3.1 Preparation and Characterization of CdS Nanosheets
To systematically investigate the impact of changes in the RHR and OHR on photocatalytic activity, it is essential to first prepare a reliable photocatalyst material with consistent energy levels and light absorption properties. We synthesized CdS NSs with uniform atomic-level thickness, ensuring a consistent bandgap and uniform light absorption. Figure 1a shows the absorption and photoluminescence (PL) spectra of the prepared CdS NSs. The absorption spectrum exhibits a distinct peak at 413 nm, indicating band-edge absorption and the successful synthesis of 6.5 ML CdS comprising seven cadmium layers and six sulfur layers.[18] The PL spectrum shows a slight band-edge emission at 430 nm due to Stokes shift, with most of the emission appearing as a broad trap-related emission from 450 nm to 800 nm, likely due to the presence of trap sites inevitably caused by the large lateral area.[19] Further analysis through XRD measurements confirmed the formation of a zincblende structure, essential for the growth of CdS in a two dimentional morphology.[20] Transmission electron microscopy (TEM) images revealed that the CdS had indeed grown into a two-dimensional nanosheet morphology. Due to high van der Waals forces, these nanosheets tend to curl, forming nanoscrolls.
For the prepared CdS NSs to be effective in hydrogen production with photocatalytic reaction, it was necessary to modify their surface from hydrophobic to hydrophilic to ensure colloidal stability in water. The original oleic acid capping ligands, which imparted hydrophobic characteristics, were replaced with S2- atomic ligands through a ligand exchange process.[18] These shorter atomic ligands facilitate charge transfer to the reactants during the photocatalytic reaction, thereby enhancing photocatalytic activity, while minus charge of S2- ensure colloidal stability. As shown in Figure S1, after the ligand exchange, the CdS NSs were successfully transferred from a non-polar solvent (hexane) to a polar solvent (NMF) and dispersed well, confirming the successful ligand exchange to S2- ligands. Significant changes in optical properties were observed in the absorption spectrum after the ligand exchange, including a red shift and broadening of the 1s peak as shown in Figure S1. The red shift is attributed to the weakening of quantum confinement due to the addition of an extra sulfur layer to the originally cadmium-terminated 6.5 ML structure.[21] The broadening is caused by electron and hole wavefunction leakage and changes in surface strain, resulting from the replacement of bulky oleic acid capping ligands with atomic ligands.[22, 23] These changes in surface strain are evident in the transformation of the morphology from curled nanoscrolls to flat sheets, as shown in Figure 1d.
3.2 Photo-deposition of Co-catalysts
To investigate how changes in the RHR rate affect the influence of different hole scavengers, we introduced Pt and Ru, known as excellent co-catalysts for hydrogen production, into the CdS NSs.[4, 9] The CdS NSs were dispersed in a solution containing a hole scavenger and metal precursors, followed by light irradiation. This process, known as photodeposition, allows the reduction of metal cations by conduction band electrons, leading to the growth of metal nanoparticles on the surface of CdS NSs, while the remaining holes in the valence band are consumed by the hole scavenger.[24] As shown in FigureS2, after the photodeposition process, we observed an additional broad absorption feature extending from the NIR to the visible region, attributed to the d-sp interband transition of the metals.[25] XPS measurements provided further confirmation of the successful growth of metal nanoparticles on CdS NSs as shown in Figure 2a, b. For Pt/CdS NSs, the Pt 4f7/2 and 4f5/2 peaks, originating from spin-orbital splitting, were observed at 72.2 eV and 75.5 eV, respectively, indicating the formation of a metallic Pt-CdS heterostructure.[26, 27] In the XPS measurements of Ru/CdS NSs, the Ru 3d5/2 peak was clearly observed at 279.6 eV.[28] Despite the overlap of the Ru 3d3/2 signal at approximately 284.8 eV with the C 1s signal around 284 eV, deconvolution successfully distinguished between the two, confirming the successful growth of metallic Ru on the CdS NSs.
3.3 Transient Absorption Spectroscopy and Bleaching Recovery
Typically, when metal nanoparticles are introduced as co-catalysts in semiconductor nanoparticles, the electrons generated in the photocatalyst are rapidly transferred to the co-catalyst, facilitating charge separation and enhancing the photocatalytic reaction.[9] To examine this electron transfer, we conducted femtosecond laser-based transient absorption experiments. In these measurements, the bleaching recovery kinetics of the 1S transition reflect the carrier population at the band edge. Specifically, for cadmium chalcogenides, the conduction band is two-fold degenerate while the valence band is six-fold degenerate, making the bleaching recovery kinetics of 1S transition sensitive to electron population.[29] Thus, analyzing the bleaching recovery kinetics of the 1S peak allows us to exclusively monitor electron dynamics.
Figures 2c-e display the transient absorption spectra of bare CdS NSs, Pt/CdS NSs, and Ru/CdS NSs as a function of delay time. For all samples, a bleaching signal near 450 nm was observed, corresponding to the band-edge position (1S transition) after ligand exchange. Over time, this bleaching gradually recovers. Notably, the bleaching recovery occurs more rapidly after the introduction of the co-catalyst, which is due to the electrons in the conduction band being transferred to the co-catalyst. To quantitatively analyze bleaching recovery, we plotted the kinetics of the bleaching recovery at the band-edge transition region, as shown in Figure 2f. This data can be fitted to a tri-exponential equation (See Table S1). For bare CdS NSs, we observed rapid decay components with time constants of a few picoseconds (τ1) and several hundred picoseconds (τ2), followed by a slower decay on the order of nanoseconds (τ3). Given that radiative recombination in nanosheets or nanoplatelets occurs on a nanosecond timescale, the fast decays are attributed to electron trapping.[30] Upon introducing metal co-catalysts, the amplitude of the fast decay component increased significantly, irrespective of the metal type, indicating electron transfer from CdS NSs to the metal. Notably, the amplitude of this fast decay component increased by 24% for Pt/CdS NSs and by 34% for Ru/CdS NSs, suggesting that Ru acts as a more efficient electron acceptor. This is due to differences in the Schottky barrier height (SBH) arising from the work function differences between the metals.[31] When a metal forms a junction with an n-type semiconductor like CdS, band bending occurs at the interface due to Fermi level equilibrium, resulting in the formation of a Schottky barrier. In spherical nanoparticles, there is no physical space to create a depletion region to induce band bending, resulting in a flat band potential and merely a shift in band-edge energy levels. However, for nanosheets, their lateral 2D morphology allows for band bending and Schottky barrier formation. Ru, with a work function of -4.5 eV, has a lower SBH compared to Pt/ which has a work function of -5.1 eV, facilitating more efficient electron transfer from CdS to Ru than to Pt.
3.4 Photocatalytic Hydrogen Evolution
To investigate the effect of different co-catalysts and hole scavengers on the hydrogen production reaction, we conducted photocatalytic experiments using bare CdS NSs and CdS NSs with Pt and Ru cocatalysts. In this experiment, CdS NSs were dispersed in water at a concentration of 10 mg/mL and exposed to 1 sun illumination at 1.5 AM as shown in Figure S3. The amount of hydrogen produced through the photocatalytic reaction was monitored in real-time using a microelectrode. To facilitate the production of valuable products from the oxidation reaction, BzOH was introduced as hole scavenger. For comparision, we also introduced mixture of NaSO4-NaSO3, which are commonly used as hole scavengers.[9]
Figure 3 illustrates the hydrogen production rate over time for experiments using various types of co-catalysts and different scavengers. Interestingly, the hydrogen production rate varied significantly depending on the type of hole scavenger and co-catalyst. In the case of bare CdS NSs, the results demonstrate that the hydrogen production rate is 870 μmol/g/h when using hole scavengers and 991 μmol/g/h when using BzOH, indicating comparable hydrogen production in both cases. When Pt was used as the co-catalyst, hydrogen production increased by approximately 3 times with BzOH as the hole scavenger, while using the Na2SO4-Na2SO3 mixture resulted in a hydrogen production rate of 4,554 μmol/g/h, which is about 5 times higher than that of bare CdS NSs, demonstrating a dramatic increase in hydrogen production. When Ru was used as the cocatalyst, hydrogen production increased compared to bare CdS NSs, but it was generally lower than that with Pt, regardless of the type of hole scavenger used. Additionally, hydrogen production rate was similar when using Na2SO4-Na2SO3 and BzOH as hole scavengers.
Understanding why such significant differences in hydrogen production occur based on the type of cocatalyst and scavenger requires a deep understanding of the photocatalytic mechanism of hydrogen production on CdS NSs in the presence of hole scavengers. Initially, without a co-catalyst, when the photocatalyst absorbs light and generates excitons, electrons are generated in the conduction band and holes in the valence band. The electrons participate in the reduction reaction to produce hydrogen, while the holes react with the hole scavenger, oxidizing it and consuming the holes. This full reduction-oxidation reaction cycle ensures that no charge remains on the photocatalyst, maintaining its neutrality and enabling a sustainable photocatalytic reaction. If the reduction reaction is slower than the oxidation reaction, the holes on the CdS NSs are quickly consumed, and the hydrogen production reaction becomes the rate-determining step (RDS) as depicted in Scheme 1a. This is the case for bare CdS NSs, which have a relatively slow hydrogen production rate due to the absence of a co-catalyst.
The redox potential of the Na2SO4-Na2SO3 mixture is -0.104 V vs. NHE, and that of BzOH is 1.98 V vs. NHE, indicating that oxidation process of Na2SO4-Na2SO3 mixture is faster than that of BzOH.[15, 16] Generally, hole scavengers with a faster oxidation half-reaction are expected to enhance photocatalytic activity, but since the reduction half-reaction for hydrogen production is the RDS, the type of scavenger does not significantly impact the hydrogen production rate, resulting in comparable hydrogen production characteristics as shown in Figure 3c.
Introducing Pt significantly increases the hydrogen production rate due to excellent catalytic properties of Pt, resulted by the Gibbs free energy for proton absorption/desorption being close to zero.[32] In the case of Pt/CdS NSs, electrons generated in the conduction band are rapidly transferred to Pt, where they are quickly consumed in the hydrogen production reaction, leaving the CdS NSs rich in holes in constrast to case of bare CdS NSs. This change shifts the RDS from the RHR to the OHR, indicating that type of scavenger is sensitive to hydrogen generation rate as depicted in Scheme 1b. Indeed, using Na2SO4-Na2SO3 as the hole scavenger with Pt/CdS NSs results in the highest hydrogen production, whereas the slower oxidation rate of BzOH leads to significantly lower hydrogen production. This demonstrates that when a co-catalyst for efficienct hydrogen production is introduced, the oxidation rate of the hole scavenger becomes critically important.
Lastly, the higher hydrogen production observed for Ru/CdS NSs compared to bare CdS NCs indicates a clear catalytic effect. However, despite the transient absorption (TA) measurements showing more efficient electron transfer from CdS NSs to Ru than to Pt, the hydrogen generation rate is lower with Ru compared with that of Pt. This discrepancy is attributed to the relatively lower catalytic properties of Ru, due to the imbalance in Gibbs free energy for proton absorption/desorption.[32] This lower catalytic activity of Ru compared to Pt suggests that, similar to bare CdS NSs, the RDS shifts back from the OHR to the RHR as shown in Scheme 1c. In fact, the hydrogen production rate remains consistent regardless of the type of hole scavenger, supporting the shift of the RDS to the RHR.
Our findings underscore the complexity of optimizing photocatalytic systems for hydrogen production. Unlike systems using conventional hole scavengers, even the most effective co-catalysts cannot prevent the OHR from becoming the RDS when BzOH is introduced, thereby limiting hydrogen production. This suggests that merely developing co-catalysts with high hydrogen production capability or enhancing charge separation is insufficient when BzOH is used. Instead, it is crucial to design photocatalysts that can significantly increase the OHR rate. Implementing photocatalysts with higher oxidation power, even with the same bandgap, can effectively shift the RDS away from the OHR. This study provides key insights into designing advanced photocatalysts that can manage varied reduction and oxidation rates, thereby optimizing both hydrogen production and the generation of valuable by-products.
3.5 Stability of photocatalyst depending on scavanger type
We conducted stability tests on bare CdS NSs, Pt/CdS NSs, and Ru/CdS NSs by using Na2SO4-Na2SO3 and BzOH as scavengers during the hydrogen genration reaction. After the photocatalytic reactions, XPS analysis was performed to assess the stability of each sample as shown in Figure 4. The XPS spectra displayed characteristic peaks around 161 eV (S 2p3/2) and 162 eV (S 2p1/2), corresponding to binding energy of Cd and S, and a higher energy peak around 165 eV, which is indicative of CdSO4 formation due to photo-corrosion of CdS NSs as a result of self-oxidation. As summarized in Figure 4d, we calculated the ratio of CdSO4 to CdS peak areas to quantify the extent of oxidation, where a higher ratio signifies more extensive oxidation. The results revealed that Pt/CdS NSs, which had the highest hydrogen production rate, exhibited the least oxidation overall. In contrast, bare CdS NSs showed significant oxidation, corresponding to their lower hydrogen production rate, leading increased residual carrier in the CdS NSs.
A significant observation was the substantial increase in oxidation when BzOH was used as hole scavenger compared to Na2SO4-Na2SO3 across all samples. This increase is attributed to the slower oxidation rate of BzOH, which causes holes to remain on the CdS surface for a longer duration, leading to increased self-oxidation. This self-oxidation was evident from the color change of the photocatalyst material. After conducting photocatalytic experiments with BzOH, the originally yellow CdS NSs sample became a slightly darker yellow due to self-oxidation, whereas there was almost no change in the samples using Na2SO4-Na2SO3 as shown in Figure S4. Consequently, ensuring the stability of CdS NSs becomes more challenging with BzOH due to its slower oxidation rate compared to faster scavengers like Na2SO4-Na2SO3. Given the necessity of rapid hole removal for stability, designing photocatalysts with high oxidation power is crucial. Lowering the valence band level to enhance the oxidation power of photocatalysts can be a key strategy. This approach not only improves hydrogen production rates but also minimizes self-oxidation