Next Article in Journal
Modified Mn/ZSM-5 for Non-Thermal Plasma Mineralization of VOCs and DFT Simulation Using a Novel Y-Type ZSM-5 Model
Next Article in Special Issue
Well-Defined Ultrasmall V-NiP2 Nanoparticles Anchored g-C3N4 Nanosheets as Highly Efficient Visible-Light-Driven Photocatalysts for H2 Evolution
Previous Article in Journal
Chitin and Silk Fibroin Biopolymers Modified by Oxone: Efficient Heterogeneous Catalysts for Knoevenagel Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Single-Atom-Based Catalysts for Photocatalytic Water Splitting on TiO2 Nanostructures

1
Department of Chemistry and Biology, Chemistry and Structure of Novel Materials, University of Siegen, Paul-Bonatz-Str. 9-11, 57076 Siegen, Germany
2
Department of Materials Science and Engineering, Institute for Surface Science and Corrosion, Friedrich-Alexander-Universität ErlangenNürnberg, 91058 Erlangen, Germany
3
Advanced Institute for Materials Research (AIMR), National University Corporation Tohoku University (TU), 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(8), 905; https://doi.org/10.3390/catal12080905
Submission received: 1 August 2022 / Revised: 11 August 2022 / Accepted: 12 August 2022 / Published: 17 August 2022

Abstract

:
H2 generation from photocatalytic water splitting is one of the most promising approaches to producing cost-effective and sustainable fuel. Nanostructured TiO2 is a highly stable and efficient semiconductor photocatalyst for this purpose. The main drawback of TiO2 as a photocatalyst is the sluggish charge transfer on the surface of TiO2 that can be tackled to a great extent by the use of platinum group materials (PGM) as co-catalysts. However, the scarcity and high cost of the PGMs is one of the issues that prevent the widespread use of TiO2/PGM systems for photocatalytic H2 generation. Single-atom catalysts which are currently the frontline in the catalysis field can be a favorable path to overcome the scarcity and further advance the use of noble metals. More importantly, single-atom (SA) catalysts simultaneously have the advantage of homogenous and heterogeneous catalysts. This mini-review specifically focuses on the single atom decoration of TiO2 nanostructures for photocatalytic water splitting. The latest progress in fabrication, characterization, and application of single-atoms in photocatalytic H2 generation on TiO2 is reviewed.

1. Introduction

Photocatalytic water splitting is the process of H2 and O2 production from water under light irradiation and in the presence of a catalyst, and follows the Equations (1)–(3) listed below.
2H+ + 2e → H2
2H2O + 2h+ → ½ O2 + 2H+
H2O → H2 + ½ O2
Basically, in the photocatalytic pathway, photon energy (from solar energy) converts to chemical energy (hydrogen) in the presence of photocatalysts. The ground-breaking work of Fujishima and Honda [1] demonstrated the feasibility of photocatalytic hydrogen generation from water, using Pt/TiO2 electrodes.
Currently, hydrogen is produced mainly from fossil fuels with a rate of ~50 billion kg per year [2]. Thus, solar water splitting can become a sustainable and renewable hydrogen production path, with the potential to mitigate the energy crisis.
Semiconductors are among the most suitable materials as photocatalysts for H2 production from water splitting or photo-reforming of H2O/alcohol mixtures, e.g., using methanol or ethanol (without or with the presence of co-catalysts). According to Takanabe [3], the process of water splitting on a semiconductor involves six major components that occur sequentially (see Figure 1), which schematically summarizes the contribution of all the steps influencing the overall water splitting. In the first step, upon irradiation, electron-hole pairs are generated. This occurs when the photon energy is larger than the bandgap of the semiconductor. The second step is charge separation, followed by the third step of diffusion of the photogenerated charge carriers. Carrier diffusion (the third step) depends mainly on carrier lifetime, mobility, and diffusion length. Electrons and holes recombination is almost zero in an ideal photocatalyst with maximum efficiency. The fourth step is carrier transportation and is confined by the following characteristics of the system, conductivity, space charge layer, and flat band potential. The fifth major step is the catalyst′s efficiency in terms of exchange current density, Tafel slope, electrocatalytic activity, activation energy, and charge transfer resistance. In the last step, mass transfer plays an important role and is under the control of mainly the iR drop in the solution, pH gradient, viscosity, and effective size of the ions in the electrolyte.
There are excellent review articles with respect to photocatalyst and photoelectrochemical water splitting, for example, refs. [4,5,6,7,8,9,10]. Of the multitude of investigated semiconductor materials, TiO2 remains one of the most investigated photocatalysts as it provides excellent photo corrosion, non-toxicity, a low cost, and, most importantly, the conduction and valence band edges [11,12]. However, the solar to hydrogen (STH) conversion efficiency is still very low for TiO2, and the insufficient energy conversion efficiency in TiO2 systems is mainly ascribed to the following factors: (i) the fast recombination of photogenerated charge carriers (electron-hole pairs), the recombination of electrons on the conduction band with the holes in the valence band causes unwanted heat or photon generation; (ii) the backward reaction, the reaction of H2 and O2 to form water, which occurs very fast; and (iii) the wide band gap of TiO2 (3.2 eV) that allows only UV light absorption for H2 production. Since only about 4% of the solar radiation energy is UV light the STH efficiency remains very low with pristine TiO2 structures.
In general, rutile TiO2 is the most stable phase among the TiO2 polymorphs, compared to anatase and brookite, and has a smaller band gap. However, anatase is the preferred polymorph in photocatalytic applications due to a higher conduction band energy and a lower electron-hole recombination rate [12,13]. Different shapes of TiO2 nanostructures have been synthesized to modify the electronic and optical properties, e.g., nanoparticles, nanobelts, nanosheets, nanoflakes, nanotubes [11]. From the different methods, the electrochemical anodization of Ti foil is a straightforward method to produce 1D nanostructures, and hydrothermal techniques as well provide scalable and convenient methods to produce 1D and 2D nanostructures. Additionally, designing anatase nanocrystals with 001 facet as the most active site using hydrothermal method is of great interest for enhancing the catalytic activities [14].
To tackle the above-mentioned issues and make photocatalytic water splitting feasible, tremendous efforts have been made to enhance the photocatalytic activity and visible light absorptions of TiO2. The following strategies prove to be the most efficient approaches to improving photocatalytic activity: (a) addition of hole scavengers (electron donors) [15,16,17,18], (b) addition of carbonate salts [19,20], (c) metal and non-metal doping [21], (d) self-doping (defect engineering) [22,23], and e) noble metal co-catalyst decoration [24]. Among the mentioned strategies, the use of noble metals as a co-catalyst for photocatalytic water splitting is one of the most efficient strategies. Platinum group metals (PGM) such as Pt, Pd, Rh, Ru, or Ir are among the least abundant elements on the earth and are thus expensive. However, they are the best-known catalysts for photocatalytic H2 evolution [25]. Therefore, there is significant interest in utilizing noble metal catalysts in heterogeneous photocatalysis. For metal particles, the ultimate size limit is the single-atom catalyst, i.e., isolated atoms dispersed on supports. In addition to their small size that leads to a higher active surface area, a characteristic is also a low coordination environment that is ascribed to unsaturated bonding on the single-atoms [26,27]. Additionally, single-atoms have a distinctive HOMO–LUMO gap due to quantum size effects [28]. Furthermore, the chemical bonding between single-atom metals and supports facilitates the charge transfer between the single-atom co-catalysts and supports [29,30].
It should be noted that the development of the single-atom catalysis field largely relies on advances in both theoretical modeling and atomic-resolution characterization techniques. The present review is specifically focused on TiO2 nanostructures for photocatalytic water splitting, using single-atom noble metals or transition metals as co-catalysts.

2. Fabrication of Single-Atom Co-Catalysts

Due to their small size, single atoms have high surface energy; thus, they often tend to agglomerate during the reaction process. One of the main goals in single-atom research is overcoming aggregation, which requires the formation of strong bonding between the single-atoms and the support material. This interaction directly influences the stability and efficiency of the reaction for which a single atom is adopted. The key to forming stable single-atom catalysts with high density is to enhance the anchoring sites. There are three types of anchoring sites for fabricating single-atom catalysts: (i) defects on the support material [31], (ii) unsaturated coordinated atoms [31], and (iii) excess O atoms on the support surface to form hollow sites [31]. Figure 2 illustrates the paths and key strategies for single-atom decoration [32], which will also be further discussed below.
In general, there are two main strategies for constructing single-atom catalysts with well-defined and separated atomistic structures, namely high vacuum physical deposition and wet chemical deposition

2.1. High Vacuum Physical Deposition

Vacuum deposition is a more suitable technique for fundamental studies mainly due to the superior control over the single-atom deposition and model catalyst formation. However, as a result of the low yield and high cost of this method, it is not yet feasible for commercial catalyst production [33,34]. Mass-selected soft-landing [35] and atomic layer deposition (ALD) [36] techniques were further developed for the fabrication of single-atom catalysts. Mass-selected soft landing (Figure 2c) includes an ion source in gas form and a mass spectrometer to filter the nanoclusters according to their mass before their deposition onto a substrate [33].
As a result, metal clusters with a well-defined number of atoms can be deposited (soft-landed) on a substrate. Kaden et al. [37] reported on the deposition of Pd clusters (Pd(n), for n = 1, 2, 4, 7, 10, 16, 20, and 25) on clean rutile TiO2 (110) to build a model catalyst for CO oxidation.
In ALD synthesis (shown schematically in Figure 2b), precursor molecules chemisorb on the support surface via gas-solid reactions and, in each reaction step, form less than a monolayer on the surface. Repeating this cycle, and depending on both the support and ALD-grown material, it is possible to form thin films, nanoparticles, and even single-atoms [38,39]. By optimizing the ALD conditions, namely temperature, counter-reactant, precursor, and pulse times, as well as selecting appropriate ALD approaches such as low-temperature selective ALD [40] or modifying the pulsing sequences [41], one can design single-atom catalysts for a wide range of catalytic applications [42].

2.2. Wet Chemical Deposition

Wet-chemical approaches are more common and used routinely to synthesize single-atom catalysts as they do not require specialized equipment. According to the integration of the components, single-atom wet chemical synthesis is categorized into bottom-up methods, where single-atoms are anchored to the substrate by the reaction of the metal complexes with the anchoring sites on the substrate surfaces, or top-down methods, where metal nanoparticles are decorated on the surface of the support followed by dispersion into single-atoms [43].
In the bottom-up approach (Figure 2d–g), very dilute metal precursors are used as a source of the single-atom where the support material is dipped inside to be decorated by single-atoms and is usually followed by a calcination step, which results in single-atom catalysts on the support surface by a reduction or activation step [44]. Impregnation [45,46], galvanic replacement [47,48], electrostatic adsorption, and photochemical adsorption [49] are among the most common bottom-up methods for single-atom catalyst fabrication.
Nanostructures are turned into smaller pieces in a top-down strategy to obtain desirable characteristics and performance. For single-atom catalyst fabrication, the high-temperature atomic migration method represents a typical top-down approach (Figure 2h). The mechanism behind this method can be explained by Ostwald ripening. This process includes detaching metal atoms from nanoparticles, diffusion of the atoms on the substrate, and attachment to bigger particles [50] or immobilizing in defects [51]. For example, Ag single-atoms were prepared starting from Ag nanoparticles at 400 °C on MnO2-based substrates [52], and Li et al. [51] showed that noble metal nanoparticles can be thermally converted to highly active and stable single-atoms (Pd, Au, Pt) at above 900 °C in an inert environment. Overall, this high-temperature atomic diffusion technique provides thermally stable and high-performance single-atom catalysts.

3. Characterization Methods for Single-Atom Co-Catalysts

Despite their extensive investigation, several key points related to the single-atom catalysts’ physicochemical properties still remain unanswered or require further investigation, especially in view of understanding and correlation these properties to their photocatalytic efficiency. This includes the reliable evaluation and characterization of the single-atoms in their environment, considering their very low amounts (low weight loadings), strong metal-support interactions, and so on, characteristics which are necessary in order to identify their amount, charge state, electronic structure, atomic configuration, bonding interactions, etc.
The characterization techniques typically employed for nanoparticle characterization (with their corresponding optimization) cannot be used to confirm the presence of single-atom catalysts, though such techniques are useful for substrate characterization or as a negative control. Nowadays, with the progress achieved in characterization techniques, it is actually feasible to observe single metal atoms or sub-nanometric metal clusters of few atoms by means of aberration-corrected electron microscopy [53] or atomic-resolution scanning tunneling microscopy (STM) [54]. Moreover, their loading and coordination environment can be further evaluated by extended X-ray absorption fine-structure spectroscopy (EXAFS), X-ray absorption near-edge spectroscopy (XANES), X-ray absorption spectroscopy (XAS), and ambient pressure X-ray photoelectron spectroscopy (XPS). Recently, focus was also given to a more thorough understanding of these characterization techniques and their application to the single-atom catalysts field, see the comprehensive reviews in refs. [28,32,55,56,57]. It is also worth mentioning that supported single atoms are usually stabilized by chemical bonding to the inorganic support (e.g., on transition metal oxides, and zeolites), as previously discussed in the fabrication section. Such SA will show a limited geometric transformation under reaction conditions, however, when supported on organic polymers with functional groups (like amine, carbonyl groups, thiol, etc.) SA may adapt their coordination environment under reaction conditions (as a result of the interaction between the single-atoms with the substrate molecules) [28]. Therefore, next we will discuss and exemplify characterization techniques for single atoms specifically decorated only on inorganic supports.

3.1. Transmission Electron Microscopy (TEM) for SA Evaluation

TEM is one of the most direct methods enabling the detailed evaluation of atomic-scale structural information of single-atom catalysts as well as their interaction with the support material. Moreover, aberration-corrected TEM either in both phase contrast TEM or sub-angstrom resolution high-angle annular dark-field STEM (HAADF-STEM) mode, was successful in imaging single-atom catalysts on various supports [58]. The advantage of HAADF-STEM lies in the fact that the imaging is based on Rutherford scattering (i.e., image intensity for the given atoms is proportional to the square of the element′s atomic number), and this results in a brighter contrast of heavy metals atoms when on low mass background supports. Current literature includes the identification of single-atom on graphene-based supports (Au [59], Fe [60], Ru [61], etc.), graphitic carbon nitride (Pt [62]), carbon paper (Pt, Ru [63]) or inorganic semiconductors such as cerium oxide (Pt [64,65], Ru [66]), alumina (Pt [67]) or TiO2 (Pt [45,68,69,70,71,72,73,74,75], Pd [71,76,77,78,79,80,81], Au [71,82,83], Ir [84,85], Rh [86]). This enabled a precise evaluation of the size and distribution of the single metal atoms, as well as their local structural information (metal species on the support) [55]. For example, HAADF-TEM images of typical Pt SAs on different inorganic semiconductors are shown in Figure 3a for Pt SAs obtained by the L-ascorbic acid (AA)-assisted reduction synthesis on CeO2. Namely, Chen et al. [64] confirmed the presence of atomically dispersed Pt catalysts throughout the CeO2 support (on the surface defects), showing also a uniform distribution of the co-catalyst on the support in the elemental mapping (Figure 3b). Similarly, Hejazi et al. [45] have shown that the Pt is present as SA on TiO2 with a uniform distribution, see Figure 3c,d, and such a disperse SA loading was achieved by a dark deposition method based on wet impregnation.
Thus, TEM enables the direct imaging of supported co-catalyst particles and can identify structural and interfacial information of both the co-catalyst metals and substrate support, but only for specific positions on the sample, which means that it can lack a broader context of the investigated sample or in bulk. Therefore, it would prove helpful to combine TEM with other characterization methods to ensure that the observation is valid for the whole of the material. The recent comprehensive reviews of Kottwitz et al. [56] and Tieu et al. [87], detailing the use of advanced electron microscopy in single-atom catalysis, are also recommended for more details.

3.2. X-ray Absorption near Edge Structure (XANES) and Extended X-ray Absorption Fine Structure (EXAFS) for SA Evaluation

XANES and EXAFS are characterization techniques reflecting the oscillatory structure of the studied samples in correlation with the X-ray absorption coefficient [57,88]. The binding energy shifts and wiggles of the single metal atoms in XANES spectra represent the near-neighbor distances and their chemical coordination [57]. For example, XANES spectra would indicate complex structures, including the formation of additional bonds (M–N, M–C, etc.) which can disrupt the precise data associated to isolated metal atoms [57]. A significant advantage of these methods is their ability to provide insight into the chemical nature and stabilization structure of single atoms on supports, as well as the investigation of bulk materials. Combined with TEM, this would lead to the overall structure of metal SA compared to local, surface information [55,89]. The absence of metal-metal interactions can be a clear indication of the absence of particles in which, in general, single supported atoms are oxidized, and thus ruling out the presence of rafts/clusters/nanoparticles.
Especially for evaluating the electronic structure of metal SA, XANES measurements are useful for characterizing the electronic and structural properties of the absorbed SA, as shown by Li et al. [90] (Figure 4a). The authors [90] attribute the white line of the Pt L3 edge to the transition of 2p electrons to unoccupied 5d valence orbitals, thus providing information related to the population of unoccupied Pt 5d orbitals. Evaluating the Pt SA on various supports (Co3O4, CeO2, ZrO2, graphene) and comparing the data with that of Pt foil and PtO2 (as references), resulted in establishing an explicit electronic correlation with the unoccupied Pt 5d states, having the following magnitude for the Pt SAs on the supports Co3O4 > ZrO2 > CeO2 > graphene [90]. The Pt SAs on the Co3O4 have shown a higher valence state of ~4+ and a larger population of unoccupied Pt 5d states [90]. Similar investigations were also performed to investigate the Pt 5d occupation state of Pt SA compared to small clusters or nanoparticles (on graphene) [91], for Pt SAs on FeOx [92] or on different polymorphs of TiO2 nanoparticles [93] and for bi-metallic SAs of PtAu [94].
Similarly, Li et al. [90] have further evaluated the EXAFS analysis of the Pt SAs on the various supports, see Figure 4b, and attributed to the coordination number of Pt–O (peak at 1.60 Å) the subsequent trend Co3O4 > CeO2 > ZrO2 > graphene, for the Pt SAs decorated different substrates, and the absence of the Pt–Pt coordination peak suggested the absence of Pt NPs or clusters. Additionally, the Pt/Co3O4 curve showed a second shell peak at 2.56 Å (attributed to Pt-Co coordination), distinct from those of Pt foil (2.37 Å) and PtO2 (2.68 Å) [90]. Similarly, the coordination of Pt SAs on different supports (TiO2 [93]) or of other SAs (Au on TiO2 [82]) can also be evaluated. For example Wan et al. [82] reported on the isolated single atomic dispersion of Au on defective TiO2, with an Au–O shell (R = 1.6 Å) and an Au–Ti shell (R = 2.4 Å), without Au–Au and Au–Cl coordination peaks (compared with Au foil and HAuCl4 as reference).
XANES and EXAFS measurements are also advantageous for investigating single-atoms on MOF, zeolites, or graphitic layers, see as examples Ye et al. [95] for Co SAs on MOF (Co–N bonding detected in EXAFS), Yin et al. [96] for Co on N-doped carbon, or Choi et al. [97] for Pt on zeolites.

3.3. X-ray Photoelectron Spectroscopy (XPS) for SA Evaluation

XPS is a well-known characterization technique, universally used for evaluating the chemical and compositional properties of surfaces, and more specifically in catalysis of the catalysts and co-catalysts. XPS is a highly surface-sensitive technique ensuring analysis of the top 10 nm of the sample. Moreover, it has shown adaptability and precision to the field of supported metal SA catalysts [56,98], and can also be used in ambient pressure mode allowing for the evaluation of catalysts under the reaction atmosphere (e.g., Simonovis et al. [99], evaluating the behavior of singly dispersed Pt atoms on the surface of Cu(111) in an ambient pressure of CO).
Please note that the trapped metal single atom, denoted as M, shows a chemical shift of δ+ due to the changed co-coordination (as a result of the oxygen coordination, see the case of Pt SAs on CeO2, where the Pt is bound to oxygen instead of Pt [100]). Compared to classical metallic nanoparticles, noble metal SAs decorated on the same support material show a shift to higher binding energies, which is related to its surface coordination—please see the work of Wan et al. [82] showing a shift in the Au4f XPS spectra of Au SAs (85 eV) compared to Au NPs (84 eV) on TiO2. Moreover, Figure 5a shows a similar shift to higher binding energies for Pt SAs/TiO2, compared to their nanoparticle counterpart. The deconvolution of the Pt4f peak of the Pt SAs (see Figure 5b) results into Ptδ+ and Pt4+ (the peaks at 72.93 and 76.28 eV correspond to Ptδ+4f7/2 and Ptδ+4f5/2, whereas the peaks at 75.04 and 78.51 eV can be attributed to Pt4+4f7/2 and Pt4+4f5/2), while for the Pt nanoparticle decoration peaks are mainly attributed to metallic Pt (4f7/2 at 71.07eV) [45].
Literature confirms that when Pt is present in the SA state, it is coordinated to its trap environment, that is mainly oxygen atoms of the oxide surface, thus forming Pt-oxide-similar Ptδ+ and Pt4+ states (with Pt4+ as evidence of a strong Pt–TiO2 interaction) [92,100,101,102,103]. Similar results were also obtained for other Pt single-atom decoration on different TiO2 supports [69,70,71,104] or on other support materials, e.g., Pt on nanosized CeO2 (no metallic Pt, only SA with +2 state) [105]. Moreso, XPS was used to evaluate the chemical state of a variety of other single-atoms co-catalysts, such as Pd [71,78], Au [71], Rh [86,106], and Ir [84], confirming the δ+ state of the single-atom co-catalyst on the evaluated support material.
Recently Wei et al. [94] reported on the synthesis of bi-metallic single-atoms on different facetted TiO2 ({001},{101}), where the metal–support interactions that originated from the coordinatively unsaturated sites contributed to the anchoring of the atomic co-catalysts—namely, onto {001}-TiO2 through Pt–O and Au–O bonds, while on {101}-TiO2 preferentially for Pt–O and Au nanoparticles. This difference was also observed in the high-resolution XPS spectra, as for the Pt4f peaks (Figure 5c), binding energies for Pt4f7/2 at 73.0 (typically attributed to Pt+4) are obtained, while for the Au4f peaks (Figure 5d) a shift of 0.2eV to more negative binding energies was observed for PtAu/{101}-TiO2 (where Au is deposited as nanoparticles). The latter may be due to the changes in the binding environment of Au, as the weaker metal-support interactions lead to agglomeration and reduce the binding energies [71,94].
With respect to the energy shift in XPS, two further aspects need to be mentioned, (i) first that both support and particle size could induce positive binding energy shifts (as shown for Pt nanostructures because of the reduction of effective atomic coordination numbers) [107], and (ii) second, in the case of comparing the different SA binding energies, careful consideration of the peaks used for calibration has to be made (especially if the C1s peak at 284.8eV is employed [108,109,110]).
As a closing remark, a combination of characterization techniques has to be used to properly identify and quantify the atomic fractions of the SA. Additional characterization techniques adapted to single-atom characterization include (a) Fourier-transform infrared spectroscopy (FTIR)—molecular groups, chemical bonding (only specific molecules can be detected) [98] or diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS)—which provides chemical and structural information related to surface species, commonly through measurement of adsorbed probe (e. g., CO) molecules in studies of catalytic reaction mechanisms, typically linearly adsorbed CO on an isolated metal atom site [56,105], (b) surface-enhanced Raman spectroscopy (SERS)—crystal structure, chemical bonding (though only working for specific elements) [98], (c) electron energy loss spectroscopy (EELS)—element and chemical state analysis [98], scanning tunneling microscopy (STM)—imagining of the supported individual single metal atoms (low temperature under ultrahigh vacuum) [28,98].

4. Photocatalytic Hydrogen Evolution on Single Atom Decorated TiO2

Sluggish charge transfer, limited photoconversion, and fast electron-hole recombination rate are the bottlenecks of TiO2 photocatalytic water splitting, and as mentioned earlier, this makes the introduction of a promising co-catalyst essential in order to improve the photocatalytic performance of TiO2. Recently, single-atom catalysts have drawn significant attention due to enhanced active reaction sites, while decreasing the loading compared to conventional nanoparticle catalysts. The active and isolated metal atoms anchored onto the photocatalysts provide more water-molecule adsorption and thus enhance photocatalytic activity. However, the aggregation of SAs during the catalytic reaction is the major issue that leads to a significant drop in the performance of the photocatalysts. Therefore, obtaining a high concentration and highly dispersed single-atom catalysts is the bottleneck in photocatalytic water splitting using SAs. The single-atom catalysts are stabilized by covalent bonding with the carrier or neighboring surface atoms or interionic interaction. In this context, the catalytic activity of a single-atom can be tuned by metal-support interaction, coordination environment, quantum size effect, and so on [111].
Noble metals such as Pt, Au, and Pd are mainly used as co-catalysts in photocatalytic water splitting, and this is due to their efficient charge separation and low activation energy. To reach a more scalable water splitting system, earth-abundant transition metals such as Cu, Ni, Co, and Fe are promising alternatives for noble metals in photocatalysis [112,113,114,115]. Utilizing noble metals and transition metals as single atom co-catalysts for producing H2 from water is known to result in higher efficiency and an overall lower cost. In this section, the most recent research on photocatalytic water splitting reactions using single-atom catalysts on TiO2 will be discussed.
Zhang et al. [116] reported on a large loading amount (1.5 wt.% of Cu) and highly dispersed copper single-atom on nanoparticle TiO2 (MIL-125(Tiv)) for photocatalytic H2 generation that exhibits a quantum efficiency of 56% under 365 nm irradiation (see Figure 6). A range of loading from 0.47 wt% to 2.57 wt% of Cu was evaluated in their experiments. The authors attributed this performance to the strongly anchored Cu single atoms on TiO2 that trigger a reversible/self-healing and continuous photocatalytic process that effectively separates photoelectrons to reduce water to H2. The results are supported with various spectroscopic and in-situ experiments as well as with DFT modelling results. Their study reveals the most efficient charge separation by Cu SA and proves the significance of the in-situ self-healing effect of Cu species during the photocatalytic reaction.
In another approach on TiO2 systems modified by Cu single-atoms, Lee et al. [117] reveal that the cooperative interplay of atomic catalysts and their environment has a significant influence on the properties and catalytic performance of the material. They suggest that the fundamental atomic-level cooperative interaction is very similar to homogeneous catalysts. Through ab initio nonadiabatic molecular dynamics, Long et al. [118] show simulations that Cu single-atoms on TiO2 are inactive for water splitting, before light irradiation. However, after illumination, the deep trap state on the Cud orbital is stabilized by the local distortion around the Cu single atom, decoupling it from the free charge thus providing a high photocatalytic hydrogen production activity.
It is also reported that bi- or multi-metallic catalytic sites show higher photocatalytic water splitting activity compared to regular single-atom catalysts. Wei et al. [94] show that bi-metallic Pt and an Au single atom on the {001}-TiO2 presented a 1000-fold increase in the H2 generation rate in comparison to pristine{001}-TiO2. They show that by regulating the coordination environment of atoms, it is possible to disperse bimetal catalytic sites on faceted semiconductors. They show an H2 evolution rate of 61.3 mmol h–1g–1, much higher than that of the TiO2 modified by Pt or Au single-atoms. They explain the mechanism by the synergistic effect of PtAu dual-atom catalysts, which can decrease the Gibbs free energies of hydrogen adsorption by optimizing the electronic states of both Pt and Au sites. Particularly, the Pt atom is activated by the Au atom and the activity of catalysts is further enhanced through the dimer interaction (see Figure 7).
In another study, Wang et al. [119] reported on the Co and Pt dual single-atom on TiO2 with ultrahigh photocatalytic hydrogen generation efficiency. They attributed this enhancement to the oxygen-coordinated Co–O–Pt dimer coupling, which is able to optimize the electronic states of Pt and Co sites in order to weaken H* binding. They further explain that the Pt single-atom activates the “mute” Co single-atom and the activity of the Pt atom itself is additionally improved by the dimer interaction. With this strategy, the authors show that the dual single-atom catalyst yields 13.4% dimer sites that lead to stable and highly active photocatalysts with an H2 generation rate of 43.467 mmol g−1 h−1 and external quantum efficiency of ≈83.4% at 365 nm.
Wu et al. [120] reported on the isolated Co atoms on TiO2 nanosheets that significantly improve the charging kinetics via a Co-O electronic coupling. With experimental investigations and theoretical calculations, the authors show that the electron transfer and the hydrogen adsorption/desorption processes are facilitated by the effective atomic scale Co-O electronic coupling.
Several experimental and theoretical studies suggest that coupling Pt single atoms with the atomic scale defects in TiO2 nanostructures is an efficient strategy for stabilizing SAs and to achieving an excellent photocatalytic H2 generation activity compared to conventional nanoparticle decoration [45,68,121,122]. The role of oxygen vacancy in single-atom catalysts expands to Pd for instance, where Wang et al. [77] showed, by experimental and theoretical studies, that the presence of Pd SAs, clusters, and VOs on TiO2 has a synergistic effect on improving the photocatalytic reaction. The authors explain the role of Pd SAs by facilitating the generation and stabilization of oxygen vacancies by the formation of the Pd–O–Ti3+ interface, at the same time Pd clusters are responsible for charge carrier separation and act as the optimum active sites for H2 generation, while surface oxygen vacancies are the preferable sites for adsorption and activation/dissociation of reactant molecules. It has also been reported that oxygen vacancies on the surface of TiO2 nanosheets act as traps for Pd single-atoms as cocatalysts for photocatalytic H2 generation [71]. More interestingly, the photocatalytic performance of Pd single-atoms outperforms that of the Pt single-atoms, and according to the DFT calculations, this performance was ascribed to the charge localization on the noble metal single-atoms implanted in the TiO2 surface [71].
Table 1 shows an overview of the photocatalytic H2 evolution rate for various SA co-catalysts on TiO2 support materials discussed above, taking also into consideration the morphology of the support material, the SA loading, light source, and reaction system used. From these data, it is evident that a clear-cut comparison between the different noble metal single-atom decorations is not feasible, as either there is a significant difference in the noble metal loading amount, the light source used, the difference in the morphology of the support material and the amount of sacrificial agent used in the reaction system.
Using first principle calculations, Yang et al. [123] shed light on the diversity and complexity of the confinement effect of transition metal SA in oxides for H2 evolution. They introduce two stable forms of Co single-atom on the (101)TiO2 surface which are interstices and Co1-substituted TiO2. They prove that the former exhibits better H2 generation activity whereas the latter demonstrates increased hydrogen spillover effect in the H2 atmosphere, producing surface O vacancies and Ti3+.
Calculating the free energy changes of hydrogen adsorption in a single-atom Ni/TiO2 system using density functional theory, Bi et al. [124] reveal that the hydrogen generation reactions most probably take place on SA Ni and its surrounding O atoms. They also prove that hydrogen activity in this system is comparable to that of Pt as the best-known catalyst for H2 evolution, and using DFT calculation, identified the active sites on the Ni single atom/TiO2(101) surface at different H coverages.

5. Conclusions and Future Outlook

Overcoming the kinetic barrier of photocatalytic water splitting half reactions, which are an H2 evolution reaction and oxygen evolution reaction, is the main challenge for converting solar light to hydrogen. The key to enhance the efficiency of the photocatalytic water splitting reaction is to increase the number and intrinsic activity of the active sites. Remarkably, reducing the size of the photocatalysts simultaneously improves the above-mentioned factors.
In this regard, research on single-atom catalysts has become a frontier in photocatalytic water splitting in the past few years and this is the direct result of the atomistic scale of the single-atoms which leads to a higher active surface area, lower loading amount compared to classical co-catalysts (thus more cost-effective), quantum size effects, facilitated charge transfer and new reaction pathways The advances achieved in both theoretical modeling and atomic-resolution characterization techniques play key roles in the development of single-atom catalysts. For example, the atomic resolution characterization techniques used for the typical nanoparticulate co-catalysts were further developed and tailored to the detection and proper evaluation of the small size of the single atoms. Currently, there is no one characterization technique from which all physico-chemical data of the single-atom loading and their interaction with the surrounding environment can be obtained. There is, however, a mix of techniques which can bring a clear and detailed overview of these properties, and this includes: (i) transmission electron microscopy (HAADF-TEM)—size, distribution, and local structural information of the single metal atoms, combined with additional elemental mapping; (ii) X-ray absorption near edge structure (XANES) and extended X-Ray absorption fine structure (EXAFS)—electronic and structural properties of the absorbed single-atom and their coordination; (iii) X-ray photoelectron spectroscopy (XPS)—coordination state and loading amount of the single-atoms; (iv) or other additional methods such as diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS), scanning tunneling microscopy (STM).
Modification of TiO2, as one of the most studied materials in photocatalysis, with single atoms as cocatalysts for photocatalytic water splitting has become an emerging field of study. The major challenges in single-atom catalysis are connected to the fact that the single-atom loading on the current photocatalysts is very low (<2 wt.%) and to the stability of the SAs, due to their high surface energy. Thus, the focus of research in this area is to enhance the stability and loading of single-atoms on the support material while avoiding the formation of/aggregation to nanoparticles, i.e., obtaining a high coverage of single-atoms on the substrate. In this respect, the interaction of single atoms with the support material has a substantial influence on the stability, loading, and overall photocatalytic activity of the developed SA/TiO2 photocatalytic system. For example, the in situ self-healing effect of Cu species on TiO2 during the photocatalytic reaction resulted in a quantum efficiency of 56% under 365 nm irradiation [116]. Remarkably, the use of bi-metallic single-atoms on a suitable support proves to be an efficient approach in designing a highly efficient photocatalyst. In general, controlling the fabrication of single atoms and operando atomic scale characterizations are the keys to achieving a better understanding of the reactions occurring on single-atoms. Therefore, rapid development in single-atom catalysts for photocatalytic water splitting and, in general, other catalytic applications, is foreseeable.

Author Contributions

Writing original draft, review and editing, S.H.; Review and funding acquisition, M.S.K.; Supervision, writing original draft, review and editing, A.M.; Supervision, writing original draft, review and editing, S.M. This manuscript was written by the contributions of all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by DFG researcher group FOR 1878 and KI 2169/2-1.

Acknowledgments

The authors acknowledge the DFG researcher group FOR 1878 and KI 2169/2-1 for funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fujishima, A.; Honda, K. TiO2 photoelectrochemistry and photocatalysis. Nature 1972, 213, 8656. [Google Scholar]
  2. Armaroli, N.; Balzani, V. The Hydrogen Issue. ChemSusChem 2011, 4, 21–36. [Google Scholar] [CrossRef] [PubMed]
  3. Takanabe, K. Photocatalytic Water Splitting: Quantitative Approaches toward Photocatalyst by Design. ACS Catal. 2017, 7, 8006–8022. [Google Scholar] [CrossRef]
  4. Osterloh, F.E. Inorganic nanostructures for photoelectrochemical and photocatalytic water splitting. Chem. Soc. Rev. 2013, 42, 2294–2320. [Google Scholar] [CrossRef]
  5. Maeda, K.; Domen, K. New Non-Oxide Photocatalysts Designed for Overall Water Splitting under Visible Light. J. Phys. Chem. C 2007, 111, 7851–7861. [Google Scholar] [CrossRef]
  6. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef]
  7. Maeda, K.; Domen, K. Photocatalytic Water Splitting: Recent Progress and Future Challenges. J. Phys. Chem. Lett. 2010, 1, 2655–2661. [Google Scholar] [CrossRef]
  8. Sivula, K.; van de Krol, R. Semiconducting materials for photoelectrochemical energy conversion. Nat. Rev. Mater. 2016, 1, 15010. [Google Scholar] [CrossRef]
  9. Ng, K.H.; Lai, S.Y.; Cheng, C.K.; Cheng, Y.W.; Chong, C.C. Photocatalytic water splitting for solving energy crisis: Myth, Fact or Busted? Chem. Eng. J. 2021, 417, 128847. [Google Scholar] [CrossRef]
  10. Gopinath, C.S.; Nalajala, N. A scalable and thin film approach for solar hydrogen generation: A review on enhanced photocatalytic water splitting. J. Mater. Chem. A 2021, 9, 1353–1371. [Google Scholar] [CrossRef]
  11. Lee, K.; Mazare, A.; Schmuki, P. One-dimensional titanium dioxide nanomaterials: Nanotubes. Chem. Rev. 2014, 114, 9385–9454. [Google Scholar] [CrossRef] [PubMed]
  12. Lan, Y.; Lu, Y.; Ren, Z. Mini review on photocatalysis of titanium dioxide nanoparticles and their solar applications. Nano Energy 2013, 2, 1031–1045. [Google Scholar] [CrossRef]
  13. Chen, H.; Nanayakkara, C.E.; Grassian, V.H. Titanium dioxide photocatalysis in atmospheric chemistry. Chem. Rev. 2012, 112, 5919–5948. [Google Scholar] [CrossRef] [PubMed]
  14. Prashant, V. Kamat TiO2 Nanostructures: Recent Physical Chemistry Advances. J. Phys. Chem. C 2012, 116, 11849–11851. [Google Scholar] [CrossRef]
  15. Bamwenda, G.R.; Tsubota, S.; Nakamura, T.; Haruta, M. Photoassisted hydrogen production from a water-ethanol solution: A comparison of activities of Au□TiO2 and Pt□TiO2. J. Photochem. Photobiol. A Chem. 1995, 89, 177–189. [Google Scholar] [CrossRef]
  16. Lee, S.G.; Lee, S.; Lee, H.-I. Photocatalytic production of hydrogen from aqueous solution containing CN as a hole scavenger. Appl. Catal. A Gen. 2001, 207, 173–181. [Google Scholar] [CrossRef]
  17. Gurunathan, K.; Maruthamuthu, P.; Sastri, M.V.C. Photocatalytic hydrogen production by dye-sensitized Pt/SnO2 and Pt/SnO2/RuO2 in aqueous methyl viologen solution. Int. J. Hydrogen Energy 1997, 22, 57–62. [Google Scholar] [CrossRef]
  18. 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]
  19. Arakawa, H.; Sayama, K. Solar hydrogen production. Significant effect of Na2CO3 addition on water splitting using simple oxide semiconductor photocatalysts. Catal. Surv. Jpn. 2000, 4, 75–80. [Google Scholar] [CrossRef]
  20. Sayama, K.; Yase, K.; Arakawa, H.; Asakura, K.; Tanaka, A.; Domen, K.; Onishi, T. Photocatalytic activity and reaction mechanism of Pt-intercalated K4Nb6O17 catalyst on the water splitting in carbonate salt aqueous solution. J. Photochem. Photobiol. A Chem. 1998, 114, 125–135. [Google Scholar] [CrossRef]
  21. Basavarajappa, P.S.; Patil, S.B.; Ganganagappa, N.; Reddy, K.R.; Raghu, A.V.; Reddy, C.V. Recent progress in metal-doped TiO2, non-metal doped/codoped TiO2 and TiO2 nanostructured hybrids for enhanced photocatalysis. Int. J. Hydrogen Energy 2020, 45, 7764–7778. [Google Scholar] [CrossRef]
  22. Hejazi, S.; Pour-Ali, S.; Killian, M.S.; Mohajernia, S. One-dimensional suboxide TiO2 nanotubes for electrodics applications. Electrochem. Commun. 2022, 136, 107246. [Google Scholar] [CrossRef]
  23. Naldoni, A.; Altomare, M.; Zoppellaro, G.; Liu, N.; Kment, Š.; Zbořil, R.; Schmuki, P. Photocatalysis with Reduced TiO2: From Black TiO2 to Cocatalyst-Free Hydrogen Production. ACS Catal. 2019, 9, 345–364. [Google Scholar] [CrossRef] [PubMed]
  24. Bumajdad, A.; Madkour, M. Understanding the superior photocatalytic activity of noble metals modified titania under UV and visible light irradiation. Phys. Chem. Chem. Phys. 2014, 16, 7146–7158. [Google Scholar] [CrossRef] [PubMed]
  25. Trasatti, S. Work function, electronegativity, and electrochemical behaviour of metals: III. Electrolytic hydrogen evolution in acid solutions. J. Electroanal. Chem. Interfacial Electrochem. 1972, 39, 163–184. [Google Scholar] [CrossRef]
  26. Van Hardeveld, R.; Hartog, F. The statistics of surface atoms and surface sites on metal crystals. Surf. Sci. 1969, 15, 189–230. [Google Scholar] [CrossRef]
  27. Lopez, N.; Janssens, T.V.W.; Clausen, B.S.; Xu, Y.; Mavrikakis, M.; Bligaard, T.; Nørskov, J.K. On the origin of the catalytic activity of gold nanoparticles for low-temperature CO oxidation. J. Catal. 2004, 223, 232–235. [Google Scholar] [CrossRef]
  28. Liu, L.; Corma, A. Metal Catalysts for Heterogeneous Catalysis: From Single Atoms to Nanoclusters and Nanoparticles. Chem. Rev. 2018, 118, 4981–5079. [Google Scholar] [CrossRef]
  29. Valden, M.; Lai, X.; Goodman, D.W. Onset of Catalytic Activity of Gold Clusters on Titania with the Appearance of Nonmetallic Properties. Science 1998, 281, 1647–1650. [Google Scholar] [CrossRef]
  30. Campbell, C.T. Electronic perturbations. Nat. Chem. 2012, 4, 597–598. [Google Scholar] [CrossRef]
  31. Wang, A.; Li, J.; Zhang, T. Heterogeneous single-atom catalysis. Nat. Rev. Chem. 2018, 2, 65–81. [Google Scholar] [CrossRef]
  32. Weon, S.; Huang, D.; Rigby, K.; Chu, C.; Wu, X.; Kim, J.-H. Environmental Materials beyond and below the Nanoscale: Single-Atom Catalysts. ACS EST Eng. 2021, 1, 157–172. [Google Scholar] [CrossRef]
  33. Liu, J. Catalysis by Supported Single Metal Atoms. ACS Catal. 2017, 7, 34–59. [Google Scholar] [CrossRef]
  34. Wang, J.; Li, Z.; Wu, Y.; Li, Y. Fabrication of Single-Atom Catalysts with Precise Structure and High Metal Loading. Adv. Mater. 2018, 30, 1801649. [Google Scholar] [CrossRef] [PubMed]
  35. Vajda, S.; White, M.G. Catalysis Applications of Size-Selected Cluster Deposition. ACS Catal. 2015, 5, 7152–7176. [Google Scholar] [CrossRef]
  36. Zhang, L.; Banis, M.N.; Sun, X. Single-atom catalysts by the atomic layer deposition technique. Natl. Sci. Rev. 2018, 5, 628–630. [Google Scholar] [CrossRef]
  37. Kaden, W.E.; Wu, T.; Kunkel, W.A.; Anderson, S.L. Electronic Structure Controls Reactivity of Size-Selected Pd Clusters Adsorbed on TiO2 Surfaces. Science 2009, 326, 826–830. [Google Scholar] [CrossRef]
  38. Van Bui, H.; Grillo, F.; van Ommen, J.R. Atomic and molecular layer deposition: Off the beaten track. Chem. Commun. 2017, 53, 45–71. [Google Scholar] [CrossRef]
  39. Miikkulainen, V.; Leskelä, M.; Ritala, M.; Puurunen, R.L. Crystallinity of inorganic films grown by atomic layer deposition: Overview and general trends. J. Appl. Phys. 2013, 113, 21301. [Google Scholar] [CrossRef]
  40. Wang, H.; Wang, C.; Yan, H.; Yi, H.; Lu, J. Precisely-controlled synthesis of Au@Pd core–shell bimetallic catalyst via atomic layer deposition for selective oxidation of benzyl alcohol. J. Catal. 2015, 324, 59–68. [Google Scholar] [CrossRef]
  41. Lu, J.; Stair, P.C. Low-Temperature ABC-Type Atomic Layer Deposition: Synthesis of Highly Uniform Ultrafine Supported Metal Nanoparticles. Angew. Chem. Int. Ed. 2010, 49, 2547–2551. [Google Scholar] [CrossRef] [PubMed]
  42. Fonseca, J.; Lu, J. Single-Atom Catalysts Designed and Prepared by the Atomic Layer Deposition Technique. ACS Catal. 2021, 11, 7018–7059. [Google Scholar] [CrossRef]
  43. Li, Z.; Wang, D.; Wu, Y.; Li, Y. Recent advances in the precise control of isolated single-site catalysts by chemical methods. Natl. Sci. Rev. 2018, 5, 673–689. [Google Scholar] [CrossRef]
  44. Chen, Y.; Ma, Z.; Tang, X. Single-Atom Heterogeneous Catalysts. In Heterogeneous Catalysts: Advanced Design, Characterization and Applications, II; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2021; pp. 103–117. [Google Scholar] [CrossRef]
  45. Hejazi, S.; Mohajernia, S.; Osuagwu, B.; Zoppellaro, G.; Andryskova, P.; Tomanec, O.; Kment, S.; Zbořil, R.; Schmuki, P. On the controlled loading of single platinum atoms as a Co-catalyst on TiO2 anatase for optimized photocatalytic H2 generation. Adv. Mater. 2020, 32, 1908505. [Google Scholar] [CrossRef] [PubMed]
  46. Moses-DeBusk, M.; Yoon, M.; Allard, L.F.; Mullins, D.R.; Wu, Z.; Yang, X.; Veith, G.; Stocks, G.M.; Narula, C.K. CO Oxidation on Supported Single Pt Atoms: Experimental and ab Initio Density Functional Studies of CO Interaction with Pt Atom on θ-Al2O3(010) Surface. J. Am. Chem. Soc. 2013, 135, 12634–12645. [Google Scholar] [CrossRef]
  47. Speck, F.D.; Kim, J.H.; Bae, G.; Joo, S.H.; Mayrhofer, K.J.J.; Choi, C.H.; Cherevko, S. Single-Atom Catalysts: A Perspective toward Application in Electrochemical Energy Conversion. JACS Au 2021, 1, 1086–1100. [Google Scholar] [CrossRef]
  48. Wang, L.; Li, H.; Zhang, W.; Zhao, X.; Qiu, J.; Li, A.; Zheng, X.; Hu, Z.; Si, R.; Zeng, J. Supported Rhodium Catalysts for Ammonia–Borane Hydrolysis: Dependence of the Catalytic Activity on the Highest Occupied State of the Single Rhodium Atoms. Angew. Chem. Int. Ed. 2017, 56, 4712–4718. [Google Scholar] [CrossRef]
  49. Liu, P.; Zhao, Y.; Qin, R.; Mo, S.; Chen, G.; Gu, L.; Chevrier, D.M.; Zhang, P.; Guo, Q.; Zang, D.; et al. Photochemical route for synthesizing atomically dispersed palladium catalysts. Science 2016, 352, 797–800. [Google Scholar] [CrossRef]
  50. Ouyang, R.; Liu, J.-X.; Li, W.-X. Atomistic Theory of Ostwald Ripening and Disintegration of Supported Metal Particles under Reaction Conditions. J. Am. Chem. Soc. 2013, 135, 1760–1771. [Google Scholar] [CrossRef]
  51. Wei, S.; Li, A.; Liu, J.-C.; Li, Z.; Chen, W.; Gong, Y.; Zhang, Q.; Cheong, W.-C.; Wang, Y.; Zheng, L.; et al. Direct observation of noble metal nanoparticles transforming to thermally stable single atoms. Nat. Nanotechnol. 2018, 13, 856–861. [Google Scholar] [CrossRef]
  52. Huang, Z.; Gu, X.; Cao, Q.; Hu, P.; Hao, J.; Li, J.; Tang, X. Catalytically Active Single-Atom Sites Fabricated from Silver Particles. Angew. Chem. Int. Ed. 2012, 51, 4198–4203. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, J.C.; Small, M.W.; Grieshaber, R.V.; Nuzzo, R.G. Recent developments and applications of electron microscopy to heterogeneous catalysis. Chem. Soc. Rev. 2012, 41, 8179. [Google Scholar] [CrossRef]
  54. Wiesendanger, R. Single-atom magnetometry. Curr. Opin. Solid State Mater. Sci. 2011, 15, 1–7. [Google Scholar] [CrossRef]
  55. Cheng, N.; Zhang, L.; Doyle-Davis, K.; Sun, X. Single-Atom Catalysts: From Design to Application. Electrochem. Energy Rev. 2019, 2, 539–573. [Google Scholar] [CrossRef]
  56. Kottwitz, M.; Li, Y.; Wang, H.; Frenkel, A.I.; Nuzzo, R.G. Single Atom Catalysts: A Review of Characterization Methods. Chem.–Methods 2021, 1, 278–294. [Google Scholar] [CrossRef]
  57. Zhang, L.; Wang, Y.; Niu, Z.; Chen, J. Single Atoms on Graphene for Energy Storage and Conversion. Small Methods 2019, 3, 1800443. [Google Scholar] [CrossRef]
  58. Fultz, B.; Howe, J. High-Resolution STEM and Related Imaging Techniques. In Transmission Electron Microscopy and Diffractometry of Materials. Graduate Texts in Physics; Springer: Berlin/Heidelberg, Germany, 2013; pp. 587–615. [Google Scholar]
  59. Wang, H.; Li, K.; Cheng, Y.; Wang, Q.; Yao, Y.; Schwingenschlögl, U.; Zhang, X.; Yang, W. Interaction between single gold atom and the graphene edge: A study via aberration-corrected transmission electron microscopy. Nanoscale 2012, 4, 2920. [Google Scholar] [CrossRef]
  60. Liu, D.; Wu, C.; Chen, S.; Ding, S.; Xie, Y.; Wang, C.; Wang, T.; Haleem, Y.A.; ur Rehman, Z.; Sang, Y.; et al. In situ trapped high-density single metal atoms within graphene: Iron-containing hybrids as representatives for efficient oxygen reduction. Nano Res. 2018, 11, 2217–2228. [Google Scholar] [CrossRef]
  61. Zhang, C.; Sha, J.; Fei, H.; Liu, M.; Yazdi, S.; Zhang, J.; Zhong, Q.; Zou, X.; Zhao, N.; Yu, H.; et al. Single-Atomic Ruthenium Catalytic Sites on Nitrogen-Doped Graphene for Oxygen Reduction Reaction in Acidic Medium. ACS Nano 2017, 11, 6930–6941. [Google Scholar] [CrossRef]
  62. 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]
  63. Wang, Z.; Yang, J.; Gan, J.; Chen, W.; Zhou, F.; Zhou, X.; Yu, Z.; Zhu, J.; Duan, X.; Wu, Y. Electrochemical conversion of bulk platinum into platinum single-atom sites for the hydrogen evolution reaction. J. Mater. Chem. A 2020, 8, 10755–10760. [Google Scholar] [CrossRef]
  64. Chen, J.; Wanyan, Y.; Zeng, J.; Fang, H.; Li, Z.; Dong, Y.; Qin, R.; Wu, C.; Liu, D.; Wang, M.; et al. Surface Engineering Protocol To Obtain an Atomically Dispersed Pt/CeO2 Catalyst with High Activity and Stability for CO Oxidation. ACS Sustain. Chem. Eng. 2018, 6, 14054–14062. [Google Scholar] [CrossRef]
  65. Bruix, A.; Lykhach, Y.; Matolínová, I.; Neitzel, A.; Skála, T.; Tsud, N.; Vorokhta, M.; Stetsovych, V.; Ševčíková, K.; Mysliveček, J.; et al. Maximum Noble-Metal Efficiency in Catalytic Materials: Atomically Dispersed Surface Platinum. Angew. Chem. Int. Ed. 2014, 53, 10525–10530. [Google Scholar] [CrossRef] [PubMed]
  66. Chen, L.; Qi, Z.; Peng, X.; Chen, J.-L.; Pao, C.-W.; Zhang, X.; Dun, C.; Young, M.; Prendergast, D.; Urban, J.J.; et al. Insights into the Mechanism of Methanol Steam Reforming Tandem Reaction over CeO2 Supported Single-Site Catalysts. J. Am. Chem. Soc. 2021, 143, 12074–12081. [Google Scholar] [CrossRef]
  67. Zhang, Z.; Zhu, Y.; Asakura, H.; Zhang, B.; Zhang, J.; Zhou, M.; Han, Y.; Tanaka, T.; Wang, A.; Zhang, T.; et al. Thermally stable single atom Pt/m-Al2O3 for selective hydrogenation and CO oxidation. Nat. Commun. 2017, 8, 16100. [Google Scholar] [CrossRef]
  68. Hu, X.; Song, J.; Luo, J.; Zhang, H.; Sun, Z.; Li, C.; Zheng, S.; Liu, Q. Single-atomic Pt sites anchored on defective TiO2 nanosheets as a superior photocatalyst for hydrogen evolution. J. Energy Chem. 2021, 62, 1–10. [Google Scholar] [CrossRef]
  69. Wu, Z.; Hwang, I.; Cha, G.; Qin, S.; Tomanec, O.; Badura, Z.; Kment, S.; Zboril, R.; Schmuki, P. Optimized Pt Single Atom Harvesting on TiO2 Nanotubes—Towards a Most Efficient Photocatalyst. Small 2022, 18, 2104892. [Google Scholar] [CrossRef]
  70. Qin, S.; Denisov, N.; Will, J.; Kolařík, J.; Spiecker, E.; Schmuki, P. A Few Pt Single Atoms Are Responsible for the Overall Co-Catalytic Activity in Pt/TiO2 Photocatalytic H2 Generation. Sol. RRL 2022, 6, 2101026. [Google Scholar] [CrossRef]
  71. 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]
  72. Han, B.; Guo, Y.; Huang, Y.; Xi, W.; Xu, J.; Luo, J.; Qi, H.; Ren, Y.; Liu, X.; Qiao, B.; et al. Strong Metal–Support Interactions between Pt Single Atoms and TiO2. Angew. Chem. Int. Ed. 2020, 59, 11824–11829. [Google Scholar] [CrossRef]
  73. Thang, H.V.; Pacchioni, G.; DeRita, L.; Christopher, P. Nature of stable single atom Pt catalysts dispersed on anatase TiO2. J. Catal. 2018, 367, 104–114. [Google Scholar] [CrossRef]
  74. Therrien, A.J.; Hensley, A.J.R.; Marcinkowski, M.D.; Zhang, R.; Lucci, F.R.; Coughlin, B.; Schilling, A.C.; McEwen, J.S.; Sykes, E.C.H. An atomic-scale view of single-site Pt catalysis for low-temperature CO oxidation. Nat. Catal. 2018, 1, 192–198. [Google Scholar] [CrossRef]
  75. Weon, S.; Suh, M.-J.; Chu, C.; Huang, D.; Stavitski, E.; Kim, J.-H. Site-Selective Loading of Single-Atom Pt on TiO2 for Photocatalytic Oxidation and Reductive Hydrodefluorination. ACS EST Eng. 2021, 1, 512–522. [Google Scholar] [CrossRef]
  76. Fujiwara, K.; Pratsinis, S.E. Single Pd atoms on TiO2 dominate photocatalytic NOx removal. Appl. Catal. B Environ. 2018, 226, 127–134. [Google Scholar] [CrossRef]
  77. Wang, T.; Tao, X.; Li, X.; Zhang, K.; Liu, S.; Li, B. Synergistic Pd Single Atoms, Clusters, and Oxygen Vacancies on TiO2 for Photocatalytic Hydrogen Evolution Coupled with Selective Organic Oxidation. Small 2021, 17, 2006255. [Google Scholar] [CrossRef]
  78. Ge, X.; Zhou, P.; Zhang, Q.; Xia, Z.; Chen, S.; Gao, P.; Zhang, Z.; Gu, L.; Guo, S. Palladium Single Atoms on TiO2 as a Photocatalytic Sensing Platform for Analyzing the Organophosphorus Pesticide Chlorpyrifos. Angew. Chem. Int. Ed. 2020, 59, 232–236. [Google Scholar] [CrossRef]
  79. Kuai, L.; Chen, Z.; Liu, S.; Kan, E.; Yu, N.; Ren, Y.; Fang, C.; Li, X.; Li, Y.; Geng, B. Titania supported synergistic palladium single atoms and nanoparticles for room temperature ketone and aldehydes hydrogenation. Nat. Commun. 2020, 11, 48. [Google Scholar] [CrossRef]
  80. Guo, Y.; Huang, Y.; Zeng, B.; Han, B.; Akri, M.; Shi, M.; Zhao, Y.; Li, Q.; Su, Y.; Li, L.; et al. Photo-thermo semi-hydrogenation of acetylene on Pd1/TiO2 single-atom catalyst. Nat. Commun. 2022, 13, 2648. [Google Scholar] [CrossRef]
  81. Lv, T.; Xiao, B.; Xia, F.; Chen, M.; Zhao, J.; Ma, Y.; Wu, J.; Zhang, J.; Zhang, Y.; Liu, Q. Insights into synergistic effect of Pd single atoms and sub-nanoclusters on TiO2 for enhanced photocatalytic H2 evolution. Chem. Eng. J. 2022, 450, 137873. [Google Scholar] [CrossRef]
  82. Wan, J.; Chen, W.; Jia, C.; Zheng, L.; Dong, J.; Zheng, X.; Wang, Y.; Yan, W.; Chen, C.; Peng, Q.; et al. Defect Effects on TiO2 Nanosheets: Stabilizing Single Atomic Site Au and Promoting Catalytic Properties. Adv. Mater. 2018, 30, 1705369. [Google Scholar] [CrossRef]
  83. 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]
  84. 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]
  85. Wang, Y.; Xu, W.; Chen, X.; Li, C.; Xie, J.; Yang, Y.; Zhu, T.; Zhang, C. Single–atom Ir1 supported on rutile TiO2 for excellent selective catalytic oxidation of ammonia. J. Hazard. Mater. 2022, 432, 128670. [Google Scholar] [CrossRef]
  86. Tang, Y.; Asokan, C.; Xu, M.; Graham, G.W.; Pan, X.; Christopher, P.; Li, J.; Sautet, P. Rh single atoms on TiO2 dynamically respond to reaction conditions by adapting their site. Nat. Commun. 2019, 10, 4488. [Google Scholar] [CrossRef] [PubMed]
  87. Tieu, P.; Yan, X.; Xu, M.; Christopher, P.; Pan, X. Directly Probing the Local Coordination, Charge State, and Stability of Single Atom Catalysts by Advanced Electron Microscopy: A Review. Small 2021, 17, 2006482. [Google Scholar] [CrossRef] [PubMed]
  88. Van der Linden, M.; van Bunningen, A.J.; Amidani, L.; Bransen, M.; Elnaggar, H.; Glatzel, P.; Meijerink, A.; de Groot, F.M.F. Single Au Atom Doping of Silver Nanoclusters. ACS Nano 2018, 12, 12751–12760. [Google Scholar] [CrossRef] [PubMed]
  89. Tyo, E.C.; Vajda, S. Catalysis by clusters with precise numbers of atoms. Nat. Nanotechnol. 2015, 10, 577–588. [Google Scholar] [CrossRef]
  90. Li, J.; Guan, Q.; Wu, H.; Liu, W.; Lin, Y.; Sun, Z.; Ye, X.; Zheng, X.; Pan, H.; Zhu, J.; et al. Highly Active and Stable Metal Single-Atom Catalysts Achieved by Strong Electronic Metal–Support Interactions. J. Am. Chem. Soc. 2019, 141, 14515–14519. [Google Scholar] [CrossRef]
  91. 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]
  92. Qiao, B.; Wang, A.; Yang, X.; Allard, L.F.; Jiang, Z.; Cui, Y.; Liu, J.; Li, J.; Zhang, T. Single-atom catalysis of CO oxidation using Pt1/FeOx. Nat. Chem. 2011, 3, 634–641. [Google Scholar] [CrossRef]
  93. Qin, S.; Denisov, N.; Sarma, B.B.; Hwang, I.; Doronkin, D.E.; Tomanec, O.; Kment, S.; Schmuki, P. Pt Single Atoms on TiO2 Polymorphs—Minimum Loading with a Maximized Photocatalytic Efficiency. Adv. Mater. Interfaces 2022, 9, 2200808. [Google Scholar] [CrossRef]
  94. Wei, T.; Ding, P.; Wang, T.; Liu, L.-M.; An, X.; Yu, X. Facet-Regulating Local Coordination of Dual-Atom Cocatalyzed TiO2 for Photocatalytic Water Splitting. ACS Catal. 2021, 11, 14669–14676. [Google Scholar] [CrossRef]
  95. Zhang, H.; Wei, J.; Dong, J.; Liu, G.; Shi, L.; An, P.; Zhao, G.; Kong, J.; Wang, X.; Meng, X.; et al. Efficient Visible-Light-Driven Carbon Dioxide Reduction by a Single-Atom Implanted Metal-Organic Framework. Angew. Chem. Int. Ed. 2016, 128, 14522–14526. [Google Scholar] [CrossRef]
  96. Yin, P.; Yao, T.; Wu, Y.; Zheng, L.; Lin, Y.; Liu, W.; Ju, H.; Zhu, J.; Hong, X.; Deng, Z.; et al. Single Cobalt Atoms with Precise N-Coordination as Superior Oxygen Reduction Reaction Catalysts. Angew. Chem. Int. Ed. 2016, 55, 10800–10805. [Google Scholar] [CrossRef]
  97. Choi, C.H.; Kim, M.; Kwon, H.C.; Cho, S.J.; Yun, S.; Kim, H.-T.; Mayrhofer, K.J.J.; Kim, H.; Choi, M. Tuning selectivity of electrochemical reactions by atomically dispersed platinum catalyst. Nat. Commun. 2016, 7, 10922. [Google Scholar] [CrossRef] [PubMed]
  98. Qi, P.; Wang, J.; Djitcheu, X.; He, D.; Liu, H.; Zhang, Q. Techniques for the characterization of single atom catalysts. RSC Adv. 2022, 12, 1216–1227. [Google Scholar] [CrossRef] [PubMed]
  99. Simonovis, J.P.; Hunt, A.; Waluyo, I. In situ ambient pressure XPS study of Pt/Cu(111) single-atom alloy in catalytically relevant reaction conditions. J. Phys. D Appl. Phys. 2021, 54, 194004. [Google Scholar] [CrossRef]
  100. Daelman, N.; Capdevila-Cortada, M.; López, N. Dynamic charge and oxidation state of Pt/CeO2 single-atom catalysts. Nat. Mater. 2019, 18, 1215–1221. [Google Scholar] [CrossRef]
  101. Lang, R.; Xi, W.; Liu, J.-C.; Cui, Y.-T.; Li, T.; Lee, A.F.; Chen, F.; Chen, Y.; Li, L.; Li, L.; et al. Non defect-stabilized thermally stable single-atom catalyst. Nat. Commun. 2019, 10, 234. [Google Scholar] [CrossRef]
  102. Huang, D.; de Vera, G.A.; Chu, C.; Zhu, Q.; Stavitski, E.; Mao, J.; Xin, H.; Spies, J.A.; Schmuttenmaer, C.A.; Niu, J.; et al. Single-Atom Pt Catalyst for Effective C–F Bond Activation via Hydrodefluorination. ACS Catal. 2018, 8, 9353–9358. [Google Scholar] [CrossRef]
  103. Tran, N.-D.; Farnesi Camellone, M.; Fabris, S. Probing the Reactivity of Pt/Ceria Nanocatalysts toward Methanol Oxidation: From Ionic Single-Atom Sites to Metallic Nanoparticles. J. Phys. Chem. C 2018, 122, 17917–17927. [Google Scholar] [CrossRef]
  104. Cha, G.; Mazare, A.; Hwang, I.; Denisov, N.; Will, J.; Yokosawa, T.; Badura, Z.; Zoppellaro, G.; Tesler, A.B.; Spiecker, E.; et al. A facile “dark”-deposition approach for Pt single-atom trapping on facetted anatase TiO2 nanoflakes and use in photocatalytic H2 generation. Electrochim. Acta 2022, 412, 140129. [Google Scholar] [CrossRef]
  105. Kottwitz, M.; Li, Y.; Palomino, R.M.; Liu, Z.; Wang, G.; Wu, Q.; Huang, J.; Timoshenko, J.; Senanayake, S.D.; Balasubramanian, M.; et al. Local Structure and Electronic State of Atomically Dispersed Pt Supported on Nanosized CeO2. ACS Catal. 2019, 9, 8738–8748. [Google Scholar] [CrossRef]
  106. Ghosh, T.K.; Nair, N.N. Rh1/γ-Al2O3 single-atom catalysis of O2 activation and CO oxidation: Mechanism, effects of hydration, oxidation state, and cluster size. ChemCatChem 2013, 5, 1811–1821. [Google Scholar] [CrossRef]
  107. Sun, Y.; Wang, Y.; Pan, J.S.; Wang, L.; Sun, C.Q. Elucidating the 4f Binding Energy of an Isolated Pt Atom and Its Bulk Shift from the Measured Surface- and Size-Induced Pt 4f Core Level Shift. J. Phys. Chem. C 2009, 113, 14696–14701. [Google Scholar] [CrossRef]
  108. Greczynski, G.; Hultman, L. Reliable determination of chemical state in X-ray photoelectron spectroscopy based on sample-work-function referencing to adventitious carbon: Resolving the myth of apparent constant binding energy of the C 1s peak. Appl. Surf. Sci. 2018, 451, 99–103. [Google Scholar] [CrossRef]
  109. Greczynski, G.; Hultman, L. The same chemical state of carbon gives rise to two peaks in X-ray photoelectron spectroscopy. Sci. Rep. 2021, 11, 11195. [Google Scholar] [CrossRef] [PubMed]
  110. Biesinger, M.C. Accessing the robustness of adventitious carbon for charge referencing (correction) purposes in XPS analysis: Insights from a multi-user facility data review. Appl. Surf. Sci. 2022, 597, 153681. [Google Scholar] [CrossRef]
  111. Zhang, D.; Li, Y.Y.; Li, Y.Y.; Zhan, S. Towards single-atom photocatalysts for future carbon-neutral application. SmartMat 2022, 1–30. [Google Scholar] [CrossRef]
  112. Barroso, M.; Cowan, A.J.; Pendlebury, S.R.; Grätzel, M.; Klug, D.R.; Durrant, J.R. The Role of Cobalt Phosphate in Enhancing the Photocatalytic Activity of α-Fe2O3 toward Water Oxidation. J. Am. Chem. Soc. 2011, 133, 14868–14871. [Google Scholar] [CrossRef]
  113. Wang, G.; He, C.-T.; Huang, R.; Mao, J.; Wang, D.; Li, Y. Photoinduction of Cu Single Atoms Decorated on UiO-66-NH2 for Enhanced Photocatalytic Reduction of CO2 to Liquid Fuels. J. Am. Chem. Soc. 2020, 142, 19339–19345. [Google Scholar] [CrossRef] [PubMed]
  114. Xiao, M.; Zhang, L.; Luo, B.; Lyu, M.; Wang, Z.; Huang, H.; Wang, S.; Du, A.; Wang, L. Molten-Salt-Mediated Synthesis of an Atomic Nickel Co-catalyst on TiO2 for Improved Photocatalytic H2 Evolution. Angew. Chem. Int. Ed. 2020, 59, 7230–7234. [Google Scholar] [CrossRef] [PubMed]
  115. Xie, J.; Jin, R.; Li, A.; Bi, Y.; Ruan, Q.; Deng, Y.; Zhang, Y.; Yao, S.; Sankar, G.; Ma, D.; et al. Highly selective oxidation of methane to methanol at ambient conditions by titanium dioxide-supported iron species. Nat. Catal. 2018, 1, 889–896. [Google Scholar] [CrossRef]
  116. 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]
  117. Lee, B.-H.; Park, S.; Kim, M.; Sinha, A.K.; Lee, S.C.; Jung, E.; Chang, W.J.; Lee, K.-S.; Kim, J.H.; Cho, S.-P.; et al. Reversible and cooperative photoactivation of single-atom Cu/TiO2 photocatalysts. Nat. Mater. 2019, 18, 620–626. [Google Scholar] [CrossRef]
  118. 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]
  119. Wang, C.; Wang, K.; Feng, Y.; Li, C.; Zhou, X.; Gan, L.; Feng, Y.; Zhou, H.; Zhang, B.; Qu, X.; et al. Co and Pt Dual-Single-Atoms with Oxygen-Coordinated Co–O–Pt Dimer Sites for Ultrahigh Photocatalytic Hydrogen Evolution Efficiency. Adv. Mater. 2021, 33, 2003327. [Google Scholar] [CrossRef]
  120. Wu, X.; Zuo, S.; Qiu, M.; Li, Y.; Zhang, Y.; An, P.; Zhang, J.J.; Zhang, H.; Zhang, J.J. Atomically defined Co on two-dimensional TiO2 nanosheet for photocatalytic hydrogen evolution. Chem. Eng. J. 2021, 420, 127681. [Google Scholar] [CrossRef]
  121. Cai, J.; Cao, A.; Wang, Z.; Lu, S.; Jiang, Z.; Dong, X.-Y.; Li, X.; Zang, S.-Q. Surface oxygen vacancies promoted Pt redispersion to single-atoms for enhanced photocatalytic hydrogen evolution. J. Mater. Chem. A 2021, 9, 13890–13897. [Google Scholar] [CrossRef]
  122. Wang, X.; Zhang, L.; Bu, Y.; Sun, W. Interplay between invasive single atom Pt and native oxygen vacancy in rutile TiO2(110) surface: A theoretical study. Nano Res. 2022, 15, 669–676. [Google Scholar] [CrossRef]
  123. Yang, K.; Zhou, G. Hydrogen evolution/spillover effect of single cobalt atom on anatase TiO2 from first-principles calculations. Appl. Surf. Sci. 2021, 536, 147831. [Google Scholar] [CrossRef]
  124. Bi, H.; Zhang, L.; Wang, Z.; Zhou, G. Identification of active sites available for hydrogen evolution of Single-Atom Ni1/TiO2 catalysts. Appl. Surf. Sci. 2022, 579, 152139. [Google Scholar] [CrossRef]
Figure 1. The parameters affecting the overall photocatalytic water splitting. The gear with the number indicates the order of the photocatalytic process to be successful for overall water splitting. The timescale for the reactions is also shown. Reprinted with permission from Ref [3], Copyright © 2017 American Chemical Society.
Figure 1. The parameters affecting the overall photocatalytic water splitting. The gear with the number indicates the order of the photocatalytic process to be successful for overall water splitting. The timescale for the reactions is also shown. Reprinted with permission from Ref [3], Copyright © 2017 American Chemical Society.
Catalysts 12 00905 g001
Figure 2. Schematic illustration of single-atom catalysts fabrication methods. Reprinted with permission from Ref [32], Copyright © 2020 American Chemical Society.
Figure 2. Schematic illustration of single-atom catalysts fabrication methods. Reprinted with permission from Ref [32], Copyright © 2020 American Chemical Society.
Catalysts 12 00905 g002
Figure 3. Pt on CeO2: (a) Cs-corrected HAADF-STEM, (b) HAADF-STEM, and corresponding elemental mappings images of CeO2-AA-Pt-cal. (c,d) Pt on TiO2: (c) HAADF-TEM image and (d) EDS map of Pt-SA-decorated TiO2 layer. Part (a,b) is adopted with permission from Ref [64] Copyright © 2018 American Chemical Society. Part (c,d) is adopted with permission from Ref [45], Copyright © 2020, John Wiley and Sons.
Figure 3. Pt on CeO2: (a) Cs-corrected HAADF-STEM, (b) HAADF-STEM, and corresponding elemental mappings images of CeO2-AA-Pt-cal. (c,d) Pt on TiO2: (c) HAADF-TEM image and (d) EDS map of Pt-SA-decorated TiO2 layer. Part (a,b) is adopted with permission from Ref [64] Copyright © 2018 American Chemical Society. Part (c,d) is adopted with permission from Ref [45], Copyright © 2020, John Wiley and Sons.
Catalysts 12 00905 g003
Figure 4. (a) XANES spectra of Pt1/Co3O4, Pt1/CeO2, Pt1/ZrO2, and Pt1/graphene single-atom catalysts as well as the Pt foil and PtO2 reference at the Pt L3-edge. (b) The corresponding K2-weighted Fourier transform spectra. With permission from Ref [90], Copyright © 2019 American Chemical Society.
Figure 4. (a) XANES spectra of Pt1/Co3O4, Pt1/CeO2, Pt1/ZrO2, and Pt1/graphene single-atom catalysts as well as the Pt foil and PtO2 reference at the Pt L3-edge. (b) The corresponding K2-weighted Fourier transform spectra. With permission from Ref [90], Copyright © 2019 American Chemical Society.
Catalysts 12 00905 g004
Figure 5. (a) XPS high-resolution spectra of SAs vs NPs for Pt on TiO2, (b) Peak fitting of the Pt SAs of (a). Bi-metallic SAs as exemplified for Pt-Au SAs on different facetted TiO2, showing the (c) Pt4f and (d) Au 4f. Part (a,b) reproduced with permission from Ref [45], Copyright © 2020, John Wiley and Sons. Part (c,d) reproduced With permission from Ref [94], Copyright © 2021 American Chemical Society.
Figure 5. (a) XPS high-resolution spectra of SAs vs NPs for Pt on TiO2, (b) Peak fitting of the Pt SAs of (a). Bi-metallic SAs as exemplified for Pt-Au SAs on different facetted TiO2, showing the (c) Pt4f and (d) Au 4f. Part (a,b) reproduced with permission from Ref [45], Copyright © 2020, John Wiley and Sons. Part (c,d) reproduced With permission from Ref [94], Copyright © 2021 American Chemical Society.
Catalysts 12 00905 g005
Figure 6. (a) The photocatalytic H2 evolution rate of TiO2 loaded with different ratios of Cu SA catalysts as a function of the irradiation time (left side) and right side—the photocatalytic activity of the ca. 1.5 wt% CuSA-TiO2 for six cyclic water splitting experiments and the last run is the activity of the sample after storing in the lab for 380 days, (b) filtered HAADF STEM image and the corresponding line scan profiles, (c) STEM-EDS mapping of Ti, O, and Cu of the fresh CuSA–TiO2. With permission from Ref [116], Copyright © 2021, Springer Nature.
Figure 6. (a) The photocatalytic H2 evolution rate of TiO2 loaded with different ratios of Cu SA catalysts as a function of the irradiation time (left side) and right side—the photocatalytic activity of the ca. 1.5 wt% CuSA-TiO2 for six cyclic water splitting experiments and the last run is the activity of the sample after storing in the lab for 380 days, (b) filtered HAADF STEM image and the corresponding line scan profiles, (c) STEM-EDS mapping of Ti, O, and Cu of the fresh CuSA–TiO2. With permission from Ref [116], Copyright © 2021, Springer Nature.
Catalysts 12 00905 g006
Figure 7. Schematic illustration of charge transfer in dual-site co-catalyzed {001} and {101}-faceted TiO2. Blue, red, gray, yellow, and brown spheres represent Ti, O, Pt, Au, and H atoms, respectively. With permission from Ref [94], Copyright © 2021 American Chemical Society.
Figure 7. Schematic illustration of charge transfer in dual-site co-catalyzed {001} and {101}-faceted TiO2. Blue, red, gray, yellow, and brown spheres represent Ti, O, Pt, Au, and H atoms, respectively. With permission from Ref [94], Copyright © 2021 American Chemical Society.
Catalysts 12 00905 g007
Table 1. Overview of the photocatalytic H2 evolution rate for SA decorated TiO2 nanostructures from literature, including the photocatalytic system (SA metal, SA loading, light source, and reactions system).
Table 1. Overview of the photocatalytic H2 evolution rate for SA decorated TiO2 nanostructures from literature, including the photocatalytic system (SA metal, SA loading, light source, and reactions system).
Type of SAType of SupportSA LoadingLight SourceH2 Evolution RateReaction SystemReference
Pt{001} TiO2 nanosheet0.47 at%325 nm3.2 mL·h−1·cm−250 vol% methanol-water[45]
PtMIL-125(Tiv)(nanoparticle)-300 W Xe lamp14 mmol·h−1·g−120 vol% methanol-water[61]
Pd{001} TiO2 nanosheet0.35 at.%365 nm500 µL·h−150 vol% methanol-water[71]
Au{001} TiO2 nanosheet0.35 at.%365 nm50 µL·h−150 vol% methanol-water[71]
Pt{001} TiO2 nanosheet0.35 at.%365 nm200 µL·h−150 vol% methanol-water[71]
Pd{001} TiO2 nanosheet 0.39 wt%Xe lamp600 µmol·h−1·g−1N,N-dimethylformamide (DMF) 50, 0.5 mL deionized water and 0.5 mL benzylamine[77]
Pt/Au{001} TiO2 nanosheet0.33 wt%300 W Xe lamp60 mmol·h1·g−110 vol% methanol-water[94]
Pt/Au{101} TiO2 nanosheet0.31 wt%300 W Xe lamp15 mmol·h−1·g−110 vol% methanol-water[94]
CuMIL-125(Tiv)(nanoparticle)1.5 wt%simulated solar light irradiation101.7 mmol·g−1·h−1H2O/methanol with a ratio of  =  1:2[116]
CuTiO2 hallow nanoparticles0.75 wt%Xe lamp17 mmol·h−1·g−1water and methanol (volume ratio of 3:1)[117]
Pt/Co{001} TiO2 nanosheetPt 0.16/Co 0.18 wt%300 W Xe lamp45 mmol·h−1·g−110 vol% methanol-water[119]
CoTiO2 nanosheets1.11 wt%300 W Xe lamp8 mmol·h−1·g−11M NaOH solution with 20% methanol[120]
PtTiO2 nanobelts1 wt%simulated solar light (AM 1.5G)90 mmol·mg−150 vol% methanol-water[112]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hejazi, S.; Killian, M.S.; Mazare, A.; Mohajernia, S. Single-Atom-Based Catalysts for Photocatalytic Water Splitting on TiO2 Nanostructures. Catalysts 2022, 12, 905. https://doi.org/10.3390/catal12080905

AMA Style

Hejazi S, Killian MS, Mazare A, Mohajernia S. Single-Atom-Based Catalysts for Photocatalytic Water Splitting on TiO2 Nanostructures. Catalysts. 2022; 12(8):905. https://doi.org/10.3390/catal12080905

Chicago/Turabian Style

Hejazi, Seyedsina, Manuela S. Killian, Anca Mazare, and Shiva Mohajernia. 2022. "Single-Atom-Based Catalysts for Photocatalytic Water Splitting on TiO2 Nanostructures" Catalysts 12, no. 8: 905. https://doi.org/10.3390/catal12080905

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