Next Article in Journal
Contamination Particle Behavior of Aerosol Deposited Y2O3 and YF3 Coatings under NF3 Plasma
Next Article in Special Issue
Tunable Perfect Narrow-Band Absorber Based on a Metal-Dielectric-Metal Structure
Previous Article in Journal
Epitaxial Versus Polycrystalline Shape Memory Cu-Al-Ni Thin Films
Previous Article in Special Issue
Ultrathin Al2O3 Coating on LiNi0.8Co0.1Mn0.1O2 Cathode Material for Enhanced Cycleability at Extended Voltage Ranges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Strategies of Anode Materials Design towards Improved Photoelectrochemical Water Splitting Efficiency

1
School of Chemical Engineering, Northwest University, Xi’an 710069, China
2
School of Materials Science and Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore
*
Authors to whom correspondence should be addressed.
Coatings 2019, 9(5), 309; https://doi.org/10.3390/coatings9050309
Submission received: 4 April 2019 / Revised: 5 May 2019 / Accepted: 6 May 2019 / Published: 9 May 2019
(This article belongs to the Special Issue Thin Films for Energy Harvesting, Conversion, and Storage)

Abstract

:
This review presents the latest processes for designing anode materials to improve the efficiency of water photolysis. Based on different contributions towards the solar-to-hydrogen efficiency, we mainly review the strategies to enhance the light absorption, facilitate the charge separation, and enhance the surface charge injection. Although great achievements have been obtained, the challenges faced in the development of anode materials for solar energy to make water splitting remain significant. In this review, the major challenges to improve the conversion efficiency of photoelectrochemical water splitting reactions are presented. We hope that this review helps researchers in or coming to the field to better appreciate the state-of-the-art, and to make a better choice when they embark on new research in photocatalytic water splitting.

1. Introduction

Nowadays, it is becoming the world’s biggest challenge to meet the increasing energy demand by developing new sustainable substitution energy and reducing carbon dioxide emission. Overall, water splitting using sunlight is identified as a promising technology for production of renewable fuels and chemicals without causing environmental pollution.
In a photoelectrochemical cell for solar water-splitting, there are three main elements; namely, an anode, a cathode, and an electrolyte. At the anode, the oxygen evolution reaction (OER) will happen, while through the hydrogen evolution reaction (HER), hydrogen forms at the cathode, as shown in reactions R1 and R2 [1].
R 1 :   2 H 2 O ( l ) 4 e + 4 H + + O 2 ( g )
R 2 :   4 H + ( aq ) + 4 e 2 H 2
The two half-reactions constitute the overall water splitting process (2H2O → 2H2 + O2), with the thermodynamic threshold energy E0 = 1.23 eV. Although both reactions are important for the overall water splitting efficiency, the process is mainly hindered by the hypokinetic four-electron water oxidation. Based on electron transfer process, OER can be divided into four reaction paths with 1.23 eV in the surface site (denoted by the * symbol) of anode materials for thermodynamic threshold energy, as shown in reactions R3 to R6 [2].
R 3 :   H 2 O +     OH + H + + e
R 4 :   OH O + H + + e
R 5 : O + H 2 O OOH + H + + e
R 6 :   OOH O 2 + H + + e
After the Honda–Fujishima effect was discovered in the early 1970s [3], intensive studies were carried out for photoelectrochemical (PEC) water splitting on semiconducting materials. The semiconducting anode materials will greatly determine the efficiency of absorbing light, suitable over-potential, and stability. It is challenging to develop efficient new anode materials or improve the ability of existing anode materials. Consequently, extensive efforts were implemented to improve the efficiency of anode materials; for example, TiO2, α-Fe2O3, WO3, and BiVO4 [4]. However, there is still lack of cheap and stable photoanode materials with sufficient efficiency for water oxidation. Although significant progress has been made by developing new materials, nanostructures, and other novel concepts over the past decades, efficiencies are still below our expectations. Therefore, to assist future development, we wish to summarize the strategies to develop anode materials for improved efficiency in PEC water splitting in a concise manner. It is up to the readers to decide how different strategies could be applied in their specific applications, or to apply them synergistically in some situations.
The solar-to-hydrogen (STH) efficiency is an important metric for benchmarking and performance evaluation. The STH for a PEC system constructed by photoelectrodes can be expressed by [5]:
STH = J P × 1.23 I 0 × 100 %
where J P is the photocurrent observed from the experiment, I 0 is the power intensity of the sunlight. The number 1.23 in the equation is the approximated value of the threshold energy for water splitting which requires the bandgap energy of any possible photoanode lager than 1.23 eV. For a photoelectrode, J P is determined by the photogeneration absorption efficiency (ηabs), the product of the injection efficiency (ηsep), and the injection efficiency (ηinj) of the photogenerated carrier to the reactant [6], which can be described by:
J P = J 0 × η a b s × η s e p × η i n j
where J 0 is the theoretical photocurrent when 100% of the photons in the solar spectrum with energies exceeding the bandgap are absorbed and converted. Theoretical STH and solar photocurrent of a photoelectrode under AM 1.5 G irradiation (100 mW·cm−2) are displayed in Figure 1 [6].
Thus, to achieve efficient PEC water splitting, several key criteria must be met, as shown in Figure 2: (1) wide range of light absorption of a material that can utilize more energy from sunlight; (2) effective generation of the charges and their transportation from the internal electrode to the photoelectric surface; (3) fast reaction of photogenerated carriers on the surface. Therefore, we will summarize recent progress of improving those efficiencies separately.
In this review, we summarize recent progress in the development of anode materials for photoelectrochemical water splitting, which are mainly based on our research work. This is carried out through a review of the strategies to enhance the light absorption, facilitate the charge separation, and enhance the surface charge injection. Finally, we share our critical views on the remaining challenges in this field.

2. Strategies to Enhance the Light Absorption

Generally, light absorption of a photoelectrode depends on its bandgap. For a known photoelectrode with a fixed bandgap, the film thickness and structure would affect its light absorption efficiency. Some reported strategies used to enhance the light absorption are summarized below.

2.1. Nanostructure Formation

Nanostructure is a good way to capture the scattered light if the light is not absorbed immediately after it reaches the surface. Nanostructure formation is a well-known strategy for silicon solar cells to reduce reflection loss and increase the energy conversion efficiency [7]. Similarly, nanostructure construction is proven effective in improving the light capture in PEC electrodes [8]. It has been reported that multiple light scattering in engineered tapered nanostructures enhances light absorption. (Figure 3) [9]. With a nanoporous Mo:BiVO4 layer on the engineering conical nanophotonic structure combined with a solar cell, a STH efficiency of 6.2% was realized. Zhou et al. reported a hierarchical BiVO4 photoelectrode, which consists of small nanoparticles and voids, could induce efficient light harvesting due to multiple light scattering [6]. Kim reported high-efficiency light absorption of GaN truncated nanocones. The truncated nanocones can trap the light within the nanostructure; thus, light loss, which is caused by surface reflection, is reduced. The TiO2 solar cell prepared by using a 1D nanowire array can have a light conversion efficiency of up to 5.02%, which is much higher than that of a simple TiO2 powder [10].

2.2. Band Engineering

To reduce the band gap of semiconductors, one commonly adopted and promising strategy is to add a foreign element (doping). Doping of foreign elements may include adding cation(s), anion(s), or adding cation(s) and anion(s) at same time.
Addition of cations usually induces a shallow valence band in the wide band gap through the hybrid valence orbital [11]. In metal oxide photocatalysts, incorporation of metal ions will introduce a donor level by hybridizing the O 2p valence orbitals with a metal ion having high-lying valence orbitals, such as Bi(III), Ag(I), Sn(II), Pb(II), or Cu(I). The crystal structure and the amount of metal incorporated greatly affect the contribution of the valence band. For example, in a MoO3 electrode, Li+, Na+, and Mg2+ can greatly reduce the gap of the band [12,13]. Thus, modulating the band structure by cation doping is considered to be effective at the atomic level.
Anion doping has been employed to shift the absorption edge from the UV to the visible region; indeed, it is one of the strategies to enhance the light absorption. A lot of attention has been paid to nonmetal doping into TiO2 to tune the light absorption, such as N doping as well as C, F, S, B doping [14]. For example, N and B doping greatly expands the light absorption to 700 nm [15]. Also, O incorporation in MoS2 is able to improve the hybridization between Mo d-orbital and S p-orbital, resulting in a much smaller band gap [16]. Similarly, replacement of O by N in Ta2O5 induces a narrowed bandgap; for example, it was observed that the absorption was extended from 320 to 500 nm by N-doping (Figure 4) [17]. Simultaneously adding cations and anions can greatly change the band structure for enhancing the light absorption; e.g., the band gap of GaN is about 3.4 eV and the band gap of ZnO is 3.2 eV (Figure 5). However, the solid solution of GaN and ZnO narrowed the band gap to around 2.58 eV with visible light response [18]. This strategy provides us with a powerful method for adjusting the bandgap of photoelectrode materials through band engineering.
Another strategy for enlarging the light absorption is by intrinsic doping, such as introducing oxygen vacancy in oxides. The typical example is black TiO2, in which lots of Ti3+ species are generated accompanied with the oxygen vacancy, being self-doping [19,20]. Doping and presence of lattice vacancies introduce midgap states which form the band tails and narrow the band gap. The light absorption of TiO2 was enlarged to around 1000 nm from the 400 nm of white TiO2. However, it is noted that the enlarged light absorption has little contribution to the enhanced performance, because the there is little visible light response for the black TiO2 (Figure 6) [21]. The enhanced performance is possibly from the better conductivity.

2.3. Dual Absorber

Wang et al. reported extraordinary PEC performance by using dual BiVO4 photoanodes made of two BiVO4 which has the same transparency (Figure 7) [22]. The second electrode can use the transmitted light unabsorbed by the first electrode. It was found that the enhanced performance was mainly because of the higher light utilization rate in the airway scope of band edge (400–520 nm). A similar strategy was employed by Kim et al. with a different absorber; BiVO4 in the front, and Fe2O3 behind BiVO4 (Figure 8) [23]. This can further utilize the solar light, because Fe2O3 has a smaller bandgap than BiVO4; thus, the transmitted light can be utilized by Fe2O3. This has increased the efficiency of hydrogen conversion to 7.7%. Wang et al. reported PEC water splitting using a hematite/Si nanowire dual-absorber system, which yielded a lower onset potential [24]. In their work, hematite was coated on the surface of Si nanowire, and the transmitted light through hematite was utilized by Si. The generated electrons of hematite combine with the holes generated on Si. Thus, holes generated on hematite are used for water oxidation, and at the same time, electrons generated on Si are used for reduction reaction. The match of the band edge positions of the two absorbers is important. A double absorber series battery was reported by Sivula et al., in which the light transmitted through the photoelectrode material is utilized by the dye-sensitized solar cells (DSSC) that provides the bias needed by the photoelectrode [25]. This concept is very useful in tandem PEC cells, which do not need additional bias and can increase the solar conversion efficiency. Besides, Chang and Ye also added Co3O4 and carbon quantum dot (CQDs) components on the surface of BiVO4 to enlarger light absorption range [26,27]. For instance, CQDs decorated BiVO4 photoanode cocatalyzed by a bi-layered Ni-FeOOH oxygen evolution catalyst have obtained an outstanding photocurrent density of 5.99 mA·cm−2 at 1.23 V vs. reversible hydrogen electrode (RHE) with an enlarged light absorption range. Noble metal nanoparticles can also enhance the absorption light scattering/trapping, such as Au nanoparticles coated BiVO4 and hematite [28,29]. However, the plasmonic effect also involves direct electron transfer mechanism (DET) and plasmon induced resonance electron transfer (PIRET) to enhance the performance except the light absorption. Currently, the full understanding of the involved physical mechanisms remains elusive, and requires more work to be clarifies [30].

3. Strategies to Improve the Charge Separation

3.1. Doping

Doping is proved to be an effective way to improve the conductivity of semiconducting compounds and thus increase the charge separation efficiency [5]. Zhao et al. reported that extrinsic Ti doping and oxygen vacancy generation (intrinsic doping) have greatly enhanced the PEC performance of hematite (Figure 9) [31]. In another example, incorporation of Mo6+ into a partial site of V5+ in BiVO4 not only has changed the crystal symmetry of BiVO4 and introduced some polarons, but also provided more free carriers (electrons) which facilitate the charge transport [32,33].
Exploration of new materials with suitable band position and narrow bandgap for water splitting is still a vital issue. Since doping is proven as an effective method to ameliorate the efficiency through improved conductivity, this method is usually combined with the design and synthesis of new PEC electrode materials. Recently, the performance of a new promising material, Zn2FeO4, for PEC water splitting was improved by Ti doping (Figure 10) [34]. Lumley et al. reported that the performance of Cu11V6O26 was enhanced after doping by Mo or W [35]. Jo et al. found the 0.5% PO4-doped BiVO4 can lower the charge transfer resistance of BiVO4 remarkably, about 30 times higher than that of BiVO4 before doping [36].
One special doping is gradient doping, which means that the dopant concentration varies from surface to the bulk. In such a case, the band bending not only occurs on the surface, but also in the bulk, which provides an electric field throughout the bulk material. It is easier for the carriers to separate and transport to the surface in a gradient electrode. Yang et al. investigated how gradient doping affects a GaAs photocathode and found that both the escape probability and the diffusion length increased when compared with uniform doping [37]. Abdi et al. synthesized a doped W gradually decreased BiVO4 from 1% to 0% between the interface of the FTO/semiconductor and semiconductor/electrolyte (Figure 11) [38]. Based on their experiment, the efficiency of charge separation increases to ~60%, while the efficiency of charge separation is ~38% for equally doped BiVO4 (Figure 12).

3.2. Nanostructure to Shorten the Diffusion Length

Besides conductivity improvement, the minority carrier’s diffusivity is another key factor influencing the charge separation efficiency. Nanostructure provides a shorter diffusion distance for the minority carrier; therefore, it has been widely applied to improve the charge separation as shown in Figure 12 [39]. Kim et al. found nanoporous morphology of BiVO4 can effectively suppress recombination, which yields an electron-hole separation efficiency as high as 90% at 1.23 V vs. RHE [40]. Zhao et al. also found enhancement of the performance by nanostructure compared with dense structure (Figure 13) and proved that the main contribution of the observed enhancement is the improved charge separation efficiency, which is caused by the easy diffusion of minority carriers to the interface [41]. Zhu et al. reported a nanorod-structured ZnFe2O4 with a new benchmark solar photocurrent of 1.0 mA·cm−2 at 1.23 V and 1.7 mA·cm−2 at 1.6 V vs. RHE (Figure 14) [42]. In general, if the average particle diameter of the nanopores is shorter than holes diffusion length, carrier recombination in bulk can be effectively inhibited. Table 1 shows holes diffusion length of some typical materials for OER.

3.3. Heterojunction

A heterojunction comprises, but is not limited to, two materials, whereby electron–hole pairs are produced by one or more semiconductor photocatalysts and then transfer to the other material, which facilitates the space charge separation [47]. The two materials should meet some requirements. Taking a photoanode as an example, the valence band of A is more positive than B; at the same time, the conduction band of photocatalyst B should be more negative than A. This would facilitate the electron and hole transfer from one to the other (Figure 15). A lot of work has reported the heterojunction for enhanced PEC performance, such as WO3/BiVO4 [48,49], Cu2O/g-C3N4 [50], Cu2O/TiO2, p-Cu2O/n-TaON [51]. It is also found that heterojunction can be formed between different facets of the same crystal, e.g., BiVO4 (010) and BiVO4 (110) surfaces. The electrons and holes can accumulate on different surfaces of BiVO4 (Figure 16) [52], which causes the space separation of electron–hole. Wang et al. found improved hydrogen production in a TiO2 photocatalyst by forming {001}-{010} “quasi” heterojunctions [53]. Another case of heterojunctions is α-Fe2O3-TiO2 heterojunction. Holes generated in TiO2 will be transferred to α-Fe2O3 with electrons preferentially accumulating in TiO2 at the same time. Under specific conditions, hybrid composites such as Fe2TiO5, Fe3TiO4, and FeTiO3 form at the interface between hematite and titania that, despite some controversial results, exhibit remarkable performances as photoanode materials in PEC devices [54].

3.4. Internal Electric Field to Improve the Charge Separation

Recent studies indicate that the internal electric field in a photocatalyst can improve the charge separation [55]. The internal electric field provides a driving force for the separation of photoinduced charge carriers. Ferroelectric materials are a big family for having internal electric field due to their spontaneous electric polarization property among various photocatalytic materials. The spontaneous polarization in ferroelectric materials are due to the noncentrosymmetric property and the polarization occurs because the positive and negative charges have different centers of symmetry [56]. Perovskite oxides (ABO3) are among the most extensively studied ferroelectrics, such as BaTiO3 [57] and BiFeO3 [58]. Rohrer et al. found that BaxSr1−xTiO3 solid solution by replacing Ba with Sr can enlarge the space-charge region and improve the activity [59]. Zhang et al. also found that incorporation of carbon into the Bi3O4Cl lattice can increase internal electric field to boost bulk charge separation [60]. Hao et al. found that double perovskite ferroelectric Bi2FeMoxNi1−xO6 thin films also possess strong ferroelectric self-polarization, which has played a crucial part in improving the photovoltaic effect [61]. Another case of charge separation is α-Fe2O3-TiO2 heterojunction. Some research works found that the Fe2O3-TiO2 heterojunction showed enhancement in the photocurrent [62,63]. However, somehow it is not consistent with our general understanding on the heterojunction, because hematite has a more positive conduction band potential than TiO2, which prohibits the electron transfer from hematite to TiO2. Further study shows that this kind of heterojunction is through iron titanates (e.g., Fe2TiO5, Fe3TiO4, and FeTiO3) formed at the interface [54].

4. Strategies to Enhance the Surface Charge Injection

4.1. Catalyst Loading

In PEC water splitting, the catalyst loading play three vital roles for improving both the activity and reliability: (1) to reduce the overpotential or activation energy; (2) to assist charge separation at the semiconductor/electrolyte interface; (3) to suppress the photocorrosion and improve the durability of the working device [64]. Transition metal or transition metal oxides are usually used as catalyst loading to achieve a higher activity and decent reaction rates [65]. Recently, many (Fe, Co, Ni)-based catalyst loading have been widely investigated. For example, cobalt phosphate (CoPi) is the star cocatalyst to lower the overpotential and enhance the performance of water oxidation (Figure 17) [66]. Hydro-oxides are also one kind of effective catalyst loading, such as FeOOH, NiOOH, and CoOOH [67,68]. Kim et al. employed FeOOH/NiOOH dual-layer catalyst loading and achieved maximum ABPE of 1.75% due to reduction of the surface recombination [40]. Zhang et al. extended the catalyst loading to metal organic framework (MOF) materials, through applying in situ synthesized poly[-Co2(benzimidazole)4] nanoparticles on the surface of BiVO4 to boost surface reaction kinetics (Figure 18) [69]. In this respect, the bilayered Ni-FeOOH oxygen evolution catalyst layer has been reported as the best catalysts for BiVO4 to date [70].

4.2. Surface Treatment and Surface Passivation

Point defects are usually found on the surface of photoelectrode. In some cases, defects play a positive role in the surface catalysis. However, in some other cases, defects serve as recombination centers. Thus, it is important to suppress the recombination on these surface centers. Recently, it was found that the surface treatment can improve the PEC performance. The mechanism for surface treatment and surface passivation is illustrated in Figure 19 [71].
Surface defect states will lead to high charge recombination rate and water oxidation with low efficiency, caused by the photogenerated holes. Surface passivation and treatment can passivate defect states, forcefully suppressing surface recombination for improving water oxidation efficiency. For example, Luo et al. found that electrochemical cyclic voltammetry surface pretreatment on Mo-doped BiVO4 can greatly enhances the photocurrent of Mo-doped BiVO4 [72]. This is mainly because the pretreatment that can remove the surface recombination center MoOx which is segregated during the preparation of the photoelectrode. After the preprocessing, MoOx on the surface is dissolved into the electrolyte (Figure 20).
Further, Wang et al. found that the electrochemical treatment is not only suitable for doped BiVO4, but also applicable to pristine BiVO4 and other materials, such as TiO2, Fe2O3 [73]. Besides the surface treatment, a thin passivation layer is also proved to be effective to enhance the surface catalysis. Originally, passivation layers were applied to semiconductor photoelectrodes to improve their photochemical or chemical stability. It was reported that even an extremely thin layer on the photoelectrodes can change the surface properties [74]. Hisatomi investigated the oxide surface layer on the onset potential of hematite photoanode, such as Al2O3, Ga2O3, and In2O3, which can greatly weaken the original potential for oxidation of water [75]. Formal et al. covered an extremely thin layer (0.1 to 2 nm) of alumina on nanostructured hematite photoanodes. The photocurrent was shifted cathodically by as much as 100 mV. Further investigation found that the cathodic shift is owing to the passivation of surface trapping states caused by the Al2O3 overlayer [76]. Besides, IrO2 and CoOx thin layers on TaON photoanodes were found to improve the durability by reducing the self-oxidation of the electrode [77,78]. In general, a suitable structure of the passivation layer for photoelectrodes can either improve the reaction kinetics, or increase the chemical or physical protection against corrosion, or both.

4.3. Active Sites

The chemical reactions for water splitting occur on the electrode surface through its active sites. The active sites, therefore, affect the performance of the surface catalysis. The low catalytic activities are mainly attributed to the relatively lower activity and smaller number of reactive sites. The activity of active sites can be reflected by the overpotential at the anode. Experimentally, it might be extremely challenging to identify the atomic surface active sites. Therefore, calculations using density functional theory (DFT) become important to reveal the overpotential at different sites, as well as to understand the atomic/molecular mechanism. There are some methods to calculate the overpotential based on different reaction process [79,80]. The well acceptable reaction process is the four intermediate steps reaction path, as shown in R3 to R6.
The overpotential tends to zero if G R 3 θ = G R 4 θ   =   G R 5 θ = G R 6 θ = 1.23 eV and this would define the ideal catalyst, as shown in Figure 21 [81].
Until now, the atomic level interaction between catalytic activity and active sites is still an outstanding problem, owing to the large difference between idealized models and real catalysts. Thus, it is necessary to identify the pivotal factors that influence the activity of reactive sites, in order to promote the catalytic properties. It is generally considered that the low coordinated steps, edges, terraces, kinks, and/or corner atoms are often the favorable catalytic reaction sites [82,83]. For example, Sun et al. suggested that the low-coordinated Co3+ atoms of porous atomically-thick Co3O4 sheets serve as the catalytically active sites [84]. Currently, there are no quantitative models to describe the relationship between the coordinated number and activity. The main reason for this difficulty is that there might be other key parameters that have not been separately verified. For example, some results show that the surface state of the atom and position of neighbor atom will greatly affect the activity of the reactive sites. Hu et al. found that oxygen vacancy in BiVO4 will induce more active sites on a BiVO4 surface, which makes the vacancy sites active for water oxidation (Figure 22) [85]. Similarly, a Ti 3d defect state in the band gap of titania also plays an important role in the surface catalysis of defect contained TiO2 [86]. Zhao et al. also revealed that W doping and W–Ti codoping into BiVO4 have led to more active sites on the surface and reduced charge transfer resistance (Figure 23) [87,88]. Theoretical calculation revealed that Ti sites become active after the codoping.

5. Conclusions and Outlook

In this review, we have highlighted the latest studies directed at developing metal oxide-based photoanodes, including BiVO4, α-Fe2O3, TiO2, and WO3 nanostructures. We have summarized the strategies to enhance the light absorption, charge separation, and surface charge injection of this class of n-type semiconductor oxide materials. Although great achievements have been obtained, significant challenges remain before commercially viable solar-driven water splitting can be realized. Among them, one challenge is to balance the contributions from factors affecting light absorption, charge separation, and charge injection. Sometimes, their contributions contradict each other; e.g., a thicker layer increases absorption efficiency, but reduces the charge separation efficiency due to a longer distance to be travelled by the charge carrier. Another challenge lies with the lack of thorough understanding of the effect of the surface-state, due to the complex roles it plays in the water splitting process. Surface states may affect the amount of stored charges, act as surface recombination centers, and pin the Fermi-level. Most of the existing models can only explain one of these potential effects and there is no commonly used model to describe the full impact of surface state. A big challenge is to combine all the advantages and improve the overall efficiency for water splitting. Usually, it is inevitable to introduce some other problems when improving one factor. For example, nanostructuring is mostly employed to enhance the charge separation and light absorption; however, more surface states will be formed which sometimes are unfavorable to the interfacial charge injection. Thus, catalyst loading is required to passivate the surface states. Then, the challenge becomes finding a suitable catalyst loading to passivate and to control its thickness through a proper deposition technology.
Recently, researchers are focusing on the formation of ultrathin layers of cocatalyst on the surface, which will provide very efficient charge injection while avoiding the recombination in the surface layer and at the interface. However, it is very difficult to precisely control the formation of a uniform cover layer of the cocatalyst on the surface with reasonable cost of production. New low-cost processing technology and materials need to be developed. Another related issue is the stability and durability from the point of view of practical application. Newly developed materials should be coated with good adhesion while having good chemical and photochemical stability. More work is needed to build a more comprehensive model to guide the design and optimization of the electrode materials for solar water splitting.

Funding

This research was funded by Ministry of Education of Singapore (RG15/16), the National Natural Science Foundation of China (Nos. 21676216, 21576224) and China Postdoctoral Science Foundation (Nos. 2014M550507; 2015T81046).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. van de Krol, R.; Grätzel, M. Photoelectrochemical Hydrogen Production; Springer: Berlin, Germany, 2012; pp. 13–69. [Google Scholar]
  2. Inoue, H.; Shimada, T.; Kou, Y.; Nabetani, Y.; Masui, D.; Takagi, S.; Tachibana, H. The water oxidation bottleneck in artificial photosynthesis: How can we get through it? an alternative route involving a two-electron process. ChemSusChem 2011, 4, 173–179. [Google Scholar] [CrossRef]
  3. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  4. Sivula, K.; Van De Krol, R. Semiconducting materials for photoelectrochemical energy conversion. Nat. Rev. Mater. 2016, 1, 15010. [Google Scholar] [CrossRef] [Green Version]
  5. Li, Z.; Luo, W.; Zhang, M.; Feng, J.; Zou, Z. Photoelectrochemical cells for solar hydrogen production: Current state of promising photoelectrodes, methods to improve their properties, and outlook. Energy Environ. Sci. 2013, 6, 347–370. [Google Scholar] [CrossRef]
  6. Zhou, Y.; Zhang, L.; Lin, L.; Wygant, B.R.; Liu, Y.; Zhu, Y.; Zheng, Y.; Mullins, C.B.; Zhao, Y.; Zhang, X. Highly efficient photoelectrochemical water splitting from hierarchical WO3/BiVO4 nanoporous sphere arrays. Nano Lett. 2017, 17, 8012–8017. [Google Scholar] [CrossRef] [PubMed]
  7. Zhao, J.; Wang, A.; Green, M.A.; Ferrazza, F. 19.8% efficient “honeycomb” textured multicrystalline and 24.4% monocrystalline silicon solar cells. Appl. Phys. Lett. 1998, 73, 1991–1993. [Google Scholar] [CrossRef]
  8. Zhao, X.; Chen, Z. Enhanced photoelectrochemical water splitting performance using morphology-controlled BiVO4 with W doping. Beilstein J. Nanotechnol. 2017, 8, 2640–2647. [Google Scholar] [CrossRef] [PubMed]
  9. Qiu, Y.; Liu, W.; Chen, W.; Zhou, G.; Hsu, P.-C.; Zhang, R.; Liang, Z.; Fan, S.; Zhang, Y.; Cui, Y. Efficient solar-driven water splitting by nanocone BiVO4-perovskite tandem cells. Sci. Adv. 2016, 2, e1501764. [Google Scholar] [CrossRef]
  10. Feng, X.; Shankar, K.; Varghese, O.K.; Paulose, M.; Latempa, T.J.; Grimes, C.A. Vertically aligned single crystal TiO2 nanowire arrays grown directly on transparent conducting oxide coated glass: Synthesis details and applications. Nano Lett. 2008, 8, 3781–3786. [Google Scholar] [CrossRef] [PubMed]
  11. Lou, S.N.; Amal, R.; Scott, J.; Ng, Y.H. Concentration-mediated band gap reduction of Bi2MoO6 photoanodes prepared by Bi3+ cation insertions into anodized MoO3 thin films: Structural, optical, and photoelectrochemical properties. ACS Appl. Energy Mater. 2018, 1, 3955–3964. [Google Scholar] [CrossRef]
  12. Mai, L.; Yang, F.; Zhao, Y.; Xu, X.; Xu, L.; Hu, B.; Luo, Y.; Liu, H. Molybdenum oxide nanowires: Synthesis & properties. Mater. Today 2011, 14, 346–353. [Google Scholar] [CrossRef]
  13. Lou, S.N.; Ng, Y.H.; Ng, C.; Scott, J.; Amal, R. Harvesting, storing and utilising solar energy using MoO3: Modulating structural distortion through pH adjustment. ChemSusChem 2014, 7, 1934–1941. [Google Scholar] [CrossRef]
  14. Asahi, R.; Morikawa, T.; Irie, H.; Ohwaki, T. Nitrogen-doped titanium dioxide as visible-light-sensitive photocatalyst: Designs, developments, and prospects. Chem. Rev. 2014, 114, 9824–9852. [Google Scholar] [CrossRef]
  15. Liu, G.; Yin, L.C.; Wang, J.; Niu, P.; Zhen, C.; Xie, Y.; Cheng, H.M. A red anatase TiO2 photocatalyst for solar energy conversion. Energy Environ. Sci. 2012, 5, 9603–9610. [Google Scholar] [CrossRef]
  16. Xie, J.; Zhang, J.; Li, S.; Grote, F.; Zhang, X.; Zhang, H.; Wang, R.; Lei, Y.; Pan, B.; Xie, Y. Controllable disorder engineering in oxygen-incorporated MoS2 ultrathin nanosheets for efficient hydrogen evolution. J. Am. Chem. Soc. 2013, 135, 17881–17888. [Google Scholar] [CrossRef] [PubMed]
  17. Morikawa, T.; Saeki, S.; Suzuki, T.; Kajino, T.; Motohiro, T. Dual functional modification by N doping of Ta2O5: P-type conduction in visible-light-activated N-doped Ta2O5. Appl. Phys. Lett. 2010, 96, 142111. [Google Scholar] [CrossRef]
  18. Maeda, K.; Takata, T.; Hara, M.; Saito, N.; Inoue, Y.; Kobayashi, H.; Domen, K. GaN: ZnO solid solution as a photocatalyst for visible-light-driven overall water splitting. J. Am. Chem. Soc. 2005, 127, 8286–8287. [Google Scholar] [CrossRef] [PubMed]
  19. Chen, X.; Liu, L.; Yu, P.Y.; Mao, S.S. Increasing solar absorption for photocatalysis with black hydrogenated titanium dioxide nanocrystals. Science 2011, 331, 746–750. [Google Scholar] [CrossRef]
  20. Naldoni, A.; Altomare, M.; Zoppellaro, G.; Liu, N.; Kment, Š.; Zbořil, R.; Schmuki, P. Photocatalysis with reduced TiO2: From black TiO2 to cocatalyst-free gydrogen production. ACS Catal. 2019, 9, 345–364. [Google Scholar] [CrossRef] [PubMed]
  21. Cui, H.; Zhao, W.; Yang, C.; Yin, H.; Lin, T.; Shan, Y.; Xie, Y.; Gu, H.; Huang, F. Black TiO2 nanotube arrays for high-efficiency photoelectrochemical water-splitting. J. Mater. Chem. A 2014, 2, 8612–8616. [Google Scholar] [CrossRef]
  22. Wang, S.; Chen, P.; Bai, Y.; Yun, J.H.; Liu, G.; Wan, L. New BiVO4 dual photoanodes with enriched oxygen vacancies for efficient solar-driven water splitting. Adv. Mater. 2018, 30, 1800486. [Google Scholar] [CrossRef] [PubMed]
  23. Kim, J.H.; Jang, J.W.; Jo, Y.H.; Abdi, F.F.; Lee, Y.H.; Van De Krol, R.; Lee, J.S. Hetero-type dual photoanodes for unbiased solar water splitting with extended light harvesting. Nat. Commun. 2016, 7, 13380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Mayer, M.T.; Du, C.; Wang, D. Hematite/Si nanowire dual-absorber system for photoelectrochemical water splitting at low applied potentials. J. Am. Chem. Soc. 2012, 134, 12406–12409. [Google Scholar] [CrossRef] [PubMed]
  25. Brillet, J.; Yum, J.-H.; Cornuz, M.; Hisatomi, T.; Solarska, R.; Augustynski, J.; Graetzel, M.; Sivula, K. Highly efficient water splitting by a dual-absorber tandem cell. Nat. Photonics 2012, 6, 824. [Google Scholar] [CrossRef]
  26. Chang, X.; Wang, T.; Zhang, P.; Zhang, J.; Li, A.; Gong, J. Enhanced surface reaction kinetics and charge separation of p–n heterojunction Co3O4/BiVO4 photoanodes. J. Am. Chem. Soc. 2015, 137, 8356–8359. [Google Scholar] [CrossRef] [PubMed]
  27. Ye, K.; Wang, Z.; Gu, J.; Xiao, S.; Yuan, Y.; Zhu, Y.; Zhang, Y.; Mai, W.; Yang, S. Carbon quantum dots as a visible light sensitizer to significantly increase the solar water splitting performance of bismuth vanadate photoanodes. Energy Environ. Sci. 2017, 10, 772–779. [Google Scholar] [CrossRef]
  28. Kim, J.K.; Shi, X.; Jeong, M.J.; Park, J.; Han, H.S.; Kim, S.H.; Guo, Y.; Heinz, T.F.; Fan, S.; Lee, C.L.; et al. Enhancing Mo:BiVO4 Solar Water Splitting with Patterned Au Nanospheres by Plasmon-Induced Energy Transfer. Adv. Energy Mater. 2018, 8, 1701765. [Google Scholar] [CrossRef]
  29. Thimsen, E.; Formal, F.L.; Grätzel, M.; Warren, S.C. Influence of plasmonic Au nanoparticles on the photoactivity of Fe2O3 electrodes for water splitting. Nano Lett. 2011, 11, 35–43. [Google Scholar] [CrossRef]
  30. Mascaretti, L.; Dutta, A.; Kment, Š.; Shalaev, V.M.; Boltasseva, A.; Zbořil, R.; Naldoni, A. Plasmon-enhanced photoelectrochemical water splitting for efficient renewable energy storage. Adv. Mater. 2019, 1805513. [Google Scholar] [CrossRef]
  31. Zhao, X.; Feng, J.; Chen, S.; Huang, Y.; Sum, T.C.; Chen, Z. New insight into the roles of oxygen vacancies in hematite for solar water splitting. Phys. Chem. Chem. Phys. 2017, 19, 1074–1082. [Google Scholar] [CrossRef]
  32. Luo, W.; Yang, Z.; Li, Z.; Zhang, J.; Liu, J.; Zhao, Z.; Wang, Z.; Yan, S.; Yu, T.; Zou, Z. Solar hydrogen generation from seawater with a modified BiVO4 photoanode. Energy Environ. Sci. 2011, 4, 4046–4051. [Google Scholar] [CrossRef]
  33. Berglund, S.P.; Rettie, A.J.E.; Hoang, S.; Mullins, C.B. Incorporation of Mo and W into nanostructured BiVO4 films for efficient photoelectrochemical water oxidation. Phys. Chem. Chem. Phys. 2012, 14, 7065–7075. [Google Scholar] [CrossRef]
  34. Guo, Y.; Zhang, N.; Wang, X.; Qian, Q.; Zhang, S.; Li, Z.; Zou, Z. A facile spray pyrolysis method to prepare Ti-doped ZnFe2O4 for boosting photoelectrochemical water splitting. J. Mater. Chem. A 2017, 5, 7571–7577. [Google Scholar] [CrossRef]
  35. Lumley, M.A.; Choi, K.S. Investigation of pristine and (Mo, W)-doped Cu11V6O26 for use as photoanodes for solar water splitting. Chem. Mater. 2017, 29, 9472–9479. [Google Scholar] [CrossRef]
  36. Jo, W.J.; Jang, J.W.; Kong, K.; Kang, H.J.; Kim, J.Y.; Jun, H.; Parmar, K.P.S.; Lee, J.S. Phosphate doping into monoclinic BiVO4 for enhanced photoelectrochemical water oxidation activity. Angew. Chem. Int. Ed. 2012, 51, 3147–3151. [Google Scholar] [CrossRef]
  37. Yang, Z.; Chang, B.; Zou, J.; Qiao, J.; Gao, P.; Zeng, Y.; Li, H. Comparison between gradient-doping GaAs photocathode and uniform-doping GaAs photocathode. Appl. Opt. 2007, 46, 7035–7039. [Google Scholar] [CrossRef]
  38. Abdi, F.F.; Han, L.; Smets, A.H.; Zeman, M.; Dam, B.; Van De Krol, R. A bismuth vanadate–cuprous oxide tandem cell for overall solar water splitting. Nat. Commun. 2013, 4, 2195. [Google Scholar] [CrossRef]
  39. Krol, R.; Liang, Y.; Schoonman, J. Solar hydrogen production with nanostructured metal oxides. J. Mater. Chem. 2008, 18, 2311–2320. [Google Scholar] [CrossRef]
  40. Kim, T.W.; Choi, K.-S. Nanoporous BiVO4 photoanodes with dual-layer oxygen evolution catalysts for solar water splitting. Science 2014, 343, 990–994. [Google Scholar] [CrossRef]
  41. Zhao, X.; Luo, W.; Feng, J.; Li, M.; Li, Z.; Yu, T.; Zou, Z. Quantitative analysis and visualized evidence for high charge separation efficiency in a solid-liquid bulk heterojunction. Adv. Energy Mater. 2014, 4, 1301785. [Google Scholar] [CrossRef]
  42. Zhu, X.; Guijarro, N.; Liu, Y.; Schouwink, P.; Wells, R.A.; Formal, F.L.; Sun, S.; Gao, C.; Sivula, K. Spinel structural disorder influences solar-water-splitting performance of ZnFe2O4 nanorod photoanodes. Adv. Mater. 2018, 30, 1801612. [Google Scholar] [CrossRef]
  43. Iwase, A.; Yoshino, S.; Takayama, T.; Ng, Y.H.; Amal, R.; Kudo, A. Water splitting and CO2 reduction under visible light irradiation using Z-Scheme systems consisting of metal sulfides, CoOx-loaded BiVO4, and a reduced graphene oxide electron mediator. J. Am. Chem. Soc. 2016, 138, 10260–10264. [Google Scholar] [CrossRef]
  44. Booshehri, A.Y.; Goh, S.C.; Hong, J.D.; Jiang, R.; Xu, R. Effect of depositing silver nanoparticles on BiVO4 in enhancing visible light photocatalytic inactivation of bacteria in water. J. Mater. Chem. A 2014, 2, 6209–6217. [Google Scholar] [CrossRef]
  45. Zhang, Z.; Wang, P. Optimization of photoelectrochemical water splitting performance on hierarchicalTiO2 nanotube arrays. Energy Environ. Sci. 2012, 5, 6506–6512. [Google Scholar] [CrossRef]
  46. Butler, M.A. Photoelectrolysis and physical properties of the semiconducting electrode WO2. J. Appl. Phys. 1977, 48, 1914–1920. [Google Scholar] [CrossRef]
  47. Moniz, S.J.; Shevlin, S.A.; Martin, D.J.; Guo, Z.-X.; Tang, J. Visible-light driven heterojunction photocatalysts for water splitting–a critical review. Energy Environ. Sci. 2015, 8, 731–759. [Google Scholar] [CrossRef]
  48. Su, J.; Guo, L.; Bao, N.; Grimes, C.A. Nanostructured WO3/BiVO4 heterojunction films for efficient photoelectrochemical water splitting. Nano Lett. 2011, 11, 1928–1933. [Google Scholar] [CrossRef]
  49. 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]
  50. Zhang, S.; Yan, J.; Yang, S.; Xu, Y.; Cai, X.; Li, X.; Zhang, X.; Peng, F.; Fang, Y. Electrodeposition of Cu2O/g-C3N4 heterojunction film on an FTO substrate for enhancing visible light photoelectrochemical water splitting. Chin. J. Catal. 2017, 38, 365–371. [Google Scholar] [CrossRef]
  51. Hou, J.; Yang, C.; Cheng, H.; Jiao, S.; Takeda, O.; Zhu, H. High-performance p-Cu2O/n-TaON heterojunction nanorod photoanodes passivated with an ultrathin carbon sheath for photoelectrochemical water splitting. Energy Environ. Sci. 2014, 7, 3758–3768. [Google Scholar] [CrossRef]
  52. Hu, J.; Chen, W.; Zhao, X.; Su, H.; Chen, Z. Anisotropic electronic characteristics, adsorption, and stability of low-index BiVO4 surfaces for photoelectrochemical applications. ACS Appl. Mater. Interfaces 2018, 10, 5475–5484. [Google Scholar] [CrossRef]
  53. Wang, D.; Kanhere, P.; Li, M.; Tay, Q.; Tang, Y.; Huang, Y.; Sum, T.C.; Mathews, N.; Sritharan, T.; Chen, Z. Improving Photocatalytic H2 Evolution of TiO2 via Formation of {001}−{010} Quasi-Heterojunctions. J. Phys. Chem. C 2013, 117, 22894–22902. [Google Scholar] [CrossRef]
  54. Kment, S.; Riboni, F.; Pausova, S.; Wang, L.; Wang, L.; Han, H.; Hubicka, Z.; Krysa, J.; Schmuki, P.; Zboril, R. Photoanodes based on TiO2 and α-Fe2O3 for solar water splitting—Superior role of 1D nanoarchitectures and of combined heterostructures. Chem. Soc. Rev. 2017, 46, 3716–3769. [Google Scholar] [CrossRef]
  55. Huang, X.; Wang, K.; Wang, Y.; Wang, B.; Zhang, L.; Gao, F.; Zhao, Y.; Feng, W.; Zhang, S.; Liu, P. Enhanced charge carrier separation to improve hydrogen production efficiency by ferroelectric spontaneous polarization electric field. Appl. Catal. B-Environ. 2018, 227, 322–329. [Google Scholar] [CrossRef]
  56. Li, L.; Salvador, P.A.; Rohrer, G.S. Photocatalysts with internal electric fields. Nanoscale 2014, 6, 24–42. [Google Scholar] [CrossRef] [Green Version]
  57. Giocondi, J.L.; Rohrer, G.S. The influence of the dipolar field effect on the photochemical reactivity of Sr2Nb2O7 and BaTiO3 microcrystals. Top. Catal. 2008, 49, 18–23. [Google Scholar] [CrossRef]
  58. Schultz, A.M.; Zhang, Y.; Salvador, P.A.; Rohrer, G.S. Effect of crystal and domain orientation on the visible-light photochemical reduction of Ag on BiFeO3. ACS Appl. Mater. Interfaces 2011, 3, 1562–1567. [Google Scholar] [CrossRef]
  59. Bhardwaj, A.; Burbure, N.V.; Gamalski, A.; Rohrer, G.S. Composition dependence of the photochemical reduction of Ag by Ba1−xSrxTiO3. Chem. Mater. 2010, 22, 3527–3534. [Google Scholar] [CrossRef]
  60. Li, J.; Cai, L.; Shang, J.; Yu, Y.; Zhang, L. Giant enhancement of internal electric field boosting bulk charge separation for photocatalysis. Adv. Mater. 2016, 28, 4059–4064. [Google Scholar] [CrossRef]
  61. Cui, X.; Li, Y.; Sun, N.; Dua, J.; Lia, X.; Yang, H.; Hao, X. Double perovskite Bi2FeMoxNi1−xO6 thin films: Novel ferroelectric photovoltaic materials with narrow bandgap and enhanced photovoltaic performance. Sol. Energy Mater. Sol. C 2018, 187, 9–14. [Google Scholar] [CrossRef]
  62. Han, H.; Riboni, F.; Karlicky, F.; Kment, S.; Goswami, A.; Sudhagar, P.; Yoo, J.; Wang, L.; Tomanec, O.; Petr, M.; et al. α-Fe2O3/TiO2 3D hierarchical nanostructures for enhanced photoelectrochemical water splitting. Nanoscale 2017, 9, 134–142. [Google Scholar] [CrossRef] [PubMed]
  63. Jeon, T.H.; Choi, W.; Park, H. Photoelectrochemical and photocatalytic behaviors of hematite-decorated Titania nanotube arrays: Energy level mismatch versus surface specific reactivity. J. Phys. Chem. C 2011, 115, 7134–7142. [Google Scholar] [CrossRef]
  64. Ran, J.; Zhang, J.; Yu, J.; Jaroniec, M.; Qiao, S.Z. Earth-abundant cocatalysts for semiconductor-based photocatalytic water splitting. Chem. Soc. Rev. 2014, 43, 7787–7812. [Google Scholar] [CrossRef] [PubMed]
  65. Maeda, K.; Domen, K. Photocatalytic water splitting: Recent progress and future challenges. J. Phys. Chem. Lett. 2010, 1, 2655–2661. [Google Scholar] [CrossRef]
  66. Zhong, D.K.; Sun, J.; Inumaru, H.; Gamelin, D.R. Solar water oxidation by composite catalyst/α-Fe2O3 photoanodes. J. Am. Chem. Soc. 2009, 131, 6086–6087. [Google Scholar] [CrossRef] [PubMed]
  67. Kim, T.W.; Ping, Y.; Galli, G.A.; Choi, K.S. Simultaneous enhancements in photon absorption and charge transport of bismuth vanadate photoanodes for solar water splitting. Nat. Commun. 2015, 6, 8769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Tang, F.; Cheng, W.; Su, H.; Zhao, X.; Liu, Q. Smoothing surface trapping states in 3D coral-like CoOOH-wrapped-BiVO4 for efficient photoelectrochemical water oxidation. ACS Appl. Mater. Interfaces 2018, 10, 6228–6234. [Google Scholar] [CrossRef]
  69. Zhang, W.; Li, R.; Zhao, X.; Chen, Z.; Law, A.W.K.; Zhou, K. A cobalt-based metal–organic framework as cocatalyst on BiVO4 photoanode for enhanced photoelectrochemical water oxidation. ChemSusChem 2018, 11, 2710–2716. [Google Scholar] [CrossRef]
  70. Shi, X.; Choi, I.Y.; Kan, Z.; Kwon, J.; Dong, Y.K.; Lee, J.K.; Sang, H.O.; Kim, J.K.; Park, J.H. Efficient photoelectrochemical hydrogen production from bismuth vanadate-decorated tungsten trioxide helix nanostructures. Nat. Commun. 2014, 5, 4775. [Google Scholar] [CrossRef] [Green Version]
  71. Liu, R.; Zheng, Z.; Spurgeon, J.; Yang, X. Enhanced photoelectrochemical water-splitting performance of semiconductors by surface passivation layers. Energy Environ. Sci. 2014, 7, 2504–2517. [Google Scholar] [CrossRef] [Green Version]
  72. Luo, W.; Li, Z.; Yu, T.; Zou, Z. Effects of surface electrochemical pretreatment on the photoelectrochemical performance of Mo-Doped BiVO4. J. Phys. Chem. C 2012, 116, 5076–5081. [Google Scholar] [CrossRef]
  73. Wang, S.; Chen, P.; Yun, J.H.; Hu, Y.; Wang, L. An electrochemically treated BiVO4 photoanode for efficient photoelectrochemical water splitting. Angew. Chem. Int. Ed. 2017, 56, 8500–8504. [Google Scholar] [CrossRef]
  74. Riha, S.C.; Klahr, B.M.; Tyo, E.C.; Seifert, S.; Vajda, S.; Pellin, M.J.; Hamann, T.W.; Martinson, A.B.F. Atomic layer deposition of a submonolayer catalyst for the enhanced photoelectrochemical performance of water oxidation with hematite. ACS Nano 2013, 7, 2396–2405. [Google Scholar] [CrossRef]
  75. Hisatomi, T.; Le Formal, F.; Cornuz, M.; Brillet, J.; Tétreault, N.; Sivula, K.; Grätzel, M. Cathodic shift in onset potential of solar oxygen evolution on hematite by 13-group oxide overlayers. Energy Environ. Sci. 2011, 4, 2512–2515. [Google Scholar] [CrossRef]
  76. Le Formal, F.; Tetreault, N.; Cornuz, M.; Moehl, T.; Gr€atzel, M.; Sivula, K. Passivating surface states on water splitting hematite photoanodes with alumina overlayers. Chem. Sci. 2011, 2, 737–743. [Google Scholar] [CrossRef]
  77. Abe, R.; Higashi, M.; Domen, K. Facile fabrication of an efficient oxynitride TaON photoanode for overall water splitting into H2 and O2 under visible light irradiation. J. Am. Chem. Soc. 2010, 132, 11828–11829. [Google Scholar] [CrossRef]
  78. Higashi, M.; Domen, K.; Abe, R. Highly stable water splitting on oxynitride TaON photoanode system under visible light irradiation. J. Am. Chem. Soc. 2012, 134, 6968–6971. [Google Scholar] [CrossRef]
  79. Yang, J.; Wang, D.; Zhou, X.; Li, C. A theoretical study on the mechanism of photocatalytic oxygen evolution on BiVO4 in aqueous solution. Chem. Eur. J. 2013, 19, 1320–1326. [Google Scholar] [CrossRef]
  80. Zhao, Z.Y. Single water molecule adsorption and decomposition on the low-index stoichiometric rutile TiO2 surfaces. J. Phys. Chem. C 2014, 118, 4287–4295. [Google Scholar] [CrossRef]
  81. Valdés, Á.; Brillet, J.; Grätzel, M.; Gudmundsdóttir, H.; Hansen, H.A.; Jónsson, H.; Klüpfel, P.; Kroes, G.J.; Formal, F.L.; Man, I.C.; et al. Solar hydrogen production with semiconductor metal oxides: New directions in experiment and theory. Phys. Chem. Chem. Phys. 2012, 14, 49–70. [Google Scholar] [CrossRef]
  82. Sun, Y.F.; Liu, Q.H.; Gao, S.; Cheng, H.; Lei, F.C.; Sun, Z.H.; Jiang, Y.; Su, H.B.; Wei, S.Q.; Xie, Y. Pits confined in ultrathin cerium (IV) oxide for studying catalytic centers in carbon monoxide oxidation. Nat. Commun. 2013, 4, 2899. [Google Scholar] [CrossRef]
  83. Sun, Y.F.; Lei, F.C.; Gao, S.; Pan, B.C.; Zhou, J.F.; Xie, Y. Atomically thin tin dioxide sheets for efficient catalytic oxidation of carbon monoxide. Angew. Chem. Int. Ed. 2013, 52, 10569. [Google Scholar] [CrossRef]
  84. Sun, Y.F.; Gao, S.; Lei, F.C.; Liu, J.W.; Liang, L.; Xie, Y. Atomically-thin non-layered cobalt oxide porous sheets for highly efficient oxygen-evolving electrocatalysts. Chem. Sci. 2014, 5, 3976–3982. [Google Scholar] [CrossRef]
  85. Hu, J.; Zhao, X.; Chen, W.; Su, H.; Chen, Z. Theoretical insight into the mechanism of photoelectrochemical oxygen evolution reaction on BiVO4 anode with oxygen vacancy. J. Phys. Chem. C 2017, 121, 18702–18709. [Google Scholar] [CrossRef]
  86. Wendt, S.; Sprunger, P.T.; Lira, E.; Madsen, G.K.H.; Li, Z.; Hansen, J.O.; Matthiesen, J.; Blekinge-Rasmussen, A.; Lægsgaard, E.; Hammer, B.; Besenbacher, F. The role of interstitial sites in the Ti3D defect state in the band gap of Titania. Science 2008, 320, 1755–1759. [Google Scholar] [CrossRef] [PubMed]
  87. Zhao, X.; Hu, J.; Chen, S.; Chen, Z. An investigation on the role of W doping in BiVO4 photoanodes used for solar water splitting. Phys. Chem. Chem. Phys. 2018, 20, 13637–13645. [Google Scholar] [CrossRef] [PubMed]
  88. Zhao, X.; Hu, J.; Wu, B.; Banerjee, A.; Chakraborty, S.; Feng, J.; Zhao, Z.; Chen, S.; Ahuja, R.; Sum, T.C.; Chen, Z. Simultaneous enhancement in charge separation and onset potential for water oxidation in a BiVO4 photoanode by W–Ti codoping. J. Mater. Chem. A 2018, 6, 16965–16974. [Google Scholar] [CrossRef]
Figure 1. Dependence of theoretical solar-to-hydrogen (STH) efficiency and solar photocurrent density of photoelectrodes on their bandgap absorption edges. Reprinted with permission from ref. [5]. Copyright 2013. The Royal Society of Chemistry.
Figure 1. Dependence of theoretical solar-to-hydrogen (STH) efficiency and solar photocurrent density of photoelectrodes on their bandgap absorption edges. Reprinted with permission from ref. [5]. Copyright 2013. The Royal Society of Chemistry.
Coatings 09 00309 g001
Figure 2. Schematic illustration of the three efficiencies of photoelectrochemical (PEC) water splitting for enhancing performance.
Figure 2. Schematic illustration of the three efficiencies of photoelectrochemical (PEC) water splitting for enhancing performance.
Coatings 09 00309 g002
Figure 3. Optical absorption and electron transport on nanoporous BiVO4 with (a) the flat substrate and (b) the conductive nanocone substrate. Reprinted with permission from ref. [9]. Copyright 2016. Science. (c) flat BiVO4 film and (d) hierarchical BiVO4 nanostructured film. Reprinted with permission from ref. [10]. Copyright 2008. American Chemical Society.
Figure 3. Optical absorption and electron transport on nanoporous BiVO4 with (a) the flat substrate and (b) the conductive nanocone substrate. Reprinted with permission from ref. [9]. Copyright 2016. Science. (c) flat BiVO4 film and (d) hierarchical BiVO4 nanostructured film. Reprinted with permission from ref. [10]. Copyright 2008. American Chemical Society.
Coatings 09 00309 g003
Figure 4. Diffuse reflectance absorption spectrum of N-Ta2O5 power (8.9 at. % N), nondoped Ta2O5, TaON, and Ta3N5 in the ultraviolet and visible light region. Reprinted with permission from ref. [17]. Copyright 2018. American Institute of Physics.
Figure 4. Diffuse reflectance absorption spectrum of N-Ta2O5 power (8.9 at. % N), nondoped Ta2O5, TaON, and Ta3N5 in the ultraviolet and visible light region. Reprinted with permission from ref. [17]. Copyright 2018. American Institute of Physics.
Coatings 09 00309 g004
Figure 5. (a) Power X-ray diffraction pattern and (b) ultraviolet light visible diffuse reflectance spectra: a, b, c, d, e stand for GaN (ref), GaN:ZnO (Zn 3.4 at. %), GaN:ZnO (Zn 6.4 at. %), GaN:ZnO (Zn 13.3 at. %), and ZnO. Reprinted with permission from ref. [18]. Copyright 2018. American Chemical Society.
Figure 5. (a) Power X-ray diffraction pattern and (b) ultraviolet light visible diffuse reflectance spectra: a, b, c, d, e stand for GaN (ref), GaN:ZnO (Zn 3.4 at. %), GaN:ZnO (Zn 6.4 at. %), GaN:ZnO (Zn 13.3 at. %), and ZnO. Reprinted with permission from ref. [18]. Copyright 2018. American Chemical Society.
Coatings 09 00309 g005
Figure 6. (a) Applied bias photon-to-current efficiency (ABPE) of the TiO2 nanotubes (TNTs) and black titania nanotube (B-TNTs) as a function of applied potential; (b) incident-photon-to-current-conversion efficiency (IPCE) spectra in the region of 300–700 nm at 0.23 V vs. Ag/AgCl. Reprinted with permission from ref. [21]. Copyright 2014. The Royal Society of Chemistry.
Figure 6. (a) Applied bias photon-to-current efficiency (ABPE) of the TiO2 nanotubes (TNTs) and black titania nanotube (B-TNTs) as a function of applied potential; (b) incident-photon-to-current-conversion efficiency (IPCE) spectra in the region of 300–700 nm at 0.23 V vs. Ag/AgCl. Reprinted with permission from ref. [21]. Copyright 2014. The Royal Society of Chemistry.
Coatings 09 00309 g006
Figure 7. (a) The BiVO4-FeOOH/NiOOH dual photoanodes coupled with s sealed perovskite solar cell; (b) I: JV curve of perovskite solar cell, II: JV curve of perovskite solar cell behind the BVO-FeOOH/NiOOH dual photoanodes, III: JV curve of BVO-FeOOH/NiOOH under AM 1.5 G illumination; (c) Stability of unassisted water splitting (0 V vs. the counter electrode Pt). Reprinted with permission from ref. [22]. Copyright 2018. Wiley.
Figure 7. (a) The BiVO4-FeOOH/NiOOH dual photoanodes coupled with s sealed perovskite solar cell; (b) I: JV curve of perovskite solar cell, II: JV curve of perovskite solar cell behind the BVO-FeOOH/NiOOH dual photoanodes, III: JV curve of BVO-FeOOH/NiOOH under AM 1.5 G illumination; (c) Stability of unassisted water splitting (0 V vs. the counter electrode Pt). Reprinted with permission from ref. [22]. Copyright 2018. Wiley.
Coatings 09 00309 g007
Figure 8. Wavelength-selective solar light absorption by hetero-type dual photoanode. Reprinted with permission from ref. [23]. Copyright 2018. Nature.
Figure 8. Wavelength-selective solar light absorption by hetero-type dual photoanode. Reprinted with permission from ref. [23]. Copyright 2018. Nature.
Coatings 09 00309 g008
Figure 9. The photocurrents of (a) pristine and (b) Ti-doped hematite before and after N2 treatment under AM 1.5 G solar simulator illumination. The N2 treatment is carried out in nitrogen gas at 600 °C for 2 h. Reprinted with permission from ref. [31]. Copyright 2017. The Royal Society of Chemistry.
Figure 9. The photocurrents of (a) pristine and (b) Ti-doped hematite before and after N2 treatment under AM 1.5 G solar simulator illumination. The N2 treatment is carried out in nitrogen gas at 600 °C for 2 h. Reprinted with permission from ref. [31]. Copyright 2017. The Royal Society of Chemistry.
Coatings 09 00309 g009
Figure 10. The photocurrents of (a) ZnFe2O4 photoanodes and Ti-doped ZnFe2O4 photoanodes in 1 M NaOH aqueous solution under AM 1.5 G illumination (100 mW·cm−2); (c) IPCE spectrum of Ti-doped ZnFe2O4 and pure ZnFe2O4 in range from 330 to 589 nm at different applied potential; (d) Mott–Schottky plots of pure Ti-doped ZnFe2O4 and ZnFe2O4 photoanodes in 1 M NaOH, the AC amplitude is 5 mV and the frequency is 1000 Hz. Reprinted with permission from ref. [34]. Copyright 2018. American Chemical Society.
Figure 10. The photocurrents of (a) ZnFe2O4 photoanodes and Ti-doped ZnFe2O4 photoanodes in 1 M NaOH aqueous solution under AM 1.5 G illumination (100 mW·cm−2); (c) IPCE spectrum of Ti-doped ZnFe2O4 and pure ZnFe2O4 in range from 330 to 589 nm at different applied potential; (d) Mott–Schottky plots of pure Ti-doped ZnFe2O4 and ZnFe2O4 photoanodes in 1 M NaOH, the AC amplitude is 5 mV and the frequency is 1000 Hz. Reprinted with permission from ref. [34]. Copyright 2018. American Chemical Society.
Coatings 09 00309 g010
Figure 11. (a) 1% W-doped BiVO4; (b) W:BiVO4 homojunction; (c) W:BiVO4 reverse homojunction; and (d) gradient-doped W:BiVO4; (e) Carrier-separation efficiency (ηsep) in different applied potential for 1% W-doped BiVO4 (black triangle), W:BiVO4 homojunction (red inverted triangle), W:BiVO4 reverse homojunction (green circle), and gradient-doped W:BiVO4 (blue square). Reprinted with permission from ref. [38]. Copyright 2018. Nature.
Figure 11. (a) 1% W-doped BiVO4; (b) W:BiVO4 homojunction; (c) W:BiVO4 reverse homojunction; and (d) gradient-doped W:BiVO4; (e) Carrier-separation efficiency (ηsep) in different applied potential for 1% W-doped BiVO4 (black triangle), W:BiVO4 homojunction (red inverted triangle), W:BiVO4 reverse homojunction (green circle), and gradient-doped W:BiVO4 (blue square). Reprinted with permission from ref. [38]. Copyright 2018. Nature.
Coatings 09 00309 g011
Figure 12. Illustration of the morphology of an α-Fe2O3 photo-anode for water splitting. The small diameter of nanowires ensures short hole diffusion path lengths. Reprinted with permission from ref. [39]. Copyright 2008. The Royal Society of Chemistry.
Figure 12. Illustration of the morphology of an α-Fe2O3 photo-anode for water splitting. The small diameter of nanowires ensures short hole diffusion path lengths. Reprinted with permission from ref. [39]. Copyright 2008. The Royal Society of Chemistry.
Coatings 09 00309 g012
Figure 13. (a) The photocurrent at 1.23 V RHE of the porous and the dense Mo doped BiVO4 films with different thickness in 0.5 M Na2SO4 aqueous solution under front (electrolyte-BiVO4 surface side) illumination with a 500 W Xe lamp equipped with a 420 nm cut-off filter at a scan rate of 30 mV·s−1; (b) IPCE at 1.23 V RHE of the porous and the dense BiVO4 films in 0.5 M Na2SO4 aqueous solution under front illumination. Reprinted with permission from [41]. Copyright 2018. Wiley.
Figure 13. (a) The photocurrent at 1.23 V RHE of the porous and the dense Mo doped BiVO4 films with different thickness in 0.5 M Na2SO4 aqueous solution under front (electrolyte-BiVO4 surface side) illumination with a 500 W Xe lamp equipped with a 420 nm cut-off filter at a scan rate of 30 mV·s−1; (b) IPCE at 1.23 V RHE of the porous and the dense BiVO4 films in 0.5 M Na2SO4 aqueous solution under front illumination. Reprinted with permission from [41]. Copyright 2018. Wiley.
Coatings 09 00309 g013
Figure 14. The performance of ZnFe2O4 (ZFO) photoanodes. (a) Linear scanning J–V curves in 1 m NaOH under 100 mW·cm−2 illumination; (b) Calculated charge separation efficiency, ηsep, and minority charge carrier injection efficiency, ηinj. Reprinted with permission from ref. [42]. Copyright 2018. Wiley.
Figure 14. The performance of ZnFe2O4 (ZFO) photoanodes. (a) Linear scanning J–V curves in 1 m NaOH under 100 mW·cm−2 illumination; (b) Calculated charge separation efficiency, ηsep, and minority charge carrier injection efficiency, ηinj. Reprinted with permission from ref. [42]. Copyright 2018. Wiley.
Coatings 09 00309 g014
Figure 15. The band structure for the different heterojunctions. CB stands for conduct band and VB stands for valance band.
Figure 15. The band structure for the different heterojunctions. CB stands for conduct band and VB stands for valance band.
Coatings 09 00309 g015
Figure 16. Schematic diagram of the heterojunction on between BiVO4 (010) and BiVO4 (110) surfaces. Reprinted with permission from ref. [52]. Copyright 2018. American Chemical Society.
Figure 16. Schematic diagram of the heterojunction on between BiVO4 (010) and BiVO4 (110) surfaces. Reprinted with permission from ref. [52]. Copyright 2018. American Chemical Society.
Coatings 09 00309 g016
Figure 17. (a) Dark (dashed) and photocurrent (solid) densities for Fe2O3 (red) and Co–Pi/Fe2O3 (blue) photoanodes, collected using simulated AM 1.5 illumination (1 sun, backside) at a scan rate of 50 mV·s−1; (b) Electronic absorption and (c) IPCE spectra for Fe2O3 and Co–Pi/Fe2O3 at 1.23 and 1 V vs. RHE, respectively. The absorption spectrum of Co–Pi on FTO without Fe2O3 is included in (b), but no photocurrent was detected for these anodes. Reprinted with permission from ref. [66]. Copyright 2018. American Chemical Society.
Figure 17. (a) Dark (dashed) and photocurrent (solid) densities for Fe2O3 (red) and Co–Pi/Fe2O3 (blue) photoanodes, collected using simulated AM 1.5 illumination (1 sun, backside) at a scan rate of 50 mV·s−1; (b) Electronic absorption and (c) IPCE spectra for Fe2O3 and Co–Pi/Fe2O3 at 1.23 and 1 V vs. RHE, respectively. The absorption spectrum of Co–Pi on FTO without Fe2O3 is included in (b), but no photocurrent was detected for these anodes. Reprinted with permission from ref. [66]. Copyright 2018. American Chemical Society.
Coatings 09 00309 g017
Figure 18. (a) Synthesis process of [Co2(bim)4]-modified BiVO4 photoanodes; (b) XRD patterns of BiVO4, Cobim/BiVO-20, Cobim/BiVO-40, Cobim/BiVO-100, and simulated [[Co2(bim)4], where the symbols asterisk, dot, and triangle correspond to the peaks from FTO, BiVO4, and [Co2(bim)4], respectively; (c) Crystal structure of [Co2(bim)4], where gray, blue, and red balls represent carbon, nitrogen, and cobalt, respectively. Reprinted with permission from ref. [69]. Copyright 2018. American Chemical Society.
Figure 18. (a) Synthesis process of [Co2(bim)4]-modified BiVO4 photoanodes; (b) XRD patterns of BiVO4, Cobim/BiVO-20, Cobim/BiVO-40, Cobim/BiVO-100, and simulated [[Co2(bim)4], where the symbols asterisk, dot, and triangle correspond to the peaks from FTO, BiVO4, and [Co2(bim)4], respectively; (c) Crystal structure of [Co2(bim)4], where gray, blue, and red balls represent carbon, nitrogen, and cobalt, respectively. Reprinted with permission from ref. [69]. Copyright 2018. American Chemical Society.
Coatings 09 00309 g018
Figure 19. Band structure of an n-type semiconductor photoelectrode experiencing corrosion (a) when contacting an aqueous electrolyte; (b) with passivation layer without corrosion. ϕox is the corrosion potential of the n-type semiconductor; and ϕred is the corrosion potential of the p-type semiconductor. Corrosion may reduce the light absorption and/or generate more surface defect states. Reprinted with permission from ref. [71] Copyright 2014. The Royal Society of Chemistry.
Figure 19. Band structure of an n-type semiconductor photoelectrode experiencing corrosion (a) when contacting an aqueous electrolyte; (b) with passivation layer without corrosion. ϕox is the corrosion potential of the n-type semiconductor; and ϕred is the corrosion potential of the p-type semiconductor. Corrosion may reduce the light absorption and/or generate more surface defect states. Reprinted with permission from ref. [71] Copyright 2014. The Royal Society of Chemistry.
Coatings 09 00309 g019
Figure 20. Transfer of photogenerated electrons in Mo-doped BiVO4 photoelectrodes illuminated from the front-side and the back-side. Reprinted with permission from ref. [72]. Copyright 2018. American Chemical Society.
Figure 20. Transfer of photogenerated electrons in Mo-doped BiVO4 photoelectrodes illuminated from the front-side and the back-side. Reprinted with permission from ref. [72]. Copyright 2018. American Chemical Society.
Coatings 09 00309 g020
Figure 21. The free energy diagram for the OER ideal catalyst. The black continuous line is the OER ideal catalyst, where the over-potential is 0 eV because all free energies of the intermediate steps equal to 1.23 V. Red lines show the actual HOO* and HO* levels for the catalysts, which have to be moved down and up, respectively, in order to reduce the over-potential. Unfortunately, a metal oxide catalyst usually changes these two levels in the same direction, as indicated by the red (move up) and blue arrows (move down). O* level is positioned between the HOO* and HO* level. Reprinted with permission from ref. [81]. Copyright 2012. The Royal Society of Chemistry.
Figure 21. The free energy diagram for the OER ideal catalyst. The black continuous line is the OER ideal catalyst, where the over-potential is 0 eV because all free energies of the intermediate steps equal to 1.23 V. Red lines show the actual HOO* and HO* levels for the catalysts, which have to be moved down and up, respectively, in order to reduce the over-potential. Unfortunately, a metal oxide catalyst usually changes these two levels in the same direction, as indicated by the red (move up) and blue arrows (move down). O* level is positioned between the HOO* and HO* level. Reprinted with permission from ref. [81]. Copyright 2012. The Royal Society of Chemistry.
Coatings 09 00309 g021
Figure 22. PEC water splitting process on a BiVO4 (010) facet with or without Ovac. Ovac will greatly affect the adsorption energy. Reprinted with permission from ref. [85]. Copyright 2017. American Chemical Society.
Figure 22. PEC water splitting process on a BiVO4 (010) facet with or without Ovac. Ovac will greatly affect the adsorption energy. Reprinted with permission from ref. [85]. Copyright 2017. American Chemical Society.
Coatings 09 00309 g022
Figure 23. Electrochemistry results in 0.5 M Na2SO4 aqueous solution under 1.5 G illumination (a) Mott-Schottky plots of pristine, W doped BiVO4 and W–Ti codoped BiVO4 at the frequency of 1 kHz; (b) Css determined from EIS (rectangle points). Inset is the equivalent circuits employed to fit the experimental EIS data; (c) EIS of pristine, W doped, and W–Ti codoped BiVO4 photoelectrodes at the applied potential of 1.23 V RHE; (d) Cyclic voltammetry scan of pristine, W doped and W–Ti codoped BiVO4 in the dark at a scan rate of 100 mV·s−1 after holding the electrode at the potential of 1.8 V RHE for 120 s; (e) Free-energy profiles of OER on different sites of BiVO4 surfaces. Reprinted with permission from ref. [88]. Copyright 2018. The Royal Society of Chemistry.
Figure 23. Electrochemistry results in 0.5 M Na2SO4 aqueous solution under 1.5 G illumination (a) Mott-Schottky plots of pristine, W doped BiVO4 and W–Ti codoped BiVO4 at the frequency of 1 kHz; (b) Css determined from EIS (rectangle points). Inset is the equivalent circuits employed to fit the experimental EIS data; (c) EIS of pristine, W doped, and W–Ti codoped BiVO4 photoelectrodes at the applied potential of 1.23 V RHE; (d) Cyclic voltammetry scan of pristine, W doped and W–Ti codoped BiVO4 in the dark at a scan rate of 100 mV·s−1 after holding the electrode at the potential of 1.8 V RHE for 120 s; (e) Free-energy profiles of OER on different sites of BiVO4 surfaces. Reprinted with permission from ref. [88]. Copyright 2018. The Royal Society of Chemistry.
Coatings 09 00309 g023
Table 1. The hole diffusion length of different materials.
Table 1. The hole diffusion length of different materials.
MaterialsHole Diffusion LengthRef.
BiVO4100 nm[43]
α-Fe2O32–4 nm[44]
TiO270 nm[45]
WO3150 nm[46]

Share and Cite

MDPI and ACS Style

Hu, J.; Zhao, S.; Zhao, X.; Chen, Z. Strategies of Anode Materials Design towards Improved Photoelectrochemical Water Splitting Efficiency. Coatings 2019, 9, 309. https://doi.org/10.3390/coatings9050309

AMA Style

Hu J, Zhao S, Zhao X, Chen Z. Strategies of Anode Materials Design towards Improved Photoelectrochemical Water Splitting Efficiency. Coatings. 2019; 9(5):309. https://doi.org/10.3390/coatings9050309

Chicago/Turabian Style

Hu, Jun, Shuo Zhao, Xin Zhao, and Zhong Chen. 2019. "Strategies of Anode Materials Design towards Improved Photoelectrochemical Water Splitting Efficiency" Coatings 9, no. 5: 309. https://doi.org/10.3390/coatings9050309

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