Next Article in Journal
Biocompatible PEO-b-PCL Nanosized Micelles as Drug Carriers: Structure and Drug–Polymer Interactions
Next Article in Special Issue
Photoactive Nanomaterials
Previous Article in Journal
Protective Spinel Coating for Li1.17Ni0.17Mn0.50Co0.17O2 Cathode for Li-Ion Batteries through Single-Source Precursor Approach
Previous Article in Special Issue
Protonation-Induced Enhanced Optical-Light Photochromic Properties of an Inorganic-Organic Phosphomolybdic Acid/Polyaniline Hybrid Thin Film
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photoactive Tungsten-Oxide Nanomaterials for Water-Splitting

by
Yerkin Shabdan
1,2,†,
Aiymkul Markhabayeva
2,†,
Nurlan Bakranov
3,4,* and
Nurxat Nuraje
5,*
1
National Laboratory Astana, Nazarbayev University, Nursultan 010000, Kazakhstan
2
Faculty of Physics and Technology, AI-Farabi Kazakh National University, Almaty 050040, Kazakhstan
3
Faculty of General Education, Kazakh-British Technical University, Almaty 050000, Kazakhstan
4
Laboratory of Engineering Profile, Satbayev University, Almaty 050000, Kazakhstan
5
Department of Chemical and Materials Engineering, Nazarbayev University, Nursultan 010000, Kazakhstan
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2020, 10(9), 1871; https://doi.org/10.3390/nano10091871
Submission received: 16 July 2020 / Revised: 25 August 2020 / Accepted: 29 August 2020 / Published: 18 September 2020
(This article belongs to the Special Issue Photoactive Nanomaterials)

Abstract

:
This review focuses on tungsten oxide (WO3) and its nanocomposites as photoactive nanomaterials for photoelectrochemical cell (PEC) applications since it possesses exceptional properties such as photostability, high electron mobility (~12 cm2 V−1 s−1) and a long hole-diffusion length (~150 nm). Although WO3 has demonstrated oxygen-evolution capability in PEC, further increase of its PEC efficiency is limited by high recombination rate of photogenerated electron/hole carriers and slow charge transfer at the liquid–solid interface. To further increase the PEC efficiency of the WO3 photocatalyst, designing WO3 nanocomposites via surface–interface engineering and doping would be a great strategy to enhance the PEC performance via improving charge separation. This review starts with the basic principle of water-splitting and physical chemistry properties of WO3, that extends to various strategies to produce binary/ternary nanocomposites for PEC, particulate photocatalysts, Z-schemes and tandem-cell applications. The effect of PEC crystalline structure and nanomorphologies on efficiency are included. For both binary and ternary WO3 nanocomposite systems, the PEC performance under different conditions—including synthesis approaches, various electrolytes, morphologies and applied bias—are summarized. At the end of the review, a conclusion and outlook section concluded the WO3 photocatalyst-based system with an overview of WO3 and their nanocomposites for photocatalytic applications and provided the readers with potential research directions.

Graphical Abstract

1. Introduction

The conversion of solar-emitted electromagnetic waves to useful forms of energy is a very promising research area in the field of renewable energy production. Although roughly 32 × 1024 J of solar energy reaches the Earth’s surface per year, only 0.001% of the incoming solar energy is used for human needs [1]. The conversion of solar light to useful forms of energy is still challenging at the scientific and engineering level in terms of energy production for the needs of human beings. Even though there are many technologies for renewable energy [2], including solar cells, solar collectors and solar fuel reactors (water-splitting), the major challenges we face are to improve efficiency and stability in the conversion of solar energy to other energy forms. Currently, one of the popular research technologies to tackle solar energy conversion is trying to convert photons into chemical energy [3] by using artificial photoelectrochemical (PEC) processes.
Metal-oxide nanomaterials have been thoroughly studied for the conversion of solar energy to hydrogen molecules due to their chemical and physical stability, optical and electronic properties, easy fabrication and low cost. They have shown good properties for use in photoelectrochemical devices such as TiO2 [4,5,6], α-Fe2O3 [7,8,9], BiVO4 [10,11,12], ZnO [13,14,15] and WO3 [16,17,18]. On the above properties of semiconductor materials, both suitable bandgap positions to generate hydrogen and oxygen gases and the ability to absorb a reasonable portion of the solar light spectrum are critical for water-splitting. However, a single metal-oxide photocatalyst cannot simultaneously satisfy all the requirements for solar-to-hydrogen-driven systems since it encounters many problems (including fast recombination of charge carriers, photo corrosion, instability in aggressive electrolytes, short lifetime of charge carriers, improper bandgap or diffusion length of photogenerated electrons and holes). As a result, most of the metal oxide semiconductors are not suitable to split water at the visible light irradiation, which occupies 54% of whole solar spectrum since they either do not have proper bandgaps or only absorb the UV light region. However, the problems stated above have been successfully addressed by introducing heterojunctions, composite nanomaterials, coupling wide band and narrow band materials, doping, surface–interface engineering, dye sensitization, etc.
Among the metal oxides, WO3 is a promising semiconductor for PEC water-splitting with favorable properties. (These properties include: suitable bandgap (~2.6 eV), good chemical stability under strong solar exposure, oxygen-evolution capability, long minority carrier diffusion length (~500 nm–6 μm [19,20]), absorption of visible light (~12%) and low cost.) The conduction band energy position of WO3 is 0.25 eV, which is not suitable for reorientation of bonds of hydrogen atoms from the aqueous phase to the gaseous (0 V vs. NHE). On the other hand, the valence band, located at 2.7 eV, is more positive than the oxidation potential of oxygen (1.23 V vs. NHE) and is suitable for oxygen evolution. Although the WO3 photocatalyst suffers from some limitations such as sluggish charge transfer [21], boosting charge separation can be achieved by modifying WO3 photoanode with numerous materials including Ag nanoparticles [22] and Au plasmonic particles [23]. Many papers have reported on using WO3 photoanodes for O2 evolution study [24,25]. During the study of hydrogen evolution from aqueous phase, various photoelectrochemical systems and configurations integrated with WO3 and its composites have been developed.
Among the published materials in this prospect, numerous amounts of work can be distinguished: Ji et al. reported a triple layer heterojunction BiVO4/WO3/SnO2 material with a perovskite solar cell [26], Liu. et al. prepared a WO3 photoanode with a tandem cell [27] and Lee used dye-sensitized solar cells to produce hydrogen with bare WO3 photoanodes [28]. Zhang fabricated the WO3@a–Fe2O3/FeOOH photoanode, which exhibits a 120 mV negative shift in onset potential and yields a photocurrent density of 1.12 mA/cm2 at 1.23 V vs. reversible hydrogen electrode (RHE) [29]. Moreover, some systems use free-particle WO3 heterostructures based on photochemical cell reactions. Despite the fact that WO3 cannot generate hydrogen, there are some publications where scientists show high photocatalytic activity for CdS-WO3 [30] and non-stoichiometric WO3−x/CdS heterostructures for efficient hydrogen generation [31].
Thus, we have conducted a literature survey on the WO3-based photocatalytic system and found a dramatic increase of publications recently (Figure 1). This indicates that the WO3 is a very important material for designing efficient photocatalytic systems. The analysis of scientific articles, reviews and conference materials found in the authoritative database revealed few review papers in the use of tungsten trioxide photocatalyst for water-splitting. As shown in Figure 1, the trend of the published papers in the WO3 photocatalytic research is increasing exponentially. Therefore, in our opinion, it is essential to present a review article to our scientific community with recent research progress of WO3 in the photocatalytic water-splitting. Although there are some review papers that included the WO3 and their water-splitting applications, from the best of our knowledge, few papers have been specifically focused on sole WO3/nanocomposites and their recent photocatalytic application.
The primary focus of our review article is to deliver the recent progress of tungsten-based photocatalytic systems that have been developed. More specifically, it discusses the morphology, crystal, doping, surface–interface engineering effect of WO3 on the heterostructured photocatalytic system, and all of the results in different conditions including electrolyte, power, applied bias, morphology, and synthesis approaches were tabulated for the researchers to check. Therefore, in this review, we try to give comprehensive information on WO3 including the physical chemistry property, crystal structure, and nanomorphology along with their composites including binary and ternary structures used in the particulate, PEC, Z-scheme and tandem configuration for effective water-splitting applications.

2. Basic Principles of the Water-Splitting Reaction

2.1. Thermodynamics of Water-Splitting

In the reaction of water-splitting, solar energy can be directly converted into chemical energy form, hydrogen gas [32,33,34]. The hydrogen acts as a green energy carrier since it possesses high energy density. When used in a fuel cell, water is the only byproduct.
As early as 1923, J. B. S. Haldane, a British scientist, proposed a concept of photocatalytic hydrogen production. Seeing that there is no naturally produced pure hydrogen on Earth, its resource is highly abundant throughout the universe. Like fossil fuels, water or biomass can be utilized to produce hydrogen or other chemical fuels. Hydrogen gas can be further used in hydrogen fuel cell-powered vehicles, which are much more environmentally friendly than the commonly used nonrenewable fuel options. Increasing the efficiency of water-splitting devices for hydrogen fuel production has a potential to decrease its dependence on using fossil fuels and importation.
There are some other approaches for hydrogen production, however, the most environmentally sustainable, “green” method is photocatalytic or photoelectrochemical water-splitting. The PEC water-splitting works similarly to a solar cell. The main difference is that it converts solar energy to a chemical bond instead of converting directly to electric power, which is beneficial to store energy for later use. PEC consists of three main components: an anode, a cathode and an electrolyte (aqueous media). At the anode, water is oxidized to generate oxygen via the oxygen evolution reaction (OER), whereas at the cathode hydrogen ions are reduced into hydrogen gas via a hydrogen evolution reaction (HER). Based on the configuration of the PEC cell, either the cathode or anode, or both, can be photoactive semiconductors which absorb light. Furthermore, water can also split via connecting a p–n junction solar cell in parallel with a photoelectrochemical cell. This process not only avoids the complicated manufacturing process, but it also reduces the system’s cost [35]. Although extensive research has been conducted using many semiconductor configurations, there is still so much that needs to be done to reach the targeted efficiency and stability goals. For a particle-based photo catalytical system, the ideal solar to hydrogen (STH) efficiency should be 10% [36,37]. This efficiency brings cheaper H2 production.
In 1972, Japanese scientists, Fujishima and Honda first studied TiO2 as a photonic material and proved that water can be decomposed under UV-light exposure [38]. Since then, scientists have been studying a variety of light-sensitive material, including all inorganic and organic dyes [39,40,41].
Decomposing water into H2 and O2 is an endothermic reaction thermodynamically (+237.2 kJ/mol). This means that additional energy is required to perform the decomposition reaction (E-1):
ΔG0 = −nFΔE0 = +237.2 kJ/mol H2
where:
F—Faraday’s constant (F = 96,485 C/mol),
n—Number of transferred electrons (n = 2)
ΔE0—standard potential of the electrochemical cell (ΔE0 = 1.229 V).
The amount of Gibbs free energy required to split a molecule of water into hydrogen and oxygen is ΔG = 237.2 kJ/mol, which is corresponded to ΔE0 1.23 eV per electron, transforming the Nernst equation under standard conditions. This means a minimum energy of 1.23 eV per electron should be supplied by the photocatalyst. This process can be written in the following two half-reactions (E-2; E-3; E-4):
Water oxidation: H2O + 2h+ → ½ O2 + 2H+ (HER)
Water reduction: 2H+ + 2e → H2 (ORE)
H2O → ½ O2 + H2 ΔG = +237.2 kJ/mol
The bandgap (Eg) is the main parameter that defines the light-harvesting ability of an absorber. Photons alone with energies higher than the bandgap can excite electrons in the valence band to the conduction band. The excess energy or the difference in the energy of the absorbed photon and the band gap energy (𝐸–𝐸𝑔), is lost as phonons. The absorption coefficient of the semiconducting materials is another parameter which shows how efficiently a photocatalyst can harness the solar spectrum. One crucial point that needs to be taken into consideration as quantifying the optimal minimum band gap value is the intrinsic loss (Eloss), associated with the solar energy conversion process. These losses are connected with the fundamental loss caused by thermodynamics because of non-ideality (kinetic losses) in the conversion process [35,38]. The former loss results from the second law of thermodynamics. In fact, the following equation shows how bandgap energy (Eg) corresponds to the change in internal energy, which is related to the change in Gibbs energy (E-5):
ΔG = ΔU + PΔV − TΔS
where U, P, V, T and S indicate the internal energy, pressure, volume, temperature and entropy, respectively. When the semiconductor absorbs photons, increasing excited states can be created in addition to ground states, increasing the entropy of the ensemble. The change in entropy or ΔSmix, occurs because there are existing excited states along with the ground states. A volume change (ΔVmix) is also caused by the mixture of excited and ground states. However, this is not true for the ideal chemical system (ΔVmix = 0). Thus, the band gap energy should be greater than the available work under ideal conditions (Gibbs energy change per electron), at least Eloss = TΔSmix with a minimum of 0.3–0.5 eV. In reality, Eloss reaches higher values (roughly 0.8 eV) as a result of kinetic losses and due to non-ideality (overpotential at the anode and cathode, reduction in resistance at the electrolyte, electron–hole pair recombination). Therefore, in order to maximize the chemical conversion efficiency, materials commonly used as photoelectrodes in PEC cells require a band gap of 2.0 to 2.25 eV [35,38].
When UV and/or visible sunlight shine onto a semiconductor photocatalyst, the semiconductor absorbs photons and excites electrons from the semiconductor’s valence band to its conduction band, leaving a hole in the valence band, i.e., electron–hole pairs (Figure 2a). This is the so-called, “photo-excited” semiconductor phase. The bandgap is the difference between the maximum valence band energy and the minimum conduction band energy. Ideally, semiconductors have a bandwidth greater than 1.23 V as well as a more negative conduction band relative to the water reduction potential and a more positive valence band relative to the water oxidation potential.
A typical one-step PEC configuration for water decomposition consists of either a photoanode or a photocathode. N-type tungsten trioxide is mostly represented as a photoanode and the basic principles of such cell can be depicted in Figure 2b. Process of PEC water decomposition is initiated via accepting light photons by photoactive materials. Then, this step is accompanied by generating excitons (electron–hole pairs) inside semiconductors. Photogenerated holes on the surface of WO3 can oxidize water while electrons flow to the Pt electrode to produce hydrogen (Figure 2b). Due to the improper positioning of the conduction and valence bands with respect to the potentials of water reduction and oxidation, external bias voltage is used to separate excitons.
Another thermodynamic precondition is the position of the band edges. For the oxidation reaction to occur, holes move from the photoelectrode to the interface between the semiconductor and the solution freely. The top edge of the valence band must be more positive than the oxidation potential of O2/H2O as seen in Figure 2b. Likewise, the reduction reaction happens if the bottom edge of the conduction band is more negative than the reduction potential of H+/H2.
Figure 3 shows the band structure and bandgap values of some semiconductors [42]. While wide band gap (Eg > 3 eV) photocatalyst can harvest only UV light (a small portion of the solar spectrum, less than 4%), its band gap can be easily engineered to absorb the visible light range via metal and nonmetal doping. Furthermore, narrow bandgap materials (e.g., WO3, Fe2O3) are not able to drive the water reduction and oxidation reactions at the same time since their bandgap energy positions are not properly positioned to the water redox potentials. Therefore, they are commonly used to construct tandem cell structures for the water-splitting reaction.

2.2. Device Requirements and Calculation of Their Efficiency

Various types of overall water-splitting techniques include: particulate systems [43], Z-schemes [44,45,46] and photoelectrochemical cells [46]. The photoelectrodes of PEC water decomposition such as a photoanode and a cathode electrode are made up of photocatalysts. However, fabrication of stable photoelectrodes under the influence of strong sunlight is still a challenging one in PECs. Right now, it is preferred to fabricate the film via direct growth on the photoelectrode, which provides a relatively stable photoactive films.
In order to design an efficient and affordable solar hydrogen production PEC system, the electrode requires low cost materials with the capability of efficient light harnessing and long term stability.
To date, large band gap semiconductors (UV-active specifically) and metal oxides have been extensively investigated for the photocatalytic water-splitting studies due to their robustness and suitable band gap energy positions. One challenge to using these materials is the limitation of solar light harnessing to a small portion of the solar spectrum.
The following formula helps us calculate the theoretical maximum photocurrent Jmax, which is proportional to the solar–hydrogen conversion efficiency (STH):
Jmax = q ∫ Φλ [1 − exp(−αλd)] dλ
where λ, q, d and αλ represent wavelength, electron charge, sample thickness and absorption coefficient under the photon flux of the AM 1.5 G solar spectrum, respectively. Considering the conversion, reflection and other losses, obtaining the goal of 10% STH conversion efficiency is very challenging.
The following efficiencies are usually reported for PEC cells. They are STH conversion efficiency, applied bias photon-current efficiency (ABPE), external quantum efficiency and internal quantum efficiency (IQE) or absorbed photons to current efficiency (APCE).
STH efficiency is commonly used to evaluate PEC device performance and is expressed in the following way:
STH = [(H2 production rate) × (Gibbs free energy per H2)]/([incident energy])
The E-8 formula can be applied to calculate ABPE:
ABPE = [Jph × (1.23 − Vb)]/Ptotal
where Jph is the photocurrent density as a bias Vb is applied and Ptotal is the total incident solar light power.
External quantum efficiency defines the photocurrent generation per incident photon flux under a certain irradiation wavelength. Solar-to-hydrogen conversion efficiency can be evaluated via applying the external quantum efficiency data over the total solar spectrum in a two-electrode system. However, applying external quantum efficiency data obtained in a three-electrode system under a bias to estimate solar-to-hydrogen conversion efficiency is not considered to be a valid method. However, it is still considered to be a useful approach for finding PEC cell material properties. The external quantum efficiency (EQE) is expressed by equation (E-9):
EQE = (Jph × hc)/(Pmono × λ)
where Jph is the photocurrent density, h is Planck’s constant, c is the light speed, Pmono is the power of calibration and monochromatic illumination, and λ is the wavelength of monochromatic light.

3. WO3 and Its Nanocomposites for Particle-Based Photocatalytic Systems

3.1. Half Reaction Systems

Nowadays, the pursuit for highly efficient photocatalytic materials to produce hydrogen fuel under the exposure of light photons is still in the active stage. In a photoelectrochemical cell, it is required to create sufficient voltage between the anode and cathode to perform the water decomposition reaction. However, most of wide bandgap semiconductor materials are not able to respond to the visible part of the spectrum. Absorption of ultraviolet radiation alone is an undesirable property of photocatalyst operating in terrestrial conditions. One of the exciting ways to solve the above contradictions is the creation of photocatalytic systems consisting of a series of photocatalysts. That is why researchers try to use photochemical systems, where water can be decomposed using colloid particles without any external voltage. Many papers have reported [47] that hydrogen can be generated, even though the efficiency is very low. Evolution of oxygen is difficult because it requires process of four electrons and four H+ transfers.
Under solar illumination, although photoexcited electrons and holes are produced, they simultaneously experience recombination and back reaction, which are competitive processes of photogeneration. Hence, most works focus only on half reactions where either H2 or O2 evolution is possible in the presence of sacrificial electron donor or acceptor.
CdS/WO3 photocatalysts produced a high hydrogen evolution rate of 369 μmol/gh with lactic acid as an electron donor [30]. Further modification of CdS/WO3 with Pd particles increased the hydrogen evolution rate to 2900 μmol/gh, 7.9-fold higher than for CdS/WO3.
Furthermore, the surface plasmon resonance (SPR) effect of non-stochiometric WO3−x was demonstrated [31] from CdS/WO3−x heterostructures photocatalysts via photoinduced electron injection for hydrogen evolution. The non-elemental metal plasmonic material WO3−x has intense SPR in the visible/NIR region (Figure 4b). Free electrons in the conduction band of WO3−x can be generated from oxygen vacancies that are results of chemical reduction during synthesis. Further excitation of electrons can happen by SPR and then they can become hot electrons for the hydrogen generation as shown in Figure 4a. Photo-excited electrons on CdS inject into conduction band of WO3−x, so that the SPR of the photocatalyst WO3−x is stable and some hot electrons participate in hydrogen evolution reactions (Figure 4c).
In addition, to choose the photoanode material for the half-reaction of water-splitting, attention should be paid to the selection of the electrolyte. For sulfide semiconductors and composites, a Na2S/Na2SO3 mixture is used as an absorbing hole agent. In type II heterojunctions, for example, WO3–NS/CdS–NR, with high conductivity, WO3 provides efficient charge collection and, therefore, reduces the rate of space charge recombination, which leads to the accumulation of holes in cadmium sulfide. An electrolyte based on Na2S/Na2SO3 provides fast hole collection, which allows the half-reaction to occur without degradation of the photoanode [48]. The effect of some electrolyte solutions on the oxidative half-reactions of WO3-based photoanodes was studied on [49]. James C. Hill and Kyoung-Shin Choi studied photo-oxidative processes in chloride solutions, acetate solutions, phosphate solutions, perchlorate solutions, sulfate solutions and solutions with K+ and Li+ cations. The electrodeposited porous WO3 layers were used as a photoanode. The results show that the presence of acetate and chloride ions suppressed the release of O2. In a phosphate solution, the release of O2 and the formation of peroxides was the main result of photooxidation. The oxidation of water in perchlorate electrolytes was accompanied by the release of O2 and the formation of peroxides. In this case, the photocurrent density in such a system was lower in comparison with phosphate electrolytes. The authors also showed that cations have a significant effect on the efficiency of conversion of the photocurrent to O2. For example, Li+ ions adsorbed on the surface of WO3 serve as blockers of water oxidation centers, while K+ ions increase oxygen evolution in perchlorate, sulfate and phosphate solutions.
The effect of tungsten trioxide layers on hydrogen reduction processes also demonstrates positive dynamics. When combining Cu2O with WO3, a semiconductor p–n junction is created and that generates the conditions for the rupture of photogenerated excitons. Thus, in the Cu2O/WO3 heterostructure, an enhancement of the half-reaction of reduction is demonstrated in comparison with the sole Cu2O photocatalyst.

3.2. Z-Schemes

Z-scheme photocatalysts for overall water-splitting are a combined system involving two photon excitation processes (Figure 5) [50]. The two–photon excitation system was proposed by Bard et al. in 1979 which mimicked natural photosynthesis [51]. A Z-scheme is composed of one H2 evolution photocatalyst and another O2 evolution photocatalyst with electron mediator. Most of Z-scheme construction was demonstrated using Pt co-catalyst loaded SrTiO3, TaON, CaTa2O2 N and BaTa2O2N for hydrogen evolution and Pt/WO3 for oxygen evolution. Photocatalytic activity depends on pH level, concentration of electron mediator and type of co-catalysts. For example, Hideki Kato showed that pH level affects photocatalytic activity of Z-scheme consisting of Pt/SrTiO3Rh–WO3–FeCl3 system [52]. The pH value was adjusted using sulfuric and perchloric acids between 1.3 and 2.5. It was shown that the best photocatalytic activity was achieved at pH 2.4 and subsequent increasing of pH led to decrease of the activity. The sulfate ions-induced formation of [Fe(H2O)5(SO4)]+ species around pH 2.4. Under 48 h of solar illumination, the Z-scheme generated both 890 and 450 μmol H2 and O2, respectively.
Yugo Miseki reported a Z-scheme system with an oxygen evolution photocatalyst of PtOx/H–Cs–WO3 [53]. The Z-scheme water-splitting efficiency with PtOx/H–Cs–WO3 was 3-fold higher than that of using PtOx/WO3. Adding Cs+ ions to the PtOx/WO3 significantly improved the oxygen evolution rate. IO3 ion was used as an electron acceptor in this work. The apparent quantum yield at 420 nm was 20% which is the best index among photocatalysts using the IO3 redox.
Another Z-scheme, containing g-C3N4–WO3 photocatalysts, demonstrated enhanced H2 evolution [54]. The high photocatalytic activity is most likely due to direct electron transfer from WO3 to g-C3N4 in the Z-scheme.
One Z-scheme system consisting of graphitic nitrite g-C3N4 and WO3 nanocomposites modified with co-catalyst Ni(OH)x showed the highest hydrogen production rate of 576 μmol/(g.h). Photogenerated electrons and holes are efficiently separated by combination of g-C3N4 and Ni(OH)x. The electron spin resonance (ESR) technique used DMPO (5,5-dimethyl-1-pyrroline N-oxide) as a trapping agent of •O2− and •OH radicals to register the change of the active oxidizing species in aqueous systems. As a result, the Z-scheme charge separation mechanism explained the high hydrogen production rate [55].
Recently, construction of Z-scheme using ZnO–WO3−x nanorods was successfully synthesized by the solvothermal method [56]. A high photocurrent value of 3.38 mA/cm2 at 1.23 V vs. RHE, which is 3.02-fold higher than pure ZnO, was obtained by an effective Z-scheme charges-transfer process. Red shift of optical absorption and better electrochemical performances were achieved by decoration of ZnO nanorods with WO3−x nanoparticles.
Sayama et al. reported [57] a stoichiometric production of H2 and O2 using a mixture of Pt-WO3 and PtSrTiO3 (Cr–Ta-doped) in NaI media. The Pt-loaded SrTiO3 (Cr–Ta-doped) produced H2 of 0.8 μmol h−1 from an aqueous NaI solution while the Pt-loaded WO3 produced O2 at an initial rate of 84 μmol h−1 in an aqueous NaIO3 solution under visible light (l > 420 nm) separately. The H2 evolution rate from the mixed photocatalyst system (1.8 μmol h−1) was higher than that from Pt–SrTiO3 (Cr–Ta-doped) alone (0.8 μmol h−1), indicating that addition of the Pt–WO3 effectively reduced the IO3− ion to I.
Even though the band position of WO3 is suitable for O2 evolution, doping WO3 with a metal can shift the energy level. Wang [58] studied electronic properties of WO3 using density functional theory (DFT) calculations with a hybrid calculation. Replacing W by Mo and Cr in the lattice can modify the bandgap of WO3 and improve absorption of visible light. The effect of replacing O atoms by S anions was simulated by substitution along the Z direction in the unit cell. The DFT results predicted that there is a decrease in energy gap (2.21 eV) as well as a conduction band shift up, which is beneficial for HER. The authors also tested the effect of doping WO3 with Ti, Zr and Gf metals, resulting in a predicted upward shift of the conduction band like the case with S anions.
Mg-doped WO3 photocatalysts [59] have been studied experimentally. The conduction band edge for p-type Mg-doped WO3 was −2.7 eV vs. Saturated calomel electrode (SCE) at pH 12, which is more negative than the reduction potential of H2. Hydrogen generation of 3 μmol/gh was achieved by doping WO3 with Mg (5–10 wt%). Doping has also been done using other metals, including Mo [59].

4. Heterostructured WO3 Nanocomposites for Photoelectrochemical Cell Systems

Photocatalytic activity of WO3 depends on the crystal structure, morphology and surface areas. High surface areas of WO3 usually increase the photo activity via providing more reaction sites. The certain morphology increases electron mobility, thus demonstrates better photocatalytic activity. For example, one dimensional WO3 demonstrates relatively high photoactivity relative to nanoparticles. In the two dimensional WO3 nanomaterials, it is very important to have optimum grain size which lead to high photoactivity. The crystal structure is critical for the photoactivity of WO3. Furthermore, monoclinic structure of WO3 offers different photocatalytic activity relative to other crystalline structure including tetragonal, etc.
Anodization [60,61], solvothermal [56], hydrothermal [62,63], spin coating [64], electrodeposition [65,66] and sol–gel [67] methods were used to fabricate different morphologies and structures.

4.1. Crystalline Structure

Many research efforts have been performed to investigate the effect of crystal structure on the tungsten photocatalytic activity. It was found that the monoclinic crystalline phase demonstrated stronger oxidation activity than other crystal phases such as hexagonal and orthorhombic. The monoclinic phase was found to be the most stable at room temperature [68,69,70,71,72]. Increase of the temperature gradually transformed WO3·0.33H2O from orthorhombic into anhydrous hexagonal and a final stable form monoclinic (Figure 6a). As the temperature transited from 400 to 500 °C, the color of the film turned into yellowish color obviously, which is corresponded to a red shift (Figure 6b). The photocurrent density increases until 500 °C, then it starts to decrease (Figure 6c). The monoclinic structure of WO3 at 500 °C showed the highest photoelectrochemical performance, on the contrary the orthorhombic WO3·0.33H2O exhibited the lowest photocurrent density [68,69,70,73,74,75].
Nayak et al. [76] used combination of a facile precipitation and solvothermal methods to fabricate WO3 nanowires. The precipitation method produced (Figure 7a–d) WO3·H2O nanoplates with an orthorhombic phase, later the solvothermal approach was used to form WO3 nanowires with a monoclinic phase (Figure 7e–h). The photocurrent density obtained from WO3 monoclinic structure was 21-fold higher than that of WO3·H2O orthorhombic phase. This enhancement was ascribed to the presence of different phases between WO3·H2O nanoplates and WO3 nanowires or the high crystallinity of WO3 nanowires, which minimized the barrier of charge transfer at the interfacial charge and enhanced the PEC performance.
The effects of crystal phase on the photocatalytic performance has been broadly explored. Park et al. [77] found that the annealing treatment reduced the surface disorder induced by water via improvement of the crystallinity or oxygen deficiencies of WO3, led to enhancement of the PEC performance. Zeng’s group explained the formation of peroxo species on the surface of WO3·H2O as it has low degree of crystallinity. As the annealing temperature of WO3 reached 500 °C, highly reactive (002) facets were formed to reduce defects, thus to minimize the recombination of electron–hole pairs [75]. The same conclusion was obtained by Su’s group [72]. From the above studies, monoclinic WO3 demonstrated higher PEC performance than that of as-prepared hydrated WO3.
Recent investigations demonstrated that with surface engineering certain crystal planes possess preferences on the photoexcited electrons and holes, which lead to either preferential oxidation or preferential reduction reactions [78]. Furthermore, photo–electrochemical efficiency has been improved via exposing the high surface crystalline surface [79,80]. Among the three crystal planes or facets of WO3 which are (200) with 1.43 J/m2, (020) with 1.54 J/m2 and (002) with (1.56 J/m2) facet of WO3, the crystal facet (002) showed preference for adsorbing the reaction species due to its highest surface energy [81]. Wang et al. confirmed this via DFT calculations [82]. The dangling O atoms of the weakest W–O bond on the (002) crystal plane of the monoclinic WO3, offer plentiful active sites for H2O and organic molecules through the hydrogen bond. Oxidization of water and degradation of organics on the (002) easily occur via consuming photo-excited holes and generating active oxygen species, which reduce the recombination of photogenerated carriers [83,84,85].
The morphology of the WO3 films can be controlled by synthesis parameters such as synthesis time, temperature and the amount of the capping agent [75]. HRTEM study revealed that annealing WO3·H2O plates transformed along with the (020) crystal face into WO3 plates with preferentially (002) facet. At annealing 500 °C, the WO3 showed 1.42 mA cm−2 at 1.23 V vs. RHE, which is relatively high current density. This is explained due to reduction of peroxo species on the surface of WO3. The high energy crystal plane of WO3 nanoplate enhanced PEC water-splitting. Zhang et al. [83] compared monoclinic WO3 nanomultilayers which has preferable (002) facet with that of WO3 nanorods and found that WO3 nanomultilayers performed higher photocurrent densities than the WO3 nanorods. These results were explained not by the specific surface area of WO3 nanorods, but the presence of highly reactive (002) facets of WO3, which contributed to the improved PEC water-splitting performance.
In addition, increasing studies have been made investigating the effect of the (002) crystal plane of 2D monoclinic WO3 on PEC water-splitting.
To enhance the PEC water-splitting performance of WO3, most of studies have been focused on engineering morphology, crystallinity, heterojunction, oxygen vacancy, doping and co-catalysts for enhancement of photocatalytic hydrogen evolution.
According to the crystalline structure of WO3, it is confirmed that the monoclinic phase of WO3 demonstrated higher OER than the hexagonal or orthorhombic phases since it is the most stable phase at room temperature and presence of highly reactive (002) facets.

4.2. Morphologic Effect

Various WO3 nanomaterials with different morphologies including nanorods [86], nanoflake [87], nanotubes [88,89], nanoplates [90] and nanoparticles [91,92] were synthesized by various methods to provide active sites for catalysis. It was found that morphology change of WO3 can significantly influence photocatalytic activity.
Ma and other authors [93] obtained nanoplates of WO3 by topological method using Na2WO4 and HBF4 and mentioned that intrinsic crystal lattice of tungsten acid plays important role to obtain morphology of final products. The crystal lattice of H2WO4 has (WO6) octahedra layers with normal direction (010) and each layer is linked to each other via hydrogen bonds. That is why H2WO4 tends to form platelike nanocrystals with (010) direction. Another factor affecting morphology is addition of directing agents for nucleation and crystal growth. Interaction of H2WO4 crystal planes and HBF4 can be reason of formation plate morphology of WO3. Meng and others [94] synthesized hierarchical structure using citric acid C6H8O7 and found that (–COOH) functional groups affect growth of nanoplate. They concluded [93,94] that uniform platelike morphology is favorable for gas sensing because it has more active sides for absorption of gas molecules. In addition, much work was done on WO3 crystal growth using fluoroboric acid [95], polyethylene glycol (PEG) [96], polyvinyl alcohol (PVA) [87].
Strategy of increasing active sides for suitable absorption of light is a way to enhance photoelectrochemical performances of photocatalysts. For example, Jiao et al. demonstrated different morphology of tungsten trioxide hydrate (3WO3·H2O) films which grown by hydrothermal method using Na2SO4, (NH4)2SO4 and CH3COONH4 as capping agents, respectively. Platelike, wedgelike and sheet like nanostructures can be obtained as shown in Figure 8 [97]. From Figure 8e it can be seen that sheet like nanostructures had the highest photocurrent density (1.15 mA/cm2) under illumination and high photocatalytic activity for photodegradation of methanol. This was in good agreement with UV-vis absorbance spectroscopy results (Figure 8d). The authors believe that the reason for high current density of sheetlike morphology can be explained by the existence of small pores among sheets. This may be beneficial for accelerating the interface electron kinetics between the sheet and electrolyte due to its large active surface area.
Davidne et al. [98] reviewed WO3 nanostructures and studied the effect of morphology on photocatalytic activity for decomposition organic dyes. Nanostructures such as nanoplate, nanoneedle, nanorods and nanowire were obtained by hydrothermal method. It was found that photocatalytic efficiency has good correlation with band gap, crystalline phase, morphology and oxidation state. Nanoneedles with hexagonal structure showed the best photocatalytic efficiency in contrast to others.
Monoclinic nanorods showed higher photocurrent density than (2.09 mA/cm2) nanoplates (1.61 mA/cm2) in the hydrogen evolution reaction (HER) [96]. Some results concluded that 1D structures have high optical, electrical, photoconductivity properties and fast charge transportation [99,100,101]. Vertically oriented nanorods and nanoneedles have remarkable PEC results [65,66,102] due to high interfacial contact area which improves redox contact area and efficient light scattering.
However, some authors believed that 2D nanostructures like nanoplates have higher specific surface area than one-dimensional (1D) materials such as nanorods and nanowires. For example, Su and others demonstrated better photoelectrochemical characteristics and optical properties of WO3 nanoflakes than WO3 nanorods [87]. Hammad et al. [103] fabricated WO3 nanorods (with a diameter 7 nm, length up to 700 nm) and WO3 nanoplates with width 700 nm on fluorine-doped tin oxide (FTO) substrate via hydrothermal treatment. Results of electrochemical spectroscopy showed that nanoplates have better contact with substrate than nanorods which led to high photocurrent density of 400 μA/cm2 over 350 μA/cm2. Through changing concentration of HCl acid, Zhou et al. [104] synthesized perpendicularly oriented WO3 nanorods and nanoplates at different amount of acid. WO3 nanoplate arrays also showed a superior photocurrent density of 1 mA/cm2 at 1.6 V vs. RHE than nanorods of 0.8 mA/cm2. A high photocurrent density may be related with a long carrier diffusion length through 2D nanostructures comparing with 1D and efficient charge transportation and separation. Contradictory results between 2D and 1D nanostructures allow them to conclude that comparing different morphologies under different conditions do not give us true information. This is because one morphology can show different results depending on its morphologic parameter such as length, thickness and diameter.
In the aspect of photocatalytic efficiency evaluation in combination with morphology, WO3–BiVO4 nanostructures have been mostly studied. Lee and others [65] fabricated WO3–BiVO4, TiO2–BiVO4, Fe2O3–BiVO4 and SnO2–BiVO4 nanostructures and showed that PEC characteristics of bare WO3 dramatically increased after modification with BiVO4. The SEM and cross-sectional images of WO3 nanorods coated with BiVO4 are presented in Figure 9. They concluded that pairing WO3 with BiVO4 creates very promising photoanodes for water oxidation than others.
Improving PEC characteristics can be achieved by also using core–shell structures. Spatial separation of photogenerated charges between the core and shell is beneficial, however, excited charges stay inside and do not react with electrolyte. Nevertheless, fast transportation of charges to surface can diminish shell thickness [105]. Rao et al. [106] synthesized core–shell nanostructures of WO3–BiVO4 to improve light absorption and charge separation. A photocurrent and an incident photon-to-current conversion efficiency reached 3.1 mA/cm2 and ~60% at 300−450 nm, respectively at 1.23 V vs. RHE under simulated sunlight.
Enhanced PEC performance was obtained by designing yolk-shell-shaped WO3/BiVO4 heterojunction which produced a photocurrent density of 2.3 mA/cm2 with the highest value of ~5.0 mA/cm2 after adding a Fe–Ni co-catalyst at a bias of 1.23 V vs. RHE under AM 1.5 illumination (100 mW/cm2) [107]. These noticeable photocurrent density results demonstrated that core–shell structures may be potentially viable for photocatalytic applications.
Nanostructures with nanoporosity have shown a better PEC activity due to their large specific surface areas, relatively higher light absorption rate and excellent charge collection efficiency [108,109,110]. A high surface area of porous nanostructures makes them promising electrode materials for electrochemical surface reactions [111,112,113]. Furthermore, the nanoporous structure creates the depletion layer and reduced diffusion distance to the photoelectrodes/electrolyte interface, which diminish recombination of electrons and holes [114,115,116]. Song et al. [117] used versatile foaming-assisted electrospinning method to produce mesoporous WO3 nanobelts which enhanced the PEC water-splitting performance compared with the as-prepared WO3 nanofiber and WO3 nanobelt samples. Shin et al. [114] used a laser ablation method to produce tree-like nanoporous WO3 photoanode for a photoelectrochemical water–oxidation performance. Both SEM and TEM image in Figure 10, show 1D treelike morphology with a thickness ~3.2 μm and many clusters with nanoporous with average size of ~60 nm. The photocurrent density of treelike porous structures was 9-fold higher (1.8 mA/cm2 at 1.23 V vs. RHE)) than dense WO3. A quantum efficiency (QE) or incident photon-to-electron conversion efficiency (IPCE) was 70% at 350–400 nm.
Fujimoto et al. [115] synthesized porous BiVO4 using the auto combustion method. Adding oxidizing agent NH4NO3 and subsequent decomposition of organic additive after heating allowed to create porous film with small crystalline BiVO4 nanoparticles during the synthesis. The optimized WO3/BiVO4 film produced a maximum IPCE value of 64% at 440 nm with photocurrent density of 3.43 mA cm2 at 1.23 V vs. RHE (under one sun illumination).
Finally, it is obvious that different morphologies as a factor of synthesis method produce different PEC results and play an important role in configuration of water-splitting devices. However, it is hard to conclude that one morphology is more beneficial than others. In fact, other factors such as substrate on which the structure is grown, electrolyte, capping agents and, etc. may be cause of change in photocatalytic activity.

4.3. Binary Structures of Hierarchical Architectures Based on WO3 Semiconductors

As mentioned above, single photocatalysts cannot satisfy all the requirements needed for water-splitting PEC systems. Therefore, scientists focus most of their attention on creating different kinds of heterostructure architectures from different various materials such as metal oxide/metal oxides [47,118,119,120], metal oxide/metals [92,121,122,123] and metal oxide/inorganic compounds [124] to create efficient systems for various spheres. Mixing several materials is a method commonly used to improve separation of charge carriers, photoelectrode stability, absorption of visible light, suitable carrier diffusion length and effective surface charge transfer.

4.3.1. Metal Oxide/Metal Oxide Binary Heterostructures

The sensitivity of TiO2 can also be obtained by modification the surface using semiconductor photocatalyst with a smaller bandgap as WO3 [125,126,127]. A WO3/TiO2 photocatalytic system was published in 1998 [128]. The photocatalytic activity of WO3 coated with TiO2 was tested for the oxidation of water using iron (III) acceptor. It was also found that iron (III) ion is preferred more than iron (II) ion as an electron acceptor for oxygen evolution.
Amorphous α-TiO2 can be used to passivate the surface of WO3 based nanostructured photoanodes. For example, Yang and et al. [129] demonstrated high photocurrent with 1.4 mA/cm2 at 0.8 V in 0.1-M Na2SO4 electrolyte using WO3 nanoflakes coated amorphous α-TiO2 films. Passivation of WO3 by α-TiO2 was realized through the O2−W6+ bonding at contact surface between WO3 and α-TiO2. Hence, passivation of surface allows to decrease recombination and improve PEC oxidation.
Hierarchical WO3/TiO2 composites for hydrogen evolution was fabricated by Momeni [130] using the anodization method. TiO2 nanotubes with a diameter ranged 80–110 nm were modified by WO3. Controlling the concentration of the Na2WO4 solution allowed them to achieve the highest amount of H2, with 2.14 mL/cm2 under 120 min of solar illumination, which is approximately 3.02-fold higher than bare samples with TiO2 nanotubes (0.71 mL/cm2). It also showed increase of photocurrent value from 0.81 to 1.61 mA/cm2 after modification proved the effectiveness of the coupled WO3/TiO2 system. The anodization method was also successfully used to prepare the hybrid WO3/TiO2 nanotube photoelectrodes [131] which showed better photo conversion efficiency, STH efficiency and H2 generation.
Many other studies highlighted that coupled WO3–TiO2 systems have better characteristics. For example, the highest photocatalytic activity of nanocomposites particles for degradation of Rhodamine B [60,61], decomposition of 1,4-dichlorobenzene (DCB) aqueous solution [132] and azo dyes [133], for effective catalytic oxidation cyclopentene to glutaraldehyde [133] were obtained.
The effectiveness of photocatalysts was also attained by engineering morphology and specific surface area of the material since electron-hole transfer occurs on the surface. According to some studies, although electron–hole pairs can be generated in volume, they can annihilate before they reach the surface due to the low diffusion length.
Most of authors have shown that dual heterostructures of WO3 and BiVO4 are effective for driving water oxidation reactions [64,134,135]. BiVO4 is an n-type semiconductor-like WO3 which has a bandgap around 2.4 eV [26,136]. The theoretical solar-to-hydrogen efficiency using this material (9.8%) is more than that of WO3 (4.8%)-based systems. Despite the fact that BiVO4 is a direct semiconductor unlike WO3, it has a poor charge transport and a short hole-diffusion length (~60 nm). Coupling WO3 with BiVO4 decreases recombination of photogenerated charge carriers and improves the efficiency of overall water-splitting systems. A dynamic of photogenerated carriers and effective charge separation of WO3–BiVO4 heterojunctions was explained by Grigioni using femtosecond transient absorption spectroscopy [137]. They determined the position of the WO3 conduction bands (+0.41 V vs. RHE) and BiVO4 (+0.02 V vs. RHE) by testing the photocatalytic reduction of thionine. The charge separation mechanism of BiVO4–WO3 system is shown in Figure 11a. When comparing the flat band position of BiVO4 in the WO3–BiVO4 composite, a shift of 170 mV is observed. This shift was explained by electron equilibrium between the two materials due to a Fermi level shift. Photoelectrons transfer from BiVO4 to WO3 while holes localize in BiVO4, so it is possible to separate photogenerated charge carriers spatially (Figure 11b).
The morphology of WO3 and the decoration method with BiVO4 are also very important issues. Chae et al. synthesized mesoporous WO3 films followed by a coating of BiVO4 to research particle sizes and shapes, as well as the effects of the photoanode thickness. Large nanoplates showed a high injection efficiency while nanospheres enhanced the charge-separation efficiency [138]. Pihosh et al. synthesized WO3–BiVO4 vertically oriented nanorods by combining the glancing-angle deposition and normal physical sputtering techniques [66]. The photocurrent density achieved 3.1 mA/cm2 at 1.23 V RHE under illumination of one sun. A nanopillar morphology of WO3–BiVO4 photoanodes fabricated by electrostatic spraying method also produced a photocurrent up to 3.2 mA/cm2 [139]. An enhanced photocurrent density of 4.55 mA/cm2 was achieved by using a WO3–BiVO4 photoanode [65]. Deposition of BiVO4 nanodots on WO3 nanorods had an increased photon to hydrogen efficiency of 80% at 1.23 V vs. RHE, which is higher than the theoretical efficiency for bare BiVO4. Rao et al. fabricated WO3–BiVO4 core shell nanowires and showed that the photoanodes demonstrated a η a r b × η s e p up to 53%. A combination of BiVO4 with more conductive WO3 leads to effective charge carrier separation and the photocurrent achieved 3.1 mA/cm2 at 1.23 V vs. RHE.
Iron is an abundant and important metal in the earth’s crust, so its use is considered economically viable. Oxidation of iron can lead to formation of the known hematite phase α-Fe2O3 which has semiconductor properties. It has good stability in most electrolytes pH > 3 and has a narrow bandgap (~2.2 eV) which can absorb 40% of the solar spectrum. Although hematite electrodes are well studied for PEC system, photoconversion efficiency is still lower than the theoretical value due to low hole mobility (~2–4 nm). Moreover, poor electrical conductivity, high recombination rate of electron-hole pairs [140] and the slow kinetics of oxygen evolution [141] limits its use. Some studies focus on binary heterostructures with WO3–hematite α-Fe2O3 photoanodes [142]. A photocurrent of 1.66 mV/cm2 was observed at 1.23 V RHE, while the photon to current efficiency was 73.7% at 390 nm. Schematic illustration of WO3-Fe2O3 composite nanosheets and bandgaps are shown in Figure 12a. The optical absorption measured by a UV-vis diffuse reflectance spectroscopy was found to be improved for the composite WO3 and Fe2O3 material (Figure 12b). Luo published enhanced electrochemical characteristics of a WO3@Fe2O3 photoelectrode compared to bare WO3 and Fe2O3 [143]. Effective photoelectrochemical splitting of seawater with Fe2O3/WO3 nanorods was achieved by Li et al. [144]. Although optical absorption is promising, the photocurrent of Fe2O3/WO3-based photoanodes is still low.

4.3.2. Metaloxide/Inorganic Compounds Heterostructures

The heterostructures formed from WO3 and sulfur components have narrow bandgaps. For example, antimony sulfide, Sb2S3, (1.7–1.9 eV), bismuth sulfite, Bi2S3, (~1.3 eV) and tungsten disulfide, WS2, (~1.3 eV) [145] are very effective. For example, Zhang [146] synthesized WO3/Sb2S3 heterostructures via a simple hydrothermal method to improve PEC performances. Tungsten trioxide nanorods and nanoplates were synthesized by controlling the concentration of acid and tungsten precursor along with subsequent growth of Sb2S3 nanoparticles. It was demonstrated, that WO3/Sb2S3 heterostructures have better electrochemical and optical characteristics than pristine WO3.
A high photocurrent of 5.95 mV/cm2 was achieved using a three-dimensional WO3/Bi2S3 heterojunction [147]. Bi2S3 is also a n-type semiconductor with bandgap 1.3 eV and has more negative conduction band edge than WO3. A WO3/Bi2S3 heterojunction was fabricated by combining of hydrothermal method, SILAR (successive ionic layer absorption and reaction) process and CBD (chemical bath deposition). Relatively high light absorption, small electron transfer impedance and high charge carrier were proved by UV-vis, EIS (Electrochemical Impedance Spectroscopy) and Mott-Shottky methods.
Despite the fact that the WO3 bandgap energy is not suitable for hydrogen evolution, it is still useful for solving problems such as high electron-hole recombination rates and poor electrical conductivity of some photocathodes [145]. A study of WO3@WS2 core–shell nanostructures fabricated by plasma assisted sublimation was published by Kumar et al. [145]. The highest achieved cathodic photocurrent was 16.2 mA/cm2 for WS2 at 0.3 V vs. RHE. Sulfurization of the WO3 surface forms a WS2 layer with a rich defect structure, resulting in a high catalytic activity.

4.3.3. Metal Oxide/Plasmon Particle Systems

The plasmonic effect induced by noble metal particles plays an important role in decorating photoelectrodes. Photoactivity of photocatalysts can be increased by enhancing light scattering and SPR [148,149]. Moreover, noble metals play a role as a co-catalyst for OER due to good electrical contact between the metal and semiconductor [150,151]. Altering the surface properties of WO3 photoanodes with plasmonic nanoparticles Au and Ag has shown enhanced visible light absorption and high photocurrent density [152]. Hu showed a high faradic efficiency of 94% for WO3@Au composites [149]. Enhanced photocurrent density and morphology of heterostructure is presented in Figure 13a,b. Modified WO3 by plasmonic Ag and Pt nanoparticles showed enhanced photocurrent of 1.13 mA/cm2 at 1.23 V vs. RHE under AM 1.5G illumination in a 0.2 M Na2SO4 solution, which is nearly 3.32 times that of bare WO3 [153]. The photocurrent density for binary systems are represented in Table 1.

4.4. Ternary Systems for Efficient Water Decomposition

A typical heterojunction between two dissimilar semiconductors comes to equilibrium without any external electric field. The result shown above is a potential difference that appears at the interface as an internal electric field. This field accelerates charge carrier drift and decreases the number of electron–hole recombination, improving the semiconductor’s photocatalytic activity [155]. The exploitation of multilayer structures in photocatalysis is considered more beneficial. An illustration of the advantages of a cascade transition of charge carriers is well explained in numerous works [156] and [157]. The authors of [156] work investigated the photocatalytic properties of a composite of CdS, TiO2 and tungsten trioxide. Since, TiO2 has the conduction band edge which is between conduction band edges of CdS and WO3, in such ternary composites electronic transitions are cascading. After generation of excitons in CdS, electrons easily migrate to TiO2 and WO3 of potential difference as shown in Figure 14.
In addition to ternary composites made only of semiconductor materials, there are also ternary structures based on semiconductors and metal complexes. An example of such ternary photocatalysts can be photonic heterostructure of CdS–Au–WO3. Cui et al. reported, that Au nanoparticles deposited between WO3 and CdS leads to form heterostructure which had photocatalytic properties superior to similar two-phase systems. It was thought that such amelioration caused by a synergistic integration of photon absorption with effective electron transfer in the heterostructure [62]. The use of doping elements—or modification by particles such as CdS quantum dots [158] in usual two-component heterojunctions—is also referred to as a ternary structure. Doping elements such as Yb [159] and Mo have been shown to suppress charge carrier recombination during photocatalysis [160], improving the efficiency of reactions. In ref. [161], high photocurrent density was achieved by doping active materials to make bilayers WO3/MoBiVO4 (BiV0.95Mo0.05O4). The photocurrent of the Mo-doped content increased by a factor of 3 and 1.5 relative to pristine photoanodes based on WO3 and BiVO4/WO3, respectively. Another promising way to create performable photocatalysts is a combination of a catalyst with heterostructures. Some examples of such formations are materials which obtained by the deposition of catalytic coatings NiOOH [162] and FeOOH [163] on the surface of two-phase WO3 structures. The deposited materials suppress both the degradation of the photoactive material and the Faradic losses [164], accelerating the photoelectrochemical reaction processes. Shouli Bai et al. also combined catalyst layers with a heterojunction by depositing NiFe bimetallic complexes onto a WO3/Fe2O3 surface [165]. Their strategy increased the photocurrent density of the ternary photoanode system to 3.0 mA cm−2, which, according to Shouli Bai, is 5 and 7-fold higher relative to that of pristine WO3 and α-Fe2O3 structures, respectively. The role of the catalyst here is to improve the absorption of holes accumulated on the electrode surface. In another study, the deposition of CoPd bimetallic nanoparticles onto the surface of WO3/α-Fe2O3 photoanodes causes a cathode shift of the initial potential, increasing the photocurrent density from 0.15 to 0.5 mA/cm2 during water oxidation at 1.23 V relative to RHE when illuminated with AM 1.5 G [166]. Substitution of iron oxide with cadmium sulfide in heterostructures based on WO3 also makes it possible to sensitively increase the photoresponse of electrodes during water decomposition. In [167] work preparation of such ternary compositions conducted via three simple hydrothermal, impregnation and photo-assisted deposition steps. Thus, authors obtained rodlike structures with a performance of photocurrent at the level of 5.85 mA/cm2 at 1.23 V (vs. RHE). Sun with co-workers explain this phenomenon by creation a larger built-in potential at interface WO3/CdS formed via impregnating appropriate CdS onto surface of WO3. This drives transport of electrons from CdS to WO3 with improvement of exciton separation. In this case, not the entire charge is torn well enough. Part of the charge recombines due to the weak involvement of holes in the valence band of cadmium sulfide in the process of water oxidation. Decoration surface of WO3/CdS heterojunctions with Co-Pi co-catalyst advances the transfer kinetics of charge advanced which is positive to suppression of charge recombination. In this case, the mechanism of improving charge transfer to the sites of redox half-reactions is also achieved by adding phosphate anions to the electrolyte. In fact, the use of various electrolytes, such as glycerol-water mixture [168], in its effect on the parameters of the transferred charges between photocatalytic coatings and a split liquid. Varieties of compositions of working electrolytes and a list of structural heterojunctions, as well as geometric schemes that receive the influence and influence of all this on the photoresponse of PEC systems are shown in Table 2. It is known that tungsten trioxide is widely used as the primary semiconductor material in three transient systems in photocatalysis. Therefore, numerous of recent works dedicated to photoinduced dye degradation processes [169] and the direct splitting of water [170,171] were carried out with exploiting WO3. Liu formed an effective three-component photoanode based on tungsten trioxide nanosheets synthesized by hydrothermal method and decorated with ZnxBi2S3+x quantum dots via layer-by-layer adsorption [172]. In obtained core/shell structure of ZnxBi2S3+x/WO3, surface ZnxBi2S3+x served as a protective layer for tungsten trioxide. Comparable photocatalytic studies conducted under visible light irradiation in 0.1-M Na2S and 0.1-M Na2SO3 aqueous solution at pH ~9 showed that ZnxBi2S3+x/WO3 composite has higher photocatalytic activity then Bi2S3/WO3.structure. The photocurrent density was 7 mA/cm2 at a bias voltage of −0.1 V. This attributed to the involvement of ZnS nanostructures with high photocatalytic properties [173]. Designing semiconducting heterocomposites via a surface–interface engineering approach showed high effectiveness for enhancing exciton separation/transportability and photoelectrochemical features. The photoactive layer of 2D BiVO4-layer/WO3 array modified with cobalt phosphate studied for oxygen evolution and showed 1.8 mA/cm2 at 1.23 V vs. RHE in a phosphate buffer under an AM1.5G sun. The result is 5 and 12-fold higher than that of bare WO3 and BiVO4 photoanodes, respectively [174]. Morphology of WO3 nanostructures affects to charge separation ability in the active layer and to charge collection efficiency in the WO3/BiVO4 heterojunction. The low-dimensional nanosphere WO3 layer showed higher photocatalytic efficiency than the platelike or rodlike ones [71].
There are three different component systems with different morphology in which the hierarchy of the band structure observed for effective water-splitting. The main types of morphology include multi-heterojunction-based photocatalysts composed of WO3 nanorods, Pt nanoparticles and TiO2 nanoparticles [175], hierarchical heterostructures with core and double shells [176], rectangular AgIn(WO4)2 nanotubes which showed excellent photocatalytic properties for decomposing water to evolve H2 [177] and linked porous structures such as WO3/porous–BiVO4/FeOOH [178]. Special attention should be paid to the three-component system made on spiral WO3 nanostructures decorated with doped Mo and BiVO4 nanoparticles [179]. Based on the assertions that 1D structures charge carriers go straight to the terminals or half-reaction centers [180], Xinjian Shi et al. used a spring morphology with an increased surface area while retaining the properties of 1D structures. As a result of the successive deposition of these structures onto conductive glass, a spiral heterocomposite WO3/(W, Mo)–BiVO4 with a length of 5.5 µm was obtained and the photocurrent density exceeded 3.9 mA/cm2. The process of making triple transitions is possible using various techniques, such as electrochemical reduction-oxidation technology [181], a hybrid synthetic method, including electrodeposition and hydrothermal treatment [182], a solvothermal-calcination process [183], an electrostatic-driven self-assembly correlated with ion-exchange process [162] and a polymer complex method [184]. Jun Lv et al. obtained photoactive LiCr(WO4)2. After heat treatment at 700 ℃ for 5 h, crystal structures of tungstate were obtained, consisting of alternating layers of WO6 and CrO6 octahedra lying on the YZ plane. The WO6 octahedra are connected by separating edges, leading to the appearance of unrelated zigzag rows along the Y-axis. On the other hand, CrO6 octahedra not related to each other. Photocatalytic tests of LiCr(WO4)2 showed that the release of H2 proceeds upon irradiation with visible light up to 540 nm [184]. Reaching the rest of the visible spectrum remains the goal. Ji Hyun Baek et al. developed a photoanode based on BiVO4/WO3/SnO2 using a simple method of applying layers on conductive glass to obtain a thin double-heterojunction coating on the order of 320 nm. A characteristic feature of BiVO4/WO3/SnO2 is the large bandwidth of visible light, which allowed the researchers to combine the photoanode with a perovskite solar cell into a tandem PEC system. This allowed the reaction to split water without applying a bias voltage [26]. In general, the development of PEC cells using independent absorbents of incident light is a promising direction, so the next section will deal tandem hydrogen production systems.

4.5. WO3-Based Tandem PEC Cells

Two strategies can produce tandem cells for photoinduced water-splitting. One strategy used is to increase the capture of photons is a tandem system of a photoanode and photocathode with n- and p-conductivities of active layers, respectively. The splitting of water under light irradiation in this type of PEC cell occurs as a result water oxidation by photogenerated holes on the photoanode surface. Meanwhile, reduction to molecular hydrogen is initiated by electrons on the photocathode surface. At the same time, negative charge carriers generated in the photoanode are directed under the action of the field, to positively charged carriers in the photocathode material, where they recombine. Robert Coridan et al. investigated the photocatalytic properties of Si/WO3 heterojunctions and Si/ITO/WO3 arrays in a core-shell manner [185]. The operation of the tandem structure depends not only on the bandgap of the semiconductors used but also on the alignment of the strip edge and the state of the electrical connection between photo absorbents. When a mismatch of conductivity levels and valence levels of semiconductors included in tandem circuits occurs, low efficiencies of PEC hydrogen evolution cells [186]. A similar effect was observed in the work of Heli Wang et al. in which they combined n-type tungsten trioxide and hematite nanorod metal oxides with p-GaInP2. It was found that even when photo electrodes are illuminated with a source with a power of 1 W/cm2, a photocurrent appears but the density values of which are rather weak. This is due to low electron mobility of in the hematite layers, short hole-diffusion length, and insufficient potential difference between the levels of the conduction band and valence semiconductors, which help reduce charge recombination [187]. Geometrical optimization of the morphology of the active layers of photoelectrodes refers to an increase the density of short circuit current [188]. Investigating the PEC properties of the tandem structure of WO3/Si, Zhuo Xing et al. concluded that it was necessary to add an intermediate layer between p and n semiconductors to reduce the number of recombinations of photoinduced charges. In [189], metal W served as an intermediate layer, resulting in a WO3/W/Si ternary structure demonstrated an increase in the photogenerated current density by a factor of 10 compared to the WO3/Si structure.
Another possible way for general water-splitting without assistance is to combine photoelectrodes with photovoltaic cells to form a tandem PEC/PV cell. In one study [190], organic molecules were used as sensitizers in a tandem PEC, which is a powerful strategy for designing hydrogen evolution systems since they allow large-scale modification of photoelectrodes by adjusting the dye redox potentials or redox mediators.
In the tandem devices shown in Figure 15, the BiVO4/WO3 photoelectrode absorbs short-wavelength photons and more extended wavelength absorbs by a dye/TiO2 electrode [185,186,187,188,189,190]. This method offers better concession between device performance, complexity and stability [3]. In addition to scientific methods, when choosing materials, morphology and hierarchy of architecture, engineering aspects related to the spatial and reciprocal arrangement of the physical elements of tandem structures are also important. For efficient use of incident photons, Pihosh Y. and colleagues produced a PEC-PV tandem system based on WO3 NRs/BiVO4+CoPi photoanode and an AlGaAsP solar cell, which were placed at 45° relative to each other using a V-shaped stand [102]. This design allows the passage of reflected light from the photoanode to the surface of the solar cell.
Thus, the characteristic differences between ternary systems and binary systems are the improvement of photocatalytic properties and corrosion resistance. As described above, tandem structures provide an operating mode for photoelectrochemical processes in a wide range of the radiation spectrum. It increases the number of components of the hydrogen evolution cells, which leads to a complication of the assembly of heterostructures and to a high cost of the obtained layers. Therefore, when choosing which or used components of tertiary composites, pay attention to postprocessing, including thermal.

5. Conclusions and Outlook

The goal of research in the field of photoinduced decomposition of water is to develop high performance photocatalytic systems with high STH efficiency. The transmission of the photocatalytic systems from the field of laboratory research to the large scale production is a key point. The principle of using semiconductor coatings based on tungsten trioxide for PEC cells is justified by the economic aspects associated with the low cost of the material, as well as with its physicochemical properties. Using nanotechnology and nanomaterials is a suitable method for addressing several of the issues listed above. Metal oxide nanoparticles can be obtained by a wide range of physical and chemical methods. They can be classified as top-down and bottom-up methods. Top-down approaches rely on physical processes, such as abrasion or ball milling. Nano powders produced in this method usually exhibit wide distribution sizes and their size, shape and morphology are difficult to control. In addition, possible structural and surface impurities can have a significant effect on surface chemistry and the catalytic properties of nanomaterials. Low-dimensional structures are most advantageous from the point of view of effective absorption of light with the generation of charge carriers, migration of charge carriers to the surfaces of the material, which fit over the exciton lifetime, as well as possessing a fairly significant semiconductor/electrolyte contact area. The mixing of semiconductors (i.e., the formation of composites) is also an accepted strategy for the development of photocatalysts that respond to radiation in the visible range. This strategy is based on a hierarchical architecture for connecting a wide-gap semiconductor with a narrow-gap semiconductor with a more negative level of the conduction band. Thus, the conduction of electrons can be introduced from a narrow-gap semiconductor into a wide band semiconductor, leading to better absorption for the mixed photocatalyst. An additional advantage of using composite semiconductor photocatalysts are that it reduces carrier recombination by facilitating electron transfer crossing interface of particles. In photocatalyst composites, semiconductor particles stay in electronic contact individually. For a successful combination of semiconductors, the certain requirements are needed to be met: the conductivity level of the narrow-gap semiconductor should be more negative than the level of the wide-gap semiconductor; the position of the conductivity level of the wide-gap semiconductor should be more negative than the recovery potential; electron injection should be quick. All of the steps can improve the characteristics of the material, as well as eliminate the influence of its shortcomings on the process of splitting water under the action of light. In any case, a review of the literature in this area indicates a special level of development in the field of photoelectrochemistry for hydrogen evolution using active materials from pure tungsten trioxide or in various compositions with it. However, so far, the complete and qualitative decomposition of water and the generation of hydrogen under the influence of sunlight has a low rate, which indicates insufficient feasibility of industrial use of existing technologies. Based on the current trend towards the creative and experimental activity of researchers in this direction, the authors of this article express deep confidence in the imminent achievement of quantum efficiency of PEC systems sufficient for universal use in human life in the near future.

Author Contributions

N.N. designed and supervised the whole manuscript. He wrote the whole manuscript together with Y.S., A.M. and N.B.; Y.S. contributed to write Section 2 and Section 4.1; A.M. wrote Section 3, Section 4.2 and Section 4.3. N.B. wrote Section 4.4 and Section 4.5; all authors wrote introduction, conclusion and outlook together. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded by Nazarbayev University FGRG grant number SEDS2020 016 and the Young Scientist Program of the Committee of Science of the Ministry of Education and Science of the Republic of Kazakhstan, grant number AP08052381.

Acknowledgments

The authors would like to express special thanks to Aygul Nuraje from BTHS for proof-reading and editing the manuscript. N.N. greatly acknowledge financial supported from Nazarbayev University. N.B. thanks the Young Scientist Program of the Ministry of Science and Education of the Republic of Kazakhstan for financial support.

Conflicts of Interest

The authors declare no potential conflict of interest.

References

  1. Gratzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef] [PubMed]
  2. Sathre, R.; Scown, C.D.; Morrow, W.R.; Stevens, J.C.; Sharp, I.D.; Ager, J.W.; Walczak, K.; Houle, F.A.; Greenblatt, J.B. Life-cycle net energy assessment of large-scale hydrogen production via photoelectrochemical water splitting. Energy Environ. Sci. 2014, 7, 3264–3278. [Google Scholar] [CrossRef]
  3. Prevot, M.S.; Sivula, K. Photoelectrochemical tandem cells for solar water splitting. J. Phys. Chem. C 2013, 117, 17879–17893. [Google Scholar] [CrossRef]
  4. Wang, G.; Wang, H.; Ling, Y.; Tang, Y.; Yang, X.; Fitzmorris, R.C.; Wang, C.; Zhang, J.Z.; Li, Y. Hydrogen-treated TiO2 nanowire arrays for photoelectrochemical water splitting. Nano Lett. 2011, 11, 3026–3033. [Google Scholar] [CrossRef]
  5. Zhang, Z.; Zhang, L.; Hedhili, M.N.; Zhang, H.; Wang, P. Plasmonic gold nanocrystals coupled with photonic crystal seamlessly on TiO2 nanotube photoelectrodes for efficient visible light photoelectrochemical water splitting. Nano Lett. 2012, 13, 14–20. [Google Scholar] [CrossRef]
  6. Zhang, Z.; Hossain, M.F.; Takahashi, T. Photoelectrochemical water splitting on highly smooth and ordered TiO2 nanotube arrays for hydrogen generation. Int. J. Hydrog. Energy 2010, 35, 8528–8535. [Google Scholar] [CrossRef]
  7. Sivula, K.; Le Formal, F.; Grätzel, M. Solar water splitting: Progress using hematite (α-Fe2O3) photoelectrodes. ChemSusChem 2011, 4, 432–449. [Google Scholar] [CrossRef]
  8. Cesar, I.; Kay, A.; Gonzalez Martinez, J.A.; Grätzel, M. Translucent thin film Fe2O3 photoanodes for efficient water splitting by sunlight: Nanostructure-directing effect of Si-doping. J. Am. Chem. Soc. 2006, 128, 4582–4583. [Google Scholar] [CrossRef]
  9. Liu, Y.; Xu, Z.; Yin, M.; Fan, H.; Cheng, W.; Lu, L.; Song, Y.; Ma, J.; Zhu, X. Enhanced photoelectrocatalytic performance of α-Fe2O3 thin films by surface plasmon resonance of Au nanoparticles coupled with surface passivation by atom layer deposition of Al2O3. Nanoscale Res. Lett. 2015, 10, 374. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, D.; Li, R.; Zhu, J.; Shi, J.; Han, J.; Zong, X.; Li, C. Photocatalytic water oxidation on BiVO4 with the electrocatalyst as an oxidation cocatalyst: Essential relations between electrocatalyst and photocatalyst. J. Phys. Chem. C 2012, 116, 5082–5089. [Google Scholar] [CrossRef]
  11. Jo, W.J.; Jang, J.W.; Kong, K.-j.; Kang, H.J.; Kim, J.Y.; Jun, H.; Parmar, K.; Lee, J.S. Phosphate doping into monoclinic BiVO4 for enhanced photoelectrochemical water oxidation activity. Angew. Chem. 2012, 124, 3201–3205. [Google Scholar] [CrossRef]
  12. Wang, D.; Jiang, H.; Zong, X.; Xu, Q.; Ma, Y.; Li, G.; Li, C. Crystal facet dependence of water oxidation on BiVO4 sheets under visible light irradiation. Chem. A Eur. J. 2011, 17, 1275–1282. [Google Scholar] [CrossRef]
  13. 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]
  14. Wolcott, A.; Smith, W.A.; Kuykendall, T.R.; Zhao, Y.; Zhang, J.Z. Photoelectrochemical study of nanostructured ZnO thin films for hydrogen generation from water splitting. Adv. Funct. Mater. 2009, 19, 1849–1856. [Google Scholar] [CrossRef]
  15. Maeda, K.; Domen, K. Solid solution of GaN and ZnO as a stable photocatalyst for overall water splitting under visible light. Chem. Mater. 2010, 22, 612–623. [Google Scholar] [CrossRef]
  16. Tacca, A.; Meda, L.; Marra, G.; Savoini, A.; Caramori, S.; Cristino, V.; Bignozzi, C.A.; Pedro, V.G.; Boix, P.P.; Gimenez, S. Photoanodes based on nanostructured WO3 for water splitting. ChemPhysChem 2012, 13, 3025–3034. [Google Scholar] [CrossRef]
  17. Hameed, A.; Gondal, M.; Yamani, Z. Effect of transition metal doping on photocatalytic activity of WO3 for water splitting under laser illumination: Role of 3d-orbitals. Catal. Commun. 2004, 5, 715–719. [Google Scholar] [CrossRef]
  18. Enesca, A.; Duta, A.; Schoonman, J. Study of photoactivity of tungsten trioxide (WO3) for water splitting. Thin Solid Films 2007, 515, 6371–6374. [Google Scholar] [CrossRef]
  19. Pala, R.A.; Leenheer, A.J.; Lichterman, M.; Atwater, H.A.; Lewis, N.S. Measurement of minority-carrier diffusion lengths using wedge-shaped semiconductor photoelectrodes. Energy Environ. Sci. 2014, 7, 3424–3430. [Google Scholar] [CrossRef] [Green Version]
  20. Coridan, R.H.; Arpin, K.A.; Brunschwig, B.S.; Braun, P.V.; Lewis, N.S. Photoelectrochemical behavior of hierarchically structured Si/WO3 core–shell tandem photoanodes. Nano Lett. 2014, 14, 2310–2317. [Google Scholar] [CrossRef] [Green Version]
  21. Wang, Y.; Shi, H.; Cui, K.; Zhang, L.; Ge, S.; Jinghua, Y. Reversible electron storage in tandem photoelectrochemical cell for light driven unassisted overall water splitting. Appl. Catal. B Environ. 2020, 275, 119094. [Google Scholar] [CrossRef]
  22. Li, Y.; Zhang, W.; Qiu, B. Enhanced surface charge separation induced by Ag nanoparticles on WO3 photoanode for photoelectrochemical water splitting. Chem. Lett. 2020, 49, 741–744. [Google Scholar] [CrossRef]
  23. Jun, J.; Ju, S.; Moon, S.; Son, S.; Huh, D.; Liu, Y.; Kim, K.; Lee, H. The optimization of surface morphology of Au nanoparticles on WO3 nanoflakes for plasmonic photoanode. Nanotechnology 2020, 31, 204003. [Google Scholar] [CrossRef]
  24. Huang, J.; Zhang, Y.; Ding, Y. Rationally designed/constructed CoOx/WO3 anode for efficient photoelectrochemical water oxidation. Acs Catal. 2017, 7, 1841–1845. [Google Scholar] [CrossRef]
  25. Wang, S.L.; Mak, Y.L.; Wang, S.; Chai, J.; Pan, F.; Foo, M.L.; Chen, W.; Wu, K.; Xu, G.Q. Visible–Near-Infrared-Light-Driven Oxygen Evolution Reaction with Noble-Metal-Free WO2–WO3 Hybrid Nanorods. Langmuir 2016, 32, 13046–13053. [Google Scholar] [CrossRef] [PubMed]
  26. Baek, J.H.; Kim, B.J.; Han, G.S.; Hwang, S.W.; Kim, D.R.; Cho, I.S.; Jung, H.S. BiVO4/WO3/SnO2 Double-Heterojunction photoanode with enhanced charge separation and visible-transparency for bias-free solar water-splitting with a perovskite solar cell. ACS Appl. Mater. Interfaces 2017, 9, 1479–1487. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, Y.; Wygant, B.R.; Mabayoje, O.; Lin, J.; Kawashima, K.; Kim, J.-H.; Li, W.; Li, J.; Mullins, C.B. Interface engineering and its effect on WO3-based photoanode and tandem cell. ACS Appl. Mater. Interfaces 2018, 10, 12639–12650. [Google Scholar] [CrossRef]
  28. Lee, W.J.; Shinde, P.S.; Go, G.H.; Ramasamy, E. Ag grid induced photocurrent enhancement in WO3 photoanodes and their scale-up performance toward photoelectrochemical H2 generation. Int. J. Hydrogen Energy 2011, 36, 5262–5270. [Google Scholar] [CrossRef]
  29. Zhang, J.; Zhu, G.; Liu, W.; Xi, Y.; Golosov, D.; Zavadski, S.; Melnikov, S. 3D core-shell WO3@ α-Fe2O3 photoanode modified by ultrathin FeOOH layer for enhanced photoelectrochemical performances. J. Alloys Compd. 2020, 154992. [Google Scholar] [CrossRef]
  30. Zhang, L.J.; Li, S.; Liu, B.K.; Wang, D.J.; Xie, T.F. Highly efficient CdS/WO3 photocatalysts: Z-scheme photocatalytic mechanism for their enhanced photocatalytic H2 evolution under visible light. ACS Catal. 2014, 4, 3724–3729. [Google Scholar] [CrossRef]
  31. Lou, Z.; Zhu, M.; Yang, X.; Zhang, Y.; Whangbo, M.-H.; Li, B.; Huang, B. Continual injection of photoinduced electrons stabilizing surface plasmon resonance of non-elemental-metal plasmonic photocatalyst CdS/WO3− x for efficient hydrogen generation. Appl. Catal. B Environ. 2018, 226, 10–15. [Google Scholar] [CrossRef]
  32. Islam, S.Z.; Reed, A.; Wanninayake, N.; Kim, D.Y.; Rankin, S.E. Remarkable enhancement of photocatalytic water oxidation in N2/Ar plasma treated, mesoporous TiO2 films. J. Phys. Chem. C 2016, 120, 14069–14081. [Google Scholar] [CrossRef]
  33. Islam, S.Z.; Rankin, S.E. Hydrazine-based synergistic Ti (III)/N doping of surfactant-templated TiO2 thin films for enhanced visible light photocatalysis. Mater. Chem. Phys. 2016, 182, 382–393. [Google Scholar] [CrossRef] [Green Version]
  34. Hou, D.; Hu, X.; Ho, W.; Hu, P.; Huang, Y. Facile fabrication of porous Cr-doped SrTiO3 nanotubes by electrospinning and their enhanced visible-light-driven photocatalytic properties. J. Mater. Chem. A 2015, 3, 3935–3943. [Google Scholar] [CrossRef]
  35. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef]
  36. Takanabe, K. Solar water splitting using semiconductor photocatalyst powders. In Solar Energy for Fuels; Springer: New York, NY, USA, 2015; pp. 73–103. [Google Scholar]
  37. Takanabe, K. Photocatalytic water splitting: Quantitative approaches toward photocatalyst by design. ACS Catal. 2017, 7, 8006–8022. [Google Scholar] [CrossRef]
  38. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  39. Youngblood, W.J.; Lee, S.-H.A.; Maeda, K.; Mallouk, T.E. Visible light water splitting using dye-sensitized oxide semiconductors. Acc. Chem. Res. 2009, 42, 1966–1973. [Google Scholar] [CrossRef] [Green Version]
  40. Maeda, K.; Eguchi, M.; Lee, S.-H.A.; Youngblood, W.J.; Hata, H.; Mallouk, T.E. Photocatalytic hydrogen evolution from hexaniobate nanoscrolls and calcium niobate nanosheets sensitized by ruthenium (II) bipyridyl complexes. J. Phys Chem. C 2009, 113, 7962–7969. [Google Scholar] [CrossRef]
  41. Islam, S.Z.; Reed, A.; Kim, D.Y.; Rankin, S.E. N2/Ar plasma induced doping of ordered mesoporous TiO2 thin films for visible light active photocatalysis. Microporous Mesoporous Mater. 2016, 220, 120–128. [Google Scholar] [CrossRef] [Green Version]
  42. 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] [PubMed]
  43. Wang, Q.; Hisatomi, T.; Jia, Q.; Tokudome, H.; Zhong, M.; Wang, C.; Pan, Z.; Takata, T.; Nakabayashi, M.; Shibata, N. Scalable water splitting on particulate photocatalyst sheets with a solar-to-hydrogen energy conversion efficiency exceeding 1%. Nat. Mater. 2016, 15, 611–615. [Google Scholar] [CrossRef] [PubMed]
  44. Iwase, A.; Ng, Y.H.; Ishiguro, Y.; Kudo, A.; Amal, R. Reduced graphene oxide as a solid-state electron mediator in Z-scheme photocatalytic water splitting under visible light. J. Am. Chem. Soc. 2011, 133, 11054–11057. [Google Scholar] [CrossRef] [PubMed]
  45. Sayama, K.; Mukasa, K.; Abe, R.; Abe, Y.; Arakawa, H. A new photocatalytic water splitting system under visible light irradiation mimicking a Z-scheme mechanism in photosynthesis. J. Photochem. Photobiol. A Chem. 2002, 148, 71–77. [Google Scholar] [CrossRef]
  46. Matsumoto, Y.; Unal, U.; Tanaka, N.; Kudo, A.; Kato, H. Electrochemical approach to evaluate the mechanism of photocatalytic water splitting on oxide photocatalysts. J. Solid State Chem. 2004, 177, 4205–4212. [Google Scholar] [CrossRef]
  47. Afroz, K.; Moniruddin, M.; Bakranov, N.; Kudaibergenov, S.; Nuraje, N. A heterojunction strategy to improve the visible light sensitive water splitting performance of photocatalytic materials. J. Mater. Chem. A 2018, 6, 21696–21718. [Google Scholar] [CrossRef]
  48. Wang, Z.; Yang, G.; Tan, C.K.; Nguyen, T.D.; Tok, A.I.Y. Amorphous TiO2 coated hierarchical WO3 Nanosheet/CdS Nanorod arrays for improved photoelectrochemical performance. Appl. Surface Sc. 2019, 490, 411–419. [Google Scholar] [CrossRef]
  49. Hill, J.C.; Choi, K.-S. Effect of Electrolytes on the Selectivity and Stability of n-type WO3 Photoelectrodes for Use in Solar Water Oxidation. J. Phys. Chem. C 2012, 116, 7612–7620. [Google Scholar] [CrossRef]
  50. Maeda, K.; Domen, K. Photocatalytic water splitting: Recent progress and future challenges. J. Phys. Chem. Lett. 2010, 1, 2655–2661. [Google Scholar] [CrossRef]
  51. Bard, A.J. Photoelectrochemistry and heterogeneous photo-catalysis at semiconductors. J. Photochem. 1979, 10, 59–75. [Google Scholar] [CrossRef]
  52. Kato, H.; Sasaki, Y.; Iwase, A.; Kudo, A. Role of iron ion electron mediator on photocatalytic overall water splitting under visible light irradiation using Z-scheme systems. Bull. Chem. Soc. Jpn. 2007, 80, 2457–2464. [Google Scholar] [CrossRef]
  53. Miseki, Y.; Fujiyoshi, S.; Gunji, T.; Sayama, K. Photocatalytic water splitting under visible light utilizing I3−/I− and IO3−/I− redox mediators by Z-scheme system using surface treated PtO x/WO3 as O2 evolution photocatalyst. Catal. Sci. Technol. 2013, 3, 1750–1756. [Google Scholar] [CrossRef]
  54. Yu, W.; Chen, J.; Shang, T.; Chen, L.; Gu, L.; Peng, T. Direct Z-scheme g-C3N4/WO3 photocatalyst with atomically defined junction for H2 production. Appl. Catal. B Environ. 2017, 219, 693–704. [Google Scholar] [CrossRef]
  55. He, K.; Xie, J.; Luo, X.; Wen, J.; Ma, S.; Li, X.; Fang, Y.; Zhang, X. Enhanced visible light photocatalytic H2 production over Z-scheme g-C3N4 nansheets/WO3 nanorods nanocomposites loaded with Ni (OH)x cocatalysts. Chin. J. Catal. 2017, 38, 240–252. [Google Scholar] [CrossRef]
  56. Chen, Y.; Wang, L.; Gao, R.; Zhang, Y.-C.; Pan, L.; Huang, C.; Liu, K.; Chang, X.-Y.; Zhang, X.; Zou, J.-J. Polarization-Enhanced direct Z-scheme ZnO-WO3-x nanorod arrays for efficient piezoelectric-photoelectrochemical Water splitting. Appl. Catal. B Environ. 2019, 259, 118079. [Google Scholar] [CrossRef]
  57. Sayama, K.; Mukasa, K.; Abe, R.; Abe, Y.; Arakawa, H. Stoichiometric water splitting into H2 and O2 using a mixture of two different photocatalysts and an IO3−/I− shuttle redox mediator under visible light irradiation. Chem. Commun. 2001, 2416–2417. [Google Scholar] [CrossRef]
  58. Wang, F.; Di Valentin, C.; Pacchioni, G. Doping of WO3 for photocatalytic water splitting: Hints from density functional theory. J. Phys. Chem. C 2012, 116, 8901–8909. [Google Scholar] [CrossRef]
  59. Hwang, D.W.; Kim, J.; Park, T.J.; Lee, J.S. Mg-doped WO3 as a novel photocatalyst for visible light-induced water splitting. Catal. Lett. 2002, 80, 53–57. [Google Scholar] [CrossRef]
  60. Higashimoto, S.; Ushiroda, Y.; Azuma, M. Electrochemically assisted photocatalysis of hybrid WO3/TiO2 films: Effect of the WO3 structures on charge separation behavior. Top. Catal. 2008, 47, 148–154. [Google Scholar] [CrossRef]
  61. Ke, D.; Liu, H.; Peng, T.; Liu, X.; Dai, K. Preparation and photocatalytic activity of WO3/TiO2 nanocomposite particles. Mater. Lett. 2008, 62, 447–450. [Google Scholar] [CrossRef]
  62. Cui, X.F.; Wang, Y.J.; Jiang, G.Y.; Zhao, Z.; Xu, C.M.; Wei, Y.C.; Duan, A.J.; Liu, J.; Gao, J.S. A photonic crystal-based CdS-Au-WO3 heterostructure for efficient visible-light photocatalytic hydrogen and oxygen evolution dagger. Rsc Adv. 2014, 4, 15689–15694. [Google Scholar] [CrossRef]
  63. Chen, B.; Zhang, Z.; Baek, M.; Kim, S.; Kim, W.; Yong, K. An antenna/spacer/reflector based Au/BiVO4/WO3/Au nanopatterned photoanode for plasmon-enhanced photoelectrochemical water splitting. Appl. Catal. B Environ. 2018, 237, 763–771. [Google Scholar] [CrossRef]
  64. Chatchai, P.; Murakami, Y.; Kishioka, S.-Y.; Nosaka, A.Y.; Nosaka, Y. Efficient photocatalytic activity of water oxidation over WO3/BiVO4 composite under visible light irradiation. Electrochim. Acta 2009, 54, 1147–1152. [Google Scholar] [CrossRef]
  65. Lee, M.G.; Kim, D.H.; Sohn, W.; Moon, C.W.; Park, H.; Lee, S.; Jang, H.W. Conformally coated BiVO4 nanodots on porosity-controlled WO3 nanorods as highly efficient type II heterojunction photoanodes for water oxidation. Nano Energy 2016, 28, 250–260. [Google Scholar] [CrossRef]
  66. Pihosh, Y.; Turkevych, I.; Mawatari, K.; Asai, T.; Hisatomi, T.; Uemura, J.; Tosa, M.; Shimamura, K.; Kubota, J.; Domen, K. Nanostructured WO3/BiVO4 photoanodes for efficient photoelectrochemical water splitting. Small 2014, 10, 3692–3699. [Google Scholar] [CrossRef]
  67. Müller, A.; Kondofersky, I.; Folger, A.; Fattakhova-Rohlfing, D.; Bein, T.; Scheu, C. Dual absorber Fe2O3/WO3 host-guest architectures for improved charge generation and transfer in photoelectrochemical applications. Mater. Res. Express 2017, 4, 016409. [Google Scholar] [CrossRef] [Green Version]
  68. Feng, X.; Chen, Y.; Qin, Z.; Wang, M.; Guo, L. Facile fabrication of sandwich structured WO3 nanoplate arrays for efficient photoelectrochemical water splitting. ACS Appl. Mater. Interfaces 2016, 8, 18089–18096. [Google Scholar] [CrossRef]
  69. Kalanur, S.S.; Hwang, Y.J.; Chae, S.Y.; Joo, O.S. Facile growth of aligned WO3 nanorods on FTO substrate for enhanced photoanodic water oxidation activity. J. Mater. Chem. A 2013, 1, 3479–3488. [Google Scholar] [CrossRef]
  70. Fan, X.; Gao, B.; Wang, T.; Huang, X.; Gong, H.; Xue, H.; Guo, H.; Song, L.; Xia, W.; He, J. Layered double hydroxide modified WO3 nanorod arrays for enhanced photoelectrochemical water splitting. Appl. Catal. A General 2016, 528, 52–58. [Google Scholar] [CrossRef]
  71. Chae, S.Y.; Lee, C.S.; Jung, H.; Joo, O.-S.; Min, B.K.; Kim, J.H.; Hwang, Y.J. Insight into charge separation in WO3/BiVO4 heterojunction for solar water splitting. ACS Appl. Mater. Interfaces 2017, 9, 19780–19790. [Google Scholar] [CrossRef]
  72. Su, J.; Zhang, T.; Wang, L. Engineered WO3 nanorods for conformal growth of WO3/BiVO4 core–shell heterojunction towards efficient photoelectrochemical water oxidation. J. Mater. Sci. Mater. Electron. 2017, 28, 4481–4491. [Google Scholar] [CrossRef]
  73. Yang, J.; Li, W.; Li, J.; Sun, D.; Chen, Q. Hydrothermal synthesis and photoelectrochemical properties of vertically aligned tungsten trioxide (hydrate) plate-like arrays fabricated directly on FTO substrates. J. Mater. Chem. 2012, 22, 17744–17752. [Google Scholar] [CrossRef]
  74. Amano, F.; Li, D.; Ohtani, B. Fabrication and photoelectrochemical property of tungsten (VI) oxide films with a flake-wall structure. Chem. Commun. 2010, 46, 2769–2771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Zeng, Q.; Li, J.; Bai, J.; Li, X.; Xia, L.; Zhou, B. Preparation of vertically aligned WO3 nanoplate array films based on peroxotungstate reduction reaction and their excellent photoelectrocatalytic performance. Appl. Catal. B Environ. 2017, 202, 388–396. [Google Scholar] [CrossRef]
  76. Nayak, A.K.; Sohn, Y.; Pradhan, D. Facile green synthesis of WO3 H2O nanoplates and WO3 nanowires with enhanced photoelectrochemical performance. Cryst. Growth Design 2017, 17, 4949–4957. [Google Scholar] [CrossRef]
  77. Park, M.; Seo, J.H.; Song, H.; Nam, K.M. Enhanced visible light activity of single-crystalline WO3 microplates for photoelectrochemical water oxidation. J. Phys. Chem. C 2016, 120, 9192–9199. [Google Scholar] [CrossRef]
  78. Yang, H.G.; Sun, C.H.; Qiao, S.Z.; Zou, J.; Liu, G.; Smith, S.C.; Cheng, H.M.; Lu, G.Q. Anatase TiO2 single crystals with a large percentage of reactive facets. Nature 2008, 453, 638–641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Wang, X.; Liu, G.; Wang, L.; Pan, J.; Lu, G.Q.M.; Cheng, H.-M. TiO2 films with oriented anatase {001} facets and their photoelectrochemical behavior as CdS nanoparticle sensitized photoanodes. J. Mater. Chem. 2011, 21, 869–873. [Google Scholar] [CrossRef]
  80. Zou, J.-P.; Wu, D.-D.; Luo, J.; Xing, Q.-J.; Luo, X.-B.; Dong, W.-H.; Luo, S.-L.; Du, H.-M.; Suib, S.L. A strategy for one-pot conversion of organic pollutants into useful hydrocarbons through coupling photodegradation of MB with photoreduction of CO2. Acs Catal. 2016, 6, 6861–6867. [Google Scholar] [CrossRef]
  81. Zhang, D.; Wang, S.; Zhu, J.; Li, H.; Lu, Y. WO3 nanocrystals with tunable percentage of (0 0 1)-facet exposure. Appl. Catal. B Environ. 2012, 123, 398–404. [Google Scholar] [CrossRef]
  82. Wang, S.; Chen, H.; Gao, G.; Butburee, T.; Lyu, M.; Thaweesak, S.; Yun, J.-H.; Du, A.; Liu, G.; Wang, L. Synergistic crystal facet engineering and structural control of WO3 films exhibiting unprecedented photoelectrochemical performance. Nano Energy 2016, 24, 94–102. [Google Scholar] [CrossRef] [Green Version]
  83. Zhang, J.; Zhang, P.; Wang, T.; Gong, J. Monoclinic WO3 nanomultilayers with preferentially exposed (002) facets for photoelectrochemical water splitting. Nano Energy 2015, 11, 189–195. [Google Scholar] [CrossRef]
  84. Dong, P.; Hou, G.; Xi, X.; Shao, R.; Dong, F. WO 3-based photocatalysts: Morphology control, activity enhancement and multifunctional applications. Environ. Sci. Nano 2017, 4, 539–557. [Google Scholar] [CrossRef]
  85. Zheng, J.Y.; Song, G.; Hong, J.; Van, T.K.; Pawar, A.U.; Kim, D.Y.; Kim, C.W.; Haider, Z.; Kang, Y.S. Facile fabrication of WO3 nanoplates thin films with dominant crystal facet of (002) for water splitting. Cryst. Growth Des. 2014, 14, 6057–6066. [Google Scholar] [CrossRef]
  86. Navarro, J.R.; Mayence, A.; Andrade, J.; Lerouge, F.; Chaput, F.; Oleynikov, P.; Bergstrom, L.; Parola, S.; Pawlicka, A. WO3 nanorods created by self-assembly of highly crystalline nanowires under hydrothermal conditions. Langmuir 2014, 30, 10487–10492. [Google Scholar] [CrossRef]
  87. Su, J.; Feng, X.; Sloppy, J.D.; Guo, L.; Grimes, C.A. Vertically aligned WO3 nanowire arrays grown directly on transparent conducting oxide coated glass: Synthesis and photoelectrochemical properties. Nano Lett. 2011, 11, 203–208. [Google Scholar] [CrossRef]
  88. Rao, P.M.; Cho, I.S.; Zheng, X. Flame synthesis of WO3 nanotubes and nanowires for efficient photoelectrochemical water-splitting. Proc. Combust. Inst. 2013, 34, 2187–2195. [Google Scholar] [CrossRef]
  89. Koo, W.-T.; Choi, S.-J.; Kim, N.-H.; Jang, J.-S.; Kim, I.-D. Catalyst-decorated hollow WO3 nanotubes using layer-by-layer self-assembly on polymeric nanofiber templates and their application in exhaled breath sensor. Sens. Actuators B Chem. 2016, 223, 301–310. [Google Scholar] [CrossRef]
  90. Chen, D.; Hou, X.; Wen, H.; Wang, Y.; Wang, H.; Li, X.; Zhang, R.; Lu, H.; Xu, H.; Guan, S. The enhanced alcohol-sensing response of ultrathin WO3 nanoplates. Nanotechnology 2009, 21, 035501. [Google Scholar] [CrossRef] [Green Version]
  91. Enferadi-Kerenkan, A.; Ello, A.S.; Do, T.-O. Synthesis, organo-functionalization, and catalytic properties of tungsten oxide nanoparticles as heterogeneous catalyst for oxidative cleavage of oleic acid as a model fatty acid into diacids. Ind. Eng. Chem. Res. 2017, 56, 10639–10647. [Google Scholar] [CrossRef]
  92. Abdullin, K.A.; Kalkozova, Z.K.; Markhabayeva, A.A.; Dupre, R.; Moniruddin, M.; Nuraje, N. Core–Shell (W@ WO3) Nanostructure to Improve Electrochemical Performance. ACS Appl. Energy Mater. 2018, 2, 797–803. [Google Scholar] [CrossRef]
  93. Ma, J.; Zhang, J.; Wang, S.; Wang, T.; Lian, J.; Duan, X.; Zheng, W. Topochemical preparation of WO3 nanoplates through precursor H2WO4 and their gas-sensing performances. J. Phys. Chem. C 2011, 115, 18157–18163. [Google Scholar] [CrossRef]
  94. Meng, D.; Wang, G.; San, X.; Song, Y.; Shen, Y.; Zhang, Y.; Wang, K.; Meng, F. Synthesis of WO3 flower-like hierarchical architectures and their sensing properties. J. Alloys Compd. 2015, 649, 731–738. [Google Scholar] [CrossRef]
  95. Adhikari, S.; Sarkar, D. Hydrothermal synthesis and electrochromism of WO3 nanocuboids. RSC Adv. 2014, 4, 20145–20153. [Google Scholar] [CrossRef]
  96. Ham, D.J.; Phuruangrat, A.; Thongtem, S.; Lee, J.S. Hydrothermal synthesis of monoclinic WO3 nanoplates and nanorods used as an electrocatalyst for hydrogen evolution reactions from water. Chem. Eng. J. 2010, 165, 365–369. [Google Scholar] [CrossRef]
  97. Jiao, Z.; Wang, J.; Ke, L.; Sun, X.W.; Demir, H.V. Morphology-tailored synthesis of tungsten trioxide (hydrate) thin films and their photocatalytic properties. ACS Appl. Mater. Interfaces 2011, 3, 229–236. [Google Scholar] [CrossRef] [Green Version]
  98. Nagy, D.; Nagy, D.; Szilágyi, I.M.; Fan, X. Effect of the morphology and phases of WO 3 nanocrystals on their photocatalytic efficiency. RSC Adv. 2016, 6, 33743–33754. [Google Scholar] [CrossRef] [Green Version]
  99. Sieb, N.R.; Wu, N.-c.; Majidi, E.; Kukreja, R.; Branda, N.R.; Gates, B.D. Hollow metal nanorods with tunable dimensions, porosity, and photonic properties. Acs Nano 2009, 3, 1365–1372. [Google Scholar] [CrossRef]
  100. Lu, X.; Wang, G.; Zhai, T.; Yu, M.; Gan, J.; Tong, Y.; Li, Y. Hydrogenated TiO2 nanotube arrays for supercapacitors. Nano Lett. 2012, 12, 1690–1696. [Google Scholar] [CrossRef]
  101. Kalanoor, B.S.; Seo, H.; Kalanur, S.S. Recent developments in photoelectrochemical water-splitting using WO3/BiVO4 heterojunction photoanode: A review. Mater. Sci. Energy Technol. 2018, 1, 49–62. [Google Scholar]
  102. Pihosh, Y.; Turkevych, I.; Mawatari, K.; Uemura, J.; Kazoe, Y.; Kosar, S.; Makita, K.; Sugaya, T.; Matsui, T.; Fujita, D.; et al. Photocatalytic generation of hydrogen by core-shell WO3/BiVO4 nanorods with ultimate water splitting efficiency. Sci. Rep. 2015, 5, 11141. [Google Scholar] [CrossRef] [Green Version]
  103. Hammad, A.; El-Bery, H.M.; El-Shazly, A.; Elkady, M. Effect of WO3 morphological structure on its photoelectrochemical properties. Int. J. Electrochem. Sci 2018, 13, 362–372. [Google Scholar] [CrossRef]
  104. Zhou, J.; Lin, S.; Chen, Y.; Gaskov, A. Facile morphology control of WO3 nanostructure arrays with enhanced photoelectrochemical performance. Appl. Surface Sci. 2017, 403, 274–281. [Google Scholar] [CrossRef]
  105. Perera, D.; Lorek, R.; Khnayzer, R.S.; Moroz, P.; O’Connor, T.; Khon, D.; Diederich, G.; Kinder, E.; Lambright, S.; Castellano, F.N. Photocatalytic activity of core/shell semiconductor nanocrystals featuring spatial separation of charges. J. Phys. Chem. C 2012, 116, 22786–22793. [Google Scholar] [CrossRef] [Green Version]
  106. Rao, P.M.; Cai, L.; Liu, C.; Cho, I.S.; Lee, C.H.; Weisse, J.M.; Yang, P.; Zheng, X. Simultaneously efficient light absorption and charge separation in WO3/BiVO4 core/shell nanowire photoanode for photoelectrochemical water oxidation. Nano Lett. 2014, 14, 1099–1105. [Google Scholar] [CrossRef]
  107. Jin, B.; Jung, E.; Ma, M.; Kim, S.; Zhang, K.; Kim, J.I.; Son, Y.; Park, J.H. Solution-processed yolk–shell-shaped WO3/BiVO4 heterojunction photoelectrodes for efficient solar water splitting. J. Mater. Chem. A 2018, 6, 2585–2592. [Google Scholar] [CrossRef]
  108. Li, Z.; Zheng, G.; Wang, J.; Li, H.; Wu, J.; Du, Y. Refining waste hardmetals into tungsten oxide nanosheets via facile method. J. Nanoparticle Res. 2016, 18, 98. [Google Scholar] [CrossRef]
  109. Tan, Y.; Wang, H.; Liu, P.; Shen, Y.; Cheng, C.; Hirata, A.; Fujita, T.; Tang, Z.; Chen, M. Versatile nanoporous bimetallic phosphides towards electrochemical water splitting. Energy Environ. Sci. 2016, 9, 2257–2261. [Google Scholar] [CrossRef]
  110. Kim, H.; Hwang, D.; Kim, Y.; Lee, J. Highly donor-doped (110) layered perovskite materials as novel photocatalysts for overall water splitting. Chem. Commun. 1999, 1077–1078. [Google Scholar] [CrossRef] [Green Version]
  111. Lee, C.Y.; Wang, L.; Kado, Y.; Killian, M.S.; Schmuki, P. Anodic nanotubular/porous hematite photoanode for solar water splitting: Substantial effect of iron substrate purity. ChemSusChem 2014, 7, 934–940. [Google Scholar] [CrossRef]
  112. Troitskaia, I.; Gavrilova, T.; Atuchin, V. Structure and micromorphology of titanium dioxide nanoporous microspheres formed in water solution. Phys. Procedia 2012, 23, 65–68. [Google Scholar] [CrossRef] [Green Version]
  113. Fan, X.; Liu, Y.; Chen, S.; Shi, J.; Wang, J.; Fan, A.; Zan, W.; Li, S.; Goddard, W.A.; Zhang, X.-M. Defect-enriched iron fluoride-oxide nanoporous thin films bifunctional catalyst for water splitting. Nat. Commun. 2018, 9, 1–11. [Google Scholar] [CrossRef] [PubMed]
  114. Shin, S.; Han, H.S.; Kim, J.S.; Park, I.J.; Lee, M.H.; Hong, K.S.; Cho, I.S. A tree-like nanoporous WO3 photoanode with enhanced charge transport efficiency for photoelectrochemical water oxidation. J. Mater. Chem. A 2015, 3, 12920–12926. [Google Scholar] [CrossRef]
  115. Fujimoto, I.; Wang, N.; Saito, R.; Miseki, Y.; Gunji, T.; Sayama, K. WO3/BiVO4 composite photoelectrode prepared by improved auto-combustion method for highly efficient water splitting. Int. J. Hydrogen Energy 2014, 39, 2454–2461. [Google Scholar] [CrossRef]
  116. Choi, K.-S. Shape effect and shape control of polycrystalline semiconductor electrodes for use in photoelectrochemical cells. J. Phys. Chem. Lett. 2010, 1, 2244–2250. [Google Scholar] [CrossRef]
  117. Song, K.; Gao, F.; Yang, W.; Wang, E.; Wang, Z.; Hou, H. WO3 mesoporous nanobelts towards efficient photoelectrocatalysts for water splitting. ChemElectroChem 2018, 5, 322–327. [Google Scholar] [CrossRef]
  118. Markhabayeva, A.A.; Moniruddin, M.; Dupre, R.; Abdullin, K.A.; Nuraje, N. Designing of WO3@ Co3O4 Heterostructure to Enhance Photoelectrochemical Performances. J. Phys. Chem. A 2020, 124, 486–491. [Google Scholar] [CrossRef]
  119. Kronawitter, C.X.; Vayssieres, L.; Shen, S.; Guo, L.; Wheeler, D.A.; Zhang, J.Z.; Antoun, B.R.; Mao, S.S. A perspective on solar-driven water splitting with all-oxide hetero-nanostructures. Energy Environ. Sci. 2011, 4, 3889–3899. [Google Scholar] [CrossRef]
  120. Moniruddin, M.; Afroz, K.; Shabdan, Y.; Bizri, B.; Nuraje, N. Hierarchically 3D assembled strontium titanate nanomaterials for water splitting application. Appl. Surface Sci. 2017, 419, 886–892. [Google Scholar] [CrossRef]
  121. Shabdan, Y.; Ronasi, A.; Coulibaly, P.; Moniruddin, M.; Nuraje, N. Engineered core-shell nanofibers for electron transport study in dye-sensitized solar cells. AIP Adv. 2017, 7, 065008. [Google Scholar] [CrossRef] [Green Version]
  122. Abdullin, K.A.; Azatkaliev, A.; Gabdullin, M.; Kalkozova, Z.K.; Mukashev, B.; Serikkanov, A. Preparation of Nanosized Tungsten and Tungsten Oxide Powders. Phys. Solid State 2018, 60, 2634–2639. [Google Scholar] [CrossRef]
  123. Lisitsyna, L.; Denisov, G.; Dauletbekova, A.; Karipbayev, Z.T.; Markhabaeva, A.; Vaganov, V.; Lisitsyn, V.M.; Akilbekov, A. Luminescence of LiF crystals doped with uranium. Proc. J. Phys. Conf. Ser. 2017, 830, 012156. [Google Scholar] [CrossRef]
  124. Moniruddin, M.; Oppong, E.; Stewart, D.; McCleese, C.; Roy, A.; Warzywoda, J.; Nuraje, N. Designing CdS-Based Ternary Heterostructures Consisting of Co-Metal and CoOx Cocatalysts for Photocatalytic H2 Evolution under Visible Light. Inorg. Chem. 2019, 58, 12325–12333. [Google Scholar] [CrossRef] [PubMed]
  125. De Tacconi, N.R.; Chenthamarakshan, C.; Rajeshwar, K.; Pauporté, T.; Lincot, D. Pulsed electrodeposition of WO3–TiO2 composite films. Electrochem. Commun. 2003, 5, 220–224. [Google Scholar] [CrossRef]
  126. Smith, W.; Wolcott, A.; Fitzmorris, R.C.; Zhang, J.Z.; Zhao, Y. Quasi-core-shell TiO2/WO3 and WO3/TiO2 nanorod arrays fabricated by glancing angle deposition for solar water splitting. J. Mater. Chem. 2011, 21, 10792–10800. [Google Scholar] [CrossRef]
  127. Bae, S.W.; Ji, S.M.; Hong, S.J.; Jang, J.W.; Lee, J.S. Photocatalytic overall water splitting with dual-bed system under visible light irradiation. Int. J. Hydrogen Energy 2009, 34, 3243–3249. [Google Scholar] [CrossRef]
  128. Ohno, T.; Tanigawa, F.; Fujihara, K.; Izumi, S.; Matsumura, M. Photocatalytic oxidation of water on TiO2-coated WO3 particles by visible light using Iron (III) ions as electron acceptor. J. Photochem. Photobiol. A Chem. 1998, 118, 41–44. [Google Scholar] [CrossRef]
  129. Yang, M.; He, H.; Zhang, H.; Zhong, X.; Dong, F.; Ke, G.; Chen, Y.; Du, J.; Zhou, Y. Enhanced photoelectrochemical water oxidation on WO3 nanoflake films by coupling with amorphous TiO2. Electrochim. Acta 2018, 283, 871–881. [Google Scholar] [CrossRef]
  130. Momeni, M.M.; Ghayeb, Y.; Davarzadeh, M. Single-step electrochemical anodization for synthesis of hierarchical WO3–TiO2 nanotube arrays on titanium foil as a good photoanode for water splitting with visible light. J. Electroanal. Chem. 2015, 739, 149–155. [Google Scholar] [CrossRef]
  131. Lai, C.W.; Sreekantan, S. Preparation of hybrid WO3–TiO2 nanotube photoelectrodes using anodization and wet impregnation: Improved water-splitting hydrogen generation performance. Int. J. Hydrogen Energy 2013, 38, 2156–2166. [Google Scholar] [CrossRef]
  132. Kwon, Y.T.; Song, K.Y.; Lee, W.I.; Choi, G.J.; Do, Y.R. Photocatalytic behavior of WO3-loaded TiO2 in an oxidation reaction. J. Catal. 2000, 191, 192–199. [Google Scholar] [CrossRef]
  133. Sajjad, A.K.L.; Shamaila, S.; Tian, B.; Chen, F.; Zhang, J. Comparative studies of operational parameters of degradation of azo dyes in visible light by highly efficient WOx/TiO2 photocatalyst. J. Hazard. Mater. 2010, 177, 781–791. [Google Scholar] [CrossRef]
  134. Hong, S.J.; Lee, S.; Jang, J.S.; Lee, J.S. Heterojunction BiVO4/WO3 electrodes for enhanced photoactivity of water oxidation. Energy Environ. Sci. 2011, 4, 1781–1787. [Google Scholar] [CrossRef] [Green Version]
  135. Xia, L.; Bai, J.; Li, J.; Zeng, Q.; Li, X.; Zhou, B. A highly efficient BiVO4/WO3/W heterojunction photoanode for visible-light responsive dual photoelectrode photocatalytic fuel cell. Appl. Catal. B Environ. 2016, 183, 224–230. [Google Scholar] [CrossRef]
  136. Santato, C.; Odziemkowski, M.; Ulmann, M.; Augustynski, J. Crystallographically oriented mesoporous WO3 films: Synthesis, characterization, and applications. J. Am. Chem. Soc. 2001, 123, 10639–10649. [Google Scholar] [CrossRef] [PubMed]
  137. Grigioni, I.; Stamplecoskie, K.G.; Selli, E.; Kamat, P.V. Dynamics of photogenerated charge carriers in WO3/BiVO4 heterojunction photoanodes. J. Phys. Chem. C 2015, 119, 20792–20800. [Google Scholar] [CrossRef]
  138. Pattengale, B.; Ludwig, J.; Huang, J. Atomic insight into the W-doping effect on carrier dynamics and photoelectrochemical properties of BiVO4 photoanodes. J. Phys. Chem. C 2016, 120, 1421–1427. [Google Scholar] [CrossRef]
  139. Mali, M.G.; Yoon, H.; Kim, M.-w.; Swihart, M.T.; Al-Deyab, S.S.; Yoon, S.S. Electrosprayed heterojunction WO3/BiVO4 films with nanotextured pillar structure for enhanced photoelectrochemical water splitting. Appl. Phys. Lett. 2015, 106, 151603. [Google Scholar] [CrossRef]
  140. Cesar, I.; Sivula, K.; Kay, A.; Zboril, R.; Graätzel, M. Influence of feature size, film thickness, and silicon doping on the performance of nanostructured hematite photoanodes for solar water splitting. J. Phys. Chem. C 2008, 113, 772–782. [Google Scholar] [CrossRef]
  141. Sadtler, B.; Demchenko, D.O.; Zheng, H.; Hughes, S.M.; Merkle, M.G.; Dahmen, U.; Wang, L.-W.; Alivisatos, A.P. Selective facet reactivity during cation exchange in cadmium sulfide nanorods. J. Am. Chem. Soc. 2009, 131, 5285–5293. [Google Scholar] [CrossRef] [Green Version]
  142. Li, Y.; Zhang, L.; Liu, R.; Cao, Z.; Sun, X.; Liu, X.; Luo, J. WO3@ α-Fe2O3 Heterojunction arrays with improved photoelectrochemical behavior for neutral ph water splitting. ChemCatChem 2016, 8, 2765–2770. [Google Scholar] [CrossRef]
  143. Luo, W.; Yu, T.; Wang, Y.; Li, Z.; Ye, J.; Zou, Z. Enhanced photocurrent–voltage characteristics of WO3/Fe2O3 nano-electrodes. J. Phys. D Appl. Phys. 2007, 40, 1091. [Google Scholar] [CrossRef]
  144. Li, Y.; Feng, J.; Li, H.; Wei, X.; Wang, R.; Zhou, A. Photoelectrochemical splitting of natural seawater with α-Fe2O3/WO3 nanorod arrays. Int. J. Hydrogen Energy 2016, 41, 4096–4105. [Google Scholar] [CrossRef]
  145. Kumar, P.; Singh, M.; Reddy, G.B. Core-Shell WO3-WS2 Nanostructured Thin Films via Plasma Assisted Sublimation and Sulfurization. ACS Appl. Nano Mater. 2019, 2, 1691–1703. [Google Scholar] [CrossRef]
  146. Zhang, J.; Liu, Z.; Liu, Z. Novel WO3/Sb2S3 heterojunction photocatalyst based on WO3 of different morphologies for enhanced efficiency in photoelectrochemical water splitting. ACS Appl. Mater. Interfaces 2016, 8, 9684–9691. [Google Scholar] [CrossRef]
  147. Wang, Y.; Tian, W.; Chen, L.; Cao, F.; Guo, J.; Li, L. Three-dimensional WO3 nanoplate/Bi2S3 nanorod heterojunction as a highly efficient photoanode for improved photoelectrochemical water splitting. ACS Appl. Mater. Interfaces 2017, 9, 40235–40243. [Google Scholar] [CrossRef]
  148. Wang, P.; Huang, B.; Dai, Y.; Whangbo, M.-H. Plasmonic photocatalysts: Harvesting visible light with noble metal nanoparticles. Phys. Chem. Chem. Phys. 2012, 14, 9813–9825. [Google Scholar] [CrossRef]
  149. Hu, D.; Diao, P.; Xu, D.; Wu, Q. Gold/WO3 nanocomposite photoanodes for plasmonic solar water splitting. Nano Res. 2016, 9, 1735–1751. [Google Scholar] [CrossRef]
  150. Warren, S.C.; Thimsen, E. Plasmonic solar water splitting. Energy Environ. Sci. 2012, 5, 5133–5146. [Google Scholar] [CrossRef]
  151. Thimsen, E.; Le Formal, F.; Gratzel, M.; Warren, S.C. Influence of plasmonic Au nanoparticles on the photoactivity of Fe2O3 electrodes for water splitting. Nano Lett. 2010, 11, 35–43. [Google Scholar] [CrossRef]
  152. Singh, T.; Müller, R.; Singh, J.; Mathur, S. Tailoring surface states in WO3 photoanodes for efficient photoelectrochemical water splitting. Appl. Surface Sci. 2015, 347, 448–453. [Google Scholar] [CrossRef]
  153. Li, Y.Z.; Liu, Z.; Guo, Z.; Ruan, M.; Li, X.; Liu, Y. Efficient WO3 photoanode modified by Pt layer and plasmonic Ag for enhanced charge separation and transfer to promote photoelectrochemical performances. ACS Sustain. Chem. Eng. 2019, 7, 12582–12590. [Google Scholar]
  154. Xu, S.; Fu, D.; Song, K.; Wang, L.; Yang, Z.; Yang, W.; Hou, H. One-dimensional WO3/BiVO4 heterojunction photoanodes for efficient photoelectrochemical water splitting. Chem. Eng. J. 2018, 349, 368–375. [Google Scholar] [CrossRef]
  155. Wu, P.D.; Liu, Z.F.; Ruan, M.N.; Guo, Z.G.; Zhao, L. Cobalt-phosphate modified Fe-Zn0.2Cd0.8S/CuSbS2 heterojunction photoanode with multiple synergistic effect for enhancing photoelectrochemical water splitting. Appl. Surface Sci. 2019, 476, 716–723. [Google Scholar] [CrossRef]
  156. Kim, H.I.; Kim, J.; Kim, W.; Choi, W. Enhanced Photocatalytic and Photoelectrochemical Activity in the Ternary Hybrid of CdS/TiO2/WO3 through the Cascadal Electron Transfer. J. Phys. Chem. C 2011, 115, 9797–9805. [Google Scholar] [CrossRef]
  157. Zeng, Q.Y.; Bai, J.; Li, J.H.; Li, L.S.; Xia, L.G.; Zhou, B.X.; Sun, Y.G. Highly-stable and efficient photocatalytic fuel cell based on an epitaxial TiO2/WO3/W nanothorn photoanode and enhanced radical reactions for simultaneous electricity production and wastewater treatment. Appl. Energy 2018, 220, 127–137. [Google Scholar] [CrossRef]
  158. Liu, C.H.; Qiu, Y.Y.; Zhang, J.; Liang, Q.; Mitsuzaki, N.; Chen, Z.D. Construction of CdS quantum dots modified g-C3N4/ZnO heterostructured photoanode for efficient photoelectrochemical water splitting. J. Photochem. Photobiol. A Chem. 2019, 371, 109–117. [Google Scholar] [CrossRef]
  159. Liew, S.L.; Zhang, Z.; Goh, T.W.G.; Subramanian, G.S.; Seng, H.L.D.; Hor, T.S.A.; Luo, H.K.; Chi, D.Z. Yb-doped WO3 photocatalysts for water oxidation with visible light. Int. J. Hydrogen Energy 2014, 39, 4291–4298. [Google Scholar] [CrossRef]
  160. Subramanyam, P.; Vinodkumar, T.; Nepak, D.; Deepa, M.; Subrahmanyam, C. Mo-doped BiVO4@reduced graphene oxide composite as an efficient photoanode for photoelectrochemical water splitting. Catal. Today 2019, 325, 73–80. [Google Scholar] [CrossRef]
  161. Zhang, K.; Shi, X.J.; Kim, J.K.; Park, J.H. Photoelectrochemical cells with tungsten trioxide/Mo-doped BiVO4 bilayers. Phys. Chem. Chem. Phys. 2012, 14, 11119–11124. [Google Scholar] [CrossRef]
  162. Tian, L.; Yang, X.F.; Cui, X.K.; Liu, Q.Q.; Tang, H. Fabrication of dual direct Z-scheme g-C3N4/MoS2/Ag3PO4 photocatalyst and its oxygen evolution performance. Appl. Surface Sci. 2019, 463, 9–17. [Google Scholar] [CrossRef]
  163. Kim, J.Y.; Youn, D.H.; Kang, K.; Lee, J.S. Highly Conformal Deposition of an Ultrathin FeOOH Layer on a Hematite Nanostructure for Efficient Solar Water Splitting. Angew. Chem. Int. Ed. 2016, 55, 10854–10858. [Google Scholar] [CrossRef] [PubMed]
  164. Lhermitte, C.R.; Verwer, J.G.; Bartlett, B.M. Improving the stability and selectivity for the oxygen-evolution reaction on semiconducting WO3 photoelectrodes with a solid-state FeOOH catalyst. J. Mater. Chem. A 2016, 4, 2960–2968. [Google Scholar] [CrossRef]
  165. Bai, S.L.; Yang, X.J.; Liu, C.Y.; Xiang, X.; Luo, R.X.; He, J.; Chen, A.F. An Integrating Photoanode of WO3/Fe2O3 Heterojunction Decorated with NiFe-LDH to Improve PEC Water Splitting Efficiency. Acs Sustain. Chem. Eng. 2018, 6, 12906–12913. [Google Scholar] [CrossRef]
  166. Davi, M.; Ogutu, G.; Schrader, F.; Rokicinska, A.; Kustrowski, P.; Slabon, A. Enhancing Photoelectrochemical Water Oxidation Efficiency of WO3/alpha-Fe2O3 Heterojunction Photoanodes by Surface Functionalization with CoPd Nanocrystals. Eur. J. Inorg. Chem. 2017, 2017, 4267–4274. [Google Scholar] [CrossRef]
  167. Sun, J.; Sun, L.; Yang, X.; Bai, S.; Luo, R.; Li, D.; Chen, A. Photoanode of coupling semiconductor heterojunction and catalyst for solar PEC water splitting. Electrochim. Acta 2020, 331, 135282. [Google Scholar] [CrossRef]
  168. Tahir, M.; Siraj, M.; Tahir, B.; Umer, M.; Alias, H.; Othman, N. Au-NPs embedded Z–scheme WO3/TiO2 nanocomposite for plasmon-assisted photocatalytic glycerol-water reforming towards enhanced H2 evolution. Appl. Surface Sci. 2020, 503, 144344. [Google Scholar] [CrossRef]
  169. Zhu, Z.F.; Yan, Y.; Li, J.Q. Preparation of flower-like BiOBr-WO3-Bi2WO6 ternary hybrid with enhanced visible-light photocatalytic activity. J. Alloys Compd. 2015, 651, 184–192. [Google Scholar] [CrossRef]
  170. Hou, Y.; Zuo, F.; Dagg, A.P.; Liu, J.K.; Feng, P.Y. Branched WO3 nanosheet array with layered C3N4 heterojunctions and CoOx Nanoparticles as a flexible photoanode for efficient photoelectrochemical water oxidation. Adv. Mater. 2014, 26, 5043–5049. [Google Scholar] [CrossRef]
  171. Yan, H.J.; Tian, C.G.; Wang, L.; Wu, A.P.; Meng, M.C.; Zhao, L.; Fu, H.G. Phosphorus-modified tungsten nitride/reduced graphene oxide as a high-performance, non-noble-metal electrocatalyst for the hydrogen evolution reaction. Angew. Chem. Int. Ed. 2015, 54, 6325–6329. [Google Scholar] [CrossRef]
  172. Liu, C.J.; Yang, Y.H.; Li, W.Z.; Li, J.; Li, Y.M.; Shi, Q.L.; Chen, Q.Y. Highly Efficient Photoelectrochemical Hydrogen Generation Using ZnxBi2S3+x Sensitized Platelike WO3 Photoelectrodes. Acs Appl. Mater. Interfaces 2015, 7, 10763–10770. [Google Scholar] [CrossRef] [PubMed]
  173. Hu, J.S.; Ren, L.L.; Guo, Y.G.; Liang, H.P.; Cao, A.M.; Wan, L.J.; Bai, C.L. Mass production and high photocatalytic activity of ZnS nanoporous nanoparticles. Angew. Chem. Int. Ed. 2005, 44, 1269–1273. [Google Scholar] [CrossRef] [PubMed]
  174. Zhang, X.L.; Wang, X.; Wang, D.F.; Ye, J.H. Conformal BiVO4-Layer/WO3-Nanoplate-Array Heterojunction Photoanode Modified with Cobalt Phosphate Cocatalyst for Significantly Enhanced Photoelectrochemical Performances. Acs Appl. Mater. Interfaces 2019, 11, 5623–5631. [Google Scholar] [CrossRef] [PubMed]
  175. Wang, C.H.; Zhang, X.T.; Yuan, B.; Wang, Y.X.; Sun, P.P.; Wang, D.; Wei, Y.A.; Liu, Y.C. Multi-heterojunction photocatalysts based on WO3 nanorods: Structural design and optimization for enhanced photocatalytic activity under visible light. Chem. Eng. J. 2014, 237, 29–37. [Google Scholar] [CrossRef]
  176. Ran, L.; Yin, L.W. Ternary Hierarchical Cu7S4/TiO2/CoCr-LDH Heterostructured Nanorod Arrays with Multiphase Reaction Interfaces for More Efficient Photoelectrochemical Water Splitting. Adv. Mater. Interfaces 2019, 6, 14. [Google Scholar] [CrossRef]
  177. Song, S.Y.; Zhang, Y.; Xing, Y.; Wang, C.; Feng, J.; Shi, W.D.; Zheng, G.L.; Zhang, H.J. Rectangular AgIn(WO4)(2) nanotubes: A promising photoelectric material. Adv. Funct. Mater. 2008, 18, 2328–2334. [Google Scholar] [CrossRef]
  178. Ma, Z.Z.; Hou, H.L.; Song, K.; Fang, Z.; Wang, L.; Gao, F.M.; Yang, Z.B.; Tang, B.; Yang, W.Y. Ternary WO3/Porous-BiVO4/FeOOH Hierarchical Architectures: Towards Highly Efficient Photoelectrochemical Performance. Chemelectrochem 2018, 5, 3660–3667. [Google Scholar] [CrossRef]
  179. Shi, X.J.; Choi, Y.; Zhang, K.; Kwon, J.; Kim, D.Y.; Lee, J.K.; Oh, S.H.; Kim, J.K.; Park, J.H. Efficient photoelectrochemical hydrogen production from bismuth vanadate-decorated tungsten trioxide helix nanostructures. Nat. Commun. 2014, 5, 8. [Google Scholar] [CrossRef] [Green Version]
  180. 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]
  181. Zhang, L.; Huang, Y.; Dai, C.H.; Liang, Q.M.; Yang, P.; Yang, H.H.; Yan, J.H. Constructing ZnO/ZnCr2O4@TiO2-NTA Nanocomposite for Photovoltaic Conversion and Photocatalytic Hydrogen Evolution. J. Electron. Mater. 2019, 48, 1724–1729. [Google Scholar] [CrossRef]
  182. Cai, J.J.; Li, S.; Qin, G.W. Interface engineering of Co3O4 loaded CaFe2O4/Fe2O3 heterojunction for photoelectrochemical water oxidation. Appl. Surface Sci. 2019, 466, 92–98. [Google Scholar] [CrossRef]
  183. Yu, C.L.; Chen, F.Y.; Zeng, D.B.; Xie, Y.; Zhou, W.Q.; Liu, Z.; Wei, L.F.; Yang, K.; Li, D.H. A facile phase transformation strategy for fabrication of novel Z-scheme ternary heterojunctions with efficient photocatalytic properties. Nanoscale 2019, 11, 7720–7733. [Google Scholar] [CrossRef] [PubMed]
  184. Lv, J.; Zhao, Z.Y.; Li, Z.S.; Ye, J.H.; Zou, Z.G. Preparation and photocatalytic property of LiCr(WO4)(2). J. Alloys Compd. 2009, 485, 346–350. [Google Scholar] [CrossRef]
  185. Coridan, R.H.; Shaner, M.; Wiggenhorn, C.; Brunschwig, B.S.; Lewis, N.S. Electrical and Photoelectrochemical Properties of WO3/Si Tandem Photoelectrodes. J. Phys. Chem. C 2013, 117, 6949–6957. [Google Scholar] [CrossRef] [Green Version]
  186. Wang, H.L.; Turner, J.A. Characterization of Hematite Thin Films for Photoelectrochemical Water Splitting in a Dual Photoelectrode Device. J. Electrochem. Soc. 2010, 157, F173–F178. [Google Scholar] [CrossRef]
  187. Wang, H.L.; Deutsch, T.; Turner, J.A. Direct water splitting under visible light with nanostructured hematite and WO3 photoanodes and a GaInP2 photocathode. J. Electrochem. Soc. 2008, 155, F91–F96. [Google Scholar] [CrossRef] [Green Version]
  188. Fountaine, K.T.; Atwater, H.A. Mesoscale modeling of photoelectrochemical devices: Light absorption and carrier collection in monolithic, tandem, Si vertical bar WO3 microwires. Opt. Express 2014, 22, A1453–A1461. [Google Scholar] [CrossRef] [Green Version]
  189. Xing, Z.; Shen, S.H.; Wang, M.; Ren, F.; Liu, Y.; Zheng, X.D.; Liu, Y.C.; Xiao, X.H.; Wu, W.; Jiang, C.Z. Efficient enhancement of solar-water-splitting by modified “Z-scheme” structural WO3-W-Si photoelectrodes. Appl. Phys. Lett. 2014, 105, 143902. [Google Scholar] [CrossRef]
  190. Sherman, B.D.; Sheridan, M.V.; Wee, K.R.; Marquard, S.L.; Wang, D.G.; Alibabaei, L.; Ashford, D.L.; Meyer, T.J. A Dye-Sensitized Photoelectrochemical Tandem Cell for Light Driven Hydrogen Production from Water. J. Am. Chem. Soc. 2016, 138, 16745–16753. [Google Scholar] [CrossRef]
  191. Shi, X.; Jeong, H.; Oh, S.J.; Ma, M.; Zhang, K.; Kwon, J.; Choi, I.T.; Choi, I.Y.; Kim, H.K.; Kim, J.K.; et al. Unassisted photoelectrochemical water splitting exceeding 7% solar-to-hydrogen conversion efficiency using photon recycling. Nat. Commun. 2016, 7, 11943. [Google Scholar] [CrossRef]
  192. Akple, M.S.; Chimmikuttanda, S.P. A ternary Z-scheme WO3-Pt-CdS composite for improved visible-light photocatalytic H-2 production activity. J. Nanoparticle Res. 2018, 20, 16. [Google Scholar] [CrossRef]
  193. Bhat, S.S.M.; Lee, S.A.; Suh, J.M.; Hong, S.-P.; Jang, H.W. Triple Planar Heterojunction of SnO2/WO3/BiVO4 with Enhanced Photoelectrochemical Performance under Front Illumination. Appl. Sci. Basel 2018, 8, 1765. [Google Scholar] [CrossRef] [Green Version]
  194. Yourey, J.E.; Kurtz, J.B.; Bartlett, B.M. Water Oxidation on a CuWO4-WO3 Composite Electrode in the Presence of Fe(CN)(6) (3-): Toward Solar Z-Scheme Water Splitting at Zero Bias. J. Phys. Chem. C 2012, 116, 3200–3205. [Google Scholar] [CrossRef]
  195. Prasad, U.; Prakash, J.; Gupta, S.K.; Zuniga, J.; Mao, Y.B.; Azeredo, B.; Kannan, A.N.M. Enhanced Photoelectrochemical Water Splitting with Er- and W-Codoped Bismuth Vanadate with WO3 Heterojunction-Based Two-Dimensional Photoelectrode. Acs Appl. Mater. Interfaces 2019, 11, 19029–19039. [Google Scholar] [CrossRef] [PubMed]
  196. Pan, Q.; Zhang, H.; Yang, Y.; Cheng, C. 3D Brochosomes-Like TiO2/WO3/BiVO4 Arrays as Photoanode for Photoelectrochemical Hydrogen Production. Small 2019, 15, 924. [Google Scholar] [CrossRef]
  197. Ma, W.; Wu, X.; Huang, K.; Wang, M.; Fu, R.; Chen, H.; Feng, S. A Co(OH)(x) nanolayer integrated planar WO3/Fe2O3 photoanode for efficient photoelectrochemical water splitting. Sustain. Energy Fuels 2019, 3, 2135–2141. [Google Scholar] [CrossRef]
  198. Salimi, R.; Alvani, A.A.S.; Mei, B.T.; Naseri, N.; Du, S.F.; Mul, G. Ag-Functionalized CuWO4/WO3 nanocomposites for solar water splitting. New J. Chem. 2019, 43, 2196–2203. [Google Scholar] [CrossRef] [Green Version]
  199. Peng, B.; Xia, M.Y.; Li, C.; Yue, C.S.; Diao, P. Network Structured CuWO4/BiVO4/Co-Pi Nanocomposite for Solar Water Splitting. Catalysts 2018, 8, 663. [Google Scholar] [CrossRef] [Green Version]
  200. Leonard, K.C.; Nam, K.M.; Lee, H.C.; Kang, S.H.; Park, H.S.; Bard, A.J. ZnWO4/WO3 Composite for Improving Photoelectrochemical Water Oxidation. J. Phys. Chem. C 2013, 117, 15901–15910. [Google Scholar] [CrossRef]
  201. Scarongella, M.; Gadiyar, C.; Strach, M.; Rimoldi, L.; Loiudice, A.; Buonsanti, R. Assembly of -Cu2V2O7/WO3 heterostructured nanocomposites and the impact of their composition on structure and photoelectrochemical properties. J. Mater. Chem. C 2018, 6, 12062–12069. [Google Scholar] [CrossRef]
  202. Li, K.Z.; Zhang, C.; Liu, A.J.; Chu, D.M.; Zhang, C.Y.; Yang, P.; Du, Y.K.; Huang, J. Mesoporous tungsten oxide modified by nanolayered manganese-calcium oxide as robust photoanode for solar water splitting. J. Colloid Interface Sci. 2018, 516, 145–152. [Google Scholar] [CrossRef] [PubMed]
  203. Li, Y.T.; Liu, Z.F.; Zhang, J.; Guo, Z.G.; Xin, Y.; Zhao, L. 1D/0D WO3/CdS heterojunction photoanodes modified with dual co-catalysts for efficient photoelectrochemical water splitting. J. Alloy. Compd. 2019, 790, 493–501. [Google Scholar] [CrossRef]
  204. Cui, Y.; Pan, L.; Chen, Y.; Afzal, N.; Ullah, S.; Liu, D.Y.; Wang, L.; Zhang, X.W.; Zou, J.J. Defected ZnWO4-decorated WO3 nanorod arrays for efficient photoelectrochemical water splitting. Rsc. Adv. 2019, 9, 5492–5500. [Google Scholar] [CrossRef] [Green Version]
  205. Ma, Z.Z.; Song, K.; Wang, L.; Gao, F.M.; Tang, B.; Hou, H.L.; Yang, W.Y. WO3/BiVO4 Type-II Heterojunction Arrays Decorated with Oxygen-Deficient ZnO Passivation Layer: A Highly Efficient and Stable Photoanode. Acs Appl. Mater. Interfaces 2019, 11, 889–897. [Google Scholar] [CrossRef]
  206. Lv, Y.; Liu, Y.K.; Chen, C.M.; Wang, T.H.; Zhang, M. Octopus tentacles-like WO3/C@CoO as high property and long life-time electrocatalyst for hydrogen evolution reaction. Electrochim. Acta 2018, 281, 1–8. [Google Scholar] [CrossRef]
  207. Yuan, K.P.; Cao, Q.; Li, X.Y.; Chen, H.Y.; Deng, Y.H.; Wang, Y.Y.; Luo, W.; Lu, H.L.; Zhang, D.W. Synthesis of WO3@ZnWO4@ZnO-ZnO hierarchical nanocactus arrays for efficient photoelectrochemical water splitting. Nano Energy 2017, 41, 543–551. [Google Scholar] [CrossRef]
  208. Lin, H.S.; Lin, L.Y. Improving Visible-light Responses and Electric Conductivities by Incorporating Sb2S3 and Reduced Graphene Oxide in a WO3 Nanoplate Array for Photoelectrochemical Water Oxidation. Electrochim. Acta 2017, 252, 235–244. [Google Scholar] [CrossRef]
  209. Jamali, S.; Moshaii, A. Improving photo-stability and charge transport properties of Cu2O/CuO for photo-electrochemical water splitting using alternate layers of WO3 or CuWO4 produced by the same route. Appl. Surface Sci. 2017, 419, 269–276. [Google Scholar] [CrossRef]
  210. Zhang, H.F.; Zhou, W.W.; Yang, Y.P.; Cheng, C.W. 3D WO3/BiVO4/Cobalt Phosphate Composites Inverse Opal Photoanode for Efficient Photoelectrochemical Water Splitting. Small 2017, 13, 8. [Google Scholar] [CrossRef]
  211. Kalanur, S.S.; Yoo, I.H.; Park, J.; Seo, H. Insights into the electronic bands of WO3/BiVO4/TiO2, revealing high solar water splitting efficiency. J. Mater. Chem. A 2017, 5, 1455–1461. [Google Scholar] [CrossRef]
  212. Gao, H.Q.; Zhang, P.; Hu, J.H.; Pan, J.M.; Fan, J.J.; Shao, G.S. One-dimenSional Z-scheme TiO2/WO3/Pt heterostructures for enhanced hydrogen generation. Appl. Surface Sci. 2017, 391, 211–217. [Google Scholar] [CrossRef]
  213. Peng, X.N.; He, C.; Liu, Q.Y.; Wang, X.N.; Wang, H.B.; Zhang, Y.J.; Ma, Q.C.; Zhang, K.; Han, Y.B.; Wang, H. Strategic Surface Modification of TiO2 nanorods by WO3 and TiCl4 for the Enhancement in Oxygen Evolution Reaction. Electrochim. Acta 2016, 222, 1112–1119. [Google Scholar] [CrossRef]
Figure 1. Statistics analysis from Web of Science indicates increase of recent publications in WO3 photocatalytic areas.
Figure 1. Statistics analysis from Web of Science indicates increase of recent publications in WO3 photocatalytic areas.
Nanomaterials 10 01871 g001
Figure 2. (a) Schematic illustration for WO3-particle-based photocatalyst system; (b) principle of photoelectrochemical water-splitting.
Figure 2. (a) Schematic illustration for WO3-particle-based photocatalyst system; (b) principle of photoelectrochemical water-splitting.
Nanomaterials 10 01871 g002
Figure 3. Bandgap positions of typical semiconductors for water-splitting. Reproduced from [42], with permission from MDPI, 2016.
Figure 3. Bandgap positions of typical semiconductors for water-splitting. Reproduced from [42], with permission from MDPI, 2016.
Nanomaterials 10 01871 g003
Figure 4. (a) Graphic illustration of charge transfer for CdS/WO3−x composite; (b) DRS spectra of WO3−x, CdS and CdS/WO3−x−10; (c) hydrogen generation for CdS nanowires, WO3−x and CdS/WO3−x−10 composites in 20 vol% lactic solution under illumination. Reproduced from [31], with permission from Elsevier, 2018.
Figure 4. (a) Graphic illustration of charge transfer for CdS/WO3−x composite; (b) DRS spectra of WO3−x, CdS and CdS/WO3−x−10; (c) hydrogen generation for CdS nanowires, WO3−x and CdS/WO3−x−10 composites in 20 vol% lactic solution under illumination. Reproduced from [31], with permission from Elsevier, 2018.
Nanomaterials 10 01871 g004
Figure 5. Schematic illustration of photocatalytic water-splitting by Z-system. C.B.—conduction band; V.B.—valence band; Eg—bandgap. Reproduced from [50], with permission from American Chemical Society, 2010.
Figure 5. Schematic illustration of photocatalytic water-splitting by Z-system. C.B.—conduction band; V.B.—valence band; Eg—bandgap. Reproduced from [50], with permission from American Chemical Society, 2010.
Nanomaterials 10 01871 g005
Figure 6. (a) Crystal-unit cells for orthorhombic WO3·0.33H2O, hexagonal WO3 and monoclinic WO3; (b) absorption peaks of WO3 films with/without heat treatment; (c) linear-sweep voltammetry of WO3 photoanodes at different temperatures under chopped illumination. Reproduced from [68], with permission from American Chemical Society, 2016.
Figure 6. (a) Crystal-unit cells for orthorhombic WO3·0.33H2O, hexagonal WO3 and monoclinic WO3; (b) absorption peaks of WO3 films with/without heat treatment; (c) linear-sweep voltammetry of WO3 photoanodes at different temperatures under chopped illumination. Reproduced from [68], with permission from American Chemical Society, 2016.
Nanomaterials 10 01871 g006
Figure 7. SEM images. (a) Stacked WO3·H2O square nanoplates synthesized at room temperature, WO3 nanowires evolving after (b) 3 h, (c) 6 h and (d) 12 h solvothermal treatment of WO3·H2O nanoplates at 200 °C; (eh) corresponding XRD patterns. Reproduced from [76], with permission from American Chemical Society, 2017.
Figure 7. SEM images. (a) Stacked WO3·H2O square nanoplates synthesized at room temperature, WO3 nanowires evolving after (b) 3 h, (c) 6 h and (d) 12 h solvothermal treatment of WO3·H2O nanoplates at 200 °C; (eh) corresponding XRD patterns. Reproduced from [76], with permission from American Chemical Society, 2017.
Nanomaterials 10 01871 g007
Figure 8. SEM images for WO3 hydrate film with sheetlike nanostructures with (a) (NH4)2SO4, (b) Na2SO4 and (c) CH3COONH4 as capping agent; (d) absorption spectra of WO3 films; (e) current-potential scans of WO3 films measured in darkness and under illumination in a 1-M H2SO4 electrolyte containing 0.1-M methanol. Reproduced from [97], with permission from American Chemical Society, 2011.
Figure 8. SEM images for WO3 hydrate film with sheetlike nanostructures with (a) (NH4)2SO4, (b) Na2SO4 and (c) CH3COONH4 as capping agent; (d) absorption spectra of WO3 films; (e) current-potential scans of WO3 films measured in darkness and under illumination in a 1-M H2SO4 electrolyte containing 0.1-M methanol. Reproduced from [97], with permission from American Chemical Society, 2011.
Nanomaterials 10 01871 g008
Figure 9. (a,b) Top and cross-section SEM images of the optimum BiVO4/70°-WO3 nanorods; (c) expanded image of BiVO4/WO3 nanorods, (d) enhancement of photocurrent density of BiVO4/WO3 nanorods (70°) nanorods; (e) schematic illustration of BiVO4/WO3 nanorods. Reproduced from [65], with permission from Elsevier, 2016.
Figure 9. (a,b) Top and cross-section SEM images of the optimum BiVO4/70°-WO3 nanorods; (c) expanded image of BiVO4/WO3 nanorods, (d) enhancement of photocurrent density of BiVO4/WO3 nanorods (70°) nanorods; (e) schematic illustration of BiVO4/WO3 nanorods. Reproduced from [65], with permission from Elsevier, 2016.
Nanomaterials 10 01871 g009
Figure 10. (a) SEM image of the WO3 photoanode; (b) TEM image of the nanoporous WO3 clusters, the inset: a high resolution TEM image; (c,d) schematic illustrations of charge transport/transfer processes. Reproduced from [114], with permission from Royal Society of Chemistry, 2015.
Figure 10. (a) SEM image of the WO3 photoanode; (b) TEM image of the nanoporous WO3 clusters, the inset: a high resolution TEM image; (c,d) schematic illustrations of charge transport/transfer processes. Reproduced from [114], with permission from Royal Society of Chemistry, 2015.
Nanomaterials 10 01871 g010
Figure 11. (a) Diagram of the band edge positions of pure WO3 and BiVO4 and (b) for a WO3–BiVO4 composite under solar irradiation. Reproduced from [137], with permission from American Chemical Society, 2015.
Figure 11. (a) Diagram of the band edge positions of pure WO3 and BiVO4 and (b) for a WO3–BiVO4 composite under solar irradiation. Reproduced from [137], with permission from American Chemical Society, 2015.
Nanomaterials 10 01871 g011
Figure 12. (a) WO3–α–Fe2O3 nanocomposites and its band diagram for PEC water-splitting (b) UV-vis absorption spectra of WO3@Fe2O3. Reproduced from ref [142], with permission from John Wiley and Sons, 2016.
Figure 12. (a) WO3–α–Fe2O3 nanocomposites and its band diagram for PEC water-splitting (b) UV-vis absorption spectra of WO3@Fe2O3. Reproduced from ref [142], with permission from John Wiley and Sons, 2016.
Nanomaterials 10 01871 g012
Figure 13. (a) Current -potential curves of WO3@Au composite with HAuCl4 concentrations of 8, 24 and 240 umol in 0.1-M Na2SO4 electrolyte and (b) SEM images of WO3@Au composites. Reproduced from ref [149] with permission from Springer Nature, 2016.
Figure 13. (a) Current -potential curves of WO3@Au composite with HAuCl4 concentrations of 8, 24 and 240 umol in 0.1-M Na2SO4 electrolyte and (b) SEM images of WO3@Au composites. Reproduced from ref [149] with permission from Springer Nature, 2016.
Nanomaterials 10 01871 g013
Figure 14. Schematic representation of a comparison of electronic transition processes in binary and ternary hybrid. Reproduced from [156], with permission from American Chemical Society, 2011.
Figure 14. Schematic representation of a comparison of electronic transition processes in binary and ternary hybrid. Reproduced from [156], with permission from American Chemical Society, 2011.
Nanomaterials 10 01871 g014
Figure 15. Schematic illustration of a PEC tandem system based on BiVO4/WO3 and DSSC. Reproduced from [191], with permission from Springer Nature, 2016.
Figure 15. Schematic illustration of a PEC tandem system based on BiVO4/WO3 and DSSC. Reproduced from [191], with permission from Springer Nature, 2016.
Nanomaterials 10 01871 g015
Table 1. Photocurrent densities of binary heterostructures.
Table 1. Photocurrent densities of binary heterostructures.
PhotoanodesMethodsMorphologyElectrolytePotentialP (mW/cm2)J (mA/cm2)/STH (%)/Gas Evolution (mL/cm2)Ref
WO3/TiO2SolvothermalNanoflake0.1-M Na2SO40.8 vs. SCE
(1.45 vs. RHE)
1001.4[129]
WO3/TiO2AnodizationNanotubes1-M NaOH0.7 vs. Ag/AgCl
(1.7 RHE)
1001.6[130]
WO3/TiO2AnodizationNanotubes1-M KOH0.6 V vs. SCE
(1.62 RHE)
1002
/3.1%/16.2
[131]
WO3−x/ZnOSolvothermal methodNanorods1-M Na2SO41.23 vs. RHE1003.38[56]
WO3/BiVO4Glancing-angle deposition/electrodepositionVertically oriented nanorods0.5-M Na2SO41.23 vs. RHE1003.1[66]
WO3/BiVO4Electrostaticspraying methodNanotextured pillar0.5-M Na2SO40.7 V vs. Ag/AgCl
(1.44 vs. RHE)
1002.1[139]
WO3/BiVO4Layer-by-layerFilm0.5-M Na2SO41.23 vs. Ag/AgCl
(1.817 vs. RHE)
1002.78[134]
WO3/BiVO4Spin coatingFilm0.5-M Na2SO41.23 V vs. Ag/AgCl
1.817 vs. RHE
1001.2[64]
WO3/BiVO4Pulsed electrodepositionNanorods0.1-M Na2SO31.23 vs. RHE1004.55[65]
WO3/BiVO4Anodic oxidationNanoporous film0.1-M KH2PO40.6 V vs. Ag/AgCl
(1.21 vs. RHE)
10002.01[135]
WO3/BiVO4ElectrospinningNanofibers0.5-M Na2SO41.23 vs. RHE1002.8[154]
WO3/Fe2O3SolvothermalNanosheets0.5-M Na2SO41.23 vs. RHE1001.66[142]
WO3/Fe2O3HydrothermalNanorods0.1-M Na2SO41.23 vs. RHE1001[144]
WO3/Fe2O3Sol–gelFilm0.1-M NaOH1.23 vs. RHE1000.7[67]
WO3/Sb2S3HydrothermalNanoplate/nanorods1-M H2SO40.8 vs. RHE1001.79[146]
WO3/Bi2S3CBDNanorods/nanoplates0.1-M Na2SO30.9 vs. RHE1005.95[147]
WO3/AuHydrothermalNanoplate0.1-M Na2SO41.23 vs. RHE1001.5[149]
Table 2. Photocurrent densities of ternary heterostructures.
Table 2. Photocurrent densities of ternary heterostructures.
Photocatalytic MaterialMethodsMorphologyElectrolytePotentialIrradiationPhotocurrent
Density
Ref
WO3–Pt–CdSCombination of wet-chemical, photodeposition and hydrothermal techniquesHollow
microspheres composed of small crystallites
0.5 M
Na2SO4
0.5 V vs. Ag/AgCl
(1.82 V vs. RHE)
Vis light0.16 μA/cm2[192]
SnO2/WO3/BiVO4Combination of electron beam deposition and metal organic decomposition techniquePlanar film0.5-M Na2SO31.23 vs. RHE100 mW/cm22.01 mA/cm2[193]
WO3/C3N 4//CoOxCombination of a hydrothermal method with wet impregnationfilm 1.23 V vs. Ag/AgCl
(1.8 V vs. RHE)
100 mW/cm25.76 mA/cm2[170]
CuWO4−WO3electrodepositionfilm0.1-M KH2PO40.618 V vs. Ag/AgCl
(1.23 vs. RHE)
100 mW/cm20.3 mA/cm2[194]
WO3/(Er, W):BiVO4spray coatingmonoclinic clinobisvanite structure0.1-M K2HPO41.23 V vs. RHE100 mW/cm24.1 ± 0.19 mA cm−2[195]
WO3/(Er, W):BiVO4spray coatingmonoclinic clinobisvanite structure0.1-M K2HPO42.3 V vs. RHE100 mW/cm27.2 ± 0.39 mA cm−2[195]
TiO2/WO3/BiVO4hydrothermalbrochosomes-like0.5-M Na2SO40.35 V vs. RHE100 mW/cm23.13 mA cm−2[196]
WO3/
Fe2O3/Co(OH)
electrospray depositionworm-like nanobars0.1-M NaOH1.23 vs. RHE 0.62 mA cm−2[197]
Ag-functionalized CuWO4/WO3electrophoretic depositionthin filmpotassium phosphate buffer solution0.62 V vs. Ag/AgCl
(1.23 V vs. RHE)
0.205 mA cm−2[198]
CuWO4/BiVO4 with Co-Pidrop-casting and
thermal annealing method
nanoflakes1.0 M of Na2SO4 with 0.1 M of sodium phosphate
buffer (pH = 7)
1.23 V vs. RHE100 mW/cm22.25 mA cm−2[199]
BiVO4/WO3/SnO2 connected with perovskite solar cell tandem deviceSpin-coatingtriple-layer planar filmpH 7 phosphate buffer electrolyte1.23 V vs. RHE100 mW/cm23.1 mA/cm2[26]
ZnWO4/WO3
Piezo-dispensingSpot Arrays0.1-M Na2SO4 at pH 70.7 V vs. Ag/AgCl
(1.31 V vs. RHE)
0.75 mA/cm2[200]
b-Cu2V2O7/WO3Seeded-growth approach 0.1-M sodium borate buffer (pH 8.2) containing 0.1-M Na2SO31.23 V vs. RHE100 mW/cm20.45 mA cm−2[201]
CaMn2O4/WO3Spin-coatingThin film0.5-M Na2SO4 solution (pH 6)1.09 V vs. RHE 1.5 × 10−3 mA cm−2[202]
Pt/WO3/AgHydrothermal method, chemical bath, photoassisted electrodepositionNanorods 100 mW/cm21.13
mA/cm2
[153]
WO3/CdS/NiOOHhydrothermal method, successive ionic layer adsorption and reaction, photo-assisted electrodepositionNanorodsd 0.2-M Na2SO4-0.1-M Na2SO31.23 V vs. RHE 1.5–2 mA/cm2[203]
ZnWO4/WO3hydrothermalNanorods0.5 M
Na2SO4
1.23 V vs. RHE100 mW/cm21.87
mA cm−2
[204]
WO3/BiVO4/ZnOdrop-casting method, atomic layer depositionNanosheets0.5-M Na2SO41.23 V vs. RHE100 mW/cm22.5–3.00 mA cm−2[205]
Au-surface/BiVO4/WO3/Au-bottomhydrothermal, sol–gel spin-coated,Nanospheres0.5
M Na2SO4
1.23 V vs. RHE 1.31 mA/cm2[63]
WO3/C@CoOhydrothermal process
and immersion method
Octopus tentacles-like1.0-M KOH55 mV (vs. RHE) 10 mA cm−2[206]
WO3@ZnWO4@ZnOlayer deposition technique and hydrothermal
process
nanosheetsmixed
aqueous solution of 0.35-M Na2S and 0.25-M NaSO3 (pH = 13.4)
1.23 V vs.
RHE
100 mW/cm2~1.57 mA/cm2[207]
WO3/rGO/Sb2S3chemical bath depositionnanoplates0.5-M Na2SO4 (pH~7)1.23 V vs. RHE 1.20 mA/cm2[208]
Cu2O/CuO/WO3Electrodeposition, spin-coatingThin film 0 V vs. RHE −1.9 mA/cm2[209]
WO3/BiVO4/Co-PiElectrodepositioncomposite inverse opals0.5-M Na2SO41.4 V
versus Ag/AgCl
(0.67 V vs. RHE)
100 mW cm−24.5 mA cm−2[210]
WO3/BiVO4/TiO2Spin-coating, wet chemistryplatelike0.1-M Na2SO41.23 V vs. RHE100 mW/cm2~3.9 mA/cm2[211]
TiO2/WO3/PtElectrospinning techniquefibers0.2-M Na2SO4 15–20×10−3 mA/cm2[212]
TiO2-TiCl4-WO3Hydrothermal
method +
Electrodeposition
nanorodsKOH1.23 V vs
NHE
100 mW/cm23.86 mW/cm2[213]

Share and Cite

MDPI and ACS Style

Shabdan, Y.; Markhabayeva, A.; Bakranov, N.; Nuraje, N. Photoactive Tungsten-Oxide Nanomaterials for Water-Splitting. Nanomaterials 2020, 10, 1871. https://doi.org/10.3390/nano10091871

AMA Style

Shabdan Y, Markhabayeva A, Bakranov N, Nuraje N. Photoactive Tungsten-Oxide Nanomaterials for Water-Splitting. Nanomaterials. 2020; 10(9):1871. https://doi.org/10.3390/nano10091871

Chicago/Turabian Style

Shabdan, Yerkin, Aiymkul Markhabayeva, Nurlan Bakranov, and Nurxat Nuraje. 2020. "Photoactive Tungsten-Oxide Nanomaterials for Water-Splitting" Nanomaterials 10, no. 9: 1871. https://doi.org/10.3390/nano10091871

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