Next Article in Journal
Hydrocracking of Polyethylene to Gasoline-Range Hydrocarbons over a Ruthenium-Zeolite Bifunctional Catalyst System with Optimal Synergy of Metal and Acid Sites
Next Article in Special Issue
Occurrence, Ecotoxicity, and Photocatalytic Remediation of Antiretroviral Drugs in Global Surface Water Matrices
Previous Article in Journal
Electro-Refinery in Organics to Produce Energy Carriers: Co-Generation of Green Hydrogen and Carboxylic Acids by Glycerol Electrooxidation Using Dimensionally Stable Anode
Previous Article in Special Issue
Eco-Friendly Mechanochemical Fabrication of Polypyrrole/Ag-ZnO Heterostructures for Enhanced Photocatalytic Degradation of Methyl Orange
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced H2 Production Efficiency in Photo-Reforming of PET Waste Plastic Using Dark-Deposited Atom/Nanocomposite Pt/TiO2 Photocatalysts

Department of Civil Engineering, New Mexico State University, 3035 S Espina St, Las Cruces, NM 88003, USA
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(4), 334; https://doi.org/10.3390/catal15040334
Submission received: 27 February 2025 / Revised: 21 March 2025 / Accepted: 29 March 2025 / Published: 31 March 2025
(This article belongs to the Special Issue Advances in Photocatalytic Degradation)

Abstract

:
Photo-reforming waste polyethylene terephthalate (PET) in alkaline aqueous solutions is a novel approach for green hydrogen production. This study focuses on improving the catalytic efficiency of Pt/TiO2 for the photo-reforming of waste PET using an innovative dark deposition method to deposit Pt single atoms on nano TiO2 (Pt/TiO2), thereby increasing the catalytic efficiency while reducing the cost of the catalyst. The precursor concentration was optimized to control the size and distribution of the Pt clusters/atoms, and the TiO2 support was annealed at different temperatures to modify the properties of Pt/TiO2. Nine Pt/TiO2 catalysts were synthesized using different Pt precursor concentrations and annealing temperatures. The catalysts were characterized to measure their morphological, crystalline, and electronic properties, as well as their hydrogen yields via PET photo-reforming. The hydrogen conversion efficiency and external quantum yield (EQY) were calculated and compared with those of traditional direct-deposited catalysts. The correlation between the different characteristics of the dark-deposited and direct-deposited catalysts and their influence on the hydrogen yield in the photo-reforming process was statistically analyzed using principal component analysis. Catalysts deposited under dark conditions exhibited 5-fold and 7-fold enhancements in hydrogen conversion efficiency and EQY, respectively, compared to conventional catalytic systems. These findings indicate that the proposed catalytic system provides a viable solution for minimizing Pt loading, reducing the cost of the catalyst, and maintaining a higher hydrogen conversion efficiency.

1. Introduction

TiO2 has been extensively studied as a benchmark catalyst for solar energy conversion to hydrogen fuel because of its compatible bandgap, low cost, and high resistance to photocorrosion [1]. However, photocatalysis-based technologies for hydrogen production using TiO2 catalysts have been challenging because of their limited light absorption ability and the rapid recombination of photoexcited electrons and holes. Numerous doping strategies, morphology control methods, defect modulation techniques, and heterostructure designs have been developed to enhance the separation of photoinduced charge carriers [2]. Cocatalyst doping enhances the catalytic efficiency by forming a Schottky barrier between the semiconductor and noble metal, while extracting photogenerated electrons and inhibiting the recombination of charge carriers [3].
Single-atom (SA) catalysts have gained attention for various catalytic reactions because of their high reactivity and selectivity, minimal noble metal usage with subsequent cost reduction, and well-defined heterostructure interfaces for efficient charge separation [4]. The main reason for the development of SA catalysts is to maximize the surface-to-volume ratio and create novel reaction pathways at the SA level. The higher catalytic activity of SA-deposited nanoparticles can be attributed to the low co-ordination environment, quantum size effect, and improved strong metal-support interactions [5]. SA catalysts have versatile applications in thermal heterogeneous catalysis, organic electrosynthesis, fuel cells, batteries, electrolysis, oxygen reduction or evolution, hydrogen (H2) evolution or oxidation, carbon dioxide reduction, and nitrogen reduction reactions [6].
The synthesis method substantially affects the creation of robust metal SA on semiconductor supports [7,8]. SA deposition onto a metal oxide surface by tuning the co-precipitation conditions while controlling the metal precursor concentration can lead to excellent catalytic activities. In addition to the co-precipitation method, incipient wetness co-impregnation, chemical vapor deposition, newer atomic layer deposition, and galvanic replacement have been reported as possible SA deposition methods [7,9]. The precipitation methods for SA catalysts can be divided into bottom-up and top-down approaches. The bottom-up processes use mononuclear metal complexes as precursors to reduce the oxidation state of the metal through subsequent reduction processes [10,11]. In contrast, top-down approaches involve energy-related methods to break down the metal into isolated particles without affecting the structure of the support material [12]. The characteristics of the support (e.g., TiO2), such as its high specific area and high density of heteroatoms or defects, can enhance the metal-support interactions. However, selecting appropriate metal precursors is crucial, and controlling precursor loading may prevent agglomeration while generating uniformly distributed SAs [11,13].
Recent studies suggest that Ag, Au, and Pt nanoparticles have capacitive properties that can trap electrons and discharge them to suitable electron acceptors [14]. Additionally, the concept of memory catalysis demonstrates the ability of catalysts to function in the dark, which depends on the number of electrons trapped during the pre-irradiation process [15]. Controlling the size and valence state of the synthesized catalysts is crucial for promoting their activity [16]. The dark deposition method proposed in this study utilizes the capacitive nature of Pt and applies the catalytic memory concept to improve the electron storage capacity and the subsequent catalytic activity in the photo-reforming process. Compared to the dark deposition method, controlling the size and valence state of traditional direct-deposited catalysts is more difficult unless impregnation, atomic layer deposition, electrodeposition, and chemical reduction methods are applied, which require rigorous operating conditions or expensive equipment [17]. Furthermore, the traditional direct deposition method results in the coexistence of Pt, PtO, and PtO2 in the composite, which reduces its electron storage ability [18]. Hence, this study compares the catalytic activity and characteristics of direct-deposited and dark-deposited Pt on TiO2.
The choice of a strong support plays a critical role because of its influence as a physical support and on the electronic configuration of the metal [19]. The annealing temperature significantly affects the photocatalytic activity and microstructural characteristics of the support material (TiO2), such as the crystallinity, morphology, specific surface area, and porous structure of the catalyst. The highest photocatalytic activity was reported for P-25 Degussa TiO2 calcined at 650 °C for 3 h [20,21]. Calcination above 700 °C causes grain growth and phase conversion from anatase to rutile, which hinders the photocatalytic activity of TiO2 [20,22]. Moreover, the calcination atmosphere was found to have a significant effect on H2 production and decreased in the order Ar > air > N2 > vacuum∼H2 [23]. The low activity resulting from vacuum or H2 calcination can be attributed to the reduced coverage of the surface hydroxyl groups and the high density of the bulk defects [23]. Annealing the TiO2 support creates defect sites that assist in establishing strong metal-support interactions and influence the electronic configuration of the deposited SAs [24,25,26,27].
The objective of this study is to develop and optimize the photo-reforming process of polyethylene terephthalate (PET) for hydrogen production by synthesizing a dark-deposited Pt single atom on TiO2 nanoparticles with highly efficient catalytic activity at a lower cost. The nature of the catalysts was customized by altering the Pt loading conditions and applying different annealing temperatures to the TiO2 support. The catalytic activity and characteristics of the dark-deposited Pt/TiO2 were compared with those of the direct-deposited catalysts by analyzing the hydrogen evolution activity during the photo-reforming process. The novelty of this study lies in the optimization of hydrogen production with improved activity of the Pt/TiO2 catalytic system utilizing minimum Pt loading via a highly efficient single-atom dark deposition method.

2. Results and Discussion

The catalysts synthesized by the dark deposition method were characterized by identifying the particle size distribution, size of the deposited Pt atoms, and their surface profiles. The crystallinity of the catalysts was investigated by calculating the crystallite size, degree of crystallinity, and average TiO2 anatase phase percentage. Furthermore, variations in the bandgap energy and actual Pt loading on the catalysts were also analyzed. The electronic properties and influence of Pt deposition on the electronic properties of the catalysts were investigated using X-ray photoelectron spectroscopy (XPS). The characteristics of the dark-deposited catalyst were compared with those of the direct-deposited catalysts and pristine TiO2. Hydrogen production using these nine dark-deposited catalysts and three direct-deposited catalysts was compared, and principal component analysis (PCA) was performed to verify the correlation between catalyst characteristics and hydrogen evolution.

2.1. Characterization of the Catalysts with Transmission Electron Microscopy

The particle size distributions and morphologies of the catalysts synthesized under different conditions were studied using transmission electron microscopy (TEM), as shown in Figures S1A, S2A, and S3A. Surface profile plots of 0.04%(w/w) Pt/TiO2 annealed at 300 °C, 400 °C, and 500 °C are shown in Figures S1B, S2B, and S3B, respectively. The TEM images obscured the morphological changes. Hence, the effect of annealing on the TiO2 particle size was analyzed based on the particle size distribution curves obtained from the TEM images, as shown in Figure 1A. In our previous study, the average particle size of pristine TiO2 was 20.52 nm in our previous study [28]. The particle size of TiO2 did not change when the direct deposition method was applied [28].
When the annealing temperature was increased to 300 °C, the average particle size of Pt/TiO2 obtained using the dark deposition method increased. The growth rate of the particles was slow because of the low-temperature variation. This crystalline growth result is consistent when the annealing temperature is increased to 400 °C for the catalyst synthesized via the dark deposition method [29,30]. However, the average particle size of Pt/TiO2 decreased when the annealing temperature was increased to 500 °C because of the increased density and reduced porosity (Figure 1A) [31]. A decrease in the particle size was observed at an annealing temperature of 500 °C, leading to a comparatively homogeneous morphology of the catalyst. Hence, the results suggest that the morphology of Pt/TiO2 was affected by the annealing temperature, thereby changing the specific surface area and catalytic activity of the catalyst.
A scanning transmission electron microscopy (STEM) image of 0.7%(w/w) Pt/TiO2 prepared using the direct deposition method is shown in Figure 1B. Pt atoms were observed as individual SA and multimers, as shown in Figure 1B, and the corresponding particle size distribution of the Pt atoms is included in Figure 1B. The average particle size of Pt atoms was observed as 0.11 nm, confirming the abundance of Pt single atoms compared to agglomerated dimers and trimers (multimers). The deposition of Pt atoms was evident on the (004) crystal plane of anatase, which was identified by the corresponding d-spacing of 0.25 nm [32]. Figure 1C shows a STEM image of 0.4%(w/w) Pt/TiO2 prepared using the dark deposition method. A comparison of the images in Figure 1B,C illustrates that the dark deposition method provides more successful Pt deposition, even at a 42% lower Pt loading. However, the agglomeration of Pt atoms was observed when the average particle size was 0.32 nm. Upon reducing the precursor concentration by a factor of 10, the atomic size of the deposited Pt was reduced threefold (0.12 nm), irrespective of the annealing temperature (Figure 1D–F). The deposited size of Pt proves the deposition of single atoms with a reduced precursor concentration of 0.01 mM, avoiding agglomeration. Furthermore, the increased annealing temperature improved the crystallinity of the support structure (TiO2), and the crystal planes are visible in Figure 1C–E.
The distribution of Pt SA on the TiO2 surface was uneven, with clusters forming predominantly at the edges and corners. These localized SA clusters exhibited greater product selectivity and catalytic activity compared to macro-sized particles or metal powders. This enhanced performance is partly attributed to the active sites formed by atoms at specific nanoscale locations. However, the uniformity of the atomic environments in SA catalysts renders them more suitable for rational design and modeling, facilitating a deeper understanding of the reaction mechanisms [33,34]. The distribution pattern also influences metal-support interactions, impacting the stability, binding site characteristics, and reaction applicability, ultimately enhancing the catalytic efficiency across various chemical reactions. Pressure-controlled metal diffusion can improve single-atom loading by minimizing aggregation, thereby boosting the catalytic activity. Furthermore, atomic layer deposition and co-deposition techniques are crucial for achieving high-density single-atom sites and further optimizing catalyst performance [35,36].

2.2. X-Ray Diffraction (XRD) Analysis of the Catalysts

The XRD patterns of the catalysts synthesized under different conditions are presented in Figure S4. The degree of crystallinity, average crystallite size, and d-spacing were calculated from the XRD profiles, and variations in the parameters with annealing temperature and precursor concentration were analyzed. Figure S5A shows the variation in the degree of crystallinity and average crystallite size under different synthesis conditions. Although greater variability in crystallinity was observed (Figure S5A), all dark-deposited catalysts exhibited a high degree of crystallinity, consistently above 85%. Furthermore, the overall degree of crystallinity increased at a lower rate with increasing annealing temperature. The highest degree of crystallinity was observed in the catalysts synthesized at 400 °C with a precursor concentration of 0.01 mM (0.04%(w/w) Pt loading). As shown in Figure S5B, the average crystallite size increased with the temperature. During the annealing of P25 at 300 °C, the primary transformation involved crystallite growth and increased crystallinity rather than significant phase changes [37]. Anatase, a metastable phase, typically transforms into rutile at high temperatures (above 600 °C), while rutile remains stable. The variation in the crystallite size was further analyzed by plotting the crystallite size distribution curves for the catalysts synthesized with a 0.01 mM precursor concentration at different annealing temperatures (Figure 2A). The average crystallite size of the respective catalysts increased from 17 to 22 nm when the temperature was increased from 300 to 500 °C. The effect of temperature on the different phases of the catalysts was studied by analyzing the crystallite sizes of the anatase and rutile crystals in the crystalline nanoparticle mixture (Figure 2B).
Interestingly, the results illustrated that the degree of crystallinity and crystallite size varied in the same subset of samples annealed at the same temperature, regardless of the experimental conditions (Figure S5). These observed variations in crystallinity and crystallite size can be attributed to the different Pt precursor concentrations, which can influence the nucleation and growth kinetics of TiO2 crystallites, with higher Pt loadings potentially influencing the crystallites of the composite material depending on the Pt interaction with TiO2 surfaces [38]. Additionally, Pt decoration can enhance localized heating, facilitating the mobility of TiO2 species and influencing Ostwald ripening, which leads to differences in crystallite size [39]. Pt can also affect the internal stress and defect healing in TiO2 crystallites, contributing to variations in crystallinity [40]. Furthermore, Pt decoration alters the surface energy, which can affect the crystallite growth rates and result in differences in the crystallinity and crystallite size [41]. Overall, these results indicate that Pt loading significantly influences the crystallization dynamics of TiO2, even when annealed under the same conditions.
Anatase TiO2 showed a lower resistance to temperature, and a higher rate of crystalline growth was observed with larger crystallite sizes ranging from 15.1 nm to 17.9 nm. However, the rutile phase of TiO2 exhibited a slightly lower rate of crystalline growth with more stable crystallite sizes ranging from 11.3 nm to 13.6 nm. Hence, the observed average crystallite growth (Section 2.1) can be attributed to the lower-temperature resistance of anatase TiO2 [42]. The degree of crystallinity and average crystallite size were at optimum conditions at 500 °C, with lower variability. Hence, the XRD results showed a trend similar to that of the particle size distribution observed in the TEM images. The average crystallite size was slightly smaller than the average particle size observed in the TEM images, indicating the presence of a combination of crystallites [43]. The variation in the d-spacing was calculated and is presented in Table S1. The d value observed in the STEM analysis was verified as 0.24 nm for the anatase (004) lattice phase. However, the anatase TiO2 percentage in the synthesized catalysts increased gradually with increasing temperature (Table 1), likely due to the transformation of the rutile phase into anatase upon increasing the temperature from 300 to 500 °C [44]. Additionally, the phase transitions of TiO2 strongly depend on temperature, with minimal changes below 300 °C, where brookite and anatase remain stable [37,45]. However, when the temperature was between 300 and 500 °C, brookite gradually converted to anatase, thereby enhancing the crystallinity.
The degree of crystallinity of P25 TiO2 was significantly improved after annealing, and the subsequent anatase percentage in the catalysts synthesized using the dark deposition method was higher than that in those synthesized using the direct deposition method. Regardless of the synthesis method (dark or direct deposition), the crystallite sizes of both anatase and rutile in pristine TiO2 increased, as shown in Figure 2B.

2.3. Inductively Coupled Plasma Optical Emission Spectroscopy (ICP-OES) Analysis

ICP-OES analysis was performed to verify the actual Pt content deposited on the TiO2. Table 2 summarizes the actual Pt loading deposited on TiO2 and compares it with the expected maximum Pt loading, assuming that all Pt in the precursor solution was deposited on the TiO2 surface. The Pt loading was overestimated in the catalysts synthesized with a precursor concentration of 0.001 mM owing to instrument limitations. The measured Pt loading increased with increasing precursor concentration in the dark deposition method, while reaching the expected loading. Furthermore, Pt loading on the TiO2 surface increased with increasing annealing temperature. The catalyst prepared with a 0.01 mM precursor concentration after annealing at 400 °C showed the most effective Pt loading using the dark deposition method.
A higher Pt loading was observed in the catalysts prepared using dark deposition than in those prepared using direct deposition. Moreover, 98% of the expected Pt loading was achieved at a precursor concentration of 0.01 mM when the dark deposition method was applied. In contrast, even at a higher precursor concentration of 0.175 mM, the maximum precursor loading was 0.029%, which was only 4% of the expected loading using the direct deposition method. These results suggest that the dark deposition method is more efficient for Pt deposition than the direct deposition method. The effect of the annealing temperature on the Pt content of the direct and dark-deposited catalysts was compared by measuring the Pt content with and without annealed TiO2. The annealing temperature increased the Pt loading when a lower precursor concentration was used in the direct deposition method. When comparing the catalysts prepared with TiO2 annealed at 500 °C using the direct and dark deposition methods, the amount of Pt was 43% and 88% higher in the catalysts prepared using the dark deposition method at 0.01 and 0.1 mM precursor loadings, respectively, proving that the dark deposition method is an efficient synthesis method for achieving higher precipitant loading on the TiO2 surface.

2.4. XPS Analysis

The electronic properties of the selected 0.04%(w/w) Pt/TiO2 were analyzed using observed X-ray photoelectron XPS. The atomic and mass concentrations of Pt on TiO2 were calculated from the XPS data. The atomic concentration calculated using XPS for dark-deposited 0.04%(w/w) Pt/TiO2 was 0.06%, and the corresponding mass concentration was 0.45%. Hence, the XPS data overestimated the mass concentration. The XPS spectra were further analyzed to understand the electronic states of the deposited Pt layer. XPS spectra of the Pt 4f and Ti 2p regions of the direct-deposited Pt with a precursor concentration of 0.175 mM and dark-deposited Pt with a precursor concentration of 0.01 mM are compared in Figure 3. The peak positions in the dark-deposited Pt/TiO2 spectra increased with noticeable shape changes. Additional small peaks were evident at ~70, ~71, and ~74 eV, corresponding to the average oxidation states of Pt0, Pt2+, and Pt4+, respectively (Figure S6) [46]. These results indicate that the Pt species in the Pt single atoms of the dark-deposited Pt/TiO2 were partially oxidized with 0 to +4 valences because of the mutual electronic interactions between the Pt atoms and the TiO2 support. However, Pt0 and Pt4+ species were not observed in the direct-deposited Pt/TiO2, and the Pt2+ species at ~72.5 eV were evident, illustrating that the oxidation state of the Pt single atoms was +2 [47]. Because surface-coordinated Ptδ+ is critical to the high activity of the Pt/TiO2 system, it is crucial to understand the nature of the Pt single atoms on the hydrogen production activity while optimizing the activity at minimum Pt loading. Overall, the XPS results suggest that the two different synthesis methods deposited different Pt species on TiO2 while producing catalysts with different activities.

2.5. UV-Vis Absorbance of the Catalysts

The variation in the absorbance of the synthesized catalysts at different annealing temperatures and precursor concentrations of 0.1, 0.01, and 0.001 mM is presented in Figure S7. Considerable variation was not evident in the absorbance of the catalysts at different annealing temperatures; therefore, the bandgap energy of the catalysts did not change under different annealing temperatures. However, the precursor concentration and Pt loading affected the bandgap energy, as shown in Table 3. The lowest bandgap energy was observed for the catalyst prepared with a 0.1 mM precursor concentration after annealing at 400 °C. The catalysts prepared with a 0.01 mM Pt precursor concentration showed a bandgap energy of 2.34 eV under all the different annealing temperatures. Overall, a slight difference in the bandgap energies was observed when the precursor loading was changed from 0.1 mM to 0.01 mM. However, considerable bandgap broadening was observed when the precursor loading was further reduced to 0.001 mM.
The bandgap values obtained using the different synthesis methods are compared in Table 3. The bandgaps of the direct-deposited catalysts were much higher than those of the dark-deposited catalysts. A narrower bandgap results in a lower energy required for electron transition and can subsequently utilize a higher applicable maximum irradiation wavelength. The maximum wavelength that can be applied in the photo-reforming process in the presence of each catalyst was calculated and is presented in Table 3. All values fall within the visible light region of the light spectrum, indicating the feasibility of photocatalytic activity under solar light. Furthermore, the natural bandgap of pristine TiO2 is ~3.2 eV, and its maximum activating wavelength is 388 nm [48]. The development of TiO2 with Pt deposition using dark and direct deposition methods improved the bandgap energy while harvesting the optimum energy from the solar spectrum. Moreover, the light-harvesting efficiency of the dark-deposited catalysts was higher than that of the direct-deposited catalysts.
The correlations between the parameters were evaluated by calculating the Pearson correlation coefficient, as summarized in Figure S7D. The precursor concentration and annealing temperature exhibited a negative Pearson correlation, confirming a decrease in the bandgap. The correlation coefficient between the annealing temperature and average bandgap was −0.24, which indicates a weak negative correlation, suggesting a neutral role for the annealing temperature in bandgap modification. However, the precursor concentration and average bandgap exhibited a moderately negative correlation, suggesting that increased precursor loading has a minor effect on bandgap reduction.

2.6. Analysis of Hydrogen Production with Dark-Deposited Catalysts

The synthesized catalysts were applied to the PET photo-reforming system under optimum conditions, as described in Section 3.3. Figure 4 shows the variations in the hydrogen yield and hydrogen production rate of the different catalysts. Hydrogen production experiments were conducted for up to 10 d, depending on the hydrogen production rate. The maximum hydrogen yield after 10 days was observed in the system that was photo-reformed with catalysts synthesized at 0.04%(w/w) Pt loading and annealed at 500 °C. The observed reaction rate after 10 days was 36 μmol/gcat/h, which was the highest rate observed after 10 days. Other catalysts synthesized with the same precursor concentration, varying from annealing temperatures of 300 °C to 400 °C, generated hydrogen at stable rates fluctuating around 12–14 μmol/gcat/h. Hence, the catalyst prepared with a 0.01 mM precursor concentration can be identified as the most stable at all considered annealing temperatures. An increase in the precursor concentration positively affected the catalysts annealed at the lowest temperature of 300 °C. A decreasing trend in the hydrogen production rate was observed in the systems with a 0.1 mM precursor concentration, which were annealed at 400 °C and 500 °C. However, the lowest hydrogen yield was observed with the catalysts prepared with a precursor concentration of 0.001 mM at all the annealing temperatures considered. In general, these results suggest that the precursor loading has a considerable impact on hydrogen production. Catalysts prepared with moderate precursor concentrations have improved catalytic activity because lower precursor concentrations deplete the active sites, whereas higher precursor dosages increase the charge-recombination efficiency.
The differences in hydrogen yield (1100, 3300, and 4400 µmol/gcat) observed for Pt SAs on P25 TiO2 annealed at 300, 400, and 500 °C, despite similar Pt loading values from ICP-OES (0.346–0.380% w/w), can be attributed to a combination of factors beyond just the degree of crystallinity. The annealing temperature affects the crystallinity and phase composition of TiO2, influencing the charge-separation efficiency and electron mobility, which directly affects the photocatalytic performance [49]. Additionally, the chemical state of Pt, as indicated by the XPS analysis, varied with annealing, with higher temperatures favoring the presence of metallic Pt (Pt0), which is more catalytically active than the oxidized states (Pt2+ and Pt4+) typically observed at lower temperatures. Furthermore, the size and dispersion of the Pt SAs, as determined by TEM analysis, play crucial roles in the catalytic performance, as higher temperatures can lead to particle sintering, reducing the active surface area and consequently affecting hydrogen production [50]. The interaction between Pt and the TiO2 support is also influenced by annealing, potentially altering the electronic properties and enhancing charge-separation efficiency at the interface. Overall, the differences in catalytic activity were driven by a combination of crystallinity, Pt chemical state, and dispersion, necessitating further correlation of the ICP-OES, XPS, and TEM data to establish definitive relationships between these factors.
Single-atom catalysts, such as isolated Pt atoms on TiO2, play a crucial role in enhancing the hydrogen evolution reaction through multiple synergistic effects. On TiO2, single Pt atoms act as proton reduction sites while promoting hydrogen spillover, which induces surface oxygen vacancies and forms Pt-O-Ti3⁺ atomic interfaces. These interfaces enhance charge separation by facilitating electron transfer from Ti3⁺ defects to Pt atoms, suppressing recombination, and boosting the photocatalytic efficiency. Furthermore, Pt deposition shifts the valence band downward, altering its potential beyond the Nernstian predictions, which enhances hydroxyl radical (•OH) generation. Additionally, Pt atoms exhibit thermoneutral hydrogen adsorption energy (ΔG_H*) that stabilizes intermediates and accelerates the hydrogen evolution reaction. These effects collectively improve charge transfer, optimize the reaction kinetics, and enhance photocatalytic hydrogen production [51,52].

2.7. Comparison of Hydrogen Production with Different Catalysts

The hydrogen yields after 48 h of photo-reforming PET with the catalysts prepared using the dark and direct deposition methods are compared in Figure 5. The catalysts prepared using the dark deposition method with a 0.001 mM precursor loading showed the lowest hydrogen yields, irrespective of the annealing temperature. All the other catalysts prepared under the dark deposition method with a precursor concentration of more than or equal to 0.01 mM showed higher photo-reforming ability with higher hydrogen yield than catalysts prepared under the direct deposition method. An increase in the hydrogen yield was observed when the annealing temperature was increased. Baseline tests were conducted using synthesized catalysts with similar Pt precursor concentrations and annealed TiO2 at 500 °C using direct and dark deposition methods, as illustrated in Figure S8. The baseline test results showed that the annealing temperature did not enhance the hydrogen production ability of the directly deposited catalysts. In addition, a higher hydrogen yield was observed after 48 h in systems with a higher precursor loading prepared using the dark deposition method. The hydrogen yield with 0.1 mM precursor loading under the direct deposition method with TiO2 annealed at 500 °C was 4.8 μmol/gcat. In contrast, dark-deposited catalysts (0.1 mM annealed at 500 °C) achieved hydrogen yields that were more than 1000 times higher than those of the direct-deposited catalysts. Hence, Pt loading proportionally increased the hydrogen yield in the direct deposition method, representing a trend similar to that of the dark deposition method.
The variations in the hydrogen conversion efficiency and external quantum yield (EQY) under different synthesis conditions using the direct and dark deposition methods are shown in Figure S10. Both the hydrogen conversion efficiency and EQY exhibited similar variations in the hydrogen yield. Hydrogen conversion and EQY were higher for the dark-deposited catalysts than for the direct-deposited catalysts. The lowest H2 conversion and EQY under the dark-deposited catalysts were observed for the catalysts synthesized with a precursor concentration of 0.001 mM, regardless of the annealing temperature. The maximum hydrogen conversion was observed in the system with catalysts synthesized with a 0.1 mM precursor concentration, annealed at 500 °C. Increased precursor loading and annealing temperatures had a strong positive influence on the hydrogen conversion efficiency. When the precursor loading increased 10 times, the EQY amplified three times for the catalyst annealed at 500 °C; however, increasing the annealing temperature by 25% increased the EQY by ~2.5 times for constant Pt loading. Hence, tuning the supporting TiO2 structure at different annealing temperatures for constant Pt loading could improve the catalysts for excellent hydrogen conversion and higher quantum yield using the dark deposition method.
The recyclability of the Pt/TiO2 photocatalysts was systematically evaluated, addressing critical challenges related to catalyst deactivation and longevity (data not shown in the present study). With the ability to regenerate, optimized catalyst regeneration strategies enhanced both catalytic performance and long-term stability. Regeneration through calcination significantly restored the catalytic performance of Pt/TiO2 [53].
Table 4 compares the hydrogen conversion and EQY of the photo-reforming process applied in this study with those of other published studies that used a similar approach. Dark-deposited catalysts exhibited extremely high hydrogen conversion efficiencies compared to those reported in the literature. The hydrogen conversion efficiency of the dark-deposited catalyst with 0.4%(w/w) Pt loading was five times higher than that of the directly deposited catalyst with 0.7%(w/w) Pt loading. Although the reduction in Pt loading subsequently reduced the hydrogen conversion efficiency, it was four times higher than that of the direct-deposited catalyst. Additionally, the reported studies with CdS/CdOx Quantum dots and NCNCNx|Ni2P showed 50% lower hydrogen conversion efficiencies than the applied novel dark-deposited catalysts. Furthermore, the EQY of the synthesized material using the dark deposition method was 7% higher than that of its counterpart, leading to a 75% better catalytic performance than CdS/CdOx Quantum dots. Key factors influencing the differences in EQY include material properties such as the bandgap, defect density, and carrier mobility, as well as charge carrier dynamics and variations in the synthesis processes.
The performance of the catalysts prepared in this study was further compared with those developed in similar studies (Table S2). The maximum hydrogen production rate achieved here is approximately three times higher than that of TiO2 nanotubes decorated with single Pt atoms used for the photo-reforming of PET, as reported by Han et al. [54]. The incorporation of different sacrificial agents, such as glycerol, along with UV irradiation with a higher flux, has significantly enhanced the hydrogen production rate with 0.44%(w/w) Pt/TiO2 [55]. Pt/TiO2 catalysts have also been used for hydrogen production by the photo-reforming of polymers, such as polyethylene and polyvinyl chloride, in alkaline and acidic solutions [56,57]. Li et al. developed Cu/TiO2 single-atom catalyst and achieved a hydrogen production rate of 193.6 μmol catalyst⁻1 h⁻1 during photo-reforming PET, demonstrating performance comparable to the catalysts in this study (Table S2) [58]. Co-doping with additional metal catalysts, C/N elements, incorporating Fe-metal organic framework, and employing synthesis methods such as sol-gel and fast precipitation, further improved the photocatalytic hydrogen evolution rate [59,60,61]. Although some previous studies reported considerably higher hydrogen production rates than this study, it is due to the differences in key operating factors such as irradiated light intensity, reactor design, process (photo-reforming or water-splitting), sacrificial agent type, catalyst dosage, pH of the reaction solution, and pretreatment methods, which all significantly impact the hydrogen production rate. Therefore, a direct comparison may not be entirely appropriate for this study.
Table 4. Comparison of catalytic performance with that reported in the literature.
Table 4. Comparison of catalytic performance with that reported in the literature.
Experimental ConditionsLight SourceHydrogen Conversion Efficiency (%)EQY (%)
Direct-deposited 0.7%(w/w) Pt/TiO2 with PET substrate (This study)UVC LED 70 mW6.41
Dark-deposited 0.4%(w/w) Pt/TiO2 annealed at 500 °C with PET substrate (This study)UVC LED 70 mW33.47
Dark-deposited 0.04%(w/w) Pt/TiO2 annealed at 500 °C with PET substrate (This study)UVC LED 70 mW25.94
CdS/CdOx Quantum dots with PET substrate [62]Simulated solar light (AM 1.5 G, 100 mW/cm2)16.64
NCNCNx|Ni2P with PET substrate [62]Simulated solar light (AM 1.5 G, 100 mW/cm2)17.1-

2.8. Principal Component Analysis (PCA)

The hydrogen production of different catalysts prepared under different conditions was analyzed in this study by considering the varying characteristics of each catalyst. The correlation between the different catalyst characteristics and hydrogen evolution ability was statistically analyzed using PCA. Nine different parameters were considered during the analysis of hydrogen production at 48 h: average particle size, average bandgap energy, precursor concentration, annealing temperature, degree of crystallinity, average crystallite size, anatase percentage, and measured Pt loading. A scree plot representing the variation in the data-capturing ability of each principal component is presented in Figure S10. The scree plot suggests that the three principal components adequately describe the data while increasing the eigenvalues to the 4th principal component (PC) and above, resulting in eigenvalues close to zero.
Figure 6 shows the loading plot created by PCA, considering the three PCs, according to the scree plot. The loading plot illustrates the influence of each characteristic on the PC. The lengths and angles of the vectors represent the correlations between the variables. The cluster of vectors formed by parameters such as the hydrogen yield, measured Pt loading, annealing temperature, and degree of crystallinity showed that they were positively correlated. The two vectors forming a 90° angle represent non-correlated parameters. Therefore, the average crystallite size, precursor concentration, and average bandgap were not correlated with each other. The average bandgap exhibited a strong negative correlation with hydrogen production. Therefore, an increased bandgap reduces hydrogen production, as discussed in Section 2.5. Furthermore, there was a strong correlation between the measured Pt loading and hydrogen production, whereas the precursor concentration was weakly correlated. Therefore, improving the synthesis process by optimizing parameters such as the measured Pt loading, annealing temperature of TiO2, degree of crystallinity, and average bandgap is crucial for increasing the hydrogen yield via the photo-reforming process. The results of the statistical analysis further support the experimental analyses of the catalyst characteristics and the hydrogen yield. Table S3 lists the calculated correlation coefficients that represent the quantitative correlation between each parameter.
PCA of the Pt loading, XPS, and UV-Vis findings extracted (Table 2 and Table 3) illustrate that the first two principal components (PC1 and PC2) in Figure 7 account for ~59.74% and 24.96% of the total variance, respectively. The expected and measured Pt loading values under different synthesis conditions, bandgap energies, and maximum absorption wavelengths of the catalysts were analyzed to identify their respective trends and correlations. The PCA (Figure 7) shows distinct clustering patterns corresponding to the synthesis methods, with dark deposition (coded as 0) and direct deposition (coded as 1) forming separate groups. The primary contributors to PC1 appear to be the synthesis method and precursor concentration, whereas PC2 represents the variations in the bandgap energies and maximum absorption wavelengths.
A clear separation was observed between the dark deposition and direct deposition methods (Figure 7), with dark deposition samples clustering closely together, indicating more consistent properties, whereas the direct deposition samples showed greater variability. The precursor concentration also influences clustering, with lower concentrations clustering tightly and higher concentrations exhibiting greater dispersion, reflecting variations in Pt loading efficiency. The annealing temperature plays a significant role, with samples treated at higher temperatures (500 °C) forming separate clusters from those annealed at lower temperatures (300 °C and 400 °C), highlighting its impact on the crystallinity and optical properties. Furthermore, samples with similar bandgap values tend to form distinct clusters, suggesting a strong correlation between the synthesis conditions and material properties. These findings indicate that the synthesis conditions significantly impact the Pt loading and optical behavior, with dark deposition resulting in more stable material properties than direct deposition.

3. Methodology

The Pt/TiO2 catalysts were synthesized using two different methods, with a detailed discussion of the dark deposition method and a summary of the traditional direct deposition method in Section 3.1 and Section 3.2, respectively. Optimized hydrogen production in the photo-reforming process via the direct deposition method was reported in our previous work [53]. The catalysts synthesized using these two methods were characterized using various advanced methods to identify their crystalline, morphological, and electronic properties, and quantify hydrogen production. Principal component analysis was performed to elucidate the statistical correlation between the characterized catalytic properties and the measured hydrogen evolution.

3.1. Dark Deposition Method of Pt/TiO2

The catalysts were synthesized using the dark deposition method by modifying P25 Degussa TiO2 (P25 with 21 nm primary particle size, ≥99.5%, Sigma-Aldrich Inc., St. Louis, MO, USA). A schematic representation of the preparation method is presented in Figure 8. The catalytic activity and characteristics of nine different catalysts were analyzed and synthesized at varying annealing temperatures and precursor concentrations. First, 0.5 g of P25 Degussa TiO2 was annealed in air at 300, 400, and 500 °C for 1 h in the dark in a temperature-controlled muffle furnace (Vulcan 3-550, DENTSPLY International Inc., York, PA, USA). The annealed TiO2 nanoparticles were immersed in 10 mL methanol (HPLC grade, ≥99.9%, Sigma-Aldrich Inc., MO, USA) solution (50%(v/v)) containing hexachloroplatinic acid (H2PtCl6, 8%(w/w) in H2O, Sigma-Aldrich Inc., MO, USA) at concentrations of 0.1, 0.01, and 0.001 mM in separate batches. The 200 mL Pyrex volumetric flasks containing the solution mixtures were sealed, covered with aluminum foil, and kept in the dark for 24 h, stirring at 300 rpm and 25 °C for dark deposition. The solution temperature was controlled using a temperature monitoring probe sensor installed on a digital hotplate magnetic stirrer (ONiLAB MS-H280-Pro, DLAB Scientific Inc., Beijing, China). The catalysts were separated via a vacuum filtering system with 0.1 µm filter papers and washed with ethanol (HPLC grade, ≥99.9%, Sigma-Aldrich Inc., MO, USA) followed by deionized (DI) water for 15 min. Finally, the samples were air-dried at room temperature for 24 h.

3.2. Direct Deposition Method

The direct deposition method was performed using hexachloroplatinic acid solutions (H2PtCl6, 8%(w/w) in H2O, Sigma-Aldrich Inc., MO, USA) in 0.075, 0.175, and 0.275 mM concentrations to prepare 0.3%(w/w), 0.7%(w/w), and 1.5%(w/w) Pt loadings on TiO2. A baseline test was conducted using similar Pt dosages to those applied in the dark deposition method with Pt precursor concentrations of 0.1, 0.01, and 0.001 mM. P25 TiO2 was added with or without annealing during the synthesis process. TiO2 was annealed at 500 °C for 1 h in a muffle furnace (Vulcan 3-550, DENTSPLY International Inc., York, PA, USA). The deposition was performed under natural light conditions using 3 M NaOH (ACS reagent, K ≤0.02%, ≥98%, Sigma-Aldrich Inc., MO, USA) as the reducing agent in a 200 mL Pyrex volumetric flask. The solution temperature was increased to 100 °C while stirring. A temperature-controlling probe sensor installed in a digital hotplate magnetic stirrer (ONiLAB MS-H280-Pro, DLAB Scientific Inc., Beijing, China) was used to maintain the temperature. The detailed preparation method is presented in our previous study [28,53], and Figure 8 summarizes the two methods of Pt deposition on TiO2.

3.3. Characterization of the Catalysts

The morphology and particle size distribution of Pt/TiO2 were analyzed using an HT7650 transmission electron microscope (TEM, Hitachi High-Tech America, Inc., Pleasanton, CA, USA). In order to conduct the analysis, 0.1 g of catalyst was dispersed in DI water and diluted 1000 times. Then, 0.1 µL of the sample was placed on a carbon-supported copper grid (Formvar/Carbon-supported Copper Grids, Sigma-Aldrich, MO, USA) after sonicating for 10 min. TEM images were analyzed using Fiji software (Version 1.54i) to understand the impact of the annealing temperature on the TiO2 particle size distribution and surface profile [63]. The catalysts prepared under dark deposition with 0.01 mM precursor concentration after annealing at 300 °C, 400 °C, and 500 °C and with 0.1 mM precursor concentration after annealing at 500 °C were further investigated using a Jeol NEOARM 200CF Transmission Electron Microscope (JEOL USA, Inc., Peabody, MA, USA) equipped with a Hitachi 3rd-order spherical aberration corrector for the scanning TEM (STEM) probe to visualize the atomic deposition of Pt particles.
The crystallinity of the synthesized catalysts was analyzed using the powder XRD method, 9430 060 03002 Empyrean Series 2 X-ray diffraction system, Malvern Panalytical Ltd., Malvern WR14 1XZ, UK). The degree of crystallinity of the catalysts prepared under various conditions was calculated using Equation (1) [64].
Degree of crystallinity % = Sum of area under crystalline peaks Total area under curve × 100
The crystallite sizes of the catalysts for each synthesis method were calculated using the Scherrer formula, as shown in Equation (2) [65].
Crystallite size = K λ β . Cos θ
where λ is the X-ray wavelength (nm), β is the peak width of the diffraction peak profile at half-maximum height resulting from the small crystallite size in radians, and K is a constant related to the crystallite shape, which is generally taken as 0.9. The value of β on the 2θ axis of the diffraction profile is expressed in radians. θ is the diffraction angle in degrees or radians because Cosθ corresponds to the same number. The crystallite size was calculated for all 2θ values, and the average crystallite size was calculated.
The d-spacing (d), which is the distance between two consecutive lattice planes, was calculated using Bragg’s law (Equation (3)).
d = n λ 2 . sin θ
where n is the diffraction order, λ is the X-ray wavelength (nm), and θ is the diffraction angle of the XRD profile.
The percentage of anatase in TiO2 after synthesis was calculated (Equation (4)) by considering the area under the characteristic peaks of anatase (101) at 25.1° and rutile (110) at 27.2°. It was assumed that P25 TiO2 consisted of only the anatase and rutile phases.
A natase percentage % = Area under Anatase 101   peak Area under Anatase 101   peak + Area under Rutile 110 peak × 100
A Perkin Elmer Avio 550 Max Inductively coupled plasma optical emission spectroscopy (ICP-OES, Perkin Elmer Avio 550 Max ICP-OES, Waltham, MA, USA) analysis following the EPA 200.7 method was used to analyze the Pt concentration in the synthesized catalysts. The samples were digested using the EPA 3051A method, without hydrogen fluoride. Samples were pre-digested for 4 h before being placed in a microwave (Milestone Ethos UP microwave, Milestone, Sorisole, Italy), followed by the determination of elemental concentrations by considering the digestion dilution factors.
XPS of the selected synthesized catalyst sample with a precursor concentration of 0.01 mM at 400 °C was collected using a Kratos Axis Ultra 165 DLD Hybrid Ultrahigh Vacuum Photoelectron Spectrometer (Kratos Analytical, Manchester, UK). The sample was mounted on carbon tape and pressed onto indium foil before analysis. The Pt 4f region was observed at a very low intensity, and a fit was performed using a Shirley background, assuming that the Pt 4f7/2 was the peak at 72.8 eV and the second peak was added corresponding to the position of the Pt 4f 5/2 with approximately 0.75 times the intensity of the first peak. This was used to calculate the percentage of the atomic composition.
The absorbance spectra in the ultraviolet (UV) and visible light range from 200 nm to 900 nm were measured using a spectrophotometer (DR6000 ultraviolet-visible spectrophotometer, Hach Company, Loveland, CO, USA). The direct and indirect bandgaps were estimated using the Tauc plot method, as described in our previous study [28]. Equation (5) was used to calculate the maximum irradiation wavelength required to initiate electron transition.
λ = h c E g
where λ is the wavelength of light (nm), h is Planck’s constant in Jule × s (Js), c is the speed of light (m/s), and Eg is the bandgap energy in J.
The Pearson correlation coefficients between parameters such as precursor concentration, annealing temperature, bandgap energy, and hydrogen yield were calculated using Origin(Pro) software [66].

3.4. Evaluation of Hydrogen Production

The hydrogen yield was measured in quartz reactors by photo-reforming of 0.5 g pretreated PET in 10 mL of an ethanol and water mixture (3:2) with 5%(w/w) NaOH. The reaction solution was diluted with 40 mL of deionized (DI) water. The synthesized catalyst (0.5 g) was added to this mixture and irradiated with light, as described in our previous studies (UV mercury vapor lamp (160W PUV-10, Zoo Med Laboratories, San Luis Obispo, CA, USA), which emitted minor peaks at 290, 315, and 335 nm, with a dominant peak at 365 nm. The light flux was measured at 60 mW/cm2) [28,67]. Hydrogen was measured at 24-h intervals using a gas chromatograph equipped with a thermal conductivity detector. Each measurement was replicated, and the standard deviation was determined.
The hydrogen conversion efficiency was calculated using stoichiometric hydrogen conversion and the photo-reforming reaction with PET in the substrate. The stoichiometric hydrogen conversion reaction in alkaline media is presented in Equations (6) and (7), assuming that hydrogen production from terephthalate is negligible [62].
C 10 H 8 O 4 ( a q ) + 2 H 2 O ( l ) C 8 H 6 O 4 ( a q ) + C 2 H 6 O 2 a q
C 2 H 6 O 2 ( a q ) + 2 H 2 O ( l ) 5 H 2 ( g ) + 2 C O 2 ( g )
H2 conversion after 48 h of irradiation was calculated using Equation (8) [62]:
H 2   C o n v e r s i o n   ( % ) = 100 × N H 2 , o b s N s u b s t r a t e , o b s N H 2 , i d e a l N s u b s t r a t e , i d e a l
where N(H2, obs) is the observed H2 yield at 48 h in moles, N(substrate, obs) is the amount of substrate used in the experiment in moles, and N(H2,ideal)/N(substrate, ideal) is the ratio of H2 in the substrate under stoichiometry derived from Equations (6) and (7).
The external quantum yield (EQY) was determined by considering the H2 yield after 1 h of light irradiation with UV-LED lamps containing eight single UV-LED components (model: KL265–50 V-SM-WD, manufactured by Crystal IS (Green Island, NY, USA); UVC LED 70 mW with a single wavelength of 265 nm). Baseline tests were conducted using water, a water-ethanol mixture, and a mixture of water, ethanol, and PET with a constant catalyst dosage. Equation (9) was used to calculate the EQY [62].
E Q Y % = 100 × k N H 2 N A h c t i r r λ I A
The number of transferred electrons, denoted as k, was assumed to be two in the baseline tests for water-splitting and 12 in the ethanol reforming. For the photo-reforming of PET (assuming ethylene glycol reforming), k was assumed to be 10, NH2 was the amount of H2 generated in moles after 1 h of reaction, NA was Avogadro’s constant (6.022 × 1023 mol−1), h was Planck’s constant (6.626 × 10−34 J s), c was the speed of light (3 × 108 m s −1), tirr was the irradiation time in seconds (s), λ was the irradiation wavelength in meters (m), I was the light intensity in watts per square meter (W/m2), and A was the irradiated area in square meters (m2).

3.5. Method of Principal Component Analysis (PCA)

PCA was performed using Origin(pro) 2024 software [66]. PCA was used to quantify the correlation between the catalyst characteristics and hydrogen production. Hence, a scree plot was created from the data concerning parameters such as the hydrogen yield at 48 h, precursor concentration, annealing temperature, average bandgap, observed Pt loading, average crystallinity, average particle size, degree of crystallinity, and anatase percentage in the catalyst to evaluate the required number of principal components. The scree plot suggests the extraction of three principal components that represent more than 90% of the data. PCA was based on the correlation matrix of the nine parameters considered. A loading plot was created based on the calculated eigenvalues and eigenvectors, depicting the nature of the correlation between each parameter and its principal component.

4. Conclusions and Future Perspectives

Pt deposition on TiO2 has been extensively studied to achieve optimum hydrogen conversion during the photo-reforming of waste PET. An innovative dark deposition method was introduced in this study using the capacitive nature of Pt and applying the memory catalyst concept. The concentration of the Pt precursor was identified as a significant parameter that controlled the size of deposited Pt particles. Therefore, the Pt precursor concentration was reduced 10 and 100 times while depositing single atoms on the TiO2 surface. Additionally, the surface defects of TiO2 play a significant role in creating strong metal-support interactions, establishing better support for the deposited Pt particles by optimizing the electronic properties of the catalytic system. The surface defects were optimized by annealing TiO2 at three different temperatures. Increasing the annealing temperature to 500 °C improved the supported Pt deposition, and a 33% higher hydrogen conversion was observed with the catalysts annealed at 500 °C with a 0.1 mM precursor concentration. The novel catalysts prepared using the dark deposition method proved to be twice as efficient as other reported conversion efficiencies with CdS/CdOx Quantum dots and NCNCNx|Ni2P. In addition, the highest EQY of 7% was observed when dark-deposited catalysts were used in the photo-reforming system. Single atoms, dimers, and trimers were observed when the precursor concentration was reduced by a factor of 10, providing a more stable hydrogen yield for 10 days with a hydrogen conversion efficiency of 26%.
The characteristics of the catalysts and hydrogen yield were compared with those of traditional direct-deposited catalysts. Owing to their different catalytic activities, significant differences in bandgap energy, particle size, crystallite size, and degree of crystallinity were observed between the direct and dark-deposited catalysts. Catalysts prepared via dark deposition exhibited higher hydrogen yields, hydrogen conversions, and external quantum yields. It can be concluded that the novel catalysts prepared by the dark deposition method are low-cost and highly efficient for the photodegradation of waste PET with minimum precursor loading. Single-atom catalysts can be fine-tuned by further increasing the annealing temperature. The effect of calcination on the stability of the synthesized catalysts should be studied. The recyclability of the catalysts and the stability of the novel catalysts synthesized via dark deposition should be studied, as these are important for large-scale economic and environmental applications. In addition, the performance of the catalysts under natural light should be studied, and the quantum yield should be calculated before the catalyst is industrialized.

5. Patents

Wang, H.; Edirisooriya E.M.N.T; Xu, P. Method of Hydrogen Manufacture. U.S. Patent Application 18/378586, 2025 [68].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15040334/s1, Figure S1: (A) TEM image and (B) surface profile plot of 0.04% Pt/TiO2 annealed at 300 °C; Figure S2: (A) TEM image and (B) surface profile plot of 0.04% Pt/TiO2 annealed at 400 °C; Figure S3: (A) TEM image and (B) surface profile plot of 0.04% Pt/TiO2 annealed at 500 °C; Figure S4: X-ray diffraction patterns of Pt/TiO2 synthesized under different conditions; Figure S5: Variation of (A) degree of crystallinity and (B) average crystallite size of catalysts synthesized under different conditions; Figure S6: Pt 4f fit of Pt/TiO2 nanopowder by X-ray photoelectron spectroscopy; Figure S7: Absorbance spectra of catalysts prepared under dark deposition method with (A) 0.1 mM, (B) 0.01 mM, and (C) 0.001 mM precursor concentrations and (D) correlation of different parameters on H2 production; Figure S8: Comparison of hydrogen yield with direct and dark-deposited catalysts synthesized with annealed TiO2 at 500 °C; Figure S9: Scree plot of the principal component analysis; Figure S10: Variation in the hydrogen conversion efficiency and eternal quantum yield (EQY) under different synthesis conditions; Table S1: Summary of d-spacing for different lattice phases; Table S2: Comparison of hydrogen production rate with similar catalysts; Table S3: Correlation matrix after principal component analysis. Refs. [54,55,56,57,58,59,60,69] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, E.M.N.T.E., P.S.S., P.X. and H.W.; methodology, E.M.N.T.E., P.S.S., P.X. and H.W.; software, E.M.N.T.E. and P.S.S.; validation, E.M.N.T.E., P.S.S., P.X. and H.W.; formal analysis, E.M.N.T.E. and P.S.S.; investigation, E.M.N.T.E., P.S.S., P.X. and H.W.; resources, E.M.N.T.E., P.S.S., P.X. and H.W.; data curation, E.M.N.T.E. and P.S.S.; writing—original draft preparation, E.M.N.T.E., P.S.S., P.X. and H.W.; writing—review and editing, E.M.N.T.E., P.S.S., P.X. and H.W.; visualization, E.M.N.T.E. and P.S.S.; supervision, P.X. and H.W.; project administration, P.X. and H.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article and its Supplementary Materials.

Conflicts of Interest

The authors declare that they have no affiliations with or involvement in any organization or entity with any financial interest in the subject matter or materials discussed in this manuscript.

References

  1. Eidsvåg, H.; Bentouba, S.; Vajeeston, P.; Yohi, S.; Velauthapillai, D. TiO2 as a Photocatalyst for Water Splitting-An Experimental and Theoretical Review. Molecules 2021, 26, 1687. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, S.; Lin, S.; Zhang, D.; Li, G.; Leung, M.K.H. Controlling charge transfer in quantum-size titania for photocatalytic applications. Appl. Catal. B Environ. 2017, 215, 85–92. [Google Scholar] [CrossRef]
  3. Lian, Z.; Wang, W.; Li, G.; Tian, F.; Schanze, K.S.; Li, H. Pt-Enhanced Mesoporous Ti3+/TiO2 with Rapid Bulk to Surface Electron Transfer for Photocatalytic Hydrogen Evolution. ACS Appl. Mater. Interfaces 2017, 9, 16959–16966. [Google Scholar] [CrossRef] [PubMed]
  4. Wei, T.; Zhu, Y.; Wu, Y.; An, X.; Liu, L.-M. Effect of single-atom cocatalysts on the activity of faceted TiO2 photocatalysts. Langmuir 2018, 35, 391–397. [Google Scholar] [CrossRef]
  5. Yang, X.F.; Wang, A.; Qiao, B.; Li, J.; Liu, J.; Zhang, T. Single-atom catalysts: A new frontier in heterogeneous catalysis. Acc. Chem. Res. 2013, 46, 1740–1748. [Google Scholar] [CrossRef]
  6. Gihoon, C.; Anca, M.; Imgon, H.; Nikita, D.; Johannes, W.; Tadahiro, Y.; Zdeněk, B.; Giorgio, Z.; Alexander, B.T.; Erdmann, S.; et al. A facile “dark”-deposition approach for Pt single-atom trapping on facetted anatase TiO2 nanoflakes and use in photocatalytic H2 generation. Electrochim. Acta 2022, 412, 140129. [Google Scholar] [CrossRef]
  7. Yu, X.; Deng, J.; Liu, Y.; Jing, L.; Hou, Z.; Pei, W.; Dai, H. Single-Atom Catalysts: Preparation and Applications in Environmental Catalysis. Catalysts 2022, 12, 1239. [Google Scholar] [CrossRef]
  8. Li, C.-f.; Pan, W.-g.; Zhang, Z.-r.; Wu, T.; Guo, R.-t. Recent Progress of Single-Atom Photocatalysts Applied in Energy Conversion and Environmental Protection. Small 2023, 19, 2300460. [Google Scholar] [CrossRef]
  9. Wang, L.; Huang, L.; Liang, F.; Liu, S.; Wang, Y.; Zhang, H. Preparation, characterization and catalytic performance of single-atom catalysts. Chin. J. Catal. 2017, 38, 1528–1539. [Google Scholar]
  10. Hiragond, C.B.; Powar, N.S.; Lee, J.; In, S.-I. Single-Atom Catalysts (SACs) for Photocatalytic CO2 Reduction with H2O: Activity, Product Selectivity, Stability, and Surface Chemistry. Small 2022, 18, 2201428. [Google Scholar] [CrossRef]
  11. Li, Z.; Wang, D.; Wu, Y.; Li, Y. Recent advances in the precise control of isolated single-site catalysts by chemical methods. Natl. Sci. Rev. 2018, 5, 673–689. [Google Scholar]
  12. Saptal, V.B.; Ruta, V.; Bajada, M.A.; Vilé, G. Single-Atom Catalysis in Organic Synthesis. Angew. Chem. Int. Ed. 2023, 62, e202219306. [Google Scholar] [CrossRef]
  13. Majeed, I.; Ali, H.; Idrees, A.; Arif, A.; Ashraf, W.; Rasul, S.; Khan, M.A.; Nadeem, M.A.; Nadeem, M.A. Understanding the role of metal supported on TiO2 in photoreforming of oxygenates. Energy Adv. 2022, 1, 842–867. [Google Scholar] [CrossRef]
  14. Choi, Y.; Koo, M.S.; Bokare, A.D.; Kim, D.H.; Bahnemann, D.W.; Choi, W. Sequential Process Combination of Photocatalytic Oxidation and Dark Reduction for the Removal of Organic Pollutants and Cr(VI) using Ag/TiO2. Environ. Sci. Technol. 2017, 51, 3973–3981. [Google Scholar] [CrossRef] [PubMed]
  15. Li, Q.; Li, Y.W.; Wu, P.; Xie, R.; Shang, J.K. Palladium Oxide Nanoparticles on Nitrogen-Doped Titanium Oxide: Accelerated Photocatalytic Disinfection and Post-Illumination Catalytic “Memory”. Adv. Mater. 2008, 20, 3717–3723. [Google Scholar] [CrossRef]
  16. Takai, A.; Kamat, P.V. Capture, store, and discharge. Shuttling photogenerated electrons across TiO2-silver interface. ACS Nano 2011, 5, 7369–7376. [Google Scholar] [CrossRef]
  17. Cai, T.; Liu, Y.; Wang, L.; Zhang, S.; Ma, J.; Dong, W.; Zeng, Y.; Yuan, J.; Liu, C.; Luo, S. “Dark Deposition” of Ag Nanoparticles on TiO2: Improvement of Electron Storage Capacity To Boost “Memory Catalysis” Activity. ACS Appl. Mater. Interfaces 2018, 10, 25350–25359. [Google Scholar] [CrossRef]
  18. Herrmann, J.M.; Disdier, J.M.; Pichat, P. Photoassisted platinum deposition on TiO2 powder using various platinum complexes. J. Phys. Chem. 1986, 90, 6028–6034. [Google Scholar]
  19. Lai, W.H.; Miao, Z.; Wang, Y.X.; Wang, J.Z.; Chou, S.L. Atomic-local environments of single-atom catalysts: Synthesis, electronic structure, and activity. Adv. Energy Mater. 2019, 9, 1900722. [Google Scholar]
  20. Porter, J.F.; Li, Y.-G.; Chan, C.K. The effect of calcination on the microstructural characteristics and photoreactivity of Degussa P-25 TiO2. J. Mater. Sci. 1999, 34, 1523–1531. [Google Scholar]
  21. He, F.; Ma, F.; Li, J.; Li, T.; Li, G. Effect of calcination temperature on the structural properties and photocatalytic activities of solvothermal synthesized TiO2 hollow nanoparticles. Ceram. Int. 2014, 40, 6441–6446. [Google Scholar]
  22. Yu, J.-G.; Yu, H.-G.; Cheng, B.; Zhao, X.-J.; Yu, J.C.; Ho, W.-K. The effect of calcination temperature on the surface microstructure and photocatalytic activity of TiO2 thin films prepared by liquid phase deposition. J. Phys. Chem. B 2003, 107, 13871–13879. [Google Scholar]
  23. Wu, N.-L.; Lee, M.-S.; Pon, Z.-J.; Hsu, J.-Z. Effect of calcination atmosphere on TiO2 photocatalysis in hydrogen production from methanol/water solution. J. Photochem. Photobiol. A Chem. 2004, 163, 277–280. [Google Scholar]
  24. Lu, G.; Linsebigler, A.; Yates, J.T., Jr. Ti3+ defect sites on TiO2 (110): Production and chemical detection of active sites. J. Phys. Chem. 1994, 98, 11733–11738. [Google Scholar]
  25. Savva, A.I.; Smith, K.A.; Lawson, M.; Croft, S.R.; Weltner, A.E.; Jones, C.D.; Bull, H.; Simmonds, P.J.; Li, L.; Xiong, H. Defect generation in TiO2 nanotube anodes via heat treatment in various atmospheres for lithium-ion batteries. Phys. Chem. Chem. Phys. 2018, 20, 22537–22546. [Google Scholar] [CrossRef]
  26. Goodman, D. “Catalytically active Au on Titania:” Yet another example of a strong metal support interaction (SMSI)? Catal. Lett. 2005, 99, 1–4. [Google Scholar]
  27. Dulub, O.; Hebenstreit, W.; Diebold, U. Imaging cluster surfaces with atomic resolution: The strong metal-support interaction state of Pt supported on TiO2(110). Phys. Rev. Lett. 2000, 84, 3646. [Google Scholar]
  28. Edirisooriya, E.M.N.T.; Senanayake, P.S.; Wang, H.B.; Talipov, M.R.; Xu, P.; Wang, H. Photo-reforming and degradation of waste plastics under UV and visible light for H2 production using nanocomposite photocatalysts. J. Environ. Chem. Eng. 2023, 11, 109580. [Google Scholar] [CrossRef]
  29. Muthee, D.K.; Dejene, B.F. Effect of annealing temperature on structural, optical, and photocatalytic properties of titanium dioxide nanoparticles. Heliyon 2021, 7, e07269. [Google Scholar] [CrossRef]
  30. Monalisha, G.; Nirab, C.A.; Suparna, B. Effect of annealing temperatures on the structural and optical properties of zinc oxide nanoparticles prepared by chemical precipitation method. Optik 2018, 158, 1006–1015. [Google Scholar] [CrossRef]
  31. Zaki, M.Y.; Sava, F.; Simandan, I.D.; Buruiana, A.T.; Mihai, C.; Velea, A.; Galca, A.C. Effect of the stacking order, annealing temperature and atmosphere on crystal phase and optical properties of Cu2SnS3. Sci. Rep. 2022, 12, 7958. [Google Scholar] [CrossRef]
  32. Zarattini, M.; Dun, C.; Isherwood, L.H.; Felten, A.; Filippi, J.; Gordon, M.P.; Zhang, L.; Kassem, O.; Song, X.; Zhang, W.; et al. Synthesis of 2D anatase TiO2 with highly reactive facets by fluorine-free topochemical conversion of 1T-TiS2 nanosheets. J. Mater. Chem. A Mater. 2022, 10, 13884–13894. [Google Scholar] [CrossRef] [PubMed]
  33. Ball, P. Single-atom catalysis: A new field that learns from tradition. Natl. Sci. Rev. 2018, 5, 690–693. [Google Scholar] [CrossRef]
  34. He, L.; Guan, C.; Bulushev, D.A.; Xiang, Q. Regulation of Metal-Support Interaction in Single-Atom Catalysis. Small 2024, 940, 2410976. [Google Scholar] [CrossRef]
  35. Al-Hilfi, S.H.; Jiang, X.; Heuer, J.; Akula, S.; Tammeveski, K.; Hu, G.; Yang, J.; Wang, H.I.; Bonn, M.; Landfester, K.; et al. Single-Atom Catalysts through Pressure-Controlled Metal Diffusion. J. Am. Chem. Soc. 2024, 146, 19886–19895. [Google Scholar] [CrossRef]
  36. Hu, Y.; Dai, L.; Li, B.; Li, Z. Single-Atom Long-Range Interaction: Basic Principles and Applications. Adv. Sustain. Syst. 2024, 8, 2300564. [Google Scholar] [CrossRef]
  37. Bhosale, R.; Hyam, R.; Dhanya, P.; Ogale, S. Chlorate ion mediated rutile to anatase reverse phase transformation in the TiO2 nanosystem. Dalton Trans. 2011, 40, 11374–11377. [Google Scholar] [CrossRef] [PubMed]
  38. Zielińska-Jurek, A.; Wei, Z.; Janczarek, M.; Wysocka, I.; Kowalska, E. Size-Controlled Synthesis of Pt Particles on TiO2 Surface: Physicochemical Characteristic and Photocatalytic Activity. Catalysts 2019, 9, 940. [Google Scholar] [CrossRef]
  39. Simonsen, S.B.; Chorkendorff, I.; Dahl, S.; Skoglundh, M.; Sehested, J.; Helveg, S. Ostwald ripening in a Pt/SiO2 model catalyst studied by in situ TEM. J. Catal. 2011, 281, 147–155. [Google Scholar] [CrossRef]
  40. Chen, W.-T.; Chan, A.; Jovic, V.; Sun-Waterhouse, D.; Murai, K.-i.; Idriss, H.; Waterhouse, G.I.N. Effect of the TiO2 Crystallite Size, TiO2 Polymorph and Test Conditions on the Photo-Oxidation Rate of Aqueous Methylene Blue. Top. Catal. 2015, 58, 85–102. [Google Scholar] [CrossRef]
  41. Zhou, Y.; Muhich, C.L.; Neltner, B.T.; Weimer, A.W.; Musgrave, C.B. Growth of Pt Particles on the Anatase TiO2 (101) Surface. J. Phys. Chem. C 2012, 116, 12114–12123. [Google Scholar] [CrossRef]
  42. Prucnal, S.; Gago, R.; Calatayud, D.G.; Rebohle, L.; Liedke, M.O.; Butterling, M.; Wagner, A.; Helm, M.; Zhou, S. TiO2 Phase Engineering by Millisecond Range Annealing for Highly Efficient Photocatalysis. J. Phys. Chem. C 2023, 127, 12686–12694. [Google Scholar] [CrossRef]
  43. Hassanzadeh-Tabrizi, S.A. Precise calculation of crystallite size of nanomaterials: A review. J. Alloys Compd. 2023, 968, 171914. [Google Scholar] [CrossRef]
  44. Kim, M.G.; Kang, J.M.; Lee, J.E.; Kim, K.S.; Kim, K.H.; Cho, M.; Lee, S.G. Effects of Calcination Temperature on the Phase Composition, Photocatalytic Degradation, and Virucidal Activities of TiO2 Nanoparticles. ACS Omega 2021, 6, 10668–10678. [Google Scholar] [CrossRef] [PubMed]
  45. Kandiel, T.A.; Robben, L.; Alkaim, A.; Bahnemann, D. Brookite versus anatase TiO2 photocatalysts: Phase transformations and photocatalytic activities. Photochem. Photobiol. Sci. 2013, 12, 602–609. [Google Scholar] [CrossRef]
  46. Shi, Y.; Ma, Z.-R.; Xiao, Y.-Y.; Yin, Y.-C.; Huang, W.-M.; Huang, Z.-C.; Zheng, Y.-Z.; Mu, F.-Y.; Huang, R.; Shi, G.-Y.; et al. Electronic metal–support interaction modulates single-atom platinum catalysis for hydrogen evolution reaction. Nat. Commun. 2021, 12, 3021. [Google Scholar] [CrossRef]
  47. Wu, Z.; Hwang, I.; Cha, G.; Qin, S.; Tomanec, O.; Badura, Z.; Kment, S.; Zboril, R.; Schmuki, P. Optimized Pt single atom harvesting on TiO2 nanotubes—Towards a most efficient photocatalyst. Small 2022, 18, 2104892. [Google Scholar] [CrossRef]
  48. Muthukrishnan, S.; Vidya, R.; Anja Olafsen, S. Band gap engineering of anatase TiO2 by ambipolar doping: A first principles study. Mater. Chem. Phys. 2023, 299, 127467. [Google Scholar] [CrossRef]
  49. Haq, S.; Rehman, W.; Waseem, M.; Meynen, V.; Awan, S.U.; Khan, A.R.; Hussain, S.; Zain-ul-Abdin; Din, S.U.; Hafeez, M.; et al. Effect of Annealing Temperature on Structural Phase Transformations and Band Gap Reduction for Photocatalytic Activity of Mesopores TiO2 Nanocatalysts. J. Inorg. Organomet. Polym. Mater. 2021, 31, 1312–1322. [Google Scholar] [CrossRef]
  50. Wu, B.; Du, F.; Wang, H.; Wu, C.; Chu, J.; Wang, X.; Xiong, S. Effects of annealing temperature of PtCu/MWCNT catalysts on their electrocatalytic performance of electrooxidation of methanol. Ionics 2022, 28, 369–382. [Google Scholar] [CrossRef]
  51. Chen, Y.; Ji, S.; Sun, W.; Lei, Y.; Wang, Q.; Li, A.; Chen, W.; Zhou, G.; Zhang, Z.; Wang, Y.; et al. Engineering the Atomic Interface with Single Platinum Atoms for Enhanced Photocatalytic Hydrogen Production. Angew. Chem. Int. Ed. 2020, 59, 1295–1301. [Google Scholar] [CrossRef]
  52. Li, X.; Su, Z.; Jiang, H.; Liu, J.; Zheng, L.; Zheng, H.; Wu, S.; Shi, X. Band Structure Tuning via Pt Single Atom Induced Rapid Hydroxyl Radical Generation toward Efficient Photocatalytic Reforming of Lignocellulose into H2. Small 2024, 20, 2400617. [Google Scholar] [CrossRef]
  53. Edirisooriya, E.M.N.T.; Senanayake, P.S.; Xu, P.; Talipov, M.R.; Wang, H. Optimization of green hydrogen evolution from low-density plastics using TiO2-based nano-photocatalysts with techno-economic and carbon footprint assessment. Nanotechnol. Environ. Eng. 2024, 9, 817–832. [Google Scholar] [CrossRef]
  54. Han, X.; Jiang, M.; Li, H.; Li, R.; Sulaiman, N.H.M.; Zhang, T.; Li, H.; Zheng, L.; Wei, J.; He, L.; et al. Upcycle polyethylene terephthalate waste by photoreforming: Bifunction of Pt cocatalyst. J. Colloid Interface Sci. 2024, 665, 204–218. [Google Scholar] [CrossRef] [PubMed]
  55. Carozo, A.M.; López-Tenllado, F.J.; Herrera-Beurnio, M.C.; Hidalgo-Carrillo, J.; Martín-Gómez, J.; Estevez, R.; Ariza-Pérez, A.; Urbano, F.J.; Marinas, A. Pt-TiO2 Systems for Enhanced Hydrogen Production from Glycerol: Direct vs Sequential Incorporation Through Photodeposition. Materials 2024, 17, 5109. [Google Scholar] [CrossRef]
  56. Pichler, C.M.; Bhattacharjee, S.; Rahaman, M.; Uekert, T.; Reisner, E. Conversion of Polyethylene Waste into Gaseous Hydrocarbons via Integrated Tandem Chemical–Photo/Electrocatalytic Processes. ACS Catal. 2021, 11, 9159–9167. [Google Scholar] [CrossRef]
  57. Kawai, T.; Sakata, T. Photocatalytic hydrogen production from water by the decomposition of poly-vinylchloride, protein, algae, dead insects, and excrement. Chem. Lett. 2006, 10, 81–84. [Google Scholar] [CrossRef]
  58. Li, Z.; Yang, Y.; Zhang, C.; Fan, W.; Li, G.; Fang, J.; Lu, L. Atomically engineering the metal-support interaction of single-atom Cu/TiO2 for efficient polyethylene terephthalate plastic photoreforming. Chem Catal. 2024, 4, 100902. [Google Scholar] [CrossRef]
  59. Zi, B.; Zheng, H.; Zhou, T.; Lu, Q.; Chen, M.; Xiao, B.; Zhang, Y.; Qiu, Z.; Sun, H.; Zhao, J.; et al. Changeable Active Sites by Pr Doping CuSA-TiO2 Photocatalyst for Excellent Hydrogen Production. Small 2024, 20, 2305779. [Google Scholar] [CrossRef]
  60. Huang, T.-C.; Chen, X.; Sisay Wolde, G.; Kuo, D.-H. Photocatalytic hydrogen production over highly oxygen deficient Cu-doped TiO2 and its composites: Insights of kinetic reaction micromechanisms. Sep. Purif. Technol. 2024, 339, 126671. [Google Scholar] [CrossRef]
  61. Liang, X.; Li, X.; Dong, Q.; Gao, T.; Cao, M.; Zhao, K.; Lichtfouse, E.; Patrocinio, A.O.T.; Wang, C. Photo- and electrochemical processes to convert plastic waste into fuels and high-value chemicals. Chem. Eng. J. 2024, 482, 148827. [Google Scholar] [CrossRef]
  62. Uekert, T. Plastic and Mixed Waste as Feedstocks for Solar-Driven H2 Production. Ph.D. Thesis, University of Cambridge, Cambridge, UK, 2021. [Google Scholar]
  63. Schindelin, J.; Arganda-Carreras, I.; Frise, E.; Kaynig, V.; Longair, M.; Pietzsch, T.; Preibisch, S.; Rueden, C.; Saalfeld, S.; Schmid, B.; et al. Fiji: An open-source platform for biological-image analysis. Nat. Methods 2012, 9, 676–682. [Google Scholar] [CrossRef]
  64. Doumeng, M.; Makhlouf, L.; Berthet, F.; Marsan, O.; Delbé, K.; Denape, J.; Chabert, F. A comparative study of the crystallinity of polyetheretherketone by using density, DSC, XRD, and Raman spectroscopy techniques. Polym. Test. 2021, 93, 106878. [Google Scholar] [CrossRef]
  65. Monshi, A.; Foroughi, M.R.; Monshi, M.R. Modified Scherrer equation to estimate more accurately nano-crystallite size using XRD. World J. Nano Sci. Eng. 2012, 2, 154–160. [Google Scholar]
  66. Origin(Pro), version 2024; OriginLab Corporation: Northampton, MA, USA, 2024.
  67. Edirisooriya, E.M.N.T.; Punhasa, S.S.; Pei, X.; Huiyao, W. Hydrogen production and value-added chemical recovery from the photo-reforming process using waste plastics. J. Environ. Chem. Eng. 2023, 11, 111429. [Google Scholar] [CrossRef]
  68. Wang, H.; Edirisooriya, E.M.N.T.; Xu, P. Method of Hydrogen Manufacture. U.S. Patent Application 18/378586, 6 February 2025. [Google Scholar]
  69. Qin, J.; Dou, Y.; Wu, F.; Yao, Y.; Andersen, H.R.; Hélix-Nielsen, C.; Lim, S.Y.; Zhang, W. In-situ formation of Ag2O in metal-organic framework for light-driven upcycling of microplastics coupled with hydrogen production. Appl. Catal. B Environ. 2022, 319, 121940. [Google Scholar] [CrossRef]
Figure 1. (A) Comparison of the particle size distribution of dark-deposited Pt/TiO2 particles after annealing at different temperatures with pristine TiO2 and (B) STEM image of 0.7%(w/w) Pt/TiO2 prepared using the direct deposition method, where the Pt single atoms are highlighted by circles. (C) Dark-deposited Pt/TiO2 synthesized with 0.4%(w/w) precursor loading and annealed at 500 °C, dark-deposited Pt/TiO2 with 0.04%(w/w) precursor loading and annealed at (D) 500 °C, (E) 400 °C, and (F) 300 °C.
Figure 1. (A) Comparison of the particle size distribution of dark-deposited Pt/TiO2 particles after annealing at different temperatures with pristine TiO2 and (B) STEM image of 0.7%(w/w) Pt/TiO2 prepared using the direct deposition method, where the Pt single atoms are highlighted by circles. (C) Dark-deposited Pt/TiO2 synthesized with 0.4%(w/w) precursor loading and annealed at 500 °C, dark-deposited Pt/TiO2 with 0.04%(w/w) precursor loading and annealed at (D) 500 °C, (E) 400 °C, and (F) 300 °C.
Catalysts 15 00334 g001
Figure 2. (A) Crystallite size distribution of 0.04%(w/w) Pt/TiO2 at different annealing temperatures and (B) variation in the crystallite size of anatase and rutile crystallites under different synthesis conditions.
Figure 2. (A) Crystallite size distribution of 0.04%(w/w) Pt/TiO2 at different annealing temperatures and (B) variation in the crystallite size of anatase and rutile crystallites under different synthesis conditions.
Catalysts 15 00334 g002
Figure 3. Comparison of XPS of (A) Pt 4f and (B) Ti2P of catalysts prepared under the direct deposition method and the dark deposition method.
Figure 3. Comparison of XPS of (A) Pt 4f and (B) Ti2P of catalysts prepared under the direct deposition method and the dark deposition method.
Catalysts 15 00334 g003
Figure 4. Variation in the hydrogen yield of catalysts synthesized at (A) 500 °C, (C) 400 °C, and (E) 300 °C and the respective hydrogen production rates with catalysts synthesized at (B) 500 °C, (D) 400 °C, and (F) 300 °C. The error bars represent the variability of the data.
Figure 4. Variation in the hydrogen yield of catalysts synthesized at (A) 500 °C, (C) 400 °C, and (E) 300 °C and the respective hydrogen production rates with catalysts synthesized at (B) 500 °C, (D) 400 °C, and (F) 300 °C. The error bars represent the variability of the data.
Catalysts 15 00334 g004
Figure 5. Comparison of hydrogen yield after 48 h with catalysts synthesized under dark and direct deposition methods under different synthesis conditions. The error bars represent the variability of the data.
Figure 5. Comparison of hydrogen yield after 48 h with catalysts synthesized under dark and direct deposition methods under different synthesis conditions. The error bars represent the variability of the data.
Catalysts 15 00334 g005
Figure 6. Loading plot of principal component analysis.
Figure 6. Loading plot of principal component analysis.
Catalysts 15 00334 g006
Figure 7. Principal Component Analysis of Pt/TiO2 catalyst synthesis under direct and dark conditions.
Figure 7. Principal Component Analysis of Pt/TiO2 catalyst synthesis under direct and dark conditions.
Catalysts 15 00334 g007
Figure 8. Schematic comparison of the synthesis methods.
Figure 8. Schematic comparison of the synthesis methods.
Catalysts 15 00334 g008
Table 1. Summary of data calculated from XRD.
Table 1. Summary of data calculated from XRD.
Synthesis ConditionDegree of Crystallinity (%)Average Crystallite Size (nm)Anatase Percentage (%)
0.1 mM annealed at 300 °C83.9620.2092
0.01 mM annealed at 300 °C94.0317.3593
0.001 mM annealed at 300 °C92.0118.1192
0.1 mM annealed at 400 °C89.4720.6691
0.01 mM annealed at 400 °C95.2921.5692
0.001 mM annealed at 400 °C88.4522.1292
0.1 mM annealed at 500 °C88.8920.5694
0.01 mM annealed at 500 °C93.1922.4094
0.001 mM annealed at 500 °C93.8918.5194
0.075 mM without annealing83.1620.5286
0.175 mM without annealing82.7421.8089
0.275 mM without annealing63.9920.5483
Pristine P25 TiO286.6518.3285
Table 2. Comparison of expected Pt loading and measured doping percentage by ICP-OES analysis.
Table 2. Comparison of expected Pt loading and measured doping percentage by ICP-OES analysis.
Synthesis MethodAnnealing TemperaturePrecursor ConcentrationExpected Maximum Pt Loading %(w/w)Measured Pt Loading by ICP-OES %(w/w)
Dark deposition method300 °C0.001 mM0.0040.011
0.010 mM0.0400.038
0.100 mM0.4000.346
400 °C0.001 mM0.0040.013
0.010 mM0.0400.039
0.100 mM0.4000.380
500 °C0.001 mM0.0040.004
0.010 mM0.0400.035
0.100 mM0.4000.362
Direct deposition method 500 °C0.001 mM0.0040.009
0.010 mM0.0400.018
0.100 mM0.4000.012
0.075 mM0.3000.024
0.175 mM0.7000.022
Without annealing0.075 mM0.3000.004
0.175 mM0.7000.029
0.275 mM1.5000.040
Table 3. Summary of the calculated bandgap energies of different catalysts.
Table 3. Summary of the calculated bandgap energies of different catalysts.
Synthesis MethodSynthesis ConditionsBandgap (eV)Maximum Wavelength (nm)
DirectIndirectAverage
Dark deposition0.1 mM 300 °C2.282.502.39520
0.1 mM 400 °C2.192.182.19567
0.1 mM 500 °C2.282.502.39520
0.01 mM 300 °C2.482.192.34531
0.01 mM 400 °C2.482.192.34531
0.01 mM 500 °C2.482.192.34531
0.001 mM 300 °C2.973.002.99415
0.001 mM 400 °C2.642.862.75451
0.001 mM 500 °C2.512.622.57483
Direct deposition0.075 mM 25 °C2.662.982.82440
0.175 mM 25 °C2.782.952.86434
0.275 mM 25 °C2.882.482.68463
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.

Share and Cite

MDPI and ACS Style

Edirisooriya, E.M.N.T.; Senanayake, P.S.; Xu, P.; Wang, H. Enhanced H2 Production Efficiency in Photo-Reforming of PET Waste Plastic Using Dark-Deposited Atom/Nanocomposite Pt/TiO2 Photocatalysts. Catalysts 2025, 15, 334. https://doi.org/10.3390/catal15040334

AMA Style

Edirisooriya EMNT, Senanayake PS, Xu P, Wang H. Enhanced H2 Production Efficiency in Photo-Reforming of PET Waste Plastic Using Dark-Deposited Atom/Nanocomposite Pt/TiO2 Photocatalysts. Catalysts. 2025; 15(4):334. https://doi.org/10.3390/catal15040334

Chicago/Turabian Style

Edirisooriya, E. M. N. Thiloka, Punhasa S. Senanayake, Pei Xu, and Huiyao Wang. 2025. "Enhanced H2 Production Efficiency in Photo-Reforming of PET Waste Plastic Using Dark-Deposited Atom/Nanocomposite Pt/TiO2 Photocatalysts" Catalysts 15, no. 4: 334. https://doi.org/10.3390/catal15040334

APA Style

Edirisooriya, E. M. N. T., Senanayake, P. S., Xu, P., & Wang, H. (2025). Enhanced H2 Production Efficiency in Photo-Reforming of PET Waste Plastic Using Dark-Deposited Atom/Nanocomposite Pt/TiO2 Photocatalysts. Catalysts, 15(4), 334. https://doi.org/10.3390/catal15040334

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop