Next Article in Journal
Optimizing Algal Oil Extraction and Transesterification Parameters through RSM, PCA, and MRA for Sustainable Biodiesel Production
Previous Article in Journal
Recent Advances in Novel Catalytic Hydrodeoxygenation Strategies for Biomass Valorization without Exogenous Hydrogen Donors—A Review
Previous Article in Special Issue
A Numerical Case Study of Particle Flow and Solar Radiation Transfer in a Compound Parabolic Concentrator (CPC) Photocatalytic Reactor for Hydrogen Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Strategies to Improve the Photocatalytic Efficiency of TiO2 for Enhanced Water Splitting to Produce Hydrogen

1
Department of Chemistry, University of Lahore, Sargodha Campus, Sargodha 40100, Pakistan
2
School of Chemistry and Chemical Engineering, Southwest University, Chongqing 400715, China
3
Department of Chemistry, Dong-A University, Busan 49315, Republic of Korea
4
Department of Chemistry, Government College University, Faisalabad 38000, Pakistan
5
Department of Chemistry, College of Science, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 11623, Saudi Arabia
6
Department of Environmental Sciences, The University of Lahore, Lahore 54000, Pakistan
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2024, 14(10), 674; https://doi.org/10.3390/catal14100674
Submission received: 25 June 2024 / Revised: 5 September 2024 / Accepted: 18 September 2024 / Published: 30 September 2024
(This article belongs to the Special Issue New Advances in Photocatalytic Hydrogen Production)

Abstract

:
Hydrogen production is one of the best solutions to the growing energy concerns, owing to its clean and sustainable assets. The current review gives an overview of various hydrogen production technologies, highlighting solar water splitting as a promising approach for its sustainable production. Moreover, it gives a detailed mechanism of the water-splitting reaction and describes the significance of titania-based catalysts for solar water splitting. It further highlights diversified strategies to improve the catalytic efficiency of TiO2 for the enhanced hydrogen production. These strategies include the doping of TiO2, dye sensitization, and the addition of co-catalysts. Doping reduces the bandgap by generating new energy levels in TiO2 and encourages visible-light absorption. Sensitization with dyes tunes the electronic states, which in turn broadens the light-absorption capacity of titania. Constructing heterojunctions reduces the charge recombination of TiO2, while co-catalysts increase the number of active sites for an enhanced reaction rate. Thus, every modification strategy has a positive impact on the stability and photocatalytic efficiency of TiO2 for improved water splitting. Lastly, this review provides a comprehensive description and future outlook for developing efficient catalysts to enhance the hydrogen production rate, thereby fulfilling the energy needs of the industrial sector.

1. Introduction

Due to greenhouse gases, climate change, decreasing fossil fuels, and increasing energy requirements, we will need a transition from a society that depends on non-renewable fuels to one that is more reliant on renewable energy in the upcoming era [1]. It is alluring to search for an alternate energy source that can replace conventional fossil fuels while not polluting the environment. Although photovoltaic cells reveal an ideal solution to the energy problem by offering a carbon-free solution, they lack a cost-effective and efficient energy-storing process that could supply energy at nighttime [2]. On the other hand, converting solar fuel into its chemical form can be a better approach for solving energy problems. In this regard, hydrogen (H2) is an excellent source of clean, highly combustible, and renewable energy. H2 energy can be produced in a variety of ways (Figure 1) [3], can be stored in different forms (gaseous, liquid, and metal hydrides), and can be transformed into a variety of fuels. The ecological effects of the energy zone are reduced because no carbon is emitted when H2 is produced from water (H2O) or any other renewable resource [4,5].

2. Hydrogen Production Methods

H2 can be produced through a variety of methods, such as biological, thermochemical, steam reforming, and coal gasification [3]. Since methane gas has the largest H2/C ratio when compared to other hydrocarbons, it has a lower risk of byproduct formation. Thus, methane is mostly used in steam reforming for H2 production. Different steps involved in the steam reforming process to produce H2 are as follows: (i) preferential oxidation reaction (POR); (ii) steam reforming; (iii) high-temperature water gas shift reaction (HT-WGSR); and (iv) low-temperature GSR (LT-WGSR) [6]. Another method for producing H2 is coal gasification, which involves partially oxidizing coal at high temperature and pressure (around 5 × 106 Pa). This oxidation produces a mixture of CO, CO2, H2, and other compounds. A mixture of CO and H2 is left only when the temperature is raised to 1273 K [7]. Biomass, i.e., animal and plant waste, is utilized in thermochemical and biological processes to produce H2. Examples of thermochemical H2 generation processes include gasification and pyrolysis, whereas biological H2 production processes include fermentation, biophotolysis, and the gas shift process [8].
Renewable energy sources include hydropower, geothermal energy, wind energy, and solar energy [9]. However, the development of an energy fountain is hampered by the direction of the wind, a shortage of storage capacity, and the construction of larger dams. H2 is generally considered a clean energy source, but this is not always the case. Carbon dioxide, a greenhouse gas that contributes to environmental damage, is produced during its production from coal, natural gas, or biomass. Therefore, there is a need for an alternative H2 source. Production of H2 from water splitting is a carbon-free solution, so this process is an ecofriendly way to address the current environmental issues. Water splitting can be categorized into three types: photobiological, thermochemical, and photocatalytic. Water and organic biophotolysis are two categories for the photobiological splitting of H2O, commonly known as biophotolysis [4]. H2O biophotolysis is the mechanism by which sunlight produces H2 through the water-splitting action of green algae or cyanobacteria (oxygenic bacteria). Enzymes such as hydrogenase and nitrogenase are responsible for this water splitting. Despite being a clean method of producing H2, there are a few known drawbacks, such as a low output and the enzyme’s potential for toxicity. One more reason that makes this process problematic is the requirement of a bioreactor, which is challenging to design [3]. Biophotolysis is a process by which organic waste is decomposed in the presence of aerobic bacteria to produce H2. Although this approach yields higher results, it is problematic due to CO2 emissions [10]. By using solar reactors, heat from sunlight is concentrated to achieve a temperature of 2273 K, which is then used to split H2O with the help of a catalyst (such as ZnO). Although it is a facile approach, it has several disadvantages. First of all, managing heat is really difficult. Secondly, it is very costly since large-scale solar concentrators are widely sought to achieve high temperatures [11].
By utilizing solar light, the photocatalytic splitting of H2O (Figure 3) is the most efficient method of producing H2 [12] without any greenhouse gas emission. Since solar energy, which has a power density of 400–1000 W/m2, is another renewable energy source [13], it can be converted in a variety of ways to produce heat or electricity. Moreover, the quantity of radiation that falls on earth in just thirty minutes can meet the world’s annual energy needs [14]. Solar photocatalytic water splitting has several advantages, such as the highest energy conversion efficiency, independent O2 and H2 emissions, and the least expensive and most environmentally beneficial approach. Reactors can be constructed on a small scale, increasing their market efficiency and lowering the need for solar power systems. Table 1 below provides a list of a few primary techniques used for H2 production in the present era [15].
Among several possible techniques, water electrolysis driven by solar radiation is a significant choice. It can be achieved by photoelectrochemical cells or by self-supported systems.

3. Mechanism of Photocatalytic Water Splitting

When light falls on a photocatalyst, a chemical reaction occurs that produces hydrogen from water [16]. Three steps make up the total photocatalytic water-splitting reaction. (i) The capture of energy above the bandgap of the semiconducting material leads to the formation of electron–hole pairs within the valence band (VB) and conduction band (CB), respectively. (ii) The holes (h+) oxidize the water, releasing H+ and O2. In order to create H2 by the reduction of h+ ions, photogenerated electrons (e) are then transferred towards the conduction band (CB). (iii) The reduction of H+ ions occur to form H2. Understanding heterogeneous photocatalytic water splitting requires an understanding of the sacrificial donor and the catalyst [17].
If a catalyst absorbs a photon of suitable energy which is equal to its bandgap, then it can split water but the efficiency of the process is limited because of the lower hydrogen yield which is ascribed to low energy absorption or charge recombination on the catalytic surface. To reduce the rate of charge recombination, a sacrificial donor is employed to increase the yield of H2 [18]. The segregation of charges requires the co-catalyst/buffer layers to be combined with the electrodes and the photocatalyst. Surface oxidation, exposed surface area, and ion absorption are all factors affecting the bandgap potential which is a key criterion for enhanced water splitting. Figure 2 shows the mechanism of photocatalytic water splitting.
The general water-splitting reactions are
H2O + 2h → 2H + 1/2O2, E° oxidation = −1.23 V
2H+ + 2e → H2, E° oxidation = 0.00 V
A minimal thermodynamic potential of 1.23 eV is required for photochemical water splitting [19]. The following are the major obstacles in photocatalytic total water splitting: (i) A large number of photocatalysts lack a suitable band alignment which is needed for overall water splitting. (ii) The most difficult step in the photocatalytic splitting of water is thought to be the O2 evolution half reaction, due to the four-electron transfer process coupled with the two-proton removal from H2O to form an O-O bond, which results in mismatched stoichiometric rates [20]. (iii) The generated H2 and O2 quickly react back to form water again on the catalytic surface, especially on non-noble metallic surfaces, which results in a low energy conversion efficiency. Currently, the maximum solar-to-hydrogen conversion capacity of CdS-based catalysts is 1.63%, which is far from real application requirements [21].
Figure 2. The stepwise mechanism for photocatalytic water splitting. Reprinted with permission from Ref. [22]. Copyright 2021, Elsevier.
Figure 2. The stepwise mechanism for photocatalytic water splitting. Reprinted with permission from Ref. [22]. Copyright 2021, Elsevier.
Catalysts 14 00674 g002

3.1. Types of Reactions

In photocatalytic water-splitting processes, H2 can be produced in two different ways:
  • Photochemical reactions;
  • Photoelectrochemical reactions.

3.1.1. Photochemical Reactions

The reactions which are conducted in photocatalytic suspensions in water are termed photochemical reactions, and the process is conducted by the immediate use of energy to cause a chemical change [23]. After absorbing enough energy, the splitting of water with a suitable photocatalyst results in the evolution of both H2 and O2. It becomes difficult to regulate a photocatalyst’s specific characteristics, like its capacity for light absorption when suspended, changes in the reaction’s pH, and the substrate’s concentration. The H2 gas evolved at the semiconductor–electrolyte junction serves as an interface to create the voltage required for water splitting. However, it is quite challenging to analyze the experimental results since these factors influence the reaction kinetics, making it rather arduous. We have shown that an artificial photosynthesis system based on photochemical diodes is capable of effective, unaided overall water splitting without the need for a sacrificial reagent [22]. A photochemical diode array demonstrated a solar-to-hydrogen efficiency of 3.3% in neutral conditions (pH 7.0), in which the total water-splitting process was carefully regulated by regulating charge carrier movement at the nanoscale [24].

3.1.2. Photoelectrochemical (PEC) Reactions

PEC water splitting is a potential remedy for the present world energy consumption and environmental issues that converts solar radiation to H2 fuel. The water oxidation reaction, which uses four electrons, limits the total efficiency of water splitting. The ideal photocatalyst for water splitting in PEC ought to be very selective, must have a narrow bandgap, and be highly active, durable, and cheap [25]. A conventional PEC setup involves an electrolyte, a photoanode, or a photocathode, sometimes both. Typically, a there is light response in an n-type semiconductor which serves as the foundation for the photoanode; it has the ability to produce electron–hole pairs when stimulated by light. In an external circuit, electrons produced in the photoanode flow to the photocathode, where they are accepted by hydrogen ions contained in the electrolyte solution to make H2. This reaction can be written as follows [26].
Photocathode: 2H+ + 2e → H2
Photoanode: 2H2O + 4H+ → O2 + 4H+
The overall reaction equation for splitting water is
2H2O → O2 + 2H2
PEC is regarded as a promising approach for H2 production to meet the energy requirements of the industrial circular system [27]. An electric current flowing through an external circuit drives the reaction in PEC. The photocatalyst is deposited to create a photoanode to carry out the reaction of water splitting in solution. The photon’s energy can be converted into electrochemical energy from solar radiation falling on the PEC surface, which causes H2O to split into O2 and H2. When electrons fall on the photocathode, H2 is created there. The scattered light is subsequently transformed into a storable form [28]. To increase the photo electrocatalytic activity of titania-based PEC cells, two methods have been presented. The first method, known as photoinduced reduction processes at the TiO2 surface, entails altering the pristine TiO2’s structure by chemical doping, surface/subsurface defects, changes in surface morphology, and particle size, all of which result in a lowering of the bandgap. The second method, referred to as “photo-induced oxidation processes at the TiO2 surface”, relies on using oxidizing substances like various alcohols that are used as an electron donors in PEC cells [29].

Working Principle of Photoelectrochemical (PEC) Reaction

The fundamental process in a photoelectrochemical (PEC) cell involves water splitting by applying an external bias to the photoelectrodes in order to convert sunlight into hydrogen fuel by using an electrolyte. PEC water splitting has four fundamental steps: Sunlight strikes the photoanode first, producing electron–hole pairs. Secondly, water oxidation is caused by photogenerated pores at the activated photoanode surface. Thirdly, the electrons produced by the incident photons move from the photoanode to the photocathode via an external cable, where they are reduced by the electrons to generate hydrogen gas at the photocathode surface [30]. The PEC cell’s photoanode is formed of an n-type semiconductor, while the photocathode is composed of a dark metal electrode. These two electrodes are separated by an ion-exchange membrane as shown in Figure 3. Whenever the semiconductors are exposed to sunlight, the n-type material is used in the photoanode to generate holes for the oxygen evolution process (OER) (Equation (6)), while electrons move to the photocathode for hydrogen evolution reaction (HER) (Equation (7)). To create a tandem cell, certain p-type materials have been created. The former is a four-electron reaction, whereas the latter is a two-electron reaction, making the former far more challenging. Equation (8) depicts the entire water-splitting reaction, which is a highly energetic reaction with a Gibbs-free energy of ∆G° = 237.2 kJ mol−1 per mole of produced H2 (using the Nernst equation as a reference, ∆E = 1.23 V) [31].
Reaction at anode
2H2O + 4h+ → O2 + 4H+O2/H2O = 1.23 V vs. RHE
Reaction at cathode
2H+ + 2e → H2H+/H2 = 0 V vs. RHE
Net reaction
2H2O → 2H2 + O2 ∆E = 1.23 V
However, as shown in Figure 4, the upper section of the valence band should be at a greater positive potential compared to the water oxidation potential into oxygen (1.23 V at pH = 0), in order to improve PEC’s water-splitting efficiency [30].

4. Desired Characteristic of the Material Used for PEC

Selecting a good photocathode and/or photoanode material is the most important step in PEC water splitting. Semiconductor materials with the following properties are necessary for an effective PEC water splitting which could serve as a photoanode or a photocathode.

4.1. Adequate Band Locations and Bandgap Energy

In total, 52% of the radiation of the solar spectrum lies in the infrared region (700–2500 nm), 43% is visible (400–700 nm), and 5% is UV (300–400 nm). Hence, to enhance the efficiency of the catalyst, a significant amount of visible light must be absorbed, which in turn depends on the semiconductor’s bandgap. Given that the O2/H2O potential is 1.23 V vs. NHE (pH = 0) and the proton reduction potential is 0 V vs. NHE, incident photons with a minimum energy of 1.23 eV, or roughly 1100 nm in wavelength of light, are needed to achieve the predicted minimum bandgap for water splitting. However, a minimum bandgap of ~1.8 eV is required, equivalent to light absorption at around 700 nm, when taking into account the thermodynamic energy losses (0.3–0.4 eV) occurring during charge carrier transit and the overpotential requirement for appropriate surface reaction kinetics (0.4–0.6 eV). The solar spectrum indicates that the maximum bandgap energy is 3.2 eV due to a sharp decrease in solar intensity below 390 nm.
A single semiconductor photoelectrode should have a bandgap energy of 1.9 eV to 3.2 eV, in order to obtain a significant photovoltage. A bandgap of 2.0 eV is theoretically required for optimal sunlight usage, in addition to the thermodynamic band position requirements [32]. In actuality, though, it has rarely been realized because the anticipated photovoltaic output is rarely reached. Even with the maximum current density, unbiased water splitting cannot be solved without a high photovoltage value (>1.61 V). Therefore, external bias or tandem devices are required to supply the additional voltage required to split water. The bandgaps and band locations of common n-type and p-type semiconductors used for PEC water splitting at pH = 0 are shown in Figure 5.
Since VH (as illustrated in Figure 6) relies on the pH of the solution (by −0.059 V per pH unit) with regard to the redox potential of the electrolyte, it is noteworthy that the band locations of oxide semiconductors fluctuate depending on the electrolyte’s pH. Moreover, the water reduction and oxidation potentials alter in a way that is comparable to the band positions (i.e., −0.059 V per unit pH), as per the Nernst equation. The broad consensus is that H+ and OH are potential-determining ions (PDIs) deposited on the semiconductor surface inside the Helmholtz layer, as indicated by the Nernstian dependence. Other compounds, including metal sulfides, may have varying PDIs, which makes the pH dependence more complex [33]. While the band locations are constant with regard to the water redox potentials, most semiconductors, and oxides in particular, exhibit the same pH dependency as shown by the Nernst equation. As a result, the relative difference between an oxide photocatalyst’s band locations and the water redox potential does not depend upon pH.
Figure 5. Band locations of different semiconductors in relation to the water-splitting redox potentials at pH = 0. Reprinted with permission from Ref. [34]. Copyright 2017, Royal society of chemistry.
Figure 5. Band locations of different semiconductors in relation to the water-splitting redox potentials at pH = 0. Reprinted with permission from Ref. [34]. Copyright 2017, Royal society of chemistry.
Catalysts 14 00674 g005

4.2. Effective Separation and Movement of Charge Carriers within the Semiconductor

Low STH efficiencies are largely caused by fast charge recombination, so an effective charge carrier separation and transport strategy is needed. This strategy depends on the material’s inherent characteristics, such as hole and electron mobility, as well as its extrinsic characteristics, such as crystallinity and nanostructure.

4.3. Stable and High Catalytic Activity

Electron–hole recombination would normally result from surface charge accumulation, which can be prevented by appropriately fast surface reaction rates. For many potential water-splitting semiconductors, especially metal sulfides, photocorrosion is a significant issue since it happens when the holes and electrons produced by photons break down the photocatalyst rather than oxidizing or reducing the water. The relative locations of the semiconductor band edges and their corresponding decomposition potentials determine these photocorrosion reactions. If the anodic decomposition potential (Epd) of the semiconductor is higher than its valence band potential, anodic photocorrosion may happen. On the other hand, if the cathodic decomposition potential (End) is lower than the semiconductor’s conduction band, cathodic photocorrosion may happen. The actual values of the decomposition potentials depend on the pH of the electrolyte; therefore, metal oxides like ZnO and BiVO4 or metal sulfides like CdS and MoS2 can easily undergo anodic photocorrosion. Nonetheless, because of their extremely slow decomposition reaction kinetics, typical photoanode materials like Fe2O3 and TiO2 are thermodynamically stable despite having an anodic deposition potential above the valence band potential.
For practical applications, the photoelectrode materials should also be inexpensive and made of elements that are abundant on Earth. This is essential to supporting the case for a cost-effective scaling up of solar-to-fuel devices [34].
  • Photocathodic materials
In PEC water splitting, the photocathode typically consists of a p-type semiconductor deposited on conductive substrates. Water is oxidized on the counter photoelectrode and reduced on the semiconductor surface when exposed to light, as depicted in Figure 7. In order to produce hydrogen, the photocatalysts’ conduction band edge needs to be more negative than the potential for hydrogen evolution from the perspective of electrochemical potential. There are a lower number of reports based on p-type semiconductors as photocathodes for PEC water splitting as compared to n-type semiconductor photoanodes [35]. Cu2O is a common p-type semiconductor employed as a photoelectrode for PEC water splitting, while it quickly degrades owing to self-reduction by photogenerated electrons [35]. In order to decrease the self-reduction of Cu2O and improve the PEC characteristics, WO3/Cu2O p-n junctions have been created. To enhance their PEC performance, p-n heterojunctions have also been produced. As illustrated in Figure 8, Colleen at al. [36] synthesized p-n Cu2O homojunction solar cells by electrochemically installing an n-Cu2O layer on a p-Cu2O layer. The produced p-Cu2O and n-Cu2O layers’ inherent doping levels were very low, and they made ohmic junctions with Cu metal. Compared to other p-n homojunctions, the greatest cell performance (a η of 1.06%, a VOC of 6.21 V an ISC of 4.07 mA cm−2, and a fill factor of 42%) was achieved. Paracchino et al. reported that the Cu2O photocathode, depicted in Figure 9, demonstrated photocurrents of up to −7.6 mA cm−2 at a potential of 0 V versus the reversible hydrogen electrode at moderate pH [37].
In another study, the Cu2O photocathode was activated for hydrogen evolution with electrodeposited Pt NPs and protected against photocathodic decomposition in water by nanolayers of Al-doped zinc oxide and titanium dioxide. A Pt-modified Si photovoltaic cell (Pt/SiPVC) has also been reported as an efficient photocathode for PEC water splitting that uses p-n radial junctions with a p-type Si substrate, and B-doped Si with the decoration of Pt has frequently been used as a p-type photocathode for PEC water splitting [38]. The cheap and plentiful Si supply makes Si-based photoelectrodes for PEC water splitting a promising option; however, their efficiency still needs to be increased.
Figure 7. Semiconductors installed on a substrate as a photoanode for PEC water splitting. Reprinted with permission from Ref. [39]. Copyright 2014, Royal society of chemistry.
Figure 7. Semiconductors installed on a substrate as a photoanode for PEC water splitting. Reprinted with permission from Ref. [39]. Copyright 2014, Royal society of chemistry.
Catalysts 14 00674 g007
Figure 8. SEM image of a p–n Cu2O homojunction. Reprinted with permission from Ref. [36]. Copyright 2012, Royal society of chemistry.
Figure 8. SEM image of a p–n Cu2O homojunction. Reprinted with permission from Ref. [36]. Copyright 2012, Royal society of chemistry.
Catalysts 14 00674 g008
Figure 9. Schematic illustration of the electrode structure of the surface-protected Cu2O electrode. Reprinted with permission from Ref. [37]. Copyright 2011, Nature.
Figure 9. Schematic illustration of the electrode structure of the surface-protected Cu2O electrode. Reprinted with permission from Ref. [37]. Copyright 2011, Nature.
Catalysts 14 00674 g009
  • Photoanodic materials
Typically, n-type semiconductors installed on conductive substrates make up the photoanode in the PEC water-splitting arrangement [39]. Under light illumination, as seen in Figure 9, electrons are transported to a counter electrode via an external circuit, while photoexcited holes build up on the surface of the photoanode semiconductors and are consumed in oxidation reactions. In order for the photoanode to produce oxygen, the photocatalysts’ valence band edge needs to be more positive than the potential for the evolution of oxygen from the perspective of electrochemical potential. One of the main benefits of PEC is that an external voltage bias can be used to make up for any inefficiencies and quicken the excited charge carrier separation process. However, zero bias is preferred once the PEC systems are well-aligned with a suitable semiconductor. This section begins with the conventional TiO2 photoanode and discusses the latest developments in TiO2 and TiO2-based hybrid photoanode systems (Figure 10).
Because of its lower cost and superior chemical and optical stability, titanium dioxide (TiO2) is the most frequently used n-type semiconductor for PEC water splitting [41]. Because of their inherent wide surface area and unidirectional charge flow, 1D titanium dioxide nanotube arrays (TNAs) have drawn a lot of interest among the diverse nanostructured TiO2 [42]. TNA has a significant deal of promise for large-scale applications and is easily manufactured by anodizing metal titanium foil or plate [43]. Via a mild hydrothermal reaction, 1D TiO2 nanowire arrays have also been coated on glass coated with a transparent conducting oxide (TCO) [44]. When using 1D nanowire arrays, N719 dye-sensitized solar cells can achieve photoconversion efficiencies of up to 5.02%, which is significantly greater than when using ordinary TiO2 powders. Since titanium dioxide has a broad bandgap of approximately 3.2 eV and is not sensitive to visible light, it exhibits a modest photoelectrochemical activity [45]. Researchers have devised a number of techniques to improve the absorption of visible light, such as doping TiO2 with metal or non-metal, creating heterojunctions, hydrogenation, or the formation of structural vacancies [46]. In the early 21st century, doping with metal or non-metal (C, N, S, B, etc.) was introduced as a practical way to reduce the bandgap of TiO2 and increase PEC efficiency [47]. The CB and VB of the acceptor can grow with metal and non-metal doping, respectively. This narrows the photocatalysts’ bandgap and enhances visible-light absorption. Park et al. synthesized high-aspect-ratio carbon-doped TiO2 nanotube arrays. Under white light illumination, the total photocurrent was over 20 times higher than that of a P-25 nanoparticulate film [48].
Another viable method for harvesting visible light is to modify a TiO2 nano-semiconductor by adding a second nano-semiconductor with a lower bandgap, creating a heterojunction. In addition to being a photosensitizer, the second nano-semiconductor creates an internal electric field across the interface. Recombination is decreased as a result of the internal potential bias, which significantly facilitates the separation and passage of excited electrons and holes over the dual photocatalyst interface. For example, p-type Cu2O and CdTe semiconductors have always been used to modify TiO2 nanotube arrays (TNAs) [49]. TiO2 is a p-type SC, while Cu2O and CdTe are p-type SCs. As a result, between them, p–n junctions may form, which helps to separate the excited holes and electrons.
Due to bandgap overlapping between the two distinct photocatalysts, certain heterojunctions can improve photoelectrochemical characteristics and promote charge carrier separation and transfer. The examples shown in Figure 11 and Figure 12 are typical examples. Since catalyst A (Bi2WO6) has a higher negative conduction band (CB) than catalyst B (TiO2), excited electrons from catalyst B (TiO2) can move swiftly to catalyst A (Bi2WO6). In comparison to catalyst B (TiO2), catalyst A’s (Bi2 WO6) valence band (VB) is more positive, and catalyst A’s (Bi2WO6) excited holes can move rapidly to catalyst B (TiO2). For effective water splitting, the excited holes and electrons can thus be swiftly separated and transported.

5. Electrolysis

A scalable, adaptable, and distributed method of producing hydrogen is water electrolysis. Electrolysis currently produces about 4% of the hydrogen in the world. Electrolysis-based hydrogen production can also serve as a buffer which helps to balance the supply and demand of electricity [51]. There are numerous ways to produce hydrogen. Because of its high purity, ease of use, and lack of contamination, water electrolysis is the most widely utilized method for the conversion of renewable energy into hydrogen fuel, yet it holds a minuscule presence in the market [52]. The most basic industrial method currently available for creating almost pure hydrogen is water electrolysis, and in the years to come, its significance is expected to increase. Water electrolysis is based on the movement of electrons, which is regulated by an external circuit. Solid oxide electrolyzers, alkaline membranes, and polymer electrolyzers are the three primary electrochemical technologies for producing hydrogen [53]. Electrolysis is the most effective technique for splitting water, which is a tried-and-true procedure. But since the process is endothermic, the required energy input is provided by electricity. As shown in Figure 13 below, a basic electrolysis unit (an electrolyzer) comprises a cathode as well as an anode submerged in an electrolyte. Typically, when an electric current is given, water splits and oxygen is released from the anode side while hydrogen is produced at the cathode through the following reaction [54].
The following are the overall anode and cathode reactions:
Reaction at anode
4OH → O2 + 2H2O
Reaction at cathode
2H2O + 2e → H2 + 2OH
Net reaction
H2O → H2 + 1/2O2 ∆H = −288 kJ mol−1
This method has the advantage of producing pure hydrogen, as it is a sulphur- and carbon-free technique [55]. However, compared to systems based on fossil fuels, electrolysis has several disadvantages, such as higher cost and energy requirements [37]. Nevertheless, electrolysis is still regarded as a potentially economical method of producing hydrogen locally due to its compactness and small-scale uses. Additionally, it can be used in existing gas stations [56]. It is crucial to maintain a balance between byproduct oxygen and oxygen demand. If the latter for hydrogen generation by water electrolysis does not exceed the potential supply of the former, a significant amount of the byproduct oxygen should be lost. The industrial gas oxygen itself is a crucial component of numerous industries, including glass melting, electric furnaces, and blast furnaces. Thus, by selling the byproduct oxygen to these industries, the minimal expense of creating hydrogen by protein exchange membrane (PEM) electrolysis will be required. On the other side, if a significant oxygen consumer uses PEM electrolysis to produce oxygen, it can sell the hydrogen which makes it a significant choice for water splitting [57].

5.1. Mechanism of Water Splitting in Alkaline Solutions

In alkaline solutions, HER activity is 2–3 times lower compared to acidic solutions and the process is highly dependent on the structure of the catalytic surface. Theoretical studies have shown that HER performance is related to H2 adsorption in acidic solutions and it is comprised of Volmer/Tafel or Volmer/Herovsky steps as shown below:
Volmer step: H+ + e → Had
Heyrovsky step: H+ + e + Had → H2
Tafel step: 2Had → H2
In alkaline solutions, it is mandatory to trade off hydroxyl and hydrogen adsorption (Had and OHad) along with water dissociation. The Volmer and Heyrovsky steps are demonstrated with water rather than protons acting as reactants.
Volmer step: H2O + e → OH + Had
Heyrovsky step: H2O + e → OH + H2
Thus, there is a need to break the strong bonding of H-O-H in alkaline solutions compared to weak bonding in the H3O+ ion. The rational design of catalysts with a higher ability to split water can enhance alkaline HER activity. There are four major factors which may affect the alkaline HER activity that primarily depends upon which reaction pathway (Heyrovsky or Volmer) is followed. These factors include the adsorption of water on catalytic surfaces, water-splitting magnitude, hydrogen binding energy, and the adsorption strength of OH groups. In HER, water adsorption is the first step in alkaline medium which is very low compared to H3O+ adsorption in acidic medium. Hence, developing stronger bonding between the metal and water could be a key solution to improve HER performance [58].

5.2. Ion-Exchange Membrane

To develop efficient, robust, and safe PEC water-splitting systems, lowering cell voltage and improving product separation are essential elements. A robust method that can prevent diffusive and convective crossover of O2 and H2, for the practical deployment of large-scale PEC water-splitting systems, is critically important to ensure the system stays below the flammability limit of 4% H2 in O2 mixture in a dynamic operation [59]. The low operating current densities of PEC water-splitting systems relative to water electrolysis systems have stricter requirements for product separations [60]. Ion-exchange membranes with a suitable permeability and thickness can provide product separation and minimize crossovers [61]. This separation barrier, however, must be ionically conductive, as to maintain charge balance across the cell. It is also crucial to improve the efficiency and economic feasibility of PEC systems by lowering the overall cell voltage. The use of membranes in the system for product separation adds to the overall cell voltage, therefore, increasing the conductivity of these membranes is key to their development [59]. As described further below, the use of ion-exchange membranes can also lead to improved catalytic performance, further reducing the cell voltage and creating ideal environments for Earth-abundant catalysts for both the oxygen evolution reaction (OER) and the hydrogen evolution reaction (HER) [62].
Ion-exchange membranes, also known as anion-exchange membranes (AEMs) or cation-exchange membranes (CEMs), are made of polymers with ionizable functional groups that enable the selective transport of anions and cations [63]. In PEC systems, monopolar membranes (CEMs or AEMs) are employed to transfer ionic currents across electrochemical cell chambers and block H2 and O2 crossing [59]. A water dissociation (WD) catalyst is often positioned at the intersection of the cation-exchange layer (CEL) and anion-exchange layer (AEL) in bipolar membranes (BPMs). At the intersection of the CEL and AEL, there is an increased WD under reverse bias as a result of a catalyst and a strong electric field. Then, distinct acidic and basic streams are created on either side of the BPM by H+ ions moving preferentially through the CEL and OH ions traveling selectively through the AEL [64]. BPMs enable the maintenance of significant pH gradients (0–14) between the cathode and anode chambers in PEC systems. Therefore, for Earth-abundant catalysts for the HER, which is kinetically more possible under acidic conditions, and the OER, which is kinetically more feasible under alkaline conditions, the pH of the catholyte and anolyte can be independently tuned [64,65]. To quantify the ion-transport parameters in these CEMs, AEMs, and BPMs, standard benchmarking techniques must be created in order to enhance ion-exchange membrane technology.

6. Hydrogen Generation Electrolyzers

Electrolysis is a chemical reaction in which a substance is broken down by an electric current flowing through it. Electrolyzers are tools that are essential to this process. Depending on the type of electrolyzer and compounds being processed, the primary purpose of an electrolyzer is to aid in the electrolysis process and generate a variety of chemicals.
A general schematic for producing hydrogen energy using electrolyzers is shown in Figure 14. In short, hydrogen-generating electrolyzers are machines in which the electrolysis process is used to split H2O into H2 and O2 with the help of electricity. Water electrolysis is the name given to this process, which uses two electrodes (an anode and a cathode) and an electrolyte (typically an alkaline solution or a polymer electrolyte membrane). Due to the external electrical potential applied to the electrodes that are dipped in water, hydrogen is generated at the cathode and oxygen is generated at the anodes. After being separated, the hydrogen gas can be gathered, kept, and used as a clean energy source [66]. The establishment of electrolyzers for hydrogen production is important for the storage of renewable energy in hydrogen fuel. Several programs such as power-to-gas and power-to-X are dedicated to store and use hydrogen as a feasible and environmentally responsible energy source [67]. Water electrolysis can also be made more affordable by utilizing renewable electricity from sources like solar cells [68].

Types of Electrolyzers

Although the basic idea behind water electrolysis is simple, it allows for the freedom to create several iterations by utilizing a variety of physiochemical and electrochemical parameters. There are various kinds of electrolyzers, and each has unique benefits and uses [69]. Four main technologies are widely used to classify electrolyzers, and these can be distinguished by the particular electrolyte and operating temperature of each technology. The selection of components and materials is heavily influenced by these distinctions. Alkaline, solid oxide electrolyzer cells (SOECs), proton-exchange membrane (PEMs), and anion-exchange membrane (AEMs) are the four main types of electrolyzers [70]. Table 2 presents an overall comparison of all electrolytic cell types that are commercially available.
The most common type of electrolysis is called alkaline water electrolysis (AWE), which uses high concentrations of KOH as the electrolyte and operates at low temperatures (60–80 °C) [72]. Furthermore, these technologies allow the use of inexpensive and non-noble materials [73]. Newer technologies such as PEM and SOECs can manufacture green hydrogen from renewable energy sources with zero carbon emission and great efficiency [70,74].
One promising path towards large-scale, renewable energy-based hydrogen production is by AEM electrolysis. However, difficulties with membrane electrode installation and reaction kinetics are primarily responsible for the efficiency of AEM electrolysis. Through focused efforts, researchers are actively working to improve AEM electrolysis performance in order to address these constraints [75].
Water electrolysis which includes the AWE, AEM, PEM, and SOEC methods offers a viable pathway for producing hydrogen with a variety of uses in the fields of catalysis and renewable energy [70,75]. However, “gray hydrogen” sourced from fossil fuels dominates the current global hydrogen energy environment. This path diverges from the initial objective of attaining net-zero emissions [76]. Because of this, current research and development initiatives are focused on producing green hydrogen in an eco-friendly manner while also trying to improve the effectiveness and performance of these electrolysis techniques for hydrogen generation at a large scale.
A water electrolysis cell’s working concept is centered on two electrodes separated by an electrolyte. The migration of generated chemical charges, such as cations (+) and anions (−), between the electrodes is facilitated by this medium. Potassium hydroxide (KOH) in a highly concentrated solution is frequently used as the electrolyte in alkaline electrolyzers to transfer OH anions. The electrodes and the generated gases are physically separated by a porous inorganic diaphragm or separator that is permeable to the KOH solution [77]. Conversely, PEM, AEM, and SOEC separate the electrodes using an electron-insulating solid electrolyte. The ion transport occurs within the PEM, AEM, or solid oxide module. Thus, these specific electrolyzers eliminate the requirement for a liquid electrolyte solution [78]. A summary of the essential elements and operating parameters for each of the four types of electrolyzers is provided in Table 3. Conditions and components that show a great deal of variation amongst different producers or research and development organizations are indicated by highlighted cells. Additionally, this presentation sheds light on less developed technologies, as seen in the solid oxide and AEM categories in particular [79].

7. Requirements for Efficient Working

The substance which is employed as a photocatalyst must meet a number of functional requirements in terms of electrochemical energy and bandgap properties. These include a higher charge segregation rate, lower charge recombination, and higher charge transference and utilization. Moreover, higher chemical stability against photocorrosion and corrosion in aqueous environments are the prerequisites for economical water splitting. Therefore, to create materials having activities in water splitting during irradiation with visible light, particular photocatalytic surface properties and bulk properties along with energy structures are required [4].
The bottom of the photocatalyst’s conduction band (CB) should be at a more negative potential than the proton reduction potential (H+/H2 = 0 V vs. the normal hydrogen electrode (NHE) at pH 0) in order to attain total water splitting. Additionally, in order to promote water oxidation, the photocatalyst’s valence band edge needs to be higher than that of the oxidation potential of water (+1.23 V vs. NHE at pH 0) [81]. Thus, based on these figures, the total energy needed to carry out the water-splitting reaction is 1.23 eV for a theoretical semiconductor’s bandgap. Theoretically, 70% of all solar photons are accessible for water splitting (Table 4) because that energy is equal to the energy of a photon with a wavelength of about 1010 nm.
There are some requirements for photocatalyst efficiency: First, it must be a semiconductor, meaning that it must have a bandgap. Secondly, the bandgap value should be around 2.0 eV to enable the catalyst to utilize a significant proportion of visible as well as UV light. Thirdly, band edge positions should be optimized to produce holes and excited electrons at the proper energy levels, where hydroxyl group anions and hydrogen cations are reducing and oxidizing, respectively. Fourth, there must be no imperfections in the structure of the photocatalyst in order to minimize the electron–holes recombination [83].
The band positions exactly meet the thermodynamic conditions needed to split water [84]. The next three factors appear to be crucial: (i) a molar ratio of 2, (ii) a turnover number (TON) that is catalytically acceptable, and (iii) a direct measurement of molecular oxygen and hydrogen that does not rely on current flow. These would represent the bare minimum needed for investigations on water splitting [85]. The production and separation of efficient light-induced carriers, as well as the adsorption of molecules of water at the catalytic center before water-splitting redox reactions, are the two main steps in an efficient photocatalytic water-splitting reaction. A directly Z-scheme heterostructure may be capable of achieving the exacting standards needed for photocatalytic water splitting [86].

8. Various Catalysts for Water Splitting

Designing a catalyst which could grab maximum solar light is the key task for enhanced photocatalytic water splitting. Excitation, bulk diffusion, and surface transmission are the three main mechanisms involved in photocatalysis. As a result, in order to be effective, a photocatalyst must fulfill several primary requirements regarding its chemical, crystallographic, and surface qualities. However, there are a few limitations in the semiconductors which must be addressed, and it is exceedingly challenging to locate a single element that can meet all of these needs. Thus, even though numerous semiconductors for water splitting have been created over the past few decades, most of them need sacrificial agents to yield hydrogen from water and work only in the UV region of the solar spectrum [87]. Numerous photocatalysts have already been designed and evaluated in recent years for both thermodynamically uphill reactions and downhill reactions, but TiO2 continues to be the most extensively studied one due to its favorable physical and chemical properties. Additionally, conducting the process in the presence of organic compounds that act as a hole scavenger stimulates an irreversible reaction. This process enhances the kinetics of the water splitting to yield a higher concentration of hydrogen [88].
For photocatalytic water splitting, many catalysts are used, each with benefits and drawbacks. Some of them perform better than others in terms of efficiency. Co-catalysts are used in photocatalytic water splitting. Catalysts are elements or compounds that speed up or improve chemical reactions. The following Table 5 shows some co-catalysts used in photocatalytic water splitting.

9. Titania for Water Splitting

In the 1970s, it was found that TiO2 reveals a lot of ideal assets which include its chemical stability, nontoxicity, and cost-effectiveness, as well as its Earth abundance. All these traits make it a suitable and ideal candidate for water splitting to generate hydrogen [90]. However, TiO2 also has a broad bandgap (3.0–3.2 eV), which lowers the possibility of visible-light absorption [91]. By increasing the number of active sites and enhancing the electrical conductivity, the structural and chemical properties of TiO2 allow for the engineering of various parameters such as the bandgap, light-absorption properties, and recombination time [92]. There are multiple polymorphs of TiO2 with varying behaviors. As seen in Figure 15, the most prevalent ones are rutile, brookite, and anatase. The most common polymorphs for photocatalytic water splitting are rutile and anatase TiO2; however, some attempts have also been made using amorphous TiO2 (aTiO2) as shown in Figure 16.
Numerous endeavors have been undertaken to incorporate dopants in order to enhance the optical absorption of TiO2. For instance, compared to pure anatase TiO2, Zhang et al. discovered that 12.5% copper (Cu)-doped anatase TiO2 displayed a wider absorption [95]. According to a theoretical analysis, the enhanced optical absorption is explained by the introduction of an unoccupied impurity continuum band by Cu at the top of the valence band. According to another theoretical analysis performed by Morgade and Cabeza in 2017, it was discovered that co-doping TiO2 with C, N and Pt, V narrows the bandgap and beneficially impacts the position of the conduction and valence band edges [96].
Our review provides a great insight about the current technologies used for water splitting in the form of electrolyzers and gives an overview of the TiO2-based catalysts used for water splitting. Additionally, it demonstrates various characterization techniques used for analyzing the charge dynamics of the photocatalyst which can be helpful in designing a suitable catalyst for water splitting. Moreover, it highlights various modification techniques used to improve the catalytic performance of TiO2 for enhancing its water-splitting aptitude to generate a higher hydrogen yield.

9.1. Photocatalytic Water-Splitting Mechanisms for the Production of Hydrogen Using TiO2

The configuration of the electrical semiconductor plays an important role in the photocatalysis process. The two bands that separate semiconductors from conductors are the conduction band (CB) and the valence band (VB). The energy difference between these two bands is known as the bandgap (Eg). The valence band would include holes and electrons if not for the activation process. The energy level of incident light must match or exceed the energy level of the bandgap of the semiconductor in order to excite the semiconductor. When photons excite electrons in the valence band of a photocatalyst to the conduction band, they leave the holes behind and form pairs of electrons and holes [97]. The water-splitting photocatalytic process of TiO2 may involve the following phases. First, as Equation (17) illustrates, when the surface is exposed to visible light, electron (e) and hole (h+) pairs are formed. When the photogenerated electrons and holes quickly join in bulk or over the semiconductor surface, the energy is released as heat or photons. Electrons and holes can mix in the way shown by Equation (18), in addition to participating in chemical reactions. Moreover, as shown in Equation (19), the photoexcited carriers are capable of surface chemical reactions, in which the holes interact with H2O molecules to produce H+ and •OH without recombining. Equations (20)–(22) ultimately demonstrate how photogenerated electrons and holes oxidize •OH and reduce H+ to generate O2 and H2 [97,98]. A simplified representation of the photocatalytic water-splitting mechanism is shown in Figure 17.
TiO2 → TiO2 (e + h+)
TiO2 (e + h+) → TiO2
h+ + H2O → H+ + •OH
2e + 2H+ → H2
OH + •OH → H2O + 1/2O2
H2O(l) → H2(g) + 1/2O2(g), (Overall reaction)
A significant positive change in the Gibbs free energy (Go = +238 KJ/mol) indicates the thermodynamic uphill reaction of water splitting into H2 and O2. It has been determined that the band structure is a thermodynamic need for water splitting. Thus, at λ > 400 nm, the photocatalyst’s bandgap energy (Eg) needs to be higher than 1.23 eV. The photocatalyst bandgap should be greater than the reduction and oxidation potential of water. The bottom of the conduction band must be more negative than the H+/H2 (0 V against the NHE), while the top of the valence band must be more positive than the O2/H2O oxidation and the bandgap of the conduction band must be compatible; thus, photoexcited holes and electrons can effectively oxidize potential (1.23 V). Therefore, 1.23–3 eV is the required bandgap range for an active semiconductor that produces hydrogen efficiently when irradiated with the visible light [100]. The development of a highly efficient semiconductor for this purpose is significantly influenced by two features of the processes and mechanism of photocatalytic water splitting using visible light: (i) The photocatalyst’s band structure need to be tuned for visible-light harvesting. (ii) The bulk or surface electron–hole recombination will not occur if the photo-induced charges in photocatalyst are successfully separated.
Numerous studies explored the idea of producing H2 from water splitting using TiO2-based photocatalysts. For example, a Pt/HS-TiO2 photocatalyst, introduced by Zhu et al. [98], follows a similar mechanism of H2 production. It has been discovered that H2 evolution is significantly enhanced by the addition of a small amount of methanol as a sacrificial reagent. Acetic acid is formed when methanol is subjected to hydroxyl radicals. Water is converted into H+ and •OH more quickly when hydroxyl radicals are consumed, and electron–hole recombination is prevented on the surface of Pt/HS-TiO2 by trapping electrons [97]. Moreover, the rate of reaction is increased when every available hole is completely utilized. In addition, water plays a crucial role in the electron–hole separation process by acting as a “sacrificial hole scavenger”. Hydrogen is produced via the reduction of H+, which is catalyzed by Pt nanoparticles on the photocatalyst surface [101].

9.2. Limitations of Solar Hydrogen Generation with TiO2

It is quite challenging to produce an effective visible-light-active photocatalyst for commercial solar H2 production to meet the energy requirements of industrial circular systems. The efficiency of TiO2-based photocatalytic water-splitting systems, which produce hydrogen from solar energy, is currently relatively low. One of the most important reasons that TiO2 nanoparticles have proven so successful in photocatalytic H2 production is their role in charge separation. It has been noted that the excited electrons and hole pairs need to be moved and separated quickly in order to start the redox reaction over the photocatalyst’s surface. Charge carriers will reassemble on the surface of the TiO2 photocatalyst or in bulk if the process proceeds slowly. On the TiO2 surface, these sluggish processes lead to charge carrier recombination, which releases energy as heat or light. Moreover, as water breaks into H2 and O2 and releases energy, the reverse reaction recombining the elements back into water may occur with little difficulty. To enhance H2 and O2 separation and manage the reverse reaction, a variety of techniques can be used. Since the bandgap of different types of TiO2 crystals ranges from 2.9 to 3.2 eV, UV light is the only light that can be utilized to generate H2 in TiO2. It is important to note that other types of materials having different bandgaps might also be appropriate for the production of H2. The efficiency of the sun’s photocatalytic hydrogen generation has the potential to be boosted because visible light provides around 50% of all energy originating from the sun, while ultraviolet (UV) light only contributes 4% [102]. Researchers are working hard to enhance photocatalytic performance and increase visible-light sensitivity in order to make solar photocatalytic H2 production a reality [103].

10. Strategies to Improve the Efficiency of the TiO2 Photocatalyst

10.1. Doping

The contribution of the catalyst to improve water splitting can be seen by comparing the efficiency of hydrogen generation with and without the catalyst. In photocatalytic processes, carrier transferring and carrier trapping are both important. Therefore, doping TiO2’s surface with metals or non-metals is a great way to boost the material’s electron–hole transmission [104]. Significant doping makes it more difficult to transport electrons and holes to the interface; as a result, metal ions are more likely to serve as recombination centers. Doping can be divided primarily into two types: (i) cationic doping and (ii) anionic doping. Both categories of doping have greatly altered TiO2. New energy levels are created in TiO2 above and below the valence band as a result of the anionic and cationic doping [105], while the latter frequently aids in altering the VB positions, reduces defects, and improves the chemical strength of TiO2. The anionic doping is ascertained to minimize the bandgap energy and charge separation and promote visible-light absorption in TiO2. The majority of the anionic doping in TiO2 has been performed with N, C, S, and P. Due to the system’s higher stability, the N doping among them demonstrated considerably increased photocatalytic activity [106]. Similar to this, numerous elements were studied for their photocatalytic properties under UV–visible light and doped at the cationic site of TiO2.

10.1.1. Anionic Doping in TiO2

The characteristics of C, N, and S-doped TiO2 nanomaterials were evaluated for visible-light absorption. It was found that TiO2-P25 exhibits the conventional band-edge absorption at about 390 nm with the 3.2 eV bandgap, as well as the same values for C and S doping being observed. However, N-doping exhibited an absorption at around 415 nm with a bandgap energy of 3.0 eV. Additionally, an intriguing aspect of their valence band X-ray photoelectrons’ spectra showed that the addition of C, S, and N to TiO2 resulted in the creation of new states (Figure 18). It was discovered that these extra states added deeper stages into the bandgap of TiO2 in the sequence of C > N > S, and they were related to the C 2p, S 3p, and N 2p orbitals [107].

N-Doped TiO2

Recent studies have found that oxygen vacancies lead directly to donor states underneath the conduction bands which could enhance the visible-light photoactivity, with substitutional nitrogen which inhibits the electron–hole pair recombination. It was hypothesized recently that the bandgap of nitrogen-doped TiO2 was reduced below the inherent bandgap edges and that the photoactivity beneath visible light was caused by substitutional nitrogen atoms. It was determined that the nitrogen-induced effect that permitted sub-bandgap excitation was caused by a single N (2p) state rather than compressing the bandgap. Optical absorption revealed the localized N states in the TiO2 lattice and substitutional doping, as well as the status of N as an anion-like (N-) ion with N-Ti-O as its chemical environment.
While oxygen vacancies were mainly responsible for absorption above 500 nm, and for absorption below 500 nm, the nitrogen states above the valence bands were primarily responsible [108].

F-Doped TiO2

It was discovered that F-doped TiO2 has a great photocatalytic efficiency for the breakdown of acetaldehyde gas when exposed to visible light and UV. The researchers came to the conclusion that F-positive doping’s effects, including the generation of oxygen vacancies, increased surface acidity, and also an increase in active sites, were responsible for the high photocatalytic activity. TiO2 thin films with carbon doping exhibited hydrophilicity when exposed to visible light. The localized C (2p) generated above the VB is thought to be the cause of the material’s reduced sensitivity to visible light, which has a lower hydrophilic coefficient than UV light.
In order to maximize the photo-cleavage of water by visible-light irradiation, it was reported that the production of a vertical grown carbon-doped TiO2 nanotube array offers high aspect ratios. Under visible-light irradiation (>420 nm), the synthesized TiO2-Cx nanotube arrays demonstrated significantly greater densities of current and more effective water splitting than pure TiO2 nanotubes. Under white-light illumination, the overall photo current was 20 times more as compared to a P-25 nanoparticle film [108].

S-Doped TiO2

S-doped TiO2 photocatalysts were made using a template-free oxidant peroxide approach and then crystallized using a hydrothermal method. It was determined that adding S to TiO2 will replace oxygen atoms with sulphur anions, causing the bandgap to narrow as shown in Figure 18. The visual absorption of TiO2 from UV to the visible area was significantly increased as a result. It was discovered that adding electron-deficient S atoms to the TiO2 crystal lattice successfully trapped photo-induced electrons and decreased charge carrier recombination. Additionally, surface oxygen vacancies created by S-doping served as a location for the photogenerated electrons to be trapped, decreasing the recombination of electron–hole pairs.
The increased visible-light photocatalytic activity of S-doped TiO2 for the degradation of methyl orange (MO) was the consequence of the harmonious effect of the effectively reduced charge carrier recombination and prolonged visible-light absorption [109].

10.1.2. Cationic Doping

In order to increase TiO2’s photocatalytic activity, a great deal of research has been performed recently on both doping the material with transitional metal ions and doping it with rare-earth metal ions. It contributes to the creation of acceptor energy levels underneath the conducting bands of the material or donor energy levels above the valence band of TiO2. Choi et al. [110] carried out a detailed study of the photoreactivity of metal ions doped into TiO2 and their results were reported. It was found that doping the material with metal ions could shift the photoresponse of TiO2 further into the visible spectrum. In a similar manner, Zhao et al. [111] synthesized bismuth (Bi)-doped TiO2 nanotubes that were created by Zhao et al. [65] and hydrothermally treated in a 10-mol·L−1 NaOH solution. They noticed that altering the Bi/Ti ratios would produce comparable outcomes. Because of their enhanced photocatalytic efficiency and increased ability to absorb light into the visible spectrum, the produced catalysts were able to produce hydrogen from a mixed glycerol/water solution more effectively than pure TiO2 nanotubes. The Bi-doped TiO2 catalyst would change the energy level of Bi ions and produce new Bi (3+x+) species. Resultantly, this leads to a decrease in the rate of electron–hole pair recombination. Noble metals, such as gold, silver, ruthenium, and platinum, have been found to be superior to other metals in the creation of H2 because of their exceptional ability to increase photocatalytic performance into the visible spectrum by using strong surface plasmon bands [83,112]. For instance, Ortiz et al. [67] created TiO2–F, an undoped TiO2, and TiO2Ag–F, a doped TiO2, using the sol-gel/solvothermal method (SGH). TiO2Ag–F was shown to exhibit an outstanding catalytic activity, producing 180 μmol of H2 per gram of catalyst during a 4 h irradiation period. One important element influencing how quickly these materials exhibit photocatalytic activity when electrons are brought to them is the impossibility of electron–hole pair recombination, which is required to create metallic Ag clusters in TiO2.
Furthermore, Ni et al. illustrated how different transition metal ions alter photocatalyst activity via diverse electron–hole migration pathways in order to discuss the doping of TiO2 with transition metals, comparable to that of noble metals, as shown in Figure 19 [97]. Because transition metal nitride materials have virtually comparable optical properties to noble metal nitride materials, they have lately found use as plasmonic materials for photocatalytic processes [113]. Transition metals transport more electrons into the TiO2 conduction band because they are more chemically stable and have a higher plasmon resonance than noble metals [114]. Consequently, the photoelectron transfer from the transition metal to the CB of TiO2 results in an efficient reduction process at the semiconductor’s surface. For example, to examine the photo-thermochemical cycle (PTC) mechanism, Xu et al. [115] produced films of TiO2 doped with iron (Fe) using a sol-gel process. Fe-doped TiO2 outperformed untreated TiO2 in the PTC mechanism’s water-splitting method for creating H2.
As previously mentioned, cationic doping effectively establishes intra-band energy levels near to the conduction band of TiO2, which causes a red shift in the system’s optical properties [116]. This effect has also been reported in cations such as transition metals, rare-earth metals, or other metals doped with TiO2. The main disadvantage of cation doping seems to be the increased number of charge carrier trapping sites (both for electrons and holes), which significantly lowers the photocatalytic efficiency.
This is due to the fact that the trapped carriers often merge back with their corresponding mobile carriers. As shown in Figure 20, the process of cationic doping fundamentally involves tuning both of the electronic structure and the Fermi level of the d-electron configurations in TiO2 in order to tailor the levels of energy for the absorption of light in the visible region.
Under visible-light irradiation, iron was utilized to dope TiO2, and its photocatalytic activity outperformed that of the Degussa P-25 commercial catalyst. Fe-doped TiO2 was discovered to have a far higher photocatalytic efficiency than Degussa P-25 when exposed to visible light. In the TiO2 lattice, Fe3+ cations served as shallow traps. After doping at a comparatively low level, the best photocatalytic characteristics were attained. The kinetics of the recombination process, which was correlated with spacing among dopant cations within the TiO2 lattice, was closely tied to this. In addition to trapping electrons and holes, Fe ions significantly increase photoactivity. Due to a reduction in the surface-active centers’ density, the highest photoactivity was seen at 0.5 weight percent of the Fe level. It was discovered that the additional Fe atoms soluble in the TiO2 phase had a rutile micro-structure with an average grain size of less than 10 nm. Compared to conventional P-25 powder, Fe-doped powder has a greater absorption threshold in the 427–496 nm range (406 nm). Fe-doped powder’s color turns from white to brilliant yellow, and the bright yellow becomes dark yellow when the Fe level is over 4.57 weight percent [106].
It is important to remember that optimal performance requires the precise management of doping levels or concentrations. If the dopant concentration is extremely high and functions as a recombination center, the photocatalytic performance can be compromised. Similar to the study of Zhao et al. [110], the transformation of anatase-phase nanotubes into titanate-phase nanobelts results from an excess of Bi ion doping in the TiO2 crystal structure. It is anticipated that this alteration will lower the catalytic efficiency in terms of hydrogen yield. Therefore, in order to optimize and improve the efficiency of charge trapping and separation, it becomes crucial to accurately control the dopant quantity [91].

10.2. Sensitization of TiO2

The excitation of a chromophore, like an organic dye, is a useful technique for filling the conduction band of a large bandgap semiconductor with electrons under visible light [117].
The two types of sensitization are composite semiconductors and dye sensitization. To use visible light for textile waste water energy conversion, dye sensitization is frequently used. The table below lists common dyes. The dyes’ wavelengths fall between 442 and 665 nm. For solar cells and photovoltaics, some dyes with redox properties and visible sensitivity to light are used. When compared to bulk TiO2 photocatalysts, highly distributed TiO2 species produced in zeolite frameworks, SiO2, B2O3, and Al2O3 matrices demonstrated a superior and distinctive photocatalytic performance [107]. Table 6 shows some commonly used dyes and their classifications.

10.2.1. Dye Sensitization

In this method, a narrow bandgap semiconductor is coupled with a higher negative conduction-band level with a wide-band semiconductor having a larger bandgap. This increases the mixed photocatalyst’s ability to absorb light by injecting conduction-band electrons into the larger bandgap from the small bandgap [72]. In this method, solar photons are used to excite the dye molecules attached to TiO2 in order to produce highly excited electrons, which are subsequently transferred towards the CB of the TiO2. The TiO2 conduction band’s electrons are then transferred to interfacial electron acceptors (such as H2O, I3−, H+, and O2) and cause a number of redox processes, as shown in Figure 21. In order to reach a high degree of photoconversion efficiency, interfacial energy transfer from the excited dye to the TiO2 conduction band and effective solar energy usage should be the critical characteristics.
The production of organic dye-sensitive TiO2 particles for the oxidation of 4, chlorophenol and Cr(VI)) pollutants under visible light (>420 nm) was described, and the impact of different anchoring groups was examined. To achieve this, they created Ru-free organic dyes in a donor–acceptor configuration with various numbers (n = 1, 2, and 3), known as D1, D2, and D3, respectively [108].
The existence of electron donors such tri-ethanol-amine (TEOA) as well as ethylene-diamine-tetra-acetic acid (EDTA) enabled researchers to discover that all three types of dyes are effective for producing hydrogen gas, with the ordering being D3-D2 > D1 (TEOA) and D3 > D2-D1 (EDTA). According to the results of the FTIR research, D1 and D3 are most likely anchored to the surface of TiO2 using bidentate modes with single and double carboxylates, accordingly. They also looked into the several ways that D2 is anchored, based upon competing electron donors. Because of the various photochemical circumstances and reaction routes, they showed that the carboxylate number becomes less significant in the sensitized pollutant oxidation. Figure 22 shows the tri-branched organic dye molecules (sensitizers D1, D2, and D3, respectively) having mono-, di-, and tri-carboxylate attaching groups.

10.2.2. Sensitization with Noble Metal Particles

To sensitize TiO2 with metallic nanoparticles is necessary by using the localized surface plasmon resonance (LSPR) phenomenon in order to increase the visible-light absorption because TiO2 can only be active in the UV region. Au, Ag, Pt, and Pd, as well as their alloys, are the main noble metal nanoparticles due to their absorption in the visible light region [118]. The LSPR effect, which serves as an electron sink, can “excite” the electrons in the CB under visible-light irradiation to produce extremely energizing “hot electrons” [119].
The “hot electrons” then travel directly to the CB of TiO2, creating a metal–semiconductor Schottky junction. The LSPR effect, meanwhile, may create a powerful local electronic field that increases its conduction electrons’ energy and makes it easier for them to transfer to and interact with electron acceptors. Resultantly, the electrons and holes produced by photons are effectively segregated [120]. It is possible to decorate TiO2 restructurings with noble metal nanoparticles using hydrothermal, electrospinning plasma sputtering, UV reduction, electrode location, and other techniques. Hollow Pt/TiO2 spheres have been created using a straightforward sol-gel process. Due to the increased donor density and decreased recombination of photogenerated electron–hole pairs, Pt/TiO2 hollow spheres exhibit a greater photocatalytic water-splitting activity than pure hollow TiO2 spheres under visible light [121]. Photocatalytic splitting by Ag/TiO2 based on the LSPR effect is shown in Figure 23.

10.3. Heterojunction Construction

One of the most successful methods to address the aforementioned problems is the creation of a heterojunction based on TiO2 by adding an additional low-bandgap conductor [122]. In hybrid systems, there are two distinct techniques used: hetero- and homojunctions. The use of a heterojunction in design is a typical strategy because it has the ability to significantly reduce serious charge recombination and maximize the use of visible light. On the contrary, the production of a homojunction (phase-junction) between two separate phases might modify the transit of the charge carriers along the inter-phase band alignment, making it a different configuration access to charge separation [123]. The variation in the photo-induced electron–hole transfer mechanism, which may be split into two types as follows, determines the heterostructure of the photocatalyst. The mechanisms are (i) the Z-schemes charge transfer mechanism, as well as (ii) the Type II heterojunctions transfer method [124]. The photocatalytic performance of a carefully designed heterojunction can remarkably enhance the efficiency of the hybrid photocatalyst as compared to the unassisted photocatalyst, according to the literature. More significantly, it has been discovered that a hybrid photocatalyst performs better for a composite heterojunction photocatalyst when two photocatalysts are mixed in a ratio of 1:1 [125].
There are three known forms of band alignment in heterojunctions as depicted in Figure 24 [87]. The valence band of SC-1 is more positive than that of SC-2 while the conduction band of SC-1 is more negative than that of SC-2 in Type I. In Type II, the valence band of SC-1 is less positive than that of SC-2, while its conduction band is more negative than that of SC-2. In Type III, both the valence band and conduction band of SC-2 are lower than those of SC-1.
In a semiconductor–semiconductor heterojunction, holes and electrons are thought to move to the less positive valence band and the less negative conduction band, respectively. In this case, SC-2 during Type I alignment, holes and electrons end up accumulating in the semiconductor with the smaller bandgap. Therefore, the chances of their recombination increase, preventing any enhancement in photocatalytic activity. Improved charge carrier separation will be attained in Type II alignment, because the electrons and holes in the two semiconductors will be separated. On the other hand, in Type III alignment, neither electrons and nor holes will be able to transport to one another. They are therefore considered to be two separate semiconductors and do not form a heterojunction. A semiconductor’s Fermi level is the total chemical potential of its electrons, and it is typical for different semiconductors to have unequal Fermi levels, so a semiconductor–semiconductor heterojunction is expected to form an electric field similar to that of a Schottky junction, where the flow of charge carriers is facilitated in one direction but hindered in another direction.
For the Type I alignment, the charge separation will not be encouraged as explained above. It is possible that just a very small internal electric field can arise because both semiconductors are intrinsic. Better charge separation is envisaged for a heterojunction with a non-negligible internal electric field. In the case where SC-1 has a higher Fermi level than SC-2, free electrons from SC-1 will flow to SC-2 during the internal electric field formation until the Fermi levels are aligned (Figure 24, Type I), resulting in the formation of a space charge region close to the interface. The internal electric field flows from SC-1 to SC-2 because of the electrostatic interaction that causes the surface of SC-2 to be negatively charged and the surface of SC-1 to be positively charged [127]. Thus, after photoexcitation, the internal electric field promotes the transfer of holes from SC-1 to SC-2 while blocking the movement of electrons in the same direction. Instead, because of the more negative conduction band of SC-1, the electrons from SC-2 will move towards SC-1 but gather at the junction [128]. In a heterojunction between an n-type SC-2 and a p-type SC-1, where the conditions are reversed (Type I-2 in Figure 25), electrons are more likely to move from SC-1 to SC-2, whereas holes from SC-2 gather at the interface. In both scenarios, it is theoretically possible to attain an improved charge separation.
For the Type II band alignment, when the Fermi level of SC-1 is greater than that of SC-2, an internal electric field is created that points from SC-1 to SC-2 (Type II-1 in Figure 25). In the space charge region for SC-1, photogenerated electrons will flow towards interior, while holes will move in the direction of interface; in the same way, the photogenerated charge carriers for SC-2 will flow in the opposite direction as for SC-1. Thus, the interface intercepts both the electrons from SC-2 and the holes from SC-1, which leads to their annihilation via recombination [129]. For catalytic reactions, however, the electrons in SC-1 and holes in SC-2 are retained without losing their reduction and oxidation ability, respectively. This kind of heterojunction is also known as the Z-scheme system because the photoexcitation and charge recombination processes show a Z-shape route in the corresponding energy diagram [130]. The internal electric field flows from SC-2 to SC-1 (Type II-2 in Figure 25), when the Fermi level of SC-2 is greater than that of SC-1 in the Type II band alignment, such as in the scenario where SC-1 is a p-type semiconductor and SC-2 is an n-type. The electrons of SC-1 will move in the direction of the interface. The electrons of SC-1 can cross the interface and move to the conduction band of SC-2 because the conduction band of SC-2 is less negative than the conduction band of SC-1. In a similar manner, the holes of SC-2 will move to the valence band of SC-1.
For the Type III-1 heterojunction, no electrons and holes can cross the junction and enter the corresponding bands of another semiconductor (Figure 25). The energy level barriers prevent the transport at the interface even though both the electrons and holes of SC-1 and SC-2 will move towards the junction. A twisted Z-scheme system can be created by installing an appropriate charge mediator to act as a bridge connecting these two semiconductors [87]. The leftover electrons and holes in a twisted Z-scheme system will ultimately wind up in the more negative conduction band and more positive valence band of the two semiconductors, respectively, similar to what a conventional Z-scheme system can achieve.
TiO2-based heterojunction semiconductors have been the subject of several studies. One such example is the solvothermal fabrication of an anatase TiO2/g-C3N4 heterojunction doped with boron. The heterojunction’s enhanced light-absorption efficiency and improved charge separation resulted in an amazing visible-light photocatalytic performance (47.3 mol h−1) [131]. Similarly, under solar light irradiation. Liu et al. investigated the Ag3PO4-TiO2 heterojunction’s band structures and its effectiveness in photocatalytic water splitting for hydrogen evolution. The determination of the Fermi levels of TiO2 and Ag3PO4, i.e., −5.09 eV and −5.95 eV, respectively, led to the evaluation of their energy band diagrams. It was revealed that the heterojunction formed with 12 wt% Ag3PO4 had the fastest rate of hydrogen evolution (44.5 μmol g−1h−1), which was 5.7 times faster than TiO2. When the heterojunction was deposited with Au nanoparticles, hydrogen generation enhanced 10.2 times (453.0 mol g−1h−1). It was discovered that the hydrogen evolution performance was consistent with the PL analysis results.
He et al. [132] successfully manipulated the nucleation and growth processes to form an edge-on heterojunction of MoS2/TiO2. This heterostructure may enhance interfacial electro-conductivity and alter the electron transfer pathway along basal planes, thereby improving the separation of photogenerated charges and, ultimately, enhancing the catalytic abilities of TiO2. This study mainly focused on the simple insertion of MoS2 into photocatalysts to generate heterojunctions, with minimal focus on the systematic design and management of nanoscale architectures to maximize light absorption, charge separation, and surface catalytic reactions. To achieve a higher photocatalytic efficiency, Sun et al. [133] developed a particulate hierarchical hollow tandem heterojunction photocatalyst to improve light utilization in a broad solar spectral range. It comprised two heterojunctions of MoS2/CdS NPs and hollow black TiO2/MoS2 nanosheets. To prevent photogenerated electron and hole recombination, the MoS2 nanosheets in this instance served as both a co-catalyst and a bridge to integrate hollow black TiO2 and CdS into a particulate tandem system (referred to as b-TiO2/MoS2/CdS). This architecture is different from the conventional single heterojunction architecture and similar to a Z-scheme structure. In order to harvest UV–Vis, Wang et al. [134] also developed a Z-scheme Cu3P/TiO2 photocatalyst that does not require a noble metal. The optimized Cu3P/TiO2 photocatalyst performs significantly better than TiO2 under solar light irradiation, with a hydrogen evolution rate of 607.5 mol h−1g−1, which is 30.4 times greater than the bare TiO2 photocatalyst. Cu3P/TiO2 exhibits enhanced charge carrier separation efficiency, as demonstrated by the PL spectra, transient photocurrent response, and electrochemical impedance spectroscopy (EIS). In this study, photoexcited electrons are enriched on Cu3P nanosheets for H2 evolution, whereas the holes on the TiO2 are left.
As discussed above, every kind of heterojunction has unique charge transfer properties. As the heterojunction and band alignment type are tightly associated, understanding the latter is crucial to understanding how charge carriers flow within heterojunctions. Assigning the band edge positions in the energy diagram makes it simple to draw the band alignment; therefore, understanding the band edge positions of the component semiconductors in a particular heterojunction is crucial in this regard. Additionally, band edge positions show the heterojunction’s redox potential; a stronger reduction strength is indicated by a more negative conduction band, while a stronger oxidation ability is indicated by a more positive valence band.

10.3.1. Type II Heterojunction

The Type II heterojunction is the one that has the most researchers’ interest out of all three typical heterojunctions. The level of the CB and VB of semiconductor-II is higher than that of semiconductor-I in Type II heterojunctions [135] as shown in Figure 26. Additionally, the band bending phenomena caused by the disparity in chemical potentials can be used to stimulate the passage of charge carriers in the opposing directions. The Type III heterojunctions’ band structure (Figure 12) is comparable to that of the Type II heterojunctions, with the exception that the staggered gap widens to the point where the bandgaps do not overlap [136].
TiO2 can be used in heterojunctions with some other semiconductors to boost redox reactions and slow down quick charge recombination, which enhances H2 production [57]. TiO2-based materials often combine the lower CB of photosystem II with the high CB of photosystem I (PS I) to generate conventional heterojunctions (PS II). The photogenerated holes on that VB of PS I will remain after the photogenerated electrons of PS I are transported to PS II under light irradiation. The recombination of the charge carrier produced by the light can be prevented in this fashion, improving the efficiency of the photocatalytic reaction. For various semiconductors, the reduction and oxidation reactions are simultaneously taking place. The redox reaction, however, takes place at a lower redox potential but also forgoes its capacity to be maximized.
To separate the photogenerated electron–hole pair and therefore increase the activity of photocatalytic water splitting, the Type II heterojunction is crucial. In Type II heterojunctions, the inner host, such as TiO2 (host) with CdS, has a wide bandgap whereas the sensitizers or outer shell typically has a small, mid bandgap and exhibits substantial visible-light absorption (sensitizers) [137]. TiO2-based Type II heterojunctions come in two different types due to the staggered band structures. Due to the effective charge separation, many conventional Type II heterojunctions were created using electrospun TiO2 NFs and a variety of semiconductors, such as metal oxides, metal sulfides, metal halides, and other materials. In Type II heterojunctions, photo-induced electrons flow from the more negative CB to a less negative CB while holes flow simultaneously in the opposite direction, leading to an all-around effective e/h+ separation [137].

10.3.2. p-Type or n-Type

An efficient method to increase photocatalytic performance for the production of hydrogen is to create a heterojunction with other n-type/p-type semiconductors. Because when a heterojunction is created, the local electric field that results causes holes to travel in the reverse direction from photogenerated electrons, improving the separation of photon generated hole and electron pairs and enhancing photocatalytic activity [138]. It is important to choose semiconductors for the heterojunction such that they have various band edge potentials and conducting kinds. This design offers the system various benefits, including an increase in (i) charge separation, (ii) charge carrier life, (iii) recombination resistance, and (iv) interfacial charge transit to the adsorption [139]. Because of the internal electric field produced by the charge carrier diffusion between the p-type and n-type semiconductors in p-n heterojunctions, photogenerated electrons and holes are ultimately driven in a separate direction, which reduces their tendency to recombine and increases their lifetime [137].
Compared to Type II, p-p, n-n, and p-n heterojunctions are proposed to maintain effective EHPs (electron–hole pairs), and these heterojunctions are obtained by combining both h+ rich and e rich semiconductors [140]. By creating an electric field at the interface, the heterojunction above speeds up the separation of the charge carriers. The potential difference of other linked semiconductors should influence the generated EHP migration from the surface (interface) to the interior bulk of PS. In n-n type and p-p type heterojunctions, the work function (Φ) difference is insufficient to produce a sufficiently strong electric field to isolate EHP [141]. Consequently, the photogenerated e and h+ can be effectively isolated by a p-n photocatalytic heterojunction due to the internal electric field produced by the majority carriers moving in opposite directions [142]. Before sunlight, the predominant charge e from n-type semiconductors tends to move to p-type, while h+ migrates from p-type semiconductors to n-type until both Fermi levels (Ef) are in equilibrium [143]. As shown in Figure 27b, under light irradiation, the formed EHPs efficiently separated due to the combined effect of the photoactive semiconductor’s band alignment and the generated charge region close to the p-n interface. Furthermore, since the CB and VB of p-type semiconductors are often higher than those of n-type semiconductors, EHP tunneling from one PS to another one is thermodynamically possible [144].
It is reported that in practical applications, certain external factors and surface properties such as surface states, surface adsorption, and interactions between the e-donor and acceptor, may cause band bending near the semiconductor junction interface, taking into account the charge migration mechanism. Allocation takes place between the bulk and surface of both semiconductors in order to align the Ef charge. If there is an excess of surface e, the bands will bend upward; if there is an accumulation of h+, the bands will bend downward. The separation and diffusion of produced charge carriers can be accelerated by band bending in semiconductor–semiconductor heterojunctions and the induced electric field resulting from charge redistribution [145]. In p-n junctions, the degree of band bending is somewhat controlled. Until their acceptor levels are dynamically in resonance with the Fermi level, the e acceptors receive charge from the semiconductor. A space charge region and related band bending are the outcome of the ground state charge transfer that occurs across the interface. The interaction between the substrate and the adsorbate is significantly weakened by this space charge area, which can also influence the charge transport characteristics across the interface. These parameters are difficult to regulate and not always well known in experiments.
Furthermore, the redox potential and surface Φ are used to determine the band bending close to the semiconductor surface. A photoactive semiconductor employed in the oxidative side often has a greater Φ than the reductive PS. When they come into contact with one another to create equilibrium, the semiconductor with the greater Φ causes downward bending towards the interface and vice versa [145]. The downward band bending transports the photoexcited e to the surface to carry out surface reduction reactions, whereas upward bending can move the h+ to speed up the oxidation reactions. The tunneling rate of photoexcited e+ and residual h+ in the p-n junction, as shown in Figure 27 is significantly affected by the direction of the internal electric field generated near the junction created by the contact between photoactive semiconductors with different Φ. Band bending may have an impact on two processes: (i) photocatalytic reactions by minimizing recombination and speeding up the tunneling rate of photoinduced EHP; and (ii) thermal absorption and also desorption of produced similar products. Numerous variables, including the kind of dopant employed, its concentration, particle size, surface phase, and chemistry affected the degree of band bending [146]. One charge carrier’s tunneling was increased by band bending to the surface of the semiconductor, either it is e or h+. This could enhance the associated e- or h+-mediated photocatalytic reaction occurring on the semiconductor’s surface. To preserve the band bending and neutrality of the photocatalytic material, the inverse charge carriers may also tunnel across the internal electric field in the opposite direction and become depleted [147].
p-n junctions based on transition metal oxide semiconductors have been developed and designed because of their exceptional stability. Because of their robust e-e interaction, configurational simplicity, and stoichiometric diversity, they provide an excellent broad range of photocatalytic applications [148]. Furthermore it has been observed that heterojunctions based on transition metal oxides exhibit excellent photostability and a highly effective recycling rate [149]. In order to get over these bottlenecks, a number of p-n transition metal oxide-based photocatalysts have been investigated.
A number of benefits are provided by efficient p-n type transition metal oxide-based heterojunctions, such as (i) more efficient charge carrier isolation; (ii) a rapid EHP tunneling rate; (iii) a longer photoexcited EHP lifespan; and (iv) the separation of locally inappropriate redox reactions. The p-n type heterojunctions based on transition metal oxides have a high photocatalytic output due to all of the previously described features [150].
There are several simple methods for fabricating p-n junctions based on transition metal oxides, all of which require readily available precursors. Choosing two photoactive semiconductor materials with h+ and e as the majority charge carriers is essential for effective synthesis procedures. Their distinct morphologies, band edge locations, Fermi level alignment, and interfacial characteristics all play an important role in the formation of an interfacial contact between the two photocatalysts. To achieve the optimum photocatalytic results, the features of metal oxide-based semiconductors should be completely utilized. Depending upon the morphological adjustments and requirements, their characteristics can vary greatly [151]. One or more steps may be involved in the fabrication strategies, depending on the specific requirements. p-n heterojunctions are created by adjusting a number of variables, including pH, temperature, surfactants, solvents, reaction time, and different kinds of transition metal oxides. A variety of techniques including solvothermal [152], sol-gel [153], self-assembly [154], hydrothermal [155], chemical-solution–deposition–decomposition [156], etc., are successfully applied to the construction of photoactive p-n heterojunctions based on transition metal oxides with appropriate morphology. The predicted characteristics and appearance of as-constructed p-n heterojunctions were analyzed using suitable characterization techniques in order to verify the effectiveness of the implemented fabrication routes.
The most facile and economical ways to modify the surface characteristics of metal oxides are by hydrothermal and sol-gel synthesis, among the other synthetic processes already described. The main benefit shared by both approaches is the achievement of stable surfaces and a large surface area. Therefore, the majority of publications have used sol-gel and hydrothermal techniques to create p-n junctions based on transition metal oxides. We examine the photodegradation ability of the NiO/TiO2 p-n junction from several research papers in order to investigate the impact of fabrication procedures on photoactivity. The two published papers discussed the synthesis of p-n heterojunctions in response to a model methylene orange pollutant [157]. Lin et al., constructed a p-NiO/n-TiO2 composite by employing the chemical-solution–deposition–decomposition technique and discovered a 92% photodegradation rate for methylene orange. In contrast, Jiang et al. [158] synthesized a NiO/TiO2 composite using the sol-gel method, which remarkably displayed reduced photooxidation (68.3%, methylene orange) and higher photoreduction (87.0%). Thus, the development of the p-n heterojunction is entirely dependent on the employed manufacturing techniques, with the final objective being the coupling of p- and n-type semiconductor photocatalysts to maximize their respective properties.
Improved illumination absorption, photocorrosion stability, charge transfer efficiency, and ultimately photocatalytic efficiency are all provided by the construction of TiO2-based p-n heterojunctions. This section contains information on various TiO2-based p-n heterojunctions to enhance the properties of UV-active TiO2 catalysts. Scientifically, under the influence of the internal electric field, excited e present in the n-type CB is directed to the VB of p-type semiconductors, promoting the EHP isolation rate. Because of its low sunlight-absorption range and extremely fast charge carrier assembly rate, onefold e-rich TiO2 is not usually a perfect semiconductor. Zeolites are effective materials with a large surface area, a low environmental impact, and unique photophysical characteristics that regulate the rate of EHP migration. Improved visible-light-responsive and extremely efficient CuO/TiO2-zeolite (Na1.89Al2Si2.88O9.68) p-n hetero nanostructures were reported (Zhao et al.) using the conventional impregnation method, providing evidence for this claim. CuO is a significant transition metal oxide semiconductor with a p-type narrow bandgap with 1.2–1.5 eV that interacts with n-type wide bandgap TiO2 with 3.2 eV. The CB, VB edge, and Ef of TiO2 photocatalysts are higher than those of a CuO semiconductor prior to linking p-type CuO and n-type TiO2 photocatalysts. During the coupling process, upshifting and downshifting of the Ef in CuO and TiO2, respectively, was noticed, until the equilibrium stage was attained. At equilibrium, an effective p-n junction with a strong inner electric field was identified. This allowed photoexcited e from the p-CuO CB to n-TiO2 and residual h+ from the n-TiO2 VB to p-CuO, respectively, to migrate quickly. The remaining h+ on TiO2 and isolated e on CuO were free to start photocatalytic redox reactions with the surface-adsorbed methylene blue (MB) dye. The photostability of an as-fabricated visible-light-responsive p-n CuO/TiO2-zeolite heterostructure was demonstrated by its efficient reuse without any degradation in its photocatalytic function [159].
Similarly, mesoporous anatase TiO2 hollow spheres having a low density, a large surface area, and crystalline nature showed exceptional optical output. For carrying out photoreactions, hybrid TiO2 nanostructures demonstrated a configurable structure and also surface morphology. Wang et al., adopted the sol-gel technique to synthesize hollow TiO2 coupled with NiO hybrid cells followed by a hydrothermal approach for H2 evolution and RhB degradation. The usage of photoactive NiO (3.55 eV) and TiO2 (3.3–3.2 eV) presents two main practical challenges: (i) wide bandgap energy, which demands high energy photons with wavelengths below 387 nm (UV area) and (ii) fast charge carrier recombination. It can successfully produce a p-n heterojunction with an n-type TiO2 semiconductor, regardless of the NiO wide bandgap. As-fabricated hollow TiO2 nanospheres had an absorption edge that was shifted from 397 to 553 nm, which increased the concentration of NiCl2-6H2O. As seen in Figure 28, anchored p-type NiO on the surface of n-type TiO2 was stimulated by visible light to produce effective charge carriers after coupling. The inner electric field at the interface between n-type TiO2 and p-type NiO was established through the migration of photoexcited e to the +ve side and photoremaining h+ to the –ve side. h+ in the VB of NiO interacted with H2O to make highly reactive ∙OH radicals, while excited e in the CB of TiO2 were trapped by O2 to produce ∙O2. Highly fluorescent 2-hydroxyterephthalic acid was produced when terephthalic acid reacted with ∙OH, as demonstrated by the increased intensity to 425 nm in the photoluminescence (PL) spectrum, which verified the production of highly active ∙OH radicals [160].
Additionally, it is examined that connecting one TiO2 molecule to another is a useful tactic for creating a nanohybrid with extensive properties. Hybrid TiO2 structures were reported to achieve a 99.57% photodegradation efficiency with effective anodization followed by the impregnation approach in research [161]. As-synthesized n-type TiO2 nanotubes (TONTs) were combined with another photoactive TiO2 that had extra aluminum doping (overlayer) to create a h+-rich TiO2 in order to take advantage of their structural benefits. Tiny p-n heterojunctions were created close to the TONT/Al: TONT interfaces when Al2O3 was deposited over the surface of TiO2, creating a significant surface area for interaction. According to the X-ray photoelectron spectroscopy (XPS) spectra, doping of Al3+ ions in the lattice formed the acceptor level near the VB maxima and replaced some of the Ti4+ ions in the host. The semiconductors’ band structures, as shown in Figure 29, showed that the p-TONT’s CB and VB were both at higher energy levels than those of the n-Al: TONT, indicating that the EHP migration followed the Type II pathway rather than the Z-scheme [161]. Thus, it can be observed from the aforementioned publications that the development of TiO2-based p-n junctions accelerated the development of their optical, physical, and chemical characteristics and significantly increased their total photocatalytic efficiency.
There are many techniques which offer the most important understanding of p-n junction faults or impurity (extrinsic) recombination states. This is a summary of some of the advanced characterization techniques, including cathodoluminescence (CL), transmission electron microscopy (TEM), electroluminescence (EL), scanning electron microscopy (SEM), electron beam-induced current (EBIC), deep-level transient spectroscopy (DLTS), scanning near field optical microscopy (SNOM), electron beam-induced current (EBIC), and secondary ion mass spectrometry (SIMS). Using electron beams, cathodoluminescence and electroluminescence are potent, costly, and labor-intensive methods for examining optoelectronic characteristics. Initially stimulating electron–hole pairs, the electron beam also aids in the local property detection of luminescence radiation. The matching defect structure distribution in the semiconductor junction is shown by the spatial distribution variation luminescence map. Defect kinds and various resistance distributions can be related to the luminescence emission wavelength.
When dealing with non-radioactive p-n junctions, the approach still loses its effectiveness. SEM or scanning tunneling microscopy (STEM) is used in conjunction with the electron beam-induced current mapping technique to examine the characteristics of defects and minority carrier diffusion lengths in p-n junction semiconductors. To measure the fluctuations in current across the p-n junction, electron beams are scanned over the entire sample; the defect states trap the electrons, lowering the current value. This technique works well for transitions that are radiative or non-radiative. Moreover, in order to convey different contrast mechanisms, NSOM uses optical excitation to scan the defect map signals transmitted through optical fibers.
Similar to this, DLTS is an extremely sensitive method that provides detailed information about electrically active flaws concealed within the forbidden gap. While the defect charge state retrieval roots the capacitance transient in accordance with its population density, the device scans at a constant capacitance value. TEM, a significant microscopic technology, was used to directly examine defect images of the p-n junctions with atomic-level resolution. Since TEM imaging uses a small quantity of the sample and does not properly validate the faults in the complete sample, its ability to confirm defects is indisputable. One sensitive surface analysis method is SIMS, which uses primary ions to sputter the sample layer. Additionally, elemental and crystalline defects are carried by secondary ion emission from the overlying surface, together with the diffusion-broadening characteristics of the p-n junctions.
The capacitance–voltage (C–V) method is another reported approach which is used for determining the breakdown voltage and doping density. However, it is discovered that this type of methodology makes data interpretation and analysis difficult. Above all, it can be challenging to calculate the accurate and real gadget properties. The p-n junction I-V method is a suitable approach that offers far more material information needed for the actual device construction and functioning. In both perfect and imperfect semiconductor crystal structures, recombination is a prevalent phenomenon. Excess recombination occurs in imperfect semiconductors because of flaws and impurities (traps) [106].

10.3.3. Z-Scheme Heterojunction

The direct Z-scheme photocatalyst idea was initially put forth in 2013 [163]. The band configuration and electron migration process of Z-scheme heterojunctions are shown in Figure 30. The electron transfer route among semiconductors is different in Z-scheme hetero junctions compared to Type II heterojunctions, although they have an identical band configuration. The path taken by electrons as they move between semiconductors resembles the letter “Z” in English [164]. The photogenerated electrons in semiconductor-II having a lower reduction ability mix with the photogenerated holes in semiconductor-I with a lower oxidation ability in the photocatalytic reaction. As a result, it is possible to preserve the photogenerated electrons in semiconductor-I having a high reduction ability and the photogenerated holes in semiconductor-II having a high oxidation ability.
Additionally, the photogenerated electron from semiconductor-II will migrate to semiconductor-I due to the electrostatic attraction between photogenerated electrons in the semiconductor-II CB and photogenerated holes in the VB of semiconductor-I, whereas in the Type II heterojunction, the electrostatic repulsion between the photogenerated electrons of semiconductor-II will prevent this from happening.

Z-Scheme System with Shuttle Redox Mediators

The redox mediated Z-scheme, also known as the PA-A/D-PS (first-generation) system, is made up of two semiconductors, PS I and II, and an A/D pair acting as an electron shuttle. Two set of charge carriers will delocalize in distinct semiconductors in such a system, causing water to divide into its component parts, H2 and O2. There is no physical contact between PS I and II, and the A/D pair controls the cascade type of vectorial electron transmission from PS II to PS I. The electrons from the CB of PS II will first reduce the electron acceptors into electron donors, leaving photogenerated holes accumulating in the VB of PS II, after the formation of electron–hole pairs as a result of photoexcitation in both PS I and II (Equation (23)). In the meantime, the photogenerated holes from the VB of PS I will transform the generated electron donors back into their oxidized state of being electron acceptors (Equation (24)). As a result, photogenerated electrons from PS II are indirectly shuttled to recombine with the help of an A/D pair, leading to effective isolation of the electrons and holes in PS I and II, respectively.
As a result, the system’s lowest VB and highest CB accommodate the photogenerated charge carriers, providing a sizable big overpotential for the Z-scheme reactions (Equations (25) and (26)) [165]. As such, it is possible to drive photocatalytic overall water splitting using this Z-scheme system while maintaining redox pairs’ generation.
A + e (CB of PS II) → D
D + h+ (VB of PS I) → A
2H2O + 4h+ (VB of PS II) → O2 + 4H+
2H+ + 2e (CB of PS I) → H2
The red dotted lines in Figure 31 reveal that the electron A/D pair in the PS-A/D-PS system can also react with the photogenerated electrons and holes in the CB of PS I and the VB of PS II, which will inevitably result in a decrease in the effective number of charge carriers. As a result, the simultaneous evolution of H2 and O2 gas in the stoichiometric ratio attributed to the backward reaction is relatively challenging for a PS-A/D-PS Z-scheme system. With respect to this issue, several surface treatments are being implemented, such as loading metal co-catalysts, depositing rutile TiO2, and exchanging Cs-H+; these are some of the attempts to hinder the backward reaction by impeding the adsorption of electron donors on PS II and electron acceptors on PS I.
At present, the redox mediators that are frequently utilized in the PS-A/D-PS Z-scheme system are Fe3+/Fe2+, IO3−/I, [Co(bpy)3] 3+/2+, [Co(phen)3] 3+/2+, and VO2+/VO2+ [166,167]. Abe et al., for example, presented early work on the photocatalytic Z-scheme system. Anatase and rutile, two distinct forms of TiO2, are interfaced via the IO3−/I redox pair as an ionic reaction mediator, serving as matching HEPs and OEPs [168]. Due to the wide range of valencies from −1 to +7, aqueous iodine is capable of acting as an electron mediator. Equations (27) and (28), however, demonstrate that I(0/1−) and I(1−/5+) are more stable during redox cycles in aqueous medium. The I3/I ionic pair is commonly used in dye-sensitized solar cells due to its facile reversible redox reaction and low photoabsorption, which only reaches about 500 nm [169]. However, because of the poor reactivity of I3 as an electron acceptor and the restricted selection of photosystems, the application of Z-schemes employing an I3/I redox cycle is highly limited [170]. The IO3−/I redox pair, on the other hand, has a standard potential that is close to the I3/I couple. Because of its electron nature, IO3−/I can thus be incorporated as an ionic redox mediator.
I3 + 2e → 3I E° versus NHE = +0.536 V
IO3 + 6e + 3H2O → I + 6OH E° vs. NHE = +0.76 V
Figure 31. Schematic band energy diagram of the PS-A/D-PS Z-scheme system. Reprinted with permission from Ref. [169]. Copyright 2020, Wiley Online Library.
Figure 31. Schematic band energy diagram of the PS-A/D-PS Z-scheme system. Reprinted with permission from Ref. [169]. Copyright 2020, Wiley Online Library.
Catalysts 14 00674 g031
Even so, the IO3−/I redox cycle’s efficacy is greatly influenced by the pH of the medium. According to the Nernst equation [171], this ionic pair shows a higher contribution under increasing pH because of the shifting of the standard potential. Under UV light, Abe’s work involved the photocatalytic gas generation of a NaI solution containing Pt-loaded anatase TiO2 (Pt-TiO2-A1) as PS I and rutile TiO2 (TiO2-R2) as PS II [172]. At pH 3, the system exhibits a slow gas evolution that is attributed to the dissociation into I3 as the primary redox product, rather than IO3, which, as previously indicated, is not capable of acting as an effective electron acceptor in photocatalysis. However, the presence of favorable IO3−/I ionic couples, as opposed to I3−/I, makes a basic solution of pH 9 the ideal environment for the IO3−/I redox system to work efficiently. The nonstoichiometric evolution of H2 and O2 is thought to be caused by the induction period of IO3− production with a certain amount of I3, which is represented by the progressive increase in activity from pH 5 to 9. In addition, the development of inefficient I3 will also cause some shielding effect because of their large extent of light absorption. Therefore, it can be said that when the pH is higher than 9, the IO3−/I redox cycle can function effectively since there is no I3 byproduct, leading to the stoichiometric evolution of H2 and O2 with high activity. Furthermore, in such a system, the ratio of IO3−/I has a major influence on the rate of gas evolution. The reverse reaction, also known as competitive oxidation, will result from the increased I concentration. For the PS-A/D-PS system to function with the IO3−/I redox mediator, an appropriate pH and NaI concentration are therefore essential.

Direct Z-Scheme Hydrogen Production

In order to build Z-scheme heterostructures using the boron-carbon nitride molecule, TiO2 has been shown to be firmly connected. In comparison to pure TiO2, the TiO2/BCN nanostructure develops a higher concentration of photocatalytic H2 generation activity and improved stability. The performance of the photocatalytic system is significantly influenced by the loading of BCN. For a 4% BCN/TiO2 nanocomposite, the maximum photocatalytic activity was attained at a hydrogen evolution rate of 19.7 µmol h−1g1, which is two times greater as compared to using pristine TiO2. The presence of middle sub-band single-electron oxygen vacancies with an innovative forbidden gap of TiO2 has been found to be advantageous for the absorption of visible light in this sort of charge transfer mechanism. Due to the creation of an ohmic contact somewhere at the interface between BCN and new TiO2, oxygen vacancies also benefit the Z-scheme charge transfer. Resultantly, the Z-scheme charge transfer in the BCN/TiO2 nanocomposites can provide an outstanding TiO2 oxidation ability and a strong BCN reduction ability [83].

10.4. Comparison of Type II Heterojunctions and Z-Scheme Heterojunctions

By using two semiconductors to produce Type II heterojunction nanocomposites, we can increase their charge efficiency and photocatalytic activity. The lifespan of photogenerated carriers can be increased by creating a heterojunction-based photocatalytic device that induces the separation of electrons and holes to two different sites [173]. The OER on O2 evolution photocatalysts (OEPs) and the HER on H2 evolution photocatalysts (HEPs) are two distinct photocatalytic reactions where gas evolution takes place. As shown in Figure 32, the migration of electrons and holes to a more electropositive CB and electronegative VB potential ascribed to the nature of charge transfer weakens the photocatalyic system’s redox ability even though heterojunction-type semiconductor nanocomposites are competent in facilitating charge separation.
In order to overcome these obstacles, the second method of photocatalysis, which is known as the Z-scheme photocatalytic system, introduces an anisotropic arrangement of two photocatalysts and an electron mediator to split water using a two-step photoexcitation process. The biomimetic artificial Z-scheme photocatalytic system, which draws inspiration from the natural photosynthesis of green plants, can concurrently meet the three demands listed below, which are lacking from both single-component photocatalysts and heterojunction-type nanocomposites: (i) an appropriate band-edge position with a sizable overpotential, (ii) small bandgap, and (iii) suppression of the recombination of electron–hole pairs [174]. Efficient water splitting can be achieved through Z-scheme photocatalysis, also referred to as a two-step photocatalytic system by the combined action of two isolated photocatalysts, one of which acts as the reduction site (photosystem I or PS I) and the other as the oxidation site (photosystem II or PS II). An electron transport chain known as a mediator is used to make it easier for electrons to flow between the two photosystems.
Z-scheme semiconductor composites, in contrast to heterojunction-type semiconductor composites, allow for a distinct electron flow profile that is ascribed to the inclusion of an electron-relaying channel. Electrons from the VB of PS I and II will be excited to the CB upon photoexcitation, leaving the photogenerated holes in the VB. The electron mediator will then move the photogenerated electrons in PS II to recombine the holes from the VB of PS I through ohmic contact [175]. Strong redox capabilities are maintained while electrons and holes can be accommodated into two distinct photocatalysts simultaneously, thanks to this unusual type of vectorial electron transfer. Thus, PS I is abundant with electrons for HER, whereas PS II is a hole-rich photocatalyst for OER. Resultantly, a Z-scheme system requires less change in Gibbs free energy to drive each photosystem than single-component photocatalysts and heterojunction-type composites [176]. It is evident from Figure 32b that the rational design of a Z-scheme provides effective charge separation in a separated position with a comparatively significant overpotential, which is more than enough to regulate the exceptional redox reaction.

10.5. Surface Plasmon Resonance (SPR) Effect

It was suggested that decorating the surface of TiO2 with noble metal particles like Au, Ag, and Pt can enhance its photocatalytic activity [177]. Three factors contribute to the enhanced photocatalytic activity of metallic particle-coated TiO2. First, by attracting the photogenerated electron and leaving a hole in TiO2, metallic nanoparticles can boost the quantum yield. Second, metallic nanoparticles’ SPR effect can boost the ability of TiO2 to absorb light and extend its sensitivity to visible light. The SPR effect is the cumulative rational oscillation of unpaired electrons on metal nanoparticles brought on by visible-light irradiation. Last but not least, SPR can strengthen the localized electric field around metal particles that interact with the coupled semiconductor and easily form electron–hole pairs in the semiconductor’s near-surface region [121].
It is well known that the LSPR (which is actually the collective vibration of the free electrons in metal nanoparticles) plays a key role in describing their optical properties. LSPR is a crucial component of biosensors because they are highly sensitive to the detection of molecularity-specific interaction analysis [178]. The aggregate vibrations with the interaction of resonant photons of metal nanoparticles of electrons like Au, Ag, and Pt are known as surface plasmon resonance. Often, plasmonic oscillation is limited to metal nanoparticles on surfaces with sizes between 10 and 100 nm [179].
LSPR, or the interface between such a conductor and a non-conductor, is where this phenomenon also occurs. Additionally, the LSPR behavior of a plasmonic material allows it to increase its absorption range into the visible region [180]. A near-field enhancement phenomenon called LSPR has the ability to capture light at the nanoscale with the highest optical absorption.
An appealing method for creating photocatalysts that can absorb visible light is to use the LSPR phenomena. Numerous studies have been conducted on the LSPR phenomena of noble metals like Au, Ag, and Pt [181]. Additionally, the compound photocatalyst’s capacity for light harvesting and the strength of the local electric field can both be enhanced with localized surface plasmon resonance of Au NPs. Plasmon-induced hot electrons can penetrate the Schottky barrier to enter semiconductors in Au–semiconductor nanocomposites, increasing the photocatalysis efficiency [182].

10.6. Co-Catalyst

One of the multiple functions of co-catalysts in photocatalytic reactions is to lower the rate of recombination [183]. This is accomplished in the case of metals by electron transfer to the co-catalyst, which happens because metals have a lower Fermi level than TiO2 [184]. By encouraging the third stage of the water-splitting process, named as reducing the activation barrier of surface reduction and oxidation reactions, co-catalysts enhance H2 evolution [185]. The mechanism of water splitting with both reduction and oxidation co-catalysts is shown in Figure 33. Pt was said to be a more successful co-catalyst for water splitting. Under UV irradiation, Maeda demonstrated stable overall water splitting into oxygen and hydrogen using rutile TiO2 nanoparticles modified with Pt nanoparticles [186]. It is interesting to note that, as discovered by Sayama and Arakawa [187], Pt-loaded anatase TiO2 photocatalyzed solely the H2 production from water and not overall water splitting, under identical conditions. When co-catalysts are introduced to the surface of TiO2, the bandgap can be reduced, allowing the photocatalytic activity to reach the visible light spectrum. The so-called surface plasmon resonance (SPR) effect is present in some metals. It is commonly known that Pt, Ag, and Cu exhibit a strong SPR effect [119].
Proton reduction into hydrogen and water oxidation to generate oxygen are the two half-reactions that normally make up water splitting into hydrogen and oxygen. The rate-limiting step between the two is water oxidation, since it requires four photogenerated holes. Furthermore, the water oxidation reaction is slower than the proton reduction reaction by approximately five orders of magnitude [188].
Ru is the least expensive element in the platinum group [189], so it is a desirable co-catalyst for industrial use. RuO2 can reside on the base photocatalyst in both its oxide and metallic form. The metallic form of Ru enhances the hydrogen evolution reaction (HER) [190], while stimulates the oxygen evolution reaction (OER), thus contributing to the overall water-splitting reaction collectively [191]. On the other hand, compared to equally active Ir NPs, Ru nanoparticles have been demonstrated to be less stable in the long run [192]. Currently, Ir and its oxides are thought to be the most stable and effective catalysts [193]. As the rate-limiting step in the water-splitting reaction, IrO2 functions as an OER catalyst [192].
While single metal co-catalysts might perform better in one of the two required reactions for water splitting, a combination of co-catalysts may work better for the overall effect [194]. Commonly, metal oxides such as RuO2 and IrO2 are considered as oxidation co-catalysts, while noble metals like Pt, Ag, and Pd are regarded as reduction co-catalysts [195]. One can further increase the hybrid material’s overall water-splitting activity by integrating both the reduction and oxidation co-catalyst on the same photocatalyst [196]. The literature has a number of studies on noble metal co-catalysts on TiO2. The activity of Pt-Ir, Ir-Ru, Pt-Ru, and Pt-Ru-Ir combinations in nanoparticles deposited over rutile TiO2 and anatase in an oxygen evolution reaction (OER) was examined by Fuentes et al. [197]. Tanaka et al. [198] synthesized TiO2 by depositing large Au nanoparticles and smaller Pt nanoparticles and checked the H2 production rate. Rutkowska and Kulesza [199] placed Pt-Ru nanoparticles onto TiO2 supported on multiwalled carbon nanotubes for ethanol oxidation. Au, Pd, Pt, Ir, Rh, Ag, and Ru were deposited onto TiO2 substrates by Lima et al. [200] in order to catalytically oxidize benzyl alcohol to benzaldehyde. Pt and Ru co-catalysts (alone and combined) were deposited by Caudillo-Flores et al. [201] onto Nb-doped TiO2 for the photocatalytic production of H2 from a methanol–water mixture.
Figure 33. Mechanism of water splitting with both reduction and oxidation co-catalysts. Reprinted with permission from Ref. [196]. Copyright 2019, Elsevier.
Figure 33. Mechanism of water splitting with both reduction and oxidation co-catalysts. Reprinted with permission from Ref. [196]. Copyright 2019, Elsevier.
Catalysts 14 00674 g033

11. Observation of Charge Carrier Kinetics in Heterojunction Structure

11.1. Transient Absorption Spectroscopy (TAS)

As mentioned in the introduction, non-radiative electron–hole recombination, which takes place before surface interactions between electrons and holes with water, is typically responsible for the moderate efficiencies of semiconductor water splitting. Transient absorption spectroscopy (TAS) is a widely used technique to assess the decay kinetics of carriers, since their lifetimes are on the scale of femtoseconds to nanoseconds, and even seconds after light absorption [202]. Although its application to PEC materials has been established, it is primarily utilized to study the kinetics of charge carriers in solar cell materials. A lock-in amplifier, a laser capable of producing femtosecond light pulses, and a device to analyze the absorption spectra as a function of time, wavelength, or applied bias make up the experiment’s setup [203]. Either the electron transient decay or the hole transient decay can be monitored with the appropriate scavenger. The technique by which an excitation (or pump) pulse promotes a fraction (0.1–10%) of the molecules to an electrically excited state is described by Barera et al. [203]. A weak probe pulse, which is one that is transmitted through the sample with a delay τ relative to the pump pulse, is one that is so weak that multiphoton/multistep processes are avoided during probing. The difference absorption spectrum can be calculated by subtracting the stimulated sample’s absorption spectrum from the ground state sample’s absorption spectrum (ΔA). A function of τ and wavelength λ, ΔA(λ,τ), is obtained by recording a ΔA spectrum at each time delay and changing the time delay (τ) between the pump and the probe. Consequently, since suppressing recombination and increasing carrier lifetimes are the primary goals of heterojunction systems, this is highly helpful for researchers studying solar fuel.
It was discovered, using nanocrystalline TiO2 (nc-TiO2) as an example, that the carrier lifetime was highly dependent on the pulse intensity and that recombination happens on the order of microseconds, while water oxidation takes place on the order of microseconds. Looking further, the absence of activity for water oxidation in N-doped TiO2 under visible excitation was attributed to the photoholes created by the rapid decay of visible light, which happens on a much faster time scale than that needed for water oxidation as compared to nc-TiO2 [204]. Pendlebury et al. used this method to study other visible-driven photoanodes and found that for α-Fe2O3, the amplitude of the long-lived hole signal is only ~10% of the initial hole signal. This suggests that most photogenerated holes still undergo rapid electron–hole recombination on a microsecond timescale, even under positive applied bias [205]. Subsequently, it was observed that the primary factor limiting water oxidation in α-Fe2O3 is recombination rather than surface kinetics [206]. As expected, the application of Co-Pi to the surface caused the observation of a cathodic shift in photocurrent as well as the development of long-lived hematite photoholes [207] as a result of the inhibition of electron–hole recombination.
The width of the space charge layer and Fermi level pinning decreased in the presence of surface catalysts, increasing the electron depletion layer’s size. When RuOx particles were added to the surface of visible-light-activated Cu2O photocathodes, the yield of long-lived (>100 μs) Cu2O electrons increased significantly (as shown in TAS, Figure 34). This increase was attributed to a decrease in fast electron–hole recombination losses caused by hole transfer from Cu2O to RuOx, which increased the spatial separation of electrons and holes and aided in the photooxidation reaction by holes [208].
Ma et al. [209] recently reported on the correlation between the applied electrical bias and the yield of long-lived (0.1–1 s) photogenerated holes in BiVO4. This correlation is attributed to the kinetic competition between water oxidation and the recombination of these surface-accumulated holes with bulk electrons across the space charge layer. Importantly, it was discovered that two mechanisms: firstly, fast (≤μs) electron–hole recombination, and secondly, the recombination of surface-accumulated holes with bulk BiVO4 electrons, limit photocurrent generation in BiVO4 photoanodes. For WO3, Cowan et al. observed that quick (<μs) electron–hole recombination dominates in the absence of an electron scavenger, and water oxidation requires the production of long-lived holes with a lifetime in the milliseconds-to-seconds time scale [210]. Nonetheless, it is evident from the literature that this method should be applied to other junction architectures as it may be useful in demonstrating charge dynamics experimentally.

11.2. Electrochemical Impedance Spectroscopy (EIS)

Along with TAS, other methods that can investigate the kinetics of charge transfer and recombination are surface photovoltaic spectroscopy, photoluminescence spectroscopy (PL), and electrochemical impedance spectroscopy (EIS). Since radiative recombination, which is typically regarded as a minority phenomenon, is necessary for PL spectroscopy, it will not be covered in depth in this paper.
Applying a small amplitude sinusoidal AC voltage, V(t), and measuring the amplitude and phase angle (with respect to the applied voltage) of the resultant current, I(t), forms the basis of electrochemical impedance spectroscopy (EIS) [211]. The impedance (Z) can then be computed using Ohm’s Law, which states that Z = V(t)/I(t). The fraction of the hole current jh that is measured in the external circuit in EIS is determined by competition between the overall rates of hole transfer and recombination (Figure 35).
The rate constants kt and kr (s−1) in the EIS response equations are used to represent the rates of recombination (cm−2 s−1) and hole transfer in terms of the surface concentration of “trapped holes”. It follows that kr should be dependent on band bending, qΔΦ, since the electron concentration at the surface is given by
nx=0 = nbulk exp (−qΔΦ/kB T)
where nbulk is determined by the doping density.
By using this method, one can learn more about the phenomenological rate constants that describe the competition between recombination and charge transfer during the oxidation of water.

11.3. Surface Photovoltage Spectroscopy

This non-destructive method measures changes in band bending at the free semiconductor surface as a function of external illumination and can yield a wealth of qualitative and quantitative data for the heterojunction photocatalysis researcher. The surface photovoltage (SPV) is defined as the illumination-induced change in the surface potential. This includes, but is not limited to, the band position’s relative locations, defect states, surface dipole, diffusion lengths, recombination rates, the degree of band bending, and flat-band potentials.

12. Theoretical Modelling of Photocatalyst Junction Structures

There will be a discontinuity in the Fermi level at the interface when two materials with different Fermi levels first come into contact. In order to reach equilibrium, this will cause an electron transfer from the material with the higher Fermi level to the one with lower Fermi level. Moreover, the accumulation of a dipole layer at the interface between two materials due to the flow of electrons between them may further alter the respective band locations [213]. Therefore, in order to comprehend and forecast the behavior of semiconductor heterojunctions for photocatalysis, a number of physical mechanisms must be modeled. These mechanisms include the relative locations of the band edges of individual materials, which set the thermodynamic limits that establish the band offsets; the electronic structure of the materials, which provides details of the charge separation characteristics of photocatalysts such as electron and hole mobility; and the junction’s nature. This information can be used to calculate the band alignment, band magnitude, and impact of any interface dipoles present. Usually, these computations can only simulate systems up to a maximum size of a few thousand atoms [214].
Due to these size limitations, and since ab initio force calculations are usually most efficiently performed using a plane wave basis set, simulations typically concentrate on crystalline materials that are representative of bulk systems. Moreover, only ground state characteristics like cohesive energy and chemical bonding are precisely calculated using DFT. The accurate determination of excited state parameters is lacking, including the bandgap and optical absorption spectra. The lack of derivative discontinuity in the exchange correlation functionals prevents DFT from achieving the accurate determination of bandgaps, and bandgaps are reported in the literature as the simple difference in Kohn–Sham eigenvalues between the CBE and the VBE [215].
There have been a number of workarounds put in place to increase DFT bandgap accuracy. For metal oxide systems, two methods are frequently employed: hybrid functionals, which incorporate a fraction of the exact Hartree–Fock exchange to induce electron localization on particular sites, and DFT+U, which treats the localization problem of DFT195 by adding an empirical on-site Coulomb term +U [216]. The time-dependent Density Functional Theory method, also known as the GW approach, is often reserved for investigations involving clusters and will not be discussed in depth in this review.
Through photocatalytic water splitting, dye-sensitized photocatalytic (DSP) devices have shown promise as a sustainable method for producing hydrogen (H2). Usually, a dye molecule is adsorbed on a semiconductor nanostructure and used as a photosensitizer in DSP systems. As illustrated in Figure 1, the dye molecules in the DSP system absorb incident light and introduce excited electrons into the semiconductor’s conduction band (CB). Reduction reactions, such as the reduction of water to produce H2, can be accelerated by these excited electrons. With the help of the electron donor, the dye molecule can be renewed for cyclic usage [217].
Significant advancements have been made in the field over the past 20 years, and research has demonstrated that DSP systems can be created to cover the visible and even near-infrared portions of the solar spectrum by carefully choosing and matching the energy levels of semiconductors and dye molecules. Even with these tremendous advancements in the industry, DSP systems continue to have serious shortcomings, including low light-absorption efficiency [218]. It has recently been shown that dye molecules in dye-sensitized photocatalytic and photovoltaic systems can have their light-absorption efficiency increased by using the plasmonic Mie resonances of metal nanostructures, such as those made of gold (Au) and silver (Ag) [219]. Moreover, Mie resonance plays an important role in the selectivity and activity of the catalysts for energy efficient photocatalysis [220].
Under Mie resonance conditions, plasmonic metal nanostructures (PMNs) can show very high absorption cross-section values, up to five orders of magnitude higher than dye molecules. Thus, through a variety of mechanisms, such as the nanoantenna effect and plasmon-induced resonance energy transfer, PMNs (such as Ag and Au) can capture a significant portion of incident light, transfer the energy into the surrounding dye molecules, and increase their light-absorption efficiency [221]. In DSP systems, these plasmonic Mie resonance-mediated effects are used to speed up dye sensitization. But these PMNs-based DSP systems also come with built-in difficulties, like incompatibilities with traditional semiconductor fabrication, expensive material requirements, and commercialization problems that become more complicated because they require both metals and semiconductors.
Here, we suggest a different strategy, dielectric Mie resonance-enhanced dye sensitization, to address the previously noted problems. Figure 36 shows this strategy schematically. The Mie resonances can be found in dielectric (ε > 0) and plasmonic (ε < 0) materials with moderate (2.5–3.5) and high (>3.5) refractive index values, respectively [222].
Therefore, DSP systems based on metal oxide semiconductors with moderate-to-high refractive index values are most suited for the dielectric Mie resonance-enhanced dye sensitization shown in Figure 36. These metal oxide semiconductors include titanium dioxide (TiO2), hematite iron oxide (α-Fe2O3), cerium (IV) oxide (CeO2), and cuprous oxide (Cu2O).
The dielectric Mie resonances of nanostructures of medium- and high-refractive-indexed metal oxide semiconductors can transfer the photonic energy into the adsorbed dye molecules via mechanisms like the resonance energy transfer, nanoantenna effect, and Mie resonance-mediated intense scattering effect, just like the plasmonic Mie resonances of PMNs. It is anticipated that the dielectric Mie resonance effects will improve the dye molecules’ light-absorption efficiency in the DSP systems (Figure 36c). One significant distinction is that a dielectric Mie resonance-based system (Figure 36c) does not require a separate light-enhancing material, in contrast to the plasmonic Mie resonance-based system (Figure 36b). This system’s primary benefit is that it only needs a dye and a dielectric semiconductor nanostructure.
In particular, the dielectric semiconductor nanostructure can function as both a source for obtaining excited electrons from the dye molecules and a Mie resonator for increasing the dye molecule’s light-absorption efficiency and promoting the reduction reaction (Figure 36c).

13. Conclusions and Future Perspectives

Hydrogen production is the best way to meet the present energy requirements, and there are various methods and sources to generate hydrogen. The best among them is solar water splitting because it is an economically viable and environmentally sustainable approach. Numerous semiconductors have been reported for solar water splitting, including titania, which shows many photocatalytic traits but faces a few limitations. A variety of modifications have been reported to upsurge its water-splitting activity, including doping, designing various heterojunctions, and sensitization. However, there are several challenges that must be addressed to enhance solar energy harvesting for improved kinetics of the water-splitting process. From our perspective, the following strategies should be implemented to improve the hydrogen production rate by TiO2:
  • Novel schemes should be developed that could maximize the light-harvesting capability of the TiO2 which in turn could offer a higher water-splitting rate to yield hydrogen.
  • Designing dual-catalytic TiO2-based setups by mimicking the natural photocatalytic system is the best way to achieve sustainable energy. In this approach, another semiconductor is coupled with TiO2 that modifies its electronic states and offers a suitable path for the enhanced kinetics of water splitting to generate hydrogen.
  • The addition of a dopant tunes the bandgap of TiO2, thus broadening its solar energy harvesting in the visible region, which is the first priority of the researchers for achieving a maximum energy input and minimal catalyst dose for the water-splitting reaction.
  • Furthermore, a thorough understanding of the underlying reaction mechanism is a prerequisite to upsurge the catalytic proficiency of TiO2 by improving its charge segregation, charge utilization, and reduced charge recombination rate.
Keeping in view the above modification strategies for titania, the more effective water-splitting setups along with highly efficient catalysts can be designed, which can prove their metal in augmenting the rate of solar-to-hydrogen conversion efficiency.

Author Contributions

T.I., conceptualization, supervision, writing—original draft, writing—review and editing; Z.E., writing—original draft, resources, methodology, writing—review and editing; A.Q., writing—original draft, investigation, writing—review and editing; Y.A., conceptualization, supervision, formal analysis (literature survey) writing—original draft; A.I., data curation, visualization, formal analysis (literature survey) funding acquisition, writing—review and editing; S.A.A.-H., funding acquisition, visualization, data curation. formal analysis (literature survey), writing—review and editing; M.A.I., data curation, writing—review and editing, formal analysis (literature survey); M.E.A.Z., funding acquisition, project administration, investigation, formal analysis, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU).

Data Availability Statement

All the data of this study are contained in manuscript. For any additional data or information needed regarding this research, please contact the corresponding authors or [email protected] (T.I.).

Acknowledgments

The authors express their gratitude to the Deanship of Scientific Research, Imam Mohammad Ibn Saud Islamic University, Saudi Arabia, for supporting this research work.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ahmad, A.; Tezcan, F.; Yerlikaya, G.; Paksoy, H.; Kardaş, G. Three dimensional rosette-rod TiO2/Bi2S3 heterojunction for enhanced photoelectrochemical water splitting. J. Alloys Compd. 2021, 868, 159133. [Google Scholar] [CrossRef]
  2. Osterloh, F.E.; Parkinson, B.A. Recent developments in solar water-splitting photocatalysis. MRS Bull. 2011, 36, 17–22. [Google Scholar] [CrossRef]
  3. Dawood, F.; Anda, M.; Shafiullah, G. Hydrogen production for energy: An overview. Int. J. Hydrogen Energy 2020, 45, 3847–3869. [Google Scholar] [CrossRef]
  4. Ishaq, T.; Yousaf, M.; Bhatti, I.A.; Batool, A.; Asghar, M.A.; Mohsin, M.; Ahmad, M. A perspective on possible amendments in semiconductors for enhanced photocatalytic hydrogen generation by water splitting. Int. J. Hydrogen Energy 2021, 46, 39036–39057. [Google Scholar] [CrossRef]
  5. Muscetta, M.; Andreozzi, R.; Clarizia, L.; Di Somma, I.; Marotta, R. Hydrogen production through photoreforming processes over Cu2O/TiO2 composite materials: A mini-review. Int. J. Hydrogen Energy 2020, 45, 28531–28552. [Google Scholar] [CrossRef]
  6. Hou, K.; Hughes, R. The kinetics of methane steam reforming over a Ni/α-Al2O catalyst. Chem. Eng. J. 2001, 82, 311–328. [Google Scholar] [CrossRef]
  7. Nowotny, J.; Sorrell, C.; Sheppard, L.; Bak, T. Solar-hydrogen: Environmentally safe fuel for the future. Int. J. Hydrogen Energy 2005, 30, 521–544. [Google Scholar] [CrossRef]
  8. Salam, M.A.; Ahmed, K.; Akter, N.; Hossain, T.; Abdullah, B. A review of hydrogen production via biomass gasification and its prospect in Bangladesh. Int. J. Hydrogen Energy 2018, 43, 14944–14973. [Google Scholar] [CrossRef]
  9. Shah, S.A.A. Feasibility study of renewable energy sources for developing the hydrogen economy in Pakistan. Int. J. Hydrogen Energy 2020, 45, 15841–15854. [Google Scholar] [CrossRef]
  10. Das, D.; Veziroglu, T.N. Advances in biological hydrogen production processes. Int. J. Hydrogen Energy 2008, 33, 6046–6057. [Google Scholar] [CrossRef]
  11. He, Y.-L.; Qiu, Y.; Wang, K.; Yuan, F.; Wang, W.-Q.; Li, M.-J.; Guo, J.-Q. Perspective of concentrating solar power. Energy 2020, 198, 117373. [Google Scholar] [CrossRef]
  12. Mallikarjuna, K.; Bari, G.A.R.; Vattikuti, S.P.; Kim, H. Synthesis of carbon-doped SnO2 nanostructures for visible-light-driven photocatalytic hydrogen production from water splitting. Int. J. Hydrogen Energy 2020, 45, 32789–32796. [Google Scholar] [CrossRef]
  13. Hassan, Q. Optimisation of solar-hydrogen power system for household applications. Int. J. Hydrogen Energy 2020, 45, 33111–33127. [Google Scholar] [CrossRef]
  14. Tahir, M.; Asiri, A.M.; Nawaz, T. A perspective on the fabrication of heterogeneous photocatalysts for enhanced hydrogen production. Int. J. Hydrogen Energy 2020, 45, 24544–24557. [Google Scholar] [CrossRef]
  15. Rusinque, B. Hydrogen Production by Photocatalytic Water Splitting under Near-UV and Visible Light Using Doped Pt and Pd TiO2. Master’s Thesis, The University of Western Ontario, London, ON, Canada, 2018. [Google Scholar]
  16. Rafique, M.; Mubashar, R.; Irshad, M.; Gillani, S.; Tahir, M.B.; Khalid, N.; Yasmin, A.; Shehzad, M.A. A comprehensive study on methods and materials for photocatalytic water splitting and hydrogen production as a renewable energy resource. J. Inorg. Organomet. Polym. Mater. 2020, 30, 3837–3861. [Google Scholar] [CrossRef]
  17. Cao, S.; Piao, L. Considerations for a more accurate evaluation method for photocatalytic water splitting. Angew. Chem. Int. Ed. 2020, 59, 18312–18320. [Google Scholar] [CrossRef]
  18. Saber, A.F.; Elewa, A.M.; Chou, H.-H.; EL-Mahdy, A.F. Donor-acceptor carbazole-based conjugated microporous polymers as photocatalysts for visible-light-driven H2 and O2 evolution from water splitting. Appl. Catal. B 2022, 316, 121624. [Google Scholar] [CrossRef]
  19. Guo, L.-J.; Luo, J.-W.; He, T.; Wei, S.-H.; Li, S.-S. Photocorrosion-limited maximum efficiency of solar photoelectrochemical water splitting. Phys. Rev. Appl. 2018, 10, 064059. [Google Scholar] [CrossRef]
  20. Guo, L.; Chen, Y.; Su, J.; Liu, M.; Liu, Y. Obstacles of solar-powered photocatalytic water splitting for hydrogen production: A perspective from energy flow and mass flow. Energy 2019, 172, 1079–1086. [Google Scholar] [CrossRef]
  21. Ning, X.; Lu, G. Photocorrosion inhibition of CdS-based catalysts for photocatalytic overall water splitting. Nanoscale 2020, 12, 1213–1223. [Google Scholar] [CrossRef]
  22. Singla, S.; Sharma, S.; Basu, S.; Shetti, N.P.; Aminabhavi, T.M. Photocatalytic water splitting hydrogen production via environmental benign carbon based nanomaterials. Int. J. Hydrogen Energy 2021, 46, 33696–33717. [Google Scholar] [CrossRef]
  23. Ni, M.; Leung, M.K.; Leung, D.Y.; 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]
  24. Chowdhury, F.A.; Trudeau, M.L.; Guo, H.; Mi, Z. A photochemical diode artificial photosynthesis system for unassisted high efficiency overall pure water splitting. Nat. Commun. 2018, 9, 1707. [Google Scholar] [CrossRef] [PubMed]
  25. Moridon, S.N.F.; Salehmin, M.I.; Mohamed, M.A.; Arifin, K.; Minggu, L.J.; Kassim, M.B. Cobalt oxide as photocatalyst for water splitting: Temperature-dependent phase structures. Int. J. Hydrogen Energy 2019, 44, 25495–25504. [Google Scholar] [CrossRef]
  26. Li, S.; Xu, W.; Meng, L.; Tian, W.; Li, L. Recent progress on semiconductor heterojunction-based photoanodes for photoelectrochemical water splitting. Small Sci. 2022, 2, 2100112. [Google Scholar] [CrossRef]
  27. Badea, G.E.; Hora, C.; Maior, I.; Cojocaru, A.; Secui, C.; Filip, S.M.; Dan, F.C. Sustainable hydrogen production from seawater electrolysis: Through fundamental electrochemical principles to the most recent development. Energies 2022, 15, 8560. [Google Scholar] [CrossRef]
  28. Ishaq, T.; Yousaf, M.; Bhatti, I.A.; Ahmad, M.; Ikram, M.; Khan, M.U.; Qayyum, A. Photo-assisted splitting of water into hydrogen using visible-light activated silver doped g-C3N4 & CNTs hybrids. Int. J. Hydrogen Energy 2020, 45, 31574–31584. [Google Scholar]
  29. Bahramian, A.; Rezaeivala, M.; He, K.; Dionysiou, D.D. Enhanced visible-light photoelectrochemical hydrogen evolution through degradation of methyl orange in a cell based on coral-like Pt-deposited TiO2 thin film with sub-2 nm pores. Catal. Today 2019, 335, 333–344. [Google Scholar] [CrossRef]
  30. Iqbal, M.Z.; Siddique, S. Recent progress in efficiency of hydrogen evolution process based photoelectrochemical cell. Int. J. Hydrogen Energy 2018, 43, 21502–21523. [Google Scholar] [CrossRef]
  31. Qiu, Y.; Pan, Z.; Chen, H.; Ye, D.; Guo, L.; Fan, Z.; Yang, S. Current progress in developing metal oxide nanoarrays-based photoanodes for photoelectrochemical water splitting. Sci. Bull. 2019, 64, 1348–1380. [Google Scholar] [CrossRef]
  32. Murphy, A.B.; Barnes, P.R.; Randeniya, L.K.; Plumb, I.C.; Grey, I.E.; Horne, M.D.; Glasscock, J.A. Efficiency of solar water splitting using semiconductor electrodes. Int. J. Hydrogen Energy 2006, 31, 1999–2017. [Google Scholar] [CrossRef]
  33. Xu, Y.; Schoonen, M.A. The absolute energy positions of conduction and valence bands of selected semiconducting minerals. Am. Mineral. 2000, 85, 543–556. [Google Scholar] [CrossRef]
  34. Jiang, C.; Moniz, S.J.; Wang, A.; Zhang, T.; Tang, J. Photoelectrochemical devices for solar water splitting–materials and challenges. Chem. Soc. Rev. 2017, 46, 4645–4660. [Google Scholar] [CrossRef] [PubMed]
  35. Hu, C.-C.; Nian, J.-N.; Teng, H. Electrodeposited p-type Cu2O as photocatalyst for H2 evolution from water reduction in the presence of WO3. Sol. Energy Mater. Sol. Cells 2008, 92, 1071–1076. [Google Scholar] [CrossRef]
  36. McShane, C.M.; Choi, K.-S. Junction studies on electrochemically fabricated p–n Cu2O homojunction solar cells for efficiency enhancement. Phys. Chem. Chem. Phys. 2012, 14, 6112–6118. [Google Scholar] [CrossRef]
  37. Paracchino, A.; Laporte, V.; Sivula, K.; Grätzel, M.; Thimsen, E. Highly active oxide photocathode for photoelectrochemical water reduction. Nat. Mater. 2011, 10, 456–461. [Google Scholar] [CrossRef]
  38. Dominey, R.N.; Lewis, N.S.; Bruce, J.A.; Bookbinder, D.C.; Wrighton, M.S. Improvement of photoelectrochemical hydrogen generation by surface modification of p-type silicon semiconductor photocathodes. J. Am. Chem. Soc. 1982, 104, 467–482. [Google Scholar] [CrossRef]
  39. Hisatomi, T.; Kubota, J.; Domen, K. Recent advances in semiconductors for photocatalytic and photoelectrochemical water splitting. Chem. Soc. Rev. 2014, 43, 7520–7535. [Google Scholar] [CrossRef]
  40. Zheng, Q.; Zhou, B.; Bai, J.; Li, L.; Jin, Z.; Zhang, J.; Li, J.; Liu, Y.; Cai, W.; Zhu, X. Self-organized TiO2 nanotube array sensor for the determination of chemical oxygen demand. Adv. Mater.—Deerfield Beach Then Weinh. 2008, 20, 1044. [Google Scholar] [CrossRef]
  41. Qin, G.; Watanabe, A. Surface texturing of TiO2 film by mist deposition of TiO2 nanoparticles. Nano-Micro Lett. 2013, 5, 129–134. [Google Scholar] [CrossRef]
  42. Su, Z.; Zhou, W.; Jiang, F.; Hong, M. Anodic formation of nanoporous and nanotubular metal oxides. J. Mater. Chem. 2012, 22, 535–544. [Google Scholar] [CrossRef]
  43. Varghese, O.K.; Paulose, M.; Grimes, C.A. Long vertically aligned titania nanotubes on transparent conducting oxide for highly efficient solar cells. Nat. Nanotechnol. 2009, 4, 592–597. [Google Scholar] [CrossRef] [PubMed]
  44. Feng, X.; Shankar, K.; Varghese, O.K.; Paulose, M.; Latempa, T.J.; Grimes, C.A. Vertically aligned single crystal TiO2 nanowire arrays grown directly on transparent conducting oxide coated glass: Synthesis details and applications. Nano Lett. 2008, 8, 3781–3786. [Google Scholar] [CrossRef] [PubMed]
  45. Trang, T.; Tu, L.; Man, T.; Mathesh, M.; Nam, N.; Thu, V. A high-efficiency photoelectrochemistry of Cu2O/TiO2 nanotubes based composite for hydrogen evolution under sunlight. Compos. Part B 2019, 174, 106969. [Google Scholar] [CrossRef]
  46. Kumar, S.G.; Devi, L.G. Review on modified TiO2 photocatalysis under UV/visible light: Selected results and related mechanisms on interfacial charge carrier transfer dynamics. J. Phys. Chem. A 2011, 115, 13211–13241. [Google Scholar] [CrossRef]
  47. Liang, R.; Hu, A.; Persic, J.; Zhou, Y.N. Palladium nanoparticles loaded on carbon modified TiO2 nanobelts for enhanced methanol electrooxidation. Nano-Micro Lett. 2013, 5, 202–212. [Google Scholar] [CrossRef]
  48. Park, J.H.; Kim, S.; Bard, A.J. Novel carbon-doped TiO2 nanotube arrays with high aspect ratios for efficient solar water splitting. Nano Lett. 2006, 6, 24–28. [Google Scholar] [CrossRef]
  49. Siripala, W.; Ivanovskaya, A.; Jaramillo, T.F.; Baeck, S.-H.; McFarland, E.W. A Cu2O/TiO2 heterojunction thin film cathode for photoelectrocatalysis. Sol. Energy Mater. Sol. Cells 2003, 77, 229–237. [Google Scholar] [CrossRef]
  50. Xu, Q.C.; Wellia, D.V.; Ng, Y.H.; Amal, R.; Tan, T.T.Y. Synthesis of porous and visible-light absorbing Bi2WO6/TiO2 heterojunction films with improved photoelectrochemical and photocatalytic performances. J. Phys. Chem. C 2011, 115, 7419–7428. [Google Scholar] [CrossRef]
  51. Zhang, C.; Greenblatt, J.B.; Wei, M.; Eichman, J.; Saxena, S.; Muratori, M.; Guerra, O.J. Flexible grid-based electrolysis hydrogen production for fuel cell vehicles reduces costs and greenhouse gas emissions. Appl. Energy 2020, 278, 115651. [Google Scholar] [CrossRef]
  52. Guo, Y.; Li, G.; Zhou, J.; Liu, Y. Comparison between hydrogen production by alkaline water electrolysis and hydrogen production by PEM electrolysis. IOP Conf. Ser. Earth Environ. Sci. 2019, 371, 042022. [Google Scholar] [CrossRef]
  53. Dincer, I.; Acar, C. Review and evaluation of hydrogen production methods for better sustainability. Int. J. Hydrogen Energy 2015, 40, 11094–11111. [Google Scholar] [CrossRef]
  54. Nikolaidis, P.; Poullikkas, A. A comparative overview of hydrogen production processes. Renew. Sustain. Energy Rev. 2017, 67, 597–611. [Google Scholar] [CrossRef]
  55. Sun, H.; Tian, C.; Fan, G.; Qi, J.; Liu, Z.; Yan, Z.; Cheng, F.; Chen, J.; Li, C.P.; Du, M. Boosting activity on Co4N porous nanosheet by coupling CeO2 for efficient electrochemical overall water splitting at high current densities. Adv. Funct. Mater. 2020, 30, 1910596. [Google Scholar] [CrossRef]
  56. Wang, Y.; Pang, Y.; Xu, H.; Martinez, A.; Chen, K.S. PEM Fuel cell and electrolysis cell technologies and hydrogen infrastructure development–a review. Energy Environ. Sci. 2022, 15, 2288–2328. [Google Scholar] [CrossRef]
  57. Campbell-Stanway, C.; Becerra, V.; Prabhu, S.; Bull, J. Investigating the role of byproduct oxygen in uk-based future scenario models for green hydrogen electrolysis. Energies 2024, 17, 281. [Google Scholar] [CrossRef]
  58. Hu, C.; Zhang, L.; Gong, J. Recent progress made in the mechanism comprehension and design of electrocatalysts for alkaline water splitting. Energy Environ. Sci. 2019, 12, 2620–2645. [Google Scholar] [CrossRef]
  59. Ardo, S.; Rivas, D.F.; Modestino, M.A.; Greiving, V.S.; Abdi, F.F.; Llado, E.A.; Artero, V.; Ayers, K.; Battaglia, C.; Becker, J.-P. Pathways to electrochemical solar-hydrogen technologies. Energy Environ. Sci. 2018, 11, 2768–2783. [Google Scholar] [CrossRef]
  60. Berger, A.; Segalman, R.; Newman, J. Material requirements for membrane separators in a water-splitting photoelectrochemical cell. Energy Environ. Sci. 2014, 7, 1468–1476. [Google Scholar] [CrossRef]
  61. Cheng, W.-H.; De La Calle, A.; Atwater, H.A.; Stechel, E.B.; Xiang, C. Hydrogen from sunlight and water: A side-by-side comparison between photoelectrochemical and solar thermochemical water-splitting. ACS Energy Lett. 2021, 6, 3096–3113. [Google Scholar] [CrossRef]
  62. Oener, S.Z.; Twight, L.P.; Lindquist, G.A.; Boettcher, S.W. Thin cation-exchange layers enable high-current-density bipolar membrane electrolyzers via improved water transport. ACS Energy Lett. 2020, 6, 1–8. [Google Scholar] [CrossRef]
  63. Luo, T.; Abdu, S.; Wessling, M. Selectivity of ion exchange membranes: A review. J. Membr. Sci. 2018, 555, 429–454. [Google Scholar] [CrossRef]
  64. Giesbrecht, P.K.; Freund, M.S. Recent advances in bipolar membrane design and applications. Chem. Mater. 2020, 32, 8060–8090. [Google Scholar] [CrossRef]
  65. Yan, Z.; Mallouk, T.E. Bipolar membranes for ion management in (photo) electrochemical energy conversion. Acc. Mater. Res. 2021, 2, 1156–1166. [Google Scholar] [CrossRef]
  66. Kim, J.S.; Kim, B.; Kim, H.; Kang, K. Recent progress on multimetal oxide catalysts for the oxygen evolution reaction. Adv. Energy Mater. 2018, 8, 1702774. [Google Scholar] [CrossRef]
  67. Zhao, H.; Lu, D.; Wang, J.; Tu, W.; Wu, D.; Koh, S.W.; Gao, P.; Xu, Z.J.; Deng, S.; Zhou, Y. Raw biomass electroreforming coupled to green hydrogen generation. Nat. Commun. 2021, 12, 2008. [Google Scholar] [CrossRef] [PubMed]
  68. Li, C.; Bao, Y.; Liu, E.; Zhao, B.; Sun, T. Recent advances of modified Ni (Co, Fe)-based LDH 2D materials for water splitting. Molecules 2023, 28, 1475. [Google Scholar] [CrossRef]
  69. Pandev, M.; Terziev, V.; Abrashev, B. The evolution of hydrogen technologies: Paving the way to a sustainable hydrogen economy. Bulg. Chem. Commun. 2023, 55, 434–440. [Google Scholar]
  70. Hua, D.; Huang, J.; Fabbri, E.; Rafique, M.; Song, B. Development of anion exchange membrane water electrolysis and the associated challenges: A review. ChemElectroChem 2023, 10, e202200999. [Google Scholar] [CrossRef]
  71. Trattner, A.; Höglinger, M.; Macherhammer, M.G.; Sartory, M. Renewable hydrogen: Modular concepts from production over storage to the consumer. Chem. Ing. Tech. 2021, 93, 706–716. [Google Scholar] [CrossRef]
  72. López-Fernández, E.; Sacedón, C.G.; Gil-Rostra, J.; Yubero, F.; González-Elipe, A.R.; de Lucas-Consuegra, A. Recent advances in alkaline exchange membrane water electrolysis and electrode manufacturing. Molecules 2021, 26, 6326. [Google Scholar] [CrossRef] [PubMed]
  73. Colli, A.N.; Girault, H.H.; Battistel, A. Non-precious electrodes for practical alkaline water electrolysis. Materials 2019, 12, 1336. [Google Scholar] [CrossRef] [PubMed]
  74. Ju, W.; Heinz, M.V.; Pusterla, L.; Hofer, M.; Fumey, B.; Castiglioni, R.; Pagani, M.; Battaglia, C.; Vogt, U.F. Lab-scale alkaline water electrolyzer for bridging material fundamentals with realistic operation. ACS Sustain. Chem. Eng. 2018, 6, 4829–4837. [Google Scholar] [CrossRef]
  75. Vincent, I.; Lee, E.-C.; Kim, H.-M. Comprehensive impedance investigation of low-cost anion exchange membrane electrolysis for large-scale hydrogen production. Sci. Rep. 2021, 11, 293. [Google Scholar] [CrossRef] [PubMed]
  76. Wu, Q.; Wang, Y.; Zhang, K.; Xie, Z.; Sun, K.; An, W.; Liang, X.; Zou, X. Advances and status of anode catalysts for proton exchange membrane water electrolysis technology. Mater. Chem. Front. 2023, 7, 1025–1045. [Google Scholar] [CrossRef]
  77. Ferriday, T.; Middleton, P.; Kolhe, M. Determining the change in performance from replacing a separator with an anion exchange membrane for alkaline water electrolysis. J. Phys. Conf. Ser. 2023, 2454, 012003. [Google Scholar] [CrossRef]
  78. Henkensmeier, D.; Najibah, M.; Harms, C.; Žitka, J.; Hnát, J.; Bouzek, K. Overview: State-of-the art commercial membranes for anion exchange membrane water electrolysis. J. Electrochem. Energy Convers. Storage 2021, 18, 024001. [Google Scholar] [CrossRef]
  79. IRENA. Making the Breakthrough: Green Hydrogen Policies and Technology Costs; International Renewable Energy Agency: Abu Dhabi, United Arab Emirates, 2021. [Google Scholar]
  80. IRENA. Scaling up Electrolysers to Meet the 1.5 C Climate Goal; International Renewable Energy Agency: Abu Dhabi, United Arab Emirates, 2020. [Google Scholar]
  81. Iwasa, N.; Teng, Z.; Ma, G.; Hisatomi, T.; Domen, K. Synthesis of Narrow Band Gap Gallium Zinc Nitride Oxide Solid Solutions for Photocatalytic Water Splitting under Visible Light. Chem. Mater. 2024, 36, 2917–2924. [Google Scholar] [CrossRef]
  82. Navarro Yerga, R.M.; Álvarez Galván, M.C.; Del Valle, F.; Villoria de la Mano, J.A.; Fierro, J.L. Water splitting on semiconductor catalysts under visible-light irradiation. ChemSusChem Chem. Sustain. Energy Mater. 2009, 2, 471–485. [Google Scholar] [CrossRef]
  83. Fajrina, N.; Tahir, M. A critical review in strategies to improve photocatalytic water splitting towards hydrogen production. Int. J. Hydrogen Energy 2019, 44, 540–577. [Google Scholar] [CrossRef]
  84. Lisovski, O.; Piskunov, S.; Zhukovskii, Y.F.; Ozolins, J. Ab initio modeling of sulphur doped TiO2 nanotubular photocatalyst for water-splitting hydrogen generation. IOP Conf. Ser. Mater. Sci. Eng. 2012, 38, 012057. [Google Scholar] [CrossRef]
  85. Idriss, H. The elusive photocatalytic water splitting reaction using sunlight on suspended nanoparticles: Is there a way forward? Catal. Sci. Technol. 2020, 10, 304–310. [Google Scholar] [CrossRef]
  86. Niu, X.; Bai, X.; Zhou, Z.; Wang, J. Rational design and characterization of direct Z-scheme photocatalyst for overall water splitting from excited state dynamics simulations. ACS Catal. 2020, 10, 1976–1983. [Google Scholar] [CrossRef]
  87. Gholipour, M.R.; Dinh, C.-T.; Béland, F.; Do, T.-O. Nanocomposite heterojunctions as sunlight-driven photocatalysts for hydrogen production from water splitting. Nanoscale 2015, 7, 8187–8208. [Google Scholar] [CrossRef] [PubMed]
  88. Bernareggi, M.; Dozzi, M.V.; Bettini, L.G.; Ferretti, A.M.; Chiarello, G.L.; Selli, E. Flame-made Cu/TiO2 and Cu-Pt/TiO2 photocatalysts for hydrogen production. Catalysts 2017, 7, 301. [Google Scholar] [CrossRef]
  89. Vijayaraghavan, T.; Reddy, N.L.; Shankar, M.; Vadivel, S.; Ashok, A. A co-catalyst free, eco-friendly, novel visible light absorbing iron based complex oxide nanocomposites for enhanced photocatalytic hydrogen evolution. Int. J. Hydrogen Energy 2018, 43, 14417–14426. [Google Scholar] [CrossRef]
  90. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  91. Wang, B.; Shen, S.; Mao, S.S. Black TiO2 for solar hydrogen conversion. J. Mater. 2017, 3, 96–111. [Google Scholar]
  92. Eidsvåg, H.; Bentouba, S.; Vajeeston, P.; Yohi, S.; Velauthapillai, D. TiO2 as a photocatalyst for water splitting—An experimental and theoretical review. Molecules 2021, 26, 1687. [Google Scholar] [CrossRef]
  93. Haggerty, J.E.; Schelhas, L.T.; Kitchaev, D.A.; Mangum, J.S.; Garten, L.M.; Sun, W.; Stone, K.H.; Perkins, J.D.; Toney, M.F.; Ceder, G. High-fraction brookite films from amorphous precursors. Sci. Rep. 2017, 7, 15232. [Google Scholar] [CrossRef]
  94. Kaur, K.; Singh, C.V. Amorphous TiO2 as a photocatalyst for hydrogen production: A DFT study of structural and electronic properties. Energy Procedia 2012, 29, 291–299. [Google Scholar] [CrossRef]
  95. Zhang, W.; Yin, J.-R.; Tang, X.-Q.; Zhang, P.; Ding, Y.-H. Density functional theory studies on the structural and physical properties of Cu-doped anatase TiO2 (101) surface. Phys. E Low-Dimens. Syst. Nanostruct. 2017, 85, 259–263. [Google Scholar] [CrossRef]
  96. Morgade, C.I.; Cabeza, G.F. First-principles study of codoping TiO2 systems capable of improving the specific surface area and the dissociation of H2O to generate H2 and O2. Comput. Mater. Sci. 2017, 127, 204–210. [Google Scholar] [CrossRef]
  97. Rafique, M.; Hajra, S.; Irshad, M.; Usman, M.; Imran, M.; Assiri, M.A.; Ashraf, W.M. Hydrogen production using TiO2-based photocatalysts: A comprehensive review. ACS Omega 2023, 8, 25640–25648. [Google Scholar] [CrossRef] [PubMed]
  98. Zhu, Z.; Kao, C.-T.; Tang, B.-H.; Chang, W.-C.; Wu, R.-J. Efficient hydrogen production by photocatalytic water-splitting using Pt-doped TiO2 hollow spheres under visible light. Ceram. Int. 2016, 42, 6749–6754. [Google Scholar] [CrossRef]
  99. Aldosari, O.F.; Hussain, I. Unlocking the potential of TiO2-based photocatalysts for green hydrogen energy through water-splitting: Recent advances, future perspectives and techno feasibility assessment. Int. J. Hydrogen Energy 2024, 59, 958–981. [Google Scholar] [CrossRef]
  100. Jing, D.; Guo, L.; Zhao, L.; Zhang, X.; Liu, H.; Li, M.; Shen, S.; Liu, G.; Hu, X.; Zhang, X. Efficient solar hydrogen production by photocatalytic water splitting: From fundamental study to pilot demonstration. Int. J. Hydrogen Energy 2010, 35, 7087–7097. [Google Scholar] [CrossRef]
  101. Abe, R.; Sayama, K.; Arakawa, H. Significant effect of iodide addition on water splitting into H2 and O2 over Pt-loaded TiO2 photocatalyst: Suppression of backward reaction. Chem. Phys. Lett. 2003, 371, 360–364. [Google Scholar] [CrossRef]
  102. Liao, C.-H.; Huang, C.-W.; Wu, J.C. Hydrogen production from semiconductor-based photocatalysis via water splitting. Catalysts 2012, 2, 490–516. [Google Scholar] [CrossRef]
  103. Hogan, B. Ensuring a Future for Australian Coal Fired Power Station; Institute of Public Affairs: Melbourne, Australia, 2016. [Google Scholar]
  104. Sayama, K.; Arakawa, H. Photocatalytic decomposition of water and photocatalytic reduction of carbon dioxide over zirconia catalyst. J. Phys. Chem. 1993, 97, 531–533. [Google Scholar] [CrossRef]
  105. Zhang, A.; Liang, Y.; Zhang, H.; Geng, Z.; Zeng, J. Doping regulation in transition metal compounds for electrocatalysis. Chem. Soc. Rev. 2021, 50, 9817–9844. [Google Scholar] [CrossRef] [PubMed]
  106. Liu, Y.; Wang, W.; Xu, X.; Veder, J.-P.M.; Shao, Z. Recent advances in anion-doped metal oxides for catalytic applications. J. Mater. Chem. A 2019, 7, 7280–7300. [Google Scholar] [CrossRef]
  107. Samsudin, E.M.; Abd Hamid, S.B. Effect of band gap engineering in anionic-doped TiO2 photocatalyst. Appl. Surf. Sci. 2017, 391, 326–336. [Google Scholar] [CrossRef]
  108. Shon, H.; Phuntsho, S.; Okour, Y.; Cho, D.-L.; Kim, K.S.; Li, H.-J.; Na, S.; Kim, J.B.; Kim, J.-H. Visible light responsive titanium dioxide (TiO2). J. Korean Ind. Eng. Chem. 2008, 19, 1–16. [Google Scholar]
  109. Choi, S.K.; Yang, H.S.; Kim, J.H.; Park, H. Organic dye-sensitized TiO2 as a versatile photocatalyst for solar hydrogen and environmental remediation. Appl. Catal. B 2012, 121, 206–213. [Google Scholar] [CrossRef]
  110. Choi, W.; Termin, A.; Hoffmann, M.R. The role of metal ion dopants in quantum-sized TiO2: Correlation between photoreactivity and charge carrier recombination dynamics. J. Phys. Chem. 2002, 98, 13669–13679. [Google Scholar] [CrossRef]
  111. Zhao, W.; Wang, X.; Sang, H.; Wang, K. Synthesis of Bi-doped TiO2 nanotubes and enhanced photocatalytic activity for hydrogen evolution from glycerol solution. Chin. J. Chem. 2013, 31, 415–420. [Google Scholar] [CrossRef]
  112. Mohammadparast, F.; Tirumala, R.T.A.; Ramakrishnan, S.B.; Dadgar, A.P.; Andiappan, M. Operando UV–Vis spectroscopy as potential in-line PAT system for size determination of functioning metal nanocatalysts. Chem. Eng. Sci. 2020, 225, 115821. [Google Scholar] [CrossRef]
  113. Ahmad, H.; Kamarudin, S.; Minggu, L.J.; Kassim, M. Hydrogen from photo-catalytic water splitting process: A review. Renew. Sustain. Energy Rev. 2015, 43, 599–610. [Google Scholar] [CrossRef]
  114. Jaafarzadeh, N.; Ghanbari, F.; Ahmadi, M. Catalytic degradation of 2,4-dichlorophenoxyacetic acid (2,4-D) by nano-Fe2O3 activated peroxymonosulfate: Influential factors and mechanism determination. Chemosphere 2017, 169, 568–576. [Google Scholar] [CrossRef]
  115. Xu, C.; Zhang, Y.; Chen, J.; Lin, J.; Zhang, X.; Wang, Z.; Zhou, J. Enhanced mechanism of the photo-thermochemical cycle based on effective Fe-doping TiO2 films and DFT calculations. Appl. Catal. B 2017, 204, 324–334. [Google Scholar] [CrossRef]
  116. Sakar, M.; Mithun Prakash, R.; Do, T.-O. Insights into the TiO2-based photocatalytic systems and their mechanisms. Catalysts 2019, 9, 680. [Google Scholar] [CrossRef]
  117. Chibac-Scutaru, A.L.; Podasca, V.-E.; Melinte, V. Symbiotic coupling of ZnO nanoparticles and coumarin photosensitizer in soft polyurethane matrices for boosting visible-light photocatalytic performance. J. Environ. Chem. Eng. 2024, 12, 112312. [Google Scholar] [CrossRef]
  118. Lv, S.; Du, Y.; Wu, F.; Cai, Y.; Zhou, T. Review on LSPR assisted photocatalysis: Effects of physical fields and opportunities in multifield decoupling. Nanoscale Adv. 2022, 4, 2608–2631. [Google Scholar] [CrossRef] [PubMed]
  119. Zada, A.; Muhammad, P.; Ahmad, W.; Hussain, Z.; Ali, S.; Khan, M.; Khan, Q.; Maqbool, M. Surface plasmonic-assisted photocatalysis and optoelectronic devices with noble metal nanocrystals: Design, synthesis, and applications. Adv. Funct. Mater. 2020, 30, 1906744. [Google Scholar] [CrossRef]
  120. Sakar, M.; Annamalai, L. Plasmon-sensitized semiconductors for photocatalysis. In Handbook of Smart Photocatalytic Materials; Elsevier: Amsterdam, The Netherlands, 2020; pp. 175–205. [Google Scholar]
  121. Wu, F.; Hu, X.; Fan, J.; Liu, E.; Sun, T.; Kang, L.; Hou, W.; Zhu, C.; Liu, H. Photocatalytic activity of Ag/TiO2 nanotube arrays enhanced by surface plasmon resonance and application in hydrogen evolution by water splitting. Plasmonics 2013, 8, 501–508. [Google Scholar] [CrossRef]
  122. Guo, R.; Zeng, D.; Xie, Y.; Ling, Y.; Zhou, D.; Jiang, L.; Jiao, W.; Zhao, J.; Li, S. Carbon nitride quantum dots (CNQDs)/TiO2 nanoparticle heterojunction photocatalysts for enhanced ultraviolet-visible-light-driven bisphenol a degradation and H2 production. Int. J. Hydrogen Energy 2020, 45, 22534–22544. [Google Scholar] [CrossRef]
  123. Li, X.; Li, Y.; Wang, H.; Miao, H.; Zhu, H.; Liu, X.; Lin, H.; Shi, G. Fabrication of a three-dimensional bionic Si/TiO2/MoS2 photoelectrode for efficient solar water splitting. ACS Appl. Energy Mater. 2020, 4, 730–736. [Google Scholar] [CrossRef]
  124. Yuan, Y.; Guo, R.-t.; Hong, L.-f.; Ji, X.-y.; Lin, Z.-d.; Li, Z.-s.; Pan, W.-g. A review of metal oxide-based Z-scheme heterojunction photocatalysts: Actualities and developments. Mater. Today Energy 2021, 21, 100829. [Google Scholar] [CrossRef]
  125. Das, A.; Kumar, P.M.; Bhagavathiachari, M.; Nair, R.G. Hierarchical ZnO-TiO2 nanoheterojunction: A strategy driven approach to boost the photocatalytic performance through the synergy of improved surface area and interfacial charge transport. Appl. Surf. Sci. 2020, 534, 147321. [Google Scholar] [CrossRef]
  126. Zhang, L.; Jaroniec, M. Toward designing semiconductor-semiconductor heterojunctions for photocatalytic applications. Appl. Surf. Sci. 2018, 430, 2–17. [Google Scholar] [CrossRef]
  127. Zhang, Z.; Yates Jr, J.T. Band bending in semiconductors: Chemical and physical consequences at surfaces and interfaces. Chem. Rev. 2012, 112, 5520–5551. [Google Scholar] [CrossRef] [PubMed]
  128. Xu, M.; Ye, T.; Dai, F.; Yang, J.; Shen, J.; He, Q.; Chen, W.; Liang, N.; Zai, J.; Qian, X. Rationally designed n–n heterojunction with highly efficient solar hydrogen evolution. ChemSusChem 2015, 8, 1218–1225. [Google Scholar] [CrossRef]
  129. Wang, X.; Liu, G.; Chen, Z.-G.; Li, F.; Wang, L.; Lu, G.Q.; Cheng, H.-M. Enhanced photocatalytic hydrogen evolution by prolonging the lifetime of carriers in ZnO/CdS heterostructures. Chem. Commun. 2009, 3452–3454. [Google Scholar] [CrossRef] [PubMed]
  130. Low, J.; Jiang, C.; Cheng, B.; Wageh, S.; Al-Ghamdi, A.A.; Yu, J. A review of direct Z-scheme photocatalysts. Small Methods 2017, 1, 1700080. [Google Scholar] [CrossRef]
  131. Chen, L.; Zhou, X.; Jin, B.; Luo, J.; Xu, X.; Zhang, L.; Hong, Y. Heterojunctions in g-C3N4/B-TiO2 nanosheets with exposed {001} plane and enhanced visible-light photocatalytic activities. Int. J. Hydrogen Energy 2016, 41, 7292–7300. [Google Scholar] [CrossRef]
  132. He, H.; Lin, J.; Fu, W.; Wang, X.; Wang, H.; Zeng, Q.; Gu, Q.; Li, Y.; Yan, C.; Tay, B.K. MoS2/TiO2 edge-on heterostructure for efficient photocatalytic hydrogen evolution. Adv. Energy Mater. 2016, 6, 1600464. [Google Scholar] [CrossRef]
  133. Sun, B.; Zhou, W.; Li, H.; Ren, L.; Qiao, P.; Li, W.; Fu, H. Synthesis of particulate hierarchical tandem heterojunctions toward optimized photocatalytic hydrogen production. Adv. Mater. 2018, 30, 1804282. [Google Scholar] [CrossRef]
  134. Wang, Q.; Xiao, L.; Liu, X.; Sun, X.; Wang, J.; Du, H. Special Z-scheme Cu3P/TiO2 hetero-junction for efficient photocatalytic hydrogen evolution from water. J. Alloys Compd. 2022, 894, 162331. [Google Scholar] [CrossRef]
  135. Zheng, Z.; Zu, X.; Zhang, Y.; Zhou, W. Rational design of type-II nano-heterojunctions for nanoscale optoelectronics. Mater. Today Phys. 2020, 15, 100262. [Google Scholar] [CrossRef]
  136. Li, R.; Li, T.; Zhou, Q. Impact of titanium dioxide (TiO2) modification on its application to pollution treatment-a review. Catalysts 2020, 10, 804. [Google Scholar] [CrossRef]
  137. He, X.; Wang, A.; Wu, P.; Tang, S.; Zhang, Y.; Li, L.; Ding, P. Photocatalytic degradation of microcystin-LR by modified TiO2 photocatalysis: A review. Sci. Total Environ. 2020, 743, 140694. [Google Scholar] [CrossRef] [PubMed]
  138. Belabed, C.; Tab, A.; Bellal, B.; Belhamdi, B.; Benrakaa, N.; Trari, M. High photocatalytic performance for hydrogen production under visible light on the hetero-junction Pani-ZnO nanoparticles. Int. J. Hydrogen Energy 2021, 46, 17106–17115. [Google Scholar] [CrossRef]
  139. Zhang, W.; He, H.; Li, H.; Duan, L.; Zu, L.; Zhai, Y.; Li, W.; Wang, L.; Fu, H.; Zhao, D. Visible-light responsive TiO2-based materials for efficient solar energy utilization. Adv. Energy Mater. 2021, 11, 2003303. [Google Scholar] [CrossRef]
  140. Lu, H.; Hao, Q.; Chen, T.; Zhang, L.; Chen, D.; Ma, C.; Yao, W.; Zhu, Y. A high-performance Bi2O3/Bi2SiO5 pn heterojunction photocatalyst induced by phase transition of Bi2O3. Appl. Catal. B 2018, 237, 59–67. [Google Scholar] [CrossRef]
  141. Sakthivel, T.; Venugopal, G.; Durairaj, A.; Vasanthkumar, S.; Huang, X. Utilization of the internal electric field in semiconductor photocatalysis: A short review. J. Ind. Eng. Chem. 2019, 72, 18–30. [Google Scholar] [CrossRef]
  142. Yao, S.; Xue, S.; Shen, X. Photocatalytic activity of cuboid WO3 rods loaded with AgCl nanoparticles under visible light irradiation. J. Nanosci. Nanotechnol. 2017, 17, 5423–5431. [Google Scholar] [CrossRef]
  143. Li, X.; Garlisi, C.; Guan, Q.; Anwer, S.; Al-Ali, K.; Palmisano, G.; Zheng, L. A review of material aspects in developing direct Z-scheme photocatalysts. Mater. Today 2021, 47, 75–107. [Google Scholar] [CrossRef]
  144. Yang, H. A short review on heterojunction photocatalysts: Carrier transfer behavior and photocatalytic mechanisms. Mater. Res. Bull. 2021, 142, 111406. [Google Scholar] [CrossRef]
  145. Bai, S.; Jiang, J.; Zhang, Q.; Xiong, Y. Steering charge kinetics in photocatalysis: Intersection of materials syntheses, characterization techniques and theoretical simulations. Chem. Soc. Rev. 2015, 44, 2893–2939. [Google Scholar] [CrossRef]
  146. Wang, H.; Zhang, L.; Chen, Z.; Hu, J.; Li, S.; Wang, Z.; Liu, J.; Wang, X. Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 5234–5244. [Google Scholar] [CrossRef] [PubMed]
  147. Cao, J.; Li, X.; Lin, H.; Chen, S.; Fu, X. In situ preparation of novel p–n junction photocatalyst BiOI/(BiO)2CO3 with enhanced visible light photocatalytic activity. J. Hazard. Mater. 2012, 239, 316–324. [Google Scholar] [CrossRef] [PubMed]
  148. Chen, C.; Cai, W.; Long, M.; Zhou, B.; Wu, Y.; Wu, D.; Feng, Y. Synthesis of visible-light responsive graphene oxide/TiO2 composites with p/n heterojunction. ACS Nano 2010, 4, 6425–6432. [Google Scholar] [CrossRef] [PubMed]
  149. Zhang, Z.; Shao, C.; Li, X.; Wang, C.; Zhang, M.; Liu, Y. Electrospun nanofibers of p-type NiO/n-type ZnO heterojunctions with enhanced photocatalytic activity. ACS Appl. Mater. Interfaces 2010, 2, 2915–2923. [Google Scholar] [CrossRef] [PubMed]
  150. Hasija, V.; Raizada, P.; Hosseini-Bandegharaei, A.; Singh, P.; Nguyen, V.-H. Synthesis and photocatalytic activity of Ni–Fe layered double hydroxide modified sulphur doped graphitic carbon nitride (SGCN/Ni–Fe LDH) photocatalyst for 2, 4-dinitrophenol degradation. Top. Catal. 2020, 63, 1030–1045. [Google Scholar] [CrossRef]
  151. Marcolongo, D.M.; Nocito, F.; Ditaranto, N.; Aresta, M.; Dibenedetto, A. Synthesis and characterization of pn junction ternary mixed oxides for photocatalytic coprocessing of CO2 and H2O. Catalysts 2020, 10, 980. [Google Scholar] [CrossRef]
  152. Mohammed, A.M.; Mohtar, S.S.; Aziz, F.; Aziz, M.; Ul-Hamid, A.; Salleh, W.N.W.; Yusof, N.; Jaafar, J.; Ismail, A.F. Ultrafast degradation of Congo Red dye using a facile one-pot solvothermal synthesis of cuprous oxide/titanium dioxide and cuprous oxide/zinc oxide pn heterojunction photocatalyst. Mater. Sci. Semicond. Process. 2021, 122, 105481. [Google Scholar] [CrossRef]
  153. Liu, W.; Chen, S.F. Visible-light activity evaluation of pn junction photocatalyst NiO/TiO2 prepared by sol-gel method. Adv. Mater. Res. 2011, 152, 441–449. [Google Scholar]
  154. Kumar, R.; El-Shishtawy, R.M.; Barakat, M.A. Synthesis and characterization of Ag-Ag2O/TiO2@ polypyrrole heterojunction for enhanced photocatalytic degradation of methylene blue. Catalysts 2016, 6, 76. [Google Scholar] [CrossRef]
  155. Prabhu, Y.T.; Rao, V.N.; Shankar, M.V.; Sreedhar, B.; Pal, U. The facile hydrothermal synthesis of CuO@ ZnO heterojunction nanostructures for enhanced photocatalytic hydrogen evolution. New J. Chem. 2019, 43, 6794–6805. [Google Scholar] [CrossRef]
  156. Zhang, X.; Wang, Y.; Liu, B.; Sang, Y.; Liu, H. Heterostructures construction on TiO2 nanobelts: A powerful tool for building high-performance photocatalysts. Appl. Catal. B 2017, 202, 620–641. [Google Scholar] [CrossRef]
  157. Chu, L.; Li, M.; Cui, P.; Jiang, Y.; Wan, Z.; Dou, S. The study of NiO/TiO2 photocatalytic activity for degradation of methylene orange. Energy Environ. Focus 2014, 3, 371–374. [Google Scholar] [CrossRef]
  158. Lin, J.; Shen, J.; Wang, R.; Cui, J.; Zhou, W.; Hu, P.; Liu, D.; Liu, H.; Wang, J.; Boughton, R.I. Nano-p–n junctions on surface-coarsened TiO2 nanobelts with enhanced photocatalytic activity. J. Mater. Chem. 2011, 21, 5106–5113. [Google Scholar] [CrossRef]
  159. Zhao, L.; Cui, T.; Li, Y.; Wang, B.; Han, J.; Han, L.; Liu, Z. Efficient visible light photocatalytic activity of p–n junction CuO/TiO2 loaded on natural zeolite. RSC Adv. 2015, 5, 64495–64502. [Google Scholar] [CrossRef]
  160. Wang, M.; Hu, Y.; Han, J.; Guo, R.; Xiong, H.; Yin, Y. TiO2/NiO hybrid shells: P–n junction photocatalysts with enhanced activity under visible light. J. Mater. Chem. A 2015, 3, 20727–20735. [Google Scholar] [CrossRef]
  161. Nair, S.B.; Joseph, J.A.; Babu, S.; Shinoj, V.; Remillard, S.K.; Shaji, S.; Philip, R.R. Influence of pn junction mechanism and alumina overlayer on the photocatalytic performance of TiO2 nanotubes. Nanotechnology 2020, 31, 275403. [Google Scholar] [CrossRef]
  162. Wang, S.; Huang, C.-Y.; Pan, L.; Chen, Y.; Zhang, X.; Zou, J.-J. Controllable fabrication of homogeneous ZnO pn junction with enhanced charge separation for efficient photocatalysis. Catal. Today 2019, 335, 151–159. [Google Scholar] [CrossRef]
  163. Yu, J.; Wang, S.; Low, J.; Xiao, W. Enhanced photocatalytic performance of direct Z-scheme gC3N4–TiO2 photocatalysts for the decomposition of formaldehyde in air. Phys. Chem. Chem. Phys. 2013, 15, 16883–16890. [Google Scholar] [CrossRef]
  164. Liu, D.; Chen, S.; Li, R.; Peng, T. Review of Z-scheme heterojunctions for photocatalytic energy conversion. Acta Phys.-Chim. Sin 2021, 37, 2010017. [Google Scholar] [CrossRef]
  165. Li, H.; Tu, W.; Zhou, Y.; Zou, Z. Z-Scheme photocatalytic systems for promoting photocatalytic performance: Recent progress and future challenges. Adv. Sci. 2016, 3, 1500389. [Google Scholar] [CrossRef]
  166. Iwase, A.; Kudo, A. Development of Ir and La-codoped BaTa2O6 photocatalysts using visible light up to 640 nm as an H2-evolving photocatalyst for Z-schematic water splitting. Chem. Commun. 2017, 53, 6156–6159. [Google Scholar] [CrossRef] [PubMed]
  167. Miseki, Y.; Fujiyoshi, S.; Gunji, T.; Sayama, K. Photocatalytic Z-scheme water splitting for independent H2/O2 production via a stepwise operation employing a vanadate redox mediator under visible light. J. Phys. Chem. C 2017, 121, 9691–9697. [Google Scholar] [CrossRef]
  168. Abe, R.; Sayama, K.; Domen, K.; Arakawa, H. A new type of water splitting system composed of two different TiO2 photocatalysts (anatase, rutile) and a IO3−/I− shuttle redox mediator. Chem. Phys. Lett. 2001, 344, 339–344. [Google Scholar] [CrossRef]
  169. Ng, B.J.; Putri, L.K.; Kong, X.Y.; Teh, Y.W.; Pasbakhsh, P.; Chai, S.P. Z-scheme photocatalytic systems for solar water splitting. Adv. Sci. 2020, 7, 1903171. [Google Scholar] [CrossRef] [PubMed]
  170. Abe, R.; Shinmei, K.; Koumura, N.; Hara, K.; Ohtani, B. Visible-light-induced water splitting based on two-step photoexcitation between dye-sensitized layered niobate and tungsten oxide photocatalysts in the presence of a triiodide/iodide shuttle redox mediator. J. Am. Chem. Soc. 2013, 135, 16872–16884. [Google Scholar] [CrossRef]
  171. Wang, Y.; Suzuki, H.; Xie, J.; Tomita, O.; Martin, D.J.; Higashi, M.; Kong, D.; Abe, R.; Tang, J. Mimicking natural photosynthesis: Solar to renewable H2 fuel synthesis by Z-scheme water splitting systems. Chem. Rev. 2018, 118, 5201–5241. [Google Scholar] [CrossRef]
  172. Abe, R. Development of a new system for photocatalytic water splitting into H2 and O2 under visible light irradiation. Bull. Chem. Soc. Jpn. 2011, 84, 1000–1030. [Google Scholar] [CrossRef]
  173. Deshpande, A.; Shah, P.; Gholap, R.; Gupta, N.M. Interfacial and physico-chemical properties of polymer-supported CdS ZnS nanocomposites and their role in the visible-light mediated photocatalytic splitting of water. J. Colloid Interface Sci. 2009, 333, 263–268. [Google Scholar] [CrossRef]
  174. Maeda, K. Z-scheme water splitting using two different semiconductor photocatalysts. ACS Catal. 2013, 3, 1486–1503. [Google Scholar] [CrossRef]
  175. Wen Teh, Y.; Wei Goh, Y.; Ying Kong, X.; Ng, B.J.; Yong, S.T.; Chai, S.P. Fabrication of Bi2WO6/Cu/WO3 all-Solid-State Z-scheme composite photocatalyst to improve CO2 photoreduction under visible light irradiation. ChemCatChem 2019, 11, 6431–6438. [Google Scholar] [CrossRef]
  176. Maeda, K.; Domen, K. Photocatalytic water splitting: Recent progress and future challenges. J. Phys. Chem. Lett. 2010, 1, 2655–2661. [Google Scholar] [CrossRef]
  177. Lincho, J.; Mazierski, P.; Klimczuk, T.; Martins, R.C.; Gomes, J.; Zaleska-Medynska, A. TiO2 nanotubes modification by photodeposition with noble metals: Characterization, optimization, photocatalytic activity, and by-products analysis. J. Environ. Chem. Eng. 2024, 12, 112990. [Google Scholar] [CrossRef]
  178. Yao, G.-Y.; Liu, Q.-L.; Zhao, Z.-Y. Studied localized surface plasmon resonance effects of Au nanoparticles on TiO2 by FDTD simulations. Catalysts 2018, 8, 236. [Google Scholar] [CrossRef]
  179. Shehzad, N.; Tahir, M.; Johari, K.; Murugesan, T.; Hussain, M. A critical review on TiO2 based photocatalytic CO2 reduction system: Strategies to improve efficiency. J. CO2 Util. 2018, 26, 98–122. [Google Scholar] [CrossRef]
  180. Khatun, F.; Abd Aziz, A.; Sim, L.C.; Monir, M.U. Plasmonic enhanced Au decorated TiO2 nanotube arrays as a visible light active catalyst towards photocatalytic CO2 conversion to CH4. J. Environ. Chem. Eng. 2019, 7, 103233. [Google Scholar] [CrossRef]
  181. Xie, S.; Zhang, H.; Liu, G.; Wu, X.; Lin, J.; Zhang, Q.; Wang, Y. Tunable localized surface plasmon resonances in MoO3− x-TiO2 nanocomposites with enhanced catalytic activity for CO2 photoreduction under visible light. Chin. J. Catal. 2020, 41, 1125–1131. [Google Scholar] [CrossRef]
  182. Chen, Q.; Zhang, M.; Li, J.; Zhang, G.; Xin, Y.; Chai, C. Construction of immobilized 0D/1D heterostructure photocatalyst Au/CuS/CdS/TiO2 NBs with enhanced photocatalytic activity towards moxifloxacin degradation. Chem. Eng. J. 2020, 389, 124476. [Google Scholar] [CrossRef]
  183. Li, G.; Huang, J.; Chen, J.; Deng, Z.; Huang, Q.; Liu, Z.; Guo, W.; Cao, R. Highly active photocatalyst of Cu2O/TiO2 octahedron for hydrogen generation. ACS Omega 2019, 4, 3392–3397. [Google Scholar] [CrossRef]
  184. Liu, X.; Iocozzia, J.; Wang, Y.; Cui, X.; Chen, Y.; Zhao, S.; Li, Z.; Lin, Z. Noble metal–metal oxide nanohybrids with tailored nanostructures for efficient solar energy conversion, photocatalysis and environmental remediation. Energy Environ. Sci. 2017, 10, 402–434. [Google Scholar] [CrossRef]
  185. Miyoshi, A.; Nishioka, S.; Maeda, K. Water splitting on rutile TiO2-based photocatalysts. Chem.–A Eur. J. 2018, 24, 18204–18219. [Google Scholar] [CrossRef]
  186. Maeda, K. Direct splitting of pure water into hydrogen and oxygen using rutile titania powder as a photocatalyst. Chem. Commun. 2013, 49, 8404–8406. [Google Scholar] [CrossRef] [PubMed]
  187. Sayama, K.; Arakawa, H. Effect of carbonate salt addition on the photocatalytic decomposition of liquid water over Pt–TiO2 catalyst. J. Chem. Soc. Faraday Trans. 1997, 93, 1647–1654. [Google Scholar] [CrossRef]
  188. Moniz, S.J.; Shevlin, S.A.; Martin, D.J.; Guo, Z.-X.; Tang, J. Visible-light driven heterojunction photocatalysts for water splitting–a critical review. Energy Environ. Sci. 2015, 8, 731–759. [Google Scholar] [CrossRef]
  189. Han, S.; Yun, Q.; Tu, S.; Zhu, L.; Cao, W.; Lu, Q. Metallic ruthenium-based nanomaterials for electrocatalytic and photocatalytic hydrogen evolution. J. Mater. Chem. A 2019, 7, 24691–24714. [Google Scholar] [CrossRef]
  190. Yamada, Y.; Shikano, S.; Fukuzumi, S. Robustness of Ru/SiO2 as a hydrogen-evolution catalyst in a photocatalytic system using an organic photocatalyst. J. Phys. Chem. C 2013, 117, 13143–13152. [Google Scholar] [CrossRef]
  191. Gu, Q.; Gao, Z.; Yu, S.; Xue, C. Constructing Ru/TiO2 heteronanostructures toward enhanced photocatalytic water splitting via a RuO2/TiO2 heterojunction and Ru/TiO2 Schottky junction. Adv. Mater. Interfaces 2016, 3, 1500631. [Google Scholar] [CrossRef]
  192. Rozman, N.; Nadrah, P.; Cornut, R.; Jousselme, B.; Bele, M.; Dražić, G.; Gaberšček, M.; Kunej, Š.; Škapin, A.S. TiO2 photocatalyst with single and dual noble metal co-catalysts for efficient water splitting and organic compound removal. Int. J. Hydrogen Energy 2021, 46, 32871–32881. [Google Scholar] [CrossRef]
  193. Loncar, A.; Moriau, L.; Stojanovski, K.; Ruiz-Zepeda, F.; Jovanovic, P.; Bele, M.; Gaberscek, M.; Hodnik, N. Ir/TiONx/C high-performance oxygen evolution reaction nanocomposite electrocatalysts in acidic media: Synthesis, characterization and electrochemical benchmarking protocol. J. Phys. Energy 2020, 2, 02LT01. [Google Scholar] [CrossRef]
  194. Gómez-Cerezo, N.; Sayago-Carro, R.; Cortés-Bazo, A.; Fernández-García, M.; Kubacka, A. PdCu deposited alloys on TiO2 for hydrogen photo-production. Catal. Today 2023, 423, 114280. [Google Scholar] [CrossRef]
  195. Zhou, X.; Dong, H. A theoretical perspective on charge separation and transfer in metal oxide photocatalysts for water splitting. ChemCatChem 2019, 11, 3688–3715. [Google Scholar] [CrossRef]
  196. Rahman, Z.U.; Wei, N.; Feng, M.; Wang, D. TiO2 hollow spheres with separated Au and RuO2 co-catalysts for efficient photocatalytic water splitting. Int. J. Hydrogen Energy 2019, 44, 13221–13231. [Google Scholar] [CrossRef]
  197. Fuentes, R.E.; Farell, J.; Weidner, J.W. Multimetallic electrocatalysts of Pt, Ru, and Ir supported on anatase and rutile TiO2 for oxygen evolution in an acid environment. Electrochem. Solid-State Lett. 2010, 14, E5. [Google Scholar] [CrossRef]
  198. Tanaka, A.; Sakaguchi, S.; Hashimoto, K.; Kominami, H. Preparation of Au/TiO2 with metal cocatalysts exhibiting strong surface plasmon resonance effective for photoinduced hydrogen formation under irradiation of visible light. ACS Catal. 2013, 3, 79–85. [Google Scholar] [CrossRef]
  199. Rutkowska, I.A.; Kulesza, P.J. Electroanalysis of ethanol oxidation and reactivity of platinum-ruthenium catalysts supported onto nanostructured titanium dioxide matrices. J. Electrochem. Soc. 2015, 163, H3052. [Google Scholar] [CrossRef]
  200. Lima, M.J.; Tavares, P.B.; Silva, A.M.; Silva, C.G.; Faria, J.L. Selective photocatalytic oxidation of benzyl alcohol to benzaldehyde by using metal-loaded g-C3N4 photocatalysts. Catal. Today 2017, 287, 70–77. [Google Scholar] [CrossRef]
  201. Caudillo-Flores, U.; Barba-Nieto, I.; Gómez-Cerezo, M.N.; Martínez-Arias, A.; Fernández-García, M.; Kubacka, A. Toward the green production of H2: Binary Pt–Ru promoted Nb-TiO2 based photocatalysts. ACS Sustain. Chem. Eng. 2019, 7, 15671–15683. [Google Scholar] [CrossRef]
  202. Gyawali, S.; Tirumala, R.T.; Andiappan, M.; Bristow, A.D. Carrier dynamics in cuprous oxide-based nanoparticles and heterojunctions. In Proceedings of the Ultrafast Phenomena and Nanophotonics XXVIII, San Francisco, CA, USA, 27 January–1 February 2024; pp. 73–79. [Google Scholar]
  203. Berera, R.; van Grondelle, R.; Kennis, J.T. Ultrafast transient absorption spectroscopy: Principles and application to photosynthetic systems. Photosynth. Res. 2009, 101, 105–118. [Google Scholar] [CrossRef]
  204. Tang, J.; Cowan, A.J.; Durrant, J.R.; Klug, D.R. Mechanism of O2 production from water splitting: Nature of charge carriers in nitrogen doped nanocrystalline TiO2 films and factors limiting O2 production. J. Phys. Chem. C 2011, 115, 3143–3150. [Google Scholar] [CrossRef]
  205. Pendlebury, S.R.; Barroso, M.; Cowan, A.J.; Sivula, K.; Tang, J.; Grätzel, M.; Klug, D.; Durrant, J.R. Dynamics of photogenerated holes in nanocrystalline α-Fe2O3 electrodes for water oxidation probed by transient absorption spectroscopy. Chem. Commun. 2011, 47, 716–718. [Google Scholar] [CrossRef]
  206. Pendlebury, S.R.; Cowan, A.J.; Barroso, M.; Sivula, K.; Ye, J.; Grätzel, M.; Klug, D.R.; Tang, J.; Durrant, J.R. Correlating long-lived photogenerated hole populations with photocurrent densities in hematite water oxidation photoanodes. Energy Environ. Sci. 2012, 5, 6304–6312. [Google Scholar] [CrossRef]
  207. Barroso, M.; Mesa, C.A.; Pendlebury, S.R.; Cowan, A.J.; Hisatomi, T.; Sivula, K.; Grätzel, M.; Klug, D.R.; Durrant, J.R. Dynamics of photogenerated holes in surface modified α-Fe2O3 photoanodes for solar water splitting. Proc. Natl. Acad. Sci. USA 2012, 109, 15640–15645. [Google Scholar] [CrossRef] [PubMed]
  208. Pastor, E.; Pesci, F.M.; Reynal, A.; Handoko, A.D.; Guo, M.; An, X.; Cowan, A.J.; Klug, D.R.; Durrant, J.R.; Tang, J. Interfacial charge separation in Cu2O/RuOx as a visible light driven CO2 reduction catalyst. Phys. Chem. Chem. Phys. 2014, 16, 5922–5926. [Google Scholar] [CrossRef] [PubMed]
  209. Ma, Y.; Pendlebury, S.R.; Reynal, A.; Le Formal, F.; Durrant, J.R. Dynamics of photogenerated holes in undoped BiVO4 photoanodes for solar water oxidation. Chem. Sci. 2014, 5, 2964–2973. [Google Scholar] [CrossRef]
  210. Pesci, F.M.; Cowan, A.J.; Alexander, B.D.; Durrant, J.R.; Klug, D.R. Charge carrier dynamics on mesoporous WO3 during water splitting. J. Phys. Chem. Lett. 2011, 2, 1900–1903. [Google Scholar] [CrossRef]
  211. Gelderman, K.; Lee, L.; Donne, S. Flat-band potential of a semiconductor: Using the Mott–Schottky equation. J. Chem. Educ. 2007, 84, 685. [Google Scholar] [CrossRef]
  212. Tahir, A.A.; Burch, H.A.; Wijayantha, K.U.; Pollet, B.G. A new route to control texture of materials: Nanostructured ZnFe2O4 photoelectrodes. Int. J. Hydrogen Energy 2013, 38, 4315–4323. [Google Scholar] [CrossRef]
  213. Steiner, K.; Chen, W.; Pasquarello, A. Band offsets of lattice-matched semiconductor heterojunctions through hybrid functionals and G0W0. Phys. Rev. B 2014, 89, 205309. [Google Scholar] [CrossRef]
  214. Shevlin, S.; Guo, Z. Density functional theory simulations of complex hydride and carbon-based hydrogen storage materials. Chem. Soc. Rev. 2009, 38, 211–225. [Google Scholar] [CrossRef]
  215. Perdew, J.P.; Levy, M. Physical content of the exact Kohn-Sham orbital energies: Band gaps and derivative discontinuities. Phys. Rev. Lett. 1983, 51, 1884. [Google Scholar] [CrossRef]
  216. Anisimov, V.I.; Zaanen, J.; Andersen, O.K. Band theory and Mott insulators: Hubbard U instead of Stoner I. Phys. Rev. B 1991, 44, 943. [Google Scholar] [CrossRef]
  217. Zhang, X.; Peng, T.; Song, S. Recent advances in dye-sensitized semiconductor systems for photocatalytic hydrogen production. J. Mater. Chem. A 2016, 4, 2365–2402. [Google Scholar] [CrossRef]
  218. Yun, S.; Vlachopoulos, N.; Qurashi, A.; Ahmad, S.; Hagfeldt, A. Dye sensitized photoelectrolysis cells. Chem. Soc. Rev. 2019, 48, 3705–3722. [Google Scholar] [CrossRef] [PubMed]
  219. Vedhanarayanan, B.; Chen, C.-C.; Lin, T.-W. Plasmon-enhanced photocatalytic hydrogen production by dual dye sensitized ternary composite of MoS3/Au core-Ag shell nanoparticles/graphene. J. Power Sources 2020, 477, 229033. [Google Scholar] [CrossRef]
  220. Addanki Tirumala, R.T.; Khatri, N.; Ramakrishnan, S.B.; Mohammadparast, F.; Khan, M.T.; Tan, S.; Wagle, P.; Puri, S.; McIlroy, D.N.; Kalkan, A.K. Tuning Catalytic Activity and Selectivity in Photocatalysis on Mie-Resonant Cuprous Oxide Particles: Distinguishing Electromagnetic Field Enhancement Effect from the Heating Effect. ACS Sustain. Chem. Eng. 2023, 11, 15931–15940. [Google Scholar] [CrossRef]
  221. Li, J.; Cushing, S.K.; Meng, F.; Senty, T.R.; Bristow, A.D.; Wu, N. Plasmon-induced resonance energy transfer for solar energy conversion. Nat. Photonics 2015, 9, 601–607. [Google Scholar] [CrossRef]
  222. Cambiasso, J.; König, M.; Cortes, E.; Schlücker, S.; Maier, S.A. Surface-enhanced spectroscopies of a molecular monolayer in an all-dielectric nanoantenna. Acs Photonics 2018, 5, 1546–1557. [Google Scholar] [CrossRef]
Figure 1. A schematic illustration of H2 generation pathways. Reprinted with permission from Ref. [3]. Copyrights 2020, Elsevier.
Figure 1. A schematic illustration of H2 generation pathways. Reprinted with permission from Ref. [3]. Copyrights 2020, Elsevier.
Catalysts 14 00674 g001
Figure 3. Schematic diagram of a photoelectrochemical cell. Reprinted with permission from Ref. [24]. Copyright 2018, Nature.
Figure 3. Schematic diagram of a photoelectrochemical cell. Reprinted with permission from Ref. [24]. Copyright 2018, Nature.
Catalysts 14 00674 g003
Figure 4. Schematic diagram of bandgap in a PEC reaction.
Figure 4. Schematic diagram of bandgap in a PEC reaction.
Catalysts 14 00674 g004
Figure 6. In three cases, (A) before equilibration between the two phases; after equilibration under dark conditions; and (C) in quasi-static equilibrium under steady-state illumination, the band energetics of an n-type semiconductor–electrolyte contact show the relationships between the electrolyte redox couple (H2O/O2 and H2/H+), the Helmholtz layer potential drop (VH), and the semiconductor work function (Φs), the electrolyte work function (ΦR), the electron quasi-Fermi level (EF,n), and the hole quasi-Fermi level (EF,p). The difference between EF,n and the electrochemical potential of the redox pair of interest—H2O/O2 for n-type semiconductors and H2/H+ for p-type semiconductors—gives the voltage (VOC) that the junction generates when illuminated. Reprinted with permission from Ref. [34]. Copyright 2017, Royal society of chemistry.
Figure 6. In three cases, (A) before equilibration between the two phases; after equilibration under dark conditions; and (C) in quasi-static equilibrium under steady-state illumination, the band energetics of an n-type semiconductor–electrolyte contact show the relationships between the electrolyte redox couple (H2O/O2 and H2/H+), the Helmholtz layer potential drop (VH), and the semiconductor work function (Φs), the electrolyte work function (ΦR), the electron quasi-Fermi level (EF,n), and the hole quasi-Fermi level (EF,p). The difference between EF,n and the electrochemical potential of the redox pair of interest—H2O/O2 for n-type semiconductors and H2/H+ for p-type semiconductors—gives the voltage (VOC) that the junction generates when illuminated. Reprinted with permission from Ref. [34]. Copyright 2017, Royal society of chemistry.
Catalysts 14 00674 g006
Figure 10. SEM images of titanium dioxide arrays (a): vertical view, (b): horizontal view. Reprinted with permission from Ref. [40]. Copyright 2008, Academia.
Figure 10. SEM images of titanium dioxide arrays (a): vertical view, (b): horizontal view. Reprinted with permission from Ref. [40]. Copyright 2008, Academia.
Catalysts 14 00674 g010
Figure 11. The overlapping in bandgaps between two different photocatalysts and the electron-trap mechanism.
Figure 11. The overlapping in bandgaps between two different photocatalysts and the electron-trap mechanism.
Catalysts 14 00674 g011
Figure 12. Schematic illustration of interfacial electron transfer mechanism between TiO2 and Bi2WO6. Reprinted with permission from Ref. [50]. Copyright 2011, American chemical society.
Figure 12. Schematic illustration of interfacial electron transfer mechanism between TiO2 and Bi2WO6. Reprinted with permission from Ref. [50]. Copyright 2011, American chemical society.
Catalysts 14 00674 g012
Figure 13. Flow diagram of the water electrolysis process. Reprinted with permission from Ref. [54]. Copyright 2017, Elsevier.
Figure 13. Flow diagram of the water electrolysis process. Reprinted with permission from Ref. [54]. Copyright 2017, Elsevier.
Catalysts 14 00674 g013
Figure 14. Hydrogen energy production diagram.
Figure 14. Hydrogen energy production diagram.
Catalysts 14 00674 g014
Figure 15. Three polymorphs of TiO2: rutile (tetragonal, P42/mmm), brookite (orthorhombic, Pbca), and anatase (tetragonal, I41/amd). Reprinted with permission from Ref. [93]. Copyright 2017, Nature.
Figure 15. Three polymorphs of TiO2: rutile (tetragonal, P42/mmm), brookite (orthorhombic, Pbca), and anatase (tetragonal, I41/amd). Reprinted with permission from Ref. [93]. Copyright 2017, Nature.
Catalysts 14 00674 g015
Figure 16. The structure of the amorphous TiO2 models with 72 atoms on the left, 96 atoms in the middle, and 216 atoms on the right. The atoms Ti and O are represented by gray and red spheres, respectively. Reprinted with permission from Ref. [94]. Copyright 2012, Elsevier.
Figure 16. The structure of the amorphous TiO2 models with 72 atoms on the left, 96 atoms in the middle, and 216 atoms on the right. The atoms Ti and O are represented by gray and red spheres, respectively. Reprinted with permission from Ref. [94]. Copyright 2012, Elsevier.
Catalysts 14 00674 g016
Figure 17. A schematic illustration of the idea behind photocatalytic water splitting (i e/h segregation, ii is e utilization and iii depicts e utilization for reduction of H+ to H2). Reprinted with permission from Ref. [99]. Copright 2023, American chemical society.
Figure 17. A schematic illustration of the idea behind photocatalytic water splitting (i e/h segregation, ii is e utilization and iii depicts e utilization for reduction of H+ to H2). Reprinted with permission from Ref. [99]. Copright 2023, American chemical society.
Catalysts 14 00674 g017
Figure 18. (A) Pure and (C, S, N)-doped TiO2 valence band (VB) XPS spectra; (B) Suggested bandgap engineering structure for all (F, N)-doped TiO2. Reprinted with permission from Ref. [107]. Copyright 2017, Elsevier.
Figure 18. (A) Pure and (C, S, N)-doped TiO2 valence band (VB) XPS spectra; (B) Suggested bandgap engineering structure for all (F, N)-doped TiO2. Reprinted with permission from Ref. [107]. Copyright 2017, Elsevier.
Catalysts 14 00674 g018
Figure 19. Electron–hole migration pathways. Reprinted with permission from Ref. [97]. Copyright 2023, American chemical society.
Figure 19. Electron–hole migration pathways. Reprinted with permission from Ref. [97]. Copyright 2023, American chemical society.
Catalysts 14 00674 g019
Figure 20. Bandgap engineering in TiO2 using (a) Fe, (b) Ce, and (c) Cu doping, demonstrating the emergence of dopant energy states beneath the conduction band of TiO2 and related carrier dynamics. Recreated with permission from Ref. [106]. Copyright 2017, Elsevier.
Figure 20. Bandgap engineering in TiO2 using (a) Fe, (b) Ce, and (c) Cu doping, demonstrating the emergence of dopant energy states beneath the conduction band of TiO2 and related carrier dynamics. Recreated with permission from Ref. [106]. Copyright 2017, Elsevier.
Catalysts 14 00674 g020
Figure 21. Illustration of the visible-light-induced H2 evolution and pollutant oxidation on organic dye-sensitized TiO2 in water. (1) Photoexcitation and relaxation of organic dye (Dye* refrers to its excited state), (2) electrons’ injection to the conduction band of TiO2 (or Pt/TiO2), (3) H2 production, (4) oxidation of dye, (5) regeneration of dye, (6) oxidation of the electron donors (ED), (7) and (9): reduction of Cr(VI) to Cr(III), (8) reduction of the dissolved O2 to the super oxide anion, and (10) oxidation of the super oxide anion.
Figure 21. Illustration of the visible-light-induced H2 evolution and pollutant oxidation on organic dye-sensitized TiO2 in water. (1) Photoexcitation and relaxation of organic dye (Dye* refrers to its excited state), (2) electrons’ injection to the conduction band of TiO2 (or Pt/TiO2), (3) H2 production, (4) oxidation of dye, (5) regeneration of dye, (6) oxidation of the electron donors (ED), (7) and (9): reduction of Cr(VI) to Cr(III), (8) reduction of the dissolved O2 to the super oxide anion, and (10) oxidation of the super oxide anion.
Catalysts 14 00674 g021
Figure 22. The structural depiction of tri-branched organic dye molecules with mono-, di-, and tri- carboxylate anchoring groups (sensitizers D1, D2, and D3). Reprinted with permission from Ref. [108]. Copyright 2012, Elsevier.
Figure 22. The structural depiction of tri-branched organic dye molecules with mono-, di-, and tri- carboxylate anchoring groups (sensitizers D1, D2, and D3). Reprinted with permission from Ref. [108]. Copyright 2012, Elsevier.
Catalysts 14 00674 g022
Figure 23. Schematic diagram of photocatalytic water splitting by Ag/TiO2 based on the LSPR effect.
Figure 23. Schematic diagram of photocatalytic water splitting by Ag/TiO2 based on the LSPR effect.
Catalysts 14 00674 g023
Figure 24. Possible band arrangements of the two semiconductors in three types of heterojunction. Reprinted with permission from Ref. [126]. Copyright 2018, Elsevier.
Figure 24. Possible band arrangements of the two semiconductors in three types of heterojunction. Reprinted with permission from Ref. [126]. Copyright 2018, Elsevier.
Catalysts 14 00674 g024
Figure 25. Kinds of heterojunction and proposed charge transfer properties as a result of the existing internal electric field. Reprinted with permission from Ref. [126]. Copyright 2018, Elsevier.
Figure 25. Kinds of heterojunction and proposed charge transfer properties as a result of the existing internal electric field. Reprinted with permission from Ref. [126]. Copyright 2018, Elsevier.
Catalysts 14 00674 g025
Figure 26. The Type II heterojunction’s band configuration and electron migration method.
Figure 26. The Type II heterojunction’s band configuration and electron migration method.
Catalysts 14 00674 g026
Figure 27. Schematic band diagrams showing a semiconductor–semiconductor junction: (a) Type II and (b) p-n junction. Schematic demonstration of EHP separation and migration processes in a p-n junction with persistent charge isolation and migration (c) Comparison of band positions with respect to introduced electronic states in p and n type junctions (d) Effect of internal electric field on tunneling rate of photoexcited charge carriers in the p-n junction. Reprinted with permission from Ref. [145]. Copyright 2015, Royal chemical society.
Figure 27. Schematic band diagrams showing a semiconductor–semiconductor junction: (a) Type II and (b) p-n junction. Schematic demonstration of EHP separation and migration processes in a p-n junction with persistent charge isolation and migration (c) Comparison of band positions with respect to introduced electronic states in p and n type junctions (d) Effect of internal electric field on tunneling rate of photoexcited charge carriers in the p-n junction. Reprinted with permission from Ref. [145]. Copyright 2015, Royal chemical society.
Catalysts 14 00674 g027
Figure 28. Illustrations of the EHP migration and isolation mechanisms involved in the photocatalytic reactions of TiO2/NiO hollow semiconductor hybrids. Regenerated with permission from Ref. [160]. Copyright 2015, Royal chemical society.
Figure 28. Illustrations of the EHP migration and isolation mechanisms involved in the photocatalytic reactions of TiO2/NiO hollow semiconductor hybrids. Regenerated with permission from Ref. [160]. Copyright 2015, Royal chemical society.
Catalysts 14 00674 g028
Figure 29. Photocatalytic mechanism of TONT/Al-TONT-40. Regenerated with permission from Ref. [162]. Copyright 2019, Elsevier.
Figure 29. Photocatalytic mechanism of TONT/Al-TONT-40. Regenerated with permission from Ref. [162]. Copyright 2019, Elsevier.
Catalysts 14 00674 g029
Figure 30. The Z-scheme heterojunction’s band configuration and electron migration method.
Figure 30. The Z-scheme heterojunction’s band configuration and electron migration method.
Catalysts 14 00674 g030
Figure 32. Schematic representation of (a) Type II heterojunction photocatalysts, and (b) Z-scheme photocatalytic system.
Figure 32. Schematic representation of (a) Type II heterojunction photocatalysts, and (b) Z-scheme photocatalytic system.
Catalysts 14 00674 g032
Figure 34. TAS spectra, band alignment, and charge transfer mechanism in (a) Co-Pi/Fe2O3, and (b) Cu2O/RuOx, representing long-lived hole (580 nm) and electron populations. Reprinted with permission from Ref. [208]. Copyright 2014, Royal society of chemistry.
Figure 34. TAS spectra, band alignment, and charge transfer mechanism in (a) Co-Pi/Fe2O3, and (b) Cu2O/RuOx, representing long-lived hole (580 nm) and electron populations. Reprinted with permission from Ref. [208]. Copyright 2014, Royal society of chemistry.
Catalysts 14 00674 g034
Figure 35. Typical Nyquist plot showing the origin of Z1 and Z2. Reprinted with permission from Ref. [212]. Ccopyright 2013, Elsevier.
Figure 35. Typical Nyquist plot showing the origin of Z1 and Z2. Reprinted with permission from Ref. [212]. Ccopyright 2013, Elsevier.
Catalysts 14 00674 g035
Figure 36. Schematic illustration of (a) dye sensitization and electron transfer into the conduction band (CB) of the semiconductor (SC), (b) localized surface plasmon resonance (LSPR)-enhanced dye sensitization, and (c) dielectric Mie resonance-enhanced dye sensitization. Reprinted with permission from Ref. [218]. Copyright 2019, Royal chemical society.
Figure 36. Schematic illustration of (a) dye sensitization and electron transfer into the conduction band (CB) of the semiconductor (SC), (b) localized surface plasmon resonance (LSPR)-enhanced dye sensitization, and (c) dielectric Mie resonance-enhanced dye sensitization. Reprinted with permission from Ref. [218]. Copyright 2019, Royal chemical society.
Catalysts 14 00674 g036
Table 1. Hydrogen production methods [15].
Table 1. Hydrogen production methods [15].
MethodTechnique UsedFeed StockTitle
PhotonicPV electrolysisWaterElectricity is produced using PV panels
PhotocatalysisWaterUsing electron–hole pairs produced by the photocatalyst, water is divided into O2 and H2
Photo-electrochemical methodWaterA hybrid cell generates voltage and current to absorb light
BiochemicalDark fermentationBiomassA hybrid cell generates voltage and current to absorb light
Electrical + ThermalHigh temperature electrolysisWaterWater splitting occurs at a high temperature using thermal and electrical energy
Hybrid thermochemical cycleWaterCycles of chemical reactions are driven by electrical and thermal energy
Coal gasificationCoalCoal is transformed into syn-gas
Reforming of fossil fuelFossil fuelFossil fuels produce H2 and CO2
ElectricalElectrolysisWaterWater is split by direct current into O2 and H2
PlasmaFossil fuelsCarbon soot and H2 are produced when natural gas is pushed through a plasma arc
ThermalThermolysisH2SH2S thermally decomposes at high temperatures
Thermochemical processesWater splittingWaterWater is oxidized by converting sunlight into electron–hole pairs
Biomass conversionBiomassConversion of fermentable hydrogen via thermo-catalysis using biomass-based materials
GasificationBiomassBiomass is transformed into syn-gas
ReformingBiomassliquid bio fuels to transform into H2
Table 2. Comparison analysis of electrolysis technologies, including PEM, AEM, alkaline, and SOEC, based on distinct characteristics and attributes [71].
Table 2. Comparison analysis of electrolysis technologies, including PEM, AEM, alkaline, and SOEC, based on distinct characteristics and attributes [71].
CharacteristicPEMAEMAlkalineSOEC
ElectrolyteProton-exchange membraneAnion-exchange membraneLiquid KOH solutionSolid oxide ceramics (e.g., YSZ)
Electrolyte MaterialAcidic/solid (polymer)Alkaline/solid (polymer)Alkaline/liquidO2 or H+ conducting/solid (ceramic)
Working PrincipleCatalysts 14 00674 i001Catalysts 14 00674 i002Catalysts 14 00674 i003Catalysts 14 00674 i004
Operating Temperature50–80 °C40–90 °C60–80 °C700–1000 °C
Electrode MaterialPlatinum group metals (PGMs)Non-precious metalsNickel-basedPerovskites, ceramics
Efficiency60–70%50–60%65–70%80–90%
DurabilityHighModerateHighLow to moderate
Water QualityHigh-purity water requiredHigh-purity water requiredLow-purity water requiredHigh-purity water required
Gas PurityHighModerate to highModerateHigh
System ComplexityHighModerateLowHigh
Start-up TimeShortShortModerateLong
CostHighModerateLowHigh
ScalabilityHighHighModerate to highLow to moderate
Commercial MaturityHighEmergingMatureResearch and development
ApplicationsMobility, small-scale hydrogenIndustrial, potential for mobilityLarge-scale industrialHigh-temperature industrial process
Table 3. Four types of water electrolyzers [80].
Table 3. Four types of water electrolyzers [80].
AlkalinePEMAEMSolid Oxide
Operating temperature70–90 °C50–80 °C40–60 °C700–850 °C
Operating pressure1–30 bar<70 bar<35 bar1 bar
ElectrolytePotassium hydroxide (KOH) 5–7 mol/LPFSA membranesDVB polymer support with KOH or NaHCO3 1 mol/LYttria-stabilized zirconia (YSZ)
SeparatorZrO2 stabilized with PPS meshSolid electrolyte (above)Solid electrolyte (above)Solid electrolyte (above)
Electrode/catalyst (oxygen side)Nickel coated perforated stainless steelIridium oxideHigh-surface-area nickel or NiFeCo alloysPerovskite-type (e.g., LSCF, LSM)
Electrode/catalyst (hydrogen side)Nickel coated perforated stainless steelPlatinum nanoparticles on carbon blackHigh surface area nickelNi/YSZ
Porous transport layer anodeNickel mesh (not always present)Platinum coated sintered porous titaniumNickel foamCoarse nickel mesh or foam
Porous transport layer cathodeNickel meshSintered porous titanium or carbon clothNickel foam or carbon clothNone
Bipolar plate anodeNickel-coated stainless steelPlatinum coated titaniumNickel-coated stainless steelNone
Bipolar plate cathodeNickel-coated stainless steelGold-coated titaniumNickel-coated stainless steelCobalt-coated stainless steel
Frames and sealingPSU, PTFE, EPDMPTFE, PSU, ETFEPTEF, siliconCeramic glass
Table 4. Distribution of energy in the terrestrial solar spectrum (AM 1.5) [82].
Table 4. Distribution of energy in the terrestrial solar spectrum (AM 1.5) [82].
Spectral RegionWavelength [nm]Energy [eV]Contribution to Total Spectrum [%]
Near-UV315–4003.93–3.092.9
Blue400–5103.09–2.4214.6
Green/yellow510–6102.42–2.0316.0
Red610–7002.03–1.7713.8
Near-IR700–9201.77–1.3423.5
Infrared920–14001.34- 0.8829.4
Table 5. List of co-catalysts used in photocatalytic water-splitting experiments [89].
Table 5. List of co-catalysts used in photocatalytic water-splitting experiments [89].
Co-CatalystsExamples
Noble metalsAu, Pt, Pd, Ru, and Ag
Transition metalsNi, Cu, and Co
Metal oxidesCuO, NiO, and Cu2O
Metal sulfidesNiS, CuS, MoS2, and WS2
Table 6. Some commonly used dyes and their classifications [107].
Table 6. Some commonly used dyes and their classifications [107].
ClassDye
ThiazinesThionine, methylene blue, new methylene blue, azure A, azure B, azure C
HiazinesToluidine blue
PhenazinesPhenosafranin, safranin-O, safranin-T, neutral red
XanthenesFluorescein, erythrosin, erythrosin B, rhodamin B, rose Bengal, pyronine Y, eosin, rhodamine 6G
AcridinesAcridine orange, proflavine, acridine yellow
Triphenyl methane derivativesFusion, crystal violet, malachite green, methyl violet
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ishaq, T.; Ehsan, Z.; Qayyum, A.; Abbas, Y.; Irfan, A.; Al-Hussain, S.A.; Irshad, M.A.; Zaki, M.E.A. Recent Strategies to Improve the Photocatalytic Efficiency of TiO2 for Enhanced Water Splitting to Produce Hydrogen. Catalysts 2024, 14, 674. https://doi.org/10.3390/catal14100674

AMA Style

Ishaq T, Ehsan Z, Qayyum A, Abbas Y, Irfan A, Al-Hussain SA, Irshad MA, Zaki MEA. Recent Strategies to Improve the Photocatalytic Efficiency of TiO2 for Enhanced Water Splitting to Produce Hydrogen. Catalysts. 2024; 14(10):674. https://doi.org/10.3390/catal14100674

Chicago/Turabian Style

Ishaq, Tehmeena, Zainab Ehsan, Ayesha Qayyum, Yasir Abbas, Ali Irfan, Sami A. Al-Hussain, Muhammad Atif Irshad, and Magdi E. A. Zaki. 2024. "Recent Strategies to Improve the Photocatalytic Efficiency of TiO2 for Enhanced Water Splitting to Produce Hydrogen" Catalysts 14, no. 10: 674. https://doi.org/10.3390/catal14100674

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