Preprint
Review

Recent Progress on Ligand Protected Metal Nanoclusters in Photocatalysis

Altmetrics

Downloads

200

Views

76

Comments

0

A peer-reviewed article of this preprint also exists.

Submitted:

05 May 2023

Posted:

06 May 2023

You are already at the latest version

Alerts
Abstract
The reckless use of non-replenishable fuels by the growing population for energy and the resultant incessant emission of hazardous gases and waste products into the atmosphere insisted scientists to fabricate materials capable of managing these global threats at once. In recent studies, photo-catalysis employed focus on utilizing renewable solar energy to initiate chemical processes with the aid of semiconductors and highly selective catalysts. A wide range of nanoparticles have showcased promising photocatalytic properties. Ultrasmall metal nanoclusters (MNCs) with sizes below 2 nm, stabilized by ligands shows discrete energy levels, exhibits unique optoelectronic properties which are vital to photocatalysis. In this review, we intend to compile information on varying photocatalytic efficiency of metal NCs with respect to changes in their properties. Review discusses about the photocatalytic activity of atomically precise ligand-protected MNCs and its hybrids in the domain of in energy conversion processes like photodegradation of dyes, Oxygen Evolution Reaction (ORR), Hydrogen Evolution Reaction (HER) and CO2 Reduction Reaction (CO2RR), and photosensitizer in solar cell.
Keywords: 
Subject: Chemistry and Materials Science  -   Physical Chemistry

1. Introduction

Over the past several decades, rising demands of the ever-growing population, excessive consumption of non-renewable resources and resulting greenhouse gas emissions and improper waste disposal have all been of great concern to mankind. The relentless efforts of scientific community towards this global crisis paved the way for more sustainable solutions like photocatalysis. Photocatalysis appears to be a perfect fit as a greener alternative to resolve energy and pollution related problems simultaneously, it aids the completion of chemical processes with the help of inexhaustible solar power, mitigates hazardous products like CO2 by photoreduction and produces cleaner fuels. In 1972, Fujishima and Honda were first to apply this strategy for hydrogen evolution reaction through photocatalysis[1]. Later, photocatalysis with semiconductors with sufficient bandgaps (i.e., TiO2, CdS, ZnO, Fe2O3 and ZnS) did turn into a promising field of study for researchers[2,3,4,5,6]. Nanotechnology has created revolutionary changes in this dimension, with the help of umpteen number of nanoparticles (NPs) which possess high surface area and plasmonic resonance. Noble metal nanoparticles such as AuNP, AgNP, PtNP, IrNP, OsNP, RhNP, RuNP are largely employed for photocatalysis for the past decades, but the efficiency is not reached the target level[7]. Apart from this, the mechanisms related information in photocatalysis driven by noble metal NPs with huge number of surface atoms did seem vague. Meanwhile, new class of zero-dimensional fluorophores, metal nanoclusters (NCs), proved dominant to their conventional nanoparticle analogues in various applications.
Metal NCs comprises of a core with ten to hundreds of atoms protected by surface ligands like thiols, proteins, peptides, enzymes, polymers, and DNA [8,9,10,11]. The emergence of this well-defined aggregates with commendable surface-to-volume ratio, fully reduced atomic core and dispersity on catalysts made them the best available choice for wide range of applications [12]. In comparison to bulk, MNCs being a sub-nanometer sized unit (< 2nm) exhibits quantum confinement effects, as it approaches Fermi wavelength of conduction electrons, thereby splits continuous density of states into discrete energy levels. As a result, metal NCs possess molecule-like behavior and acts as a missing link between atoms/molecules and metal nanoparticles [13]. The atomic level precision of metal NCs assisted in confining size and to analyze involved mechanisms quite understandable. Due to these factors, they have been assigned few other titles too such as Quantum clusters (QCs) and Monolayer protected clusters [14,15]. Initially, gold MNCs were the focus of study due to the simple synthetic routes, commendable stability and novel optical properties presented, even in the absence of plasmonic resonance, are all attributed to their shift in valency from Au (III) to Au (I)/Au(0) in Au NCs [16].
The physiochemical properties of nanoclusters are highly influenced by the size, metal core composition, assembly architecture and surface components [17,18,19]. In addition, metal NCs are known for their attractive optical characteristics including tunable luminescence, HOMO–LUMO transitions, substantial Stokes shift, two photon absorption, photostability, magnetism, chirality and biocompatibility [20,21]. Their photostability, biocompatibility and low cytotoxicity in turn caused them to be chosen for biomedical applications [22]. Monolayer protected clusters are less aggressive and relatively more stable compared to gas phase clusters containing unsatisfied valence electrons in its free state. Currently, further studies on the factors and their impacts on optical properties set forth new developments in catalysis, biosensing, bioimaging, gene therapy and drug delivery [13,23,24,25,26]. Advancements in the preparation of polar or organic solvent soluble NCs, penetrable sized clusters, computational strategies to gain knowledge on surface atom arrangements and instrumentational techniques can build the future of metal NCs. Tailoring cluster’s distinctive optical and electronic properties by enhancing parameters like formal charges, geometry, metal composition and ligand, play pivotal role in in photocatalysis [27]. Despite these, metal NCs with fully reduced metal atom cores have relatively low-band gap, thus it indeed inhibits photo-corrosion appreciably. Therefore, metal NCs can be utilized, as photosensitizer and cocatalyst, in energy intensive processes like photodegradation of dyes, ORR, HER and CO2RR, photosensitizer in solar cell [28,29,30,31]. (Figure 1)
A comprehensive review on metal nanomaterials for heterogeneous catalysis has been written by Liu and Corma [12]. Jianping Xie and colleauges highlighted in a minireview the major important characteristics of MNCs which are vital to photo and electro catalysis [27]. This review aims to update the literature and recent advances with a focus on photocatalysis particularly from 2016 onwards, since a comprehensive review on catalytic applications of ligand-protected, atomically precise MNCs was published in Coordination Chemistry Journal in 2016 to which the readers are referred to [32].

1.1. Chemical Composition and structural properties

The central metal atom, ligands, charge states and composition play significant role in deciding the physiochemical properties of MNCs. In general, MNCs expressed by the molecular formula [Mn (SR)m]^q (n- represents the number of metal atoms, m- represents the thiolate ligands (staple mottif), and q denotes the net charge of the cluster) [33]. Engineering metal, ligand and charge state on an atomic level can alter the performances and physicochemical properties of MNCs. In addition, the size (~2nm) and structure can be adjusted with atomic precision for new application possibilities in a variety of disciplines [34]. By analogy with terminology employed in protein field, metal NCs is composed of primary, secondary and tertiary structures. whereas metallic core is the primary structure and the repetitive local structural motifs serve as a bridge between the core and the ligands. These motifs are organic ligands and surprisingly, the length or size and structure of these motifs can be easily manipulated, resulting in a wide range of metal NCs. The protective shell's exterior structure is made up of spatial ligands. It has been highlighted that insight of the crystal structure of the material is very crucial as it reveals the information on the atomicity of the core, nature of the Au–S linkage, chirality of the nanoclusters, the arrangements of the ligands around the metal core, and so on [35]. In the recent times, crystal structures of some of the gold nanoclusters such as Au102SR44,[36] Au25SR18,[36] Au38SR24,[37] Au36SR24,[38] [Au24 (PPh3 )10 (SR)5 Cl2]+,[39] Au28SR20,[40] (where SR= thiolate) were resolved using various analytical techniques [41].
The molecular weight of MNCs is identified using mass spectrometry based methods, like matrix-assisted laser-desorption ionization-time of flight mass spectrometry (MALDI-TOF-MS) and electrospray ionization mass spectrometry (ESI-MS) [42,43]. For MALDI analysis traditional weak organic acid used as matrices to recognize the molecular mass of the core of the metal NCs; Even though it’s a sophisticated analytical instrument, it also has some limitations. The weak organic acid matrices employed in MALDI do not eliminate ligand fragmentation, The mass accuracy of MS equipment is also decreased by the presence of significant chemical noise and/or an unresolved isotopic pattern, leads to the possibility of numerous molecular formulae being nearly isobaric and the potential for impurities or ligand fragmentation to produce significant chemical noise that makes it difficult to assign atomic information. For instance, isotope-resolved mass spectrometry has been pushed forwards by Antoine’s group to unravel the molecular formula of ultrasmall NCs [44].

Synthesis of Metal Nanocluster

Gas phase cluster is the first reported nanocluster, where the MNCs are formed by evaporation, and it is unprotected. This unprotected clusters are observed to be very reactive to form larger sized particles. Therefore, proper stabilization techniques should be used for MNC synthesis. The selection of ligand is an important step in the controlled synthesis of the MNCs. The first ligand chosen for the synthesis of MNCs was phosphine due to its high affinity towards metal ions. Followed by phosphine stabilized cluster, various thiol stabilized NCs were reported due to the relatively high affinity of Sulphur towards metal ion. Due to the insolubility of organic thiol stabilized NCs in water, water-soluble thiols were introduced to synthesize the NCs. Later new possibilities has raised to create quantum clusters using macromolecule template. In this green synthesis macromolecules such as various DNA, polymers and dendrimers, have been used for cluster stabilization and protection. Reaction duration, pH, temperature, type of ligand, template structure, reducing agent concentration, and Au3+/ligand ratio are crucial synthetic parameters for influencing the structure, size, surface characteristics, oxidation state, and consequently the optical properties of MNCs. It is necessary that the template or ligand used for the stabilization should have a high binding affinity towards the metals (Au, Ag, Pt etc) in order to prepare highly stable AuNCs with high monodispersed (Figure 2) [45]. Figure 2 depicts core-shell nature of MNCs and can be prepared by selecting appropriate capping ligands [46]. The nanoclusters can be synthesised using different ways such as chemical reduction, photoreduction, hydrothermal, biomineralization and etching, etc[47].
Xie et al. first time developed a biomineralization strategy for the synthesis of highly luminescent gold nanoclusters using BSA as protecting and reducing agent [48]. The red luminescent AuNCs is made up of 25 gold atoms (Au25). The same strategy is used for the synthesis of various other protein protected metal NCs [49,50]. Thus, in general, the nanocluster synthesis can be categorized as Top-down approach, bottom-up approach and Inter cluster conversion approach.
Bottom-up method uses metal salt, ligands and reducing agents are used as precursors. In this method, nanoclusters are formed by wet chemical reduction of metal salts with a suitable reducing agent. In the first stage, a Metal(I)_thiolate complex formed by reacting metal salt with thiolate ligand. Then, Metal(I)_thiolate complex is treated with a reducing agent such as sodium borohydride (NaBH4) or ascorbic acid to reduce M(I) to M(0), and produce M(0)@M(I)-based NCs [51]. In some circumstances, the stabilizing ligand itself serves as a reducing agent, which eliminates the need for a second reducing agent. This technique of synthesis is known as the biomineralization method. Other than wet chemical reduction, photoreduction method is also employed to produce luminescent metal nanoclusters, which in light initiate the reduction reaction.
In Top-Down approach, smaller sized MNCs can be prepared from larger sized metal nanoparticle/MNCs via chemical etching. Here, chemical etching carried out between excess ligands and larger metal nanoclusters to obtain metal NCs with smaller sizes. This usually done in solution phase either in one solution or in the interface of two solution. Based on the use of etching agent, etching method is classified into two categories, ligand etching and solvent etching. One of the major advantage of nanoclusters produced by etching method is the controlled size focusing [52].
In the case of Inter-Cluster conversion method, the NCs formed via seed-mediated synthesis, cluster conversion, metal exchange ligand exchange, and motif exchange. In this process, nanoclusters are used as the starting material, and through the adjustment kinetic or thermodynamic parameters the structure of the nanoclusters are changed. The most common methods for cluster conversion is the ligand exchange reaction [53].Ligand exchange reaction (LER) is a widely used technique for modifying nanoclusters after their creation. The adaptability of the gold and Sulphur interphase makes this possible. The ability to change the size and phase of clusters, as well as the ability to impart fluorescence on nanoclusters for biological labelling purposes, are some advantages of ligand exchange reactions. They can also increase the enantiomeric excess of already chiral clusters and give chirality to nonchiral clusters. In this way, LER broadens the range of MNCs by forming distinctive and precise nanoclusters [54]
Similarly, an electrochemical method can be used to produce a large quantity of high-quality Au25(SR)18 0 NC’s [55]. This method of crystallisation can aid in determining the structure of new NCs, enabling a deeper comprehension of their molecular physiochemical characteristics.

1.2. Key physico-chemical properties of MNCs in photocatalysis

Metal NCs with specific structural design and resultant exceptional physiochemical, electronic and optical properties gave birth to a plethora of distinctive materials capable of driving photochemical reactions. Especially, tailoring their optical and electronic parameters could effectively alter their light harvesting abilities and would render us a method to manipulate certain photocatalytic processes, by inducing electron-hole pairs generation, in order to attain maximum quantum yield. For photocatalysis, the major criteria to choose a materials is strong absorption of solar light [56]. Figure 3 shows the absorption spectrum and solar cells excited with AM 1.5 (100 mW cm−2) of Au25SG18. The overlapping region between these two curves is a key component of photocatalysis. For gold nanoclusters, efficient light energy conversion necessitates a slow rate of excited state relaxation. Indeed, Au18SG14 exhibited the highest electron transfer rate and longest excited state lifetime of the NC series [56].

1.3. Optical Properties

Metal nanoparticles, having size range ~3 nm to 100 nm, obey the plasmonic regime where a single surface plasmon resonance band dominates the optical spectrum and which has a far higher number of atoms than the number of surface molecules that stabilize it (see Figure 4). In such regime (plasmonic regime), nanoparticles interact with the incident light and surface plasmon resonance occurs, causing conduction band electrons to collectively oscillate and forms characteristic peaks (gold nanoparticle shows surface plasmon resonance at 520 nm) [57,58]. On the contrary, the NCs have a very small number of metal atoms (such as 2, 8, 18, 25, 55 etc.), results in an optical spectrum with several bands, distinctive energy levels, and quantum behavior [59].
Figure 4 depicts the electrical transitions between discrete energy levels in metal NCs [61,62]. As the particle diameter decreases, SPR peaks disappear gradually as shown in the Figure 4. In the Kohn-Sham orbital energy diagram of Au25SR18 nanoclusters indicates HOMO composed of triply degenerated 5d10 atomic orbitals and doubly degenerated LUMO, which are 6sp atomic orbitals of gold [63] (Figure 4). The peak ‘a’ formed due to the HOMO (sp) to LUMO (sp), peak ‘b’ is a mixed type band (sp to sp and d to sp) transition. The peak ‘c’ is formed due to (d to sp) interband transition (Figure 4).
The photoluminescence property of metal NCs opens up possibilities and applications in variety of fields including sensing, imaging etc. [64,65].The luminescence in MNCs are depends on its geometric and electrical structures of the metal core, as well as the ligand shell, alteration in the metal core or ligands, one can tune its optical property. Such nanoclusters can be thought of as a "multi-shell system," comprising a metallic core containing metal-metal bond, a metal-ligand staple motif as interface, and surface ligand molecules (see Figure 4). As observed in metal complexes, metal nanocluster also exhibits the similar charge transfer or electron transfer process. These three shells can interact with each other via ligand to metal core or ligand to metal-metal charge transfer or by directly donating delocalized electrons from the ligands' electron-rich groups [11]. Following visible or near-UV absorption, certain broad tendencies in de-excitation routes have been discovered. The experimental and theoretical data from the literature were used to produce the energy diagram in Figure 5. It is possible for transitions from molecular orbitals with strong ligand contributions to orbitals with high metal character (LMCT) and from metal-to-metal electronic transitions to cause near-ultraviolet and visible absorption. Clusters with a core of metal atoms have an initial decay pathway that could result in an emission in the visible spectrum. All clusters have a charge-transfer component with a long-lived decay [66]. Furthermore, intersystem crossings are associated with multiple energy transfers (reinforced ISC), may result in an overall increase in Photoluminescence (PL) emission and longer PL lifetimes [59]. The luminescence property is affected by temperature in such a way that, on decreasing temperature the HOMO-LUMO gap increases [51].
Two-photon Absorption: Two-photon-absorbing (TPA) photocatalyst uses near-infrared low-energy photons [67] for photocatalysis. The unique Nonlinear optical (NLO) properties of sub-nanometer core sized clusters of metal nanoclusters exhibits outstanding characteristics [21]. Compared to the conventional organic dyes, NMNCs shows superior two-photon absorption cross section (TPA). Although these two-photon processes in small numbered MNCs are well established, the basic photophysical mechanisms behind them are still poorly understood. Generalizations based on complementary theoretical and experimental studies have been made possible by their two-photon absorption properties [68]. It is reported that, the TPA characteristic of AuNCs and AgNCs can be enhanced by the concept of “ligand-core NLO-phore” [69,70]. The different electron or charge transfer process between metallic core and ligand are essential for boosting the transition dipole moments leading to enhanced TPA cross sections (as exemplified in Figure 6 with Au10L10 nanoclusters).
Chirality: The photocatalytic activity induced by chiral metallic particles varies with the helicity of the illumination light [71]. MNCs such as gold quantum clusters display fascinating chiral properties [72]. Moreover, the chirality depends mainly on three factors, (a) fundamental chirality is induced first by atomic packing mode, (b) the geometrical isomerisms of the surrounding motifs and (c) the natural chiral characteristics of the protective ligands and their arrangements covering the metal core [51] (Figure 7). Commonly, the SR-Au-SR units of AuNCs display chirality, a chiral center at each Sulphur atom and can exist in either a (R,R)/(S,S) trans-configuration or a (R,S) cis-configuration

1.4. Stability

Metal clusters shielded by thiolates or polymers are attractive possibilities for nanoscale devices [74]. For instance, among various GSH-protected AuNCs, Au25SR18 is the most stable one [75]. The isolated Au55 clusters exhibit an amazing resistance to oxidation, even when exposed to oxygen atoms and radicals produced by oxygen plasmas, [76]. When the particle size is decreased to 1.6 nm, the metallic component's spectral weight increases rapidly, indicating that these particles are more oxidation-resistant. This resistance is much more for the case of naked Au55 clusters. In addition, ligands influence on the stability of metal NCs such that thermal stability of captopril-protected Au25 is more when compared to GSH-protected AuQCs [77].
NMR measurements [Figure 8(A, B and C, D)] of equal amounts of glutathione and captopril after heating for 2 hrs. at 80ºC showed no significant spectral changes for captopril, while new peaks were observed for glutathione due to degradation. This proves the enhanced stability of Au25(Capt)18 over Au25(SG)18.
Doping the central metallic core Au25 with other elements like Cu, Pd alters the geometric structure and increased the resistance to deterioration, respectively [78]. To illustrate this, Negishi, Y., Kurashige and co-workers [79] doped the Au25(SG)18 with Pd leading to the formation of Pd1Au24(SC12H25)18. The Pd doped gold nanoclusters exhibited increased thermodynamic stability and stability against degradation.
The spectral profile evolves with time, and after 30 days, it resembles that of Pd1Au24(SC12H25)18 rather closely and the ESI MS spectrum shows just a single peak that may be ascribed to Pd1Au24(SC12H25)18 [80]. The strong interaction energy between Pd and Au24(SC12H25)18 was proved by the DFT calculation of Jiang et al. Consequently, with the Au24(SC12H25)18 frame, Pd develops an intermetallic structure and on the strength of this Pd1Au24(SC12H25)18 presents higher thermodynamic stability than Au25(SC12H25)18. On the other hand, doping the Au25(SC2H4Ph)18 with Copper reduced its overall nanocluster stability [81]. Doping with copper formed Cu~1Au~24(SC2H4Ph)18. In comparison to Au25(SC2H4Ph)18, the optical absorption spectra moved toward lower energy, suggesting that the HOMO-LUMO gap has shrunk.
From the electronic density of states study of Au25(SR)181- reveals its electronic state of Au13 core (8 electrons) protected by six [(SR)3Au2] complexes [82]. The superior stability for Au25(RS)18 is achieved by structural toughness and an 8 e- shell of delocalized Au(6s) electrons for the anion. The gold nanocluster super atom model expands the jellium model [83] to explain multiple ligand-protected AuNCs with closed valence-electron shells (such as, 2, 8, 18, 20, 34, 40, etc.). The metal core is regarded as a single atom in the super atom hypothesis. The extraordinarily stability of these nanoclusters emerge from the regular closure of outer electronic shells [84]. Even if strong structural and electronic stabilization occurs for Au25(RS)18, the nature of protecting ligands affect the stability of the cluster. The role of isomer were studied using para, meta, and ortho MBA stabilized nanoclusters [85]. Figure 9 shows the isomeric effect of MBA on the stability of Au25(MBA)18NCs. The steric hindrance of the carboxylic groups plays major role. m-MBA and o-MBA ligands stabilized nanocluster shows noticeable steric hindrance at the gold core's surface, which leading to significantly lowering the binding energy required for the fragmentation of Au-S bond

2. Application of Nanoclusters in Photocatalysis

Over the course of time, research on relevance of cleaner fuels, alternate potential energy sources as well as tools to alleviate pollution, have been gainful. Research on visible light photocatalysis over the past decade draws much attention in this line. Semiconductor heterogeneous nanotechnology based photocatalysts lasted long enough and showcased their prominence in this specialty. At present, material chemists are busy preparing competent monometallic or hybrid entities by combining conventional semiconductors with suitable pre-catalyst/co-catalysts to bring about the best results in photocatalysis. Metal NCs have proved their supremacy in multiple economical and sustainable processes of photocatalysis such as photodegradation of organic pollutants, photocatalytic H2 splitting, photoreduction of CO2, oxidation and hydrogenation reactions etc.,
ROS e-/h+ fast recombination of e-/h+ of semiconductors, noble metal/metallic/non-metallic moieties on semiconductors support extending photocatalytic activity towards visible range.
The following section discusses about the role of metal NCs in photocatalytic as cocatalyst as well as catalyst in various light conversion processes.

2.1. Photodegradation of organic pollutants

Globalization led industrialization, to satisfy the needs of world’s population, has led to the steep rise in harmful pollutants in the environment at an alarming rate. The organic pollutants from textile industries majorly constitute azo dyes while inorganic pollutants hold oxidized heavy metals. These under-treated materials containing carcinogenic effluents are being dumped into the soil and water, have already created damages beyond repair to humans as well as to the aquatic flora and fauna. Initiating action plans to safely remove them without the formation of other stable secondary pollutants demanded scrutiny. Methyl Orange (MO), Rhodamine B (RB), Malachite Green (MG) and Methylene blue (MB), some of the commonly found dyes, that have been present in industrial effluents. Reactive oxygen species (ROS) such as superoxide radicals (O2-.) and hydroxyl radicals (OH-.) are the prime contributors and starting materials in the degradation of these toxic materials [86].
Titania (TiO2), ZnO, SiO2, Nb2O5 and so on have been reported widely as a solution for pollution initiated by organic pollutants through photodegradation [87,88]. These semiconductor systems ensure photostability, low cost, mere toxicity and essential level of oxidative ability. However, the disadvantages of these semiconductors remain a barrier to photocatalytic reactions due to the large bandgap that they offer and it indeed requires UV irradiation for the electron-hole generation, this restricts their use as effective catalyst. As the research studies progressed, metal nanoparticle incorporated versions semiconductors became the focal point. Metal NCs linked to semiconductor systems as support have profoundly influenced the works in photocatalytic degradation of cationic and anionic dye pollutants as they could enable us to engineer the bandgap width [89]
Zhu and co-workers prepared AuNCs coupled with toroid structured per-6-thio-β-cyclodextrins placed on TiO2 surface (TiO2-Au NCs@β-CD) provided a better space for the Au cores to interact with the incoming pollutants through a host-guest interaction trap [90]. This synergistic effect between the nanocluster metallic core, peculiar ligand cavities and support system increased the photodegradation rate of methyl orange (MO) to 98% in 10 minutes. Concurrently, the integrated material’s rate constant values for photodegradation (1st cycle=0.31min-1, 5th cycle= 0.15min-1) was phenomenal in comparison with TiO2(1st cycle= 0.12min-1) even after five cycles. Sharma and coworkers studied of Au-TiO2 conjugated nano assembly were used as light harvester under visible light [91].
The mechanism for photocatalytic degradation of dyes using Au-TiO2 can explained by following reaction pathways
Au + hν → Au•
Au• + TiO2 → h+(Au) + e-(TiO2)
e-(TiO2) + O2 → TiO2 + •O2-
•O2- + H+ → HOO•
HOO• + e-(TiO2) + H+ → H2O2 + TiO2
H2O2 + e-(TiO2) → •OH + OH- + TiO2
h+(Au) + H2O → Au + H+ + •OH
h+(Au) + OH- → Au + • OH-
Organic pollutants + O2-• or •OH- → CO2 + H2O +….
Gowswami et.al created colloidal nanocomposite material with TiO2/Nb2O5 conjugated to silver NCs with captopril as ligand. They have varied the niobium loading in the incorporated product, to investigate the adhesion of the material towards cationic and anionic dyes as well as their photodegradation capability [92]. As per the analysis, the mentioned nanocomposite with 48.1% niobium loading and Ag NCs resulted in ternary junction which narrowed down the recombination rates and the surface acidity caused the Ag NCs ligands end groups to polarize COO-, thereby attracting the cationic dyes for effective 100% degradation with high photostability. Spectroscopic techniques validate the fact that Ag NCs are the reason behind absorption in visible light region and themselves act as oxidative sites, also initiates electron transfer from valence band of the NCs to conduction band of nanocomposites. Samai’s group made polyethelene imine template Ag NCs incorporated with CeO2 nanoparticle for degrading Acridin Red dye photocatalytically [93] (Figure 10). XPS mechanistic work and radical experiments pin down their radical pathway leading to 42% to 80% degradation in 2 hours with 1.07% and 3.10% Ag NCs loading respectively. The reduction potential and two oxidational potential vs NHE values were recorded with Cyclic Voltammogram to be -0.64V, 0.07V and 1.03V respectively (Figure 10).
ZnO nanoparticle decorated with Ag NCs were established as a prospective candidate for Orange II (OII) dye degradation under both UV and white light by Rodriguez and coworkers [94]. Atomic Force Microscopy (AFM) confirmed the deposition of NCs moiety on ZnO rather than substitution of metallic core itself. Additionally, the optimal loading range of Ag on ZnO was found to be 1.3% w/w and it certainly ensures ample number of interaction sites for the pollutants to approach ZnO nanoparticles (Figure 11). Figure 11 shows the spectroscopic and morphological characterization of AgNC decorated ZnO nanoparticle. Vilar-Vidal et.al formulated green emitting stable closed shell electronic structured and recyclable CuNCs to degrade MB under UV and Visible light irradiation [95]. UV-Vis data and administered multiple photoluminescence emission peaks were an indication of mixture of clusters and LDI-TOF spectrometry approves proposed structural formulas of Cu NCs: [Cu18(CH3COO)(OH)]-2 and [Cu34O2(CH3COO)3N(C4H9)3Na]-2.
Light steered preparation of 3-Mercaptopropyl trimethoxy silane (MPTS) stabilized Au NCs and their performance in photodegradation of MB were examined by Zhou et.al [96]. It was observed that the nanoclusters decolorizes the dye progressively and color fades completely in 60 mins.
Man Cao and his coworkers synthesized a self-assembled silver nanocluster for photocatalytic degradation of sulphur mustard simulant (2-chloroethyl ethyl sulfide, CEES), a toxic vesicant against human proteins and DNA, can cause skin blisters, eye and respiratory system irritation, and even fatal damage. The silver cluster assembled material is prepared using a photosensitizer (5,10,15,20-tetra(4-pyridyl)porphyrin (TPyP) as the organic linker and which linked with 12-core silver chalcogenolate cluster to form [Ag12(St Bu)6(CF3COO)3(TPyP)]n, (designated as Ag12TPyP). They reported 98 % degradation of CEES with 1% loading [97,98].
Wen and co-workers developed a highly stable core-shell type catalyst for photo redox reactions. The photostability of nanocluster were improved by loading it on SiO2 sphere by utilizing multifunctional branched poly-ethylenimine (BPEI) as a surface charge modifying, reducing and stabilizing agent. Then TiO2 were coted SiO2-Au GSH clusters-BPEI to form SiO2-Au GSH clusters-BPEI@TiO2 core- shell structure further significantly improve the photocatalytic efficiency for dye degradation of organic dye (Rh B) [99].

2.2. Oxidation and hydrogenation processes

Application of NCs on oxidation and reduction processes are still underway. Oxidation and hydrogenation reactions in most reported cases specifically depend on their electronic structures as electron-hole separation remains the key to it. Researchers have carried out typical oxidation-reduction reactions like high selectivity styrene oxidation [100], aerobic oxidation of amines to imines [101], cyclohexane or phenol derivatives [102], water splitting [103], catalytic hydrogenation [104], pollutants treatment [105], in the presence of metal NCs, especially gold NCs (AuNCs). In 1987, Haruta and co-workers were the first to initiate oxidation of CO with few atom AuNCs along with α-Fe2O3, NiO and Co3O4 at low temperatures[106]. After two decades, mechanistic view of photocatalysis of Au25 nanoclusters with TiO2 as support under visible and near infrared emissions were illustrated by Kogo and co-workers [107]. They attached AuNCs to TiO2 so that the excited electrons could be transferred to the conduction band of TiO2 with ease and can aid in reduction of Ag+ in the counter electrode and the generated holes (h+) could perform oxidation of donors (phenol derivatives or formic acid) in working electrode (Figure 12).
Similarly, Zhu et.al did a comparative study to figure out the capabilities of a set of superatoms to oxidize styrene- Au25(SR)18, Au38(SR)24 and Au144(SR)60 with diameters of 1nm, 1.3nm and 1.6nm respectively [108]. However, their studies reassured the size and electronic structure dependence of photocatalytic abilities and Au25 NCs gave the highest average of ~27 ± 1.0% as smaller the superatom, higher is the HOMO- LUMO gap (1.3eV). Later, Chen & group loaded [Au25(PPh3)10(SR)5Cl2] on P25 to covert benzylamines to imines, TOF was recorded to be 152 2h-1 for 4-methylbenzamine and other amines too showed appreciable conversions (73-99%) [101]. The attempts to learn the possible conversion route using Time Dependent Density Functional Theory (TD-DFT) calculations, Fourier-transform ion cyclotron resonance mass spectrometry (FT-IC-MS) with Electrospray ioinization (ESI) and trapping intermediates with scavengers (K2S2O8 and ammonium oxalate) confirmed the presence of +2 charge on the cluster. Hence the photocatalytically induced electron could lead to oxygen radical formation and thereby persuade conversions. Studies with organic supports too have gained momentum over the years. Gold nanocluster with organic supports (Polyvinylpyrrolidone (PVP) & Polyperoxyacetic acid (PAA)) were analyzed by H.Tsunoyama and co-workers and X-ray photoelectron spectroscopy (XPS) & X-ray absorption near edge structure (XANES) reinforced the existence of negatively charged cluster which led to the production of superoxo- or peroxo- radicals required to jump start the photocatalytic reactions [109]. Their activities and recyclability were keenly observed for CO oxidation. In addition, these set of reactions enlightened research community about the prime role of nanoclusters in oxidation and reduction reactions.
Hamoud and co-workers studied photocatalytic activity Ag NCs encapsulated into zeolite ( ZX-Bi zeolite) for photooxidation of methanol under visible light. They have found that the Ag/ZX-Bi exhibits very low activity, compared to activated sample at 200 °C (Ag/ZX-Bi_200) [110].

2.3. Photocatalytic H2 production

The demand for green energy sources better suited for the environment is rising due to the reduction and environmental pollution generated by conventional energy resources. While the distinct qualities shape, Hydrogen gas as an acceptable renewable energy source for a sustainable future [111]. Ahluwalia and coworkers studied the fuel economy of Hydrogen-fuelled fuel cell (H2-FCV) vehicles and common gasoline-fuelled passenger cars in 2003 [112]. In 1972, Honda and Fujishima et al. first time discovered that water can be divided into H2 and O2 in presence of light. There after various photocatalyst were developed and studied for photocatalytic water splitting reaction. A. Fujishima and K. Honda, Nature, 1972, 238, [36,37].
Catalyst based on the metals such as Ti, Cobalt, Nickel, Iron, Molybdenum are widely used for as catalyse for the electrochemical H2 production [113]. Thereafter, the catalytic efficiency of metal oxides, metal nitrides were enhanced by incorporating suitable cocatalyst on these semiconducting materials. The Co-catalytic activity of metal nanoclusters was studied in the recent past due to its excellent optical, electronic and catalytic activity. The nanoclusters were incorporated into various semiconductors to enhance the catalytic activity by suppressing the electron-hole recombination rate.
Kamat et al studied the photoelectrochemical and photocatalytic production of H2 using GSH stabilised AuNCs-TiO2 film hybrid system under visible light irradiation [113].
Liu et al carried out a detailed investigation on the photocatalytic activity of AuNCs as cocatalyst in highly ordered nanoporous layer-covered TiO2 nanotube arrays (NP-TNTAs). The photocatalytic activity of NP-TNTAs/AuNCs analysed by monitoring photodegradation of organic dyes, photocatalytic reduction of aromatic nitro compounds and photoelectrochemical water splitting [114].
The quantum confinement of MNCs enables the charge transfer, easy adsorption and desorption of intermediates and thereby fastens the photocatalytic hydrogen evolution reactions (HER) [29].
Glutathione stabilised AuNCs (Au25(SG)18) were incorporated on BaLa4Ti4O15 used for photocatalytic water splitting. The catalytic activity of sub nanometer sized Au- BaLa4Ti4O15 of compared with larger sized gold nanoparticle and found 2.6 times higher catalytic activity for AuNC composite [115]. Similarly, AuNCs loaded on SrTiO3 were studied for HER and found the enhanced catalytic activity of SrTiO3 in presence AuNCs as Cocatalyst [116]. Recently Hanieh Mousavi and Co-workers studied the photocatalytic production of H2 using AuNCs as Co-catalyst. They have prepared a nanocomposite (Au101NC-AlSrTiO3 -rGO) containing AuNCs, RGO and AlSrTiO3 and the composite showed enhanced photocatalytic water splitting under UV light irradiation [117].
Sub-nanosized clusters of silver adsorbed on specific sites of Gold nanorods (GNRs) disturb the growth symmetry of Au facets and leads to anisotropy [118]. In the presence of hole scavenger namely ethanol and in the absence of electron scavengers like O2, the photoelectrons accumulate in GNRs with Ag clusters. Moreover, on low UV light irradiation, Ag3 clusters at a concentration of 0.43µg within the GNRs shows high H2 production efficiency of 10%. Similarly, loading of 1 wt% of sub- nm Au clusters in CdS uplifts the photocatalytic H2 production about 35 times than the unmodified CdS under visible light [119]. The comparison of sub-nm Au loaded CdS with Pd/CdS and Rh/CdS of similar size revealed that the 3-nm sized Au/CdS as a better co-catalyst.
Likewise monolayer niobate (HTi2NbO7) nanosheets with Pt nanoclusters proved themselves as potential photocatalysts for high H2 production [120]. Monolayer niobate nanosheets enable the charge separation between photoelectron and holes and displays 5 times higher photocatalytic H2 production than that observed in its layered form. 1 wt% Pt nanoclusters loaded to niobate nanosheets via photoreduction presented increased activity due to the close contact between the HTi2NbO7 nanosheets and Pt nanocluster. On light irradiation electrons migrated to the conduction band moves to surface of nanosheets. Since Pt have a higher work function than niobate nanosheets, these electrons in the surface of nanosheets transfers to Pt and form H2.
Moreover, by monitoring nanocluster size, the photocatalytic activity of the whole system can be varied. Heiz, Feldmann and co-workers modified cadmium sulfide (CdS) nanorod with a series of Pt NCs such as Pt8NC, Pt22NC, Pt34NC, Pt46NC, Pt68NC and found that, Pt46/CdS exhibits the highest activity for photocatalytic water splitting, due to the well-known quantum confinement effect., were the bandgap increases with a decrease of the NC size [121]. Rongchao Jin and his coworkers reported a detailed review on the effect of size of nanocluster in photocatalysis [122] While Cheng and co-workers[118] revealed the effective co-catalytic effect of sub nano sized Pt-Au alloy clusters in photocatalytic H2 evolution. The synergistic effect of 0.5 wt% of both Pt and Au clusters dispersed in TiO2 (Pt-Au/T) creates increased charge separation and 80.1 µmol h-1 of H2 is evolved with a quantum efficiency 50% at 365nm. As like HER, similar strategy have been established for enhanced efficiency for OER, by using nanocluster as cocatalyst [123]. Au25 NC- CoSe2 composite were studied for OER activity found enhanced OER activity in presence of Au25/CoSe2 obtained current density of 10 mA cm–2 at small overpotential of ∼0.43 V (cf. CoSe2: ∼0.52 V). The ligand and gold cluster size can also tune the catalytic performance of the composites.
The photocatalytic activity of Co-catalyst was studied by monitoring doping of heteroatom on the nanoclusters. Negishi et al. demonstrated that doping of Pt on Au25 NC enhances the water splitting activity, while Pd doping reduces the water splitting activity. They proposed that the doping position plays the role in the catalytic activity. The doped Pd is located at the surface of the metal-cluster cocatalyst, whereas Pt is located at the interface between the metal-cluster cocatalyst and the photocatalyst [124]. Yang et al. also illustrated the effect of heteroatom doping on the photocatalytic activity of PtAg24-loaded graphitic carbon nitride (PtAg24/g-C3N4) and found that PtAg24/g-C3N4 shows higher efficiency for photocatalytic H2 production than Ag25/g-C3N4 alone [125]. In addition to the above studies, scientist are exploring the efficiency of photocatalytic water using composites including metal NCs with g-C3N4 nanosheets,TiO2, etc [126].

2.4. Photocatalytic CO2 reduction

Nowadays, reducing greenhouse gas emissions from various Photovoltaic (PV) systems has become one of the scientific community's primary concerns [127]. Carbon dioxide is one of the chief greenhouse gases that influence the heat content of earth’s atmosphere [1]. With a focus on reducing carbon dioxide emission, novel technologies are adopted for the production of commodity chemicals by using carbon dioxide as feedstock [128].The conversion of carbon dioxide to value-added chemicals or to other hydrocarbon fuels such as methane, ethylene and carbon monoxide by utilizing energy from non-fossil resources such as solar energy not only increases the carbon recycling but also assists in fuel production [129,130]. Solar-driven transformation of carbon dioxide to useful products could be achieved through two major approaches such as photocatalytic and electrochemical CO2 reduction processes [131]. In 1978, M. Halmann utilized p-type semiconductors for photo-electrochemical reduction of carbon dioxide [132]. Methanol and Carbon monoxide obtained from the conversion of carbon dioxide are identified as useful feedstocks [133]. The former is utilized for synthesising other hydrocarbon fuels while latter is used for Fischer–Tropsch syntheses.
CO2 is a thermodynamically stable molecule and catalysts assist the electrochemical CO2 reduction reactions and aid in achieving desired product [134]. MNCs possess ultrafine structure, electronic and optical properties [34] and function both as electrocatalyst and photocatalysts [27]. Various features of MNCs such as size, core, composition, surface ligands, charge state, geometry influences its electro and photocatalytic behaviors [27]. Colombo Jr and coworkers studied femtosecond electron-hole recombination in TiO2- nanoclusters and explained the intracluster dynamics [135]. The study demonstrates the steps involved in electron trapping, recombination and formation of long-lived species. Kauffman et. al studied the reversible weak interaction between CO2 and Au25(SC2H4Ph)18- clusters [136]. The electrochemical reduction of CO2 using Au25 catalyst in aqueous 0.1 M KHCO3 showed maximum CO production at -1.0 V with 100% Faradaic efficiency. The electrochemical CO2 reduction performance of silver nanoclusters confined in bovine serum albumin (AgNC@BSA) is enhanced via polyoxometalates [α-SiW12O40]4- [137].
The presence of suitable photocatalyst possessing features such as high light absorptive power, convenient catalytic sites and low activation energy enhances the photocatalytic reduction of CO2 [138]. The small-scale size of MNCs of about 2nm, interfacial surface, energy gap, tunable chemical properties, quantum confinement are the advantages of ultrafine MNCs for CO2 reduction over metal nanoparticles [139]. Titanium dioxides or titania are widely used semiconductors and have broad range of applications including photoreduction of CO2 [140]. Doping and decoration with other elements or metal ions strengthens its photocatalytic activity by reducing its band gap. In addition to these elements, attaching nanoclusters to TiO2 makes it a potent visible light photocatalyst [141]. Inserting Ti3+ ions into the TiO2 creates isolated states in the presence of UV and visible light. The electrons are trapped in these states and due to recombination with charge carriers’ photocatalytic activity is decreased. On combining Ti3+ introduced TiO2 with nanoclusters of Cu (II) oxides (i.e., Cu (II)-TiO2@Ti3+), electrons from the isolated states of Ti3+ moves to surface of Cu (II) nanoclusters. Hence, the photocatalytic activity of Cu (II)-TiO2@Ti3+ under visible light is raised and decomposed gaseous 2-propanol (IPA) completely to yield ~18 µmol CO2 generation. On UV irradiation, copper oxide (CuxO) nanoclusters incorporated in strontium titanate nanorod thin film [142] and niobate sheets [143] showed selective CO production form photocatalytic CO2 reduction.
Product selectivity towards CH4 and CO is shown by brookite TiO2 quasi nanocubes (BTN) on surface decoration with Cu-nanoclusters (Cu-BTN) under xenon lamp irradiation [144]. XRD diffraction peaks exhibits the presence of Cu-nanoclusters only on the surface of BTN. The total consumed electron number (TCEN) is utilized for examining overall photocatalytic CO2 reduction. At a 1.5 % optimum concentration of Cu-nanoclusters in BTN, maximum photoactivity with TCEN of 150.9 µmolg-1h-1 and highest production rate of 4.23 µmolg-1h-1 CO and 17.81 µmolg-1h-1 CH4 is observed (Figure 13).
In situ DRIFTS IR spectra suggested CO32- intermediate for CO formation and HCO3- for CH4 (Figure 14a). Cui at.el worked on the role of bridging ligand and metal ions grafted to Au nanoclusters in photocatalytic conversion of CO2 to CO [145]. The functionalization of L-cysteine with glutathione protected Au nanocluster (Au-GSH NCs) helps in binding of metal cations such as Fe2+ Co2+ Ni2+ and Cu2+ and thus improves the selective CO production. Under visible light along with CO2 and H2O, Co2+ cation within Au nanocluster (Auc-C-Co) at optimum concentration of 4 µmol exhibited maximum CO production of 3.45 µmol.gcat-1h-1. Similarly, via 3-mercaptopropionic acid (MPA), Co2+ is grafted to surface of Au nanocluster (Auc-MPA-Co) and exhibited high-rise in photocatalytic activity compared to (Auc-C-Co) through strong interlinkage between S-metal cation (Figure 15).
Zhang and co-workers [143] developed a quasi-ternary complex consisting of polymethacrylic acid stabilized Ag nanoclusters (AgNCs-PMAA), carbon monoxide dehydrogenase (CODH) and TiO2 nanoparticles and under visible light CO2 reduction results generation of CO with 20 s-1 turnover frequency at 25℃ . Jiang, Y et. al [146] upgraded the chemical stability of gold nanoclusters (Au-NCs) by combining with metal organic framework (UiO-68) through N-Heterocyclic carbene stabilizing ligands (NHC) and denoted as AuNC@UiO-68-NHC. The photocatalytic activity AuNC@UiO-68-NHC is observed to be enhanced as a result of strong covalent bond formation between AuNCs and UiO-68 facilitated by Au-NHC bridges. This enabled easy movement of excited electrons from the Fermi levels of Au-NCs to conduction band of UiO-68-NHC and thus the recombination of photogenerated electrons and holes are reduced (Figure 16) [146].
The selective production of CO and presence of CH4 and H2 as side product is observed. Billo and coworkers developed effective photocatalyst with dual sites by making oxygen vacancies in Ni-nanoclusters loaded in black TiO2 (Ni/TiO2[Vo]) [147]. Ni and oxygen defective sites acts as dual sites, lessen the C-O bond strength and supports the separation of charge carriers (Figure 17). On light irradiation from halogen lamp, Ni/TiO2[Vo] produced 10 µmol.g-cat-1 of acetaldehyde which is 18 times higher than TiO2. Thus the study refers to a different approach for enhancing photocatalytic CO2 reduction by introducing active dual sites in photocatalyst.
Recently, Mohamad El-Roz et al de prepared a silver nanocluster-based catalyst for the conversion of formic acid to CO2 and H2 under visible light irradiation. Here the nanocluster was incorporated into nano sized zeolite crystal 145.
Table 1. Summary of Literature Reports on the use of nanocluster in Catalysis.
Table 1. Summary of Literature Reports on the use of nanocluster in Catalysis.
Sl No Photocatalyst Co catalyst Application Efficiency Reference
1 TiO2 AuNC Photodegradation of dyes 98% degradation in 10 minutes exposure. [90]
2 TiO2/Nb2O5 AgNC Photodegradation of dyes. 100% degradation [92]
3 CeO2 AgNC Photodegradation of dyes. 80% degradation in 2 hours [93]
4 ZnO nanoparticle AgNC Photodegradation of dyes. 100% degradation in 1 hr [94]
5 CuNC No Cocatalyst Photodegradation of dyes. 100% degradation in 1 69hr [95]
6 AuNC No cocatalyst Photodegradation of dyes. 100% degradation in 1 hr [96]
7 AgNC-assembled materials No cocatalyst Photodegradation of dyes. 98% degradation [98]
8 TiO2 core- shell SiO2-Au GSH clusters Photodegradation of organic dyes. 99.1% degradation in 0.5 hr [99]
9 TiO2 Au25 NC Oxidation of phenol derivatives and ferrocyanide and reduction of Ag+, Cu2+ and oxygen [107]
10 Au25NC No Catalyst oxidation of styrene and hydrogenation of α,β-unsaturated ketone 27 ± 1.0% [108]
11 TiO2 AuNC Oxidation of benzylamines to imines 73-99% [101]
12 Zeolite (ZX-Bi zeolite) Ag NC Photooxidation of methanol 49.60 mmol·g–1·cm–2 after 12 h of reaction [110]
13 TiO2 AuNC Production of H2 0.3 mmol of hydrogen/h/g [148]
14 TiO2 nanotube arrays (NP-TNTAs). AuNC Photodegradation of organic dyes, photocatalytic reduction of aromatic nitro compounds and photoelectrochemical water splitting. [114]
15 BaLa4Ti4O15 AuNCs Photocatalytic water splitting 190 µmol/h [115]
16 SrTiO3 AuNC Hydrogen evolution reaction 41.2 µmol/h of H [116]
17 AlSrTiO3 and rGO AuNCs photocatalytic production of H2, photocatalytic water splitting 385 ± 22 nmol h−1 [117]
18 Gold nanorods (GNRs) AgNCs Hydrogen evolution reaction 10% [118]
19 Monolayer niobate (HTi2NbO7) Pt NC Higher H2 production 10 μmol h−1 [120]
20 Modified cadmium sulfide (CdS) nanorod Pt NC Photocatalytic water splitting 1.5‰ h−1 [121]
21 BaLa4Ti4O15 Au24Pd NCs and Au24Pt NCs photocatalytic H2 evolution 100-150 µmolh-1 [124]
22 Graphitic carbon nitride (g-C3N4) PtAg NC photocatalytic H2 production 39.7 µmolh-1 [125]
23 Brookite TiO2 quasi nanocubes CuNCS Photocatalytic CO2 reduction 150.9 μmol g−1 h−1 [144]
24 Metal cations -Fe2+, Co2+, Ni2+ and Cu2+ Au NCs Photocatalytic CO2 reduction 3.54 µmol⋅gcat−1⋅h−1 [145]
25 carbon monoxide dehydrogenase (CODH) and TiO2 nanoparticles Ag NC Photocatalytic CO2 reduction turnover frequency of 20 s−1 [149]
26 Au-NCs Photocatalytic CO2 reduction 57.6 μmol g-1h-1 [146]
27 TiO2 Ni-NCs photocatalytic CO2 reduction 10 µmol g-cat−1 [147]
28 AgNC No Cocatalyst reforming of formic acid to H2
and CO2
99% selectivity [150]

3. Conclusion and Future Perspective

Metal NCs are an under-explored class of nanomaterials with unimaginable grade of potential. The review summarizes recent progress in the metal nanocluster based photocatalyst for various photocatalytic reactions, correlating it to their structure and properties to boost single metal, alloy hybrid system’s and nanocomposite systems effectiveness. However, research on metal NCs hasn’t surpassed embryonic stage yet. In particular, preventing aggregation, effective capping agents or stabilizers, optical & electronic structures, metal-support bonding, mechanistic learning, in situ characterization techniques, structural model theories and appreciable yield yearn for more attention. Gold and AgNCs still dominate a major fraction of scientific publications in this domain, works on other transition metals could turn out to be more productive and cost-effective. Recyclability of metal NCs in photocatalysis with adequate stability after multiple cycles have to be extensively analyzed. At the same time, computer simulations could open the world of possibilities of metal NCs to us and recent additions to the software have proven their effectiveness. Of note, NCs can plausibly be the answer for low carbon fuels, a milestone to sustainable development and an affordable photocatalyst.

Acknowledgements

Meegle S Mathew thanks University Grant omission (UGC)DR DS Kothari Fellowship for financial support. In addition, the authors deeply acknowledge French National Centre for Scientific Research (CNRS) for funding through International Emerging Actions (IEA) between Institut Lumière Matière, CNRS, France and Mahatma Gandhi University, India.

References

  1. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [CrossRef]
  2. Bakbolat, B.; Daulbayev, C.; Sultanov, F.; Beissenov, R.; Umirzakov, A.; Mereke, A.; Bekbaev, A.; Chuprakov, I. Recent Developments of TiO2-Based Photocatalysis in the Hydrogen Evolution and Photodegradation: A Review. Nanomaterials 2020, 10, 1790. [CrossRef]
  3. Cheng, L.; Xiang, Q.; Liao, Y.; Zhang, H. CdS-Based Photocatalysts. Energy Environ. Sci. 2018, 11, 1362–1391. [CrossRef]
  4. Ong, C.B.; Ng, L.Y.; Mohammad, A.W. A Review of ZnO Nanoparticles as Solar Photocatalysts: Synthesis, Mechanisms and Applications. Renewable and Sustainable Energy Reviews 2018, 81, 536–551. [CrossRef]
  5. Hitam, C.N.C.; Jalil, A.A. A Review on Exploration of Fe2O3 Photocatalyst towards Degradation of Dyes and Organic Contaminants. Journal of Environmental Management 2020, 258, 110050. [CrossRef]
  6. Lee, G.-J.; Wu, J.J. Recent Developments in ZnS Photocatalysts from Synthesis to Photocatalytic Applications — A Review. Powder Technology 2017, 318, 8–22. [CrossRef]
  7. Li, J.; Lou, Z.; Li, B. Nanostructured Materials with Localized Surface Plasmon Resonance for Photocatalysis. Chinese Chemical Letters 2022, 33, 1154–1168. [CrossRef]
  8. Song, X.-R.; Goswami, N.; Yang, H.-H.; Xie, J. Functionalization of Metal Nanoclusters for Biomedical Applications. Analyst 2016, 141, 3126–3140. [CrossRef]
  9. Goswami, N.; Zheng, K.; Xie, J. Bio-NCs – the Marriage of Ultrasmall Metal Nanoclusters with Biomolecules. Nanoscale 2014, 6, 13328–13347. [CrossRef]
  10. Basu, S.; Paul, A.; Antoine, R. Controlling the Chemistry of Nanoclusters: From Atomic Precision to Controlled Assembly. Nanomaterials 2021, 12, 62. [CrossRef]
  11. Wu, Z.; Jin, R. On the Ligand’s Role in the Fluorescence of Gold Nanoclusters. Nano Lett. 2010, 10, 2568–2573. [CrossRef]
  12. Liu, L.; Corma, A. Metal Catalysts for Heterogeneous Catalysis: From Single Atoms to Nanoclusters and Nanoparticles. Chem. Rev. 2018, 118, 4981–5079. [CrossRef]
  13. Du, X.; Jin, R. Atomically Precise Metal Nanoclusters for Catalysis. ACS Nano 2019, 13, 7383–7387. [CrossRef]
  14. Khandelwal, P.; Poddar, P. Fluorescent Metal Quantum Clusters: An Updated Overview of the Synthesis, Properties, and Biological Applications. J. Mater. Chem. B 2017, 5, 9055–9084. [CrossRef]
  15. Templeton, A.C.; Wuelfing, W.P.; Murray, R.W. Monolayer-Protected Cluster Molecules. Acc. Chem. Res. 2000, 33, 27–36. [CrossRef]
  16. Qu, X.; Li, Y.; Li, L.; Wang, Y.; Liang, J.; Liang, J. Fluorescent Gold Nanoclusters: Synthesis and Recent Biological Application. Journal of Nanomaterials 2015, 2015, 1–23. [CrossRef]
  17. Yang, T.-Q.; Peng, B.; Shan, B.-Q.; Zong, Y.-X.; Jiang, J.-G.; Wu, P.; Zhang, K. Origin of the Photoluminescence of Metal Nanoclusters: From Metal-Centered Emission to Ligand-Centered Emission. Nanomaterials 2020, 10, 261. [CrossRef]
  18. Antoine, R. Supramolecular Gold Chemistry: From Atomically Precise Thiolate-Protected Gold Nanoclusters to Gold-Thiolate Nanostructures. Nanomaterials 2020, 10, 377. [CrossRef]
  19. Kolay, S.; Bain, D.; Maity, S.; Devi, A.; Patra, A.; Antoine, R. Self-Assembled Metal Nanoclusters: Driving Forces and Structural Correlation with Optical Properties. Nanomaterials 2022, 12, 544. [CrossRef]
  20. Su, Y.; Xue, T.; Liu, Y.; Qi, J.; Jin, R.; Lin, Z. Luminescent Metal Nanoclusters for Biomedical Applications. Nano Res. 2019, 12, 1251–1265. [CrossRef]
  21. Bonačić-Koutecký, V.; Antoine, R. Enhanced Two-Photon Absorption of Ligated Silver and Gold Nanoclusters: Theoretical and Experimental Assessments. Nanoscale 2019, 11, 12436–12448. [CrossRef]
  22. Genji Srinivasulu, Y.; Yao, Q.; Goswami, N.; Xie, J. Interfacial Engineering of Gold Nanoclusters for Biomedical Applications. Mater. Horiz. 2020, 7, 2596–2618. [CrossRef]
  23. Antoine, R.; Maysinger, D.; Sancey, L.; Bonačić-Koutecký, V. Open Questions on Proteins Interacting with Nanoclusters. Commun Chem 2022, 5, 47. [CrossRef]
  24. Combes, G.F.; Vučković, A.-M.; Perić Bakulić, M.; Antoine, R.; Bonačić-Koutecky, V.; Trajković, K. Nanotechnology in Tumor Biomarker Detection: The Potential of Liganded Nanoclusters as Nonlinear Optical Contrast Agents for Molecular Diagnostics of Cancer. Cancers 2021, 13, 4206. [CrossRef]
  25. Rudzińska, M.; Daglioglu, C.; Savvateeva, L.V.; Kaci, F.N.; Antoine, R.; Zamyatnin Jr, A.A. Current Status and Perspectives of Protease Inhibitors and Their Combination with Nanosized Drug Delivery Systems for Targeted Cancer Therapy. DDDT 2021, Volume 15, 9–20. [CrossRef]
  26. Borghei, Y.; Hosseinkhani, S.; Ganjali, M.R. Bridging from Metallic Nanoclusters to Biomedical in Understanding Physicochemical Interactions at the Nano–Bio Interface. Part & Part Syst Charact 2022, 39, 2100202. [CrossRef]
  27. Chai, O.J.H.; Liu, Z.; Chen, T.; Xie, J. Engineering Ultrasmall Metal Nanoclusters for Photocatalytic and Electrocatalytic Applications. Nanoscale 2019, 11, 20437–20448. [CrossRef]
  28. Liang, H.; Liu, B.-J.; Tang, B.; Zhu, S.-C.; Li, S.; Ge, X.-Z.; Li, J.-L.; Zhu, J.-R.; Xiao, F.-X. Atomically Precise Metal Nanocluster-Mediated Photocatalysis. ACS Catal. 2022, 12, 4216–4226. [CrossRef]
  29. Munir, A.; Joya, K.S.; Ul haq, T.; Babar, N.; Hussain, S.Z.; Qurashi, A.; Ullah, N.; Hussain, I. Metal Nanoclusters: New Paradigm in Catalysis for Water Splitting, Solar and Chemical Energy Conversion. ChemSusChem 2019, 12, 1517–1548. [CrossRef]
  30. Kauffman, D.R.; Thakkar, J.; Siva, R.; Matranga, C.; Ohodnicki, P.R.; Zeng, C.; Jin, R. Efficient Electrochemical CO 2 Conversion Powered by Renewable Energy. ACS Appl. Mater. Interfaces 2015, 7, 15626–15632. [CrossRef]
  31. Al Dosari, H.M.; Ayesh, A.I. Nanocluster Production for Solar Cell Applications. Journal of Applied Physics 2013, 114, 054305. [CrossRef]
  32. Fang, J.; Zhang, B.; Yao, Q.; Yang, Y.; Xie, J.; Yan, N. Recent Advances in the Synthesis and Catalytic Applications of Ligand-Protected, Atomically Precise Metal Nanoclusters. Coordination Chemistry Reviews 2016, 322, 1–29. [CrossRef]
  33. Serhan, M.; Jackemeyer, D.; Long, M.; Sprowls, M.; Diez Perez, I.; Maret, W.; Chen, F.; Tao, N.; Forzani, E. Total Iron Measurement in Human Serum With a Novel Smartphone-Based Assay. IEEE J. Transl. Eng. Health Med. 2020, 8, 1–9. [CrossRef]
  34. Yao, Q.; Yuan, X.; Chen, T.; Leong, D.T.; Xie, J. Engineering Functional Metal Materials at the Atomic Level. Adv. Mater. 2018, 30, 1802751. [CrossRef]
  35. Mathew, A.; Pradeep, T. Noble Metal Clusters: Applications in Energy, Environment, and Biology. Part. Part. Syst. Charact. 2014, 31, 1017–1053. [CrossRef]
  36. Jadzinsky, P.D.; Calero, G.; Ackerson, C.J.; Bushnell, D.A.; Kornberg, R.D. Structure of a Thiol Monolayer-Protected Gold Nanoparticle at 1.1 Å Resolution. Science 2007, 318, 430–433. [CrossRef]
  37. Qian, H.; Eckenhoff, W.T.; Zhu, Y.; Pintauer, T.; Jin, R. Total Structure Determination of Thiolate-Protected Au 38 Nanoparticles. J. Am. Chem. Soc. 2010, 132, 8280–8281. [CrossRef]
  38. Zeng, C.; Qian, H.; Li, T.; Li, G.; Rosi, N.L.; Yoon, B.; Barnett, R.N.; Whetten, R.L.; Landman, U.; Jin, R. Total Structure and Electronic Properties of the Gold Nanocrystal Au 36 (SR) 24. Angew. Chem. Int. Ed. 2012, 51, 13114–13118. [CrossRef]
  39. Das, A.; Li, T.; Nobusada, K.; Zeng, Q.; Rosi, N.L.; Jin, R. Total Structure and Optical Properties of a Phosphine/Thiolate-Protected Au 24 Nanocluster. J. Am. Chem. Soc. 2012, 134, 20286–20289. [CrossRef]
  40. Zeng, C.; Li, T.; Das, A.; Rosi, N.L.; Jin, R. Chiral Structure of Thiolate-Protected 28-Gold-Atom Nanocluster Determined by X-Ray Crystallography. J. Am. Chem. Soc. 2013, 135, 10011–10013. [CrossRef]
  41. Jin, R.; Zeng, C.; Zhou, M.; Chen, Y. Atomically Precise Colloidal Metal Nanoclusters and Nanoparticles: Fundamentals and Opportunities. Chem. Rev. 2016, 116, 10346–10413. [CrossRef]
  42. Lu, Y.; Chen, W. Application of Mass Spectrometry in the Synthesis and Characterization of Metal Nanoclusters. Anal. Chem. 2015, 87, 10659–10667. [CrossRef]
  43. Chakraborty, P.; Pradeep, T. The Emerging Interface of Mass Spectrometry with Materials. NPG Asia Mater 2019, 11, 48. [CrossRef]
  44. Comby-Zerbino, C.; Dagany, X.; Chirot, F.; Dugourd, P.; Antoine, R. The Emergence of Mass Spectrometry for Characterizing Nanomaterials. Atomically Precise Nanoclusters and Beyond. Mater. Adv. 2021, 2, 4896–4913. [CrossRef]
  45. Halawa, M.I.; Lai, J.; Xu, G. Gold Nanoclusters: Synthetic Strategies and Recent Advances in Fluorescent Sensing. Materials Today Nano 2018, 3, 9–27. [CrossRef]
  46. Chen, L.-Y.; Wang, C.-W.; Yuan, Z.; Chang, H.-T. Fluorescent Gold Nanoclusters: Recent Advances in Sensing and Imaging. Anal. Chem. 2015, 87, 216–229. [CrossRef]
  47. Hayyan, M.; Hashim, M.A.; AlNashef, I.M. Superoxide Ion: Generation and Chemical Implications. Chem. Rev. 2016, 116, 3029–3085. [CrossRef]
  48. Xie, J.; Zheng, Y.; Ying, J.Y. Protein-Directed Synthesis of Highly Fluorescent Gold Nanoclusters. J. Am. Chem. Soc. 2009, 131, 888–889. [CrossRef]
  49. Le Guével, X.; Hötzer, B.; Jung, G.; Hollemeyer, K.; Trouillet, V.; Schneider, M. Formation of Fluorescent Metal (Au, Ag) Nanoclusters Capped in Bovine Serum Albumin Followed by Fluorescence and Spectroscopy. J. Phys. Chem. C 2011, 115, 10955–10963. [CrossRef]
  50. Ling, S.; Cui, X.; Zhang, X.; Liu, B.; He, C.; Wang, J.; Qin, W.; Zhang, Y.; Gao, Y.; Bai, G. Glutathione-Protected Gold Nanocluster Decorated Cadmium Sulfide with Enhanced Photostability and Photocatalytic Activity. Journal of Colloid and Interface Science 2018, 530, 120–126. [CrossRef]
  51. Liu, Z.; Wu, Z.; Yao, Q.; Cao, Y.; Chai, O.J.H.; Xie, J. Correlations between the Fundamentals and Applications of Ultrasmall Metal Nanoclusters: Recent Advances in Catalysis and Biomedical Applications. Nano Today 2021, 36, 101053. [CrossRef]
  52. Li, Y.; Zaluzhna, O.; Xu, B.; Gao, Y.; Modest, J.M.; Tong, Y.J. Mechanistic Insights into the Brust−Schiffrin Two-Phase Synthesis of Organo-Chalcogenate-Protected Metal Nanoparticles. J. Am. Chem. Soc. 2011, 133, 2092–2095. [CrossRef]
  53. Meng, X.; Xu, Q.; Wang, S.; Zhu, M. Ligand-Exchange Synthesis of Selenophenolate-Capped Au25 Nanoclusters. Nanoscale 2012, 4, 4161. [CrossRef]
  54. Wang, Y.; Bürgi, T. Ligand Exchange Reactions on Thiolate-Protected Gold Nanoclusters. Nanoscale Adv. 2021, 3, 2710–2727. [CrossRef]
  55. Antonello, S.; Dainese, T.; Pan, F.; Rissanen, K.; Maran, F. Electrocrystallization of Monolayer-Protected Gold Clusters: Opening the Door to Quality, Quantity, and New Structures. J. Am. Chem. Soc. 2017, 139, 4168–4174. [CrossRef]
  56. Stamplecoskie, K.G.; Swint, A. Optimizing Molecule-like Gold Clusters for Light Energy Conversion. J. Mater. Chem. A 2016, 4, 2075–2081. [CrossRef]
  57. Robinson, D. Synthesis and Characterization of Metal Nanoclusters Stabilized by Dithiolates. [CrossRef]
  58. Kreibig, U.; Vollmer, M. Optical Properties of Metal Clusters; Springer Series in Materials Science; Springer Berlin Heidelberg: Berlin, Heidelberg, 1995; Vol. 25; ISBN 978-3-642-08191-0.
  59. Antoine, R.; Bonačić-Koutecký, V. Liganded Silver and Gold Quantum Clusters. Towards a New Class of Nonlinear Optical Nanomaterials; SpringerBriefs in Materials; Springer International Publishing: Cham, 2018; ISBN 978-3-319-64742-5.
  60. Jin, R.; Higaki, T. Open Questions on the Transition between Nanoscale and Bulk Properties of Metals. Commun Chem 2021, 4, 28. [CrossRef]
  61. Huang, X.; Li, Z.; Yu, Z.; Deng, X.; Xin, Y. Recent Advances in the Synthesis, Properties, and Biological Applications of Platinum Nanoclusters. Journal of Nanomaterials 2019, 2019, 1–31. [CrossRef]
  62. Muhammed, M.A.H.; Verma, P.K.; Pal, S.K.; Kumar, R.C.A.; Paul, S.; Omkumar, R.V.; Pradeep, T. Bright, NIR-Emitting Au 23 from Au 25 : Characterization and Applications Including Biolabeling. Chem. Eur. J. 2009, 15, 10110–10120. [CrossRef]
  63. Zhu, M.; Aikens, C.M.; Hollander, F.J.; Schatz, G.C.; Jin, R. Correlating the Crystal Structure of A Thiol-Protected Au 25 Cluster and Optical Properties. J. Am. Chem. Soc. 2008, 130, 5883–5885. [CrossRef]
  64. Shang, L.; Dong, S.; Nienhaus, G.U. Ultra-Small Fluorescent Metal Nanoclusters: Synthesis and Biological Applications. Nano Today 2011, 6, 401–418. [CrossRef]
  65. Mathew, M.S.; Joseph, K. Green Synthesis of Gluten-Stabilized Fluorescent Gold Quantum Clusters: Application As Turn-On Sensing of Human Blood Creatinine. ACS Sustainable Chem. Eng. 2017, 5, 4837–4845. [CrossRef]
  66. Bertorelle, F.; Wegner, D.; Bakulić, M.P.; Fakhouri, H.; Comby-Zerbino, C.; Sagar, A.; Bernadó, P.; Resch-Genger, U.; Koutecky, V.B.; Guével, X.L.; et al. Tailoring NIR-II Photoluminescence of Single Thiolated Au25 Nanoclusters by Selective Binding to Proteins; In Review, 2021;
  67. Tang, J.-H.; Han, G.; Li, G.; Yan, K.; Sun, Y. Near-Infrared Light Photocatalysis Enabled by a Ruthenium Complex-Integrated Metal–Organic Framework via Two-Photon Absorption. iScience 2022, 25, 104064. [CrossRef]
  68. Russier-Antoine, I.; Bertorelle, F.; Calin, N.; Sanader, Ž.; Krstić, M.; Comby-Zerbino, C.; Dugourd, P.; Brevet, P.-F.; Bonačić-Koutecký, V.; Antoine, R. Ligand-Core NLO-Phores: A Combined Experimental and Theoretical Approach to the Two-Photon Absorption and Two-Photon Excited Emission Properties of Small-Ligated Silver Nanoclusters. Nanoscale 2017, 9, 1221–1228. [CrossRef]
  69. Antoine, R. Ligand-Core NLO-Phores: Two-Photon Absorption and Two-Photon Excited Emission Properties of Atomically Precise Clusters of Gold and Silver. In Molecular Spectroscopy—Experiment and Theory; Koleżyński, A., Król, M., Eds.; Challenges and Advances in Computational Chemistry and Physics; Springer International Publishing: Cham, 2019; Vol. 26, pp. 139–160 ISBN 978-3-030-01354-7.
  70. Sanader, Ž.; Krstić, M.; Russier-Antoine, I.; Bertorelle, F.; Dugourd, P.; Brevet, P.-F.; Antoine, R.; Bonačić-Koutecký, V. Two-Photon Absorption of Ligand-Protected Ag 15 Nanoclusters. Towards a New Class of Nonlinear Optics Nanomaterials. Phys. Chem. Chem. Phys. 2016, 18, 12404–12408. [CrossRef]
  71. Hao, C.; Xu, L.; Ma, W.; Wu, X.; Wang, L.; Kuang, H.; Xu, C. Unusual Circularly Polarized Photocatalytic Activity in Nanogapped Gold-Silver Chiroplasmonic Nanostructures. Adv. Funct. Mater. 2015, 25, 5816–5822. [CrossRef]
  72. Schaaff, T.G.; Knight, G.; Shafigullin, M.N.; Borkman, R.F.; Whetten, R.L. Isolation and Selected Properties of a 10.4 KDa Gold:Glutathione Cluster Compound. J. Phys. Chem. B 1998, 102, 10643–10646. [CrossRef]
  73. Bertorelle, F.; Russier-Antoine, I.; Calin, N.; Comby-Zerbino, C.; Bensalah-Ledoux, A.; Guy, S.; Dugourd, P.; Brevet, P.-F.; Sanader, Ž.; Krstić, M.; et al. Au 10 (SG) 10 : A Chiral Gold Catenane Nanocluster with Zero Confined Electrons. Optical Properties and First-Principles Theoretical Analysis. J. Phys. Chem. Lett. 2017, 8, 1979–1985. [CrossRef]
  74. Tsukuda, T.; Häkkinen, H. Introduction. In Frontiers of Nanoscience; Elsevier, 2015; Vol. 9, pp. 1–7 ISBN 978-0-08-100086-1.
  75. Shichibu, Y.; Negishi, Y.; Tsunoyama, H.; Kanehara, M.; Teranishi, T.; Tsukuda, T. Extremely High Stability of Glutathionate-Protected Au25 Clusters Against Core Etching. Small 2007, 3, 835–839. [CrossRef]
  76. Boyen, H.-G.; Kästle, G.; Weigl, F.; Koslowski, B.; Dietrich, C.; Ziemann, P.; Spatz, J.P.; Riethmüller, S.; Hartmann, C.; Möller, M.; et al. Oxidation-Resistant Gold-55 Clusters. Science 2002, 297, 1533–1536. [CrossRef]
  77. Tracy, J.B.; Kalyuzhny, G.; Crowe, M.C.; Balasubramanian, R.; Choi, J.-P.; Murray, R.W. Poly(Ethylene Glycol) Ligands for High-Resolution Nanoparticle Mass Spectrometry. J. Am. Chem. Soc. 2007, 129, 6706–6707. [CrossRef]
  78. Kumar, S.; Jin, R. Water-Soluble Au25(Capt)18 Nanoclusters: Synthesis, Thermal Stability, and Optical Properties. Nanoscale 2012, 4, 4222. [CrossRef]
  79. Negishi, Y.; Kurashige, W.; Niihori, Y.; Iwasa, T.; Nobusada, K. Isolation, Structure, and Stability of a Dodecanethiolate-Protected Pd1Au24 Cluster. Phys. Chem. Chem. Phys. 2010, 12, 6219. [CrossRef]
  80. Jiang, D.; Dai, S. From Superatomic Au 25 (SR) 18  to Superatomic M@Au 24 (SR) 18 q Core−Shell Clusters. Inorg. Chem. 2009, 48, 2720–2722. [CrossRef]
  81. Negishi, Y.; Munakata, K.; Ohgake, W.; Nobusada, K. Effect of Copper Doping on Electronic Structure, Geometric Structure, and Stability of Thiolate-Protected Au 25 Nanoclusters. J. Phys. Chem. Lett. 2012, 3, 2209–2214. [CrossRef]
  82. Akola, J.; Walter, M.; Whetten, R.L.; Häkkinen, H.; Grönbeck, H. On the Structure of Thiolate-Protected Au 25. J. Am. Chem. Soc. 2008, 130, 3756–3757. [CrossRef]
  83. Tlahuice-Flores, A.; Muñoz-Castro, A. Bonding and Properties of Superatoms. Analogs to Atoms and Molecules and Related Concepts from Superatomic Clusters. Int J Quantum Chem 2019, 119, e25756. [CrossRef]
  84. De Heer, W.A.; Knight, W.D.; Chou, M.Y.; Cohen, M.L. Electronic Shell Structure and Metal Clusters. In Solid State Physics; Elsevier, 1987; Vol. 40, pp. 93–181 ISBN 978-0-12-607740-7.
  85. Bertorelle, F.; Russier-Antoine, I.; Comby-Zerbino, C.; Chirot, F.; Dugourd, P.; Brevet, P.-F.; Antoine, R. Isomeric Effect of Mercaptobenzoic Acids on the Synthesis, Stability, and Optical Properties of Au 25 (MBA) 18 Nanoclusters. ACS Omega 2018, 3, 15635–15642. [CrossRef]
  86. Das, K.; Roychoudhury, A. Reactive Oxygen Species (ROS) and Response of Antioxidants as ROS-Scavengers during Environmental Stress in Plants. Front. Environ. Sci. 2014, 2. [CrossRef]
  87. Sacco, O.; Stoller, M.; Vaiano, V.; Ciambelli, P.; Chianese, A.; Sannino, D. Photocatalytic Degradation of Organic Dyes under Visible Light on N-Doped TiO 2 Photocatalysts. International Journal of Photoenergy 2012, 2012, 1–8. [CrossRef]
  88. Guo, W.; Yang, L.; Lu, J.; Gao, P.; Li, W.; Feng, Z. An Accurate Growth Mechanism and Photocatalytic Degradation Rhodamine B of Crystalline Nb2O5 Nanotube Arrays. Catalysts 2020, 10, 1480. [CrossRef]
  89. Goswami, T.; Singh, M.; Reddy, K.M.; Mishra, A.K. Facile Synthesis of Ag-TiO 2 Hybrid Nanocluster:A Comprehensive Experimental and Computational Insight into the Role of Surface Ligands on Enhanced Visible Light Photo-Catalysis. ChemistrySelect 2018, 3, 10892–10899. [CrossRef]
  90. Zhu, H.; Goswami, N.; Yao, Q.; Chen, T.; Liu, Y.; Xu, Q.; Chen, D.; Lu, J.; Xie, J. Cyclodextrin–Gold Nanocluster Decorated TiO 2 Enhances Photocatalytic Decomposition of Organic Pollutants. J. Mater. Chem. A 2018, 6, 1102–1108. [CrossRef]
  91. Sharma, V.; Kumar, S.; Krishnan, V. Clustered Au on TiO 2 Snowman-Like Nanoassemblies for Photocatalytic Applications. ChemistrySelect 2016, 1, 2963–2970. [CrossRef]
  92. Goswami, T.; Reddy, K.M.; Bheemaraju, A. Silver Nanocluster Anchored TiO 2 /Nb 2 O 5 Hybrid Nanocomposite as Highly Efficient and Selective Visible-Light Sensitive Photocatalyst. ChemistrySelect 2019, 4, 6790–6799. [CrossRef]
  93. Samai, B.; Chall, S.; Mati, S.S.; Bhattacharya, S.C. Role of Silver Nanoclusters in the Enhanced Photocatalytic Activity of Cerium Oxide Nanoparticles. Eur. J. Inorg. Chem. 2018, 2018, 3224–3231. [CrossRef]
  94. González-Rodríguez, J.; Fernández, L.; Bava, Y.B.; Buceta, D.; Vázquez-Vázquez, C.; López-Quintela, M.A.; Feijoo, G.; Moreira, M.T. Enhanced Photocatalytic Activity of Semiconductor Nanocomposites Doped with Ag Nanoclusters Under UV and Visible Light. Catalysts 2019, 10, 31. [CrossRef]
  95. Vilar-Vidal, N.; Rey, J.R.; López Quintela, M.A. Green Emitter Copper Clusters as Highly Efficient and Reusable Visible Degradation Photocatalysts. Small 2014, 10, 3632–3636. [CrossRef]
  96. Zhou, S.; Duan, Y.; Wang, F.; Wang, C. Fluorescent Au Nanoclusters Stabilized by Silane: Facile Synthesis, Color-Tunability and Photocatalytic Properties. Nanoscale 2017, 9, 4981–4988. [CrossRef]
  97. Veziroglu, S.; Obermann, A.-L.; Ullrich, M.; Hussain, M.; Kamp, M.; Kienle, L.; Leißner, T.; Rubahn, H.-G.; Polonskyi, O.; Strunskus, T.; et al. Photodeposition of Au Nanoclusters for Enhanced Photocatalytic Dye Degradation over TiO 2 Thin Film. ACS Appl. Mater. Interfaces 2020, 12, 14983–14992. [CrossRef]
  98. Cao, M.; Pang, R.; Wang, Q.-Y.; Han, Z.; Wang, Z.-Y.; Dong, X.-Y.; Li, S.-F.; Zang, S.-Q.; Mak, T.C.W. Porphyrinic Silver Cluster Assembled Material for Simultaneous Capture and Photocatalysis of Mustard-Gas Simulant. J. Am. Chem. Soc. 2019, 141, 14505–14509. [CrossRef]
  99. Weng, B.; Lu, K.-Q.; Tang, Z.; Chen, H.M.; Xu, Y.-J. Stabilizing Ultrasmall Au Clusters for Enhanced Photoredox Catalysis. Nat Commun 2018, 9, 1543. [CrossRef]
  100. Al-Shankiti, B.; Al-Maksoud, W.; Habeeb Muhammed, M.A.; Anjum, D.H.; Moosa, B.; Basset, J.-M.; Khashab, N.M. Ligand-Free Gold Nanoclusters Confined in Mesoporous Silica Nanoparticles for Styrene Epoxidation. Nanoscale Adv. 2020, 2, 1437–1442. [CrossRef]
  101. Chen, H.; Liu, C.; Wang, M.; Zhang, C.; Luo, N.; Wang, Y.; Abroshan, H.; Li, G.; Wang, F. Visible Light Gold Nanocluster Photocatalyst: Selective Aerobic Oxidation of Amines to Imines. ACS Catal. 2017, 7, 3632–3638. [CrossRef]
  102. Liu, L.; Li, H.; Tan, Y.; Chen, X.; Lin, R.; Yang, W.; Huang, C.; Wang, S.; Wang, X.; Liu, X.Y.; et al. Metal-Support Synergy of Supported Gold Nanoclusters in Selective Oxidation of Alcohols. Catalysts 2020, 10, 107. [CrossRef]
  103. Kawawaki, T.; Kataoka, Y.; Ozaki, S.; Kawachi, M.; Hirata, M.; Negishi, Y. Creation of Active Water-Splitting Photocatalysts by Controlling Cocatalysts Using Atomically Precise Metal Nanoclusters. Chem. Commun. 2021, 57, 417–440. [CrossRef]
  104. Li, G.; Zeng, C.; Jin, R. Thermally Robust Au 99 (SPh) 42 Nanoclusters for Chemoselective Hydrogenation of Nitrobenzaldehyde Derivatives in Water. J. Am. Chem. Soc. 2014, 136, 3673–3679. [CrossRef]
  105. Mu, J.; Yang, J.-L.; Zhang, D.-W.; Jia, Q. Progress in Preparation of Metal Nanoclusters and Their Application in Detection of Environmental Pollutants. Chinese Journal of Analytical Chemistry 2021, 49, 319–329. [CrossRef]
  106. Haruta, M.; Kobayashi, T.; Sano, H.; Yamada, N. Novel Gold Catalysts for the Oxidation of Carbon Monoxide at a Temperature Far Below 0 °C. Chem. Lett. 1987, 16, 405–408. [CrossRef]
  107. Kogo, A.; Sakai, N.; Tatsuma, T. Photocatalysis of Au25-Modified TiO2 under Visible and near Infrared Light. Electrochemistry Communications 2010, 12, 996–999. [CrossRef]
  108. Zhu, Y.; Qian, H.; Zhu, M.; Jin, R. Thiolate-Protected Aun Nanoclusters as Catalysts for Selective Oxidation and Hydrogenation Processes. Adv. Mater. 2010, 22, 1915–1920. [CrossRef]
  109. Tsunoyama, H.; Ichikuni, N.; Sakurai, H.; Tsukuda, T. Effect of Electronic Structures of Au Clusters Stabilized by Poly( N -Vinyl-2-Pyrrolidone) on Aerobic Oxidation Catalysis. J. Am. Chem. Soc. 2009, 131, 7086–7093. [CrossRef]
  110. Hamoud, H.I.; Douma, F.; Lafjah, M.; Djafri, F.; Lebedev, O.; Valtchev, V.; El-Roz, M. Size-Dependent Photocatalytic Activity of Silver Nanoparticles Embedded in ZX-Bi Zeolite Supports. ACS Appl. Nano Mater. 2022, 5, 3866–3877. [CrossRef]
  111. Jain, I.P. Hydrogen the Fuel for 21st Century. International Journal of Hydrogen Energy 2009, 34, 7368–7378. [CrossRef]
  112. Ahluwalia, R.K.; Wang, X.; Rousseau, A.; Kumar, R. Fuel Economy of Hydrogen Fuel Cell Vehicles. Journal of Power Sources 2004, 130, 192–201. [CrossRef]
  113. Thoi, V.S.; Sun, Y.; Long, J.R.; Chang, C.J. Complexes of Earth-Abundant Metals for Catalytic Electrochemical Hydrogen Generation under Aqueous Conditions. Chem. Soc. Rev. 2013, 42, 2388–2400. [CrossRef]
  114. Xiao, F.-X.; Hung, S.-F.; Miao, J.; Wang, H.-Y.; Yang, H.; Liu, B. Metal-Cluster-Decorated TiO 2 Nanotube Arrays: A Composite Heterostructure toward Versatile Photocatalytic and Photoelectrochemical Applications. Small 2015, 11, 554–567. [CrossRef]
  115. Negishi, Y.; Mizuno, M.; Hirayama, M.; Omatoi, M.; Takayama, T.; Iwase, A.; Kudo, A. Enhanced Photocatalytic Water Splitting by BaLa4Ti4O15 Loaded with ∼1 Nm Gold Nanoclusters Using Glutathione-Protected Au25 Clusters. Nanoscale 2013, 5, 7188. [CrossRef]
  116. Kurashige, W.; Kumazawa, R.; Mori, Y. Au25 Cluster-Loaded SrTiO3 Water-Splitting Photocatalyst; Preparation and Elucidation of the Effect of Cocatalyst Refinement on Photocatalytic Activity. J. Mater. Appl.
  117. Mousavi, H.; Small, T.D.; Sharma, S.K.; Golovko, V.B.; Shearer, C.J.; Metha, G.F. Graphene Bridge for Photocatalytic Hydrogen Evolution with Gold Nanocluster Co-Catalysts. Nanomaterials 2022, 12, 3638. [CrossRef]
  118. Attia, Y.A.; Buceta, D.; Blanco-Varela, C.; Mohamed, M.B.; Barone, G.; López-Quintela, M.A. Structure-Directing and High-Efficiency Photocatalytic Hydrogen Production by Ag Clusters. J. Am. Chem. Soc. 2014, 136, 1182–1185. [CrossRef]
  119. Shen, P.; Zhao, S.; Su, D.; Li, Y.; Orlov, A. Outstanding Activity of Sub-Nm Au Clusters for Photocatalytic Hydrogen Production. Applied Catalysis B: Environmental 2012, 126, 153–160. [CrossRef]
  120. Wang, H.; Luo, S.; Song, Y.; Shi, Y.; Wang, Z.; Guo, B.; Wu, L. Enhanced Photocatalytic Hydrogen Evolution over Monolayer HTi2NbO7 Nanosheets with Highly Dispersed Pt Nanoclusters. Applied Surface Science 2020, 511, 145501. [CrossRef]
  121. Schweinberger, F.F.; Berr, M.J.; Döblinger, M.; Wolff, C.; Sanwald, K.E.; Crampton, A.S.; Ridge, C.J.; Jäckel, F.; Feldmann, J.; Tschurl, M.; et al. Cluster Size Effects in the Photocatalytic Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 13262–13265. [CrossRef]
  122. Li, S.; Du, X.; Liu, Z.; Li, Y.; Shao, Y.; Jin, R. Size Effects of Atomically Precise Gold Nanoclusters in Catalysis. Precision Chemistry 2023, 1, 14–28. [CrossRef]
  123. Naveen, M.H.; Khan, R.; Bang, J.H. Gold Nanoclusters as Electrocatalysts: Atomic Level Understanding from Fundamentals to Applications. Chem. Mater. 2021, 33, 7595–7612. [CrossRef]
  124. Kurashige, W.; Hayashi, R.; Wakamatsu, K.; Kataoka, Y.; Hossain, S.; Iwase, A.; Kudo, A.; Yamazoe, S.; Negishi, Y. Atomic-Level Understanding of the Effect of Heteroatom Doping of the Cocatalyst on Water-Splitting Activity in AuPd or AuPt Alloy Cluster-Loaded BaLa 4 Ti 4 O 15. ACS Appl. Energy Mater. 2019, 2, 4175–4187. [CrossRef]
  125. Du, X.L.; Wang, X.L.; Li, Y.H.; Wang, Y.L.; Zhao, J.J.; Fang, L.J.; Zheng, L.R.; Tong, H.; Yang, H.G. Isolation of Single Pt Atoms in a Silver Cluster: Forming Highly Efficient Silver-Based Cocatalysts for Photocatalytic Hydrogen Evolution. Chem. Commun. 2017, 53, 9402–9405. [CrossRef]
  126. Méndez-Medrano, M.G.; Kowalska, E.; Lehoux, A.; Herissan, A.; Ohtani, B.; Rau, S.; Colbeau-Justin, C.; Rodríguez-López, J.L.; Remita, H. Surface Modification of TiO 2 with Au Nanoclusters for Efficient Water Treatment and Hydrogen Generation under Visible Light. J. Phys. Chem. C 2016, 120, 25010–25022. [CrossRef]
  127. Tawalbeh, M.; Al-Othman, A.; Kafiah, F.; Abdelsalam, E.; Almomani, F.; Alkasrawi, M. Environmental Impacts of Solar Photovoltaic Systems: A Critical Review of Recent Progress and Future Outlook. Science of The Total Environment 2021, 759, 143528. [CrossRef]
  128. Kondratenko, E.V.; Mul, G.; Baltrusaitis, J.; Larrazábal, G.O.; Pérez-Ramírez, J. Status and Perspectives of CO2 Conversion into Fuels and Chemicals by Catalytic, Photocatalytic and Electrocatalytic Processes. Energy Environ. Sci. 2013, 6, 3112. [CrossRef]
  129. Roy, S.C.; Varghese, O.K.; Paulose, M.; Grimes, C.A. Toward Solar Fuels: Photocatalytic Conversion of Carbon Dioxide to Hydrocarbons. ACS Nano 2010, 4, 1259–1278. [CrossRef]
  130. Alper, E.; Yuksel Orhan, O. CO2 Utilization: Developments in Conversion Processes. Petroleum 2017, 3, 109–126. [CrossRef]
  131. Du, C.; Wang, X.; Chen, W.; Feng, S.; Wen, J.; Wu, Y.A. CO2 Transformation to Multicarbon Products by Photocatalysis and Electrocatalysis. Materials Today Advances 2020, 6, 100071. [CrossRef]
  132. Halmann, M. Photoelectrochemical Reduction of Aqueous Carbon Dioxide on P-Type Gallium Phosphide in Liquid Junction Solar Cells. Nature 1978, 275, 115–116. [CrossRef]
  133. Olah, G.A. Beyond Oil and Gas: The Methanol Economy. Angew. Chem. Int. Ed. 2005, 44, 2636–2639. [CrossRef]
  134. Lu, Q.; Jiao, F. Electrochemical CO2 Reduction: Electrocatalyst, Reaction Mechanism, and Process Engineering. Nano Energy 2016, 29, 439–456. [CrossRef]
  135. Philip Colombo, D.; Roussel, K.A.; Saeh, J.; Skinner, D.E.; Cavaleri, J.J.; Bowman, R.M. Femtosecond Study of the Intensity Dependence of Electron-Hole Dynamics in TiO2 Nanoclusters. Chemical Physics Letters 1995, 232, 207–214. [CrossRef]
  136. Kauffman, D.R.; Alfonso, D.; Matranga, C.; Qian, H.; Jin, R. Experimental and Computational Investigation of Au 25 Clusters and CO 2 : A Unique Interaction and Enhanced Electrocatalytic Activity. J. Am. Chem. Soc. 2012, 134, 10237–10243. [CrossRef]
  137. Guo, S.-X.; MacFarlane, D.R.; Zhang, J. Bioinspired Electrocatalytic CO 2 Reduction by Bovine Serum Albumin-Capped Silver Nanoclusters Mediated by [ α -SiW 12 O 40 ] 4−. ChemSusChem 2016, 9, 80–87. [CrossRef]
  138. Corma, A.; Garcia, H. Photocatalytic Reduction of CO2 for Fuel Production: Possibilities and Challenges. Journal of Catalysis 2013, 308, 168–175. [CrossRef]
  139. Qin, L.; Ma, G.; Wang, L.; Tang, Z. Atomically Precise Metal Nanoclusters for (Photo)Electroreduction of CO2: Recent Advances, Challenges and Opportunities. Journal of Energy Chemistry 2021, 57, 359–370. [CrossRef]
  140. Linsebigler, A.L.; Lu, G.; Yates, J.T. Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 1995, 95, 735–758. [CrossRef]
  141. Liu, M.; Qiu, X.; Miyauchi, M.; Hashimoto, K. Cu(II) Oxide Amorphous Nanoclusters Grafted Ti 3+ Self-Doped TiO 2 : An Efficient Visible Light Photocatalyst. Chem. Mater. 2011, 23, 5282–5286. [CrossRef]
  142. Shoji, S.; Yin, G.; Nishikawa, M.; Atarashi, D.; Sakai, E.; Miyauchi, M. Photocatalytic Reduction of CO2 by Cu O Nanocluster Loaded SrTiO3 Nanorod Thin Film. Chemical Physics Letters 2016, 658, 309–314. [CrossRef]
  143. Yin, G.; Nishikawa, M.; Nosaka, Y.; Srinivasan, N.; Atarashi, D.; Sakai, E.; Miyauchi, M. Photocatalytic Carbon Dioxide Reduction by Copper Oxide Nanocluster-Grafted Niobate Nanosheets. ACS Nano 2015, 9, 2111–2119. [CrossRef]
  144. Jin, J.; Luo, J.; Zan, L.; Peng, T. One-Pot Synthesis of Cu-Nanocluster-Decorated Brookite TiO 2 Quasi -Nanocubes for Enhanced Activity and Selectivity of CO 2 Photoreduction to CH 4. ChemPhysChem 2017, 18, 3230–3239. [CrossRef]
  145. Cui, X.; Wang, J.; Liu, B.; Ling, S.; Long, R.; Xiong, Y. Turning Au Nanoclusters Catalytically Active for Visible-Light-Driven CO 2 Reduction through Bridging Ligands. J. Am. Chem. Soc. 2018, 140, 16514–16520. [CrossRef]
  146. Jiang, Y.; Yu, Y.; Zhang, X.; Weinert, M.; Song, X.; Ai, J.; Han, L.; Fei, H. N-Heterocyclic Carbene-Stabilized Ultrasmall Gold Nanoclusters in a Metal-Organic Framework for Photocatalytic CO 2 Reduction. Angew. Chem. Int. Ed. 2021, 60, 17388–17393. [CrossRef]
  147. Billo, T.; Fu, F.-Y.; Raghunath, P.; Shown, I.; Chen, W.-F.; Lien, H.-T.; Shen, T.-H.; Lee, J.-F.; Chan, T.-S.; Huang, K.-Y.; et al. Ni-Nanocluster Modified Black TiO 2 with Dual Active Sites for Selective Photocatalytic CO 2 Reduction. Small 2018, 14, 1702928. [CrossRef]
  148. Chen, Y.-S.; Kamat, P.V. Glutathione-Capped Gold Nanoclusters as Photosensitizers. Visible Light-Induced Hydrogen Generation in Neutral Water. J. Am. Chem. Soc. 2014, 136, 6075–6082. [CrossRef]
  149. Zhang, L.; Can, M.; Ragsdale, S.W.; Armstrong, F.A. Fast and Selective Photoreduction of CO 2 to CO Catalyzed by a Complex of Carbon Monoxide Dehydrogenase, TiO 2 , and Ag Nanoclusters. ACS Catal. 2018, 8, 2789–2795. [CrossRef]
  150. El-Roz, M.; Telegeiev, I.; Mordvinova, N.E.; Lebedev, O.I.; Barrier, N.; Behilil, A.; Zaarour, M.; Lakiss, L.; Valtchev, V. Uniform Generation of Sub-Nanometer Silver Clusters in Zeolite Cages Exhibiting High Photocatalytic Activity under Visible Light. ACS Appl. Mater. Interfaces 2018, 10, 28702–28708. [CrossRef]
Figure 1. Applications of Metal NCs in photocatalysis.
Figure 1. Applications of Metal NCs in photocatalysis.
Preprints 72790 g001
Figure 2. shows a schematic representation of the core-shell structure for fluorescent MNCs and the ligands used for the synthesis.
Figure 2. shows a schematic representation of the core-shell structure for fluorescent MNCs and the ligands used for the synthesis.
Preprints 72790 g002
Figure 3. UV-Vis absorption spectrum of Au25SG18, and solar flux for AM 1.5. (Solar radiation by using a Xe lamp). Reprinted with permission from [56]. Copyright 2012 Royal Society of Chemistry.
Figure 3. UV-Vis absorption spectrum of Au25SG18, and solar flux for AM 1.5. (Solar radiation by using a Xe lamp). Reprinted with permission from [56]. Copyright 2012 Royal Society of Chemistry.
Preprints 72790 g003
Figure 4. Explains the major differences between MNCs and Metal nanoparticles by Size, and optical absorption. Reprinted with permission from [60]Copyright 2021 Nature.
Figure 4. Explains the major differences between MNCs and Metal nanoparticles by Size, and optical absorption. Reprinted with permission from [60]Copyright 2021 Nature.
Preprints 72790 g004
Figure 5. Schematic representation of the relaxation pathways of ligand-stabilized atomically precise gold nanoclusters.
Figure 5. Schematic representation of the relaxation pathways of ligand-stabilized atomically precise gold nanoclusters.
Preprints 72790 g005
Figure 6. Shows the TD-DFT Two photon absorption spectrum for Au10L10 (L-−SCH2CH(NH2)CH3). Bottom panel of figure shows the leading excitations responsible for the characteristic features of TPA.
Figure 6. Shows the TD-DFT Two photon absorption spectrum for Au10L10 (L-−SCH2CH(NH2)CH3). Bottom panel of figure shows the leading excitations responsible for the characteristic features of TPA.
Preprints 72790 g006
Figure 7. Schematic representation of hierarchical structure of AuNCs and emergence of chirality. Reprinted with permission from [73] Copyright 2021 Nature.
Figure 7. Schematic representation of hierarchical structure of AuNCs and emergence of chirality. Reprinted with permission from [73] Copyright 2021 Nature.
Preprints 72790 g007
Figure 8. The NMR spectra of captopril before (A) and after heating (B) and glutathione before (C) and after heating (D)respectively Reprinted with permission from [78] Copyright 2007 Royal Society of Chemistry.
Figure 8. The NMR spectra of captopril before (A) and after heating (B) and glutathione before (C) and after heating (D)respectively Reprinted with permission from [78] Copyright 2007 Royal Society of Chemistry.
Preprints 72790 g008
Figure 9. a) diagram shows the synthesis protocol of different structural MBA isomers stabilized Au25(MBA)18 NCs. b) ESI mass spectra Au25 clusters c) collision induced dissociation breakdown curves for the 4– charge state of o/m/p-MBA stabilized. Reprinted with permission from [85] Copyright 2018 American Chemical Society.
Figure 9. a) diagram shows the synthesis protocol of different structural MBA isomers stabilized Au25(MBA)18 NCs. b) ESI mass spectra Au25 clusters c) collision induced dissociation breakdown curves for the 4– charge state of o/m/p-MBA stabilized. Reprinted with permission from [85] Copyright 2018 American Chemical Society.
Preprints 72790 g009
Figure 10. Dye degradation of ZnO nanoparticle decorated with AgNCs. Reprinted with permission from [94] Copyright 2019 MDPI.
Figure 10. Dye degradation of ZnO nanoparticle decorated with AgNCs. Reprinted with permission from [94] Copyright 2019 MDPI.
Preprints 72790 g010
Figure 11. A) UV Visible B) Photoluminescence C) &D) AFM results of ZnO nanoparticle decorated with Ag NCs . . Reprinted with permission from [94] Copyright 2019 MDPI.
Figure 11. A) UV Visible B) Photoluminescence C) &D) AFM results of ZnO nanoparticle decorated with Ag NCs . . Reprinted with permission from [94] Copyright 2019 MDPI.
Preprints 72790 g011
Figure 12. Photocatalytic oxidation ability. Reprinted with permission from [107] Copyright 2010 Elsevier.
Figure 12. Photocatalytic oxidation ability. Reprinted with permission from [107] Copyright 2010 Elsevier.
Preprints 72790 g012
Figure 13. CO/CH4 production rates (a) and the corresponding total consumed electron numbers (TCEN) (b) for the CO2 photoreduction over pristine BTN and Cu-BTN production with different Cu loading contents during the initial 2 h of irradiation. Reprinted with permission from [144] Copyright 2017 American Chemical Society.
Figure 13. CO/CH4 production rates (a) and the corresponding total consumed electron numbers (TCEN) (b) for the CO2 photoreduction over pristine BTN and Cu-BTN production with different Cu loading contents during the initial 2 h of irradiation. Reprinted with permission from [144] Copyright 2017 American Chemical Society.
Preprints 72790 g013
Figure 14. In situ DRIFTS IR spectra of H2O (a) or CO2/H2O vapor (b) interaction with pristine BTN, 1.5% Cu-BTN, and 0.5% Cu/BTN. Reprinted with permission from [144] Copyright 2017 American Chemical Society.
Figure 14. In situ DRIFTS IR spectra of H2O (a) or CO2/H2O vapor (b) interaction with pristine BTN, 1.5% Cu-BTN, and 0.5% Cu/BTN. Reprinted with permission from [144] Copyright 2017 American Chemical Society.
Preprints 72790 g014
Figure 15. (a) Average production rates of CH4, CO and H2 in light drven CO2 reduction with H2O in the presence of TEOA, catalyzed by 10-mg Auc-MPA-Co grafted with 4 µmol Co2+ (b) The binding energy of S-Co bond in the coordination of Co with L-cys and MPA. Reprinted with permission from [145] Copyright 2018 American Chemical Society.
Figure 15. (a) Average production rates of CH4, CO and H2 in light drven CO2 reduction with H2O in the presence of TEOA, catalyzed by 10-mg Auc-MPA-Co grafted with 4 µmol Co2+ (b) The binding energy of S-Co bond in the coordination of Co with L-cys and MPA. Reprinted with permission from [145] Copyright 2018 American Chemical Society.
Preprints 72790 g015
Figure 16. (a) CO evolution period of photocatalytic CO2 reduction using UiO-68-NHC, Au-NC@UiO-68-NHC, UiO-68-NH2 and Au/ UiO-68-NH2 as photocatalysts upon AM 1.5G irradiation. (b) Time courses of photocatalytic CO2 reduction on Au-NC@ UiO-68-NHC under AM 1.5 G irradiation for 12h, with evacuation every 4 h (dashed line). . Reprinted with permission from [146] Copyright 2021 Wiely.
Figure 16. (a) CO evolution period of photocatalytic CO2 reduction using UiO-68-NHC, Au-NC@UiO-68-NHC, UiO-68-NH2 and Au/ UiO-68-NH2 as photocatalysts upon AM 1.5G irradiation. (b) Time courses of photocatalytic CO2 reduction on Au-NC@ UiO-68-NHC under AM 1.5 G irradiation for 12h, with evacuation every 4 h (dashed line). . Reprinted with permission from [146] Copyright 2021 Wiely.
Preprints 72790 g016
Figure 17. (a) Band edge positions. (b) Schematic illustration of photocatalytic CO2 reduction mechanism of Ni/ TiO2[Vo]. Adapted from Ref 146.
Figure 17. (a) Band edge positions. (b) Schematic illustration of photocatalytic CO2 reduction mechanism of Ni/ TiO2[Vo]. Adapted from Ref 146.
Preprints 72790 g017
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

© 2024 MDPI (Basel, Switzerland) unless otherwise stated