Open access peer-reviewed chapter

On the Limits of Photocatalytic Water Splitting

Written By

Bahar Ipek and Deniz Uner

Submitted: 21 March 2019 Reviewed: 19 August 2019 Published: 16 October 2019

DOI: 10.5772/intechopen.89235

From the Edited Volume

Water Chemistry

Edited by Murat Eyvaz and Ebubekir Yüksel

Chapter metrics overview

1,815 Chapter Downloads

View Full Metrics

Abstract

The major drawbacks on the limited H2 and O2 evolution activities of one-step photocatalytic water splitting systems are given here with the emphasis on charge recombination, back-oxidation reactions, and mass transfer limitations. Suppression of these unwanted phenomena is shown to be possible with the usage of small crystal-sized photocatalysts with low defect concentrations, presence of phase junctions, selection of co-catalyst that would be active for H2 evolution but inactive for O2 reduction, coating of the co-catalyst or the whole photocatalyst with selectively permeable nanolayers, and usage of photocatalytic systems with high solid–liquid and liquid–gas surface areas. The mass transfer limitations are shown to be important especially in the liquid–gas interfaces for agitated and suspended systems with estimated H2 transfer rates in the range of ∼200–8000 μmol/h.

Keywords

  • hydrogen production
  • photocatalyst
  • water splitting
  • mass transfer
  • back-oxidation

1. Introduction

Hydrogen gas is one of the best alternatives to fossil fuels since it has a high gravimetric energy density (142 MJ/kg) and it produces zero carbon upon combustion. Hydrogen is also used as a major reactant in environmentally important reactions such as carbon dioxide hydrogenation to methanol [1] or ammonia production (Haber-Bosch reaction) [2]. For hydrogen to be used as a clean energy source, its production via renewable ways is of great importance. It is conventionally produced via steam reforming of methane and fossil fuels (energy intensive, ΔH0rxn = 206 kJ/mol, 700–1100°C [3]) and coal gasification, which results in significant amounts of carbon dioxide production. The renewable ways for carbon-free production include biological sources (microalgae and cyanobacteria) and electrolysis of water using wind energy and photovoltaic cells as electricity generation sources. In addition to the mentioned renewable ways, photocatalytic water splitting/oxidation is a promising alternative, in which solar energy is used as the driving force to split water molecules to hydrogen and oxygen on the surface of a catalyst. This renewable production method of hydrogen is advantageous over other renewable methods due to the free source of energy and lower cost of the photocatalysts when compared to that of photovoltaic cells or wind turbines. Solar-driven catalytic (photocatalytic) reactions are considered to be of fundamental importance to the catalysis community since the solar energy is inexhaustible; i.e., the solar energy absorbed by the lands and oceans on an hourly basis (432 EJ/h or 120,000 TW [4]) is comparable to the Earth’s yearly energy consumption (reaching 575 EJ/year or 18 TW in 2017). However, the solar-to-hydrogen energy conversion efficiency value for photocatalytic water splitting systems is much lower (targeted to be 10%, currently reaching 1% [5]) than that of photovoltaic-assisted electrolysis (reaching 30% [6]) due to the major drawbacks in the one-step photocatalytic water splitting systems. Herein, we firstly introduce photocatalytic water splitting systems and give the major developments in materials such as visible light utilization and corresponding H2 and O2 production activity values (in Section 2). Then in Section 3, we discuss the causes of the low efficiencies in photocatalytic water splitting systems and the recent approaches in preventing energy efficiency-lowering factors such as inefficient visible light utilization, charge recombination, back-oxidation reactions, and mass transfer limitations.

Advertisement

2. Photocatalysis and water splitting

The first report on water splitting via harvesting photon energy is authored by Fujishima and Honda using a photoelectrochemical cell with a TiO2 photoelectrode [7]. Following this first report suggesting the oxidation of water molecule via photo-generated holes on TiO2 surface with the aid of small electrical voltage, photocatalytic water splitting on powder photocatalyst particles is demonstrated by other authors in the late twentieth century [8, 9, 10, 11, 12, 13, 14, 15]. Metal-loaded semiconductors (such as Pt/TiO2) are described as “short-circuited photoelectrochemical cells” that provide both the oxidizing centers and the reduction centers on the same catalyst (see Figure 1) [16].

Figure 1.

Schematic representation of photocatalytic water splitting on metal-loaded semiconductor particle systems: (1) light absorption and charge excitation from valence band to conduction band, (2) transfer of the photo-generated electrons and holes to the catalyst surface, (3) surface redox reactions, and (4) charge recombination.

Photocatalytic reactions are initiated by absorption of light having an energy higher than (or equal to) the bandgap of the photocatalysts that consist of semiconductor materials. This bandgap energy should be larger than 1.23 V for overall water oxidation reaction, for which the maximum of the valence band and the minimum of the conduction band should be located at proper potentials for the oxygen and hydrogen evolution reactions to occur. To illustrate, the minimum of the conduction band energy level should be located at a more negative potential than 0 V vs. NHE, at pH = 0 for H2 evolution (Eq. (1)), and the maximum of the valence band should be at a more positive potential than 1.23 V vs. NHE at pH = 0 for oxygen evolution reaction (Eq. (2)):

4H + + 4e 2H 2 E1
2H 2 O + 4 h + 4H + + O 2 E2

Following the light absorption, photoexcited electrons are transferred to the conduction band, while a positively charged charge carrier (hole) is generated at the valence band. These charge carriers are then transferred to the catalyst surface (step 2 in Figure 1) to be utilized in surface redox reactions, unless they recombine in the bulk or on the surface (step 4). Ultimately, electrons and holes reduce/oxidize the adsorbed species on the catalyst surface (step 3), the products of which should then be desorbed from the surface to complete the overall process.

2.1 Semiconductors

TiO2, having a large bandgap (anatase: 3.2 eV), is the most commonly used photocatalyst due to its photostability, nontoxicity, and high activity (upon UV radiation λ < 387 nm). Following the report on water oxidation reaction [7], various photochemical reaction activities of TiO2 such as carbon dioxide reduction with H2O [17, 18, 19], alkene and alkyne hydrogenation [20, 21], CH3Cl oxidation [22], 1-octanol degradation [23], phenol degradation [24], surfactant degradation [25], and more have been reported. Detailed reviews on TiO2-based materials and photocatalytic performances can be found in literature [26, 27, 28].

As photostable and active TiO2 is, UV light requirement to activate the large bandgap of TiO2 motivated research for visible light active semiconductors as well as bandgap engineering for TiO2 such as nonmetal ion doping (N [29], C [30], F [31], S [32]). Substitution of lattice oxygen atoms by these anions is reported to shift the valence band level upward and narrow the bandgap to as low as 2.25 eV (∼550 nm) with 16.5% N doping [33].

Similar to TiO2, oxides of other transition metals with d0 (such as Ti4+, Zr4+, Nb5+, Ta5+, and W6+ [34, 35]) and d10 electronic configurations (such as Ga3+, In3+, Ge4+, Sn4+, and Sb5 [36, 37, 38]) are shown to possess large bandgap energies (>3 eV) due to the maximum valence band levels consisting O2p orbitals located near 3 V (vs. NHE at pH = 0). These d0 and d10 metal oxide catalysts are reported to show remarkable one-step photocatalytic water splitting activity under UV light irradiation [39] reaching 71% quantum yield with photocatalysts such as Al-doped SrTiO3 [40] or Zn-doped Ga2O3 [41]. The H2 and O2 evolution activity under UV radiation and the apparent quantum yields of some of these materials are given in Table 1. The apparent quantum yield is defined as the number of reacted electrons and holes divided by the number of incident photons on the photocatalysts. Table 1 is not intended to cover the whole range of particulate catalysts in literature but rather to give a selection of examples. A wider selection of d0 and d10 metal oxide particulate catalysts’ one-step water oxidation activity and apparent quantum yields can be found in the works of Kudo et al., Chen et al., and Domen et al. [39, 42, 43].

Semiconductor Co-catalyst Bandgap (eV) H2 activity (μmol/h) O2 activity (μmol/h) AQY (%) Reference
La2Ti2O7:Ba NiOx 3.26 5000 50 [44]
SrTiO3:Al Rh2−yCryO3 3.2 550* 280* 30 at 300 nm [35]
SrTiO3:Al (200–500 nm) Rh2−yCryO3 3.2 1372* 683* 56 at 365 nm [45]
SrTiO3:Al MoOy/RhCrOx 3.2 1800* 900* 69 at 365 nm [40]
NaTaO3 NiO 4.0 3390 1580 20 at 270 nm [46]
NaTaO3:La NiO 4.1 19,800 9700 56 at 270 nm [34]
Ga2O3:Zn NiO 4.4 4100 2200 20 [47]
Ga2O3:Zn Rh0.5Cr1.5O3 4.4 32,000 16,000 71 at 254 nm [41]

Table 1.

H2 and O2 evolution activity of d0 and d10 metal oxide particulate catalysts under UV light irradiation.

0.1 g of photocatalyst is used instead of 1 g.


The most remarkable upgrades in the apparent quantum yields are achieved by material engineering such as (i) doping the metal oxides/perovskites with cations having lower valences, (ii) decreasing the crystal sizes to submicron levels, and (iii) loading with H2/O2 evolution co-catalysts.

2.2 Co-catalysts

An important addition to the light-harvesting semiconductors is H2 evolution/O2 evolution co-catalysts on the surface. The early co-catalysts that have been widely used included the noble metals and transition metal oxides such as Pt [12, 13], Rh [10], Ru [48], Au [49, 50], and NiOx [11] that mainly promote the hydrogen evolution, and CoOx [51] and Fe [52], Mn [52], RuO2 [53], and IrO2 [54] that accelerate the oxygen evolution. These metals are considered to act as charge carrier sinks that suppress electron–hole pair recombination as well as increasing the reaction kinetics by lowering the activation energy of the redox reactions. Co-catalysts are also known to inhibit photodegradation of the photocatalysts such as oxysulfides and oxynitrides by generated holes due to the effective extraction of these holes by the co-catalysts [55, 56].

Following the works of noble metal co-catalysts, Domen et al. showed water splitting activity on SrTiO3 photocatalyst together with the co-catalyst NiO [57, 58], which became the choice of H2 evolution co-catalyst for many d0 and d10 metal oxides such as La2Ti2O7:Ba [44], La4CaTi5O17 [59], Rb4Nb6O17 [60], NaTaO3 [46], and Ga2O3:Zn [47]. The photocatalyst stability of NiO-loaded K2La2Ti3O10 is reported to increase by addition of a second co-catalyst, Cr, using a co-impregnation method [61]. Based on the promoting effect of Cr, a systematic study of Cr and various transition metals (such as Fe, Co, Ni, Cu, Ru, Rh, Pd, Ag, Pt) on (Ga1-xZnx)(N1-xOx) has been conducted [62], from which core-shell structures of core Rh nanoislands and shell Cr2O3 structures (10–30 nm in size) are found to promote H2 and O2 evolution reactions to significant levels [63].

2.3 Visible light utilization

The alterations to the semiconductors such as doping with low-valence cations, reducing particle sizes to submicron levels and obtaining a high degree of crystallinity help with the overall water splitting activity. However, activation of these photocatalysts uses a narrow portion of the solar spectrum (4%); i.e., the UV light sustains as a problem due to the large bandgap energies of these materials. To enable visible light utilization of the d0- and d10-type oxide semiconductors and to split water into H2 and O2 via one-step excitation, the valence band levels should be shifted upward without changing the conduction band potentials. One approach to do this is to use oxynitrides to make use of N2p states that lie at a more negative potential than O2p states. Emerging LaMgTa1−xO1+3xN2−3x [64] and Ga1−xZnxN1−xOx [65] oxynitrides are representatives of visible light active overall water splitting catalysts. Using N doping, the absorbed light wavelength can be increased up to 500 nm on solid solutions of GaN:ZnO (Ga1−xZnxN1−xOx) [66] and up to 600 nm on solid solutions of LaTaON2 and LaMg2/3Ta1/3O3 (LaMgTa1−xO1+3xN2−3x) [64]. Other examples include LaScxTa1−xO1+2xN2−2x [67] and CaTaO2N [68] in which La or Ta sites are replaced by Ca and Sc that alters O/N ratios due to charge compensation, which in turn results in valence band energy level shift.

Some examples of visible light active photocatalyst and their H2 and O2 evolution activity are given in Table 2. As it can be seen from the table, one-step water splitting quantum yields are quite lower when compared to those of the UV-activated photocatalysts (Table 1). The exceptions to the low activity are reported by Rh2-yCryO3 (Rh 1.0 wt%, Cr 1.5 wt%)-loaded (Ga1−xZnx)(N1−xOx) photocatalyst [63], multiband InGaN/GaN nanowire arrays [69], and monodisperse 4 nm graphite nanoparticle-deposited C3N4 catalysts [70].

Semiconductor Co-catalyst Bandgap (eV) H2 activity (μmol/h) O2 activity (μmol/h) AQY (%) Reference
SrTiO3:Rh,Sb IrO2 4.4 1.9 0.1 at 420 nm [71]
g-C3N4 Pt-CoOx 2.8 ∼8.5 ∼3.5 0.3 at 405 nm [72]
CDots-C3N4 2.74 46 16 at 420 nm [70]
Bi1-xInxV1-xMoxO4 RuO2 2.5 17 3.2 at 420 nm [73]
BiYWO6 RuO2 2.7 4.1 1.8 0.17 at 420 nm [74]
LaMg1/3Ta2/3O2N RhCrOx 22 11 0.18 at 440 nm [75]
(Zn0.18Ga0.82)(N0.82O0.18) Rh2-yCryO3 2.64 927 460 5.9 at 420 nm [76]
GaN:Mg/InGaN:Mg Rh/Cr2O3 2.22 38 21 12.3 at 400 nm [69]

Table 2.

H2 and O2 evolution activity of one-step water splitting catalysts under visible light irradiation.

An alternative way to cover both oxidation and reduction reactions with semiconductors that could be activated under visible light radiation is to utilize two individual photocatalysts with an electron transfer mediator to obtain two-step excitation known as the two-step water oxidation (“Z-scheme system,” see Figure 2). In this system, O2 evolution photocatalysts oxidize the water molecules to O2, while the photo-generated electron is transferred to the mediator to reduce the electron acceptor (such as Fe3+ ions or IO3 ions). Then, the reduced mediator is oxidized by donating its electron to the H2 evolution photocatalyst. At the same time, the photo-generated electrons in the H2 evolution photocatalyst reduce H+s to H2.

Figure 2.

Schematic diagram for photocatalytic water splitting using a two-step photoexcitation system.

The semiconductors used in this two-step water splitting process should be selected based on the energy levels of their corresponding valence or conduction band maximum/minimum that would enable O2/H2O oxidation and H+/H2 reduction. As H2 evolution and O2 evolution reactions are realized at separate photocatalysts, these semiconductors could have bandgap energy values lower than 3 eV that would enable visible light utilization such as Pt- or RuO2-loaded WO3 (Eg ∼ 2.8 eV) or oxynitrides such as TaON (Eg ∼ 2.4 eV) or Rh-doped SrTiO3 (Eg ∼ 2.4 eV). Examples of these materials and systems can be seen in Table 3. The detailed reviews on two-step photocatalytic water splitting can be found elsewhere [83].

H2 photocatalyst O2 photocatalyst Mediator H2 activity (μmol/h) O2 activity (μmol/h) AQY (%) Reference
Pt/SrTiO3:Rh BiVO4 Fe3+/Fe2+ 15 7.2 0.4 at 420 nm [77]
Pt/SrTiO3:Cr/Ta PtOx/WO3 IO3/I ∼16 ∼8 1 at 420 nm [78]
Pt/TaON PtOx/WO3 IO3/I ∼16.5 ∼8 0.5 at 420 nm [79]
Pt/ZrO2/TaON PtOx/WO3 IO3/I 52 27 6.3 at 420 nm [80]
Ru/SrTiO3:Rh BiVO4 Fe3+/Fe2+ 88 44 4.2 at 420 nm [81]
Pt/MgTa2O6-xNy/ TaON PtOx/WO3 IO3/I 108 55 6.8 at 420 nm [82]

Table 3.

Z-scheme-type photocatalysts for water splitting without sacrificial agents.

Advertisement

3. Drawbacks on photocatalytic activity

There are numerous and challenging processes that need to be realized for photocatalytic evolution of H2 and O2 (Table 4) via a thermodynamically unfavorable reaction (Eq. (3)):

H 2 O H 2 + 1 / 2 O 2 ΔG 0 = 237 kJ / mol E3

These processes include (i) excitation of the semiconductor photocatalyst with photon having higher energies than the bandgap energy of the material, (ii) transfer of the photo-generated electrons and holes to the reaction sites on the surface, (iii) utilization of these charge carriers in the oxidation/reduction reactions, and (iv) desorption of the products from the surface of the photocatalyst to the liquid/gas medium.

As the timescale of these processes varies, recombination of the electrons and holes in the bulk or on the surface happens more frequently than the rate of the chemical oxidation/reduction reactions. Recombination is therefore considered to be one of the main reasons limiting the photocatalytic activity. Together with the recombination events, realization of back-oxidation reactions (Eq. (4)) on noble metals and the rate-limiting mass transfer events are the major drawbacks in an efficient photocatalytic process:

H 2 + 1 / 2O 2 H 2 O ΔG 0 = 237 kJ / mol E4

Natural photosynthesis yields a much higher rate of O2 evolution (see Table 4) when compared to artificial water splitting due to improved charge carrier and mass transfer events. From this comparison, it is clear that the photocatalytic systems still need to be perfected to compete with the nature’s intricate design.

3.1 Charge recombination

Due to the presence of the multiple processes, the overall photocatalytic reactions are extremely complicated. In order to obtain an efficient photocatalytic performance, the photo-generated charges must be transferred to the surface reaction sites as rapidly as possible while preventing recombination or trapping of these charge carriers. It is reported by Leytner and Hupp that 60% of the trapped electron–hole pairs recombine with a timescale of about 25 ns while releasing heat of 154 kJ/mol [85]. As the defects such as vacancies and dislocations are considered as recombination sites, higher crystallinity of the photocatalysts is often aimed to decrease the recombination rates. From diffusion point of view, the shorter distances for the charge carriers to the surface reaction centers are also aimed to prevent the recombination. Shorter pathways are achieved via smaller crystal/particle sizes of the photocatalysts. More than two times of increase in the H2 and O2 evolution rates on Al-doped SrTiO3 photocatalyst (reaching an apparent quantum yield of 56% [45]) as the particle size drops from few micrometers to 200 nm is a direct evidence of the effect of the particle size. Another method for reducing the charge recombination is to make use of phase junctions. One example is the α-β-phase junction of Ga2O3, which results in enhanced interfacial charge transfer, charge separation, and therefore enhanced water splitting activity [86]. Loading the photocatalysts with co-catalysts such as noble metals or transition metal oxides to accelerate the reduction/oxidation reactions is a commonly employed method. These co-catalysts are known to enhance the charge migration from the semiconductor depending on the alignment of the potentials of the semiconductor and the co-catalyst. As these co-catalysts accelerate the desired H2 evolution and O2 evolution reactions, they can also increase the rates of undesired secondary reactions such as hydrogen oxidation or oxygen reduction to water reactions.

3.2 Back-oxidation reactions

Introduction of one-step photocatalysts for overall water splitting combined the H2 evolution and O2 evolution sites on the same catalyst surface. This design of a photocatalytic system that realizes both charge trapping and reduction/oxidation reactions on the same surface not only accelerated the charge recombination but also allowed secondary reactions on these reduction/oxidation centers. When fast removal of the products, i.e., H2 and O2, is not provided and there are no barriers that prevent interaction of these products with highly active sites, the reaction of H2 and O2 on the photocatalyst surface to produce H2O (2H2 + O2 → H2O) is highly probable. And back-oxidation of the produced H2 is considered to be one of the main reasons for observed low photocatalytic water splitting activity values.

As early as 1985, Sato and coworkers realized the importance of back-oxidation of H2 with O2 to produce H2O. They have realized that the metal-loaded photocatalysts, mainly Pt- or Pd-loaded TiO2, can oxidize H2 with O2 easily under the same photocatalytic water oxidation conditions. They have reported first-order reaction rate constants in the range of 0.23–0.51 h−1 for Pt, 0.32–1.8 h−1 for Pd, and 0.2–0.3 h−1 for Rh, suggesting the least active metal for back-oxidation reaction to be Rh [10]. Later in 2000, Anpo and coworkers investigated back-oxidation reaction on Pt/TiO2 systems under dark conditions and observed increased back-oxidation rate with increasing Pt loading (up to 0.1 wt.% [87]). While Pt is active for H2 evolution (Eq. (1)), it is also notoriously active for dark H2–O2 recombination reaction (Eq. (4)) even at room temperature [88]. In order to prevent H2–O2 recombination reaction, the Pt surface is modified with F ions for Pt/TiO2 catalyst, and the reaction rate decreased from 2 to 0.3 h−1 upon F modification [89]. The inhibition mechanism is suggested to be due to the occupation of the H2 surface adsorption sites on Pt by F atoms.

Another modification to the noble metal surfaces is reported by Lercher et al., in which CO is chemisorbed on the Rh co-catalyst for GaN:ZnO semiconductor. Chemisorbed molecular layer of CO suppressed the back-oxidation reaction by selective metal poisoning of the back-oxidation sites by CO. While H2 evolution rates of 28 μmol/h are achieved (75 mg photocatalyst, 300 W Xe lamp [90]), significant CO oxidation to CO2 is also observed.

The back-oxidation reaction-inhibiting effects of the nanolayer coating on noble metals are shown on Rh/Cr2O3-loaded GaN:ZnO photocatalysts. Rh/Cr2O3 core-shell structure [91] is formed by photodeposition of Rh and reduction of CrO42− by electrons coming from Rh upon radiation, resulting in few nanometer thickness of Cr2O3 layer (2–3 nm, see Figure 3). Hydrated Cr2O3 nanolayer is reported to selectively permeate protons for H2 evolution reaction [92], whereas it hinders O2 permeation from the layer inhibiting O2 reduction reaction (Eq. (5)) on Rh sites [93]. The same effect is also valid for Cr2O3-coated Pt catalyst (GaN:ZnO). Back-oxidation rates on Pt-loaded GaN:ZnO photocatalyst decreased significantly from ∼105 × 10−12 molecules/s to ∼8 × 10−12 molecules/s, while photocatalytic H2 evolution rate increased from ∼5 × 10−12 molecules/s to ∼30 × 10−12 molecules/s upon Cr2O3 coating [93]. Apart from the oxygen-blocking role of the Cr2O3 nanolayer, much lower back-oxidation rate of Rh-loaded GaN:ZnO when compared to Pt-loaded GaN:ZnO (11 × 10−12 molecules/s vs. 105 × 10−12 molecules/s) explains the significant photocatalytic H2 evolution activity on Rh/Cr2O3-loaded GaN:ZnO (130 × 10−12 molecules/s). Lower back-oxidation rate of Rh-loaded GaN:ZnO could be related to the low-oxygen reduction reaction (Eq. 5) activity of Rh when compared to Pt [94]:

Figure 3.

HR-TEM images of GaN:ZnO photodeposited with (A) Rh and (B) Rh/Cr2O3. Reprinted with permission from [55]. Copyright 2007 American Chemical Society. (C) Schematic representation of O2 and H2 evolution reactions with inhibited O2 permeation and O2 reduction reaction on core-shell-type co-catalysts.

O 2 + 4H + + 4e 2H 2 O E5

Similar selective permeability concept is considered to be the case for Ni/NiO core-shell structures deposited on various photocatalysts such as SrTiO3 or NaTaO3 [34, 57]. In these systems, in addition to the back-oxidation reaction impeding effect of NiO layer on Ni [58], low-oxygen reduction activity of NiOx catalysts when compared to Pt can also be considered to be effective for improved water splitting activity.

Coatings of the whole photocatalyst instead of the co-catalyst by oxyhydroxides of Ti, Nb, and Ta are reported on Rh-loaded SrTiO3:Sc photocatalyst. Surface nanolayer not only suppressed back-oxidation reactions but also prevented the access of sacrificial agents such as ethanol to the photocatalyst surface, resulting in nearly stoichiometric H2/O2 ratios [95]. Surface nanolayer coatings on the whole photocatalysts have proven to also prevent photodecomposition (N2 evolution) of oxynitride photocatalysts while increasing the overall water splitting activity [64, 68].

Prevention of the secondary reactions such as H2 oxidation or O2 reduction reaction to H2O is found to be essential for improving the overall water splitting activity and the apparent quantum yield values (reaching apparent quantum yield value of 69% under irradiation at 365 nm [40]). In addition to the reduced back-oxidation rates, complementary measures such as decreasing the charge recombination rates and enhancing the product transfer rates away from the surface (increasing the mass transfer rates) are necessary for increased photocatalytic water splitting activity.

3.3 Mass transfer limitations

Mass transfer limitations especially in the slurry photocatalytic systems can be the most overlooked problem in the photocatalytic field. To complete the photocatalytic reaction cycle, adsorption of the reactants, reduction/oxidation of the reactants, desorption of the products, and transfer of the products from the photocatalyst surface to the gas phase need to be realized. When the rates of the mass transfer of the products from the surface are slower than the reduction/oxidation rates, produced H2 and O2 would stay longer on the surface, resulting in promotion of back-oxidation reactions. Moreover, when the mass transfer rates are slower than the reaction kinetics, the apparent H2 and O2 evolution rates in the gas phase will be limited by the mass transfer rates.

Experimental evidence for mass transfer limitations in agitated systems is presented in a previous publication [96]. In a batch slurry reactor, where the catalyst particles are suspended via agitation, observed H2 evolution rates for UV-irradiated Pt/TiO2 photocatalyst showed improvement with increasing stirring rates up to 900 rpm (Figure 4a). This improvement is a direct indication of mass transfer limitations on the solid–liquid and gas–liquid interfaces as the turbulence in the liquid and therefore boundary layers are affected by increasing stirring rates. In another experiment, the effect of liquid volume is investigated by varying catalyst weight and liquid volumes (keeping the catalyst concentration constant). H2 evolution rates on an hour basis (μmol H2/h) are found the same regardless of the liquid volume (or catalyst weight) above 62.5 ml (Figure 4b) as the H2 evolution rate per gram and hour basis decreased as liquid volume increased. Similar H2 evolution rates regardless of the catalyst weight indicate significant mass transfer limitations in the liquid–gas interface.

Figure 4.

(a) Effect of stirring rate on photocatalytic hydrogen evolution with methanol as sacrificial agent, with 0.5 wt% Pt/TiO2, 250 ml deionized water, 2 ml methanol, () 900 rpm and () 350 rpm. (b) Observed hydrogen evolution rates in the gas phase with changing liquid volume, CH3OH/H2O:1/125 (v/v) and Ccatalyst: 1 g/L for each case. Adapted from [96].

Mass transfer limitations for different photocatalytic reaction systems are analyzed by different groups. For immobilized photocatalyst systems, the importance of internal mass transfer resistance is emphasized [97]. In another investigation, severe mass transfer limitations are observed in the product separator (liquid–gas interface) for a fluidized bed/separator system, in which modification of the liquid–gas surface area enhanced the H2 evolution rates by 350% [98].

To prevent mass transfer limitations in the photocatalytic tests and to report actual kinetic rates; stirring rates, liquid levels, and mass transfer areas should be designed carefully. To design these parameters, approximate mass transfer rates should be known. Here, we present a sample calculation for H2 mass transfer rate in a slurry reactor containing 0.5 g TiO2 photocatalyst having a surface area of 40 m2/g inside an agitated glass reactor having 200 ml liquid volume and a tank diameter of 7 cm.

Mass transfer resistances in a gas–liquid–solid multiphase photocatalytic systems involve the internal mass transfer, mass transfer from the solid catalyst particles to liquid (Eq. 6), transfer from the liquid bulk to the liquid interface (Eq. 7), and transfer from the liquid–gas interface to the gas phase (Eq. 8). Photocatalysts such as perovskites and TiO2 are known to be nonporous (unless mesoporous versions are prepared on purpose [99, 100]) and have surface area values between 5 and 50 m2/g. For nonporous photocatalysts, the internal mass transfer limitations can be discarded (Eq. 9). Hence, the H2 mass transfer rate equation will have a form containing the mass transfer resistances from the solid–liquid and liquid–gas interfaces as seen in Eq. 8:

r H 2 , S = k s a S C S C L E6
r H 2 , L = k L a L C L C L , i E7
r H 2 , G = k G a G C G , i C G E8
r H 2 = C H 2 , s H C H 2 , g 1 k s a S + 1 k L a L + H k G a G E9

The mass transfer limitations coming from the solid–liquid and liquid–gas interfaces may play important role depending on the photocatalytic reactor type. The most often used photocatalytic reactor systems such as slurry reactors have solid–liquid and liquid–gas phase interfaces that suspend its catalysts by agitation using an impeller or a magnetic stirrer. The convection mass transfer coefficient for solid–liquid interface of such a system could be estimated using Eq. 10 suggested by Armenante and Kirwan for agitated tanks using Kolmogorov’s theory for Reynold’s number calculation to consider the effect of solid particle size [101]:

Sh = k s d p D H 2 H 2 O = 2 + 0.52 Re 0.52 Sc 1 / 3 E10

where k s is the convection mass transfer coefficient from solid to liquid in (m/s), d p is the particle diameter in (m), D H 2 H 2 O is the diffusion coefficient of H2 in liquid water (m2/s), and Sc is the Schmidt number. A rough estimation for k s for such system can be found in Table 5.

Process Timescale
Light absorption and electron and hole generation Semiconductor e + h + fs
Photo-generated electron and hole transfer to the surface and trapping h VB + h trap +
e CB e trap
200 fs
50 ps
Recombination of charge carriers e tr + h tr + recombination >20 ns
Interfacial charge transfer e CB + O 2 O 2 10–100 μs
Observed O2 evolution* 2 H 2 O + 4 h + O 2 + 4 H + 37 s*
Photosynthesis of H2O oxidation 2 H 2 O + 4 D + TA O 2 + 4 H + + DTA 1.59 ms [84]

Table 4.

The processes occurring in photocatalytic water splitting on TiO2 and their timescales [27] and the references therein.

Based on 16,000 μmol O2/g/h O2 evolution rate on Rh0.5Cr1.5O3-doped Ga2O3:Zn upon illumination at 254 nm [41], assuming 10 m2/g surface area and 1015 sites/cm2 site density.


Parameter Unit Value
Particle size, d p μm 1
Density of water, ρ L kg/m3 997 at 25°C
Viscosity of water, μ L Pa s 890 × 10−6 at 25°C
Kinematic viscosity, ν m2/s 8.93 × 10−7
Schmidt number Sc = ν D H 2 H 2 O 141
Energy density, ε = Power mass of liquid m2/s3 25 for 5 W stirrer, 200 g solution
Reynold’s number Re = ε 1 / 3 d p 4 / 3 ν 0.033
Sherwood number, Sh = k s d p D H 2 H 2 O = 2 + 0.52 Re 0.52 Sc 1 / 3 2.46
Diffusion coefficient of H2 in water m2/s 6.30*10−9 at 25 °C
Convection mass transfer coefficient for solid, k s m/s 0.015
k s a S m3/s 0.31

Table 5.

Convection mass transfer coefficient calculation for solid–liquid transfer and parameters used in the calculation.

The creation of air/inert bubbles in the continuous phase (water) due to agitation could be considered as the transfer mechanism of produced H2 from the liquid phase to the gas phase. In such systems, comparing the mass transfer resistance from liquid to interface and interface to gas, it can be assumed that nearly all of the mass transfer resistance comes from the liquid side of the interface [102], leaving Eq. 9 as Eq. 11:

r H 2 = C H 2 , s C H 2 , Li 1 k s a S + 1 k L a L E11

The liquid side mass transfer coefficient for such a system could then be calculated using Calderbank and Moo-Young correlation for rising small bubbles of gas in continuous liquid phase (Eq. 12) [103]:

Sh = k L d b D H 2 H 2 O = 2 + 0.31 Ra 1 / 3 where Ra = d b 3 ρ L ρ G g μ L D H 2 H 2 O E12

The first term on Eq. 12 is the molecular diffusion term, whereas the second term is for the rise of the bubbles due to gravitational forces independent of the agitation. With estimations on the bubble size and gas holdup of such a system (given in Table 6), the mass transfer coefficient and k L a L term are calculated to be 2.5 × 10−4 m/s and 2.2 × 10−6 m3/s, rendering liquid–gas mass transfer resistance way more important than solid–liquid resistance.

Parameter Unit Value
Bubble size, d b μm 700
Density of water, ρ L kg/m3 997 at 25°C
Density of air, ρ G kg/m3 1.18 at 25°C
Viscosity of water, μ L Pa s 890 × 10−6 at 25°C
Raleigh number Ra = d b 3 ρ L ρ G g μ L D H 2 H 2 O 5.97 × 105
Sherwood number, Sh = k L d b D H 2 H 2 O = 2 + 0.31 Ra 1 / 3 28.1
Diffusion coefficient of H2 in water m2/s 6.3 × 10−9 at 25°C
Convection mass transfer coefficient for liquid, k L m/s 2.5 × 10−4
Liquid–gas bubble contact area, a L = 6 V G d b = 6 ϕ d b V L where ϕ = V G V L m2 0.008 (gas holdup, ϕ , assumed to be 0.005)
k L a L m3/s 2.2 × 10−6

Table 6.

Convection mass transfer coefficient calculation for liquid–gas transfer and parameters used in the calculation.

The overall mass transfer coefficient and the mass transfer rate from solid to the gas phase can be calculated with the estimated k s a S and k L a L values. As the concentration of H2 in the gas phase will be negligible ( C G ∼0), the liquid phase interface can also be assumed to be equal to zero ( C H 2 , Li ∼0) with negligible gas phase resistance. Therefore, from Eq. 11, the rate of H2 mass transfer can be calculated by assuming H2 concentration at the catalyst surface and the gas holdup ratio in the liquid. The rate of H2 mass transfer values for 200 ml of water and 0.5 g of catalyst are calculated in the range of ∼200–8000 ?mol/h (see Figure 5) for a surface H2 adsorption capacity range of 50–400 ?mol/g (H2 chemisorption on 0.1% Pt/TiO2 is reported to be ∼400 ?mol/g at room temperature [104]) and gas holdup ratio between 0.001 and 0.005.

Figure 5.

Calculated H2 mass transfer rate values (μmol H2/h) for gas holdup values of 0.001 and 0.005 and H2 adsorption capacity values in the range of 50–400 μmol/g. The liquid volume is taken as 200 ml and catalyst weight is taken as 0.5 g.

The H2 concentration on the solid surface and the gas–liquid contact area are not easy to estimate. However, for the limited gas–liquid area, the photocatalytic reaction rates above the calculated mass transfer rate will be suppressed due to the limiting mass transfer rates. Therefore, special care must be given for the UV-irradiated photocatalytic systems, in which observed H2 and O2 evolution rates are found to be close to the calculated mass transfer rates here (see Table 1).

These studies show that for each type of photocatalytic system that contains limited gas–liquid contact area or immobilized photocatalyst, mass transfer limitations should not be underestimated, and not only the materials but also the systems should be improved for better photocatalytic efficiencies.

Advertisement

4. Future of photocatalytic H2 evolution

The literature examples of photocatalytic water splitting activities show improvements in visible light utilization, charge separation, and prevention of back-oxidation reactions via fine tuning of photocatalyst materials that enabled more efficient water splitting systems. The efficiency of these systems working under sunlight is better defined with solar-to-hydrogen energy conversion efficiency (STH), i.e., hydrogen production rate times the Gibbs free energy for generating 1 mole of H2 divided by the power of incident sunlight (Eq. 13):

STH = mmol H 2 / s 237 10 3 J / mol P mW cm 2 A cm 2 × 100 E13

The estimated STH required for the particulate photocatalytic systems to be economically compatible with current H2 production technologies is 10% [104]. However, the highest STH values obtained with current developed photocatalysts are in the order of 1–2% (1.8% at 400–475 nm using Rh/Cr2O3-loaded GaN:Mg/InGaN:Mg photocatalyst [69] and 2% at 420 nm using CDots-C3N4 [70]). The STH conversion efficiency depends both on the catalytic activity and the extent of the utilization of sunlight that depends on the bandgap of the semiconductor. With current photocatalysts having absorption edges around 500 nm, even 100% apparent quantum yields would not guarantee 10% STH values [106]. For a photocatalyst to show 10% STH values, it should have absorption edges at least at 600 nm with apparent quantum yields around 60%. Under the light of these calculations, it can be said that the present photocatalysts having adsorption edge values around 450 nm and quantum yields around 10% are far from being utilized in commercial systems. In order to achieve targeted STH values, the photocatalysts with lower bandgap energies such as (oxy)nitrides and (oxy)sulfides should be improved for H2 evolution activities while ensuring their thermal stability and photostability.

Large-scale photocatalytic water splitting reactors are implemented with current low STH values as of 2015. The first example of large-scale photocatalytic water splitting utilized Pt-loaded C3N4 photocatalyst with sacrificial electron donor triethanolamine in a flat-panel-type photocatalytic reactor system in 2015 [107]. The solar-to-hydrogen conversion efficiency is reported to be 0.12%, for which the photocatalytic activity is monitored for 30 days. In such systems, where a sacrificial reagent such as triethanolamine or methanol is irreversibly oxidized at a more negative potential than water (thermodynamically more favorable) at the oxidation centers, the photo-generated charges can be more efficiently separated, thus increasing H2 evolution rates. However, in those systems, hydrogen production is not solely due to the water splitting; as the carbon- and hydrogen-containing “sacrificial agents” are being oxidized at the oxidation centers, they produce hydrogen as well as aldehydes, carboxylic acid, and carbon dioxide [108].

Another large-scale photoreactor is reported by Domen et al., who used Al-doped and RhCrOx-loaded SrTiO3 photocatalyst sheets in their 1 × 1 m water splitting panel [45]. The achieved STH value under simulated sunlight is 0.6% at 331 K and limited to a maximum value of 1.4% due to the large bandgap energy of the photocatalyst (3.2 eV). Improved STH values (reaching 1.1%) are shown to be possible on a two-step excitation system in 2016 using photocatalyst sheets having smaller bandgap energy values such as Mo-doped BiVO4 (2.4 eV) and La- and Rh-doped SrTiO3 [5].

Advertisement

5. Conclusion

The developments in the photocatalytic water splitting reactions are explained here with the emphasis on the one-step photocatalysis systems. The early photocatalyst improvements with bandgap engineering, co-catalyst usage, and size reductions are shown to contribute to the increased visible light-driven H2 evolution activity values. The main drawbacks in the present systems are discussed to be the charge recombination, back-oxidation reactions of the products into water, and mass transfer limitations especially in the three-phase systems. Using defect-free small crystals of photocatalysts and making use of phase junctions or metal co-catalysts are suggested to decrease charge recombination rates. Back-oxidation of H2 into water or oxygen reduction reaction to water is expected in many noble metal-containing particulate photocatalyst systems. The prevention of these unwanted secondary reactions is shown to be possible to some extent by modification of the noble metal surfaces. Some examples of these modifications are anion coating, partial adsorption of a poison, or nanolayer coating of the co-catalyst or the whole photocatalyst. Selective permeation property of the nanolayer coatings such as Cr2O3 is reported to suppress the back-oxidation rates, resulting in enhanced H2 and O2 evolution rates. Possible mass transfer limitations, limiting the observed rates in three-phase systems, are predicted especially in the liquid–gas interfaces. The literature examples attracted attention for the liquid–gas interfaces in suspended systems and internal mass transfer limitations for the immobilized photocatalyst systems. It is concluded that, in addition to the required developments in activities with suppression of charge recombination, back-oxidation, and mass transfer limitations, future of the photocatalytic systems would necessitate active and stable photocatalysts with narrower bandgap energies (to be activated at >600 nm) for achieving targeted 10% solar-to-hydrogen energy conversion efficiency value.

References

  1. 1. Centi G, Perathoner S. Opportunities and prospects in the chemical recycling of carbon dioxide to fuels. Catalysis Today. 2009;148:191-205
  2. 2. Erisman JW, Sutton MA, Galloway J, Klimont Z, Winiwarter W. How a century of ammonia synthesis changed the world. Nature Geoscience. 2008;1:636-639
  3. 3. Aasberg-Petersen K, Dybkjær I, Ovesen CV, Schjødt NC, Sehested J, Thomsen SG. Natural gas to synthesis gas—Catalysts and catalytic processes. Journal of Natural Gas Science and Engineering. 2011;3:423-459
  4. 4. Morton O. A new day dawning?: Silicon valley sunrise. Nature. 2006;32:19-22
  5. 5. Wang Q, Hisatomi T, Jia Q, Tokudome H, Zhong M, Wang C, et al. Scalable water splitting on particulate photocatalyst sheets with a solar-to-hydrogen energy conversion efficiency exceeding 1%. Nature Materials. 2016;15:611-615
  6. 6. Jia J, Seitz LC, Benck JD, Huo Y, Chen Y, Ng JWD, et al. Solar water splitting by photovoltaic-electrolysis with a solar-to-hydrogen efficiency over 30%. Nature Communications. 2016;7:1-6
  7. 7. Fujishima A, Honda K. Electrochemical photolysis of water at a semiconductor electrode. Nature. 1972;238:37-38
  8. 8. Sato S, White JM. Photodecomposition of water over Pt/TiO2 catalysts. Chemical Physics Letters. 1980;72:83-86
  9. 9. Wagner FT, Somorjai GA. Photocatalytic hydrogen production from water on Pt-free SrTiO3 in alkali hydroxide solutions. Nature. 1980;285:559-560
  10. 10. Yamaguti K, Sato S. Photolysis of water over metallized powdered titanium dioxide. Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases. 1985;81:1237-1246
  11. 11. Kudo A, Domen K, Ken-ichi M, Onishi T. Photocatalytic activities of TiO2 loaded with NiO. Chemical Physics Letters. 1987;133:517-519
  12. 12. Sayama K, Arakawa H. Effect of carbonate salt addition on the photocatalytic decomposition of liquid water over Pt-TiO2 catalyst. Journal of the Chemical Society, Faraday Transactions. 1997;93:1647-1654
  13. 13. Tabata S, Nishida H, Masaki Y, Tabata K. Stoichiometric photocatalytic decomposition of pure water in Pt/TiO2 aqueous suspension system. Catalysis Letters. 1995;34:245-249
  14. 14. Moon SC, Mametsuka H, Suzuki E, Anpo M. Stoichiometric decomposition of pure water over Pt-loaded Ti/B binary oxide under UV-radiation. Chemistry Letters. 1998;27:117-118
  15. 15. Sayama K, Arakawa H. Effect of Na2CO3 addition on photocatalytic decomposition of liquid water over various semiconductor catalysts. Journal of Photochemistry and Photobiology A: Chemistry. 1994;77:243-247
  16. 16. Bard AJ. Photoelectrochemistry. Science (80-). 1980;207:139-144
  17. 17. Anpo M, Yamashita H, Ichihashi Y, Ehara S. Photocatalytic reduction of CO2 with H2O on various titanium oxide catalysts. Journal of Elcetroanalytical Chemistry. 1995;396:21-26
  18. 18. Saladin F, Forss L, Kamber I. Photosynthesis of CH4 at a TiO2 surface from gaseous H2O and CO2. Journal of the Chemical Society, Chemical Communications. 1995:533-534
  19. 19. Yamashita H, Nishiguchi H, Kamada N, Anpo M, Teraoka Y, Hatano H, et al. Photocatalytic reduction of CO2 with H2O on TiO2 and Cu/TiO2 catalysts. Research on Chemical Intermediates. 1994;20:815-823
  20. 20. Anpo M, Aikawa N, Kubokawa Y. Photocatalytic hydrogenation of alkynes and alkenes with water over TiO2. Pt-loading effect on the primary processes. The Journal of Physical Chemistry. 1984;88:3998-4000
  21. 21. Anpo M, Nakaya H, Kodama S, Kubokawa Y, Domen K, Onishi T. Photocatalysis over binary metal oxides. Enhancement of the photocatalytic activity of titanium dioxide in titanium-silicon oxides. The Journal of Physical Chemistry. 1986;90:1633-1636
  22. 22. Choi W, Termin A, Hoffmann MR. The role of metal ion dopants in quantum-sized TiO2: Correlation between photoreactivity and charge carrier recombination dynamics. The Journal of Physical Chemistry. 1994;98:13669-13679
  23. 23. Yamashita H, Ichihashi Y, Harada M, Stewart G, Fox MA, Anpo M. Photocatalytic degradation of 1-octanol on anchored titanium oxide and on TiO2 powder catalysts. Journal of Catalysis. 1996;158:97-101
  24. 24. Trillas M, Pujol M, Domènech X. Phenol photodegradation over titanium dioxide. Journal of Chemical Technology and Biotechnology. 1992;55:85-90
  25. 25. Hidaka H, Zhao J. Photodegradation of surfactants catalyzed by a TiO2 semiconductor. Colloids and Surfaces. 1992;67:165-182
  26. 26. Kitano M, Matsuoka M, Ueshima M, Anpo M. Recent developments in titanium oxide-based photocatalysts. Applied Catalysis A: General. 2007;325:1-14
  27. 27. Schneider J, Matsuoka M, Takeuchi M, Zhang J, Horiuchi Y, Anpo M, et al. Understanding TiO2 photocatalysis: Mechanisms and materials. Chemical Reviews. 2014;114:9919-9986
  28. 28. Gaya UI, Abdullah AH. Heterogeneous photocatalytic degradation of organic contaminants over titanium dioxide: A review of fundamentals, progress and problems. Journal of Photochemistry and Photobiology C: Photochemistry Reviews. 2008;9:1-12
  29. 29. Asahi R, Morikawa T, Ohwaki T, Aoki K, Taga Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science (80- ). 2001;293:269-271
  30. 30. Parayil SK, Kibombo HS, Wu CM, Peng R, Baltrusaitis J, Koodali RT. Enhanced photocatalytic water splitting activity of carbon-modified TiO2 composite materials synthesized by a green synthetic approach. International Journal of Hydrogen Energy. 2012;37:8257-8267
  31. 31. Xing M, Qi D, Zhang J, Chen F, Tian B, Bagwas S, et al. Super-hydrophobic fluorination mesoporous MCF/TiO2 composite as a high-performance photocatalyst. Journal of Catalysis. 2012;294:37-46
  32. 32. Periyat P, Pillai SC, Mccormack DE, Colreavy J, Hinder SJ. Improved high-temperature stability and sun-light-driven photocatalytic activity of sulfur-doped anatase TiO2. Journal of Physical Chemistry C. 2008;112:7644-7652
  33. 33. Kitano M, Funatsu K, Matsuoka M, Ueshima M, Anpo M. Preparation of nitrogen-substituted TiO2 thin film photocatalysts by the radio frequency magnetron sputtering deposition method and their photocatalytic reactivity under visible light irradiation. The Journal of Physical Chemistry B. 2006;110:25266-25272
  34. 34. Kato H, Asakura K, Kudo A. Highly efficient water splitting into H2 and O2 over lanthanum-doped NaTaO3 photocatalysts with high crystallinity and surface nanostructure. Journal of the American Chemical Society. 2003;125:3082-3089
  35. 35. Ham Y, Hisatomi T, Goto Y, Moriya Y, Sakata Y, Yamakata A, et al. Flux-mediated doping of SrTiO3 photocatalysts for efficient overall water splitting. Journal of Materials Chemistry A. 2016;4:3027-3033
  36. 36. Ikarashi K, Sato J, Kobayashi H, Saito N, Nishiyama H, Inoue Y. Photocatalysis for water decomposition by RuO2-dispersed ZnGa2O4 with d10 configuration. The Journal of Physical Chemistry B. 2002;106:9048-9053
  37. 37. 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. The Journal of Physical Chemistry. B. 2001;105:6061-6063
  38. 38. Inoue Y. Photocatalytic water splitting by RuO2-loaded metal oxides and nitrides with d0- and d10-related electronic configurations. Energy & Environmental Science. 2009;2:364-386
  39. 39. Kudo A, Miseki Y. Heterogeneous photocatalyst materials for water splitting. Chemical Society Reviews. 2009;38:253-278
  40. 40. Chiang TH, Lyu H, Hisatomi T, Goto Y, Takata T, Katayama M, et al. Efficient photocatalytic water splitting using Al-doped SrTiO3 coloaded with molybdenum oxide and rhodium-chromium oxide. ACS Catalysis. 2018;8:2782-2788
  41. 41. Sakata Y, Hayashi T, Yasunaga R, Yanaga N, Imamura H. Remarkably high apparent quantum yield of the overall photocatalytic H2O splitting achieved by utilizing Zn ion added Ga2O3 prepared using dilute CaCl2 solution. Chemical Communications. Royal Society of Chemistry. 2015;51:12935-12938
  42. 42. Chen X, Shen S, Guo L, Mao SS. Semiconductor-based photocatalytic hydrogen generation. Chemical Reviews. 2010;110:6503-6570
  43. 43. Chen S, Takata T, Domen K. Particulate photocatalysts for overall water splitting. Nature Reviews Materials. 2017;2:1-17
  44. 44. Kim J, Hwang DW, Kim HG, Bae SW, Lee JS, Li W, et al. Highly efficient overall water splitting through optimization of preparation and operation conditions of layered perovskite photocatalysts. Topics in Catalysis. 2005;35:295-303
  45. 45. Goto Y, Hisatomi T, Wang Q, Higashi T, Ishikiriyama K, Maeda T, et al. A particulate photocatalyst water-splitting panel for large-scale solar hydrogen generation. Joule. 2018;2:509-520
  46. 46. Kato H, Kudo A. Water splitting into H2 and O2 on alkali tantalate photocatalysts ATaO3 (A = Li, Na, and K). The Journal of Physical Chemistry B. 2001;105:4285-4292
  47. 47. Sakata Y, Matsuda Y, Yanagida T, Hirata K, Imamura H, Teramura K. Effect of metal ion addition in a Ni supported Ga2O3 photocatalyst on the photocatalytic overall splitting of H2O. Catalysis Letters. 2008;125:22-26
  48. 48. Inoue Y, Niiyama T, Asai Y, Sato K. Stable photocatalytic activity of BaTi4O9 combined with ruthenium oxide for decomposition of water. Journal of Chemical Society, Chemical Communications. 1992:579-580
  49. 49. Iwase A, Kato H, Kudo A. Nanosized Au particles as an efficient cocatalyst for photocatalytic overall water splitting. Catalysis Letters. 2006;108:7-10
  50. 50. Murdoch M, Waterhouse GIN, Nadeem MA, Metson JB, Keane MA, Howe RF, et al. The effect of gold loading and particle size on photocatalytic hydrogen production from ethanol over Au/TiO2 nanoparticles. Nature Chemistry. 2011;3:489-492
  51. 51. Li R, Han H, Zhang F, Wang D, Li C. Highly efficient photocatalysts constructed by rational assembly of dual-cocatalysts separately on different facets of BiVO4. Energy & Environmental Science. 2014;7:1369-1376
  52. 52. Liu L, Ji Z, Zou W, Gu X, Deng Y, Gao F, et al. In situ loading transition metal oxide clusters on TiO2 nanosheets as co-catalysts for exceptional high photoactivity. ACS Catalysis. 2013;3:2052-2061
  53. 53. Maeda K, Wang X, Nishihara Y, Lu D, Antonietti M, Domen K. Photocatalytic activities of graphitic carbon nitride powder for water reduction and oxidation under visible light. Journal of Physical Chemistry C. 2009;113:4940-4947
  54. 54. Youngblood WJ, Lee S-HA, Kobayashi Y, Hernandez-Pagan EA, Hoertz PG, Moore TA, et al. Photoassisted overall water splitting in a visible light-absorbing dye-sensitized photoelectrochemical cell. Journal of the American Chemical Society. 2009;131:926-927
  55. 55. Maeda K, Domen K. New non-oxide photocatalysts designed for overall water splitting under visible light. Journal of Physical Chemistry C. 2007;111:7851-7861
  56. 56. Kamata K, Maeda K, Lu D, Kako Y, Domen K. Synthesis and photocatalytic activity of gallium-zinc-indium mixed oxynitride for hydrogen and oxygen evolution under visible light. Chemical Physics Letters. 2009;470:90-94
  57. 57. Domen K, Naito S, Soma M, Onishi T, Tamaru K. Photocatalytic decomposition of water vapour on an NiO–SrTiO3 catalyst. Journal of the Chemical Society, Chemical Communications. 1980:543-544
  58. 58. Domen K, Kudo A, Onishi T, Kosugi N, Kuroda H. Photocatalytic decomposition of water into H2 and O2 over NiO-SrTiO3 powder. 1. Structure of the catalysts. Journal of Physical Chemistry A. 1986;90:292-295
  59. 59. Kim HG, Hwang DW, Kim J, Kim YG, Lee JS. Highly donor-doped (1 1 0) layered perovskite materials as novel photocatalysts for overall water splitting. Chemical Communications. 1999;2:1077-1078
  60. 60. Sayama K, Arakawa H, Domen K. Photocatalytic water splitting on nickel intercalated A4TaxNb6−xO17 (A = K, Rb). Catalysis Today. 1996;28:175-182
  61. 61. Thaminimulla CTK, Takata T, Hara M, Kondo JN, Domen K. Effect of chromium addition for photocatalytic overall water splitting on Ni-K2La2Ti3O10. Journal of Catalysis. 2000;196:362-365
  62. 62. Maeda K, Teramura K, Saito N, Inoue Y, Domen K. Improvement of photocatalytic activity of (Ga1−xZnx)(N1−xOx) solid solution for overall water splitting by co-loading Cr and another transition metal. Journal of Catalysis. 2006;243:303-308
  63. 63. 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. Angewandte Chemie International Edition. 2006;45:7806-7809
  64. 64. Pan C, Takata T, Nakabayashi M, Matsumoto T, Shibata N, Ikuhara Y, et al. A complex perovskite-type oxynitride: The first photocatalyst for water splitting operable at up to 600 nm. Angewandte Chemie International Edition. 2015;54:2955-2959
  65. 65. 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. Journal of the American Chemical Society. 2012;134:8254-8259
  66. 66. Maeda K, Teramura K, Takata T, Hara M, Saito N, Toda K, et al. Overall water splitting on (Ga1−xZnx)(N1−xOx) solid solution photocatalyst: Relationship between physical properties and photocatalytic activity. The Journal of Physical Chemistry B. 2005;109:20504-20510
  67. 67. Pan C, Takata T, Kumamoto K, Khine Ma SS, Ueda K, Minegishi T, et al. Band engineering of perovskite-type transition metal oxynitrides for photocatalytic overall water splitting. Journal of Materials Chemistry A. 2016;4:4544-4552
  68. 68. Xu J, Pan C, Takata T, Domen K. Photocatalytic overall water splitting on the perovskite-type transition metal oxynitride CaTaO2N under visible light irradiation. Chemical Communications. 2015;51:7191-7194
  69. 69. Kibria MG, Nguyen HPT, Cui K, Zhao S, Liu D, Guo H, et al. One-step overall water splitting under visible light using multiband InGaN/GaN nanowire heterostructures. ACS Nano. 2013;7:7886-7893
  70. 70. Liu J, Liu Y, Liu N, Han Y, Zhang X, Huang H, et al. Metal-free efficient photocatalyst for stable visible water splitting via a two-electron pathway. Science (80-). 2015;347:970-974
  71. 71. Asai R, Nemoto H, Jia Q, Saito K, Iwase A, Kudo A. A visible light responsive rhodium and antimony-codoped SrTiO3 powdered photocatalyst loaded with an IrO2 cocatalyst for solar water splitting. Chemical Communications. 2014;50:2543-2546
  72. 72. Zhang G, Lan ZA, Lin L, Lin S, Wang X. Overall water splitting by Pt/g-C3N4 photocatalysts without using sacrificial agents. Chemical Science. 2016;7:3062-3066
  73. 73. Jo WJ, Kang HJ, Kong K-J, Lee YS, Park H, Lee Y, et al. Phase transition-induced band edge engineering of BiVO4 to split pure water under visible light. Proceedings of the National Academy of Sciences. 2015;112:13774-13778
  74. 74. Liu H, Yuan J, Shangguan W, Teraoka Y. Visible-light-responding BiYWO3 solid solution for stoichiometric photocatalytic water splitting. Journal of Physical Chemistry C. 2008;112:8521-8523
  75. 75. Pan C, Takata T, Domen K. Overall water splitting on the transition-metal oxynitride photocatalyst LaMg1/3Ta2/3O2N over a large portion of the visible-light spectrum. Chemistry: A European Journal. 2016;22:1854-1862
  76. 76. Maeda K, Teramura K, Domen K. Effect of post-calcination on photocatalytic activity of (Ga1−xZnx)(N1−xOx) solid solution for overall water splitting under visible light. Journal of Catalysis. 2008;254:198-204
  77. 77. Kato H, Hori M, Konta R, Shimodaira Y, Kudo A. Construction of Z-scheme type heterogeneous photocatalysis systems for water splitting into H2 and O2 under visible light irradiation. Chemistry Letters. 2004;33:1348-1349
  78. 78. Abe R, Sayama K, Sugihara H. Development of new photocatalytic water splitting into H2 and O2 using two different semiconductor photocatalysts and a shuttle redox mediator IO3/I. The Journal of Physical Chemistry. B. 2005;109:16052-16061
  79. 79. Abe R, Higashi M, Domen K. Overall water splitting under visible light through a two-step photoexcitation between TaON and WO3 in the presence of an iodate-iodide shuttle redox mediator. ChemSusChem. 2011;4:228-237
  80. 80. Maeda K, Higashi M, Lu D, Abe R, Domen K. Efficient nonsacrificial water splitting through two-step photoexcitation by visible light using a modified oxynitride as a hydrogen evolution photocatalyst. Journal of the American Chemical Society. 2010;132:5858-5868
  81. 81. Kato H, Sasaki Y, Shirakura N, Kudo A. Synthesis of highly active rhodium-doped SrTiO3 powders in Z-scheme systems for visible-light-driven photocatalytic overall water splitting. Journal of Materials Chemistry A. 2013;1:12327-12333
  82. 82. Chen S, Qi Y, Hisatomi T, Ding Q, Asai T, Li Z, et al. Efficient visible-light-driven Z-scheme overall water splitting using a MgTa2O6−xNy/TaON heterostructure photocatalyst for H2 evolution. Angewandte Chemie, International Edition. 2015;54:8498-8501
  83. 83. Wang Y, Suzuki H, Xie J, Tomita O, Martin DJ, Higashi M, et al. Mimicking natural photosynthesis: Solar to renewable H2 fuel synthesis by Z-scheme water splitting systems. Chemical Reviews. 2018;118:5201-5241
  84. 84. Haumann M, Liebisch P, Muller C, Barra M, Grabolle M, Dau H. Photosynthetic O2 formation tracked by time-resolved X-ray experiments. Science (80- ). 2005;310:1019-1021
  85. 85. Leytner S, Hupp JT. Evaluation of the energetics of electron trap states at the nanocrystalline titanium dioxide/aqueous solution interface via time-resolved photoacoustic spectroscopy. Chemical Physics Letters. 2000;330:231-236
  86. 86. Wang X, Xu Q, Li M, Shen S, Wang X, Wang Y, et al. Photocatalytic overall water splitting promoted by an α-β phase junction on Ga2O3. Angewandte Chemie International Edition. 2012;51:13089-13092
  87. 87. Yoshida Y, Matsuoka M, Moon SC, Mametsuka H, Suzuki E, Anpo M. Photocatalytic decomposition of liquid-water on the Pt-loaded TiO2 catalysts: Effects of the oxidation states of Pt species on the photocatalytic reactivity and the rate of the back reaction. Research on Chemical Intermediates. 2000;26:567-574
  88. 88. Sanap KK, Varma S, Dalavi D, Patil PS, Waghmode SB, Bharadwaj SR. Variation in noble metal morphology and its impact on functioning of hydrogen mitigation catalyst. International Journal of Hydrogen Energy. 2011;36:10455-10467
  89. 89. Wang M, Li Z, Wu Y, Ma J, Lu G. Inhibition of hydrogen and oxygen reverse recombination reaction over Pt/TiO2 by F− ions and its impact on the photocatalytic hydrogen formation. Journal of Catalysis. 2017;353:162-170
  90. 90. Berto TF, Sanwald KE, Byers JP, Browning ND, Gutiérrez OY, Lercher JA. Enabling overall water splitting on photocatalysts by CO-covered noble metal co-catalysts. Journal of Physical Chemistry Letters. 2016;7:4358-4362
  91. 91. Maeda K, Teramura K, Lu D, Saito N, Inoue Y, Domen K. Roles of Rh/Cr2O3 (core/shell) nanoparticles photodeposited on visible-light-responsive (Ga1−xZnx)(N1−xOx) solid solutions in photocatalytic overall water splitting. Journal of Physical Chemistry C. 2007;111:7554-7560
  92. 92. Yoshida M, Takanabe K, Maeda K, Ishikawa A, Kubota J, Sakata Y, et al. Role and function of noble-metal/Cr-layer core/shell structure cocatalysts for photocatalytic overall water splitting studied by model electrodes. Journal of Physical Chemistry C. 2009;113:10151-10157
  93. 93. Dionigi F, Vesborg PCK, Pedersen T, Hansen O, Dahl S, Xiong A, et al. Suppression of the water splitting back reaction on GaN:ZnO photocatalysts loaded with core/shell cocatalysts, investigated using a μ-reactor. Journal of Catalysis. 2012;292:26-31
  94. 94. Nørskov JK, Rossmeisl J, Logadottir A, Lindqvist L, Kitchin JR, Bligaard T, et al. Origin of the overpotential for oxygen reduction at a fuel-cell cathode. The Journal of Physical Chemistry B. 2004;108:17886-17892
  95. 95. Takata T, Pan C, Nakabayashi M, Shibata N, Domen K. Fabrication of a core-shell-type photocatalyst via photodeposition of group IV and V transition metal oxyhydroxides: An effective surface modification method for overall water splitting. Journal of the American Chemical Society. 2015;137:9627-9634
  96. 96. Ipek B, Uner D. Artificial photosynthesis from a chemical engineering perspective. In: Najafpour M, editor. Artificial Photosynthesis. IntechOpen; 2012. pp. 13-36
  97. 97. Chen D, Li F, Ray AK. External and internal mass transfer effect on photocatalytic degradation. Catalysis Today. 2001;66:475-485
  98. 98. Reilly K, Wilkinson DP, Taghipour F. Photocatalytic water splitting in a fluidized bed system: Computational modeling and experimental studies. Applied Energy. 2018;222:423-436
  99. 99. Niu B, Wang X, Wu K, He X, Zhang R. Mesoporous titanium dioxide: Synthesis and applications in photocatalysis, energy and biology. Materials (Basel). 2018;11:1-23
  100. 100. Sangle AL, Singh S, Jian J, Bajpe SR, Wang H, Khare N, et al. Very high surface area mesoporous thin films of SrTiO3 grown by pulsed laser deposition and application to efficient photoelectrochemical water splitting. Nano Letters. 2016;16:7338-7345
  101. 101. Armenante PM, Kirwan DJ. Mass transfer to microparticles in agitated systems. Chemical Engineering Science. 1989;44:2781-2796
  102. 102. Escudero JC, Simarro R, Cervera-March S, Gimenez J. Rate-controlling steps in a three-phase (solid-liquid-gas) photoreactor: A phenomenological approach applied to hydrogen photoproduction using Pt-TiO2 aqueous suspensions. Chemical Engineering Science. 1989;44:583-593
  103. 103. Calderbank PH, Moo-Young MB. The continuous phase heat and mass-transfer properties of dispersions. Chemical Engineering Science. 1961;16:39-54
  104. 104. Uner D, Tapan NA, Özen I, Üner M. Oxygen adsorption on Pt/TiO2 catalysts. Applied Catalysis A: General. 2003;251:225-234
  105. 105. Pinaud BA, Benck JD, Seitz LC, Forman AJ, Chen Z, Deutsch TG, et al. Technical and economic feasibility of centralized facilities for solar hydrogen production via photocatalysis and photoelectrochemistry. Energy & Environmental Science. 2013;6:1983-2002
  106. 106. Hisatomi T, Takanabe K, Domen K. Photocatalytic water-splitting reaction from catalytic and kinetic perspectives. Catalysis Letters. 2015;145:95-108
  107. 107. Schröder M, Kailasam K, Borgmeyer J, Neumann M, Thomas A, Schomäcker R, et al. Hydrogen evolution reaction in a large-scale reactor using a carbon nitride photocatalyst under natural sunlight irradiation. Energy Technology. 2015;3:1014-1017
  108. 108. Schneider J, Bahnemann DW. Undesired role of sacrificial reagents in photocatalysis. Journal of Physical Chemistry Letters. 2013;4:3479-3483

Written By

Bahar Ipek and Deniz Uner

Submitted: 21 March 2019 Reviewed: 19 August 2019 Published: 16 October 2019