Rugma
T. P.
,
Michael N.
Pillay
and
C. W.
Liu
*
Department of Chemistry, National Dong Hwa University, Hualien, 97401, Taiwan, Republic of China. E-mail: chenwei@gms.ndhu.edu.tw
First published on 7th May 2025
Photocatalytic hydrogen production offers a sustainable approach for utilising light energy, providing a promising solution to global energy challenges. The efficiency of this process relies on developing photocatalysts with broad light responsiveness and effective charge carrier separation capabilities. Atomically precise metal nanoclusters (NCs) have emerged as a highly favourable class of materials for this role due to their unique atomic arrangements, ultrasmall size, quantum confinement effects, and plenty of surface-active sites. These exceptional properties endow NCs with semiconductor-like behaviour, allowing for the generation of electrons and holes under light excitation, thus driving the hydrogen production reaction. Moreover, their robust light-absorption properties across the UV to near-IR spectrum, coupled with tuneable optical properties controlled by their composition and structure, promise NCs as next-generation photocatalysts. This review explores recent developments in the application of NCs for photocatalytic hydrogen production, emphasising strategies to enhance charge carrier separation and transfer efficiency, as well as photostability. The discussion also highlights the challenges and future opportunities in using NCs for efficient hydrogen production.
Due to their extremely small size, numerous surface sites with unsaturated coordination, distinctive atomic arrangements, and pronounced quantum confinement effects, NCs have attracted significant interest in the field of catalysis.9–11 Their electronic distribution itself is a major property that alters their catalytic efficiency.12 In addition, the pronounced quantum confinement effect in these tiny clusters means that even a single atom's addition or removal can dramatically alter their catalytic activity. Furthermore, the precise composition, structure, and surface environment of NCs allows for in-depth exploration of catalytic mechanisms and the development of detailed structure–property relationships.13 For effective photocatalysis, an ideal catalyst must possess strong light absorption capabilities, efficiently generate photoinduced electron–hole pairs, and facilitate rapid transfer and separation of these charge carriers.14,15 The surface plasmon resonance (SPR) in conventional NPs results in distinctive optical absorption properties, usually restricted to the visible light spectrum. This limitation arises because the collective oscillation of electrons on the NP surface, characteristic of SPR, only interacts with specific wavelengths.16,17 In contrast, NCs exhibit a much broader light absorption spectrum, extending from the ultraviolet (UV) to the near-infrared (NIR) regions. This wide absorption range makes NCs particularly advantageous for applications requiring enhanced light-harvesting capabilities across a diverse range of wavelengths, including those beyond the visible spectrum.18,19
Ideally, NCs can function as single photocatalysts, capable of generating electrons and holes under light excitation due to their small-band-gap semiconductor properties. However, practical challenges, such as their ultrasmall size leading to high surface energy and the tendency for charge carrier recombination, often make it difficult to achieve efficient photocatalysis when they are used alone. As a result, NCs are more commonly employed as co-catalysts in experimental setups. In this role, they enhance the performance of semiconductor-based photocatalysts by facilitating charge separation and transfer, which improves overall photocatalytic efficiency. The utilisation of NCs in catalytic applications has seen rapid growth, with significant advancements in recent years. However, their use in photocatalysis has not yet caught up in other catalytic fields like electrocatalysis.20–23 To address the challenges posed by NCs, such as rapid charge carrier recombination and poor photostability, several synthesis methods have been explored to create NC–porous support composites, which help prevent aggregation and enhance photocatalytic performance. Recent studies have proposed coupling NCs with other materials to form heterostructures, thereby enhancing photostability and extending charge carrier lifetimes.24,25
Some recent reviews have explored the electrocatalytic and photocatalytic applications of NCs,26,27 but their role as co-catalysts in photocatalytic reactions has not been thoroughly examined. In addition, these reviews have broadly addressed the wide-ranging applications of NCs, but our focus is on photocatalytic hydrogen production from water due to its potential for simplicity and scalability. Unlike more complex systems, water-splitting using powder photocatalysts offers a facile, cost-effective approach that can be easily adapted for large-scale hydrogen production. Therefore, we believe it is essential to provide a detailed review of these studies to shed light on the recent progress, challenges, and potential of powder-based photocatalysts in hydrogen production. Additionally, the review has addressed the current challenges and future opportunities in this field. By summarising the fundamental principles of photocatalytic hydrogen production and discussing various NC-based heterostructures, this review offers a comprehensive understanding of the key role that NCs play in advancing photocatalytic hydrogen production.
2H2O → 2H2 + O2 (E > 1.23 eV) | (1) |
The mechanism of photocatalytic water splitting for hydrogen production generally consists of three main steps: (1) light absorption by photocatalysts, which generates electron (e−)–hole (h+) pairs; (2) separation and movement of these charges to the surface of the photocatalyst; and (3) surface reactions for water reduction and oxidation (Fig. 1a).30 Semiconductor photocatalysts typically have a conduction band (CB) and a valence band (VB) separated by a band gap (Eg). When the photon energy is equal to or greater than the band gap energy, the photocatalyst absorbs light, leading to the excitation of electrons from the VB to the CB, leaving holes in the VB. The electrons in the CB drive the reduction reaction to produce hydrogen, while the holes in the VB drive the oxidation process. For the reactions to occur, the CB edge must be more negative than the H+/H2 redox potential (0 V vs. NHE), and the VB edge must be more positive than the O2/H2O redox potential (Fig. 1(b)).2 Without suitable active sites on the photocatalyst surface, electron–hole pairs can rapidly recombine within the material. Studies have shown that the recombination of photogenerated charges in the catalyst bulk happens very quickly, within picoseconds to milliseconds. Therefore, to ensure efficient charge separation and transfer, cocatalysts must quickly capture the electrons and holes and facilitate the surface reduction reactions, which are vital for successful photocatalytic water splitting.
Photocatalytic water splitting typically involves two types of systems: photochemical and photoelectrochemical.31 The most basic method is photochemical water splitting, using a semiconductor slurry. In this setup, a photocatalytic system contains an aqueous suspension of semiconductor particles exposed to light. Upon irradiation, the semiconductor absorbs photons, generating electrons in the conduction band and holes in the valence band. These charge carriers then move to the catalyst's surface, where they participate in the splitting of water molecules into hydrogen and oxygen at the active sites.32 However, one limitation of this approach is that the band gap photocatalyst must align with the water reduction and oxidation potentials, which restricts the available material choices. Additionally, comparing the hydrogen production rates of various catalysts is challenging due to the various light sources and reactor designs used in different studies. Another challenge is the simultaneous production of hydrogen and oxygen in the same reactor, which complicates the separation of the two gases. Photocatalytic water splitting can be performed using either a single catalyst or a dual-catalyst configuration, such as a tandem or Z-scheme or any of the charge transfer pathways illustrated in Fig. 2. The Z-scheme approach mimics natural photosynthesis, where the two chlorophyll centres in photosystems II and I carry out the oxidation and reduction reactions, respectively. Similarly, in a Z-scheme, water splitting occurs across two different catalysts, with each semiconductor only needing to match the potential for each half-reaction.33 This allows for a broader selection of materials compared to single-catalyst systems. To further boost photocatalytic performance, cocatalysts (such as noble metals or metal oxides) can be added to the semiconductor to serve as electron or hole sinks, facilitating water reduction and oxidation while minimising charge recombination. The water oxidation catalyst produces oxygen using the holes generated in the valence band, while the water reduction catalyst utilises the electrons for the reduction process.
![]() | ||
Fig. 2 (a) One-step photoexcitation process and (b) two-step photoexcitation (Z-scheme) process for overall water-splitting. Reproduced with permission from ref. 2. Copyright 2021 Elsevier. (c) Four types of conventional semiconductor heterojunctions. |
Taking this further, NCs, due to their ultra-small size and quantum confinement effects, offer unique properties compared to traditional metal nanoparticles. Metal NCs can serve as highly active cocatalysts due to their enhanced surface area and a higher density of reactive sites, which improve the transfer and separation of charge carriers at the semiconductor interface. This, combined with the tuneable surface properties of NCs, allows for more precise control over catalytic activity. Furthermore, NCs can exhibit strong quantum size effects, where even small changes in particle size can drastically affect their electronic properties, further optimising photocatalytic efficiency. Thus, the transition from traditional metal nanoparticles to NCs could lead to breakthroughs in designing more efficient and cost-effective photocatalytic systems. The following sections provide a summary of the recent progress in utilising NCs as co-catalysts for photocatalytic hydrogen production reactions.
Catalyst | Co-catalyst | Light source | Amount of catalyst used (g) | Sacrificial agent (vol %) | H2 activity (mmol h−1 g−1) | QY (%) | Ref. |
---|---|---|---|---|---|---|---|
Pt/Pt5/CdS | Pt | 300 W Xe lamp (λ > 420 nm) | 0.005 | TEOA (20) | 13 | 25.03 (400 nm) | 50 |
Au25(Cys)18/g-C3N4 | Au25(Cys)18 | 300 W Xe lamp (λ > 420 nm) | 0.02 | TEOA (10) | 0.32 | 0.20 (420 nm) | 52 |
MoSe2/CdSe/Au25 (GS)18 | Au25(GS)18 | 300 W Xe lamp (λ > 420 nm) | 0.01 | Lactic acid (10) | 0.14 | 53 | |
Cr2O3/Au25/BaLa4Ti4O15 | Au25(PET)18 | 400 W Hg lamp | 0.05 | — | ∼5 | 54 | |
Au25/BaLa4Ti4O15 | Au25 | 400 W Hg lamp | 0.5 | — | 0.50 | 55 | |
Au25–Cr2O3–BaLa4Ti4O15 | Au25(GS)18 | 400 W Hg lamp | 0.5 | Methanol (10) | 4.69 | 6.30 (270 nm) | 56 |
CdS@Au/Ti3−xC2Ty | CdS@Au | 300 W Xe lamp (λ > 420 nm) | 0.1 | Lactic acid (10) | 5.37 | 16.70 (420 nm) | 57 |
Ag44(SR)30/TiO2 | Ag44(SR)30 | 300 W Xe lamp | 0.02 | Methanol (20) | 7.40 | 59 | |
Pt1Ag28-BTT/CoP | Pt1Ag28 | 300 W Xe lamp (λ > 420 nm) | 0.005 | TEOA (15) | 240 | 25.77 (400 nm) | 60 |
Au12Ag32/TiO2 | Au12Ag32 | Xe lamp | 0.005 | Methanol (20) | 6.81 | 61 | |
UiO-66-NH2-Au25(L-Cys)18 | Au25(L-Cys)18 | 300 W Xe lamp (λ > 420 nm) | 0.005 | TEOA (10) | 17.02 | 62 | |
AuAg24@UiO-66-NH2 | AuAg24 | 300 W Xe lamp (λ > 380 nm) | 0.005 | TEA (—) | 3.60 | 63 | |
Ni6(SR)12/TiO2 | Ni6(SR)12 | 300 W Xe lamp | 0.02 | Methanol (20) | 5.60 | 7.80 (365 nm) | 65 |
Ni12(SR)24/g-C3N4 | Ni12(SR)24 | 300 W Xe lamp | 0.01 | TEOA (15) | 3.00 | 66 | |
Ni6(SC2H4Ph)12/g-C3N4 | Ni6(SC2H4Ph)12 | 300 W Xe lamp (λ > 400 nm) | 0.02 | TEOA (20) | 5.87 | 8.90 (400 nm) | 67 |
Cu20/TiO2 | Cu20 | 300 W Xe lamp | — | TEOA (—) | 13.00 | 2.32 (390 nm) | 68 |
Cu8-MOF | — | 300 W Xe lamp (λ > 380 nm) | 0.002 | TEOA (—) | 14.10 | 69 |
Maximizing the utilisation of photo-excited electrons is crucial for achieving high-efficiency photocatalytic hydrogen production. Among the various materials explored, Pt-based compounds stand out as highly effective co-catalysts due to their exceptional capacity to facilitate the separation of photogenerated electron–hole pairs. Lu et al. made a significant contribution to this field, developing a potent Pt5/CdS photocatalyst by attaching Pt5(GS)10 (GS = glutathione) clusters onto CdS nanorods.50 This structure was analysed using X-ray absorption fine structure (XAFS) spectroscopy with synchrotron radiation, confirming the presence of monoatomic platinum distributed across the surface of the CdS nanorods to form Pt–S4 active sites (Fig. 3d). This structure promoted a marked enhancement in photocatalytic performance. When exposed to visible light (λ > 400 nm), the Pt5(GS)10 loaded CdS photocatalyst achieved a hydrogen production rate of 13 mmol h−1 g−1 as shown in Fig. 3(a), which significantly surpassed the hydrogen production activity of bare CdS nanorods. Furthermore, ultrafast transient absorption (TA) spectroscopy revealed that the atomically dispersed platinum effectively extracted photo-generated electrons from the CdS nanorods, thereby promoting efficient separation of the photo-excited charge carriers (Fig. 3(c)). This process was key to the enhanced photocatalytic activity observed in the Pt5/CdS system. In addition, DFT calculations showed that the single Pt cocatalyst in Pt5/CdS possessed an optimal ΔGH* value close to zero. Apparently, the isolated single Pt atoms present on the surface of CdS nanorods formed new active sites and improved the intrinsic activity of CdS. The photocatalytic hydrogen production activity of the composite was also investigated after calcination. Although a physical mixture of CdS NRs and Pt5(GS)10 clusters initially demonstrated a significant increase in hydrogen evolution activity, this performance was further improved with thermal treatments. The catalyst's cycling stability was notably enhanced, as shown in Fig. 3(b). The quantum efficiency was calculated to be 25.08% at 400 nm.
![]() | ||
Fig. 3 (a) H2 evolution activity as a function of loading Pt5(GS)10, (b) the cycling stability of the Pt5/CdS catalysts before (red) and after photocatalytic reaction (black), (c) plausible photocatalytic mechanism of the Pt5/CdS photocatalyst, and (d) XAFS analysis (black: measured; red: fitted) of Pt5/CdS and a probable coordination configuration of Pt atoms in Pt5/CdS (inset). Reproduced with permission from ref. 50. Copyright 2022 Royal Society of Chemistry. |
In 2012, Shen et al.51 developed gold-modified CdS photocatalysts, revealing impressive hydrogen production. This work was the first to demonstrate the potential of gold particles for photocatalytic hydrogen production. Building on these developments, Wang et al.52 reported a heterostructure composed of Au25(L-Cys)18 NCs (L-Cys = L-cysteine) integrated with graphitic carbon nitride (g-C3N4) via a wet-impregnation method. The Au25(L-Cys)18 NCs were uniformly dispersed on the surface of g-C3N4, significantly enhancing photocatalytic hydrogen production under visible light. The highly dispersed Au25 NCs formed numerous junctions at the metal–semiconductor interface, optimising the active site distribution. This led to a remarkable increase in hydrogen production, reaching 320 μmol h−1 g−1 under visible light irradiation in an aqueous solution containing triethanolamine as a hole scavenger. The enhancement in photocatalytic performance was attributed to the effective band alignment between the metal clusters and the semiconductor, which facilitated the efficient separation of photogenerated electron–hole pairs at the interface. Moreover, the hybrid photocatalysts demonstrated excellent stability over three consecutive recycling tests. The AQY of the system, measured using monochromatic light at a wavelength of 420 ± 10 nm, was determined to be 0.2%. Recently, Yan et al.53 developed a Z-type heterojunction photocatalyst using Au25(GS)18 clusters with controlled interfacial charge. The MoSe2/CdSe/Au25 (M/C/A) catalyst was formed by modifying MoSe2 nanosheets with mercaptoacetic acid to introduce a negative charge, while CdSe quantum dots were treated with 2-aminoethanethiol to impart a positive charge. These modified components were self-assembled into a binary heterostructure, which was then combined with Au25 clusters, forming a ternary heterostructure (Fig. 4(a)). The photocatalytic hydrogen production results showed negligible activity for MoSe2 and CdSe alone and significantly improved activity for the MoSe2/CdSe (M/C) heterostructure (Fig. 4(b1–b3)). A comparison with a mechanically mixed catalyst revealed that the presence of all three components, with charge regulation, was key for maximising hydrogen production. The improved performance was attributed to MoSe2 accelerating electron transfer from CdSe and reducing electron–hole recombination. The electron transfer mechanism follows a Z-scheme: under visible light, electrons in Au25(GS)18 are excited and recombine with CdSe holes, while the remaining electrons transfer to MoSe2, enhancing the catalytic efficiency as shown in Fig. 4(c).
![]() | ||
Fig. 4 (a) Synthesis route of MoSe2/CdSe/Au25, (b) hydrogen production activities of MoSe2, CdSe and Mx/C, and (c) photocatalytic mechanism over the MoSe2/CdSe/Au25 heterostructure. Reproduced with permission from ref. 53. Copyright 2023 Wiley-VCH. |
In NC-based heterostructures, ligands at the interface often hinder interaction between the reactants and the NCs, resulting in reduced catalytic activity. While calcination can remove some of these ligands, excessive calcination may cause NC aggregation, leading to a loss of size-dependent catalytic properties. Kawawaki et al.54 investigated the effects of temperature on ligand removal and NC size changes, using BaLa4Ti4O15-supported 2-phenylethane thiolate-protected Au25(SR)18 NCs as a model. The desorption of ligands occurs in three stages as the temperature increases: (1) dissociation of ligands from the NC surface, (2) adsorption of the resulting compounds onto the support, and (3) desorption of these compounds from the support. The phenomenon that occurs during the calcination of Au25(PET, p-MBA)18/BaLa4Ti4O15 is shown in Fig. 5(a). In this system (i.e. Au25/BaLa4Ti4O15 and Cr2O3/Au25/BaLa4Ti4O15), the Au NCs only act as the co-catalyst and not as light absorbers (Fig. 5(b)). By carefully controlling the calcination process, the authors developed a highly active water-splitting photocatalyst with good stability, as shown in Fig. 5(c). Even earlier, Negishi et al.55,56 explored the photocatalytic water-splitting capabilities of gold NCs. They deposited Au25(GS)18 onto BaLa4Ti4O15 by removing the ligand at 300 °C under vacuum conditions. The photocatalytic activity of Au25-BaLa4Ti4O15 was found to be 2.6 times higher compared to AuNP-BaLa4Ti4O15 when exposed to light.
![]() | ||
Fig. 5 (a) Phenomenon that occurs during calcination of Au25(PET, p-MBA)18/BaLa4Ti4O15, (b) schematic illustration of photocatalytic water splitting using a one-step photoexcitation system where Au NCs only act as a co-catalyst, and (c) photocatalytic water-splitting activities of the composites. Reproduced with permission from ref. 54. Copyright 2021 Wiley-VCH. |
Alongside traditional wide-band-gap metal oxides, several new materials with photocatalytic hydrogen production activity have been recently developed. Li et al.57 used the reductive Ti vacancies in Ti3xC2Ty MXene to create a core–shell structure to form a ternary structure of CdS/Au/MXene. At the interface, they formed two Schottky barriers, facilitating charge transfer from CdS to Au NCs and from Au NCs to MXene. The optimised Au NC and MXene loadings on CdS achieved a hydrogen production rate of 5371 mmol g−1 h−1, which exceeded that of bare CdS. In addition to enhancing light absorption by serving as a photosensitiser, the NCs also acted as a small-band-gap semiconductor, enabling the formation of a heterojunction. Both Au and MXene, known for their excellent electrical conductivity, cooperatively improved the efficiency of charge separation and transfer in CdS, resulting in a greater number of electrons available for photocatalytic reactions. The apparent quantum efficiency (AQE) at 420 nm for bare CdS and the composite photocatalysts were 0.3% and 16.7%, respectively. Polymers have also proved successful in guiding the controlled assembly of NCs into specific configurations. The surface properties of these polymers often complement those of the surface ligands on NCs. Building on this, Bera et al.58 grafted gold NCs (≈2 nm) and super clusters (SCs) (≈100 nm) on two-dimensional polydopamine (PDA). AuSCs@PDA showed enhanced photostability, lower charge transfer resistance, and higher photocurrent responses than AuNCs@PDA, AuSCs, and PDA NPs, with the highest hydrogen evolution rate of 3.20 mmol g−1 h−1.
Wang et al.59 developed a type II photosystem by combining TiO2 with atomically precise Ag44(SR)30 NCs (SR = thiolate). The composite exhibited a hydrogen generation rate of 7.4 mmol h−1 g−1, which is ten times greater than that of pure TiO2 nanoparticles under the same conditions (Fig. 6(b)). In addition, the composite shows good stability, with 83% of the photocatalytic activity remaining after five cycles (Fig. 6(c)). This notable enhancement in activity is attributed to the extended photoresponse and improved charge carrier separation and transport. Ultrafast transient absorption (TA) spectroscopy revealed that Ag44 NCs function as both light absorbers and small-band-gap semiconductors under UV/vis light irradiation. Specifically, under visible light, the NCs act solely as light absorbers, while under UV/vis light, they also enhance charge separation efficiency as small-band-gap semiconductors (Fig. 6(a)). Thus, this type II photosystem enables the NCs to act as co-catalysts rather than just a photosensitiser.
![]() | ||
Fig. 6 (a) Schematic energy diagram of TiO2 before (black) and after (orange) deposition of NCs, (b) hydrogen evolution activity under UV/vis light irradiation, and (c) recyclability test of the composite. Reproduced with permission from ref. 59. Copyright 2020 Wiley-VCH. |
Zhu et al.60 constructed a Z-scheme heterojunction based on Pt1Ag28-BTT/CoP (BTT = 1,3,5-benzenetrithiol) through the Co–S interface (Fig. 7(a)). This heterostructure shows a marked improvement in photocatalytic hydrogen production, achieving a rate of 24.89 mmol h−1 g−1 compared to bare CoP. Additionally, the stable and well-integrated interface contributes to the long-term stability of the nanocomposite, with Pt1Ag28-BTT/CoP maintaining nearly consistent catalytic activity over five recycling tests (Fig. 7(b and c)). The composite also achieved an apparent quantum yield of 25.77% at 420 nm and retained approximately 100% activity. The improved photocatalytic hydrogen production performance is attributed to the strong internal electric field and the efficient Co–S interfacial charge transfer. The inclusion of central Pt atoms in Pt1Ag28-BTT improves the band alignment with CoP, while sulfur modification on the surface creates an effective charge transport pathway, boosting photocatalytic efficiency, as shown in Fig. 7(d).
![]() | ||
Fig. 7 (a) Schematic of fabrication of the Pt1Ag28-BTT/CoP heterojunction, (b) wavelength dependent hydrogen production and AQY of Pt1Ag28-BTT/CoP, (c) cycling stability of Pt1Ag28-BTT/CoP, and (d) charge transfer mechanism in Pt1Ag28-BTT/CoP. Reproduced with permission from ref. 60. Copyright 2024 Springer. |
Bootharaju et al.61 synthesized Au12Ag32(SePh)30 (SePh = phenylselenolate) NCs with a unique structure comprising an Au icosahedral core and an Ag dodecahedral shell using the galvanic exchange technique. These Au12Ag32(SePh)30 NCs demonstrated superior stability and near-infrared-II photoluminescence compared to Ag44(SePh)30 NCs, their homometallic counterpart. Taking advantage of the oxygen vacancies in TiO2, the Au12Ag32 NCs were anchored onto the TiO2 surface. The well-aligned energy levels between the Au12Ag32 (SePh)30 NCs and TiO2 facilitated better separation of photogenerated charge carriers (Fig. 8(c and d)). The Au12Ag32/TiO2 composite showed a remarkable increase in photocatalytic hydrogen production, reaching 6810 μmol g−1 h−1, approximately 6.2 times higher than that of Ag44/TiO2, as displayed in Fig. 8(a). The external quantum efficiency of Au12Ag32/TiO2 was determined to be 0.89%, 0.52%, and 0.96% for photoexcitation at 365 nm, 380 nm, and 400 nm, respectively. The Au12Ag32/TiO2 catalysts maintained stability for up to four months, with only a minor decrease (approximately 6%) in photocatalytic activity (Fig. 8(b)), highlighting the importance of tailoring NCs to improve H2 production and stability.
![]() | ||
Fig. 8 (a) Photocatalytic activity of the NCs and the composite photocatalysts, (b) H2 production by a fresh and aged Au12Ag32/TiO2 sample; inset bar diagram showing nearly intact photocatalytic activity of Au12Ag32/TiO2 clusters after four months of storage, and (c and d) photocatalytic H2 generation mechanism under solar irradiation. Reproduced with permission from ref. 61. Copyright 2023 Wiley-VCH. |
The extremely small size of NCs contributes to increased photocatalytic activity but also leads to limited stability. Some porous materials can serve as supports by accumulating these NCs on their surface, thereby reducing aggregation. Metal–organic frameworks (MOFs) and covalent organic frameworks (COFs) are examples of some widely explored porous functional supports made up of several ligands linked by covalent bonds. Their structurally adjustable nature allows for the rational design and precise control of NC deposition. Yao et al.62 developed a strategy to stabilise Au25(L-Cys)18 NCs on porous UiO-66-NH2 by creating covalent amide bonds (Fig. 9(a)). The strong metal–support interaction resulted in enhanced photocatalytic hydrogen production activity compared to the mechanically mixed UiO-66-NH2/Au25(PET)18 (PET = 2-phenylethanethiolate). As shown in Fig. 9(b), the covalently bonded UiO-66-NH2-Au25(L-Cys)18 demonstrated the best hydrogen production rate, 90 times higher than that of its bare counterpart. Moreover, the covalently bonded UiO-66-NH2-Au25(L-Cys)18 showed excellent stability (Fig. 9(c)), whereas the physically mixed UiO-66-NH2/Au25(PET)18 exhibited a significant decrease in catalytic performance during the second cycle. The improved performance is attributed to the covalent bond between UiO-66-NH2 and Au25(L-Cys)18, which strengthens the metal–support interaction, facilitating charge transfer and reducing charge carrier recombination.
![]() | ||
Fig. 9 (a) Different synthesis procedures for loading Au25 NCs on UiO-66-NH2, (b) photocatalytic H2 production activities of UiO-66-NH2, UiO-66-NH2-Au25(L-Cys)18 and UiO-66-NH2/Au25(PET)18, and (c) stability runs of UiO-66-NH2-Au25(L-Cys)18 and UiO-66-NH2/Au25(PET)18. Reproduced with permission from ref. 62. Copyright 2023 Springer Nature. |
Recently, Wang et al.63 integrated M1Ag24(BT)x (M = Ag, Pd, Pt or Au; BT = benzenethiolate) NCs into UIO-66-NH2via electrostatic interactions (Fig. 10(a)) to evaluate the influence of the central dopant on photocatalytic performance. The authors state that the incorporation of heteroatoms into NCs could contribute to the redistribution of the surface charge and change the physicochemical properties of the NCs. MAg24 NCs are negatively charged whereas UiO-66-NH2 is positively charged, as confirmed by zeta potential analysis. Therefore, MAg24 can be electrostatically attracted onto UiO-66-NH2 to form MAg24@UiO-66-NH2. This integration led to a 5.6- and 6.4-fold increase in the photocatalytic hydrogen production rate which is far superior to that of AgNPs@UiO-66-NH2 and Ag25 supported on UiO-66-NH2 respectively, under 380 nm light in a CH3CN-H2O medium with TEOA as the sacrificial agent (Fig. 10(b)). Additionally, they demonstrate outstanding photocatalytic recyclability as shown in Fig. 10(c). X-ray photoelectron spectroscopy (XPS) under dark and illuminated conditions confirmed the existence of a Z-scheme charge transfer between M1Ag24 NCs and UiO-66-NH2 (Fig. 10(d)). The atomically precise structure allows for a detailed exploration of structure–property relationships. The AQE of AuAg24@UiO-66-NH2, measured under 380 nm light irradiation, is approximately 0.53%.
![]() | ||
Fig. 10 (a) Synthesis route to fabricate MAg24@UiO-66-NH2, (b) photocatalytic hydrogen production activities of MAg24@UiO-66-NH2, (c) photocatalytic cycling performance of AuAg24@UiO-66-NH2, and (d) Z-scheme heterojunction formation at the MAg24–MOF interface. Reproduced with permission from ref. 63. Copyright 2024 Wiley-VCH. |
In 2013, Kagalwala et al.64 showcased the impressive photocatalytic hydrogen production activity of Ni-based clusters. They employed Ni6(SR)12 clusters as the photocatalyst, in the presence of triethylamine (TEA) scavenger and a photosensitizer. Their system achieved a turnover frequency (TOF) of 970 h−1. This study highlighted that Ni-based clusters can be good candidates for photocatalytic hydrogen production reactions. Later, Tian et al.65 incorporated Ni6(BT)12 clusters onto TiO2, creating a composite catalyst that markedly improved photocatalytic hydrogen production under simulated sunlight in a water–methanol medium. The Ni6/TiO2 photocatalyst retained its structure after the photocatalytic process as confirmed by the diffusion reflectance spectroscopy (DRS), X-ray photoelectron spectroscopy (XPS) and powder X-ray diffraction (PXRD) studies that were carried out before and after the photocatalytic reactions. Ni6(SR)12 exhibited distinct absorption peaks in the UV-visible spectrum (Fig. 11(a)). Fig. 11(c) illustrates the photocatalytic hydrogen production efficiency of catalysts with varying Ni6 loadings, with greater cluster content leading to improved hydrogen production activity. It was observed that visible light significantly improves the hydrogen evolution efficiency of the Ni6/TiO2 composite. However, visible light alone does not activate the photocatalytic activity of the Ni6/TiO2 composite, as the amount of hydrogen produced under these conditions remains negligible even after 8 hours of light exposure (Fig. 11(b)). A dynamic electron transfer process from TiO2 to Ni6(SCH2Ph)12 NCs was proposed, where the photogenerated electrons of TiO2 transfer to the near HOMO occupied orbitals of Ni6(SCH2Ph)12 (Fig. 11(d)). Such a charge transfer process is similar to the Z-scheme mechanism in heterojunction photocatalysis. The AQE for hydrogen evolution by the Ni6/TiO2 photocatalyst is 7.8% at 365 nm, although this high performance is dependent on the presence of a sacrificial reagent, with no significant hydrogen evolution observed without it.
![]() | ||
Fig. 11 (a) UV/vis absorption spectrum of Ni6(SR)12, (b) H2 activity over Ni6/TiO2 under different light irradiation, (c) hydrogen evolution activity with varying cluster loadings under a solar simulator, (d) charge transfer mechanism in Ni6/TiO2, (e) Ni–S toroidal geometry, and (f) the full molecular structure of Ni12(SR)24 (atom notations: Ni, blue; yellow, S; gray, C; and white, H). Reproduced with permission from ref. 65 and 66. Copyright 2021 Elsevier and American Chemical Society. |
Later, they examined the influence of the passivation layer on photocatalytic performance, using g-C3N4, which has delocalized π bonds, incorporating Ni12(SR)24 as the cocatalyst. The Ni12 cluster consisted of a ring structure, protected by methylbenzene thiolate ligands (Fig. 11e and f). The cluster's absorption in the UV-visible range primarily resulted from electronic transition within the Ni–S ring structure and from the cluster core to the benzene ring's π-bonds in the ligands. Upon exposure to UV light, electrons migrated from the core of the ring structure to the benzene ring present in the ligand. Neither Ni12 nor g-C3N4 exhibited photocatalytic activity in their bare form, but Ni12-modified g-C3N4 demonstrated a high hydrogen evolution rate, which increased with higher cluster loading.66 In comparison, Wei et al.67 utilized a smaller hexanuclear Ni cluster of the type [Ni6(SC2H4Ph)12] (Ni6) as a cocatalyst supported on g-C3N4 nanosheets. The electrostatic interaction between the Ni6 clusters and g-C3N4 nanosheets led to the formation of a stable Ni6/g-C3N4 hybrid photocatalyst. Under visible light irradiation, in the presence of the triethanolamine scavenger, the photocatalytic hydrogen evolution activity of Ni6/g-C3N4 significantly increased, reaching 5.87 mmol h−1 g−1. Upon illumination, photogenerated electrons were transferred from Ni6 to the surface of g-C3N4, resulting in a substantial improvement over bare g-C3N4. The AQY of the Ni6/g-C3N4 catalyst was determined to be 8.9% at 400 nm.
Cu-based clusters have also demonstrated remarkable catalytic properties for photochemical H2 production. Cao et al.68 introduced [Cu20O1(C20H24O2)12(CH3COO)6] (UJN-Cu20), stabilized by ethinylestradiol ligands and acetate anions (Fig. 12(a)). They combined UJN-Cu20 with TiO2 nanosheets to create a highly efficient composite for hydrogen evolution photocatalysis. When tested with varying loadings of UJN-Cu20, and under a 300 W xenon lamp simulating sunlight, the UJN-Cu20@TiO2 composite significantly outperformed the bare materials, which produced only trace amounts of hydrogen. The optimal H2 production rate of 13 mmol g−1 h−1 was achieved with 2% UJN-Cu20 loading (Fig. 12(b)). The photocurrent response curve as shown in Fig. 12(d) revealed that while both TiO2 and UJN-Cu20 showed weak individual responses, the composite exhibited a substantial increase in photoelectric response. This indicated that the combination of UJN-Cu20 clusters and TiO2 greatly enhanced the separation of photogenerated electrons and holes. A Z-type photocatalytic system was constructed, significantly promoting this separation process, as shown in the electron transfer mechanism diagram (Fig. 12(c)).
![]() | ||
Fig. 12 (a) The intrinsically chiral structure of O@Cu20 in UJN-Cu20, (b) photocatalytic activities of UJN-Cu20@TiO2 containing different wt% UJN-Cu20, (c) photocatalytic H2 production mechanism of UJN-Cu20@TiO2, and (d) photocurrent measurements of TiO2, UJN-Cu20 and UJN-Cu20@TiO2. Reproduced with permission from ref. 68. Copyright 2021 Royal Society of chemistry. |
Xu et al.69 fabricated Cu8(SN)4(tBuS)4 (SN = 4-(4-pyridinyl) thiazole-2-thiol) with a well-defined structure. They then utilized this cluster as a metal node to construct a copper-based MOF. In this setup, SN acted as an organic linker/Janus type ligand with dual coordination sites, with the thiolate and pyridyl directing the self-assembly of the solid-state framework (Fig. 13(a)). The MOF, when tested with fluorescein and triethylamine as photosensitizers and sacrificial agents, exhibited good photocatalytic hydrogen production performance (Fig. 13b). Furthermore, the Cu-MOF demonstrated relatively good stability, with performance maintained over three cycles (Fig. 13c). The addition of fluorescein significantly improved the photocurrent response of the Cu-MOF, suggesting effective electron transfer from the photosensitizer to the catalyst and efficient separation and movement of photogenerated electrons. Compared to Cu8(SN)4(tBuS)4 units, the ordered porous framework of the Cu-MOF facilitated more efficient charge transfer, contributing to its superior photocatalytic activity.
![]() | ||
Fig. 13 (a) Synthesis procedure of copper cluster-based MOFs via secondary assembly of copper cluster units, (b) photocatalytic H2 evolution performance of CCMOF and the Cu8SN4 cluster, and (c) cycling stability of CCMOF. Reproduced with permission from ref. 69. Copyright 2023 Royal Society of chemistry. |
Cao et al.70 developed a series of anion templated clusters by introducing [MoOS3]2− units into the CuI clusters to form a series of atomically precise MoVI–CuI bimetallic clusters of [Cu6(MoOS3)2(C6H5(CH2)S)2(P(C6H4-R)3)4]·xCH3CN ((C6H5(CH2)S = benzyl mercaptan; P(C6H4-R)3 = triphenylphosphine; R = H, CH3, or F)). This modification not only enhances the structural stability of the Cu clusters but also introduces additional active sites. When these clusters were combined with Fe3O4, the resulting composite catalyst showed remarkable catalytic performance and stability in photocatalytic hydrogen production. The improved performance is due to the electronic effects of the ligands, which fine-tune the HOMO–LUMO levels of the clusters. This innovative copper cluster design provides a significant boost to photocatalytic hydrogen production. Moreover, three similar Cu–Mo heterometal clusters, incorporating [MoOS3]2− units and Cu(I) thiolate clusters with various substituents (CH3, H, or F), were engineered to adjust the electron effects of the stabilizing ligands. The photocatalytic hydrogen evolution activity of these clusters is closely related to their LUMO and HOMO energy levels, which control the efficiency of photo-generated charge carrier transfer in the photocatalytic process.
Zhang et al.71 incorporated thiomolybdate [Mo2S12]2− NCs onto TiO2via a simple impregnation method. The composite photocatalyst showed an improved hydrogen evolution rate when exposed to light. The optimized composite achieves a hydrogen evolution rate of 213.1 μmol h−1 g−1, approximately 51 times greater than that of pure TiO2. The authors attributed the enhanced photocatalytic activity to the close contact formed between [Mo2S12]2− and TiO2 which enhances the separation of electron–hole pairs thus extending the charge carrier lifetime. Additionally, the abundant bridging sulfur atoms in [Mo2S12]2− serve as active sites for hydrogen evolution, further enhancing the hydrogen production rate. Wang et al.72 developed supertetrahedral T4 NCs of the type [Cd3In17Se31]5, which functioned as both visible light absorber and co-catalysts to aid in charge separation. When combined with TiO2, the [Cd3In17Se31]5/TiO2 composite demonstrated a significant enhancement in photocatalytic hydrogen production, reaching 328.2 μmol g−1 h−1 in a reaction solution containing triethanolamine as the sacrificial agent. The [Cd3In17Se31]5/TiO2 composite showed good stability, maintaining its performance with minimal change after 50 hours of photocatalytic hydrogen production testing. Wu et al.73 successfully incorporated a series of ultra-small Cd4−xZnxIn16S35 (x: 0–4) metal chalcogenide supertetrahedral NCs onto g-C3N4 to create a 0D/2D heterostructure for photocatalytic hydrogen production. The MCSNs, with their adjustable Cd ratios, allowed fine-tuning of the energy band structure, leading to improved electron–hole pair separation and reduced charge transfer resistance. The optimized Cd1Zn3/g-C3N4 heterostructure achieved a hydrogen production rate of 288 μmol g−1 h−1, approximately seven times higher than that of pure g-C3N4.
In some cases, the stability of NCs when exposed to light remains a major challenge. Key issues include the detachment of ligands, the nature of any detached fragments, and how these changes affect the catalytic performance. One approach is the development of advanced surface ligands that can protect NCs from photodegradation without compromising their catalytic or optical properties. Ligands play a crucial role in both the morphology and surface interactions. One primary function of the ligand is to provide solubility and improved dispersion in composites or the catalytic system. This could be a key area for atomically precise NCs, providing extremely uniform catalysts or composites. In addition, water-soluble NCs could be a beneficial aspect to explore both from a performance-based standpoint and in alignment with the sustainable energy goals. These water-soluble NCs can improve stability and dispersion in aqueous systems, which are ideal for photocatalytic water splitting. Research into the design of NCs with enhanced thermal and chemical stability will be essential for pushing the boundaries of their performance in real-world applications, where long-term stability is critical.
This journal is © The Royal Society of Chemistry 2025 |