Houkui
Xiang
ab,
Wanqi
Liu
c,
Zhaokai
Zeng
d,
Jingxian
Chen
d,
Beile
Lin
a,
Guoliang
Zhang
d,
Xuhui
Wei
e,
Songbo
Zhou
ab,
Tongqiang
Xiong
*ab,
Pengfei
Li
*c and
Libing
Qian
*d
aSchool of Automation, Hubei University of Science and Technology, Xianning 437100, China. E-mail: Xiongtongqiang@hbust.edu.cn
bKey Laboratory of Optoelectronic Sensing and Intelligent Control, Hubei University of Science and Technology, Xianning 437100, China
cSchool of Nuclear Science and Technology, Harbin Engineering University, Harbin 150001, China. E-mail: lipf2022@hrbeu.edu.cn
dHubei Key Laboratory of Radiation Chemistry and Functional Materials, School of Nuclear Technology and Chemistry & Biology, Hubei University of Science and Technology, Xianning 437100, China. E-mail: qianlb@hbust.edu.cn
eCollege of Chemistry and Chemical Engineering, Anqing Normal University, Anqing 246133, China
First published on 6th October 2025
Photothermal catalytic methanol (CH3OH) dehydrogenation is a sustainable pathway for the co-production of hydrogen (H2) and formaldehyde (HCHO) under mild reaction conditions. However, this process is hindered by the inefficient interfacial electron transfer at the semiconductor–cocatalyst (SC) interface, which is exacerbated by charge leakage. Using TiO2 as a model, we systematically investigated the individual effects of irradiation intensity, wavelegth, and temperature on interfacial electron-transfer dynamics. Our findings reveal that increasing the photon flux at a constant temperature effectively reduces the SC barrier and accelerates the electron-transfer kinetics, with little impact on the overall charge-transfer duration. Although short-wavelength light exhibits limited penetration in aqueous media, reducing the wavelength enhances SC interfacial electron-transfer. Conversely, raising the temperature under a fixed photon flux not only lowers the SC barrier but also markedly shortens the total charge transfer time. To address the temperature-induced charge leakage, Sn was doped into the TiO2 lattice. The optimised Pt/3% Sn–TiO2 catalyst, employed in a fixed-bed reactor designed to overcome poor UV penetration, achieves an H2 production rate of 82.86 mmol g−1 h−1 under 30 W UV illumination (254 nm) at 80 °C, representing a 4.08-fold enhancement over Pt/TiO2. In a 20 h stability test, the system maintained average production rates of 88.12 mmol g−1 h−1 for H2 and 84.13 mmol g−1 h−1 for HCHO, achieving an apparent quantum efficiency (AQY) of 40.02%. Under 320 W UV irradiation, the catalyst sustained 413.5 mmol g−1 h−1 H2 generation over 6 h with an AQY of 25.16%. This work presents a promising strategy for designing high-performance catalysts by enhancing SC interfacial electron transfer, thereby significantly improving the photothermal catalytic efficiency.
Green foundation1. Photocatalytic hydrogen evolution coupled with organic synthesis offers a promising green chemical pathway. However, its application is limited by low efficiency. Herein, we conducted a kinetic analysis of the effects of irradiation intensity and temperature on interfacial electron transfer during the reaction, aiming to establish a novel theoretical basis for enhancing the efficiency of integrated photocatalytic H2 production–organic synthesis systems.2. Our research focuses on improving the efficiency of photocatalytic systems to facilitate their practical implementation. Using this technology, we successfully mineralised low-concentration benzene gas (∼10 mg m−3 in air) at a flow rate of 100 L min−1, with emissions meeting stringent regulatory standards. 3. This study provides new insights into interfacial charge transfer under photothermal synergistic effects, providing theoretical guidance for designing advanced photothermal catalysts. |
As an example, formaldehyde (HCHO) remains indispensable in polymer manufacturing, textile production, and agricultural applications, despite its well-documented occupational health risks.12,13 Currently, industrial HCHO production predominantly relies on two thermocatalytic processes: (1) methanol (CH3OH) partial oxidation over Cu (Ag or Fe2O3–MoO3) catalysts at 200–300 °C (CH3OH (g) + 1/2 O2 (g) → HCHO (g) + H2O (g), ΔH° 298 = −157.0 kJ mol−1), and (2) methane (CH4) oxidation using Fe2O3 (or Cr2O3) at 400–800 °C (CH4 (g) + O2 (g) → HCHO (g) + H2O (g), ΔH° 298 = −275.8 kJ mol−1).14–16 Although these methods are widely adopted, they suffer from intrinsic limitations, such as the inability to produce H2 simultaneously. Photocatalytic methanol dehydrogenation (CH3OH (g) → HCHO (g) + H2 (g), ΔH° 298 = +85.1 kJ mol−1) thus presents a thermodynamically challenging yet promising alternative for concurrent H2 and HCHO production under mild conditions.17 However, similar to other photocatalytic H2 generation systems coupled with organic oxidation reactions, this approach is constrained by relatively low efficiency, which presents significant barriers to its large-scale industrial application.18,19
For water splitting, the slow oxygen-evolution reaction is often identified as the primary rate-determining step.6,20 However, even when easily oxidisable organic molecules such as lactic acid (LA) or CH3OH are used as hole scavengers and efficient reduction cocatalysts such as Pt, Au, Pd, Ag, or MoS2 are employed, the H2-production efficiency remains unsatisfactory (Fig. 1a).21–24 In such cases, the reduction half-reaction becomes the key bottleneck limiting the photocatalytic efficiency (Fig. S1). Previous studies have indicated that poor interfacial electron transfer between the semiconductor and cocatalyst (SC) is one of the main reasons.25–27 Typically, electron transfer from the semiconductor to the cocatalyst proceeds via two primary mechanisms: thermionic emission and charge recombination (Fig. 1b).25,28 In thermionic emission, the high work function of the cocatalyst creates a significant Schottky barrier upon contact with the semiconductor, severely impeding photogenerated electron transfer.29,30 Meanwhile, in the charge recombination process, excessive surface states not only exacerbate charge leakage but also serve as recombination centres for photogenerated electrons and holes, leading to ineffective charge separation.
To address the issue of insufficient SC interfacial electron transfer caused by high barriers, we doped tin (Sn) doping into the titanium dioxide (TiO2) in a controlled manner. This modification enhances semiconductor electron concentration, reduces the SC barriers, and minimises charge leakage. These improvements establish a strong positive correlation between Pt/Sn–TiO2 and temperature, which is crucial for enhancing photothermal catalytic efficiency. Under mild conditions, Pt/3% Sn–TiO2 exhibits efficient photothermal catalytic CH3OH dehydrogenation for HCHO production. Specifically, within 20 h at 80 °C under a 30 W low-pressure mercury lamp, the average H2 and HCHO yields were 88.12 and 84.13 mmol g−1 h−1, respectively, with an average apparent quantum efficiency (AQY) of 40.02%. Using a 320 W low-pressure mercury lamp, the average H2 production rate reached 413.5 mmol g−1 h−1 within 6 h at 80 °C, with an average AQY of 25.16%. This work provides new insights into the catalyst design and coupling strategy for photocatalytic H2 production via organic molecule oxidation.
Pt/TiO2 and Pt/Sn–TiO2 powders were fabricated using the photo-deposition method (Fig. 2). For this process, 0.3 g of (Sn–)TiO2 was placed in a quartz reactor and suspended in an aqueous solution (20 vol% CH3OH + 30 μL 0.1 g mL−1 H2PtCl6) via sonication. The suspension was degassed with high-purity argon (Ar, 99.999%) for 15 min to eliminate dissolved oxygen and then exposed to continuous irradiation from a 300 W UV high-pressure lamp (model: CEM-M500, Beijing Mircoenerg, Beijing, China) for 2 h. Throughout irradiation, an Ar gas atmosphere was maintained under magnetic stirring conditions. Afterward, the resulting powder was isolated by centrifugation, washed multiple times with pure water, and dried at 60 °C for 12 h.
The amount of HCHO produced was determined using the acetylacetone spectrophotometric method.17,22,31,32 Specifically, the procedure involved adding 15 mL of chromogenic reagent solution (comprising 2 mL of CH3COCH2COCH3, 100 g of CH3COONH4, and 15 of mL CH3COOH, with the pH adjusted to 6.0 using NaOH) to a 50 mL volumetric flask. Subsequently, 20 μL of the reaction solution obtained after CH3OH catalytic oxidation was introduced into the flask. The solution was diluted with distilled water to a final volume of 50 mL. Afterward, the flask was placed in a drying oven maintained at 60 °C for 1 h. Finally, the HCHO concentration was quantified by measuring the UV–vis absorbance at 414 nm (Rayleigh UV1801, Beijing Beifen-Ruili Analytical Instrument (Group), Beijing, China) based on pre-established calibration curves. Additionally, gas chromatography–mass spectrometry (GC–MS, Agilent 5977C GC/MSD, Santa Clara, California, USA) and high-performance liquid chromatography (HPLC, PolyPak H column, Fuli L75, Fuli Instruments, China) were employed to analyse the products formed from CH3OH after a 12 h reaction period in a fixed-bed reactor.
The band gap structures of TiO2 and Sn–TiO2 were characterised using UPS and UV–vis spectroscopy. The work function (Φ) was determined from the secondary electron cutoff energy (ESECE) measured using UPS under He-Iα excitation (hν = 21.22 eV):33,34
| Φ = hν − ESECE | (1) |
For pristine TiO2, the UPS spectrum (applied bias: −5 V) yielded ESECE = 17.79 eV (Fig. S8), resulting in a work function of ΦTiO2 = 21.22–17.79 = 3.43 eV, corresponding to the position of the Fermi energy level (EF). Additionally, the valence band (VB) level was extracted from the UPS data and identified at 2.53 eV below the Fermi level. The optical bandgap (Eg) was derived from the UV–vis data using the Tauc equation for indirect transitions:35,36
| (αhν)2 = A(hν − Eg) | (2) |
The photocatalytic CH3OH reaction was analysed using in situ diffuse reflectance infrared Fourier-transform spectroscopy (DRIFTS) with a Bruker INVENIO S spectrometer (Karlsruhe, Germany). The experimental setup included a Harrick diffuse reflectance in situ cell and an MCT detector. Before the analysis, the empty reaction chamber was heated to 200 °C and maintained at that temperature for 0.5 h under an air atmosphere to eliminate residual impurities, followed by cooling to room temperature. The Pt/3% Sn–TiO2 photocatalyst was introduced into the chamber, which was subsequently sealed tightly. The system was purged with Ar gas at a flow rate of 30 mL min−1 for 15 min, followed by heating to 150 °C and holding at this temperature for 0.5 h to remove any impurity gases adsorbed on the catalyst surface. After cooling the system to 50 °C, Ar was used as the carrier gas to introduce the vapour of a 20 vol% aqueous CH3OH solution into the reaction chamber via a bubbler at a flow rate of 30 mL min−1 for 15 min, ensuring complete adsorption saturation on the catalyst surface. The chamber was sealed again, and the dark adsorption state was recorded as the background spectrum. A 3 W LED UV lamp (365 nm) was used to irradiate the sample through the observation window, and time-resolved IR absorbance spectra were continuously collected at predefined intervals over a 1 h period to monitor the reaction progress. A UV lamp was continuously illuminated throughout the experiment.
During the measurement, 15 vol% LA (or 0.5 M Na2SO4) aqueous solution was used for open-circuit potential decay (OCP) and voltammetry measurements, with illumination provided by a 300 W high-pressure lamp (model: CEM-M500, Beijing Mircoenergy). Additionally, a 0.5 M Na2SO4 aqueous solution without any hole extractors was used as the electrolyte for the Mott–Schottky analysis. The donor density (Nd) concentration was calculated as:37,38
![]() | (3) |
The hydrogen evolution (HE) behaviour of Pt/TiO2 and Pt/Sn–TiO2 under varying light intensities were investigated in an aqueous LA solution system (Fig. 3). The experimental procedure was as follows. First, 10 mg of the catalyst was dispersed into a solution consisting of 10 mL of LA and 60 mL of deionised water. The resulting mixture was then purged with high-purity Ar gas and stirred for 15 min to ensure complete deoxygenation. Subsequently, the light source was activated and, continuous stirring and Ar gas purging were maintained throughout the H2 production test. The gas flow rate was controlled by adjusting the valve, and the flow rate of the outlet gas mixture (Ar and H2) was measured using a bubble flowmeter. Simultaneously, the H2 concentration in the mixed gas was periodically analysed using gas chromatography. The reaction temperature was precisely maintained using an external circulating water bath.
The AQY of Pt/TiO2 and Pt/Sn–TiO2 under varying solar light intensities can be determined by integrating the wavelength range from the UV region to 382 nm (band gap: 3.25 eV), as defined by the cut-off edge obtained from UV–vis spectroscopy and normalized to the standard solar light spectrum (ASTM G173-03 (2020)).39,40 Under one standard solar irradiation intensity (100 mW cm−2, AM 1.5G), the photon flux density (N1) that (Sn–)TiO2 can absorb is 1.204 × 1020 photons per m2 per s. Using this value, the AQY can be subsequently calculated.41–43
![]() | (4) |
Here, nH2 represents the number of evolved H2 molecules (mol s−1), NA denotes the Avogadro constant (6.02 × 1023 mol−1), N1 is the number of incident photons (m−2 s−1), and A is the effective area of the reactor (m2).
Taking Pt/TiO2 as an example, under standard solar light intensity (100 mW cm−2, AM 1.5G) at 20 °C, its H2 production rate is 3.258 mmol h−1 g−1 (based on 10 mg of catalyst). Given that the reactor area was 0.001963 m2 (diameter: 5 cm), the AQY was calculated as follows:
The calculation process for AQY under other irradiation intensity conditions follows was similar, involving multiplication of the photon concentration by a corresponding factor. For example, at 300 mW cm−2 (AM 1.5G), the number of effective incident photons was 3N1.
Photothermal oxidation of CH3OH was conducted in a custom-designed tubular fixed-bed reactor.44,45 A slurry consisting of 0.2 g of (Sn–)TiO2 dispersed in 5 mL of anhydrous ethanol was sonicated for 15 min and uniformly brush-coated onto a low-pressure mercury lamp (Philips TUV 30 W/G30, 254 nm, Poland). The catalyst-coated lamp was air-dried for 2 h to form a homogeneous thin film before being assembled into the reactor.
For Pt deposition, a 1 wt% Pt cocatalyst was photo-deposited by irradiating 20 μL of a cycle of 1 wt% H2PtCl6 aqueous solution under 254 nm UV light (30 W, effective power: 12 W × 0.9) for 2 h. Before irradiation, the reactor was purged with Ar gas for 30 min to remove dissolved oxygen, and an inert atmosphere was maintained throughout the process by a continuous flow of Ar gas. The residual solvent was subsequently removed by purging with dry compressed air at room temperature for 1 h. The reactor was then reconfigured with 1 L of 20 vol% CH3OH aqueous solution. The pre-treatment step involved circulating deoxygenated CH3OH at 47 mL min−1 for 30 min using a peristaltic pump. Under continuous Ar bubbling and irradiation, the reaction temperature was stabilised using a heating tape. H2 evolution was quantified hourly via gas chromatography, and the HCHO concentration was measured hourly using the acetylacetone spectrophotometric method.22,46 The AQY was determined using the following equations:45,47,48
![]() | (5) |
Here, nH2 represents the number of evolved H2 molecules (mol s−1), N is the number of incident photons, NA denotes Avogadro's constant (6.02 × 1023 mol−1), h is Planck's constant (6.62 × 10−34 J s), c is the speed of light (3 × 108 m s−1), P signifies the effective power of the UV lamp used in the experiment (12 W), η indicates the effective utilisation rate (0.9), and λ represents the wavelength (254 nm). By substituting the aforementioned constants into the formula, we obtain
![]() | (6) |
As an example, the AQY of Pt/TiO2 at 20 °C (the nH2 was 9.889 mmol g−1 h−1, 0.2 g catalyst) was calculated to be 4.81%.
![]() | ||
| Fig. 4 XRD patterns of TiO2 doped with varying amounts of Sn obtained at a scanning rate of 5° min−1. The high-resolution images of the (101) crystal plane (inset) were obtained at a reduced scanning rate of 1° min−1 (a). In panels b, c, and d, the high-resolution XPS spectra of Ti 2p, O 1s (OL: lattice oxygen; OH: hydroxyl oxygen; OA: adsorbed oxygen),56,57 and Sn 3d are presented for both pure TiO2 and 3% Sn-doped TiO2. | ||
Furthermore, the XPS patterns showed a blue shift in the Ti 2p peak after Sn doping, accompanied by enhanced hydroxyl oxygen signals and the emergence of a Sn 3d signal. These spectral changes indicate that Sn doping significantly alters the coordination environments of Ti and O.50,53 Specifically, the replacement of balanced Ti–O–Ti bonds with Sn–O–Ti bonds, driven by the higher electronegativity of Sn (1.96) than that of Ti (1.54),53 induces significant alterations in the local electron density. Such modifications may affect the photoelectric properties of the material. Consistent with previous reports, Sn4+ doping enhances carrier mobility in TiO2, suppresses recombination rates, and improves the OCP.54,55
![]() | (7) |
Here, τOCP-dual and τOCP can be obtained by monitoring the OCP decay:58,59
![]() | (8) |
An analysis of the electron-transfer time constants under varying irradiation intensities (Fig. S4) shows that for pristine TiO2 at 180 mW cm−2, the average time constant increased by factors of 5.25 and 6.65 compared with those at 88 mW cm−2 and 35 mW cm−2, respectively (Fig. S4a). This indicates that a higher irradiation intensity leads to slower electron-transfer kinetics. Although stronger illumination promoted the entry of more electrons into the CB (producing a more negative photovoltage at 180 mW cm−2 in Fig. 5a), the kinetic limitations at the TiO2/electrolyte interface hindered efficient electron transfer to the solution. After Pt deposition, the photovoltage remained intensity dependent (Fig. 5b), but the catalytic properties of Pt significantly enhanced the electron transfer efficiency. At 180 mW cm−2, Pt/TiO2 shows only ∼1.22 and ∼1.33 higher time constants than those at 88 mW cm−2 and 35 mW cm−2 (Fig. S4b), respectively, demonstrating a weakened dependence on the light intensity. Notably, Fig. 5c shows that τSC decreases with increasing light intensity. Therefore, while pristine TiO2 exhibits a strong negative correlation between light intensity and electron transfer kinetics, Pt deposition mitigates this relationship by accelerating SC interfacial transfer, enabling higher reaction rates under high-intensity irradiation conditions.
Furthermore, band-pass filters were employed to isolate specific wavelengths—254 nm (∼2.0 mW cm−2), 313 nm (∼2.4 mW cm−2), and 365 nm (∼283.5 mW cm−2)—as individual light sources for evaluating the OCP behaviour of TiO2 and Pt/TiO2 electrodes (Fig. S5a and S5b). The results show that although the light intensity at 313 nm is considerably lower than that at 365 nm, the difference in photovoltage between these wavelengths is relatively small (0.023 V (vs. Ag/AgCl) for TiO2 and 0.001 V (vs. Ag/AgCl) for Pt/TiO2). This can be attributed to the relatively small difference in wavelength, which resulted in similar penetration depths. Despite its lower intensity, 313 nm light generates higher-energy excited electrons that overcome the SC barrier more effectively, thereby facilitating charge transfer.
Under equivalent light intensity conditions (vs. 313 nm), a shorter-wavelength (254 nm) can theoretically excite higher-energy electrons, which should overcome the SC barrier. However, owing to its limited penetration capacity (Fig. S6), light traveling through the solution to the electrode surface is significantly attenuated. Consequently, the photovoltage generated under 254 nm illumination was substantially lower than that under 365 nm and 313 nm. Further analysis of the interfacial charge-transfer time constant (Fig. S5c and S5d) revealed that, for both TiO2 and Pt/TiO2, despite a 118.1-fold difference in light intensity between 365 and 313 nm, the change in the time constant remained marginal (∼1.57-fold). By contrast, under 254 nm illumination, the charge-transfer time constants for Pt/TiO2 decreased by a factor of 2.41 compared with that at 313 nm. As illustrated in Fig. 5d, shorter wavelengths correspond to smaller SC interfacial charge-transfer time constants, indicating a higher probability of successful charge transfer. This is because, although fewer short-wavelength photons reach the electrode surface, the excited electrons possess higher energy, allowing them more easily overcome the SC barrier. Based on the schematic illustrating the influence of irradiation intensity and wavelength on the SC barrier (Fig. 5e), both a higher light intensity and shorter wavelength contribute to the excitation of more high-energy electrons, thereby promoting more efficient SC interfacial electron transfer.
After Pt deposition, the influence of the temperature on the photovoltage diminished (Fig. 6b), indicating that nearly all the additional electrons generated by the temperature increase were rapidly transferred to Pt via the SC interface and subsequently to the solution (Fig. S8b and S9a). Further analysis of the SC electron transfer time constant reveals that increasing the temperature from 10 °C to 50 °C reduces the time constant by an average factor of 3.51 (Fig. 6c), confirming that electron transfer kinetics at the interface are markedly enhanced with rising temperature (Fig. 6d).
Analysis of SC interfacial charge transfer time constants (Fig. 7d) reveals that Sn doping enhances the transfer rate by factors of 1.46 (at 10 °C), 3.64 (at 30 °C), and 2.81 (at 50 °C). Combined with the increased electron concentrations observed in the Sn-doped samples (Fig. S7b), these findings indicate that Sn doping not only mitigates surface electron leakage but also enables photogenerated electrons to rapidly traverse the reduced SC potential barrier (Fig. S10), achieving efficient charge separation (Fig. 7e and Fig. S11). Consequently, Pt/1% Sn–TiO2 delivered significantly enhanced photothermal catalytic HE performance in LA solution.
Moreover, under identical temperature conditions in both the LA solution and 0.5 M Na2SO4 solution, Sn–TiO2 showed the lowest photocurrent density, while Pt/Sn–TiO2 showed the highest (Fig. 8b and Fig. S11). This confirms that Sn doping passivates TiO2, whereas Pt loading facilitates charge separation. Importantly, for all electrode materials, the photocurrent densities in the LA solution were significantly higher than those in 0.5 M Na2SO4 at equivalent temperatures, indicating that LA provides a more favourable environment for the HER. Overall, these findings indicate that Pt/Sn–TiO2 exhibits superior photothermal catalytic performance for HER.
The results indicate that, the temperature dependence of Pt/TiO2 under varying light intensities is not entirely positively correlated (dashed regions, Fig. 9b and c), which can be attributed to changes in the interfacial charge-transfer mechanisms. Under low-intensity irradiation (Fig. 9a and Fig. S12a), the elevated SC barrier restricts electron transfer primarily to thermally enhanced charge recombination.25 Since recombination accelerates with temperature,63 HER efficiency increases accordingly. At intermediate intensities (300 and 500 mW cm−2), the increased carrier generation partially lowers the SC barrier. While the overall transfer kinetics remained constrained, the dominant mechanism shifted toward thermionic emission. The associated charge leakage intensifies with temperature (Fig. 6a and S8a), outweighing the recombination benefits and suppressing HE efficiency, which is consistent with the observed OCP characteristics. When the irradiation intensity is further increased (Fig. 9d), the diminished SC barrier shifts the rate-limiting step to the cocatalyst–solution (CS) electrochemical barrier. Considering the temperature-dependent electrocatalytic enhancement caused by Pt (Fig. S9a), elevated temperatures significantly enhanced HER efficiency under high irradiation intensities.
In additional, the temperature-dependent HE behaviours of Pt/TiO2 at varying wavelengths also was investigated (Fig. S12b). The results showed that under constant temperature and normalised light intensity, both the HE rates and AQY increased with longer wavelengths. Specifically, 365 nm irradiation yielded the highest H2 production rate at 80 °C, with an AQY of 28.30%, followed by 313 nm (AQY, 3.02%) and 254 nm (AQY, 2.73%). This trend is likely due to the differences in light penetration at different wavelengths (Fig. S6). Over the temperature range of 20–80 °C, the HE rates under 254, 313, and 365 nm increased by factors of 13.49, 3.07, and 1.37, respectively, indicating that shorter wavelengths lead to a more pronounced thermal enhancement effect and stronger photo-thermal synergy, possibly owing to faster SC interfacial charge transfer under short-wavelength irradiation (Fig. 5d and Fig. S5d).
After Sn doping, Sn not only introduced surface states but also effectively suppressed charge leakage (Fig. S7 and S9b). Consequently, under both weak and strong light conditions, the photothermal catalytic HE performance of Pt/1% Sn–TiO2 mostly surpasses that of Pt/TiO2 (Fig. 9). Specifically, at light intensities of 50, 100, 300, and 500 mW cm−2 (AM 1.5G), the HE rate of Pt/Sn–TiO2 consistently exceeds that of Pt/TiO2 at the same temperature. Moreover, as the temperature increases, the HE effect of Pt/1% Sn–TiO2 becomes more pronounced, which can be attributed to the inhibition of charge leakage by Sn doping. Furthermore, under incident light intensities of 300 mW cm−2 (AM 1.5G) and 500 mW cm−2 (AM 1.5G), no decrease in the HE rates owing to temperature increase was observed. Instead, a stable positive correlation between the HE rates and temperature was observed. At the same temperature, the higher the incident irradiation intensity, the greater the HE rate. Notably, as the irradiation intensity increased to 1000 mW cm−2 (AM 1.5G), the gradual reduction of the SC barrier weakened the charge suppression effect caused by Sn doping (Fig. 9d). This explains the only slight increase in the HE rates between of the doped and undoped materials as the temperature increased. Although the HE rate of Pt/1% Sn–TiO2 reached its maximum value (36.279 mmol g−1 h−1) at 80 °C, the corresponding AQY was the lowest (only 5.14%). This phenomenon can be attributed to light scattering, reduced light absorption, and effective photon loss.65,66 At higher light intensities, these losses significantly diminish the utilisation efficiency of the effective photons, thereby impacting the quantum efficiency.
As shown in Fig. 10b, under irradiation by a 30 W low-pressure mercury lamp (254 nm, with an effective photon flux of 1.38 × 1019 photons per s), the HE rates of all catalysts increased with rising temperature. Notably, when the Sn doping level was 3%, the catalyst achieved the highest H2 production rate, reaching 82.86 mmol g−1 h−1 at 80 °C (below methanol–water azeotrope boiling point, ∼85 °C at 1 atm (ref. 68)), with a corresponding AQY of 40.27% (Fig. 10c), representing a 4.08-fold increase compared with the performance of Pt/TiO2. Meanwhile, the catalytic efficiency of Pt/3% Sn–TiO2 within the temperature range under 30 W UV light irradiation can reach up to 29.45–39.36 times that under dark conditions (Fig. S13a). Subsequently, experiments were conducted under higher irradiation intensity conditions (320 W, 254 nm, with an effective photon flux of 1.34 × 1020 photons per s1) (Fig. S13b). The average HE rate reached up to 413.5 mmol g−1 h−1 at 80 °C within 6 h, with the corresponding average AQY reaching 25.16%. Despite a nearly 10-fold increase in the incident irradiation intensity, the HE efficiency only improved by ∼5.0 times, resulting in a lower AQY compared with that observed under 30 W conditions. This discrepancy can be primarily attributed to the fact that under high irradiation intensities, the SC interfacial electron transfer process is predominantly governed by thermionic emission, which is accompanied by intensified charge leakage during this process.
Given that the AQY was higher at a power of 30 W, we conducted a stability test at 80 °C using the Pt/3% Sn–TiO2 catalyst (Fig. 10d). Although the AQY decreased after 10 h, during the subsequent 20 h reaction process, the average yields of H2 and HCHO were 88.12 and 84.13 mmol g−1 h−1, respectively, with an average quantum yield of 40.02%. These results indicate that Pt/3% Sn–TiO2 has excellent potential for practical applications in photothermal catalytic CH3OH reforming. Moreover, these production rates rank among the leading values in the field of photocatalytic dehydrogenation of CH3OH to HCHO (Fig. 10e). Post-reaction characterisation of the catalyst using XRD and TEM (Fig. S14), performed according to literature protocols,69,70 indicated no significant changes in the phase composition or microstructure, confirming its high structural stability under these reaction conditions.
Analysis of the results reveals that the characteristic peaks of CH3OH remained relatively stable throughout the reaction, showing slight attenuation (e.g., 2948 cm−1) over time (Fig. 11c). This can be attributed to the continuous replenishment of CH3OH in the reaction chamber. During the experiment, Ar gas was introduced into the reaction system by bubbling it through a 20 vol% CH3OH solution, after which the system was sealed for detection. If the system had been purged with inert gas at this stage to remove the physically adsorbed species, the amount of CH3OH remaining on the catalyst surface would likely have been insufficient to produce detectable reaction-related signals. In addition, several intermediate products were observed during the reaction, including methoxy (2921, 2820, 1143, and 1055 cm−1) and formate species (1649, 1560, and 1355 cm−1). The gradual increase (e.g., 1143, 1560, and 1355 cm−1) in the peak intensity indicates that the reaction progressed continuously. Moreover, the signals corresponding to the final products CO2 (2353 and 2319 cm−1) and CO (2046 cm−1) became increasingly pronounced as the reaction proceeded. By contrast, functional groups associated with the target product HCHO were barely detected, indicating that under gas-phase reaction conditions, HCHO is highly susceptible to further photocatalytic oxidation, a finding consistent with previously reported literature.71,73,74 A long-term photothermal catalytic experiment conducted under liquid-phase conditions (Fig. 10e) revealed that HCHO production declined significantly after reacting for 10 h, potentially owing to its subsequent oxidation.
To further investigate the products of the liquid-phase reaction, the CH3OH solution obtained after 12 h of reaction was analyzed using GC–MS and HPLC (Fig. S15). The mass spectrum (Fig. S15a) exhibited a weak signal corresponding to the HCHO fragment ion at m/z 29, whereas intense peaks arise at m/z 89, 119, 149, and 179. As reported in the literature,75–78 these signals are characteristic of oligomeric HCHO fragments, indicating that HCHO exists predominantly in polymerised form in aqueous solution, with only trace amounts present as free molecules. Additional ion peaks at m/z 45 and 59 are tentatively assigned to HCOOH and methyl formate, respectively, consistent with the DRIFTS results and previous reports.79 The presence of HCHO was also confirmed by HPLC analysis (Fig. S15b).
Fig. 11e shows the reaction mechanism for the photothermal catalytic dehydrogenation of CH3OH to HCHO. Upon photoexcitation of TiO2, the oxidation half-reaction involving photogenerated holes (h+) proceeds more rapidly than the reduction half-reaction involving photogenerated electrons (e−). This kinetic imbalance led to accumulation of negative charges on the catalyst surface. Although Pt serves as an effective reduction cocatalyst that facilitates electron transfer from the semiconductor to the solution, its efficiency remains insufficient to match the rapid oxidation process (Fig. S1 and S16). To address this imbalance, photothermal synergy leverages the combined effects of irradiation intensity, light wavelength, and temperature to significantly enhance interfacial electron transfer. Specifically, under higher irradiation flux, shorter wavelengths generate higher-energy photoelectrons that can overcome the semiconductor energy barrier more effectively, thereby improving charge separation and migration. Although elevated temperatures may promote some degree of charge leakage, they concurrently accelerate interfacial charge-transfer kinetics. The overall effect is the enhancement of charge-transfer efficiency. Consequently, the catalytic performance demonstrates a moderate positive correlation with temperature.
However, in conventional dispersed-catalyst systems, the limited penetration depth of short-wavelength light restricts photon absorption by the photocatalyst, thereby limiting the overall efficiency. Furthermore, increasing the temperature exacerbates charge leakage, counteracting the benefits of thermal activation and yielding only a marginal improvement in the activity of Pt/TiO2 with increasing temperature. Sn doping effectively addresses these limitations by suppressing charge leakage, increasing the electron density, reducing the SC barrier, and facilitating electron transfer at the SC interface. Concurrently, the adoption of a fixed-bed reactor mitigates the poor transmittance of short-wavelength light through liquid media. Through synergistic optimisation of both catalyst and reactor designs, Pt/3% Sn–TiO2 exhibited excellent performance in the photothermal catalytic dehydrogenation of CH3OH to HCHO (Fig. 10e).
In the proposed reaction pathway, adsorbed CH3OH molecules on the catalyst surface initially undergo C–H bond cleavage via hydroxyl radicals (˙OH) generated by h+, forming surface-bound –CH3O and H2O.79,80 The –CH3O species are subsequently dehydrogenated through oxidation by h+, yielding coordinated HCHO and releasing atomic H. The H atoms are adsorbed and bridged on adjacent Pt sites, where they combine with other electron-acquired H atoms to form H2.81–83 HCHO molecules desorb into the solution and gradually polymerise into oligomeric species as the concentration increases.75–77 Simultaneously, some adsorbed –CH3O groups may couple with neighbouring HCHO to form methyl formate.84 Adsorbed HCHO may also further oxidise by adjacent ˙OH radicals, resulting in the formation of HCOOH or other by-products.85
Original data are available from the authors upon reasonable request.
During the preparation of this study the authors used AI-assisted technologies [https://ai.youdao.com/] to enhance the readability and language of the manuscript.
| This journal is © The Royal Society of Chemistry 2026 |