Tokuhisa
Kawawaki
ab,
Yutaro
Mori
a,
Kosuke
Wakamatsu
a,
Shuhei
Ozaki
a,
Masanobu
Kawachi
a,
Sakiat
Hossain
a and
Yuichi
Negishi
*ab
aDepartment of Applied Chemistry, Faculty of Science, Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8601, Japan. E-mail: negishi@rs.tus.ac.jp; Fax: +81-3-5261-4631; Tel: +81-3-5228-9145
bResearch Institute for Science & Technology, Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8601, Japan
First published on 27th July 2020
In recent years, research on the use of metal nanoparticles (NPs) and nanoclusters (NCs) synthesized by liquid-phase reduction in water-splitting photocatalysts has been actively conducted. Water-splitting photocatalysts have been attracting attention because they can produce hydrogen (H2), which is attractive as a next-generation energy source, from solar energy and water. However, further improvement of water-splitting photocatalysts is required for their practical use in society. Recent studies have demonstrated that the active sites (cocatalysts) of water-splitting photocatalysts can be controlled using the advanced NP/NC syntheses and structural modulation techniques established in the fields of colloid, NP, and NC chemistry and thereby highly active water-splitting photocatalysts can be developed. If such research progresses further, it is expected that a transition to a new society using H2 as the main energy source will become possible. However, such applied research has just started and examples of such research are currently limited. The purpose of this review is to introduce the importance of controlled colloidal NPs/NCs in research on water-splitting photocatalysis to readers by summarizing the existing research. We hope that this review will raise interest in the application of metal NPs/NCs in water-splitting photocatalysis and that a society actively addressing energy and environmental problems will become a reality as soon as possible.
Metal nanoparticles (NPs) (Fig. 1),1,2 which are aggregates of metal atoms, play a central role in such nanotechnology research. Research on metal NPs was initiated by Faraday in the 1840s in his investigation of metal colloids.3 Since the 1980s, the research on metal colloids has progressed substantially and the expression “metal NPs” has been coined.4–23 In the last two decades, research on metal NPs has increased explosively, and the techniques to synthesize metal NPs have advanced dramatically. In recent years, it has become possible to control not only the size of metal NPs but also their geometrical structure. Metal NPs composed of coinage metals such as gold (Au), silver (Ag), and copper (Cu) exhibit localized surface plasmon resonance absorption in the visible region (Fig. 1),23–27 which makes these NPs even more attractive. Because of these properties, metal NPs can be used in surface-enhanced Raman spectroscopy.28 Therefore, studies on the use of metal NPs to detect trace amounts of molecules are being conducted for applications.29 In addition, techniques to control the arrangement of metal NPs have been established,30 which has opened the way to apply metal NPs in the field of electronic devices. Furthermore, metal NPs are also expected to be useful in biotechnology applications such as drug delivery systems and diagnosis (Fig. 2).31
Metal nanoclusters (NCs), as shown in Fig. 1, are smaller in size than metal NPs and are also important materials in nanotechnology.32,33 There is no clear definition of the boundary between metal NPs and metal NCs. However, when the particle size is less than 2 nm, the materials are generally called metal NCs (Fig. 1). Such ultrafine metal NCs possess electronic and geometrical structures that are different from those of both the corresponding bulk metal and metal NPs, which leads to the appearance of new physical/chemical properties and functions (Fig. 1).34–45 Furthermore, because the physical/chemical properties and functions of NCs strongly depend on the number of constituent atoms, if the number of constituent atoms in metal NCs can be controlled, numerous functions can be realized by one metal element. When multiple elements are used in NCs, it is possible to access further various functions.46–56
Gas-phase experiments played a leading role in research on metal NCs in the 1980s and 90s.57–69 Also, the synthesis of metal NCs composed of gold, palladium (Pd), and platinum (Pt) started at that time.70–78 However, it was only after 2000 that research on these NCs began to increase explosively.79 In 2005, the first precise synthesis method for thiolate (SR)-protected metal NCs was established.80 Since then, numerous noble metal and alloy NCs have been precisely synthesized using SR, phosphine, and alkyne ligands.81–109 Since 2007, it has become possible to determine the geometrical structure of NCs by single-crystal X-ray diffraction (XRD) analysis.110–114 Thus, at present, inorganic chemists precisely synthesize SR-protected metal NCs as organic chemists synthesize organic molecules. It is possible to obtain a deep understanding of the structure–property relationship of metal NCs with precisely determined geometrical structure.82–114 The use of these well-defined NCs as catalysts, chemical sensors, photosensitizers, and solar cell components is currently being studied (Fig. 2).34,79,115–122
In this way, the syntheses and applications are being actively investigated for metal NPs and metal NCs at present. Multiple reviews have been published on the recent research development of such metal NPs and NCs.34,79,115–122 Readers hoping to obtain comprehensive knowledge of the synthesis techniques of metal NPs/NCs, their geometrical structures, and potential applications should refer to these reviews.
Fig. 3 Possible scheme for large-scale H2 production via solar water splitting. Reproduced with permission from ref. 124. Copyright 2010 American Chemical Society. |
Most H2 is currently produced by steam reforming of fossil resources. However, this method releases carbon dioxide as a byproduct and consumes fossil resources. Therefore, if H2 continues to be produced by this method, it will not provide a solution for both energy and environmental problems. The water-splitting photocatalytic reaction123 has been proposed as a clean and renewable H2 production method (Fig. 3).124 Using this reaction, it is possible to produce H2 from sunlight and water, which are available in almost unlimited quantities on the earth. This is different from the case of the water-splitting electrocatalytic reaction125–128 in which electric power is consumed to proceed the reaction (thus, the combination with the solar cells is indispensable in this case129,130). However, although water-splitting photocatalysts have been attracting attention for many years, further improvement is still required to realize their practical use.
Water-splitting semiconductor photocatalysts are often composed of semiconductor photocatalysts and metal NP/NC cocatalysts that work as reaction sites (Fig. 4).131 The cocatalyst plays a role in promoting the photocatalytic reaction, and it has been shown that control of the cocatalyst particle size and dispersion improves photocatalytic activity. However, it is difficult to control the particle size and chemical composition of cocatalysts loaded using conventional photodeposition and impregnation methods because the metal NPs/NCs are grown on the surface of the photocatalyst in these methods (Fig. 5A).132,133 To overcome these problems and produce highly active water-splitting photocatalysts suitable for practical use, it is essential to introduce new techniques for the preparation of water-splitting photocatalysts.
Fig. 5 Comparison of (A) conventional and (B) recent cocatalyst loading methods. As conventional methods, (a) impregnation and (b) photodeposition methods are shown in (A). |
As described in Section 1.1, techniques to control the particle size distribution of metal NPs fabricated by liquid-phase synthesis have already been established. When such controlled metal NPs are adsorbed on a photocatalyst and the surface protective organic molecules (ligand, polymer, etc.) are removed, it is possible to obtain controlled metal NP-loaded photocatalysts with high water-splitting activity (Fig. 5B).132,133 When precisely controlled metal NCs are used as the precursor, it is possible to regulate the cocatalyst on a photocatalyst with atomic precision.134–136 For such precise metal NCs, detailed information about electronic and geometrical structures can be obtained by various high-resolution measurements and theoretical calculations, as shown in the previous studies on the supported metal NPs/NCs.118,137 Therefore, loading the cocatalyst controlled with atomic accuracy would lead to the clear design guidelines for high activation of water-splitting photocatalysts. If appropriate metal NCs are designed and fabricated based on the knowledge obtained in this way, the activity of water-splitting photocatalysts is expected to be further enhanced.
It should be noted that there are many studies using NPs/NCs not as active sites but as light absorption sites in water-splitting photocatalysis.138–144 A number of reviews on the use of NPs/NCs as photosensitizers have already been published,37,115 so this review does not cover such research.
Fig. 6 Relationship between the semiconductor band structure and redox potentials of water splitting. Reproduced with permission from ref. 145–147, and 160. Copyright 2009 Royal Society of Chemistry, Copyright 2020 Royal Society of Chemistry, Copyright 2018 Royal Society of Chemistry, and Copyright 2010 American Chemical Society. |
On the other hand, to realize the practical application of water-splitting photocatalysis, it is necessary to use visible light, which accounts for about 40% of solar energy. As described in Section 2.1, to completely decompose water, the positions of the CB and VB of a photocatalyst must be at appropriate energies to enable the HER and OER, respectively. Furthermore, to suppress recombination of excited electrons and holes, the photocatalyst needs to have high crystallinity and large specific surface area. Because of these requirements, only a limited number of photocatalysts for one-step photoexcitation systems that can completely decompose water with visible light have been reported.151
Z-scheme systems include semiconductor photocatalysts that proceed a half reaction of water splitting (HER or OER). Therefore, compared with the case for one-step photoexcitation systems, there are many photocatalysts available, and it is possible to use a longer wavelength of sunlight. Additionally, because the HER and OER occur on different photocatalysts, it is possible to suppress the reverse reaction between the generated H2 and O2 by using a two-cell reaction tube with an ion-exchange membrane. Also, an operation to separate the generated gas is unnecessary. However, in Z-scheme systems, the reverse reaction involving a redox couple occurs, which is different from the case for one-step photoexcitation systems. Furthermore, because two photons are required for one reaction in overall water splitting, Z-scheme systems have the disadvantage of lower theoretical conversion efficiency than that of one-step photoexcitation systems.
The impregnation method (Fig. 5A(a)) was first used for loading the cocatalysts on the photocatalysts by Domen et al. in 1980.172 Since then, this method has been used as an effective method for loading the cocatalysts. In this method, the photocatalyst and precursor metal salt of the cocatalyst are thoroughly mixed using a mortar and pestle and then calcined to deposit fine cocatalyst particles on the photocatalyst surface. In the photodeposition method (Fig. 5A(b)), a photocatalyst is dispersed in a solution containing a precursor metal salt of the cocatalyst and then light is irradiated on the photocatalyst. The cocatalyst is deposited on the photocatalyst surface through the reduction or oxidation of the metal salt by the electrons or holes generated via photoexcitation.173 Using the photodeposition method, it is possible to preferentially load the cocatalyst particles on a specific crystal plane of the photocatalyst, which become the reaction sites on the photocatalyst surface.174
In the catalyst systems fabricated using these two methods, the metal NP/NCs necessarily have a size distribution because the metal atoms are aggregated on the photocatalyst (Fig. 5A).175 Also, it is not possible to apply methods such as size separation and size convergence to these loaded metal NPs/NCs, unlike metal NPs/NCs dispersed in a solution. Therefore, it is difficult to load cocatalysts with a uniform size on photocatalysts using these conventional methods.
In the liquid-phase adsorption method, first, metal NPs/NCs are adsorbed on the photocatalyst by mixing the metal NPs/NCs and photocatalyst in a solvent. Because metal oxide photocatalysts generally have numerous surface hydroxyl groups (–OH) in water, metal NPs/NCs with protective organic molecules with dissociative functional groups (such as –CO2H, –SO3H, or –NH3 groups) can be adsorbed on the photocatalyst surface at a high adsorption rate via hydrogen bond formation between the protective organic molecules and photocatalyst surface.176 Even when the protective organic molecules do not have dissociative functional groups, if they contain phenyl groups, the metal NPs/NCs can be adsorbed on the photocatalyst at a relatively high adsorption rate via dipole–dipole interactions.177 After adsorption, the protective organic molecules are removed from the metal NPs/NCs by calcination, photocatalytic oxidation/reduction by light irradiation, or ozone oxidation. This approach has attracted much attention in recent years as a novel method to load photocatalysts with cocatalysts with controlled particle size and chemical composition (Fig. 5B).
Fig. 7 Geometrical structures of photocatalysts introduced in this review. (A) GaN:ZnO, (B) SrTiO3, (C) BaLa4Ti4O15, (D) TiO2, (E) g-C3N4, (F) CdS, and (G) WO3. |
Semiconductor | Cocatalyst | Light source | Reactant solution | Ref. | ||
---|---|---|---|---|---|---|
Kinds | Loading amount | Size | ||||
GaN:ZnO | Cr2O3/Rh NPs | 0.3–0.4 wt% | 1.9 ± 0.6 nm | 450 W Hg lamp (NaNO2 aq. filter) | Distilled water, 400 mL | 176 |
GaN:ZnO | Cr2O3/Rh NPs (adsorption) | 2.5 wt% | 1.9 ± 0.6 nm | 450 W Hg lamp NaNO2 aq. filter) | Distilled water, 370–400 mL | 181 |
Cr2O3/Rh NPs (photodeposition) | 1.0 wt% | >2–3 nm | ||||
Cr2O3/Rh NPs (impregnation) | 0.25 wt% | >2–3 nm | ||||
GaN:ZnO | Rh NPs | 0.1–0.3 wt% | 1.5 ± 0.3 nm | 450 W Hg lamp (NaNO2 aq. filter) | H2SO4 aq. (pH 4.5), 400 mL | 182 |
0.1–0.3 wt% | 3.8 ± 0.8 nm | |||||
0.1–0.3 wt% | 6.6 ± 1.1 nm | |||||
GaN:ZnO | Cr2O3/Rh NPs (shell/core), Mn3O4 NPs | Rh: 0.75 wt%, Cr: 0.31 wt%, Mn: 0.05 wt% | Cr2O3/Rh: >2–3 nm, Mn3O4: 9.2 ± 0.4 nm | 300 W Xe lamp (λ > 420 nm) | Pure water, 100 mL | 184 |
GaN:ZnO | Cr2O3/Rh NPs, Mn3O4 NPs | Rh: 0.75 wt%, Cr: 0.31 wt%, Mn3O4: 0.2 wt% | 300 W Xe lamp (λ > 420 nm) | H2O, 100 mL | 185 | |
Cr2O3/Rh NPs, IrO2 NPs | Rh: 0.75 wt%, Cr: 0.31 wt%, IrO2: 0.3 wt% | |||||
Cr2O3/Rh NPs, RuO2 NPs | Rh: 0.75 wt%, Cr: 0.31 wt%, RuO2: 0.2 wt% | |||||
SrTiO3 | Cr2O3/Rh NPs, CoxMn3−xO4 NPs, (Co/(Co + Mn) ratio of 40 mol%) | Sum of Co and Mn: 0.05 wt%, Rh: 0.3 wt% | CoxMn3−xO4: 9.0 ± 0.8 nm | 300 W Xe lamp (λ > 300 nm) | H2O, 100 mL | 186 |
Mn: 0.05 wt%, Rh: 0.3 wt% | 9.0 ± 0.5 nm | |||||
BaLa4Ti4O15 | Au25 | 0.1 wt% | 1.2 ± 0.3 nm | 400 W Hg lamp | Distilled water, 350 mL | 187 |
Au NPs | 0.5 wt% | 10–30 nm | ||||
SrTiO3 | Au25 | 0.1 wt% | ∼1.2 nm | 400 W Hg lamp | Distilled water, 350 mL | 188 |
Au NPs | 0.5 wt% | 9.5 ± 3.3 nm | ||||
BaLa4Ti4O15 | Au10 | 0.1 wt% | 0.89 ± 0.19 nm | 400 W Hg lamp | Distilled water, 350 mL | 189 |
Au15 | 0.1 wt% | 0.95 ± 0.21 nm | ||||
Au18 | 0.1 wt% | 1.12 ± 0.22 nm | ||||
Au25 | 0.1 wt% | 1.19 ± 0.28 nm | ||||
Au39 | 0.1 wt% | 1.53 ± 0.26 nm | ||||
BaLa4Ti4O15 | Au25 | 0.1 wt% | 1.1 ± 0.2 nm | 400 W Hg lamp | Distilled water, 350 mL | 190 |
Cr2O3/Au25 | Cr: 0.5 wt%, Au: 0.1 wt% | 1.1 ± 0.3 nm | ||||
BaLa4Ti4O15 | Au25 | 0.1 wt% | 1.24 ± 0.21 nm | 400 W Hg lamp | Distilled water, 350 mL | 202 |
Au24Pd | 0.1 wt% | 1.11 ± 0.19 nm | ||||
Au24Pt | 0.1 wt% | 1.11 ± 0.19 nm | ||||
BaLa4Ti4O15 | Rh2−xCrxO3 NCs | Rh: 0.09 wt%, Cr: 0.10 wt% | 1.3 ± 0.3 nm | 400 W Hg lamp | Distilled water, 350 mL | 208 |
Rh2−xCrxO3 NCs (impregnation) | Rh: 0.10 wt%, Cr: 0.15 wt% | 3.0 ± 2.3 nm | ||||
TiO2 | PtO NCs | 0.5 wt% | 1.0 ± 0.3 | 300 W Xe lamp (λ > 300 nm) | Methanol aq. | 211 |
Pt NCs | 1.0 wt% | 2.0 ± 0.5 nm |
Fig. 8 (A) Procedural flow of the proposed liquid-phase adsorption method: (a) Rh NPs stabilized by organic ligand molecules before cation exchange. (b) Stabilized Rh NPs after cation exchange. (c) Electrostatic adsorption on GaN:ZnO catalyst. (d) Removal of organic ligand. (B) TEM images of Rh NPs loaded on GaN:ZnO (a and b) before and (c and d) after coating with Cr2O3. (C) Time course of overall water splitting over Cr2O3/Rh/GaN:ZnO prepared by liquid-phase adsorption and photodeposition. Reaction conditions: catalyst, 0.15 g; distilled water, 400 mL; light source, high-pressure mercury lamp (450 W) via aqueous NaNO2 solution filter to cut UV light; reaction vessel, Pyrex inner-irradiation vessel. Almost the same amount of Rh (0.3–0.4 wt%) was loaded on each catalyst. Reproduced with permission from ref. 176. Copyright 2009 The Royal Society of Chemistry. |
In 2010, the same group further developed such research. In this study, various noble metals (Rh, Pd, and Pt) and metal oxides (NiOx, RuO2, and Rh2O3) were used as HER cocatalysts for GaN:ZnO photocatalysts.181 Both conventional methods (photodeposition and impregnation methods; Fig. 5A) and liquid-phase adsorption (Fig. 5B) were used to load NPs on the photocatalyst. A Cr2O3 shell was also formed on the cocatalyst surface. The highest activity was obtained when Rh was used as the core metal and the cocatalysts were loaded by liquid-phase adsorption. In the liquid-phase adsorption method, Rh cocatalysts with fine particle size can be loaded with high dispersion. Additionally, it is presumed that Rh cocatalysts loaded by liquid-phase adsorption have more metallic properties than those loaded by conventional methods. The photocatalyst with Rh NPs loaded by liquid-phase adsorption was considered to show high water-splitting activity because of these factors. This study also revealed that the morphology of the Cr2O3 shell depended on the valence state of the core Rh NPs and pH of the solution during the photodeposition of the Cr2O3 layer. Based on this knowledge, they formed a Cr2O3 shell with appropriate thickness on metallic Rh NPs at an appropriate pH (3.0–7.5) (Fig. 9), which yielded an efficient water-splitting photocatalyst.
Fig. 9 TEM images of Cr2O3/Rh NPs (shell/core)/GaN:ZnO. The core Rh NPs were loaded by (A) photodeposition, (B) impregnation, and (C) liquid-phase adsorption. Reproduced with permission from ref. 181. Copyright 2010 Wiley-VCH. |
In 2013, to further functionalize the Rh NP cocatalyst, Teranishi and colleagues investigated the correlation between the particle size of Rh NPs and water-splitting activity.182 In this study, Rh NPs were synthesized by polyol reduction using polyvinylpyrrolidone (PVP, Mw: 10000 or 40000) as a protective polymer. The effects of the pH and temperature of the reaction solution (ethylene glycol solution) on the particle size of the resulting PVP-protected Rh NPs (PVP–Rh NPs) were investigated. The nucleation rate of PVP–Rh NPs was controlled by the pH. The results obtained in this experiment were in good agreement with those of a theoretical calculation reported by Goia and co-workers.183 The reaction temperature also strongly affected the nucleation rate of PVP–Rh NPs. Based on these findings, they selected the appropriate pH and reaction temperature conditions and fabricated size-controlled PVP–Rh NPs (1.6 ± 0.3, 2.7 ± 0.3, and 5.1 ± 0.5 nm). Each sample of PVP–Rh NPs was stirred in ethanol with a GaN:ZnO photocatalyst to induce cocatalyst adsorption. Then, PVP was removed from the NPs by calcination at 673 K (Fig. 10A). A transmission electron microscope (TEM) image of the calcined sample revealed that although slight aggregation occurred on the GaN:ZnO photocatalyst, the Rh NPs still maintained their high monodispersity (1.5 ± 0.3, 3.8 ± 0.8, and 6.6 ± 1.1 nm). Then, a Cr2O3 layer was formed on the Rh NPs by photodeposition. Measurements of the water-splitting activity of the series of photocatalysts revealed that their water-splitting activity increased with decreasing cocatalyst size (Fig. 10B). They attributed this phenomenon to the following effects caused by the miniaturization of Rh NPs: (i) increased relative surface area of smaller Rh NPs; (ii) improved charge separation; and (iii) increased active sites for the HER.
Fig. 10 (A) TEM images of three kinds of PVP–Rh NPs (a–c) as-synthesized, (d–f) adsorbed on GaN:ZnO, and (g–i) loaded on GaN:ZnO after calcination. (B) Initial rates of H2 and O2 evolution over GaN:ZnO modified with different-sized Cr2O3/Rh (shell/core) NPs. Black and white symbols/bars indicate H2 and O2, respectively. Reaction conditions: catalyst, 0.15 g; H2SO4 aq. (pH 4.5), 400 mL; light source, high-pressure Hg lamp (450 W) through an NaNO2 aq. filter to cut UV light; reaction vessel, Pyrex inner-irradiation vessel. Reproduced with permission from ref. 183. Copyright 2013 American Chemical Society. |
The above examples demonstrated that control of the HER cocatalyst is an effective approach to improve the water-splitting activity of photocatalysts. In 2010, Domen et al.184 showed that co-loading of an OER cocatalyst with an HER cocatalyst led to even higher water-splitting activity. In this study, GaN:ZnO was used as the photocatalyst, Cr2O3/Rh (shell/core) NPs were employed as the HER cocatalyst, and Mn3O4 NPs as the OER cocatalyst (Fig. 11A). First, MnO NPs (9.2 ± 0.4 nm) were synthesized by a liquid-phase reduction method as a precursor of the OER cocatalyst. The MnO NPs were adsorbed on the GaN:ZnO photocatalyst and then calcined at 673 K to form Mn3O4 NPs. Next, Cr2O3/Rh NPs were loaded on the GaN:ZnO photocatalyst as an HER cocatalyst. TEM measurements confirmed that the two types of cocatalysts were loaded on the photocatalyst without covering each other (Fig. 11B). The water-splitting activities of the photocatalyst loaded with only Cr2O3/Rh NPs and that loaded with both Cr2O3/Rh NPs and Mn3O4 NPs were measured. The results revealed that: (i) an HER cocatalyst (Cr2O3/Rh) is necessary for H2 production; (ii) the photocatalyst co-loaded with both HER and OER cocatalysts (Cr2O3/Rh NPs and Mn3O4 NPs) showed higher activity than the photocatalyst loaded with only the HER cocatalyst (Cr2O3/Rh NPs) (Fig. 11C); and (iii) water-splitting activity depended on the loading amount of OER cocatalyst (Mn3O4 NPs). It was noted that such a co-loading method of both HER and OER cocatalysts is also applicable to cases where NPs composed of different metals are used as cocatalysts.
Fig. 11 (A) Proposed reaction mechanism for visible light-driven overall water splitting on GaN:ZnO modified with Mn3O4 and Cr2O3/Rh (shell/core) NPs. CB: conduction band, VB: valence band, e−: electron, h+: hole. (B) TEM images of GaN:ZnO modified with Mn3O4 and Cr2O3/Rh (shell/core) NPs: (a) Cr2O3/Rh NPs, (b) Mn3O4 NPs. (C) Time courses of H2 and O2 evolution using modified GaN:ZnO catalysts under visible light (λ > 420 nm). Mn loading: 0.05 wt%. Reproduced with permission from ref. 184. Copyright 2010 Wiley-VCH. |
Thus, control of the cocatalyst is effective for both the HER and OER. However, in the early stages of research on cocatalysts, it was unclear whether the HER or OER was the rate-limiting step in the water-splitting reaction. In 2014, Domen and colleagues conducted a study to determine the rate-limiting step of water splitting.185 In this study, GaN:ZnO was again used as the photocatalyst, along with Mn3O4 NPs, RuO2 NPs, or IrO2 NPs as the OER cocatalyst and Cr2O3/Rh NPs as the HER cocatalyst (Fig. 12A). Evaluation of the water-splitting activity of each photocatalyst revealed that their activities were almost independent of the type of OER cocatalyst (Fig. 12B). Also, the amount of loaded OER cocatalyst to achieve high activity was much smaller than that of the HER cocatalyst. These results suggested that the HER was the rate-limiting step of water splitting by the photocatalyst used in this study. Based on these findings, they also optimized the HER cocatalyst. Specifically, GaN:ZnO photocatalysts co-loaded with RuO2 NPs and Cr2O3/Rh NPs were fabricated by both liquid-phase adsorption of Rh NPs (3.3 ± 0.8 nm) and photodeposition of Rh NPs. They measured the water-splitting activities of both types of Cr2O3/Rh NPs + RuO2 NPs/GaN:ZnO photocatalysts. The photocatalyst with Cr2O3/Rh NP cocatalyst loaded using liquid-phase adsorption showed higher water-splitting activity than that with the cocatalyst NPs loaded using the photodeposition method. These results indicate that controlling the HER cocatalyst is very important for enhancing the water-splitting activity of the GaN:ZnO photocatalyst.
Fig. 12 (A) TEM images of (a and b) Mn3O4 NPs/GaN:ZnO and (c and d) Cr2O3/Rh NPs + Mn3O4 NPs/GaN:ZnO (a and c) before and (b and d) after calcination of the photocatalyst at 873 K. (B) Photocatalytic activity of GaN:ZnO co-loaded with different OER cocatalysts and Cr2O3/Rh NPs. Circles, triangles, and squares indicate the loading of Mn3O4, IrO2, and RuO2 NPs, respectively. Closed and open symbols denote H2 and O2, respectively. Reproduced with permission from ref. 185. Copyright 2014 Wiley-VCH. |
When SrTiO3 was used as the photocatalyst instead of GaN:ZnO solid solution, control of the OER cocatalyst also proved effective to improve water-splitting activity. In 2018, Teranishi et al.186 synthesized CoxMn3−xO4 NPs (Co/(Co + Mn) = 0–40 mol%) by doping Co into Mn3O4 NPs using a liquid-phase reduction method (Fig. 13A). The obtained CoxMn3−xO4 NPs were loaded on an SrTiO3 photocatalyst as OER cocatalysts after Cr2O3/Rh NPs were loaded as HER cocatalysts (Fig. 13B). TEM analysis confirmed that the particle size (∼9 nm) of the CoxMn3−xO4 NPs was maintained after loading on SrTiO3. The water-splitting activity of the photocatalyst increased with the amount of Co of CoxMn3−xO4. The photocatalyst loaded with CoxMn3−xO4 NPs with a Co/(Co + Mn) ratio of 40 mol% showed water-splitting activity that was 1.8 times higher than that of the catalyst loaded with Mn3O4 NPs (Fig. 13B and C). To investigate the origin of such activity enhancement, they examined the electrocatalytic activity of CoxMn3−xO4-loaded BiVO4 photocatalysts in the OER. The results revealed that the efficiency of hole transfer from the photocatalyst to the cocatalyst was not increased by Co doping. Based on these results, they attributed the improvement of water-splitting activity to the increased OER activity on the cocatalyst surface induced by Co doping.
Fig. 13 (A) TEM images of CoxMn1−xO NPs (∼9 nm) with different Co ratios of (a) 0, (b) 10, (c) 20, (d) 30, and (e) 40 mol% Co. (B) Time course of overall water splitting using Cr2O3/Rh + CoxMn3−xO4/SrTiO3 under light irradiation (λ > 300 nm). The amount of loaded CoxMn3−xO4 was 0.05 wt% calculated from the sum of Co and Mn. (C) Photocatalytic activities of Cr2O3/Rh/SrTiO3 and Cr2O3/Rh + CoxMn3−xO4/SrTiO3 in overall water splitting. Co ratios were 0, 10, 20, 30, and 40 mol% Co. The loading amount was 0.05 wt%, as calculated from the sum of Co and Mn. Reaction conditions: catalyst, 0.1 g; aqueous solution, 100 mL; light source, 300 W Xe lamp (λ > 300 nm); reaction vessel, Pyrex top-irradiation vessel. Reproduced with permission from ref. 186. Copyright 2018 The Royal Society of Chemistry. |
Since 2013, Negishi's group has published several papers on the use of atomically precise metal NCs as precursors of HER cocatalysts. First, in 2013, they succeeded in loading Au25 NC with high dispersion on the surface of a UV light-responsive BaLa4Ti4O15 photocatalyst.187 In this experiment, first, Au25 NC protected by glutathionate (SG), Au25(SG)18, were synthesized precisely. Then, Au25(SG)18 was adsorbed on BaLa4Ti4O15. Thereafter, the ligands of Au25(SG)18 were removed by calcination, providing Au25 NC-loaded BaLa4Ti4O15 (denoted as Au25 NC/BaLa4Ti4O15). Various evaluations of Au25 NC/BaLa4Ti4O15 confirmed that the Au25 NCs were well dispersed on BaLa4Ti4O15 (Fig. 14A), and that most of the ligands were removed by calcination. The water-splitting activity of Au25 NC/BaLa4Ti4O15 was 2.6 times higher than that of BaLa4Ti4O15 loaded with Au NPs with a diameter of 10–30 nm by the photodeposition method (this sample is denoted as Au NPs/BaLa4Ti4O15) (Fig. 14B). These results indicate that the use of the fine Au NC cocatalyst is indeed effective for enhancing the water-splitting activity of the photocatalyst. In another paper, Negishi's group reported that such an improvement in water-splitting activity caused by the ultra-miniaturization of the Au cocatalyst also occurred when SrTiO3 was used as a photocatalyst.188 This result demonstrates that loading the fine Au NC cocatalyst (more widely metal NCs120) is also possible for the other photocatalysts when using SG as a ligand of NCs.
Fig. 14 (A) (a) TEM image and (b) size distribution of adsorbed NCs estimated from the TEM image of Au25 NC/BaLa4Ti4O15. (B) Time course of water splitting over Au25 NC/BaLa4Ti4O15 photocatalyst (0.1 wt% Au) prepared by the present method (red) and Au NPs/BaLa4Ti4O15 photocatalyst (0.5 wt% Au) prepared by the conventional photodeposition method (black). Reaction conditions: photocatalyst, 0.5 g; distilled water, 350 mL; light source, high-pressure Hg lamp (400 W), inner irradiation cell made of quartz. Reproduced with permission from ref. 187. Copyright 2013 The Royal Society of Chemistry. |
In 2015, Negishi's group synthesized a series of Aun(SG)m NCs (n = 10, 15, 18, 25, and 39) with controlled numbers of constituent atoms with atomic precision and used the NCs as cocatalyst precursors.189 They investigated the correlation between cocatalyst size and water-splitting activity. In this study, when Aun(SG)m NCs (n = 22, 29, and 33) were used as precursors, the aggregation of Au NCs on the photocatalyst surface occurred during adsorption or calcination. It has been revealed that Aun(SG)m NCs (n = 22, 29, and 33) are metastable species which are kinetically trapped during the NC formation and decompose (release of SG or Au-SG oligomers) in aqueous solution in a shorter time when compared with Aun(SG)m NCs (n = 10, 15, 18, 25, and 39).81 It was that for Aun(SG)m NCs (n = 22, 29, and 33), part of the Aun(SG)m NCs dissociate during the stirring process and Au-SG oligomers, formed from the dissociated products, adsorb conjunctively with the Aun(SG)m NCs. The existence of such Au-SG oligomers was considered to promote cohesion of clusters on BaLa4Ti4O15 during calcination. These results indicate that it is difficult to load Aun(SG)m NCs that are unstable in solution on the BaLa4Ti4O15 photocatalyst while maintaining the number of constituent atoms of NCs. This means that it is essential to use metal NCs that are stable in solution as a cocatalyst precursor to achieve precise loading of metal NCs (Fig. 15A). Then, the correlation between cocatalyst size and water-splitting activity was investigated by evaluating the water-splitting activities of Aun NCs/BaLa4Ti4O15 photocatalysts prepared using Aun(SG)m NCs (n = 10, 15, 18, 25, and 39), which are stable in solution,81 as cocatalyst precursors. It was found that the water-splitting activity increased as the size of the Aun NC cocatalyst decreased (Fig. 15B). Because the increase in activity with cocatalyst size decrease was modest, it was considered that the change in activity between the catalysts was mainly caused by the change in the proportion of Au NC surface atoms with miniaturization (i.e., Au atoms that react with hydrogen). Conversely, the difference in water-splitting activity between Aun NCs/BaLa4Ti4O15 and Au NPs/BaLa4Ti4O15 could not be fully explained by the difference in the number of surface atoms (Fig. 15B). From various analyses, it was found that the activity per Au atom on the surface of Au10 NCs/BaLa4Ti4O15 was only about 15–25% of that of Au NPs/BaLa4Ti4O15. Therefore, it was concluded that the main factor behind the improved activity caused by the ultra-miniaturization of the HER cocatalyst (Fig. 15B) was the increase of the number of surface Au atoms with an efficiency exceeding the decrease in activity per Au atom.
Fig. 15 (A) Schematic of the size control of Aun NCs loaded on photocatalysts. (B) Effect of cluster size on water-splitting activity studied using Aun NCs/BaLa4Ti4O15 and Au NPs/BaLa4Ti4O15 photocatalysts. The average values obtained from four measurements are plotted herein. (C) Comparison of the photocatalytic activities of BaLa4Ti4O15, Au25(SG)18 NC/BaLa4Ti4O15, and Au25 NC/BaLa4Ti4O15. Reproduced with permission from ref. 189. Copyright 2015 American Chemical Society. |
In the same paper, Negishi et al.189 also examined the importance of ligand removal in such composite systems. Multiple studies have reported that for composites composed of Aun(SG)m NCs and semiconductor photocatalysts, electron transfer occurs even when the ligands are present.139 Indeed, Au25(SG)18 NC/BaLa4Ti4O15 that had not been calcined showed higher water-splitting activity than that of BaLa4Ti4O15 without the NC cocatalysts (Fig. 15C). This indicates that electron transfer occurs between the Au25(SG)18 NC and BaLa4Ti4O15 without removing the ligands, and thus Au25(SG)18 NC also function as a cocatalyst. However, the water-splitting activity of Au25(SG)18 NC/BaLa4Ti4O15 was about 23% of that of Au25 NC/BaLa4Ti4O15 with the ligands removed (Fig. 15C). This means that the presence of the ligands lowers the efficiency of the electron transfer between Au25 NC and BaLa4Ti4O15 or decreases the activity of individual Au atoms. These results clarified that ligand removal by calcination is very important to obtain composite photocatalysts with high water-splitting activity.
In this way, the use of the fine metal NCs as a cocatalyst is effective to improve the water-splitting activity of photocatalysts. However, the ORR, which is a reverse reaction of water splitting, proceeds simultaneously with the HER on the Au25 NC surface.188 Therefore, to more effectively utilize the high surface area unique to fine metal NCs and thereby obtain highly active photocatalysts, it is necessary to form a shell that suppresses the ORR on the surface of Au25 NC. In 2018, Kurashige and co-workers attempted to form a Cr2O3 shell on Au25 NC.190 In the case of the Cr2O3/Rh NPs described in Section 4.1, the Cr2O3 shell was formed by photodeposition. However, in the case of Au25 NC/BaLa4Ti4O15, light irradiation induced the aggregation of Au25 NC on BaLa4Ti4O15. Therefore, it was difficult to use photodeposition to form a Cr2O3 shell on Au25 NC loaded on BaLa4Ti4O15 while maintaining the size of the NCs.
Surface science research has revealed that when a metal oxide supporting metal NPs is heated under H2 or O2 atmosphere, a strong metal–support interaction (SMSI) is induced and thereby an oxide film is formed on the metal NPs.191–194 Kurashige et al. attempted to use such an SMSI effect to form a Cr2O3 shell on the Au25 NC surface (Fig. 16A). Specifically, a Cr2O3 layer was deposited on BaLa4Ti4O15 by photodeposition to form Cr2O3/BaLa4Ti4O15 before loading Au25 NC (Fig. 16A). Then, Au25(SG)18 was adsorbed on Cr2O3/BaLa4Ti4O15, and both ligand removal and surface protection by the SMSI effect were achieved by calcination. A TEM image of the photocatalyst after calcination showed the formation of a thin film with a thickness of about 0.7–0.9 nm around particles with a diameter of about 1 nm (Fig. 16B). This indicates that the Au25 NC were covered with a chromium oxide layer during calcination (Fig. 16A). Some of the chromium oxide layer of Cr2O3/Au25 NC/BaLa4Ti4O15 was oxidized to a highly oxidized state during calcination. The photocatalyst was irradiated with UV light to reduce the highly oxidized chromium oxide to Cr2O3 and thereby the desired Cr2O3/Au25 NC/BaLa4Ti4O15 was obtained. The water-splitting activity of Cr2O3/Au25 NC/BaLa4Ti4O15 was about 19 time higher than that of Au25 NC/BaLa4Ti4O15 (Fig. 16C). This indicates that the formation of the Cr2O3 shell is effective at suppressing the ORR on the surface of the Au NC cocatalyst. The formation of such a Cr2O3 shell also suppressed aggregation of the cocatalyst during light irradiation. This indicates that the Cr2O3 shell formed by the method established by Kurashige and colleagues improves not only the water-splitting activity but also the stability of the cocatalyst on the photocatalyst surface.
Fig. 16 (A) Schematic of the preparation of Cr2O3/Au25 NC/BaLa4Ti4O15. (B) TEM images of (a) Au25 NC/BaLa4Ti4O15 and (b) Cr2O3/Au25 NC/BaLa4Ti4O15 after UV irradiation for 10 h. (C) Comparison of rates of H2 and O2 evolution by photocatalytic water splitting over Au25 NC/BaLa4Ti4O15 and Cr2O3/Au25 NC/BaLa4Ti4O15 (0.5 wt% Cr). Averages of values obtained from several experiments are shown. Reproduced with permission from ref. 190. Copyright 2018 American Chemical Society. |
It is expected that photocatalysts with higher water-splitting activity could be realized by heteroatom doping of fine Au NC cocatalysts. In previous studies on the effect of heteroatom doping on photocatalyst activity, experiments have been conducted using photocatalysts with distributions of the particle size and doping ratio (chemical composition) of the cocatalysts.195–201 To obtain a deep understanding of the effect of heteroatom doping on photocatalytic activity and thereby establish clear design guidelines for photocatalyst activation, it is essential to study composite photocatalysts on which the cocatalysts have strictly controlled chemical composition. Recently, Kurashige et al.202 attempted to use Au24Pd and Au24Pt NCs, in which one Au atom of the Au25 NC is substituted with Pd or Pt, as an HER cocatalyst. Au24Pd(SR)18 and Au24Pt(SR)18 NCs can be precisely synthesized only when a hydrophobic ligand is used.202 However, the metal NCs protected by hydrophobic ligands could not strongly interact with the hydrophilic surface of BaLa4Ti4O15, which led to poor adsorption on the photocatalyst. Then, they replaced some of the ligands of Au24Pd(SR)18 and Au24Pt(SR)18 NCs with a hydrophilic ligand, which allowed them to adsorb on BaLa4Ti4O15 at a high adsorption rate (Fig. 17A). The ligands on the NC surface were removed by calcination (Fig. 17A).
Fig. 17 (A) Schematic of the experimental procedure. (B) Proposed structures of Au24M NC/BaLa4Ti4O15 with M of (a) Au, (b) Pd, and (c) Pt before (upper) and during (lower) the water-splitting reaction. (C) Rates of photocatalytic evolution of H2 and O2 by water splitting over Au24M NC/BaLa4Ti4O15 with M of Au, Pd, and Pt. (D) Time course of water splitting over Cr2O3/(Au24Pt)1–3 NCs/BaLa4Ti4O15 and Au24Pt NC/BaLa4Ti4O15. Reproduced with permission from ref. 202. Copyright 2019 American Chemical Society. |
Investigation of the photocatalyst loaded with cocatalyst NCs controlled with atomic precision revealed the following three points about the heteroatom doping of the Au cocatalyst: (i) the Pd atom was located on the surface of the metal NC and the Pt atom was located at the interface between the metal NC and photocatalyst (Fig. 17B); (ii) Pd doping induced a decrease of water-splitting activity and Pt doping caused water-splitting activity to increase (Fig. 17C); (iii) these opposite effects of Pd and Pt heteroatom doping are strongly related to the location of the doped heteroatom (Fig. 17B). The results also showed that combining Pt doping and surface protection of the cocatalyst with a Cr2O3 shell increased the activity and stability of the photocatalysts to a greater extent than only Pt doping (Fig. 17D).
In the above series of research, Au was used as the base element of the NC cocatalyst. It has been predicted that Rh has higher catalytic activity than Au with respect to the HER based on volcano plots for hydrogen adsorption and desorption.203 Therefore, it is expected that a highly active water-splitting photocatalyst could be obtained by loading Rh and Cr oxide NPs/NCs as an HER cocatalyst on the photocatalyst. Indeed, Domen and co-workers reported that a photocatalyst loaded with Rh(III)–Cr(III) mixed oxide NPs (Rh2−xCrxO3; particle size = 10–30 nm) showed higher water-splitting activity than that of photocatalysts loaded with metal NPs composed of other elements.204–207 If ultrafine Rh2−xCrxO3 NCs could be loaded on the photocatalyst as an HER cocatalyst, it might be possible to activate the photocatalyst further. Very recently, Kurashige et al.208 attempted to load fine Rh2−xCrxO3 NCs as an HER cocatalyst on BaLa4Ti4O15 by modifying the method shown in Fig. 16A. Unfortunately, the precise synthesis of Rh2−xCrxO3 NCs has not been reported. Therefore, in this experiment, a complex containing Rh2(SG)2 as the main component was used as a precursor. First, Rh2(SG)2 was adsorbed on BaLa4Ti4O15 covered with a Cr2O3 film, and then the resulting photocatalyst was calcined. These operations resulted in loading of Rh2−xCrxO3 NCs on the photocatalyst (Fig. 18A). Various structural analyses showed that about six Rh2(SG)2 complexes aggregated during adsorption and that the Rh and Cr2O3 films formed solid solutions during calcination (Fig. 18B). Monodisperse Rh2−xCrxO3 NCs with a particle size of 1.3 ± 0.3 nm were loaded on BaLa4Ti4O15 by this method (Fig. 18C). The obtained photocatalyst exhibited an apparent quantum yield of 16% under 270 nm excitation, which is the highest achieved for BaLa4Ti4O15 to date (Fig. 18D). These results indicate that loading Rh2−xCrxO3 NCs by this method is an effective approach to improve the water-splitting activity of the BaLa4Ti4O15 photocatalyst. In principle, this method is applicable to the other photocatalysts. Rh2−xCrxO3 has already proved a useful cocatalyst in many water-splitting photocatalysts.204–207,209,210 In the future, it is expected that high quantum yields can be achieved for many water-splitting photocatalysts using this technique.
Fig. 18 (A) Schematic of the experimental procedure to form (a) BaLa4Ti4O15, (b) Cr2O3/BaLa4Ti4O15, (c) Rh–SG/Cr2O3/BaLa4Ti4O15, (d) Rh2−xCrxOy NCs/BaLa4Ti4O15, and (e) Rh2−xCrxO3 NCs/BaLa4Ti4O15. Rh2−xCrxOy NCs indicates Rh2−xCrxO3 NCs including highly oxidized Cr (>+3). (B) Elemental mapping line analysis by STEM measurements for Rh2−xCrxOy NCs/BaLa4Ti4O15. (C) TEM image of Rh2−xCrxOy NCs/BaLa4Ti4O15. The red circles indicate the Rh2−xCrxO3 NCs. (D) Comparison of gas evolution rates over different photocatalysts (a) Rh2−xCrxO3 (1.3 nm) NCs/BaLa4Ti4O15 (0.09 wt% Rh and 0.10 wt% Cr), (b) Cr2O3/Au25 NC/BaLa4Ti4O15 (0.10 wt% Au and 0.50 wt% Cr), (c) NiOx/Ni NPs/BaLa4Ti4O15 (0.50 wt% Ni), and (d) Rh2−xCrxO3 (3.0 nm) NPs/BaLa4Ti4O15 (0.10 wt% Rh and 0.15 wt% Cr). Reproduced with permission from ref. 208. Copyright 2020 Wiley-VCH. |
In the above research, the Cr2O3 shell on the cocatalyst surface suppressed the ORR, which is one of the reverse reactions of water splitting. When Pt NPs were used as a cocatalyst, the hydrogen oxidation reaction (HOR), which is also a reverse reaction of water splitting, proceeds simultaneously with the HER. In 2013, Wang and colleagues showed that platinum oxide (PtO) NCs could suppress the HOR.211 In their study, UV-responsive TiO2 was used as the photocatalyst and PtO or Pt NCs were used as the HER cocatalyst. To load PtO NCs on the photocatalyst, TiO2{001} nanosheets and poly(methacrylic acid) were dispersed in an aqueous solution containing chloroplatinic acid. Fine PtO NCs were loaded on TiO2 by injecting an aqueous solution of NaBH4 into the vigorously stirring reaction solution to form PtO NCs/TiO2.212 The polymer in PtO NCs/TiO2 was removed by washing the sample with ethanol several times. Pt NCs were loaded on TiO2 to form Pt NCs/TiO2 using the same mixing method without polymer. Scanning TEM (STEM) images revealed that the loaded PtO NCs (Fig. 19A) and Pt NCs had particle diameters of 1.0 ± 0.3 and 2.0 ± 0.5 nm, respectively. The HER and HOR activities of the obtained photocatalysts were evaluated (Fig. 19B). It was revealed that PtO NCs/TiO2 showed high HER activity and suppressed HOR activity. Similar HOR suppression was not observed for Pt NCs/TiO2 (Fig. 19B). To clarify the reasons for this difference, they conducted density functional theory (DFT) calculations using Pt8O8/TiO2 and Pt12/TiO2 as models (Fig. 19C). The results revealed that the reaction between H and O occurred easily on Pt12/TiO2 because the adsorption energies of Pt–H and Pt–O were large. Conversely, for Pt8O8/TiO2, it was shown that it was difficult for the HOR to occur because the adsorption energies of H and O were smaller than those of Pt12/TiO2 (Fig. 19D). In this study, because PtO NCs were synthesized and loaded on TiO2 in a single reaction system, it is difficult to judge whether size-controlled PtO NCs were synthesized and subsequently loaded on TiO2 or the size of PtO NCs loaded on TiO2 was controlled by the coexistence of polymer. In addition, the number of constituent atoms of the PtO and Pt NCs was not controlled with atomic precision. However, this study is important because it revealed a different approach to inhibit reverse reactions of water splitting.
Fig. 19 (A) Representative STEM image clearly showing isolated three-dimensional PtO NCs (bright spots). (B) (a) Reaction time profiles of the HOR with H2 and O2 on PtO NCs/TiO2 and Pt NCs/TiO2 photocatalysts under UV-visible light irradiation (λ > 300 nm). (b) H2 evolution and undesirable oxidation in methanol aqueous solution during three cycles of light irradiation (λ > 300 nm, 2 h) followed by dark conditions (light off, 2 h) on PtO NCs/TiO2 and Pt NCs/TiO2. (C) (a) Structure of the anatase TiO2(001) surface. (b) Optimized Pt8O8 NCs adsorbed on the TiO2(001) surface (Pt8O8 NCs/TiO2) and (c) optimized Pt12 NCs/TiO2. (d) Transition state structure of H*–H* coupling on Pt8O8 NCs/TiO2 in the liquid phase, which contained two layers of water molecules above the Pt8O8 NCs. (e) Standard Gibbs free energy profile of the HER in aqueous solution on Pt8O8 NCs/TiO2. (f) Standard Gibbs free energy profile of H2 reacting with O2 on Pt8O8 NCs/TiO2 and Pt12 NCs/TiO2 surfaces in the gas phase. Dark blue, gray, white, and red balls represent Pt, Ti, H, and O atoms, respectively. (D) Both PtO NCs and m-Pt NCs cocatalysts acted as H2 evolution sites on the host photocatalyst surface. The undesirable H2 reverse-reaction was suppressed by the PtO NC cocatalyst but facilitated by the Pt NC cocatalyst. Reproduced with permission from ref. 211. Copyright 2013 Nature Publishing Group. |
In this way, the use of metal NCs with controlled fine size as a cocatalyst can induce an increase in water-splitting activity because of their large surface area. Such precisely controlled cocatalysts make it possible to obtain a deep understanding of the effects of cocatalyst size and heteroatom doping on photocatalyst activity, as well as the origins of these effects. The findings obtained from these studies are expected to lead to clear design guidelines for the development of highly active and stable water-splitting photocatalysts.
Semiconductor | Cocatalyst | Light source | Reactant solution | Ref. | ||
---|---|---|---|---|---|---|
Kinds | Loading amount | Size | ||||
TiO2 | Pt NPs {100} | 0.5 wt% | 4.7 ± 1.6 nm | 300 W Xe lamp (λ > 420 nm) | 15% (v/v) TEOA aq. with Eosin Y | 214 |
Pt NPs {100/111} | 0.5 wt% | 6.0 ± 1.5 nm | ||||
Pt NPs {111} | 0.5 wt% | 6.5 ± 2.0 nm | ||||
TiO2 | Au NPs {100} | 0.5 wt% | 9.2 nm | 300 W Xe lamp (λ > 420 nm) | 15% (v/v) TEOA aq. with Eosin Y | 215 |
Au NPs {100/111} | 0.5 wt% | 9.5 nm | ||||
Au NPs {111} | 0.5 wt% | 8.5 nm | ||||
g-C3N4 | Rh NPs | 0.25 wt% | 4.1 ± 0.8 nm | Xe lamp (420 < λ < 630 nm) | 10% methanol aq. | 216 |
0.23 wt% | 5.9 ± 1.1 nm | |||||
0.34 wt% | 7.8 ± 1.4 nm | |||||
0.27 wt% | 9.1 ± 1.7 nm | |||||
g-C3N4/Al2Si2O5(OH)4 | Ni(OH)2 NPs | 1 wt% | <100 nm | 300 W Xe lamp (λ > 400 nm) | 10 vol% methanol aq. | 217 |
CdS | Pt–Pd NPs {100} (Pt:Pd = 2:1) | 0.5 wt% | 7.9 nm | 300 W Xe lamp (λ > 420 nm) | 1.0 M (NH4)2SO3 aq. | 218 |
Pt–Pd NPs {111} (Pt:Pd = 2:1) | 0.5 wt% | 5.2 nm | ||||
TiO2 nanotubes | CuO NPs (A–C) | Cu/Ti ratio: 9 atom% | 3 nm | 400 W Hg lamp | 10 vol% methanol aq. | 219 |
CuO NPs (WI) | Cu/Ti ratio: 9.6 atom% | |||||
CdS | Pt NCs | 1 wt% | 0.9 ± 0.1 nm | 300 W Xe lamp (λ > 420 nm) | Na2S and Na2SO3 aq. | 220 |
Pt NPs | 1 wt% | 5 nm | ||||
g-C3N4 | Ag25 | 0.2 wt% | 1.5 ± 0.25 nm | Visible light (λ > 420 nm) | 10 vol% TEOA aq. | 221 |
PtAg24 | 0.2 wt% | 1.5 ± 0.25 nm | ||||
g-C3N4 | Au25 | 0.18 wt% | Visible light (λ > 420 nm) | 10 vol% TEOA aq. | 224 | |
0.49 wt% | ||||||
0.96 wt% | ||||||
ZIF-8/TiO2 | Au25 Re complex | 1.2 nm | 300 W Xe lamp (λ > 420 nm) | H2O (1 mL), TEOA (1 mL), acetonitrile (4 mL), [Ru(bpy)3]Cl2·6H2O (10.0 μmol) | 225 | |
Au NPs Re complex | ∼10 nm |
Fig. 20 (A) TEM images of (a) Pt{100}, (b) Pt{100/111}, and (c) Pt{111} NPs, respectively. The insets are schematics of the corresponding Pt NPs. (B) H2 evolution rates of Eosin Y (4.0 × 10−4 M)-sensitized Pt{100} NPs/TiO2, Pt{100/111} NPs/TiO2, and Pt{111} NPs/TiO2 photocatalysts from 100 mL of 15% (v/v) TEOA aqueous solution under visible light irradiation (λ > 420 nm). (C) Apparent quantum yield of the HER for the above three photocatalysts. Light source: 300 W Xe lamp with either a cut-off filter of 420 nm or band-pass filter. (D) Schematic of the different energy levels of Pt{100} (solid curve) and Pt{111} (dotted curve) facets. Reproduced with permission from ref. 214. Copyright 2013 American Chemical Society. |
In 2014, Lu's group also conducted similar research using Au NPs.215 In this experiment, first, cubic-shaped Au{100} NPs, truncated cubic-shaped Au{100/111} NPs, and octahedral-shaped Au{111} NPs were synthesized by liquid-phase reduction. Each type of Au NPs was adsorbed on TiO2 by stirring the Au NPs and TiO2 in water. A series of photocatalytic HER activity measurements revealed that the photocatalyst using Au{111} NPs as a cocatalyst (denoted as Au{111} NPs/TiO2) possessed the highest HER activity of the samples. Photoluminescence measurements and model calculations indicated that Au{111} NPs/TiO2 showed high HER activity because of the same reasons as Pt{111} NPs/TiO2.
As shown in the above examples, to control the cocatalyst, it is very important to select an appropriate element, refine the particle size, improve the dispersibility, form an alloy, form a shell to suppress reverse reactions of water splitting, and control shell thickness. In 2014, Hensen et al.216 showed that the valence state of the Rh NP surface had also a large effect on the HER activity of a complex system composed of Rh NPs and visible light-responsive g-C3N4 photocatalyst. They synthesized PVP–Rh NPs with sizes of 4.1 ± 0.8, 5.9 ± 1.1, 7.8 ± 1.4, and 9.1 ± 1.7 nm and adsorbed them on g-C3N4 to form PVP–Rh NPs/g-C3N4 (Fig. 21A). Then, PVP on the Rh NP surface was removed by ozone oxidation. In general, removal of the protective ligand induces an increase in HER activity. However, the Rh NPs/g-C3N4 samples obtained after ozone oxidation showed lower HER activity than PVP–Rh NPs/g-C3N4 (Fig. 21B). When the samples after ozone oxidation were calcined under H2 flow, their activity increased. These results indicated that the activity decreased when the surface of the Rh NP cocatalyst was oxidized by ozone, whereas the activity increased when the Rh NPs were reduced by calcination under flowing H2. Thus, it was clarified that the high metallicity of the cocatalyst surface is important to improve the HER activity of Rh NPs/g-C3N4 photocatalysts.
Fig. 21 (A) Representative TEM images and particle size distributions of (a) PVP–Rh (4.1 nm) NPs/g-C3N4, (b) PVP–Rh (5.9 nm) NPs/g-C3N4, (c) PVP–Rh (7.8 nm) NPs/g-C3N4 and (d) PVP–Rh (9.1 nm) NPs/g-C3N4. (B) Photocatalytic hydrogen production rates as a function of time for (a) pure g-C3N4, (b) PVP–Rh (4.1 nm) NPs/g-C3N4, (c) PVP–Rh (5.9 nm) NPs/g-C3N4, (d) PVP–Rh (7.8 nm) NPs/g-C3N4 and (e) PVP–Rh (9.1 nm) NPs/g-C3N4. Reproduced with permission from ref. 216. Copyright 2014 Elsevier Ltd. |
Recently, Hojamberdiev et al.217 also studied the HER activity of g-C3N4. In their study, Ni(OH)2, which is a relatively inexpensive HER cocatalyst, and halloysite (Al2Si2O5(OH)4) as a hole trapping agent were co-loaded on g-C3N4 (Fig. 22A). The HER activity of the obtained photocatalyst was measured in an aqueous solution containing 10 vol% methanol as a sacrificial agent. The HER activity of the catalyst loaded with 1 wt% Ni(OH)2 was about 40 times higher than that of g-C3N4 alone. This indicates that co-loading Ni(OH)2 and halloysite is an effective approach to improve the HER activity of g-C3N4. Ni(OH)2 loaded on a photocatalyst promoted charge separation by capturing photoexcited electrons, and the negative charges on the halloysite surface captured holes generated by photoexcitation. This behavior led to the extremely high HER activity of the catalyst co-loaded with Ni(OH)2 and halloysite (Fig. 22B). A model calculation was performed to evaluate the adsorption affinity of water and methanol molecules on the catalyst surface. The results showed that the combination of g-C3N4, halloysite, and Ni(OH)2 promoted the adsorption of water and methanol on the cocatalyst surface.
Fig. 22 (A) TEM and high-resolution TEM images of (a) g-C3N4, (b) Ni(OH)2, (c) halloysite, and (d) Ni(OH)2/g-C3N4 halloysite nanocomposites prepared with 1 wt% Ni(OH)2. (B) Schematic representation of the separation and transfer of photogenerated charge carriers involved in photocatalytic H2 evolution over the developed nanocomposite. Reproduced with permission from ref. 217. Copyright 2019 Elsevier Ltd. |
In the above research on HER cocatalysts, mono-metal NPs were used. In 2016, Yao and colleagues reported a study using alloy NPs as a cocatalyst.218 In this study, visible light-responsive CdS was used as the photocatalyst and Pt–Pd alloy NPs were used as the cocatalyst. The loaded Pt–Pd NPs were Pt–Pd nanocubes with an {100} crystal plane (Fig. 23A) and Pt–Pd nano-octahedra with the {111} crystal plane. Both types of NPs were synthesized with multiple compositions (Pt:Pd = 1:1, 2:1, 3:1, and 1:2). The HER activity of the obtained photocatalysts was measured in aqueous solutions containing ammonium sulfite as a sacrificial agent. The catalyst loaded with Pt–Pd NPs with a Pd:Pt ratio of 2:1 showed the highest activity. This finding was attributed to the optimized hydrogen adsorption/desorption energy on the cocatalyst surface with a Pt:Pd ratio of 2:1. In addition, comparison of cocatalysts with different crystal planes revealed that the Pt–Pd nanocubes with the {100} crystal plane had higher activity than that of the Pt–Pd nano-octahedra with the {111} crystal plane (Fig. 23B). The {100} crystal plane has a lower atomic density than that of the {111} crystal plane. They interpreted that the Pt–Pd nanocubes showed higher activity than the Pt–Pd nano-octahedra because the electron transfer from the photocatalyst to the cocatalyst occurred more efficiently on the {100} crystal plane than on the {111} crystal plane due to the lower atomic density of the former than the latter.
Fig. 23 (A) (a) TEM and high-resolution TEM images and (b) particle size distribution of Pt–Pd (1:1) alloy nanocubes (Pt–Pd (1:1) nanocubes). (B) (a) Irradiation time course for H2 evolution over Pt–Pd (1:1) nanocubes/CdS, Pt–Pd (1:1) nano-octahedra/CdS, and bare CdS photocatalysts. (b) Photocatalytic turnover frequencies (TOF) of Pt–Pd nanocubes/CdS and Pt–Pd nano-octahedra/CdS photocatalysts. Reproduced with permission from ref. 218. Copyright 2016 American Chemical Society. |
The shape and crystal plane of the photocatalyst also affect HER activity. In 2011, Sun et al.219 reported that the HER activity of TiO2 was improved when its morphology was modified to form cylindrical TiO2 nanotubes (TNT) (Fig. 24A(a)). In this study, CuO NPs were used as an HER cocatalyst. CuO NPs were loaded on TiO2 and TNT by adsorption and calcination (denoted as A–C; this is the same as liquid-phase adsorption) using Cu(NO3)2 as a precursor to form CuO NPs(A–C)/TiO2 and CuO NPs(A–C)/TNT, respectively (Fig. 24A(b)). For comparison, CuO NPs were also loaded on TiO2 and TNT by wet impregnation (WI) to give CuO NPs(WI)/TiO2 and CuO NPs(WI)/TNT, respectively (Fig. 24A(c)). The HER activity of this series of photocatalysts was measured in an aqueous solution containing methanol as a sacrificial agent. The results revealed that the HER activity of the photocatalysts increased in the following order: CuO NPs(A–C)/TNT > CuO NPs(WI)/TNT > CuO NPs(A–C)/TiO2 > CuO NPs(WI)/TiO2 (Fig. 24B). CuO NPs(A–C)/TNT showed the highest activity, which was higher than that of the corresponding photocatalyst using noble metal Pt/Ni NPs as a cocatalyst. Because TNTs are cylindrical, they have a high specific surface area (Fig. 24C). In addition, adsorption and calcination loaded cocatalyst NPs with higher dispersibility than that in the case of wet impregnation, and NP aggregation was less likely to occur in the former samples than in the latter. The high HER activity of CuO NPs(A–C)/TNT was ascribed to these factors.
Fig. 24 (A) High-resolution TEM images of (a) TNT, (b) TNT-A–C, and (c) TNT-WI photocatalysts. (B) Time courses of H2 evolution over the photocatalysts in (A) under irradiation. Inset: average H2 evolution rates over 5 h of reaction. (C) Schematic diagram of charge transfer in the CuO NPs/TNT photocatalyst under irradiation. Reproduced with permission from ref. 219. Copyright 2011 Elsevier Ltd. |
Fig. 25 (A) TEM image of 5 wt% Pt NCs loaded on Al2O3. The inset is the particle size distribution determined from the TEM image, showing an average Pt NC diameter of 0.9 ± 0.1 nm. (B) Rate of H2 evolution of CdS under different conditions. (C) Contours of electrostatic surface potential of Pt38 NC/CdS on cutting planes that are normal to the (b) z-axis and (c) x-axis, as highlighted in blue in the structural model in (a). z- and x-axes are along the [110] and [0001] directions, respectively. Potential energies are in eV. Reproduced with permission from ref. 220. Copyright 2015 The Royal Society of Chemistry. |
In the above study, although the cocatalyst (∼0.9 nm) was in the size range of NCs, it is not clear whether the number of constituent atoms was controlled with atomic precision. In recent years, there have been several studies using NCs with atomic precision as cocatalyst precursors. In 2017, Yang et al.221 studied the HER activity of photocatalyst systems using [Ag25(SPhMe2)18](PPh4) (SPhMe2 = 2,4-dimethylbenzenethiolate; PPh4 = tetraphenylphosphine) and [PtAg24(SPhMe2)18](PPh4)2 as cocatalyst precursors. [Ag25(SPhMe2)18](PPh4) and [PtAg24(SPhMe2)18](PPh4)2 were synthesized by the methods reported by Bakr et al.222 and Zhu et al.,223 respectively. Each type of NC was adsorbed on g-C3N4 by stirring for 12 h in a toluene/dichloromethane mixture with g-C3N4. The samples were calcined at 150 °C under flowing Ar for 2 h to yield Ag25 NC/g-C3N4 and PtAg24 NC/g-C3N4. Both types of NCs had a particle size of about 1 nm. The removal of the ligands was confirmed by photoelectron spectroscopy. X-ray photoelectron spectroscopy and X-ray absorption fine structure analyses confirmed that Pt and Ag were in a zero-valent oxidation state in the NCs loaded on the photocatalyst. The photocatalysts were dispersed in an aqueous solution containing TEOA as a sacrificial agent and irradiated with visible light (>420 nm) to induce the HER. It was found that PtAg24 NCs/g-C3N4 generated four times as much hydrogen as Ag25/g-C3N4. The HER activity of PtAg24 NC/g-C3N4 was 330 times higher than that of g-C3N4 without a cocatalyst (Fig. 26A and B). Open-circuit voltage decay (OCVD) measurements revealed that PtAg24 NC/g-C3N4 had a longer carrier lifetime than that of Ag25 NC/g-C3N4 (Fig. 26C). The high HER activity of PtAg24 NC/g-C3N4 was ascribed to effective trapping of the photoexcited electrons by Pt, which suppressed the recombination of electrons and holes (Fig. 26D). Thus, electron transfer was promoted by substituting one Ag atom of Ag25 with Pt, thereby improving the HER activity of the resulting NC-loaded photocatalyst.
Fig. 26 Photocatalytic H2 evolution performance. (A) Comparison of the photocatalytic H2 evolution activities of g-C3N4, Ag NPs/g-C3N4, Ag25 NC/g-C3N4, and PtAg24 NC/g-C3N4. (B) Cycling runs of photocatalytic H2 evolution over g-C3N4, Ag25 NC/g-C3N4, and PtAg24 NC/g-C3N4. (C) Transient OCVD measurements. The inset shows the average lifetimes of the photogenerated carriers (τn) obtained from the OCVD measurements. (D) Proposed photocatalytic H2 evolution mechanism. Reproduced with permission from ref. 221. Copyright 2017 The Royal Society of Chemistry. |
In 2018, Fang et al.224 also succeeded in improving the HER activity of g-C3N4 using precisely controlled metal NCs as a cocatalyst. They used Au25(Cys)18 with L-cysteine (Cys) as a ligand as a cocatalyst. First, Au25(Cys)18 (0.18, 0.49, or 0.96 wt%) and g-C3N4 were stirred in ethanol for 1 h to adsorb Au25(Cys)18 on g-C3N4. The obtained series of photocatalysts was dispersed in aqueous solutions containing TEOA as a sacrificial agent and then the HER was induced by irradiating the aqueous solution with visible light (>420 nm). The photocatalyst with 0.96 wt% Au25(Cys)18 exhibited the highest HER activity of the samples (Fig. 27A and B). Photoelectrochemical and photoluminescence measurements showed that the photocatalyst loaded with 0.96 wt% Au25(Cys)18 promoted charge separation to the greatest extent of the catalyst series. Based on these results, they concluded that the adsorption of Au25(Cys)18 not only provided HER active sites, but also induced effective interfacial charge transfer between the cocatalyst and g-C3N4 (Fig. 27C and D).
Fig. 27 (A) Plots of photocatalytic H2 production under visible-light irradiation by g-C3N4 (GCN), Au25(Cys)18 NCs/GCN, Au NPs/GCN, and Pt NPs/GCN photocatalysts. (B) Rate of H2 evolution by various Au25(Cys)18 NCs/GCN photocatalysts under visible light. (C) Energy diagram and (D) schematic diagram of photocatalytic H2 evolution by Au25(Cys)18 NCs/GCN photocatalysts. Reproduced with permission from ref. 224. Copyright 2018 American Chemical Society. |
Thus, the precisely controlled NCs function as an effective HER cocatalysts. In addition, as described in Section 4.2, when precisely controlled NCs are used as a HER cocatalyst, it is easy to analyze the structure of the cocatalyst and clarify the origin of the improved activity. To utilize the advantages of such precise NCs, it is necessary to suppress aggregation of the cocatalyst on the photocatalyst surface. Section 4.2 revealed that the formation of a Cr2O3 shell suppressed both cocatalyst aggregation and the reverse reaction of water splitting. In 2019, Shi et al.225 revealed that encapsulating precisely controlled NCs with metal–organic frameworks (MOFs) can also suppress the aggregation of NCs. In this study, Au25 was used as a precisely controlled NC and ZIF-8 was used as the MOF. First, Au25(SG)18 (∼1.2 nm) was encapsulated into ZIF-8 by a coordination-assisted self-assembly strategy. Next, TiO2 crystals were grown on the ZIF-8 surface by a hydrothermal method. Finally, a rhenium (Re) complex was adsorbed on the surface of the TiO2 crystals to form ZIF-8/Au25(SG)18 NC/TiO2-ReP with an Au content of 0.5 wt% (Fig. 28A). In this photocatalyst, Au25(SG)18 was used as an HER cocatalyst and the Re complex was used as a cocatalyst for the CO2 reduction reaction. TEM measurements confirmed that during the preparation of ZIF-8/Au25(SG)18 NC/TiO2-ReP, the aggregation of Au25(SG)18 NC hardly occurred. The obtained photocatalyst was dispersed in an aqueous solution containing TEOA as a sacrificial agent and irradiated with visible light (>420 nm) to induce the HER and CO2 reduction reaction. The HER activity of ZIF-8/Au25(SG)18 NC/TiO2-ReP was about twice that of the case using Au NPs (∼10 nm) as the HER cocatalyst. The turnover frequency of ZIF-8/Au25(SG)18 NC/TiO2-ReP was 87 h−1 (based on an Au content of 0.5 wt%) and its apparent quantum yield at 500 nm was 2.06%. No decrease in activity was observed even when the light irradiation was continued for 7.5 h. In addition, high-angle annular dark field-STEM-energy-dispersive X-ray spectroscopy measurements indicated that the monodispersity of the Au25(SG)18 NC in ZIF-8/Au25(SG)18 NC/TiO2-ReP was maintained after the photocatalytic reaction (Fig. 28B). These results demonstrated that encapsulating the cocatalyst with an MOF is also an effective way to produce stable and highly active photocatalysts.
Fig. 28 (A) Fabrication of ZIF-8/Au25(SG)18 NC/TiO2-ReP and ZIF-8/Au NPs/TiO2-ReP composite photocatalysts by hydrothermal growth of a TiO2 shell followed by grafting of RePH molecules, and the photocatalytic processes of CO2 reduction and H2 generation over ZIF-8/Au25(SG)18 NC/TiO2-ReP under visible-light irradiation. (B) TEM images of (a) ZIF-8/Au25(SG)18 NC, (b) ZIF-8/Au25(SG)18 NC/TiO2, (c) ZIF-8/Au25(SG)18 NC/TiO2-ReP, (a′) ZIF-8/Au NPs, (b′) ZIF-8/Au NPs/TiO2, and (c′) ZIF-8/Au NPs/TiO2-ReP, and (d) high-resolution TEM image and (e) elemental maps of ZIF-8/Au25(SG)18 NC/TiO2-ReP. Green: Ti, purple: O, red: Au, pink: Re, yellow: Zn. Reproduced with permission from ref. 225. Copyright 2019 Royal Society of Chemistry. |
Fig. 29 (A) TEM image of as-prepared Co3O4 NPs. (B) Wavelength dependence O2 evolution rate over 3 wt% Co3O4 NPs/g-C3N4. The inset is the O2 evolution curves under visible light (λ > 455 nm). Reproduced with permission from ref. 227. Copyright 2017 Royal Society of Chemistry. |
Semiconductor | Cocatalyst | Light source | Reactant solution | Ref. | ||
---|---|---|---|---|---|---|
Kinds | Loading amount | Size | ||||
g-C3N4 | Co3O4 NPs | 3 wt% | ∼3 nm | 300 W Xe lamp (λ > 455 nm) | 0.01 M AgNO3 aq. | 227 |
Thus, spectroscopy, SPM, and DFT calculations are powerful tools for elucidating the phenomenon occurred at the interface of NPs/NCs-photocatalysis. In addition to these methods, electrochemical measurements are also useful for predicting the appropriate cocatalysts121,122,244 because the overpotential obtained by electrochemical measurements are strongly related to the activation energy of each reaction. At present, these measurements and analyses have not been necessarily conducted for the photocatalysts described in Section 4. However, these measurements and analyses should be conducted also for the studies using the advanced water-splitting photocatalysts (Section 4). The results would provide a deeper understanding of their interfaces and enable us to predict the appropriate elements, size, and geometric structure of the cocatalysts suitable for obtaining highly active photocatalysts.
(i) The use of precise synthesis and structure control techniques (nanotechnology) established in the fields of colloid, NP, and NC chemistry are effective for controlling the cocatalysts of water-splitting photocatalysts to obtain highly active water-splitting photocatalysts. At present, the kinds of ligands of the NPs/NCs and the photocatalysts used for liquid-phase adsorption method are limited to several species because this type of study just started in recent years. However, considering the mechanism (Section 3.2), other ligand-protected metal NPs/NCs245 and photocatalysts131 seem to be also applicable to the liquid-phase adsorption method.
(ii) To control a cocatalyst, it is very important to select appropriate elements, refine the particle size, improve dispersibility, form an alloy, form a shell with controlled thickness that suppresses the reverse reaction, expose a suitable crystal face, and control the charge state.
(iii) The formation of a shell that suppresses the reverse reaction is effective to enhance both the activity and stability of the cocatalyst.
(iv) Although the above-mentioned approaches provide an overall guidance to enhance cocatalyst activity, the most effective means to achieve high activation differs slightly depending on the combination of photocatalyst and cocatalyst.
(v) In one-step photoexcitation systems for overall water splitting, control of both HER and OER cocatalysts is effective for enhancing the overall activity of water-splitting photocatalysts.
(vi) For Z-scheme catalysts, reduction of the mediator may be rate-determining. Thus, control of the cocatalyst that promotes mediator reduction is an effective approach to increase the rate of this reaction.
(vii) Using precisely controlled NC cocatalysts provides a deeper understanding of geometrical structure and electronic state of the cocatalyst, and thereby greater knowledge of the main factors influencing photocatalyst activation.
These findings are expected to be useful for researchers in the field of water-splitting photocatalysis, as well as those interesting in the applications of controlled NPs/NCs.
(i) Close collaboration between photocatalyst chemists and NP/NC chemists. To enhance the activity of water-splitting photocatalysts, it is necessary to improve both the semiconductor photocatalyst and cocatalyst. Regarding semiconductor photocatalysts, photocatalysis researchers would continue to make substantial improvements. On the other hand, the recent studies revealed that the use of precise synthesis and structure control techniques established in the fields of colloid, NP, and NC chemistry is very effective at raising cocatalyst activity. Therefore, researchers in the field of NPs/NCs should also conduct thorough research on enhancing the activity of cocatalysts. Effective approaches to maximize activation depend on the specific combination of photocatalyst and cocatalyst. If NCs/NPs researchers continue to conduct cocatalyst research using commercial photocatalysts, the results would not lead to the high activation of the most advanced photocatalysts. To develop highly active water-splitting photocatalysts for practical use, researchers in both fields need to collaborate more closely in future research.
(ii) Selective loading of cocatalysts on the optimal crystal plane. The crystal planes of semiconductor photocatalysts show different transport rates of photogenerated electrons and holes. If HER and OER cocatalysts could be selectively loaded on crystal planes with rapid electron and hole transport, electrons and holes could be efficiently used in the relevant reactions. In fact, research on photocatalysts in which a cocatalyst was loaded by photodeposition demonstrated that such selective cocatalyst deposition is an effective approach to increase activation.246 In the case of photodeposition, the HER cocatalyst is formed by reduction of metal ions and the OER cocatalyst is formed by oxidation of metal ions. Therefore, each cocatalyst is inevitably loaded on a crystal plane where electrons and holes can easily reach the active sites. Even for colloidal NPs/NCs, if such crystal plane-selective loading can be achieved, higher activity should be obtained compared with that provided by the present method. One of the methods seems to adsorb the colloidal metal NPs/NCs on the photocatalyst after a particular crystal-plane which is undesired site are selectively covered with a surfactant.
(iii) Control of the geometrical structure of loaded NPs/NCs. The geometrical structure of loaded NPs/NCs appears to show some variation. In the case of NPs, the size distribution occurs at the synthesis stage. In the case of NCs, there is no variation in size (chemical composition) and geometrical structure at the synthesis stage, but the NCs after loading sometimes seem to have varied geometrical structure. To identify the geometrical structures of NPs/NCs that lead to high activity and selectively load NPs/NCs with such geometrical structures as cocatalysts, it is necessary to establish a new method for controlling the geometrical structure of the loaded NPs/NCs. For NPs/NCs dispersed in solution, post-treatment methods for size247 and structure convergence55 have been reported. In the future, it is expected that post-treatment to achieve size and structure convergence will also be established for loaded NPs/NCs, which would realize the loading of NPs/NCs with controlled geometrical structure.
(iv) Elucidation of the calcination process. Protective organic molecules on colloidal metal NPs/NCs are often removed by calcination. However, the presence or absence of protective organic molecules is generally determined only by X-ray photoelectron spectroscopy and/or X-ray absorption fine structure measurements. In the case of NCs, although NCs are synthesized while controlling the chemical composition at the atomic/molecular level, ambiguity remains about the elimination of the protective ligands. For ultrafine NCs, the remaining protective ligands should strongly affect the electronic/geometrical structure of the NCs. To obtain reliable design guidelines for improving the activity of cocatalysts, it is necessary to determine whether or not the protective ligands remain on the NCs at the atomic/molecular level. Thus, it is necessary to gain a deep understanding of the calcination mechanism and to further improve the structural analysis techniques for loaded NCs. If the calcination mechanism can be clarified at the molecular level, it will not be necessary to consider the protective ligands remaining after calcination and will also be possible to conduct calcination at an appropriate temperature to suppress cocatalyst aggregation on the photocatalyst. This understanding is also necessary for better control of the geometrical structure of loaded NCs.
(v) Utilization of various metal NPs/NCs. In recent years, it has become possible to synthesize various monodisperse NPs/NCs with controlled size and structure. Inorganic chemists can now synthesize metal NCs as precisely as organic chemists synthesize organic molecules, and the types of synthesized NCs are steadily increasing. However, the types of NPs/NCs used in the study of water-splitting photocatalysts are limited. In the future, more of the existing NPs/NCs should be used in water-splitting photocatalysis research. This would expand knowledge of the major factors affecting cocatalyst activity and increase the possibility of producing photocatalysts with high water-splitting activity.
(vi) Theoretical calculations of real systems. For understanding the main factors influencing the activity, theoretical research using the actual geometrical structures is required. At present, deep understanding of the chemical composition and geometrical structure of loaded NPs/NCs and their loading sites on photocatalysts have not been obtained experimentally. In the future, it is expected that the structural analysis techniques for loaded NCs will be improved and thereby enable theoretical calculations to be carried out using actual geometrical structures obtained by such experiments. If such a progress would be achieved, it became possible to predict appropriate photocatalyst systems by DFT calculations.
We hope that by overcoming these current limitations of water-splitting photocatalysts, a society able to solve energy and environmental problems will emerge as soon as possible.
This journal is © The Royal Society of Chemistry 2020 |