Next Article in Journal
Indoor Building Reconstruction from Occluded Point Clouds Using Graph-Cut and Ray-Tracing
Next Article in Special Issue
Triple Planar Heterojunction of SnO2/WO3/BiVO4 with Enhanced Photoelectrochemical Performance under Front Illumination
Previous Article in Journal
Optical Assembling of Micro-Particles at a Glass–Water Interface with Diffraction Patterns Caused by the Limited Aperture of Objective
Previous Article in Special Issue
Conformal Titanyl Phosphate Surface Passivation for Enhancing Photocatalytic Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ferroelectric Materials: A Novel Pathway for Efficient Solar Water Splitting

Department of Electrical Engineering, Gachon University, Seongnam 13120, Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2018, 8(9), 1526; https://doi.org/10.3390/app8091526
Submission received: 27 July 2018 / Revised: 24 August 2018 / Accepted: 25 August 2018 / Published: 1 September 2018
(This article belongs to the Special Issue Nanomaterials for Solar Water Splitting)

Abstract

:
Over the past few decades, solar water splitting has evolved into one of the most promising techniques for harvesting hydrogen using solar energy. Despite the high potential of this process for hydrogen production, many research groups have encountered significant challenges in the quest to achieve a high solar-to-hydrogen conversion efficiency. Recently, ferroelectric materials have attracted much attention as promising candidate materials for water splitting. These materials are among the best candidates for achieving water oxidation using solar energy. Moreover, their characteristics are changeable by atom substitute doping or the fabrication of a new complex structure. In this review, we describe solar water splitting technology via the solar-to-hydrogen conversion process. We will examine the challenges associated with this technology whereby ferroelectric materials are exploited to achieve a high solar-to-hydrogen conversion efficiency.

1. Introduction

In order to effectively address energy-related problems such as global warming, which is caused by the emissions of air pollutants including carbon dioxide (CO2), there is a growing interest in many kinds of renewable energy (such as solar, wind power, hydroelectric energy, and biomass), and their related applications as an alternative to fossil fuels [1,2].
Among the various alternative energy sources, solar energy is an almost unlimited and clean energy source, given that the Sun has been providing energy in the form of light and heat to the Earth for more than four billion years. Until now, solar energy has been considered as an alternative energy option that could help address several problems of global importance such as an insufficient energy supply, environmental pollution, and fossil fuel resource depletion. Moreover, most regions of our planet including the atmosphere, oceans, and soil receive approximately 3850 zetta joules (ZJ = 1021 J) of energy every year (~0.539 ZJ of energy was consumed worldwide in 2010). Considering only the amount of energy received from the Sun, solar energy is one of best candidates as a substitution for fossil fuels. Even though many researchers have been actively involved in alternative energy research and related technologies have been developed over the years, no more than 1% of the energy produced by the Sun is used by our planet. Therefore, we need to concentrate on the efficiently of solar energy use, whether indirectly or directly [3,4]. In particular, for a solar energy carrier, three main points have to be considered. (1) How much cheaper and abundant are chemical reactants? (2) Is the carrier easily stored and transported? (3) In the process, is carbon dioxide generated and emitted?
On the other hand, hydrogen energy and related applications are considered to be excellent alternative renewable energy carriers due to their storability, transportability, and convertibility to energy on demand. Hydrogen gas is a clean, nontoxic, colorless, and eco-friendly source (does not release carbon dioxide) that could be obtained by splitting naturally occurring water into hydrogen and oxygen. Moreover, hydrogen could be used in various energy applications such as hydrogen fuel cells (HFCs), combustible fuel, and synthetic natural materials. It is known that when the gas is used for electricity production such as in the case of hydrogen-powered fuel cells, it has three times the efficiency of electricity generation (~65%) compared to conventional combustion-based power plants (~35%) [5,6,7].
In order for hydrogen to be classified as an environmentally friendly alternative energy source, it should be produced using appropriate processes that use carbon-free environmentally friendly materials instead of fossil fuels. Compared with other methods of hydrogen production (biomass, thermal decomposition of natural gas, water electrolysis, and thermochemical cycles), water-splitting technology has been proposed as the most suitable method for the eco-friendly generation of hydrogen gas from water. It should be noted that solar energy has been used as an energy conversion source of solar thermal and photovoltaics starting from solar light. In particular, water splitting using photovoltaics based on solar energy is ideal for the generation of hydrogen gas from water [8,9].
This review will address solar water-splitting technology (such as water splitting in photosynthesis, photoelectrochemical water splitting, and photocatalytic water splitting), including the systems involved and hydrogen production from water and solar energy. Moreover, we will introduce the recent progress in solar water splitting using ferroelectric materials, and summarize how each material plays an important role in achieving efficient solar water splitting.

2. Overview of Solar Water Splitting

In a solar water-splitting system, hydrogen is produced from the semiconductor–electrolyte interface where sunlight is absorbed in the depletion layer and electron–hole pairs are separated to drive the chemical reaction at the semiconductor–electrolyte interface, Light energy is used to directly dissociate water molecules into hydrogen and oxygen through the following steps: light absorbance, conversion of light to exciton (electron–hole pair), and chemical reaction (separated hydrogen). Water splitting via solar energy is achieved through three kinds of hydrogen production systems [8,9,10,11,12]: (1) particulate photocatalysis (PC) systems, (2) photoelectrochemical (PEC) systems, and (3) photovoltaic–photoelectrochemical (PV–PEC) systems, as shown in Figure 1. In this section, we describe in more detail the mechanism for solar energy conversion to hydrogen fuel, and the as-mentioned three approaches for hydrogen generation are introduced.

2.1. Basic Mechanism for Solar Energy to Hydrogen Conversion System

The concept of photocatalytic water splitting (PWS) was introduced by hydrogen production via photoelectrochemical water splitting with a rutile TiO2 (anode) as the photocatalyst, and a platinum (cathode). This was first reported by Fujishima and Honda in 1972 [13]. Hydrogen could be generated directly from water, and solar light could be generated through the solar energy conversion process.
In PWS systems, photocatalysts play an important role in the conversion of solar energy to hydrogen in response to visible light in the water-splitting system. The hydrogen gas production process involves a photocatalytic reaction such as the direct conversion of solar energy into hydrogen gas in a semiconductor with an energy bandgap that is positioned between their energy band structures.
A photocatalytic water-splitting reaction via semiconductor-based photocatalysts is conducted in terms of charge carrier generation, separation, transport, and transfer. This refers to the following three-step process [10,11,12,13,14].
Step 1: Light (photons) is absorbed near the surface of the semiconducting materials with energies greater than their bandgap energy (e.g., ~1.7 eV). Excited electrons (excitons) and holes are generated inside the semiconducting materials by bandgap excitation.
Step 2: The electrons and holes henceforth referred to as photogenerated electrons and holes are separated in the bandgap and drift to the surface of the semiconducting materials, causing the release of oxygen from water (catalytic water oxidation reaction).
Step 3: The photogenerated electrons combine with H+ via the metal electrode (catalytic water reduction reaction).
Figure 2 shows the principle of photocatalytic water-splitting reactions with photoelectrons and holes that are generated by the absorption of light. The photoelectrons and holes reduce and oxidize water on the surface of the semiconductors as the two half-reactions of water splitting via the following reactions [10,11,12].
Applsci 08 01526 i001
The water-splitting reaction is thermodynamically nonspontaneous, and therefore is an uphill reaction. As such, the net Gibbs free energy (ΔG ≈ 237 kJ/mol) needs to be increased. Namely, if a photocatalyst has a bandgap that is larger than the energy required for water splitting (~1.23 eV), then its conduction and valence band edge should contain the oxidation and reduction potential of water [8].
It is clear that the O2/H2O redox potential difference is 1.23 V. Therefore, the requirement of a bandgap is larger than the required energy for water splitting (1.23 eV). For more details, the valence band potential must be more positive than the O2/H2O redox potential of 1.23 V versus Normal hydrogen electrode (NHE, pH = 0) to permit water oxidation, and the conduction band must be more negative than the H+/H2 redox potential of 0 V versus NHE to facilitate water reduction [8,9,10].

2.2. Solar Energy to Hydrogen Conversion System for Solar Water Splitting

Among the previously mentioned hydrogen production systems, i.e., the PC, PEC, and PV–PEC, the PC offers reasonable solar-to-hydrogen efficiency, a low process cost, simplicity of design, and a large-scale method for water splitting. Moreover, the PC reaction could be performed in the homogeneous phase without transparent electrodes and directional illumination. However, a PC system should be inserted in a gas separator because both hydrogen and oxygen gases are produced during the water-splitting process. Therefore, an enclosed reaction system is required on a large-scale [10,14,15,16]. However, the PEC systems do not require gas separation due to their structure, which entails one or two conductive electrodes and a small bias. The two different gases are sequentially generated and remain at the opposite electrodes.
A PEC cell system is composed of an electrolyte, an n-type semiconductor (anode), and a p-type semiconductor (cathode) with a semiconductor/liquid junction, as shown in Figure 3. Photogenerated charge carriers are separated, and minor carriers (holes and electrons originated from p-/n-type semiconductor electrodes) move within the semiconductor–liquid interface. This allows for the separate release of oxygen and hydrogen gases during the water-splitting reaction.
In electrode–liquid electrolyte systems, the photovoltaic and photoelectrochemical reaction behavior is mainly conducted in three steps: (1) light absorption, (2) water reduction and oxidation, and (3) the recombination of electrons and holes. When the system is exposed to photons with sufficiently high light energy, they are induced to form electron–hole pairs, resulting in chemical reactions such as oxidation-reduction reactions. During these reactions, photogenerated currents and voltages are observed inside the system. Such electrical phenomena can be observed in both photovoltaic and photoelectrochemical systems. The photovoltaic (PV) process is so-called because the conversion of light into electricity occurs in a photoelectrode through a process of conversion photons. In the case of a photoelectrochemical process, the generated photons induce a chemical reaction in the electrodes. Since both photovoltaic and photoelectrochemical processes are related to light absorbance and the conversion of photons to induce chemicals reactions, various semiconducting materials can be used to convert solar energy to hydrogen. They have been extensively used in the past for solar water splitting [17,18].
Until now, several different concepts have been proposed and reported from different research groups for solar water splitting. Figure 3 shows a schematic diagram of an enhanced PEC and PV–PEC hybrid water splitting system, which was proposed to improve solar energy to hydrogen efficiency. The PV/PEC hybrid cell consists of a photoelectrode PEC cell and a PV cell. The PV cells play an important role in supplying a potential to the PV cell. The PV cell assists the PEC cell to develop a potential for photogenerated electrons, allowing water reduction at the surface of the photocathode [19,20].

2.3. Strategy for High Solar-to-Hydrogen (STH) Efficiencies

The energy conversion efficiency in solar water-splitting cells is usually evaluated by various methods such as standard solar-to-hydrogen (STH) conversion efficiency ( η S T H ), quantum efficiencies such as the incident photon-to-current efficiency (IPCE), photon-to-current efficiency (ABPE), and the absorbed photon-to-current conversion efficiency (APCE) [21,22]. Among these, we will further examine the standard η S T H and ABPE in greater detail.
The standard η S T H is the ratio between the total energy generated and the total energy input in the sunlight irradiation of AM 1.5 G (air mass at a solar zenith angle of 48.2°), which can be calculated using following the equation:
  η S T H   = T o t a l   E n e r g y   g e n e r a t e d T o t a l   e n e r y   i n p u t = Δ G × r H 2 P s u n × S
where ΔG is the Gibbs free energy (237 kJ/mol), r H is the rate of hydrogen production in moles per second, Psun is the incident light intensity (100 mW/cm2) and S is the illuminated area of the photoelectrode (cm2).
The ABPE can be calculated from the J–V curve of the photoelectrode while an external bias is applied to the two electrodes.
  ABPE   ( % )   = [ J   ( m A / c m 2 )   × ( 1.23 V b i a s ( V ) P i n   ( m A / c m 2 ) ] A M   1.5 G × 100   %

2.4. Strategy to Improve High STH Efficiency for Water Splitting

Solar water splitting with high STH efficiencies has been developed over the last few years. Recently, Jia et al. [23] reported on solar water splitting by photovoltaic–electrolysis STH with an efficiency over 30% for two days using two polymer electrolyte membrane electrolyzers in series with one InGaP/GaAs/GaInNAs(Sb) triple junction solar cell.
To achieve high STH efficiency for solar water splitting, we should carefully select the semiconducting materials. The bandgap of the semiconducting materials, including the electrode, is very important in the solar water-splitting process. Most photons cannot be absorbed by the semiconductor if they have energy that is lower than the bandgap energy. Therefore, the bandgap of the selected semiconducting materials must be chosen in the range of at least ~1.23–2.0 eV to absorb light in the visible range, which facilitates a high photocatalytic effect. The semiconducting materials used for the cells should have strong catalytic activity and stability regardless of the oxidation/reduction reaction. With regard to the process, the semiconductor is easily oxidized and reduced by excited electrons and holes. Since their photoreproductive capability decreases, the cell’s STH efficiency also decreases. Additionally, the materials used for cells should be easily modified and readily available at low cost [23,24,25].

3. Solar Energy Conversion of Ferroelectric Materials

3.1. Ferroelectric Effect and Materials

The basic concept of ferroelectricity was introduced by Rochelle salt in 1920: if an external electric field is applied to a ferroelectric material, the dipoles in the crystalline or polycrystalline structure are induced to produce a ferroelectric phenomenon with spontaneous polarization and alignment with the external field. Even when the electric field is switched off, the material maintains a spontaneous polarization. This is called the ferroelectric phenomenon. Furthermore, the spontaneous polarization is reversed by the application of an electric field. Ferroelectrics exhibit a phase transition phenomenon. Namely, the polarization of the dipoles exhibits interactive electric dipole moments with and without an applied electric field. When the ferroelectric is at a higher temperature than the phase transition temperature, the spontaneous polarization is lost due to thermal fluctuations.
When the spontaneous polarization is reversed, it causes the ferroelectrics to exhibit the property of piezoelectricity. Ferroelectric materials can be categorized into two types: the order–disorder type and the displacement type, based on the physical mechanism used to generate the ferroelectric characteristics.
Ferroelectric materials, which are mostly used as condenser and as semiconductor elements, are a cubic system with a perovskite structure. When the crystal structure of the perovskite exceeds the Curie temperature Tc (~120 °C), similar to a ferroelectric crystal of the ABO3 perovskite structure, it changes its form to a paraelectric phase in which the crystal structure is the cubic symmetry [26,27].

3.2. Ferroelectric Photovoltaics—Phenomenology of Ferroelectric Solar Cells

The photovoltaic effect in a semiconducting material occurs when an asymmetry in the electric potential occurs throughout the materials due to a flow of photogenerated electrons and holes. A photocurrent occurs as a result of the absorption of photons with high energy, which leads to the transfer of electrons from the valence band (VB) to the conduction band (CB).
Conventional photovoltaic phenomenon could be observed in interfaced doped semiconductors such the p–n junction structure when the following three processes should be involved: high light absorption, generation of the electron–hole pairs, and separation of the free charges.
Usually, hole (positive charge) and electron (negative charge) carriers are generated by high energy photons. Over time, the excited carriers return to the ground-state level, and the internally absorbed energies are released as light or heat. While most commercial solar cells are based on the p–n junction principle between two different metallic contacts, they are limited by the so-called Shockley–Queisser limit, which prevents any single p–n junction solar cell from converting more than 33.7% of the incident light [23]. Their heterojunction structure is limited with regard to the selection of potential materials due to dopant issues and the mismatched lattice of the structure.
Ferroelectric materials have been recently proposed as excellent potential candidates to solve these problems. The phenomenon of a photovoltaic effect in a ferroelectric has been observed in BaTiO3, LiNbO3, and Pb(Zr,Ti)O3 with weak pyroelectric currents (photovoltaic current above Tc). However, ferroelectrics have a significant disadvantage of large band gaps in the range of ~3–4 eV, which is challenging for applications as photovoltaic materials. Their large band gaps values result in the degradation of efficient conversion because small current densities and large open circuit photovoltages are typical. In recent years, the higher efficiency of narrower bandgap ferroelectrics in ferroelectric solar cells such as those of BiFeO3 (BFO, 2.2–2.7 eV) and multiferroic Bi12FeCrO6 (BFCO, 1.9–2.1 eV), has been reported.
Ferroelectric photovoltaic effects are divided into photoferroic phenomena including the bulk photovoltaic effect (BPE) and the anomalous photovoltaic effect (APE). Photoferroic phenomena could be demonstrated in hybrid organic–inorganic halide perovskites such as CH3NH3PbI3 (MAPI) to fabricate high-efficiency photovoltaic devices. Ferroelectric domains as internal junctions can induce the generation of photoexcited electrons and hole pairs, and decrease recombination by dividing charged carriers. The effect could produce a higher open circuit voltage (VOC) and the current−voltage hysteresis observed in perovskite solar cells. The photoferroic system has an intricate interface between the photoresponse and the ferroelectric phase stability, single/multidomain, and distributions. The photoferroic system has recently achieved power efficacies in excess of 8% power for Bi2FeCrO6 [26]. The larger bandgap (~3 eV) in the polar materials shows their potential for development as light absorbers [25,26,27].
The photovoltaic effect (BPE) is a phenomenon that is observed in an open circuit voltage under dark illumination. The BPE generates a photovoltaic current that results in an anomalous photovoltaic effect (APE), as a type of bulk photovoltaic effect that occurs between semiconductors and insulators as a result of an open-circuit voltage induced by light. Sometimes, the generated voltage approaches almost thousands of volts.
The mechanisms of the anomalous photovoltaic effect (APE) are schematically represented in Figure 4. The generated photovoltage in the anomalous photovoltaic effect (APE) is limited by its bandgap. This results in a photoresponse of the heterojunctions, the photo-Dember effect, and grain boundaries. As shown in Figure 4, its mechanisms could be explained by three categories as follows: (a) the photo-Dember effect, (b) p–n homojunction domains, (c) and ferroelectric domains [27,28].

4. Ferroelectric Materials for Photoelectrochemical (PEC) Water-Splitting Devices

4.1. General Requirements for Photoelectrode Materials

The choice of appropriate materials for photoelectrodes plays a decisive role in the PEC activity for water splitting. Conventional materials, including semiconductors, can be categorized into wide-bandgap and narrow band-gap materials. Traditionally, wide-bandgap materials (TiO2, SrTiO3, BaTiO3, and WO3) are considerably stable in an electrolyte solution, but they can have the limited absorbance range in the ultraviolet region. Even though narrow bandgap materials (Cu2O, InAs, and CuO) can absorb much more visible light irradiation in the solar spectrum, they have a high possibility of being damaged from strong electrolytic environments. However, we do not have many options for selecting suitable materials for photoelectrodes (photocathodes and photoanodes) of PEC. For photocatalytic water splitting, the selected materials must have bandgaps of approximately 2 eV (at least 1.23 eV) as a result of thermodynamic energy and kinetic loss. Figure 5 shows the bandgap of various materials. Therefore, it is imperative that the following general requirements are met prior to the fabrication of high-performance PECs [12,29,30].
(i) Suitable bandgap energy and band edge positions. The electrolysis of water is composed of two reactions occurring at different electrodes. There is a reduction reaction generating hydrogen with H+/H2 redox potential of 0 V vs. NHE at the photocathode, whilst the O2/H2O redox potential of 1.23 V vs. NHE (pH = 0) in an oxidation reaction at the photoanode; thus, the bandgap of the materials needs to be larger than 1.23 V in order to be able to occur water split. In addition, the thermodynamic energy losses (0.3–0.4 V) and an overpotential (0.4–0.6 V) should also be considered. Consequently, a minimum bandgap of ~1.8 eV is required. Furthermore, visible light is very limited below 390 nm, and this limits the upper value of the bandgap energy to 3.2 eV. Therefore, materials with a bandgap energy in the range between 1.9–3.2 eV are promising candidates for the generation of large photovoltages. When a photoelectrode material that meets these requirements with respect to the bandgap value absorbs incoming photons, electrons are excited to the conduction band (CB) and leave holes in the valence band (VB). The CB edge must be more negative than the H+/H2 redox potential of 0 V versus normal hydrogen electrodes (NHEs) (pH = 0) to induce H2 production, while the VB edge value must be more positive than the O2/H2O redox potential of 1.23 V versus NHE to generate water oxidation.
(ii) Efficient charge separation and transfer. Low efficiency in PEC devices is mostly attributed to charge recombination and inefficient carrier transportation. As a result, many studies have been performed with the objective of addressing these challenges to achieve enhanced performance on PEC devices. Both the intrinsic properties (hole and electron mobility) and the extrinsic properties (crystallinity, nanostructure) affect the separation and transfer of the photogenerated carriers.
(iii) Robust catalytic activity and stability. Generally, there are energy barriers for the transfer of electrons or holes, and these barriers exhibit energy losses in the reaction. Appropriately rapid surface reaction kinetics can overcome these barriers and suppress electron–hole recombination. In PEC systems, the hydrogen evolution reaction (HER) generates H2 from H+ (in acid) or H2O (in base), whereas oxygen evolution reaction produces O2 from H2O (in acidic) and OH (in base) [31,32]. Photocorrosion is a major problem for long-term PEC cells because of water splitting, which leads to the decomposition and fluctuation of photocurrent in a short time interval, especially in harsh conditions (strongly acidic or alkaline). However, a high ionic environment (low or high pH) is essential to minimize other challenges such as ohmic losses and localized pH gradient overpotentials in the PEC systems of separate photoelectrodes [33,34]. Therefore, ferroelectric materials with intrinsically stable properties even in harsh electrolyte are potential candidates for the generation of new PEC systems. Furthermore, the suitable alignment between the band edge potentials of the candidate materials and the relative decomposition potentials can promote stable PEC devices. Lastly, cost-effective materials along with earth-abundant elements are also required for a practical approach.

4.2. Ferroelectric-Based Materials in Photocathodes for Hydrogen

Promising photocathodes are generally p-type semiconductors and must meet some of the aforementioned requirements to generate the required cathodic current for water reduction. Single metallic oxides such as Cu2O, CuO, and NiO are at the forefront of the p-type semiconductors that have been investigated as photocathodes for solar water splitting. However, these materials are hindered by some major shortcomings such as photocorrosion [35,36], low conduction band position [37,38,39] and wide bandgap (3.6–4.0 eV), along with the low potential of the valence band [40,41]. As a result, the PEC devices fabricated from conventional single semiconductors exhibit a limited photocurrent and low stability, thereby reducing their PEC performance. Therefore, the possibility of incorporating ferroelectrics into conventional oxide-based photocathodes is a novel and facile way to enhance their PEC activity. In this section, the review is focused on ferroelectric-based materials for photocathodes, which include ferroelectric ternary/quaternary metal oxides.

4.2.1. Ferroelectric Oxide Perovskites

Ternary Metal Oxides
(1) BiFeO3
BiFeO3 (BFO) is a well-known multiferroic material with a bandgap in the range of 2.2–2.7 eV, which has been researched extensively because it possesses a wide range of intriguing properties [42,43,44,45,46,47,48,49,50,51]. Many studies have been reported on the significant effect of polarization on band bending in BFO, which plays a crucial role in the photocatalytic activity for PEC water splitting. This enhanced charge separation is attributed to the intimate relation between band bending and the separation of photoexcited electron–hole pairs in the space charge region [52]. Depending on the fabrication method, BFO can exhibit properties of either an n or p-type semiconducting material, thereby playing a role as either an anodic or cathodic electrode in PEC water splitting [42,45,47,51].
Although BFO presents robust ferroelectricity and has attracted significant interest from researchers involved in PEC water-splitting research, there are two challenges that limit the tremendous potential of BFO as a photoelectrode material. The first issue is the low photocathodic current due to the rapid recombination of electron–hole pairs, which is required to achieve a better PEC performance. Another challenge is the rapid degradation during the PEC reaction [53,54]. Undoubtedly, the noble metal Pt is by far the best-performing photocathode for H2 evolution to date [55]. Theoretically, a combination of BFO and a noble metal such as Pt is considered a good choice to facilitate H2 generation reaction. This can be explained by the reduced charge recombination and enhanced electric field that is attributed to the Schottky barrier due to the different work functions between BFO and Pt as well as the localized surface plasmon resonance (LSPR) effect of these noble metals [56,57]. However, contact between Pt and BFO impedes the transfer of photoexcited electrons from an electrode surface to the electrolyte due to the Schottky upward barrier. Fortunately, the injection of a buffer layer between Pt and BFO has been discovered to overcome this shortcoming.
Gu et al. [53] inserted a porous carbon layer sandwiched between Pt and BFO to obtain a ITO/BFO/carbon/Pt photocathode in order to inhibit the formation of a BFO/Pt Schottky upward barrier. As a result, drastically enhanced photocathodic performance is achieved with a photocurrent density (Jo) and onset potential (Vop) up to −235.4 µA/cm2 and 1.19 V versus reversible hydrogen electrode (RHE), respectively, while the values for a photocathode without a carbon interlayer are only −61.6 µA/cm2 and 0.83 V versus RHE, respectively. More importantly, the BFO films were macroscopically poled to investigate the effect of the ferroelectric polarization on the PEC performance, and the results are reported in Figure 6. After poling, the Jo and Vop values increased significantly, which is attributed to the effective separation of the photogenerated carriers in the BFO films, thereby enhancing the PEC activity of the photocathodes.
The novel configuration with carbon resulted in a reduced charge recombination and the facilitation of charge transfer by the removal of the upward barrier between Pt and BFO. Based on this result, an amorphous TiO2 was chosen as a buffer layer to insert between BFO and Pt with the similar desire for the Schottky barrier [58]. As expected, the PEC yielded a Jo of −460 µA/cm2 at 0 V versus RHE and Vop of 1.25 V versus RHE under 100 mW/cm2 Xe-lamp illumination. More importantly, the role of TiO2 in stabilizing the photocathode is apparent with a 10-h continuous effective PEC reaction in an acid solution. Meanwhile, most of the BFO photocathodic performance is degradable within only 2 h of continuous reaction, even when measured using the neutral Na2SO4 electrolyte [53,54]. The limitation of the performance of PEC cells used BFO as photocathodes could be overcome by an interlayer TiO2 as protection layer to avoid direct contact with its electrolyte. This interpretation could show a promising pathway to reduce cathodic photocorrosion at the semiconductor/electrolyte interface by employing a variety of protection layers on the photocathodes. To utilize a wide range of the visible light spectrum, tandem PEC systems composed of both photoanodes and photocathodes have been investigated in many recent studies [59,60]. Among many tandem architectures, the photocathodes based on the crystal Si-pn+ were constructed with an improvement in PEC activity [61,62]. However, the efficiency is far from the desired value because of the remaining photoexcited electrons inside the holes of Si. In a recent report, Cheng et al. [54] overcame his shortcoming to fabricate a Si-pnn+/ITO/Au/BFO hybrid photocathode with enhanced photocurrent and onset potential by utilizing the local surface plasmon resonance (LSPR) effect of Au nanoparticles and the depolarization electric field of BFO.
(2) LaFeO3
LaFeO3 (LFO) is an n-type semiconducting perovskite oxide that has emerged as an attractive candidate for photoelectrochemical water splitting because of its many advantages, such as high stability under illumination in an aqueous environment, and abundant quantity of rare-earth element oxides [59,60]. The bandgap of LFO is estimated to be from 2.0 eV to 2.6 eV, which is small enough to utilize a portion of the visible spectrum light [39].
Although LFO has stood out as a promising perovskite oxide, it still has some drawbacks that limit its use as an effective photocathode for hydrogen production [63,64,65,66,67]. The LFO photocathode in PEC shows low photoresponse characteristics because of the lack of a sophisticated preparative technique [68]. The p-type LFO film, which is prepared by more sophisticated fabrication techniques such as pulsed laser deposition (PLD) and atomic layer deposition (ALD) as a photocathode coupling with an n-type Fe2O3 photoanode for stable water splitting, has assisted in the suppression of the aforementioned limitation [63,69]. High-quality LFO films and good back contacts resulted in the generation of a photocurrent density at 0 V versus RHE of 64.5 µA/cm2 under AM 1.5 G irradiation for 100 nm of LFO, while the reaction time was maintained for 120 h without any obvious decline of the rate of oxygen and hydrogen production [63]. In addition, under external bias, the Fe2O3-LaFeO3 system shows outstanding performance in both the development of the gas evolution rates and stability, even in alkaline conditions, in comparison with an Fe2O3-Pt system. Furthermore, the modified LFO photocathode with high performance, especially its long-term stability and low cost, could be considered one of the promising candidates to replace the normal Pt photocathode in PEC systems.
Although the low photoresponse problem was solved by PLD instead of the sol-gel spin-coating method, the value of the photocurrent onset potential was ~1.0 V versus RHE, which is lower than the standard potential of 1.23 V for water electrolysis. Recent efforts to enhance the photoelectrochemical behavior of LFO has yielded many positive results. Díez-Garcia et al. [70] synthesized metal-doped LFO thin film electrodes using the sol-gel method, resulting in a significant improvement in efficiency, with an onset potential as high as 1.4 V versus RHE. The enhancement of the efficiency in PCE devices is attributed to an increase of both majority carrier density and mobility by doping with a relatively large amount of Mg2+ and Zn2+ (5% with respect to the iron atoms).
In addition, it is reported that metal doping is an effective strategy for enhancing the p-type nature of the ternary oxides [71]. In particular, LFOs have been shown to exhibit both cathodic and anodic photocurrents depending on the specific conditions [72]. Although sophisticated fabrication techniques such as PLD can improve the PEC response of LFO photocathodes, the high cost and challenging technical requirements are still primary impediments. Recently, novel research in the preparation of nanostructured LFO thin films using an inexpensive spray-pyrolysis method resulted in outstanding results [73]. The fabricated photoelectrode yielded a photocurrent density of 0.16 mA/cm2 at 0.26 V versus RHE, which is by far a more drastic enhancement compared to the PLD technique [63].
Interestingly, this is the first time that spontaneous hydrogen generation has been achieved using PEC water splitting without any external bias voltage using LFO as a single photoelectrode material. Admittedly, LFO is a potential candidate for a photocathode because of its stable characteristics in aqueous solutions, as well as its very positive onset potential values. However, further research should be performed to improve the slow kinetics of charge transfer, which favors recombination at surface trap states.
(3) SrTiO3
Among the various ternary metal oxides, SrTiO3 (STO), which is a stable n-type semiconductor with a bandgap value of 3.25 eV [74], has been utilized as a promising photoelectrode for splitting water into hydrogen and oxygen, with many outstanding properties. However, a wide bandgap limits light absorption to the UV region.
Since the first report by Wrighton et al. [75] on the usefulness of STO for light-assisted photocatalytic water splitting, numerous studies have focused on bandgap engineering [76,77,78] and the tuning of the charge separation ability [79,80,81] to overcome the aforementioned drawbacks and achieve effective photoelectrochemical water splitting. It is expected that a heterojunction of STO with a small bandgap material should be able to narrow the overall bandgap and help suppress the limitation of visible light absorption.
TiO2 is a conventional n-type semiconductor, which is considered as the most promising photoanode material [82,83,84,85], and Cu2O is one of the most investigated p-type semiconductors for PEC water-splitting. It is reported that loading Cu2O particles on TiO2 nanotube arrays can significantly improve visible light absorption compared with pure TiO2 nanotubes [86]. Meanwhile, there is a similar valence and conduction band position between STO and TiO2, with a conduction band edge that is about 200 mV higher than TiO2 [87]. Recently, Cu2O has been combined with STO with controllable thickness to form a Cu2O/STO heterojunction photoelectrode, which has been proven to be effective in driving the separation of charge carriers by an electric field generated at the Cu2O/STO interface [88]. Under illumination, the Cu2O/STO photoelectrode presented a p-type photocurrent used for hydrogen generation. As expected, the Cu2O/STO heterojunction with an optimal thickness of 343 nm exhibited a photocurrent density of 2.52 mA/cm2 at 0.8 V versus saturated calomel electrode (SCE), which is 25 times higher than that of pristine Cu2O (0.10 mA/cm2 at 0.8 V versus SCE), as presented in Figure 7. This result is attributed to the synergy of broadening solar absorption and improved charge transportation in the Cu2O/STO heterojunction. In another report, the Cu2O/STO heterojunction photocathode was investigated using a less complicated method in which Cu2O nanoparticles (NPs) were loaded onto the surface of STO nanocubes (NCs) through a facile deposition—precipitation technique [89].
(4) PbTiO3
PbTiO3 (PTO) is a visible light active titania-based perovskite oxide that is well-known for its high ferroelectricity. It is a good candidate as a photocatalytic material in water splitting due to its bandgap value of 2.75 eV and suitable band structure, which assists in charge transfer and separation [90,91,92]. However, there are a limited number of studies on the application of PTO as a photoelectrode material with demonstrated higher photocatalytic activity for hydrogen evolution, especially in the case of PEC devices.
Recently, transition metal-doped PTO was investigated as a photocathode material for improving photogenerated electron transfer on the photocathode/electrolyte interface [92,93]. In particular, the 1 wt % Cu-doped PTO photoelectrode in an aqueous methanol solution under visible light irradiation (λ ≥ 400 nm) presented 2.5 times higher photocatalytic performance compared to PTO without Cu loading using the same configuration [92].
In another report, Hu et al. [93] modified the ITO/PTO photocathode using both Fe (III) doping and grafting, which resulted in a photocurrent increases of up to 220 µA/cm2 in comparison with 38 µA/cm2 of pure PTO photocathode. These enhancements were attributed to the change of band positions, which facilitated the transfer of photoinduced charge carriers. A similar phenomenon was observed for a fabricated Ag–Pt bimetallic catalyst on a ferroelectric PTO photocathode surface deposited on ITO quartz glass. As a result, the photocurrent density increased significantly from 60 µA/cm2 for pure PTO electrode to 202 µA/cm2 for the Ag–Pt electrode one under 0 V versus SCE, and with a 100 mW/cm2 Xe lamp illumination [94].
Due to the superior performance of PTO in comparison with other TiO2-based perovskite-type oxide materials (BaTiO3, SrTiO3) and their stability, some innovative synthesis methods have been introduced such as microwave-assisted synthesis [95,96] and nonhydrolytic sol-gel [97]. A promising and effective PTO photocathode with higher photocurrent efficiency has been achieved using such facile and cost-effective synthesis methods.
(5) YFeO3
YFeO3 is one of the ferrite perovskites featuring a bandgap energy of approximately 2.3–2.4 eV. Although several studies have been performed on the utilization of YFeO3 as a photocatalyst for water splitting [98,99,100,101], research on YFeO3 as a photocathode material in PEC devices remains limited. YFeO3 thin films are prepared by two different methodologies: nanoparticle thin film electrode via an ionic liquid protocol [68], and compact thin film electrode via a sol-gel method. Both types of electrodes have been evaluated for their PEC activity. They show exhibited cathodic photocurrent responses with an onset potential of 1.05 V versus RHE with complex dynamic features that should be addressed to application in PEC water splitting [102].

Quaternary Metal Oxides

Quaternary metal oxides, especially double perovskite materials featuring ferroelectric behavior, have been identified as promising candidates for solar energy to electricity conversion because of their efficient charge separation, which results from suitable control of the polarization-induced internal electric field [103,104]. Bi2FeCrO6 (BFCO) exhibiting multiferroic properties plays a role as a photocathode with a narrow bandgap (1.9–2.1 eV), and an appropriate conduction band position, which shows a twofold increase of the photocurrent density after negative poling, as presented in Figure 8 [105]. Recently, Shen et al. performed many studies on Pb(Zr,Ti)O3 (PZT) and revealed that ferroelectric PZT films deposited on indium tin oxide (ITO)-coated quartz glass [106,107] along with PZT films decorated with Ag nanoparticles on an ITO coated Si-pn+ junction [108] can function as stable and effective photocathodes for water splitting. More importantly, polarization switching can drive the photogenerated electrons transfer process, which then results in significantly enhanced PEC activity.

4.3. Ferroelectric-Based Materials in Photoanodes for Water

Photoanode materials, which possess bandgap values that utilize a wide range of visible light absorption, high carrier mobility, and the possibility of efficient charge transport and separation, generally stem from n-type semiconductors. In addition, they are cost-effective materials that have a long-term stability in aqueous solution, and are therefore considered promising candidates for PEC water splitting. In this section, ferroelectric-based materials are incorporated in a photoanodic configuration with various nanostructures via optimal fabrication techniques.

4.3.1. Ferroelectric Oxide Perovskites

(1) BaTiO3
BaTiO3 (BTO) is an n-type semiconductor, and its ferroelectricity was discovered for the first time in 1945 by Wul and Goldman [109]. Subsequently, the relation between photocurrent efficiency and the change in the electrode potential for the photooxidation of water was first investigated by Kennedy [110] in 1976. This was a precursor to advanced studies on this promising ferroelectric material. Apart from possessing a favorable structure for water splitting [111], BTO is considered as a potential photoelectrode material because of its high stability in aqueous solution along with a suitable band edge position, and controllable electronic properties through lattice defect modification or oxygen stoichiometry [112,113]. However, the only disadvantage of BTO is a relatively large bandgap of 3.2 eV, which limits its absorption spectrum to the ultraviolet region [114].
Therefore, many previous reports [115,116] were focused on doping BTO with Fe to extend its photoresponse for effective PEC activity. More importantly, it was reported that the ferroelectric polarization in BTO could enhance PEC performance in comparison with pristine TiO2. This is a result of the facilitation of the separation of photogenerated electron–hole pairs along with the effectiveness of tuning of the electronic band structure (i.e., upward band bending) in heterojunction-based PEC devices of TiO2/BTO core/shell nanowire (NW) arrays as photoanodes [117]. However, the contribution of the ferroelectric polarization effect plays a predominant role, which is evidenced by numerous different measurements such as the PE hysteresis loop, dynamic contact electrostatic force microscopy (DC-EFM) characterization, and electric poling by different directions. As a result, the positive polarization switched by external electric field poling at the TiO2/BTO (5 nm of thickness) interface yielded an optimal photocurrent density of water oxidation (1.30 mA/cm2) that was 67% higher than that of a photoanode without BTO (0.78 mA/cm2). This result hints at the tremendous potential of ferroelectric photoelectrodes in enhancing PEC performance.
(2) BiVO4
Bismuth vanadate (BiVO4) is an n-type semiconductor (bandgap value ~2.5 eV), which has been widely applied as a photoanode material in PEC cells because of its large absorbance range in the visible spectrum along with a suitable conduction band structure [118,119]. Based on its bandgap energy, BiVO4 is able to generate a photocurrent density of ~7.5 mA/cm2 and a theoretical solar-to-hydrogen conversion efficiency of 9.2% under AM 1.5 G conditions [120,121]. However, BiVO4 suffers from poor surface catalytic reactivity and substantial recombination losses, which prevent it from achieving the theoretical maximum. With the aim of addressing these drawbacks, many strategies such as doping, morphology control, and the construction of heterojunction structures have been investigated in PEC devices containing BiVO4 photoelectrodes.
Choi et al. [122] prepared nanoporous BiVO4 photoanodes with dual-layer oxygen evolution catalysts (OEC), which is reported to alleviate bulk carrier recombination at the BiVO4/OEC junction by creating a more favorable Helmholtz layer potential drop at the OEC/electrolyte junction. PEC devices containing BiVO4/FeOOH/NiOOH photoanodes have been shown to yield a photocurrent density of 2.73 mA/cm2 at a potential as low as 0.6 V versus RHE without any extrinsic doping and composition tuning. Subsequently, this author’s group performed further mild annealing treatment of nanoporous BiVO4 under N2 flow that produced nitrogen doping and the generation of oxygen vacancies [123]. This is attributed to the enhanced major carrier density as well as the major carrier mobility of BiVO4. The photocurrrent was drastically improved to 4.16 ± 0.41 mA/cm2, and the NiOOH/FeOOH/N-BiVO4 photoanode manifested an applied biased photo-to-current efficiency (ABPE) of 2.0% under a bias of 0.6 V. In further investigations of the same BiVO4/Fe(Ni)OOH tandem structure, further enhancement of the photocurrent density up to 5.82 ± 0.36 mA/cm2 at 1.23 V versus RHE was achieved by Mo-doping with a concentration of 3% on a nanocone FTO/BiVO4 substrate [124].
In other studies [125,126,127,128], because WO3 has well-known properties such as its low cost, high chemical stability, and good charge transport activity, BiVO4 was coupled with this material to overcome the moderate charge transport feature of PEC cells fabricated from pure BiVO4 to produce PEC cells with robust performance. Nevertheless, these studies at potentials as low as 0.6 V versus RHE still demonstrate charge separation efficiencies of less than 60%.
Kuang et al. [129] reported on a non-doped nanostructured BiVO4 photoanode with a bimetallic NiFe-(oxy)hydroxide/borate(NiFeOx-Bi) oxygen evolution catalyst as an efficient oxidation co-catalyst to achieve a solar energy conversion efficiency in excess of 2%. More importantly, a very high of incident photo-to-current efficiency (IPCE) of approximately 80% was achieved at a potential as low as 0.6 V versus RHE under front irradiation up to 460 nm, which is nearly twice of that of a previous nanoporous BiVO4 electrode [122] without compromising other beneficial properties. Recently, Lee et al. [130] fabricated epitaxial BiVO4 on a thin γ-WO3 template layer deposited on a SrTiO3 (001) substrate by pulsed laser deposition (PLD).
Although PLD was applied in the fabrication of a BiVO4 photoanode electrode in PEC cells for the first time by Liu and Yan [131] in the last decade, it drew less attention at the time because of its very low photocurrent values, regardless of the high-quality films. However, the effect of the thin γ-WO3 template layer on the effective charge transfer and increased active surface area of BiVO4 accounted for the significantly enhanced photocurrent density (2.20 mA/cm2) at 1.23 V versus RHE, which is approximately 10 times higher than that of bare BiVO4. This study highlighted that PLD combined with suitable BiVO4 photoanode materials is a facile and versatile technique for producing good-quality electrodes for PEC water splitting.
(3) SrTiO3
Titanium dioxide (TiO2) was the first oxide used as a photoanode material for PEC water splitting and was discovered by Fujishima-Honda [13]. Until now, it has been one of the most attractive materials for PEC cells. However, TiO2 absorbs only in the UV part of the spectrum because of its large bandgap [132], which exhibits low efficiency in water-splitting reactions when TiO2 is used as a single photoanode [133,134,135,136].
SrTiO3 (STO) is a perovskite oxide that shows considerable ferroelectricity [137,138] and possesses a favorable band edge structure that overlaps with the water redox potentials. In addition, the conduction band edge is more negative than TiO2, as presented in Figure 9 [139]. Similar to TiO2, STO responds only to the UV region and a negligible portion of the visible light. This has prevented it from being a good single photoanode candidate for PEC cells. As a result, STO has been doped with dopants, or coupled with TiO2, which the aim of shifting the Fermi level of the dual-phase STO/TiO2 composite [87], for application as a potential candidate for improving photoconversion efficiency.
Limited reports have focused on TO thin films with dopants such as (Cr, Rh, Ir, and Nb) [140,141,142,143,144]. However, a few of these studies have reported some drawbacks such as non-uniform doping and the insignificant effect of such dopants, or even the resulting instability because of the formation of IrO2 [145]. Recently, further investigation into STO/TiO2 photoanode materials has revealed some desirable outcomes [87,139,145,146,147,148]. Wysmulek et al. [136] studied a durable eutectic system made up of TiO2 and STO as an active photoanode material for PEC cells. Under 600 mW/cm2 of solar irradiation, the TiO2/STO eutectic photoelectrode yielded a photocurrent density of up to 8.5 mA/cm2 at 1.5 V versus NHE and stabilized after 30 h of testing. This result not only shows an improvement in the photocurrent density, it also showed a better long-term operation compared with Ir-doped STO (24 h) [145]. Thus, eutectic composite-based photoelectrodes are untapped components for PEC water splitting. In addition, the spontaneous electric polarization in STO can simultaneously enhance charge separation and hole transportation in TiO2/STO core-shell nanowires (NWs) [146]. Compared to the TiO2/BaTiO3 ferroelectric PEC system [138], STO ferroelectric material shows better charge mobility, which will facilitate hole migration inside the ferroelectric layer. Interestingly, the core shell with an optimal STO thickness of 10 nm generates the highest photocurrent density of 1.43 mA/cm2 and has the charge-separation efficiency of 87.7% at 1.23 V versus RHE. This corresponds to an 83% and 79% improvement in comparison with pristine TiO2 NWs. This study has paved the way for the application of semiconducting ferroelectric materials to further advance the development of ferroelectric PEC systems.
(4) BiFeO3
The BiFeO3 (BFO) ferroelectric material shows unique photovoltaic effects that result from large spontaneous polarization. This prominent characteristic develops a high built-in potential that effectively enhances the separation and drift of photogenerated carriers for application in PEC devices composed of BFO-containing photoelectrodes. In comparison with photocathodic materials, there are few reports regarding the application of BFO as photoanodic materials in PEC water splitting.
Very recently, Song et al. [4] reported on epitaxial BFO thin-film photoanodes with different crystallographic orientations including (111)pc, (110)pc, (001)pc, and the consequence of the ferroelectric domain structures was also investigated. To investigate the difference among various crystallographic orientations on ferroelectric properties and the effect of different polarization states on photovoltaic performance, ferroelectric P–E hysteresis loop and polarization switching measurements were performed, respectively, as shown in Figure 10. The results revealed that in the absence of polarization, the charge energy band bending, which induces inefficient charge separation, was not observed. In contrast, an optimal PEC performance with (111)pc BFO thin film under a downward polarization state was revealed. Previously, enhanced efficiency in a polycrystalline BFO photoanode also exhibited a similar mechanism when switching the polling bias from +8 V to −8 V [149]. This BFO thin film was based on the cost-effective technology of spin coating, instead of growing epitaxial BFO thin film using more sophisticated techniques including radio frequency (RF) and pulsed laser deposition (PLD) [150,151]. However, this limits their application in large-scale technology.
(5) PbTiO3
Although a large number of studies have been performed based on PbTiO3-containing materials as photocathodes in PEC cells for hydrogen production, photoanodes composed of PbTiO3 (PTO) are increasingly of concern. Specifically, there are numerous reports on nanostructured PTO because of their superior PEC properties, which are related to their randomly shaped particles [152,153]. Moreover, many useful approaches based on heterojunctions from different semiconductors have resulted in improved photoactivity [127,154,155,156,157,158].
Based on these two interpretations, Jang et al. [159] synthesized vertically aligned core–shell PTO@TiO2 heterojunction nanotube arrays fabricated on FTO glass by a three-step process that can help suppress the opaque photoanode problem reported in a previous study [91]. As expected, the directional charge transport in one-dimensional (1D) nanostructure and additional heterojunction effects between PTO and TiO2 favor charge separation and enhanced PEC efficiency in PTO@TiO2 photoanode-based systems. Recently, there was a report on the synthesis of nanotubular PTO-based photoanodes in PEC devices for the first time [160]. The investigation was performed with two synthetic electrodes of 1D nanotube arrays of PTO (NT–PTO) and Pt-dots@PTO nanotube arrays (PNT–PTO) in comparison with two reference electrodes made of PTO nanotubes (no Pt dots) and PTO powder (no nanotubes). The result presents that the NT–PTO photoanode achieved the highest photocurrent density of 64 µA/cm2 at 1.05 V versus RHE. In comparison, the PTO powder-based photoelectrode exhibited a meager value of 8 µA/cm2 at the same RHE potential. The enhanced PEC performance was elucidated by reducing the electron–hole recombination through the isolated oxidation sites at the external surface of the Pt-dots@PTO nanotubular structure.

4.3.2. Ferroelectric Chalcogenides

(1) CdS
Cadmium sulfide (CdS) is the most common ferroelectric chalcogenide investigated in PEC water splitting because of its narrow energy bandgap of 2.4 eV and suitable band edge position [157,158]. Despite these advantages, the CdS-based PEC devices showed a low performance because of inefficient charge transfer and separation [161,162,163,164,165].
Uniform carbon-coated CdS core–shell nanostructures are potentially useful as a facile and novel approach to overcome this challenge [166]. The enhanced performance is elucidated by the substantially improved electron transfer, which results from the high electrical conductivity of the coated carbon layer on CdS. CdS widely couples with ZnO nanowires and nanorods as an ideal heterostructure in PEC devices that helps to suppress internal charge recombination and enhance solar-to-chemical conversion [167,168,169]. Accordingly, the development of ZnO/CdS nanotube arrays outperforms the previous structures with a larger specific surface, which results in a superior ability for charge extraction, especially in the case of nanoparticle plasmon metal addition [170,171].
Besides the common nanostructured CdS, a rare report on the versatile hierarchical structure of CdS film was investigated recently [172]. Zang et al. [173] grew screw-like SnO2 nanosheets on rod-like single-crystalline SnO2 nanowires deposited with CdS quantum dots in PEC cells that tend to move toward efficient PEC water splitting. This structure is a combination of both the fast charge transport path of 1D nanostructures and the high porosity and light absorption of 2D nanosheets, which will extend the perspective of high-performance PEC cells in the near future.
(2) ZnS
Zinc sulfide (ZnS) is regarded as one of the most important II–VI semiconductors, and is an appealing candidate for water splitting because of its own outstanding properties, which include an ability to generate mobile photoexcited charge carriers [174], fast electron transfer because of high CB potential, and high catalytic activity for H2 generation under conditions without a co-catalyst [175,176]. However, the performance of pure ZnS-based photoelectrode in water-splitting devices is limited, because absorption is only in the UV region due to the wide bandgap value of 3.66 eV of this material [177]. Furthermore, overcoming the high charge recombination is also a serious challenge [178,179].
As a result, many modifications have been investigated for ZnS-based electrodes to increase visible light absorption. To enhance photocatalytic activity, transition metal-doped ZnS (Cu, Ni, Mo) or doping with GaN were investigated as potentially efficient photocatalysts to generate H2 under visible light irradiation [180,181,182]. However, due to the remaining challenges of dopant introduction [181,183,184,185,186] an alternative method by controlling defects in nanostructured ZnS has been exploited through PLD to enhance the overall PEC properties of ZnS [187]. The fabrication of ZnS-based heterojunctions with other semiconductors for more advanced designs has been considered to enhance PEC performance for water splitting.
Typically, the ZnO/ZnS heterostructure has drawn tremendous attention because of the extended visible absorption that it affords and its appropriate alignment [188,189,190,191,192]. Among ZnO/ZnS heterostructures with different morphologies such as nanowires [193] and nanorings [194] synthesized by means of chemical processes, the PEC behavior is scarcely introduced by anodization, which facilitates electron–hole separation, and consequently, the enhancement of photoelectrochemical activity for water splitting [195]. Moreover, the design of sandwich-structured ZnO/ZnS photoanodes with a third component, such as noble metals Au or graphite-like carbon nitride (C3N4), have been demonstrated with significantly superior PEC activity compared to those of a pristine ZnO photoanode [196,197]. Nevertheless, the remaining challenge is the wide bandgap of the ZnO/ZnS heterojunction, which still limits large-range visible light absorption. The systems that result from coupling ZnS with narrow bandgap chalcogenide CdS exhibit much higher performance than those of every single material-based system [198,199,200,201,202].

4.3.3. Hybrid Halide Ferroelectric CH3NH3PbI3 Perovskite Tandem System Approach

Tandem cell configuration is a facile and novel approach to address the limitation of single or heterojunction PEC devices for water splitting. Organic–inorganic hybrid perovskite CH3NH3PbI3 has attracted tremendous attention in perovskite solar cells (PSCs) because of their superb light-harvesting characteristics, large electron/hole diffusion lengths, and high crystallinity [203,204,205,206,207]. Based on the latest research, the certified highest efficiency of PSCs is 22.1% fabricated on limited to very small areas (~1 cm2) [208]. Ferroelectric domains have been observed in CH3NH3PbI3, despite the centrosymmetric structure of this perovskite material, which is ascribed to the reduced symmetry of molecular CH3NH3+ dipoles [209,210].
Recently, Luo et al. [211] demonstrated two perovskite solar cells connected in series as a tandem cell that can serve as a separated and external power supply for water splitting, as presented in Figure 11. Although this device can achieve a high photocurrent density of ~10 mA/cm2, the electrodes were not directly settled in the photoreactions. Moreover, the rapid fluctuation of the photocurrent due to the instability of the perovskite was a challenge that could be addressed by encapsulation techniques to achieve intriguing PEC tandem systems. More recently, Da et al. reported on a CH3NH3PbI3-based photoanode coated by an ultrathin Ni layer for the first time with a much enhanced photocurrent density of 12 mA/cm2, which is attributed to the improved photoabsorption of CH3NH3PbI3 [212]. In their report, a Ni top layer played a decisive role in the enhancement of the photocurrent density and assisted in the suppression of the instability of perovskite PEC tandem systems in water; this evokes many promising perspectives in the development of hybrid perovskite-based tandem cells.

5. Conclusions and Outlook

Interest in solar water splitting is on the rise, which is in line with a general trend of increasing environmental awareness. Based on state-of-the-art reports, amazing progress in solar water splitting has been demonstrated, including the upgrade of systems, and the development of new electrode materials and their structure. To perform high-efficiency solar water splitting, the materials used in the different systems have to satisfy three conditions. They should have: an appropriate bandgap energy, a suitable position of the bandgap, and chemical stability. The ideal material has a bandgap of approximately ~2.0–2.2 eV, which can absorb about 40% of the total incident sunlight. The conduction band (CB) and valence band (VB) of the materials should be higher—between the water reduction level and below the oxidation level of water—in order to induce the water decomposition reaction. At present, no materials have been identified that meet all of the requirements for water decomposition reactions.
In this review, we introduced various ferroelectric materials for solar water splitting with high STH efficiency. Ferroelectric materials have been known to be very stable and strong for chemical and physical applications, and their bandgap could be narrowed by doping and/or substitution. Even though solar water splitting systems with ferroelectric materials are the newest technology and have attracted important new research into hydrogen generation, until now, there has not yet been much research performed on ferroelectric materials for use in water-splitting systems. Solar water-splitting systems with ferroelectrics are expected as significant advances in the process of hydrogen generation, with high STH efficiencies compared to other materials.

Author Contributions

C.W.B. had the main idea for the article. S.K., N.T.N. performed the literature search and the writing of the manuscript in consultation with C.W.B. All authors discussed the concept of review and commented on the manuscript.

Funding

This research was supported by the Korea Electric Power Corporation. (Grant number: R17XA05-10) and Creative Materials Discovery Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (2017M3D1A1040828), and by the grants from the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT, and Future Planning (NRF-2016R1C1B1014649)

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lopes, T.; Dias, P.; Andrade, L.; Mendes, A. An innovative photoelectrochemical lab device for solar water splitting. Sol. Energy Mater. Sol. Cells 2014, 128, 399–410. [Google Scholar] [CrossRef] [Green Version]
  2. Ngoh, S.K.; Njomo, D. An overview of hydrogen gas production from solar energy. Renew. Sustain. Energy Rev. 2012, 16, 6782–6792. [Google Scholar] [CrossRef]
  3. Sun, J.W.; Zhong, D.K.; Gamelin, D.R. Composite photoanodes for photoelectrochemical solar water splitting. Energy Environ. Sci. 2010, 3, 1252–1261. [Google Scholar] [CrossRef]
  4. Song, J.; Kim, T.L.; Lee, J.; Cho, S.Y.; Cha, J.; Jeong, S.Y.; An, H.; Kim, W.S.; Jung, Y.S.; Park, J.; et al. Domain-engineered BiFeO3 thin-film photoanodes for highly enhanced ferroelectric solar water splitting. Nano Res. 2018, 11, 642–655. [Google Scholar] [CrossRef]
  5. Pan, H.; Zhu, S.; Lou, X.; Mao, L.; Lin, J.; Tian, F.; Zhang, D. Graphene-based photocatalysts for oxygen evolution from water. RSC Adv. 2015, 5, 6543–6552. [Google Scholar] [CrossRef]
  6. Zhao, J.; Wang, X.; Xu, Z.C.; Loo, J.S.C. Hybrid catalysts for photoelectrochemical reduction of carbon dioxide: A prospective review on semiconductor/metal complex co-catalyst systems. J. Mater. Chem. A 2014, 2, 15228–15233. [Google Scholar] [CrossRef]
  7. Jang, J.S.; Kim, H.G.; Lee, J.S. Heterojunction semiconductors: A strategy to develop efficient photocatalytic materials for visible light water splitting. Catal. Today 2012, 185, 270–277. [Google Scholar] [CrossRef]
  8. Vilanova, A.; Lopes, T.; Spenke, C.; Wullenkord, M.; Mendes, A. Optimized photoelectrochemical tandem cell for solar water splitting. Energy Storage Mater. 2018, 13, 175–188. [Google Scholar] [CrossRef]
  9. Tentu, R.D.; Basu, S. Photocatalytic water splitting for hydrogen production. Curr. Opin. Electrochem. 2017, 5, 56–62. [Google Scholar] [CrossRef]
  10. Li, R. Latest progress in hydrogen production from solar water splitting via photocatalysis, photoelectrochemical, and photovoltaic-photoelectrochemical solutions. Chin. J. Catal. 2017, 38, 5–12. [Google Scholar] [CrossRef]
  11. Ismail, A.A.; Bahnemann, D.W. Photochemical splitting of water for hydrogen production by photocatalysis: A review. Sol. Energy Mater. Sol. Cells 2014, 128, 85–101. [Google Scholar] [CrossRef]
  12. Jiang, C.R.; Moniz, S.J.A.; Wang, A.G.; Zhang, T.; Tang, J.W. Photoelectrochemical devices for solar water splitting-materials and challenges. Chem. Soc. Rev. 2017, 46, 4645–4660. [Google Scholar] [CrossRef] [PubMed]
  13. Honda, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nat. Commun. 1972, 238, 37–38. [Google Scholar]
  14. Alfaifi, B.Y.; Ullah, H.; Alfaifi, S.; Tahir, A.A.; Mallick, T.K. Photoelectrochemical Solar Water Splitting: From Basic Principles to Advanced Devices. Veruscr. Funct. Nanomater. 2018, 2, 1–26. [Google Scholar] [CrossRef]
  15. Chen, Z.; Dinh, D.; Miller, E. Chapter 3 Experimental Considerations. In Photoelectrochemical Water Splitting, 1st ed.; Springer: New York, NY, USA, 2013; pp. 17–44. ISBN 978-1-4614-8298-7. [Google Scholar]
  16. Paillard, C.; Bai, X.F.; Infante, I.C.; Guennou, M.; Geneste, G.; Alexe, M.; Kreisel, J.; Dkhil, B. Photovoltaics with ferroelectrics: Current status and beyond. Adv. Mater. 2016, 28, 5153–5168. [Google Scholar] [CrossRef] [PubMed]
  17. Wick, R.; Tilley, S.D. Photovoltaic and photoelectrochemical solar energy conversion with Cu2O. J. Phys. Chem. C 2015, 119, 26243–26257. [Google Scholar] [CrossRef]
  18. Tablero, C. Photovoltaic application of the multiferroic Bi2FeCrO6 double perovskite. Sol. Energy 2016, 137, 173–178. [Google Scholar] [CrossRef]
  19. Archer, M.D. Photovoltaics and photoelectrochemistry: Similarities and differences. Physica E 2002, 14, 61–64. [Google Scholar] [CrossRef]
  20. Lopez-Varo, P.; Bertoluzzi, L.; Bisquert, J.; Alexe, M.; Coll, M.; Huang, J.S.; Juan, A.J.T.; Kirchartz, T.; Nechache, R.; Rosei, F.; et al. Physical aspects of ferroelectric semiconductors for photovoltaic solar energy conversion. Phys. Rep. 2016, 653, 1–40. [Google Scholar] [CrossRef] [Green Version]
  21. Spanier, J.E.; Fridkin, V.M.; Rappe, A.M.; Akbashev, A.R.; Polemi, A.; Qi, Y.B.; Gu, Z.Q.; Young, S.M.; Hawley, C.J.; Imbrenda, D.; et al. Power conversion efficiency exceeding the shockley-queisser limit in a ferroelectric insulator. Nat. Photonics 2016, 10, 611–616. [Google Scholar] [CrossRef]
  22. Bonke, S.A.; Wiechen, M.; MacFarlane, D.R.; Spiccia, L. Renewable fuels from concentrated solar power: Towards practical artificial photosynthesis. Energy Environ. Sci. 2015, 8, 2791–2796. [Google Scholar] [CrossRef]
  23. Jia, J.Y.; Seitz, L.C.; Benck, J.D.; Huo, Y.J.; Chen, Y.S.; Ng, J.W.D.; Bilir, T.; Harris, J.S.; Jaramillo, T.F. Solar water splitting by photovoltaic-electrolysis with a solar-to-hydrogen efficiency over 30%. Nat. Commun. 2016, 7, 1–6. [Google Scholar] [CrossRef] [PubMed]
  24. Chu, S.; Li, W.; Yan, Y.; Hamann, T.; Shin, I.; Wang, D.; Mi, Z. Roadmap on solar water splitting: Current status and future prospects. Nano Futures 2017, 1, 1–29. [Google Scholar] [CrossRef]
  25. Jacobsson, T.J.; Fjallstrom, V.; Edoff, M.; Edvinsson, T. Sustainable solar hydrogen production: From photoelectrochemical cells to pv-electrolyzers and back again. Energy Environ. Sci. 2014, 7, 2056–2070. [Google Scholar] [CrossRef]
  26. Tamirat, A.G.; Rick, J.; Dubale, A.A.; Su, W.N.; Hwang, B.J. Using hematite for photoelectrochemical water splitting: A review of current progress and challenges. Nanoscale Horiz. 2016, 1, 243–267. [Google Scholar] [CrossRef]
  27. Butler, K.T.; Frost, J.M.; Walsh, A. Ferroelectric materials for solar energy conversion: Photoferroics revisited. Energy Environ. Sci. 2015, 8, 838–848. [Google Scholar] [CrossRef] [Green Version]
  28. Khan, M.A.; Nadeem, M.A.; Idrissn, H. Ferroelectric polarization effect on surface chemistry and photo-catalytic activity: A review. Surf. Sci. Rep. 2016, 71, 1–31. [Google Scholar] [CrossRef]
  29. Li, L.; Salvador, P.A.; Rohrer, G.S. Photocatalysts with internal electric fields. Nanoscale 2014, 6, 24–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Sivula, K.; Krol, R.V.D. Semiconducting materials for photoelectrochemical energy conversion. Nat. Rev. Mater. 2016, 1, 1–17. [Google Scholar] [CrossRef]
  31. 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]
  32. Burke, M.S.; Enman, L.J.; Batchellor, A.S.; Zou, S.; Boettcher, S.W. Oxygen Evolution Reaction Electrocatalysis on Transition Metal Oxides and (Oxy)hydroxides: Activity Trends and Design Principles. Chem. Mater. 2015, 27, 7549–7558. [Google Scholar] [CrossRef]
  33. Jin, J.; Walczak, K.; Singh, M.R.; Karp, C.; Lewis, N.S.; Xiang, C. An experimental and modeling/simulation-based evaluation of the efficiency and operational performance characteristics of an integrated, membrane-free, neutral pH solar-driven water-splitting system. Energy Environ. Sci. 2014, 7, 3371–3380. [Google Scholar] [CrossRef] [Green Version]
  34. Seger, B.; Castelli, I.E.; Vesborg, P.C.K.; Jacobsen, K.W.; Hansen, O.; Chorkendorff, I. 2-Photon tandem device for water splitting: Comparing photocathode first versus photoanode first designs. Energy Environ. Sci. 2014, 7, 2397–2413. [Google Scholar] [CrossRef] [Green Version]
  35. Hara, M.; Kondo, T.; Komoda, M.; Ikeda, S.; Shinohara, K.; Tanaka, A.; Kondo, J.N.; Domen, K. Cu2O as a photocatalyst for overall water splitting under visible light irradiation. Chem. Commun. 1998, 357–358. [Google Scholar] [CrossRef]
  36. Jongh, P.E.D.; Vanmaekelbergh, D.; Kelly, J.J. Cu2O: A catalyst for the photochemical decomposition of water? Chem. Commun. 1999, 1069–1070. [Google Scholar] [CrossRef]
  37. Chauhan, D.; Satsangi, V.R.; Dass, S.; Shrivastav, R. Preparation and characterization of nanostructured CuO thin films for photoelectrochemical splitting of water. Bull. Mater. Sci. 2006, 29, 709–716. [Google Scholar]
  38. Koffyberg, F.P.; Benko, F.A. A photoelectrochemical determination of the position of the conduction and valence band edges of p-type CuO. J. Appl. Phys. 1982, 53, 1173–1177. [Google Scholar] [CrossRef]
  39. Hsu, Y.K.; Yu, C.H.; Lin, H.H.; Chen, Y.C.; Lin, Y.G. Template synthesis of copper oxide nanowires for photoelectrochemical hydrogen generation. J. Electroanal. Chem. 2013, 704, 19–23. [Google Scholar] [CrossRef]
  40. Gupta, R.K.; Ghosh, K.; Kahol, P.K. Fabrication and characterization of NiO/ZnO p-n junctions by pulsed laser deposition. Physica E 2009, 41, 617–620. [Google Scholar] [CrossRef]
  41. Dong, Y.M.; Chen, Y.M.; Jiang, P.P.; Wang, G.L.; Wu, X.M.; Wu, R.X.; Zhang, C. Efficient and stable MoS2/CdSe/NiO photocathode for photoelectrochemical hydrogen generation from water. Chem. Asian J. 2015, 10, 1660–1667. [Google Scholar] [CrossRef] [PubMed]
  42. Yilmaz, P.; Yeo, D.; Chang, H.; Loh, L.; Dunn, S. Perovskite BiFeO3 thin film photocathode performance with visible light activity. Nanotechnology 2016, 27, 1–9. [Google Scholar] [CrossRef] [PubMed]
  43. Yi, H.T.; Choi, T.; Choi, S.G.; Oh, Y.S.; Cheong, S.W. Mechanism of the switchable photovoltaic effect in ferroelectric BiFeO3. Adv. Mater. 2011, 23, 3403–3407. [Google Scholar] [CrossRef] [PubMed]
  44. Ren, Y.; Nan, F.; You, L.; Zhou, Y.; Wang, Y.Y.; Wang, J.L.; Su, X.D.; Shen, M.R.; Fang, L. Enhanced photoelectrochemical performance in reduced graphene oxide/BiFeO3 heterostructures. Small 2017, 13, 1–7. [Google Scholar] [CrossRef] [PubMed]
  45. Quynh, L.T.; Van, C.N.; Bitla, Y.; Chen, J.W.; Do, T.H.; Tzeng, W.Y.; Liao, S.C.; Tsai, K.A.; Chen, Y.C.; Wu, C.L.; et al. Self-assembled BiFeO3-ε-Fe2O3 vertical heteroepitaxy for visible light photoelectrochemistry. Adv. Energy Mater. 2016, 6. [Google Scholar] [CrossRef]
  46. Liu, Q.; Zhou, Y.; You, L.; Wang, J.L.; Shen, M.R.; Fang, L. Enhanced ferroelectric photoelectrochemical properties of polycrystalline BiFeO3 film by decorating with ag nanoparticles. Appl. Phys. Lett. 2016, 108, 1–5. [Google Scholar] [CrossRef]
  47. Ji, W.; Yao, K.; Lim, Y.F.; Liang, Y.C.; Suwardi, A. Epitaxial ferroelectric BiFeO3 thin films for unassisted photocatalytic water splitting. Appl. Phys. Lett. 2013, 103, 1–4. [Google Scholar] [CrossRef]
  48. Hauser, A.J.; Zhang, J.; Mier, L.; Ricciardo, R.A.; Woodward, P.M.; Gustafson, T.L.; Brillson, L.J.; Yang, F.Y. Characterization of electronic structure and defect states of thin epitaxial BiFeO3 films by UV-visible absorption and cathodoluminescence spectroscopies. Appl. Phys. Lett. 2008, 92, 1–3. [Google Scholar] [CrossRef]
  49. Gupta, S.; Medwal, R.; Limbu, T.B.; Katiyar, R.K.; Pavunny, S.P.; Tomar, M.; Morell, G.; Gupta, V.; Katiyar, R.S. Graphene/semiconductor silicon modified BiFeO3/indium tin oxide ferroelectric photovoltaic device for transparent self-powered windows. Appl. Phys. Lett. 2015, 107. [Google Scholar] [CrossRef]
  50. Cho, S.; Jang, J.W.; Zhang, W.R.; Suwardi, A.; Wang, H.Y.; Wang, D.W.; MacManus-Driscoll, J.L. Single-crystalline thin films for studying intrinsic properties of BiFeO3-SrTiO3 solid solution photoelectrodes in solar energy conversion. Chem. Mater. 2015, 27, 6635–6641. [Google Scholar] [CrossRef]
  51. Chen, X.Y.; Yu, T.; Gao, F.; Zhang, H.T.; Liu, L.F.; Wang, Y.M.; Li, Z.S.; Zou, Z.G.; Liu, J.M. Application of weak ferromagnetic BiFeO3 films as the photoelectrode material under visible-light irradiation. Appl. Phys. Lett. 2007, 91. [Google Scholar] [CrossRef]
  52. Mohan, S.; Subramanian, B.; Bhaumik, I.; Gupta, P.K.; Jaisankar, S.N. Nanostructured Bi1−xGdxFeO3—A multiferroic photocatalyst on its sunlight driven photocatalytic activity. RSC Adv. 2014, 4, 16871–16878. [Google Scholar] [CrossRef]
  53. Gu, S.L.; Zhou, X.X.; Zheng, F.G.; Fang, L.; Dong, W.; Shen, M.R. Improved photocathodic performance in Pt catalyzed ferroelectric BiFeO3 films sandwiched by a porous carbon layer. Chem. Commun. 2017, 53, 7052–7055. [Google Scholar] [CrossRef] [PubMed]
  54. Cheng, X.R.; Shen, H.Y.; Dong, W.; Zheng, F.G.; Fang, L.A.; Su, X.D.; Shen, M.R. Nano-au and ferroelectric polarization mediated Si/ITO/BiFeO3 tandem photocathode for efficient H2 production. Adv. Mater. Interfaces 2016, 3, 1–7. [Google Scholar] [CrossRef]
  55. Kang, D.H.; Kim, T.W.; Kubota, S.R.; Cardiel, A.C.; Cha, H.G.; Choi, K.S. Electrochemical synthesis of photoelectrodes and catalysts for use in solar water splitting. Chem. Rev. 2015, 115, 12839–12887. [Google Scholar] [CrossRef] [PubMed]
  56. Atwater, H.A.; Polman, A. Plasmonics for improved photovoltaic devices. Nat. Mater. 2010, 9, 205–213. [Google Scholar] [CrossRef] [PubMed]
  57. Takai, A.; Kamat, P.V. Capture, store, and discharge. Capture, store, and discharge. Shuttling photogenerated electrons across TiO2-silver interface. ACS Nano 2011, 5, 7369–7376. [Google Scholar] [CrossRef] [PubMed]
  58. Shen, H.Y.; Zhou, X.X.; Dong, W.; Su, X.D.; Fang, L.A.; Wu, X.; Shen, M.R. Dual role of TiO2 buffer layer in Pt catalyzed BiFeO3 photocathodes: Efficiency enhancement and surface protection. Appl. Phys. Lett. 2017, 111. [Google Scholar] [CrossRef]
  59. Prévot, M.S.; Sivula, K. Photoelectrochemical tandem cells for solar water splitting. J. Phys. Chem. C 2013, 117, 17879–17893. [Google Scholar] [CrossRef]
  60. Hu, S.; Xiang, C.X.; Haussener, S.; Berger, A.D.; Lewis, N.S. An analysis of the optimal band gaps of light absorbers in integrated tandem photoelectrochemical water-splitting systems. Energy Environ. Sci. 2013, 6, 2984–2993. [Google Scholar] [CrossRef]
  61. Fan, R.L.; Dong, W.; Fang, L.A.; Zheng, F.G.; Su, X.D.; Zou, S.; Huang, J.; Wang, X.S.; Shen, M.R. Stable and efficient multi-crystalline n(+)p silicon photocathode for H2 production with pyramid-like surface nanostructure and thin Al2O3 protective layer. Appl. Phys. Lett. 2015, 106. [Google Scholar] [CrossRef]
  62. Fan, R.L.; Min, J.W.; Li, Y.; Su, X.D.; Zou, S.; Wang, X.S.; Shen, M.R. N-type silicon photocathodes with Al-doped rear p(+) emitter and Al2O3-coated front surface for efficient and stable H2 production. Appl. Phys. Lett. 2015, 106. [Google Scholar] [CrossRef]
  63. Yu, Q.; Meng, X.G.; Wang, T.; Li, P.; Liu, L.Q.; Chang, K.; Liu, G.G.; Ye, J.H. A highly durable p-LaFeO3/n-Fe2O3 photocell for effective water splitting under visible light. Chem. Commun. 2015, 51, 3630–3633. [Google Scholar] [CrossRef] [PubMed]
  64. Wheeler, G.P.; Choi, K.S. Photoelectrochemical properties and stability of nanoporous p-type LaFeO3 photoelectrodes prepared by electrodeposition. ACS Energy Lett. 2017, 2, 2378–2382. [Google Scholar] [CrossRef]
  65. Ding, J.L.; Lu, X.M.; Shu, H.M.; Xie, J.M.; Zhang, H. Microwave-assisted synthesis of perovskite ReFeO3 (Re: La, sm, eu, gd) photocatalyst. Mater. Sci. Eng. B 2010, 171, 31–34. [Google Scholar] [CrossRef]
  66. Parida, K.M.; Reddy, K.H.; Martha, S.; Das, D.P.; Biswal, N. Fabrication of nanocrystalline LaFeO3: An efficient sol-gel auto-combustion assisted visible light responsive photocatalyst for water decomposition. Int. J. Hydrogen Energy 2010, 35, 12161–12168. [Google Scholar] [CrossRef]
  67. Tijare, S.N.; Joshi, M.V.; Padole, P.S.; Mangrulkar, P.A.; Rayalu, S.S.; Labhsetwar, N.K. Photocatalytic hydrogen generation through water splitting on nano-crystalline LaFeO3 perovskite. Int. J. Hydrogen Energy 2012, 37, 10451–10456. [Google Scholar] [CrossRef]
  68. Celorrio, V.; Bradley, K.; Weber, O.J.; Hall, S.R.; Fermin, D.J. Photoelectrochemical properties of LaFeO3 nanoparticles. ChemElectroChem 2014, 1, 1667–1671. [Google Scholar] [CrossRef]
  69. Peng, Q.; Wang, J.; Feng, Z.J.; Du, C.; Wen, Y.W.; Shan, B.; Chen, R. Enhanced photoelectrochemical water oxidation by fabrication of p-LaFeO3/n-Fe2O3 heterojunction on hematite nanorods. J. Phys. Chem. C 2017, 121, 12991–12998. [Google Scholar] [CrossRef]
  70. Diez-Garcia, M.I.; Gómez, R. Metal doping to enhance the photoelectrochemical behavior of LaFeO3 photocathodes. ChemSusChem 2017, 10, 2457–2463. [Google Scholar] [CrossRef] [PubMed]
  71. Perkins, J.D.; Paudel, T.R.; Zakutayev, A.; Ndione, P.F.; Parilla, P.A.; Young, D.L.; Lany, S.; Ginley, D.S.; Zunger, A.; Perry, N.H.; et al. Inverse design approach to hole doping in ternary oxides: Enhancing p-type conductivity in cobalt oxide spinels. Phys. Rev. B 2011, 84, 1–8. [Google Scholar] [CrossRef]
  72. Peng, Q.; Wang, J.; Wen, Y.W.; Shan, B.; Chen, R. Surface modification of LaFeO3 by co-pi electrochemical deposition as an efficient photoanode under visible light. RCS Adv. 2016, 6, 26192–26198. [Google Scholar] [CrossRef]
  73. Pawar, G.S.; Tahir, A.A. Unbiased spontaneous solar fuel production using stable LaFeO3 photoelectrode. Sci. Rep. 2018, 8, 1–9. [Google Scholar] [CrossRef] [PubMed]
  74. Cardona, M. Optical Properties and Band Structure of SrTiO3 and BaTiO3. Phys. Rev. 1965, 140, A651–A655. [Google Scholar] [CrossRef]
  75. Wrighton, M.S.; Ellis, A.B.; Wolczanski, P.T.; Morse, D.L.; Abrahamson, H.B.; Ginley, D.S. Strontium titanate photoelectrodes. Efficient photoassisted electrolysis of water at zero applied potential. J. Am. Chem. Soc. 1976, 98, 2774–2779. [Google Scholar] [CrossRef]
  76. Yu, H.G.; Irie, H.; Hashimoto, K. Conduction band energy level control of titanium dioxide: Toward an efficient visible-light-sensitive photocatalyst. J. Am. Chem. Soc. 2010, 132, 6898–6899. [Google Scholar] [CrossRef] [PubMed]
  77. Ouyang, S.X.; Ye, J.H. β-AgAl1−xGaxO2 solid-solution photocatalysts: Continuous modulation of electronic structure toward high-performance visible-light photoactivity. J. Am. Chem. Soc. 2011, 133, 7757–7763. [Google Scholar] [CrossRef] [PubMed]
  78. Wang, D.F.; Pierre, A.; Kibria, M.G.; Cui, K.; Han, X.G.; Bevan, K.H.; Guo, H.; Paradis, S.; Hakima, A.R.; Mi, Z.T. Wafer-level photocatalytic water splitting on GaN nanowire arrays grown by molecular beam epitaxy. Nano Lett. 2011, 11, 2353–2357. [Google Scholar] [CrossRef] [PubMed]
  79. Nolan, M. Surface modification of TiO2 with metal oxide nanoclusters: A route to composite photocatalytic materials. Chem. Commun. 2011, 47, 8617–8619. [Google Scholar] [CrossRef] [PubMed]
  80. Tada, H.; Jin, Q.; Nishijima, H.; Yamamoto, H.; Fujishima, M.; Okuoka, S.; Hattori, T.; Sumida, Y.; Kobayashi, H. Titanium(iv) dioxide surface-modified with iron oxide as a visible light photocatalyst. Angew. Chem. Int. Ed. 2011, 50, 3501–3505. [Google Scholar] [CrossRef] [PubMed]
  81. Yu, H.; Irie, H.; Shimodaira, Y.; Hosogi, Y.; Kuroda, Y.; Miyauchi, M.; Hashimoto, K. An efficient visible-light-sensitive Fe(iii)-grafted TiO2 photocatalyst. J. Phys. Chem. C 2010, 114, 16481–16487. [Google Scholar] [CrossRef]
  82. Ni, M.; Leung, M.K.H.; Leung, D.Y.C.; Sumathy, K. A review and recent developments in photocatalytic water-splitting using TiO2 for hydrogen production. Renew. Sustain. Energy Rev. 2007, 11, 401–425. [Google Scholar] [CrossRef]
  83. Roy, P.; Berger, S.; Schmuki, P. TiO2 nanotubes: Synthesis and applications. Angew. Chem. Int. Ed. 2011, 50, 2904–2939. [Google Scholar] [CrossRef] [PubMed]
  84. Lee, K.; Mazare, A.; Schmuki, P. One-dimensional titanium dioxide nanomaterials: Nanotubes. Chem. Rev. 2014, 114, 9385–9454. [Google Scholar] [CrossRef] [PubMed]
  85. Tian, J.; Zhao, Z.H.; Kumar, A.; Boughton, R.I.; Liu, H. Recent progress in design, synthesis, and applications of one-dimensional TiO2 nanostructured surface heterostructures: A review. Chem. Soc. Rev. 2014, 43, 6920–6937. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, J.Y.; Zhu, H.L.; Zheng, S.K.; Pan, F.; Wang, T.M. TiO2 film/Cu2O microgrid heterojunction with photocatalytic activity under solar light irradiation. ACS Appl. Mater. Interfaces 2009, 1, 2111–2114. [Google Scholar] [CrossRef] [PubMed]
  87. Zhang, J.; Bang, J.H.; Tang, C.C.; Kamat, P.V. Tailored TiO2-SrTiO3 heterostructure nanotube arrays for improved photoelectrochemical performance. ACS Nano 2010, 4, 387–395. [Google Scholar] [CrossRef] [PubMed]
  88. Sharma, D.; Upadhyay, S.; Satsangi, V.R.; Shrivastav, R.; Waghmare, U.V.; Dass, S. Improved photoelectrochemical water splitting performance of Cu2O/SrTiO3 heterojunction photoelectrode. J. Phys. Chem. C 2014, 118, 25320–25329. [Google Scholar] [CrossRef]
  89. Liu, C.B.; Li, P.; Wu, G.L.; Luo, B.F.; Lin, S.; Ren, A.; Shi, W.D. Enhanced photoelectrochemical and photocatalytic activity by Cu2O/SrTiO3 p-n heterojunction via a facile deposition-precipitation technique. RSC Adv. 2015, 5, 33938–33945. [Google Scholar] [CrossRef]
  90. Arney, D.; Watkins, T.; Maggard, P.A. Effects of particle surface areas and microstructures on photocatalytic H2 and O2 production over PbTiO3. J. Am. Ceram. Soc. 2011, 94, 1483–1489. [Google Scholar] [CrossRef]
  91. Yang, Y.; Wang, X.H.; Zhong, C.F.; Sun, C.K.; Yao, G.F.; Li, L.T. Synthesis and growth mechanism of lead titanate nanotube arrays by hydrothermal method. J. Am. Ceram. Soc. 2008, 91, 3388–3390. [Google Scholar] [CrossRef]
  92. Reddy, K.H.; Parida, K. Fabrication, characterization, and photoelectrochemical properties of Cu-doped PbTiO3 and its hydrogen production activity. ChemCatChem 2013, 5, 3812–3820. [Google Scholar] [CrossRef]
  93. Hu, Y.X.; Dong, W.; Zheng, F.G.; Fang, L.; Shen, M.R. Fe(iii) doped and grafted PbTiO3 film photocathode with enhanced photoactivity for hydrogen production. Appl. Phys. Lett. 2014, 105, 1–4. [Google Scholar] [CrossRef]
  94. Wang, Y.; Dong, W.; Zheng, F.; Fang, L.; Shen, M. Improved Photocathodic Properties of Ferroelectric PbTiO3 Films Using Ag-Pt Bimetallic Catalyst. Energy Environ. Focus 2015, 4, 95–100. [Google Scholar] [CrossRef]
  95. Tabari, T.; Singh, D.; Calisan, A.; Ebadi, M.; Tavakkoli, H.; Caglar, B. Microwave assisted synthesis of La1−xCaxMnO3 (x = 0, 0.2 and 0.4): Structural and capacitance properties. Ceram. Int. 2017, 43, 15970–15977. [Google Scholar] [CrossRef]
  96. Jamali, S.S.; Singh, D.; Tavakkoli, H.; Kaveh, F.; Tabari, T. Microwave-assisted synthesis of nanostructured perovskite-type oxide with efficient photocatalytic activity against organic reactants in gaseous and aqueous phases. Mater. Sci. Semicond. Process 2017, 64, 47–54. [Google Scholar] [CrossRef]
  97. Tabari, T.; Ebadi, M.; Singh, D.; Caglar, B.; Yagci, M.B. Efficient synthesis of perovskite-type oxide photocathode by nonhydrolytic sol-gel method with an enhanced photoelectrochemical activity. J. Alloys Compd. 2018, 750, 248–257. [Google Scholar] [CrossRef]
  98. Tang, P.S.; Chen, H.F.; Cao, F.; Pan, G.X. Magnetically recoverable and visible-light-driven nanocrystalline YFeO3 photocatalysts. Catal. Sci. Technol. 2011, 1, 1145–1148. [Google Scholar] [CrossRef]
  99. Zhang, Y.W.; Yang, J.X.; Xu, J.F.; Gao, Q.Y.; Hong, Z.L. Controllable synthesis of hexagonal and orthorhombic YFeO3 and their visible-light photocatalytic activities. Mater. Lett. 2012, 81, 1–4. [Google Scholar] [CrossRef]
  100. Chen, Y.Y.; Yang, J.; Wang, X.L.; Feng, F.Y.; Zhang, Y.M.; Tang, Y. Synthesis YFeO3 by salt-assisted solution combustion method and its photocatalytic activity. J. Ceram. Soc. Jpn. 2014, 122, 146–150. [Google Scholar] [CrossRef]
  101. Lu, X.M.; Xie, J.M.; Shu, H.M.; Liu, J.; Yin, C.Q.; Lin, J.M. Microwave-assisted synthesis of nanocrystalline YFeO3 and study of its photoactivity. Mater. Sci. Eng. B 2007, 138, 289–292. [Google Scholar] [CrossRef]
  102. Díez-García, M.I.; Celorrio, V.; Calvillo, L.; Tiwari, D.; Gomez, R.; Fermin, D.J. YFeO3 photocathodes for hydrogen evolution. Electrochim. Acta 2017, 246, 365–371. [Google Scholar] [CrossRef]
  103. Glass, A.M.; von der Linde, D.; Auston, D.H.; Negran, T.J. Excited state polarization, bulk photovoltaic effect and the photorefractive effect in electrically polarized media. J. Electron. Mater. 1975, 4, 915–943. [Google Scholar] [CrossRef]
  104. Fridkin, V.M. Bulk photovoltaic effect in noncentrosymmetric crystals. Crystallogr. Rep. 2001, 46, 654–658. [Google Scholar] [CrossRef]
  105. Li, S.; AlOtaibi, B.; Huang, W.; Mi, Z.T.; Serpone, N.; Nechache, R.; Rosei, F. Epitaxial Bi2FeCrO6 multiferroic thin film as a new visible light absorbing photocathode material. Small 2015, 11, 4018–4026. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, M.; Jiang, C.X.; Dong, W.; Zheng, F.G.; Fang, L.; Su, X.D.; Shen, M.R. Composition dependence of the photochemical reduction of Ag+ by as-grown Pb(ZrxTi1−x)O3 films on indium tin oxide electrode. Appl. Phys. Lett. 2013, 103, 1–4. [Google Scholar] [CrossRef]
  107. Wang, C.Y.; Cao, D.W.; Zheng, F.G.; Dong, W.; Fang, L.; Su, X.D.; Shen, M.R. Photocathodic behavior of ferroelectric Pb(Zr,Ti)O3 films decorated with silver nanoparticles. Chem. Commun. 2013, 49, 3769–3771. [Google Scholar] [CrossRef] [PubMed]
  108. Cheng, X.R.; Dong, W.; Zheng, F.G.; Fang, L.; Shen, M.R. Enhanced photocathodic behaviors of Pb(Zr0.20Ti0.80)O3 films on Si substrates for hydrogen production. Appl. Phys. Lett. 2015, 106, 1–4. [Google Scholar] [CrossRef]
  109. Wul, B.; Goldman, J.M. Ferroelectric switching in BaTiO3 ceramics. CR Acad. Sci. URSS 1946, 51, 21. [Google Scholar]
  110. Kennedy, J.H.; Frese, K.W. Photo-oxidation of water at Barium Titanate electrodes. J. Electrochem. Soc. 1976, 123, 1683–1686. [Google Scholar] [CrossRef]
  111. Zielinska, B.; Borowiak-Palen, E.; Kalenczuk, R.J. Photocatalytic hydrogen generation over alkaline-earth titanates in the presence of electron donors. Int. J. Hydrogen Energy 2008, 33, 1797–1802. [Google Scholar] [CrossRef]
  112. Dholam, R.; Patel, N.; Adami, M.; Miotello, A. Hydrogen production by photocatalytic water-splitting using Cr- or Fe-doped TiO2 composite thin films photocatalyst. Int. J. Hydrogen Energy 2009, 39, 5337–5346. [Google Scholar] [CrossRef]
  113. Toupin, J.; Strub, H.; Kressmann, S.; Boudot, M.; Artero, V.; Laberty-Robert, C. Engineering n-p junction for photoelectrochemical hydrogen production. Phys. Chem. Chem. Phys. 2017, 19, 30675–30682. [Google Scholar] [CrossRef] [PubMed]
  114. Maruska, H.P.; Ghosh, A.K. Transition-metal dopants for extending the response of titanate photoelectrolysis anodes. Sol. Energy Mater. Sol. Cells 1979, 1, 237–247. [Google Scholar] [CrossRef]
  115. Upadhyay, R.K.; Sharma, D. Fe doped BaTiO3 sensitized by Fe3O4 nanoparticles for improved photoelectrochemical response. Mater. Res. Express 2018, 5, 1–9. [Google Scholar] [CrossRef]
  116. Upadhyay, S.; Shrivastava, J.; Solanki, A.; Choudhary, S.; Sharma, V.; Kumar, P.; Singh, N.; Satsangi, V.R.; Shrivastav, R.; Waghmare, U.V.; et al. Enhanced photoelectrochemical response of BaTiO3 with Fe doping: Experiments and first-principles analysis. J. Phys. Chem. C 2011, 115, 24373–24380. [Google Scholar] [CrossRef]
  117. Yang, W.G.; Yu, Y.H.; Starr, M.B.; Yin, X.; Li, Z.D.; Kvit, A.; Wang, S.F.; Zhao, P.; Wang, X.D. Ferroelectric polarization-enhanced photoelectrochemical water splitting in TiO2-BaTiO3 core-shell nanowire photoanodes. Nano Lett. 2015, 15, 7574–7580. [Google Scholar] [CrossRef] [PubMed]
  118. Kudo, A.; Omori, K.; Kato, H. A novel aqueous process for preparation of crystal form-controlled and highly crystalline BiVO4 powder from layered vanadates at room temperature and its photocatalytic and photophysical properties. J. Am. Chem. Soc. 1999, 121, 11459–11467. [Google Scholar] [CrossRef]
  119. Park, Y.; McDonald, K.J.; Choi, K.S. Progress in bismuth vanadate photoanodes for use in solar water oxidation. Chem. Soc. Rev. 2013, 42, 2321–2337. [Google Scholar] [CrossRef] [PubMed]
  120. Abdi, F.F.; van de Krol, R. Nature and light dependence of bulk recombination in co-pi-catalyzed BiVO4 photoanodes. J. Phys. Chem. C 2012, 116, 9398–9404. [Google Scholar] [CrossRef]
  121. Tokunaga, S.; Kato, H.; Kudo, A. Selective preparation of monoclinic and tetragonal BiVO4 with scheelite structure and their photocatalytic properties. Chem. Mater. 2001, 13, 4624–4628. [Google Scholar] [CrossRef]
  122. Kim, T.W.; Choi, K.S. Nanoporous BiVO4 Photoanodes with Dual-Layer Oxygen Evolution Catalysts for Solar Water Splitting. Science 2014, 343, 990–994. [Google Scholar] [CrossRef] [PubMed]
  123. Kim, T.W.; Ping, Y.; Galli, G.A.; Choi, K.S. Simultaneous enhancements in photon absorption and charge transport of bismuth vanadate photoanodes for solar water splitting. Nat. Commun. 2015, 6, 1–10. [Google Scholar] [CrossRef] [PubMed]
  124. Qiu, Y.C.; Liu, W.; Chen, W.; Chen, W.; Zhou, G.M.; Hsu, P.C.; Zhang, R.F.; Liang, Z.; Fan, S.S.; Zhang, Y.G.; et al. Efficient solar-driven water splitting by nanocone BiVO4-perovskite tandem cells. Sci. Adv. 2016, 2, 1–8. [Google Scholar] [CrossRef] [PubMed]
  125. Rao, P.M.; Cai, L.L.; Liu, C.; Cho, I.S.; Lee, C.H.; Weisse, J.M.; Yang, P.D.; Zheng, X.L. Simultaneously efficient light absorption and charge separation in WO3/BiVO4 core/shell nanowire photoanode for photoelectrochemical water oxidation. Nano Lett. 2014, 14, 1099–1105. [Google Scholar] [CrossRef] [PubMed]
  126. Pihosh, Y.; Turkevych, I.; Mawatari, K.; Asai, T.; Hisatomi, T.; Uemura, J.; Tosa, M.; Shimamura, K.; Kubota, J.; Domen, K.; et al. Nanostructured WO3/BiVO4 photoanodes for efficient photoelectrochemical water splitting. Small 2014, 10, 3692–3699. [Google Scholar] [CrossRef] [PubMed]
  127. Hong, S.J.; Lee, S.; Jang, J.S.; Lee, J.S. Heterojunction BiVO4/WO3 electrodes for enhanced photoactivity of water oxidation. Energy Environ. Sci. 2011, 4, 1781–1787. [Google Scholar] [CrossRef]
  128. Van, C.N.; Do, T.H.; Chen, J.W.; Tzeng, W.Y.; Tsai, K.A.; Song, H.L.; Liu, H.J.; Lin, Y.C.; Chen, Y.C.; Wu, C.L.; et al. WO3 mesocrystal-assisted photoelectrochemical activity of BiVO4. NPG Asia Mater. 2017, 9. [Google Scholar] [CrossRef]
  129. Kuang, Y.B.; Jia, Q.X.; Nishiyama, H.; Yamada, T.; Kudo, A.; Domen, K. A front-illuminated nanostructured transparent BiVO4 photoanode for > 2% efficient water splitting. Adv. Energy Mater. 2016, 6. [Google Scholar] [CrossRef]
  130. Song, J.; Cha, J.; Lee, M.G.; Jeong, H.W.; Seo, S.; Yoo, J.A.; Kim, T.L.; Lee, J.; No, H.; Kim, D.H.; et al. Template-engineered epitaxial BiVO4 photoanodes for efficient solar water splitting. J. Mater. Chem. A 2017, 5, 18831–18838. [Google Scholar] [CrossRef]
  131. Liu, Z.K.; Yan, F. The application of bismuth-based oxides in organic-inorganic hybrid photovoltaic devices. J. Am. Ceram. Soc. 2012, 95, 1944–1948. [Google Scholar] [CrossRef]
  132. Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  133. Mohapatra, S.K.; Misra, M. Enhanced photoelectrochemical generation of hydrogen from water by 2,6-dihydroxyantraquinone-functionalized titanium dioxide nanotubes. J. Phys. Chem. C 2007, 111, 11506–11510. [Google Scholar] [CrossRef]
  134. Mohapatra, S.K.; Misra, M.; Mahajan, V.K.; Raja, K.S. Design of a Highly Efficient Photoelectrolytic Cell for Hydrogen Generation by Water Splitting:  Application of TiO2−xCx Nanotubes as a Photoanode and Pt/TiO2 Nanotubes as a Cathode. J. Phys. Chem. C 2007, 111, 8677–8685. [Google Scholar] [CrossRef]
  135. Ruan, C.M.; Paulose, M.; Varghese, O.K.; Grimes, C.A. Enhanced photo electrochemical-response in highly ordered TiO2 nanotube-arrays anodized in boric acid containing electrolyte. Sol. Energy Mater. Sol. Cells 2006, 90, 1283–1295. [Google Scholar] [CrossRef]
  136. Mor, G.K.; Shankar, K.; Paulose, M.; Varghese, O.K.; Grimes, C.A. Enhanced Photocleavage of Water Using Titania Nanotube Arrays. Nano Lett. 2005, 5, 191–195. [Google Scholar] [CrossRef] [PubMed]
  137. Haeni, J.H.; Irvin, P.; Chang, W.; Uecker, R.; Reiche, P.; Li, Y.L.; Choudhury, S.; Tian, W.; Hawley, M.E.; Craigo, B.; et al. Room-temperature ferroelectricity in strained SrTiO3. Nature 2004, 430, 758–761. [Google Scholar] [CrossRef] [PubMed]
  138. Lee, J.H.; Selloni, A. TiO2/Ferroelectric Heterostructures as Dynamic Polarization-Promoted Catalysts for Photochemical and Electrochemical Oxidation of Water. Phys. Rev. Lett. 2014, 112, 196102. [Google Scholar] [CrossRef] [PubMed]
  139. Wysmulek, K.; Sar, J.; Osewski, P.; Orlinski, K.; Kolodziejak, K.; Trenczek-Zajac, A.; Radecka, M.; Pawlak, D.A. A SrTiO3-TiO2 eutectic composite as a stable photoanode material for photoelectrochemical hydrogen production. Appl. Catal. B 2017, 206, 538–546. [Google Scholar] [CrossRef]
  140. Kawasaki, S.; Nakatsuji, K.; Yoshinobu, J.; Komori, F.; Takahashi, R.; Lippmaa, M.; Mase, K.; Kudo, A. Epitaxial Rh-doped SrTiO3 thin film photocathode for water splitting under visible light irradiation. Appl. Phys. Lett. 2012, 101, 1–4. [Google Scholar] [CrossRef]
  141. Jiao, Z.B.; Chen, T.; Xiong, J.Y.; Wang, T.; Lu, G.X.; Ye, J.H.; Bi, Y.P. Visible-light-driven photoelectrochemical and photocatalytic performances of Cr-doped SrTiO3/TiO2 heterostructured nanotube arrays. Sci. Rep. 2013, 3, 1–6. [Google Scholar] [CrossRef] [PubMed]
  142. Modak, B.; Ghosh, S.K. Enhancement of Visible Light Photocatalytic Activity of SrTiO3: A Hybrid Density Functional Study. J. Phys. Chem. C 2015, 119, 23503–23514. [Google Scholar] [CrossRef]
  143. Sangle, A.L.; Singh, S.; Jian, J.; Bajpe, S.R.; Wang, H.Y.; Khare, N.; MacManus-Driscoll, J.L. Very high surface area mesoporous thin films of SrTiO3 grown by pulsed laser deposition and application to efficient photoelectrochemical water splitting. Nano Lett. 2016, 16, 7338–7345. [Google Scholar] [CrossRef] [PubMed]
  144. Cen, J.J.; Wu, Q.Y.; Yan, D.H.; Tao, J.; Kisslinger, K.; Liu, M.Z.; Orlov, A. Photoelectrochemical water splitting with a SrTiO3:Nb/SrTiO3 n+-n homojunction structure. Phys. Chem. Chem. Phys. 2017, 19, 2760–2767. [Google Scholar] [CrossRef] [PubMed]
  145. Kawasaki, S.; Takahashi, R.; Yamamoto, T.; Kobayashi, M.; Kumigashira, H.; Yoshinobu, J.; Komori, F.; Kudo, A.; Lippmaa, M. Photoelectrochemical water splitting enhanced by self-assembled metal nanopillars embedded in an oxide semiconductor photoelectrode. Nat. Commun. 2016, 7, 1–6. [Google Scholar] [CrossRef] [PubMed]
  146. Wu, F.; Yu, Y.H.; Yang, H.; German, L.N.; Li, Z.Q.; Chen, J.G.; Yang, W.G.; Huang, L.; Shi, W.M.; Wang, L.J.; et al. Simultaneous Enhancement of Charge Separation and Hole Transportation in a TiO2-SrTiO3 Core-Shell Nanowire Photoelectrochemical System. Adv. Mater. 2017, 29, 1–7. [Google Scholar] [CrossRef] [PubMed]
  147. Scanlon, D.O.; Dunnill, C.W.; Buckeridge, J.; Shevlin, S.A.; Logsdail, A.J.; Woodley, S.M.; Catlow, C.R.A.; Powell, M.J.; Palgrave, R.G.; Parkin, I.P.; et al. Band alignment of rutile and anatase TiO2. Nat. Mater. 2013, 12, 798–801. [Google Scholar] [CrossRef] [PubMed]
  148. Marschall, R. Semiconductor Composites: Strategies for Enhancing Charge Carrier Separation to Improve Photocatalytic Activity. Adv. Funct. Mater. 2014, 24, 2421–2440. [Google Scholar] [CrossRef]
  149. Cao, D.W.; Wang, Z.J.; Nasori Wen, L.Y.; Mi, Y.; Lei, Y. Switchable Charge-Transfer in the Photoelectrochemical Energy-Conversion Process of Ferroelectric BiFeO3 Photoelectrodes. Angew. Chem. Int. Ed. 2014, 126, 11027–11031. [Google Scholar] [CrossRef] [PubMed]
  150. Ji, W.; Yao, K.; Liang, Y.C. Bulk Photovoltaic Effect at Visible Wavelength in Epitaxial Ferroelectric BiFeO3 Thin Films. Adv. Mater. 2010, 22, 1763–1766. [Google Scholar] [CrossRef] [PubMed]
  151. Bhatnagar, A.; Chaudhuri, A.R.; Kim, Y.H.; Hesse, D.; Alexe, M. Role of domain walls in the abnormal photovoltaic effect in BiFeO3. Nat. Commun. 2013, 4, 1–8. [Google Scholar] [CrossRef] [Green Version]
  152. Jun, H.; Im, B.; Kim, J.Y.; Im, Y.O.; Jang, J.W.; Kim, E.S.; Kim, J.Y.; Kang, H.J.; Hong, S.J.; Lee, J.S. Photoelectrochemical water splitting over ordered honeycomb hematite electrodes stabilized by alumina shielding. Energy Environ. Sci. 2012, 5, 6375–6382. [Google Scholar] [CrossRef] [Green Version]
  153. Kim, J.Y.; Magesh, G.; Youn, D.H.; Jang, J.W.; Kubota, J.; Domen, K.; Lee, J.S. Single-crystalline, wormlike hematite photoanodes for efficient solar water splitting. Sci. Rep. 2013, 3, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Zhong, M.; Ma, Y.H.; Oleynikov, P.; Domen, K.; Delaunay, J.J. A conductive ZnO–ZnGaON nanowire-array-on-a-film photoanode for stable and efficient sunlight water splitting. Energy Environ. Sci. 2014, 7, 1693–1699. [Google Scholar] [CrossRef]
  155. Hou, Y.; Zuo, F.; Dagg, A.; Feng, P.Y. A Three-Dimensional Branched Cobalt-Doped α-Fe2O3 Nanorod/MgFe2O4 Heterojunction Array as a Flexible Photoanode for Efficient Photoelectrochemical Water Oxidation. Angew. Chem. Int. Ed. 2013, 52, 1248–1252. [Google Scholar] [CrossRef] [PubMed]
  156. Kumar, S.G.; Rao, K.S.R.K. Physics and chemistry of CdTe/CdS thin film heterojunction photovoltaic devices: Fundamental and critical aspects. Energy Environ. Sci. 2014, 7, 45–102. [Google Scholar] [CrossRef]
  157. Wang, M.Y.; Sun, L.; Lin, Z.Q.; Cai, J.H.; Xie, K.P.; Lin, C.J. p–n Heterojunction photoelectrodes composed of Cu2O-loaded TiO2 nanotube arrays with enhanced photoelectrochemical and photoelectrocatalytic activities. Energy Environ. Sci. 2013, 6, 1211–1220. [Google Scholar] [CrossRef]
  158. Jang, J.S.; Hwang, D.W.; Lee, J.S. CdS-AgGaS2 photocatalytic diodes for hydrogen production from aqueous Na2S/Na2SO3 electrolyte solution under visible light (lambda ≥ 420 nm). Catal. Today 2007, 120, 174–181. [Google Scholar] [CrossRef]
  159. Jang, J.S.; Ahn, C.W.; Won, S.S.; Kim, J.H.; Choi, W.; Lee, B.S.; Yoon, J.H.; Kim, H.G.; Lee, J.S. Vertically Aligned Core–Shell PbTiO3@TiO2 Heterojunction Nanotube Array for Photoelectrochemical and Photocatalytic Applications. J. Phys. Chem. C 2017, 121, 15063–15070. [Google Scholar] [CrossRef]
  160. Ahn, C.W.; Borse, P.H.; Kim, J.H.; Kim, J.Y.; Jang, J.S.; Cho, C.R.; Yoon, J.H.; Lee, B.S.; Bae, J.S.; Kim, H.G.; et al. Effective charge separation in site-isolated Pt-nanodot deposited PbTiO3 nanotube arrays for enhanced photoelectrochemical water splitting. Appl. Catal. B 2018, 224, 804–809. [Google Scholar] [CrossRef]
  161. Li, C.M.; Ahmed, T.; Ma, M.G.; Edvinsson, T.; Zhu, J.F. A facile approach to Zno/Cds nanoarrays and their photocatalytic and photoelectrochemical properties. Appl. Catal. B 2013, 138, 175–183. [Google Scholar] [CrossRef]
  162. Liu, Z.; Wang, Y.; Wang, B.; Li, Y.B.; Liu, Z.C.; Han, J.H.; Guo, K.Y.; Li, Y.J.; Cui, T.; Han, L.; et al. PEC electrode of ZnO nanorods sensitized by CdS with different size and its photoelectric properties. Int. J. Hydrogen Energy 2013, 38, 10226–10234. [Google Scholar] [CrossRef]
  163. Xu, Y.; Zhao, W.W.; Xu, R.; Shi, Y.M.; Zhang, B. Synthesis of ultrathin CdS nanosheets as efficient visible-light-driven water splitting photocatalysts for hydrogen evolution. Chem. Commun. 2013, 49, 9803–9805. [Google Scholar] [CrossRef] [PubMed]
  164. Pareek, A.; Purbia, R.; Paik, P.; Hebalkar, N.Y.; Kim, H.G.; Borse, P.H. Stabilizing effect in nano-titania functionalized CdS photoanode for sustained hydrogen generation. Int. J. Hydrogen Energy 2014, 39, 4170–4180. [Google Scholar] [CrossRef]
  165. Pareek, A.; Paik, P.; Borse, P.H. Nanoniobia Modification of CdS Photoanode for an Efficient and Stable Photoelectrochemical Cell. Langmuir 2014, 30, 15540–15549. [Google Scholar] [CrossRef] [PubMed]
  166. Han, S.C.; Pu, Y.C.; Zheng, L.X.; Hu, L.F.; Zhang, J.Z.; Fang, X.S. Uniform carbon-coated CdS core–shell nanostructures: Synthesis, ultrafast charge carrier dynamics, and photoelectrochemical water splitting. J. Mater. Chem. A 2016, 4, 1078–1086. [Google Scholar] [CrossRef]
  167. Zhao, K.; Yan, X.Q.; Gu, Y.S.; Kang, Z.; Bai, Z.M.; Cao, S.Y.; Liu, Y.C.; Zhang, X.H.; Zhang, Y. Self-Powered Photoelectrochemical Biosensor Based on CdS/RGO/ZnO Nanowire Array Heterostructure. Small 2016, 12, 245–251. [Google Scholar] [CrossRef] [PubMed]
  168. Wang, G.M.; Yang, X.Y.; Qian, F.; Zhang, J.Z.; Li, Y. Double-Sided CdS and CdSe Quantum Dot Co-Sensitized ZnO Nanowire Arrays for Photoelectrochemical Hydrogen Generation. Nano Lett. 2010, 10, 1088–1092. [Google Scholar] [CrossRef] [PubMed]
  169. Bu, Y.Y.; Chen, Z.Y.; Li, W.B.; Yu, J.Q. High-efficiency photoelectrochemical properties by a highly crystalline CdS-sensitized ZnO nanorod array. ACS Appl. Mater. Interfaces 2013, 5, 5097–5104. [Google Scholar] [CrossRef] [PubMed]
  170. Yuan, Y.B.; Xiao, Z.G.; Yang, B.; Huang, J.S. Arising applications of ferroelectric materials in photovoltaic devices. J. Mater. Chem. A 2014, 2, 6027–6041. [Google Scholar] [CrossRef] [Green Version]
  171. Wei, R.B.; Kuang, P.Y.; Cheng, H.; Chen, Y.B.; Long, J.Y.; Zhang, M.Y.; Liu, Z.Q. Plasmon-Enhanced Photoelectrochemical Water Splitting on Gold Nanoparticle Decorated ZnO/CdS Nanotube Arrays. ACS Sustain. Chem. Eng. 2017, 5, 4249–4257. [Google Scholar] [CrossRef]
  172. Guo, X.; Zhu, J.J.; Wei, H.X.; Lee, S.T.; Li, Y.Q.; Tang, J.X. Facile approaching hierarchical CdS films as electrode toward photoelectrochemical water splitting. Nanotechnology 2015, 26, 1–9. [Google Scholar] [CrossRef] [PubMed]
  173. Zhang, Z.M.; Gao, C.T.; Wu, Z.M.; Han, W.H.; Wang, Y.L.; Fu, W.B.; Li, X.D.; Xie, E.Q. Toward efficient photoelectrochemical water-splitting by using screw-like SnO2 nanostructures as photoanode after being decorated with CdS quantum dots. Nano Energy 2016, 19, 318–327. [Google Scholar] [CrossRef]
  174. Tran, T.K.; Park, W.; Tong, W.; Kyi, M.M.; Wagner, B.K.; Summers, C.J. Photoluminescence properties of ZnS epilayers. J. Appl. Phys. 1997, 81, 2803–2809. [Google Scholar] [CrossRef]
  175. Zhang, J.; Yu, J.G.; Jaroniec, M.; Gong, J.R. Noble Metal-Free Reduced Graphene Oxide-ZnxCd1−xS Nanocomposite with Enhanced Solar Photocatalytic H2-Production Performance. Nano Lett. 2012, 12, 4584–4589. [Google Scholar] [CrossRef] [PubMed]
  176. Hong, Y.P.; Zhang, J.; Wang, X.; Wang, Y.J.; Lin, Z.; Yu, J.G.; Huang, F. Influence of lattice integrity and phase composition on the photocatalytic hydrogen production efficiency of ZnS nanomaterials. Nanoscale 2012, 4, 2859–2862. [Google Scholar] [CrossRef] [PubMed]
  177. Meier, J.H.; Meier, K. Photochemical production of hydrogen with Zinc Sulfide suspensions. J. Phys. Chem. 1984, 88, 5903–5913. [Google Scholar]
  178. Hu, J.S.; Ren, L.L.; Guo, Y.G.; Liang, H.P.; Cao, A.M.; Wan, L.J.; Bai, C.L. Mass Production and High Photocatalytic Activity of ZnS Nanoporous Nanoparticles. Angew. Chem. Int. Ed. 2005, 44, 1269–1273. [Google Scholar] [CrossRef] [PubMed]
  179. Youn, H.C.; Barral, S.; Fendler, J.H. Dihexadecyl Phosphate, Vesicle-Stabilized and in Situ Generated Mixed CdS and ZnS Semiconductor Particles. Preparation and Utilization for Photosensitized Charge Separation and Hydrogen Generation. J. Phys. Chem. 1988, 92, 6320–6327. [Google Scholar] [CrossRef]
  180. Kudo, A.; Sekizawa, M. Photocatalytic H2 evolution under visible light irradiation on Zn1−xCuxS solid solution. Catal. Lett. 1999, 58, 241–243. [Google Scholar] [CrossRef]
  181. Kudo, A.; Sekizawa, M. Photocatalytic H2 evolution under visible light irradiation on Ni-doped ZnS photocatalyst. Chem. Commun. 2000, 0, 1371–1372. [Google Scholar] [CrossRef]
  182. Hassan, M.A.; Kang, J.H.; Johar, M.A.; Ha, J.S.; Ryu, S.W. High-performance ZnS/GaN heterostructure photoanode for photoelectrochemical water splitting applications. Acta Mater. 2018, 146, 171–175. [Google Scholar] [CrossRef]
  183. Herrmann, J.M.; Disdier, J.; Pichat, P. Effect of chromium doping on the electrical and catalytic properties of powder titania under UV and visible illumination. Chem. Phys. Lett. 1984, 108, 618–622. [Google Scholar] [CrossRef]
  184. Kaneva, N.V.; Dimitrov, D.T.; Dushkin, C.D. Effect of nickel doping on the photocatalytic activity of ZnO thin films under UV and visible light. Appl. Surf. Sci. 2011, 257, 8113–8120. [Google Scholar] [CrossRef]
  185. Bellam, J.B.; Ruiz-Preciado, M.A.; Edely, M.; Szade, J.; Jouanneaux, A.; Kassiba, A.H. Visible-light photocatalytic activity of nitrogen-doped NiTiO3 thin films prepared by a co-sputtering process. RSC Adv. 2015, 5, 10551–10559. [Google Scholar] [CrossRef]
  186. Hajjaji, A.; Trabelsi, K.; Atyaoui, A.; Gaidi, M.; Bousselmi, L.; Bessais, B.; El Khakani, M.A. Photocatalytic activity of Cr-doped TiO2 nanoparticles deposited on porous multicrystalline silicon films. Nanoscale Res. Lett. 2014, 9, 1–6. [Google Scholar] [CrossRef] [PubMed]
  187. Kurnia, F.; Ng, Y.H.; Amal, R.; Valanoor, N.; Hart, J.N. Defect engineering of ZnS thin films for photoelectrochemical water-splitting under visible light. Sol. Energy Mater. Sol. Cells C 2016, 153, 179–185. [Google Scholar] [CrossRef]
  188. Rai, S.C.; Wang, K.; Ding, Y.; Marmon, J.K.; Bhatt, M.; Zhang, Y.; Zhou, W.L.; Wang, Z.L. Piezo-phototronic Effect Enhanced UV/Visible Photodetector Based on Fully Wide Band Gap Type-II ZnO/ZnS Core/Shell Nanowire Array. ACS Nano 2015, 9, 6419–6427. [Google Scholar] [CrossRef] [PubMed]
  189. Liu, C.H.; Qiu, Y.Y.; Wang, F.; Li, L.Z.; Liang, Q.; Chen, Z.D. Electrodeposition of ZnO nanoflake-based photoanode sensitized by carbon quantum dots for photoelectrochemical water oxidation. Ceram. Int. 2017, 43, 5329–5333. [Google Scholar] [CrossRef]
  190. Tian, W.; Zhang, C.; Zhai, T.; Li, S.L.; Wang, X.; Liu, J.W.; Jie, X.; Liu, D.Q.; Liao, M.Y.; Koide, Y.S.; et al. Flexible Ultraviolet Photodetectors with Broad Photoresponse Based on Branched ZnS-ZnO Heterostructure Nanofilms. Adv. Mater. 2014, 26, 3088–3093. [Google Scholar] [CrossRef] [PubMed]
  191. Wang, M.D.; Chen, C.X.; Qin, H.Y.; Zhang, L.; Fang, Y.T.; Liu, J.B.; Meng, L. Construction of FeS2-Sensitized ZnO@ZnS Nanorod Arrays with Enhanced Optical and Photoresponse Performances. Adv. Mater. Interfaces 2015, 2, 1–10. [Google Scholar] [CrossRef]
  192. Tian, W.; Zhai, T.Y.; Zhang, C.; Li, S.L.; Wang, X.; Liu, F.; Liu, D.Q.; Cai, X.K.; Tsukagoshi, K.; Golberg, D.; et al. Low-Cost Fully Transparent Ultraviolet Photodetectors Based on Electrospun ZnO-SnO2 Heterojunction Nanofibers. Adv. Mater. 2013, 25, 4625–4630. [Google Scholar] [CrossRef] [PubMed]
  193. Kushwaha, A.; Aslam, M. Zns shielded ZnO nanowire photoanodes for efficient water splitting. Electrochim. Acta 2014, 130, 222–231. [Google Scholar] [CrossRef]
  194. Wu, X.; Jiang, P.; Ding, Y.; Cai, W.; Xie, S.; Wang, Z.L. Mismatch strain induced formation of ZnO/ZnS heterostructured rings. Adv. Mater. 2007, 19, 2319–2323. [Google Scholar] [CrossRef]
  195. Sánchez-Tovar, R.; Fernández-Domene, R.M.; Montañés, M.T.; Sanz-Marco, A.; Garcia-Antón, J. ZnO/ZnS heterostructures for hydrogen production by photoelectrochemical water splitting. RSC Adv. 2016, 6, 30425–30435. [Google Scholar] [CrossRef] [Green Version]
  196. Liu, Y.C.; Gu, Y.S.; Yan, X.Q.; Kang, Z.; Lu, S.N.; Sun, Y.H.; Zhang, Y. Design of sandwich-structured ZnO/ZnS/Au photoanode for enhanced efficiency of photoelectrochemical water splitting. Nano Res. 2015, 8, 2891–2900. [Google Scholar] [CrossRef]
  197. Liu, C.H.; Qiu, Y.Y.; Wang, F.; Wang, K.; Liang, Q.; Chen, Z.D. Design of Core–Shell-Structured ZnO/ZnS Hybridized with Graphite-Like C3N4 for Highly Efficient Photoelectrochemical Water Splitting. Adv. Mater. Interfaces 2017, 4, 1–11. [Google Scholar] [CrossRef]
  198. Huang, J.D.; Liu, J.Y.; Han, K.L. Hybrid functionals studies of structural and electronic properties of ZnxCd((1−x))S and (ZnxCd1−x)(SexS1−x) solid solution photocatalysts. Int. J. Hydrogen Energy 2012, 37, 17870–17881. [Google Scholar] [CrossRef]
  199. Jang, G.G.; Jacobs, C.B.; Ivanov, I.N.; Joshi, P.C.; Iii, H.M.M.; Kidder, M.; Armstrong, B.L.; Datskos, P.G.; Graham, D.E.; Moon, J.W. In situ capping for size control of monochalcogenide (ZnS, CdS and SnS) nanocrystals produced by anaerobic metal-reducing bacteria. Nanotechnology 2015, 26, 1–5. [Google Scholar] [CrossRef] [PubMed]
  200. Qutub, N.; Pirzada, B.M.; Umar, K.; Mehraj, O.; Muneer, M.; Sabir, S. Synthesis, characterization and visible-light driven photocatalysis by differently structured CdS/ZnS sandwich and core-shell nanocomposites. Physica E 2015, 74, 74–86. [Google Scholar] [CrossRef]
  201. Xu, J.Y. Preparation of ZnS-CdS nanocomposite for photoelectrochemical hydrogen production. Int. J. Electrochem. Sci. 2017, 12, 2253–2261. [Google Scholar] [CrossRef]
  202. Zhang, P.; Guan, B.Y.; Yu, L.; Lou, X.W. Facile Synthesis of Multi-shelled ZnS-CdS Cages with Enhanced Photoelectrochemical Performance for Solar Energy Conversion. CHEM 2018, 4, 162–173. [Google Scholar] [CrossRef]
  203. Im, J.H.; Jang, I.H.; Pellet, N.; Gratzel, M.; Park, N.G. Growth of CH3NH3PbI3 cuboids with controlled size for high-efficiency perovskite solar cells. Nat. Nanotechnol. 2014, 9, 927–932. [Google Scholar] [CrossRef] [PubMed]
  204. Burschka, J.; Pellet, N.; Moon, S.J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M.K.; Gratzel, M. Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature 2013, 499, 316–320. [Google Scholar] [CrossRef] [PubMed]
  205. Mei, A.; Li, X.; Liu, L.F.; Ku, Z.L.; Liu, T.F.; Rong, Y.G.; Xu, M.; Hu, M.; Chen, J.Z.; Yang, Y.; et al. A hole-conductor-free, fully printable mesoscopic perovskite solar cell with high stability. Science 2014, 345, 295–298. [Google Scholar] [CrossRef] [PubMed]
  206. Li, L.; Shi, C.W.; Deng, X.L.; Wang, Y.Q.; Ni, L.L. High-crystallinity and large-grain CH3NH3PbI3 thin films for efficient TiO2 nanorod. IET Micro Nano Lett. 2018, 13, 131–134. [Google Scholar] [CrossRef]
  207. Dong, Q.F.; Fang, Y.J.; Shao, Y.C.; Mulligan, P.; Qiu, J.; Cao, L.; Huang, J.S. Electron-hole diffusion lengths > 175 μm in solution-grown CH3NH3PbI3 single crystals. Science 2015, 347, 967–970. [Google Scholar] [CrossRef] [PubMed]
  208. NREL. Available online: https://www.nrel.gov/pv/assets/images/efficiency-chart.png 2016 (accessed on 4 July 2017).
  209. Frost, J.M.; Butler, K.T.; Brivio, F.; Hendon, C.H.; van Schilfgaarde, M.; Walsh, A. Atomistic Origins of High-Performance in Hybrid Halide Perovskite Solar Cells. Nano Lett. 2014, 14, 2584–2590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Baikie, T.; Fang, Y.N.; Kadro, J.M.; Schreyer, M.; Wei, F.X.; Mhaisalkar, S.G.; Graetzel, M.; White, T.J. Synthesis and crystal chemistry of the hybrid perovskite (CH3NH3)PbI3 for solid-state sensitised solar cell applications. J. Mater. Chem. A 2013, 1, 5628–5641. [Google Scholar] [CrossRef]
  211. Luo, J.S.; Im, J.H.; Mayer, M.T.; Schreier, M.; Nazeeruddin, M.K.; Park, N.G.; Tilley, S.D.; Fan, H.J.; Gratzel, M. Water photolysis at 12.3% efficiency via perovskite photovoltaics and Earth-abundant catalysts. Science 2014, 345, 1593–1596. [Google Scholar] [CrossRef] [PubMed]
  212. Da, P.M.; Cha, M.Y.; Sun, L.; Wu, Y.Z.; Wang, Z.S.; Zheng, G.F. High-performance perovskite photoanode enabled by Ni passivation and catalysis. Nano Lett. 2015, 15, 3452–3457. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The solutions for solar hydrogen via water splitting. (a) Particulate photocatalytic (PC) water splitting system, (b) Photoelectrochemical (PEC) water splitting system and (c) Photovoltaic–photoelectrochemical hybrid (PV–PEC) system [10]. Copyright 2017, Elsevier.
Figure 1. The solutions for solar hydrogen via water splitting. (a) Particulate photocatalytic (PC) water splitting system, (b) Photoelectrochemical (PEC) water splitting system and (c) Photovoltaic–photoelectrochemical hybrid (PV–PEC) system [10]. Copyright 2017, Elsevier.
Applsci 08 01526 g001
Figure 2. Principle of photocatalytic water-splitting (PC) reactions [7]. Copyright 2014, Elsevier.
Figure 2. Principle of photocatalytic water-splitting (PC) reactions [7]. Copyright 2014, Elsevier.
Applsci 08 01526 g002
Figure 3. Schematic diagram of photoelectrochemical (PEC) photovoltaic (PV)–PEC hybrid water splitting system; (a) PV/PEC hybrid cell, (b) A photoelectrode PEC cell with anode and cathode combined in parallel, and (c) Two photoelectrodes combined in series [3]. Copyright 2010, Royal Society of Chemistry.
Figure 3. Schematic diagram of photoelectrochemical (PEC) photovoltaic (PV)–PEC hybrid water splitting system; (a) PV/PEC hybrid cell, (b) A photoelectrode PEC cell with anode and cathode combined in parallel, and (c) Two photoelectrodes combined in series [3]. Copyright 2010, Royal Society of Chemistry.
Applsci 08 01526 g003
Figure 4. Conventional models for the polycrystalline anomalous photovoltaic effect [27]. Copyright 2015, Royal Society of Chemistry.
Figure 4. Conventional models for the polycrystalline anomalous photovoltaic effect [27]. Copyright 2015, Royal Society of Chemistry.
Applsci 08 01526 g004
Figure 5. The bandgap of various materials including ferroelectrics and the semiconductor.
Figure 5. The bandgap of various materials including ferroelectrics and the semiconductor.
Applsci 08 01526 g005
Figure 6. Photocurrent potential (J–V) curves of different BiFeO3 (BFO)-based photocathodes before (a) and after (b) positively poling the BFO film [53]. Copyright 2017, Royal Society of Chemistry.
Figure 6. Photocurrent potential (J–V) curves of different BiFeO3 (BFO)-based photocathodes before (a) and after (b) positively poling the BFO film [53]. Copyright 2017, Royal Society of Chemistry.
Applsci 08 01526 g006
Figure 7. (a) Energy band diagram of Cu2O and SrTiO3 before and after formation of the p–n junction, (b) Photocurrent density vs. applied potential curve for (A) pristine Cu2O and (C) 343 nm of Cu2O/SrTiO3 heterojunction [88]. Copyright 2014, American Chemical Society.
Figure 7. (a) Energy band diagram of Cu2O and SrTiO3 before and after formation of the p–n junction, (b) Photocurrent density vs. applied potential curve for (A) pristine Cu2O and (C) 343 nm of Cu2O/SrTiO3 heterojunction [88]. Copyright 2014, American Chemical Society.
Applsci 08 01526 g007
Figure 8. (a) Variations of the current density with applied voltage and (b) Simplified energy band diagrams of the PEC cell based on Bi2FeCrO6 (BFCO) thin film without polarization (left) and either negatively poled (middle) or positively poled (right) [105]. Copyright 2015, Wiley Online Library.
Figure 8. (a) Variations of the current density with applied voltage and (b) Simplified energy band diagrams of the PEC cell based on Bi2FeCrO6 (BFCO) thin film without polarization (left) and either negatively poled (middle) or positively poled (right) [105]. Copyright 2015, Wiley Online Library.
Applsci 08 01526 g008
Figure 9. SrTiO3–TiO2 eutectic composite-based PEC cell. (a) A simplified PEC cell employing a eutectic system as the photoactive anode material. (b) Charge-carrier separation mechanism at the phase boundary, with bandgap positions and values with respect to electrochemical potential (E) vs. NHE based on [139,147,148]. Figure 9 [139] from Copyright 2017 Elsevier.
Figure 9. SrTiO3–TiO2 eutectic composite-based PEC cell. (a) A simplified PEC cell employing a eutectic system as the photoactive anode material. (b) Charge-carrier separation mechanism at the phase boundary, with bandgap positions and values with respect to electrochemical potential (E) vs. NHE based on [139,147,148]. Figure 9 [139] from Copyright 2017 Elsevier.
Applsci 08 01526 g009
Figure 10. (a) P–E hysteresis loop for a 400-nm-thick epitaxial BiFeO3 thin film with (001)pc BFO at 10 kHz and 50 kHz frequencies and (110)pc, (111)pc BFO thin film at 10 kHz. (b) Polarization switching measurement of the 50-nm thick (111)pc BFO thin film photoanodes in different polarization states. The inset shows the magnified portion for the onset potential [4]. Copyright 2018, Springer.
Figure 10. (a) P–E hysteresis loop for a 400-nm-thick epitaxial BiFeO3 thin film with (001)pc BFO at 10 kHz and 50 kHz frequencies and (110)pc, (111)pc BFO thin film at 10 kHz. (b) Polarization switching measurement of the 50-nm thick (111)pc BFO thin film photoanodes in different polarization states. The inset shows the magnified portion for the onset potential [4]. Copyright 2018, Springer.
Applsci 08 01526 g010
Figure 11. Combination of the perovskite tandem cell with NiFe DLH/Ni foam electrodes for water splitting, (a) Schematic diagram of the water-splitting device, (b) A generalized energy schematic of the perovskite tandem cell for water splitting [211]. Copyright from 2014, American Association for the Advancement of Science.
Figure 11. Combination of the perovskite tandem cell with NiFe DLH/Ni foam electrodes for water splitting, (a) Schematic diagram of the water-splitting device, (b) A generalized energy schematic of the perovskite tandem cell for water splitting [211]. Copyright from 2014, American Association for the Advancement of Science.
Applsci 08 01526 g011

Share and Cite

MDPI and ACS Style

Kim, S.; Nguyen, N.T.; Bark, C.W. Ferroelectric Materials: A Novel Pathway for Efficient Solar Water Splitting. Appl. Sci. 2018, 8, 1526. https://doi.org/10.3390/app8091526

AMA Style

Kim S, Nguyen NT, Bark CW. Ferroelectric Materials: A Novel Pathway for Efficient Solar Water Splitting. Applied Sciences. 2018; 8(9):1526. https://doi.org/10.3390/app8091526

Chicago/Turabian Style

Kim, Sangmo, Nguyen Thi Nguyen, and Chung Wung Bark. 2018. "Ferroelectric Materials: A Novel Pathway for Efficient Solar Water Splitting" Applied Sciences 8, no. 9: 1526. https://doi.org/10.3390/app8091526

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