Next Article in Journal
Effect of the Calcination Temperature of LaNiO3 on the Structural Properties and Reaction Performance of Catalysts in the Steam Reforming of Methane
Next Article in Special Issue
Synthesis of Rape Pollen-Fe2O3 Biohybrid Catalyst and Its Application on Photocatalytic Degradation and Antibacterial Properties
Previous Article in Journal
Synthesis of Hollow Mesoporous Silica Nanospheroids with O/W Emulsion and Al(III) Incorporation and Its Catalytic Activity for the Synthesis of 5-HMF from Carbohydrates
Previous Article in Special Issue
Core/Shell Nitrogen-Doped TiO2@SiO2 Nano-Catalyst as an Additive in Photocatalytic Paint for Gaseous Acetaldehyde Decomposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cocatalysts for Photocatalytic Overall Water Splitting: A Mini Review

1
International Research Center for Renewable Energy, State Key Laboratory of Multiphase Flow in Power Engineering, Xi’an Jiaotong University, Xi’an 710049, China
2
Key Laboratory of Subsurface Hydrology and Ecological Effects in Arid Region, Ministry of Education, School of Water and Environment, Chang’an University, Xi’an 710064, China
*
Authors to whom correspondence should be addressed.
Submission received: 30 December 2022 / Revised: 2 February 2023 / Accepted: 3 February 2023 / Published: 5 February 2023

Abstract

:
Photocatalyst overall water splitting is usually restricted by low carrier separation efficiency and a slow surface reaction rate. Cocatalysts provide a satisfactory solution to significantly improve photocatalytic performance. In this review, some recent advances in cocatalysts for photocatalytic overall water splitting are gathered and divided into groups. Firstly, the loading method of the cocatalyst is introduced. Then, the role of the cocatalyst applied for the photocatalytic overall water splitting process is further discussed. Finally, the key challenges and possible research directions of photocatalytic overall water splitting are proposed. This review is expected to promote research on the design of efficient cocatalysts in photocatalytic systems for overall water splitting.

1. Introduction

With the rapid development of society, greenhouse gases and pollutants released by the massive combustion of traditional fossil fuels have caused global warming and environmental pollution. Therefore, it is urgent to find renewable energy to substitute fossil energy for carbon-neutral development. Solar energy is safe, convenient, and low-cost [1,2]. In the development and utilization of solar energy, photocatalysis has attracted extensive attention [3,4,5,6,7,8]. Photocatalysis utilizes light energy as an energy source to catalyze various reactions through catalysts to satisfy different requirements. At present, the more common applications of photocatalysis include photocatalytic water splitting [4,5], photocatalytic CO2 reduction [7,9], photocatalytic organic synthesis [10], photocatalytic degradation [11,12], etc. Hydrogen, a clean and non-polluting energy with high energy density, is available for storing and transporting. More importantly, water is the only combustion product. Since Honda and Fujishima first reported the phenomenon of photoelectrochemical water splitting on TiO2 electrodes in 1972 [13], the splitting of water into hydrogen using semiconductor photocatalysts has emerged as a promising strategy to produce hydrogen fuel using sunlight and water [3,6,10,14,15,16,17]. Due to the distinctive merits of no additives and bias-free, photocatalytic overall water splitting (POWS) into hydrogen and oxygen is considered the most promising method [4,5,18,19].
In general, POWS involves three main steps (Figure 1): (1) A photon with energy greater than the bandgap energy of the photocatalytic semiconductor is absorbed by the semiconductor, and excites an electron from the valence band to the conduction band, leaving a hole in the valence band. Such electrons and holes are referred to as photogenerated electron-hole pairs or photogenerated carriers. Since photo-generation of electrons/holes relies on excitation of the semiconductor with light irradiation, photocatalysts with a wide range of light responses and good light absorption coefficient should be developed, which could be realized by tailing modification of photocatalysts via doping or solid-solution construction [20]. (2) Photogenerated carriers separate and migrate towards the surface redox-active sites. During the migration process of carriers to the surface of the photocatalyst, some electrons and holes will recombine, and only the electrons and holes that successfully reach the redox-active sites can participate in the redox reaction. As such, accelerated charge separation should be realized via building hetero/homo-junctions [15,21]. (3) Photogenerated carriers reaching the redox-active sites are consumed in the process of photocatalytic hydrogen and oxygen production. Due to the high activation energy of water splitting, not all the photogenerated carriers reaching the photocatalyst surface could participate in the water splitting reaction timely, which, consequently, results in surface recombination, finally greatly limiting the efficiency of POWS [22,23]. Therefore, various strategies, including loading cocatalysts, have been employed to boost the surface redox reaction.
For facilitating the surface redox reaction, various methods of loading cocatalysts on the photocatalyst have been developed, such as impregnation [24,25], photodeposition [26,27], and so on [28,29,30,31]. The cocatalysts supported by different loading methods exhibit variations, including loading sites, loading amounts, and particle sizes, which play a significant role in photocatalytic performance. Therefore, summarizing the various loading methods of cocatalysts and the influence on photocatalyst efficiency is highly desired for developing novel photocatalytic systems.
Appropriate cocatalysts loaded on semiconductor photocatalysts can act as reaction sites and catalyze the progress of the reaction, promoting charge separation and migration driven by the interfacial junction formed between the cocatalyst and the semiconductor photocatalyst [23,32]. Noble metals with large work function (Table 1) and suitable adsorption-desorption energy, such as Pt [33,34], Pd [35,36], Rh [27,37], Au [38,39], Ag [31,40,41], etc., have been widely applied on photocatalysts to improve photocatalytic performance. To reduce costs, many non-noble metal and non-metal cocatalysts have also been developed to replace noble metals for hydrogen production, such as Ni [42,43], Cu [44,45], Co [46], C [47,48], etc.
Oxygen production is a 4-electron process, which is considered as the rate-limiting step in the POWS reaction. Introducing an appropriate oxygen-generating cocatalyst can reduce the activation energy barrier of the oxygen-generating reaction and simultaneously accelerate the desorption of the generated oxygen, thereby achieving a higher oxygen generation rate. Common noble metal oxygen-generating cocatalysts include IrO2 [49] and RuO2 [50,51,52], while non-noble metal cocatalysts include MnOx [53,54], NiOx [55], CoOx [25,56], NiCo2O4 [29], etc.
For POWS, the reverse hydrogen-oxygen recombination reaction is an important factor hindering photocatalytic activity. In this regard, researchers have developed various cocatalysts to isolate the contact between oxygen and hydrogen production sites, such as CrOx [27,57], NiOx [58,59], etc. Another way to prevent the reverse reaction of hydrogen-oxygen recombination is by depositing hydrogen-generating and oxygen-generating cocatalysts into the different sites of the photocatalyst and successfully separating the hydrogen-generating and oxygen-generating sites to suppress the reverse reaction [26,49,60,61]. Another problem with POWS is the oxidation of photocatalysts and cocatalysts by the holes or produced oxygen, also known as photocorrosion. For this reason, researchers have developed a series of cocatalysts to inhibit photocorrosion, primarily a protective shell that keeps out oxygen [62,63,64,65].
In this review, the preparation methods of several common cocatalysts are introduced first and the advantages of various methods are then comparatively discussed to facilitate the subsequent design and preparation of cocatalysts. Focusing first on the different roles of cocatalysts, the cocatalysts for POWS are then categorized and the roles of various cocatalysts in the process of POWS are summarized. Hopefully, this paper will serve as a link for the development of promoters for POWS.

2. Cocatalysts with Various Loading Methods

Various methods of loading cocatalysts on the photocatalyst have been developed. The loading methods can be mainly divided into two categories. One is cocatalysts that are directly grown on the surface of the photocatalyst using precursors, referred to here as the direct addition-type method. The other is loading the as-synthesized cocatalyst on the surface of the photocatalyst, referred to as the composite-type method. Direct addition-type methods, such as solvothermal treatment, precipitation, immersion, photo-deposition, etc., could ensure sufficient and tight interfacial contact between cocatalyst and photocatalyst semiconductors, thereby facilitating charge separation and transfer. Compound-type methods involve synthesizing the desired cocatalyst and then loading on the photocatalyst by impregnation, grinding, physical mixing, and other methods. For Compound-type synthesis, the microstructure of the cocatalyst can be precisely controlled and the synthesis conditions are not limited by the photocatalyst semiconductor.

2.1. Direct Addition-Type Cocatalysts

Direct addition-type cocatalysts refer to those cocatalysts directly grown on the surface of photocatalyst with corresponding precursors, which contributes to the formation of a tight interface between the cocatalyst and the photocatalyst. For this reason, the cocatalyst loaded by direct addition-type methods can effectively reduce the resistance at the interface. Moreover, the cocatalyst loading site, particle size, and distribution are easily controlled via this loading method, and special structures such as core–shell structures can be formed through layer-by-layer loading. As reported by Liu et al. in 2020, the prepared C quantum dots were added to the raw materials and then used a hydrothermal method to prepare a C-loaded CoP photocatalyst. The POWS efficiency was 4 times higher than that of the physical mixture of C quantum dots and CoP [48]. The research shows that the product obtained from an in situ growth method have relatively tight chemical bonds, thus leading to better performance.

2.1.1. Impregnation

The impregnation-calcination method is one of the most common methods applied to the cocatalysts deposition on photocatalyst semiconductors (Figure 2a) [23,66]. Photocatalysts were first immersed into the precursor solution of the cocatalyst. During the following slow evaporation process of the solvent, the supersaturated cocatalyst precursor continued to precipitate and evenly disperse on the surface of the photocatalyst. The metal compounds attached to the photocatalyst surface were pyrolyzed into metal or metal oxide nanoparticles at different temperatures/atmospheres. Afterward, a photocatalyst with a uniformly supported cocatalyst was obtained. Due to the interactions within high-temperature treatment, the cocatalyst prepared by the impregnation-calcination method was tightly combined with the photocatalyst.
Domen’s group first reported that a mixed oxide cocatalyst of rhodium and chromium was supported on (Ga1−xZnx)(N1−xOx) by the impregnation method [67]. As the sources of rhodium and chromium, Na3RhCl6·2H2O and Cr(NO3)3·9H2O were loaded on the photocatalyst by immersion method. After calcination, RhCrOx particles with a diameter of 10–20 nm and good crystallization were formed, which were uniformly distributed on the surface of (Ga1−xZnx)(N1−xOx). The apparent quantum yield (AQY) of the samples at POWS achieved was 2.5% at 420–440 nm, which is about an order of magnitude higher than previously reported. The impregnation method can also be compliant with molecular cluster cocatalysts. For example, Seo et al. loaded the synthesized [MO3S4]4+ molecular clusters on the NaTaO3 surface via the impregnation method and obtained molecular cluster cocatalysts with a diameter of about 1 nm [68]. The hydrogen production rate of POWS is 28 times higher than that of pure NaTaO3.
In addition to oxide cocatalyst, the immersion calcination method can also obtain phosphide cocatalyst by changing the calcination atmosphere. Xue et al. successfully loaded Co and Ni precursor on PCN by the impregnation method and then employed NaH2PO2·H2O as P source to obtain CoxNiyP cocatalyst through a high-temperature phosphorization process [69]. The formation of the P+-Pδ-Co/Niδ+ chemical bridge between the cocatalyst and the photocatalyst has been demonstrated to significantly enhance charge transfer and separation, as well as H-H bond formation and H2 desorption. Then, the H2 production rate reached 239.3 μmol h−1 g−1 for POWS.

2.1.2. Photodeposition

Photochemical deposition has been widely applied to the load of metals and their oxides on photocatalyst semiconductors as cocatalysts (Figure 2b) [70,71,72]. A typical photodeposition process can be described as follows: The photocatalyst is excited by light to generate photogenerated electrons and holes, which migrate to the reduction and oxidation active sites on the surface of the photocatalyst, respectively. The photogenerated electrons migrated to the reduction active site will preferentially react with metal cations in solution to reduce metal cations into metal elements. Likewise, during photodeposition of metal oxide cocatalysts, metal cations are oxidized to metal oxides by photogenerated holes that migrate to the oxidation active sites. Therefore, the photodeposition method can selectively load metal cocatalysts on the active sites to shorten the photogenerated carrier migration distance and improve the carrier separation efficiency, and the particle size can be adjusted by means of metal precursor solution concentration, deposition time, different sacrificial agents, etc., which is difficult to achieve by other methods.
Domen’s group reported that the excellent POWS performance obtained by photodepositing RhCrOx on SrTiO3 surface was about 2 times as that obtained from impregnation method [72]. Subsequent studies by this group showed that the photodeposited rhodium and chromium oxides exhibited a core–shell structure with metallic Rh core and Rh(III)-Cr(III) mixed oxide as core and shell, respectively. During the experiment, the authors found that the concentration of K2CrO4 solution during the photo-deposition process could directly influence the valence state of deposited Cr. The POWS performance of the core–shell structure supported by photodeposition method is 5.8 times than that of the mixed oxide of Rh(III) and Cr(III) [71].
After the metal is deposited on the photocatalyst surface by using the photodeposition method, the metal compound can also be obtained by further processing. In our previous work, Co was deposited on the surface of Al: STO by the photodeposition method and then phosphatized at a high temperature to obtain CoP [70]. The photodeposition process makes the contact between the cocatalyst and the photocatalyst more intimate than that prepared by the impregnation method. Thus, the activity of CoP loaded on SrTiO3 by the photodeposition method was nearly 1.3 times higher than that by the impregnation method.

2.1.3. Electrostatic Adsorption and Low-Temperature Heat Treatment

Electrostatic adsorption and low-temperature heat treatment is also a common method for loading cocatalysts. In particular, this method is commonly used when it is desired to support single-atom cocatalysts. Single-atom metal loading not only maximizes the utilization of metal, but also promotes the separation and transfer of charge carriers. Xiao et al. embedded sodium copper chlorophyllin salt into the melamine-based supramolecular precursor and then the pyrolysis was controlled to form a single-atom Cu-loaded C3N4 [73]. The visible-light photocatalytic hydrogen production rate was 30 times higher than that of bulk C3N4. Melamine was hydrolyzed and self-assembled in the presence of phosphoric acid, while chlorophyll sodium copper salt was inserted into melamine due to large layer spacing. During thermal decomposition, Cu was anchored to a N atom via a Cu-Nx bond with N in C3N4 planes. As an excellent cocatalyst for hydrogen production, single-atom Pt can be loaded on the surface of g-C3N4 [74], CdS [75], and TiO2 [76] through electrostatic adsorption and low-temperature heat treatment, respectively. The corresponding photocatalytic performance was 8.6 times, 7.69 times and 591 times higher than that by nanoparticles loading, respectively. It is worth noting that the reports of single-atom cocatalysts mainly focus on photocatalytic hydrogen production and there are no relevant reports on single-atom cocatalysts to achieve POWS.

2.1.4. Others

In addition to the above-mentioned methods of loading cocatalysts, there are other direct addition-type methods such as hydrothermal treatment, controlled ink-jet printing, and so on. MoS2-RuO2 modified CdS with MoS2 at the ends and RuO2 on the sidewalls of CdS nanowires were prepared by Qiu et al. [77]. In the process of hydrothermal treatment, the amino group on the surface of CdS concentrates the precursor of MoS2 on the two ends of the CdS nanowire. When RuO2 is further loaded, the interaction between Ru and amino group loads RuO2 on the sidewalls of the nanowire. Thereby, the photocatalyst realized the separation of hydrogen and oxygen production sites. Consequently, the POWS efficiency is 10 times higher than that with unseparated sites. In 2021, A. Aguirre-Astrain et al. decorated the photocatalyst with semiconductor oxides using the controlled ink-jet printing method [28]. Firstly, the metal precursor solution was printed on the surface of the BaV2O6 film and then the metal oxide-modified BaV2O6 photocatalyst was obtained after high-temperature calcination. Ultimately, the POWS activity of CuO/CuO2-modified BaV2O6 was 30 times higher than that of pure BaV2O6.

2.2. The Composite Type of Cocatalysts

Although cocatalyst directly synthesized on the photocatalyst has the above advantages, it is necessary to ensure that the photocatalyst does not change in the process of synthesis. Therefore, the synthesis conditions of direct addition-type methods are restricted by the inherent quality of the photocatalyst. However, synthesizing the cocatalyst first and then loading it provides another choice to prepare the cocatalyst, which allows us to synthesize the desired cocatalyst without restriction.

2.2.1. Grinding

A non-noble metal Mn2Co2C@C/Mn2N0.86 composite cocatalyst composed of Mn2N0.86 and N-doped graphitized carbon-coated Mn2Co2C nanoparticles were fabricated by Zhou et al. and loaded on SrTiO3 and g-C3N4 by a mechanical grinding method [78]. During the illumination process, Mn2N0.86 was hydrolyzed and formed a MnOOH and Mn2Co2C dual cocatalyst through an in situ growing method. Thereinto, Mn2Co2C is used for hydrogen production and MnOOH is used for oxygen production; finally, POWS was achieved and the corresponding AQY reached 2.54% and 1.45% at 400 and 420 nm, respectively. The FeNi inter-metallic compound nanoparticles coated with N-doped graphitized carbon prepared by pyrolysis were also loaded on the g-C3N4 surface by grinding [79]. The secondary charge transfer in FeNi alloys can promote carrier separation and accelerate hydrogen desorption, which is 283 times as high as that of primary g-C3N4. As a result, a high AQY up to 24.78% (400 nm) was obtained.

2.2.2. Calcination

Garcia Esparza et al. used mesoporous g-C3N4 as a template and carbon source to synthesize tungsten carbide nanocrystals with different structures and compositions using tungsten precursors at different temperatures. The cocatalyst was loaded onto sodium-doped SrTiO3 by the immersion method and POWS was achieved [80]. Recently, a new hybrid cocatalyst CoOx-Mo2N was prepared by a hydrothermal method and loaded by calcining the mixture of CoOx, Mo2N and Ge3N4 in N2 flow [81]. Among them, Mo2N is used for hydrogen production and CoOx is used for oxygen production, and the combination of them realizes POWS. The activity is 1.5 times higher than that by Ge3N4 loaded with the spatial separated dual cocatalysts CoOx and Mo2N.

3. Application of Cocatalysts for POWS

In photocatalytic water splitting, cocatalysts play various roles in enhancing the photocatalytic performance of semiconductor photocatalysts (Figure 3). First, the cocatalyst can trap photogenerated carriers to enhance the carrier transfer and separation efficiency. Secondly, the cocatalyst can provide reaction sites, enhance the adsorption and desorption of surface species, and reduce the overpotential of the surface reaction, thus increasing the H2 and O2 generation rate. For POWS, some cocatalysts also have unique functions, such as inhibiting the reverse reaction of H2 and O2, or inhibiting the photocorrosion of the photocatalyst and the cocatalyst. In this section, the roles of various cocatalysts in POWS will be discussed for designing an efficient POWS system.

3.1. Promote the Water Splitting Reaction

3.1.1. Hydrogen-Production Cocatalysts

The band gap of a photocatalyst realizing POWS should not be less than 1.23 eV [82], while the potential of conduction band bottom should be more negative than H+/H2 (0 V vs. NHE at pH = 0) and the potential of valence band top should be more positive than O2/H2O (1.23 V vs. NHE at pH = 0). Beyond that, the oxidation and reduction reaction on the surface of the photocatalyst requires a certain overpotential. Even with the appropriate potential to split water, many semiconductor photocatalysts cannot produce hydrogen/oxygen without cocatalysts, which is mainly restrained by the severe carrier recombination and slow surface reaction rate. Appropriate cocatalysts loaded on semiconductor photocatalysts can act as reaction sites and catalyze the progress of the reaction, thus promoting charge separation and accelerating the reaction rate. Many materials have been developed as cocatalysts used for hydrogen production, among which NiOx and RhCrOx are widely used for POWS.
As early as 1980, Domen et al. achieved POWS by anchoring the NiO cocatalyst on SrTiO3 [83]. Through a series of comparison experiments, it was found that the NiO cocatalyst on the surface of SrTiO3 was composed of metallic Ni particles and a layer of NiO, in which hydrogen was generated on Ni and oxygen was generated on NiO [42]. Moreover, in-depth studies on the state of Ni-NiO cocatalysts in photocatalytic reactions were carried out by XPS, TEM, etc. The results show that the Ni(OH)2 on the surface will be oxidized to NiOOH when illuminated, and when the illumination is stopped, the NiOOH will undergo a disproportionation reaction to form embedded Ni(OH)2 and Ni particles [58].
RhCrOx as a cocatalyst for POWS was first reported by Maeda et al. in 2006. The RhCrOx was supported on (Ga1−xZnx)(N1−xOx) surface through the impregnation-calcination method [67]. Subsequently, they developed a photo-deposition method to prepare Rh-CrOx core–shell structure hydrogen production cocatalysts [57]. Since then, RhCrOx and Rh-CrOx core–shell structures have been widely used in various photocatalysts to achieve POWS [26,30,37,50,72,84,85,86,87,88,89,90]. The role of each component in the RhCrOx cocatalyst was explored through electron spin resonance spectral and the results show that RhOx can accelerate the hydrogen production reaction process and promote the production of superoxide radical and singlet oxygen in competition with the hydrogen production reaction, while CrOx inhibits the production of these two species [90]. Although CrOx is not able to accelerate the hydrogen production reaction, the co-action with RhOx enhances the hydrogen production performance (Figure 4a). The Al-doped SrTiO3 loaded with RhCrOx and MoOy prepared by the impregnation-calcination method realized POWS and an AQY of 69% was achieved at 365 nm [91].
A Pd@Pt core–shell structure cocatalyst was successfully prepared. Additionally, it was further deposited on TiO2 nanosheets to improve the efficiency of Pt cocatalysts on POWS by reducing the load amount of Pt, and finally obtained an enhanced performance of 7.2 times when compared with Pt-TiO2 [92]. Pd@Pt cocatalyst has improved the capabilities of charge trapping and accelerated charge separation through interfacial charge polarization, while optimizing lattice structure and enhancing the adsorption capacity of H2O, thereby enhancing the photocatalytic performance and simultaneously reducing the load amount of Pt (Figure 4b).
Other materials such as WxC [80] and RuOx [49] have also been developed for hydrogen production cocatalysts for POWS catalysts.

3.1.2. Oxygen-Production Cocatalysts

Oxygen production reaction is generally considered as the rate-limiting step due to its 4-electron process. Accordingly, increasing the oxygen production reaction rate can effectively improve photocatalytic efficiency [3,14,93,94]. Appropriate oxygen-generating cocatalyst loading can effectively reduce the activation energy and increase the oxygen-generating reaction rate [32,95,96,97,98,99].
As a common cocatalyst for oxygen generation, RuO2 has been widely used in POWS [51,52,100,101,102]. For the first time, β-Ge3N4 supported by RuO2 realized POWS for non-oxide photocatalysts (Figure 5a) [51]. Besides RuO2, Yuan et al. used the photo-deposition method to deposit MnOx on the Z-scheme photocatalyst composed of γ-MnS and Cu7S4 to achieve POWS [53]. Oxygen generation cocatalyst MnOx and hydrogen production cocatalyst PtS were selectively deposited on a ZnIn2S4/WO3 Z-scheme photocatalyst with an AQY of 0.5% at 420 nm [103].
Cobalt oxide phosphate is a common electrocatalyst for water oxidation, which was introduced into the photocatalytic system by Li et al. in 2013 [104]. The cobalt-oxide-phosphate cocatalyst was loaded on mpg-CNx using the direct photo-deposition method and impregnation-photo-assisted oxidative conversion method, which was very efficient for the photocatalytic water oxidation reaction with electron scavenger. With hole scavenger, the cobalt-oxide-phosphate will be reduced to a cobalt-oxo/hydroxy-phosphate compound, an HER electrocatalyst. In the POWS process, part of the cobalt-oxide-phosphate is converted into cobalt-oxo/hydroxy-phosphate compound, thus achieving POWS. Other Co species, such as CoOx, Co(OH)2, etc. were also used as efficient oxygen-evolution cocatalysts in the POWS system [105,106].
Figure 5. (a) SEM image of RuO2-loaded β-Ge3N4. Reproduced with permission from Ref. [51]. Copyright 2005, American Chemistry Society. (b) OER rate with AgNO3 of fabricated photocatalysts. (c) Wavelength-dependent AQY (orange dots) of Pt·Ni(OH)2/C3N4. (b,c) Reproduced with permission from Ref. [107]. Copyright 2019, Elsevier.
Figure 5. (a) SEM image of RuO2-loaded β-Ge3N4. Reproduced with permission from Ref. [51]. Copyright 2005, American Chemistry Society. (b) OER rate with AgNO3 of fabricated photocatalysts. (c) Wavelength-dependent AQY (orange dots) of Pt·Ni(OH)2/C3N4. (b,c) Reproduced with permission from Ref. [107]. Copyright 2019, Elsevier.
Catalysts 13 00355 g005
NiP was obtained by the in situ photo-oxidation Ni(OH)2 method and, subsequently, selectively photo-deposited Pt to prepare Pt clusters decorated with Ni(OH)2 nanoparticles as an oxygen-generating cocatalyst [107]. The close contact between nanosized Pt and Ni(OH)2 produces a strong Ptσ+-Oσ− interaction, which is beneficial to the adsorption of H2O, and, at the same time, greatly reduces the activation energy of the H-O-H bond and promotes oxygen production (Figure 5b), and the AQY for POWS can reach 4.2% at 420 nm (Figure 5c).

3.2. Suppressed Backward Reactions

3.2.1. Separate Hydrogen and Oxygen Production Sites

In 2012, Townsend et al. supported NiOx on the surface of SrTiO3 by the impregnation-calcination method and achieved excellent POWS performance [42]. It has been shown that NiOx is composed of Ni metal particles and NiO particles, and neither Ni or NiO can achieve POWS. NiO can be generated by photodeposition after loading Ni. The surface photovoltaics prove that Ni acts as a hydrogen production site. Comparative experiments of deposited Pt confirm that NiO is not a hydrogen-producing site. Therefore, it is concluded that Ni is a hydrogen-generating cocatalyst, and NiO is an oxygen-generating cocatalyst. The Ni and NiO are supported at different sites to prevent the reverse reaction, thereby achieving excellent POWS performance. Subsequently, the photo-deposition method was used to selectively deposit MnOx and Pt on the (110) surface and the (010) surface of BiVO4, respectively, which was acted as an oxygen-generating and hydrogen-generating cocatalyst. Consequently, an AQY of 5.5% at 420 nm at POWS was realized [53]. The oxygen-production cocatalyst reduced graphene oxide and hydrogen-production cocatalyst Ag nanoparticles were loaded on different parts of the surface of SnS2 nanosheets, respectively (Figure 6) [31]. This structure realizes the separation of hydrogen production and oxygen production sites, improves the carrier separation efficiency, and suppresses the reverse reaction. Thus, POWS was obtained under visible light irradiation.

3.2.2. Oxygen Barrier

In the core–shell structure cocatalyst, the shell can effectively prevent the contact between H2 generated on the core and O2 in the environment, thus inhibiting the occurrence of reverse reaction (Figure 7). Maeda et al. utilized a photo-deposition method to prepare Rh-Cr2O3 core–shell cocatalysts (Figure 8a) [57]. The study showed that Rh core was the hydrogen-producing site and Cr2O3 shell isolated the contact between the hydrogen-producing site and O2, preventing the reverse reaction. On this basis, a Cr2O3 shell covered on Pt by Qi et al. also played a significant role in preventing the reverse reaction [108].
Ni-NiO core–shell structure was reported to be loaded on Ta2O5, achieving POWS [59]. It is shown that Ni has excellent catalytic performance for hydrogen production, but at the same time, Ni also acts as the reverse reaction site of H2 and O2, which limits the further improvement of the POWS performance. The researchers found through experiments that H2 generated on the Ni surface can penetrate the NiO shell, while H2O and O2 cannot. This isolates the generation sites of H2 and O2, inhibits the reverse reaction, and enhances the POWS performance (Figure 8b).

3.2.3. Other Means of Inhibiting Recombination

In 2018, Wang et al. found in the POWS experiment of Pt-loaded TiO2 that halogen atoms can be adsorbed on the surface of the Pt cocatalyst, which decreases the adsorption and activation degree of hydrogen and oxygen molecules on Pt, thereby suppressing the reverse reaction (Figure 8c) [109,110].

3.3. Restrain Photocorrosion

A large number of holes and generated O2 during the POWS process will cause severe photocorrosion of photocatalyst semiconductors, which limits the commercial application of photocatalysts [27,50,63,111].

3.3.1. Restrain the Photocorrosion of Photocatalyst

Thaminimulla et al. found that depositing a small amount of Cr oxide on the surface of Ni-deposited K2La2Ti3O10 by the impregnation method could improve the POWS activity and stability of the photocatalyst [43]. Recently, an Al2O3 shell was loaded on the surface of CdS as a cocatalyst and isolated CdS from generated O2, which effectively inhibited the photocorrosion of CdS and achieved an AQY of 0.11% at 430 nm [112]. Inductively coupled plasma emission spectrometer (ICP) analysis showed that before loading Al2O3, the post-reaction solution contained a large amount of Cd2+, indicating that there was a severe photocorrosion phenomenon in the reaction. Nevertheless, after loading Al2O3, the change of Cd2+ content in the solution before and after the reaction was almost negligible, so the Al2O3 shell on CdS effectively inhibits the photocorrosion of CdS (Figure 9a,b). Similarly, the Ni2P shell loaded on the surface of CdS as a cocatalyst also achieved a similar effect and achieved an AQY of 3.89% at 430 nm for POWS (Figure 9c,d) [62]. Recently, uniform CdS nanorods protected by ultrathin NiOOH achieved POWS and exhibited good photostability over 25 h [113]. Another example is removing oxygen-production sites from a photocatalyst to a cocatalyst. CoOx was loaded as a cocatalyst providing oxygen-production sites, which could extract holes from a GaN/InGaN photocatalyst. As a consequence, impressive stability of >580 h was achieved for POWS [114,115].

3.3.2. Restrain the Photocorrosion of Cocatalyst

Zhang et al. found that the Ni@NiO core–shell structure cocatalyst deposited on the surface of TiO2 can only generate hydrogen but not oxygen in the POWS reaction [64]. The research shows that the Ni@NiO core–shell structure cocatalyst deposited on the surface of TiO2 will gradually be corroded and dissolved during the hydrogen production process, leaving the NiO shell, so POWS cannot be achieved (Figure 10a,b). The generation of H2 is a chemical reaction in the Ni corrosion process. Therefore, inhibiting the photocorrosion of the cocatalyst is also one of the factors to consider to improve the performance of POWS.
In 2015, Busser et al. loaded Cu on Ga2O3 to replace the noble metal Rh, and simultaneously used CrOx and MoOx to suppress the photoreduction of CuO, which effectively improved the stability of the CuO cocatalyst [45]. On this basis, another layer of CrOx was deposited on the surface of the Ni/NiO cocatalyst and CrOx significantly inhibited the dissolution of Ni [116]. In 2019, with the quantum efficiency of Al-SrTiO3 loaded with RhCrOx as high as 50%, Lyu et al. further loaded CoOOH and TiO2, which effectively inhibited the dissolution of Cr and improved the stability of the photocatalyst, maintaining POWS activity for over 1000 h [65].

4. Conclusions and Perspectives

The low surface redox reaction rate is one of the vital influencing factors for poor POWS efficiency. Loading cocatalysts is a modification strategy to boost the surface redox reaction. Therefore, numerous cocatalysts have been developed and applied for POWS. This review summarizes the recent progress made in the field of cocatalysts in enhancing POWS performance in regard to loading methods and functions (Table 2). Although highly active and stable cocatalysts as well as various loading means have been developed, numerous challenges to achieving efficient and stable POWS still remain. Thus, significant efforts should be devoted to the following aspects:
(1)
Explore suitable low-cost cocatalysts for hydrogen and oxygen production. Currently, the most efficient cocatalysts for hydrogen production and oxygen production are still noble metals and their oxides, while non-noble metal cocatalysts still need to improve their performance and stability. Therefore, it is urgent to develop efficient and stable non-noble-metal cocatalysts to reduce the costs of the catalytic system. Many transition metal-based cocatalysts have been developed as an alternative to noble metal cocatalysts, which is conducive to reducing the preparation cost of photocatalytic systems effectively. Although the efficiency of non-noble metal-based cocatalysts still needs to be improved, non-noble metal-based cocatalysts are widely used. For example, non-noble metals and their phosphides can be used as cocatalysts for hydrogen production, and their oxides can be used as cocatalysts for oxygen production, the inhibition of reverse reactions, and the prevention of photocorrosion, etc. Transition metal-based cocatalysts that could promise a comparable performance as noble metal are yet to be developed.
(2)
Develop a single-atom cocatalyst. It is widely acknowledged that single-atom metal cocatalysts could maximize the utilization of metal and reduce the costs of a catalyst system. Moreover, single-atom photocatalysts with unique unsaturated coordination environments also exhibit excellent performance in various reactions. However, although single-atom cocatalysts have been widely used in the field of photocatalysis, it is of great significance to develop single-atom cocatalysts for POWS.
(3)
Develop new methods for loading cocatalysts. The activity of cocatalyst-modified photocatalysts is highly dependent on the methods of cocatalyst preparation and loading, which determine the properties of the cocatalyst and the interface between the cocatalyst and the base photocatalyst. However, the currently developed methods cannot be simply adopted on different photocatalysts and different precursors of cocatalysts. Therefore, it is necessary to develop new general/universal strategies for cocatalyst loading, especially those low-cost methods, for instance, simply blending non-noble-metal cocatalysts with photocatalysts in the reaction solution, which is suitable for large-scale commercial applications.
(4)
Comprehensively configure different types of cocatalysts to improve the performance of the catalytic system. For POWS, accelerating the formation of hydrogen and oxygen, inhibiting backward reactions, and reducing photocorrosion are significant factors that need to be considered. It is difficult for a single cocatalyst to achieve all the above functions, so how to rationally configure cocatalysts with different functions is worth further study.
Table 2. The most promising cocatalyst for POWS catalyst with H2 or O2 production.
Table 2. The most promising cocatalyst for POWS catalyst with H2 or O2 production.
Loading MethodPhotocatalystH2-Production CocatalystO2-Production CocatalystLight SourceActivity (μmol h−1 g−1)AQY (%)Stability at Least (h)Ref.
ImpregnationSrTiO3NiO (Hg)0.098 (H2)
0.049 (O2)
100[83]
(Ga1−xZnx)(N1−xOx)Rh2-yCryO3 λ > 200 nm (Xe)1760 (H2)
880 (O2)
[72]
SrTiO3Ni@NiOx AM 1.5G18 (H2)
9 (O2)
[58]
STO:AlGraphene-RhCrOx (Xe) 6375 (H2)
3083 (O2)
0.67 (360 nm)16[89]
(Ga1−xZnx)(N1−xOx)Rh2-yCryO3 367 nm (LED) 0.16 (367 nm)20[87]
STO-AlRhCrOx (Xe)3288 (H2)
166.24 (O2)
12[88]
ZrO2/TaONCr2O3/RuOxIrO2/RuOxλ > 300 nm (Xe) 10[49]
mpg-CNx CoPiλ > 400 nm (Xe)13.6 (H2)
6.6 (O2)
15[104]
ß-Ge3N4 RuO2(Hg)933 (H2)9 (300 nm)25[51]
Impregnation-calcinationAl:STOMoOy-RhCrOx λ > 300 nm (Xe) 69 (365 nm)16[91]
(Ga1−xZnx)(N1−xOx)RhCrOx λ > 400 nm (Hg)3080 (H2)
1540 (O2)
51 (430 nm)35[67]
Hydrothermal -impregnationTiO2Pd@Pt (Xe)601.7 (H2) [92]
Photodeposition(Ga1−xZnx)(N1−xOx)Rh-CrOx λ > 400 nm (Hg)1153 (H2)
593 (O2)
[57]
SrTiO3NiOx λ > 250 nm (Xe)28 (H2)
14 (O2)
[42]
Ta2O5Rh-CrOx λ > 250 nm (Hg)204 (H2)
91 (O2)
[37]
Ga2O3Rh λ > 250 nm (Hg)249 (H2)
126 (O2)
[37]
GaN:ZnORh2-yCryO3 λ > 400 nm (Xe)17.5 (H2)
7 (O2)
2160[50]
Y2Ti2O5S2Rh/Cr2O3IrO2λ > 420 nm (Xe) 5.3 ± 0.3 (H2)
2.3 ± 0.1 (O2) (450 nm)
20[26]
(Zn1+xGe)(N2Ox)Rh2-yCryO3 λ > 200 nm (Xe)53.3 (H2)
25.7 (O2)
[72]
SrTiO3Rh2-yCryO3 λ > 200 nm (Xe)3533.3 (H2)
1788.7 (O2)
[72]
Ca2Nb2O7Rh2-yCryO3 λ > 200 nm (Xe)500 (H2)
216.7 (O2)
[72]
ß-Ga2O3Rh2-yCryO3 λ > 200 nm (Xe)4156.7 (H2)
1950 (O2)
[72]
H-Bi0.5Y0.5VO4Rh@Rh2O3 λ > 300 nm (Xe)516 (H2)
258.5 (O2)
16[85]
GaN:ZnOMn3O4 + Rh/Cr2O3 λ > 420 nm (Xe) 12[30]
InGaN/GaNRh/Cr2O3 λ > 300 nm (Xe)37.94 (H2)
20.78 (O2)
(μmol h−1)
1.86 (400 nm)20[86]
γ-MnS/Cu7S4 MnOx(Xe)718 (H2)18.8 (420 nm)20[53]
ZnIn2S4/WO3PtSMnOx(Xe)14.85 (H2)
5.6 (O2)
0.5 (420 nm)24[103]
Photodeposition-adsorptionY2Ti2SO5Cr2O3-RhIrO2λ > 420 nm (Xe) 0.36 (420 nm) 0.23 (500 nm)
0.05 (600 nm)
0.007 (STH)
22[84]
Photodeposition-impregnationCNN/BDCNNPtCo(OH)2λ > 300 nm (Xe)62.9 (H2)
31.25 (O2)
11.76 (420 nm) 1.16 (STH)24[106]
hydrothermal -photodepositionC3N4PtNi(OH)2(Xe)1330 (H2)
632 (O2)
4.2 (420 m)18[107]

Author Contributions

Conceptualization, X.G. and S.Z.; formal analysis, L.T. and X.G.; investigation, L.T.; resources, L.T.; writing—original draft preparation, L.T.; writing—review and editing, A.D. and J.Q.; funding acquisition, X.G. and S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work is funded by the National Natural Science Foundation of China (No. 51906197 and 52142604), the China Postdoctoral Science Foundation (No. 2020M673386 and 2020T130503), the Natural Science Basic Research Program of Shaanxi Province (No. 2019JCW-10), the Natural Science Foundation of Shaanxi Province (No. 2023-JC-QN-0618), and the Fundamental Research Funds for the Central Universities.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Robu, V.; Flynn, D. Outsmart supply dips in renewable energy. Nature 2017, 544, 161. [Google Scholar] [CrossRef] [PubMed]
  2. Lewis, N.S. Toward cost-effective solar energy use. Science 2007, 315, 798–801. [Google Scholar] [CrossRef] [PubMed]
  3. Moniz, S.J.A.; 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]
  4. Chen, S.; Takata, T.; Domen, K. Particulate photocatalysts for overall water splitting. Nat. Rev. Mater. 2017, 2, 17050. [Google Scholar] [CrossRef]
  5. Wang, L.; Zhang, Y.; Chen, L.; Xu, H.; Xiong, Y. 2D polymers as emerging materials for photocatalytic overall water splitting. Adv. Mater. 2018, 30, 1801955. [Google Scholar] [CrossRef] [PubMed]
  6. Nadeem, M.A.; Khan, M.A.; Ziani, A.A.; Idriss, H. An overview of the photocatalytic water splitting over suspended particles. Catalysts 2021, 11, 60. [Google Scholar] [CrossRef]
  7. Bian, J.; Zhang, Z.; Liu, Y.; Chen, E.; Tang, J.; Jing, L. Strategies and reaction systems for solar-driven CO2 reduction by water. Carbon Neutrality 2022, 1, 5. [Google Scholar] [CrossRef]
  8. Liu, F.; Xue, F.; Si, Y.; Chen, G.; Guan, X.; Lu, K.; Liu, M. Functionalized Cd0.5Zn0.5S chalcogenide nanotwins enabling Z-scheme photocatalytic water splitting. ACS Appl. Nano Mater. 2021, 4, 759–768. [Google Scholar] [CrossRef]
  9. Wang, Q.; Pan, Z. Advances and challenges in developing cocatalysts for photocatalytic conversion of carbon dioxide to fuels. Nano Res. 2022, 15, 10090–10109. [Google Scholar] [CrossRef]
  10. Xiong, L.; Tang, J. Strategies and challenges on selectivity of photocatalytic oxidation of organic substances. Adv. Energy Mater. 2021, 11, 2003216. [Google Scholar] [CrossRef]
  11. Zhao, W.; Adeel, M.; Zhang, P.; Zhou, P.; Huang, L.; Zhao, Y.; Ahmad, M.A.; Shakoor, N.; Lou, B.; Jiang, Y.; et al. A critical review on surface-modified nano-catalyst application for the photocatalytic degradation of volatile organic compounds. Environ. Sci. Nano 2022, 9, 61–80. [Google Scholar] [CrossRef]
  12. Zheng, Y.; Hou, X.; Li, Q.; Fang, Z.; Yang, T.; Liang, T.; Duan, X.; Shang, M.; Yang, W. Electrostatic interaction assisted synthesis of a CdS/BCN heterostructure with enhanced photocatalytic effects. J. Mater. Chem. C 2020, 8, 1803–1810. [Google Scholar] [CrossRef]
  13. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef]
  15. Guan, X.; Zong, S.; Shen, S. Homojunction photocatalysts for water splitting. Nano Res. 2022, 15, 10171–10184. [Google Scholar] [CrossRef]
  16. Luévano-Hipólito, E.; Torres-Martínez, L.M.; Ávila-López, M.a. Visible-light-driven CO2 reduction and H2 evolution boosted by 1D Cu2O/CuO heterostructures. J. Phys. Chem. Solids 2022, 170, 110924. [Google Scholar] [CrossRef]
  17. Zhou, L.; Yang, T.; Fang, Z.; Zhou, J.; Zheng, Y.; Guo, C.; Zhu, L.; Wang, E.; Hou, X.; Chou, K.-C.; et al. Boosting of water splitting using the chemical energy simultaneously harvested from light, kinetic energy and electrical energy using N doped 4H-SiC nanohole arrays. Nano Energy 2022, 104, 107876. [Google Scholar] [CrossRef]
  18. Wang, Z.; Li, C.; Domen, K. Recent developments in heterogeneous photocatalysts for solar-driven overall water splitting. Chem. Soc. Rev. 2019, 48, 2109–2125. [Google Scholar] [CrossRef]
  19. Liu, F.; Xue, F.; Zhang, C.; Lu, K.; Chen, G.; Ye, X.; Chen, W.; Guo, L.; Liu, M. Noble-metal-free chalcogenide nanotwins for efficient and stable photocatalytic pure water splitting by surface phosphorization and cocatalyst modification. Mater. Today Energy 2022, 30, 101180. [Google Scholar] [CrossRef]
  20. Zhang, W.; Peng, Q.; Shi, L.; Yao, Q.; Wang, X.; Yu, A.; Chen, Z.; Fu, Y. Merging single-atom-dispersed iron and graphitic carbon nitride to a joint electronic system for high-efficiency photocatalytic hydrogen evolution. Small 2019, 15, 1905166. [Google Scholar] [CrossRef]
  21. Zhao, D.; Guan, X.; Shen, S. Design of polymeric carbon nitride-based heterojunctions for photocatalytic water splitting: A review. Environ. Chem. Lett. 2022, 20, 3505–3523. [Google Scholar] [CrossRef]
  22. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. Dual cocatalysts in TiO2 photocatalysis. Adv. Mater. 2019, 31, 1807660. [Google Scholar] [CrossRef]
  23. Zhao, G.; Xu, X. Cocatalysts from types, preparation to applications in the field of photocatalysis. Nanoscale 2021, 13, 10649–10667. [Google Scholar] [CrossRef] [PubMed]
  24. Townsend, T.K.; Browning, N.D.; Osterloh, F.E. Nanoscale strontium titanate photocatalysts for overall water splitting. ACS Nano 2012, 6, 7420–7426. [Google Scholar] [CrossRef] [PubMed]
  25. Yuan, Y.-J.; Chen, D.; Yang, S.; Yang, L.-X.; Wang, J.-J.; Cao, D.; Tu, W.; Yu, Z.-T.; Zou, Z.-G. Constructing noble-metal-free Z-scheme photocatalytic overall water splitting systems using MoS2 nanosheet modified CdS as a H2 evolution photocatalyst. J. Mater. Chem. A 2017, 5, 21205–21213. [Google Scholar] [CrossRef]
  26. Wang, Q.; Nakabayashi, M.; Hisatomi, T.; Sun, S.; Akiyama, S.; Wang, Z.; Pan, Z.; Xiao, X.; Watanabe, T.; Yamada, T.; et al. Oxysulfide photocatalyst for visible-light-driven overall water splitting. Nat. Mater. 2019, 18, 827–832. [Google Scholar] [CrossRef]
  27. Takata, T.; Jiang, J.; Sakata, Y.; Nakabayashi, M.; Shibata, N.; Nandal, V.; Seki, K.; Hisatomi, T.; Domen, K. Photocatalytic water splitting with a quantum efficiency of almost unity. Nature 2020, 581, 411–414. [Google Scholar] [CrossRef]
  28. Aguirre-Astrain, A.; Luévano-Hipólito, E.; Torres-Martínez, L.M. Integration of 2D printing technologies for AV2O6 (A=Ca, Sr, Ba)-MO (M=Cu, Ni, Zn) photocatalyst manufacturing to solar fuels production using seawater. Int. J. Hydrogen Energy 2021, 46, 37294–37310. [Google Scholar] [CrossRef]
  29. Cheng, C.; Mao, L.; Shi, J.; Xue, F.; Zong, S.; Zheng, B.; Guo, L. NiCo2O4 nanosheets as a novel oxygen-evolution-reaction cocatalyst in situ bonded on the g-C3N4 photocatalyst for excellent overall water splitting. J. Mater. Chem. A 2021, 9, 12299–12306. [Google Scholar] [CrossRef]
  30. Maeda, K.; Xiong, A.; Yoshinaga, T.; Ikeda, T.; Sakamoto, N.; Hisatomi, T.; Takashima, M.; Lu, D.; Kanehara, M.; Setoyama, T.; et al. Photocatalytic overall water splitting promoted by two different cocatalysts for hydrogen and oxygen evolution under visible light. Angew. Chem. Int. Ed. 2010, 49, 4096–4099. [Google Scholar] [CrossRef]
  31. Liu, Y.; Zhou, X.; Zhou, Y.; Chen, G. Construction of Ag decorated 2D rGO/SnS2 nanostructure towards synergistically enabling overall water splitting. Chem. Eng. J. 2022, 433, 133198. [Google Scholar] [CrossRef]
  32. Zhao, H.; Jian, L.; Gong, M.; Jing, M.; Li, H.; Mao, Q.; Lu, T.; Guo, Y.; Ji, R.; Chi, W.; et al. Transition-metal-based cocatalysts for photocatalytic water splitting. Small Struct. 2022, 3, 2100229. [Google Scholar] [CrossRef]
  33. Cao, S.; Chan, T.-S.; Lu, Y.-R.; Shi, X.; Fu, B.; Wu, Z.; Li, H.; Liu, K.; Alzuabi, S.; Cheng, P.; et al. Photocatalytic pure water splitting with high efficiency and value by Pt/porous brookite TiO2 nanoflutes. Nano Energy 2020, 67, 104287. [Google Scholar] [CrossRef]
  34. Remiro-Buenamañana, S.; Cabrero-Antonino, M.; Martínez-Guanter, M.; Álvaro, M.; Navalón, S.; García, H. Influence of co-catalysts on the photocatalytic activity of MIL-125(Ti)-NH2 in the overall water splitting. Appl. Catal. B 2019, 254, 677–684. [Google Scholar] [CrossRef]
  35. Li, W.; Chu, X.-S.; Wang, F.; Dang, Y.-Y.; Liu, X.-Y.; Ma, T.-H.; Li, J.-Y.; Wang, C.-Y. Pd single-atom decorated CdS nanocatalyst for highly efficient overall water splitting under simulated solar light. Appl. Catal. B 2022, 304, 121000. [Google Scholar] [CrossRef]
  36. Kosco, J.; McCulloch, I. Residual Pd enables photocatalytic H2 evolution from conjugated polymers. ACS Energy Lett. 2018, 3, 2846–2850. [Google Scholar] [CrossRef]
  37. Menze, J.; Mei, B.; Weide, P.; Muhler, M. Oxidative photo-deposition of chromia: Tuning the activity for overall water splitting of the Rh/CrOx co-catalyst system. J. Mater. Chem. A 2017, 5, 17248–17252. [Google Scholar] [CrossRef]
  38. Ivanova, I.; Kandiel, T.A.; Cho, Y.-J.; Choi, W.; Bahnemann, D. Mechanisms of photocatalytic molecular hydrogen and molecular oxygen evolution over La-doped NaTaO3 particles: Effect of different cocatalysts and their specific activity. ACS Catal. 2018, 8, 2313–2325. [Google Scholar] [CrossRef]
  39. Iwase, A.; Kato, H.; Kudo, A. The effect of Au cocatalyst loaded on La-doped NaTaO3 on photocatalytic water splitting and O2 photoreduction. Appl. Catal. B 2013, 136–137, 89–93. [Google Scholar] [CrossRef]
  40. Yan, C.; Xue, X.; Zhang, W.; Li, X.; Liu, J.; Yang, S.; Hu, Y.; Chen, R.; Yan, Y.; Zhu, G.; et al. Well-designed Te/SnS2/Ag artificial nanoleaves for enabling and enhancing visible-light driven overall splitting of pure water. Nano Energy 2017, 39, 539–545. [Google Scholar] [CrossRef]
  41. Irie, H.; Yoda, M.; Takashima, T.; Osaki, J. Silver cocatalyst-concentration dependence of overall water splitting performance over silver-inserted solid-state heterojunction photocatalyst composed of zinc rhodium oxide and bismuth vanadium oxide. Appl. Catal. B 2021, 284, 119744. [Google Scholar] [CrossRef]
  42. Townsend, T.K.; Browning, N.D.; Osterloh, F.E. Overall photocatalytic water splitting with NiOx–SrTiO3—A revised mechanism. Energy Environ. Sci. 2012, 5, 9543–9550. [Google Scholar] [CrossRef]
  43. Thaminimulla, C.T.K.; Takata, T.; Hara, M.; Kondo, J.N.; Domen, K. Effect of chromium addition for photocatalytic overall water splitting on Ni–K2La2Ti3O10. J. Catal. 2000, 196, 362–365. [Google Scholar] [CrossRef]
  44. Wei, D.; Tan, Y.; Wang, Y.; Kong, T.; Shen, S.; Mao, S.S. Function-switchable metal/semiconductor junction enables efficient photocatalytic overall water splitting with selective water oxidation products. Sci. Bull. 2020, 65, 1389–1395. [Google Scholar] [CrossRef] [PubMed]
  45. Busser, G.W.; Mei, B.; Weide, P.; Vesborg, P.C.K.; Stührenberg, K.; Bauer, M.; Huang, X.; Willinger, M.-G.; Chorkendorff, I.; Schlögl, R.; et al. Cocatalyst designing: A regenerable molybdenum-containing ternary cocatalyst system for efficient photocatalytic water splitting. ACS Catal. 2015, 5, 5530–5539. [Google Scholar] [CrossRef]
  46. Qiao, S.; Guo, J.; Wang, D.; Zhang, L.; Hassan, A.; Chen, T.; Feng, C.; Zhang, Y.; Wang, J. Core-shell cobalt particles Co@CoO loaded on nitrogen-doped graphene for photocatalytic water-splitting. Int. J. Hydrogen Energy 2020, 45, 1629–1639. [Google Scholar] [CrossRef]
  47. Mateo, D.; Esteve-Adell, I.; Albero, J.; Primo, A.; García, H. Oriented 2.0.0 Cu2O nanoplatelets supported on few-layers graphene as efficient visible light photocatalyst for overall water splitting. Appl. Catal. B 2017, 201, 582–590. [Google Scholar] [CrossRef]
  48. Liu, Y.; Zhao, Y.; Sun, Y.; Cao, J.; Wang, H.; Wang, X.; Huang, H.; Shao, M.; Liu, Y.; Kang, Z. A 4e-–2e- cascaded pathway for highly efficient production of H2 and H2O2 from water photo-splitting at normal pressure. Appl. Catal. B 2020, 270, 118875. [Google Scholar] [CrossRef]
  49. Maeda, K.; Lu, D.; Domen, K. Direct water splitting into hydrogen and oxygen under visible light by using modified TaON photocatalysts with d0 electronic configuration. Chem. Eur. J. 2013, 19, 4986–4991. [Google Scholar] [CrossRef]
  50. Ohno, T.; Bai, L.; Hisatomi, T.; Maeda, K.; Domen, K. Photocatalytic water splitting using modified GaN:ZnO solid solution under visible light: Long-time operation and regeneration of activity. J. Am. Chem. Soc. 2012, 134, 8254–8259. [Google Scholar] [CrossRef]
  51. Sato, J.; Saito, N.; Yamada, Y.; Maeda, K.; Takata, T.; Kondo, J.N.; Hara, M.; Kobayashi, H.; Domen, K.; Inoue, Y. RuO2-loaded β-Ge3N4 as a non-oxide photocatalyst for overall water splitting. J. Am. Chem. Soc. 2005, 127, 4150–4151. [Google Scholar] [CrossRef] [PubMed]
  52. Sato, J.; Saito, N.; Nishiyama, H.; Inoue, Y. New photocatalyst group for water decomposition of RuO2-loaded p-block metal (In, Sn, and Sb) oxides with d10 configuration. J. Phys. Chem. B 2001, 105, 6061–6063. [Google Scholar] [CrossRef]
  53. Yuan, Q.; Liu, D.; Zhang, N.; Ye, W.; Ju, H.; Shi, L.; Long, R.; Zhu, J.; Xiong, Y. Noble-metal-free janus-like structures by cation exchange for Z-scheme photocatalytic water splitting under broadband light irradiation. Angew. Chem. Int. Ed. 2017, 56, 4206–4210. [Google Scholar] [CrossRef]
  54. Pan, J.; Chen, Z.; Wang, P.; Wang, P.; Yu, Q.; Zhao, W.; Wang, J.; Zhu, M.; Zheng, Y.; Li, C. The overall water splitting of CdS/Ti3+-SrTiO3 core–shell heterojunction via OER enhancement of MnOx nanoparticles. Chem. Eng. J. 2021, 424, 130357. [Google Scholar] [CrossRef]
  55. Zhang, Q.; Li, Z.; Wang, S.; Li, R.; Zhang, X.; Liang, Z.; Han, H.; Liao, S.; Li, C. Effect of redox cocatalysts location on photocatalytic overall water splitting over cubic NaTaO3 semiconductor crystals exposed with equivalent facets. ACS Catal. 2016, 6, 2182–2191. [Google Scholar] [CrossRef]
  56. Qi, Y.; Zhao, Y.; Gao, Y.; Li, D.; Li, Z.; Zhang, F.; Li, C. Redox-based visible-light-driven Z-scheme overall water splitting with apparent quantum efficiency exceeding 10%. Joule 2018, 2, 2393–2402. [Google Scholar] [CrossRef]
  57. Maeda, K.; Teramura, K.; Lu, D.; Saito, N.; Inoue, Y.; Domen, K. Noble-metal/Cr2O3 core/shell nanoparticles as a cocatalyst for photocatalytic overall water splitting. Angew. Chem. Int. Ed. 2006, 45, 7806–7809. [Google Scholar] [CrossRef]
  58. Han, K.; Kreuger, T.; Mei, B.; Mul, G. Transient behavior of Ni@NiOx functionalized SrTiO3 in overall water splitting. ACS Catal. 2017, 7, 1610–1614. [Google Scholar] [CrossRef] [PubMed]
  59. Liu, Q.; Zhang, L.; Crozier, P.A. Structure–reactivity relationships of Ni–NiO core–shell co-catalysts on Ta2O5 for solar hydrogen production. Appl. Catal. B 2015, 172-173, 58–64. [Google Scholar] [CrossRef]
  60. Zhang, J.; Bai, T.; Huang, H.; Yu, M.-H.; Fan, X.; Chang, Z.; Bu, X.-H. Metal–organic-framework-based photocatalysts optimized by spatially separated cocatalysts for overall water splitting. Adv. Mater. 2020, 32, 2004747. [Google Scholar] [CrossRef]
  61. Wang, D.; Hisatomi, T.; Takata, T.; Pan, C.; Katayama, M.; Kubota, J.; Domen, K. Core/shell photocatalyst with spatially separated co-catalysts for efficient reduction and oxidation of water. Angew. Chem. Int. Ed. 2013, 52, 11252–11256. [Google Scholar] [CrossRef]
  62. Zhen, W.; Ning, X.; Yang, B.; Wu, Y.; Li, Z.; Lu, G. The enhancement of CdS photocatalytic activity for water splitting via anti-photocorrosion by coating Ni2P shell and removing nascent formed oxygen with artificial gill. Appl. Catal. B 2018, 221, 243–257. [Google Scholar] [CrossRef]
  63. Chen, S.; Vequizo, J.J.M.; Pan, Z.; Hisatomi, T.; Nakabayashi, M.; Lin, L.; Wang, Z.; Kato, K.; Yamakata, A.; Shibata, N.; et al. Surface modifications of (ZnSe)0.5(CuGa2.5Se4.25)0.5 to promote photocatalytic Z-scheme overall water splitting. J. Am. Chem. Soc. 2021, 143, 10633–10641. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, L.; Liu, Q.; Aoki, T.; Crozier, P.A. Structural evolution during photocorrosion of Ni/NiO core/shell cocatalyst on TiO2. J. Phys. Chem. C 2015, 119, 7207–7214. [Google Scholar] [CrossRef]
  65. Lyu, H.; Hisatomi, T.; Goto, Y.; Yoshida, M.; Higashi, T.; Katayama, M.; Takata, T.; Minegishi, T.; Nishiyama, H.; Yamada, T.; et al. An Al-doped SrTiO3 photocatalyst maintaining sunlight-driven overall water splitting activity for over 1000 h of constant illumination. Chem. Sci. 2019, 10, 3196–3201. [Google Scholar] [CrossRef]
  66. Kawawaki, T.; Mori, Y.; Wakamatsu, K.; Ozaki, S.; Kawachi, M.; Hossain, S.; Negishi, Y. Controlled colloidal metal nanoparticles and nanoclusters: Recent applications as cocatalysts for improving photocatalytic water-splitting activity. J. Mater. Chem. A 2020, 8, 16081–16113. [Google Scholar] [CrossRef]
  67. Maeda, K.; Teramura, K.; Lu, D.; Takata, T.; Saito, N.; Inoue, Y.; Domen, K. Photocatalyst releasing hydrogen from water. Nature 2006, 440, 295. [Google Scholar] [CrossRef]
  68. Seo, S.W.; Park, S.; Jeong, H.-Y.; Kim, S.H.; Sim, U.; Lee, C.W.; Nam, K.T.; Hong, K.S. Enhanced performance of NaTaO3 using molecular co-catalyst [Mo3S4]4+ for water splitting into H2 and O2. Chem. Commun. 2012, 48, 10452–10454. [Google Scholar] [CrossRef]
  69. Xue, F.; Si, Y.; Wang, M.; Liu, M.; Guo, L. Toward efficient photocatalytic pure water splitting for simultaneous H2 and H2O2 production. Nano Energy 2019, 62, 823–831. [Google Scholar] [CrossRef]
  70. Zong, S.; Tian, L.; Guan, X.; Cheng, C.; Shi, J.; Guo, L. Photocatalytic overall water splitting without noble-metal: Decorating CoP on Al-doped SrTiO3. J. Colloid Interface Sci. 2022, 606, 491–499. [Google Scholar] [CrossRef]
  71. Maeda, K.; Lu, D.; Teramura, K.; Domen, K. Simultaneous photodeposition of rhodium–chromium nanoparticles on a semiconductor powder: Structural characterization and application to photocatalytic overall water splitting. Energy Environ. Sci. 2010, 3, 470–477. [Google Scholar] [CrossRef]
  72. Maeda, K.; Lu, D.; Teramura, K.; Domen, K. Direct deposition of nanoparticulate rhodium–chromium mixed-oxides on a semiconductor powder by band-gap irradiation. J. Mater. Chem. 2008, 18, 3539–3542. [Google Scholar] [CrossRef]
  73. Xiao, X.; Gao, Y.; Zhang, L.; Zhang, J.; Zhang, Q.; Li, Q.; Bao, H.; Zhou, J.; Miao, S.; Chen, N.; et al. A promoted charge separation/transfer system from Cu single atoms and C3N4 layers for efficient photocatalysis. Adv. Mater. 2020, 32, 2003082. [Google Scholar] [CrossRef] [PubMed]
  74. Li, X.; Bi, W.; Zhang, L.; Tao, S.; Chu, W.; Zhang, Q.; Luo, Y.; Wu, C.; Xie, Y. Single-atom Pt as co-catalyst for enhanced photocatalytic H2 evolution. Adv. Mater. 2016, 28, 2427–2431. [Google Scholar] [CrossRef]
  75. Wu, X.; Zhang, H.; Dong, J.; Qiu, M.; Kong, J.; Zhang, Y.; Li, Y.; Xu, G.; Zhang, J.; Ye, J. Surface step decoration of isolated atom as electron pumping: Atomic-level insights into visible-light hydrogen evolution. Nano Energy 2018, 45, 109–117. [Google Scholar] [CrossRef]
  76. Chen, Y.; Ji, S.; Sun, W.; Lei, Y.; Wang, Q.; Li, A.; Chen, W.; Zhou, G.; Zhang, Z.; Wang, Y.; et al. Engineering the atomic interface with single platinum atoms for enhanced photocatalytic hydrogen production. Angew. Chem. Int. Ed. 2020, 59, 1295–1301. [Google Scholar] [CrossRef]
  77. Qiu, B.; Cai, L.; Zhang, N.; Tao, X.; Chai, Y. A ternary dumbbell structure with spatially separated catalytic sites for photocatalytic overall water splitting. Adv. Sci. 2020, 7, 1903568. [Google Scholar] [CrossRef] [PubMed]
  78. Zhou, X.; Li, J.; Cai, X.; Gao, Q.; Zhang, S.; Yang, S.; Wang, H.; Zhong, X.; Fang, Y. In situ photo-derived MnOOH collaborating with Mn2Co2C@C dual co-catalysts boost photocatalytic overall water splitting. J. Mater. Chem. A 2020, 8, 17120–17127. [Google Scholar] [CrossRef]
  79. Chen, S.; Zhou, X.; Liao, J.; Yang, S.; Zhou, X.; Gao, Q.; Zhang, S.; Fang, Y.; Zhong, X.; Zhang, S. FeNi intermetallic compound nanoparticles wrapped with N-doped graphitized carbon: A novel cocatalyst for boosting photocatalytic hydrogen evolution. J. Mater. Chem. A 2020, 8, 3481–3490. [Google Scholar] [CrossRef]
  80. Garcia-Esparza, A.T.; Cha, D.; Ou, Y.; Kubota, J.; Domen, K.; Takanabe, K. Tungsten carbide nanoparticles as efficient cocatalysts for photocatalytic overall water splitting. ChemSusChem 2013, 6, 168–181. [Google Scholar] [CrossRef] [PubMed]
  81. Ma, B.; Dang, Y.; Li, D.; Wang, X.; Lin, K.; Wang, W.; Zhou, X.; Chen, Y.; Xie, T.; Zhang, X.; et al. A Yin-Yang hybrid co-catalyst (CoOx-Mo2N) for photocatalytic overall water splitting. Appl. Catal. B 2021, 298, 120491. [Google Scholar] [CrossRef]
  82. Zhao, B.; Wang, Y.; Yao, X.; Chen, D.; Fan, M.; Jin, Z.; He, Q. Photocatalysis-mediated drug-free sustainable cancer therapy using nanocatalyst. Nat. Commun. 2021, 12, 1345. [Google Scholar] [CrossRef] [PubMed]
  83. Domen, K.; Naito, S.; Soma, M.; Onishi, T.; Tamaru, K. Photocatalytic decomposition of water vapour on an NiO–SrTiO3 catalyst. J. Chem. Soc. Chem. Commun. 1980, 543–544. [Google Scholar] [CrossRef]
  84. Zhang, G.; Wang, X. Oxysulfide semiconductors for photocatalytic overall water splitting with visible light. Angew. Chem. Int. Ed. 2019, 58, 15580–15582. [Google Scholar] [CrossRef] [PubMed]
  85. Wei, Z.; Zhu, Y.; Guo, W.; Liu, J.; Fang, W.; Jiang, Z.; Shangguan, W. Enhanced twisting degree assisted overall water splitting on a novel nano-dodecahedron BiVO4-based heterojunction. Appl. Catal. B 2020, 266, 118664. [Google Scholar] [CrossRef]
  86. Kibria, M.G.; Nguyen, H.P.T.; Cui, K.; Zhao, S.; Liu, D.; Guo, H.; Trudeau, M.L.; Paradis, S.; Hakima, A.-R.; Mi, Z. One-step overall water splitting under visible light using multiband InGaN/GaN nanowire heterostructures. ACS Nano 2013, 7, 7886–7893. [Google Scholar] [CrossRef] [PubMed]
  87. Dionigi, F.; Vesborg, P.C.K.; Pedersen, T.; Hansen, O.; Dahl, S.; Xiong, A.; Maeda, K.; Domen, K.; Chorkendorff, I. Gas phase photocatalytic water splitting with Rh2−yCryO3/GaN:ZnO in μ-reactors. Energy Environ. Sci. 2011, 4, 2937–2942. [Google Scholar] [CrossRef]
  88. Sanwald, K.E.; Berto, T.F.; Jentys, A.; Camaioni, D.M.; Gutiérrez, O.Y.; Lercher, J.A. Kinetic coupling of water splitting and photoreforming on SrTiO3-based photocatalysts. ACS Catal. 2018, 8, 2902–2913. [Google Scholar] [CrossRef]
  89. Mateo, D.; García-Mulero, A.; Albero, J.; García, H. N-doped defective graphene decorated by strontium titanate as efficient photocatalyst for overall water splitting. Appl. Catal. B 2019, 252, 111–119. [Google Scholar] [CrossRef]
  90. Zhang, X.; Li, Z.; Ta, N.; Han, H. Unique properties of RhCrOx cocatalyst regulating reactive oxygen species formation in photocatalytic overall water splitting. ACS Sustain. Chem. Eng. 2022, 10, 4059–4064. [Google Scholar] [CrossRef]
  91. Chiang, T.H.; Lyu, H.; Hisatomi, T.; Goto, Y.; Takata, T.; Katayama, M.; Minegishi, T.; Domen, K. Efficient photocatalytic water splitting using Al-doped SrTiO3 coloaded with molybdenum oxide and rhodium–chromium oxide. ACS Catal. 2018, 8, 2782–2788. [Google Scholar] [CrossRef]
  92. Bai, S.; Yang, L.; Wang, C.; Lin, Y.; Lu, J.; Jiang, J.; Xiong, Y. Boosting photocatalytic water splitting: Interfacial charge polarization in atomically controlled core–shell cocatalysts. Angew. Chem. Int. Ed. 2015, 54, 14810–14814. [Google Scholar] [CrossRef]
  93. Bard, A.J.; Fox, M.A. Artificial photosynthesis: Solar splitting of water to hydrogen and oxygen. Acc. Chem. Res. 1995, 28, 141–145. [Google Scholar] [CrossRef]
  94. Li, X.; Wang, W.; Dong, F.; Zhang, Z.; Han, L.; Luo, X.; Huang, J.; Feng, Z.; Chen, Z.; Jia, G.; et al. Recent advances in noncontact external-field-assisted photocatalysis: From fundamentals to applications. ACS Catal. 2021, 11, 4739–4769. [Google Scholar] [CrossRef]
  95. Yang, J.; Wang, D.; Han, H.; Li, C. Roles of cocatalysts in photocatalysis and photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. [Google Scholar] [CrossRef]
  96. Deng, X.; Tüysüz, H. Cobalt-oxide-based materials as water oxidation catalyst: Recent progress and challenges. ACS Catal. 2014, 4, 3701–3714. [Google Scholar] [CrossRef]
  97. 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]
  98. Soni, V.; Xia, C.; Cheng, C.K.; Nguyen, V.-H.; Nguyen, D.L.T.; Bajpai, A.; Kim, S.Y.; Le, Q.V.; Khan, A.A.P.; Singh, P.; et al. Advances and recent trends in cobalt-based cocatalysts for solar-to-fuel conversion. Appl. Mater. Today 2021, 24, 101074. [Google Scholar] [CrossRef]
  99. Zhao, H.; Mao, Q.; Jian, L.; Dong, Y.; Zhu, Y. Photodeposition of earth-abundant cocatalysts in photocatalytic water splitting: Methods, functions, and mechanisms. Chinese J. Catal. 2022, 43, 1774–1804. [Google Scholar] [CrossRef]
  100. Inoue, Y.; Kubokawa, T.; Sato, K. Photocatalytic activity of sodium hexatitanate, Na2Ti6O13, with a tunnel structure for decomposition of water. J. Chem. Soc. Chem. Commun. 1990, 1298–1299. [Google Scholar] [CrossRef]
  101. Ogura, S.; Kohno, M.; Sato, K.; Inoue, Y. Photocatalytic properties of M2Ti6O13 (M=Na, K, Rb, Cs) with rectangular tunnel and layer structures: Behavior of a surface radical produced by UV irradiation and photocatalytic activity for water decomposition. Phys. Chem. Chem. Phys. 1999, 1, 179–183. [Google Scholar] [CrossRef]
  102. Inoue, Y.; Asai, Y.; Sato, K. Photocatalysts with tunnel structures for decomposition of water. Part 1.—BaTi4O9, a pentagonal prism tunnel structure, and its combination with various promoters. J. Chem. Soc. Faraday Trans. 1994, 90, 797–802. [Google Scholar] [CrossRef]
  103. Ding, Y.; Wei, D.; He, R.; Yuan, R.; Xie, T.; Li, Z. Rational design of Z-scheme PtS-ZnIn2S4/WO3-MnO2 for overall photo-catalytic water splitting under visible light. Appl. Catal. B 2019, 258, 117948. [Google Scholar] [CrossRef]
  104. Lee, R.-L.; Tran, P.D.; Pramana, S.S.; Chiam, S.Y.; Ren, Y.; Meng, S.; Wong, L.H.; Barber, J. Assembling graphitic-carbon-nitride with cobalt-oxide-phosphate to construct an efficient hybrid photocatalyst for water splitting application. Catal. Sci. Technol. 2013, 3, 1694–1698. [Google Scholar] [CrossRef]
  105. Guan, X.; Tian, L.; Zhang, Y.; Shi, J.; Shen, S. Photocatalytic water splitting on BiVO4: Balanced charge-carrier consumption and selective redox reaction. Nano Res. 2022; in press. [Google Scholar] [CrossRef]
  106. Zhao, D.; Wang, Y.; Dong, C.-L.; Huang, Y.-C.; Chen, J.; Xue, F.; Shen, S.; Guo, L. Boron-doped nitrogen-deficient carbon nitride-based Z-scheme heterostructures for photocatalytic overall water splitting. Nat. Energy 2021, 6, 388–397. [Google Scholar] [CrossRef]
  107. Sun, S.; Feng, Y.; Pan, L.; Zhang, X.; Zou, J.-J. Integrating Pt@Ni(OH)2 nanowire and Pt nanoparticle on C3N4 with fast surface kinetics and charge transfer towards highly efficient photocatalytic water splitting. Appl. Catal. B 2019, 259, 118028. [Google Scholar] [CrossRef]
  108. Qi, Y.; Chen, S.; Cui, J.; Wang, Z.; Zhang, F.; Li, C. Inhibiting competing reactions of iodate/iodide redox mediators by surface modification of photocatalysts to enable Z-scheme overall water splitting. Appl. Catal. B 2018, 224, 579–585. [Google Scholar] [CrossRef]
  109. Wang, M.; Zhen, W.; Tian, B.; Ma, J.; Lu, G. The inhibition of hydrogen and oxygen recombination reaction by halogen atoms on over-all water splitting over Pt-TiO2 photocatalyst. Appl. Catal. B 2018, 236, 240–252. [Google Scholar] [CrossRef]
  110. Wang, M.; Liu, H.; Ma, J.; Lu, G. The activity enhancement of photocatalytic water splitting by F- pre-occupation on Pt(100) and Pt(111) co-catalyst facets. Appl. Catal. B 2020, 266, 118647. [Google Scholar] [CrossRef]
  111. Nishiyama, H.; Yamada, T.; Nakabayashi, M.; Maehara, Y.; Yamaguchi, M.; Kuromiya, Y.; Nagatsuma, Y.; Tokudome, H.; Akiyama, S.; Watanabe, T.; et al. Photocatalytic solar hydrogen production from water on a 100-m2 scale. Nature 2021, 598, 304–307. [Google Scholar] [CrossRef]
  112. Ning, X.; Zhen, W.; Wu, Y.; Lu, G. Inhibition of CdS photocorrosion by Al2O3 shell for highly stable photocatalytic overall water splitting under visible light irradiation. Appl. Catal. B 2018, 226, 373–383. [Google Scholar] [CrossRef]
  113. Chen, W.; Huang, G.-B.; Song, H.; Zhang, J. Efficient and stable charge transfer channels for photocatalytic water splitting activity of CdS without sacrificial agents. J. Mater. Chem. A 2020, 8, 20963–20969. [Google Scholar] [CrossRef]
  114. Guan, X.; Chowdhury, F.A.; Wang, Y.; Pant, N.; Vanka, S.; Trudeau, M.L.; Guo, L.; Vayssieres, L.; Mi, Z. Making of an industry-friendly artificial photosynthesis device. ACS Energy Lett. 2018, 3, 2230–2231. [Google Scholar] [CrossRef]
  115. Zhou, P.; Navid, I.A.; Ma, Y.; Xiao, Y.; Wang, P.; Ye, Z.; Zhou, B.; Sun, K.; Mi, Z. Solar-to-hydrogen efficiency of more than 9% in photocatalytic water splitting. Nature 2023, 613, 66–70. [Google Scholar] [CrossRef]
  116. Han, K.; Haiber, D.M.; Knöppel, J.; Lievens, C.; Cherevko, S.; Crozier, P.; Mul, G.; Mei, B. CrOx-mediated performance enhancement of Ni/NiO-Mg:SrTiO3 in photocatalytic water splitting. ACS Catal. 2021, 11, 11049–11058. [Google Scholar] [CrossRef]
Figure 1. Schematic descriptions for photocatalytic overall water splitting.
Figure 1. Schematic descriptions for photocatalytic overall water splitting.
Catalysts 13 00355 g001
Figure 2. Schematic illustrations of (a) impregnation and (b) photodeposition method.
Figure 2. Schematic illustrations of (a) impregnation and (b) photodeposition method.
Catalysts 13 00355 g002
Figure 3. Various roles of cocatalyst for POWS.
Figure 3. Various roles of cocatalyst for POWS.
Catalysts 13 00355 g003
Figure 4. (a) Diagram of the mechanism by which RhCrOx promotes photocatalytic performance: RhOx could facilitate both the desired hydrogen evolution reaction (HER) and undesired reactions of superoxide radicals and singlet oxygen formation. On the contrary, CrOx inhibits formation of these two ROS although it could not promote the POWS reaction. Reproduced with permission from Ref. [90]. Copyright 2022, American Chemistry Society. (b) Photocatalytic H2 evolution from water with the samples as catalysts under the same irradiation conditions. Reproduced with permission from Ref. [92]. Copyright 2022, Wiley.
Figure 4. (a) Diagram of the mechanism by which RhCrOx promotes photocatalytic performance: RhOx could facilitate both the desired hydrogen evolution reaction (HER) and undesired reactions of superoxide radicals and singlet oxygen formation. On the contrary, CrOx inhibits formation of these two ROS although it could not promote the POWS reaction. Reproduced with permission from Ref. [90]. Copyright 2022, American Chemistry Society. (b) Photocatalytic H2 evolution from water with the samples as catalysts under the same irradiation conditions. Reproduced with permission from Ref. [92]. Copyright 2022, Wiley.
Catalysts 13 00355 g004
Figure 6. Morphology and microstructure analysis of rGO/SnS2/Ag. (a) SEM image and (b) TEM image, (c) HRTEM lattice fringes. (d) rGo diffraction patterns, (e) rGO/SnS2 diffraction patterns, (f) rGO/SnS2/Ag diffraction patterns. Reproduced with permission from Ref. [31]. Copyright 2022, Elsevier.
Figure 6. Morphology and microstructure analysis of rGO/SnS2/Ag. (a) SEM image and (b) TEM image, (c) HRTEM lattice fringes. (d) rGo diffraction patterns, (e) rGO/SnS2 diffraction patterns, (f) rGO/SnS2/Ag diffraction patterns. Reproduced with permission from Ref. [31]. Copyright 2022, Elsevier.
Catalysts 13 00355 g006
Figure 7. Core–shell structured cocatalyst.
Figure 7. Core–shell structured cocatalyst.
Catalysts 13 00355 g007
Figure 8. (a) HR-TEM images of Rh-loaded (Ga1−xZnx)(N1−xOx) after photodeposition of the Cr shell. Reproduced with permission from Ref. [57]. Copyright 2006, Wiley. (b) H2 production rate vs. reaction time of samples oxidized at different temperatures (different oxide shell thickness). Reproduced with permission from Ref. [59]. Copyright 2015, Elsevier. (c) H2 and O2 recombination rates for Pt-TiO2, F-Pt-TiO2, Cl-Pt-TiO2, Br-Pt-TiO2 and I-Pt-TiO2. Reproduced with permission from Ref. [109]. Copyright 2018, Elsevier.
Figure 8. (a) HR-TEM images of Rh-loaded (Ga1−xZnx)(N1−xOx) after photodeposition of the Cr shell. Reproduced with permission from Ref. [57]. Copyright 2006, Wiley. (b) H2 production rate vs. reaction time of samples oxidized at different temperatures (different oxide shell thickness). Reproduced with permission from Ref. [59]. Copyright 2015, Elsevier. (c) H2 and O2 recombination rates for Pt-TiO2, F-Pt-TiO2, Cl-Pt-TiO2, Br-Pt-TiO2 and I-Pt-TiO2. Reproduced with permission from Ref. [109]. Copyright 2018, Elsevier.
Catalysts 13 00355 g008
Figure 9. (a) The cadmium ion concentration changes in the solution of CdS NPs and Pt/CdS@Al2O3 composite with light irradiation time (b) Cycling runs for the photocatalytic hydrogen evolution activity under visible light over various samples: (a) CdS NPs, (b) Pt/CdS, (c) Pt/CdS@Al2O3 with artificial gill and (d) Pt/CdS@Al2O3 without artificial gill. (a,b) Reproduced with permission from Ref. [112]. Copyright 2018, Elsevier. (c) HRTEM images of 10Ni2P@CdS and 1Pt@CdS. (d) The concentration of Cd2+ for each run in the reaction solution of CdS and 10Ni2P@CdS under visible light irradiation. (c,d) Reproduced with permission from Ref. [62]. Copyright 2018, Elsevier.
Figure 9. (a) The cadmium ion concentration changes in the solution of CdS NPs and Pt/CdS@Al2O3 composite with light irradiation time (b) Cycling runs for the photocatalytic hydrogen evolution activity under visible light over various samples: (a) CdS NPs, (b) Pt/CdS, (c) Pt/CdS@Al2O3 with artificial gill and (d) Pt/CdS@Al2O3 without artificial gill. (a,b) Reproduced with permission from Ref. [112]. Copyright 2018, Elsevier. (c) HRTEM images of 10Ni2P@CdS and 1Pt@CdS. (d) The concentration of Cd2+ for each run in the reaction solution of CdS and 10Ni2P@CdS under visible light irradiation. (c,d) Reproduced with permission from Ref. [62]. Copyright 2018, Elsevier.
Catalysts 13 00355 g009
Figure 10. (a) Partial void/shell structure areas of deactivated materials show cleavages in between TiO2 and Ni metal. (b) Void/shell structures after 50% deactivation. Reproduced with permission from Ref. [64]. (a,b) Copyright 2015, American Chemistry Society.
Figure 10. (a) Partial void/shell structure areas of deactivated materials show cleavages in between TiO2 and Ni metal. (b) Void/shell structures after 50% deactivation. Reproduced with permission from Ref. [64]. (a,b) Copyright 2015, American Chemistry Society.
Catalysts 13 00355 g010
Table 1. The values of work function (Φm) of different noble metals and non-noble metals (eV).
Table 1. The values of work function (Φm) of different noble metals and non-noble metals (eV).
Noble MetalsΦmNon-Noble MetalΦm
Pt5.65Fe4.50
Pd5.12Co5.00
Rh4.98Ni5.15
Au5.12Cu4.65
Ag4.26
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tian, L.; Guan, X.; Zong, S.; Dai, A.; Qu, J. Cocatalysts for Photocatalytic Overall Water Splitting: A Mini Review. Catalysts 2023, 13, 355. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13020355

AMA Style

Tian L, Guan X, Zong S, Dai A, Qu J. Cocatalysts for Photocatalytic Overall Water Splitting: A Mini Review. Catalysts. 2023; 13(2):355. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13020355

Chicago/Turabian Style

Tian, Li, Xiangjiu Guan, Shichao Zong, Anna Dai, and Jingkuo Qu. 2023. "Cocatalysts for Photocatalytic Overall Water Splitting: A Mini Review" Catalysts 13, no. 2: 355. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13020355

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