Next Article in Journal
Materials Testing for the Development of Biocompatible Devices through Vat-Polymerization 3D Printing
Next Article in Special Issue
Protonation-Induced Enhanced Optical-Light Photochromic Properties of an Inorganic-Organic Phosphomolybdic Acid/Polyaniline Hybrid Thin Film
Previous Article in Journal
TinO2n−1 Suboxide Phases in TiO2/C Nanocomposites Engineered by Non-hydrolytic Sol–Gel with Enhanced Electrocatalytic Properties
Previous Article in Special Issue
Influence of Metal Oxide Particles on Bandgap of 1D Photocatalysts Based on SrTiO3/PAN Fibers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Developments of TiO2-Based Photocatalysis in the Hydrogen Evolution and Photodegradation: A Review

1
Faculty of Chemistry and Chemical Technology, al-Farabi Kazakh National University, Almaty 050000, Kazakhstan
2
Laboratory of Energy-Intensive and Nanomaterials, Institute of Combustion Problems, Almaty 050000, Kazakhstan
3
Department of Engineering Physics, Satbayev University, Almaty 050013, Kazakhstan
4
Laboratory of EPR Spectroscopy Named after Y.V. Gorelkinsky, LLP “Institute of Physics and Technology”, Almaty 050032, Kazakhstan
5
Faculty of Physics and Technology, al-Farabi Kazakh National University, Almaty 050000, Kazakhstan
6
Frank Laboratory of Neutron Physics, Joint Institute for Nuclear Research (JINR), Dubna 141980, Russia
7
Faculty of Physics and Technical Sciences, L.N. Gumilyov Eurasian National University, Astana 010000, Kazakhstan
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(9), 1790; https://doi.org/10.3390/nano10091790
Submission received: 7 August 2020 / Revised: 28 August 2020 / Accepted: 29 August 2020 / Published: 9 September 2020
(This article belongs to the Special Issue Photoactive Nanomaterials)

Abstract

:
The growth of industrialization, which is forced to use non-renewable energy sources, leads to an increase in environmental pollution. Therefore, it is necessary not only to reduce the use of fossil fuels to meet energy needs but also to replace it with cleaner fuels. Production of hydrogen by splitting water is considered one of the most promising ways to use solar energy. TiO2 is an amphoteric oxide that occurs naturally in several modifications. This review summarizes recent advances of doped TiO2-based photocatalysts used in hydrogen production and the degradation of organic pollutants in water. An intense scientific and practical interest in these processes is aroused by the fact that they aim to solve global problems of energy conservation and ecology.

Graphical Abstract

1. Introduction

At present, hydrogen is regarded as the fuel of the future. Compared to carbon fuel, hydrogen is considered a renewable and environmentally friendly source of energy. There are various methods for producing hydrogen on an industrial scale. However, all known methods are characterized by high energy consumption, which makes the process of hydrogen production on a large scale disadvantageous. The production of hydrogen by photocatalytic water splitting is technologically simple, and the outcoming gases are environmentally friendly.
TiO2 is a wide-bandgap semiconductor. In nature, TiO2 is usually found in three different crystalline structures: rutile, anatase, and brookite. TiO2 in anatase form is the most widespread photocatalyst for hydrogen evolution [1]. However, it cannot be used in the spectrum of visible light, since its bandgap (Eg) for different crystalline phases (anatase—3.2 eV, rutile—3.0 eV, and brookite—3.3 eV) is in the UV region. Also, the efficiency of photocatalysis in addition to the bandgap depends on many other factors [2]. In the 1970s, Fujishima and Honda studied the photoelectrochemical splitting of water, where a TiO2-based electrode demonstrated the ability to split water under the influence of ultraviolet radiation [3]. Photocatalysis is a complex reaction consisting of the processes starting from light absorption to generate charge carriers up to surface catalytic reactions due to which gas is formed (Figure 1). This process only requires a photocatalyst (TiO2), which is not consumed during the entire process, water, sunlight or ultraviolet radiation. During the absorption of light (≥ E g ) by the photocatalyst, an excited electron ( e C B ) transfers from the valence band into the conduction band. This transition of the electron leads to the generation of a positively charged carrier-hole ( h V B + ) in the valence band (Equation (1)). Also, these charge carriers can recombine among themselves (Equation (2)) [4,5].
T i O 2   h v E g   h V B + +   e C B
h V B + +   e C B   r e c o m b i n a t i o n   E n e r g y
Another important application of TiO2-based photocatalysts is based on their ability to discolor and completely decompose organic dyes contained in water. In addition, there are a number of works in which TiO2-based photocatalysts were used to neutralize harmful to the atmosphere gases [6,7,8,9]. Chemical stability, easy accessibility, nontoxicity and the ability to oxidize under the influence of radiation, allow TiO2-based photocatalysts to solve the main global problems associated with pollution of the environment and need for renewable energy [10,11]. Currently, TiO2 as a photocatalyst is commercially produced in powder form. The most common brand of TiO2 based photocatalyst are P-25 Degussa/Evonik, TiO2 nanofibers from Kertak, TiO2 Millennium PC-500, TiO2 Hombikat UV-100, Sigma Aldrich TiO2, TiO2 PK-10, and P90 Aeroxide.
This paper aims to provide a brief overview of articles published starting from 2017 on new research and developments of TiO2 based photocatalysts with significant advances. In this review, attention is also paid to the study on the mechanism of photocatalytic processes, factors affecting the activity of photocatalysts, and new techniques used to increase the activity of TiO2 based photocatalysts during the splitting of water with the evolution of hydrogen and decomposition of organic compounds utilized for water purification.

2. Factors Affecting the Efficiency of Photocatalysis and Techniques Used to Improve the Efficiency of TiO2-Based Photocatalysts

2.1. Lifetime of Photogenerated Charge Carriers

The activity of photocatalysts strongly depends on the lifetime of photogenerated electron-hole pairs. An important role is played by the rate at which charge carriers can reach the surface of the photocatalysts. The results of spectroscopic studies show that the time intervals between redox reactions or recombination involving charge carriers are extremely short, resulting in a significant reduction of the photocatalytic activity of TiO2. In the case of recombination of charge carriers in a sufficiently fast interval (<0.1 ns), the photocatalytic activity of the semiconductor is not observed. For example, the lifetime of electron-hole pairs of ~250 ns (TiO2) is considered relatively long [12]. Thus, it can be concluded that a high recombination rate and barriers, that prevent the transfer of charge carriers to the semiconductor surface, reduce photoactivity, despite the high concentration of initially photogenerated pairs. In this regard, to avoid their recombination, it becomes necessary to use cocatalysts in order to increase the lifetimes of electrons and holes.

2.2. The Particle Size of Photocatalyst

Compared to microparticles, TiO2 nanoparticles have, generally, a higher photocatalytic activity [13,14]. This is due to the small diameter of the nanoparticles, in which the charge requires minimal effort to transfer to the surface. If the particle size decreases, the distance that photogenerated electrons and holes need to travel to the surface where the reactions take place is reduced, thereby reducing the probability of recombination. For TiO2 photocatalyst microparticles, the penetration depth of UV rays is limited and amounts to about ~100 nm. This means that the inner part of the TiO2 photocatalyst microparticle remains in a passive state. [15]. Figure 2 shows the scheme of light absorption by nano- and microparticles of TiO2. This is one of the reasons for the increased interest in nanosized particles of TiO2.
As shown in Figure 2, a decrease in the particle size of the photocatalyst to nanoscale facilitates the absorption of light by the entire volume of particles. However, there is a limitation regarding the minimum sizes to which it is desirable to reduce the particles of the photocatalyst, due to the onset of quantum effects. They become significant at particle sizes less than 2 nm for both anatase and rutile, and finally, this leads to a change in the bandgap. An increase in the size of the photocatalyst crystals leads to a decrease in the recombination of the electron-hole pair at the defects of the crystal lattice and to an increase in photocatalytic activity. For example, in [16], nanoparticles with a size of 25 nm were found to be more productive than nanoparticles of 15 nm. On the other hand, the bandgap is directly proportional to the size of the photocatalyst crystals. That is why it is necessary to find the optimal crystal size of TiO2 and control it in the process of its obtaining.
To increase the photoactivity of TiO2 in the visible region of the spectrum, the spectral region of its absorption should be expanded. There are several approaches for sensitizing TiO2 to visible light: doping with cationic and anionic elements or metal nanoparticles. Elements of the 3d- and 2p-groups are often used as additives to reduce the value of the bandgap of the photocatalyst [17,18].

2.3. Doping with Cations

The essence of cationic doping is the introduction of metal cations into the crystal structure of TiO2 at the position of Ti4+ ions. Rare-earth, noble, and transition metal cations can be used [19]. Doping with cations significantly expands the absorption spectrum of TiO2, increases the redox potential of the formed radicals, and increases quantum efficiency by reducing the degree of recombination of electrons and holes. The nature and concentration of the dopant change the charge distribution on the TiO2 surface and affects the process of photo corrosion and photocatalytic activity [20]. However, an increase in the absorption of visible light does not always lead to an increase in the activity of the photocatalyst. As a result of doping with cations, a certain number of defects appear in the TiO2 structure, which can act as charge recombination centers, this leading to a decrease in photocatalytic activity even under the influence of UV light.
In [21], Kryzhitsky et al. show the change in the activity of photocatalytic properties of rutile and anatase forms of nanocrystalline TiO2 depending on the nature of metal-based dopants. According to the results of the study, it is found that doping does not significantly change the bandgap of rutile, while in the case of doping anatase with iron and chromium, its bandgap narrows significantly. As a result of doping with metals, the photocatalytic activity of anatase (A) increases in the following order: A < A/Co < A/Cu < A/Fe. In the case of rutile (R), its photocatalytic activity decreases in the following order: R > R/Co > R/Cu > R/Fe > R/Cr. According to the authors, the decrease in the photoactivity of TiO2 may be associated with the inhibitory effect of impurity cations.

2.4. Doping with Anionic Elements

Over the past few years, it has been shown that TiO2 samples doped with nonmetallic elements (nitrogen, carbon, sulfur, boron, phosphorus, and fluorine) in the anionic positions of TiO2, demonstrate high photoactivity in the UV and visible regions of the solar spectrum [22,23]. Among all the anions, carbon, nitrogen, and fluorine caused the most significant interest [24,25,26]. The substitution of oxygen atoms to carbon leads to the formation of new levels (C2p) above the ceiling of the valence band of TiO2 (O2p), which reduces the bandgap and shifts the absorption spectrum. The inclusion of carbon in TiO2 can also lead to the formation of carbon compounds on the surface of the photocatalyst, which acts as absorption centers of visible radiation [27].
Doping with nitrogen atoms is the most popular way to improve the photocatalytic performance of TiO2. Introduction of nitrogen into the TiO2 structure contributes to a significant shift of the absorption spectrum into the visible region of the solar spectrum, a change in the refractive index, an increase in hardness, electrical conductivity, elastic modulus, and photocatalytic activity in regard to visible light [28,29]. Upon substitution of anions, a new level is formed above the valence band of TiO2 [30]. As shown in Figure 3, the presence of nitrogen leads to a change of the bandgap Eg1 (TiO2) > Eg2 (N-doped TiO2), thus contributing to the absorption of photons of light with lower energy.
Doping with nitrogen in the oxygen position is difficult since the ionic radius of nitrogen (1.71 Å) is much larger than that for oxygen (1.4 Å). Another auspicious element in the anionic positions of TiO2 is fluorine atoms [31]. Unlike nitrogen, fluorine atoms easily replace oxygen due to the close ion radius (1.33 Å for F and 1.4 Å for O2−). The increase in photocatalytic activity is mainly associated with an improvement in the degree of crystallinity of TiO2 due to doping with fluorine [32]. It has been determined that crystallinity and the specific surface area also affect the photoactivity of TiO2 [33,34]. The crystalline modification of TiO2 in comparison with amorphous TiO2 has significantly fewer defects, which reduce the possibility of recombination processes and contributes to the efficient movement of photogenerated charge carriers in the semiconductor. Since redox reactions occur on the surface of TiO2, one of the main requirements for photocatalysts is the presence of a developed specific surface area. However, the presence of a developed specific surface area implies a large number of defects in the structure and a low degree of crystallinity and, as a result, reduces photocatalytic activity. Therefore, to increase photocatalytic activity, it is important to find a balance between the above factors.

2.5. Doping/Loading with Metal Nanoparticles

The application of metal nanoparticles is another alternative approach to the modification of photocatalysts. A review of the recent literature shows that metals (Co, Pt, Ag, Au, Pd, Ni, Cu, Eu, Fe, etc.) significantly increase the photocatalytic activity of TiO2 [35,36,37]. The low location of the Fermi level of these metals compared to TiO2 can lead to the movement of electrons from the TiO2 structure to metal particles deposited on its surface. This helps to avoid the recombination of charge carriers, since the holes remain in the valence band of TiO2. This is also beneficial for avoiding the recombination of charge carriers since the holes remain in the valence band of TiO2. A number of conducted investigations indicate that the properties of these photocatalysts depend on the dispersion of metal particles [38,39]. Enhanced photocatalytic properties of metals appear when their size decreases <2.0 nm [40]. Despite the foregoing, too high concentration of metal particles can block the surface of TiO2 and prevent the absorption of photons, leading to a decrease in the efficiency of the photocatalyst.

3. The Utilization of Photocatalysts Based on TiO2

3.1. Hydrogen Evolution

Hydrogen can be produced by using nanoscale TiO2 based photocatalysts with various morphologies in the form of nanowires, nanospheres, nanorods, nanotubes, and nanosheets [41]. Table 1 lists some TiO2 nanocomposites with different structures, as well as their photocatalytic characteristics.
P. Melián et al. [42] demonstrated that loading of nanosized TiO2 microspheres with Au and Pt metals increases the hydrogen evolution twice. In the case of using Au, the maximum hydrogen evolution was determined at 1.5 wt.% content of Au in the sample with the yield of hydrogen 1118 µmol h−1, while for Pt its optimal content in the sample was 0.27 wt.% with the yield of hydrogen 2125 µmol h−1. The excess of dopants may result in decrease of photocatalyst activity due to possible complete coverage of TiO2 surface, thus, hindering the light to be absorbed. A literature review also showed that the difference in the optimal ratio for each metal could be associated with the formation of recombination centers on the semiconductor surface by metal particles [51,52].
A positive effect on the rate of H2 production was found when using sacrificial agents acting as electron donors (hole scavengers) during photoreforming, in which the hydroxyl radical is consumed by the sacrificial agents. In general, there are two types of sacrificial reagents: organic and inorganic based electron donors. Among organic electron donors, the most effective are water–alcohol mixtures, in particular, methanol > ethanol > ethylene glycol > glycerol [53,54,55,56,57,58,59]. However, an increase in the concentration of the sacrificial agent does not always lead to an increase in the yield of hydrogen. Y.-K. Park et al. [60] showed that the rate of hydrogen evolution also increases depending on the concentration of methanol (Figure 4a). At low concentrations, the rate of hydrogen formation in solutions is proportional to the concentration of methanol, while at higher concentrations, it approaches to a constant value [61]. Nevertheless, after adding a certain amount of methanol and ethanol, a further increase in its concentration leads to a decrease in the rate of hydrogen evolution (Figure 4b). The yield of hydrogen during photoreforming has a maximum output during 80−90 min and after it decreases. This is due to the formation of a significant amount of methane and ethane, during which photogenerated electrons ( e C B ) and holes ( h V B + ) are consumed [54].
The amount of dopant also has a significant effect on the efficiency of light absorption by the photocatalyst and on its photocatalytic activity. For example, Udayabhanu et al. [62] prepared Cu-TiO2/CuO nanocomposites containing different amount of Cu. The color of the obtained samples, depending on the concentration of Cu-TiO2/CuO (CUT) from 1 to 4 mol% (the samples were as named CUT 1, CUT 2, CUT 3, and CUT 4) changes from light green to dark green (Figure 5). To evaluate the photocatalytic activity of hydrogen production, scientists compared the activities of obtained nanocomposites with conventional TiO2. Sunlight was used as the source of radiation, and glycerol was chosen as a sacrificial agent. According to the results, the CUT 3 sample based on Cu-TiO2/CuO possessed the highest yield of hydrogen (10.453 µmol h−1 g−1 of H2 under sunlight and 4.714 µmol h−1 g−1 of H2 under visible light). Nevertheless, this type of photocatalyst is effective only for hydrogen production, since it has shown low efficiency in the decomposition of organic dye and metal detoxification in water.
The photocatalyst efficiency is affected not only by the nature of the alloying element but also by its concentration. For example, X. Xing et al. in [63] demonstrated the dependence of the yield of hydrogen on the light intensity and the concentration of the dopant. From Figure 6a, it is clear that the increase of concentrations of each photocatalyst (pure TiO2 and Au/TiO2) results in increase of the rate of hydrogen production. However, the hydrogen generation rate for Au/TiO2 photocatalyst at the same light intensity is 18–21 times higher than that for pure TiO2. The explanation to this is that Au can limit charge carriers’ recombinations and, what is more, the visible light absorption of photocatalyst is enhanced by the localized surface plasmon resonance effect of Au nanosized particles. The influence of the light intensity (from 1 to 9 kW/m2) and the duration of exposure on the rate of hydrogen generation for Au/TiO2 nanoparticles with a concentration of 1 g/L is also shown in Figure 6b.
The photocatalytic properties of the material can be improved by creating composites based on different photocatalysts. For example, E.-C. Su et al. [64] obtained a composite photocatalyst based on Pt/N-TiO2/SrTiO3-TiO2 in the form of nanotubes using a two-stage hydrothermal process (SrTiO3 is also a photocatalyst with a bandgap of 3.2 eV with a perovskite-type structure [65]). The obtained results showed that this composite photocatalyst is able to operate under sunlight with the rate of hydrogen evolution up to 3873 µmol/h/g.
In practice, photocatalytic reactions are mainly carried out at room temperature. It is found that increasing temperature has a positive effect on the activity of some photocatalysts, which makes it relevant to develop new photocatalysts based on the anatase form of TiO2 with a thermostable phase.

3.2. Photodegradation

Organic dyes are used on a large scale in modern industries. Wastewater polluted with such substances subsequently leads to environmental problems [66,67]. This is due to the fact that most organic dyes consist of biodegradable aromatic structures and azo-groups. Adsorption and catalytic oxidation remain the most effective among the methods for treating wastewater from dyes [68,69,70]. For photocatalytic degradation of organic compounds, in most cases, TiO2, in the doped form, is used.
Graphene has a large specific surface area and electrical conductivity. The presence of such properties in graphene is of interest to scientists in the preparation of a graphene/nano-TiO2-based photocatalyst [71,72,73]. After irradiation of a TiO2-based photocatalyst, electrons move into the graphene structure. Based on the author’s results [74], it helps to avoid recombination between charge carriers. TiO2/graphene-based photocatalyst can be obtained by creating either a chemical bond between them or via the approach of their mechanically mixing [75,76]. Due to existence of chemical bond, the electron is transferred unhindered from the photocatalyst to graphene, reducing the probability of recombination. This is the main explanation for the increased activity of the photocatalyst with graphene.
Composites based on TiO2/graphene attract scientists’ attention not only as highly efficient photocatalysts due to light absorption in a wide range and charge separation, but also taking in account its high adsorption capacity to pollutants. However, TiO2/graphene-based composite produced by the hydrothermal method cannot serve as an effective photocatalyst. The reason is the agglomeration of graphene layers, which adversely affects the adsorption and photocatalytic properties. The solution to this problem is described in [77], where graphene-based aerogel was used as an auxiliary material for TiO2 [78,79]. Aerogel was obtained by thermal reduction of graphene oxide. Table 2 lists some reports on the use of composite TiO2/graphene based photocatalysts to remove various organic compounds (pollutants and dyes) from water.
Before testing photocatalytic activity, it was necessary to saturate the materials in the dark conditions. If saturation is not performed at dark conditions, a decrease in the concentration of the target pollutant or the color intensity change of the used in experiment dye will be associated not only with photocatalysis, but also with adsorption and photocatalytic degradation, so the effectiveness of the photocatalyst will be incorrect. X. Sun et al. [77] compared two samples of a photocatalyst of the same mass: hydrothermally obtained TiO2-reduced graphene oxide (rGO), and aerogel based on TiO2-rGO (Figure 7a). As can be seen, despite the equal mass of both samples, the TiO2-rGO based aerogel has a larger specific volume than that of the second sample. For further characterization of their adsorption, both samples were used to adsorb methylene blue in the dark. During the experiment, the absorption intensity was analyzed. The results showed that it took 2 min for TiO2-rGO based aerogel to achieve adsorption saturation, while for TiO2-rGO powder, it took more than 10 min (Figure 7b). Figure 7c,d demonstrates the color change of methylene blue, which proves the high absorption coefficient of visible light by TiO2-rGO aerogels. According to the authors, the adsorption rate also affects the efficiency of photocatalysts.
Photocatalytic activity directly depends on the number of active centers on the surface of the photocatalyst. Photolithography is a block of technological processes of photochemical technology aimed at creating the relief in the film, as well as a film of metal deposited on a substrate [87,88]. Using this method, a group of scientists [89] managed to obtain TiO2 films with lattice, square, and hexagonal structures (Figure 8) and investigate the influence of a such surface textures on the photocatalysis. Obtained results showed that the activity of photocatalyst in the form of a film is not improved by increasing the values of specific surface area. Surface texture also has an effect on mass transfer during photocatalysis.
To evaluate the activity of photocatalysts, scientists conducted an experiment, in which the dye degradation occurred in water as a result of its exposure with UV (254 nm) on methyl orange. As shown in Figure 9, the microstructured TiO2 films exhibit a more active photocatalytic activity than TiO2 films with a flat surface. According to scientists, this is due to the presence of reaction centers on the surface of microstructured films. The highest photocatalytic activity was observed for TiO2 films with a square microstructure. The experimental results showed that the efficiency of photocatalysts is influenced not only by the surface area, but also by the type of their microstructure. It was revealed, that the TiO2 film with a grating structure, despite its low specific surface area, possessed photoactive properties similar to the TiO2 film with a hexagonal structure [89].
A significant role during photocatalysis is played by the surface microstructure, which influences the mass transfer of degradable organic compounds by the diffusion to the surface of the photocatalyst. It is important to take into account the fact that the fluid flowing over the surface with protrusions meets more resistance from the side of the walls, compared to the fluid flowing through the flat surface. As a result, such protrusions can adversely affect the degradation efficiency of organic pollutants.
TiO2 doped with Fe2O3 is one of the best-known effective photocatalysts. Due to the narrow bandgap of Fe2O3 (2.2 eV), its doping leads to a redshift of the light response of the photocatalyst. The phenomenon of decreasing of photoactivity of TiO2 based photocatalyst doped with non-metals during heating, the high cost of some metals and the availability of Fe2O3 in large quantities increases the attractiveness of Fe2O3 over other dopants [90]. J.-J. Zhang et al. [91] used Fe2O3 nanoparticles, which served as a doping agent for obtaining the TiO2/graphene aerogel (GA) based photocatalyst with a 3D structure (Figure 10). Due to the narrow bandgap (2.0 eV), Fe2O3 can easily generate electron-hole pairs, thereby contributing to the photodegradation of rhodamine B even in visible light. The results showed that the aqueous solution containing rhodamine B (RhB) was purified to 97.7% (Figure 10 b). Fe3O4 can also be used as a doping agent for the photocatalysis [92]. For example, F. Soltani-Nezhad et al. [93] presented a method for producing a GO/Fe3O4/TiO2-NiO-based photocatalyst, which is able to efficiently degrade imidacloprid (pesticide).
To increase the efficiency of the process of degradation of organic pollutants by photocatalysis, attempts have been made to combine radiation with ultrasonic cavitation. S. Rajoriya et al. [94] reported the photodegradation of 4-acetamidophenol to 91% using Sm (samarium) and N-doped TiO2 photocatalysts, in which a combination of UV radiation, hydrodynamic cavitation, and ultrasound was applied. In another work, the use of N and Cu-doped TiO2@CNTs in sono-photocatalysis for purification of pharmaceutical wastewater is discussed [95]. The results showed that when using a commercial Xe lamp (50 W) and ultrasound for pharmaceutical wastewaters treatment, the photocatalyst removal efficiency within 180 min were 100, 93, and 89% for sulfamethoxazole, chemical oxygen demand, and total organic carbon, respectively.
The great interest in the use of ultrasonic action during photocatalysis is justified by the enhancement of electronic excitations, which leads to an increase in the density of pairs of charged particles. Under the influence of ultrasound, the aggregate photocatalysts are dispersed, contributing to the rapid renewal and expansion of the boundaries of heterogeneous reactions, which improves the mass transfer and the course of chemical reactions [96,97].

4. Conclusions

In this review, information on the main mechanisms of water splitting and decomposition of organic compounds by TiO2-based photocatalysts was collected and analyzed. The efficiency of TiO2 doping by various components, which significantly increases the photocatalytic activity, is shown. In addition, other factors, such as the bandgap, charge carrier recombination, the use of sacrificial agents, the size and type of photocatalyst structures, surface morphology, photocatalyst concentration in solution, radiation source power, etc., which can affect the photocatalysis were analyzed. Despite the reports of scientists on the development of effective photocatalysts capable of operating under sunlight, their use is still not economically viable. Therefore, due to the growth of environmental problems and the limited availability of hydrocarbon fuels, investigations in the field of hydrogen production by “artificial photosynthesis” will undoubtedly increase.

Author Contributions

Conceptualization, B.B. and C.D.; writing—original draft preparation, F.S., R.B., A.B., I.C. and B.B.; writing—review and editing, A.U. and A.M.; project administration, R.B. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by Ministry of Education and Science of the Republic of Kazakhstan № AP05135273. The name of the focal area of the development of science, on which the application is submitted: Energy and Machine industry. The name of the specialized research area, on which the application is filed, type of study: Alternative energy and technology: renewable energy sources, nuclear and hydrogen energy, other energy sources, applied research.

Acknowledgments

Askhat Bekbaev is grateful to al-Farabi Kazakh National University for financial support in the frame of the postdoctoral fellowship program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Samokhvalov, A. Hydrogen by photocatalysis with nitrogen codoped titanium dioxide. Renew. Sustain. Energy Rev. 2017, 72, 981–1000. [Google Scholar] [CrossRef]
  2. Phoon, B.L.; Lai, C.W.; Juan, J.C.; Show, P.-L.; Pan, G.-T. Recent developments of strontium titanate for photocatalytic water splitting application. Int. J. Hydrogen Energy 2019, 44, 14316–14340. [Google Scholar] [CrossRef]
  3. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  4. Xiao, N.; Li, S.; Li, X.; Ge, L.; Gao, Y.; Li, N. The roles and mechanism of cocatalysts in photocatalytic water splitting to produce hydrogen. Chin. J. Catal. 2020, 41, 642–671. [Google Scholar] [CrossRef]
  5. Katal, R.; Masudy-Panah, S.; Tanhaei, M.; Farahani, M.H.; Jiangyong, H. A review on the synthesis of the various types of anatase TiO2 facets and their applications for photocatalysis. Chem. Eng. J. 2020, 384, 123384. [Google Scholar] [CrossRef]
  6. Bingham, M.; Mills, A. Photonic efficiency and selectivity study of M (M = Pt, Pd, Au and Ag)/TiO2 photocatalysts for methanol reforming in the gas phase. J. Photochem. Photobiol. A Chem. 2020, 389, 112257. [Google Scholar] [CrossRef]
  7. Paz, Y. Application of TiO2 photocatalysis for air treatment: Patents’ overview. Appl. Catal. B Environ. 2010, 99, 448–460. [Google Scholar] [CrossRef]
  8. Shayegan, Z.; Lee, C.-S.; Haghighat, F. TiO2 photocatalyst for removal of volatile organic compounds in gas phase—A review. Chem. Eng. J. 2018, 334, 2408–2439. [Google Scholar] [CrossRef]
  9. Verbruggen, S.W. TiO2 photocatalysis for the degradation of pollutants in gas phase: From morphological design to plasmonic enhancement. J. Photochem. Photobiol. C Photochem. Rev. 2015, 24, 64–82. [Google Scholar] [CrossRef]
  10. Kou, J.; Lu, C.; Wang, J.; Chen, Y.; Xu, Z.; Varma, R.S. Selectivity Enhancement in Heterogeneous Photocatalytic Transformations. Chem. Rev. 2017, 117, 1445–1514. [Google Scholar] [CrossRef] [Green Version]
  11. Parrino, F.; Bellardita, M.; García-López, E.I.; Marcì, G.; Loddo, V.; Palmisano, L. Heterogeneous Photocatalysis for Selective Formation of High-Value-Added Molecules: Some Chemical and Engineering Aspects. ACS Catal. 2018, 8, 11191–11225. [Google Scholar] [CrossRef]
  12. Yang, Q.; Dong, L.; Su, R.; Hu, B.; Wang, Z.; Jin, Y.; Wang, Y.; Besenbacher, F.; Dong, M. Nanostructured heterogeneous photo-catalysts for hydrogen production and water splitting: A comprehensive insight. Appl. Mater. Today 2019, 17, 159–182. [Google Scholar] [CrossRef]
  13. Vajda, K.; Saszet, K.; Kedves, E.Z.; Kása, Z.; Danciu, V.; Baia, L.; Magyari, K.; Hernádi, K.; Kovács, G.; Pap, Z. Shape-controlled agglomeration of TiO2 nanoparticles. New insights on polycrystallinity vs. single crystals in photocatalysis. Ceram. Int. 2016, 42, 3077–3087. [Google Scholar] [CrossRef] [Green Version]
  14. Li, D.; Song, H.; Meng, X.; Shen, T.; Sun, J.; Han, W.; Wang, X. Effects of Particle Size on the Structure and Photocatalytic Performance by Alkali-Treated TiO2. Nanomaterials 2020, 10, 546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Nagaraj, G.; Irudayaraj, A.; Josephine, R.L. Tuning the optical band Gap of pure TiO2 via photon induced method. Optik 2019, 179, 889–894. [Google Scholar]
  16. Allen, N.S.; Vishnyakov, V.; Kelly, P.J.; Kriek, R.J.; Mahdjoub, N.; Hill, C. Characterisation and photocatalytic assessment of TiO2 nano-polymorphs: Influence of crystallite size and influence of thermal treatment on paint coatings and dye fading kinetics. J. Phys. Chem. Solids 2019, 126, 131–142. [Google Scholar] [CrossRef] [Green Version]
  17. Sinhmar, A.; Setia, H.; Kumar, V.; Sobti, A.; Toor, A.P. Enhanced photocatalytic activity of nickel and nitrogen codoped TiO2 under sunlight. Environ. Technol. Innov. 2020, 18, 100658. [Google Scholar] [CrossRef]
  18. Liu, X.; Li, Y.; Wei, Z.; Shi, L. A Fundamental DFT Study of Anatase (TiO2) Doped with 3d Transition Metals for High Photocatalytic Activities. J. Wuhan Univ. Technol. Mater. Sci. Ed. 2018, 33, 403–408. [Google Scholar] [CrossRef]
  19. Benjwal, P.; De, B.; Kar, K.K. 1-D and 2-D morphology of metal cation co-doped (Zn, Mn) TiO2 and investigation of their photocatalytic activity. Appl. Surf. Sci. 2018, 427, 262–272. [Google Scholar] [CrossRef]
  20. Shen, Q.; Wang, Y.; Xue, J.; Gao, G.; Liu, X.; Jia, H.; Li, Q.; Xu, B. The dual effects of RGO films in TiO2/CdSe heterojunction: Enhancing photocatalytic activity and improving photocorrosion resistance. Appl. Surf. Sci. 2019, 481, 1515–1523. [Google Scholar] [CrossRef]
  21. Kernazhitsky, L.; Shymanovska, V.; Gavrilko, T.; Naumov, V.; Kshnyakin, V.; Khalyavka, T. A comparative study of optical absorption and photocatalytic properties of nanocrystalline single-phase anatase and rutile TiO2 doped with transition metal cations. J. Solid State Chem. 2013, 198, 511–519. [Google Scholar] [CrossRef]
  22. Song, E.; Kim, Y.-T.; Choi, J. Anion additives in rapid breakdown anodization for nonmetal-doped TiO2 nanotube powders. Electrochem. Commun. 2019, 109, 106610. [Google Scholar] [CrossRef]
  23. Fadlallah, M.M. Magnetic, electronic, optical, and photocatalytic properties of nonmetal- and halogen-doped anatase TiO2 nanotubes. Phys. E Low Dimens. Syst. Nanostruct. 2017, 89, 50–56. [Google Scholar] [CrossRef]
  24. Varnagiris, S.; Medvids, A.; Lelis, M.; Milcius, D.; Antuzevics, A. Black carbon-doped TiO2 films: Synthesis, characterization and photocatalysis. J. Photochem. Photobiol. A Chem. 2019, 382, 111941. [Google Scholar] [CrossRef]
  25. Li, G.; Zou, B.; Feng, S.; Shi, H.; Liao, K.; Wang, Y.; Wang, W.; Zhang, G. Synthesis of N-Doped TiO2 with good photocatalytic property. Phys. B Condens. Matter 2020, 588, 412184. [Google Scholar] [CrossRef]
  26. Li, H.; Qiu, L.; Bharti, B.; Dai, F.; Zhu, M.; Ouyang, F.; Lin, L. Efficient photocatalytic degradation of acrylonitrile by Sulfur-Bismuth co-doped F-TiO2/SiO2 nanopowder. Chemosphere 2020, 249, 126135. [Google Scholar] [CrossRef]
  27. He, D.; Li, Y.; Wu, J.; Yang, Y.; An, Q. Carbon wrapped and doped TiO2 mesoporous nanostructure with efficient visible-light photocatalysis for NO removal. Appl. Surf. Sci. 2017, 391, 318–325. [Google Scholar] [CrossRef]
  28. Boningari, T.; Inturi, S.N.R.; Suidan, M.; Smirniotis, P.G. Novel one-step synthesis of nitrogen-doped TiO2 by flame aerosol technique for visible-light photocatalysis: Effect of synthesis parameters and secondary nitrogen (N) source. Chem. Eng. J. 2018, 350, 324–334. [Google Scholar] [CrossRef]
  29. Jiang, J.X.; Zhang, Q.Q.; Li, Y.H.; Li, L. Three-dimensional network graphene aerogel for enhancing adsorption and visible light photocatalysis of nitrogen-doped TiO2. Mater. Lett. 2019, 234, 298–301. [Google Scholar] [CrossRef]
  30. Marques, J.; Gomes, T.D.; Forte, M.A.; Silva, R.F.; Tavares, C.J. A new route for the synthesis of highly-active N-doped TiO2 nanoparticles for visible light photocatalysis using urea as nitrogen precursor. Catal. Today 2019, 326, 36–45. [Google Scholar] [CrossRef]
  31. Du, M.; Qiu, B.; Zhu, Q.; Xing, M.; Zhang, J. Fluorine doped TiO2/mesocellular foams with an efficient photocatalytic activity. Catal. Today 2019, 327, 340–346. [Google Scholar] [CrossRef]
  32. Li, C.; Sun, Z.; Ma, R.; Xue, Y.; Zheng, S. Fluorine doped anatase TiO2 with exposed reactive (001) facets supported on porous diatomite for enhanced visible-light photocatalytic activity. Microporous Mesoporous Mater. 2017, 243, 281–290. [Google Scholar] [CrossRef]
  33. Wei, J.Q.; Chen, X.J.; Wang, P.F.; Han, Y.B.; Xu, J.C.; Hong, B.; Jin, H.X.; Jin, D.F.; Peng, X.L.; Li, J.; et al. High surface area TiO2/SBA-15 nanocomposites: Synthesis, microstructure and adsorption-enhanced photocatalysis. Chem. Phys. 2018, 510, 47–53. [Google Scholar] [CrossRef]
  34. Hafez, H.S. Synthesis of highly-active single-crystalline TiO2 nanorods and its application in environmental photocatalysis. Mater. Lett. 2009, 63, 1471–1474. [Google Scholar] [CrossRef]
  35. Vargas Hernández, J.; Coste, S.; García Murillo, A.; Carrillo Romo, F.; Kassiba, A. Effects of metal doping (Cu, Ag, Eu) on the electronic and optical behavior of nanostructured TiO2. J. Alloys Compd. 2017, 710, 355–363. [Google Scholar] [CrossRef]
  36. Belošević-Čavor, J.; Koteski, V.; Umićević, A.; Ivanovski, V. Effect of 5d transition metals doping on the photocatalytic properties of rutile TiO2. Comput. Mater. Sci. 2018, 151, 328–337. [Google Scholar] [CrossRef]
  37. Unal, F.A.; Ok, S.; Unal, M.; Topal, S.; Cellat, K.; Şen, F. Synthesis, characterization, and application of transition metals (Ni, Zr, and Fe) doped TiO2 photoelectrodes for dye-sensitized solar cells. J. Mol. Liq. 2020, 299, 112177. [Google Scholar] [CrossRef]
  38. Liu, L.; Zhang, X.; Yang, L.; Ren, L.; Wang, D.; Ye, J. Metal nanoparticles induced photocatalysis. Natl. Sci. Rev. 2017, 4, 761–780. [Google Scholar] [CrossRef] [Green Version]
  39. Yao, G.Y.; Zhao, Z.Y.; Liu, Q.L.; Dong, X.D.; Zhao, Q.M. Theoretical calculations for localized surface plasmon resonance effects of Cu/TiO2 nanosphere: Generation, modulation, and application in photocatalysis. Sol. Energy Mater. Sol. Cells 2020, 208, 110385. [Google Scholar] [CrossRef]
  40. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S.C. Photocatalytic hydrogen production using metal doped TiO2: A review of recent advances. Appl. Catal. B Environ. 2019, 244, 1021–1064. [Google Scholar] [CrossRef]
  41. Xiu, Z.; Guo, M.; Zhao, T.; Pan, K.; Xing, Z.; Li, Z.; Zhou, W. Recent advances in Ti3+ self-doped nanostructured TiO2 visible light photocatalysts for environmental and energy applications. Chem. Eng. J. 2020, 382, 123011. [Google Scholar] [CrossRef]
  42. Pulido Melián, E.; Nereida Suárez, M.; Jardiel, T.; Calatayud, D.G.; Del Campo, A.; Doña-Rodríguez, J.M.; Araña, J.; González Díaz, O.M. Highly photoactive TiO2 microspheres for photocatalytic production of hydrogen. Int. J. Hydrogen Energy 2019, 44, 24653–24666. [Google Scholar] [CrossRef]
  43. Huang, G.; Liu, X.; Shi, S.; Li, S.; Xiao, Z.; Zhen, W.; Liu, S.; Wong, P.K. Hydrogen producing water treatment through mesoporous TiO2 nanofibers with oriented nanocrystals. Chin. J. Catal. 2020, 41, 50–61. [Google Scholar] [CrossRef]
  44. Chen, W.; Wang, Y.; Liu, S.; Gao, L.; Mao, L.; Fan, Z.; Shangguan, W.; Jiang, Z. Non-noble metal Cu as a cocatalyst on TiO2 nanorod for highly efficient photocatalytic hydrogen production. Appl. Surf. Sci. 2018, 445, 527–534. [Google Scholar] [CrossRef] [Green Version]
  45. Gao, H.; Zhang, P.; Zhao, J.; Zhang, Y.; Hu, J.; Shao, G. Plasmon enhancement on photocatalytic hydrogen production over the Z-scheme photosynthetic heterojunction system. Appl. Catal. B Environ. 2017, 210, 297–305. [Google Scholar] [CrossRef]
  46. El-Bery, H.M.; Matsushita, Y.; Abdel-moneim, A. Fabrication of efficient TiO2-RGO heterojunction composites for hydrogen generation via water-splitting: Comparison between RGO, Au and Pt reduction sites. Appl. Surf. Sci. 2017, 423, 185–196. [Google Scholar] [CrossRef]
  47. Wu, L.; Shi, S.; Li, Q.; Zhang, X.; Cui, X. TiO2 nanoparticles modified with 2D MoSe2 for enhanced photocatalytic activity on hydrogen evolution. Int. J. Hydrogen Energy 2019, 44, 720–728. [Google Scholar] [CrossRef]
  48. Xing, X.; Zhu, H.; Zhang, M.; Xiao, L.; Li, Q.; Yang, J. Effect of heterojunctions and phase-junctions on visible-light photocatalytic hydrogen evolution in BCN-TiO2 photocatalysts. Chem. Phys. Lett. 2019, 727, 11–18. [Google Scholar] [CrossRef]
  49. Li, F.; Xiao, X.; Zhao, C.; Liu, J.; Li, Q.; Guo, C.; Tian, C.; Zhang, L.; Hu, J.; Jiang, B. TiO2-on-C3N4 double-shell microtubes: In-situ fabricated heterostructures toward enhanced photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2020, 572, 22–30. [Google Scholar] [CrossRef] [PubMed]
  50. Zhang, C.; Zhou, Y.; Bao, J.; Fang, J.; Zhao, S.; Zhang, Y.; Sheng, X.; Chen, W. Structure regulation of ZnS@g-C3N4/TiO2 nanospheres for efficient photocatalytic H2 production under visible-light irradiation. Chem. Eng. J. 2018, 346, 226–237. [Google Scholar] [CrossRef]
  51. Melián, E.P.; López, C.R.; Méndez, A.O.; Díaz, O.G.; Suárez, M.N.; Doña Rodríguez, J.M.; Navío, J.A.; Fernández Hevia, D. Hydrogen production using Pt-loaded TiO2 photocatalysts. Int. J. Hydrogen Energy 2013, 38, 11737–11748. [Google Scholar] [CrossRef]
  52. Vaiano, V.; Lara, M.A.; Iervolino, G.; Matarangolo, M.; Navio, J.A.; Hidalgo, M.C. Photocatalytic H2 production from glycerol aqueous solutions over fluorinated Pt-TiO2 with high {001} facet exposure. J. Photochem. Photobiol. A Chem. 2018, 365, 52–59. [Google Scholar] [CrossRef]
  53. Yasuda, M.; Matsumoto, T.; Yamashita, T. Sacrificial hydrogen production over TiO2-based photocatalysts: Polyols, carboxylic acids, and saccharides. Renew. Sustain. Energy Rev. 2018, 81, 1627–1635. [Google Scholar] [CrossRef]
  54. López, C.R.; Melián, E.P.; Ortega Méndez, J.A.; Santiago, D.E.; Doña Rodríguez, J.M.; González Díaz, O. Comparative study of alcohols as sacrificial agents in H2 production by heterogeneous photocatalysis using Pt/TiO2 catalysts. J. Photochem. Photobiol. A Chem. 2015, 312, 45–54. [Google Scholar] [CrossRef]
  55. Velázquez, J.J.; Fernández-González, R.; Díaz, L.; Pulido Melián, E.; Rodríguez, V.D.; Núñez, P. Effect of reaction temperature and sacrificial agent on the photocatalytic H2-production of Pt-TiO2. J. Alloys Compd. 2017, 721, 405–410. [Google Scholar] [CrossRef]
  56. Kampouri, S.; Stylianou, K.C. Dual-Functional Photocatalysis for Simultaneous Hydrogen Production and Oxidation of Organic Substances. ACS Catal. 2019, 9, 4247–4270. [Google Scholar] [CrossRef]
  57. Zhang, S.; Wang, L.; Liu, C.; Luo, J.; Crittenden, J.; Liu, X.; Cai, T.; Yuan, J.; Pei, Y.; Liu, Y. Photocatalytic wastewater purification with simultaneous hydrogen production using MoS2 QD-decorated hierarchical assembly of ZnIn2S4 on reduced graphene oxide photocatalyst. Water Res. 2017, 121, 11–19. [Google Scholar] [CrossRef]
  58. Bellardita, M.; García-López, E.I.; Marcì, G.; Nasillo, G.; Palmisano, L. Photocatalytic Solar Light H2 Production by Aqueous Glucose Reforming: Photocatalytic Solar Light H2 Production by Aqueous Glucose Reforming. Eur. J. Inorg. Chem. 2018, 2018, 4522–4532. [Google Scholar] [CrossRef] [Green Version]
  59. Yao, Y.; Gao, X.; Li, Z.; Meng, X. Photocatalytic Reforming for Hydrogen Evolution: A Review. Catalysts 2020, 10, 335. [Google Scholar] [CrossRef] [Green Version]
  60. Park, Y.K.; Kim, B.J.; Jeong, S.; Jeon, K.J.; Chung, K.H.; Jung, S.C. Characteristics of hydrogen production by photocatalytic water splitting using liquid phase plasma over Ag-doped TiO2 photocatalysts. Environ. Res. 2020, 188, 109630. [Google Scholar] [CrossRef]
  61. Wu, G.; Chen, T.; Su, W.; Zhou, G.; Zong, X.; Lei, Z.; Li, C. H2 production with ultra-low CO selectivity via photocatalytic reforming of methanol on Au/TiO2 catalyst. Int. J. Hydrogen Energy 2008, 33, 1243–1251. [Google Scholar] [CrossRef]
  62. Udayabhanu; Reddy, N.L.; Shankar, M.V.; Sharma, S.C.; Nagaraju, G. One-pot synthesis of Cu-TiO2/CuO nanocomposite: Application to photocatalysis for enhanced H2 production, dye degradation & detoxification of Cr (VI). Int. J. Hydrogen Energy 2020, 45, 7813–7828. [Google Scholar]
  63. Xing, X.; Tang, S.; Hong, H.; Jin, H. Concentrated solar photocatalysis for hydrogen generation from water by titania-containing gold nanoparticles. Int. J. Hydrogen Energy 2020, 45, 9612–9623. [Google Scholar] [CrossRef]
  64. Su, E.C.; Huang, B.S.; Lee, J.T.; Wey, M.Y. Excellent dispersion and charge separation of SrTiO3-TiO2 nanotube derived from a two-step hydrothermal process for facilitating hydrogen evolution under sunlight irradiation. Sol. Energy 2018, 159, 751–759. [Google Scholar] [CrossRef]
  65. Sultanov, F.; Daulbayev, C.; Bakbolat, B.; Daulbayev, O.; Bigaj, M.; Mansurov, Z.; Kuterbekov, K.; Bekmyrza, K. Aligned composite SrTiO3/PAN fibers as 1D photocatalyst obtained by electrospinning method. Chem. Phys. Lett. 2019, 737, 136821. [Google Scholar] [CrossRef]
  66. Sultanov, F.; Bakbolat, B.; Mansurov, Z.; Pei, S.-S.; Ebrahim, R.; Daulbayev, C.; Urazgaliyeva, A.; Tulepov, M. Spongy Structures Coated with Carbon Nanomaterialsfor Efficient Oil/Water Separation. Eurasian Chem. Technol. J. 2017, 19, 127. [Google Scholar] [CrossRef] [Green Version]
  67. Sultanov, F.; Bakbolat, B.; Daulbaev, C.; Urazgalieva, A.; Azizov, Z.; Mansurov, Z.; Tulepov, M.; Pei, S.S. Sorptive Activity and Hydrophobic Behavior of Aerogels Based on Reduced Graphene Oxide and Carbon Nanotubes. J. Eng. Phys. Thermophys. 2017, 90, 826–830. [Google Scholar] [CrossRef]
  68. Luo, L.; Li, J.; Dai, J.; Xia, L.; Barrow, C.J.; Wang, H.; Jegatheesan, J.; Yang, M. Bisphenol A removal on TiO2–MoS2–reduced graphene oxide composite by adsorption and photocatalysis. Process. Saf. Environ. Prot. 2017, 112, 274–279. [Google Scholar] [CrossRef]
  69. Nguyen, C.H.; Tran, M.L.; Tran, T.T.V.; Juang, R.S. Enhanced removal of various dyes from aqueous solutions by UV and simulated solar photocatalysis over TiO2/ZnO/rGO composites. Sep. Purif. Technol. 2020, 232, 115962. [Google Scholar] [CrossRef]
  70. Kumar, K.Y.; Saini, H.; Pandiarajan, D.; Prashanth, M.K.; Parashuram, L.; Raghu, M.S. Controllable synthesis of TiO2 chemically bonded graphene for photocatalytic hydrogen evolution and dye degradation. Catal. Today 2020, 340, 170–177. [Google Scholar] [CrossRef]
  71. Rajender, G.; Kumar, J.; Giri, P.K. Interfacial charge transfer in oxygen deficient TiO2-graphene quantum dot hybrid and its influence on the enhanced visible light photocatalysis. Appl. Catal. B Environ. 2018, 224, 960–972. [Google Scholar] [CrossRef]
  72. Ali, M.H.H.; Al-Afify, A.D.; Goher, M.E. Preparation and characterization of graphene—TiO2 nanocomposite for enhanced photodegradation of Rhodamine-B dye. Egypt. J. Aquat. Res. 2018, 44, 263–270. [Google Scholar] [CrossRef]
  73. Ton, N.N.T.; Dao, A.T.N.; Kato, K.; Ikenaga, T.; Trinh, D.X.; Taniike, T. One-pot synthesis of TiO2/graphene nanocomposites for excellent visible light photocatalysis based on chemical exfoliation method. Carbon 2018, 133, 109–117. [Google Scholar] [CrossRef]
  74. Lettieri, S.; Gargiulo, V.; Pallotti, D.K.; Vitiello, G.; Maddalena, P.; Alfè, M.; Marotta, R. Evidencing opposite charge-transfer processes at TiO2/graphene-related materials interface through a combined EPR, photoluminescence and photocatalysis assessment. Catal. Today 2018, 315, 19–30. [Google Scholar] [CrossRef]
  75. Shahbazi, R.; Payan, A.; Fattahi, M. Preparation, evaluations and operating conditions optimization of nano TiO2 over graphene based materials as the photocatalyst for degradation of phenol. J. Photochem. Photobiol. A Chem. 2018, 364, 564–576. [Google Scholar] [CrossRef]
  76. Khan, H.; Jiang, Z.; Berk, D. Molybdenum doped graphene/TiO2 hybrid photocatalyst for UV/visible photocatalytic applications. Sol. Energy 2018, 162, 420–430. [Google Scholar] [CrossRef]
  77. Sun, X.; Ji, S.; Wang, M.; Dou, J.; Yang, Z.; Qiu, H.; Kou, S.; Ji, Y.; Wang, H. Fabrication of porous TiO2-RGO hybrid aerogel for high-efficiency, visible-light photodegradation of dyes. J. Alloys Compd. 2020, 819, 153033. [Google Scholar] [CrossRef]
  78. Sultanov, F.R.; Mansurov, Z.A.; Daulbayev, C.; Urazgaliyeva, A.A.; Bakbolat, B.; Pei, S.-S. Study of Sorption Capacity and Surface Morphology of Carbon Nanomaterials/Chitosan Based Aerogels. Eur. Chem. Technol. J. 2016, 18, 19. [Google Scholar] [CrossRef] [Green Version]
  79. Sultanov, F.R.; Pei, S.S.S.; Auyelkhankyzy, M.; Smagulova, G.; Lesbayev, B.T.; Mansurov, Z.A. Aerogels Based on Graphene Oxide with Addition of Carbon Nanotubes: Synthesis and Properties. Eur. Chem. Technol. J. 2014, 16, 265–269. [Google Scholar] [CrossRef]
  80. Ali, M.H.H.; Al-Qahtani, K.M.; El-Sayed, S.M. Enhancing photodegradation of 2,4,6 trichlorophenol and organic pollutants in industrial effluents using nanocomposite of TiO2 doped with reduced graphene oxide. Egypt. J. Aquat. Res. 2019, 45, 321–328. [Google Scholar] [CrossRef]
  81. Nadimi, M.; Ziarati Saravani, A.; Aroon, M.A.; Ebrahimian Pirbazari, A. Photodegradation of methylene blue by a ternary magnetic TiO2/Fe3O4/graphene oxide nanocomposite under visible light. Mater. Chem. Phys. 2019, 225, 464–474. [Google Scholar] [CrossRef]
  82. Park, K.; Ali, I.; Kim, J.O. Photodegradation of perfluorooctanoic acid by graphene oxide-deposited TiO2 nanotube arrays in aqueous phase. J. Environ. Manag. 2018, 218, 333–339. [Google Scholar] [CrossRef] [PubMed]
  83. Al Kausor, M.; Chakrabortty, D. Facile fabrication of N-TiO2/Ag3PO4@GO nanocomposite toward photodegradation of organic dye under visible light. Inorg. Chem. Commun. 2020, 116, 107907. [Google Scholar] [CrossRef]
  84. Zhou, Q.; Wang, M.; Tong, Y.; Wang, H.; Zhou, X.; Sheng, X.; Sun, Y.; Chen, C. Improved photoelectrocatalytic degradation of tetrabromobisphenol A with silver and reduced graphene oxide-modified TiO2 nanotube arrays under simulated sunlight. Ecotoxicol. Environ. Saf. 2019, 182, 109472. [Google Scholar] [CrossRef] [PubMed]
  85. Zhao, Y.; Wang, G.; Li, L.; Dong, X.; Zhang, X. Enhanced activation of peroxymonosulfate by nitrogen-doped graphene/TiO2 under photo-assistance for organic pollutants degradation: Insight into N doping mechanism. Chemosphere 2020, 244, 125526. [Google Scholar] [CrossRef]
  86. Chen, F.; An, W.; Li, Y.; Liang, Y.; Cui, W. Fabricating 3D porous PANI/TiO2-graphene hydrogel for the enhanced UV-light photocatalytic degradation of BPA. Appl. Surf. Sci. 2018, 427, 123–132. [Google Scholar] [CrossRef]
  87. Ghasemi, A.; Azzouz, R.; Laipple, G.; Kabak, K.E.; Heavey, C. Optimizing capacity allocation in semiconductor manufacturing photolithography area—Case study: Robert Bosch. J. Manuf. Syst. 2020, 54, 123–137. [Google Scholar] [CrossRef]
  88. Daulbaev, C.B.; Dmitriev, T.P.; Sultanov, F.R.; Mansurov, Z.A.; Aliev, E.T. Obtaining Three-Dimensional Nanosize Objects on a “3D Printer + Electrospinning” Machine. J. Eng. Phys. Thermophys. 2017, 90, 1115–1118. [Google Scholar] [CrossRef]
  89. Liu, J.; Liu, H.; Zuo, X.; Wen, F.; Jiang, H.; Cao, H.; Pei, Y. Micro-patterned TiO2 films for photocatalysis. Mater. Lett. 2019, 254, 448–451. [Google Scholar] [CrossRef]
  90. Wang, Q.; Jin, R.; Zhang, M.; Gao, S. Solvothermal preparation of Fe-doped TiO2 nanotube arrays for enhancement in visible light induced photoelectrochemical performance. J. Alloys Compd. 2017, 690, 139–144. [Google Scholar] [CrossRef]
  91. Zhang, J.J.; Qi, P.; Li, J.; Zheng, X.C.; Liu, P.; Guan, X.X.; Zheng, G.P. Three-dimensional Fe2O3-TiO2-graphene aerogel nanocomposites with enhanced adsorption and visible light-driven photocatalytic performance in the removal of RhB dyes. J. Ind. Eng. Chem. 2018, 61, 407–415. [Google Scholar] [CrossRef]
  92. Lesbayev, A.B.; Smagulova, G.T.; Kim, S.; Prikhod’ko, N.G.; Manakov, S.M.; Guseinov, N.; Mansurov, Z.A. Solution-Combustion Synthesis and Characterization of Fe3O4 Nanoparticles. Int. J. Self Propag. High Temp. Synth. 2018, 27, 195–197. [Google Scholar] [CrossRef]
  93. Soltani-nezhad, F.; Saljooqi, A.; Shamspur, T.; Mostafavi, A. Photocatalytic degradation of imidacloprid using GO/Fe3O4/TiO2-NiO under visible radiation: Optimization by response level method. Polyhedron 2019, 165, 188–196. [Google Scholar] [CrossRef]
  94. Rajoriya, S.; Bargole, S.; George, S.; Saharan, V.K.; Gogate, P.R.; Pandit, A.B. Synthesis and characterization of samarium and nitrogen doped TiO2 photocatalysts for photo-degradation of 4-acetamidophenol in combination with hydrodynamic and acoustic cavitation. Sep. Purif. Technol. 2019, 209, 254–269. [Google Scholar] [CrossRef]
  95. Isari, A.A.; Hayati, F.; Kakavandi, B.; Rostami, M.; Motevassel, M.; Dehghanifard, E. N, Cu co-doped TiO2@functionalized SWCNT photocatalyst coupled with ultrasound and visible-light: An effective sono-photocatalysis process for pharmaceutical wastewaters treatment. Chem. Eng. J. 2020, 392, 123685. [Google Scholar] [CrossRef]
  96. Wu, D.; Wang, X.; Wang, H.; Wang, F.; Wang, D.; Gao, Z.; Wang, X.; Xu, F.; Jiang, K. Ultrasonic-assisted synthesis of two dimensional BiOCl/MoS2 with tunable band gap and fast charge separation for enhanced photocatalytic performance under visible light. J. Colloid Interface Sci. 2019, 533, 539–547. [Google Scholar] [CrossRef]
  97. Ding, J.; Liu, Q.; Zhang, Z.; Liu, X.; Zhao, J.; Cheng, S.; Zong, B.; Dai, W.L. Carbon nitride nanosheets decorated with WO3 nanorods: Ultrasonic-assisted facile synthesis and catalytic application in the green manufacture of dialdehydes. Appl. Catal. B Environ. 2015, 165, 511–518. [Google Scholar] [CrossRef]
Figure 1. The mechanism of photocatalytic water splitting on TiO2 based photocatalyst.
Figure 1. The mechanism of photocatalytic water splitting on TiO2 based photocatalyst.
Nanomaterials 10 01790 g001
Figure 2. Light absorption by micro- and nanoparticles of TiO2.
Figure 2. Light absorption by micro- and nanoparticles of TiO2.
Nanomaterials 10 01790 g002
Figure 3. Changes in the band gap of TiO2 upon doping with nitrogen atoms.
Figure 3. Changes in the band gap of TiO2 upon doping with nitrogen atoms.
Nanomaterials 10 01790 g003
Figure 4. (a) Rate of hydrogen evolution from photocatalysis of aqueous methanol solution on Ag/TiO2 photocatalysts. This figure is reprinted from [60], with permission from Elsevier, 2020; (b) H2 production patterns for 24.47 M (100% v/v) of methanol and 17.06 M (100% v/v) of ethanol. This figure is reprinted from [54], with permission from Elsevier, 2015.
Figure 4. (a) Rate of hydrogen evolution from photocatalysis of aqueous methanol solution on Ag/TiO2 photocatalysts. This figure is reprinted from [60], with permission from Elsevier, 2020; (b) H2 production patterns for 24.47 M (100% v/v) of methanol and 17.06 M (100% v/v) of ethanol. This figure is reprinted from [54], with permission from Elsevier, 2015.
Nanomaterials 10 01790 g004
Figure 5. Colour of synthesized pristine and Cu-TiO2/CuO composite nanopowders. This figure is reprinted from [62], with permission from Elsevier, 2020.
Figure 5. Colour of synthesized pristine and Cu-TiO2/CuO composite nanopowders. This figure is reprinted from [62], with permission from Elsevier, 2020.
Nanomaterials 10 01790 g005
Figure 6. (a) Hydrogen evolution of pure TiO2 nanoparticles and Au/TiO2 nanoparticles at the same light intensity of 5 kW/m2; (b) Hydrogen evolution in 1–9 kW/m2 light intensities of 1 g/L Au/TiO2 solutions. Both figures are reprinted from [63], with permission from Elsevier, 2020.
Figure 6. (a) Hydrogen evolution of pure TiO2 nanoparticles and Au/TiO2 nanoparticles at the same light intensity of 5 kW/m2; (b) Hydrogen evolution in 1–9 kW/m2 light intensities of 1 g/L Au/TiO2 solutions. Both figures are reprinted from [63], with permission from Elsevier, 2020.
Nanomaterials 10 01790 g006
Figure 7. (a) Photograph of 80 mg TiO2-reduced graphene oxide (rGO) powder (left) and aerogel (right); (b) methylene blue (MB) dark adsorption results of the TiO2-rGO aerogel/powder; color changing of MB solution during dark adsorption with TiO2-rGO powder (c), and aerogel (d). All figures are reprinted from [77], with permission from Elsevier, 2020.
Figure 7. (a) Photograph of 80 mg TiO2-reduced graphene oxide (rGO) powder (left) and aerogel (right); (b) methylene blue (MB) dark adsorption results of the TiO2-rGO aerogel/powder; color changing of MB solution during dark adsorption with TiO2-rGO powder (c), and aerogel (d). All figures are reprinted from [77], with permission from Elsevier, 2020.
Nanomaterials 10 01790 g007
Figure 8. SEM images of (a) planar TiO2 film; (b) grating-structured TiO2 film; (c) square-structured TiO2 film; (d) hexagon-structured TiO2 film. All figures are reprinted from [89], with permission from Elsevier, 2019.
Figure 8. SEM images of (a) planar TiO2 film; (b) grating-structured TiO2 film; (c) square-structured TiO2 film; (d) hexagon-structured TiO2 film. All figures are reprinted from [89], with permission from Elsevier, 2019.
Nanomaterials 10 01790 g008
Figure 9. Photocatalytic degradation of MO under ultraviolet light (254 nm) irradiation. C and C0 represent the real-time and initial concentration of methyl orange solution. This figure is reprinted from [89], with permission from Elsevier, 2019.
Figure 9. Photocatalytic degradation of MO under ultraviolet light (254 nm) irradiation. C and C0 represent the real-time and initial concentration of methyl orange solution. This figure is reprinted from [89], with permission from Elsevier, 2019.
Nanomaterials 10 01790 g009
Figure 10. (a) The rate constants for the adsorption and the collective removal of rhodamine B (RhB) over TiO2-GA and the Fe2O3-TiO2-GA composites; (b) stability of the Fe2O3-TiO2-GA (25%) composites in the removal of RhB dye. Both figures are reprinted from [91], with permission from Elsevier, 2018.
Figure 10. (a) The rate constants for the adsorption and the collective removal of rhodamine B (RhB) over TiO2-GA and the Fe2O3-TiO2-GA composites; (b) stability of the Fe2O3-TiO2-GA (25%) composites in the removal of RhB dye. Both figures are reprinted from [91], with permission from Elsevier, 2018.
Nanomaterials 10 01790 g010
Table 1. TiO2-based photocatalysts with different structures utilized for hydrogen evolution under splitting water mixtures.
Table 1. TiO2-based photocatalysts with different structures utilized for hydrogen evolution under splitting water mixtures.
PhotocatalystStructureLight SourceSacrificial AgentsEvolution H2Ref.
Au/TiO2
Pt/TiO2
microspheres15 W fluorescent tubes
max = 365 nm)
25% vol. methanol1118 µmol h−1
2125 µmol h−1
[42]
TiO2nanofibers300 W Xenon lamp10% vol. methanol3200 μmol h−1 g−1[43]
Cu/TiO2nanorods300 W Xe lamp
(λ > 300 nm)
20% vol. methanol 1023.8 μmol h−1[44]
TiO2/WO3/Aunanofibers300 W Xe arc lamp35% vol. methanol269.63 µmol h−1[45]
M/TiO2/rGO
M = Au or Pt
nanoparticles300 W Xenon lamp
(λ > 300 nm)
20% vol. methanol670 µmol h−1[46]
MoSe2/TiO2nanoparticlesXe arc lamp
(PLS-SXE300)
10% vol. methanol 4.9 μmol h−1[47]
BCN-TiO2nanosheets&
nanoparticles
300 W xenon lamp with a UV-cutoff filter
(λ ≥ 420 nm)
20% vol. triethanolamine68.54 μmol h−1 g−1[48]
TiO2/C3N4double-shell microtubes300 W xenon lamp20% vol. methanol10.1 mmol h−1 g−1[49]
ZnS@g-C3N4/TiO2nanospheres300 W Xenon lamp
(λ > 400 nm)
10% vol. triethanolamine422 μmol h−1 g−1[50]
Table 2. Removal percentage of some organic pollutants by TiO2/graphene based photocatalysts.
Table 2. Removal percentage of some organic pollutants by TiO2/graphene based photocatalysts.
PhotocatalystOrganic PollutantLight SourceIrradiation TimeEfficiencyRef.
TiO2@rGO2,4,6 trichlorophenolMercury lamp
(11 W)
180 min90%[80]
TiO2/Fe3O4/GOMethylene blueHalogen lamp
(500 W)
90 min76%[81]
GO/TiO2
nanotubes
Perfluorooctanoic acidUV lamp
(8 W)
240 min97%[82]
N-TiO2/Ag3PO4@GOAcid Blue 25Halogen bulb (250 W) 20 min98%[83]
Ag and rGO modified TiO2Tetrabromobisphenol AXenon light (500 W)80 min99.6%[84]
N-doped graphene/TiO2Bisphenol AMercury lamp (300 W) 60 min100%[85]
3D polyaniline/TiO2/rGO hydrogelBPAMercury lamp (500 W)40 min100%[86]

Share and Cite

MDPI and ACS Style

Bakbolat, B.; Daulbayev, C.; Sultanov, F.; Beissenov, R.; Umirzakov, A.; Mereke, A.; Bekbaev, A.; Chuprakov, I. Recent Developments of TiO2-Based Photocatalysis in the Hydrogen Evolution and Photodegradation: A Review. Nanomaterials 2020, 10, 1790. https://doi.org/10.3390/nano10091790

AMA Style

Bakbolat B, Daulbayev C, Sultanov F, Beissenov R, Umirzakov A, Mereke A, Bekbaev A, Chuprakov I. Recent Developments of TiO2-Based Photocatalysis in the Hydrogen Evolution and Photodegradation: A Review. Nanomaterials. 2020; 10(9):1790. https://doi.org/10.3390/nano10091790

Chicago/Turabian Style

Bakbolat, Baglan, Chingis Daulbayev, Fail Sultanov, Renat Beissenov, Arman Umirzakov, Almaz Mereke, Askhat Bekbaev, and Igor Chuprakov. 2020. "Recent Developments of TiO2-Based Photocatalysis in the Hydrogen Evolution and Photodegradation: A Review" Nanomaterials 10, no. 9: 1790. https://doi.org/10.3390/nano10091790

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop