Next Article in Journal
Inhibition of Urea Hydrolysis in Human Urine for Resource and Energy Recovery: Pharmaceuticals and Their Metabolites as Co-Existing Anticatalyzers
Previous Article in Journal
Evaluation of Aqueous Enzymatic Extraction for Obtaining Oil from Thevetia peruviana Seeds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Integration of g-C3N4 with SnS: Unlocking Enhanced Photocatalytic Efficiency and Electrochemical Stability for Dual-Functional Applications

by
Aya Ahmed
1,
Farid M. Abdel-Rahim
1,
Fatemah H. Alkallas
2,
Amira Ben Gouider Trabelsi
2,
Shoroog Alraddadi
3 and
Abdelaziz M. Aboraia
1,*
1
Physics Department, Faculty of Science, Al-Azhar University, Assiut Branch, Assiut 71524, Egypt
2
Department of Physics, College of Science, Princess Nourah Bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
3
Department of Physics, University College in AlJumum, Umm Al-Qura University, P.O. Box 715, Makkah 21955, Saudi Arabia
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(7), 629; https://doi.org/10.3390/catal15070629 (registering DOI)
Submission received: 23 May 2025 / Revised: 19 June 2025 / Accepted: 24 June 2025 / Published: 27 June 2025

Abstract

The synthesis of graphitic carbon nitride (g-C3N4) coupled with tin sulfide (SnS) has been identified as an effective method for improving the photocatalytic and electrochemical performance of SnS, a promising material for environmental and energy-related applications. In this study, we focused on how g-C3N4 influences the structural, optical, electrochemical, and functional properties of SnS. XRD and FTIR confirmed the formation of SnS/g-C3N4 heterostructure, while surface morphology analysis by SEM showed proper dispersion of SnS particles over g-C3N4 with a good interface contact. The SnS/g-C3N4 composite itself demonstrated improved photocatalytic performance, with the degradation rate of methylene blue reaching approximately 94% under visible light irradiation compared to the moderate activity of SnS. This enhancement can be credited to the successful charge carrier separation enabled by the type II heterojunction created between SnS and g-C3N4. Moreover, the composite presented improved electrochemical activity with a specific capacitance of 1340 F·g−1 at a scan rate of 10 A·g−1 and good cycling stability, where the capacitance was 92% after 5000 cycles. As such, these SnS/g-C3N4 composites suggest the specific application of this class of material in photocatalytic degradation as well as energy storage, putting forward new effective resolutions to environmental and energy issues.

Graphical Abstract

1. Introduction

The Industrial Revolution triggered rapid industrial expansion, which produced tremendous wastewater containing heavy metals and organic pollutants, threatening both environmental sustainability and the societal development of humans [1,2]. Semiconductor-based photocatalysts demonstrate great promise as efficient tools to solve these problems because they reduce heavy metals and degrade organic pollutants simultaneously, thus creating dual protection against industrial wastewater [3,4]. TiO2 functions as a semiconductor known for photocatalysis, which researchers have thoroughly studied for environmental pollutant degradation through its effective breakdown of organic and inorganic contaminants [5,6,7]. Practical use of this material faces two key barriers because the separation of electrons and holes remains slow, and its capacity to absorb visible light spectra is insufficient [8,9]. Contemporary photocatalysis research has established novel visible-light-responsive photocatalysts with high activity as its central focus for overcoming present challenges.
Semantic materials receive extensive research interest in renewable energy technology because they serve fundamental functions within photocatalytic and electrochemical processes [10,11]. Tin monosulfide (SnS) has proven itself as an outstanding candidate semiconductor material because of its narrow bandgap abilities, which match with solar-driven hydrogen production, environmental remediation, and energy storage demands [12,13]. The practical usage of SnS is restricted by fast charge carrier recombination, as well as restricted light absorption from solar radiation and unstable crystal structures when operating [14,15]. Research has attempted to address performance issues in SnS by combining it with relevant materials which improve its functional properties [16]. Scientists consider graphitic carbon nitride (g-C3N4) to be an exceptional metal-free polymer semiconductor because of its adaptable electronic structure, together with its chemical durability and visible-light activating capability [17,18]. g-C3N4 demonstrates excellent properties for simultaneous charge separation and expanded light absorption, which makes it suitable for cooperative application with SnS [19,20]. Some studies exist regarding the combined effect of SnS and g-C3N4 on their photocatalytic and electrochemical performance capabilities. Research examines the bonding patterns and physical transformations that occur between SnS elements and g-C3N4 networks in order to develop both energy conversion and energy storage applications.
Photocatalytic applications making use of SnS have received extensive study because of its earth-abundant chemical arrangement and suitable bandgap range of 1.3–1.7 eV [14,15,21]. The limiting factors for SnS performance are its weak electron–hole separation ability and its tendency to undergo photocorrosion [22]. Research has investigated the use of SnS as a lithium-ion battery anode material through electrochemistry, yet structural collapse during cycling, with poor electrical conductivity, stands as a significant challenge [23,24].
The extensive application of g-C3N4 in photocatalysis happens because it shows effective visible-light activity together with chemical resistance. The photophysical properties of charge carriers improved through the combination of ZnO or MoS2 with g-C3N4 appeared in recent research to boost photocatalytic activity [25,26,27]. The combination of g-C3N4 materials with other elements helps develop supercapacitor electrodes that possess better capacitance performance with improved cycling stability. Investigations mostly study metal oxides and transition metal dichalcogenides instead of metal sulfides, including SnS, although they show promise for research in this area. Some studies unite g-C3N4 with other sulfides (including CdS), yet they primarily optimize photocatalytic activity rather than electrochemical performance. No investigation evaluates how SnS/g-C3N4 combinations function as integrated solutions for light-driven and charging applications.
Previous research has not investigated the performance of this composite material for hydrogen evolution reactions or supercapacitive storage applications, yet these tests would enable its use in integrated energy systems. Design strategies remain limited by the lack of thorough investigations about how charges transfer between components under combined operational stress conditions involving light and electrochemical cycling. This research develops SnS/g-C3N4 composites through controlled in-situ synthesis, followed by optoelectronic characterization and tandem energy evaluation to build a complete framework for multiple-functional semiconductor development.

2. Results and Discussion

2.1. X-Ray Diffraction (XRD)

The XRD patterns of pristine SnS and SnS coated with different concentrations of g-C3N4 composites are presented in Figure 1a. The diffraction peaks of SnS correspond to the orthorhombic phase (JCPDS No. 14-0620), with prominent reflections observed at 2θ = ~32.03°, 37°, and 39.2°, assigned to the (111), (131), and (221) planes, respectively, as demonstrated in Figure 1a. For g-C3N4, a characteristic peak at 27.4° is evident, corresponding to the interlayer stacking of conjugated aromatic systems. In the SnS/g-C3N4 composite, both sets of peaks are present, indicating successful integration of g-C3N4 with SnS. Notably, a slight shift in the SnS peaks is observed, suggesting strong interfacial interaction between the two components, as shown in Figure 1b. The reduction in the crystallinity of SnS in the composite indicates that g-C3N4 suppresses particle growth, favoring the formation of nanoscale structures that enhance catalytic activity. The inverse relation between g-C3N4 dosage and SnS crystal size and strain behavior arises from the controlled nucleation process, together with interfacial strain dissipation mechanisms and defect reduction strategies, according to Table 1. The modified composite structures improve functional characteristics through g-C3N4’s multipurpose role in semiconductor systems.
The FTIR spectra of the samples reveal characteristic features of both SnS and g-C3N4. For pristine SnS, peaks in the range of 451–600 cm−1 correspond to the Sn–S stretching vibrations, as demonstrated in Figure 2a. In g-C3N4, the peaks at 1254 and 1633 cm−1 are associated with the stretching vibrations of C–N and C=N bonds, while the broad peak at ~3250 cm−1 corresponds to N–H stretching in terminal amine groups. In the SnS/g-C3N4 composite, these peaks are preserved with slight shifts, indicating successful coupling of the materials. The intensity reduction of the N–H peak suggests an interaction between the amine groups of g-C3N4 and the surface of SnS, which contributes to enhanced charge transfer at the interface, as shown in Figure 2b. In the composite materials, the XRD peaks attributed to g-C3N4 progressively intensify from pure SnS to 0.075 g-C3N4, indicating an increased incorporation of g-C3N4 within the matrix. This trend corroborates the earlier XRD analysis. For the SNS sample, a distinct yet moderate absorption peak is observed at 510 cm−1, which corresponds to the vibrational mode of the Sn–S bond.
The g-C3N4/SnS nanocomposite features were examined using field-emission scanning electron microscopy (FESEM). The microscopic evaluation of SnS demonstrated an irregular particulate formation with well-defined crystalline structures as depicted in Figure 3a. The heterostructure g-C3N4/SnS nanocomposites presented themselves through uniformly anchored irregular SnS nanoparticles measuring 60–120 nm in size across a g-C3N4 surface (Figure 3b–d). The g-C3N4/SnS nanocomposite features were investigated using field-emission scanning electron microscopy (FESEM). The microscopic evaluation of SnS demonstrated an irregular particulate formation with well-defined crystalline structures as depicted in Figure 3a. The heterostructure g-C3N4/SnS nanocomposites presented themselves through uniformly anchored irregular SnS nanoparticles measuring 60–120 nm in size across a g-C3N4 surface (Figure 3b–d). EDS analysis provided validation of the elemental substances that make up the nanocomposites. The marker signals in the EDS spectra (Figure 4a–f) proved the existence of Sn, S, C, and N elements, which matched the composition of g-C3N4/SnS hybrid materials.

2.2. Optical Properties

The bandgap energy of semiconductors was determined using the Tauc equation [28,29,30]:
α h v n = A h v E g
The absorption coefficient term α contains Planck’s constant h, while the photon frequency is represented by ν, and the bandgap energy appears as Eg in the relation alongside constants A and n. The bandgap energy for SnS and SnS coated with different concentrations of g-C3N4 was determined from Tauc plots showing (α)2 versus to have values of 2.24 eV and 2.13 eV (Figure 5a–d). Incorporating g-C3N4 into SnS leads to a decrease in the energy band gap relative to pure SnS. The absorption spectra of pure SnS and its SnS/g-C3N4 nanocomposites exhibit a noticeable redshift, demonstrating a reduction in their respective band gaps. Specifically, pure SnS has a band gap of 2.24 eV, while the SnS/g-C3N4 composite shows a narrowed band gap of 2.17 eV, and the SnS/0.075 g-C3N4 composite further decreases to 2.13 eV. This progressive reduction in band gap energy highlights that the integration of g-C3N4 effectively lowers the electronic barrier of the SnS matrix, enhancing its optical absorption properties.

2.3. Photocatalytic Performance

The photocatalytic performance of the materials was tested for the degradation of methylene blue (MB) under visible light. The pristine SnS had moderate degradation efficiency because of a high rate of recombination of the charge carriers. The g-C3N4 alone had low activity in terms of degradation due to its wide bandgap. Compared with SnS and g-C3N4 in the form of a physical mixture, the SnS/g-C3N4 composite exhibited higher photocatalytic activity and removed about 94% of MB within 90 min. This enhancement is due to the addition of a type II heterojunction, which increases the ability to separate photogenerated charges by a large margin. The photocatalytic mechanism of the g-C3N4/SnS composite depends on the transfer of electrons in the conduction band of g-C3N4 to SnS and holes in the valence band of SnS for oxidation processes. It is also said that more surface area and better light absorption are other advantages of the composite, as shown in Figure 6a–e. Among these, the SnS@0.075 g-C3N4 sample possessed higher photocatalytic capability, possibly because of the larger specific surface area.
All SnS coated with g-C3N4 composites exhibit significantly higher photodegradation performance compared to pristine SnS under visible light irradiation, indicating that the addition of a small amount of g-C3N4 enhances the photocatalytic activity of SnS. As the g-C3N4 content increases from 0 to 0.075 wt%, the MB degradation efficiency rises sharply, with the 0.075 g-C3N4 sample achieving the highest performance, degrading 94% of the MB solution within 90 min, as demonstrated in Figure 7b.
A linearized representation of MB photocatalytic degradation occurs with SnS@g-C3N4 photocatalysts, as shown in Figure 8a. The degradation process shows pseudo-first-order behavior because the plots remain linear throughout irradiation time. Analysis of Figure 8a data enabled the determination of apparent rate constants (k) for SnS@ g-C3N4 composites with different g-C3N4 contents presented in Table 2. The rate constant achieved by SnS@0.075 g-C3N4 reached 0.028 min−1, which indicates a 1.33-fold increase over the rate constant of pure SnS (0.021 min−1), as shown in Figure 9. When heterojunctions are built using proper g-C3N4 amounts, photocatalytic efficiency experiences a noteworthy boost. A photocatalyst’s ability to reuse constitutes a fundamental method for determining maintenance duration alongside implementation potential. The durability of SnS@0.075 g-C3N4 became measurable after engaging the catalyst in five subsequent degradation cycles while using visible light irradiation. The separated photocatalyst was reused in subsequent cycles without the addition of new catalyst materials. Under visible light illumination, the SnS@0.075 g-C3N4 photocatalyst displayed stable performance in five consecutive cycles, which led to a 94% degradation of MB in the last run. Prolonged exposure to light results in marginal performance loss of the catalyst because nanoparticle agglomeration occurs together with g-C3N4 structural damage from photon bombardment.

2.4. Electrochemical (Supercapacitor) Performance

The classical three-electrode system with 6 mol L−1 KOH as an electrolyte is used to probe the performance of SnS coated with x wt.% g-C3N4 (x = 0, 0.025, 0.05, and 0.075) as anode materials. As shown in Figure 10a–d, cyclic voltammetry (CV) curves for SnS coated with x wt.% g-C3N4 (x = 0, 0.25, 0.5 and 0.75) were measured at 10, 20, 30, 40, 50, 60, 70, 80, 90, and 100 mV·s−1. The current density and voltametric curve area increase proportionally with the scan rate. The anodic peak shifts in a positive potential direction while the cathodic peak shifts in a negative potential direction when the scan rate increases. The potential separation enables better efficiency in electronic charge transfer and improves electron transport kinetics. The material’s electrochemical stability, alongside its reversible behavior and enhanced capacitive properties, is reflected through the symmetric profiles shown in all curves.
The galvanostatic charge–discharge (GCD) measurements for SnS with different x wt.% g-C3N4 (x = 0, 0.025, 0.05, and 0.075) nanocomposite electrodes follow a current density of 10 A·g−1 to determine their anode capacitance level, as shown in Figure 11a–d. All electrodes exhibited dome-shaped curves while charging and discharging, thus proving that the electrode materials behave capacitively in agreement with CV curves. The charging/discharging tests of pristine SnS and SnS/g-C3N4 active electrodes took place across diverse densities from 10 A·g−1 to 20 A·g−1, as demonstrated in Figure 10a–d. When increasing the density of the electrode material, the charging/discharging time shortens, thereby decreasing the specific capacitance of the electrode material through interchanges between electrodes and ions. The capacitance of the anode material reached its maximum level at lower densities because cations could easily penetrate its surface.
Electrochemical impedance spectroscopy (EIS) was performed to analyze the charge transfer dynamics and electron diffusion at the electrode–electrolyte interface. Figure 12a presents Nyquist plots for pristine SnS2 and SnS2/g-C3N4 composites (with varying g-C3N4 loadings), measured across a frequency range of 10 mHz–100 kHz. The high-frequency intercept on the real axis, termed equivalent series resistance (Rs), reflects the combined resistance of the electrolyte, active material, and electrode–electrolyte contact. Notably, SnS2@0.05 g-C3N4 exhibited the lowest Rs among all samples, highlighting its potential for high-power-density applications. In the low-frequency region, the vertical Warburg impedance (W) correlates with ion diffusion kinetics. The steeper slopes of composite materials compared to pure SnS2 indicate reduced diffusion resistance and a diffusion-dominated charge transfer mechanism. Furthermore, the charge transfer resistance (Rct) decreased significantly in SnS2/g-C3N4 composites, confirming enhanced interfacial charge transport efficiency. Specific capacitance values for SnS2 and SnS2/g-C3N4 were derived from CV and GCD measurements. At a current density of 20 A·g−1, SnS2@0.075 g-C3N4 retained a capacitance of 1145 F·g−1, far exceeding the 315 F·g−1 of pure SnS2 (Figure 12b,c). Figure 12d compares energy and power densities, revealing a trade-off: energy density decreased slightly with increasing power density. The exceptional cycling stability of SnS2/g-C3N4—attributed to the heterostructure formed during composite synthesis—mitigates the volume fluctuations in SnS2 during charge–discharge cycles, thereby extending its operational lifespan.

3. Experimental Technique

3.1. Materials and Reagents

Tin(II) (99.999%, Sigma Aldrich, St Louis, MO, USA), sulfide (S), (99.999%, Sigma Aldrich, St Louis, MO, USA), thiourea (CH4N2S), (99.999%, Loba Co., Mumbai, India), graphitic carbon nitride (g-C3N4) powder, ethanol (C2H5OH), and distilled water were used in these experiments.

3.2. Procedure Synthesis

Tin(II) sulfide (SnS) was synthesized under controlled laboratory conditions using high-purity tin (Sigma Aldrich, St Louis, MO, USA) and sulfur (Sigma Aldrich, St Louis, MO, USA). The stoichiometric quantities of tin and sulfur were weighed, with approximately 2 g of the mixture loaded into a pre-cleaned quartz ampoule (70 mm length and 9 mm diameter). The ampoule was flame-sealed under a high vacuum of 1 × 10−5 Torr to ensure an oxygen-free environment. The sealed ampoule was placed in a custom-built horizontal rotary furnace and heated incrementally at a rate of 100 °C·h−1 to 950 °C. The temperature was maintained at 950 °C for 48 h with continuous rotation to achieve a homogeneous melt. Subsequently, the ampoule was cooled gradually to room temperature over 24 h [31]. The synthesized SnS was recovered by carefully breaking the quartz ampoule. Analytical-grade urea served as the precursor for synthesizing graphitic carbon nitride (g-C3N4) via a single-step pyrolysis approach. In a standard procedure, 10 g of urea was transferred into a covered crucible and subjected to thermal treatment at 550 °C for 3 h in a muffle furnace, with a ramp rate of 10 °C·min−1. Following the reaction, the furnace was then naturally cooled to ambient temperature. The resulting powder exhibited a pale yellow hue. This template-free, additive-free synthesis method is self-sustaining, eliminating the need for external agents or structural templates. To fabricate SnS coated with different concentrations of g-C3N4, specific amounts of g-C3N4 (0.025, 0.05, and 0.075 g) were admixed with 0.5g of SnS and 50 mL of ethanol via an ultrasonic probe. The dispersion was sonicated for 30 min to obtain a good dispersion of g-C3N4 nanosheets under constant stirring to form a homogeneous solution. Then, the prepared SnS solution was slowly poured onto the g-C3N4 dispersion while sonicating. A high-power ultrasonic homogenizer (power = 700 W) with a 0.6 mL horn was employed for 1 h to ensure that the SnS was uniformly deposited on the g-C3N4 surface. During the ultrasonic process, the temperature was maintained below 50 °C so that the materials did not aggregate or undergo thermal degradation. After the ultrasonic treatment, the suspension was allowed to settle and then the mixture was spun at 5000 rpm for 10 min. After precipitation, it was washed using ethanol and distilled water several times to eliminate other impurities and unreacted precursors. After the preparation of the SnS/g-C3N4 composite, the collected composite was dried in an oven at 60 °C for 12 h.

3.3. Characterization

The synthesized SnS/g-C3N4 composite can be characterized using techniques such as X-ray diffraction (XRD) for structural analysis, scanning electron microscopy (SEM) for morphology, FTIR, electrochemical performance “CS305”, and UV-Vis spectroscopy for optical properties. The ratio of SnS to g-C3N4 can be varied to optimize the performance of the composite for specific applications. Ultrasonic energy enhances the interaction between SnS and g-C3N4, ensuring uniform coating and strong interfacial bonding. This ultrasonic-assisted synthesis is a simple and efficient method to prepare SnS/g-C3N4 composites with enhanced photocatalytic and electrochemical properties.

3.4. Evaluation of Photocatalytic Activity in SnS Coated with g-C3N4

Photocatalytic tests were carried out in a specially designed photoreactor equipped with a 500 W tungsten halogen linear lamp emitting 9500 lumens of visible light. A dye solution was prepared in deionized water, and 0.025 g of either pure SnS or SnS coated with g-C3N4 photocatalysts were added to the mixture. To ensure uniform dispersion of the catalysts, the suspension underwent ultrasonication in the dark for 30 min. During irradiation, the reaction mixture was continuously stirred and aerated with ambient air. At defined time intervals, 5 mL aliquots were extracted, and the photocatalyst particles were removed via centrifugation. The degradation of methylene blue (MB) was tracked by measuring the absorbance at its maximum wavelength (λₘₐ) using a UV–Visible spectrophotometer. A decrease in absorbance at 665 nm correlated with the decolorization of the dye, indicating degradation efficiency. The removal efficiency of the dye through photocatalytic action was calculated using a standard formula for degradation percentage:
R e m o v a l   p r c e n t a g e = C 0 C t C 0 × 100
Following a comprehensive assessment of the photocatalytic degradation mechanism, the initial (C0) and residual (Ct) pollutant concentrations were precisely quantified. After catalyst recovery through centrifugation, the material was reintroduced into the dye solution for subsequent reuse testing.

3.5. Electrochemical Performance

The electrode materials were analyzed using a three-electrode system, where the working electrode comprised the active material, an Ag/AgCl electrode served as the reference electrode, and a platinum foil functioned as the counter electrode. For electrode fabrication, a uniform black paste was prepared by blending PVDF (10%), the active material (80%), and carbon black (10%). This paste was uniformly coated onto nickel foam substrates and subsequently dried in a vacuum oven at 60 °C for 10 h. Electrochemical characterization, including impedance spectroscopy (EIS), galvanostatic charge–discharge (GCD), and cyclic voltammetry (CV), was conducted using a Corrtest CS350 electrochemical workstation (manufactured in China) within a 6 mol L−1 KOH electrolyte environment.

4. Conclusions

The production of SnS/g-C3N4 composite materials for photocatalysts and supercapacitors occurred through a simple melt mixing method. The photocatalytic and capacitive properties of composite materials exceed those of pure SnS, especially when SnS@0.075 g-C3N4 functions as a photocatalyst and the SnS@0.05 g-C3N4 provides a specific capacitance of 1340 F·g−1 at 10 A·g−1 because of SnS/g-C3N4 synergistic action. The conductivity enhancement of the SnS/g-C3N4 composite results from the high-quality heterostructured system and the presence of g-C3N4 nitrogen, which promotes the composite material’s conductive properties. The g-C3N4 functions as both a conducting support to disperse SnS and prevents its aggregation, which normally causes cycle life degradation. The SnS/g-C3N4 composite reveals excellent potential for future energy-storage applications.

Author Contributions

Conceptualization, A.B.G.T., S.A. and A.M.A.; methodology, A.A., F.H.A. and A.M.A.; validation, F.M.A.-R., A.B.G.T. and S.A.; formal analysis, A.A., F.H.A. and S.A.; investigation, A.A., F.M.A.-R. and A.B.G.T.; resources, F.M.A.-R., F.H.A. and S.A.; data curation, F.M.A.-R.; writing—original draft, A.A., F.H.A., A.B.G.T. and A.M.A.; writing—review and editing, A.M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Princess Nourah bint Abdulrahman University Researchers Supporting Project (PNURSP2025R223), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Data Availability Statement

The data will be available on reasonable request.

Acknowledgments

The authors express their gratitude to Princess Nourah bint Abdulrahman University Researchers Supporting Project (PNURSP2025R223), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ahmed, J.; Thakur, A.; Goyal, A. Industrial wastewater and its toxic effects. In Chemistry in the Environment: Biological Treatment of Industrial Wastewater; Shah, M.P., Ed.; The Royal Society of Chemistry: London, UK, 2021; pp. 1–14. [Google Scholar]
  2. Singh, A. A review of wastewater irrigation: Environmental implications. Resour. Conserv. Recycl. 2021, 168, 105454. [Google Scholar] [CrossRef]
  3. Ethiraj, S.; Samuel, M.S.; Indumathi, S.M. A comprehensive review of the challenges and opportunities in microalgae-based wastewater treatment for eliminating organic, inorganic, and emerging pollutants. Biocatal. Agric. Biotechnol. 2024, 60, 103316. [Google Scholar] [CrossRef]
  4. Yasasve, M.; Manjusha, M.; Manojj, D.; Hariharan, N.; Preethi, P.S.; Asaithambi, P.; Karmegam, N.; Saravanan, M. Unravelling the emerging carcinogenic contaminants from industrial waste water for prospective remediation by electrocoagulation—A review. Chemosphere 2022, 307, 136017. [Google Scholar] [CrossRef]
  5. Mahlambi, M.M.; Ngila, C.J.; Mamba, B.B.; Zhang, X. Recent Developments in Environmental Photocatalytic Degradation of Organic Pollutants: The Case of Titanium Dioxide Nanoparticles—A Review. J. Nanomater. 2015, 2015, 790173. [Google Scholar] [CrossRef]
  6. Li, R.; Li, T.; Zhou, Q. Impact of Titanium Dioxide (TiO2) Modification on Its Application to Pollution Treatment—A Review. Catalysts 2020, 10, 804. [Google Scholar] [CrossRef]
  7. Jabbar, Z.H.; Graimed, B.H. Recent developments in industrial organic degradation via semiconductor heterojunctions and the parameters affecting the photocatalytic process: A review study. J. Water Process. Eng. 2022, 47, 102671. [Google Scholar] [CrossRef]
  8. Zhu, S.; Wang, D. Photocatalysis: Basic Principles, Diverse Forms of Implementations and Emerging Scientific Opportunities. Adv. Energy Mater. 2017, 7, 1700841. [Google Scholar] [CrossRef]
  9. Di, J.; Xiong, J.; Li, H.; Liu, Z. Ultrathin 2D photocatalysts: Electronic-structure tailoring, hybridization, and applications. Adv. Mater. 2018, 30, 1704548. [Google Scholar] [CrossRef]
  10. Batool, M.; Sanumi, O.; Jankovic, J. Application of artificial intelligence in the materials science, with a special focus on fuel cells and electrolyzers. Energy AI 2024, 18, 100424. [Google Scholar] [CrossRef]
  11. Mak, C.H.; Han, X.; Du, M.; Kai, J.-J.; Tsang, K.F.; Jia, G.; Cheng, K.-C.; Shen, H.-H.; Hsu, H.-Y. Heterogenization of homogeneous photocatalysts utilizing synthetic and natural support materials. J. Mater. Chem. A 2021, 9, 4454–4504. [Google Scholar] [CrossRef]
  12. Carminati, S.A.; Rodríguez-Gutiérrez, I.; de Morais, A.; da Silva, B.L.; Melo, M.A.; Souza, F.L.; Nogueira, A.F. Challenges and prospects about the graphene role in the design of photoelectrodes for sunlight-driven water splitting. RSC Adv. 2021, 11, 14374–14398. [Google Scholar] [CrossRef]
  13. Lah, N.A.C. Late transition metal nanocomplexes: Applications for renewable energy conversion and storage. Renew. Sustain. Energy Rev. 2021, 145, 111103. [Google Scholar] [CrossRef]
  14. Norton, K.J.; Alam, F.; Lewis, D.J. A Review of the Synthesis, Properties, and Applications of Bulk and Two-Dimensional Tin (II) Sulfide (SnS). Appl. Sci. 2021, 11, 2062. [Google Scholar] [CrossRef]
  15. Pejjai, B.; Reddy, V.R.M.; Gedi, S.; Park, C. Status review on earth-abundant and environmentally green Sn-X (X = Se, S) nanoparticle synthesis by solution methods for photovoltaic applications. Int. J. Hydrogen Energy 2017, 42, 2790–2831. [Google Scholar] [CrossRef]
  16. Zhou, T.; Pang, W.K.; Zhang, C.; Yang, J.; Chen, Z.; Liu, H.K.; Guo, Z. Enhanced Sodium-Ion Battery Performance by Structural Phase Transition from Two-Dimensional Hexagonal-SnS2 to Orthorhombic-SnS. ACS Nano 2014, 8, 8323–8333. [Google Scholar] [CrossRef] [PubMed]
  17. Ong, W.-J.; Tan, L.-L.; Ng, Y.H.; Yong, S.-T.; Chai, S.-P. Graphitic carbon nitride (g-C3N4)-based photocatalysts for artificial photosynthesis and environmental remediation: Are we a step closer to achieving sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  18. Bhanderi, D.; Lakhani, P.; Modi, C.K. Graphitic carbon nitride (g-C3N4) as an emerging photocatalyst for sustainable environmental applications: A comprehensive review. RSC Sustain. 2023, 2, 265–287. [Google Scholar] [CrossRef]
  19. Ai, L.; Shi, R.; Yang, J.; Zhang, K.; Zhang, T.; Lu, S. Efficient combination of g-C3N4 and CDs for enhanced photocatalytic performance: A review of synthesis, strategies, and applications. Small 2021, 17, 2007523. [Google Scholar] [CrossRef]
  20. Wen, Y.; Sun, Y.; Liu, Y.; Zhao, M.; Zhao, L. Green synthesis of 2% g-C3N4/SnS2-V3/CQD1 composite photocatalyst from waste plant soot for efficient U(VI) removal: Mechanistic insights. Chem. Eng. J. 2024, 494, 153247. [Google Scholar] [CrossRef]
  21. Hegde, S.; Bose, R.S.C.; Surendra, B.; Vinoth, S.; Murahari, P.; Ramesh, K. SnS-Nanocatalyst: Malachite green degradation and electrochemical sensor studies. Mater. Sci. Eng. B 2022, 283, 115818. [Google Scholar] [CrossRef]
  22. Xie, X.; Ge, P.; Xue, R.; Lv, H.; Xue, W.; Liu, E. Enhanced photocatalytic H2 evolution and anti-photocorrosion of sulfide photocatalyst by improving surface reaction: A review. Int. J. Hydrogen Energy 2023, 48, 24264–24284. [Google Scholar] [CrossRef]
  23. Ying, H.; Han, W.Q. Metallic Sn-based anode materials: Application in high-performance lithium-ion and sodium-ion batteries. Adv. Sci. 2017, 4, 1700298. [Google Scholar] [CrossRef]
  24. Mei, S.; An, W.; Fu, J.; Guo, W.; Feng, X.; Li, X.; Gao, B.; Zhang, X.; Huo, K.; Chu, P.K. Hierarchical micro-flowers self-assembled from SnS monolayers and nitrogen-doped graphene lamellar nanosheets as advanced anode for lithium-ion battery. Electrochimica Acta 2020, 331, 135292. [Google Scholar] [CrossRef]
  25. Jo, W.-K.; Lee, J.Y.; Selvam, N.C.S. Synthesis of MoS2 nanosheets loaded ZnO–g-C3N4 nanocomposites for enhanced photocatalytic applications. Chem. Eng. J. 2016, 289, 306–318. [Google Scholar] [CrossRef]
  26. Nemiwal, M.; Zhang, T.C.; Kumar, D. Recent progress in g-C3N4, TiO2 and ZnO based photocatalysts for dye degradation: Strategies to improve photocatalytic activity. Sci. Total. Environ. 2021, 767, 144896. [Google Scholar] [CrossRef]
  27. Kumar, S.G.; Kavitha, R.; Manjunatha, C. Review and perspective on rational design and interface engineering of g-C3N4/ZnO: From type-II to step-scheme heterojunctions for photocatalytic applications. Energy Fuels 2023, 37, 14421–14472. [Google Scholar] [CrossRef]
  28. Aboraia, A.M.; Al-Omoush, M.; Solayman, M.; Saad, H.M.H.; Khabiri, G.; Saad, M.; Alsulaim, G.M.; Soldatov, A.V.; Ismail, Y.A.M.; Gomaa, H. A heterostructural MoS2QDs@UiO-66 nanocomposite for the highly efficient photocatalytic degradation of methylene blue under visible light and simulated sunlight. RSC Adv. 2023, 13, 34598–34609. [Google Scholar] [CrossRef]
  29. El-Maaref, A.A.; Hasaneen, M.F.; Alraddadi, S.; Ismail, Y.A.M.; Aboraia, A.M. The effect of rGO on the enhancement of photocatalytic activity of the CdS nanorods. J. Mater. Sci. Mater. Electron. 2024, 35, 2289. [Google Scholar] [CrossRef]
  30. Aboraia, A.M.; Almohammedi, A.; Alraddadi, S.; Taha, S.A.; Saad, M.; Sharaf, I.; Ismail, Y.A.M. Comparison study between as-synthesized ZnO and ZnO derived from ZiF-8 metalorganic framework in removing methylene blue. Mod. Phys. Lett. B 2024, 38, 2450221. [Google Scholar] [CrossRef]
  31. Kumar, G.G.; Reddy, K.; Nahm, K.S.; Angulakshmi, N.; Stephan, A.M. Synthesis and electrochemical properties of SnS as possible anode material for lithium batteries. J. Phys. Chem. Solids 2012, 73, 1187–1190. [Google Scholar] [CrossRef]
Figure 1. (a) The XRD patterns of SnS coated with different concentrations of g-C3N4. (b) Zoom 2 theta of SnS coated with different concentrations of g-C3N4 at plane (1 1 1).
Figure 1. (a) The XRD patterns of SnS coated with different concentrations of g-C3N4. (b) Zoom 2 theta of SnS coated with different concentrations of g-C3N4 at plane (1 1 1).
Catalysts 15 00629 g001
Figure 2. (a) The FTIR spectrum of pristine SnS and (b) the spectra of pure SnS, SnS coated with different concentrations of g-C3N4, and g-C3N4.
Figure 2. (a) The FTIR spectrum of pristine SnS and (b) the spectra of pure SnS, SnS coated with different concentrations of g-C3N4, and g-C3N4.
Catalysts 15 00629 g002
Figure 3. FSEM photos of (a) pure SnS, (b) SnS@0.025wt.% g-C3N4, (c) SnS@0.05wt.% g-C3N4, and (d) SnS@0.075wt.% g-C3N4.
Figure 3. FSEM photos of (a) pure SnS, (b) SnS@0.025wt.% g-C3N4, (c) SnS@0.05wt.% g-C3N4, and (d) SnS@0.075wt.% g-C3N4.
Catalysts 15 00629 g003
Figure 4. (a) FSEM SnS@g-C3N4, (b) full elemental mapping of SnS@g-C3N4, (c) carbon element, (d) nitrogen, (e) sulfur, and (f) tin.
Figure 4. (a) FSEM SnS@g-C3N4, (b) full elemental mapping of SnS@g-C3N4, (c) carbon element, (d) nitrogen, (e) sulfur, and (f) tin.
Catalysts 15 00629 g004
Figure 5. Optical energy of (a) pure SnS, (b) SnS@0.025 g-C3N4, (c) SnS@0.05 g-C3N4, and (d) SnS@0.075 g-C3N4.
Figure 5. Optical energy of (a) pure SnS, (b) SnS@0.025 g-C3N4, (c) SnS@0.05 g-C3N4, and (d) SnS@0.075 g-C3N4.
Catalysts 15 00629 g005
Figure 6. UV-Visible spectrum of MB for photocatalyst (a) SnS, (b) g-C3N4, (c) SnS@0.025wt% g-C3N4, (d) SnS@0.05wt% g-C3N4, and (e) SnS@0.075wt% g-C3N4.
Figure 6. UV-Visible spectrum of MB for photocatalyst (a) SnS, (b) g-C3N4, (c) SnS@0.025wt% g-C3N4, (d) SnS@0.05wt% g-C3N4, and (e) SnS@0.075wt% g-C3N4.
Catalysts 15 00629 g006
Figure 7. (a) C/C0 and (b) efficiency of pure SnS and SnS covered with different concentrations of g-C3N4.
Figure 7. (a) C/C0 and (b) efficiency of pure SnS and SnS covered with different concentrations of g-C3N4.
Catalysts 15 00629 g007
Figure 8. (a) Linearized kinetic plots of methylene blue (MB) degradation, expressed as (−ln (Ct/C0)) versus time (t). (b) Cycling stability tests of the photocatalyst for MB degradation under visible light over five consecutive cycles.
Figure 8. (a) Linearized kinetic plots of methylene blue (MB) degradation, expressed as (−ln (Ct/C0)) versus time (t). (b) Cycling stability tests of the photocatalyst for MB degradation under visible light over five consecutive cycles.
Catalysts 15 00629 g008
Figure 9. The degradation kinetics of MB for pure SnS and coated with different amounts of g-C3N4.
Figure 9. The degradation kinetics of MB for pure SnS and coated with different amounts of g-C3N4.
Catalysts 15 00629 g009
Figure 10. CV of (a) pure SnS, (b) SnS@0.025 g-C3N4, (c) SnS@0.05 g-C3N4, and (d) SnS@0.075 g-C3N4.
Figure 10. CV of (a) pure SnS, (b) SnS@0.025 g-C3N4, (c) SnS@0.05 g-C3N4, and (d) SnS@0.075 g-C3N4.
Catalysts 15 00629 g010
Figure 11. The GCD of (a) pure SnS, (b) SnS@0.025 g-C3N4, (c) SnS@0.05 g-C3N4, and (d) SnS@0.075 g-C3N4.
Figure 11. The GCD of (a) pure SnS, (b) SnS@0.025 g-C3N4, (c) SnS@0.05 g-C3N4, and (d) SnS@0.075 g-C3N4.
Catalysts 15 00629 g011
Figure 12. EIS of the pure SnS and SnS covered with g-C3N4 composites (a). The specific capacitances of the pure SnS and SnS covered with g-C3N4 composites calculated by CV (b). GCD curves at different current densities (c). The Ragone plots (d).
Figure 12. EIS of the pure SnS and SnS covered with g-C3N4 composites (a). The specific capacitances of the pure SnS and SnS covered with g-C3N4 composites calculated by CV (b). GCD curves at different current densities (c). The Ragone plots (d).
Catalysts 15 00629 g012
Table 1. The calculated crystalline size and strain of SnS coated with different concentrations of g-C3N4.
Table 1. The calculated crystalline size and strain of SnS coated with different concentrations of g-C3N4.
SampleCrystalline Size (d)Strain (ε) × 10−3Dislocation Density (δ) × 10−4
SnS44.4051.55255.07
SnS@0.025wt% g-C3N446.691.15754.58
SnS@0.05wt% g-C3N437.50.82757.11
SnS@0.075wt% g-C3N428.720.497512.1
Table 2. The efficiency and constant (K rate) of pure SnS and coated with different amounts of g-C3N4.
Table 2. The efficiency and constant (K rate) of pure SnS and coated with different amounts of g-C3N4.
SampleEfficiencyK, min−1R2
SnS87.5%0.0210.98
SnS@0.025 g-C3N488.1%0.0240.97
SnS@0.05 g-C3N492.5%0.0260.92
SnS@0.075 g-C3N494%0.0280.93
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

Ahmed, A.; Abdel-Rahim, F.M.; Alkallas, F.H.; Trabelsi, A.B.G.; Alraddadi, S.; Aboraia, A.M. Synergistic Integration of g-C3N4 with SnS: Unlocking Enhanced Photocatalytic Efficiency and Electrochemical Stability for Dual-Functional Applications. Catalysts 2025, 15, 629. https://doi.org/10.3390/catal15070629

AMA Style

Ahmed A, Abdel-Rahim FM, Alkallas FH, Trabelsi ABG, Alraddadi S, Aboraia AM. Synergistic Integration of g-C3N4 with SnS: Unlocking Enhanced Photocatalytic Efficiency and Electrochemical Stability for Dual-Functional Applications. Catalysts. 2025; 15(7):629. https://doi.org/10.3390/catal15070629

Chicago/Turabian Style

Ahmed, Aya, Farid M. Abdel-Rahim, Fatemah H. Alkallas, Amira Ben Gouider Trabelsi, Shoroog Alraddadi, and Abdelaziz M. Aboraia. 2025. "Synergistic Integration of g-C3N4 with SnS: Unlocking Enhanced Photocatalytic Efficiency and Electrochemical Stability for Dual-Functional Applications" Catalysts 15, no. 7: 629. https://doi.org/10.3390/catal15070629

APA Style

Ahmed, A., Abdel-Rahim, F. M., Alkallas, F. H., Trabelsi, A. B. G., Alraddadi, S., & Aboraia, A. M. (2025). Synergistic Integration of g-C3N4 with SnS: Unlocking Enhanced Photocatalytic Efficiency and Electrochemical Stability for Dual-Functional Applications. Catalysts, 15(7), 629. https://doi.org/10.3390/catal15070629

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