can satisfy full charge compensation. In addition, recent experimental and theoretical studies show that hydrogenation on the surface of TiO2 can lead to a structural disorder and introduce mid-gap states, which is responsible for the improved solar-driven photocatalytic activity [176,177]. In summary, the influence of hydrogen impurity on the structural stability and electronic property of TiO2 is one interesting topic, and further work is worthy to be done.

3.10. Co-Doping

Introducing impurity states in the bandgap by nonmetal or metal doping is effective to extend the optical absorption edge of TiO2 into the visible light region; however, the electron–hole recombination center, which generally refers to the impurity-related states near the middle of the bandgap, inhibits the photocatalytic efficiency of TiO2. Therefore, narrowing bandgaps without creating mid-gap states is necessary to maximize the photocatalytic performance of TiO2 under the visible light irradiation [178,179,180,181]. To this aim, co-doping can be an effective approach because it has three important effects:
(1) Co-doping can eliminate impurity states in the bandgap via charge compensation between different dopants, which can reduce the number of recombination centers and promote the separation of electron–hole pairs.
(2) Co-doping can facilitate the mixing between the impurity states and VB (CB) by adjusting the position of impurity states in the bandgap, and narrow the bandgap effectively.
(3) Co-doping can reduce the formation energy of the combination defect with respect to the monodoped TiO2, and thus improving the solution of ideal dopants in TiO2 or other metal oxides becomes possible.
Herein, we divided the co-doping into three classes, i.e., anion–anion co-doping, anion–cation co-doping and cation–cation co-doping, according to the ionic type (anions or cations) of dopants, and summarized the recent theoretical research progress of the co-doped TiO2, mainly from the viewpoint of charge-compensation.

3.10.1. Anion–Anion Co-Doping

One typical example of anion–anion co-doping is (N, F) co-doped TiO2 [60,150,151,182,183,184], in which N and F both replace O ions. In N-anion monodoped TiO2, one N at an O site introduces an accept level in the bandgap and exists as a N2− ion, while in F-anion monodoped TiO2, one F at O site introduces one donor level by donating one additional electron into the TiO2 lattice. As a result, in (N, F) co-doped TiO2, this additional electron can occupy the N 2p accept level, and then all the N impurity states become occupied. These occupied N 2p states are located just above the VB, and do not act as the electron–hole recombination center, thus improving the visible light photocatalytic activity of TiO2 with respect to that of the N-anion monodoped TiO2 [60,185]. In addition, (N, F) co-doping reduces the formation energy with respect to that of the N monodoped TiO2, and thus a higher nitrogen concentration doping can be realized. Similar anion–anion charge compensation effect is expected to occur in (N, Cl), (N, Br), (N, I), (P, F), (P, Cl), (P, Br), and (P, I) co-doped TiO2.

3.10.2. Anion–Cation Co-Doping

(N, H i n t ) co-doped TiO2 is a typical example of the anion–cation co-doping, in which N replaces O while H locates at an interstitial site. In this system, one electron donated by an interstitial H occupy the N 2p acceptor level, and the bandgap narrows approximately 0.26 eV, much larger than the N-anion monodoped TiO2 (0.04 eV) [171].
Another two kinds of anion–cation co-doping combinations are (N, X) and (C,Y)[174,179,186,187], in which X can be Ta5+ or Nb5+ while Y (Herein Y is not element yttrium.) can be W6+ or Mo6+. It is known that N dopant exists as N2− (s2p5) ion at an O site [173], and introduces one acceptor level in the bandgap, while the pentavalent Ta (Nb) cation at a Ti site donates one electron into TiO2 lattice, introducing one donor level in the bandgap. As a result, the charge compensation can occur in (N, Ta) [174,186], and (N, Nb) co-doped TiO2, and the electron–hole recombination center (i.e., the impurity-related states near the middle of the bandgap) might be removed, thus promoting the rate of electron–hole separation and improving its photocatalytic efficiency. Similarly, the C dopant exists as C2− (s2p4) ion at an O site [175], and introduces two acceptor levels in the bandgap. In contrast, the hexavalent W (Mo) cation at a Ti site donates two electrons into TiO2 lattice, i.e., introducing two donor levels. Therefore, in (C, W) and (C, Mo) co-doped TiO2 [179,187], the two acceptor levels introduced by C dopant become occupied by capturing the two electrons donated by W and Mo, and C 2p impurity states are more close to the VB, reducing the number of electron–hole recombination centers. However, it is noted that the full charge compensation only occurs when the doped anion and cation are bonded together, though this configuration generally possesses the lowest total energy [187,188]. This is reasonable because the doped anion and cation can form a strong bond by a direct charge transfer.
Another kind of anion–cation co-doping, such as (N, Mo) co-doping, has been proposed [188], in which two N anions are bonded to one hexavalent Mo cation, forming a near-linear N-Mo-N unit. For convenience, we refer it to as (N-Mo-N) co-doping. In this case, two electrons donated by one hexavalent Mo cation at a Ti site can compensate two holes introduced by two N anions, thus achieving full charge compensation. Liu et al. further studied the difference of the electronic structures between non-passivated and passivated (N, Mo) co-doping [188]. The local geometrical configurations of N and Mo dopants for these two kinds of (N, Mo) co-doped TiO2 are shown in Figure 6a,b. Their calculated TDOS and PDOS plots are shown in Figure 6c. For non-passivated (N, Mo) co-doping, one N anion is bonded to one Mo cation, and the doping ratio between N and Mo is 1:1. In this case, one hole introduced by one N cannot compensate two electrons donated by one Mo, thus leading to n-type conductivity. In passivated (N, Mo) co-doped TiO2, i.e., (N-Mo-N) co-doped system, the two electrons are fully compensated by the two holes introduced by the two N anions, and the N 2p acceptor levels are removed and do not act as an electron–hole recombination centers, thus improving the photocatalytic efficiency. Similar to the case of (N-Mo-N) co-doped TiO2, full charge compensation is also expected in (N-W-N), (Ta-C-Ta), and (Nb-C-Nb) co-doped TiO2.

3.10.3. Cation–Cation Co-Doping

Cation–cation co-doping has also been proposed to narrow the bandgap of TiO2, and two criteria of choosing cations are suggested [180]. First, the cation should have a closed-shell electronic configuration like d0 or d10. Second, to keep the semiconductor characteristic of TiO2, co-doping with cations Ax+ and By+ should satisfy a simple rule, i.e., x + y = 8. On the basis of the criteria, the cation–cation co-doping combination of (Mo6+, Zn2+/Cd2+) and (Ta5+, Ga3+/In3+) is expected to cause an effective bandgap narrowing without introducing gap states. In these two kinds of cation–cation co-doped TiO2, the number of the holes introduced by Zn2+/Cd2+ (Ga3+/In3+) equals to that of electrons introduced by Mo6+ (Ta5+), and thus full charge compensation can be obtained. In addition, the electronic configuration of d0 or d10 of the cations can guarantee that the d states of dopants either are fully unoccupied or occupied, and thus no d-orbital-related impurity states can be created in the bandgap. First-principles calculations show that charge compensation does occur in the (Ta5+, Ga3+/In3+) and (Mo6+, Zn2+/Cd2+) co-doped TiO2, and with respect to cation monodoped TiO2, the mid-gap states are passivated.
In summary, co-doping can be an effective approach to narrow the bandgap of semiconductor photocatalyst and get rid of the electron–hole recombination centers through the full charge compensation. To this aim, the number of the holes on acceptor levels should be equal to that of the electrons on donor levels. Therefore, we can define the following equation to be a basic rule for co-doping:
N a × N h = N d × N e .
N a and N d represent the number of acceptor atoms and donor atoms, respectively. N h represents the number of introduced holes per acceptor atom and N e represents the number of donated electrons per donor atom. It is noted that the Ti (O) ions in TiO2 and SrTiO3 have same chemical state, and thus the co-doping combinations mentioned above can also be applied in SrTiO3 [189,190,191]. For a reference to choose an ideal combination of dopants for co-doping in TiO2 and SrTiO3, we list the common acceptor (donor) atoms as well as the number of the holes (electrons) per acceptor (donor) atom in Table 2.

4. Interface

4.1. TiO2/Perovskite Interface

The organic–inorganic halide perovskite solar cells have received significant attention in the recent years because of the low cost, high power conversion efficiency, and the flexibility of halide perovskite materials. In the perovskite solar cells, TiO2 is widely used as electron transport layers and plays a significant role in achieving high efficiency and stability of photovoltaic devices. One major reason is the conduction band of TiO2 is lower than that of the halide perovskite such as MAPbI3 and MASnI3 (MA = CH3NH3, methylammonium) [16,17]. Therefore, first-principles computational studies of TiO2/perovskite interfaces have been an emerging topic recently [16,17,18,19,20,21,22,23,24].
De Angelis’s research team has extensively studied the material interfaces between the anatase TiO2 (denoted as aTiO2) and tetragonal MAPbI3 and MAPbI 3 x Clx perovskite from first-principles calculations [18,21,194,195,196]. To build materials interface models, the first step is generally to choose the most stable surfaces. The authors considered pseudo-cubic (001) and tetragonal (110) perovskite surfaces because they found the (001) surfaces of the cubic and tetragonal phases have an identical topology [21]. That is to say, the (001) surface of the tetragonal phase is essentially same with one of the three equivalent surfaces of the cubic phase. Their calculations show that the isolated (110) surface is energetically more favorable than the (001) surface, with a total energy difference of approximately 0.7 and 0.1 eV for MAPbI3 and MAPbI 3 x Clx, respectively. Interestingly, after “depositing” the perovskite on the aTiO2 (101) surface, the (110) perovskite surface are further stabilized against the (001) surface from the energetic analysis; moreover, the interfacial Cl atoms further increases the interfacial binding energy to TiO2 in MAPbI 3 x Clx compared to that in MAPbI3. Later, a combined angle-resolved X-ray photoelectron spectroscopy (AR-XPS) and first-principles calculations by the same research team revealed more details at the MAPbI 3 x Clx/TiO2 interface. They found that Cl is preferentially located at TiO2 interface rather than the bulk perovskite [194], and the interfacial Cl is also found to induce the band bending that creates a directional “electron funnel”, improving the charge collection efficiency of the photovoltaic devices. This conclusion is also consistent with the previous computational study on Cl-doped bulk TiO2 in which Cl is thermodynamically favorable to replace O [45].
The strong interfacial interaction between TiO2 and perovskite is essentially attributed to the charge transfer across the TiO2/perovskite interface. To quantify the charge transfer, De Angelis et al. carried out a charge displacement analysis, as shown by calculated charge displacement curves for TiO2/MAPbI3 and TiO2/MAPbI 3 x Clx interfaces in Figure 7 [21]. It clearly shows a strong charge displacement variation across the interfacial region in both cases, indicating a strong interaction between TiO2 and perovskite and a strong polarization. Furthermore, there is a slight increase in charge donation to TiO2 in the case of MAPbI 3 x Clx, suggesting an increased charge accumulation at the TiO2/MAPbI 3 x Clx interface compared to TiO2/MAPbI3 interface. This charge displacement analysis well explains the stronger binding energy for TiO2/MAPbI 3 x Clx than that for TiO2/MAPbI3. This work indicates that the charge displacement analysis is one effective approach to visualize the charge displacement from the non-interacting two materials fragments to the interacting heterointerface.
In 2019, Sultana et al. reported first-principles computational studies for the MAPbI3 perovskite interfaces with TiO2, ZnO, and SnO2 [22]. The interface models were built by placing tetragonal MAPbI3 along its [001] direction on top of rutile (001) TiO2, wurtzite (1010) ZnO, and rutile (110) SnO2 slabs. To understand the interfacial charge transport properties of these systems, the authors calculated their charge difference density using the following equation.
Δ ρ = ρ i n t e r f a c e ρ M A P b I ρ M O
The authors analyzed the charge transfer from charge density difference plot and found that the charge accumulation and depletion mainly occur around the interfacial Pb and O atoms, indicating that the charge transfer mostly occurs at the interface. Among the three materials interfaces, the TiO2-based interface has the highest charge transfer based on the Bader charge analysis, thus suggesting that TiO2-based interface has the highest efficiency for the perovskite solar cells. However, the authors found that TiO2 and ZnO have a stronger binding energy with MAPbI3 than SnO2, but the strong binding may lead to decomposition of CH3NH3 molecules. The authors thus conclude that SnO2 can be a feasible replacement for TiO2 and ZnO to improve the stability of perovskite solar cells.
The defect properties at the TiO2/perovskite interface were also studied from first-principles calculations. In 2017, Haruyama et al. reported the defect properties at the a TiO2(101)/tMAPbI3(110) interface using first-principles computational approach, in which anatase TiO2 (001) was placed on the MAPbI3 (110) plane [20]. They found that the vacancy defects in the TiO2 layer create undesired defect levels within the bandgap, which serves as hole traps and recombination centers; while most of the vacancy defects in the MAPbI3 layer produces no additional states, thus having no influence on the electron–hole separation rates. This computational work indicates a possible route to further improve the performance of perovskite solar cells via interface modification.
Yang et al. studied relative stability and charge transport properties of different interfaces between the lead-free MASnI3 and TiO2, including anatase and rutile phases from first-principles calculations [24], in which (001) surface of MASnI3 and (001) surfaces of anatase and rutile were used to build models. The authors considered two types of terminations of MASnI3 and built four interface models—MAI/A, MA/R, SnI2/A, and SnI2/R, in which A and R represents anatase and rutile phases, respectively, as shown in Figure 8. The authors found that SnI2/A interface is more stable than the other three systems and this interface also has a better separation of photoinduced electron–hole pairs from the analysis of the plane-averaged electrostatic potential and density of states. Similarly, by examining four types of interfaces between MAPbI3/TiO2 from first-principles calculations, Gent et al. also concluded that PbI2/A interface has a stronger interaction but PbI2/R interface model is most efficient for charge separation from the electrostatic potential analysis, which was attributed to the better lattice and atoms arrangement match between MAPbI3 and rutile TiO2 [19].
In addition to the extensively studied materials interfaces between the hybrid halide perovskites and TiO2, it is noted that the materials heterointerfaces between the classical inorganic perovskite oxides such as SrTiO3 and TiO2 has also attracted increasing attention because of the emerging interfacial properties and/or enhanced functionalities [23,197]. One major consideration is that the formed heterointerface has a better stability compared to the isolated bulk materials [23,198]. In 2017, Kitchin’s research team explored the possibility to stabilize the epitaxial thin films of anatase and rutile TiO2 on perovskite (Sr/Ba)TiO3 by studying relevant bulk and interface energy terms using first-principles calculations [23]. The authors first identified four potential coherent epitaxial interfaces—(001)A/(001)(S/B)TO, (102)A/(011)(S/B)TO, (112)A/(111)(S/B)TO, and (100)R/(111)(S/B)TO—based on the experimental findings, volumetric formation and strain, and areal substrate, and then calculated their interface energies. Their computational models predict a decrease order of anatase stability from (001) to (011) to (111) perovskite substrates, which is in a general agreement with the experimental epitaxial stability investigations. In contrast, rutile is energetically more favorable than anatase on the (111) perovskite substrate. This work indicates that first-principles calculations of interface energy is a promising approach to predict epitaxial polymorph stability and to search for coherent epitaxial interfaces toward the stability of target materials even beyond TiO2 polymorphs.
The TiO2 polymorphic structures exhibit different photocatalytic properties and the synthesis of the specific TiO2 polymorph with target materials properties are desired. One possible way to stabilize the metastable polymorphs is to use the epitaxial synthesis approach, i.e., materials heterointerfaces.

4.2. TiO2/BiOI Interface

Bismuth oxyhalide BiOX (X = Cl, Br, and I) has a layer structure and exhibits promising catalytic properties for degrading pollutants under visible light irradiation because of their low bandgaps [22,23,24,194,195]. However, high electron–hole recombination rates in this class of materials severely inhibit their photocatalytic applications. A heterostructure that combines two different semiconductors could transfer charge carriers between a high energy band and a low energy band, which could effectively separate photoinduced electron–hole pairs, and thus greatly enhance the photocatalytic efficiency of semiconductor heterostructure. In a very recent study [199], Qu et al. studied BiOI/TiO2 heterostructure using first-principles electronic structure calculations, in which the heterostructure model was built by “depositing” (001) BiOI slab on the (001) rutile TiO2 slab. In this work, two types of materials interfaces including 1I/TiO2 and BiO/TiO2 were modeled, as BiOI has two types of surface terminations along the [001] direction. Note that the authors chose the average value of the two separate lattice parameters of the BiOI and TiO2 as the lattice parameters of the new heterostructure in order to achieve a high lattice match between BiOI (001) and TiO2 (001) surfaces, leading to a lattice mismatch of 6.63%. The calculated formation energy is −2.21 eV/A and −2.25 eV for 1I/TiO2 and BiO/TiO2 models, respectively, suggesting that both two models are likely to be formed and the BiO/TiO2 is slightly more stable than 1I/TiO2. More interestingly, the bandgap of 1I/TiO2 heterostructure reduces to 0.28 eV while BiO/TiO2 heterostructure changes into an n-type semiconductor, see Figure 9a. Moreover, as shown in Figure 9c, the built-in potential of the heterostructure can effectively separate the photoinduced electron–hole pair across the interface, thus potentially improving the photocatalytic performance.

4.3. TiO2/RuO2 Interface

Ruthenium oxide (RuO2) has various technological applications such as being used as electrodes, supercapacitors, and catalysts because of its promising properties [11]. It has a rutile-like structure and exhibits an excellent electrical conductivity. In recent years, TiO2/RuO2 heterostructure has been proposed as one efficient photocatalyst for water splitting and decomposition of organics [11,12,13,14,15], which is attributed to the high separation rate of photogenerated electron–hole pairs. In 2015, Wei et al. studied the energetic and electronic properties of the TiO2/RuO2 (110) heterostructure in which RuO2 was “deposited” on the rutile TiO2 (110) surface using first-principles DFT calculations [200]. The calculated interfacial energy shows that the TiO2/RuO2 interface is exothermic, indicating a strong bonding interaction between TiO2 and RuO2. The calculated projected DOS plots for the TiO2/RuO2 heterostructure show that the valence and conduction band of TiO2 bend downward toward RuO2. Interestingly, upon introducing oxygen vacancies at the interface, there exists an upward band bending toward RuO2, in agreement with the experimental finding [11].
Ruthenium oxide (RuO2) has various technological applications such as being used as electrodes, supercapacitors, and catalysts because of its promising properties [11]. It has a rutile-like structure and exhibits an excellent electrical conductivity. In recent years, TiO2/RuO2 heterostructure has been proposed as one efficient photocatalyst for water splitting and decomposition of organics [11,12,13,14,15], which is attributed to the high separation rate of photogenerated electron–hole pairs. In 2015, Wei et al. studied the energetic and electronic properties of the TiO2/RuO2 (110) heterostructure in which RuO2 was “deposited” on the rutile TiO2 (110) surface using first-principles DFT calculations [200]. The calculated interfacial energy shows that the TiO2/RuO2 interface is exothermic, indicating a strong bonding interaction between TiO2 and RuO2. The calculated projected DOS plots for the TiO2/RuO2 heterostructure show that the valence and conduction band of TiO2 bend downward toward RuO2. Interestingly, upon introducing oxygen vacancies at the interface, there exists an upward band bending toward RuO2, in agreement with the experimental finding [11].
For convenience, a brief summary of these materials interfaces and their critical materials properties are listed in Table 3. However, it is worth mentioning that, in addition to the material interfaces discussed above, there exist many other types of TiO2-based materials interfaces such as TiO2/SnO2 [201], TiO2/WS2 [202], TiO2/WSe2 [203], TiO2/Mo2 [204], TiO2/ZnS [205], TiO2/g-C3N4 [206], TiO2/MnOx [207], TiO2/Mo3 [208], TiO2/C60 [209], and TiO2-based plasmonic composite [210]. The central idea of these heterostructures is to enhance the charge separation rate of the photoinduced electron–hole pairs via the interfacial charge transfer mechanism, which generally requires the appropriate band alignment. Among these interfaces, it is particularly worth mentioning that plasmonic composites that consist of plasmonic nanoparticles such as noble metals [210,211] and TiO2 are emerging as one promising photocatalytic system because of their efficient light absorption. Details on this new class of photocatalytic composites can be found in recent review articles [212,213,214,215,216,217,218,219,220].

5. Conclusions and Outlook

In summary, this review covered first-principles computational and theoretical understanding on the structural, energetic, electronic, and optical properties of TiO2-based nanocluster, bulk, and material interface. A fundamental understanding and computational design of novel TiO2-based materials structures and/or composites is of great importance for developing highly efficient photocatalytic and photovoltaic applications, in which the first-principles computational approach is expected to play an increasing role. Herein, we discuss several possible future research directions from the viewpoint of first-principles calculations:
(i) Nonmetal or metal doping is one effective approach to tune electronic properties of TiO2, which, however, will also significantly influence the anatase-to-rutile phase transition. A clear computational and theoretical understanding of the phase-transition mechanism will be useful to further optimize the performance of TiO2 in photocatalytic and other applications. In this regard, future efforts can be devoted to the development of novel computational and analytical approaches based on first-principles calculations.
(ii) Increasing the number of photoinduced electron–hole pairs is one effective way to achieve high photocatalytic performance of TiO2. Researchers have made great efforts to tune the bandgap of TiO2 via doping engineering or defect engineering, which essentially utilizes light absorption property of TiO2 for producing electron–hole pairs. Instead of utilizing generated electron–hole pairs in TiO2, the emerging TiO2-based plasmonic composites that utilize the light absorption of plasmonic nanoparticles instead of TiO2 are one promising approach to achieve high photocatalytic performance. However, the current plasmonic nanoparticles are typically noble metal particles, and thus future direction could be to search for low-cost alternative plasmonic nanoparticles that can work with TiO2 and exhibit comparable and even superior photocatalytic properties.
(iii) Increase the separation rate of photoinduced electron–hole pairs is another approach to improve the phtocatalytic activity of TiO2-based catalysts. Although several TiO2-based materials interfaces have been studied from first-principles, future efforts can be on the design of novel TiO2-based material interfaces such as heterostructures or core-shell structures that can significantly promote the separation rate of electron–hole pairs.
(iv) The emerging high-throughput computational materials design approach is becoming one powerful tool to significantly accelerate the materials discovery and development process, which has been successfully applied in the design of bulk materials [221,222,223,224], material interfaces [225,226], and organic–inorganic hybrid materials with target properties [227,228]. It is thus expected that the application of such an approach in the design of TiO2-based photocatalysts is likely to yield fruitful results and worth further exploring.

Funding

This work was supported by the National Science foundation of China under Grant 51972195. K.Y. acknowledges Academic Senate General Campus Research Grant Committee at University of California San Diego.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript.
DFTDensity Functional Theory
GGAGeneralized Gradient Approximation
DOSDensity of States
VBValence Band
VBMVB Maximum
CBConduction Band
CBMCB Minimum
2,5-DMBQDimethylbenzoquinones
4-BBA4-bromobenzaldehyde
XPSX-ray Photoelectron Spectroscopy
AR-XPSAngle-resolved XPS
UVUltraviolet

References

  1. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  2. Linsebigler, A.L.; Lu, G.; Yates, J.T., Jr. Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  3. Hoffmann, M.R.; Martin, S.T.; Wonyong, C.; Bahnemann, D.W. Environmental Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  4. Chen, X.; Mao, S.S. Titanium Dioxide Nanomaterials: Synthesis, Properties, Modifications, and Applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, L.; Jiang, Y.; Zhao, H.; Chen, J.; Cheng, J.; Yang, K.; Li, Y. Engineering Coexposed and 001 and 101 Facets in Oxygen-Deficient TiO2 Nanocrystals for Enhanced CO2 Photoreduction under Visible Light. ACS Catal. 2016, 6, 1097–1108. [Google Scholar] [CrossRef]
  6. Nie, X.; Zhuo, S.; Gloria, M.; Sohlberg, K. Doping of TiO2 Polymorphs for Altered Optical and Photocatalytic Properties. Int. J. Photoenergy 2009, 2009, 294042. [Google Scholar] [CrossRef] [Green Version]
  7. Zaleska, A. Doped-TiO2: A Review. Recent Patents Eng. 2008, 2, 157–164. [Google Scholar] [CrossRef]
  8. Henderson, M.A. A Surface Science Perspective on Photocatalysis. Surf. Sci. Rep. 2011, 66, 185–297. [Google Scholar] [CrossRef]
  9. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef] [PubMed]
  10. Khan, S.U.M.; Al-Shahry, M.; William, B.I., Jr. Efficient Photochemical Water Splitting by a Chemically Modified n-TiO2. Science 2002, 297, 2243–2245. [Google Scholar] [CrossRef] [PubMed]
  11. Uddin, M.T.; Nicolas, Y.; Olivier, C.; Toupance, T.; Müller, M.M.; Kleebe, H.J.; Rachut, K.; Ziegler, J.; Klein, A.; Jaegermann, W. Preparation of RuO2/TiO2 Mesoporous Heterostructures and Rationalization of Their Enhanced Photocatalytic Properties by Band Alignment Investigations. J. Phys. Chem. C 2013, 117, 22098–22110. [Google Scholar] [CrossRef]
  12. Ismail, A.A.; Robben, L.; Bahnemann, D.W. Study of the Efficiency of UV and Visible-Light Photocatalytic Oxidation of Methanol on Mesoporous RuO2-TiO2 Nanocomposites. ChemPhysChem 2011, 12, 982–991. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, F.; Kundu, S.; Vidal, A.B.; Graciani, J.; Ramirez, P.J.; Senanayake, S.D.; Stacchiola, D.; Evans, J.; Liu, P.; Sanz, J.F.; et al. Determining the Behavior of RuO(x) Nanoparticles in Mixed-Metal Oxides: Structural and Catalytic Properties of RuO2/TiO2(110) Surfaces. Angew. Chem. Int. Ed. Engl. 2011, 50, 10198–10202. [Google Scholar] [CrossRef] [PubMed]
  14. Kundu, S.; Vidal, A.B.; Yang, F.; Ramà rez, P.J.; Senanayake, S.D.; Stacchiola, D.; Evans, J.; Liu, P.; Rodriguez, J.A. Special Chemical Properties of RuOx Nanowires in RuOx/TiO2(110): Dissociation of Water and Hydrogen Production. J. Phys. Chem. C 2012, 116, 4767–4773. [Google Scholar] [CrossRef]
  15. Xiang, G.; Shi, X.; Wu, Y.; Zhuang, J.; Wang, X. Size Effects in Atomic-Level Epitaxial Redistribution Process of RuO2 Over TiO2. Sci. Rep. 2012, 2, 801. [Google Scholar] [CrossRef] [PubMed]
  16. Guo, Y.; Xue, Y.B.; Li, X.C.; Li, C.B.; Song, H.X.; Niu, Y.S.; Liu, H.; Mai, X.M.; Zhang, J.X.; Guo, Z.H. Effects of Transition Metal Substituents on Interfacial and Electronic Structure of CH3NH3PbI3/TiO2 Interface: A First-Principles Comparative Study. Nanomaterials 2019, 9, 966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Bernal, C.; Yang, K. First-Principles Hybrid Functional Study of the Organic-Inorganic Perovskites CHNH3SnBr3 and CH3NH3SnI3. J. Phys. Chem. C 2014, 118, 24383–24388. [Google Scholar] [CrossRef]
  18. Akbari, A.; Hashemi, J.; Mosconi, E.; De Angelis, F.; Hakala, M. First Principles Modelling of Perovskite Solar Cells Based on TiO2 and Al2O3: Stability and Interfacial Electronic Structure. J. Mater. Chem. A 2017, 5, 2339–2345. [Google Scholar] [CrossRef]
  19. Geng, W.; Tong, C.J.; Liu, J.; Zhu, W.J.; Lau, W.M.; Liu, L.M. Structures and Electronic Properties of Different CH3NH3PbI3/TiO2 Interface: A First-Principles Study. Sci. Rep. 2016, 6, 20131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Haruyama, J.; Sodeyama, K.; Hamada, I.; Han, L.; Tateyama, Y. First-Principles Study of Electron Injection and Defects at the TiO2/CH3NH3PbI3 Interface of Perovskite Solar Cells. J. Phys. Chem. Lett. 2017, 8, 5840–5847. [Google Scholar] [CrossRef] [PubMed]
  21. Mosconi, E.; Ronca, E.; De Angelis, F. First-Principles Investigation of the TiO2/Organohalide Perovskites Interface: The Role of Interfacial Chlorine. J. Phys. Chem. Lett. 2014, 5, 2619–2625. [Google Scholar] [CrossRef] [PubMed]
  22. Sultana, N.; Al Amin, A.; Metin, D.Z.; Gaston, N. Unveiling the Structures and Electronic Properties of CH3NH3PbI3 Interfaces with TiO2, ZnO, and SnO2: A First-Principles Study. J. Mater. Sci. 2019, 54, 13594–13608. [Google Scholar] [CrossRef]
  23. Xu, Z.N.; Salvador, P.; Kitchin, J.R. First-Principles Investigation of the Epitaxial Stabilization of Oxide Polymorphs: TiO2 on (Sr,Ba)TiO3. ACS Appl. Mater. Interfaces 2017, 9, 4106–4118. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, Z.Z.; Wang, Y.X.; Liu, Y.Y. Stability and Charge Separation of Different CH3NH3SnI3/TiO2 Interface: A First-Principles Study. Appl. Sur. Sci. 2018, 441, 394–400. [Google Scholar] [CrossRef]
  25. Mohajeri, A.; Haghshenas, F. Global Reactivity and Site Selectivity of (TiO2) Nanoclusters (n = 5–10) Toward Hydrogen Peroxide. Mater. Chem. Phys. 2016, 183, 326–333. [Google Scholar] [CrossRef]
  26. Bai, F.Y.; Ni, S.; Ren, Y.; Tang, Y.Z.; Zhao, Z.; Pan, X.M. DFT Analysis on the Removal of Dimethylbenzoquinones in Atmosphere and Water Environments: Center Dot OH-Initiated Oxidation and Captured by (TiO2)n Clusters (n = 1–6). J. Hazard. Mater. 2020, 386, 121636. [Google Scholar] [CrossRef] [PubMed]
  27. Qu, Z.W.; Kroes, G.J. Theoretical Study of the Electronic Structure and Stability of Titanium Dioxide Clusters (TiO2)n with n = 1–9. J. Phys. Chem. B 2006, 110, 8998–9007. [Google Scholar] [CrossRef] [PubMed]
  28. Qu, Z.w.; Kroes, G.J. Theoretical Study of Stable, Defect-Free (TiO2)n Nanoparticles with n = 10–16. J. Phys. Chem. C 2007, 111, 16808–16817. [Google Scholar] [CrossRef]
  29. Qu, Z.W.; Zhu, H. Do Anionic Titanium Dioxide Nano-Clusters Reach Bulk Band Gap? A Density Functional Theory Study. J. Comput. Chem. 2010, 31, 2038–2045. [Google Scholar] [CrossRef] [PubMed]
  30. Hernandez, A.B.; Hernandez, W.I.; Cid, A.A.P.; Garcia, J.H.C.; Villanueva, M.S. Prediction, and Physic-Chemical Properties of (TiO2)n n = 15–20 Clusters and Their Possible Catalytic Application: A DFT Study. Comput. Mater. Sci. 2019, 162, 228–235. [Google Scholar] [CrossRef]
  31. Arab, A.; Ziari, F.; Fazli, M. Electronic Structure and Reactivity of (TiO2)n (n = 1–10) Nano-Clusters: Global and Local Hardness Based DFT Study. Comput. Mater. Sci. 2016, 117, 90–97. [Google Scholar] [CrossRef]
  32. Joo, P.H.; Cheng, J.; Yang, K. Size Effects and Odd–Even Effects in MoS2 Nanosheets: First-Principles Studies. Phys. Chem. Chem. Phys. 2017, 19, 29927–29933. [Google Scholar] [CrossRef] [PubMed]
  33. Nollet, H.; Roels, M.; Lutgen, P.; Van der Meeren, P.; Verstraete, W. Removal of PCBs from Wastewater Using Fly Ash. Chemosphere 2003, 53, 655–665. [Google Scholar] [CrossRef]
  34. Gan, H.L.; Peng, L.; Gu, F.L. A DFT Study on the Mechanism of Photoselective Catalytic Reduction of 4-Bromobenzaldehyde in Different Solvents Employing an OH-Defected TiO2 Cluster Model. Phys. Chem. Chem. Phys. 2017, 19, 27755–27764. [Google Scholar] [CrossRef] [PubMed]
  35. Andreev, A.S.; Kuznetsov, V.N.; Chizhov, Y.V. Atomic Hydrogen Activated TiO2 Nanocluster: DFT Calculations. J. Phys. Chem. C 2012, 116, 18139–18145. [Google Scholar] [CrossRef]
  36. Morikawa, T.; Asahi, R.; Ohwaki, T.; Aoki, K.; Taga, Y. Band-Gap Narrowing of Titanium Dioxide by Nitrogen Doping. Jpn. J. Appl. Phys 2001, 40, L561–L563. [Google Scholar] [CrossRef]
  37. Van de Walle, C.G.; Neugebauer, J. First-Principles Calculations for Defects and Impurities: Applications to III-nitrides. J. Appl. Phys. 2004, 95, 3851–3879. [Google Scholar] [CrossRef]
  38. Di Valentin, C.; Pacchioni, G.; Selloni, A. Theory of Carbon Doping of Titanium Dioxide. Chem. Mater. 2005, 17, 6656–6665. [Google Scholar] [CrossRef]
  39. Livraghi, S.; Paganini, M.C.; Giamello, E.; Selloni, A.; Di Valentin, C.; Pacchioni, G. Origin of Photoactivity of Nitrogen-Doped Titanium Dioxide Under Visible Light. J. Am. Chem. Soc. 2006, 128, 15666–15671. [Google Scholar] [CrossRef] [PubMed]
  40. Yang, K.; Dai, Y.; Huang, B.; Han, S. Theoretical Study of N-Doped TiO2 Rutile Crystals. J. Phys. Chem. B 2006, 110, 24011–24014. [Google Scholar] [CrossRef] [PubMed]
  41. Di Valentin, C.; Finazzi, E.; Pacchioni, G.; Selloni, A.; Livraghi, S.; Paganini, M.C.; Giamello, E. N-doped TiO2: Theory and Experiment. Chem. Phys. 2007, 339, 44–56. [Google Scholar] [CrossRef]
  42. Yang, K.; Dai, Y.; Huang, B. Study of Nitrogen-Concentration Influence on N-doped TiO2 Anatase from First-Principles. J. Phys. Chem. C 2007, 111, 12086–12090. [Google Scholar] [CrossRef]
  43. Yang, K.; Dai, Y.; Huang, B. Understanding Photocatalytic Activity of S- and P-doped TiO2 Under Visible Light from First-Principles. J. Phys. Chem. C 2007, 111, 18985–18994. [Google Scholar] [CrossRef]
  44. Yang, K.; Dai, Y.; Huang, B. First-Principles Calculations for Geometrical Structures and Electronic Properties of Si-doped TiO2. Chem. Phys. Lett. 2008, 456, 71–75. [Google Scholar] [CrossRef]
  45. Yang, K.; Dai, Y.; Huang, B.; Whangbo, M.H. Density Functional Characterization of the Band Edges, the Band Gap States, and the Preferred Doping Sites of Halogen-Doped TiO2. Chem. Mater. 2008, 20, 6528–6534. [Google Scholar] [CrossRef]
  46. Long, R.; Dai, Y.; Meng, G.; Huang, B. Energetic and Electronic Properties of X- (Si, Ge, Sn, Pb) Doped TiO2 from First-Principles. Phys. Chem. Chem. Phys. 2009, 11, 8165–8172. [Google Scholar] [CrossRef] [PubMed]
  47. Shi, W.; Chen, Q.; Xu, Y.; Wu, D.; Huo, C.f. Investigation of the Silicon Concentration Effect on Si-doped Anatase TiO2 by First-Principles Calculation. J. Solid State Chem. 2011, 184, 1983–1988. [Google Scholar] [CrossRef]
  48. Matsushima, S.; Takehara, K.; Yamane, H.; Yamada, K.; Nakamura, H.; Arai, M.; Kobayashi, K. First-Principles Energy Band Calculation for Undoped and S-doped TiO2 with Anatase Structure. J. Phys. Chem. Solids 2007, 68, 206–210. [Google Scholar] [CrossRef]
  49. Zheng, J.W.; Bhattcahrayya, A.; Wu, P.; Chen, Z.; Highfield, J.; Dong, Z.; Xu, R. The Origin of Visible Light Absorption in Chalcogen Element (S, Se, and Te)-Doped Anatase TiO2 Photocatalysts. J. Phys. Chem. C 2010, 114, 7063–7069. [Google Scholar] [CrossRef]
  50. Yu, J.C.; Ho, W.; Yu, J.; Yip, H.; Wong, P.K.; Zhao, J. Efficient Visible-Light-Induced Photocatalytic Disinfection on Sulfur-Doped Nanocrystalline Titania. Environ. Sci. Technol. 2005, 39, 1175–1179. [Google Scholar] [CrossRef] [PubMed]
  51. Ohno, T.; Akiyoshi, M.; Umebayashi, T.; Asai, K.; Mitsui, T.; Matsumura, M. Preparation of S-doped TiO2 Photocatalysts and Their Photocatalytic Activities Under Visible Light. Appl. Catal. A Gen. 2004, 265, 115–121. [Google Scholar] [CrossRef]
  52. Ohno, T.; Mitsui, T.; Matsumura, M. Photocatalytic Activity of S-doped TiO2 Photocatalyst Under Visible Light. Chem. Lett. 2003, 32, 364. [Google Scholar] [CrossRef] [Green Version]
  53. Umebayashi, T.; Yamaki, T.; Itoh, H.; Asai, K. Band Gap Narrowing of Titanium Dioxide by Sulfur Doping. Appl. Phys. Lett. 2002, 81, 454–456. [Google Scholar] [CrossRef]
  54. Umebayashi, T.; Yamaki, T.; Tanaka, S.; Asai, K. Visible Light-Induced Degradation of Methylene Blue on S-doped TiO2. Chem. Lett. 2003, 32, 330. [Google Scholar] [CrossRef]
  55. Long, R.; Dai, Y.; Huang, B. Structural and Electronic Properties of Iodine-Doped Anatase and Rutile TiO2. Comput. Mater. Sci. 2009, 45, 223–228. [Google Scholar] [CrossRef]
  56. Ma, X.; Dai, Y.; Wei, W.; Huang, B.; Whangbo, M.H. Insights Into How Fluorine-Adsorption and n-Type Doping Affect the Relative Stability of the (001) and (101) Surfaces of TiO2: Enhancing the Exposure of More Active but Thermodynamically Less Stable (001). J. Phys. Chem. Lett. 2015, 6, 1876–1882. [Google Scholar] [CrossRef] [PubMed]
  57. Ma, X.; Dai, Y.; Guo, M.; Huang, B. Relative Photooxidation and Photoreduction Activities of the 100, 101, and 001 Surfaces of Anatase TiO2. Langmuir 2013, 29, 13647–13654. [Google Scholar] [CrossRef] [PubMed]
  58. Ma, X.; Dai, Y.; Guo, M.; Huang, B. Insights Into the Role of Surface Distortion in Promoting the Separation and Transfer of Photogenerated Carriers in Anatase TiO2. J. Phys. Chem. C 2013, 117, 24496–24502. [Google Scholar] [CrossRef]
  59. Reeves, K.G.; Dambournet, D.; Laberty-Robert, C.; Vuilleumier, R.; Salanne, M. A First-Principles Computational Comparison of Defect-Free and Disordered, Fluorinated Anatase TiO2 (001) Interfaces with Water. RSC Adv. 2020, 10, 8982–8988. [Google Scholar] [CrossRef] [Green Version]
  60. Di Valentin, C.; Finazzi, E.; Pacchioni, G.; Selloni, A.; Livraghi, S.; Czoska, A.M.; Paganini, M.C.; Giamello, E. Density Functional Theory and Electron Paramagnetic Resonance Study on the Effect of N-F Codoping of TiO2. Chem. Mater. 2008, 20, 3706–3714. [Google Scholar] [CrossRef]
  61. Geng, H.; Yin, S.; Yang, X.; Shuai, Z.; Liu, B. Geometric and Electronic Structures of the Boron-Doped Photocatalyst TiO2. J. Phys. Condens. Matter 2006, 18, 87–96. [Google Scholar] [CrossRef]
  62. Finazzi, E.; Di Valentin, C.; Pacchioni, G. Boron-Doped Anatase TiO2: Pure and Hybrid DFT Calculations. J. Phys. Chem. C 2008, 113, 220–228. [Google Scholar] [CrossRef]
  63. Lee, J.; Park, J.; Cho, J. Electronic Properties of N- and C-doped TiO2. Appl. Phys. Lett. 2005, 87, 011904. [Google Scholar] [CrossRef]
  64. Yang, K.; Dai, Y.; Huang, B. Origin of the Photoactivity in B-doped Anatase and Rutile TiO2 from First Principles. Phys. Rev. B 2007, 76, 195201. [Google Scholar] [CrossRef]
  65. Yang, K.; Dai, Y.; Huang, B.; Whangbo, M.H. Density Functional Characterization of the Visible-Light Absorption in Substitutional C-anion and C-cation Doped TiO2. J. Phys. Chem. C 2009, 113, 2624–2629. [Google Scholar] [CrossRef]
  66. Kilic, C.; Zunger, A. N-Type Doping of Oxides by Hydrogen. Appl. Phys. Lett. 2002, 81, 73–75. [Google Scholar] [CrossRef] [Green Version]
  67. Peacock, P.W.; Robertson, J. Behavior of Hydrogen in High Dielectric Constant Oxide Gate Insulators. Appl. Phys. Lett. 2003, 83, 2025–2027. [Google Scholar] [CrossRef]
  68. Robertson, J.; Peacock, P.W. Doping and Hydrogen in Wide Gap Oxides. Thin Solid Film. 2003, 445, 155–160. [Google Scholar] [CrossRef]
  69. Xiong, K.; Robertson, J.; Clark, S.J. Behavior of Hydrogen in Wide Band Gap Oxides. J. Appl. Phys. 2007, 102, 083710-13. [Google Scholar] [CrossRef] [Green Version]
  70. 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]
  71. Lindgren, T.; Mwabora, J.M.; Avendaño, E.; Jonsson, J.; Hoel, A.; Granqvist, C.G.; Lindquist, S.E. Photoelectrochemical and Optical Properties of Nitrogen Doped Titanium Dioxide Films Prepared by Reactive DC Magnetron Sputtering. J. Phys. Chem. B 2003, 107, 5709–5716. [Google Scholar] [CrossRef]
  72. Diwald, O.; Thompson, T.L.; Zubkov, T.; Goralski, E.G.; Walck, S.D.; Yates, J.T. Photochemical Activity of Nitrogen-Doped Rutile TiO2(110) in Visible Light. J. Phys. Chem. B 2004, 108, 6004–6008. [Google Scholar] [CrossRef]
  73. Sakthivel, S.; Janczarek, M.; Kisch, H. Visible Light Activity and Photoelectrochemical Properties of Nitrogen-Doped TiO2. J. Phys. Chem. B 2004, 108, 19384–19387. [Google Scholar] [CrossRef]
  74. Torres, G.R.; Lindgren, T.; Lu, J.; Granqvist, C.G.; Lindquist, S.E. Photoelectrochemical Study of Nitrogen-Doped Titanium Dioxide for Water Oxidation. J. Phys. Chem. B 2004, 108, 5995–6003. [Google Scholar] [CrossRef]
  75. Matsui, H.; Tabata, H.; Hasuike, N.; Harima, H.; Mizobuchi, B. Epitaxial Growth and Characteristics of N-doped Anatase TiO2 Films Grown Using a Free-Radical Nitrogen Oxide Source. J. Appl. Phys. 2005, 97, 123511. [Google Scholar] [CrossRef]
  76. Nakano, Y.; Morikawa, T.; Ohwaki, T.; Taga, Y. Deep-Level Optical Spectroscopy Investigation of N-doped TiO2 Films. Appl. Phys. Lett. 2005, 86, 132104. [Google Scholar] [CrossRef]
  77. Batzill, M.; Morales, E.H.; Diebold, U. Influence of Nitrogen Doping on the Defect Formation and Surface Properties of TiO2 Rutile and Anatase. Phys. Rev. Lett. 2006, 96, 026103. [Google Scholar] [CrossRef]
  78. Kitano, M.; Funatsu, K.; Matsuoka, M.; Ueshima, M.; Anpo, M. Preparation of Nitrogen-Substituted TiO2 Thin Film Photocatalysts by the Radio Frequency Magnetron Sputtering Deposition Method and Their Photocatalytic Reactivity Under Visible Light Irradiation. J. Phys. Chem. B 2006, 110, 25266–25272. [Google Scholar] [CrossRef]
  79. Wang, Y.; Feng, C.; Jin, Z.; Zhang, J.; Yang, J.; Zhang, S. A Novel N-doped TiO2 with High Visible Light Photocatalytic Activity. J. Mater. Chem. A 2006, 260, 1–3. [Google Scholar] [CrossRef]
  80. Borrás, A.; López, C.; Rico, V.; Gracia, F.; González-Elipe, A.R.; Richter, E.; Battiston, G.; Gerbasi, R.; McSporran, N.; Sauthier, G.; et al. Effect of Visible and UV Illumination on the Water Contact Angle of TiO2 Thin Films with Incorporated Nitrogen. J. Phys. Chem. C 2007, 111, 1801–1808. [Google Scholar] [CrossRef]
  81. Cong, Y.; Zhang, J.; Chen, F.; Anpo, M. Synthesis and Characterization of Nitrogen-Doped TiO2 Nanophotocatalyst with High Visible Light Activity. J. Phys. Chem. C 2007, 111, 6976–6982. [Google Scholar] [CrossRef]
  82. Livraghi, S.; Czoska, A.M.; Paganini, M.C.; Giamello, E. Preparation and Spectroscopic Characterization of Visible Light Sensitized N Doped TiO2 (rutile). J. Solid State Chem. 2009, 182, 160–164. [Google Scholar] [CrossRef]
  83. Ihara, T.; Miyoshi, M.; Iriyama, Y.; Matsumoto, O.; Sugihara, S. Visible-Light-Active Titanium Oxide Photocatalyst Realized by an Oxygen-Deficient Structure and by Nitrogen Doping. Appl. Catal. B Environ. 2003, 42, 403–409. [Google Scholar] [CrossRef]
  84. Zhao, Z.; Liu, Q. Mechanism of Higher Photocatalytic Activity of Anatase TiO2 Doped with Nitrogen Under Visible-Light Irradiation from Density Functional Theory Calculation. J. Phys. D Appl. Phys. 2008, 41, 025105. [Google Scholar] [CrossRef]
  85. Okato, T.; Sakano, T.; Obara, M. Suppression of Photocatalytic Efficiency in Highly N-doped Anatase Films. Phys. Rev. B 2005, 72, 115124. [Google Scholar] [CrossRef]
  86. Di Valentin, C.; Pacchioni, G.; Selloni, A. Origin of the Different Photoactivity of N-doped Anatase and Rutile TiO2. Phys. Rev. B 2004, 70, 085116. [Google Scholar] [CrossRef]
  87. Di Valentin, C.; Pacchioni, G.; Selloni, A.; Livraghi, S.; Giamello, E. Characterization of Paramagnetic Species in N-Doped TiO2 Powders by EPR Spectroscopy and DFT Calculations. J. Phys. Chem. B 2005, 109, 11414–11419. [Google Scholar] [CrossRef]
  88. Xu, T.H.; Song, C.L.; Liu, Y.; Han, G.R. Band Structures of TiO2 Doped with N, C and B. J. Zhejiang Univ. Sci. B 2006, 7, 299–303. [Google Scholar] [CrossRef] [Green Version]
  89. Ma, X.; Dai, Y.; Huang, B. Origin of the Increased Photocatalytic Performance of TiO2 Nanocrystal Composed of Pure Core and Heavily Nitrogen-Doped Shell: A Theoretical Study. ACS Appl. Mater. Interfaces 2014, 6, 22815–22822. [Google Scholar] [CrossRef]
  90. Lin, Z.; Orlov, A.; Lambert, R.M.; Payne, M.C. New Insights Into the Origin of Visible Light Photocatalytic Activity of Nitrogen-Doped and Oxygen-Deficient Anatase TiO2. J. Phys. Chem. B 2005, 109, 20948–20952. [Google Scholar] [CrossRef]
  91. 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]
  92. Kuznetsov, V.N.; Serpone, N. Visible Light Absorption by Various Titanium Dioxide Specimens. J. Phys. Chem. B 2006, 110, 25203–25209. [Google Scholar] [CrossRef] [PubMed]
  93. Graciani, J.; Ortega, Y.; Fdez Sanz, J. Carbon Doping of the TiO2 (110) Rutile Surface. A Theoretical Study Based on DFT. Chem. Mater. 2009, 21, 1431–1438. [Google Scholar] [CrossRef]
  94. Long, R.; English, N.J.; Dai, Y. First-Principles Study of S Doping at the Rutile TiO2 (110) Surface. J. Phys. Chem. C 2009, 113, 17464–17470. [Google Scholar] [CrossRef] [Green Version]
  95. Long, R.; English, N.J. Energetic and Electronic Properties of P Doping at the Rutile TiO2 (110) Surface from First Principles. J. Phys. Chem. C 2009, 113, 9423–9430. [Google Scholar] [CrossRef] [Green Version]
  96. Kavan, L.; Grätzel, M.; Gilbert, S.E.; Klemenz, C.; Scheel, H.J. Electrochemical and Photoelectrochemical Investigation of Single-Crystal Anatase. J. Am. Chem. Soc. 1996, 118, 6716–6723. [Google Scholar] [CrossRef]
  97. Pascual, J.; Camassel, J.; Mathieu, H. Resolved Quadrupolar Transition in TiO2. Phys. Rev. Lett. 1977, 39, 1490–1493. [Google Scholar] [CrossRef] [Green Version]
  98. Diwald, O.; Thompson, T.L.; Goralski, E.G.; Walck, S.D.; Yates, J.T. The Effect of Nitrogen Ion Implantation on the Photoactivity of TiO2 Rutile Single Crystals. J. Phys. Chem. B 2004, 108, 52–57. [Google Scholar] [CrossRef]
  99. Umebayashi, T.; Yamaki, T.; Yamamoto, S.; Miyashita, A.; Tanaka, S. Sulfur-Doping of Rutile-Titanium Dioxide by Ion Implantation: Photocurrent Spectroscopy and First-Principles Band Calculation Studies. J. Appl. Phys. 2003, 93, 5156–5160. [Google Scholar] [CrossRef]
  100. Ho, W.; Yu, J.C.; Lee, S. Low-Temperature Hydrothermal Synthesis of S-doped TiO2 with Visible Light Photocatalytic Activity. J. Solid State Chem. 2006, 179, 1171–1176. [Google Scholar] [CrossRef]
  101. Li, H.; Zhang, X.; Huo, Y.; Zhu, J. Supercritical Preparation of a Highly Active S-Doped TiO2 Photocatalyst for Methylene Blue Mineralization. Environ. Sci. Technol. 2007, 41, 4410–4414. [Google Scholar] [CrossRef] [PubMed]
  102. Sun, H.; Bai, Y.; Cheng, Y.; Jin, W.; Xu, N. Preparation and Characterization of Visible-Light-Driven Carbon-Sulfur-Codoped TiO2 Photocatalysts. Ind. Eng. Chem. Res. 2006, 45, 4971–4976. [Google Scholar] [CrossRef]
  103. Ohno, T. Preparation of Visible Light Active S-doped TiO2 Photocatalysts and Their Photocatalytic Activities. Water Sci. Technol. 2004, 49, 159–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Yamamoto, T.; Yamashita, F.; Tanaka, I.; Matsubara, E.; Muramatsu, A. Electronic States of Sulfur Doped TiO2 by First Principles Calculations. Mater. Trans. 2004, 45, 1987–1990. [Google Scholar] [CrossRef]
  105. Sathish, M.; Sankaran, M.; Viswanathan, B.; Viswanath, R.P. DFT Studies on Anionic Hetero Atom (N Or/and S) Substitution in TiO2. Indian J. Chem. A 2007, 46, 895–898. [Google Scholar]
  106. Tian, F.; Liu, C. DFT Description on Electronic Structure and Optical Absorption Properties of Anionic S-Doped Anatase TiO2. J. Phys. Chem. B 2006, 110, 17866–17871. [Google Scholar] [CrossRef]
  107. Shi, Q.; Yang, D.; Jiang, Z.; Li, J. Visible-Light Photocatalytic Regeneration of NADH Using P-doped TiO2 Nanoparticles. J. Mol. Catal. Enzym. 2006, 43, 44–48. [Google Scholar] [CrossRef]
  108. Yu, H.F. Photocatalytic Abilities of Gel-Derived P-doped TiO2. J. Phys. Chem. Solids 2007, 68, 600–607. [Google Scholar] [CrossRef]
  109. Lin, L.; Lin, W.; Zhu, Y.; Zhao, B.; Xie, Y. Phosphor-Doped Titania-Novel Photocatalyst Active in Visible Light. Chem. Lett. 2005, 34, 284. [Google Scholar] [CrossRef]
  110. Yu, J.C.; Zheng, Z.; Zhao, J. Synthesis and Characterization of Phosphated Mesoporous Titanium Dioxide with High Photocatalytic Activity. Chem. Mater. 2003, 15, 2280–2286. [Google Scholar] [CrossRef]
  111. Jung, K.Y.; Park, S.B.; Ihm, S.K. Local Structure and Photocatalytic Activity of B2O3-SiO2/TiO2 Ternary Mixed Oxides Prepared by Sol-Gel Method. Appl. Catal. B Environ. 2004, 51, 239–245. [Google Scholar] [CrossRef]
  112. Chen, D.; Yang, D.; Wang, Q.; Jiang, Z. Effects of Boron Doping on Photocatalytic Activity and Microstructure of Titanium Dioxide Nanoparticles. Ind. Eng. Chem. Res. 2006, 45, 4110–4116. [Google Scholar] [CrossRef]
  113. Zhao, W.; Ma, W.; Chen, C.; Zhao, J.; Shuai, Z. Efficient Degradation of Toxic Organic Pollutants with Ni2O3/TiO2−xBx Under Visible Irradiation. J. Am. Chem. Soc. 2004, 126, 4782–4783. [Google Scholar] [CrossRef] [PubMed]
  114. Zhukov, V.P.; Zainullina, V.M.; Chulkov, E.V. Ab Initio Approach Struct. Electron. Opt. Prop. B- C- N-Doped Anatase. Int. J. Mod. Phys. B 2010, 24, 6049–6067. [Google Scholar] [CrossRef]
  115. Yang, K.; Dai, Y.; Huang, B. Density Functional Study of Boron-Doped Anatase TiO2. J. Phys. Chem. C 2010, 114, 19830–19834. [Google Scholar] [CrossRef]
  116. Pankove, J.I. Optical Processes in Semiconductors; Dover Publications: New York, NY, USA, 1975; p. 189. [Google Scholar]
  117. Moss, T.S. The Interpretation of the Properties of Indium Antimonide. Proc. Phys. Soc. B 1954, 67, 775. [Google Scholar] [CrossRef]
  118. Burstein, E. Anomalous Optical Absorption Limit in InSb. Phys. Rev. 1954, 93, 632–633. [Google Scholar] [CrossRef]
  119. Jin, H.; Dai, Y.; Wei, W.; Huang, B. Density Functional Characterization of B Doping at Rutile TiO2 (110) Surface. J. Phys. D Appl. Phys. 2008, 41, 195411. [Google Scholar] [CrossRef]
  120. Feng, N.; Zheng, A.; Wang, Q.; Ren, P.; Gao, X.; Liu, S.B.; Shen, Z.; Chen, T.; Deng, F. Boron Environments in B-doped and (B, N)-codoped TiO2 Photocatalysts: A Combined Solid-State NMR and Theoretical Calculation Study. J. Phys. Chem. C 2011, 115, 2709–2719. [Google Scholar] [CrossRef]
  121. Shen, M.; Wu, Z.; Huang, H.; Du, Y.; Zou, Z.; Yang, P. Carbon-Doped Anatase TiO2 Obtained from TiC for Photocatalysis Under Visible Light Irradiation. Mater. Lett. 2006, 60, 693–697. [Google Scholar] [CrossRef]
  122. Hsu, S.W.; Yang, T.S.; Chen, T.K.; Wong, M.S. Ion-Assisted Electron-Beam Evaporation of Carbon-Doped Titanium Oxide Films as Visible-Light Photocatalyst. Thin Solid Film. 2007, 515, 3521–3526. [Google Scholar] [CrossRef]
  123. Wang, S.H.; Chen, T.K.; Rao, K.K.; Wong, M.S. Nanocolumnar Titania Thin Films Uniquely Incorporated with Carbon for Visible Light Photocatalysis. Appl. Catal. B Environ. 2007, 76, 328–334. [Google Scholar] [CrossRef]
  124. Wong, M.S.; Wang, S.H.; Chen, T.K.; Weng, C.W.; Rao, K.K. Co-Sputtered Carbon-Incorporated Titanium Oxide Films as Visible Light-Induced Photocatalysts. Surf. Coat. Technol. 2007, 202, 890–894. [Google Scholar] [CrossRef]
  125. Wu, G.; Nishikawa, T.; Ohtani, B.; Chen, A. Synthesis and Characterization of Carbon-Doped TiO2 Nanostructures with Enhanced Visible Light Response. Chem. Mater. 2007, 19, 4530–4537. [Google Scholar] [CrossRef]
  126. Choi, Y.; Umebayashi, T.; Yoshikawa, M. Fabrication and Characterization of C-doped Anatase TiO2 Photocatalysts. J. Mater. Sci. 2004, 39, 1837–1839. [Google Scholar] [CrossRef]
  127. Irie, H.; Watanabe, Y.; Hashimoto, K. Carbon-Doped Anatase TiO2 Powders as a Visible-Light Sensitive Photocatalyst. Chem. Lett. 2003, 32, 772–773. [Google Scholar] [CrossRef]
  128. Irie, H.; Washizuka, S.; Hashimoto, K. Hydrophilicity on Carbon-Doped TiO2 Thin Films Under Visible Light. Thin Solid Film. 2006, 510, 21–25. [Google Scholar] [CrossRef]
  129. Huang, Y.; Ho, W.; Lee, S.; Zhang, L.; Li, G.; Yu, J.C. Effect of Carbon Doping on the Mesoporous Structure of Nanocrystalline Titanium Dioxide and Its Solar-Light-Driven Photocatalytic Degradation of NOx. Langmuir 2008, 24, 3510–3516. [Google Scholar] [CrossRef]
  130. Yang, X.; Cao, C.; Hohn, K.; Erickson, L.; Maghirang, R.; Hamal, D.; Klabunde, K. Highly Visible-Light Active C- and V-doped TiO2 for Degradation of Acetaldehyde. J. Catal. 2007, 0, 1–7. [Google Scholar] [CrossRef]
  131. Mohapatra, S.; Misra, M.; Mahajan, V.; Raja, K. A Novel Method for the Synthesis of Titania Nanotubes Using Sonoelectrochemical Method and Its Application for Photoelectrochemical Splitting of Water. J. Catal. 2007, 246, 362–369. [Google Scholar] [CrossRef]
  132. Mohapatra, S.K.; Misra, M.; Mahajan, V.K.; Raja, K.S. Design of a Highly Efficient Photoelectrolytic Cell for Hydrogen Generation by Water Splitting: Application of TiO2−xCx Nanotubes as a Photoanode and Pt/TiO2 Nanotubes as a Cathode. J. Phys. Chem. C 2007, 111, 8677–8685. [Google Scholar] [CrossRef]
  133. Park, J.H.; Kim, S.; Bard, A.J. Novel Carbon-Doped TiO2 Nanotube Arrays with High Aspect Ratios for Efficient Solar Water Splitting. Nano Lett. 2006, 6, 24–28. [Google Scholar] [CrossRef] [PubMed]
  134. Xu, C.; Killmeyer, R.; Gray, M.L.; Khan, S.U. Enhanced Carbon Doping of n-TiO2 Thin Films for Photoelectrochemical Water Splitting. Electrochem. Commun. 2006, 8, 1650–1654. [Google Scholar] [CrossRef]
  135. Kamisaka, H.; Adachi, T.; Yamashita, K. Theoretical Study of the Structure and Optical Properties of Carbon-Doped Rutile and Anatase Titanium Oxides. J. Chem. Phys. 2005, 123, 084704. [Google Scholar] [CrossRef] [PubMed]
  136. Ren, W.; Ai, Z.; Jia, F.; Zhang, L.; Fan, X.; Zou, Z. Low Temperature Preparation and Visible Light Photocatalytic Activity of Mesoporous Carbon-Doped Crystalline TiO2. Appl. Catal. B Environ. 2007, 69, 138–144. [Google Scholar] [CrossRef]
  137. Tian, F.; Liu, C.; Zhang, D.; Fu, A.; Duan, Y.; Yuan, S.; Yu, J.C. On the Origin of the Visible-Light Activity of Titanium Dioxide Doped with Carbonate Species. ChemPhysChem 2010, 11, 3269–3272. [Google Scholar] [CrossRef]
  138. Oh, S.M.; Kim, S.S.; Lee, J.E.; Ishigaki, T.; Park, D.W. Effect of Additives on Photocatalytic Activity of Titanium Dioxide Powders Synthesized by Thermal Plasma. Thin Solid Film. 2003, 435, 252–258. [Google Scholar] [CrossRef]
  139. Yan, X.; He, J.; Evans, D.G.; Duan, X.; Zhu, Y. Preparation, Characterization and Photocatalytic Activity of Si-doped and Rare Earth-Doped TiO2 from Mesoporous Precursors. Appl. Catal. B Environ. 2005, 55, 243–252. [Google Scholar] [CrossRef]
  140. Ozaki, H.; Iwamoto, S.; Inoue, M. Enhanced Visible Light Sensitivity of Nitrogen-Doped Nanocrystalline Si-Modified Titania Prepared by the Glycothermal Method. Chem. Lett. 2005, 34, 1082–1083. [Google Scholar] [CrossRef]
  141. Ozaki, H.; Iwamoto, S.; Inoue, M. Effect of the Addition of a Small Amount of Vanadium on the Photocatalytic Activities of N- and Si- Co-Doped Titanias Under Visible-Light Irradiation. Catal. Lett. 2007, 113, 95–98. [Google Scholar] [CrossRef]
  142. Chang, J.; Jiang, Z.Y.; Zhang, Z.Y.; Lin, Y.M.; Tian, P.L.; Zhou, B.; Chen, L. Theoretical Studies of Photocatalytic Behaviors of Isoelectronic C/Si/Ge/Sn-doped TiO2: DFT+U. Appl. Surf. Sci. 2019, 484, 1304–1309. [Google Scholar] [CrossRef]
  143. Hattori, A.; Tada, H. High Photocatalytic Activity of F-Doped TiO2 Film on Glass. J. Sol-Gel Sci. Technol. 2001, 22, 47. [Google Scholar] [CrossRef]
  144. Wu, G.; Wang, J.; Thomas, D.F.; Chen, A. Synthesis of F-Doped Flower-Like TiO2 Nanostructures with High Photoelectrochemical Activity. Langmuir 2008, 24, 3503–3509. [Google Scholar] [CrossRef] [PubMed]
  145. Yu, J.C.; Yu, J.; Ho, W.; Jiang, Z.; Zhang, L. Effects of F-Doping on the Photocatalytic Activity and Microstructures of Nanocrystalline TiO2 Powders. Chem. Mater. 2002, 14, 3808–3816. [Google Scholar] [CrossRef]
  146. Su, Y.; Zhang, X.; Han, S.; Chen, X.; Lei, L. F-B-codoping of Anodized TiO2 Nanotubes Using Chemical Vapor Deposition. Electrochem. Commun. 2007, 9, 2292–2299. [Google Scholar] [CrossRef]
  147. Yamaki, T.; Sumita, T.; Yamamoto, S. Formation of TiO2−xFx Compounds in Fluorine-Implanted TiO2. J. Mater. Sci. Lett. 2002, 21, 33–35. [Google Scholar] [CrossRef]
  148. Yamaki, T.; Umebayashi, T.; Sumita, T.; Yamamoto, S.; Maekawa, M.; Kawasuso, A.; Itoh, H. Fluorine-Doping in Titanium Dioxide by Ion Implantation Technique. Nucl. Instrum. Methods Phys. Res. Sect. B 2003, 206, 254–258. [Google Scholar] [CrossRef]
  149. Li, D.; Haneda, H.; Hishita, S.; Ohashi, N.; Labhsetwar, N.K. Fluorine-Doped TiO2 Powders Prepared by Spray Pyrolysis and Their Improved Hotocatalytic Activity for Decomposition of Gas-Phase Acetaldehyde. J. Fluor. Chem. 2005, 126, 69–77. [Google Scholar] [CrossRef]
  150. Li, D.; Haneda, H.; Hishita, S.; Ohashi, N. Visible-Light-Driven N-F-Codoped TiO2 Photocatalysts. 1. Synthesis by Spray Pyrolysis and Surface Characterization. Chem. Mater. 2005, 17, 2588–2595. [Google Scholar] [CrossRef]
  151. Li, D.; Haneda, H.; Hishita, S.; Ohashi, N. Visible-Light-Driven N-F-Codoped TiO2 Photocatalysts. 2. Optical Characterization, Hotocatalysis, and Potential Application to Air Purification. Chem. Mater. 2005, 17, 2596–2602. [Google Scholar] [CrossRef]
  152. Li, D.; Haneda, H.; Labhsetwar, N.K.; Hishita, S.; Ohashi, N. Visible-Light-Driven Photocatalysis on Fluorine-Doped TiO2 Powders by the Creation of Surface Oxygen Vacancies. Chem. Phys. Lett. 2005, 401, 579–584. [Google Scholar] [CrossRef]
  153. Zhou, J.K.; Lv, L.; Yu, J.; Li, H.L.; Guo, P.Z.; Sun, H.; Zhao, X.S. Synthesis of Self-Organized Polycrystalline F-doped TiO2 Hollow Microspheres and Their Photocatalytic Activity Under Visible Light. J. Phys. Chem. C 2008, 112, 5316–5321. [Google Scholar] [CrossRef]
  154. Huang, D.; Liao, S.; Liu, J.; Dang, Z.; Petrik, L. Preparation of Visible-Light Responsive N-F-codoped TiO2 Photocatalyst by a Sol-Gel-Solvothermal Method. J. Photochem. Photobiol. A 2006, 184, 282–288. [Google Scholar] [CrossRef]
  155. Huang, D.; Liao, S.; Quan, S.; Liu, L.; He, Z.; Wan, J.; Zhou, W. Preparation of Anatase F Doped TiO2 Sol and Its Performance for Photodegradation of Formaldehyde. J. Mater. Sci. 2007, 42, 8193–8202. [Google Scholar] [CrossRef]
  156. Tang, J.; Quan, H.; Ye, J. Photocatalytic Properties and Photoinduced Hydrophilicity of Surface-Fluorinated TiO2. Chem. Mater. 2007, 19, 116–122. [Google Scholar] [CrossRef]
  157. Luo, H.; Takata, T.; Lee, Y.; Zhao, J.; Domen, K.; Yan, Y. Photocatalytic Activity Enhancing for Titanium Dioxide by Co-Doping with Bromine and Chlorine. Chem. Mater. 2004, 16, 846–849. [Google Scholar] [CrossRef]
  158. Hong, X.; Wang, Z.; Cai, W.; Lu, F.; Zhang, J.; Yang, Y.; Ma, N.; Liu, Y. Visible-Light-Activated Nanoparticle Photocatalyst of Iodine-Doped Titanium Dioxide. Chem. Mater. 2005, 17, 1548–1552. [Google Scholar] [CrossRef]
  159. Liu, G.; Chen, Z.; Dong, C.; Zhao, Y.; Li, F.; Lu, G.; Cheng, H. Visible Light Photocatalyst: Iodine-Doped Mesoporous Titania with a Bicrystalline Framework. J. Phys. Chem. B 2006, 110, 20823–20828. [Google Scholar] [CrossRef]
  160. Guo, M.L.; Zhang, X.D.; Liang, C.T.; Jia, G.Z. Mechanism of Visible Photoactivity of F-doped TiO2. Chin. Phys. Lett. 2010, 27, 057103. [Google Scholar] [CrossRef]
  161. Czoska, A.M.; Livraghi, S.; Chiesa, M.; Giamello, E.; Agnoli, S.; Granozzi, G.; Finazzi, E.; Di Valentin, C.; Pacchioni, G. The Nature of Defects in Fluorine-Doped TiO2. J. Phys. Chem. C 2008, 112, 8951–8956. [Google Scholar] [CrossRef]
  162. Finazzi, E.; Di Valentin, C.; Pacchioni, G.; Selloni, A. Excess Electron States in Reduced Bulk Anatase TiO2: Comparison of Standard GGA, GGA+U, Hybrid DFT Calc. J. Chem. Phys. 2008, 129, 154113. [Google Scholar] [CrossRef] [PubMed]
  163. Long, R.; English, N.J. Electronic Properties of F/Zr Co-Doped Anatase TiO2 Photocatalysts from GGA+U Calc. Chem. Phys. Lett. 2010, 498, 338–344. [Google Scholar] [CrossRef]
  164. Sato, J.; Kobayashi, H.; Inoue, Y. Photocatalytic Activity for Water Decomposition of Indates with Octahedrally Coordinated D10 Configuration. II. Roles Geom. Electron. Struct. J. Phys. Chem. B 2003, 107, 7970–7975. [Google Scholar] [CrossRef]
  165. Long, M.; Cai, W.; Wang, Z.; Liu, G. Correlation of Electronic Structures and Crystal Structures with Photocatalytic Properties of Undoped, N-doped and I-doped TiO2. Chem. Phys. Lett. 2006, 420, 71–76. [Google Scholar] [CrossRef]
  166. Kuhn, W.; Wagner, H.P.; Stanzl, H.; Wolf, K.; Worle, K.; Lankes, S.; Betz, J.; Worz, M.; Lichtenberger, D.; Leiderer, H.; et al. The MOVPE Growth and Doping of ZnTe. Semicond. Sci. Technol. 1991, 6, A105–A108. [Google Scholar] [CrossRef]
  167. Pan, H.; Zhang, Y.W.; Shenoy, V.B.; Gao, H. Effects of H-, N-, and (H, N)-Doping on the Photocatalytic Activity of TiO2. J. Phys. Chem. C 2011, 115, 12224–12231. [Google Scholar] [CrossRef]
  168. Chen, W.P.; Wang, Y.; Chan, H.L.W. Hydrogen: A Metastable Donor in TiO2 Single Crystals. Appl. Phys. Lett. 2008, 92, 112907. [Google Scholar] [CrossRef] [Green Version]
  169. Herklotz, F.; Lavrov, E.V.; Weber, J. Infrared Absorption of the Hydrogen Donor in Rutile TiO2. Phys. Rev. B 2011, 83, 235202. [Google Scholar] [CrossRef]
  170. Yang, K.; Dai, Y.; Huang, B.; Feng, Y. Density Functional Characterization of the Antiferromagnetism in Oxygen-Deficient Anatase and Rutile TiO2. Phys. Rev. B 2010, 81, 033202. [Google Scholar] [CrossRef] [Green Version]
  171. Mi, L.; Xu, P.; Shen, H.; Wang, P.N.; Shen, W. First-Principles Calculation of N:H Codoping Effect on Energy Gap Narrowing of TiO2. Appl. Phys. Lett. 2007, 90, 171909. [Google Scholar] [CrossRef]
  172. Li, N.; Yao, K.L.; Li, L.; Sun, Z.Y.; Gao, G.Y.; Zhu, L. Effect of Carbon/Hydrogen Species Incorporation on Electronic Structure of anatase-TiO2. J. Appl. Phys. 2011, 110, 073513. [Google Scholar] [CrossRef]
  173. Yang, K.; Dai, Y.; Huang, B.; Whangbo, M.H. Density Functional Studies of the Magnetic Properties in Nitrogen Doped TiO2. Chem. Phys. Lett. 2009, 481, 99–102. [Google Scholar] [CrossRef]
  174. Long, R.; English, N.J. Band Gap Engineering of (N,Ta)-codoped TiO2: A First-Principles Calculation. Chem. Phys. Lett. 2009, 478, 175–179. [Google Scholar] [CrossRef] [Green Version]
  175. Yang, K.; Dai, Y.; Huang, B.; Whangbo, M.H. On the Possibility of Ferromagnetism in Carbon-Doped Anatase TiO2. Appl. Phys. Lett. 2008, 93, 132507. [Google Scholar] [CrossRef]
  176. 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] [PubMed]
  177. Lu, J.; Dai, Y.; Jin, H.; Huang, B. Effective Increasing of Optical Absorption and Energy Conversion Efficiency of Anatase TiO2 Nanocrystals by Hydrogenation. Phys. Chem. Chem. Phys. 2011, 13, 18063–18068. [Google Scholar] [CrossRef] [PubMed]
  178. Yu, H.; Irie, H.; Hashimoto, K. Conduction Band Energy Level Control of Titanium Dioxide: Toward an Efficient Visible-Light-Sensitive Photocatalyst. J. Am. Chem. Soc. 2010, 132, 6898–6899. [Google Scholar] [CrossRef]
  179. Gai, Y.; Li, J.; Li, S.S.; Xia, J.B.; Wei, S.H. Design of Narrow-Gap TiO2: A Passivated Codoping Approach for Enhanced Photoelectrochemical Activity. Phys. Rev. Lett. 2009, 102, 036402. [Google Scholar] [CrossRef]
  180. Long, R.; English, N.J. Tailoring the Electronic Structure of TiO2 by Cation Codoping from Hybrid Density Functional Theory Calculations. Phys. Rev. B 2011, 83, 155209. [Google Scholar] [CrossRef]
  181. Zhu, Y.; Dai, Y.; Lai, K.; Huang, B. Synergistic Modification of Electronic and Photocatalytic Properties of TiO2 Nanotubes by Implantation of Au and N Atoms. ChemPhysChem 2013, 14, 2800–2807. [Google Scholar] [CrossRef]
  182. Wu, G.; Wen, J.; Nigro, S.; Chen, A. One-Step Synthesis of N- and F-codoped Mesoporous TiO2 Photocatalysts with High Visible Light Activity. Nanotechnology 2010, 21, 085701. [Google Scholar] [CrossRef] [PubMed]
  183. Zhang, X.; Zhang, J.; Cui, H. Adsorption Mechanism of SF6 Decomposition Components Onto N, F-co-doped TiO2: A DFT Study. J. Fluor. Chem. 2018, 213, 18–23. [Google Scholar] [CrossRef]
  184. Salehi-Abar, P.; Kazempour, A. Effect of N and F Doping on the Electronic Properties of Rutile TiO2 Quantum Dot Solar Cells: A First Principle Study. Chem. Phys. Lett. 2017, 673, 56–61. [Google Scholar] [CrossRef]
  185. Lu, J.; Jin, H.; Dai, Y.; Yang, K.; Huang, B. Effect of Electronegativity and Charge Balance on the Visible-Light-Responsive Photocatalytic Activity of Nonmetal Doped Anatase TiO2. Int. J. Photoenergy 2012, 2012. [Google Scholar] [CrossRef] [Green Version]
  186. Obata, K.; Irie, H.; Hashimoto, K. Enhanced Photocatalytic Activities of Ta, N Co-Doped TiO2 Thin Films Under Visible Light. Chem. Phys. 2007, 339, 124–132. [Google Scholar] [CrossRef]
  187. Long, R.; English, N.J. Synergistic Effects on Band Gap-Narrowing in Titania by Codoping from First-Principles Calculations. Chem. Mater. 2010, 22, 1616–1623. [Google Scholar] [CrossRef] [Green Version]
  188. Liu, H.; Lu, Z.; Yue, l.; Liu, J.; Gan, Z.; Shu, C.; Zhang, T.; Shi, J.; Xiong, R. (Mo + N) Codoped TiO2 for Enhanced Visible-Light Photoactivity. Appl. Sur. Sci. 2011, 257, 9355–9361. [Google Scholar] [CrossRef]
  189. Wei, W.; Dai, Y.; Guo, M.; Yu, L.; Huang, B. Density Functional Characterization of the Electronic Structure and Optical Properties of N-doped, La-doped, and N/La-codoped SrTiO3. J. Phys. Chem. C 2009, 113, 15046–15050. [Google Scholar] [CrossRef]
  190. Wei, W.; Dai, Y.; Guo, M.; Yu, L.; Jin, H.; Han, S.; Huang, B. Codoping Synergistic Effects in N-doped SrTiO3 for Higher Energy Conversion Efficiency. Phys. Chem. Chem. Phys. 2010, 12, 7612–7619. [Google Scholar] [CrossRef]
  191. Sulaeman, U.; Yin, S.; Sato, T. Visible Light Photocatalytic Properties of Ta and N Codoped SrTiO3 Nanoparticles. Int. J. Opt. 2010, 2010. [Google Scholar] [CrossRef] [Green Version]
  192. Fang, Y.; Xu, T.; Zhang, Y.; Kong, X.; Liu, J.; Cui, S.; Wang, D. Structural, Electronic and Optical Properties of La, C-Codoped TiO2 Investigated by First Principle Calculations. J. Phys. Chem. Solids 2019, 132, 121–129. [Google Scholar] [CrossRef]
  193. Tao, J.G.; Guan, L.X.; Pan, J.S.; Huan, C.H.A.; Wang, L.; Kuo, J.L. Possible Room Temperature Ferromagnetism of Li-doped Anatase TiO2: A First-Principles Study. Phys. Lett. A 2010, 374, 4451–4454. [Google Scholar] [CrossRef]
  194. Colella, S.; Mosconi, E.; Pellegrino, G.; Alberti, A.; Guerra, V.L.; Masi, S.; Listorti, A.; Rizzo, A.; Condorelli, G.G.; De Angelis, F.; et al. Elusive Presence of Chloride in Mixed Halide Perovskite Solar Cells. J. Phys. Chem. Lett. 2014, 5, 3532–3538. [Google Scholar] [CrossRef] [PubMed]
  195. De Angelis, F. Modeling Materials and Processes in Hybrid/Organic Photovoltaics: From Dye-Sensitized to Perovskite Solar Cells. Acc. Chem. Res. 2014, 47, 3349–3360. [Google Scholar] [CrossRef] [PubMed]
  196. Roiati, V.; Mosconi, E.; Listorti, A.; Colella, S.; Gigli, G.; De Angelis, F. Stark Effect in Perovskite/TiO2 Solar Cells: Evidence of Local Interfacial Order. Nano Lett. 2014, 14, 2168–2174. [Google Scholar] [CrossRef] [Green Version]
  197. Cheng, J.; Luo, J.; Yang, K. Comparison Studies of Interfacial Electronic and Energetic Properties of LaAlO3/TiO2 and TiO2/LaAlO3 Heterostructures from First-Principles Calculations. ACS Appl. Mater. Interfaces 2017, 9, 7682–7690. [Google Scholar] [CrossRef] [PubMed]
  198. Chen, Y.; Lei, Y.; Li, Y.; Yu, Y.; Cai, J.; Chiu, M.H.; Rao, R.; Gu, Y.; Wang, C.; Choi, W.; et al. Strain Engineering and Epitaxial Stabilization of Halide Perovskites. Nature 2020, 577, 209–215. [Google Scholar] [CrossRef]
  199. Qu, Z.; Su, Y.L.; Sun, L.; Liang, F.; Zhang, G.H. Study of the Structure, Electronic and Optical Properties of BiOI/Rutile-TiO2 Heterojunction by the First-Principle Calculation. Materials 2020, 13, 323. [Google Scholar] [CrossRef] [Green Version]
  200. Wei, W.; Dai, Y.; Huang, B.; Li, X.; Nägele, F.; Over, H.; Whangbo, M.H.; Jacob, T. Density Functional Characterization of the Electronic Structures and Band Bending of Rutile RuO2/TiO2(110) Heterostructures. J. Phys. Chem. C 2015, 119, 12394–12399. [Google Scholar] [CrossRef]
  201. Zhou, Y.D.; Liu, Q.L.; Yang, C.; Zhao, Z.Y. Interfacial Micro-Structure and Properties of TiO2/SnO2 Heterostructures with Rutile Phase: A DFT Calculation Investigation. Appl. Surf. Sci. 2018, 451, 258–271. [Google Scholar] [CrossRef]
  202. Pak, Y.; Park, W.; Alaal, N.; Kumaresan, Y.; Aravindh, S.A.; Mitra, S.; Xin, B.; Min, J.W.; Kim, H.; Lim, N.; et al. Enhanced Photoresponse of WS2 Photodetectors through Interfacial Defect Engineering Using a TiO2 Interlayer. ACS Appl. Electron. Mater. 2020, 2, 838–845. [Google Scholar] [CrossRef] [Green Version]
  203. Abbasi, A.; Sardroodi, J.J. Investigation of the Adsorption of Ozone Molecules on TiO2/WSe2 Nanocomposites by DFT Computations: Applications to Gas Sensor Devices. Appl. Surf. Sci. 2018, 436, 27–41. [Google Scholar] [CrossRef]
  204. Zeng, X.; Xiao, X.; Zhang, W.; Wan, C.; Wang, H. Interfacial Charge Transfer and Mechanisms of Enhanced Photocatalysis of an Anatase TiO2(001)-MoS2-graphene Nanocomposite: A First-Principles Investigation. Comput. Mater. Sci. 2017, 126, 43–51. [Google Scholar] [CrossRef]
  205. Liberto, G.D.; Tosoni, S.; Pacchioni, G. Charge Carriers Cascade in a Ternary TiO2/TiO2/ZnS Heterojunction: A DFT Study. ChemCatChem 2020, 12, 2097–2105. [Google Scholar] [CrossRef]
  206. Zhao, Y.; Lin, Y.; Wang, G.; Jiang, Z.; Zhang, R.; Zhu, C. Electronic and Optical Performances of (Cu, N) Codoped TiO2/g-C3N4 Heterostructure Photocatalyst: A Spin-Polarized DFT+U Study. Sol. Energy 2018, 162, 306–316. [Google Scholar] [CrossRef]
  207. Wei, L.; Cui, S.; Guo, H.; Ma, X. Study on the Role of Mn Species in Low Temperature SCR on MnOx/TiO2 through Experiment and DFT Calculation. Mol. Catal. 2018, 445, 102–110. [Google Scholar] [CrossRef]
  208. Meng, Q.; Fan, L.; Zhu, L.; Xu, N.; Zhang, Q. Electronic and Optical Properties of α-MoO3/TiO2 Heterostructures: A DFT Study. Int. J. Quantum Chem. 2018, 118, e25681. [Google Scholar] [CrossRef]
  209. Long, R.; Dai, Y.; Huang, B. Fullerene Interfaced with a TiO2(110) Surface May Not Form an Efficient Photovoltaic Heterojunction: First-Principles Investigation of Electronic Structures. J. Phys. Chem. Lett. 2013, 4, 2223–2229. [Google Scholar] [CrossRef]
  210. Ma, X.; Dai, Y.; Yu, L.; Huang, B. New Basic Insights Into the Low Hot Electron Injection Efficiency of Gold-Nanoparticle-Photosensitized Titanium Dioxide. ACS Appl. Mater. Interfaces 2014, 6, 12388–12394. [Google Scholar] [CrossRef]
  211. Jin, C.; Dai, Y.; Wei, W.; Ma, X.; Li, M.; Huang, B. Effects of Single Metal Atom (Pt, Pd, Rh and Ru) Adsorption on the Photocatalytic Properties of Anatase TiO2. Appl. Surf. Sci. 2017, 426, 639–646. [Google Scholar] [CrossRef]
  212. Ma, X.C.; Dai, Y.; Yu, L.; Huang, B.B. Energy Transfer in Plasmonic Photocatalytic Composites. Light Sci. Appl. 2016, 5, e16017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Clavero, C. Plasmon-Induced Hot-Electron Generation at Nanoparticle/metal-Oxide Interfaces for Photovoltaic and Photocatalytic Devices. Nat. Photonics 2014, 8, 95–103. [Google Scholar] [CrossRef]
  214. Cheng, H.; Fuku, K.; Kuwahara, Y.; Mori, K.; Yamashita, H. Harnessing Single-Active Plasmonic Nanostructures for Enhanced Photocatalysis Under Visible Light. J. Mater. Chem. A 2015, 3, 5244–5258. [Google Scholar] [CrossRef]
  215. Lou, Z.; Wang, Z.; Huang, B.; Dai, Y. Synthesis and Activity of Plasmonic Photocatalysts. ChemCatChem 2014, 6, 2456–2476. [Google Scholar] [CrossRef]
  216. Zhang, X.; Chen, Y.L.; Liu, R.S.; Tsai, D.P. Plasmonic Photocatalysis. Rep. Prog. Phys. 2013, 76, 046401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. 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]
  218. Hou, W.; Cronin, S.B. A Review of Surface Plasmon Resonance-Enhanced Photocatalysis. Adv. Funct. Mater. 2013, 23, 1612–1619. [Google Scholar] [CrossRef]
  219. Jiang, R.; Li, B.; Fang, C.; Wang, J. Metal/Semiconductor Hybrid Nanostructures for Plasmon-Enhanced Applications. Adv. Mater. 2014, 26, 5274–5309. [Google Scholar] [CrossRef]
  220. Brongersma, M.L.; Halas, N.J.; Nordlander, P. Plasmon-Induced Hot Carrier Science and Technology. Nat. Nanotechnol. 2015, 10, 25–34. [Google Scholar] [CrossRef]
  221. Yang, K.; Setyawan, W.; Wang, S.; Nardelli, M.B.; Curtarolo, S. A Search Model for Topological Insulators with High-Throughput Robustness Descriptors. Nat. Mater. 2012, 11, 614–619. [Google Scholar] [CrossRef]
  222. Curtarolo, S.; Hart, G.L.W.; Nardelli, M.B.; Mingo, N.; Sanvito, S.; Levy, O. The High-Throughput Highway to Computational Materials Design. Nat. Mater. 2013, 12, 191–201. [Google Scholar] [CrossRef] [PubMed]
  223. Yang, K. High-Throughput Design of Functional Materials Using Materials Genome Approach. Chin. Phys. B 2018, 27, 128103. [Google Scholar] [CrossRef]
  224. Cai, Y.; Xie, W.; Teng, Y.T.; Harikesh, P.C.; Ghosh, B.; Huck, P.; Persson, K.A.; Mathews, N.; Mhaisalkar, S.G.; Sherburne, M.; et al. High-Throughput Computational Study of Halide Double Perovskite Inorganic Compounds. Chem. Mater. 2019, 31, 5392–5401. [Google Scholar] [CrossRef] [Green Version]
  225. Yang, K.; Nazir, S.; Behtash, M.; Cheng, J. High-Throughput Design of Two-Dimensional Electron Gas Systems Based on Polar/Nonpolar Perovskite Oxide Heterostructures. Sci. Rep. 2016, 6, 34667. [Google Scholar] [CrossRef] [PubMed]
  226. Cheng, J.; Yang, K. Design of Two-Dimensional Electron Gas Systems via Polarization Discontinuity from Large-Scale First-Principles Calculations. J. Mater. Chem. C 2018, 6, 6680–6690. [Google Scholar] [CrossRef]
  227. Li, Y.; Yang, K. High-Throughput Computational Design of Organic-Inorganic Hybrid Halide Semiconductors Beyond Perovskites for Optoelectronics. Energy Environ. Sci. 2019, 12, 2233–2243. [Google Scholar] [CrossRef]
  228. Li, Y.; Maldonado-Lopez, D.; Ríos Vargas, V.; Zhang, J.; Yang, K. Stability Diagrams, Defect Tolerance, and Absorption Coefficients of Hybrid Halide Semiconductors: High-Throughput First-Principles Characterization. J. Chem. Phys. 2020, 152, 084106. [Google Scholar] [CrossRef]
Figure 1. Illustration of adsorption configurations of 2,5-DMBQ onto the surface of (TiO2)n (n = 1–6) clusters. The left panel (a) shows the absorption pattern of the quinone ring on to the TiO2 clusters and the right panel (b) shows the absorption pattern of the ketonic O (H) of the 2,5-DMBQ absorbed on the Ti (O) atoms of TiO2 clusters. Reprinted with permission from Reference [26]. Copyright (2020) Elsevier.
Figure 1. Illustration of adsorption configurations of 2,5-DMBQ onto the surface of (TiO2)n (n = 1–6) clusters. The left panel (a) shows the absorption pattern of the quinone ring on to the TiO2 clusters and the right panel (b) shows the absorption pattern of the ketonic O (H) of the 2,5-DMBQ absorbed on the Ti (O) atoms of TiO2 clusters. Reprinted with permission from Reference [26]. Copyright (2020) Elsevier.
Catalysts 10 00972 g001
Figure 2. Illustration of (a) geometrical structure, (b) electronic structure diagram, and (c) spin density surface of Ti8O16 cluster. Reprinted with permission from the authors of [35]. Copyright (2012) American Chemical Society.
Figure 2. Illustration of (a) geometrical structure, (b) electronic structure diagram, and (c) spin density surface of Ti8O16 cluster. Reprinted with permission from the authors of [35]. Copyright (2012) American Chemical Society.
Catalysts 10 00972 g002
Figure 3. (Color online) Calculated formation energies of substitutional nonmetal-doped TiO2 as a function of the oxygen chemical potential ( μ O ). (a) B- and carbon group- (C, Si, and Ge) doped TiO2. (b) Nitrogen group- (N, P, As, and Sb) doped TiO2. (c) Chalcogen- (S, Se, and Te) doped TiO2. (d) Halogen- (F, Cl, Br, and I) doped TiO2. X@O and X@Ti represent substitutional X for O and substitutional X for Ti doped structure, respectively. The Pauling electronegativities of these nonmetal dopants are given inside the parentheses.
Figure 3. (Color online) Calculated formation energies of substitutional nonmetal-doped TiO2 as a function of the oxygen chemical potential ( μ O ). (a) B- and carbon group- (C, Si, and Ge) doped TiO2. (b) Nitrogen group- (N, P, As, and Sb) doped TiO2. (c) Chalcogen- (S, Se, and Te) doped TiO2. (d) Halogen- (F, Cl, Br, and I) doped TiO2. X@O and X@Ti represent substitutional X for O and substitutional X for Ti doped structure, respectively. The Pauling electronegativities of these nonmetal dopants are given inside the parentheses.
Catalysts 10 00972 g003
Figure 4. (Color online) (a) Total and projected density of states (DOS) plots calculated for undoped anatase TiO2. (b) Total DOS plot calculated for C-cation-doped anatase TiO2. (c) Projected DOS plots calculated for the C s and C 2p states of C-cation-doped anatase TiO2. (d) Zoomed-in views of the projected DOS plots of the C 2s/2p states and those of the O 2s/2p states for one of the two O atoms forming the linear O-C-O unit. The energy zero represents the valence band maximum (VBM) of the undoped and C-cation-doped anatase TiO2, the vertical dashed line the conduction band minimum (CBM) of undoped anatase TiO2, and the vertical dotted line the CBM of C-cation-doped anatase TiO2. Reprinted with permission from Reference [65]. Copyright (2009) American Chemical Society.
Figure 4. (Color online) (a) Total and projected density of states (DOS) plots calculated for undoped anatase TiO2. (b) Total DOS plot calculated for C-cation-doped anatase TiO2. (c) Projected DOS plots calculated for the C s and C 2p states of C-cation-doped anatase TiO2. (d) Zoomed-in views of the projected DOS plots of the C 2s/2p states and those of the O 2s/2p states for one of the two O atoms forming the linear O-C-O unit. The energy zero represents the valence band maximum (VBM) of the undoped and C-cation-doped anatase TiO2, the vertical dashed line the conduction band minimum (CBM) of undoped anatase TiO2, and the vertical dotted line the CBM of C-cation-doped anatase TiO2. Reprinted with permission from Reference [65]. Copyright (2009) American Chemical Society.
Catalysts 10 00972 g004
Figure 5. (Color online) Comparison of the calculated VBM and CBM positions of X-doped anatase TiO2 with X@O (X = F, Cl, Br, and I) from the DFT calculations with the corresponding experimental values of undoped TiO2. The VBM and CBM values are given with respect to the normal hydrogen electrode (NHE) potential. Reprinted with permission from the authors of [45]. Copyright (2008) American Chemical Society.
Figure 5. (Color online) Comparison of the calculated VBM and CBM positions of X-doped anatase TiO2 with X@O (X = F, Cl, Br, and I) from the DFT calculations with the corresponding experimental values of undoped TiO2. The VBM and CBM values are given with respect to the normal hydrogen electrode (NHE) potential. Reprinted with permission from the authors of [45]. Copyright (2008) American Chemical Society.
Catalysts 10 00972 g005
Figure 6. (Color online) Illustration of relative positions of substitutional N anion and Mo cation in TiO2: (a) Non-passivated doping and (b) passivated doping, and calculated TDOSs of (Mo + N) co-doped TiO2 (Ti 1 x Mo x O 2 y N y ): (c) non-passivated and (d) passivated. The red dash lines represent the Fermi level. Reprinted with permission from the authors of [188]. Copyright (2011) Elsevier.
Figure 6. (Color online) Illustration of relative positions of substitutional N anion and Mo cation in TiO2: (a) Non-passivated doping and (b) passivated doping, and calculated TDOSs of (Mo + N) co-doped TiO2 (Ti 1 x Mo x O 2 y N y ): (c) non-passivated and (d) passivated. The red dash lines represent the Fermi level. Reprinted with permission from the authors of [188]. Copyright (2011) Elsevier.
Catalysts 10 00972 g006
Figure 7. Illustration of (a) calculated charge displacement and (b) geometrical structure for TiO2/MAPbI3 (red) and TiO2/MAPbI 3 x Cl x (blue). Reprinted with permission from the authors of [21]. Copyright (2014) American Chemical Society.
Figure 7. Illustration of (a) calculated charge displacement and (b) geometrical structure for TiO2/MAPbI3 (red) and TiO2/MAPbI 3 x Cl x (blue). Reprinted with permission from the authors of [21]. Copyright (2014) American Chemical Society.
Catalysts 10 00972 g007
Figure 8. Relaxed geometrical structures of (a) MAI/R, (b) MAI/A, (c) SnI2/R, and (d) SnI2/A TiO2/MASnI3 interface. Reprinted with permission from the authors of [24]. Copyright (2018) Elsevier.
Figure 8. Relaxed geometrical structures of (a) MAI/R, (b) MAI/A, (c) SnI2/R, and (d) SnI2/A TiO2/MASnI3 interface. Reprinted with permission from the authors of [24]. Copyright (2018) Elsevier.
Catalysts 10 00972 g008
Figure 9. Calculated electronic band structure of (a) 1I-BiOI/TiO2 and (b) BiO-BiOI/TiO2 heterojunction and (c) electrostatic potential of 1I-BiOI/TiO2 model. Reprinted with permission from the authors of [199]. Copyright (2018) Elsevier.
Figure 9. Calculated electronic band structure of (a) 1I-BiOI/TiO2 and (b) BiO-BiOI/TiO2 heterojunction and (c) electrostatic potential of 1I-BiOI/TiO2 model. Reprinted with permission from the authors of [199]. Copyright (2018) Elsevier.
Catalysts 10 00972 g009
Table 2. List of the acceptor atoms and donor atoms in TiO2. The numbers in the third column represent the introduced holes per acceptor atom ( N h ) or donated electrons per donor atom ( N e ). H i n t and Li i n t represent the interstitial H and Li ions, respectively.
Table 2. List of the acceptor atoms and donor atoms in TiO2. The numbers in the third column represent the introduced holes per acceptor atom ( N h ) or donated electrons per donor atom ( N e ). H i n t and Li i n t represent the interstitial H and Li ions, respectively.
Acceptor Atom N h
AnionN 2 , P 2 1[43,173]
C 2 2[175]
B 2 3[115]
CationB 3 + , Al 3 + , Ga 3 + , In 3 + , La 3 + 1[115,192]
Be 2 + , Mg 2 + , Ca 2 + , Sr 2 + , Ba 2 + 2[115]
Li + , Na + , K + , Rb + , Cs + 3[115,193]
Donor atom N e
AnionF , Cl , Br , I 1[45,161]
CationH i n t , Li i n t 1[66,67,68,171]
I 5 + , Nb 5 + , Ta 5 + 1[45,174]
Mo 6 + , W 6 + 2[180]
Table 3. Summary of TiO2-based materials interfaces along with their critical interfacial properties. a TiO2 and r TiO2 denote anatase and rutile TiO2, respectively.
Table 3. Summary of TiO2-based materials interfaces along with their critical interfacial properties. a TiO2 and r TiO2 denote anatase and rutile TiO2, respectively.
InterfacePropertiesRef.
a TiO2/MAPbI3 &
a TiO2/MAPbI 3 x Clx
(110) perovskite surface was stabilized against (001) surface after
“depositing” TiO2 for both materials interfaces;
interfacial Cl atoms increase the interfacial binding energy.
[21]
r TiO2/MAPbI3Higher interfacial charge transfer rate than ZnO- and SnO2-based interface.[22]
a TiO2/MASnI3 &
r TiO2/MASnI3
a TiO2/SnI2 interface is energetically most favorable among the
four considered systems and has the highest electron–hole separation rate.
[24]
a TiO2/MAPbI3 &
r TiO2/MAPbI3
a TiO2/PbI2 has stronger interfacial interaction but r TiO2/PbI2 model
is most efficient for electron–hole separation.
[19]
a TiO2/(Ba,Sr)TiO2Stability of anatase TiO 2 decreases in the order
from (001) to (011) to (111) perovskite substrates.
[23]
r TiO2/BiOX
(X = Cl,Br, and I)
BiO/TiO2 interface is more stable than 1I/TiO2 interface;
band gap of 1I/TiO2 interface reduced by 0.28 eV
while BiO/TiO2 exhibits an n-type conductivity.
[199]
r TiO2/RuO2Strong bonding interaction at the interface;
oxygen vacancies at the interface changes the band bending direction.
[200]

Share and Cite

MDPI and ACS Style

Yang, K.; Dai, Y.; Huang, B. Review of First-Principles Studies of TiO2: Nanocluster, Bulk, and Material Interface. Catalysts 2020, 10, 972. https://doi.org/10.3390/catal10090972

AMA Style

Yang K, Dai Y, Huang B. Review of First-Principles Studies of TiO2: Nanocluster, Bulk, and Material Interface. Catalysts. 2020; 10(9):972. https://doi.org/10.3390/catal10090972

Chicago/Turabian Style

Yang, Kesong, Ying Dai, and Baibiao Huang. 2020. "Review of First-Principles Studies of TiO2: Nanocluster, Bulk, and Material Interface" Catalysts 10, no. 9: 972. https://doi.org/10.3390/catal10090972

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