Next Article in Journal
Surface Modifications of 2D-Ti3C2O2 by Nonmetal Doping for Obtaining High Hydrogen Evolution Reaction Activity: A Computational Approach
Previous Article in Journal
Catalytic Reaction of Carbon Dioxide with Methane on Supported Noble Metal Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Developments in the Use of Heterogeneous Semiconductor Photocatalyst Based Materials for a Visible-Light-Induced Water-Splitting System—A Brief Review

by
Prabhakarn Arunachalam
1,2,*,
Keiji Nagai
3,
Mabrook S. Amer
1,2,4,
Mohamed A. Ghanem
1,2,
Rajabathar Jothi Ramalingam
2 and
Abdullah M. Al-Mayouf
1,2,4
1
Electrochemical Sciences Research, Chemistry Department, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
2
Department of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
3
Division of Photoenergy Conversion Materials, Laboratory for Chemistry and Life Science, Institute of Innovative Research, Tokyo Institute of Technology, R1-26 Suzukake-dai, Midori-ku, Yokohama 226-8503, Japan
4
K.A.CARE Energy Research and Innovation, Riyadh 11454, Saudi Arabia
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(2), 160; https://doi.org/10.3390/catal11020160
Submission received: 3 December 2020 / Revised: 6 January 2021 / Accepted: 11 January 2021 / Published: 25 January 2021

Abstract

:
Visible-light-driven photoelectrochemical (PEC) and photocatalytic water splitting systems featuring heterogeneous semiconductor photocatalysts (oxynitrides, oxysulfides, organophotocatalysts) signify an environmentally friendly and promising approach for the manufacturing of renewable hydrogen fuel. Semiconducting electrode materials as the main constituents in the PEC water splitting system have substantial effects on the device’s solar-to-hydrogen (STH) conversion efficiency. Given the complication of the photocatalysis and photoelectrolysis methods, it is indispensable to include the different electrocatalytic materials for advancing visible-light-driven water splitting, considered a difficult challenge. Heterogeneous semiconductor-based materials with narrower bandgaps (2.5 to 1.9 eV), equivalent to the theoretical STH efficiencies ranging from 9.3% to 20.9%, are recognized as new types of photoabsorbents to engage as photoelectrodes for PEC water oxidation and have fascinated much consideration. Herein, we spotlight mainly on heterogenous semiconductor-based photoanode materials for PEC water splitting. Different heterogeneous photocatalysts based materials are emphasized in different groups, such as oxynitrides, oxysulfides, and organic solids. Lastly, the design approach and future developments regarding heterogeneous photocatalysts oxide electrodes for PEC applications and photocatalytic applications are also discussed.

1. Introduction

The accessibility of an economical, clean, and copious energy source is one of the foremost challenges for humankind in the 21st century. Effective utilization of solar energy and converting it to use fuels that are more appropriate for storage and transportation is mandatory. Solar power provides a clean, renewable energy source with minimal ecological impact since it is a decentralized and virtually unlimited resource. In detail, the solar power reaching the earth’s surface is equivalent to that supplied by 130 million power plants of 500 megawatts [1,2,3]. However, the forthcoming use of solar energy as a major source to satisfy the worldwide requirements depends on accomplishing important milestones of a fundamental and technological nature. The feasibility of the enlargement of solar energy at the terawatt scale depends on finding cost-efficient compatible solutions to these difficulties [4,5,6].
The manufacture of solar hydrogen (H2) through PEC water splitting as a proficient and renewable approach to arrest recurrent solar irradiance in the method of chemical fuels has drained remarkable consideration since the revolutionary report by Fujishima and Honda [7]. PEC water photoelectrolysis reaction is determined by solar energy generates an ecological approach to store energy in the covalent bonds of H2 [8,9]. During earlier stages of research, stoichiometric water splitting was not successful in a theoretical way because of several difficulties, the simultaneous, reverse process, unstable nature of photocatalyst, poor variation in selecting photocatalyst and co-catalyst materials, etc. However, significant advancement has been executed, and many complications and hurdles have been overcome [9,10,11,12,13,14,15]. In this context, substantial research efforts have been carried out to upsurge the STH efficiency of those solar-driven photocatalysts. Thermodynamically, the conversion of one molecule of H2O into H2 and ½ O2 is an uphill reaction with a great positive change in free energy change (ΔG° = 238 kJ/mol). In order to efficiently transform solar photons into usable fuels, n-type semiconductor materials can absorb a major portion of the spectral area, with appropriate energy band levels, and lower overpotential to execute a water photoelectrolysis reactions are mandatory. Extensive research efforts have been completed to promote viable n-type semiconducting electrodes that have emerged as photoanodes [14,15].
The oxygen evolution reaction (OER) is mutually kinetically and thermodynamically more exciting, and it hurts from severe oxidizing circumstances [16]. Moreover, the PEC water splitting system is one of the promising methods to examine the electron transfer process in producing oxygen (O2) in water oxidation reaction to use solar energy [17,18,19,20,21]. In this regard, several oxide-based electrode materials have been investigated [22,23,24,25,26,27,28,29,30,31,32,33,34,35]. On the other hand, all these metal oxides have absorption activity restricted to the UV spectral region due to its wider band gap (>3.0 eV). This is a big drawback since the UV region comprises only 3–5% of all incident solar energy. Further, n-type oxide semiconductor materials such as Fe2O3, BiVO4, and WO3 photoelectrode materials have been extensively applied to gather the visible-light photons for highly proficient as well as steady PEC water splitting system and involve a large external bias [36,37,38,39,40,41,42,43,44,45].
It is challenging to develop cost-effective and steady photoelectrodes with STH efficiencies of over 10%. In this regard, a supreme semiconductor light absorber must satisfy a minimum of three simple necessities: (i) its bandgap must be narrow (varied between 1.8 and 2.4 eV) to display a strong light absorption, and its band energy levels must overlap water redox potentials for naturally splitting water into H2 and O2; (ii) it must possess great carrier mobilities, and thereby permit quick transport of photoinduced hole-electron pairs from the bulk to the surface; (iv) its surface assembly must be active for water adsorption and successive reactions; (v) it must display lower charge-transfer resistance at the interface. Moreover, highly stable, cost-effective, and ecofriendly photoelectrode materials are mandatory for solar photolysis.
Considering all these conditions for a material to carry out water photolysis (energy levels of the conduction band (CB) and valence band (VB) concerning oxidation and reduction potential of water), most heterogeneous semiconductor-based photocatalysts can also reduce the water. Figure 1 shows the energy for numerous oxide and heterogeneous photocatalysts oxide semiconductors with the water redox reactions, illustrating this fact.
The definite processes in PEC water photolysis are defined in Figure 2. In order to execute this, several steps are needed to gratify the execution of the overall PEC water photolysis reactions. These steps comprise (i) the capture of visible-light photons; (ii) separation of carriers; (iii) transportation of the separated electron/hole pairs through the charge transfer; which need to be catalytically active; and (iv) surface reactions. Both the potential of the excited charge carriers and suitable water oxidation kinetics are mandatory for proficient water photoelectrolysis reactions.
In recent years, organic semiconductors (OSCs) with heterojunctions have generally been utilized to construct cost-efficient and effective photocatalysts for photoenergy conversion systems [46,47,48]. In some cases, photocatalytic reactions via organic p–n bilayers react to the whole visible-light energy for the oxidative elimination of organic pollutants. In another instance, an organic p-n bilayer of phthalocyanine [MPc (M = H2 or Zn), p-type] and perylene derivatives (3,4,9,10-perylenetetracarboxyl-bisbenzimidazole, PTCBI) or Pt-loaded fullerene (C60, n-type) can act as a photocathode, whereas a reducing power can be produced at the C60/water interface via a sequence of photophysical events in the bilayer [49,50,51,52]. Moreover, it was recently evidenced that an organic p-n bilayer can engage near-infrared energy to carry out photocatalysis [53]. In recent years, our group employed a wet-chemical process of a reprecipitation through a polar solvent in creating biphasic nanocrystals of aluminum phthalocyanine (AlPc)/cobalt phthalocyanine (CoPc) and C60 [54,55,56], whereas earlier reprecipitation developed by Kasai et al. provide a single component of organic crystals whose size, however, is manageable among nanometers to micrometers [57,58,59].
(Oxy) nitrides are new kinds of materials recently employed as photoelectrode materials for water splitting systems. Oxynitride based photocatalyst tends to possess narrower bandgaps than the equivalent metal oxides since the M-N bond in oxynitrides has potential energy greater than that of the M-O bond in metal oxides. In particular, it must be noted that these kinds of materials vary from photocatalysts doped with N or sulfur [60]. Moreover, (Oxy)nitrides are well-known for their solidity against anodic dissolution in alkaline media, providing them appropriate materials for water splitting applications.61 Comparatively, certain (Oxy)nitrides photocatalysts can replace metal-oxide-based materials to absorb visible-light photons for water photolysis reactions [61,62,63]. Similarly, an organic solid/water interface in a p/n organic bilayer was established to be proficient of photoinduced redox reactions combined with photoconduction of the generated hole/electron pairs in the interior [50,64,65], whereas wide-range visible-light energy (λ < 750 nm) is accessible for the photogenerated redox reactions at the interface mentioned above.
It is projected that long-lasting struggles in this area will result in more fascinating approaches towards emerging durable and high-efficiency heterogeneous photocatalysts based electrodes for water oxidation shortly. This focused review illustrates recent findings in promoting and advancing heterogeneous photocatalyst materials for light-driven water splitting reactions. In this review, heterogeneous photocatalyst materials, such as organophotocatalysts, (Oxy)nitrides, oxysulfides, and other heterogeneous photocatalysts, are presented for solar-driven photocatalytic water splitting, concentrating on recent developments on heterogeneous photocatalyst materials. An outlook for the future growth of heterogeneous photocatalysts-based photoanodes is also projected.

2. (Oxy)nitrides Based Materials

Comparatively, metal nitride/oxynitride based photocatalysts are subjected to photodissociation and have not been employed for water photolysis reaction [66]. In particular, oxynitrides are exceptionally favorable photocatalysts due to their narrower bandgap and subsequent superior visible-light photons related to oxide-based materials. For instance, TaON, Ta3N5, and Ba(Ca, Sr)TaO2N are the distinct classes of (Oxy)nitrides-based families. TaON, whose bandgap is varied from 1.9 to 2.5 eV and energy levels overlap with water redox potentials, is accepted as a more suitable material for developing highly active photoanodes.
The energy level positons of both TaON and Ta3N5 are more appropriate for water oxidation and reduction in visible-light conditions. Balaz and co-workers reported the CB and VB’s absolute positions of various kinds of characteristic TaON perovskites ATaO2N (A = Ca, Sr, and Ba) and PrTaON2 [67]. Moreover, TaON and Ta3N5 have been widely examined as solar-driven photoelectrode materials in water redox processes [68]. Particularly, above 10% quantum yield has been accomplished with TaON and Ta3N5 photocatalysts [69,70,71]. Domen and co-workers mostly investigated oxynitride based photoelectrodes as an alternative to oxide material, which is likely to arrest the incoming photons for water oxidation reactions [72,73,74]. In this regard, creating highly active oxynitride-based photoanodes is significant in achieving a high-efficiency PEC water photoelectrolysis process.
Electronic structures of oxynitrides were theoretically investigated in terms of the framework of density functional theory [75,76]. These results illustrate that the N2p orbital has greater potential energy than the O2p orbital; it might be remarkable to engage a metal (Oxy)nitride as a photocatalytic material. In this context, (Oxy)nitrides display distinctive optical features, owing to O2/N3 ions’ interplay in the system [75,76]. Comparatively, (Oxy)nitrides of the metal system cause a red-shift in the material than oxide-based materials. This might be owing to incorporating nitrogen (N) into the lattices, whereas the extra energy levels in the host material band considerably affect the optical transition in the system [77]. Particularly, this originated owing to the lesser electronegativity of N to that of O. For instance, Figure 3 compares the schematic representation of the oxide-based NaTaO3 and (Oxy)nitride BaTaO2N, and both of them possess a similar perovskite structure [78].
Generally, the most vital parameter affecting photocatalysts’ visible-light-driven water oxidation efficiency is the amount of living photoinduced holes directed by the morphology and size of a photocatalyst [79]. In general, the photocatalytic performance can be promoted by means of photoelectrodes with huge surface area, higher dispersion, high crystallinity, reduced defect density, well-defined morphologies, and precisely exposed facets, permitting to decrease the amount of recombination sites and to highly photochemically active sites. In particular, the morphological and structural modification of oxynitrides considerably affects the photocatalytic activity for water oxidation. Also, the flux-assisted nitridation method is beneficial for enhancing the crystallinity, modifying the morphological as well as surface modification of oxynitrides [80,81]. It is also evidenced that the cocatalyst–photoelectrode interface and co-catalysts dispersion over electrodes play a significant part in oxynitride photocatalysis. For instance, Maegli et al. reported the controlled morphological and structural changes of LaTiO2N/CoOx photoelectrodes for photocatalytic O2 oxidation [82]. They also demonstrated that the flux method improved the crystallinity and higher nitrogen contents in LaTiO2N than if prepared without flux. In particular, the skeletal and edgy morphology has higher CoOx catalyst dispersion on the LaTiO2N electrode contained unsaturated bonds, thereby improving the photocatalytic O2 evolution. Moreover, Kodera et al. demonstrated that flux treatment on BaNbO2N photoelectrodes greatly influenced the morphological natures [83]. Prior to flux action, the BaNbO2N morphological features were uncertain, with an accumulation of particle sizes varied from a few 10’ s nm to 100′ s nm. After the flux treatment, a noticeable modification in morphology happened. Figure 4 displays FE-SEM images of BaNbO2N photocatalysts before and after flux action. After flux treatment, a cubic morphology was clearly noticed. Similar flux-influenced morphological modifications are also demonstrated for BaTaO2N crystal structures [84]. Thus, the flux-aided nitridation process is mandatory to promote oxynitride’s photocatalytic features by enhancing crystallinity, tailoring the morphology and surface modifications, reducing the defect density.
Different (Oxy)nitrides materials have been considered to replace the oxide catalysts to capture the light-photons to produce H2 and O2 from water at the stoichiometric ratio [85,86,87,88,89,90]. Afterwards, these kinds of (Oxy)nitride-based photoelectron materials are well-known for PEC water oxidation, with nominal input [91,92]. A large quantity of such (Oxy)nitride photoanodes have been reported to date. Moreover, few of them attained superior incident photon-to-current efficiency (IPCE) efficiency of >10% with an applied bias lesser than the thermodynamically needed bias (1.23 V) for water photoelectrolysis. Further, in the case of wavelength, light photons with wavelengths up to 600 nm are accessible for PEC water oxidation with Ta3N5 [93] and LaTiO2N [94]. In this regard, no narrower photoanodes (<2.0 eV~equivalent to 600 nm) are available to oxidize water without any externally applied bias has been demonstrated till now. Furthermore, while decreasing the photoanodes’ bandgap, reducing the driving powers for water redox reactions makes water electrolysis reactions more challenging.
Both the superior performances and stability of (oxy)nitrides for PEC water electrolysis rely on the effective transfer of the photoinduced holes from the photoelectrodes to reactants adsorbed on the surface to induce the water oxidation for OER. Building a catalytic active surface morphology for PEC water electrolysis is extremely required to advance surface charge transfer. In particular, the catalytic surface must provide the roles of gathering photoinduced holes, permitting ample adsorption of water molecules and making it a lower reaction kinetic barrier for the oxidation reaction. Despite its ability to capture solar energy and chemical inertness, the PEC behavior of (oxy)nitrides is bizarrely restricted due to its deprived photon absorption, the greater recombination rate of photogenerated carriers sluggish OER kinetics. To rectify these drawbacks, (oxy)nitrides photoanodes must be loaded with co-catalyst materials to promote light photon absorption. During the last decades, numerous works have been carried out to advance (oxy) nitride-based photoanodes with appropriate co-catalysts to efficiently utilize visible light photons [63,92]. Moreover, much attention has been paid to develop (oxy)nitride-based photoanodes loaded with cost-effective transition metal oxides. Nishimura al. demonstrated the TaON based (oxy)nitrides along with co-catalyst photoanodes and attained IPCE of 76% at 400 nm with a nominal external bias [92]. Subsequently, different (oxy)nitrides, such as LaTaO2N [94], SrNbO2N [95,96,97], and BaNbO2N [98], among others, have been newly advanced as photoanodes [99,100,101,102,103].
Photoelectrodes with catalytic surface modification of several co-catalysts for water electrolysis, namely IrO2, CoOx and “Co-Pi” have been deeply examined [102,103,104,105,106,107,108,109]. In particular, Abe and co-workers reported incorporating IrO2.nH2O nanoparticles over porous TaON photoanodes and have achieved the IPCE of 76% in an aq.Na2SO4 solution [109]. Further, it is the first report to demonstrate the PEC water splitting on TaON (oxy)nitride with a scannable production of O2, as well as measured anodic photocurrent. Notably, Higashi et al. reported a methodology of pre-depositing earth plentiful CoOx co-catalysts over TaON particles via impregnation before the electrophoretic deposition (EPD) of TaON particles [106]. Recent years, (Oxy)nitrides photoelectrode materials, namely LaTiO2N/CoOx, BaTaO2N/BaZrO3, and BaNbO2N [110,111,112,113], have been developed to exploit the visible-light photons absorption with the supporting sacrificial reagents. Much research has mainly focused on creating the visible-light active narrower bandgap semiconductors photoanodes (<2 eV). In this regard, Zhang et al. reported LaTaO2N photocatalysts materials prepared through the flux technique. It also revealed enriched PEC behavior for the water oxidation reaction [114]. Quite recently, new kinds of rare earth hafnium (RHfO2N) (Oxy)nitride were demonstrated for the photocatalytic water redox reactions [77,115]. The RHfO2N perovskites (R = La, Nd and Sm) signified the GdFeO3-type structure with a bandgap energy of 3.35, 3.40, and 2.85 eV. These hafnium-based materials photocatalytic results revealed RHfO2N oxynitrides that have been successfully applied towards overall water splitting reactions. Also, Takata et al. have demonstrated and activated LaMgxTa1−xO1+3xN2−3x photocatalysts for photocatalytic water splitting by incorporating the co-catalyst of RhCrOx [116].
Niobium-based oxynitrides are of specific interest in water splitting owing to their narrow bandgap of 1.7 eV, while it was evidenced to be satisfactory for photocatalytic water photodissociation [117]. Maeda et al. first reported the narrower bandgap semiconductor of FTO/SrNbO2N photoanodes (<2.0 eV) substrate via electrophoretic deposition (EPD) process and examined for water splitting reactions under neutral conditions.118 IrO2 incorporated SrNbO2N/FTO electrodes chronoamperometric (CA) investigation showed significantly improved PEC water oxidation behavior compared to SrNbO2N/FTO as presented in Figure 5A. During the CA analysis, an external bias of 1.55 VRHE was engaged under steady-state light irradiance (λ > 420 nm). The enhanced performance was accredited to the substantial role of adsorbed IrO2 colloids over SrNbO2N particles to suppress the material’s oxidative self-decomposition and thereby enhance the selectivity of photoinduced holes toward water oxidation. This is an efficient approach to improving the PEC performances in the lower potential region, thereby promoting water oxidation. The authors further evidenced the water oxidation by evaluating the stoichiometric generation of H2 and O2 upon irradiation and was shown in Figure 5B. Moreover, in terms of evolved H2 and O2, the authors estimated that 47% of the visible-light current might be credited to photogenerated water oxidation by IrO2/SrNbO2N.
In recent years, we also demonstrated the PEC activity of cobalt phosphate (CoPi)/La(Ta,Nb)O2N and CoPi/ZnTaO2N electrodes for PEC water oxidation in alkaline conditions [99,119], which showed up to a nearly four-fold advancement in the PEC activities at a lower oxidation potential. In particular, the loading of the CoPi over (Oxy)nitride photoanodes might promote the charge separation as well as the carrier collection created at the electrode surface, thereby improving PEC features of the photoanodes). We have also recently reported the CoPi loaded ZnLaTaON2 photoanodes for PEC water oxidation performances under alkaline solution [120]. In all these cases, the photodeposition (PD) of CoPi onto (oxy)nitrides photoanodes under irradiation conditions were evidenced to promote the OER at the photocatalysts [121]. Further, the CoPi co-catalyst layer was incorporated over (oxy)nitrides by employing the in-situ PD approach to promote the OER half-reaction kinetics. We have also evidence that the CoPi deposition’s magnitude was determined by the electrodes’ morphology, time, and loading process method. The thicker CoPi incorporation over (Oxy)nitrides photoanodes resulted in the poorer PEC behavior for water oxidation. These facts can be illustrated that the photocatalytic water oxidation reaction happens at the CoPi/electrolyte interface. In the thicker CoPi, the photoinduced holes need to transfer between more number CoPi molecules and a CoPi/electrolyte interface. Subsequently, a decline in photocurrent is witnessed. However, in thin CoPi layer over (oxy)nitrides, associated cobalt ions on (Oxy)nitrides surface easily capture the photoinduced holes to yield the active catalytic species of cobalt (Co4+) mandatory for a water electrolysis reaction [122,123]. Likewise, amorphous Co(OH)x has been used as a co-catalytic material for water oxidation reaction and can enhance the PEC stability as well as performances of Ta3N5 photoanodes [122,123,124,125], even though it is not as efficient as CoPi. One of the major roles for superior performance of CoPi co-catalyst in comparison with Co(OH)x is the whole coverage of Co-Pi on the Ta3N5 photoanodes. Generally, in Co(OH)x is incorporated over Ta3N5 photoanodes surface via the co-precipitation process; thereby, it is hard to obtain the homogeneous dispersion by PD method. Thus, the PD methodologies benefit from permitting the comparatively homogeneous wrapping of CoPi over the (Oxy)nitride surface of the photoanodes. These results show that the incorporation of CoPi over (Oxy)nitride has been an efficient route to enhance the PEC current gain via the PEC water oxidation reaction.
Like CoPi electrocatalysts, layered double hydroxides (LDHs) have been widely investigated as catalytic materials for water oxidation [126,127]. Particularly, LDHs attained by a solution exfoliation process revealed greater OER performances and improved durability compared to noble metal/metal oxide electrocatalysts. Mostly, LDHs co-catalysts were engaged to promote the PEC performances for water electrolysis reaction [128]. In this regard, Wang et al. reported the Ni–Fe LDHs incorporated Ta3N5 photoanodes array and are significantly enhanced stability and PEC performances for water oxidation [129]. Further, the author’s group deposited the Co(OH)x and CoPi electrocatalysts on sequence over Ni–Fe LDH/Ta3N5 photoanodes, further enriched the photocurrent. Liao et al. associated the effect of various co-catalysts (Co(OH)x, Co3O4 and IrO2) and electrolytes (NaOH and Na2SO4) on the photocurrent and superior durability of photoanodes comprising of nanoparticulate Ta3N5 [130].
Allam and co-workers recently reported a novel photoanode material of Ti-Pd mixed (Oxy)nitride films prepared through anodization method and attained a photocurrent density of 1.9 mA/cm2 and a nearly five-fold enhancement in the efficiency compared to pure TiO2 nanotubes [131]. Similarly, titanium oxynitride (TiON) emerges as a favourable candidate for light absorption and appropriate energy level positions for water splitting [132,133]. In particular, Soliman et al. fabricated the TiON nanotube arrays via NH3 nitridation from anodized Ti foil. Very recently, Soliman et al. demonstrated the highly ordered Ag nanoparticles incorporated TiON nanotubes arrays for solar-driven water splitting [134]. Further, the author group investigated PEC properties of Ag/TiON nanotubes arrays resulting in an exceptionally superior photocurrent of 14 mA/cm2 at 1.0 V SCE (Figure 6b), which is consistent with the absorption spectra (Figure 6a). The observed improvement of photocurrent over TiON counterparts is credited to co-operative features of Ag nanoparticles incorporation, N-doping, and the distinctive morphological characteristics of the fabricated NTs arrays. In detail, the IPCE measurements of Ag/TiON nanotube arrays with the potential between 0.2 V and 0.4 V (Figure 6d) evidences the maximum IPCE of 41% in the wavelength ranging between 450 and 510 nm and is comparatively superior to TiON nanotubes arrays (Figure 6c).
Table 1 summarizes studies on different (Oxy)nitride-based electrode materials for solar light-driven water splitting reactions. The electrodes presented in Table 1 and the co-catalysts’ deposition significantly improved the PER performance with higher durability for solar-water splitting. In recent years, (Oxy)nitrides-based electrode materials have gained much consideration. They must be investigated in numerous ways, namely morphological features, co-doping strategies, defect induced features, surface activation, co-catalyst loading, etc. These examinations on the (Oxy)nitride photoelectrode materials basically lead to a more detailed understanding of their effective usage in the water-splitting system.

3. Oxysulfide Based Photocatalyst

Oxysulfides are new kinds of fascinating photocatalytic materials in heterogeneous based semiconductors photocatalysts. Similar to (Oxy)nitrides, the main idea of combining sulfur into oxides is basically the same as that for N incorporation. Particularly, the oxysulfides recognized as stable photocatalyst form than its corresponding metal oxides. Its energy level positions are appropriate for water photoelectrolysis reactions. In particular, oxysulfides comprise S as a constituent element that forms the tops of the VB. Thus, the sulfur loading usually reduces the bandgap compared to its counterparts of the corresponding oxide by developing shallower filled energy states than O 2p states. Thereby, photoinduced holes can travel smoothly in the VB of the material, predominantly beneficial for water oxidation concerning the 4-electron transfer. Till now, overall water splitting through an independent oxysulfide is yet to be developed. In contrast, few oxysulfides based photocatalysts are applied to PEC water electrolysis reaction with bias assistance, tandem style, or Z-scheme, water splitting.
Ishikawa et al. reported the stable oxysulfide of Sm2Ti2S2O5 with a wide absorption band in the visible-light region (λ < 600 nm) and band edge potential are more appropriate for overall water splitting [146]. These oxysulfide photocatalysts can be attained by annealing the mixture of equivalent precursors under vacuum and subsequently processed for sulfurization through H2S gas [147]. Density function theory calculations further evidence that the S3p atomic orbitals constitute the upper region of the VB of Sm2Ti2S2O5 and provide a significant influence in lowering the bandgap energy to its counterpart of metal oxide (Sm2Ti2O7). Further, metal sulfides such as CdS are not appropriate to engage as a catalytic material for water oxidation due to the integral unstable nature of these electrode materials in the reaction through the photoinduced generation of holes [148,149]. In contrast, Sm2Ti2S2O5 performed as a relatively stable photoelectrode material for water electrolysis reaction. Similar to (Oxy)nitride, the photooxidation performance of the oxysulfides will be improved by incorporating the colloidal IrO2 over Sm2Ti2S2O5, thereby increasing the quantum efficiency for O2 generation to 1.1% [150,151]. Moreover, physical characterization evidence that the decomposition of oxysulfide in bulk and at the surface, in precise the oxidation of S2− species, does not happen throughout the photooxidation of water. Quite recently, Ma et al. reported the Sm2Ti2S2O5 via novel flux method and demonstrated its photocatalytic activity is promoted by incorporating IrO2 co-catalyst and Pt over oxysulfides. Further, Z-scheme visible-light assisted water splitting into H2, and O2 was evidenced by applying the oxysulfides photocatalyst for H2 production combined with an H-CS-WO3 photocatalyst (shown in Figure 7) [152]. These results conclude that the oxysulfide Sm2Ti2S2O5 behaves as a highly stable photoelectrode material for water redox reactions under visible-light illumination. Similarly, other kinds of Ln2Ti2S2O5 (Ln: Pr, Nd, Gd, Tb, Dy, Ho, and Er) oxysulfides with similar layered perovskite structure are recognized as active and stable electrode materials for water photoelectrolysis reactions with supporting electrolyte comprising a sacrificial electron donor or acceptor [153]. Likewise, Zhang et al. developed that Ln2Ti2S2O5 compounds (absorption edge at ca. 600 nm) evidenced photocatalytic behavior for water electrolysis reaction with appropriate support sacrificial reagents [154]. Moreover, this is the first Ln2Ti2S2O5 oxysulfide to behave as a photocatalyst for the water oxidation reaction. Similarly, Ogisu and co-workers developed the visible-light responsive La5In3S9O3 and La3GaS5O14 oxysulfide photocatalysts and investigated water electrolysis reaction with a proper hole or electron scavenger(s) under irradiation conditions [155,156,157].
These oxysulfides based electrodes reveal remarkable PEC behaviors. Usually, p-type semiconductors behave as photocathodes, and generate H2 at the boundary between the semiconductor and the electrolyte. In particular, most of the sulfides are naturally an n-type semiconductor, while Cu+ -containing sulfides tend to behave as p-type semiconductors. In this regard, Liu et al. reported the Cu+-containing oxysulfide La5Ti2CuS5O7 photocathode via the particle transfer method. In contrast, doped oxysulfides show an 8-fold enhancement compared to their counterparts for PEC water splitting reaction [157]. Likewise, Ma et al. reported the La5Ti2CuS5O7 oxysulfides enabling to act both as photocathodes and photoanodes by tuning its work functions and employing them as electrodes in the water-splitting reaction [158]. Similarly, Hisatomi et al. reported the new kinds of oxysulfides La5Ti2Cu1−xAgxS5O7 photocathodes, functioning at the operative potential for PEC H2 evolution under solar-light irradiance [159].

4. Other Heterogeneous Photocatalysts

As an alternative to oxides, Yoshimizu attempted new kinds of silicide based photocatalyst (Iron disilicide β-FeSi2) composed of abundant, durable, and non-toxic elements [160]. In particular, the author investigated the β-FeSi2 catalyst material for H2 evolution catalysts because it possesses appropriate conduction band energy levels and succeeded an obvious quantum efficiency of 24% upto 950 nm support of a sacrificial agent. Similarly, Akiyama et al. succeeded in fabricating the semiconducting β-FeSi2/Si composite through the vapor-liquid-solid method. They investigated the photocatalytic H2 behavior with the support of the formaldehyde as a sacrificial agent [161]. As an H2 evolution photocatalyst, β-FeSi2 is projected to be sensitive to near-infrared light (>1300 nm), which is the longest wavelength of light to be used. Recently, Akiyama et al. fabricated the β-FeSi2/SiC composite powder and evolved H2 under the visible light photons with methyl alcohol as a sacrificial reagent [162].
Similar to heterogeneous photocatalysts, metal-free photocatalysts are favorable photocatalytic materials that have not been widely investigated. On this phenomenon, Wang et al. demonstrated a new class of metal-free graphitic C3N4 (g-C3N4) based material and employed as photocatalysts for water splitting reactions under visible-light photons [163,164]. In this regard, few kinds of metal-free photocatalytic materials are presently available: g-C3N4, C3N3S3 [165], a-sulfur [166], red phosphorus [167], black phosphorus [168], and organic ones as below. However, all these metal-free photocatalysts require scarce/expensive noble metal co-catalysts to be active for PEC water splitting reactions [169]. Fortunately, boron carbide and hexagonal boron carbon nitride (h-BCN) are fascinating new kinds of proficient metal-free and light-active photocatalysts for water splitting reaction. Liu et al. developed the boron carbides as photocatalytic materials for H2 production. They achieved H2 production of 14.5 µmol h−1 g−1 without any co-catalysts [170].

5. Organic Photocatalyst (Organophotocatalyst)

Generally, organic materials have run after inorganic ones that exhibit high electromagnetic properties due to high electron density and various elements such as conductions, magnetism, optics, etc. Organophotocatalysis based system exhibits an innovative technique for water splitting reactions exploiting energy over the whole range of visible light.
Primary reports designated photocathodes made from conducting polymers, namely polyacetylene [171,172,173], polypyrrole [174], polyaniline [175,176], poly(3-methylthiophene) [177] or poly(3-hexylthiophene), P3HT [178,179], phthalocyanine [180,181,182,183,184]. Organic photocatalyst without electrode also has a long history using poly(p-phenylene) [185,186,187] etc. Conversely, these photocatalytic materials did not comprise any junction between OSCs and charge separation from the Schottky junction between p-type OSC and electrolyte and kinetically low efficiency.
The closest research field of organic photovoltaic (OPV) devices to organic photocatalyst becomes more than 14% power conversion efficiency from the sunlight [188], and organic photocatalysts would have promising high efficiency in the near future if the stability problem is solved. Besides general merits in organic materials such as fabrication process varieties, flexibility, low density, and solubility, there are great merits of huge variety to control electronic properties in light absorption with a pi-system, conducting polymers nanostructures, e.g., synthetic metals or phase separations of bulk heterojunction [189,190,191,192,193,194,195]. As shown, the first p-n junction OPV [196] exhibited unidirectional electron-hole transfer path at both interfaces and the potential difference is derived by the photo illumination as visible-light-induced charge creation is succeeded at the boundary between the two OSCs as shown in Figure 8. Upon absorption of visible-light photons in one of the OSCs, an exciton is created and then transported to the junction with another OSC. At this point, the exciton dissociates, resultant in the arrival of a hole in one OSC called the donor, and an electron in the other OSC called the acceptor. In photovoltaic systems, charge carriers then transport to the electrodes where charges have been collected. These concepts are useful to design an efficient PEC system. By choosing a bilayer system, photoinduced charge separation becomes efficient in the case of PEC also [52]. Furthermore, various modifications gave more efficiency, such as p-n interface [197] and liquid/electrolyte interface [198]. Low cost and water-based synthesis of p-n junction like nanoparticles are possible to mass productively [46,47,48]. By combining polymer substrate, macroscopic separation of redox site is possible [49], and multilayerization of the photocatalyst film [47] exhibit more efficient utilization of natural light. The utilization of longer wavelength reaching to near-infrared is also attractive to utilize the sunlight [48,199]. Very recently, lateral charge separation to the substrate direction was discovered for step shape electrode using Kelvin probe force microscopy [200]. It is noteworthy that phthalocyanine has another role as a catalyst. One of the evidential phenomena is P3HT/PCBM photoanode [201,202], where coating with co-catalyst phthalocyanine enhanced photoanodic current. A recent observation by Kelvin force probe microscopy reveals a dipole formation between the co-catalyst and bulk heterojunction (BHJ), suggesting efficient charge separation for multi-electron transfer [202].
The main concern is the arrangement of the two different things since excitons can only diffuse in the range of few tens of nm for the duration of their lifetime, i.e., beforehand spontaneous recombination happened. As a result, the two OSCs must be either loaded as two very thin layers, probably with an interlayer comprising both constituents, resulting in a mixture of the two materials creating molecular p-n junctions all over the bulk layer BHJ. More importantly, BHJ is made between a polymer or an array of small molecules, performing as the light absorber, and a C60 derivative [195] typically. In order to employ such organic junctions in direct PEC water splitting, there are similarities in the continuous processes as above, while dissimilar conditions must be seen at the interface between the organic layer with a liquid electrolyte.
OER has a more crucial issue that strongly relates to stability and efficiency problems due to multielectron oxidation and co-existing thermodynamically easy path of photoanode etching. However, OER in an aqueous medium was validated with organic polymer catalysts/electrolyte solution junctions, obtaining photocurrent densities in the range of a few microamperes [203,204,205], or more bias potential [206]. Saline oxidation is one of the smart systems as alternative oxidation of chloride ions [207,208]. Coating of the metal oxide layer is another choice for long term stability [209,210,211,212,213,214].
Pt nanoparticles are good HER catalyst and were incorporated over the uppermost of an organic p/n bilayer electrode of small moieties (metal-free phthalocyanine, H2Pc/C60), photocurrent values grasped several cents of µA.cm−2 and the faradic efficiency for H2 evolution in aqueous media reached nearly 90% (Figure 9) [45]. H2 generation at Pt-attached C60/water interface utilizes electrons from the CB of the C60 layer. While the lowest of the CB of C60 is less negative than H2 evolution, the production of 2-electron reduced C602− species was demonstrated spectroscopically after a 20-min induction period.
The organic electrode presented in Figure 10 comprises photoelectrodes based on OSC to oxidize thiol [215], based on TiO2 to oxidize water [216], hydrazine oxidation [217] or borane-ammonia [218]. By employing photoelectrodes, H2 can be produced without additional electric bias. In this case, concomitant H2 evolution takes place at the Pt electrode without any input. Our research group recently investigated the photocatalysis system comprising an organic p/n bilayer for the stoichiometric degradation of hydrazine into N2 and H2 [218]. In this regard, these works illustrate the H2 generation by reducing power photoinduced at the PTCBI in an organic p/n bilayer.
Our research group recently reported the PEC system containing an organic-based photocathode of a p/n bilayer (ITO/ZnPc/C60–Pt) alongside a WO3 photoanode, where a stoichiometric generation of H2 and O2 was initiate to happen (Figure 11) [220]. In this system, oxidation and reduction powers can be markedly produced at the tungsten oxide photoanode and organo-photocathode, correspondingly. In this context, electron transport from tungsten oxide CB to the hole continued p-type VB in the organic p/n bilayer can proficiently happen for executing the PEC method, accordingly effecting a high concentration of carriers, which is readily existent for rate-limiting O2 evolution at photoanodes for effective charge separation. Finally, combining the two different water splitting materials is an efficient approach to developing a Z-scheme photocatalytic system.
The recent trend to use OSM is to be covered with a thin inorganic layer, for example, amorphous molybdenum sulfide (a-MoSx) as HER catalyst as an alternative of Pt [221,222]. Photocathodes fabricated via spray deposition of a combination of a-MoSx and TiO2 on the BHJ layer obtained a photocurrent density of 200 µA·cm−2 at 0 VRHE and up to 300 µA·cm−2 at –0.4 VRHE. Another inorganic layer is TiOx layer on P3HT:PCBM BHJ layer and the catalytic film [223]. Moreover, metallic Al/Ti interfacial layers resulting in an enhancement in the photocurrent nearly about the 8 mA·cm−2 at 0 V vs. RHE. A 50-nm-thick C60 layer also functions as an interfacial layer, with a current density attaining 1 mA·cm−2 at 0 V vs. RHE [224]. Also, reduced graphene oxide, WO3 and NiO, CuI, polyaniline underlayers are effective for persistent PEC features in acidic aqueous electrolytes [225,226,227,228,229,230]. The use of α-sexithiophene (α-6T) and chloroboron subphthalocyanine (SubPc) combined with C60 as the electron extracting layer gave more H2 evolution [231]. MoS3 coating gave 3 mW/cm2 at 0 V vs. RHE [225]. Tandem cells are known to obtain high Open circuit voltage (Voc) in a solar cell. For water splitting, such high Voc is necessary, at least 1.23 V theoretically. Applications by the use of high Voc with the tandem structure are reported [231,232,233,234,235,236,237,238]. There are several review papers about using advanced organic and carbon-based materials [239,240,241,242,243,244,245,246,247,248], while the present paper focused only on the heterogeneous organic semiconductor photocatalysts.
Among these heterogeneous photocatalysts, the conjugated microporous polymers (CMPs) have been recognized as photocathodes for H2 evolution from water to support a sacrificial electron donor. Sprick et al. also investigated various kinds of CMPs by varying monomer linker length and were investigated for photocatalytic H2 evolution reactions [248]. Like CMPs, covalent organic frameworks (COFs) are also an emergent class of crystalline polymers comprised of organic units connected via covalent bonds to produce porous networks. Very recently, Sick et al. reported first-time COFs as a new kinds of porous electrode materials of BDT-ETTA (benzo[1,2-b:4,5-b′]-dithiophene-2,6-dicarboxaldehyde (BDT); 1,1′,2,2′-tetra-p-aminophenylethylene, ETTA) for water-splitting reaction [249]. Further, these materials satisfy the necessities for light-driven water reduction application, which comprise effective light-harvesting, appropriate energy level positions, and stability.

6. Conclusions and Future Outlook

Energy harvesting technology from the limitless solar irradiation source has been recognized as a sustainable solution for the global energy crisis. In this review, the visible-light-driven water-splitting system has significantly improved since H2 is a water-splitting reaction expected to support future societies. In this context, a series of heterogenous photocatalyst materials have been examined for PEC water splitting reaction. Special consideration has been paid to organophotocatalysts based electrodes for the applications of solar-driven water splitting reactions. (Oxy)nitrides, oxysulfides, and organophotocatalysts have the potential to acts as stable photoelectrodes for water redox reaction under visible-light illumination, have been shown to PEC behavior without noticeable degradation. In some cases, heterogenous semiconductor photocatalyst systems need to feature organophotocatalysts along with a simple conventional oxide and were evidenced to be effective for the stoichiometric decomposition of water into H2 and O2. Over the past few decades, the available photocatalysts for visible-light-driven water splitting has mostly comprised of metal oxides. The development of heterogenous photocatalysts attempting to attain the same role has probably given a boost to research on photocatalytic material for solar energy conversion.
In summary, over the past decades, (oxy)nitride and oxysulfides based photoelectrodes have been faced a high-speed advancement in terms of both the stable electrochemical nature as well as the activity of PEC water oxidation. Understanding the mechanism assembled by investigating this kind of photoanodes will cover a solid path for advancing other types of proficient photoanodes. In all these systems, the photoelectrode materials, which cannot merely contribute to the complete decomposition of water, can play active parts. The usage of these novel kinds of heterogeneous photocatalyst based materials and suitable electrocatalyst materials have more advantages in gathering solar energy. PEC water decomposition is already recognized as the most promising methodologies for obtaining solar hydrogen. In this perspective, developing new kinds of heterogeneous photocatalyst will be an efficient strategy for creating a commercially viable water-splitting system.
During the past few years, considerable research efforts have applied photoelectrochemistry for other reactions besides water splitting. In this context, the conversion of CO2 into value-added organic molecules has gained enormous interest in recent years. Carbon dioxide reduction reaction (CO2RR) is far more complicated than water splitting. Still, it permits the opportunity of extracting the greenhouse gas CO2 from the atmosphere and turning it into value-added fuels. Commercially, products such as methane or ethanol generation via CO2RR have been explored [250,251]. Conversely, the major limits for CO2RR are finding suitable photoelectrodes and robust semiconductors to be kept in direct contact with the reaction, as the higher required reductive potentials (>1 V) improve materials corrosion [252,253]. Despite the described potentialities for the usage of photoelectrochemistry in H2 production and CO2RR, it has not been possible to demonstrate an application capable of competing economically with conventional procedures that are frequently used e.g., methane reforming permits obtaining H2 at € 3/kg. In this regard, to be capable of developing alternate methods built on the usage of renewable energies such as photoelectrocatalysis, it is essential to promote efficient and robust devices created using cost-effective most-abundant electrode materials (like heterogeneous photocatalyst), demanding optimization approaches like the tandem arrangement of photoelectrodes to relax the stringent conditions required for effective PEC value-added solar fuel production. Indeed, PEC water splitting development will help to identify and adapt photoelectrodes for other electrochemical reactions.

Author Contributions

P.A., M.A.G., A.M.A.-M., and K.N. Initiated and planned the subject and conceived the subject; P.A. perceived the subject, discussed related articles, wrote this manuscript collaboratively, and led this work. M.S.A. and R.J.R. assisted in writing. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Deputyship for Research and Innovation, “Ministry of Education” in Saudi Arabia for funding this research work through the Project number IFKSURP-025, Kingdom of Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research and Innovation, “Ministry of Education” in Saudi Arabia for funding this research work through the Project number IFKSURP-025, Kingdom of Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lewis, N.S.; Crabtree, G. Basic Research Needs for Solar Energy Utilization: Report of the Basic Energy Sciences Workshop on Solar Energy Utilization, April 18–21, 2005; US Department of Energy, Office of Basic Energy Science: Washington, DC, USA, 2005.
  2. Walter, M.G.; Warren, E.L.; McKone, J.R.; Boettcher, S.W.; Mi, Q.; Santori, E.A.; Lewis, N.S. Solar water splitting cells. Chem. Rev. 2010, 110, 6446–6473. [Google Scholar] [CrossRef] [PubMed]
  3. Herron, J.A.; Kim, J.; Upadhye, A.A.; Huber, G.W.; Maravelias, C.T. A general framework for the assessment of solar fuel technologies. Energy Environ. Sci. 2015, 8, 126–157. [Google Scholar] [CrossRef]
  4. Lewis, N.S. Toward cost-effective solar energy use. Science 2007, 315, 798–801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Balzani, V.; Pacchioni, G.; Prato, M.; Zecchina, A. Solar-driven chemistry: Towards new catalytic solutions for a sustainable world. Rendiconti Lincei. Sci. Fisiche E Nat. 2019, 30, 443–452. [Google Scholar]
  6. McKone, J.R.; Lewis, N.S.; Gray, H.B. Will solar-driven water-splitting devices see the light of day? Chem. Mater. 2014, 26, 407–414. [Google Scholar] [CrossRef]
  7. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  8. Xing, J.; Fang, W.Q.; Zhao, H.J.; Yang, H.G. Inorganic photocatalysts for overall water splitting. Chem. Asian J. 2012, 7, 642–657. [Google Scholar] [CrossRef]
  9. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef]
  10. Lianos, P. Review of recent trends in photoelectrocatalytic conversion of solar energy to electricity and hydrogen. Appl. Catal. B 2017, 210, 235–254. [Google Scholar] [CrossRef]
  11. Acar, C.; Dincer, I.; Zamfirescu, C. A review on selected heterogeneous photocatalysts for hydrogen production. Int. J. Energy Res. 2014, 38, 1903–1920. [Google Scholar] [CrossRef]
  12. Arunachalam, P.; Al Mayouf, A.M. Photoelectrochemical Water Splitting. In Noble Metal-Metal Oxide Hybrid Nanoparticles; Woodhead Publishing: Cambridge, UK, 2019; pp. 585–606. [Google Scholar]
  13. Jafari, T.; Moharreri, E.; Amin, A.S.; Miao, R.; Song, W.; Suib, S.L. Photocatalytic water splitting—The untamed dream: A review of recent advances. Molecules 2016, 21, 900. [Google Scholar] [CrossRef] [PubMed]
  14. Ru Ng, A.Y.; Boruah, B.; Chin, K.F.; Modak, J.M.; Soo, H.S. Photoelectrochemical Cells for Artificial Photosynthesis: Alternatives to Water Oxidation. ChemNanoMat 2020, 6, 185–203. [Google Scholar] [CrossRef]
  15. Deyab, N.M.; Salem, K.E.; Mokhtar, A.M.; Ramadan, M.; Steegstra, P.; Hubin, A.; Allam, N.K. Electrochemical Fabrication of Ternary Black Ti-Mo-Ni Oxide Nanotube Arrays for Enhanced Photoelectrochemical Water Oxidation. ChemistrySelect 2020, 5, 12151–12158. [Google Scholar] [CrossRef]
  16. Tang, J.; Durrant, J.R.; Klug, D.R. Mechanism of photocatalytic water splitting in TiO2. Reaction of water with photoholes, importance of charge carrier dynamics, and evidence for four-hole chemistry. J. Am. Chem. Soc. 2008, 130, 13885–13891. [Google Scholar] [CrossRef] [PubMed]
  17. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef]
  18. Takata, T.; Hitoki, G.; Kondo, J.; Hara, M.; Kobayashi, H.; Domen, K. Visible-light-driven photocatalytic behavior of tantalum-oxynitride and nitride. Res. Chem. Intermed. 2007, 33, 13–25. [Google Scholar] [CrossRef]
  19. Malathi, A.; Arunachalam, P.; Grace, A.N.; Madhavan, J.; Al-Mayouf, A.M. A robust visible-light driven BiFeWO6/BiOI nanohybrid with efficient photocatalytic and photoelectrochemical performance. Appl. Surf. Sci. 2017, 412, 85–95. [Google Scholar] [CrossRef]
  20. Arunachalam, P.; Amer, M.S.; Ghanem, M.A.; Al-Mayouf, A.M.; Zhao, D. Activation effect of silver nanoparticles on the photoelectrochemical performance of mesoporous TiO2 nanospheres photoanodes for water oxidation reaction. Int. J. Hydrog. Energy 2017, 42, 11346–11355. [Google Scholar] [CrossRef]
  21. Amer, M.S.; Arunachalam, P.; Al-Mayouf, A.M.; Prasad, S.; Alshalwi, M.N.; Ghanem, M.A. Mesoporous tungsten trioxide photoanodes modified with nitrogen-doped carbon quantum dots for enhanced oxygen evolution photo-reaction. Nanomaterials 2019, 9, 1502. [Google Scholar] [CrossRef] [Green Version]
  22. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338. [Google Scholar] [CrossRef]
  23. Watanabe, T.; Fujishima, A.; Honda, K.-I. Photoelectrochemical reactions at SrTiO3 single crystal electrode. Bull. Chem. Soc. Jpn. 1976, 49, 355–358. [Google Scholar] [CrossRef] [Green Version]
  24. Ghanem, M.A.; Arunachalam, P.; Amer, M.S.; Al-Mayouf, A.M. Mesoporous titanium dioxide photoanodes decorated with gold nanoparticles for boosting the photoelectrochemical alkali water oxidation. Mater. Chem. Phys. 2019, 213, 56–66. [Google Scholar] [CrossRef]
  25. Melián, E.P.; López, C.R.; Méndez, A.O.; Díaz, O.G.; Suárez, M.N.; Rodríguez, J.D.; Navío, J.; Hevia, D.F. Hydrogen production using Pt-loaded TiO2 photocatalysts. Int. J. Hydrog. Energy 2013, 38, 11737–11748. [Google Scholar] [CrossRef]
  26. Yuan, Y.; Zhang, X.; Liu, L.; Jiang, X.; Lv, J.; Li, Z.; Zou, Z. Synthesis and photocatalytic characterization of a new photocatalyst BaZrO3. Int. J. Hydrog. Energy 2008, 33, 5941–5946. [Google Scholar] [CrossRef]
  27. Amer, M.S.; Ghanem, M.A.; Al-Mayouf, A.M.; Arunachalam, P.; Khdary, N.H. Low-loading of oxidized platinum nanoparticles into mesoporous titanium dioxide for effective and durable hydrogen evolution in acidic media. Arab. J. Chem. 2020, 13, 2257–2570. [Google Scholar] [CrossRef]
  28. Amer, M.S.; Ghanem, M.A.; Al-Mayouf, A.M.; Arunachalam, P. Low-Symmetry Mesoporous Titanium Dioxide (lsm-TiO2) Electrocatalyst for Efficient and Durable Oxygen Evolution in Aqueous Alkali. J. Electrochem. Soc. 2018, 165, H300–H309. [Google Scholar] [CrossRef]
  29. Hara, Y.; Takashima, T.; Kobayashi, R.; Abeyrathna, S.; Ohtani, B.; Irie, H. Silver-inserted heterojunction photocatalyst consisting of zinc rhodium oxide and silver antimony oxide for overall pure-water splitting under visible light. Appl. Catal. B Environ. 2017, 209, 663–668. [Google Scholar] [CrossRef]
  30. Amer, M.S.; Ghanem, M.A.; Arunachalam, P.; Al-Mayouf, A.M.; Hadadi, S.M. Bifunctional Electrocatalyst of Low-Symmetry Mesoporous Titanium Dioxide Modified with Cobalt Oxide for Oxygen Evolution and Reduction Reactions. Catalysts 2019, 9, 836. [Google Scholar] [CrossRef] [Green Version]
  31. Hara, S.; Yoshimizu, M.; Tanigawa, S.; Ni, L.; Ohtani, B.; Irie, H. Hydrogen and oxygen evolution photocatalysts synthesized from strontium titanate by controlled doping and their performance in two-step overall water splitting under visible light. J. Phys. Chem. C 2011, 116, 17458–17463. [Google Scholar] [CrossRef] [Green Version]
  32. Manikandan, M.; Tanabe, T.; Li, P.; Ueda, S.; Ramesh, G.V.; Kodiyath, R.; Ariga, K. Photocatalytic water splitting under visible light by mixed-valence Sn3O4. ACS Appl. Mater. Interfaces 2014, 6, 3790–3793. [Google Scholar] [CrossRef]
  33. Kamimura, S.; Higashi, M.; Abe, R.; Ohno, T. Fabrication of a porous ZnRh2O4 photocathode for photoelectrochemical water splitting under visible light irradiation and a significant effect of surface modification by ZnO necking treatment. J. Mater. Chem. A 2016, 4, 6116–6123. [Google Scholar] [CrossRef] [Green Version]
  34. Iwashina, K.; Iwase, A.; Nozawa, S.; Adachi, S.I.; Kudo, A. Visible-Light-Responsive CuLi1/3Ti2/3O2 Powders Prepared by a Molten CuCl Treatment of Li2TiO3 for Photocatalytic H2 Evolution and Z-Schematic Water Splitting. Chem. Mater. 2016, 28, 4677–4685. [Google Scholar] [CrossRef]
  35. Tateno, H.; Miseki, Y.; Sayama, K. Photoelectrochemical dimethoxylation of furan via a bromide redox mediator using a BiVO4/WO3 photoanode. Chem. Commun. 2017, 53, 4378–4381. [Google Scholar] [CrossRef] [PubMed]
  36. Amano, F.; Li, D.; Ohtani, B. Fabrication and photoelectrochemical property of tungsten (VI) oxide films with a flake-wall structure. Chem. Commun. 2010, 46, 2769–2771. [Google Scholar] [CrossRef] [Green Version]
  37. Hu, Y.-S.; Kleiman-Shwarsctein, A.; Stucky, G.D.; McFarland, E.W. Improved photoelectrochemical performance of Ti-doped α-Fe2O3 thin films by surface modification with fluoride. Chem. Commun. 2009, 2652–2654. [Google Scholar] [CrossRef]
  38. Sayama, K.; Nomura, A.; Zou, Z.; Abe, R.; Abe, Y.; Arakawa, H. Photoelectrochemical decomposition of water on nanocrystalline BiVO4 film electrodes under visible light. Chem. Commun. 2003, 2908–2909. [Google Scholar] [CrossRef]
  39. Shaddad, M.N.; Cardenas-Morcoso, D.; Arunachalam, P.; Garcia-Tecedor, M.; Ghanem, M.A.; Bisquert, J.; Al-Mayouf, A.; Gimenez, S. Enhancing the Optical Absorption and Interfacial Properties of BiVO4 with Ag3PO4 Nanoparticles for Efficient Water Splitting. J. Phys. Chem. C 2018, 122, 11608–11615. [Google Scholar] [CrossRef]
  40. Malathi, A.; Madhavan, J.; Ashokkumar, M.; Arunachalam, P. A review on BiVO4 photocatalyst: Activity enhancement methods for solar photocatalytic applications. Appl. Catal. A: Gen. 2018, 555, 47–74. [Google Scholar]
  41. Malathi, A.; Vasanthakumar, V.; Arunachalam, P.; Madhavan, J.; Ghanem, M.A. A low cost additive-free facile synthesis of BiFeWO6/BiVO4 nanocomposite with enhanced visible-light induced photocatalytic activity. J. Colloid Interface Sci. 2017, 506, 553–563. [Google Scholar] [CrossRef]
  42. Hou, T.F.; Johar, M.A.; Boppella, R.; Hassan, M.A.; Patil, S.J.; Ryu, S.W.; Lee, D.W. Vertically aligned one-dimensional ZnO/V2O5 core-shell hetero-nanostructure for photoelectrochemical water splitting. J. Energy Chem. 2020, 49, 262–274. [Google Scholar] [CrossRef]
  43. Ghosh, M.; Liu, J.; Chuang, S.S.; Jana, S.C. Fabrication of hierarchical V2O5 nanorods on TiO2 nanofibers and their enhanced photocatalytic activity under visible light. ChemCatChem 2018, 10, 3305–3318. [Google Scholar] [CrossRef]
  44. 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 Environ. 2018, 226, 373–383. [Google Scholar] [CrossRef]
  45. Li, F.; Zhao, W.; Leung, D.Y. Enhanced photoelectrocatalytic hydrogen production via Bi/BiVO4 photoanode under visible light irradiation. Appl. Catal. B Environ. 2019, 258, 117954. [Google Scholar] [CrossRef]
  46. Nagai, K.; Abe, T.; Kaneyasu, Y.; Yasuda, Y.; Kimishima, I.; Iyoda, T.; Imaya, H. A full-spectrum visible-light-responsive organophotocatalyst film for removal of trimethylamine. ChemSusChem 2011, 4, 727–730. [Google Scholar] [CrossRef]
  47. Nagai, K.; Yasuda, Y.; Iyoda, T.; Abe, T. Multilayerization of Organophotocatalyst Films that Efficiently Utilize Natural Sunlight in a One-Pass-Flow Water Purification System. ACS Sustain. Chem. Eng. 2013, 1, 1033–1039. [Google Scholar] [CrossRef]
  48. Mendori, D.; Hiroya, T.; Ueda, M.; Sanyoushi, M.; Nagai, K.; Abe, T. Novel photocatalytic material of organic p–n bilayer responsive to near-infrared energy. Appl. Catal. B Environ. 2017, 205, 514–518. [Google Scholar] [CrossRef]
  49. Abe, T.; Tobinai, S.; Taira, N.; Chiba, J.; Itoh, T.; Nagai, K. Molecular hydrogen evolution by organic p/n bilayer film of phthalocyanine/fullerene in the entire visible-light energy region. J. Phys. Chem. C 2011, 115, 7701–7705. [Google Scholar] [CrossRef]
  50. Abe, T.; Nagai, K.; Sekimoto, K.; Tajiri, A.; Norimatsu, T. Novel characteristics at a fullerene/water interface in an organic bilayer photoelectrode of phthalocyanine/fullerene. Electrochem. Commun. 2005, 7, 1129–1132. [Google Scholar] [CrossRef]
  51. Abe, T.; Hiyama, Y.; Fukui, K.; Sahashi, K.; Nagai, K. Efficient p-zinc phthalocyanine/n-fullerene organic bilayer electrode for molecular hydrogen evolution induced by the full visible-light energy. Int. J. Hydrog. Energy 2015, 40, 9165–9170. [Google Scholar] [CrossRef]
  52. Abe, T.; Nagai, K.; Kaneko, M.; Okubo, T.; Sekimoto, K.; Tajiri, A.; Norimatsu, T. A Novel and Efficient System of a Visible-Light-Responsive Organic Photoelectrocatalyst Working in a Water Phase. ChemPhysChem 2004, 5, 716–720. [Google Scholar] [CrossRef]
  53. Zhang, S.; Sakai, R.; Abe, T.; Iyoda, T.; Norimatsu, T.; Nagai, K. Photoelectrochemical and photocatalytic properties of biphasic organic p-and n-type semiconductor nanoparticles fabricated by a reprecipitation process. ACS Appl. Mater. Interfaces 2011, 3, 1902–1909. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, S.; Arunachalam, P.; Abe, T.; Iyoda, T.; Nagai, K. Photocatalytic decomposition of N-methyl-2-pyrrolidone, aldehydes, and thiol by biphase and p/n junction-like organic semiconductor composite nanoparticles responsive to nearly full spectrum of visible light. J. Photochem. Photobiol. A Chem. 2012, 244, 18–23. [Google Scholar] [CrossRef]
  55. Arunachalam, P.; Zhang, S.; Abe, T.; Komura, M.; Iyoda, T.; Nagai, K. Weak visible light (mW/cm2) organophotocatalysis for mineralization of amine, thiol and aldehyde by biphasic cobalt phthalocyanine/fullerene nanocomposites prepared by wet process. Appl. Catal. B Environ. 2016, 193, 240–247. [Google Scholar] [CrossRef]
  56. Kasai, H.; Nalwa, H.S.; Oikawa, H.; Okada, S.; Matsuda, H.; Minami, N.; Kakuta, A.; Ono, K.; Mukoh, A.; Nakanishi, H. A novel preparation method of organic microcrystals. Jpn. J. Appl. Phys. 1992, 31, L1132. [Google Scholar] [CrossRef]
  57. Tachikawa, T.; Chung, H.R.; Masuhara, A.; Kasai, H.; Oikawa, H.; Nakanishi, H.; Fujitsuka, M.; Majima, T. In situ and ex situ observations of the growth dynamics of single perylene nanocrystals in water. J. Am. Chem. Soc. 2006, 128, 15944–15945. [Google Scholar] [CrossRef]
  58. Baba, K.; Kasai, H.; Nishida, K.; Nakanishi, H. Poly (N-isopropylacrylamide)-based thermoresponsive behavior of fluorescent organic nanocrystals. Jpn. J. Appl. Phys. 2011, 50, 010202. [Google Scholar] [CrossRef]
  59. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  60. Hitoki, G.; Takata, T.; Kondo, J.; Hara, M.; Kobayashi, H.; Domen, K. (Oxy) nitrides as new photocatalysts for water splitting under visible light irradiation. Electrochemistry 2002, 70, 463–465. [Google Scholar] [CrossRef] [Green Version]
  61. Abe, R.; Takata, T.; Sugihara, H.; Domen, K. The use of TiCl4 treatment to enhance the photocurrent in a TaON photoelectrode under visible light irradiation. Chem. Lett. 2005, 34, 1162–1163. [Google Scholar] [CrossRef]
  62. Mucha, N.R.; Som, J.; Choi, J.; Shaji, S.; Gupta, R.K.; Meyer, H.M.; Kumar, D. High-Performance Titanium Oxynitride Thin Films for Electrocatalytic Water Oxidation. ACS Appl. Energy Mater. 2020, 3, 8366–8374. [Google Scholar] [CrossRef]
  63. Nishimura, N.; Raphael, B.; Maeda, K.; Le Gendre, L.; Abe, R.; Kubota, J.; Domen, K. Effect of TiCl4 treatment on the photoelectrochemical properties of LaTiO2N electrodes for water splitting under visible light. Thin Solid Film. 2010, 518, 5855–5859. [Google Scholar] [CrossRef]
  64. Abe, T.; Nagai, K.; Sekimoto, K.; Tajiri, A.; Norimatsu, T. Novel photocathodic characteristics of organic bilayer composed of a phthalocyanine and a perylene derivative in a water phase containing a redox molecule. J. Electroanal. Chem. 2005, 583, 327–332. [Google Scholar] [CrossRef]
  65. Abe, T.; Nagai, K.; Ichinohe, H.; Shibata, T.; Tajiri, A.; Norimatsu, T. An efficient oxidation at photofunctional interface of phthalocyanine in combination with fullerene. J. Electroanal. Chem. 2007, 599, 65–71. [Google Scholar] [CrossRef]
  66. Gerischer, H.; Luebke, M. Competition between Photocorrosion and Photooxidation of Redox Systems at n-Type Semiconductor Electrodes. Ber. Der Bunsenges. Für Phys. Chem. 1983, 87, 123–128. [Google Scholar] [CrossRef]
  67. Balaz, S.; Porter, S.H.; Woodward, P.M.; Brillson, L.J. Electronic structure of tantalum oxynitride perovskite photocatalysts. Chem. Mater. 2013, 25, 3337–3343. [Google Scholar] [CrossRef]
  68. Zhang, P.; Zhang, J.; Gong, J. Tantalum-based semiconductors for solar water splitting. Chem. Soc. Rev. 2014, 43, 4395–4422. [Google Scholar] [CrossRef]
  69. Hara, M.; Hitoki, G.; Takata, T.; Kondo, J.N.; Kobayashi, H.; Domen, K. TaON and Ta3N5 as new visible light driven photocatalysts. Catal. Today 2003, 78, 555–560. [Google Scholar] [CrossRef]
  70. Hitoki, G.; Takata, T.; Kondo, J.N.; Hara, M.; Kobayashi, H.; Domen, K. An oxynitride, TaON, as an efficient water oxidation photocatalyst under visible light irradiation (λ ≤ 500 nm). Chem. Commun. 2002, 1698–1699. [Google Scholar] [CrossRef]
  71. Hitoki, G.; Ishikawa, A.; Takata, T.; Kondo, J.N.; Hara, M.; Domen, K. Ta3N5 as a novel visible light-driven photocatalyst (λ < 600 nm). Chem. Lett. 2002, 31, 736–737. [Google Scholar]
  72. Kaneko, M.; Fujii, M.; Hisatomi, T.; Yamashita, K.; Domen, K. Regression model for stabilization energies associated with anion ordering in perovskite-type oxynitrides. J. Energy Chem. 2019, 36, 7–14. [Google Scholar] [CrossRef] [Green Version]
  73. Kawashima, K.; Hojamberdiev, M.; Yubuta, K.; Domen, K.; Teshima, K. Synthesis and visible-light-induced sacrificial photocatalytic water oxidation of quinary oxynitride BaNb0. 5Ta0. 5O2N crystals. J. Energy Chem. 2018, 27, 1415–1421. [Google Scholar] [CrossRef] [Green Version]
  74. Takata, T.; Pan, C.; Domen, K. Recent progress in oxynitride photocatalysts for visible-light-driven water splitting. Sci. Technol. Adv. Mater. 2015, 16, 033506. [Google Scholar] [CrossRef] [PubMed]
  75. Fang, C.; Orhan, E.; De Wijs, G.; Hintzen, H.; De Groot, R.; Marchand, R.; Saillard, J.Y. The electronic structure of tantalum (oxy) nitrides TaON and Ta3N5. J. Mater. Chem. 2001, 11, 1248–1252. [Google Scholar] [CrossRef] [Green Version]
  76. Fang, C.; De Wijs, G.; Orhan, E.; De With, G.; De Groot, R.; Hintzen, H.; Marchand, R. Local structure and electronic properties of BaTaO2N with perovskite-type structure. J. Phys. Chem. Solids 2003, 64, 281–286. [Google Scholar] [CrossRef]
  77. Sakar, M.; Prakash, R.M.; Shinde, K.; Balakrishna, G.R. Revisiting the materials and mechanism of metal oxynitrides for photocatalysis. Int. J. Hydrog. Energy 2020, 45, 7691–7705. [Google Scholar] [CrossRef]
  78. Maeda, K.; Domen, K. New non-oxide photocatalysts designed for overall water splitting under visible light. J. Phys. Chem. C 2007, 111, 7851–7861. [Google Scholar] [CrossRef]
  79. Amano, F.; Ishinaga, E.; Yamakata, A. Effect of particle size on the photocatalytic activity of WO3 particles for water oxidation. J. Phys. Chem. C 2013, 117, 22584–22590. [Google Scholar] [CrossRef]
  80. Kawashima, K.; Hojamberdiev, M.; Wagata, H.; Yubuta, K.; Vequizo, J.J.M.; Yamakata, A.; Oishi, S.; Domen, K.; Teshima, K. NH3-assisted flux mediated direct growth of LaTiO2N crystallites for visible-light-induced water splitting. J. Phys. Chem. C 2015, 119, 15896–15904. [Google Scholar] [CrossRef]
  81. Abeysinghe, D.; Skrabalak, S.E. Toward shape-controlled metal oxynitride and nitride particles for solar energy applications. ACS Energy Lett. 2018, 3, 1331–1344. [Google Scholar] [CrossRef]
  82. Maegli, A.E.; Pokrant, S.; Hisatomi, T.; Trottmann, M.; Domen, K.; Weidenkaff, A. Enhancement of photocatalytic water oxidation by the morphological control of LaTiO2N and cobalt oxide catalysts. J. Phys. Chem. C 2014, 118, 16344–16351. [Google Scholar] [CrossRef]
  83. Kodera, M.; Katayama, M.; Hisatomi, T.; Minegishi, T.; Domen, K. Effects of flux treatment on morphology of single-crystalline BaNbO2N particles. CrystEngComm 2016, 18, 3186–3190. [Google Scholar] [CrossRef] [Green Version]
  84. Hojamberdiev, M.; Kawashima, K.; Hisatomi, T.; Katayama, M.; Hasegawa, M.; Domen, K.; Teshima, K. Distinguishing the effects of altered morphology and size on the visible light-induced water oxidation activity and photoelectrochemical performance of BaTaO2N crystal structures. Faraday Discuss. 2019, 215, 227–241. [Google Scholar] [CrossRef] [PubMed]
  85. Hirayama, N.; Nakata, H.; Wakayama, H.; Nishioka, S.; Kanazawa, T.; Kamata, R.; Ebato, Y.; Kosaku, K.; Kumagai, H.; Yamakata, A.; et al. Solar-Driven Photoelectrochemical Water Oxidation over an n-Type Lead–Titanium Oxyfluoride Anode. J. Am. Chem. Soc. 2019, 141, 17158–17165. [Google Scholar] [CrossRef] [PubMed]
  86. Lawley, C.; Nachtegaal, M.; Stahn, J.; Roddatis, V.; Döbeli, M.; Schmidt, T.J.; Lippert, T. Examining the surface evolution of LaTiOxNy an oxynitride solar water splitting photocatalyst. Nat. Commun. 2020, 11, 1728. [Google Scholar] [CrossRef] [Green Version]
  87. Wang, X.; Maeda, K.; Lee, Y.; Domen, K. Enhancement of photocatalytic activity of (Zn1 + xGe)(N2Ox) for visible-light-driven overall water splitting by calcination under nitrogen. Chem. Phys. Lett. 2008, 457, 134–136. [Google Scholar] [CrossRef]
  88. 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]
  89. Ishikawa, A.; Takata, T.; Kondo, J.N.; Hara, M.; Domen, K. Electrochemical behavior of thin Ta3N5 semiconductor film. J. Phys. Chem. B 2004, 108, 11049–11053. [Google Scholar] [CrossRef]
  90. Le Paven-Thivet, C.; Ishikawa, A.; Ziani, A.; Le Gendre, L.; Yoshida, M.; Kubota, J.; Tessier, F.; Domen, K. Photoelectrochemical properties of crystalline perovskite lanthanum titanium oxynitride films under visible light. J. Phys. Chem. C 2009, 113, 6156–6162. [Google Scholar] [CrossRef]
  91. Konta, R.; Ishii, T.; Kato, H.; Kudo, A. Photocatalytic activities of noble metal ion doped SrTiO3 under visible light irradiation. J. Phys. Chem. B 2004, 108, 8992–8995. [Google Scholar] [CrossRef]
  92. 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. J. Am. Chem. Soc. 2003, 125, 3082–3089. [Google Scholar] [CrossRef]
  93. Feng, X.; LaTempa, T.J.; Basham, J.I.; Mor, G.K.; Varghese, O.K.; Grimes, C.A. Ta3N5 nanotube arrays for visible light water photoelectrolysis. Nano Lett. 2010, 10, 948–952. [Google Scholar] [CrossRef] [PubMed]
  94. Si, W.; Pergolesi, D.; Haydous, F.; Fluri, A.; Wokaun, A.; Lippert, T. Investigating the behavior of various cocatalysts on LaTaON 2 photoanode for visible light water splitting. Phys. Chem. Chem. Phys. 2017, 19, 656–662. [Google Scholar] [CrossRef]
  95. Lin, G.; Wang, R.; Cao, T.; Yuan, L.; Xu, X. Zr modified SrNbO2N as active photocatalyst for water oxidation under visible light illumination. Inorg. Chem. Front. 2020, 7, 2629–2636. [Google Scholar] [CrossRef]
  96. Oka, D.; Hirose, Y.; Kaneko, M.; Nakao, S.; Fukumura, T.; Yamashita, K.; Hasegawa, T. Anion-Substitution-Induced Nonrigid Variation of Band Structure in SrNbO3–x N x (0 ≤ x ≤ 1) Epitaxial Thin Films. ACS Appl. Mater. Interfaces 2018, 10, 35008–35015. [Google Scholar] [CrossRef] [PubMed]
  97. Sun, X.; Liu, G.; Xu, X. Defect management and efficient photocatalytic water oxidation reaction over Mg modified SrNbO 2 N. J. Mater. Chem. A 2018, 6, 10947–10957. [Google Scholar] [CrossRef]
  98. Hisatomi, T.; Katayama, C.; Moriya, Y.; Minegishi, T.; Katayama, M.; Nishiyama, H.; Yamada, T.; Domen, K. Photocatalytic oxygen evolution using BaNbO2N modified with cobalt oxide under photoexcitation up to 740 nm. Energy Environ. Sci. 2013, 6, 3595–3599. [Google Scholar] [CrossRef] [Green Version]
  99. Arunachalam, P.; Shaddad, M.N.; Ghanem, M.A.; Al-Mayouf, A.M.; Weller, M.T. Zinc tantalum Oxynitride (ZnTaO2N) photoanode modified with cobalt phosphate layers for the photoelectrochemical oxidation of alkali water. Nanomaterials 2018, 8, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Landsmann, S.; Maegli, A.E.; Trottmann, M.; Battaglia, C.; Weidenkaff, A.; Pokrant, S. Design guidelines for high-performance particle-based photoanodes for water splitting: Lanthanum titanium oxynitride as a model. ChemSusChem 2015, 8, 3451–3458. [Google Scholar] [CrossRef]
  101. Wang, J.; Wang, X.; Liu, B.; Li, X.; Cao, M. Facile synthesis of SrNbO2N nanoparticles with excellent visible-light photocatalytic performances. Mater. Lett. 2015, 152, 131–134. [Google Scholar] [CrossRef]
  102. Li, Y.; Nagato, K.; Delaunay, J.-J.; Kubota, J.; Domen, K. Fabrication of highly ordered Ta2O5 and Ta3N5 nanorod arrays by nanoimprinting and through-mask anodization. Nanotechnology 2013, 25, 014013. [Google Scholar] [CrossRef] [Green Version]
  103. Li, Y.; Takata, T.; Cha, D.; Takanabe, K.; Minegishi, T.; Kubota, J.; Domen, K. Vertically Aligned Ta3N5 Nanorod Arrays for Solar-Driven Photoelectrochemical Water Splitting. Adv. Mater. 2013, 25, 125–131. [Google Scholar] [CrossRef] [PubMed]
  104. Li, Y.; Zhang, L.; Torres-Pardo, A.; González-Calbet, J.M.; Ma, Y.; Oleynikov, P.; Terasaki, O.; Asahina, S.; Shima, M.; Cha, D. Cobalt phosphate-modified barium-doped tantalum nitride nanorod photoanode with 1.5% solar energy conversion efficiency. Nat. Commun. 2013, 4, 2566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Higashi, M.; Domen, K.; Abe, R. Fabrication of efficient TaON and Ta3N5 photoanodes for water splitting under visible light irradiation. Energy Environ. Sci. 2011, 4, 4138–4147. [Google Scholar] [CrossRef]
  106. Higashi, M.; Domen, K.; Abe, R. Highly stable water splitting on oxynitride TaON photoanode system under visible light irradiation. J. Am. Chem. Soc. 2012, 134, 6968–6971. [Google Scholar] [CrossRef] [PubMed]
  107. Theerthagiri, J.; Thiagarajan, K.; Senthilkumar, B.; Khan, Z.; Senthil, R.A.; Arunachalam, P.; Ashokkumar, M. Synthesis of hierarchical cobalt phosphate nanoflakes and their enhanced electrochemical performances for supercapacitor applications. ChemistrySelect 2017, 2, 201–210. [Google Scholar] [CrossRef]
  108. Duraisamy, N.; Arshid, N.; Kandiah, K.; Iqbal, J.; Arunachalam, P.; Dhanaraj, G.; Ramesh, S. Development of asymmetric device using Co 3 (PO 4) 2 as a positive electrode for energy storage application. J. Mater. Sci. Mater. Electron. 2019, 30, 7435–7446. [Google Scholar] [CrossRef]
  109. Haydous, F.; Si, W.; Guzenko, V.A.; Waag, F.; Pomjakushina, E.; El Kazzi, M.; Lippert, T. Improved Photoelectrochemical Water Splitting of CaNbO2N Photoanodes by CoPi Photodeposition and Surface Passivation. J. Phys. Chem. C 2018, 123, 1059–1068. [Google Scholar] [CrossRef] [Green Version]
  110. Abe, R.; Higashi, M.; Domen, K. Facile fabrication of an efficient oxynitride TaON photoanode for overall water splitting into H2 and O2 under visible light irradiation. J. Am. Chem. Soc. 2010, 132, 11828–11829. [Google Scholar] [CrossRef]
  111. Hisatomi, T.; Katayama, C.; Teramura, K.; Takata, T.; Moriya, Y.; Minegishi, T.; Katayama, M.; Nishiyama, H.; Yamada, T.; Domen, K. The effects of preparation conditions for a BaNbO2N photocatalyst on its physical properties. ChemSusChem 2014, 7, 2016–2021. [Google Scholar] [CrossRef]
  112. Zhang, F.; Yamakata, A.; Maeda, K.; Moriya, Y.; Takata, T.; Kubota, J.; Teshima, K.; Oishi, S.; Domen, K. Cobalt-modified porous single-crystalline LaTiO2N for highly efficient water oxidation under visible light. J. Am. Chem. Soc. 2012, 134, 8348–8351. [Google Scholar] [CrossRef]
  113. Yokoyama, D.; Hashiguchi, H.; Maeda, K.; Minegishi, T.; Takata, T.; Abe, R.; Kubota, J.; Domen, K. Ta3N5 photoanodes for water splitting prepared by sputtering. Thin Solid Film. 2011, 519, 2087–2092. [Google Scholar] [CrossRef]
  114. Zhang, L.; Song, Y.; Feng, J.; Fang, T.; Zhong, Y.; Li, Z.; Zou, Z. Photoelectrochemical water oxidation of LaTaON2 under visible-light irradiation. Int. J. Hydrog. Energy 2014, 39, 7697–7704. [Google Scholar] [CrossRef]
  115. Black, A.P.; Suzuki, H.; Higashi, M.; Frontera, C.; Ritter, C.; De, C.; Fuertes, A. New rare earth hafnium oxynitride perovskites with photocatalytic activity in water oxidation and reduction. Chem. Commun. 2018, 54, 1525–1528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Takata, T.; Pan, C.; Domen, K. Design and development of oxynitride photocatalysts for overall water splitting under visible light irradiation. ChemElectroChem 2016, 3, 31–37. [Google Scholar] [CrossRef]
  117. Siritanaratkul, B.; Maeda, K.; Hisatomi, T.; Domen, K. Synthesis and Photocatalytic Activity of Perovskite Niobium Oxynitrides with Wide Visible-Light Absorption Bands. ChemSusChem 2011, 4, 74–78. [Google Scholar] [CrossRef]
  118. Maeda, K.; Higashi, M.; Siritanaratkul, B.; Abe, R.; Domen, K. SrNbO2N as a water-splitting photoanode with a wide visible-light absorption band. J. Am. Chem. Soc. 2011, 133, 12334–12337. [Google Scholar] [CrossRef]
  119. Arunachalam, P.; Al-Mayouf, A.; Ghanem, M.A.; Shaddad, M.N.; Weller, M.T. Photoelectrochemical oxidation of water using La(Ta, Nb)O2N modified electrodes. Int. J. Hydrog. Energy 2016, 41, 11644–11652. [Google Scholar] [CrossRef]
  120. Shaddad, M.N.; Arunachalam, P.; Al-Mayouf, A.M.; Ghanem, M.A.; Alharthi, A.I. Enhanced photoelectrochemical oxidation of alkali water over cobalt phosphate (Co-Pi) catalyst-modified ZnLaTaON2 photoanodes. Ionics 2018, 25, 737–745. [Google Scholar] [CrossRef]
  121. Walsh, D.; Sanchez-Ballester, N.M.; Ting, V.P.; Hall, S.R.; Terry, L.R.; Weller, M.T. Visible light promoted photocatalytic water oxidation: Effect of metal oxide catalyst composition and light intensity. Catal. Sci. Technol. 2015, 5, 4760–4764. [Google Scholar] [CrossRef] [Green Version]
  122. Kanan, M.W.; Nocera, D.G. In situ formation of an oxygen-evolving catalyst in neutral water containing phosphate and Co2+. Science 2008, 321, 1072–1075. [Google Scholar] [CrossRef] [Green Version]
  123. Al-Sharif, M.S.; Arunachalam, P.; Abiti, T.; Amer, M.S.; Al-Shalwi, M.; Ghanem, M.A. Mesoporous cobalt phosphate electrocatalyst prepared using liquid crystal template for methanol oxidation reaction in alkaline solution. Arab. J. Chem. 2020, 13, 2873–2882. [Google Scholar] [CrossRef]
  124. Zhen, C.; Wang, L.; Liu, G.; Lu, G.Q.M.; Cheng, H.-M. Template-free synthesis of Ta3N5 nanorod arrays for efficient photoelectrochemical water splitting. Chem. Commun. 2013, 49, 3019–3021. [Google Scholar] [CrossRef] [PubMed]
  125. Roy, P.; Berger, S.; Schmuki, P. TiO2 nanotubes: Synthesis and applications. Angew. Chem. Int. Ed. 2011, 50, 2904–2939. [Google Scholar] [CrossRef] [PubMed]
  126. Song, F.; Hu, X. Exfoliation of layered double hydroxides for enhanced oxygen evolution catalysis. Nat. Commun. 2014, 5, 4477. [Google Scholar] [CrossRef] [PubMed]
  127. Gong, M.; Li, Y.; Wang, H.; Liang, Y.; Wu, J.Z.; Zhou, J.; Wang, J.; Regier, T.; Wei, F.; Dai, H. An advanced Ni–Fe layered double hydroxide electrocatalyst for water oxidation. J. Am. Chem. Soc. 2013, 135, 8452–8455. [Google Scholar] [CrossRef]
  128. Shao, M.; Ning, F.; Wei, M.; Evans, D.G.; Duan, X. Hierarchical nanowire arrays based on ZnO core-layered double hydroxide shell for largely enhanced photoelectrochemical water splitting. Adv. Funct. Mater. 2014, 24, 580–586. [Google Scholar] [CrossRef]
  129. Wang, L.; Dionigi, F.; Nguyen, N.T.; Kirchgeorg, R.; Gliech, M.; Grigorescu, S.; Strasser, P.; Schmuki, P. Tantalum nitride nanorod arrays: Introducing Ni–Fe layered double hydroxides as a cocatalyst strongly stabilizing photoanodes in water splitting. Chem. Mater. 2015, 27, 2360–2366. [Google Scholar] [CrossRef]
  130. Liao, M.; Feng, J.; Luo, W.; Wang, Z.; Zhang, J.; Li, Z.; Yu, T.; Zou, Z. Co3O4 nanoparticles as robust water oxidation catalysts towards remarkably enhanced photostability of a Ta3N5 photoanode. Adv. Funct. Mater. 2012, 22, 3066–3074. [Google Scholar] [CrossRef]
  131. Allam, N.K.; Poncheri, A.J.; El-Sayed, M.A. Vertically oriented Ti–Pd mixed oxynitride nanotube arrays for enhanced photoelectrochemical water splitting. ACS Nano 2011, 5, 5056–5066. [Google Scholar] [CrossRef]
  132. Chen, X.; Burda, C. Photoelectron spectroscopic investigation of nitrogen-doped titania nanoparticles. J. Phys. Chem. B 2004, 108, 15446–15449. [Google Scholar] [CrossRef]
  133. Kim, D.; Fujimoto, S.; Schmuki, P.; Tsuchiya, H. Nitrogen doped anodic TiO2 nanotubes grown from nitrogen-containing Ti alloys. Electrochem. Commun. 2008, 10, 910–913. [Google Scholar] [CrossRef]
  134. Soliman, K.A.; Zedan, A.F.; Khalifa, A.; El-Sayed, H.A.; Aljaber, A.S.; AlQaradawi, S.Y.; Allam, N.K. Silver nanoparticles-decorated titanium oxynitride nanotube arrays for enhanced solar fuel generation. Sci. Rep. 2017, 7, 1913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Ce’line, M.L.; Alexandra, E.M.; Kevin, S.; Takashi, H.; Nicolas, X.; Eugenio, H.O.; Songhak, Y.; Anke, W.; Rosendo, S.; Michael, G. LaTiO2N/In2O3 photoanodes with improved performance for solar water splitting. Chem. Commun. 2012, 48, 820–882. [Google Scholar]
  136. Gujral, S.S.; Simonov, A.; Higashi, M.; Fang, X.; Abe, R.; Spiccia, L. Highly Dispersed Cobalt Oxide on TaON as Efficient Photoanodes for Long-Term Solar Water Splitting. ACS Catal. 2016, 6, 3404–3417. [Google Scholar] [CrossRef]
  137. Higashi, M.; Domen, K.; Abe, R. Fabrication of an Efficient BaTaO2N Photoanode Harvesting a Wide Range of Visible Light for Water Splitting. J. Am. Chem. Soc. 2013, 135, 10238–10241. [Google Scholar] [CrossRef]
  138. Seo, J.; Hisatomi, T.; Nakabayashi, M.; Shibata, N.; Minegishi, T.; Katayama, M.; Domen, K. Efficient Solar-Driven Water Oxidation over Perovskite-Type BaNbO2N Photoanodes Absorbing Visible Light up to 740 nm. Adv. Energy Mater. 2018, 8, 1800094. [Google Scholar] [CrossRef]
  139. Urabe, H.; Hisatomi, T.; Minegishi, T.; Kubota, J.; Domen, K. Photoelectrochemical Properties of SrNbO2N Photoanodes for Water Oxidation Fabricated by the Particle Transfer Method. Faraday Discuss. 2014, 176, 213–223. [Google Scholar] [CrossRef]
  140. Ueda, K.; Minegishi, T.; Clune, J.; Nakabayashi, M.; Hisatomi, T.; Nishiyama, H.; Katayama, M.; Shibata, N.; Kubota, J.; Yamada, T.; et al. Photoelectrochemical Oxidation of Water Using BaTaO2N Photoanodes Prepared by Particle Transfer Method. J. Am. Chem. Soc. 2015, 137, 2227–2230. [Google Scholar] [CrossRef]
  141. Kodera, M.; Urabe, H.; Katayama, M.; Hisatomi, T.; Minegishi, T.; Domen, K. Effects of Flux Synthesis on SrNbO2N Particles for Photoelectrochemical Water Splitting. J. Mater. Chem. A 2016, 4, 7658–7664. [Google Scholar] [CrossRef] [Green Version]
  142. Seo, J.; Nakabayashi, M.; Hisatomi, T.; Shibata, N.; Minegishi, T.; Katayama, M.; Domen, K. The effects of annealing barium niobium oxynitride in argon on photoelectrochemical water oxidation activity. J. Mater. Chem. A 2019, 7, 493–502. [Google Scholar] [CrossRef]
  143. Zili, M.; Aleksander, J.; Janine, G.; Anna, R.; Thomas, T.; Tetyana, M.B.; Geoffroy, H.; Andrew, J.P.; Richar, D.; Piotr, K.; et al. Exploring the Origins of Improved Photocurrent by Acidic Treatment for Quaternary Tantalum-based Oxynitride Photoanodes on the Example of CaTaO2N. J. Phys. Chem. C 2020, 124, 152–160. [Google Scholar]
  144. Maoqi, C.; Hongmei, L.; Kang, L.; Junhua, H.; Hao, P.; Junwei, F.; Min, L. Vertical SrNbO2N Nanorod Arrays for Solar-Driven Photoelectrochemical Water Splitting. Solar RRL 2020, 202000448. [Google Scholar] [CrossRef]
  145. Yingchen, Y.; Zirui, L.; Weisheng, L.; Yichen, W.; Rong, L.; Chao, Q.; Liping, Z. Enhanced photoelectrochemical water-splitting performance of SrNbO2N photoanodes using flux-assisted synthesis method and surface defect management. Sustain. Energy Fuels 2020, 4, 1674–1680. [Google Scholar]
  146. Ishikawa, A.; Takata, T.; Kondo, J.N.; Hara, M.; Kobayashi, H.; Domen, K. Oxysulfide Sm2Ti2S2O5 as a stable photocatalyst for water oxidation and reduction under visible light irradiation (λ ≤ 650 nm). J. Am. Chem. Soc. 2002, 124, 13547–13553. [Google Scholar] [CrossRef]
  147. Ishikawa, A.; Yamada, Y.; Takata, T.; Kondo, J.N.; Hara, M.; Kobayashi, H.; Domen, K. Novel synthesis and photocatalytic activity of oxysulfide Sm2Ti2S2O5. Chem. Mater. 2003, 15, 4442–4446. [Google Scholar] [CrossRef]
  148. Williams, R. Becquerel photovoltaic effect in binary compounds. J. Chem. Phys. 1960, 32, 1505–1514. [Google Scholar] [CrossRef]
  149. Ellis, A.B.; Kaiser, S.W.; Bolts, J.M.; Wrighton, M.S. Study of n-type semiconducting cadmium chalcogenide-based photoelectrochemical cells employing polychalcogenide electrolytes. J. Am. Chem. Soc. 1977, 99, 2839–2848. [Google Scholar] [CrossRef]
  150. Harriman, A.; Thomas, J.; Milward, G. Catalytic and structural properties of iridium-iridium dioxide colloids. New J. Chem. 1987, 11, 757–762. [Google Scholar]
  151. Harriman, A.; Pickering, I.J.; Thomas, J.M.; Christensen, P.A. Metal oxides as heterogeneous catalysts for oxygen evolution under photochemical conditions. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1988, 84, 2795–2806. [Google Scholar] [CrossRef]
  152. Ma, G.; Chen, S.; Kuang, Y.; Akiyama, S.; Hisatomi, T.; Nakabayashi, M.; Shibata, N.; Katayama, M.; Minegishi, T.; Domen, K. Visible light-driven Z-scheme water splitting using oxysulfide H2 evolution photocatalysts. J. Phys. Chem. Lett. 2016, 7, 3892–3896. [Google Scholar] [CrossRef]
  153. Ishikawa, A.; Takata, T.; Matsumura, T.; Kondo, J.N.; Hara, M.; Kobayashi, H.; Domen, K. Oxysulfides Ln2Ti2S2O5 as stable photocatalysts for water oxidation and reduction under visible-light irradiation. J. Phys. Chem. B 2004, 108, 2637–2642. [Google Scholar] [CrossRef]
  154. Zhang, F.; Maeda, K.; Takata, T.; Domen, K. Modification of oxysulfides with two nanoparticulate cocatalysts to achieve enhanced hydrogen production from water with visible light. Chem. Commun. 2010, 46, 7313–7315. [Google Scholar] [CrossRef] [PubMed]
  155. Ogisu, K.; Ishikawa, A.; Teramura, K.; Toda, K.; Hara, M.; Domen, K. Lanthanum–indium oxysulfide as a visible light driven photocatalyst for water splitting. Chem. Lett. 2007, 36, 854–855. [Google Scholar] [CrossRef]
  156. Ogisu, K.; Ishikawa, A.; Shimodaira, Y.; Takata, T.; Kobayashi, H.; Domen, K. Electronic band structures and photochemical properties of La−Ga-based oxysulfides. J. Phys. Chem. C 2008, 112, 11978–11984. [Google Scholar] [CrossRef]
  157. Liu, J.; Hisatomi, T.; Ma, G.; Iwanaga, A.; Minegishi, T.; Moriya, Y.; Katayama, M.; Kubota, J.; Domen, K. Improving the photoelectrochemical activity of La5Ti2CuS5O7 for hydrogen evolution by particle transfer and doping. Energy Environ. Sci. 2014, 7, 2239–2242. [Google Scholar] [CrossRef] [Green Version]
  158. Ma, G.; Suzuki, Y.; Singh, R.B.; Iwanaga, A.; Moriya, Y.; Minegishi, T.; Liu, J.; Hisatomi, T.; Nishiyama, H.; Katayama, M. Photoanodic and photocathodic behaviour of La5Ti2Cu5O7 electrodes in the water splitting reaction. Chem. Sci. 2015, 6, 4513–4518. [Google Scholar] [CrossRef] [Green Version]
  159. Hisatomi, T.; Okamura, S.; Liu, J.; Shinohara, Y.; Ueda, K.; Higashi, T.; Katayama, M.; Minegishi, T.; Domen, K. La5Ti2Cu1−xAgxS5O7 photocathodes operating at positive potentials during photoelectrochemical hydrogen evolution under irradiation of up to 710 nm. Energy Environ. Sci. 2015, 8, 3354–3362. [Google Scholar] [CrossRef] [Green Version]
  160. Yoshimizu, M.; Kobayashi, R.; Saegusa, M.; Takashima, T.; Funakubo, H.; Akiyama, K.; Yoshihisa, M.; Irie, H. Photocatalytic hydrogen evolution over β-iron silicide under infrared-light irradiation. Chem. Commun. 2015, 51, 2818–2820. [Google Scholar] [CrossRef]
  161. Akiyama, K.; Motoizumi, Y.; Funakubo, H.; Irie, H.; Matsumoto, Y. Metal–organic chemical vapor deposition growth of β-FeSi2/Si composite powder via vapor–liquid–solid method and its photocatalytic properties. Jpn. J. Appl. Phys. 2016, 55, 06HC02. [Google Scholar] [CrossRef]
  162. Akiyama, K.; Motoizumi, Y.; Okuda, T.; Funakubo, H.; Irie, H.; Matsumoto, Y. Synthesis and Photocatalytic Properties of Iron Disilicide/SiC Composite Powder. MRS Adv. 2017, 2, 471–476. [Google Scholar] [CrossRef]
  163. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef] [PubMed]
  164. Liu, J.; Liu, Y.; Liu, N.; Han, Y.; Zhang, X.; Huang, H.; Lifshitz, Y.; Lee, S.T.; Zhong, J.; Kang, Z. Metal-free efficient photocatalyst for stable visible water splitting via a two-electron pathway. Science 2015, 347, 970–974. [Google Scholar] [CrossRef] [PubMed]
  165. Zhang, Z.; Long, J.; Yang, L.; Chen, W.; Dai, W.; Fu, X.; Wang, X. Organic semiconductor for artificial photosynthesis: Water splitting into hydrogen by a bioinspired C3N3S3 polymer under visible light irradiation. Chem. Sci. 2011, 2, 1826–1830. [Google Scholar] [CrossRef]
  166. Liu, G.; Niu, P.; Yin, L.; Cheng, H.M. α-Sulfur crystals as a visible-light-active photocatalyst. J. Am. Chem. Soc. 2012, 134, 9070–9073. [Google Scholar] [CrossRef]
  167. Wang, F.; Ng, W.K.H.; Jimmy, C.Y.; Zhu, H.; Li, C.; Zhang, L.; Li, Q. Red phosphorus: An elemental photocatalyst for hydrogen formation from water. Appl. Catal. B Environ. 2012, 111, 409–414. [Google Scholar] [CrossRef]
  168. Zhu, M.; Sun, Z.; Fujitsuka, M.; Majima, T. Z-Scheme Photocatalytic Water Splitting on a 2D Heterostructure of Black Phosphorus/Bismuth Vanadate Using Visible Light. Angew. Chem. Int. Ed. 2018, 57, 2160–2164. [Google Scholar] [CrossRef] [PubMed]
  169. Hou, Y.; Abrams, B.L.; Vesborg, P.C.; Björketun, M.E.; Herbst, K.; Bech, L.; Setti, A.M.; Damsgaard, C.D.; Pedersen, T.; Hansen, O.; et al. Bioinspired molecular co-catalysts bonded to a silicon photocathode for solar hydrogen evolution. Nat. Mater. 2011, 10, 434–438. [Google Scholar] [CrossRef] [Green Version]
  170. Liu, J.; Wen, S.; Hou, Y.; Zuo, F.; Beran, G.J.O.; Feng, P. Boron Carbides as Efficient, Metal-Free, Visible-Light-Responsive Photocatalysts. Angew. Chem. Int. Ed. 2013, 52, 3241–3245. [Google Scholar] [CrossRef]
  171. Chen, S.N.; Heeger, A.J.; Kiss, Z.; MacDiarmid, A.G.; Gau, S.C.; Peebles, D.L. Polyacetylene, (CH)x: Photoelectrochemical solar cell. Appl. Phys. Lett. 1980, 36, 96–98. [Google Scholar] [CrossRef]
  172. Yamase, T.; Harada, H.; Ikawa, T.; Ikeda, S.; Shirakawa, H. A Photoelectrochemical Study of Polyacetylene, (CH)x. Bull. Chem. Soc. Jpn. 1981, 54, 2817–2818. [Google Scholar] [CrossRef] [Green Version]
  173. Aizawa, M.; Watanabe, S.; Shinohara, H.; Shirakawa, H. Photodoping of polyacetylene films. J. Chem. Soc. Chem. Commun. 1985, 62–63. [Google Scholar] [CrossRef]
  174. Kaneko, M.; Okuzumi, K.; Yamada, A. Photocurrent generation by a liquid-junction poly (pyrrole) film. J. Electroanal. Chem. Interfacial Electrochem. 1985, 183, 407–410. [Google Scholar] [CrossRef]
  175. Genies, E.M.; Lapkowski, M. Influence of light on the electrochemical behavior of polyaniline films. Synth. Met. 1988, 24, 69–76. [Google Scholar] [CrossRef]
  176. Hursán, D.; Kormányos, A.; Rajeshwar, K.; Janáky, C. Polyaniline films photoelectrochemically reduce CO2 to alcohols. Chem. Commun. 2016, 52, 8858–8861. [Google Scholar] [CrossRef] [PubMed]
  177. Glenis, S.; Tourillon, G.; Garnier, F. Photoelectrochemical properties of thin films of polythiophene and derivatives: Doping level and structure effects. Thin Solid Film. 1984, 122, 9–17. [Google Scholar] [CrossRef]
  178. El-Rashiedy, O.A.; Holdcroft, S. Photoelectrochemical properties of poly (3-alkylthiophene) films in aqueous solution. J. Phys. Chem. 1996, 100, 5481–5484. [Google Scholar] [CrossRef]
  179. Suppes, G.; Ballard, E.; Holdcroft, S. Aqueous photocathode activity of regioregular poly (3-hexylthiophene). Polym. Chem. 2013, 4, 5345–5350. [Google Scholar] [CrossRef]
  180. Schlettwein, D.; Jaeger, N.I. Identification of the Mechanism in the Photoelectrochemical Reduction of Oxygen on the Surface of a Molecular Semiconductor. J. Phys. Chem. 1993, 97, 3333–3337. [Google Scholar] [CrossRef]
  181. Schlettwein, D.; Armstrong, N.R. Correlation of frontier orbital positions and conduction type of molecular semiconductors as derived from UPS in combination with electrical and photoelectrochemical experiments. J. Phys. Chem. 1994, 98, 11771–11779. [Google Scholar] [CrossRef]
  182. Schlettwein, D.; Kaneko, M.; Yamada, A.; Wöhrle, D.; Jaeger, N.I. Light-induced dioxygen reduction at thin film electrodes of various porphyrins. J. Phys. Chem. 1991, 95, 1748–1755. [Google Scholar] [CrossRef]
  183. Yanagi, H.; Kanbayashi, Y.; Schlettwein, D.; Woehrle, D.; Armstrong, N.R. Photoelectrochemical investigations on naphthalocyanine derivatives in thin films. J. Phys. Chem. 1994, 98, 4760–4766. [Google Scholar] [CrossRef]
  184. Yanagida, S.; Kabumoto, A.; Mizumoto, K.; Pac, C.; Yoshino, K. Poly (p-phenylene)-catalysed photoreduction of water to hydrogen. J. Chem. Soc. Chem. Commun. 1991, 474–475. [Google Scholar] [CrossRef]
  185. Shibata, T.S.; Kabumoto, A.; Hiragami, T.S.; Ishitani, O.; Pac, C.; Yanagida, S. Novel visible-light-driven photocatalyst. Poly (p-phenylene)-catalyzed photoreductions of water, carbonyl compounds, and olefins. J. Phys. Chem. 1990, 94, 2068–2076. [Google Scholar] [CrossRef]
  186. Matsuoka, S.; Fujii, H.; Yamada, T.; Pac, C.; Ishida, A.; Takamuku, S.; Kusaba, M.; Nakashima, N.; Yanagida, S. Photocatalysis of oligo (p-phenylenes): Photoreductive production of hydrogen and ethanol in aqueous triethylamine. J. Phys. Chem. 1991, 95, 5802–5808. [Google Scholar] [CrossRef]
  187. Maruo, K.; Yamada, K.; Wada, Y.; Yanagida, S. Visible-light induced photocatalysis of partially fluorinated poly (p-phenylene) and related linear phenylene derivatives. Bull. Chem. Soc. Jpn. 1993, 66, 1053–1064. [Google Scholar] [CrossRef]
  188. Li, S.; Ye, L.; Zhao, W.; Yan, H.; Yang, B.; Liu, D.; Li, W.; Ade, H.; Hou, J. A Wide Band Gap Polymer with a Deep Highest Occupied Molecular Orbital Level Enables 14.2% Efficiency in Polymer Solar Cells. J. Am. Chem. Soc. 2018, 140, 7159–7167. [Google Scholar] [CrossRef]
  189. Green, M.A.; Emery, K.; Hishikawa, Y.; Warta, W.; Dunlop, E.D. Solar cell efficiency tables (version 46). Prog. Photovoltaics Res. Appl. 2015, 23, 1–9. [Google Scholar] [CrossRef]
  190. Thompson, B.C.; Fréchet, J.M. Polymer–fullerene composite solar cells. Angew. Chem. Int. Ed. 2008, 47, 58–77. [Google Scholar] [CrossRef]
  191. Günes, S.; Neugebauer, H.; Sariciftci, N.S. Conjugated polymer-based organic solar cells. Chem. Rev. 2007, 107, 1324–1338. [Google Scholar] [CrossRef]
  192. Matsuo, Y.; Sato, Y.; Niinomi, T.; Soga, I.; Tanaka, H.; Nakamura, E. Columnar structure in bulk heterojunction in solution-processable three-layered pin organic photovoltaic devices using tetrabenzoporphyrin precursor and silylmethyl fullerene. J. Am. Chem. Soc. 2009, 131, 16048–16050. [Google Scholar] [CrossRef]
  193. Hiramoto, M.; Fujiwara, H.; Yokoyama, M. Three-layered organic solar cell with a photoactive interlayer of codeposited pigments. Appl. Phys. Lett. 1991, 58, 1062–1064. [Google Scholar] [CrossRef]
  194. Peumans, P.; Uchida, S.; Forrest, S.R. Efficient bulk heterojunction photovoltaic cells using small-molecular-weight organic thin films. Nature 2003, 425, 158–162. [Google Scholar] [CrossRef] [PubMed]
  195. Sariciftci, N.S.; Smilowitz, L.; Heeger, A.J.; Wudl, F. Photoinduced electron transfer from a conducting polymer to buckminsterfullerene. Science 1992, 258, 1474–1476. [Google Scholar] [CrossRef] [PubMed]
  196. Tang, C.W. Two-layer organic photovoltaic cell. Appl. Phys. Lett. 1986, 48, 183–185. [Google Scholar] [CrossRef]
  197. Abe, T.; Ichinohe, H.; Kakuta, S.; Nagai, K. Organic photoanode of fullerene/phthalocyanine working in the water phase with respect to preparation methods of the bilayer film. Jpn. J. Appl. Phys. 2010, 49, 015101. [Google Scholar] [CrossRef]
  198. Abe, T.; Tanno, Y.; Ebina, T.; Miyakushi, S.; Nagai, K. Enhanced photoanodic output at an organic p/n bilayer in the water phase by means of the formation of whiskered phthalocyanine. ACS Appl. Mater. Interfaces 2013, 5, 1248–1253. [Google Scholar] [CrossRef]
  199. Ahmad, M.F.; Abe, T.; Musgrave, C.S.; Nagai, K. Study of Co-deposition Photoelectrode of Perylene Derivative and Phthalocyanine in Comparison with Its Bilayer Focusing on Charge Transfer Complex and Kinetic Analysis. Electrochemistry 2018, 235–242. [Google Scholar] [CrossRef] [Green Version]
  200. Ahmad, M.F.; Suzuki, M.; Abe, T.; Nagai, K. Enhanced oxidation power in photoelectrocatalysis based on a micrometer-localized positive potential in a terrace hetero p–n junction. NPG Asia Mater. 2018, 10, 630–641. [Google Scholar] [CrossRef]
  201. Abe, T.; Ichikawa, M.; Hikage, T.; Kakuta, S.; Nagai, K. Relationship between the morphology of poly (3-hexylthiophene)/methanofullerene composite and its photoelectrode characteristics in the water phase. Chem. Phys. Lett. 2012, 549, 77–81. [Google Scholar] [CrossRef]
  202. Nagai, K.; Kuwabara, T.; Ahmad, M.-A.; Nakano, M.; Karakawa, M.; Taima, T.; Takahashi, K. High performance photoanodic catalyst prepared from an active organic photovoltaic cell -high potential gain from visible light. Chem. Commun. 2019, 55, 12491–12494. [Google Scholar] [CrossRef] [Green Version]
  203. Nagai, K.; Abe, T. Full-Spectrum-Visible-Light Photocatalyst Based on the Active Layer of Organic Solar Cell-Towards Water Splitting and Volatile Molecule Degradation. Kobunshi Ronbunshu 2013, 7, 459–475. [Google Scholar] [CrossRef]
  204. Abe, T.; Nagai, K.; Kabutomori, S.; Kaneko, M.; Tajiri, A.; Norimatsu, T. An Organic Photoelectrode Working in the Water Phase: Visible-Light-Induced Dioxygen Evolution by a Perylene Derivative/Cobalt Phthalocyanine Bilayer. Angew. Chem. Int. Ed. 2006, 45, 2778–2781. [Google Scholar] [CrossRef] [PubMed]
  205. Abe, T.; Nagai, K.; Ogiwara, T.; Ogasawara, S.; Kaneko, M.; Tajiri, A.; Norimatsu, T. Wide visible light-induced dioxygen evolution at an organic photoanode coated with a noble metal oxide catalyst. J. Electroanal. Chem. 2006, 587, 127–132. [Google Scholar] [CrossRef]
  206. Yang, S.; Fan, L.; Yang, S. Significantly enhanced photocurrent efficiency of a poly (3-hexylthiophene) photoelectrochemical device by doping with the endohedral metallofullerene Dy@ C82. Chem. Phys. Lett. 2004, 388, 253–258. [Google Scholar] [CrossRef]
  207. Bornoz, P.; Prévot, M.S.; Yu, X.; Guijarro, N.; Sivula, K. Direct light-driven water oxidation by a ladder-type conjugated polymer photoanode. J. Am. Chem. Soc. 2015, 137, 15338–15341. [Google Scholar] [CrossRef] [PubMed]
  208. Lanzarini, E.; Antognazza, M.R.; Biso, M.; Ansaldo, A.; Laudato, L.; Bruno, P.; Metrangolo, P.; Resnati, G.; Ricci, D.; Lanzani, G. Polymer-based photocatalytic hydrogen generation. J. Phys. Chem. C 2012, 116, 10944–10949. [Google Scholar] [CrossRef]
  209. Chu, S.; Li, W.; Yan, Y.; Hamann, T.; Shih, I.; Wang, D.; Mi, Z. Roadmap on solar water splitting: Current status and future prospects. Nano Futures 2017, 1, 022001. [Google Scholar] [CrossRef]
  210. Mezzetti, A.; Fumagalli, F.; Alfano, A.; Iadicicco, D.; Antognazza, M.R.; di Fonzo, F. Stable hybrid organic/inorganic photocathodes for hydrogen evolution with amorphous WO 3 hole selective contacts. Faraday Discuss. 2017, 198, 433–448. [Google Scholar] [CrossRef]
  211. Wang, L.; Yan, D.; Shaffer, D.W.; Ye, X.; Layne, B.H.; Concepcion, J.J.; Liu, M.; Nam, C.Y. Improved Stability and Performance of Visible Photoelectrochemical Water Splitting on Solution-Processed Organic Semiconductor Thin Films by Ultrathin Metal Oxide Passivation. Chem. Mater. 2018, 30, 324–335. [Google Scholar] [CrossRef]
  212. Francàs, L.; Burns, E.; Steier, L.; Cha, H.; Solà-Hernández, L.; Li, X.; Tuladhar, P.S.; Bofill, R.; Garcia-Anton, J.; Sala, X.; et al. Rational design of a neutral pH functional and stable organic photocathode. Chem. Commun. 2018, 54, 5732–5735. [Google Scholar] [CrossRef] [Green Version]
  213. Sun, T.; Song, J.; Jia, J.; Li, X.; Sun, X. Real roles of perylenetetracarboxylic diimide for enhancing photocatalytic H2-production. Nano Energy 2016, 26, 83–89. [Google Scholar] [CrossRef]
  214. Li, J.X.; Li, Z.J.; Ye, C.; Li, X.B.; Zhan, F.; Fan, X.B.; Li, J.; Chen, B.; Tao, Y.; Tung, C.H.; et al. Visible light-induced photochemical oxygen evolution from water by 3, 4, 9, 10-perylenetetracarboxylic dianhydride nanorods as an n-type organic semiconductor. Catal. Sci. Technol. 2016, 6, 672–676. [Google Scholar] [CrossRef]
  215. Kirner, J.T.; Stracke, J.J.; Gregg, B.A.; Finke, R.G. Visible-light-assisted photoelectrochemical water oxidation by thin films of a phosphonate-functionalized perylene diimide plus CoO x cocatalyst. ACS Appl. Mater. Interfaces 2014, 6, 13367–13377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Abe, T.; Chiba, J.; Ishidoya, M.; Nagai, K. Organophotocatalysis system of p/n bilayers for wide visible-light-induced molecular hydrogen evolution. RSC Adv. 2012, 2, 7992–7996. [Google Scholar] [CrossRef]
  217. Abe, T.; Fukui, K.; Kawai, Y.; Nagai, K.; Kato, H. A water splitting system using an organo-photocathode and titanium dioxide photoanode capable of bias-free H2 and O2 evolution. Chem. Commun. 2016, 52, 7735–7737. [Google Scholar] [CrossRef] [Green Version]
  218. Abe, T.; Taira, N.; Tanno, Y.; Kikuchi, Y.; Nagai, K. Decomposition of hydrazine by an organic fullerene–phthalocyanine p–n bilayer photocatalysis system over the entire visible-light region. Chem. Commun. 2014, 50, 1950–1952. [Google Scholar] [CrossRef]
  219. Abe, T.; Tanno, Y.; Taira, N.; Nagai, K. Efficient organo-photocatalysis system of an n-type perylene derivative/p-type cobalt phthalocyanine bilayer for the production of molecular hydrogen from hydrazine. RSC Adv. 2015, 5, 46325–46329. [Google Scholar] [CrossRef] [Green Version]
  220. Kawai, Y.; Nagai, K.; Abe, T. A visible-light-induced photoelectrochemical water-splitting system featuring an organo-photocathode along with a tungsten oxide photoanode. RSC Adv. 2017, 7, 34694–34698. [Google Scholar] [CrossRef] [Green Version]
  221. Bourgeteau, T.; Tondelier, D.; Geffroy, B.; Brisse, R.; Laberty-Robert, C.; Campidelli, S.; Bettignies, R.D.; Artero, V.; Palacin, S.; Jousselme, B. A H2-evolving photocathode based on direct sensitization of MoS3 with an organic photovoltaic cell. Energy Environ. Sci. 2013, 6, 2706–2713. [Google Scholar] [CrossRef] [Green Version]
  222. Tran, P.D.; Tran, T.V.; Orio, M.; Torelli, S.; Truong, Q.D.; Nayuki, K.; Sasaki, Y.; Chiam, S.Y.; Yi, R.; Honma, I.; et al. Coordination polymer structure and revisited hydrogen evolution catalytic mechanism for amorphous molybdenum sulfide. Nat. Mater. 2016, 15, 640. [Google Scholar] [CrossRef]
  223. Haro, M.; Solis, C.; Molina, G.; Otero, L.; Bisquert, J.; Gimenez, S.; Guerrero, A. Toward stable solar hydrogen generation using organic photoelectrochemical cells. J. Phys. Chem. C 2015, 119, 6488–6494. [Google Scholar] [CrossRef] [Green Version]
  224. Steier, L.; Bellani, S.; Rojas, H.C.; Pan, L.; Laitinen, M.; Sajavaara, T.; Fonzo, F.D.; Grätzel, M.; Antognazza, M.R.; Mayer, M.T. Stabilizing organic photocathodes by low-temperature atomic layer deposition of TiO2. Sustain. Energy Fuels 2017, 1, 1915–1920. [Google Scholar] [CrossRef]
  225. Bourgeteau, T.; Tondelier, D.; Geffroy, B.; Brisse, R.; Cornut, R.; Artero, V.; Jousselme, B. Enhancing the Performances of P3HT: PCBM–MoS3-Based H2-Evolving Photocathodes with Interfacial Layers. ACS Appl. Mater. Interfaces 2015, 7, 16395–16403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Bourgeteau, T.; Tondelier, D.; Geffroy, B.; Brisse, R.; Campidelli, S.; Cornut, R.; Jousselme, B. All solution-processed organic photocathodes with increased efficiency and stability via the tuning of the hole-extracting layer. J. Mater. Chem. A 2016, 4, 4831–4839. [Google Scholar] [CrossRef]
  227. Bellani, S.; Antognazza, M.R.; Bonaccorso, F. Carbon-Based Photocathode Materials for Solar Hydrogen Production. Adv. Mater. 2018, 1801446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Fumagalli, F.; Bellani, S.; Schreier, M.; Leonardi, S.; Rojas, H.C.; Ghadirzadeh, A.; Tullii, G.; Savoini, A.; Marra, G.; Meda, L.; et al. Hybrid organic-inorganic H2-evolving photocathodes: Understanding the route towards high performance organic photoelectrochemical water splitting. J. Mater. Chem. A 2016, 4, 2178–2187. [Google Scholar] [CrossRef]
  229. Belarb, E.; Blas-Ferrando, V.M.; Haro, M.; Maghraoui-Meherzi, H.; Gimenez, S. Electropolymerized polyaniline: A promising hole selective contact in organic photoelectrochemical cells. Chem. Eng. Sci. 2016, 154, 143–149. [Google Scholar] [CrossRef]
  230. Rojas, H.C.; Bellani, S.; Fumagalli, F.; Tullii, G.; Leonardi, S.; Mayer, M.T.; Schriever, M.; Grätzel, M.; Lanzani, G.; Fonzo, F.D.; et al. Polymer-based photocathodes with a solution-processable cuprous iodide anode layer and a polyethyleneimine protective coating. Energy Environ. Sci. 2016, 9, 3710–3723. [Google Scholar] [CrossRef] [Green Version]
  231. Morozan, A.; Bourgeteau, T.; Tondelier, D.; Geffroy, B.; Jousselme, B.; Artero, V. Noble metal-free hydrogen-evolving photocathodes based on small molecule organic semiconductors. Nanotechnology 2016, 27, 355401. [Google Scholar] [CrossRef] [Green Version]
  232. Gao, Y.; Le Corre, V.M.; Gaïtis, A.; Neophytou, M.; Hamid, M.A.; Takanabe, K.; Beaujuge, P.M. Homo-Tandem Polymer Solar Cells with VOC > 1.8 V for Efficient PV-Driven Water Splitting. Adv. Mater. 2016, 28, 3366–3373. [Google Scholar] [CrossRef]
  233. You, J.; Chen, C.C.; Hong, Z.; Yoshimura, K.; Ohya, K.; Xu, R.; Ye, S.; Gao, J.; Li, G.; Yang, Y. 10.2% power conversion efficiency polymer tandem solar cells consisting of two identical sub-cells. Adv. Mater. 2013, 25, 3973–3978. [Google Scholar] [CrossRef] [PubMed]
  234. Esiner, S.; van Eersel, H.; Wienk, M.M.; Janssen, R.A. Triple junction polymer solar cells for photoelectrochemical water splitting. Adv. Mater. 2013, 25, 2932–2936. [Google Scholar] [CrossRef] [PubMed]
  235. Esiner, S.; van Pruissen, G.W.; Wienk, M.M.; Janssen, R.A. Optimized light-driven electrochemical water splitting with tandem polymer solar cells. J. Mater. Chem. A 2016, 4, 5107–5114. [Google Scholar] [CrossRef] [Green Version]
  236. Esiner, S.; Willems, R.E.; Furlan, A.; Li, W.; Wienk, M.M.; Janssen, R.A. Photoelectrochemical water splitting in an organic artificial leaf. J. Mater. Chem. 2015, 3, 23936–23945. [Google Scholar] [CrossRef] [Green Version]
  237. Chen, S.; Zhang, G.; Liu, J.; Yao, H.; Zhang, J.; Ma, T.; Yan, H. An All-Solution Processed Recombination Layer with Mild Post-Treatment Enabling Efficient Homo-Tandem Non-fullerene Organic Solar Cells. Adv. Mater. 2017, 29, 1604231. [Google Scholar] [CrossRef]
  238. Esiner, S.; van Eersel, H.; van Pruissen, G.W.; Turbiez, M.; Wienk, M.M.; Janssen, R.A. Water splitting with series-connected polymer solar cells. ACS Appl. Mater. Interfaces 2016, 8, 26972–26981. [Google Scholar] [CrossRef]
  239. Windle, C.D.; Chandrasekaran, S.; Kumagai, H.; Sahara, G.; Nagai, K.; Abe, T.; Chavarot-Kerlidou, M.; Ishitani, O.; Artero, V. Molecular Design of Photocathode Materials for Hydrogen Evolution and Carbon Dioxide Reduction, Chapter 10. In Molecular Technology: Energy Innovation; Yamamoto, H., Kato, T., Eds.; Wiley: Hoboken, NJ, USA, 2018; pp. 251–286. ISBN 9783527341. [Google Scholar]
  240. Li, C.; Liu, M.; Pschirer, N.G.; Baumgarten, M.; Mullen, K. Polyphenylene-based materials for organic photovoltaics. Chem. Rev. 2010, 110, 6817–6855. [Google Scholar] [CrossRef]
  241. Steier, L.; Holliday, S. A bright outlook on organic photoelectrochemical cells for water splitting. J. Mater. Chem. A 2018, 6, 21809–21826. [Google Scholar] [CrossRef] [Green Version]
  242. Kirner, J.T.; Finke, R.G. Water-oxidation photoanodes using organic light-harvesting materials: A review. J. Mater. Chem. A 2017, 5, 19560–19592. [Google Scholar] [CrossRef]
  243. Rudolf, M.; Kirner, S.V.; Guldi, D.M. A multicomponent molecular approach to artificial photosynthesis–the role of fullerenes and endohedral metallofullerenes. Chem. Soc. Rev. 2016, 45, 612–630. [Google Scholar] [CrossRef] [Green Version]
  244. Queyriaux, N.; Kaeffer, N.; Morozan, A.; Chavarot-Kerlidou, M.; Artero, V. Molecular cathode and photocathode materials for hydrogen evolution in photoelectrochemical devices. J. Photochem. Photobiol. C Photochem. Rev. 2015, 25, 90–105. [Google Scholar] [CrossRef]
  245. Artero, V.; Chavarot-Kerlidou, M.; Fontecave, M. Splitting water with cobalt. Angew. Chem. Int. Ed. 2011, 50, 7238–7266. [Google Scholar] [CrossRef] [PubMed]
  246. Wen, J.; Xie, J.; Chen, X.; Li, X. A review on g-C3N4-based photocatalysts. Appl. Surf. Sci. 2017, 391, 72–123. [Google Scholar] [CrossRef]
  247. Chen, Y.Z.; Li, W.H.; Li, L.; Wang, L.N. Progress in organic photocatalysts. Rare Met. 2018, 37, 1–12. [Google Scholar] [CrossRef]
  248. Sprick, R.S.; Bonillo, B.; Sachs, M.; Clowes, R.; Durrant, J.R.; Adams, D.J.; Cooper, A.I. Extended conjugated microporous polymers for photocatalytic hydrogen evolution from water. Chem. Commun. 2016, 52, 10008–10011. [Google Scholar] [CrossRef] [Green Version]
  249. Sick, T.; Hufnagel, A.G.; Kampmann, J.; Kondofersky, I.; Calik, M.; Rotter, J.M.; Böhm, D. Oriented films of conjugated 2D covalent organic frameworks as photocathodes for water splitting. J. Am. Chem. Soc. 2018, 140, 2085–2092. [Google Scholar] [CrossRef] [Green Version]
  250. Kumaravel, V.; Bartlett, J.; Pillai, S.C. Photoelectrochemical conversion of carbon dioxide (CO2) into fuels and value-added products. ACS Energy Lett. 2020, 5, 486–519. [Google Scholar] [CrossRef] [Green Version]
  251. Galan-Mascaros, J.R. Photoelectrochemical solar fuels from carbon dioxide, water and sunlight. Catal. Sci. Technol. 2020, 10, 1967–1974. [Google Scholar] [CrossRef]
  252. Montoya, J.H.; Seitz, L.C.; Chakthranont, P.; Vojvodic, A.; Jaramillo, T.F.; Nørskov, J.K. Materials for solar fuels and chemicals. Nat. Mater. 2017, 16, 70–81. [Google Scholar] [CrossRef]
  253. Kalamaras, E.; Maroto-Valer, M.M.; Shao, M.; Xuan, J.; Wang, H. Solar carbon fuel via photoelectrochemistry. Catal. Today 2018, 317, 56–75. [Google Scholar] [CrossRef]
Figure 1. Energy diagram with the conduction and valence band positions for different semiconductor materials. Water oxidation (H2O/O2) and reduction potentials (H+/H2) are also shown.
Figure 1. Energy diagram with the conduction and valence band positions for different semiconductor materials. Water oxidation (H2O/O2) and reduction potentials (H+/H2) are also shown.
Catalysts 11 00160 g001
Figure 2. A schematic energy illustration for a photoanodes (n-type semiconductor). Several phases of process are displayed, specifically: (i) light-absorption; (ii) charge-transfer; (iii) transportation of charge-induced electron-hole pairs; and (iv) surface reactions.
Figure 2. A schematic energy illustration for a photoanodes (n-type semiconductor). Several phases of process are displayed, specifically: (i) light-absorption; (ii) charge-transfer; (iii) transportation of charge-induced electron-hole pairs; and (iv) surface reactions.
Catalysts 11 00160 g002
Figure 3. Schematic representation of electronic energy band positions of a metal oxide (NaTaO3) and metal (oxy)nitride (BaTaO2N). Reprinted with permission from Reference K. Maeda and K. Domen, J. Phys. Chem. C, 2007, 111, 7851. Copyright 2007 American Chemical Society [78].
Figure 3. Schematic representation of electronic energy band positions of a metal oxide (NaTaO3) and metal (oxy)nitride (BaTaO2N). Reprinted with permission from Reference K. Maeda and K. Domen, J. Phys. Chem. C, 2007, 111, 7851. Copyright 2007 American Chemical Society [78].
Catalysts 11 00160 g003
Figure 4. FE-SEM photographs of obtained BaNbO2N particles (a) before and after flux treatment with (b) LiCl, (c) NaCl, (d) KCl, (e) RbCl, and (f) CsCl. Scale bars represent 1 μm [83].
Figure 4. FE-SEM photographs of obtained BaNbO2N particles (a) before and after flux treatment with (b) LiCl, (c) NaCl, (d) KCl, (e) RbCl, and (f) CsCl. Scale bars represent 1 μm [83].
Catalysts 11 00160 g004
Figure 5. (A) Photocurrent-time plots in aq. 0.5 M Na2SO4 (pH ≈ 6) for SrNbO2N photoelectrodes (6 cm2) with and without colloidal IrO2 at +1.55 VRHE under visible-light conditions (λ > 420 nm). (B) and its equivalent time progresses of gas evolution in PEC reactions by an IrO2-loaded electrode. Reprinted with permission from Reference K. Maeda, M. Higashi et al. J. Am. Chem. Soc., 2011, 133, 12334. Copyright 2018 American Chemical Society [118].
Figure 5. (A) Photocurrent-time plots in aq. 0.5 M Na2SO4 (pH ≈ 6) for SrNbO2N photoelectrodes (6 cm2) with and without colloidal IrO2 at +1.55 VRHE under visible-light conditions (λ > 420 nm). (B) and its equivalent time progresses of gas evolution in PEC reactions by an IrO2-loaded electrode. Reprinted with permission from Reference K. Maeda, M. Higashi et al. J. Am. Chem. Soc., 2011, 133, 12334. Copyright 2018 American Chemical Society [118].
Catalysts 11 00160 g005
Figure 6. (a) UV-Vis absorption analysis of NTs, air annealed TiO2 NTs, and Ag incorporated TiON NTs (Ag/TiON), (b) LSV plots under the irradiation of TiO2, TiON and Ag/TiON, (c) the IPCE spectra of the samples, and (d) the IPCE of Ag/TiON electrodes under externally applied bias [134].
Figure 6. (a) UV-Vis absorption analysis of NTs, air annealed TiO2 NTs, and Ag incorporated TiON NTs (Ag/TiON), (b) LSV plots under the irradiation of TiO2, TiON and Ag/TiON, (c) the IPCE spectra of the samples, and (d) the IPCE of Ag/TiON electrodes under externally applied bias [134].
Catalysts 11 00160 g006
Figure 7. Schematic representation of the oxysulfide system employed for visible-light-driven Z-scheme water splitting into H2 evolution in conjunction with Cs loaded WO3 photocatalysts. Reprinted with permission from Reference G. Ma, S. Chen et al. J. Phys. Chem. Lett, 2016, 7, 3892. Copyright 2016 American Chemical Society [154].
Figure 7. Schematic representation of the oxysulfide system employed for visible-light-driven Z-scheme water splitting into H2 evolution in conjunction with Cs loaded WO3 photocatalysts. Reprinted with permission from Reference G. Ma, S. Chen et al. J. Phys. Chem. Lett, 2016, 7, 3892. Copyright 2016 American Chemical Society [154].
Catalysts 11 00160 g007
Figure 8. Schematic view of p-n junction photovoltaic cell (a), PEC (b), and photocatalyst (c). There are five common processes, while the final methods of carrier collections are different. Reprinted with permission from Kobunshi Ronbunshu., vol. 70, page 461. Copyright 2013, The Society of Polymer Science, Japan [202].
Figure 8. Schematic view of p-n junction photovoltaic cell (a), PEC (b), and photocatalyst (c). There are five common processes, while the final methods of carrier collections are different. Reprinted with permission from Kobunshi Ronbunshu., vol. 70, page 461. Copyright 2013, The Society of Polymer Science, Japan [202].
Catalysts 11 00160 g008
Figure 9. Schematic representation of visible-light-assisted H2 generation in the organic ITO/H2Pc/C60-Pt system. H2 evolves at Pt-loaded C60, after a 20 min period during which C602− is formed. Reprinted with permission from The Journal of Physical Chemistry C., vol. 115, page 7704. Copyright 2011, American Chemical Society [45].
Figure 9. Schematic representation of visible-light-assisted H2 generation in the organic ITO/H2Pc/C60-Pt system. H2 evolves at Pt-loaded C60, after a 20 min period during which C602− is formed. Reprinted with permission from The Journal of Physical Chemistry C., vol. 115, page 7704. Copyright 2011, American Chemical Society [45].
Catalysts 11 00160 g009
Figure 10. Twin-compartment cell utilized for one active electrode photocatalysis and the structures of the chemicals. Reproduced from Ref. [219] with permission from the Royal Society of Chemistry [219].
Figure 10. Twin-compartment cell utilized for one active electrode photocatalysis and the structures of the chemicals. Reproduced from Ref. [219] with permission from the Royal Society of Chemistry [219].
Catalysts 11 00160 g010
Figure 11. A schematic representation of the water-splitting system comprising an organo-photocathode and WO3 photoanodes. Reproduced from Ref. [220] with permission from the Royal Society of Chemistry [220].
Figure 11. A schematic representation of the water-splitting system comprising an organo-photocathode and WO3 photoanodes. Reproduced from Ref. [220] with permission from the Royal Society of Chemistry [220].
Catalysts 11 00160 g011
Table 1. Review on different kinds of (Oxy)nitrides based photoelectrodes with co-catalysts and their photoelectrochemical features for water splitting reactions.
Table 1. Review on different kinds of (Oxy)nitrides based photoelectrodes with co-catalysts and their photoelectrochemical features for water splitting reactions.
ElectrodesCocatalystsSynthesis MethodPhotocurrent(j)Efficiency
IPCE
(@1.23 V vs. RHE)
StabilityReference
LaTiOxNyIrO2magnetron sputtering70 µA cm−2 at +1.0 V vs.
Ag/AgCl
-30 min[90]
LaTiO2NIn2O3Electrophoretic deposition (EPD)0.61 mA cm−2 at 1.23 VRHE,3.2% at 400 nm30 min[135]
SrNbO2NIrO2Polymer Complex(PC)/EPD0.35 mA cm−2 at +1.2 V vs.
RHE
0.2%2h[118]
TaONCoOx/TiO2screen-printing0.7 mA cm−2 at +1.2 V vs.
RHE
27% at with λ = 410 nm15 h[136]
BaTaO2NCoOx/RhOxEPD-10% at 600 nm60 min[137]
BaNbO2NFeOx/Co(OH)xparticle transfer (PT)5.2 mA cm−2 at 1.23 VRHE,38% at with 460 nm« 30 min[138]
SrNbO2NCoOxmagnetron
sputtering
4.1 mA cm−2 at +1.2 V vs.
RHE
10% λ = 400 nm--[139]
BaTaO2NCoPT4.2 mA cm−2 at 1.2 VRHE24% with λ = 400 nm6 h[140]
SrNbO2NCoPiflux treatment/PT method1.5 mA cm−2 at 1.23 VRHE-2 h[141]
BaNbO2N-Flux treatment/PT5 mA cm−2 at 1.23 VRHE--[142]
CaNbO2NCo-Pi/Al2O3solid state/PC70 μA cm−2 at 1.23 VRHE-30 min[109]
CaTaO2N-ANiBEPD40 μA cm−2 at 1.23 V versus RHE--[143]
SrNbO2N-NRs-PT1.3 mA cm−2 at 1.23 VRHE,-15 min[144]
SrNbO2NCo3O4EPD2.0 mA cm−2 at 1.23 VRHE12% with 420 nm30 min[145]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Arunachalam, P.; Nagai, K.; Amer, M.S.; Ghanem, M.A.; Ramalingam, R.J.; Al-Mayouf, A.M. Recent Developments in the Use of Heterogeneous Semiconductor Photocatalyst Based Materials for a Visible-Light-Induced Water-Splitting System—A Brief Review. Catalysts 2021, 11, 160. https://doi.org/10.3390/catal11020160

AMA Style

Arunachalam P, Nagai K, Amer MS, Ghanem MA, Ramalingam RJ, Al-Mayouf AM. Recent Developments in the Use of Heterogeneous Semiconductor Photocatalyst Based Materials for a Visible-Light-Induced Water-Splitting System—A Brief Review. Catalysts. 2021; 11(2):160. https://doi.org/10.3390/catal11020160

Chicago/Turabian Style

Arunachalam, Prabhakarn, Keiji Nagai, Mabrook S. Amer, Mohamed A. Ghanem, Rajabathar Jothi Ramalingam, and Abdullah M. Al-Mayouf. 2021. "Recent Developments in the Use of Heterogeneous Semiconductor Photocatalyst Based Materials for a Visible-Light-Induced Water-Splitting System—A Brief Review" Catalysts 11, no. 2: 160. https://doi.org/10.3390/catal11020160

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