Next Article in Journal
The Isorhamnetin-Containing Fraction of Philippine Honey Produced by the Stingless Bee Tetragonula biroi Is an Antibiotic against Multidrug-Resistant Staphylococcus aureus
Previous Article in Journal
Raman Enhancement of Nanoparticle Dimers Self-Assembled Using DNA Origami Nanotriangles
Previous Article in Special Issue
An Insight into the Excitation States of Small Molecular Semiconductor Y6
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

TiO2 as a Photocatalyst for Water Splitting—An Experimental and Theoretical Review

by
Håkon Eidsvåg
1,*,
Said Bentouba
1,
Ponniah Vajeeston
2,
Shivatharsiny Yohi
3 and
Dhayalan Velauthapillai
1,*
1
Department of Computing, Mathematics and Physics, Western Norway University of Applied Sciences, Inndalsveien 28, Box 5063, N-5009 Bergen, Norway
2
Center for Materials Science and Nanotechnology, Department of Chemistry, University of Oslo, Box 1033 Blindern, N-0315 Oslo, Norway
3
Department of Chemistry, Faculty of Science, University of Jaffna, Sir. Pon, Ramanathan Rd, Jaffna 40000, Sri Lanka
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(6), 1687; https://doi.org/10.3390/molecules26061687
Submission received: 2 February 2021 / Revised: 9 March 2021 / Accepted: 10 March 2021 / Published: 17 March 2021
(This article belongs to the Special Issue Advanced Materials for Solar Energy)

Abstract

:
Hydrogen produced from water using photocatalysts driven by sunlight is a sustainable way to overcome the intermittency issues of solar power and provide a green alternative to fossil fuels. TiO2 has been used as a photocatalyst since the 1970s due to its low cost, earth abundance, and stability. There has been a wide range of research activities in order to enhance the use of TiO2 as a photocatalyst using dopants, modifying the surface, or depositing noble metals. However, the issues such as wide bandgap, high electron-hole recombination time, and a large overpotential for the hydrogen evolution reaction (HER) persist as a challenge. Here, we review state-of-the-art experimental and theoretical research on TiO2 based photocatalysts and identify challenges that have to be focused on to drive the field further. We conclude with a discussion of four challenges for TiO2 photocatalysts—non-standardized presentation of results, bandgap in the ultraviolet (UV) region, lack of collaboration between experimental and theoretical work, and lack of large/small scale production facilities. We also highlight the importance of combining computational modeling with experimental work to make further advances in this exciting field.

1. Introduction

Over the last years, there has been a steadily increasing focus on clean, renewable energy sources as a priority to hinder the irreversible climate change the world is facing and to meet the continuously growing energy demand [1]. One hour of solar energy can satisfy the energy consumption of the whole world for a year [2]. Hence, direct harvesting of solar light and its conversion into electrical energy with photovoltaic cells or chemical energy by photoelectrochemical reactions are the most relevant technologies to overcome this challenge. Conventionally, both technologies rely on the collection of light in semiconductor materials with appropriate bandgaps matching the solar spectrum, and thus providing a high-energy conversion efficiency.
Unfortunately, the technology has drawbacks, which prevent it from overtaking non-renewable energy as the main energy source. A major issue is the uneven power distribution caused by varying solar radiation and a lack of proper storage alternatives. As a solution to this problem, the focus is moving toward research on storage options for the produced electricity, which we can divide into mechanical and electrochemical storage systems. For example, in Oceania, pumped hydroelectricity (mechanical) is the most common storage system for excess electricity [3]. Different batteries (lithium–ion, sodium–sulfur (S), vanadium, etc.), hydrogen fuel cells, and supercapacitors are the current focus areas for electrochemical storage [3]. There are several reasons for choosing hydrogen as a way to store solar energy, namely, (1) there is a high abundance of hydrogen from renewable sources; (2) it is eco-friendly when used; (3) hydrogen has a high-energy yield, and (4) it is easy to store as either a gas or a liquid [4,5,6].
The high energy yield and ease of storage make hydrogen viable as fuel for the long transport sector; airplanes, cruise ships, trailers, and cargo ships [7,8]. The realization of a green energy shipping fleet could alone yearly cut 2.5% of global greenhouse emissions (GHG) [9]. However, to succeed in this strategy, hydrogen must be produced in a clean and renewable way.
As water splitting got the attention of the researchers in the 1970s, titanium dioxide became the most prominent photocatalyst used [10]. There are several good reasons for this: low cost, chemical stability, earth abundance, and nontoxicity [11]. However, TiO2 also sports a wide bandgap (3.0–3.2 eV), which reduces the potential for absorption of visible light [11]. Due to TiO2s structural and chemical properties, it is possible to engineer the bandgap, light absorption properties, recombination time, etc. by increasing the active sites and improving the electrical conductivity [12]. TiO2 exists in several different polymorphs that all behave differently. The most common ones are rutile, brookite, and anatase as shown in Figure 1. Rutile and anatase TiO2 are the most used polymorphs for photocatalytic water splitting; nevertheless, some attempts with amorphous TiO2 (aTiO2) have been made as shown in Figure 2.
Figure 1. Crystal structures of TiO2 rutile (tetragonal, P42/mmm), brookite (orthorhombic, Pbca), and anatase (tetragonal, I41/amd) polymorphs. Reused with permission from [13].
Figure 1. Crystal structures of TiO2 rutile (tetragonal, P42/mmm), brookite (orthorhombic, Pbca), and anatase (tetragonal, I41/amd) polymorphs. Reused with permission from [13].
Molecules 26 01687 g001
Figure 2. The structure of 72-atom (left), 96-atom (middle), and 216-atoms (right) models of amorphous TiO2. The red and grey spheres represent O and Ti atoms respectively. Reused with permission from [14]. Copyright 2012, with permission from Elsevier.
Figure 2. The structure of 72-atom (left), 96-atom (middle), and 216-atoms (right) models of amorphous TiO2. The red and grey spheres represent O and Ti atoms respectively. Reused with permission from [14]. Copyright 2012, with permission from Elsevier.
Molecules 26 01687 g002
Several attempts have been made to introduce dopants to improve the optical absorption of TiO2. For example, Zhang et al. found that 12.5% copper (Cu)-doped anatase TiO2 showed a broader absorption peak than pure anatase titanium dioxide [15]. Through a theoretical study, they found that Cu introduces an unoccupied impurity continuum band at the top of the valence band, which explains the improved optical absorption. Another theoretical study, conducted by Morgade and Cabeza in 2017, shows that co-doping of TiO2 with (Pt, V) and (C, N) narrows the bandgap and favorably modifies the position of the valence and conduction band edges [16]. Our study will provide an insight into current theoretical and computational studies carried out on water splitting using either pure or doped TiO2 semiconductors. In addition, we will compare it with the state of art experimental studies conducted within the field. Our aim is to help bridge the gap between theoretical simulations and experimental research. The two approaches complement each other and when combined could support the field moving forward toward the realization of the hydrogen economy. The theoretical study allows testing of the properties of thousands of different materials with different parameters to gain an understanding of how and why certain dopants and material combinations work. However, the computational models are worked out using perturbation theory, which lowers the overall accuracy of the results. At the same time, an experimental study is important to verify the theoretical results and find the best methods to synthesize the materials in practice.

2. Solar-Driven Hydrogen Production

Most of the commercial production of hydrogen stems from four sources: natural gas, coal, oil, and electrolysis. Of these, steam reforming alone stands for 48% of the world’s hydrogen production, while coal contributes 18%, oil 30%, and electrolysis 4% [17]. The first three hydrogen production processes are energy-consuming and use non-renewable energy sources, which is unattractive for environment protection and climate change [18,19]. However, the production of hydrogen by electrolysis requires only water and electrical current. To have green hydrogen, produced friendly to the environment, we propose to use renewable energy sources—wind, hydro, and solar power—to produce the electric current needed for the electrolysis of water. Solar power is ideal due to the high amount of incoming energy. There are several functional methods used in driving the electrolysis process, i.e., thermochemical water splitting [20], photo-biological water splitting [21], and photocatalytic water splitting [22]. Furthermore, photocatalytic water splitting (PWS) is considered the best option, due to the following reasons: (1) PWS has a good solar-hydrogen conversion efficiency, (2) it has a low production cost, (3) oxygen and hydrogen can easily be separated during the PWS process, and (4) hydrogen electrolysis could be used on both small- and large-scale facilities [4,22,23].

2.1. Photocatalytic Water Splitting (PWS)

The photocatalytic process splits water (H2O) into hydrogen (H2) and oxygen (O2) in the presence of a catalyst and natural light; it is an artificial photosynthesis method. Figure 3 shows a schematic illustration of the major steps involved in the process of photocatalytic water splitting. In the first step (1), electron–hole pairs are generated in the presence of irradiation. This is carried out by utilizing the semiconducting nature of the photocatalyst to excite electrons from the valence band (VB) to the conduction band (CB). Photons with energies larger than the bandgap can excite electrons from the VB to the CB. The second step (2) consists of charge separation and migration of the photogenerated electron-hole pairs. Ideally, all electrons and holes reach the surface without recombination to maximize the efficiency of the photocatalyst. In the final step (3), the electrons, which move from the CB to the surface of the catalyst participate in a reduction reaction and generate hydrogen, and the holes diffuse from the VB to the surface of the photocatalyst involved in an oxidation reaction to form oxygen. In general, the efficiency of the catalyst can be enhanced by including dopants or co-catalysts that include metals or metal oxides, such as Pt, NiO, and RuO2, which can act as the active sites via enhancing electron mobility. The redox and oxidations reactions on the surface of the photocatalyst are described by the following equations [24]:
Oxidation :   H 2 O + 2 h + 2 H + + 1 2 O 2 .
Reduction :   2 H + + 2 e H 2 .
Overall   reaction :   H 2 O + 1.23   e V H 2 + 1 2 O 2 .
Figure 3. Schematic illustration of the main photocatalytic steps for a semiconductor photocatalyst: (1) light is absorbed to generate electron–hole pairs; (2) migration of excited carriers (electrons and holes) to the surface; (3) surface reaction to produce hydrogen with electrons and oxygen with holes. Reproduced with permission from [25], Copyright 2013, with permission from Elsevier.
Figure 3. Schematic illustration of the main photocatalytic steps for a semiconductor photocatalyst: (1) light is absorbed to generate electron–hole pairs; (2) migration of excited carriers (electrons and holes) to the surface; (3) surface reaction to produce hydrogen with electrons and oxygen with holes. Reproduced with permission from [25], Copyright 2013, with permission from Elsevier.
Molecules 26 01687 g003
The process of water splitting is highly endothermic and requires a Gibbs free energy of 1.23 eV per electron, which corresponds to light with a wavelength of 1008 nm. This means that the photocatalyst must have a bandgap > 1.23 eV, or else the electrons will not have enough energy to start the reaction. In practice, this limit should be 1.6 eV to 1.8 eV due to some overpotentials [24]. Naturally, it should not be too high either, as that would reduce the amount of visible light the photocatalyst can absorb. This means that it is important to find suitable catalysts with a bandgap between 1.6–2.2 eV to ensure maximum absorption of the incoming light. Another important factor regarding the efficiency of a photocatalyst is the recombination time, i.e., the time it takes for an electron to recombine with a hole. If recombination occurs before the electrons can reach the surface and interact with the water molecules, the energy gets wasted and no redox reaction takes place. Unfortunately, there are only a few materials with sufficient recombination time and a satisfactory bandgap that have been identified. However, a recent study by Takata et. al. demonstrates that it is possible to achieve water splitting without any charge recombination losses [26]. With SrTiO3 as the photocatalyst loaded with Rh, Cr, and Co as cocatalysts, they achieved an external quantum efficiency up to 96% at wavelengths between 350 nm and 360 nm [26]. This is equivalent to having an internal quantum efficiency of almost unity. The requirements for having an efficient photocatalyst can be summarized in the solar–hydrogen conversion efficiency (STH) equation [27] as follows:
η STH = η A × η CS × η CT × η CR .  
The STH conversion efficiency depends on (1) the efficiencies of light absorption (ηA), (2) charge separation (ηCS), (3) charge transport (ηCT), and (4) charge collection/reaction efficiency (ηCR). The efficiency of the photocatalyst depends on several factors and they are elaborated in the following section.

2.2. Important Aspects of Photocatalytic Efficiency for Nanomaterials

There are several ways to improve and modify the fundamental properties of a photocatalyst by focusing on its shape, size, order, uniformity, and morphology.

2.2.1. Crystallinity

Research has shown that the crystallinity of the material affects its optoelectronic properties [28,29]. Structures with a high crystallinity perform better than amorphous variations of the same material. The increase in crystallinity reduces the number of defects in the structures and thus decreases the electron-hole recombination sites, which leads to a better catalytic activity [30,31,32,33]. Liu et al. studied the effect of crystalline TiO2 nanotubes against that of amorphous TiO2 nanotubes and found that better photocurrent properties were attained with the crystalline structures due to the lower amount of electron-hole recombination [34]. In another study, enhanced hydrogen production was obtained using extremely ordered nanotubular TiO2 arrays [35].

2.2.2. Dimensionality

Nanomaterials can be classified into four different categories depending on their dimensionality—zero-dimensional (0D), one-dimensional (1D), two-dimensional (2D), and three-dimensional (3D) [36,37]. Zero-dimensional (0D) nanostructures used in PWS are primarily quantum dots (QDs) and hollow shells. In general, QDs are used to decorate the photocatalyst because they increase the visible light absorption and reduce the electron-hole recombination [38,39,40]. One-dimensional (1D) structures include nanorods, nanotubes, and nanowires, which are all attractive for photocatalysts. It is found that nanorod and nanowire arrays result in a more efficient photogenerated electron transport and collection [41,42,43]. On the other hand, nanotubes have a higher surface area for redox reactions compared to nanorods or nanowires although they have less material for light absorption [44,45]. Two-dimensional (2D) nanostructures have a high surface area and a small thickness that reduces the travel distance for generated holes. This results in efficient light harvesting. Lastly, 3D nanostructures are promising candidates for PWS because they can be designed into high-performance photoanodes [27]. In general, it is possible to design and create nanostructures that cater to specific tasks.

2.2.3. Temperature and Pressure

Temperature and pressure during the production phase will affect the resulting properties of a photocatalyst. Research shows that by varying the pressure, the STH performance of the catalyst will change [46]. Another research group found that by using a low-temperature thermal treatment process the charge transfer resistance could be reduced [47].

2.2.4. Size

As mentioned TiO2 exist in three phases, anatase (tetragonal; a = 3.7845 Å; c = 9.5143 Å), rutile (tetragonal; a = 4.5937 Å; c = 2.9587 Å), and brookite (orthorhombic; a = 5.4558 Å; b = 9.1819 Å; c = 5.1429 Å). Among the three different crystalline phases of TiO2, anatase exhibits the highest stability for particle size less than 11 nm, whereas rutile shows thermodynamic stability for particle size greater than 35 nm, and brookite is stable in the size range of 11–35 nm. The size of the nanomaterials and cocatalyst can alter the overall performance of the system. Smaller particles are dominated by electrokinetics and are thus more suited for photocatalysis. Alternatively, larger particles are better suited for photoelectrochemical (PEC) water splitting because they have a lower electron–hole recombination rate [48]. The size of the particles also influences the electron-hole recombination time. In larger particles the travel distance to the active sites on the surface becomes longer, thus increasing the probability for electron-hole recombination. This probability is decreased in smaller particles due to the shorter migration distance [49,50].

2.2.5. Bandgap

The bandgap is one of the most important properties of the photocatalyst. It is defined as the energy needed for an electron to move from the valence band maximum (VBM) to the conduction band minimum (CBM) in a semiconductor. In addition to a fitting bandgap, the CBM must be more negative than the redox potential of H+/H2 (0 V vs. normal hydrogen electrode (NHE)), while the VBM must be more positive than the redox potential of O2/H2O (1.23 V). Therefore, the theoretical minimum bandgap for water splitting is 1.23 eV. Nanomaterials are used to tune the band positions and the bandgap toward the appropriate range of 1.6 eV to 2.2 eV [51,52,53].

2.2.6. pH Dependency

The pH value of the solution in which the photocatalyst is placed affects the end STH efficiency [54]. It will similarly affect the stability and lifetime of the catalyst. Photoelectrochemical water splitting is very dependent on the pH of the electrolyte solution, which determines the net total charge adsorbed at the surface of the catalyst. The migration of ions during the reactions may weaken the surface of the electrode. The electrode incorporated with nanomaterials exhibits better stability in different pH conditions, however, it was evident that the stability was further improved when the solution is buffered [55,56,57].

2.2.7. Light

It is important that the light source be specified, as semiconductors doped with nanomaterials can absorb both infrared and UV light in addition to visible light [58].

2.3. Theoretical Methods

Numerical studies of electronic, optical and mechanical characterization of TiO2 polymorphs are performed as ab initio calculations within the framework of density functional theory (DFT). However, the calculation model and details vary between the researchers depending on the calculation tool/code chosen; for example, Vienna ab initio simulation package (VASP) [59,60,61,62,63,64], CASTEP [65], CRYSTAL [66,67], and GPAW [68]. In general, the interaction between the core and the valence electrons is described by the projector augmented-wave method [69,70]. The electron properties are calculated by G0W0 [71,72], HSE06 [73,74], or by the generalized gradient approximation (GGA, which is less accurate but faster) [75]. It is possible to calculate the stress tensor by applying a set of strains to the crystal, which leads to the elastic constants (e.g., VASPKIT [76]). Moreover, it is possible to calculate the real space force constants of the supercell, and then to evaluate this with appropriate software (e.g., Phonopy [77]) in order to find the phonon frequencies for dynamical stability, heat capacity, free energy, and entropy analysis.

2.4. Experimental Methods

Several synthesis methods are used for the synthesis of TiO2 materials depending on the end application and experiment performed. The methods can roughly be divided into thermal reactions, deposition methods, sol–gel, Micelle, and electromagnetic methods. In general, the thermal methods are heterogeneous reactions in the presence of aqueous solvents or mineralizers under high pressure and temperature [78,79,80]. Deposition methods (e.g., chemical vapor deposition and electrodeposition) are primarily used to create thin-film materials or coatings on a substrate [81,82,83]. Sol–gel methods utilize the conversion of a liquid solution (sol) into a solid gel phase, in which the nanoparticles are formed by hydrolysis and condensation [78,84]. Micelles are the long-chain molecules made by surfactants, which contain a hydrophilic head and a hydrophobic chain. Self-assembling of these amphiphilic molecules forms an organized structure in solution [85]. The ultrasound technique can be used for preparations of a wide range of nanomaterials, especially high-surface-area transition metals, carbides, oxides, alloys, and colloids [84,86]. However, if the goal is to create a dielectric material, then the microwave method can be used [78,87].

3. Theoretical Research

The amount of theoretical and computational research has increased over the last years due to improved accuracy of the models and increased computing power. As with experimental work, most of the focus is on various dopants for TiO2. The majority of research is conducted on anatase because that phase seems most promising for water splitting. However, rutile TiO2 does have some interesting properties.

3.1. Metal Dopants

A large part of the conducted research is devoted to metal dopants and their contribution to a broadened bandgap of photocatalyst TiO2. In a comprehensive study, Pan et al. investigated how noble metals could enhance the catalytic activity of anatase TiO2 for hydrogen evolution reaction [88]. They proposed three different structural models for hydrogenated anatase TiO2, as seen in Figure 4, and proposed the preferred location (H1 in Figure 4a) of the hydrogen atom when TiO2 was noble metal doped. This is because of the strongly localized hybridization between hydrogen and TiO2 [88]. Based on these findings, they showed that anatase TiO2 was easy to hydrogenate, and the introduced hydrogen could improve the electronic transport between the conduction band and valence band near the Fermi level [88]. In general, silver (Ag)- and gold (Au)-doping are more thermodynamically stable than that of platinum (Pt)-, palladium (Pd)- and ruthenium (Ru)-doping. The band structures for noble metal-doped TiO2 are shown in Figure 5, and it is clearly seen that the introduction of dopants reduces the bandgap of TiO2. However, Ag-doping seems to be the best option for noble metal doping of TiO2 as the other dopants reduce the bandgap below 1.23 eV [88].
Y. Zhang et al. doped (001) anatase TiO2 with Pt, cobalt (Co), and Ru [89], which led to surface-localized states that enhanced the electron transfer at the surface. However, another study by S.T Zhang et al. showed that to achieve a stable interface between supported Run (n = 1–10, 20, 22) clusters and TiO2, n > 6 was needed [90]. Interestingly enough, this is among the preferred geometries for TiO2 [90].
Metal dopants decrease the bandgap, and as shown by Jin et al., Pt, Pd, rodhium (Rh), and Ru single atom doping significantly reduces the work function of the compound [91].
Another option to increase the optical absorption is to use iron (Fe) or nickel (Ni) as dopants because they will induce impurity states in the forbidden region [92]. Electrons with energy less than the bandgap can use these impurity states as steps when moving from the valence band to the conduction band. Especially co-doping of TiO2 with Fe and Ni results in higher absorption and reduced electron–hole recombination according to Lin et al. [92]. Ghuman et al. also showed that Fe2+ doped aTiO2 adsorbs water better than pristine aTiO2 and that it also has a better photocatalytic effect [93].
Introducing Cu and/or N dopants will also create isolated states in the bandgap, and therefore, TiO2 doped with these dopants function better than pure anatase TiO2 [94]. Assadi et al. showed that the improved photocatalytic activity in Cu/TiO2 was because of effective bandgap narrowing and increased charge transfer (electronic interactions) and not surface chemistry [95]. Wei Zhang et al. found that the stability of Cu doped TiO2 depends on which oxygen atoms that is replaced with Cu atoms. [15] They observed a blueshift in absorption for anatase TiO2 (101) compared to bulk TiO2, while in Cu-doped bulk anatase TiO2 they observed a redshift in optical absorption [15]. Co-doped SrTiO3 has a narrower bandgap compared to that of pure TiO2 according to Sikam et al. and this is due to states being formed in the gap [96]. In addition, they also found that co-doping resulted in magnetism due to inequality of spin down and spin up states [96]. Ghuman et al. looked into the difference between monodoping and co-doping using nitrogen (N) and niobium (Nb) on amorphous TiO2 [97]. They found that monodoping reduces the bandgap but it also increases the number of recombination centers [97]. Charge compensated co-doping, on the other hand, reduces the bandgap with 0.4 eV and suppresses the recombination effect by eliminating band gap states [97]. S and Nb co-doping of anatase TiO2 resulted in a bandgap of 2.15 eV as shown by Ren et al. [98].
However, not all states in the bandgap are appreciated; Gao et al. looked into how Mg doping could reduce the shallow defect states under the CBM in TiO2, increasing the photocatalytic effect [99].
Although several different dopants have been proposed and tested, not all of them are viable to be incorporated into TiO2. Chen et al. used cerium (Ce), praseodymium (Pr), europium (Eu), and gadolinium (Gd) dopants to see how they would incorporate with TiO2 [100]. They found that Ce was the easiest among these, while Pr and Gd had low substitutional energy and should be able to be incorporated into TiO2 [100]. Eu on the other hand was difficult to incorporate with TiO2 [100]. Ce, Pr, and Eu monodoping should move the light absorption more toward/into the visible light region [100]. Another approach when using lanthanides could be ion triple doping as showcased by Li et al. [101]. Through co- and triple-doping they improved the oxidizing ability, light absorption, and charge carrier separation of their photocatalyst (Bi2MoO6) [101].

3.2. Non-Metal Dopants

In an attempt to find non-metal dopant alternatives Shi et al. co-doped anatase TiO2 with C and neodymium (Nd) and found tuned band gaps, which were lower than that of pure TiO2: C@O and Nd@Ti, 2.372 Ev, and carbon (C) and Nd @ TiO2, 2.850 eV [102]. The main function of the C and Nd dopants is to enhance the intensity of light absorption and to extend the optical absorption range into the visible light region respectively. This is clearly seen from Figure 6, in which the optical absorption spectra of TiO2 do not extend into the visible light region, while both C doping and C and Nd doping extend the absorption range into the visible light region and effectively enhancing the efficiency of the photocatalysts. Experimental studies shown in Figure 6c confirmed these findings. By using graphite carbon spheres on TiO2 Jiang et al. introduced isolated energy levels between the VBM and the CBM [103]. These intermediate bandgaps increase the number of electrons that could be excited from the VB to the CB.
Gurkan et al. found that selenium (Se) doping introduces localized mid-gap levels that increase the visible light photocatalytic effect [104]. However, no significant change in the position of the band edges was observed. A study by Zhao et al. showed that N, Co, Co–N, Co–2N, and Co–3N dopants all increased the optical absorption rate compared to that of undoped TiO2 [105]. Unfortunately, Co–N and Co–3N shift the CBM below the H+/H2 reduction potential, which means that it is not useful for water splitting [105]. Pengfei Wang et al. revealed that carbonate could be incorporated into mesoporous TiO2 and significantly improve the visible light hydrogen evolution [106]. In addition, the intimate homo-junctions between anatase and rutile phases and graphite carbon on the surface of TiO2 can significantly help promote the separation of charge carriers [106].
Figure 6. Theoretical and experimental optical absorption spectra for pure and doped TiO2 structures. (a) Computational obtained spectra for substitutional doping, (b) computational obtained spectra for TiO2 doped with the dopants at different interstitial sites, and (c) experimental obtained optical absorption spectra for C, Nd, and C+Nd doped TiO2 from [107]. Reprinted with permission from [102]. Copyright 2017, with permission from Elsevier.
Figure 6. Theoretical and experimental optical absorption spectra for pure and doped TiO2 structures. (a) Computational obtained spectra for substitutional doping, (b) computational obtained spectra for TiO2 doped with the dopants at different interstitial sites, and (c) experimental obtained optical absorption spectra for C, Nd, and C+Nd doped TiO2 from [107]. Reprinted with permission from [102]. Copyright 2017, with permission from Elsevier.
Molecules 26 01687 g006
In an attempt to decrease the bandgap of TiO2, Zongyan et al. used boron (B), C, N, fluorine (F), phosphorus (P), S, and chlorine (Cl) as dopants on anatase TiO2 [108]. The tuned bandgaps for TiO2 with the dopants (B, C, N, P and S) were 2.72 eV, 2.44 eV, 2.74 eV, 2.38 eV, and 2.59 eV, respectively [108]. The improvements are due to new impurity energy levels, which red-shift their fundamental absorption edges to the visible light region [108]. In addition, they found that higher dipole moments could lead to easier separation of the photoexcited electron-hole pairs, which would increase the photocatalytic effect [108].
There has been some research on TiO2/g–C3N4 heterostructure photocatalysts. For example, Yali Zhao et al. co-doped TiO2/g–C3N4 with Cu and N [109]. This resulted in an obvious narrowing of the bandgap compared with pure TiO2, co-doping induces impurity states of N 2p and hybridized states of Cu 3d and N 2p in the bandgap of TiO2/g–C3N4 [109]. Yanming Lin et al. used TiO2/g–C3N4 heterostructure through interfacial coupling for H2 production. The calculated band gap was significantly reduced compared to pure TiO2 [110]. TiO2/g–C3N4 heterostructure also has a higher CBM energy, which provides the photoexcited electrons with stronger reducing power to produce more hydrogen per unit time compared with TiO2 [110].

3.3. Rutile

Although most researchers focus on anatase TiO2, there has been some development on rutile TiO2 as well. Atanelov et al. doped rutile TiO2 with C and it seems to have worse photocatalytic performances than pure TiO2, this is due to C–O dimes creating mid-band states [111]. However, N doped rutile TiO2 introduced no mid-bandgap states [111]. Ghuman et al. used Rh to dope rutile TiO2 and found that it had a lower bandgap but with more recombination centers [112]. By using Rh–Nb (charge compensated doping) as co-dopants on rutile TiO2, they obtained no isolated bandgap states that might act as a recombination center [112]. Moreover, the bandgap was decreased by 0.5 eV, which makes it a better photocatalyst [112].

3.4. Nanotubes

Research on different TiO2 nanostructures has led to some interesting results. Lisovski et al. doped TiO2 nanotubes with sulfur and achieved a narrower bandgap compared to that of pure TiO2 [113]. The band edges were also close to the limits for efficient water splitting [113]. In another article, Lisovski et al. present nitrogen (N), S, and S-and-N doping of six-layer (101) anatase TiO2 nanotubes [114]. They found that monodoping with N or S reduces the photocatalytic effect of the nanotubes [114]. However, co-doping with N and S could improve the photocatalytic activity, although it depends on the defect concentration [114]. Dmitry Bocharov et al. used TiO2 (4,4) nanotubes doped with scandium (Sc), depicted in Figure 7, and found that they were a good candidate with a bandgap of 1.8–1.9 eV [115]. Working on the same variation of nanotubes, E.P. D’yachkov et al. found that doping with Nb, molybdenum (Mo), technetium (Tc), and Pd leads to bandgaps around 2 eV [116]. This is a significant decrease from around 4 eV for undoped (4,4) TiO2 nanotubes [116]. Lisovski et al. also investigated if the arrangement of bandgap edges would change when going from bulk to nanotubes. They found that this only happened if the diameter of the nanotubes was small, i.e., the internal strain was extremely large [117].

3.5. Pure TiO2

Naturally, some focus had been on pure TiO2 and its different states, looking into which one was most suited for photocatalytic activity.
Ma et al. found that the (101) facet had higher activity compared to other facets of TiO2 [118], which was an important breakthrough. If we have pristine conditions and water molecules, water splitting can be expected both with rutile and anatase TiO2 according to Deak et al. [119]. This is generally not the case because the surface will contain bridging OH+ and the terminal OH (dissociated water). In these conditions, anatase TiO2 is the best for water splitting [119]. However, by increasing the OH+/OH ratio, one can increase the driving force for water splitting [119]. Hanaor et al. [120] looked into the phase stability of anatase and rutile TiO2, with and without doping. Pure TiO2 has a more stable rutile phase than anatase and the transformation from anatase TiO2 to rutile TiO2 is irreversible [120]. F, Si, Fe, and Al as dopants work as inhibitors for the transformation process, and they slow down the process from anatase to rutile [120]. Alghamdi et al. report no sign of overlapping HOMO levels between H2O2 and TiO2 rutile (110) surface; this in addition to the high adsorption energy could explain why water splitting is slow [121].

3.6. Collected Data

In Table 1 we have tabulated the bandgap, photocurrent density, hydrogen, and oxygen production rate from the articles discussed in this section.
As Table 1 clearly illustrates, the major focus for theoretical studies had been on bandgap calculations and alterations, but these numerical results had deviations from the results from experimental research. The majority of dopants introduced intermediate bandgaps to the structure, thus reducing the needed phonon energy for excitation. We find that by choosing the correct dopant, the effective bandgap can be lowered to 2.15 eV by co-doping TiO2 with sulfur and niobium [98]. Other promising candidates are N-doped aTiO2 (2.25 eV) [97] and the more complex g-C3N4/TiO2 (2.21 eV) [110]. A general problem with the computational studies we have reviewed is the lack of a successful descriptor of the hydrogen evolution reaction (HER) activity, however, the hydrogen adsorption free energy, ΔGH, has shown promises [122].

4. Experimental Research

Recent advances in fabrication techniques have made it possible to deposit ultra-thin films, various-sized nanoparticles, and to create nanowires, nanorods, nanobelts, etc. This has made it possible to utilize interesting properties of nanostructure and improve TiO2 photocatalysts.

4.1. Metal Dopants

An interesting phenomenon that could be exploited to increase the solar absorption is the surface plasmon resonance (SPR) effect and localized surface plasmon resonance (LSPR) effect, where metal nanoparticles absorb incoming light outside the bandgap of the catalyst. The generated electrons will then be transferred to the surface of the photocatalyst and take part in the oxidation and reduction of the water molecules. To achieve this, one must add metal nanoparticles to the surface of the photocatalyst and let them absorb incoming radiation. Several attempts utilizing metal dopants on TiO2 photocatalysts have revealed an increase in light absorption due to SPR/LSPR.
Zhao Li et al. worked on aluminum-doped TiO2, and they report an increased PEC efficiency due to the LSPR effect [123]. Interestingly enough, they also found that an ultrathin (0.8–2.5 nm) layer of Al2O3 is formed naturally, which works as a protective layer against Al NPs corrosion and in reducing the surface charge recombination [123].
A similar increase in light absorption due to SPR and LSPR can be found using Co, Ni, titanium nitride (TiN), Au, Cu, or Ag dopants [124,125,126,127,128,129]. Nickel was also found to improve the separation of electron–hole pairs [125], while TiN assisted with charge generation, separation transportation, and injection efficiency [126]. Another advantage of the Ag nanoparticles is that they lower the charge carrier recombination rate [129]. In their research on Au dopants, Jinse Park et al. used ZnO–TiO2 nanowires and found that the nanowires themselves excel in charge separation and transportation [127]. Shuai Zhang et al. used a Cu doped TiO2 nanowire film, which showcased clear improvement in photocurrent density due to the unique architecture [128].
In a similar experiment, Jie Liu et al. used Co3O4 quantum dots on TiO2 nanobelts and achieved H2 and O2 production rates of 41.8 and 22.0 µmol/hg [130]. The QDs favored transfer and accommodation of photo-generated electrons, in addition, to inhibit the recombination of charge carriers [130].
Doping could also induce a Schottky junction in the photocatalyst, which could help increase the charge transfer and help separate the photogenerated electrons and holes. He et al. showed this, using TiO2 nanowire decorated with Pd NPs and achieving a photocurrent density of 1.4 mA/cm2 [131]. The use of platinum within TiO2 based photocatalysts is well known, and Lichao Wang et al. showed that by creating a Pt/TiO2 photocatalyst, an H2 production rate of 7410 µmol/gh is achievable [132].
Complex dopants have also been used on TiO2, in addition to nanostructures. This makes it possible to combine the properties of the various dopants on TiO2. In an attempt to increase the hydrogen production of TiO2, Ejaz Hussain et al. doped TiO2 with Pd–BaO NPs [133]. They achieved an H2 production of 29.6 mmol/hg in a solution of 5% ethanol and 95% water [133]. Hussain et al. took advantage of the inherent high catalytic activity of the Pd nanoparticles, and the fact that barium oxide (BaO) enhances the electron transfer from the semiconductor band to the Pd centers [133]. In a similar approach, cadmium sulfide (CdS) was incorporated into a TiO2 photoanode [134]. This utilized the suppression of electron-hole recombination and efficient charge separation/diffusion due to the nanorod structure, in addition to the SPR effect from the dopants [134]. Instead of doping TiO2 only with CdS, it could be combined with tin (IV) oxide (SnO2) nanosheets. The reason for this is that TiO2 reduces the charge recombination between Cds and SnO2 [135]. Thus, the number of electrons and holes reaching the surface and participating in the reduction or oxidation process increases.
It is also possible to dope TiO2 with Ti3+ and Ni to improve the overall efficiency; in fact, Ti3+/Ni co-doped TiO2 nanotubes have a bandgap of 2.84 eV [136]. This is roughly 12% narrower than that of pure TiO2 and could be explained by the SPR effect.
Lately, there has been some research devoted to black titanium dioxide. Mengqiao Hu et al. used Ti3+ self-doped mesoporous black TiO2/SiO2/g–C3N4 sheets [137]. The system has a bandgap of ~2.25 eV and photocatalytic hydrogen evolution of 572.6 µmol/gh. This is all due to the unique mesoporous framework enhancing the adsorption of pollutants and favoring the mass transfer, Ti3+ self-doping reducing the bandgap, and extending the photoresponse to the visible light region [137].
Through modifying the TiO2 NPs with 2D molybdenum disulfide (MoSe2), Lulu Wu et al. achieved a hydrogen production rate of 5.13 µmol/h for samples with 0.1 wt.% MoSe2 [138]. They created MoSe2 nanosheets, which were then combined with TiO2 nanoparticles to create an efficient photocatalyst (Figure 8), by taking advantage of the wide light response and rapid charge migration ability of 2D nanosheets MoSe2. A slightly different approach would be to wrap rutile TiO2 nanorods with amorphous Ta2OxNy to achieve an optical bandgap of 2.86 eV with band edge positions suitable for water splitting [139].
Bismuth vanadate (BiVO4), iron (III) oxide/hematite (Fe2O3), and bismuth ferrite (BiFeO3) are materials with interesting properties for solar-driven water splitting. They all have low bandgaps, which could help with visible light absorption, and are both simple and inexpensive materials [140,141,142]
Xin Wu et al. utilized BiFeO3 (BFO) on top of TiO2 and found a photocurrent density as high as 11.25 mA/cm2, 20 times higher than that of bare TiO2 [143]. The improvement is mainly due to the heterostructure of BFO/TiO2 and the ferroelectric polarization due to the introduction of BFO, which could lead to upward bending at the interface and thus effectively drive the separation and transportation of photogenerated carriers [143].
Bismuth vanadate is most often used together with a dopant. For example, Jia et al. used W to dope TiO2/BiVO4 nanorods and obtained a bandgap of 2.4 eV [144]. In addition, Wengfeng Zhou et al. synthesized an ultrathin Ti/TiO2/BiVO4 nanosheet heterojunction [145]. It had an enhanced photocatalytic effect due to the formation of a built-in electric field in the heterojunction between TiO2 and BiVO4 [145]. Using Co, Pi Quan Liu et al. modified a TiO2/BiVO4 composite photoanode, which shows improved visible light absorption and a more efficient charge transfer relay [146]. By combining FeOOH/TiO2/BiVO4, Xiang Yin et al. created a photoanode that led to a hydrogen production rate of 2.36 µmol/cm2 after testing for 2.5 h [147].
However, it is possible to use BiVO4 without a dopant BiV because O4 and TiO2 naturally complement each other. It allows for the exploitation of the excellent absorption properties of BiVO4 to produce highly reductive electrons through TiO2 sensitization under visible light [148]. Another example of this is how Ahmad Radzi et al. deposited BiVO4 on TiO2 to increase PEC efficiency [149].
Hematite is usually combined with more complex structures, for example, 3d ordered urchin-like TiO2@Fe2O3 arrays [150]. Using these arrays Chai et al. reported a photocurrent density of 1.58 mA/cm2 at 1.23 V vs. reversible hydrogen electrode (RHE) [150]. This is a clear improvement compared to pristine TiO2.
A different approach is to use amorphous Fe2O3 with TiO2 and silicon (Si). With this method, Zhang et al. achieved a photocurrent density of 3.5 mA/cm2 at 1.23 V vs. RHE [151]. It is also possible to use TiO2 as the dopant on hematite. Fan Feng et al. decorated a hematite PEC with TiO2 at the grain boundaries [152] that increased the charge carrier density and improved the charge separation efficiency, resulting in a photocurrent density of 2.90 mA/cm2 at 1.23 V vs. RHE [152].
However, one could also use a simple TiO2/Fe2O3 heterojunction, which Deng et al. found improved the photocurrent density due to improved separation and transfer of photogenerated carriers [153].
A few studies are also reported on on metal-organic frameworks (MOFs) in cooperation with TiO2.
Yoon et al. coated TiO2 nanorods (NRs) with NH2–MIL-125(Ti) and achieved a photocurrent density of 1.62 mA/cm2 [154]. The high photocurrent can be explained by several factors: the large surface area and crystallinity of TiO2 NRs, which leads to effective light absorption and charge transport. Or the moderate bandgap of MIL(125)–NH2, the uniform and conformal coating of the MIL layer, and the efficient charge carrier separation and transportation through the type (II) band alignment of TiO2 and MIL(125)–NH2.

4.2. Non-Metal Dopants

Metal dopants could act as recombination centers for electrons and holes and thus lowering the overall efficiency of the photocatalyst [155]. Thus, a large number of research studies have been going on toward doping TiO2 with non-metal dopants, for example, Si, S, C, F, and N.
Yang Lu et al. doped TiO2 nanowires with earth-abundant Si and achieved an 18.2 times increase in the charge carrier density, which was better than in N and Ti (III) doped TiO2 [156]. The increase in visible light photocatalytic activity is due to the enhanced electron transport, because of higher charge-carrier density, longer electron lifetime, and larger diffusion coefficient in Si-doped TiO2 NWs [156]. High-quality graphene could be of use in water splitting as quantum dots on rutile TiO2 nanoflowers because they are highly luminescent and can absorb UV and visible light up to wavelengths of 700 nm [157]. Bellamkonda et al. found that multiwalled carbon nanotubes–graphene–TiO2 (CNT–GR–TiO2) could achieve a hydrogen production rate of 29 mmol/hg (19 mmol/hg for anatase TiO2) [158]. They also had an estimated solar energy conversion efficiency of 14.6% and a bandgap of 2.79 eV, which was due to the generation of Ti3+ and oxygen vacancies within the composite [158]. TiO2 absorbs UV light due to its inherent large bandgap, Qiongzhi Gao et al. [159] utilized this and doped TiO2 with hydrogenated F. The hydrogen-treated F atoms increased both UV and visible light absorption. When TiO2 is doped with sulfur, an abundant element, a bandgap of 2.15 eV can be expected [160]. N and lanthanum (La) co-doping of TiO2 does not reduce the bandgap, but the photocatalytic effect is seen to be enhanced due to an increase in surface nitrogen and oxygen vacancies [161].

4.3. Improved Production Methods

As discussed, although there are several options in order to dope TiO2 with metals and non-metals, the research community has devoted much energy to improving the characteristics of TiO2 photocatalysts through appropriate synthesis conditions. It is possible to achieve improved mechanical strength, enhanced composite stability, surface area, roughness, and fill factor for TiO2 by using branched multiphase TiO2 [162].
Treating TiO2 with Ar/NH3 during the fabrication process, which could improve the density of the charge carrier and broaden the photon absorption both in the UV and visible light regions [163]. An increase in density of states at the surface and a 2.5-fold increase in photocurrent density at 1.23 vs. RHE could be achieved by anodizing and annealing TiO2 during the fabrication process [164]. Ning Wei et al. showed that by controlling TiO2 core shells, it was possible to achieve a bandgap of 2.81 eV, and had an H2 evolution rate of 49.2 µmol/(h cm) [165].
Huali Huang et al. looked into the effect of annealing the atmosphere on the performance of TiO2 NR [46]. Oxygen, air, nitrogen, and argon were used as the different atmospheres. The same rutile phase was observed, but it resulted in different H2 activities. Samples annealed in argon showed the highest photocurrent density of 0.978 mA/cm2 at 1.23 V vs. RHE [46], an increase of 124.8% compared to the oxygen annealed samples. It was found that the density of oxygen vacancies in the samples increased with the decrease in oxygen in the annealing atmosphere [46]. The increase in oxygen vacancies enhances visible light absorption and increases the electron conductivity (inhibits recombination of the charge carriers) [46].
Aleksander et al. examined what would happen if the substrate, which TiO2 was fabricated on was changed [166]. The authors lowered the optical reflection by using black silicon, which in turn increased the light collection [166]. They also found that the addition of noble metals could induce SPR in the visible light region [166].
By combining improved production methods and doping/co-doping with metals/non-metals, TiO2 could be realized as an efficient photocatalyst. Fu et. al. showed that by controlling the HCl concentration during the synthesis process, it was possible to synthesize well-crystallized rutile TiO2 nanorods with special aspect ratios [167]. They proposed a process, as shown in Figure 9, for synthesizing rutile TiO2 with different aspect ratios. Rutile TiO2 nanorods with small aspect ratios were formed by placing titanium tetrachloride (TiCl4) in liquid hydrochloric acid (HCl) before undertaking hydrothermal treatment. The key factors were the presence of Cl and H+ at high temperature and pressure. For the synthesis of nanorods with medium/large aspect ratios titanium butoxide (TBOT) was added dropwise to HCL (aq.)/NaCl (aq.) This created rutile/anatase crystal seeds, which were placed in HCl (aq.) for the final growth process. They concluded that with decreasing aspect ratios, the photocatalytic water splitting activity would increase for TiO2 nanorods [167].

4.4. Collected Data

We present in Table 2 the obtained bandgap, photocurrent density, H2, and O2 production rate values from experimental studies that are reviewed here.
We see from the data presented in Table 2 that it is possible to adjust the properties of TiO2 photocatalysts by doping or through structural changes. For example, Elbakkay et al. achieved a bandgap of 2.15 eV using an S–TiO2/S-reduced graphene oxide catalyst [160]. Both theoretical and experimental studies point at sulfur as a possible dopant for TiO2 that could drastically reduce the effective bandgap of the photocatalyst. In general, theoretical studies tend to focus on simpler structures and doping of TiO2, while experimental research has moved on toward more complex structures of TiO2 that consist of several layers, materials, and nanostructures.
It is clear from Table 2 that the various solutions and potentials are used for measuring HER, and presenting photocurrent density in different ways is an issue that makes the direct comparison difficult. This hampers the evaluation of the most promising TiO2 structure for photocatalytic activity. Likewise, the theoretical studies use different models and approximations that make it difficult to compare the numerical results for different TiO2 structures. Combined theoretical and experimental study along with the establishment of standards, for example, for measuring H2 production, would help the path to develop TiO2 photocatalysts towards commercial realization.
Table 2. This table displays the bandgap, photocurrent density, and H2 production rate experimentally achieved for doped/modified TiO2 structures in articles reviewed in this study.
Table 2. This table displays the bandgap, photocurrent density, and H2 production rate experimentally achieved for doped/modified TiO2 structures in articles reviewed in this study.
NanomaterialBandgap
[eV]
Photocurrent Density at 1.23 V vs. RHE [mA/cm2]H2 Production Rate
1.5G Sunlight Bias at 1.23 vs. RHE
O2 Production Rate
1.5G Sunlight Bias at 1.23 vs. RHE
Ref.
TiO2@Fe2O3/TiO22.21.58, and 3.6 at 1.6 V vs. RHENANA[150]
α-Ta2OxNy enwrapped TiO2 rutile nanorods2.861.32244.2 mmol/m2h112.7 mmol/m2h[139]
Ag-TiO2-NR052.640.08 and 0.10 mA/cm2 at 1.2 and 1.6 V vs. RHENANA[129]
W-TiO2/BiVO4 nanorods2.42.541 µmol/hNA[144]
Branched multiphase TiO23.040.95NANA[162]
Co3O4 quantum dots on TiO23.070.000541.8 µmol/h/g22.0 µmol/h/g[130]
Co-Pi modified 3D TiO2/BiVO4NA4.96 at 0.63 V vs. Ag/AgClNANA[146]
Co doped TiO2 nanotubes2.881.0NANA[124]
Controllable TiO2 core shells2.813.8849.2 µmol/cm2h25.2 µmol/cm2h[165]
A-Fe2O3/TiO2/SiNA3.5NANA[151]
Al@TiO2NANANANA[123]
Si-doped TiO2 nanowiresNANANANA[156]
Three-layer (SiO2, Al2O3, and TiO2) structure with Au particles for LSPRNANANANA[168]
BiFeO3/TiO2NA11.25NANA[143]
Graphene QDs decorated rutile TiO2 nanoflowersNA~0.32 at 0.5 V vs. Ag/AgClNANA[157]
Hierarchical TiO2/Fe2O3NA1.79NANA[153]
CNT-GR-TiO22.79NA29 mmol/h/gNA[158]
SnO2 nanosheets with TiO2 and CdS QDNA4.7 at 0V vs. Ag/AgClNANA[135]
TiO2 nanotubes treated with Ar/NH3+NA1 at 1.18 V vs. RHENANA[163]
TiO2 nanowire decorated with PdNA1.4NANA[131]
NH2-MIL-125(Yi) on TiO2 nanorodsNA1.62NANA[154]
Ni-doped TiO2 nanotubesNA0.93 at 0 V vs. Ag/AgClNANA[125]
N doped La/TiO22.96–2.99NA8.25 µmol/h/gNA[161]
TiN boosted N doped TiO2NA3.12NANA[126]
CuO@TiO2 nanowiresNA0.56NANA[128]
Pd-BaO NPs on TiO2NANA29.6 mmol/h/gNA[133]
S-TiO2/S-RGO2.153.36 at 1 V vs. Ag/AgClNANA[160]
Anodized and H2 annealed TiO2NA2.5 fold TiO2NANA[164]
TiO2 NPs modified with 2D MoSe2NANA5.12 µmol/hNA[138]
Ultrathin Ti/TiO2/BiVO4NA5.8 µa/cm2 at 0.5 V vs. Ag/AgClNANA[145]
TiO2 on black SiNANANANA[166]
ZnO-TiO2 core-shell nanowires decorated with Au NPsNA1.63NANA[127]
TiO2/CdS system2.2530 mA/cm2 (at 1 V vs. Ag/AgCl) under 1.5 AM1.3 mmol/cm2hNA[134]
FeOOH/TiO2/BiVO4NA3.212.36 µmol/cm21.09 µmol/cm2h[147]
hematite PEC decorated with TiO2 at the grain boundariesNA2.90NANA[152]
the effect of annealing atmosphere on the performance of TiO2 NRNA0.978NANA[46]
Ti3+/Ni co-doped TiO2 nanotubes2.840.87NANA[136]
Hydrogenated F-doped TiO23.0NA3.76 mmol/h/gNA[159]
BiVO4 deposited on TiO2NA35 µ under 100 mW/cm2 in 0.5M Na2SO4NANA[149]
BiVO4 used together with TiO2NA~0.3 at 1.0 V vs. RHENANA[148]
Pt/TiO2(anatase) photocatalystNANA7410 µmol/h/g5096 µmol/h/g[132]
Ti3+ self-doped mesoporous black TiO2/SiO2/g-C3N4 sheets~2.25NANANA[137]
Rutile TiO2 nanorods with small aspect ratioNANA1229 µmol/h/g549 µmol/h/g[167]
Rutile TiO2 nanorods with medium aspect ratioNANA783 µmol/h/g369 µmol/h/g[167]
Rutile TiO2 nanorods with large aspect ratioNANA549 µmol/h/g252 µmol/h/g[167]
As we can observe from the results presented so far, great steps have been taken to make TiO2 photocatalysts for water splitting a viable technology for green hydrogen production. However, even though the progress has been rapid over the past decade there are still obstacles in the way to large-scale production facilities.
Coordinated theoretical and experimental study of TiO2 structures for enhancing the electronic, optical, and physical properties will help achieve the goal of efficient low-cost photocatalysts for water splitting.
In general, during the synthesis of the photocatalysts, there is uncertainty in the exact composition and structure of the compound [169,170]. This is especially prevalent for doping and the location of the dopants in the compound. For example, if the dopants are too deep or too shallow (on the surface), they will behave as recombination centers and thus reduce the overall solar to hydrogen efficiency [171]. The selection of deposition techniques for TiO2 structures will have an effect on the performance as these techniques have advantages and disadvantages. For example, chemical vapor deposition (CVD) and physical vapor deposition (PVD) produce homogenous and flexible microstructure and super hard materials, but the drawbacks are challenges with deposition rate, maintenance cost, and the size of the component. Thermal spraying has the advantage of fast deposition rates, large components, and ease of exploitation, but does not produce coatings of the same quality as CVD, electrodeposition, or PVD [172]. In other words, there is always the dilemma of choosing the correct deposition method and figuring out how it could affect the performance of the material. The research community employs a variety of techniques to characterize the material, infrared spectroscopy, Raman spectroscopy, scanning electron microscopy (SEM), X-ray spectroscopy, etc. Because the reviewed studies employ different characterization methods to verify, for example, the bandgap of TiO2 structures, a comparison of results in different studies becomes challenging. Another challenge is that experiments are not carried out under standardized conditions (e.g., with constant irradiance and homogeneous light distribution), the assessment of the real progress achieved with modified TiO2 is often difficult. Furthermore, comparisons between pristine TiO2 and enhanced photocatalysts are frequently biased because samples selected as reference materials present a relatively low photoactivity [173].
Computational modeling and simulations can help relieve some of these issues, although it comes with its own limitations. One advantage of theoretical simulations is that it is possible to create the exact structure you want to work on. Thus, we can investigate specific attributes and properties by fine-tuning the structure and composition of the compound. Moreover, the calculation of the H2 and O2 production is not based on measurement conditions, which makes results easier to compare.

4.5. Production Facilities

The aforementioned challenges are not the only ones that the research community faces in using TiO2 as the photocatalytic material for water splitting. Unfortunately, there is still a lack of scalable systems that could produce hydrogen in an economically feasible manner, even with an efficient catalyst [174]. Pinaud et. al. proposed and discussed the economic feasibility of four different designs of photocatalytic water splitting plants and the schematics are shown in Figure 10 [175].
Type 1 is a single bed particle suspension reactor and it is the simplest of the four. It consists of a low-lying horizontal plastic bag containing a slurry of photoactive particles in an electrolyte. The plastic bag is designed to allow light to penetrate, while it retains the electrolyte, photoactive particles, and evolved gases.
The Type 2 reactor is a dual bed with particle suspension, and it is similar to that of the Type 1 reactor. However, the biggest difference is that separate beds are used for H2 and O2 production.
The third option, named Type 3 reactor, is a fixed panel array, which consists of an integral planar electrode with multiple photoactive layers sandwiched between two electrodes. The entire system is within a transparent plastic electrolyte reservoir. The final alternative is the Type 4 reactor, which is a tracking concentrator array that uses an offset parabolic cylinder array to focus sunlight on a linear PEC cell receiver and has two-axis steering to track the daily movement of the sun [175].
In general, it was found that the key component for realizing these designs was to improve the solar to hydrogen efficiencies [175]. However, there are also other limitations, such as safety issues with the H2 and O2 gas mixture, how to split and collect the H2 and O2 gasses, a lack of general understanding of how the photocatalyst particle works, the mechanical integrity of the plastic bags, etc. [175].

5. Conclusions and Long-Term Outlook

Even though considerable progress has been made in the development of solar-driven water splitting with TiO2 as the photocatalyst, we believe there are four major challenges the research community must tackle before it becomes a viable technology.
The first challenge is the lack of a standard way to express the hydrogen production rates with varying photocatalytic materials. Research groups have been presenting these generations’ rates in different ways that make the comparison challenging. As seen in the literature presented here, the measured hydrogen evolution rates depend on specific details of the experimental setup, such as the spectrum of the light source, the light intensity at the sample, co-catalyst selection, size of the potential, and type and selection of the solution. Suggestions and solutions for standard experimental setups are also needed. However, reporting the apparent quantum yield (AQY) instead of only the gas evolution could be a part of the solution [176,177]. This will help bridge the gap between experimental and theoretical results. Computational modeling has some of the same challenges as the model, assumptions, approximations, and software used will affect the numerical results that make the comparison of results demanding. The missing piece here is calculations of the hydrogen and oxygen evolution rates, and an alternative can be studying the Volmer reaction, the Tafel reaction or the Herovsky reaction, and the Gibbs free energy [178,179].
The second challenge is the material TiO2 and its wide bandgap of 3.2 eV, which is in the ultraviolet section of the visible light spectrum. This means that 97% of the energy coming from the sun is not usable for TiO2 photocatalysts. There have been several attempts to lower the bandgap of TiO2, both experimentally and theoretically, which have been successful. However, the most successful experimental works are based on complex nanostructures or layered structures that are difficult and expensive to create at a larger scale. Computational modeling has in general focused more on various dopants and doping percentages. Unfortunately, the best results are seen when using noble metals or expensive metals. However, sulfur doping could be a solution to this problem. Combined TiO2 with other earth-abundant materials as MoS2 or WS2 could be better photocatalysts in the future [180,181].
The third challenge is the lack of research combining experimental research with theoretical simulations to optimize the characteristics of TiO2 structures for photocatalytic applications. This is a weakness in the current research as theoretical modeling and simulation could work as a great screening tool for the experimentalists, reducing their workload. Theoretical research could also help with the fundamental understanding of the process involved in solar-driven water splitting. By combining the two methods, it is easier to see the inner workings of the photocatalyst and to determine where improvement is needed. Moreover, computational work requires experimental verification and for validating the numerical results.
The final challenge is the lack of scalable systems that could produce hydrogen in an economically feasible manner, even with an efficient catalyst [174]. There is currently a lack of scalable and functional production facilities. The ones that do exist have a solar-to-hydrogen efficiency of 1.8% [182], indicating that further research is needed before photocatalytic water splitting is competitive with other hydrogen production methods. The main factor for the low efficiencies reported for photocatalytic hydrogen production is the low solar-to-hydrogen rates of the photocatalyst itself. An interesting idea is to look into combining TiO2 with perovskites due to the latter’s excellent optoelectronic properties. Naturally, water-insoluble perovskites combined with TiO2 could be the missing link that solar-driven water splitting needs.

Author Contributions

Conceptualization, H.E. and D.V.; methodology, H.E.; writing—original draft preparation, H.E.; writing—review and editing, H.E., S.B., P.V., S.Y. and D.V.; supervision, D.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

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

Acknowledgments

The computations/simulations/[SIMILAR] were performed on resources provided by UNINETT Sigma2—the National Infrastructure for High Performance Computing and Data Storage in Norway (project: NN2867K).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. OECD/IEA. World Energy Outlook. 2016. Available online: http://www.iea.org/t&c (accessed on 14 May 2018).
  2. Dincer, I.; Zamfirescu, C. Sustainable Energy Systems and Applications, 1st ed.; Springer: Berlin/Heidelberg, Germany, 2012; p. XVII. [Google Scholar] [CrossRef]
  3. Tong, Y.; Liang, J.; Liu, H.K.; Dou, S.X. Energy storage in Oceania. Energy Storage Mater. 2019, 20, 176–187. [Google Scholar] [CrossRef]
  4. Jafari, T.; Moharreri, E.; Amin, A.S.; Miao, R.; Song, W.; Suib, S.L. Photocatalytic water splitting-the untamed dream: A Review of recent advances. Molecules 2016, 21, 900. [Google Scholar] [CrossRef]
  5. Sharma, P.; Kolhe, M.L. Review of sustainable solar hydrogen production using photon fuel on artificial leaf. Int. J. Hydrog. Energy 2017, 42, 22704–22712. [Google Scholar] [CrossRef]
  6. Pulido Melián, E.; González Díaz, O.; Ortega Méndez, A.; López, C.R.; Nereida Suárez, M.; Doña Rodríguez, J.M.; Navío, J.A.; Fernández Hevia, D.; Pérez Peña, J. Efficient and affordable hydrogen production by water photo-splitting using TiO2-based photocatalysts. Int. J. Hydrog. Energy 2013, 38, 2144–2155. [Google Scholar] [CrossRef]
  7. Bicer, Y.; Dincer, I. Clean fuel options with hydrogen for sea transportation: A life cycle approach. Int. J. Hydrog. Energy 2018, 43, 1179–1193. [Google Scholar] [CrossRef]
  8. van Biert, L.; Godjevac, M.; Visser, K.; Aravind, P.V. A review of fuel cell systems for maritime applications. J. Power Sources 2016, 327, 345–364. [Google Scholar] [CrossRef] [Green Version]
  9. International Maritime Organization. Third IMO Greenhouse Gas Study 2014. Available online: http://www.imo.org/en/OurWork/Environment/PollutionPrevention/AirPollution/Pages/Greenhouse-Gas-Studies-2014.aspx (accessed on 14 May 2018).
  10. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37. [Google Scholar] [CrossRef] [PubMed]
  11. Wang, B.; Shen, S.; Mao, S.S. Black TiO2 for solar hydrogen conversion. J. Mater. 2017, 3, 96–111. [Google Scholar] [CrossRef]
  12. Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef]
  13. Haggerty, J.E.S.; Schelhas, L.T.; Kitchaev, D.A.; Mangum, J.S.; Garten, L.M.; Sun, W.; Stone, K.H.; Perkins, J.D.; Toney, M.F.; Ceder, G.; et al. High-fraction brookite films from amorphous precursors. Sci. Rep.-Uk 2017, 7, 15232. [Google Scholar] [CrossRef] [PubMed]
  14. Kaur, K.; Singh, C.V. Amorphous TiO2 as a photocatalyst for hydrogen production: A DFT study of structural and electronic properties. Energy Proc. 2012, 29, 291–299. [Google Scholar] [CrossRef] [Green Version]
  15. Zhang, W.; Yin, J.-R.; Tang, X.-Q.; Zhang, P.; Ding, Y.-H. Density functional theory studies on the structural and physical properties of Cu-doped anatase TiO2 (101) surface. Phys. E 2017, 85, 259–263. [Google Scholar] [CrossRef]
  16. Morgade, C.I.N.; Cabeza, G.F. First-principles study of codoping TiO2 systems capable of improving the specific surface area and the dissociation of H2O to generate H2 and O2. Comput. Mater. Sci. 2017, 127, 204–210. [Google Scholar] [CrossRef]
  17. Press, R.J.; Santhanam, K.S.V.; Miri, M.J.; Bailey, A.V.; Takacs, G.A. Introduction to hydrogen technology. ChemSusChem 2009, 3, 432. [Google Scholar]
  18. Hoang, D.L.; Chan, S.H.; Ding, O.L. Kinetic and modelling study of methane steam reforming over sulfide nickel catalyst on a gamma alumina support. Chem. Eng. J. 2005, 112, 1–11. [Google Scholar] [CrossRef]
  19. Ozcan, H.; Dincer, I. Thermodynamic analysis of a combined chemical looping-based trigeneration system. Energy Convers. Manag. 2014, 85, 477–487. [Google Scholar] [CrossRef]
  20. Steinfeld, A. Solar hydrogen production via a two-step water-splitting thermochemical cycle based on Zn/ZnO redox reactions. Int. J. Hydrog. Energy 2002, 27, 611–619. [Google Scholar] [CrossRef]
  21. Akkerman, I.; Janssen, M.; Rocha, J.; Wijffels, R.H. Photobiological hydrogen production: Photochemical efficiency and bioreactor design. Int. J. Hydrog. Energy 2002, 27, 1195–1208. [Google Scholar] [CrossRef]
  22. Liao, C.-H.; Huang, C.-W.; Wu, J.C.S. Hydrogen production from semiconductor-based photocatalysis via water splitting. Catalysts 2012, 2, 490. [Google Scholar] [CrossRef] [Green Version]
  23. Kudo, A. Photocatalysis and solar hydrogen production. Pure Appl. Chem. 2007, 79, 1917. [Google Scholar] [CrossRef]
  24. Jang, J.S.; Kim, H.G.; Lee, J.S. Heterojunction semiconductors: A strategy to develop efficient photocatalytic materials for visible light water splitting. Catal. Today 2012, 185, 270–277. [Google Scholar] [CrossRef]
  25. Moriya, Y.; Takata, T.; Domen, K. Recent progress in the development of (oxy)nitride photocatalysts for water splitting under visible-light irradiation. Coordin. Chem. Rev. 2013, 257, 1957–1969. [Google Scholar] [CrossRef]
  26. Takata, T.; Jiang, J.; Sakata, Y.; Nakabayashi, M.; Shibata, N.; Nandal, V.; Seki, K.; Hisatomi, T.; Domen, K. Photocatalytic water splitting with a quantum efficiency of almost unity. Nature 2020, 581, 411–414. [Google Scholar] [CrossRef]
  27. Cho, S.; Jang, J.-W.; Lee, K.-H.; Lee, J.S. Research Update: Strategies for efficient photoelectrochemical water splitting using metal oxide photoanodes. APL Mater. 2014, 2. [Google Scholar] [CrossRef] [Green Version]
  28. Naik, V.M.; Haddad, D.; Naik, R.; Benci, J.; Auner, G.W. Optical properties of anatase, rutile and amorphous phases of TiO2 thin films grown at room temperature by RF magnetron sputtering. Mrs Proc. 2002, 755, DD11.12. [Google Scholar] [CrossRef]
  29. Liu, H.; Hsu, Y.; Su, H.; Huang, R.; Hou, F.; Tu, G.; Liu, W. A Comparative study of amorphous, anatase, rutile, and mixed phase TiO2 films by mist chemical vapor deposition and ultraviolet photodetectors applications. IEEE Sens. J. 2018, 18, 4022–4029. [Google Scholar] [CrossRef]
  30. Furube, A.; Asahi, T.; Masuhara, H.; Yamashita, H.; Anpo, M. Charge carrier dynamics of standard TiO2 catalysts revealed by femtosecond diffuse reflectance spectroscopy. J. Phys. Chem. B 1999, 103, 3120–3127. [Google Scholar] [CrossRef]
  31. Ohtani, B.; Ogawa, Y.; Nishimoto, S.-I. Photocatalytic activity of amorphous−anatase mixture of Titanium(IV) oxide particles suspended in aqueous solutions. J. Phys. Chem. B 1997, 101, 3746–3752. [Google Scholar] [CrossRef] [Green Version]
  32. Stone, V.F.; Davis, R.J. Synthesis, Characterization, and Photocatalytic Activity of Titania and Niobia Mesoporous Molecular Sieves. Chem. Mater. 1998, 10, 1468–1474. [Google Scholar] [CrossRef]
  33. Tanaka, K.; Capule, M.F.V.; Hisanaga, T. Effect of crystallinity of TiO2 on its photocatalytic action. Chem. Phys. Lett. 1991, 187, 73–76. [Google Scholar] [CrossRef]
  34. Liu, N.; Albu, S.P.; Lee, K.; So, S.; Schmuki, P. Water annealing and other low temperature treatments of anodic TiO2 nanotubes: A comparison of properties and efficiencies in dye sensitized solar cells and for water splitting. Electrochim. Acta 2012, 82, 98–102. [Google Scholar] [CrossRef]
  35. Gong, J.; Lai, Y.; Lin, C. Electrochemically multi-anodized TiO2 nanotube arrays for enhancing hydrogen generation by photoelectrocatalytic water splitting. Electrochim. Acta 2010, 55, 4776–4782. [Google Scholar] [CrossRef]
  36. Tiwari, J.N.; Tiwari, R.N.; Kim, K.S. Zero-dimensional, one-dimensional, two-dimensional and three-dimensional nanostructured materials for advanced electrochemical energy devices. Prog. Mater. Sci. 2012, 57, 724–803. [Google Scholar] [CrossRef]
  37. Joy, J.; Mathew, J.; George, S.C. Nanomaterials for photoelectrochemical water splitting—Review. Int. J. Hydrog. Energy 2018, 43, 4804–4817. [Google Scholar] [CrossRef]
  38. Kumar, P.; Devi, P.; Jain, R.; Shivaprasad, S.M.; Sinha, R.K.; Zhou, G.; Nötzel, R. Quantum dot activated indium gallium nitride on silicon as photoanode for solar hydrogen generation. Commun. Chem. 2019, 2, 4. [Google Scholar] [CrossRef] [Green Version]
  39. Basu, K.; Zhang, H.; Zhao, H.; Bhattacharya, S.; Navarro-Pardo, F.; Datta, P.K.; Jin, L.; Sun, S.; Vetrone, F.; Rosei, F. Highly stable photoelectrochemical cells for hydrogen production using a SnO2–TiO2/quantum dot heterostructured photoanode. Nanoscale 2018, 10, 15273–15284. [Google Scholar] [CrossRef] [PubMed]
  40. Liu, Z.; Wu, J.; Zhang, J. Quantum dots and plasmonic Ag decorated WO3 nanorod photoanodes with enhanced photoelectrochemical performances. Int. J. Hydrog. Energy 2016, 41, 20529–20535. [Google Scholar] [CrossRef]
  41. Beermann, N.; Vayssieres, L.; Lindquist, S.E.; Hagfeldt, A. Photoelectrochemical studies of oriented nanorod thin films of hematite. J. Electrochem. Soc. 2000, 147, 2456–2461. [Google Scholar] [CrossRef]
  42. Khan, S.U.M.; Sultana, T. Photoresponse of n-TiO2 thin film and nanowire electrodes. Sol. Energy Mat. Sol. Cells 2003, 76, 211–221. [Google Scholar] [CrossRef]
  43. Fitch, A.; Strandwitz, N.C.; Brunschwig, B.S.; Lewis, N.S. A comparison of the behavior of single crystalline and nanowire array ZnO photoanodes. J. Phys. Chem. C 2013, 117, 2008–2015. [Google Scholar] [CrossRef] [Green Version]
  44. Zhang, J.Z. Metal oxide nanomaterials for solar hydrogen generation from photoelectrochemical water splitting. Mrs Bull. 2011, 36, 48–55. [Google Scholar] [CrossRef]
  45. Varghese, O.K.; Grimes, C.A. Appropriate strategies for determining the photoconversion efficiency of water photoelectrolysis cells: A review with examples using titania nanotube array photoanodes. Sol. Energy Mat. Sol. Cells 2008, 92, 374–384. [Google Scholar] [CrossRef]
  46. Huang, H.; Hou, X.; Xiao, J.; Zhao, L.; Huang, Q.; Chen, H.; Li, Y. Effect of annealing atmosphere on the performance of TiO2 nanorod arrays in photoelectrochemical water splitting. Catal. Today 2019, 330, 189–194. [Google Scholar] [CrossRef]
  47. Nie, Q.; Yang, L.; Cao, C.; Zeng, Y.; Wang, G.; Wang, C.; Lin, S. Interface optimization of ZnO nanorod/CdS quantum dots heterostructure by a facile two-step low-temperature thermal treatment for improved photoelectrochemical water splitting. Chem. Eng. J. 2017, 325, 151–159. [Google Scholar] [CrossRef]
  48. Pokrant, S.; Dilger, S.; Landsmann, S.; Trottmann, M. Size effects of cocatalysts in photoelectrochemical and photocatalytic water splitting. Mater. Today Energy 2017, 5, 158–163. [Google Scholar] [CrossRef]
  49. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef] [PubMed]
  50. Liu, X.; Gu, S.; Zhao, Y.; Zhou, G.; Li, W. BiVO4, Bi2WO6 and Bi2MoO6 photocatalysis: A brief review. J. Mater. Sci. Technol. 2020, 56, 45–68. [Google Scholar] [CrossRef]
  51. Vishwakarma, A.K.; Tripathi, P.; Srivastava, A.; Sinha, A.S.K.; Srivastava, O.N. Band gap engineering of Gd and Co doped BiFeO3 and their application in hydrogen production through photoelectrochemical route. Int. J. Hydrog. Energy 2017, 42, 22677–22686. [Google Scholar] [CrossRef]
  52. Yin, W.-J.; Tang, H.; Wei, S.-H.; Al-Jassim, M.M.; Turner, J.; Yan, Y. Band structure engineering of semiconductors for enhanced photoelectrochemical water splitting: The case of TiO2. Phys. Rev. B 2010, 82, 045106. [Google Scholar] [CrossRef] [Green Version]
  53. Wang, J.; Sun, H.; Huang, J.; Li, Q.; Yang, J. Band Structure Tuning of TiO2 for Enhanced Photoelectrochemical Water Splitting. J. Phys. Chem. C 2014, 118, 7451–7457. [Google Scholar] [CrossRef]
  54. Zhang, G.; Zhang, W.; Minakata, D.; Chen, Y.C.; Crittenden, J.; Wang, P. The pH effects on H2 evolution kinetics for visible light water splitting over the Ru/(CuAg)0.15In0.3Zn1.4S2 photocatalyst. Int. J. Hydrog. Energy 2013, 38, 11727–11736. [Google Scholar] [CrossRef]
  55. Fekete, M.; Riedel, W.; Patti, A.F.; Spiccia, L. Photoelectrochemical water oxidation by screen printed ZnO nanoparticle films: Effect of pH on catalytic activity and stability. Nanoscale 2014, 6, 7585–7593. [Google Scholar] [CrossRef]
  56. Momeni, M.M.; Ghayeb, Y. Visible light-driven photoelectrochemical water splitting on ZnO–TiO2 heterogeneous nanotube photoanodes. J. Appl. Electrochem. 2015, 45, 557–566. [Google Scholar] [CrossRef]
  57. Varadhan, P.; Fu, H.-C.; Priante, D.; Retamal, J.R.D.; Zhao, C.; Ebaid, M.; Ng, T.K.; Ajia, I.; Mitra, S.; Roqan, I.S.; et al. Surface Passivation of GaN Nanowires for Enhanced Photoelectrochemical Water-Splitting. Nano Lett. 2017, 17, 1520–1528. [Google Scholar] [CrossRef] [PubMed]
  58. Ismail, A.A.; Bahnemann, D.W. Photochemical splitting of water for hydrogen production by photocatalysis: A review. Sol. Energy Mat. Sol. Cells 2014, 128, 85–101. [Google Scholar] [CrossRef]
  59. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  60. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comp. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  61. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  62. Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal-amorphous-semiconductor transition in germanium. Phys. Rev. B 1994, 49, 14251–14269. [Google Scholar] [CrossRef] [PubMed]
  63. Kresse, G.; Hafner, J. Norm-conserving and ultrasoft pseudopotentials for first-row and transition elements. J. Phys.-Condens Mat. 1994, 6, 8245–8257. [Google Scholar] [CrossRef]
  64. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  65. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Krist.-Cryst. Mater. 2005, 220, 567–570. [Google Scholar] [CrossRef] [Green Version]
  66. Dovesi, R.; Erba, A.; Orlando, R.; Zicovich-Wilson, C.M.; Civalleri, B.; Maschio, L.; Rérat, M.; Casassa, S.; Baima, J.; Salustro, S.; et al. Quantum-mechanical condensed matter simulations with CRYSTAL. Wires Comput. Mol. Sci. 2018, 8, e1360. [Google Scholar] [CrossRef]
  67. Dovesi, R.; Pascale, F.; Civalleri, B.; Doll, K.; Harrison, N.M.; Bush, I.; D’Arco, P.; Noël, Y.; Rérat, M.; Carbonnière, P.; et al. The CRYSTAL code, 1976–2020 and beyond, a long story. J. Chem. Phys. 2020, 152, 204111. [Google Scholar] [CrossRef]
  68. Mortensen, J.J.; Hansen, L.B.; Jacobsen, K.W. Real-space grid implementation of the projector augmented wave method. Phys. Rev. B 2005, 71, 035109. [Google Scholar] [CrossRef] [Green Version]
  69. Kan, M.; Wang, J.Y.; Li, X.W.; Zhang, S.H.; Li, Y.W.; Kawazoe, Y.; Sun, Q.; Jena, P. Structures and Phase Transition of a MoS2 Monolayer. J. Phys. Chem. C 2014, 118, 1515–1522. [Google Scholar] [CrossRef]
  70. Xia, Z.; Tao, Y.; Pan, Z.; Shen, X. Enhanced photocatalytic performance and stability of 1T MoS2 transformed from 2H MoS2 via Li intercalation. Results Phys. 2019, 12, 2218–2224. [Google Scholar] [CrossRef]
  71. Liu, P.; Kaltak, M.; Klimeš, J.; Kresse, G. Cubic scaling GW: Towards fast quasiparticle calculations. Phys. Rev. B 2016, 94, 165109. [Google Scholar] [CrossRef] [Green Version]
  72. Ramberger, B.; Sukurma, Z.; Schäfer, T.; Kresse, G. RPA natural orbitals and their application to post-Hartree-Fock electronic structure methods. J. Chem. Phys. 2019, 151, 214106. [Google Scholar] [CrossRef] [Green Version]
  73. Heyd, J.; Scuseria, G.E. Efficient hybrid density functional calculations in solids: Assessment of the Heyd-Scuseria-Ernzerhof screened Coulomb hybrid functional. J. Chem. Phys. 2004, 121, 1187–1192. [Google Scholar] [CrossRef] [Green Version]
  74. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef] [Green Version]
  75. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Wang, V.; Xu, N.; Liu, J.C.; Tang, G.; Geng, W.-T. VASPKIT: A User-friendly Interface Facilitating High-throughput Computing and Analysis Using VASP Code. arXiv 2019, arXiv:1908.08269. [Google Scholar]
  77. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef] [Green Version]
  78. Kumar, A. Different methods used for the synthesis of TiO2 based nanomaterials: A review. Am. J. Nano Res. Appl. 2018, 6, 1. [Google Scholar] [CrossRef]
  79. Nagaraju, G.; Tharamani, C.; Chandrappa, G.; Livage, J. Hydrothermal synthesis of amorphous MoS2nanofiber bundles via acidification of ammonium heptamolybdate tetrahydrate. Nanoscale Res. Lett. 2007, 2, 461–468. [Google Scholar] [CrossRef] [Green Version]
  80. Bai, J.; Meng, T.; Guo, D.; Wang, S.; Mao, B.; Cao, M. Co9S8@MoS2 Core–Shell Heterostructures as Trifunctional Electrocatalysts for Overall Water Splitting and Zn–Air Batteries. ACS Appl. Mater. Interfaces 2018, 10, 1678–1689. [Google Scholar] [CrossRef]
  81. Lee, J.-H.; Choi, H.-S.; Lee, J.-H.; Kim, Y.-J.; Suh, S.-J.; Chi, C.-S.; Oh, H.-J. Fabrication of titania nanotubular film with metal nanoparticles. J. Cryst. Growth 2009, 311, 638–641. [Google Scholar] [CrossRef]
  82. Patil, U.M.; Kulkarni, S.B.; Deshmukh, P.R.; Salunkhe, R.R.; Lokhande, C.D. Photosensitive nanostructured TiO2 grown at room temperature by novel “bottom-up” approached CBD method. J. Alloy. Compd. 2011, 509, 6196–6199. [Google Scholar] [CrossRef]
  83. Wu, J.-M.; Hayakawa, S.; Tsuru, K.; Osaka, A. Nanocrystalline Titania Made from Interactions of Ti with Hydrogen Peroxide Solutions Containing Tantalum Chloride. Cryst. Growth Des. 2002, 2, 147–149. [Google Scholar] [CrossRef]
  84. AlHammad, M.S. Nanostructure hydroxyapatite based ceramics by sol gel method. J. Alloy. Compd. 2016, 661, 251–256. [Google Scholar] [CrossRef]
  85. Kluson, P.; Luskova, H.; Cajthaml, T.; Šolcová, O. Non thermal preparation of photoactive titanium (IV) oxide thin layers. Thin Solid Film. 2006, 495, 18–23. [Google Scholar] [CrossRef]
  86. Arami, H.; Mazloumi, M.; Khalifehzadeh, R.; Sadrnezhaad, S.K. Sonochemical preparation of TiO2 nanoparticles. Mater. Lett. 2007, 61, 4559–4561. [Google Scholar] [CrossRef]
  87. Amaresh, S.; Karthikeyan, K.; Jang, I.C.; Lee, Y.S. Single-step microwave mediated synthesis of the CoS2 anode material for high rate hybrid supercapacitors. J. Mater. Chem. A 2014, 2, 11099–11106. [Google Scholar] [CrossRef]
  88. Pan, Y.; Wen, M. Noble metals enhanced catalytic activity of anatase TiO2 for hydrogen evolution reaction. Int. J. Hydrog. Energy 2018, 43, 22055–22063. [Google Scholar] [CrossRef]
  89. Zhang, Y.; Kilin, D.S. Computational modeling of wet TiO2(001) anatase surfaces functionalized by transition metal doping. Int. J. Quantum Chem. 2012, 112, 3867–3873. [Google Scholar] [CrossRef]
  90. Zhang, S.-T.; Li, C.-M.; Yan, H.; Wei, M.; Evans, D.G.; Duan, X. Density Functional Theory Study on the Metal–Support Interaction between Ru Cluster and Anatase TiO2(101) Surface. J. Phys. Chem. C 2014, 118, 3514–3522. [Google Scholar] [CrossRef]
  91. 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]
  92. Lin, Y.; Jiang, Z.; Zhu, C.; Zhang, R.; Hu, X.; Zhang, X.; Zhu, H.; Lin, S.H. The electronic structure, optical absorption and photocatalytic water splitting of (Fe + Ni)-codoped TiO2: A DFT + U study. Int. J. Hydrog. Energy 2017, 42, 4966–4976. [Google Scholar] [CrossRef]
  93. Ghuman, K.K. Mechanistic insights into water adsorption and dissociation on amorphous -based catalysts. Sci. Technol. Adv. Mater. 2018, 19, 44–52. [Google Scholar] [CrossRef] [Green Version]
  94. Koteski, V.; Belošević-Čavor, J.; Umićević, A.; Ivanovski, V.; Toprek, D. Improving the photocatalytic properties of anatase TiO2(101) surface by co-doping with Cu and N: Ab initio study. Appl. Surf. Sci. 2017, 425, 1095–1100. [Google Scholar] [CrossRef]
  95. Assadi, M.H.N.; Hanaor, D.A.H. The effects of copper doping on photocatalytic activity at (101) planes of anatase TiO2: A theoretical study. Appl. Surf. Sci. 2016, 387, 682–689. [Google Scholar] [CrossRef] [Green Version]
  96. Sikam, P.; Moontragoon, P.; Sararat, C.; Karaphun, A.; Swatsitang, E.; Pinitsoontorn, S.; Thongbai, P. DFT calculation and experimental study on structural, optical and magnetic properties of Co-doped SrTiO3. Appl. Surf. Sci. 2018, 446, 92–113. [Google Scholar] [CrossRef]
  97. Ghuman, K.K.; Singh, C.V. Effect of doping on electronic structure and photocatalytic behavior of amorphous TiO2. J. Phys.-Condens Mat. 2013, 25, 475501. [Google Scholar] [CrossRef] [Green Version]
  98. Ren, D.; Li, H.; Cheng, X. Tailoring the electronic and optical properties of anatase TiO2 by (S, Nb) co-doping from a DFT plus U calculation. Solid State Commun. 2015, 223, 54–59. [Google Scholar] [CrossRef]
  99. Gao, L.; Li, Y.; Ren, J.; Wang, S.; Wang, R.; Fu, G.; Hu, Y. Passivation of defect states in anatase TiO2 hollow spheres with Mg doping: Realizing efficient photocatalytic overall water splitting. Appl. Catal. B-Environ. 2017, 202, 127–133. [Google Scholar] [CrossRef]
  100. Chen, W.; Yuan, P.; Zhang, S.; Sun, Q.; Liang, E.; Jia, Y. Electronic properties of anatase TiO2 doped by lanthanides: A DFT+U study. Phys. B 2012, 407, 1038–1043. [Google Scholar] [CrossRef]
  101. Li, H.; Li, W.; Liu, X.; Ren, C.; Miao, X.; Li, X. Engineering of Gd/Er/Lu-triple-doped Bi2MoO6 to synergistically boost the photocatalytic performance in three different aspects: Oxidizability, light absorption and charge separation. Appl. Surf. Sci. 2019, 463, 556–565. [Google Scholar] [CrossRef]
  102. Shi, H.; Lin, Y.; Jiang, Z.; Su, Y.; Ding, X.; Zhang, X.; Zhu, H.; Zhang, R. Enhanced optical absorption and photocatalytic activity of anatase TiO2 through C Nd-codoped: A DFT+ U calculations. J. Phys. Chem. Solids 2017, 109, 70–77. [Google Scholar] [CrossRef]
  103. Jiang, Z.; Wan, W.; Wei, W.; Chen, K.; Li, H.; Wong, P.K.; Xie, J. Gentle way to build reduced titanium dioxide nanodots integrated with graphite-like carbon spheres: From DFT calculation to experimental measurement. Appl. Catal. B-Environ. 2017, 204, 283–295. [Google Scholar] [CrossRef] [Green Version]
  104. Gurkan, Y.Y.; Kasapbasi, E.; Cinar, Z. Enhanced solar photocatalytic activity of TiO2 by selenium(IV) ion-doping: Characterization and DFT modeling of the surface. Chem. Eng. J. 2013, 214, 34–44. [Google Scholar] [CrossRef]
  105. Zhao, Y.; Wang, W.; He, L. The effects of Co/N dopants on the electronic, redox potential, optical, and photocatalytic water-splitting properties of TiO2: First principles calculations. Chem. Phys. Lett. 2017, 685, 108–113. [Google Scholar] [CrossRef]
  106. Wang, P.; Zhou, Q.; Xia, Y.; Zhan, S.; Li, Y. Understanding the charge separation and transfer in mesoporous carbonate-doped phase-junction TiO2 nanotubes for photocatalytic hydrogen production. Appl. Catal. B-Environ. 2018, 225, 433–444. [Google Scholar] [CrossRef]
  107. Wu, X.; Yin, S.; Dong, Q.; Guo, C.; Kimura, T.; Matsushita, J.-I.; Sato, T. Photocatalytic properties of Nd and C Codoped TiO2 with the whole range of visible light absorption. J. Phys. Chem. C 2013, 117, 8345–8352. [Google Scholar] [CrossRef]
  108. Zongyan, Z.; Xiang, Z.; Juan, Y.; Qingju, L. Effects of nonmetal doping on electronic structures and optical property of anatase TiO2 from first-principles calculations. Rare Met. Mat. Eng. 2015, 44, 1568–1574. [Google Scholar] [CrossRef]
  109. 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]
  110. Lin, Y.; Shi, H.; Jiang, Z.; Wang, G.; Zhang, X.; Zhu, H.; Zhang, R.; Zhu, C. Enhanced optical absorption and photocatalytic H2 production activity of g-C3N4/TiO2 heterostructure by interfacial coupling: A DFT+ U study. Int. J. Hydrog. Energy 2017, 42, 9903–9913. [Google Scholar] [CrossRef]
  111. Atanelov, J.; Gruber, C.; Mohn, P. The electronic and magnetic structure of p -element (C,N) doped rutile-TiO2: A hybrid DFT study. Comput. Mater. Sci. 2015, 98, 42–50. [Google Scholar] [CrossRef] [Green Version]
  112. Ghuman, K.K.; Singh, C.V. A DFT + U study of (Rh, Nb)-codoped rutile TiO2. J. Phys.-Condens Mat. 2013, 25, 085501. [Google Scholar] [CrossRef] [Green Version]
  113. Lisovski, O.; Chesnokov, A.; Piskunov, S.; Bocharov, D.; Zhukovskii, Y.F.; Wessel, M.; Spohr, E. Ab initio calculations of doped TiO2 anatase (101) nanotubes for photocatalytical water splitting applications. Mat. Sci. Semicon. Proc 2016, 42, 138–141. [Google Scholar] [CrossRef]
  114. Lisovski, O.; Piskunov, S.; Zhukovskii, Y.F.; Ozolins, J. Ab initiomodeling of sulphur doped TiO2 nanotubular photocatalyst for water-splitting hydrogen generation. Iop Conf. Ser.-Mat. Sci. 2012, 38. [Google Scholar] [CrossRef] [Green Version]
  115. Bocharov, D.; Piskunov, S.; Zhukovskii, Y.F.; Spohr, E.; D’Yachkov, P.N. First principles modeling of 3 d -metal doped three-layer fluorite-structured TiO2(4,4) nanotube to be used for photocatalytic hydrogen production. Vacuum 2017, 146, 562–569. [Google Scholar] [CrossRef]
  116. D’yachkov, E.P.; Bochkov, I.A.; Zaluev, V.A.; D’yachkova, P.N. Electronic properties of titanium dioxide nanotubes doped with 4d metals. Russ. J. Inorg. Chem. 2017, 62, 1048–1050. [Google Scholar] [CrossRef]
  117. Lisovski, O.; Piskunov, S.; Zhukovskii, Y.F.; Bocharov, D. Quantum chemical simulations of titanium dioxide nanotubes used for photocatalytic water splitting. J. Surf. Investig. 2017, 11, 78–86. [Google Scholar] [CrossRef]
  118. Ma, S.; Song, W.; Liu, B.; Zhong, W.; Deng, J.; Zheng, H.; Liu, J.; Gong, X.-Q.; Zhao, Z. Facet-dependent photocatalytic performance of TiO2: A DFT study. Appl. Catal. B-Environ. 2016, 198, 1–8. [Google Scholar] [CrossRef]
  119. Deák, P.; Kullgren, J.; Aradi, B.; Frauenheim, T.; Kavan, L. Water splitting and the band edge positions of TiO2. Electrochim. Acta 2016, 199, 27–34. [Google Scholar] [CrossRef]
  120. Hanaor, D.A.H.; Assadi, M.H.N.; Li, S.; Yu, A.; Sorrel, C.C. Ab Initio Study of Phase Stability in Doped TiO2. Comput. Mech. 2012, 50, 185–194. [Google Scholar] [CrossRef] [Green Version]
  121. Alghamdi, H.; Idriss, H. Study of the modes of adsorption and electronic structure of hydrogen peroxide and ethanol over TiO2 rutile (110) surface within the context of water splitting. Surf. Sci. 2017. [Google Scholar] [CrossRef]
  122. Tsai, C.; Chan, K.; Nørskov, J.K.; Abild-Pedersen, F. Theoretical insights into the hydrogen evolution activity of layered transition metal dichalcogenides. Surf. Sci. 2015, 640, 133–140. [Google Scholar] [CrossRef] [Green Version]
  123. Li, Z.; Shi, L.; Franklin, D.; Koul, S.; Kushima, A.; Yang, Y. Drastic enhancement of photoelectrochemical water splitting performance over plasmonic Al@TiO2 heterostructured nanocavity arrays. Nano Energy 2018, 51, 400–407. [Google Scholar] [CrossRef]
  124. Venturini, J.; Bonatto, F.; Guaglianoni, W.C.; Lemes, T.; Arcaro, S.; Alves, A.K.; Bergmann, C.P. Cobalt-doped titanium oxide nanotubes grown via one-step anodization for water splitting applications. Appl. Surf. Sci. 2019, 464, 351–359. [Google Scholar] [CrossRef]
  125. Dong, Z.; Ding, D.; Li, T.; Ning, C. Ni-doped TiO2 nanotubes photoanode for enhanced photoelectrochemical water splitting. Appl. Surf. Sci. 2018, 443, 321–328. [Google Scholar] [CrossRef]
  126. Li, Y.; Wang, J.-G.; Fan, Y.; Sun, H.; Hua, W.; Liu, H.; Wei, B. Plasmonic TiN boosting nitrogen-doped TiO2 for ultrahigh efficient photoelectrochemical oxygen evolution. Appl. Catal. B-Environ. 2019, 246, 21–29. [Google Scholar] [CrossRef]
  127. Park, J.; Desmukh, P.R.; Sohn, y.; Shin, W.G. ZnO-TiO2 core-shell nanowires decorated with Au nanoparticles for plasmon-enhanced photoelectrochemical water splitting. J. Alloy. Compd. 2019, 787, 1310–1319. [Google Scholar] [CrossRef]
  128. Zhang, S.; Cao, X.-b.; Wu, J.; Zhu, L.-w.; Gu, L. Preparation of hierarchical CuO@TiO2 nanowire film and its application in photoelectrochemical water splitting. T. Nonferr. Metal. Soc. 2016, 26, 2094–2101. [Google Scholar] [CrossRef]
  129. Sawant, S.Y.; Sayed, M.S.; Han, T.H.; Karim, M.R.; Shim, J.J.; Cho, M.H. Bio-synthesis of finely distributed Ag nanoparticle-decorated TiO2 nanorods for sunlight-induced photoelectrochemical water splitting. J. Ind. Eng. Chem. 2019, 69, 48–56. [Google Scholar] [CrossRef]
  130. Liu, J.; Ke, J.; Li, Y.; Liu, B.J.; Wang, L.D.; Xiao, H.N.; Wang, S.B. Co3O4 quantum dots/TiO2 nanobelt hybrids for highly efficient photocatalytic overall water splitting. Appl. Catal. B-Environ. 2018, 236, 396–403. [Google Scholar] [CrossRef]
  131. He, J.; Wang, M.; Wu, X.; Sun, Y.; Huang, K.; Chen, H.; Gao, L.; Feng, S. Influence of controlled Pd nanoparticles decorated TiO2 nanowire arrays for efficient photoelectrochemical water splitting. J. Alloy. Compd. 2019, 785, 391–397. [Google Scholar] [CrossRef]
  132. Wang, L.; Cao, S.; Guo, K.; Wu, Z.; Ma, Z.; Piao, L. Simultaneous hydrogen and peroxide production by photocatalytic water splitting. Chin. J. Catal. 2019, 40, 470–475. [Google Scholar] [CrossRef]
  133. Hussain, E.; Majeed, I.; Nadeem, M.A.; Iqbal, A.; Chen, Y.; Choucair, M.; Jin, R.; Nadeem, M.A. Remarkable effect of BaO on photocatalytic H2 evolution from water splitting via TiO2 (P25) supported palladium nanoparticles. J. Environ. Chem. Eng. 2019, 7, 102729. [Google Scholar] [CrossRef]
  134. Yoo, I.-H.; Kalanur, S.S.; Seo, H. A nanoscale p–n junction photoelectrode consisting of an NiOx layer on a TiO2/CdS nanorod core-shell structure for highly efficient solar water splitting. Appl. Catal. B-Environ. 2019, 250, 200–212. [Google Scholar] [CrossRef]
  135. Wang, Z.W.; Li, X.L.; Tan, C.K.; Qian, C.; Grimsdale, A.C.; Tok, A.I.Y. Highly porous SnO2 nanosheet arrays sandwiched within TiO2 and CdS quantum dots for efficient photoelectrochemical water splitting. Appl. Surf. Sci. 2019, 470, 800–806. [Google Scholar] [CrossRef]
  136. Dong, Z.; Ding, D.; Li, T.; Ning, C. Facile preparation of Ti3+/Ni co-doped TiO2 nanotubes photoanode for efficient photoelectrochemical water splitting. Appl. Surf. Sci. 2019, 480, 219–228. [Google Scholar] [CrossRef]
  137. Hu, M.; Xing, Z.; Cao, Y.; Li, Z.; Yan, X.; Xiu, Z.; Zhao, T.; Yang, S.; Zhou, W. Ti3+ self-doped mesoporous black TiO2/SiO2/g-C3N4 sheets heterojunctions as remarkable visible-lightdriven photocatalysts. Appl. Catal. B-Environ. 2018, 226, 499–508. [Google Scholar] [CrossRef]
  138. Wu, L.L.; Shi, S.; Li, Q.D.; Zhang, X.Y.; Cui, X.L. TiO2 nanoparticles modified with 2D MoSe2 for enhanced photocatalytic activity on hydrogen evolution. Int. J. Hydrog. Energy 2019, 44, 720–728. [Google Scholar] [CrossRef]
  139. Zhang, H.W.; Ma, L.; Ming, J.T.; Liu, B.Q.; Zhao, Y.B.; Hou, Y.D.; Ding, Z.X.; Xu, C.; Zhang, Z.Z.; Long, J.L. Amorphous Ta2OxNy-enwrapped TiO2 rutile nanorods for enhanced solar photoelectrochemical water splitting. Appl. Catal. B-Environ. 2019, 243, 481–489. [Google Scholar] [CrossRef]
  140. Kaur, G.; Pandey, O.P.; Singh, K. Optical, structural, and mechanical properties of different valence-cation-doped bismuth vanadate oxides. Phys. Status Solidi A 2012, 209, 1231–1238. [Google Scholar] [CrossRef]
  141. Yang, S.Y.; Seidel, J.; Byrnes, S.J.; Shafer, P.; Yang, C.H.; Rossell, M.D.; Yu, P.; Chu, Y.H.; Scott, J.F.; Ager Iii, J.W.; et al. Above-bandgap voltages from ferroelectric photovoltaic devices. Nat. Nanotechnol. 2010, 5, 143. [Google Scholar] [CrossRef] [Green Version]
  142. Kay, A.; Cesar, I.; Grätzel, M. New Benchmark for Water Photooxidation by Nanostructured α-Fe2O3 Films. J. Am. Chem. Soc. 2006, 128, 15714–15721. [Google Scholar] [CrossRef] [PubMed]
  143. Wu, X.; Li, H.; Wang, X.; Jiang, L.; Xi, J.; Du, G.; Ji, Z. Ferroelectric enhanced photoelectrochemical water splitting in BiFeO3/TiO2 composite photoanode. J. Alloy. Compd. 2019, 783, 643–651. [Google Scholar] [CrossRef]
  144. Jia, Y.L.; Wang, Z.H.; Ma, Y.; Liu, J.L.; Shi, W.B.; Lin, Y.H.; Hu, X.; Zhang, K. Boosting interfacial charge migration of TiO2/BiVO4 photoanode by W doping for photoelectrochemical water splitting. Electrochim. Acta. 2019, 300, 138–144. [Google Scholar] [CrossRef]
  145. Zhou, W.F.; Jiang, T.F.; Zhao, Y.; Xu, C.; Pei, C.G.; Xue, H.G. Ultrathin Ti/TiO2/BiVO4 nanosheet heterojunction arrays for photoelectrochemical water oxidation. J. Alloy. Compd. 2019, 777, 1152–1158. [Google Scholar] [CrossRef]
  146. Liu, Q.; Mo, R.; Li, X.L.; Yang, S.; Zhong, J.X.; Li, H.X. Cobalt phosphate modified 3D TiO2/BiVO4 composite inverse opals photoanode for enhanced photoelectrochemical water splitting. Appl. Surf. Sci. 2019, 464, 544–551. [Google Scholar] [CrossRef]
  147. Yin, X.; Liu, Q.; Yang, Y.H.; Liu, Y.; Wang, K.K.; Li, Y.M.; Li, D.W.; Qiu, X.Q.; Li, W.Z.; Li, J. An efficient tandem photoelectrochemical cell composed of FeOOH/TiO2/BiVO4 and Cu2O for self-driven solar water splitting. Int. J. Hydrog. Energy 2019, 44, 594–604. [Google Scholar] [CrossRef]
  148. Polo, A.; Grigioni, I.; Dozzi, M.V.; Selli, E. Sensitizing effects of BiVO4 and visible light induced production of highly reductive electrons in the TiO2/BiVO4 heterojunction. Catal. Today 2018. [Google Scholar] [CrossRef]
  149. Radzi, A.A.S.M.; Safaei, J.; Teridi, M.A.M. Photoelectrochemical enhancement from deposition of BiVO4 photosensitizer on different thickness layer TiO2 photoanode for water splitting application. Nano-Struct. Nano-Objects 2019, 18. [Google Scholar] [CrossRef]
  150. Chai, X.B.; Zhang, H.F.; Pan, Q.; Bian, J.L.; Chen, Z.F.; Cheng, C.W. 3D ordered urchin-like TiO2@Fe2O3 arrays photoanode for efficient photoelectrochemical water splitting. Appl. Surf. Sci. 2019, 470, 668–676. [Google Scholar] [CrossRef]
  151. Zhang, W.; Chen, H.; Zhang, L.; Zhang, P.; Dong, E.; Ma, J.; Wang, G. Design and fabrication of an α-Fe2O3 TiO2 Si 3D hierarchical photoanode for improved photoelectrochemical water splitting. J. Alloy. Compd. 2019, 773, 597–604. [Google Scholar] [CrossRef]
  152. Feng, F.; Li, C.; Jian, J.; Qiao, X.; Wang, H.; Jia, L. Boosting hematite photoelectrochemical water splitting by decoration of TiO2 at the grain boundaries. Chem. Eng. J. 2019, 368, 959–967. [Google Scholar] [CrossRef]
  153. Deng, J.; Zhuo, Q.; Lv, X. Hierarchical TiO2 Fe2O3 heterojunction photoanode for improved photoelectrochemical water oxidation. J. Electroanal. Chem. 2019, 835, 287–292. [Google Scholar] [CrossRef]
  154. Yoon, J.W.; Kim, D.H.; Kim, J.H.; Jang, H.W.; Lee, J.H. NH2-MIL-125(Ti)/TiO2 nanorod heterojunction photoanodes for efficient photoelectrochemical water splitting. Appl. Catal. B-Environ. 2019, 244, 511–518. [Google Scholar] [CrossRef]
  155. Yalçın, Y.; Kılıç, M.; Çınar, Z. The Role of Non-Metal Doping in TiO2 Photocatalysis. J. Adv. Oxid. Technol. 2010, 13, 281. [Google Scholar] [CrossRef]
  156. Lu, Y.; Wei, Z.; Salke, N.P.; Yu, L.; Yan, H. Enhanced electron transport in rutile TiO2 nanowires via H2S-assisted incorporation of dissolved silicon for solar-driven water splitting. Appl. Catal. B-Environ. 2019, 244, 767–772. [Google Scholar] [CrossRef]
  157. Bayat, A.; Saievar-Iranizad, E. Graphene quantum dots decorated rutile TiO2 nanoflowers for water splitting application. J. Energy Chem. 2018, 27, 306–310. [Google Scholar] [CrossRef] [Green Version]
  158. Bellamkonda, S.; Thangavel, N.; Hafeez, H.Y.; Neppolian, B.; Rao, G.R. Highly active and stable multi-walled carbon nanotubes-graphene-TiO2 nanohybrid: An efficient non-noble metal photocatalyst for water splitting. Catal. Today 2019, 321, 120–127. [Google Scholar] [CrossRef]
  159. Gao, Q.; Si, F.; Zhang, S.; Fang, Y.; Chen, X.; Yang, S. Hydrogenated F-doped TiO2 for photocatalytic hydrogen evolution and pollutant degradation. Int. J. Hydrog. Energy 2019, 44, 8011–8019. [Google Scholar] [CrossRef]
  160. Elbakkay, M.H.; El Rouby, W.M.A.; El-Dek, S.I.; Farghali, A.A. S-TiO2/S-reduced graphene oxide for enhanced photoelectrochemical water splitting. Appl. Surf. Sci. 2018, 439, 1088–1102. [Google Scholar] [CrossRef]
  161. Dubnova, L.; Zvolska, M.; Edelmannova, M.; Matejova, L.; Reli, M.; Drobna, H.; Kustrowski, P.; Koci, K.; Capek, L. Photocatalytic decomposition of methanol-water solution over N-La/TiO2 photocatalysts. Appl. Surf. Sci. 2019, 469, 879–886. [Google Scholar] [CrossRef]
  162. Liu, X.; Cao, X.E.; Liu, Y.; Li, X.; Wang, M.; Li, M. Branched multiphase TiO2 with enhanced photoelectrochemical water splitting activity. Int. J. Hydrog. Energy 2018, 43, 21365–21373. [Google Scholar] [CrossRef]
  163. Xu, Y.; Ahemd, R.; Klein, D.; Cap, S.; Freedy, K.; McDonnel, S.; Zangari, G. Improving photo-oxidation activity of water by introducing Ti3+ in self-ordered TiO2 nanotube arrays treated with Ar NH3. J. Power Sour. 2019, 414, 242–249. [Google Scholar] [CrossRef]
  164. Xu, Y.; Lin, Q.Y.; Ahmed, R.; Hoglund, E.R.; Zangari, G. Synthesis of TiO2-based nanocomposites by anodizing and hydrogen annealing for efficient photoelectrochemical water oxidation. J. Power Sour. 2019, 410, 59–68. [Google Scholar] [CrossRef]
  165. Wei, N.; Liu, Y.; Feng, M.; Li, Z.X.; Chen, S.G.; Zheng, Y.B.; Wang, D.A. Controllable TiO2 core-shell phase heterojunction for efficient photoelectrochemical water splitting under solar light. Appl. Catal. B-Environ. 2019, 244, 519–528. [Google Scholar] [CrossRef]
  166. Alexander, F.; AlMheiri, M.; Dahal, P.; Abed, J.; Rajput, N.S.; Aubry, C.; Viegas, J.; Jouiad, M. Water splitting TiO2 composite material based on black silicon as an efficient photocatalyst. Sol. Energy Mat. Sol. Cells 2018, 180, 236–242. [Google Scholar] [CrossRef]
  167. Fu, B.; Wu, Z.; Cao, S.; Guo, K.; Piao, L. Effect of aspect ratios of rutile TiO2 nanorods on overall photocatalytic water splitting performance. Nanoscale 2020, 12, 4895–4902. [Google Scholar] [CrossRef] [PubMed]
  168. Abed, J.; AlMheiri, M.; Alexander, F.; Rajput, N.S.; Viegas, J.; Jouiad, M. Enhanced solar absorption of gold plasmon assisted TiO2-based water splitting composite. Sol. Energy Mat. Sol. Cells 2018, 180, 228–235. [Google Scholar] [CrossRef]
  169. Pérez-Larios, A.; Lopez, R.; Hernández-Gordillo, A.; Tzompantzi, F.; Gómez, R.; Torres-Guerra, L.M. Improved hydrogen production from water splitting using TiO2–ZnO mixed oxides photocatalysts. Fuel 2012, 100, 139–143. [Google Scholar] [CrossRef]
  170. Wang, Q.; Huang, J.; Sun, H.; Zhang, K.-Q.; Lai, Y. Uniform carbon dots@TiO2 nanotube arrays with full spectrum wavelength light activation for efficient dye degradation and overall water splitting. Nanoscale 2017, 9, 16046–16058. [Google Scholar] [CrossRef] [PubMed]
  171. Leung, D.Y.; Fu, X.; Wang, C.; Ni, M.; Leung, M.K.; Wang, X.; Fu, X. Hydrogen production over titania-based photocatalysts. ChemSusChem 2010, 3, 681–694. [Google Scholar] [CrossRef]
  172. Makhlouf, A.S.H.; Tiginyanu, I. Nanocoatings and Ultra-Thin Films; Woodhead Publishing: Cambridge, UK, 2011; p. 448. [Google Scholar]
  173. Hernández-Alonso, M.; Fresno, F.; Suárez, S.; Coronado, J. Development of alternative photocatalysts to TiO2: Challenges and opportunities. Energy Environ. Sci. 2009, 2, 1231–1257. [Google Scholar] [CrossRef]
  174. Goto, Y.; Hisatomi, T.; Wang, Q.; Higashi, T.; Ishikiriyama, K.; Maeda, T.; Sakata, Y.; Okunaka, S.; Tokudome, H.; Katayama, M.; et al. A Particulate Photocatalyst Water-Splitting Panel for Large-Scale Solar Hydrogen Generation. Joule 2018, 2, 509–520. [Google Scholar] [CrossRef] [Green Version]
  175. Pinaud, B.A.; Benck, J.D.; Seitz, L.C.; Forman, A.J.; Chen, Z.; Deutsch, T.G.; James, B.D.; Baum, K.N.; Baum, G.N.; Ardo, S.; et al. Technical and economic feasibility of centralized facilities for solar hydrogen production via photocatalysis and photoelectrochemistry. Energy Environ. Sci. 2013, 6, 1983–2002. [Google Scholar] [CrossRef] [Green Version]
  176. Han, L.; Lin, M.; Haussener, S. Reliable Performance Characterization of Mediated Photocatalytic Water-Splitting Half Reactions. ChemSusChem 2017, 10, 2158–2166. [Google Scholar] [CrossRef] [Green Version]
  177. Kisch, H.; Bahnemann, D. Best Practice in Photocatalysis: Comparing Rates or Apparent Quantum Yields? J. Phys. Chem. Lett. 2015, 6, 1907–1910. [Google Scholar] [CrossRef] [PubMed]
  178. Liang, Z.; Zhong, X.; Li, T.; Chen, M.; Feng, G. DFT Study on the Hydrogen Evolution Reaction for Different Facets of Co2P. ChemElectroChem 2019, 6, 260–267. [Google Scholar] [CrossRef]
  179. Skúlason, E.; Karlberg, G.S.; Rossmeisl, J.; Bligaard, T.; Greeley, J.; Jónsson, H.; Nørskov, J.K. Density functional theory calculations for the hydrogen evolution reaction in an electrochemical double layer on the Pt(111) electrode. Phys. Chem. Chem. Phys. 2007, 9, 3241–3250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Parzinger, E.; Miller, B.; Blaschke, B.; Garrido, J.A.; Ager, J.W.; Holleitner, A.; Wurstbauer, U. Photocatalytic Stability of Single- and Few-Layer MoS2. ACS Nano 2015, 9, 11302–11309. [Google Scholar] [CrossRef] [Green Version]
  181. Li, Z.; Meng, X.; Zhang, Z. Recent development on MoS2-based photocatalysis: A review. J. Photoch. Photobio. C 2018, 35, 39–55. [Google Scholar] [CrossRef]
  182. Zhao, Y.; Ding, C.; Zhu, J.; Qin, W.; Tao, X.; Fan, F.; Li, R.; Li, C. A Hydrogen Farm Strategy for Scalable Solar Hydrogen Production with Particulate Photocatalysts. Angew. Chem. Int. Edit. 2020, 59, 9653–9658. [Google Scholar] [CrossRef] [PubMed]
Figure 4. Structural model for (a) possible locations of H dopants and (b) hydrogenated noble-metal doped TiO2. Reprinted with permission from [88]. Copyright 2018, with permission from Elsevier.
Figure 4. Structural model for (a) possible locations of H dopants and (b) hydrogenated noble-metal doped TiO2. Reprinted with permission from [88]. Copyright 2018, with permission from Elsevier.
Molecules 26 01687 g004
Figure 5. Calculated band structure, using generalized gradient approximation (GGA) with Perdew-Burke-Ernzerhof (PBE) exchange-energy, in the Brillouin zone for, (a) Ag-doping, (b) Au-doping, (c) Pt-doping, (d) Pd-doping, and (e) Ru-doping. Reprinted with permission from [88]. Copyright 2018, with permission from Elsevier.
Figure 5. Calculated band structure, using generalized gradient approximation (GGA) with Perdew-Burke-Ernzerhof (PBE) exchange-energy, in the Brillouin zone for, (a) Ag-doping, (b) Au-doping, (c) Pt-doping, (d) Pd-doping, and (e) Ru-doping. Reprinted with permission from [88]. Copyright 2018, with permission from Elsevier.
Molecules 26 01687 g005
Figure 7. Structural models of non-optimized (a) pristine and (b) doped armchair type (4,4) fluorite-structured TiO2 nanotubes. The red balls represent O, grey balls are TiO2, and the turquoise balls represent the 3d-metal dopants substituted for Ti atoms. The nanotube diameter (dnt) is 0.84 nm and the dopant concentration is considered to be 4.17%. Reprinted with permission from [115]. Copyright 2017, with permission from Elsevier.
Figure 7. Structural models of non-optimized (a) pristine and (b) doped armchair type (4,4) fluorite-structured TiO2 nanotubes. The red balls represent O, grey balls are TiO2, and the turquoise balls represent the 3d-metal dopants substituted for Ti atoms. The nanotube diameter (dnt) is 0.84 nm and the dopant concentration is considered to be 4.17%. Reprinted with permission from [115]. Copyright 2017, with permission from Elsevier.
Molecules 26 01687 g007
Figure 8. SEM images of MoSe2 with TiO2 nanoparticles synthesised using a simple hydrothermal method; (a,b) 0.025%, (c,d) 0,05%, and (e,f) 0.1% mass ratio of MoSe2:TiO2. Reprinted with permission from [138]. Copyright 2019, with permission from Elsevier.
Figure 8. SEM images of MoSe2 with TiO2 nanoparticles synthesised using a simple hydrothermal method; (a,b) 0.025%, (c,d) 0,05%, and (e,f) 0.1% mass ratio of MoSe2:TiO2. Reprinted with permission from [138]. Copyright 2019, with permission from Elsevier.
Molecules 26 01687 g008
Figure 9. Schematic illustration of the synthesis of rutile TiO2 with specific (small, medium, and large) aspect ratios. Reprinted with permission from [167]. Copyright 2018, with permission from RSC.
Figure 9. Schematic illustration of the synthesis of rutile TiO2 with specific (small, medium, and large) aspect ratios. Reprinted with permission from [167]. Copyright 2018, with permission from RSC.
Molecules 26 01687 g009
Figure 10. Schematic of the four different reactors. (a) Cross section of the Type 1 reactor showing the particle slurry contained by the baggies and separated by the driveway, (b) Type 2 reactor cross section with separate oxygen and hydrogen baggies connected by a porous bridge, (c) Type 3 reactor utilizing a photoelectrochemical (PEC) cell instead of photocatalytic water splitting (PWS) being directed toward the sun and (d) Type 4 reactor design that combines the PEC cell with an offset parabolic solar concentrator. Drawing not to scale. Reused with permission from [175]. Copyright 2013, with permission from Royal Society of Chemistry.
Figure 10. Schematic of the four different reactors. (a) Cross section of the Type 1 reactor showing the particle slurry contained by the baggies and separated by the driveway, (b) Type 2 reactor cross section with separate oxygen and hydrogen baggies connected by a porous bridge, (c) Type 3 reactor utilizing a photoelectrochemical (PEC) cell instead of photocatalytic water splitting (PWS) being directed toward the sun and (d) Type 4 reactor design that combines the PEC cell with an offset parabolic solar concentrator. Drawing not to scale. Reused with permission from [175]. Copyright 2013, with permission from Royal Society of Chemistry.
Molecules 26 01687 g010
Table 1. This table displays the theoretical bandgap values for different doped TiO2 materials and for a few of these the H2 production rate is presented if published.
Table 1. This table displays the theoretical bandgap values for different doped TiO2 materials and for a few of these the H2 production rate is presented if published.
NanomaterialBandgap [eV]Ref.
Ag doped TiO22.312[88]
Au doped TiO20.996[88]
Pt doped TiO20.754[88]
Pd doped TiO20.363[88]
Ru doped TiO20.176[88]
Wet TiO2 (001)1.8571[89]
Pt doped Wet TiO2 (001)1.4546[89]
Ru doped Wet TiO2 (001)0.1636[89]
Co doped Wet TiO2 (001)0.0539[89]
Ru clusters on TiO2NA[90]
Anatase TiO23.05[91]
Pt adsorbed on TiO23.06[91]
Pd adsorbed on TiO23.05[91]
Rh adsorbed on TiO22.80[91]
Ru adsorbed on TiO23.10[91]
Anatase TiO22.98[92]
Rutile TiO22.78[92]
aTiO2NA[93]
Cu + N co-doped TiO2NA[94]
Cu doped anatase TiO2 (101)NA[95]
Cu doped anatase TiO2 (101)NA[15]
Co-doped SrTiO33.07[96]
N-doped aTiO22.25[97]
S-doped anatase TiO22.33[98]
Nb-doped anatase TiO22.25[98]
(S, Nb)-doped anatase TiO22.15[98]
TiO2 hollow spheres doped with MgNA (H2 production rate: 850 µmol/h/g. O2 production rate: 425 µmol/h/g)[99]
TiO2 doped by lanthanidesNA[100]
C@O-doped TiO23.019[102]
C@gap-doped TiO23.021[102]
Nd@Ti-doped TiO23.032[102]
Nd@gap-doped TiO22.353[102]
C@O&Nd@Ti-doped TiO22.372[102]
C&Nd@gap-doped TiO22.850[102]
TiO2-X2.6 (H2 production rate: 46.9 µmol/h/g)[103]
g-CS@TiO2-X2.5 (H2 production rate: 255.2 µmol/h/g)[103]
g-CS+TiO2-X2.3 (H2 production rate: 68.3 µmol/h/g)[103]
Se(IV) ion doped TiO22.85[104]
N-doped TiO23.06[105]
Co-doped TiO22.92[105]
Co-1N-doped TiO22.91[105]
Co-2N-doped TiO22.90[105]
Co-3N-doped TiO22.92[105]
Mesoporous carbonate-doped phase-junction TiO2 nanotubes2.69-2.92 (H2 production rate: 6108 µmol/h/g)[106]
B-doped TiO22.40[108]
S-doped TiO22.23[108]
C-doped TiO22.53[108]
P-doped TiO22.30[108]
N-doped TiO22.51[108]
F-doped TiO22.61[108]
Cl-doped TiO22.34[108]
N-TiO22.94[109]
Cu-TiO23.22[109]
(Cu, N)-TiO22.96[109]
TiO2/g-C3N42.34[109]
N-TiO2/g-C3N42.31[109]
Cu-TiO2/g-C3N42.23[109]
(Cu, N)-TiO2/g-C3N42.26[109]
g-C3N4/TiO22.21[110]
(C, N)-doped rutile TiO22.59[111]
Rh, Nb co-doped TiO2NA[112]
S, N, or S+N doped TiO2 anatase (101) nanotubes2.78–4.32[113]
S-doped TiO22.72[114]
Sc-doped three-layer fluorite structured TiO24.00[115]
V-doped three-layer fluorite structured TiO23.95[115]
Cr-doped three-layer fluorite structured TiO23.98[115]
Mn-doped three-layer fluorite structured TiO23.66[115]
Fe-doped three-layer fluorite structured TiO23.39[115]
Co-doped three-layer fluorite structured TiO24.01[115]
Ni-doped three-layer fluorite structured TiO24.20[115]
Cu-doped three-layer fluorite structured TiO24.20[115]
Zn-doped three-layer fluorite structured TiO23.60[115]
4d metals doped TiO2 nanotubes2–4[116]
Three-layer TiO2 (101) nanotubes3.83[117]
Six-layer TiO2 (101) nanotubes4.17[117]
Nine-layer TiO2 (001) nanotubes3.95[117]
Six-layer TiO2 (001) nanotubes4.15[117]
Facer dependency of TiO2NA[118]
TiO2NA[119]
Phase stability in TiO2NA[120]
Rutile TiO2NA[121]
aTiO22.70–2.85[14]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Eidsvåg, H.; Bentouba, S.; Vajeeston, P.; Yohi, S.; Velauthapillai, D. TiO2 as a Photocatalyst for Water Splitting—An Experimental and Theoretical Review. Molecules 2021, 26, 1687. https://doi.org/10.3390/molecules26061687

AMA Style

Eidsvåg H, Bentouba S, Vajeeston P, Yohi S, Velauthapillai D. TiO2 as a Photocatalyst for Water Splitting—An Experimental and Theoretical Review. Molecules. 2021; 26(6):1687. https://doi.org/10.3390/molecules26061687

Chicago/Turabian Style

Eidsvåg, Håkon, Said Bentouba, Ponniah Vajeeston, Shivatharsiny Yohi, and Dhayalan Velauthapillai. 2021. "TiO2 as a Photocatalyst for Water Splitting—An Experimental and Theoretical Review" Molecules 26, no. 6: 1687. https://doi.org/10.3390/molecules26061687

Article Metrics

Back to TopTop