Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Precipitation synthesis and characterization of SnO2@g-C3N4 heterojunctions for enhanced photocatalytic H2 production

Rizwan Khanab, Shah Sawar Ahmadb, Hasnain Ihsanb, Syeda Sheeza Nadeemc, Syed Zulfiqarb and Ferry Anggoro Ardy Nugroho*ad
aDepartment of Physics, Faculty of Mathematics and Natural Sciences, Universitas Indonesia, Depok 16424, Indonesia. E-mail: f.a.a.nugroho@sci.ui.ac.id
bDepartment of Physics, Abdul Wali Khan University Mardan, Mardan, Pakistan
cDepartment of Chemistry, Faculty of Mathematics and Natural Sciences, Universitas Indonesia, Depok 16424, Indonesia
dInstitute for Advanced Sustainable Materials Research and Technology, Faculty of Mathematics and Natural Sciences, Universitas Indonesia, Depok 16424, Indonesia

Received 27th May 2025 , Accepted 22nd July 2025

First published on 28th July 2025


Abstract

This study reports the development of SnO2@g-C3N4 heterojunctions, a hybrid semiconductor photocatalyst with varying mass percent ratios using a facile precipitation method for hydrogen (H2) production. The synergistic effect between the SnO2 nanoparticles and g-C3N4 sheets suppresses the charge recombination and enhances carrier separation, leading to improved photocatalytic activity. The nanocomposites demonstrate increased hydrogen production across all composites, with SC-20 sample (i.e., 80% SnO2 and 20% g-C3N4) achieving the highest H2 production rate of 287.7 μmol g−1 h−1, that is, 1.87-fold and 1.63-fold higher than that of SnO2 and of g-C3N4 counterparts, respectively. Furthermore, the nanocomposites maintain excellent photostability. Specifically, SC-20 achieves approximately 1500 μmol H2 evolution per 5 hour-cycle. The facile precipitation-based synthesis and enhanced photocatalytic activity of the SnO2@g-C3N4 nanocomposite position it as a reliable, cost-effective, and sustainable candidate for solar-driven hydrogen production and other clean energy applications.


Introduction

Population growth and rapid industrialization have significantly increased the global demand for energy, which is predominantly met by finite, non-renewable petroleum resources.1–3 The continuous depletion of these energy reserves, combined with the environmental damage caused by their combustion, has accelerated global warming and climate change, creating an energy crisis that threatens both energy security and environmental sustainability.4–8 To address these critical issues, relevant stakeholders have prioritized the development of alternative energy sources that are environmentally friendly, sustainable, and cost-effective. In this context, hydrogen energy has emerged as a promising solution to mitigate the energy crisis and to reduce environmental pollution due to its clean nature and high energy content.9

Prominent approaches to produce hydrogen include semiconductor-based photocatalytic water splitting. In particular, such a method has gained significant attention as a renewable and eco-friendly approach, as it not only provides a clean energy source but also reduces ecological pollution by decreasing reliance on fossil fuels.10,11 Over the past few decades, various semiconductor materials explored for this purpose, specifically metal oxide materials including SrTiO3, TiO2, ZrO2, Ta2O5, ZnO, WO3, and SnO2, have been systematically studied for their photocatalytic properties.1,4,5,12–14 However, these materials face significant limitations. For instance, TiO2 and SnO2, with a large bandgap energy, absorb only UV light, utilizing merely 4% of sunlight. On the other hand, materials like ZnO are prone to photo-corrosion under illumination, while WO3 is inactive for H2 production due to its low conduction band edge potential.13–15

In response, several composite photocatalysts have been developed to address these limitations, among which graphitic carbon nitride (g-C3N4) emerged as a promising candidate.16,17 g-C3N4 is a metal-free polymeric semiconductor with a suitable bandgap of 2.7 eV, enabling efficient sunlight absorption and charge carrier excitation.18,19 Additionally, its high chemical and thermal stability, attributed to its polymeric structure and degree of polymerization, render it a robust material for photocatalytic applications.20,21 Furthermore, g-C3N4 is cost-effective and can be readily synthesized through the simple thermal decomposition of urea.22 As a result, it has been extensively explored for photocatalytic hydrogen evolution,23,24 and a variety of g-C3N4-based heterostructures have been reported, including BiOCl/g-C3N4,25 and CdS/g-C3N4,26,27 and metal oxide-based ones such as Bi2WO6/g-C3N4,28,29 TiO2/g-C3N4,30,31 ZnO/g-C3N4,32,33 and TaON/g-C3N4,34 which have shown promise in suppressing charge recombination and enhancing photocatalytic performance through the formation of heterojunction.35–39

Among metal oxide-based heterostructures, tin dioxide (SnO2) is notable due to its non-toxicity, low cost, and excellent optical, physical and photoelectrochemical properties,40,41 which understandably finds wide range applications in, e.g., energy storage, gas sensing, solar cells, photocatalysis, electronics, and electrochemical cells.42–44 In the context of photocatalysts, previous studies have shown that SnO2 outperforms ZnO and TiO2 as an electron acceptor, making it a more appealing candidate for such a system.45 With these properties, SnO2 thus constitutes a rational choice to be coupled with g-C3N4 to enhance its photocatalytic performance.46 In recent years, several studies have reported the synthesis of SnO2/g-C3N4 heterostructures for photocatalytic hydrogen production using various methods, including ultrasonic-assisting deposition,47 sol–gel,48 hydrothermal,49,50 and solid-phase methods.51,52 While these techniques offer advantages like crystallinity, good interfacial contact, and promising H2 evolution performance, they are often time consuming, require high temperature and pressure, and involve toxic solvents and specialized equipment, making them less scalable and environmentally friendly.53 Therefore, developing simple, low-cost, scalable and eco-friendly techniques to synthesize SnO2@g-C3N4 photocatalyst for hydrogen evolution is highly desirable.

As a response, in this study we report the successful synthesis of SnO2@g-C3N4 hybrid photocatalyst via a simple and cost-effective precipitation method, which is widely recognized for producing metal oxide nanoparticles under mild conditions (ambient temperature and pressure) without the use of toxic gases.54–57 To the best of our knowledge, no previous studies reported the use of precipitation-based synthesis for constructing SnO2/g-C3N4 heterojunctions for photocatalytic H2 production. To address this gap, we systematically varied the mass ratios of SnO2 to g-C3N4, and we found SC-20 sample (i.e., 80% SnO2 and 20% g-C3N4) to deliver the highest and stable photocatalytic hydrogen production, indicating an optimal interface for charge transfer. Aided with various materials characterization, the improved hydrogen production is attributed to the enhancement in the photo-induced charge carrier separation and suppressed charge recombination. This performance positions our synthesis strategy to be comparable with other complex methods employing noble metal catalysts, highlighting its potential for further optimization and practical applications.

Experimental section

Chemicals

All chemicals, including urea (CH4N2O), SnCl2·2H2O and NH3 were purchased from Sigma-Aldrich without further purification. DI water was produced using water distillation apparatus DU-L4 MEDILAB.

Synthesis of g-C3N4 nanosheets

g-C3N4 nanosheets were prepared using the bath sonication method. First, 10 g of urea (CH4N2O) was heated in a muffle furnace at 550 °C for 3 h to produce bulk g-C3N4, which was subsequently ground into a powder. To convert this into nanosheets, 0.05 g of g-C3N4 powder was sonicated in 50 mL of distilled water for 90 min. The material was washed seven times with distilled water and dried at 70 °C for 24 h, yielding the desired light-yellow g-C3N4 nanosheets.

Preparation of SnO2 nanoparticles

SnO2 nanoparticles were synthesized via precipitation. First, 2.5 g of SnCl2·2H2O was dissolved in 50 mL DI water under stirring. A diluted ammonia solution (4 mL of 35% ammonia in 26 mL DI water) was added dropwise under continuous stirring to the SnCl2·2H2O solution until the mixture turned milky at pH 9. The precipitate was washed five times with DI water by centrifugation, dried at 60 °C for 24 h, ground into a powder, and annealed at 400 °C for 2 h.

Preparation of SnO2@g-C3N4 nanocomposites

SnO2@g-C3N4 nanocomposites with varying mass ratios were synthesized via a precipitation method as illustrated in Fig. 1. Initially, a measured quantity of g-C3N4 was dispersed in DI water, followed by the dissolution of SnCl2·2H2O under continuous stirring. A 35% ammonia solution was added dropwise to the mixture under stirring until a milky precipitate formed at pH 9. The precipitate was washed five times with DI water, dried at 60 °C for 24 h, ground into a fine powder, and annealed at 400 °C for 2 h. The prepared composites were labeled as SC-X, where X denotes the at% of the g-C3N4, that is, SC-10 comprises 90% SnO2 and 10% g-C3N4, SC-20 does 80% SnO2 and 20% g-C3N4, SC-30 does 70% SnO2 and 30% g-C3N4, and SC-40 does 60% SnO2 and 40% g-C3N4.
image file: d5ra03721b-f1.tif
Fig. 1 Step-by-step schematic illustration of synthesis of SnO2@g-C3N4 nanocomposites. First, the pre-synthesized g-C3N4 nanosheets are dispersed in 50 mL of deionized (DI) water. Subsequently, SnCl2·2H2O and NH3 are added to the solution, followed by stirring and pH adjustment. After washing and drying, the resulting precipitates are then annealed. The final product, SnO2@g-C3N4 composites, appears as light-yellow sheets with black dots on the surface.

Physiochemical characterizations

Various characterization techniques were employed to analyze the properties of the prepared samples. XRD (JDX-3532, JEOL) with Cu Kα radiation (λ = 1.5418 Å) at 40 kV and 30 mA confirmed the tetragonal structure for SnO2, hexagonal structure for g-C3N4, and the coexistence of both in SnO2@g-C3N4 composites. FTIR (Cary 630, Agilent Technologies) identified Sn–O and C–N vibrational modes in SnO2 and g-C3N4, respectively. SEM (JSM-6490A, JEOL) revealed sheet-like morphology for g-C3N4, particle morphology for SnO2, and heterojunction formation in the composites. UV-vis (PerkinElmer) spectra showed enhanced absorbance in the visible region for the composites. PL (LS45 PerkinElmer) investigated the lower recombination of photo-generated electrons.

Photoelectrochemical measurements

Photocurrent (PC) measurements and electrochemical impedance spectroscopy (EIS) were performed using a CHI-660E electrochemical workstation with platinum counter electrode, Ag/AgCl reference electrode, and 0.5 M Na2SO4 electrolyte solution.

Photocatalytic H2 generation measurements

Photocatalytic hydrogen production was evaluated in a flask irradiated by a 300 W xenon lamp (CEL HXF300), equipped with a cutoff filter (λ > 420 nm). For the experiment, 20 mg of the prepared sample was dispersed in 80 mL of aqueous solution containing 10% methanol as the hole scavenger, via ultrasonication. Prior to irradiation, the system was purged with nitrogen (N2) for 15 min to eliminate any residual oxygen. The reaction was carried out under continuous stirring and illumination for a period of 5 h, and the amount of hydrogen produced was quantified using gas chromatography (GC-2014, Shimadzu).

Results and discussions

As the first step of our study, we investigate the structural properties and the molecular composition of our prepared samples using X-ray diffraction (XRD). Fig. 2a shows the obtained XRD patterns of pure g-C3N4 and SnO2, their composites, as well as the corresponding JCPDS references for g-C3N4 and SnO2. The diffraction pattern of pure g-C3N4 shows a characteristic peak at 2θ = 27.5° corresponding to the (002) reflection of its graphitic stacking structure, which matches perfectly with the corresponding standard reference (ICSD 01-087-1526), confirming the hexagonal structure of pure g-C3N4.58 In the case of pure SnO2, observed peaks at 26.4°, 34°, 38°, 51.5°, and 65° correspond to (110), (101), (200), (211) and (301) planes, respectively, corroborating the tetragonal rutile structure of the SnO2 (JCPDS 41-1445).59 Examining the patterns of the composites (SC-10 to SC-40), the coexistence of both hexagonal g-C3N4 and tetragonal SnO2 phases is confirmed, in that the corresponding patterns comprise the characteristic peaks of the constituents.60 This finding thus corroborates the successful formation of heterojunction structures, notably without introducing any impurity phase. In detail, the intensity of the SnO2 diffraction peaks gradually decrease with increasing g-C3N4 content from SC-10 to SC-30. Interestingly, among the composites, SC-40 shows a diffraction pattern closely resembling with pure g-C3N4, likely due to the higher g-C3N4 content in SC-40.
image file: d5ra03721b-f2.tif
Fig. 2 X-ray diffraction pattern of g-C3N4, SnO2, and their composites, alongside its standard reference. The coexistence of both g-C3N4 and SnO2 signature peaks in the composite samples confirms their successful formation.

Moreover, FTIR spectroscopy was performed to analyze the chemical bonding and functional groups of the samples. As shown in Fig. 3, the pristine g-C3N4 exhibit peaks at 1243–1637 cm−1, which correspond to C–N and C[double bond, length as m-dash]N stretching vibrations, at 808 cm−1, which is attributed to the tri-s-triazine ring structure, and at 3180–3331 cm−1, which belongs to N–H stretching mode.58,61 For pure SnO2, a broad peak at 659 cm−1 represents the Sn–O stretching mode in Sn–O–Sn.62 To this end, the absence of additional peaks in the spectra of pure g-C3N4 and SnO2 confirms their high purity, consistent with the XRD data above. In the case of the composites (SC-10 to SC-30), the distinct vibrational modes of both g-C3N4 and SnO2 are observed, corroborating the successful formation of the nanocomposites. The FTIR spectrum of SC-40, however, closely resembles that of g-C3N4, again indicating a dominant presence of g-C3N4 in this composition.


image file: d5ra03721b-f3.tif
Fig. 3 FTIR spectra of pure g-C3N4, SnO2, and their composites in the range of 500–4000 cm−1. The characteristic C–N, C[double bond, length as m-dash]N, and Sn–O bonds, along with some additional bonds, are observed in pure g-C3N4 and SnO2, as well as their coexistence in the composites.

Last, Scanning Electron Microscopy (SEM) was employed to analyze the morphology of the synthesized samples. As shown in Fig. 4a and b, pure g-C3N4 exhibits a sheet-like morphology, though the sheets are not distinctly visible due to agglomeration. Meanwhile, Fig. 4c and d confirm the formation of well-defined SnO2 nanoparticles with average diameters around 94.5 nm (Fig. S1). When integrated into a heterostructure, the SnO2 nanoparticles are expected to spread uniformly on the g-C3N4 sheets (Fig. 4e and f). In addition, energy dispersive X-ray (EDX) analysis confirmed the elemental composition and purity of g-C3N4, SnO2 and SnO2@g-C3N4 nanocomposite (Fig. S2). Specifically, the g-C3N4 spectra displays distinct peaks corresponding to elements C and N, while SnO2 shows peaks attributed to elements Sn and O. Finally, the SnO2@g-C3N4 composite exhibits peaks Sn, O, C and N, confirming no impurity and creation of highly pure composite.


image file: d5ra03721b-f4.tif
Fig. 4 Heterostructure morphology. SEM image of (a and b) g-C3N4, (c and d) SnO2 and (e and f) SnO2@g-C3N4 nanocomposites. g-C3N4 assumes sheet-like morphology, while SnO2 does particle-like morphology. In the composite form, both these morphologies are observed.

Next, we proceed to investigate the photoctalytic H2 production performance of the synthesized heterostructures, along with those of their pure counterparts. To this end, we performed the photocatalytic reaction under visible light irradiation for 5 h (Fig. 5a). First we note that the hydrogen production of pure g-C3N4 is higher than that of SnO2, throughout the reaction. Converting into production rate, pristine g-C3N4 exhibits H2 production of 175.98 μmolg−1 h−1 compared to 160.84 μmolg−1 h−1 of SnO2 (Fig. 5b). This behavior can be explained by the narrower bandgap, and thus higher visible light absorption, of g-C3N4 compared to SnO2, as we will show later. Notably, and of our interest here, all four composites exhibit greater hydrogen generation than pure SnO2 and g-C3N4, indicating the formation of an effective heterojunction that promotes interfacial charge transfer and suppresses photogenerated charge recombination. Among the composites, the champion sample is SC-20, whose production reaches 287.7 μmolg−1 h−1, which is 1.63 times higher than g-C3N4 and 1.79 times higher than SnO2. This performance enhancement can be attributed to the optimal ratio between SnO2 and g-C3N4, which ensures sufficient interfacial contact for charge while maintaining visible light absorption. From the results in the composites it is evident that as the amount of SnO2 increases, the hydrogen production increases, reaching a maximum at SC-20. Surprisingly, further increasing the SnO2 content leads to a decline in hydrogen production at SC-10. This decline can be attributed to the shielding effect, as the g-C3N4 nanosheets become fully covered by SnO2 nanoparticles. Since SnO2 is a wide-bandgap semiconductor, this coverage inhibits the effective absorption of visible light, thereby reducing hydrogen production.63,64 Furthermore, the hydrogen production level in SC-40 is comparable to that of pure g-C3N4, as the high g-C3N4 content dominates. This observation is consistent with the XRD pattern of the SC-40 above. Establishing that SC-20 outperforms all other composites, in the subsequent discussion we will focus on SC-20, alongside pure g-C3N4 and SnO2 for further analysis regarding the enhancement mechanism.


image file: d5ra03721b-f5.tif
Fig. 5 Photocatalytic H2 production performance of the as prepared SnO2, g-C3N4, and their composites. (a) H2 production over 5 h, showing SC-20 achieving the highest performance; (b) H2 production rate per hour, where SC-20 outperforms pure SnO2, g-C3N4, and other composites, a clear increasing trend is observed from SC-40 to SC-20, reaching a value of 287.7 μmolg−1 h−1. However, a further increase in SnO2 content (SC-10) results in a decline in H2 production, indicating SC-20 as the optimal composition.

To better understand the enhanced hydrogen evolution performance of SC-20 compared to its counterparts, additional verification of its optical properties is necessary. To assess the light-harvesting capabilities, UV-vis spectroscopy was conducted in the wavelength range of 200–800 nm. The UV-vis spectra of pure SnO2, g-C3N4, and SnO2@g-C3N4 nanocomposites, along with their corresponding Tauc plots, are presented in Fig. 6a–d, respectively. Pure g-C3N4 shows strong absorption in the visible region due to its narrow bandgap (2.62 eV), while SnO2 exhibit absorption predominantly in the UV region, attributed to its wider bandgap (3.29 eV). Notably, SC-20 sample demonstrates a broader absorption across both visible and UV regions. The Tauc plot of the composite SC-20 reveals an apparent bandgap of approximately 2.75 eV, which represents the integrated optical response of the heterostructure. This apparent bandgap likely arises from the synergistic interaction between SnO2 and g-C3N4, leading to enhanced photoresponse and improved photocatalytic activity.65,66


image file: d5ra03721b-f6.tif
Fig. 6 Optical properties of g-C3N4, SnO2, and SC-20 composite. UV-vis spectra of (a) g-C3N4, (b) SnO2 and (c) SC-20 composite, along with their corresponding Tauc's plots shown as in (d). The band gaps were estimated to be 2.62 eV for g-C3N4, 3.29 eV for SnO2, and 2.75 eV for SC-20. The slight increase in band gap of SC-20 compared to g-C3N4 is due to incorporation of SnO2, while maintaining visible-light absorption, indicating successful heterojunction formation.

Furthermore, to evaluate the separation efficiency and photoresponse of the photocatalysts, time-resolved photocurrent measurements were carried out under visible light illumination. In a photocurrent analysis, the sample is sandwiched between electrodes and then excited by a light pulse to generate charges. The photo-generated charges produce current on the electrodes, measured under light-on and light-off. The light-on phase represents charge buildup on electrodes, while light-off phase shows charge decay. As shown in Fig. 7a, SC-20 heterojunction exhibits the highest and most stable photocurrent density compared to pure SnO2 and g-C3N4. The sharp photocurrent spikes under light-on and rapid decay during light-off indicate fast charge generation and relatively good carrier mobility. This enhanced photocurrent response in SC-20 can be attributed to improved charge separation and interfacial coupling between SnO2 and g-C3N4, which facilitate directional charge transfer and reduce recombination losses. To further investigate charge transport dynamics, electrochemical impedance spectroscopy (EIS) was performed. In the Nyquist plots shown in Fig. 7b, SC-20 displays the smallest semicircular arc radius, indicating the lowest charge transfer resistance among the tested samples, followed by g-C3N4 and SnO2. The decreased arc radius for SC-20 validates the formation of an efficient heterojunction, supporting the photocurrent findings. Moreover, photoluminescence (PL) spectroscopy was employed to assess photo-generated charge carrier recombination behavior in photocatalysts. In Fig. 7c SC-20 displayed significantly quenched photoluminescence intensity, compared to pure SnO2 and g-C3N4. This quenching is the evidence of enhanced suppression of charge carrier recombination in the composite. These results collectively indicate that SC-20 heterostructure exhibits enhanced charge carrier separation, faster interfacial electron transport, and reduced recombination rates compared to the pure components.


image file: d5ra03721b-f7.tif
Fig. 7 Photocurrent, EIS, and PL analyses of SnO2, g-C3N4, and SC-20. (a) Transient photocurrent response of pure SnO2, g-C3N4, and SC-20 composite under chopped visible light irradiaion, shows that SC-20 exhibits a higher photocurrent compared to pure SnO2 and g-C3N4, attributed to the synergistic effect in SC-20. (b) Nyquist plots of SC-20, g-C3N4, and SnO2, illustrating their electrochemical impedance properties. SC-20 exhibits the smallest semi-circle, confirming the low charge transfer resistance in composite as compared to pure counterparts. (c) PL spectra of SC-20, g-C3N4, and SnO2. SC-20 shows the lowest PL, indicating low charge recombination under illumination of visible light.

Last, to evaluate the long-term photostability and practical viability of our SnO2@g-C3N4 (SC-20) nanocomposite, a four-cycle photocatalytic hydrogen evolution test was conducted, with each cycle lasting 5 h under visible light irradiation (Fig. 8). Throughout all four cycles, the H2 evolution rate remained nearly constant, suggesting excellent structural and photoelectrochemical stability. The consistent performance indicates that the heterojunction interface effectively prevents photocorrosion and supports stable charge separation over extended illumination periods. Specifically, SC-20 achieves approximately 1500 μmol H2 evolution per 5 hour-cycle, demonstrating negligible performance loss. This robustness surpasses that of many conventional such as TiO2, ZnO, Fe2O3, and WO3 based photocatalysts, which typically exhibit a noticeable decline in activity due to rapid charge accumulation, poor stability and limited visible light absorption.67–70 The photocyclic results thus highlight the material's potential for real PEC applications requiring long-term operation without catalyst regeneration or replacement.


image file: d5ra03721b-f8.tif
Fig. 8 Photostability of SC-20. Time-dependent H2 evolution of SC-20 over four consecutive 5-hour cycles under visible light irradiation. Our best composite SC-20 maintains nearly constant H2 production rate across all four cycles, demonstrating its strong photostability.

Comparative study

Having established the enhanced performance of our heterostructures compared to their pure counterparts, it is then interesting to benchmark them against other works reported in the literature employing similar heterostructures, however using different synthesis routes. As shown in Table 1, our composite, synthesized via a simple precipitation method, achieved a hydrogen evolution of 287.7 μmolg−1 h−1, positioning it within the middle range of previously reported works. Notably, the heterostructures incorporating Pt as a co-catalyst or utilizing more advanced nanostructures such as quantum dots or nanodots have shown superior performance, often exceeding 1300 μmolg−1 h−1. However, these methods often need precise control over nanostructures formation and multistep complicated synthesis methods, which can limit large-scale production. In conclusion, while advanced nanostructures and Pt incorporation boost H2 evolution, our coprecipitation-based SnO2@g-C3N4 nanocomposites offers a competitive balance of photocatalytic performance, synthesis simplicity, and cost-effectiveness. These findings suggest that further enhancements of photocatalytic activity may be possible through the incorporation of co-catalysts like Pt or quantum dots, as well as through refined interface engineering strategies.
Table 1 H2 evolution performance of SnO2/g-C3N4-based photocatalysts reported in the literature
Photocatalyst Preparation method Amount of H2 gas evolved References
g-C3N4/SnO2 Solvent evaporation followed by calcination 132 μmol h−1 71
Pt g−1-C3N4/SnO2 One pot pyrolysis 241 μmol h−1 g−1 72
SnO2@g-C3N4nanocomposites Coprecipitation 287.7 μmol h−1g−1 Our work
Pt g−1-C3N4/SnO2 Simple calcination 627[thin space (1/6-em)]μmol h−1 73
g-C3N4/SnO2–Pt Physical mixing 900 μmol h−1 g−1 74
C3N4–SnO2–Pt Hydrothermal 1060 μmol h−1 g−1 75
SnO2 QDs/g-C3N4 Thermal decomposition 1305.4 μmol h−1 g−1 76
SnO2 nanodots/g-C3N4 One-step polymerization 1398.2 μmol h−1 g−1 77


Photocatalytic mechanism of SnO2@g-C3N4

To close the discussion, a Z-scheme heterojunction mechanism is proposed based on band structure alignment and charge transfer behavior to explain the enhanced photocatalytic activity of SnO2@g-C3N4 composite. As illustrated in Fig. 9, g-C3N4 is a visible light-driven photocatalyst with bandgap ≈ 2.62 eV (cf. Fig. 6d) while the photoresponse of SnO2 is limited to UV-region with a bandgap ≈ 3.29 eV (cf. Fig. 6d). Furthermore, the conduction band position of g-C3N4 is (−1.12 eV vs. NHE) while SnO2 have (0.05 eV vs. NHE).78,79 upon solar illumination, both SnO2 and g-C3N4 are excited to generate electron–hole pairs. In the Z-scheme configuration, photogenerated electrons in the conduction band of SnO2 recombine with holes in the valence band of g-C3N4 at the heterojunction interface. This unique charge transfer pathway preserves the highly energetic electrons in the CB of g-C3N4 and the strong oxidizing holes in the VB of SnO2, effectively enhancing the overall redox capability of the heterojunction.60 As a result, the photogenerated electrons in the conduction band of g-C3N4 take participation in the reduction of H+ ions to produce H2. Meanwhile, photogenerated holes in the valence band of SnO2 oxidize water molecules, generating O2. The synergistic effect between these two components enhances overall photocatalytic performance. This mechanism aligns well with the experimental findings, including increased photocurrent response, reduced charge transfer resistance (EIS), and quenched PL intensity all of which confirm efficient charge separation and transport in the SnO2@g-C3N4 heterostructure.
image file: d5ra03721b-f9.tif
Fig. 9 Probable mechanism of the SnO2@g-C3N4 photocatalyst: sunlight excites electrons in the valence band of g-C3N4, which transfer to the SnO2 conduction band, driving the reduction of H+ ions to H2. Simultaneously, holes oxidize water to generate O2. The diagram highlights the synergistic effect of combining SnO2 and g-C3N4, to enhance photocatalytic performance.

In addition to improved and charge carrier dynamics, the hydrogen adsorption on the catalyst surface is a critical factor influencing overall HER efficiency. Literature-based Gibbs free energy (ΔG_H*) studies provide further thermodynamic insight into this aspect. ΔG_H* is a key thermodynamic parameter for evaluating a material's catalytic activity toward the hydrogen evolution reaction (HER).80 It reflects how favorably hydrogen atoms bind to the catalyst surface, which in turn governs the balance between adsorption and desorption steps. Ideally, an effective HER catalyst should have a ΔG_H* value close to zero, ensuring that hydrogen atoms can adsorb and desorb efficiently. If ΔG_H* is too positive, hydrogen adsorption is unfavorable, leading to sluggish HER kinetics. If it is too negative, hydrogen binds too strongly, hindering H2 release.81 For pristine g-C3N4, previous theoretical studies have reported a ΔG_H* value of approximately −0.54 eV, indicating relatively strong hydrogen binding and potentially slow desorption.82 In contrast, SnO2 surfaces have shown ΔG_H* values ranging from 0.85 eV to 2.37 eV at different crystallographic orientations and adsorption sites, suggesting unfavorable hydrogen adsorption that may limit the generation of active hydrogen intermediates.83,84 Although the Gibbs free energy of hydrogen atom adsorption for the SnO2@g-C3N4 heterojunction itself has not yet been theoretically investigated, the significantly improved experimental hydrogen evolution performance observed in this study and previously explored studies suggests that interfacial charge redistribution may help tune ΔG_H* closer to the thermoneutral range. This highlights the need for future theoretical studies to better understand the active sites and mechanisms of hydrogen adsorption in such heterostructures.

Conclusion

In summary, we have successfully synthesized SnO2@g-C3N4 heterojunctions using a cost-effective and environmentally friendly precipitation method. Four composites with varying mass ratios of SnO2 to g-C3N4 (SC-10, SC-20, SC-30, and SC-40) were prepared to evaluate their photocatalytic performance for hydrogen production under solar illumination. Among these, SC-20 (80% SnO2, 20% g-C3N4) exhibited the best performance, achieving the highest H2 production rate (287.7 μmolg−1 h−1) due to its optimal heterojunction structure, which effectively suppressed charge recombination and enhanced electron transfer to the conduction band. Compared to previously reported works, employing complex synthetic routes or noble metal co-catalysts, our approach offers a practical balance of photocatalytic efficiency, material simplicity, and stability. Importantly, the excellent photostability of our SnO2@g-C3N4 composite, demonstrated over extended cycling, further reinforces the potential of this system for sustainable solar-to-fuel applications. Furthermore, this synthesis platform can be enhanced by introducing additional functional components such as Pt co-catalyst or nanostructure modifiers to further improve charge carrier dynamics and overall efficiency.

Data avaibility

All data generated or analyzed during this study are included in the article and its ESI.

Author contributions

R. K.: investigation, formal analysis, data curation, validation, visualization, writing – original draft. S. S. A.: investigation. H. I.: investigation. S. S. N.: writing – original draft. S. Z.: conceptualization, data curation, methodology, validation, supervision, resources. F. A. A. N.: data curation, validation, supervision, resources, funding acquisition, writing – review and editing.

Conflicts of interest

The authors declare no conflict of interest.

Acknowledgements

We acknowledge funding by the Faculty of Mathematics and Natural Sciences, Universitas Indonesia, under Publication Grant scheme 2025.

References

  1. A. Kumar, et al., Recent progress in advanced strategies to enhance the photocatalytic performance of metal molybdates for H2 production and CO2 reduction, J. Alloys Compd., 2024, 971, 172665 CrossRef CAS .
  2. Y. Zheng, J. Liu, J. Liang, M. Jaroniec and S. Z. Qiao, Graphitic carbon nitride materials: controllable synthesis and applications in fuel cells and photocatalysis, Energy Environ. Sci., 2012, 5, 6717–6731 RSC .
  3. S. Patial, et al., Tunable photocatalytic activity of SrTiO3 for water splitting: Strategies and future scenario, J. Environ. Chem. Eng., 2020, 8, 103791 CrossRef CAS .
  4. S. Hou, Y. Li, W. Li, X. Ma and Y. Fan, Novel noble-metal-free NiCo-LDH/CdSe S-type heterojunction with built-in electric field for high-efficiency photocatalytic H2 production, J. Taiwan Inst. Chem. Eng., 2024, 156, 105394 CrossRef CAS .
  5. F. Liu, et al., Vacancy engineering mediated hollow structured ZnO/ZnS S-scheme heterojunction for highly efficient photocatalytic H2 production, Chin. J. Catal., 2024, 64, 152–165 CrossRef CAS .
  6. W. Shi, et al., Construction of ZrC@ ZnIn2S4 core–shell heterostructures for boosted near-infrared-light driven photothermal-assisted photocatalytic H2 evolution, Chem. Eng. J., 2023, 474, 145690 CrossRef CAS .
  7. J. Lu, et al., Construction of S-scheme heterojunction catalytic nanoreactor for boosted photothermal-assisted photocatalytic H2 production, Appl. Surf. Sci., 2024, 642, 158648 CrossRef CAS .
  8. Z. Yan, et al., Photocatalysis for synergistic water remediation and H2 production: A review, Chem. Eng. J., 2023, 472, 145066 CrossRef CAS .
  9. S. V. P. Vattikuti, P. A. K. Reddy, J. Shim and C. Byon, Visible-Light-Driven Photocatalytic Activity of SnO2-ZnO Quantum Dots Anchored on g-C3N4 Nanosheets for Photocatalytic Pollutant Degradation and H2 Production, ACS Omega, 2018, 3, 7587–7602 CrossRef CAS PubMed .
  10. M. R. Hoffmann, S. T. Martin, W. Choi and D. W. Bahnemann, Environmental Applications of Semiconductor Photocatalysts, Chem. Rev., 1995, 95, 69–96 CrossRef CAS .
  11. X. Chen, S. Shen, L. Guo and S. S. Mao, Semiconductor-based photocatalytic hydrogen generation, Chem. Rev., 2010, 110, 6503–6570 CrossRef CAS .
  12. R. Singh, Different anticipated criteria to achieve novel and efficient photocatalysis via green ZnO: scope and challenges, Int. J. Environ. Sci. Technol., 2022, 19, 9209–9242 CrossRef CAS .
  13. S. Martha, P. Chandra Sahoo and K. M. Parida, An overview on visible light responsive metal oxide based photocatalysts for hydrogen energy production, RSC Adv., 2015, 5, 61535–61553 RSC .
  14. K. Mallikarjuna, G. A. K. M. Rafiqul Bari, S. V. P. Vattikuti and H. Kim, Synthesis of carbon-doped SnO2 nanostructures for visible-light-driven photocatalytic hydrogen production from water splitting, Int. J. Hydrogen Energy, 2020, 45, 32789–32796 CrossRef CAS .
  15. K. Hashimoto, H. Irie and A. Fujishima, TiO2 photocatalysis: A historical overview and future prospects, Jpn. J. Appl. Phys., 2005, 44, 8269–8285 CrossRef CAS .
  16. L. Mao, et al., Simultaneous bulk and surface modifications of g-C3N4 via supercritical CO2-assisted post-treatment towards enhanced photocatalytic activity, Appl. Catal. B Environ. Energy, 2025, 362, 124712 CrossRef CAS .
  17. D. Liu, et al., Constructing asymmetric dual active sites of Ag single atoms and nitrogen defects on carbon nitride for enhanced photocatalytic H2O2 production, J. Mater. Sci. Technol., 2025, 223, 56–65 CrossRef CAS .
  18. T. Li, et al., Synthesis of g-C3N4/SmVO4 composite photocatalyst with improved visible light photocatalytic activities in RhB degradation, Appl. Catal., B, 2013, 129, 255–263 CrossRef CAS .
  19. W. Iqbal, et al., One-step large-scale highly active gC3N4 nanosheets for efficient sunlight-driven photocatalytic hydrogen production, Dalton Trans., 2017, 46, 10678 RSC .
  20. G. Dong, Y. Zhang, Q. Pan and J. Qiu, A fantastic graphitic carbon nitride (g-C3N4) material: Electronic structure, photocatalytic and photoelectronic properties, J. Photochem. Photobiol. C Photochem. Rev., 2014, 20, 33–50 CrossRef CAS .
  21. I. Papailias, et al., Chemical vs. thermal exfoliation of g-C3N4 for NOx removal under visible light irradiation, Appl. Catal., B, 2018, 239, 16–26 CrossRef CAS .
  22. Y. He, et al., Z-scheme SnO2−x/g-C3N4 composite as an efficient photocatalyst for dye degradation and photocatalytic CO2 reduction, Sol. Energy Mater. Sol. Cells, 2015, 137, 175–184 CrossRef CAS .
  23. A. Mishra, et al., Graphitic carbon nitride (g–C3N4)–based metal-free photocatalysts for water splitting: A review, Carbon, 2019, 149, 693–721 CrossRef CAS .
  24. Y. Zhu, D. Zhang, L. Gong, L. Zhang and Z. Xia, Catalytic activity origin and design principles of graphitic carbon nitride electrocatalysts for hydrogen evolution, Front. Mater., 2019, 6, 439322 Search PubMed .
  25. Y. Bai, P. Q. Wang, J. Y. Liu and X. J. Liu, Enhanced photocatalytic performance of direct Z-scheme BiOCl–g-C3N4 photocatalysts, RSC Adv., 2014, 4, 19456–19461 RSC .
  26. J. Zhang, et al., Efficient visible-light photocatalytic hydrogen evolution and enhanced photostability of core/shell CdS/g-C3N4 nanowires, ACS Appl. Mater. Interfaces, 2013, 5, 10317–10324 CrossRef CAS .
  27. L. Ge, et al., Synthesis and efficient visible light photocatalytic hydrogen evolution of Polymeric g-C3N4 coupled with CdS quantum dots, J. Phys. Chem. C, 2012, 116, 13708–13714 CrossRef CAS .
  28. J. Wang, X. Lian, S. Chen, H. Li and K. Xu, Effect of Bi2WO6/g-C3N4 composite on the combustion and catalytic decomposition of energetic materials: An efficient catalyst with g-C3N4 carrier, J. Colloid Interface Sci., 2022, 610, 842–853 CrossRef CAS .
  29. S. Qi, R. Zhang, Y. Zhang, X. Liu and H. Xu, Preparation and photocatalytic properties of Bi2WO6/g-C3N4, Inorg. Chem. Commun., 2021, 132, 108761 CrossRef CAS .
  30. X. Du, X. Bai, L. Xu, L. Yang and P. Jin, Visible-light activation of persulfate by TiO2/g-C3N4 photocatalyst toward efficient degradation of micropollutants, Chem. Eng. J., 2020, 384, 123245 CrossRef CAS .
  31. H. Yan and H. Yang, TiO2–g-C3N4 composite materials for photocatalytic H2 evolution under visible light irradiation, J. Alloys Compd., 2011, 509, 26–29 CrossRef .
  32. Y. He, Y. Wang, L. Zhang, B. Teng and M. Fan, High-efficiency conversion of CO2 to fuel over ZnO/g-C3N4 photocatalyst, Appl. Catal., B, 2015, 168–169, 1–8 CAS .
  33. W. Liu, M. Wang, C. Xu, S. Chen and X. Fu, Significantly enhanced visible-light photocatalytic activity of g-C3N4 via ZnO modification and the mechanism study, J. Mol. Catal. A:Chem., 2013, 369, 9–15 CrossRef .
  34. S. C. Yan, S. B. Lv, Z. S. Li and Z. G. Zou, Organic-inorganic composite photocatalyst of g-C3N4 and TaON with improved visible light photocatalytic activities, Dalton Trans., 2010, 39, 1488–1491 RSC .
  35. P. Mary Rajaitha, et al., Graphitic carbon nitride nanoplatelets incorporated titania based type-II heterostructure and its enhanced performance in photoelectrocatalytic water splitting, Appl. Sci., 2020, 2, 572 CAS .
  36. M. Ramachandra, S. Devi Kalathiparambil Rajendra Pai, J. Resnik Jaleel UC and D. Pinheiro, Improved Photocatalytic Activity of g-C3N4/ZnO: A Potential Direct Z-Scheme Nanocomposite, ChemistrySelect, 2020, 5, 11986–11995 CrossRef CAS .
  37. D. Li, et al., Synthesis of a g-C3N4-Cu2O heterojunction with enhanced visible light photocatalytic activity by PEG, J. Colloid Interface Sci., 2018, 531, 28–36 CrossRef CAS .
  38. T. Zhang, et al., A facile one-pot and alkali-free synthetic procedure for binary SnO2/g-C3N4 composites with enhanced photocatalytic behavior, Mater. Sci. Semicond. Process., 2020, 115, 105112 CrossRef CAS .
  39. A. R. Fareza, F. A. A. Nugroho, F. F. Abdi and V. Fauzia, Nanoscale metal oxides–2D materials heterostructures for photoelectrochemical water splitting—a review, J. Mater. Chem. A, 2022, 10, 8656–8686 RSC .
  40. H. Wang and A. L. Rogach, Hierarchical SnO2 nanostructures: Recent advances in design, synthesis, and applications, Chem. Mater., 2014, 26, 123–133 CrossRef CAS .
  41. Y. Li, et al., Rapid fabrication of SnO2 nanoparticle photocatalyst: computational understanding and photocatalytic degradation of organic dye, Inorg. Chem. Front., 2018, 5, 3005–3014 RSC .
  42. B. Xiong Wen Lou, C. Ming Li and L. A. Archer, Designed Synthesis of coaxial SnO2@ carbon hollow nanospheres for highly reversible lithium storage, Adv. Mater., 2009, 21, 2536–2539 CrossRef .
  43. J. Pan, R. Ganesan, H. Shen and S. Mathur, Plasma-modified SnO2 nanowires for enhanced gas sensing, J. Phys. Chem. C, 2010, 114, 8245–8250 CrossRef CAS .
  44. J. Pan, et al., Heteroepitaxy of SnO2 nanowire arrays on TiO2 single crystals: Growth patterns and tomographic studies, J. Phys. Chem. C, 2011, 115, 15191–15197 CrossRef CAS .
  45. M. T. Uddin, et al., Nanostructured SnO2-ZnO heterojunction photocatalysts showing enhanced photocatalytic activity for the degradation of organic dyes, Inorg. Chem., 2012, 51, 7764–7773 CrossRef CAS PubMed .
  46. K. N. Van, et al., Facile construction of S-scheme SnO2/g-C3N4 photocatalyst for improved photoactivity, Chemosphere, 2022, 289, 133120 CrossRef CAS PubMed .
  47. R. Yin, et al., SnO2/g-C3N4 photocatalyst with enhanced visible-light photocatalytic activity, J. Mater. Sci., 2014, 49, 6067–6073 CrossRef CAS .
  48. L. Peng, R. rong Zheng, D. wei Feng, H. Yu and X. ting. Dong, Synthesis of eco-friendly porous g-C3N4/SiO2/SnO2 composite with excellent visible-light responsive photocatalysis, Arab. J. Chem., 2020, 13, 4275–4285 CrossRef CAS .
  49. Y. Zhang, J. Liu, X. Chu, S. Liang and L. Kong, Preparation of g–C3N4–SnO2 composites for application as acetic acid sensor, J. Alloys Compd., 2020, 832, 153355 CrossRef CAS .
  50. K. N. Van, et al., Facile construction of S-scheme SnO2/g-C3N4 photocatalyst for improved photoactivity, Chemosphere, 2022, 289, 133120 CrossRef CAS PubMed .
  51. Y. He, et al., Z-scheme SnO2−x/g-C3N4 composite as an efficient photocatalyst for dye degradation and photocatalytic CO2 reduction, Sol. Energy Mater. Sol. Cells, 2015, 137, 175–184 CrossRef CAS .
  52. X. Wang and P. Ren, Flower-like SnO2/g-C3N4 heterojunctions: The face-to-face contact interface and improved photocatalytic properties, Adv. Powder Technol., 2018, 29, 1153–1157 CrossRef CAS .
  53. W. Ren, et al., Recent progress in SnO2/g-C3N4 heterojunction photocatalysts: Synthesis, modification, and application, J. Alloys Compd., 2022, 906, 164372 CrossRef CAS .
  54. A. K. Atul, S. K. Srivastava, A. K. Gupta and N. Srivastava, Synthesis and characterization of NiO nanoparticles by chemical co-precipitation method: an easy and cost-effective approach, Braz. J. Phys., 2022, 52, 2 CrossRef CAS .
  55. M. J. Ndolomingo, N. Bingwa and R. Meijboom, Review of supported metal nanoparticles: synthesis methodologies, advantages and application as catalysts, J. Mater. Sci., 2020, 55, 6195–6241 CrossRef CAS .
  56. G. N. Kokila, C. Mallikarjunaswamy and V. L. Ranganatha, A review on synthesis and applications of versatile nanomaterials, Inorg. Nano-Met. Chem., 2022, 54, 942–971 CrossRef .
  57. M. Aminzai, M. Yildirim and E. Y. Talanta, Metallic nanoparticles unveiled: Synthesis, characterization, and their environmental, medicinal, and agricultural applications, Talanta, 2024, 280, 126790 CrossRef CAS PubMed .
  58. P. Agale, V. Salve, S. Arade, S. Balgude and P. More, Tailoring structural and chemical properties of ZnO@ g-C3N4 nanocomposites through Sr doping: Insights from multi technique characterization, Solid State Sci., 2025, 166, 107960 CrossRef CAS .
  59. J. Cao, et al., Calcination Method Synthesis of SnO2/g-C3N4 Composites for a High-Performance Ethanol Gas Sensing Application, Nanomaterials, 2017, 7, 98 CrossRef PubMed .
  60. V. Salve, et al., Enhanced photocatalytic activity of SnO2@g-C3N4 heterojunctions for methylene blue and bisphenol-A degradation: effect of interface structure and porous nature, RSC Adv., 2025, 15, 15651–15669 RSC .
  61. L. Mao, et al., Simultaneous bulk and surface modifications of g-C3N4 via supercritical CO2-assisted post-treatment towards enhanced photocatalytic activity, Appl. Catal. B Environ. Energy, 2025, 362, 124712 CrossRef CAS .
  62. Y. Zang, L. Li, X. Li, R. Lin and G. Li, Synergistic collaboration of g-C3N4/SnO2 composites for enhanced visible-light photocatalytic activity, Chem. Eng. J., 2014, 246, 277–286 CrossRef CAS .
  63. H. Wu, et al., A facile one-step strategy to construct 0D/2D SnO2/g-C3N4 heterojunction photocatalyst for high-efficiency hydrogen production performance from water splitting, Int. J. Hydrogen Energy, 2020, 45, 30142–30152 CrossRef CAS .
  64. A. Zada, M. Khan, M. N. Qureshi, S. Y. Liu and R. Wang, Accelerating Photocatalytic Hydrogen Production and Pollutant Degradation by Functionalizing g-C3N4 With SnO2, Front. Chem., 2020, 7, 941 CrossRef .
  65. Y. Zang, L. Li, X. Li, R. Lin and G. Li, Synergistic collaboration of g-C3N4/SnO2 composites for enhanced visible-light photocatalytic activity, Chem. Eng. J., 2014, 246, 277–286 CrossRef CAS .
  66. H. Ji, et al., Construction of SnO2/graphene-like g-C3N4 with enhanced visible light photocatalytic activity, RSC Adv., 2017, 7, 36101–36111 RSC .
  67. H. Khan and M. U. H. Shah, Modification strategies of TiO2 based photocatalysts for enhanced visible light activity and energy storage ability: A review, J. Environ. Chem. Eng., 2023, 11, 111532 CrossRef CAS .
  68. K. Hasibur Rahman, A. Kumar Kar and K. C. Chen, Highly active ZnO/Fe3+-TiO2 photocatalysts for visible-light photodegradation application and its colour change behaviour by d-d transition, Mater. Sci. Eng., B, 2024, 305, 117394 CrossRef CAS .
  69. Q. Wang, et al., Hollow spherical WO3/TiO2 heterojunction for enhancing photocatalytic performance in visible-light, J. Water Proc. Eng., 2021, 40, 101943 CrossRef .
  70. A. Raub, R. Bahru and S. Nashruddin, Advances of nanostructured metal oxide as photoanode in photoelectrochemical (PEC) water splitting application, Heliyon, 2024, 10, e39079 CrossRef PubMed .
  71. A. Zada, M. Khan, M. N. Qureshi, S. Liu and R. Wang, Accelerating Photocatalytic Hydrogen Production and Pollutant Degradation by Functionalizing g-C3N4 With SnO2, Front. Chem., 2020, 7, 941 CrossRef PubMed .
  72. C. Cai, et al., Facile one-pot pyrolysis preparation of SnO2/g-C3N4 composites for improved photocatalytic H2 production, J. Chem. Technol. Biotechnol., 2022, 97, 2921–2931 CrossRef CAS .
  73. M. Ismael, E. Elhaddad and M. Wark, Construction of SnO2/g-C3N4 composite photocatalyst with enhanced interfacial charge separation and high efficiency for hydrogen production and Rhodamine B degradation, Colloids Surf. A Physicochem. Eng. Asp., 2022, 638, 128288 CrossRef CAS .
  74. Y. Zang, L. Li, X. Li, R. Lin and G. Li, Synergistic collaboration of g-C3N4/SnO2 composites for enhanced visible-light photocatalytic activity, Chem. Eng. J., 2014, 246, 277–286 CrossRef CAS .
  75. X. Wang, M. Xue, X. Li, L. Qin and S.-Z. Kang, Boosting the photocatalytic H2 production performance and stability of C3N4 nanosheets via the synergistic effect between SnO2 nanoparticles and Pt nanoclusters, Inorg. Chem. Commun., 2021, 133, 108976 CrossRef CAS .
  76. S. V. P. Vattikuti, H. P. K. Sudhani, M. A. Habila, P. Rosaiah and J. Shim, SnO2 Quantum Dot-Decorated g-C3N4 Ultrathin Nanosheets: A Dual-Function Photocatalyst for Pollutant Degradation and Hydrogen Evolution, Catalysts, 2024, 14, 824 CrossRef CAS .
  77. H. Wu, et al., A facile one-step strategy to construct 0D/2D SnO2/g-C3N4 heterojunction photocatalyst for high-efficiency hydrogen production performance from water splitting, Int. J. Hydrogen Energy, 2020, 45, 30142–30152 CrossRef CAS .
  78. A. Seza, et al., Novel microwave-assisted synthesis of porous g-C3N4/SnO2 nanocomposite for solar water-splitting, Appl. Surf. Sci., 2018, 440, 153–161 CrossRef CAS .
  79. K. Zhu, et al., Facile fabrication of g-C3N4/SnO2 composites and ball milling treatment for enhanced photocatalytic performance, J. Alloys Compd., 2019, 802, 13–18 CrossRef CAS .
  80. J. K. Nørskov, et al., Trends in the Exchange Current for Hydrogen Evolution, J. Electrochem. Soc., 2005, 152, J23 CrossRef .
  81. N. T. Suen, et al., Electrocatalysis for the oxygen evolution reaction: recent development and future perspectives, Chem. Soc. Rev., 2017, 46, 337–365 RSC .
  82. Y. Zheng, et al., Hydrogen evolution by a metal-free electrocatalyst, Nat. Commun., 2014, 5, 3783 CrossRef .
  83. E. German, C. Pistonesi and V. Verdinelli, A DFT study of H2 adsorption on Pdn/SnO2 (110) surfaces (n = 1–10), Eur. Phys. J. B, 2019, 92, 98 CrossRef .
  84. R. B. Goncalves, Z. Chen, K. W. Chapman, R. Q. Snurr and J. T. Hupp, Experimental and theoretical investigation of hydrogen sorption by SnO2 nanostructures in a metal–organic framework scaffold, Mol. Phys., 2025, e2499204 CrossRef .

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra03721b

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.