Next Article in Journal
Visible-Light-Sensitive Polymerizable and Polymeric Triazine-Based Photoinitiators with Enhanced Migration Stability
Next Article in Special Issue
ZrO2-Based Photocatalysts for Wastewater Treatment: From Novel Modification Strategies to Mechanistic Insights
Previous Article in Journal
Photochemical Water Splitting via Transition Metal Oxides
Previous Article in Special Issue
Chromium (III) Ions Were Extracted from Wastewater Effluent Using a Synergistic Green Membrane with a BinaryCombination of D2EHPA and Kerosene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Single-Atom Transition Metal Photocatalysts for Hydrogen Evolution Reactions

Department of Chemical and Biological Engineering, Gachon University, Seongnam-si 13120, Gyeonggi-do, Korea
*
Author to whom correspondence should be addressed.
Submission received: 8 September 2022 / Revised: 18 October 2022 / Accepted: 21 October 2022 / Published: 24 October 2022

Abstract

:
Hydrogen is one of the potential fuels that is easily stored in ammonia compounds and reacts with oxygen in an environmentally friendly manner, producing water and transferring a significant amount of heat for powering mechanical facilities or transportation. Recently, single-atom photocatalysts have attracted significant attention owing to their ability to produce clean fuels or reduce gaseous pollution, thereby contributing to the preservation of our planet. Utilizing metals composed of a single atom on a semiconductor platform can improve the active sites, thereby increasing the efficiency of the hydrogen evolution reaction. This review focuses on the use of single-atom transition metals as photocatalysts in a solar-powered water-splitting system that produces hydrogen gas. The approach to synthesis, reaction mechanism, and current performance of these materials is exhaustively discussed. In addition, the main challenges and improvement strategies are highlighted.

Graphical Abstract

1. Introduction

The impending depletion and negative environmental effects of fossil fuels necessitate a vast supply of renewable energy sources for human life and industry. Hydrogen (H2) is considered as a potential renewable energy source owing to its abundance, high energy density, and environmental friendliness [1,2,3]. In addition, hydrogen plays an important role in numerous applications, including fuel, oil refinery, fertilizer, and metal refining. Figure 1 illustrates solar hydrogen production and its applications. Using solar energy, the photocatalyst absorbs photons to produce excited electrons that can combine with a proton to produce hydrogen gas. This approach is sustainable and can satisfy the demand for a large-scale, low-cost hydrogen generation system [4,5]. H2 possesses the lightest weight, the strongest reaction with oxygen that produces water, and a tremendous amount of heat energy, allowing it to operate machinery or generate electricity via a converter [6]. The majority of hydrogen gas is produced by methane steam reforming or oil reforming, but water splitting is significantly less expensive, requires fewer facilities, consumes renewable solar energy, and emits non-toxic by-products [7,8,9]. Therefore, the development of this sustainable energy pathway is the optimal choice for the energy economy of the future.
Protons (H+) receive an electron to produce hydrogen, resulting in the formation of hydrogen gas and formation of a bubble on the surface of materials [10]. This reaction depends on the redox potential of H+/H2, and the active proton source depends on the pH of the electrolyte, where the protons are from the free H+ in acidic media or the H2O molecule in basic media [11,12,13,14]. For example, in the acidic media (pH < 5), the protons are directly absorbed on the surface of material, receive the electrons, and are reduced to hydrogen. Meanwhile, in basic media (pH > 8), the protons are separated from water molecules on the surface of active material. Therefore, the water splitting potential depends on the pH level; the HER potential is 0 V vs. standard hydrogen electrode (SHE) at pH = 0, while it is −0.83 V vs. SHE at pH = 14 [15,16]. This change can be calculated as a function of pH; thus, to simplify the reaction, we can mention the behaviour of the adsorbed H+ (namely, Hads). Accordingly, the adsorption ability determines the performance of the hydrogen evolution reaction (HER). The main active material in single-atom photocatalysts is the light harvester, a semiconductor, which absorbs visible light or UV light and then excites the formation of electron-hole pairs, producing an electron source for HER [17]. Excited electrons with higher energy than the band gap of the semiconductor will transfer to the conduction band, while the holes remain in the valence band or join in an oxidation reaction [18]. By selecting a semiconductor with a conduction band close to the H+/H2 reduction level of ~4.5 eV (or 0 V vs. SHE), protons can receive electrons from host materials with relative ease. This is how conventional photocatalysts function [19,20].
Recently, single-metal atom (SMA) photocatalysts have attracted considerable interest [21,22,23]. The term of SMA photocatalyst indicates the decoration of a very tiny metal particle in atomic scale on the host photocatalyst (semiconductors). The single metal functions as a co-catalyst on the surface of semiconductors, increasing the number of active sites, receiving electrons from the light-harvesting process, and efficiently reducing the amount of precious metal. Typically, precious metals such as Pt and Au have a low work function that is close to the H+/H2 reduction level; as a result, they have been utilized as hydrogen catalysts in electrochemical and photoelectrochemical hydrogen generation [24]. Due to the Schottky contact between the metal atom and semiconductor host material, the electron can easily transfer to a lower potential and is able to donate electrons to Habs, thereby forming hydrogen molecules [25,26]. Common semiconductor hosts include perovskites, TiO2, and NiO, as well as graphitic carbon nitride (g-C3N4) [27]. Meanwhile, to reduce the expense of precious metals, numerous types of abundant transition metals such as Pt, Au, Ni, V, and Fe are used to produce SMA photocatalysts [28,29,30,31].
In this review, the use of transition metal as a single-atom photocatalyst is examined. In addition, the fundamentals and predictions based on theoretical calculations are discussed, and the evaluation, recent developments, and future prospects are discussed.

2. Fundamentals

Typically, the SMA photocatalyst consists of a single metal atom or tiny metal nanoparticle anchored on a host material, as shown in Figure 2a. The host material is semiconductor-type material, which itself can act as a photocatalyst or light harvesting material [21]. As mentioned above, the requirement of active material for HER should have a conduction band close to ~4.5 eV. However, it is hard to tune a band gap of semiconductors as designed. For metal, the conduction band and valence band are overlapped or have no band gap. Therefore, the electron and hole are easily formed in normal conditions. As per previous reports, the work function of metal can be understood by the “activity descriptors” [32]. Thus, Pt with the work function of ~5.5 eV seems to be the best element to effectively perform HER [24]. In fact, many other factors of transition metal affect HER performance, which will be discussed later. To operate the HER, the transition metal nano particle should be anchored on semiconductor material, as described in the schematic in Figure 2a; the host material acts as the light harvester, then gives the generated electrons to metal NPs. The tiny metal NPs receive the electron and perform the HER, producing H2 gas. Due to the natural catalytic activity and the increasing surface area of metal nanoparticles, photocatalytic properties of SMA can enhance outstanding photocatalytic properties of bare semiconductor material [33].
To fully comprehend the mechanism of hydrogen absorption and reduction on the surface of a transition metal, the density functional theory (DFT) calculation can be used to gain insight into the electron structure, particularly the d-orbital of the transition metal, which determines the conducting electron energy [10,36,37,38]. After adsorption on the material surface, the proton forms a bonding state that can be described as a 1s-d hydrogen-transition metal bond. Hammer et al. reported that the density of one-electron states (DOS) of four typical metals in this group, including Pt, Au, Ni, and Cu, were calculated to predict the chemisorbed hydrogen on their surfaces [39]. Au and Cu have electrons in their H 1s-d antibonding states, while those of Ni and Pt are empty. It indicates that metals interact more strongly with Hads of Ni and Pt than with Hads of Au and Cu. However, the H 1s-d bonding in both these transition metals shows a strong intensity, indicating they can highly absorb the hydrogen atom. Therefore, the position and occupation of d-orbitals determine the hydrogen adsorption properties of metal surfaces. In HER, the bond-breaking barrier and the electron transfer process are also essential [40]. Metals appear to have the lowest bond-breaking barrier, allowing them to readily transfer electrons and activate hydrogen in HER. Thus, the use of metal atoms such as Pt, Au, Pd, Co, Ni, and Cu in photocatalyst systems will reduce the massive electron transfer effort of the semiconductor, resulting in more efficient hydrogen production.
The ability to absorb hydrogen can be described by a physical quantity called Gibbs free energy ( Δ G H * ) [41,42]. The value of Δ G H * is more negative, indicating that hydrogen bonds to the surface more strongly. The greater the positive value, the more difficult it is to absorb hydrogen. Thus, the material with the closest value to zero for Δ G H * is the best material for interacting with hydrogen atoms with the lowest activation energy, as it can absorb and release the proton and hydrogen gas, respectively, with ease. Figure 2b depicts the typical free energy diagram of Ni metal at various sites [34]. The Ni 1-zigzag edge has the lowest   Δ G H * value of −0.01 eV, indicating HER’s high activity. In contrast, the other coordinated sites exhibit either strong or very weak hydrogen bonds, making hydrogen absorption difficult. By repeating the calculation with other transition metals, such as Mn, Fe, Co, Ni, Cu, and Pd, Gao et al. obtained a clear picture of these materials in which, if the right coordinated sites are exposed in the bonding with photocatalyst active sites, they are both highly active in HER [34]. Figure 2c displays the volcano plots of Gibbs free energy for the majority of transition metals (Pt, Au, Pd, Rh, Ir, Ni, Co, Mo, Nb, W, Ag), which were summarized by Yao et al. using previously reported data [35,43,44,45,46]. It is easy to evaluate transition metals for hydrogen adsorption based on this graph. However, as stated previously, the Gibbs free energy of a metallic surface is highly dependent on its coordinated sites. Therefore, a graph with numerous coordinated sites must be completely filled to provide an overview of transition metal in HER.

3. Single-Atom Metal Photocatalyst Preparation

To synthesize single-atom photocatalysts, numerous techniques can be used, which can be categorized into three approaches: Absorption, deposition, and refining method (as shown in Figure 3). In the absorption method, metallic salts were absorbed by a structure, such as a metal organic framework (MOF), a polymer structure, or a mixture of precursors of host materials. Zhang et al., for instance, fabricated a few types of SMAs, including Cu, Co, Ni, Fe, Mn, Zn, and Pt, by diffusing these ions into a MIL-125-Ti, which is an MOF structure of Ti with terephthalic acid, and then reducing this structure to SMA on TiO2 host structure as a photocatalyst for hydrogen production [47]. Cha et al. used TiO2 dissolved in Pt, Pd, and Au chlorides [48]. The absorption of these ions on the surface of TiO2 produces 2–5 nm-sized particles that are considered atomic-scale catalysts. The atomic deposition technique permits a single atom to exist on the surface of materials within seconds. Pan et al. investigated the electrostatic deposition method for SMA Ni particle on ZnIn2S4, as an example [49]. Due to the large surface area of ZnIn2S4, Ni2+ ions are readily attracted to the material via electronic attraction. By hydrothermally reducing Ni2+, the SMA Ni particle was uniformly decorated on the surface of the host material. Meanwhile, Zhou et al. employed a three-electrode system and a solution of H2PtCl6 as a Pt source for electrodeposition on Ni/NiO/Ag nanowire [50]. The amount of Pt particles can be manipulated by adjusting the time and current/voltage of the electrodes system. In the final step of the refining method, metal ions were reduced through a natural decomposition process under light/dark conditions (with light-sensitive compounds of Pt, Ag, and Au) or through the use of reduction agents such as hydrogen or calcination with a reducing agent. Vile et al. applied sodium borohydride to C3N4 nanosheets to reduce Ni2+ [51]. Additionally, Zhang et al. used a high annealing temperature of ~850 °C to decompose cobalt phthalocyanine in a composition containing carbon black in the presence of nitrogen [52]. Moreover, Zhou et al. used the self-decomposition of IrCl3 in the absence of light to produce single-atom Ir on TiO2 nanotubes [53]. Table 1 provides a summary of additional methods. Approaching the synthesis of SMA photocatalysts may not be a simple process, but it is not an impediment to obtaining the desired photocatalysts.

4. Single-Atom Photocatalysts

4.1. TiO2 Base

TiO2 is a well-known photocatalyst with a wide band gap of ~3.2 eV that absorbs ultraviolet light intensely (UV) [58,59]. In addition, it has been used globally for numerous applications, including batteries, ceramic compounds, colorants, and solar cells, owing to its low cost, low toxicity to the environment, and ease of processing [60,61,62]. As a hydrogen photocatalyst, TiO2 shows a low efficiency due to the ultrafast recombination of excited electron–hole pairs [63]. Correspondingly, the morphology and size of TiO2 can be altered to improve its catalytic performance [62,64]. However, to reach the large-scale production requirement, there is still a gap to overcome. Numerous strategies aim to modify the TiO2 surface, including the use of semiconductors with a smaller band gap and the combination of metal particles and graphene, among others. Both received a tremendous performance boost from TiO2. In particular, the use of SMA on photocatalysts opens a new route for functionalizing the TiO2 surface with tiny metal particles.
Figure 4a,b demonstrates that Yi et al. synthesized single-atom Co on N-doped graphene composited with TiO2 nanobelts, resulting in a significant improvement [65]. The rate of H2 generation could reach ~677 μmolh−1 g−1, which is comparable to the performance of Pt-decorated TiO2 samples (~741 μmolh−1 g−1). Meanwhile, the hydrogen production efficiency of unmodified TiO2 and N-doped graphene/TiO2 is poor and low, respectively. Using a similar approach, Cha et al. compared the effect of using the noble metals Pt, Pd, and Au on the surface of TiO2, as depicted in Figure 4c,d [48]. The size of tiny noble metals was between 2 and 5 nm. By adjusting the concentration of noble metal chlorides, it is possible to observe the effect of SMA on TiO2 host material, which demonstrates that a single atom can enhance the performance of photocatalysts by increasing the number of active sites, without obstructing the light to the host material compared to larger nanoparticles. The best performance as an SMA photocatalyst was recorded for Pd, which has a hydrogen production rate of ~600 μLh−1. Zhou et al. deposited Ir on TiO2 nanotubes using a dark deposition technique, loading the nanotubes with Ir nanoparticles measuring ~2.5 nm in size [53]. The SMA on TiO2 photocatalyst performs significantly better than the conventionally prepared co-catalyst TiO2, achieving a high turnover frequency of ~4 × 106 h−1. Zhang et al. reported the SMA Cu on TiO2, as depicted in Figure 5, after selecting a material with a lower cost [47]. The trapped Cu2+ ion in MIL125-Ti was easily converted to SMA Cu/TiO2 via high-temperature calcination. By comparing various types of transition metals, such as Co, Ni, Fe, Mn, Zn, and Pt, the author concluded that single atom Cu (~2 nm) on TiO2 possesses a high H2 evolution rate of ~100 mmolh−1 g−1, which is higher than that of SMA Pt on TiO2 (~80 mmolh−1 g−1). The significant improvement of these SMA photocatalysts can be attributed to two factors. First, the new approach using MIL-125-Ti demonstrates that an MOF can absorb other ions efficiently if their ionic size and repulsion/attraction force are suitable. In this work, Cu is abundantly absorbed by the network, resulting in the formation of 1–2 atomic-sized uniform single atoms. Cu is a well-known metal that is utilized in HERs effectively due to its compatibility with the H+/H2 redox potential. The disintegration of MIL-125-Ti also generates a large number of Ti vacancies, thereby increasing the specific surface area and exposing sites for Cu atoms. Therefore, the use of SMA in conjunction with a highly porous host material can increase the efficiency of hydrogen generation.

4.2. g-C3N4 Base

Graphitic carbon nitride (g-C3N4), a two-dimensional layered material that is a member of the graphene-like family, is an efficient metal-free photocatalyst material [66]. In contrast to graphene, which has a zero band gap, g-C3N4 has a band gap of ~2–3 eV [67,68]. Consequently, g-C3N4 has a broad light absorption band. In addition, g-C3N4 has a simple synthesis method, high physical and chemical stability, and an active band gap in relation to the redox potential of water-splitting reaction [69]. Therefore, g-C3N4 materials can be utilized as photocatalysts for HER. However, due to its low conductivity and rapid recombination of excited charge carriers, g-C3N4 is ineffective as a photocatalyst [70]. Accordingly, surface engineering of g-C3N4 is required to overcome the limitation and exploit the semiconducting properties effectively. Due to the abundance of nitrogen atoms in its structure, the surface of g-C3N4 can be easily modified by covalent or noncovalent bonding with a variety of functional groups [71]. During its early use as a photocatalyst, the metal decoration in the surface of g-C3N4 has garnered considerable attraction [72,73]. The emergence of SMA properties on the host photocatalyst material g-C3N4 can have a substantial impact on their hydrogen production performance. Cao et al. utilized a single atom of noble Pt metal on g-C3N4 as an effective photocatalyst in HER, as depicted in Figure 6a–b [74]. The Pt atoms are widely distributed on the host surface materials as individual atoms, thereby enhancing the active site and photocatalytic properties. Pt is used not only to increase the active surface area but also to improve the electronic conducting property, resulting in outstanding catalytic properties. The hydrogen production rate of SA Pt on g-C3N4 is significantly greater than that of bare g-C3N4 and greater than 13 times that of Pt nanoparticles of larger size (a few nanometers) on g-C3N4. In addition to V, Co, Cu, and Fe, other transition metals are used in the single atomic concept of g-C3N4. To display a wider graph of SMA on g-C3N4, however, the synthesis approach still has some limitations. Li et al. fabricated SMA Co metal on P-doped g-C3N4 (Co/P/CN) using hexahydrate triethanolamine as a reducing agent during the calcination process in a microwave reactor, as depicted in Figure 6c,d [75]. The rate of H2 production was measured at ~3730 μmolh−1 g−1. The addition of Co and P atoms to g-C3N4 increases its active surface area. Consequently, the presence of P-N and Co-N bindings causes a change in the band gap structure, which is reduced from ~2.78 eV for bare g-C3N4 to ~2.58 eV for Co/P/CN, as depicted in Figure 6c. Therefore, the photogenerated hole–electron pairs could easily separate and contribute to the HER. Similarly, Wang et al. reduced V ion during the preparation of g-C3N4, resulting in a single atom of V on g-C3N4 photocatalyst (SAVCN) for hydrogen evolution reaction (Figure 6e,f) [55]. Additionally, the simulations demonstrated that the V on SAVCN decreases the Δ G H * from −1.11 to −0.34 eV, which is closer to 0 eV and therefore better for hydrogen adsorption. Under blue and green LED lights, the H2 production capacity of the SAVCN catalyst was measured and found to be ~5.0 and 3.0 mmolh−1 g−1, respectively. Consequently, the SMA improved the light absorption, hydrogen adsorption, and band gap structure of the host material to facilitate the HER.

4.3. Other Host Materials

In addition to TiO2 and g-C3N4, which are the traditional photocatalyst host materials, there are numerous other semiconductor materials whose band gap matches the requirements for a photocatalyst material [76]. However, their use is contingent on their influence on low-cost production, simple processing, and environmentally friendly properties, which are not adequately accounted for by their benefits and drawbacks. Graphene quantum dots, for instance, are one of the most promising candidates; however, large-scale production and uniformity are still being evaluated [23]. Due to their layered structure, photo-stability, low band gap of ~2.5 eV, and low toxicity, ZnIn2S4 semiconductor materials have recently attracted the attention of researchers [77,78,79]. To date, Pan et al. investigated the SMA Ni on ZnIn2S4 materials for HER, as shown in Figure 7a,b [49]. Preparing a ZnIn2S4 semiconductor with sulphur vacancies and synergizing it with SMA Ni on the surface enabled the photocatalytic evolution of hydrogen. The evolution rate with 0.9 wt.% Ni on ZnIn2S4 was approximately 89.4 μmolh−1. Shi et al. fabricated SMA Pt on hexagonal ZnIn2S4 [80]. Calculating a new Δ G H * at −0.23 eV for the third hydrogen atoms demonstrated that the presence of Pt increases the hydrogen adsorption of the photocatalyst. Under solar light, the rate of hydrogen production was approximately 30 mmolg−1 h−1, and under visible light, it was greater than 16 mmolg−1 h−1. These outcomes suggest ZnIn2S4 bearing SMA coating is also a possible photocatalyst for hydrogen production.
Organometal halide perovskites (OHP) are well-known in the solar cell industry due to their tuneable band gap, broad light absorption band, long carrier diffusion length, and simple synthesis [83]. In particular, the photocatalytic splitting of hydroiodic acid (HI) in aqueous solution promotes the solar-powered production of hydrogen [84,85,86,87,88,89,90]. Zhou et al. identified Cs2SnI6 perovskites combined with SMA Pt as a promising candidate for the treatment of HER [91]. The Pt-I3 sites were proposed to have a strong metal-support interaction effect, enhancing the photocatalytic performance of HER. Wu et al. also utilized a single Pt atom on a formamidinium (FA)-based lead perovskite containing Br and I atoms [81]. The FA perovskites have a small band gap and halide mixture, which increases their stability and light absorption, as illustrated in Figure 7c. Pt on FA perovskites exhibited a Δ G H * of ~0.04 eV, which is suitable for the adsorption of hydrogen atoms and the subsequent release of hydrogen gas. Approximately 700 μmolh−1 of hydrogen was produced. From light-harvesting materials, OHP with SMA could be a candidate for hydrogen production and other photocatalytic applications.
In addition to conventional semiconductors, the 2D structure, high stability, and catalytic properties of covalent organic frameworks (COFs) make them attractive for photocatalyst systems [92,93,94,95,96,97,98,99,100]. High catalytic performance is exhibited by COFs containing transition metals when reducing the toxic organic chemical CO2 or producing hydrogen. For example, Zhong et al. synthesized single Ni site COFs for CO2 and H2 production [95]. The COFs were produced using a solvothermal method involving 1,3,5-triformylphloroglucinol and 5,5′-diamino-2,2′-bipyridine, resulting in a material that is stable and more selective for CO2 reduction than HER. In COFs, the selected metal and framework play a greater role in determining their catalytic properties. Dong et al. used a Pt single atom in β-ketoenamine-linked COFs (Pt-SA-TPa-1-COFs), as shown in Figure 7d,e [82]. The 2D Pt-SA-TPa-1-COFs exhibit numerous active sites and a low band gap of ~2.0 eV, resulting in a strong absorption of visible light from ~550 nm to the near IR band. Consequently, the calculation of Δ G H * reveals a low energy of –0.092 eV, activating the hydrogen adsorption property and encouraging the release of H2 following production. Pt-SA-TPa-1-COFs can generate hydrogen at a rate of 719 μmolh−1 g−1. With an SMA on the surface, the development of semiconductors is accelerating. Accordingly, numerous semiconductor laboratories are employing this approach to identify the most effective catalyst activator for OER, HER, CO2 reduction, and organic reduction. They can be metal oxide material, MOF, metal sulphides, metal selenides, etc. [50,101,102,103,104,105,106].
To evaluate the performance of different materials, Table 2 shows the summary of the single atom photocatalyst for hydrogen evolution reaction with various metals and host materials. It can be clearly recognized that in each host material, the different transition metal shows the different behaviour. In general, the presence of transition metal can improve the catalytic properties of host material. This is because free electron can be easily generated and transferred from host material to give it to Hads. Pt is the most common metal for enhancing catalytic performance, due to its natural catalytic properties [74]. However, the abundance and price of Pt limits its commerciality. Therefore, Cu, Ni, and Co are the next candidates for Pt replacement [47,107]. Moreover, the host materials in nano scale, such as nanosheets, nanobelts, or particles, are generally utilized. It is worth noting that a high surface area is a prerequisite for a light harvester. In Table 2, the TiO2 is derived from MIL125-Ti structure, which can achieve a high surface area from the collapse of this framework. Therefore, it can boost the hydrogen production up to ~100 mmol g−1 h−1 [47]. Moreover, the use of Cu in TiO2 shows a better performance in comparison to Pt. This indicates that the suitable structure and metal can generate a synergistic effect for their catalytic properties. Therefore, host materials in the right form with metals are promising to be a future active material for hydrogen production.

5. Conclusions and Outlook

The use of single-atom photocatalyst has reached a milestone in terms of its physical and chemical properties, thereby enhancing the performance of the catalyst in HER. In addition, it allows any type of light harvester to become a photocatalyst for hydrogen production, CO2 reduction, oxygen evolution reaction (OER), or organic oxidation. In order to obtain the superior photocatalyst, various heterogeneous or homogeneous methods are utilized. Various in/ex situ experimental/computational techniques, such as transmission electron microscopy, X-ray photoelectron spectroscopy, X-ray absorption spectroscopy, and DFT calculation, gradually reveal the structure and electronic properties. The mechanism of the photocatalyst is partially revealed. Recorded hydrogen production rates range from hundreds of micro to millimoles per hour. Accordingly, these are significant steps toward the future of producing clean, environmentally friendly fuels.
Despite numerous advancements, the SMA photocatalyst still faces a number of obstacles and requires additional research in areas such as selectivity, metal coordination investigation in practical experiments, stability, and controllable processes. For instance, computational research readily identified the transition metals (Mn, Fe, Co, Ni, Cu, Pd, and Pt) with distinct Gibbs free energy coordination for hydrogen absorption; however, there is still a gap where experimental work can determine the behaviour of transition metals with various coordination [34]. In addition, computational work is necessary to predict and comprehend the behaviour of single-metal atoms on various types of materials. It has been suggested that OHP are highly stable; however, there is a lack of experimental evidence demonstrating how long the material can maintain continuous photocatalytic performance. The combination of light-harvesting materials with varying absorption ranges can be an effective strategy for using solar light to drive photocatalytic reactions.
In addition, the OER reaction is also a boost-up key for hydrogen evolution reaction due to its counter reaction to HER [110,111]. As the balancing system, the generated electron–hole couple should be effectively separated and join in HER and OER. Therefore, the rate of the overall system is determined based on the rate of HER and OER. Unlike HER, OER works based on oxidation by the valence band of semiconductors [112,113]. Recently, the single atom photocatalyst also shows a great improvement for OERs; therefore, with the same concept, SMA photocatalysts are promising for practical and sustainable environmental devices [114,115]. Solar driven HER and OER by two separated electrodes will be the ideal system, efficiently producing H2 and O2. However, they both face the same hurdles to overcome: full investigation, material preparation, and theoretical predictions. The rapid rise of single atoms to the forefront of photocatalyst research demonstrates a promising future for the use of a brand new catalyst material to address a variety of issues, including energy, the environment, and healthcare.

Author Contributions

T.P.N.: Conceptualization, visualization, writing, review, and editing. I.T.K.: project administration, funding acquisition, review, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by a National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (NRF-2022R1F1A1062928). This research was also supported by the Basic Science Research Capacity Enhancement Project through a Korea Basic Science Institute (National Research Facilities and Equipment Center) grant funded by the Ministry of Education (2019R1A6C1010016).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Turner, J.; Sverdrup, G.; Mann, M.K.; Maness, P.-C.; Kroposki, B.; Ghirardi, M.; Evans, R.J.; Blake, D. Renewable hydrogen production. Int. J. Energy Res. 2008, 32, 379–407. [Google Scholar] [CrossRef] [Green Version]
  2. Dubouis, N.; Grimaud, A. The hydrogen evolution reaction: From material to interfacial descriptors. Chem. Sci. 2019, 10, 9165–9181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Nguyen, T.P.; Tuan Nguyen, D.M.; Tran, D.L.; Le, H.K.; Vo, D.-V.N.; Lam, S.S.; Varma, R.S.; Shokouhimehr, M.; Nguyen, C.C.; Le, Q.V. MXenes: Applications in electrocatalytic, photocatalytic hydrogen evolution reaction and CO2 reduction. Mol. Catal. 2020, 486, 110850. [Google Scholar] [CrossRef]
  4. Wang, X.; Jiang, X.; Sharman, E.; Yang, L.; Li, X.; Zhang, G.; Zhao, J.; Luo, Y.; Jiang, J. Isolating hydrogen from oxygen in photocatalytic water splitting with a carbon-quantum-dot/carbon-nitride hybrid. J. Mater. Chem. A 2019, 7, 6143–6148. [Google Scholar] [CrossRef]
  5. Lin, S.; Huang, H.; Ma, T.; Zhang, Y. Photocatalytic Oxygen Evolution from Water Splitting. Adv. Sci. 2021, 8, 2002458. [Google Scholar] [CrossRef] [PubMed]
  6. Dawood, F.; Anda, M.; Shafiullah, G.M. Hydrogen production for energy: An overview. Int. J. Hydrogen Energy 2020, 45, 3847–3869. [Google Scholar] [CrossRef]
  7. Holladay, J.D.; Hu, J.; King, D.L.; Wang, Y. An overview of hydrogen production technologies. Catal. Today 2009, 139, 244–260. [Google Scholar] [CrossRef]
  8. Strmcnik, D.; Lopes, P.P.; Genorio, B.; Stamenkovic, V.R.; Markovic, N.M. Design principles for hydrogen evolution reaction catalyst materials. Nano Energy 2016, 29, 29–36. [Google Scholar] [CrossRef] [Green Version]
  9. Nikolaidis, P.; Poullikkas, A. A comparative overview of hydrogen production processes. Renew. Sustain. Energy Rev. 2017, 67, 597–611. [Google Scholar] [CrossRef]
  10. Santos, E.; Lundin, A.; Pötting, K.; Quaino, P.; Schmickler, W. Model for the electrocatalysis of hydrogen evolution. Phys. Rev. B 2009, 79, 235436. [Google Scholar] [CrossRef]
  11. Crawford, S.; Thimsen, E.; Biswas, P. Impact of Different Electrolytes on Photocatalytic Water Splitting. J. Electrochem. Soc. 2009, 156, H346. [Google Scholar] [CrossRef]
  12. Karimi Estahbanati, M.R.; Mahinpey, N.; Feilizadeh, M.; Attar, F.; Iliuta, M.C. Kinetic study of the effects of pH on the photocatalytic hydrogen production from alcohols. Int. J. Hydrogen Energy 2019, 44, 32030–32041. [Google Scholar] [CrossRef]
  13. Huang, B.; Rao, R.R.; You, S.; Hpone Myint, K.; Song, Y.; Wang, Y.; Ding, W.; Giordano, L.; Zhang, Y.; Wang, T.; et al. Cation- and pH-Dependent Hydrogen Evolution and Oxidation Reaction Kinetics. JACS Au 2021, 1, 1674–1687. [Google Scholar] [CrossRef]
  14. Zhang, G.; Zhang, W.; Minakata, D.; Chen, Y.; Crittenden, J.; Wang, P. The pH effects on H2 evolution kinetics for visible light water splitting over the Ru/(CuAg)0.15In0.3Zn1.4S2 photocatalyst. Int. J. Hydrogen Energy 2013, 38, 11727–11736. [Google Scholar] [CrossRef]
  15. Anantharaj, S.; Noda, S. Amorphous Catalysts and Electrochemical Water Splitting: An Untold Story of Harmony. Small 2020, 16, 1905779. [Google Scholar] [CrossRef]
  16. Jerkiewicz, G. Standard and Reversible Hydrogen Electrodes: Theory, Design, Operation, and Applications. ACS Catal. 2020, 10, 8409–8417. [Google Scholar] [CrossRef]
  17. Xia, Y.; Sayed, M.; Zhang, L.; Cheng, B.; Yu, J. Single-atom heterogeneous photocatalysts. Chem Catalysis 2021, 1, 1173–1214. [Google Scholar] [CrossRef]
  18. Zhao, W.; Chen, Z.; Yang, X.; Qian, X.; Liu, C.; Zhou, D.; Sun, T.; Zhang, M.; Wei, G.; Dissanayake, P.D.; et al. Recent advances in photocatalytic hydrogen evolution with high-performance catalysts without precious metals. Renew. Sustain. Energy Rev. 2020, 132, 110040. [Google Scholar] [CrossRef]
  19. Gellé, A.; Moores, A. Water splitting catalyzed by titanium dioxide decorated with plasmonic nanoparticles. Pure Appl. Chem. 2017, 89, 1817–1827. [Google Scholar] [CrossRef]
  20. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef]
  21. Xia, B.; Zhang, Y.; Ran, J.; Jaroniec, M.; Qiao, S.-Z. Single-Atom Photocatalysts for Emerging Reactions. ACS Central Science 2021, 7, 39–54. [Google Scholar] [CrossRef] [PubMed]
  22. Hasija, V.; Patial, S.; Raizada, P.; Aslam Parwaz Khan, A.; Asiri, A.M.; Van Le, Q.; Nguyen, V.-H.; Singh, P. Covalent organic frameworks promoted single metal atom catalysis: Strategies and applications. Coord. Chem. Rev. 2022, 452, 214298. [Google Scholar] [CrossRef]
  23. Wang, Q.; Li, J.; Tu, X.; Liu, H.; Shu, M.; Si, R.; Ferguson, C.T.J.; Zhang, K.A.I.; Li, R. Single Atomically Anchored Cobalt on Carbon Quantum Dots as Efficient Photocatalysts for Visible Light-Promoted Oxidation Reactions. Chem. Mater. 2020, 32, 734–743. [Google Scholar] [CrossRef] [Green Version]
  24. Li, C.; Baek, J.-B. Recent Advances in Noble Metal (Pt, Ru, and Ir)-Based Electrocatalysts for Efficient Hydrogen Evolution Reaction. ACS Omega 2020, 5, 31–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Li, X.-H.; Baar, M.; Blechert, S.; Antonietti, M. Facilitating room-temperature Suzuki coupling reaction with light: Mott-Schottky photocatalyst for C-C-coupling. Sci. Rep. 2013, 3, 1743. [Google Scholar] [CrossRef] [Green Version]
  26. Yan, F.; Wang, Y.; Zhang, J.; Lin, Z.; Zheng, J.; Huang, F. Schottky or Ohmic Metal–Semiconductor Contact: Influence on Photocatalytic Efficiency of Ag/ZnO and Pt/ZnO Model Systems. ChemSusChem 2014, 7, 101–104. [Google Scholar] [CrossRef]
  27. Cheng, C.; Fang, W.-H.; Long, R.; Prezhdo, O.V. Water Splitting with a Single-Atom Cu/TiO2 Photocatalyst: Atomistic Origin of High Efficiency and Proposed Enhancement by Spin Selection. JACS Au 2021, 1, 550–559. [Google Scholar] [CrossRef]
  28. Ge, Z.; Fu, B.; Zhao, J.; Li, X.; Ma, B.; Chen, Y. A review of the electrocatalysts on hydrogen evolution reaction with an emphasis on Fe, Co and Ni-based phosphides. J. Mater. Sci. 2020, 55, 14081–14104. [Google Scholar] [CrossRef]
  29. Wang, X.; Pan, H.; Sun, M.; Zhang, Y. Au single atom-anchored WO3/TiO2 nanotubes for the photocatalytic degradation of volatile organic compounds. J. Mater. Chem. A 2022, 10, 6078–6085. [Google Scholar] [CrossRef]
  30. Hou, T.; Peng, H.; Xin, Y.; Wang, S.; Zhu, W.; Chen, L.; Yao, Y.; Zhang, W.; Liang, S.; Wang, L. Fe Single-Atom Catalyst for Visible-Light-Driven Photofixation of Nitrogen Sensitized by Triphenylphosphine and Sodium Iodide. ACS Catal. 2020, 10, 5502–5510. [Google Scholar] [CrossRef]
  31. Xiong, X.; Mao, C.; Yang, Z.; Zhang, Q.; Waterhouse, G.I.N.; Gu, L.; Zhang, T. Photocatalytic CO2 Reduction to CO over Ni Single Atoms Supported on Defect-Rich Zirconia. Adv. Energy Mater. 2020, 10, 2002928. [Google Scholar] [CrossRef]
  32. Łosiewicz, B.; Popczyk, M.; Napłoszek, I.; Budniok, A. Intermetallic Compounds as Catalysts in the Reaction of Electroevolution/Absorption of Hydrogen. Solid State Phenom. 2015, 228, 16–22. [Google Scholar] [CrossRef]
  33. Xue, Z.-H.; Luan, D.; Zhang, H.; Lou, X.W. Single-atom catalysts for photocatalytic energy conversion. Joule 2022, 6, 92–133. [Google Scholar] [CrossRef]
  34. Gao, G.; Bottle, S.; Du, A. Understanding the activity and selectivity of single atom catalysts for hydrogen and oxygen evolution via ab initial study. Catal. Sci. Technol. 2018, 8, 996–1001. [Google Scholar] [CrossRef]
  35. Zheng, Y.; Jiao, Y.; Jaroniec, M.; Qiao, S.Z. Advancing the Electrochemistry of the Hydrogen-Evolution Reaction through Combining Experiment and Theory. Angew. Chem. Int. Ed. 2015, 54, 52–65. [Google Scholar] [CrossRef]
  36. Yang, C.; Zhao, Z.-Y.; Wei, H.-T.; Deng, X.-Y.; Liu, Q.-J. DFT calculations for single-atom confinement effects of noble metals on monolayer g-C3N4 for photocatalytic applications. RSC Adv. 2021, 11, 4276–4285. [Google Scholar] [CrossRef]
  37. Zhang, C.; Qin, D.; Zhou, Y.; Qin, F.; Wang, H.; Wang, W.; Yang, Y.; Zeng, G. Dual optimization approach to Mo single atom dispersed g-C3N4 photocatalyst: Morphology and defect evolution. Appl. Catal. B 2022, 303, 120904. [Google Scholar] [CrossRef]
  38. Song, W.; Lv, X.; Gao, Y.; Wang, L. Photocatalytic HER Performance of TiO2-supported Single Atom Catalyst Based on Electronic Regulation: A DFT Study. Chem. Res. Chin. Univ. 2022, 38, 1025–1031. [Google Scholar] [CrossRef]
  39. Hammer, B.; Norskov, J.K. Why gold is the noblest of all the metals. Nature 1995, 376, 238–240. [Google Scholar] [CrossRef]
  40. Santos, E.; Schmickler, W. d-Band Catalysis in Electrochemistry. ChemPhysChem 2006, 7, 2282–2285. [Google Scholar] [CrossRef]
  41. He, T.; Zhang, C.; Du, A. Single-atom supported on graphene grain boundary as an efficient electrocatalyst for hydrogen evolution reaction. Chem. Eng. Sci. 2019, 194, 58–63. [Google Scholar] [CrossRef]
  42. Hu, J.; Zhang, C.; Meng, X.; Lin, H.; Hu, C.; Long, X.; Yang, S. Hydrogen evolution electrocatalysis with binary-nonmetal transition metal compounds. J. Mater. Chem. A 2017, 5, 5995–6012. [Google Scholar] [CrossRef]
  43. Zheng, Y.; Jiao, Y.; Zhu, Y.; Li, L.H.; Han, Y.; Chen, Y.; Du, A.; Jaroniec, M.; Qiao, S.Z. Hydrogen evolution by a metal-free electrocatalyst. Nat. Commun. 2014, 5, 3783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Nørskov, J.K.; Bligaard, T.; Logadottir, A.; Kitchin, J.R.; Chen, J.G.; Pandelov, S.; Stimming, U. Trends in the Exchange Current for Hydrogen Evolution. J. Electrochem. Soc. 2005, 152, J23. [Google Scholar] [CrossRef] [Green Version]
  45. Jaramillo, T.F.; Jørgensen, K.P.; Bonde, J.; Nielsen, J.H.; Horch, S.; Chorkendorff, I. Identification of Active Edge Sites for Electrochemical H2 Evolution from MoS2 Nanocatalysts. Science 2007, 317, 100–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Greeley, J.; Nørskov, J.K.; Kibler, L.A.; El-Aziz, A.M.; Kolb, D.M. Hydrogen Evolution Over Bimetallic Systems: Understanding the Trends. ChemPhysChem 2006, 7, 1032–1035. [Google Scholar] [CrossRef]
  47. Zhang, Y.; Zhao, J.; Wang, H.; Xiao, B.; Zhang, W.; Zhao, X.; Lv, T.; Thangamuthu, M.; Zhang, J.; Guo, Y.; et al. Single-atom Cu anchored catalysts for photocatalytic renewable H2 production with a quantum efficiency of 56%. Nat. Commun. 2022, 13, 58. [Google Scholar] [CrossRef]
  48. Cha, G.; Hwang, I.; Hejazi, S.; Dobrota, A.S.; Pašti, I.A.; Osuagwu, B.; Kim, H.; Will, J.; Yokosawa, T.; Badura, Z.; et al. As a single atom Pd outperforms Pt as the most active co-catalyst for photocatalytic H2 evolution. iScience 2021, 24, 102938. [Google Scholar] [CrossRef]
  49. Pan, J.; Zhang, G.; Guan, Z.; Zhao, Q.; Li, G.; Yang, J.; Li, Q.; Zou, Z. Anchoring Ni single atoms on sulfur-vacancy-enriched ZnIn2S4 nanosheets for boosting photocatalytic hydrogen evolution. J. Energy Chem. 2021, 58, 408–414. [Google Scholar] [CrossRef]
  50. Zhou, K.L.; Wang, Z.; Han, C.B.; Ke, X.; Wang, C.; Jin, Y.; Zhang, Q.; Liu, J.; Wang, H.; Yan, H. Platinum single-atom catalyst coupled with transition metal/metal oxide heterostructure for accelerating alkaline hydrogen evolution reaction. Nat. Commun. 2021, 12, 3783. [Google Scholar] [CrossRef]
  51. Vilé, G.; Sharma, P.; Nachtegaal, M.; Tollini, F.; Moscatelli, D.; Sroka-Bartnicka, A.; Tomanec, O.; Petr, M.; Filip, J.; Pieta, I.S.; et al. An Earth-Abundant Ni-Based Single-Atom Catalyst for Selective Photodegradation of Pollutants. Solar RRL 2021, 5, 2100176. [Google Scholar] [CrossRef]
  52. Zhang, P.; Zhan, X.; Xu, L.; Fu, X.; Zheng, T.; Yang, X.; Xu, Q.; Wang, D.; Qi, D.; Sun, T.; et al. Mass production of a single-atom cobalt photocatalyst for high-performance visible-light photocatalytic CO2 reduction. J. Mater. Chem. A 2021, 9, 26286–26297. [Google Scholar] [CrossRef]
  53. Zhou, X.; Hwang, I.; Tomanec, O.; Fehn, D.; Mazare, A.; Zboril, R.; Meyer, K.; Schmuki, P. Advanced Photocatalysts: Pinning Single Atom Co-Catalysts on Titania Nanotubes. Adv. Funct. Mater. 2021, 31, 2102843. [Google Scholar] [CrossRef]
  54. Yang, J.; Sun, Z.; Yan, K.; Dong, H.; Dong, H.; Cui, J.; Gong, X.; Han, S.; Huang, L.; Wen, J. Single-atom-nickel photocatalytic site-selective sulfonation of enamides to access amidosulfones. Green Chem. 2021, 23, 2756–2762. [Google Scholar] [CrossRef]
  55. Wang, K.; Jiang, L.; Xin, T.; Li, Y.; Wu, X.; Zhang, G. Single-atom V-N charge-transfer bridge on ultrathin carbon nitride for efficient photocatalytic H2 production and formaldehyde oxidation under visible light. Chem. Eng. J. 2022, 429, 132229. [Google Scholar] [CrossRef]
  56. Sun, S.; Zhang, G.; Gauquelin, N.; Chen, N.; Zhou, J.; Yang, S.; Chen, W.; Meng, X.; Geng, D.; Banis, M.N.; et al. Single-atom Catalysis Using Pt/Graphene Achieved through Atomic Layer Deposition. Sci. Rep. 2013, 3, 1775. [Google Scholar] [CrossRef] [Green Version]
  57. Liang, J.; Song, Q.; Wu, J.; Lei, Q.; Li, J.; Zhang, W.; Huang, Z.; Kang, T.; Xu, H.; Wang, P.; et al. Anchoring Copper Single Atoms on Porous Boron Nitride Nanofiber to Boost Selective Reduction of Nitroaromatics. ACS Nano 2022, 16, 4152–4161. [Google Scholar] [CrossRef]
  58. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef]
  59. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef]
  60. Song, T.; Paik, U. TiO2 as an active or supplemental material for lithium batteries. J. Mater. Chem. A 2016, 4, 14–31. [Google Scholar] [CrossRef]
  61. Cho, H.; Joo, H.; Kim, H.; Kim, J.-E.; Kang, K.-S.; Jung, H.; Yoon, J. Enhanced Photoelectrochemical Activity of TiO2 Nanotubes Decorated with Lanthanide Ions for Hydrogen Production. Catalysts 2022, 12, 866. [Google Scholar] [CrossRef]
  62. Son, Y.J.; Kang, J.S.; Yoon, J.; Kim, J.; Jeong, J.; Kang, J.; Lee, M.J.; Park, H.S.; Sung, Y.-E. Influence of TiO2 Particle Size on Dye-Sensitized Solar Cells Employing an Organic Sensitizer and a Cobalt(III/II) Redox Electrolyte. J. Phys. Chem. C 2018, 122, 7051–7060. [Google Scholar] [CrossRef]
  63. Nam, Y.; Lim, J.H.; Ko, K.C.; Lee, J.Y. Photocatalytic activity of TiO2 nanoparticles: A theoretical aspect. J. Mater. Chem. A 2019, 7, 13833–13859. [Google Scholar] [CrossRef]
  64. Shayegan, Z.; Lee, C.-S.; Haghighat, F. TiO2 photocatalyst for removal of volatile organic compounds in gas phase – A review. Chem. Eng. J. 2018, 334, 2408–2439. [Google Scholar] [CrossRef] [Green Version]
  65. Yi, L.; Lan, F.; Li, J.; Zhao, C. Efficient Noble-Metal-Free Co-NG/TiO2 Photocatalyst for H2 Evolution: Synergistic Effect between Single-Atom Co and N-Doped Graphene for Enhanced Photocatalytic Activity. ACS Sustain. Chem. Eng. 2018, 6, 12766–12775. [Google Scholar] [CrossRef]
  66. Cui, J.; Liang, S.; Zhang, J. A multifunctional material of two-dimensional g-C4N3/graphene bilayer. Phys. Chem. Chem. Phys. 2016, 18, 25388–25393. [Google Scholar] [CrossRef]
  67. Chen, J.; Fang, S.; Shen, Q.; Fan, J.; Li, Q.; Lv, K. Recent Advances of Doping and Surface Modifying Carbon Nitride with Characterization Techniques. Catalysts 2022, 12, 962. [Google Scholar] [CrossRef]
  68. Zuluaga, S.; Liu, L.H.; Shafiq, N.; Rupich, S.M.; Veyan, J.F.; Chabal, Y.J.; Thonhauser, T. Structural band-gap tuning in g-C3N4. Phys. Chem. Chem. Phys. 2015, 17, 957–962. [Google Scholar] [CrossRef] [Green Version]
  69. Hong, Y.; Liu, E.; Shi, J.; Lin, X.; Sheng, L.; Zhang, M.; Wang, L.; Chen, J. A direct one-step synthesis of ultrathin g-C3N4 nanosheets from thiourea for boosting solar photocatalytic H2 evolution. Int. J. Hydrogen Energy 2019, 44, 7194–7204. [Google Scholar] [CrossRef]
  70. Antil, B.; Kumar, L.; Ranjan, R.; Shenoy, S.; Tarafder, K.; Gopinath, C.S.; Deka, S. One-Dimensional Multichannel g-C3N4.7 Nanostructure Realizing an Efficient Photocatalytic Hydrogen Evolution Reaction and Its Theoretical Investigations. ACS Appl. Energy Mater. 2021, 4, 3118–3129. [Google Scholar] [CrossRef]
  71. Majdoub, M.; Anfar, Z.; Amedlous, A. Emerging Chemical Functionalization of g-C3N4: Covalent/Noncovalent Modifications and Applications. ACS Nano 2020, 14, 12390–12469. [Google Scholar] [CrossRef]
  72. Wang, X.; Chen, X.; Thomas, A.; Fu, X.; Antonietti, M. Metal-Containing Carbon Nitride Compounds: A New Functional Organic–Metal Hybrid Material. Adv. Mater. 2009, 21, 1609–1612. [Google Scholar] [CrossRef]
  73. Thomas, A.; Fischer, A.; Goettmann, F.; Antonietti, M.; Müller, J.-O.; Schlögl, R.; Carlsson, J.M. Graphitic carbon nitride materials: Variation of structure and morphology and their use as metal-free catalysts. J. Mater. Chem. 2008, 18, 4893–4908. [Google Scholar] [CrossRef] [Green Version]
  74. Cao, Y.; Wang, D.; Lin, Y.; Liu, W.; Cao, L.; Liu, X.; Zhang, W.; Mou, X.; Fang, S.; Shen, X.; et al. Single Pt Atom with Highly Vacant d-Orbital for Accelerating Photocatalytic H2 Evolution. ACS Appl. Energy Mater. 2018, 1, 6082–6088. [Google Scholar] [CrossRef]
  75. Li, W.; Li, W.; Guo, Z.; Song, Y.; Tang, S.; Ma, Y.; Xing, X.; Wang, Q. Synthesis of Atomically Thin g-C3N4 Nanosheets via Supercritical CO2 Doping with Single-Atom Cobalt for Photocatalytic Hydrogen Evolution. ACS Appl. Mater. Interfaces 2021, 13, 52560–52570. [Google Scholar] [CrossRef]
  76. Serpone, N.; Emeline, A.V. Semiconductor Photocatalysis — Past, Present, and Future Outlook. J. Phys. Chem. Lett. 2012, 3, 673–677. [Google Scholar] [CrossRef]
  77. Li, W.; Lin, Z.; Yang, G. A 2D self-assembled MoS2/ZnIn2S4 heterostructure for efficient photocatalytic hydrogen evolution. Nanoscale 2017, 9, 18290–18298. [Google Scholar] [CrossRef]
  78. Jiao, X.; Chen, Z.; Li, X.; Sun, Y.; Gao, S.; Yan, W.; Wang, C.; Zhang, Q.; Lin, Y.; Luo, Y.; et al. Defect-Mediated Electron-Hole Separation in One-Unit-Cell ZnIn2S4 Layers for Boosted Solar-Driven CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 7586–7594. [Google Scholar] [CrossRef]
  79. Yang, W.; Zhang, L.; Xie, J.; Zhang, X.; Liu, Q.; Yao, T.; Wei, S.; Zhang, Q.; Xie, Y. Enhanced Photoexcited Carrier Separation in Oxygen-Doped ZnIn2S4 Nanosheets for Hydrogen Evolution. Angew. Chem. Int. Ed. 2016, 55, 6716–6720. [Google Scholar] [CrossRef]
  80. Shi, X.; Dai, C.; Wang, X.; Hu, J.; Zhang, J.; Zheng, L.; Mao, L.; Zheng, H.; Zhu, M. Protruding Pt single-sites on hexagonal ZnIn2S4 to accelerate photocatalytic hydrogen evolution. Nat. Commun. 2022, 13, 1287. [Google Scholar] [CrossRef]
  81. Wu, Y.; Wu, Q.; Zhang, Q.; Lou, Z.; Liu, K.; Ma, Y.; Wang, Z.; Zheng, Z.; Cheng, H.; Liu, Y.; et al. An organometal halide perovskite supported Pt single-atom photocatalyst for H2 evolution. Energy Environ. Sci. 2022, 15, 1271–1281. [Google Scholar] [CrossRef]
  82. Dong, P.; Wang, Y.; Zhang, A.; Cheng, T.; Xi, X.; Zhang, J. Platinum Single Atoms Anchored on a Covalent Organic Framework: Boosting Active Sites for Photocatalytic Hydrogen Evolution. ACS Catal. 2021, 11, 13266–13279. [Google Scholar] [CrossRef]
  83. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef] [PubMed]
  84. Chen, G.; Wang, P.; Wu, Y.; Zhang, Q.; Wu, Q.; Wang, Z.; Zheng, Z.; Liu, Y.; Dai, Y.; Huang, B. Lead-Free Halide Perovskite Cs3Bi2xSb2–2xI9 (x ≈ 0.3) Possessing the Photocatalytic Activity for Hydrogen Evolution Comparable to that of (CH3NH3)PbI3. Adv. Mater. 2020, 32, 2001344. [Google Scholar] [CrossRef] [PubMed]
  85. Wang, X.; Wang, H.; Zhang, H.; Yu, W.; Wang, X.; Zhao, Y.; Zong, X.; Li, C. Dynamic Interaction between Methylammonium Lead Iodide and TiO2 Nanocrystals Leads to Enhanced Photocatalytic H2 Evolution from HI Splitting. ACS Energy Lett. 2018, 3, 1159–1164. [Google Scholar] [CrossRef]
  86. Wang, H.; Wang, X.; Chen, R.; Zhang, H.; Wang, X.; Wang, J.; Zhang, J.; Mu, L.; Wu, K.; Fan, F.; et al. Promoting Photocatalytic H2 Evolution on Organic–Inorganic Hybrid Perovskite Nanocrystals by Simultaneous Dual-Charge Transportation Modulation. ACS Energy Lett. 2019, 4, 40–47. [Google Scholar] [CrossRef]
  87. Guo, Y.; Liu, G.; Li, Z.; Lou, Y.; Chen, J.; Zhao, Y. Stable Lead-Free (CH3NH3)3Bi2I9 Perovskite for Photocatalytic Hydrogen Generation. ACS Sustain. Chem. Eng. 2019, 7, 15080–15085. [Google Scholar] [CrossRef]
  88. Romani, L.; Malavasi, L. Solar-Driven Hydrogen Generation by Metal Halide Perovskites: Materials, Approaches, and Mechanistic View. ACS Omega 2020, 5, 25511–25519. [Google Scholar] [CrossRef]
  89. Ricciarelli, D.; Kaiser, W.; Mosconi, E.; Wiktor, J.; Ashraf, M.W.; Malavasi, L.; Ambrosio, F.; De Angelis, F. Reaction Mechanism of Photocatalytic Hydrogen Production at Water/Tin Halide Perovskite Interfaces. ACS Energy Lett. 2022, 7, 1308–1315. [Google Scholar] [CrossRef]
  90. Yu, J.; Xu, X. Expediting H2 Evolution over MAPbI3 with a Nonnoble Metal Cocatalyst Mo2C under Visible Light. Energy Mater. Adv. 2022, 2022, 9836095. [Google Scholar] [CrossRef]
  91. Zhou, P.; Chen, H.; Chao, Y.; Zhang, Q.; Zhang, W.; Lv, F.; Gu, L.; Zhao, Q.; Wang, N.; Wang, J.; et al. Single-atom Pt-I3 sites on all-inorganic Cs2SnI6 perovskite for efficient photocatalytic hydrogen production. Nat. Commun. 2021, 12, 4412. [Google Scholar] [CrossRef]
  92. Ding, S.Y.; Gao, J.; Wang, Q.; Zhang, Y.; Song, W.G.; Su, C.Y.; Wang, W. Construction of covalent organic framework for catalysis: Pd/COF-LZU1 in Suzuki-Miyaura coupling reaction. J. Am. Chem. Soc. 2011, 133, 19816–19822. [Google Scholar] [CrossRef]
  93. Cui, K.; Zhong, W.; Li, L.; Zhuang, Z.; Li, L.; Bi, J.; Yu, Y. Well-Defined Metal Nanoparticles@Covalent Organic Framework Yolk–Shell Nanocages by ZIF-8 Template as Catalytic Nanoreactors. Small 2019, 15, 1804419. [Google Scholar] [CrossRef]
  94. Lu, S.; Hu, Y.; Wan, S.; McCaffrey, R.; Jin, Y.; Gu, H.; Zhang, W. Synthesis of Ultrafine and Highly Dispersed Metal Nanoparticles Confined in a Thioether-Containing Covalent Organic Framework and Their Catalytic Applications. J. Am. Chem. Soc. 2017, 139, 17082–17088. [Google Scholar] [CrossRef]
  95. Zhong, W.; Sa, R.; Li, L.; He, Y.; Li, L.; Bi, J.; Zhuang, Z.; Yu, Y.; Zou, Z. A Covalent Organic Framework Bearing Single Ni Sites as a Synergistic Photocatalyst for Selective Photoreduction of CO2 to CO. J. Am. Chem. Soc. 2019, 141, 7615–7621. [Google Scholar] [CrossRef]
  96. Pachfule, P.; Acharjya, A.; Roeser, J.; Langenhahn, T.; Schwarze, M.; Schomacker, R.; Thomas, A.; Schmidt, J. Diacetylene Functionalized Covalent Organic Framework (COF) for Photocatalytic Hydrogen Generation. J. Am. Chem. Soc. 2018, 140, 1423–1427. [Google Scholar] [CrossRef]
  97. Stegbauer, L.; Schwinghammer, K.; Lotsch, B.V. A hydrazone-based covalent organic framework for photocatalytic hydrogen production. Chem. Sci. 2014, 5, 2789–2793. [Google Scholar] [CrossRef] [Green Version]
  98. Chai, S.; Chen, X.; Zhang, X.; Fang, Y.; Sprick, R.S.; Chen, X. Rational design of covalent organic frameworks for efficient photocatalytic hydrogen peroxide production. Environ. Sci. Nano 2022, 9, 2464–2469. [Google Scholar] [CrossRef]
  99. Li, C.; Liu, J.; Li, H.; Wu, K.; Wang, J.; Yang, Q. Covalent organic frameworks with high quantum efficiency in sacrificial photocatalytic hydrogen evolution. Nat. Commun. 2022, 13, 2357. [Google Scholar] [CrossRef]
  100. Li, X.; Kawai, K.; Fujitsuka, M.; Osakada, Y. COF-based photocatalyst for energy and environment applications. Surf. Interfaces 2021, 25, 101249. [Google Scholar] [CrossRef]
  101. Wang, S.; Zhang, J.-J.; Zong, M.-Y.; Xu, J.; Wang, D.-H.; Bu, X.-H. Energy Level Engineering: Ru Single Atom Anchored on Mo-MOF with a [Mo8O26(im)2]4– Structure Acts as a Biomimetic Photocatalyst. ACS Catal. 2022, 12, 7960–7974. [Google Scholar] [CrossRef]
  102. Ma, X.; Yong, X.; Jian, C.-c.; Zhang, J. Transition Metal-Functionalized Janus MoSSe Monolayer: A Magnetic and Efficient Single-Atom Photocatalyst for Water-Splitting Applications. J. Phys. Chem. C 2019, 123, 18347–18354. [Google Scholar] [CrossRef]
  103. Han, A.; Zhou, X.; Wang, X.; Liu, S.; Xiong, Q.; Zhang, Q.; Gu, L.; Zhuang, Z.; Zhang, W.; Li, F.; et al. One-step synthesis of single-site vanadium substitution in 1T-WS2 monolayers for enhanced hydrogen evolution catalysis. Nat. Commun. 2021, 12, 709. [Google Scholar] [CrossRef] [PubMed]
  104. Wang, X.; Zhang, Y.; Si, H.; Zhang, Q.; Wu, J.; Gao, L.; Wei, X.; Sun, Y.; Liao, Q.; Zhang, Z.; et al. Single-Atom Vacancy Defect to Trigger High-Efficiency Hydrogen Evolution of MoS2. J. Am. Chem. Soc. 2020, 142, 4298–4308. [Google Scholar] [CrossRef] [PubMed]
  105. Jiang, K.; Liu, B.; Luo, M.; Ning, S.; Peng, M.; Zhao, Y.; Lu, Y.-R.; Chan, T.-S.; de Groot, F.M.F.; Tan, Y. Single platinum atoms embedded in nanoporous cobalt selenide as electrocatalyst for accelerating hydrogen evolution reaction. Nat. Commun. 2019, 10, 1743. [Google Scholar] [CrossRef] [Green Version]
  106. Luo, Y.; Zhang, S.; Pan, H.; Xiao, S.; Guo, Z.; Tang, L.; Khan, U.; Ding, B.-F.; Li, M.; Cai, Z.; et al. Unsaturated Single Atoms on Monolayer Transition Metal Dichalcogenides for Ultrafast Hydrogen Evolution. ACS Nano 2020, 14, 767–776. [Google Scholar] [CrossRef]
  107. Lv, S.; Pei, M.; Liu, Y.; Si, Z.; Wu, X.; Ran, R.; Weng, D.; Kang, F. An isolation strategy to anchor atomic Ni or Co cocatalysts on TiO2(A) for photocatalytic hydrogen production. Nano Res. 2022, 15, 5848–5856. [Google Scholar] [CrossRef]
  108. Liu, G.; Lv, H.; Zeng, Y.; Yuan, M.; Meng, Q.; Wang, Y.; Wang, C. Single-Atom Pd–N3 Sites on Carbon-Deficient g-C3N4 for Photocatalytic H2 Evolution. Trans. Tianjin Univ. 2021, 27, 139–146. [Google Scholar] [CrossRef]
  109. Guan, Z.; Pan, J.; Li, Q.; Li, G.; Yang, J. Boosting Visible-Light Photocatalytic Hydrogen Evolution with an Efficient CuInS2/ZnIn2S4 2D/2D Heterojunction. ACS Sustain. Chem. Eng. 2019, 7, 7736–7742. [Google Scholar] [CrossRef]
  110. Ding, C.; Shi, J.; Wang, Z.; Li, C. Photoelectrocatalytic Water Splitting: Significance of Cocatalysts, Electrolyte, and Interfaces. ACS Catal. 2017, 7, 675–688. [Google Scholar] [CrossRef]
  111. Li, D.; Liu, Y.; Shi, W.; Shao, C.; Wang, S.; Ding, C.; Liu, T.; Fan, F.; Shi, J.; Li, C. Crystallographic-Orientation-Dependent Charge Separation of BiVO4 for Solar Water Oxidation. ACS Energy Lett. 2019, 4, 825–831. [Google Scholar] [CrossRef]
  112. Sun, H.; Hua, W.; Li, Y.; Wang, J.-G. Conformal coating of superhydrophilic metal-organic complex toward substantially improved photoelectrochemical water oxidation. Chem. Eng. J. 2022, 427, 131004. [Google Scholar] [CrossRef]
  113. Sun, H.; Hua, W.; Liang, S.; Li, Y.; Wang, J.-G. A self-adaptive semiconductor–liquid junction for highly active and stable solar water splitting. J. Mater. Chem. A 2022, 10, 20414–20423. [Google Scholar] [CrossRef]
  114. Lee, W.H.; Ko, Y.-J.; Kim, J.-Y.; Min, B.K.; Hwang, Y.J.; Oh, H.-S. Single-atom catalysts for the oxygen evolution reaction: Recent developments and future perspectives. Chem. Commun. 2020, 56, 12687–12697. [Google Scholar] [CrossRef]
  115. Yin, J.; Jin, J.; Lu, M.; Huang, B.; Zhang, H.; Peng, Y.; Xi, P.; Yan, C.H. Iridium Single Atoms Coupling with Oxygen Vacancies Boosts Oxygen Evolution Reaction in Acid Media. J. Am. Chem. Soc. 2020, 142, 18378–18386. [Google Scholar] [CrossRef]
Figure 1. Solar-driven hydrogen production and hydrogen applications.
Figure 1. Solar-driven hydrogen production and hydrogen applications.
Catalysts 12 01304 g001
Figure 2. (a) The structure of single transition metal atom (SMA) photocatalyst and scheme of SMA photocatalyst in hydrogen evolution reaction. (b) Gibbs free energy for hydrogen evolution reaction (HER) at different coordination sites of Ni. Reproduced with permission from ref. [34] with permission from the Royal Society of Chemistry. (c) Dependence of exchange current density (j0) on the Δ G H *   of the surface of various metals. Reproduced with permission from ref. [35] Copyright 2015, Wiley-VCH.
Figure 2. (a) The structure of single transition metal atom (SMA) photocatalyst and scheme of SMA photocatalyst in hydrogen evolution reaction. (b) Gibbs free energy for hydrogen evolution reaction (HER) at different coordination sites of Ni. Reproduced with permission from ref. [34] with permission from the Royal Society of Chemistry. (c) Dependence of exchange current density (j0) on the Δ G H *   of the surface of various metals. Reproduced with permission from ref. [35] Copyright 2015, Wiley-VCH.
Catalysts 12 01304 g002
Figure 3. Synthesis of the single-metal atom photocatalysts.
Figure 3. Synthesis of the single-metal atom photocatalysts.
Catalysts 12 01304 g003
Figure 4. (a) Schematic for photocatalytic reaction of TiO2 nanobelts on N-graphene, (b) H2 evolution rate of TiO2 nanobelts and composites. Reprinted with permission from ref. [65] Copyright 2018, American Chemical Society; (c) illustration of noble metal on anatase (001) TiO2 nanosheets and (d) photocatalytic H2 evolution rate of noble metal on TiO2 nanosheets. Reproduced with permission from ref. [48] Copyright 2021, Elsevier.
Figure 4. (a) Schematic for photocatalytic reaction of TiO2 nanobelts on N-graphene, (b) H2 evolution rate of TiO2 nanobelts and composites. Reprinted with permission from ref. [65] Copyright 2018, American Chemical Society; (c) illustration of noble metal on anatase (001) TiO2 nanosheets and (d) photocatalytic H2 evolution rate of noble metal on TiO2 nanosheets. Reproduced with permission from ref. [48] Copyright 2021, Elsevier.
Catalysts 12 01304 g004
Figure 5. (a) The photocatalytic H2 evolution rate of bare TiO2 and SMA TiO2 with different metals; (b) H2 production of bare TiO2 and SMA TiO2 with different amount of Cu; (c) with 1.5 wt.% Cu SA-TiO2 for 380 days; (d) the photocatalytic H2 evolution mechanism on Cu SA-TiO2; and (e) the corresponding schematic representation of the formation of copper SMA in the lattice of TiO2, corresponds to their electron images. Reproduced with permission from ref. [47] Copyright 2022, Springer Nature.
Figure 5. (a) The photocatalytic H2 evolution rate of bare TiO2 and SMA TiO2 with different metals; (b) H2 production of bare TiO2 and SMA TiO2 with different amount of Cu; (c) with 1.5 wt.% Cu SA-TiO2 for 380 days; (d) the photocatalytic H2 evolution mechanism on Cu SA-TiO2; and (e) the corresponding schematic representation of the formation of copper SMA in the lattice of TiO2, corresponds to their electron images. Reproduced with permission from ref. [47] Copyright 2022, Springer Nature.
Catalysts 12 01304 g005
Figure 6. (a) Illustration of the photocatalytic mechanism on Pt SMA on g-C3N4; (b) H2 evolution of Pt SMA on g-C3N4 compared with Pt nanoparticle and bare C3N4 samples. Reprinted with permission from ref. [74] Copyright 2018, American Chemical Society. (c) Electronic band structures and (d) H2 evolution of the as-prepared g-C3N4 and SMA g-C3N4 samples. Reproduced with permission from ref. [75] Copyright 2021, American Chemical Society. (e) Spherical aberration-corrected HAADF-STEM image of the SMA on ultrathin C3N4 (SAVCN) and (f) hydrogen production rate under blue/green LED of polymeric C3N4 (PCN) and SAVCN samples. Reproduced with permission from ref. [55] Copyright 2022, Elsevier.
Figure 6. (a) Illustration of the photocatalytic mechanism on Pt SMA on g-C3N4; (b) H2 evolution of Pt SMA on g-C3N4 compared with Pt nanoparticle and bare C3N4 samples. Reprinted with permission from ref. [74] Copyright 2018, American Chemical Society. (c) Electronic band structures and (d) H2 evolution of the as-prepared g-C3N4 and SMA g-C3N4 samples. Reproduced with permission from ref. [75] Copyright 2021, American Chemical Society. (e) Spherical aberration-corrected HAADF-STEM image of the SMA on ultrathin C3N4 (SAVCN) and (f) hydrogen production rate under blue/green LED of polymeric C3N4 (PCN) and SAVCN samples. Reproduced with permission from ref. [55] Copyright 2022, Elsevier.
Catalysts 12 01304 g006
Figure 7. (a) H2 evolution rates of ZnIn2S4 with different weight ratios of Ni co-catalyst and (b) HRTEM image showing Ni single atom (0.9%) on ZnIn2S4. Reproduced with permission from ref. [49] Copyright 2021, Elsevier. (c) Calculated Gibbs free energy of H* adsorption for Pt, FA perovskite with and without Pt SMA. Reproduced with permission from ref. [81] Copyright 2022, Royal Society of Chemistry. (d) Calculated Gibbs free energy of H* adsorption with an illustration of TpPa-1-COF; and (e) H2 evolution rate of TpPa-1-COF with different amounts of Pt SMA. Reproduced with permission from ref. [82] Copyright 2021, American Chemical Society.
Figure 7. (a) H2 evolution rates of ZnIn2S4 with different weight ratios of Ni co-catalyst and (b) HRTEM image showing Ni single atom (0.9%) on ZnIn2S4. Reproduced with permission from ref. [49] Copyright 2021, Elsevier. (c) Calculated Gibbs free energy of H* adsorption for Pt, FA perovskite with and without Pt SMA. Reproduced with permission from ref. [81] Copyright 2022, Royal Society of Chemistry. (d) Calculated Gibbs free energy of H* adsorption with an illustration of TpPa-1-COF; and (e) H2 evolution rate of TpPa-1-COF with different amounts of Pt SMA. Reproduced with permission from ref. [82] Copyright 2021, American Chemical Society.
Catalysts 12 01304 g007
Table 1. Summarization of synthesis approaches for SMA photocatalyst.
Table 1. Summarization of synthesis approaches for SMA photocatalyst.
Single-Atom MetalHost MaterialsSynthesis MethodRef
Cu, Co, Ni, Fe, Mn, Zn, PtTiO2Absorption on MOF and calcination[47]
NiTiO2Refining by hydrothermal[54]
NiZnIn2S4Deposition: electrostatic deposition + hydrothermal[49]
Pd, Pt, AuTiO2Absorption: Immersing in salts[48]
NiC3N4Refining by sodium borohydride[51]
CoN-CarbonRefining by calcination[52]
PtNi/NiO on Ag NWsElectrodeposition[50]
VC3N4Refining by calcination[55]
IrTiO2Deposition in dark[53]
PtGrapheneDeposition: Atomic layer deposition[56]
CuBNRefining by calcination[57]
Table 2. Comparison of various kinds of catalyst with/without single transition metal atom for hydrogen evolution reaction.
Table 2. Comparison of various kinds of catalyst with/without single transition metal atom for hydrogen evolution reaction.
Metal AtomHost MaterialsPhotocurrent (μA cm–2)H2 Evolution Rate (mmol g−1·h−1)Reference
CoTiO2 nanobelt~83~0.677[65]
-~28~0.0217
CoTiO2 nanosheets (NSs)n/a~2.9[107]
Nin/a~1.0
-n/a~0.04
CuTiO2 (derived from MIL125-Ti)~3.0~101.7 [47]
Ptn/a~95.0
-~1.0~4.2
Ptg-C3N4 NSsn/a~0.042[74]
-n/a~0.001
CoP-doped C3N4 ultrathin NSs6.0~3.7[75]
-4.0~0.4
PdCarbon deficient g-C3N4n/a~2.8[108]
-n/a~0.115
VPolymeric C3N44.6~5.0[55]
-1.7~1.5
NiSulfur-vacancy-enrich ZnI2S4~22.5~0.089[49]
-~10.0~0.04
PtZnI2S4 NSs~17.00.35[80]
-~4.00.02
-CuInS2/ZnIn2S4~0.75~0.34[109]
PtFAPbBr3−xIx~12.0~0.7[81]
-~3.0~0.05
-Cs3Bi0.6Sb1.4I9~20n/a[84]
PtTpPa COFn/a~0.72[82]
-n/a0.015
-TP-BDDA COFn/a0.32[96]
-TP-DTP COFn/a0.03
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nguyen, T.P.; Kim, I.T. Single-Atom Transition Metal Photocatalysts for Hydrogen Evolution Reactions. Catalysts 2022, 12, 1304. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12111304

AMA Style

Nguyen TP, Kim IT. Single-Atom Transition Metal Photocatalysts for Hydrogen Evolution Reactions. Catalysts. 2022; 12(11):1304. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12111304

Chicago/Turabian Style

Nguyen, Thang Phan, and Il Tae Kim. 2022. "Single-Atom Transition Metal Photocatalysts for Hydrogen Evolution Reactions" Catalysts 12, no. 11: 1304. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12111304

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