Next Article in Journal
Aerosol-Deposited 8YSZ Coating for Thermal Shielding of 3YSZ/CNT Composites
Previous Article in Journal
Evaluation of Magnetron Sputtered TiAlSiN-Based Thin Films as Protective Coatings for Tool Steel Surfaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Decorating TiO2 Nanoparticle Thin Film with SnSx (x < 1): Preparation, Characterization, and Photocatalytic Activity

1
Department of Biology, Ecology, and Earth Science, University of Calabria, 87036 Arcavacata di Rende, CS, Italy
2
Department of Physics, University of Calabria, 87036 Arcavacata di Rende, CS, Italy
*
Author to whom correspondence should be addressed.
Coatings 2024, 14(9), 1185; https://doi.org/10.3390/coatings14091185
Submission received: 5 August 2024 / Revised: 8 September 2024 / Accepted: 10 September 2024 / Published: 12 September 2024

Abstract

:
We report a study on the SnSx (x < 1) decoration of porous TiO2 nanoparticle thin films using the ionic layer adsorption and reaction (ILAR) method. UV-vis absorption measurements revealed a direct bandgap of 1.40–2.10 eV for SnSx (with x = 0.85) and 3.15 eV for TiO2. Degradation of rhodamine B molecules in aqueous solutions shows that coating with a Sn-to-Ti molar ratio of 2% improves the efficiency of the photocatalytic performance of titanium dioxide, but excessive coverage decreases it. We interpret the observed behavior as due to a delicate balance of many competing factors. The formation of intimate interfaces guaranteed by the ILAR growth technique and a nearly optimal alignment of conduction band edges facilitate electron transfer, reducing electron–hole recombination rates. However, the valence hole transfer from TiO2 to SnS reduces the oxidative potential, which is crucial in the degradation mechanism.

1. Introduction

Water is utilized in innumerable chemical processes in biochemical, pharmaceutical, petrochemical, and other industries. Degrading and decomposing various organic and inorganic contaminants in the produced wastewater is paramount for human health and environmental concerns. As a sustainable and unselective process, heterogeneous photocatalysis employing solar light as a photon source has gained increasing attention in recent decades [1,2,3,4]. Titanium dioxide is one of the most widely used photocatalysts. Upon absorption of ultraviolet (UV) light with photon energy larger than the band gap value (3.2 eV), electron–hole pairs are generated. If they survive recombination, they can migrate to the surface and produce hydroxyl radicals and superoxide anions through redox reactions involving water and oxygen molecules. These oxidative and reductive species can decompose and mineralize various pollutants, such as residuals of medicines, pesticides, herbicides, dyes, phenols, hydrocarbons, and heavy metals, into CO2, H2O, and other simpler and harmless compounds [1,2,4,5].
Extensive research has focused on broadening solar light absorption range and diminishing electron–hole recombination rate to improve the photocatalytic efficacy of TiO2. Doping TiO2 with foreign species can reduce the bandgap, while creating appropriate nano-heterojunctions can facilitate charge transfer [2]. In this context, SnS is one of the best choices among IV–VI semiconductors. It possesses a bandgap of only 1.1–2.1 eV and an electron affinity similar to TiO2 [6].
Synthesis of SnS nanoparticles can employ different methods, such as precipitation, solvothermal, microemulsion, and (successive) ionic layer adsorption and reaction ((S)ILAR) [7,8,9,10]. The most common and straightforward route to producing thin films is wet chemistry [11,12]. A supported thin film is necessary for practical applications of SnS@TiO2 nanoparticles as photocatalysts. One effective method for forming such nano-heterojunctions with intimate contact involves coating a porous TiO2 nanoparticle (NP) thin film with Sn2+ and S2− precursor solutions using the SILAR method, followed by post-process annealing [13,14,15].
In this study, we report on the fabrication and characterization of SnSx-decorated TiO2 NP thin films and their photocatalytic activity toward rhodamine B (RhB) molecules, taken as a representative organic material. We demonstrate that, by selecting appropriate S2− and Sn2+ precursor solution concentrations, it is possible to improve the photocatalytic performance of TiO2.

2. Materials and Methods

TiO2 nanoparticles (P25) were purchased from Evonik (Essen, Germany). All chemicals, including polyethylene glycol (PEG-20000, Merck, Taufkirchen, Germany), Triton X-100 (Union Carbide Company, Houston, TX, USA), acetylacetone (>99%, Sigma-Aldrich/Merck, St. Louis. MI, USA), anhydrous stannous chloride (SnCl2, 98%, Alfa Aesar/Thermo Scientific, Waltham, MA, USA), sodium sulfide nonahydrate (Na2S·9H2O, Sigma-Aldrich/Merck), and rhodamine B (Thermo Scientific), were used as received without further purification.
TiO2 sol (5 g in 25 mL distilled water), added with PEG-20000, acetylacetone, and Triton X-100, was sonicated for an extended period to break up large aggregates. Spin coating the sol onto a 0.15μm fluorine-doped tin oxide (FTO)-covered glass substrate (2.5 × 2.5 cm2) and subsequent calcination at 450 °C for 1 h yield a uniform thin layer. Further details can be found elsewhere [16]. Repeating the procedure several times could produce films with varying desired thicknesses. TiO2 layers, with an average coverage of 2.2 ± 0.2 mg cm−2, were routinely grown.
The coating of TiO2 layers with SnSx was achieved using the ILAR method. In a typical reaction scheme, the sample was first immersed in the SnCl2 aqueous solution with a molar concentration of [Sn2+] = Snμμμ (in mM), then rinsed and submerged into distilled water to remove unbounded Sn2+ ions, immersed in the Na2S solution with a concentration of [S2−] = Sννν (in mM), and again washed. The immersion time was typically 30 s for each, unless otherwise specified, and the washing lasted 3 min. The so-prepared samples follow the SnμμμSννν nomination scheme. Both precursor solutions were magnetically stirred and kept at 75 °C. This adsorption and reaction process could be reiterated for several cycles. The samples were immediately heated at 180 °C in an oven with a nitrogen flow of 800 mL/min for 90 min. The SnSx@TiO2 films were brownish, but the hue depended on the amount and the composition. For heavily coated films, the color was nearly black.
Scanning electron microscopy (SEM) was performed with an ultra-high-resolution ZEISS Cross-Beam 350 (Zeiss, Oberkochen, Germany). For energy dispersive X-ray analysis (EDX), a beam of 15 keV was used to analyze a large surface area of 1.5 × 2.0 mm2. X-ray diffraction (XRD) was employed to investigate the crystalline structure of the NP thin films using a Bruker D8 Advance diffractometer with Cu Kα radiation (λ = 1.5406 ) (Bruker, Billerica, MA, USA).
X-ray photoelectron spectroscopy (XPS) was performed in a UHV chamber with a base pressure of 1 × 10−9 Torr utilizing a hemispherical electron energy analyzer with a constant pass energy of 100 eV. The excitation source was the Cu Kα emission. We calibrated the binding energy (BE) scale by setting graphite’s C 1 s peak to 284.3 eV and corrected the spectra for the analyzer transmission factor.
The absolute total UV-Vis-NIR (200–1250 nm) diffusive reflectance and transmittance were measured with a Cary-5E hemispherical spectrophotometer. According to the Tauc method, it is possible to determine the band gap energy Eg of a semiconductor with the following equation:
( A h ν ) β = B ( h ν E g )
with A being the absorbance, B a constant, the photon energy, and β = 2 for direct band transitions.
We assessed the photocatalytic performance of the SnSx@TiO2 NP thin films by monitoring the change in concentration of RhB in an aqueous solution. The reaction occurred in a 600 mL glass beaker with an 8.2 cm inner diameter. The reaction cell contained a magnetically stirred 50 mL RhB solution with a concentration of 1 × 10−5 M and a pH of 8.5. A 300 W Xe lamp served as the source of photons, and the light, diaphragmed and incident through a JS1-type UV-transparent quartz cover, irradiated only the sample surface (ϕ = 2 cm). The radiance flux density was 21 mW cm−2, measured with a power meter, and the optical path was 7 mm.
To determine the concentration of RhB using the Beer–Lambert method, an aliquot of 1.5 mL of the solution was withdrawn at specific time intervals to measure the absorbance spectrum with a UV-vis spectrometer. The transmittance signal in the 553–555 nm range was integrated, with water as a reference. After the analysis in less than 1 min, the sampled solution was promptly returned to the reaction cell.

3. Results

Figure 1 illustrates SEM images of a pure TiO2 thin film (A and C) and one coated with SnSx Sn100S600 (B and D). The TiO2 film was uniform and porous. PEG-20000 contained in the sol evaporated during the sintering, leaving large pores in the film. The particle size ranged from approximately 14 nm to 30 nm, though many particles formed aggregates. The coating of SnSx using the ILAR method did not make any significant morphology changes except for a slight increase in the NP size due to the intimate growth of tiny SnS on TiO2.
The cross-sectional SEM images, as a typical one shown in Figure 2, revealed an average film thickness of about 11 μm. The TiO2 layer mean coverage was 2.2 ± 0.2 mg cm−2. By using the density of single crystalline TiO2 of 4.23 g cm−3, we could easily estimate the thickness of a compact film as 2.2 mg cm−2/4.23 g cm−3 = 5.3 × 10−4 cm = 5.3 μm. Consequently, the film porosity was roughly 50%.
As illustrated in Figure 3, EDX detected no Sn and Si signals for pure TiO2, sample Sn000S000, indicating that the film was thick enough to shield the underlying FTO-glazed glass substrate. For sample Sn010S060, the measured atomic ratio between Sn and Ti, χ, was 0.03, whereas that between S and Sn, x, was 0.85. For Sn100S600, χ increased to 0.08 but x remained the same.
We explored the impact of the concentration ratio of precursor solutions on the characteristics of SnSx NPs, keeping the [Sn2+] constant (10 mM) and varying [S2−] from 10 to 100 mM. The results, summarized in Table 1, show that as the ratio α = [S2−]/[Sn2+] increased from 1 to 6, more S2− ions reacted to the adsorbed Sn2+, resulting in a rise of x from 0.20 to 0.85. The lack of perfectly stochiometric SnS nanoparticle formation aligns with the findings of Nengzi et al., likely due to the oxidation of some Sn2+ ions either during the formation process or during the heating stage [13]. We observed a similar trend when extending the immersion time in the S2− precursor solution, maintaining the same molar concentrations [S2−] = [Sn2+] = 10 mM. However, the maximum value of x was only 0.75. In the following, we focus on the cases of α = 6 for several [Sn2+] and the same immersion time (30 s), for which x remained constant at 0.85 ± 0.05. These SnSx NPs could be considered either SnO-doped, SnO2-doped, or mixed.
Figure 4 displays the XRD patterns of various SnSx@TiO2 films. The bottom-most curve is of P25 TiO2 thin film annealed at 450 °C for 1 h. It mainly contains anatase polymorph, about 10% of the rutile phase, and the SnO2 of the underlying FTO. Decoration with Sn020S120 did not yield discernible SnS structures (curve b). However, in the case of Sn100S600, features in 2θ between 30 and 32° became visible (curve c). Heavy SnSx decoration using Sn100S600 (5 cycles) resulted in two well-distinct peaks at about 2θ = 30.8° and 31.8°, identified as SnS (curve d).
In Figure 5, we present a set of XPS spectra of Ti 2p, O 1s, Sn 3d, and S 2p levels for pristine TiO2 (Sn000S000) and SnSx-decorated TiO2 (Sn100S600) films. The line shape of the Ti 2p spectrum did not show any appreciable changes upon the decoration of SnSx. The same was true also for O 1s. The Sn 3d5/2 and 3d3/2 peak energies were 486.5 eV and 495 eV, respectively, and the S 2p BE was 261.2 eV. These XPS spectra confirm the formation of ZnS [13,15,17,18].
The upper panel of Figure 6 shows a series of absorbance curves, with α = 6, obtained using UV-vis total diffusive reflection and transmission spectroscopy for various NP thin films. In the visible range, absorption gradually increased as more SnSx NPs coated on TiO2, broadening the absorption spectrum. The lower panel displays the Tauc plot of (Ahν)2 as a function of the photon energy for determining the direct bandgap energies Eg of the SnSx@TiO2 coating layers. The curve for a pure TiO2 film yielded Eg(TiO2) = 3.15 eV, which agrees with previous reports. It was still possible to infer this bandgap value even after SnSx decoration. The absorption of the FTO-covered glass substrate was null for < 4 eV. All Tauc curves have been vertically offset for clarity. No absorption from SnSx was discernible for Sn001S006. With increasing precursor concentration, Eg(SnSx) decreased from 2.10 to 1.40 eV (see Table 2).
We investigated the effects of decorating TiO2 with SnSx on the photocatalytic activity of thin films. The upper panel of Figure 7 depicts the time evolution of the RhB absorption for sample Sn005S030. The middle panel illustrates the evolution of RhB concentration change over time C(t)/C(0) under Xe lamp illumination for α = 6 (x = 0.85). The figure also shows the direct photolysis curve for a bare glass sample. At low coverages, the SnSx coating seemed to speed up the degradation process compared to the pure TiO2 film. However, at higher concentrations, there was a decrease in photocatalytic efficiency. It is important to note that the concentration change follows zero-order kinetics. The linear fitting of C(t)/C(0) − t yielded the RhB degradation rate constant κ. In the lower panel of Figure 7, we plot κ versus the S-to-Sn ratio χ, whose highest value corresponded to x = 2%.

4. Discussion

The XRD, EDX, XPS, and UV-vis absorption results all confirm that the synthesized thin films using the ILAR method were of SnSx@TiO2 NPs. The large Eg(SnSx) values observed at small χ can be related to the quantum size effect of the SnSx NPs [6,19]. The continuous evolution of the bandgap energies reported in Table 2 supports this interpretation.
It is well known that a homogeneous RhB solution exhibits considerably high stability under visible light excitation [20]. Photolysis occurs only under UV irradiation but with a negligible quantum efficiency [20]. This explains why the direct photolysis of this dye is a highly inefficient process (see Figure 7).
We roughly estimate the conduction band minimum (CBM) and the valence band maximum (VBM) of SnS by using the following empirical formula:
CBM(SnS) = X(SnS) − Eg(SnS)/2 + E0
VBM(SnS) = CBM(SnS) + Eg(SnS)
The Mulliken electronegativity, X(SnS) = 5.17 eV, is estimated as the geometric mean of the Sn and S electronegativities, with values of 4.3 and 6.22 eV, respectively. E0 = −4.44 eV is the vacuum level relative to the normal hydrogen electrode (NHE). With a bandgap energy of Eg(SnS) = 2.00 eV for SnS, the CBM(SnS) and VBM(SnS) potentials are approximately −0.27 V and 1.73 V, respectively, on the NHE scale. When the bandgap energy is reduced to 1.40 eV, the potentials shift to 0.03 V and 1.43 V, agreeing with previously reported values [10,21]. Anatase TiO2 has a CBM around −3.95 V relative to the vacuum level or −0.50 V relative to the NHE [22].
Figure 8 schematically depicts the potential diagram of the VBM and CBM of both TiO2 and SnS, the ground state and the first excited state of rhodamine B, *OH radical, as well as the relevant redox reactions. It is worthwhile mentioning that the bandgap energy of SnS reported in the literature varies for several tenths of eV, depending on the fabrication procedure and other factors [6].
In the presence of a semiconductor catalyst, the photo-induced degradation process can start through two primary reaction pathways: photoexcitation of either adsorbed dye molecules or semiconductor photocatalyst. When a RhB molecule on the surface absorbs light, the excited electron may recombine back to the ground state. Alternatively, it may transfer to an adjacent adsorbed oxygen molecule, forming an *O2 superoxide ion, or to the conduction band of the semiconductor catalyst. Due to the proximity of the *O2, the excited electron has a high probability of transferring back to the (RhB)+ ion, leading to the regeneration of the dye (O2 sensitization) [23]. If the charge injected in the conduction band diffuses away, then (RhB)+ can initiate the N-deethylation and subsequent decomposition [23]. It is important to note that the excitation of RhB is a resonance process with the maximum cross-section being at 564 nm so that only light with a wavelength in a very narrow range can be absorbed. The fact that the photodegradation of RhB by TiO2 under visible light illumination is negligible indicates that this mechanism is not operative for TiO2.
In photocatalysis, when a photon with energy greater than the band gap is absorbed, an electron in the valence band is excited into the conduction band. If the photogenerated charges survive the recombination, the electron can migrate to the surface and diffuse in the semiconductor matrix, reducing an adsorbed oxygen molecule to a superoxide anion. Meanwhile, the hole left in the valence band can either directly oxidize a dye molecule or react with an adsorbed OH or H2O molecule, generating a strong oxidant hydroxyl radical *OH [5]. *O2 and *OH are fundamental components in the degradation and decomposition mechanisms of nearly all organic and inorganic pollutant molecules [1,2,5]. The specific pathways and mechanisms for photodegradation and decomposition of RhB have been well described in the literature [20,23,24].
SnS has a bandgap in the visible range, allowing it to absorb a significant portion of the solar light spectrum. Its photocatalytic activity has been extensively investigated for various dyes and pollutant molecules [13,14,15,17]. The efficacy of photocatalysis depends on several critical factors, including the actual band gap energy, the positions of band edges relative to the relevant redox potentials, the rate of electron–hole recombination, and the energy levels and degradation pathways specific to the molecules involved. Additionally, the performance is influenced by the crystalline structure, morphology, loading amount, particle size, and specific surface area of the photocatalyst nanoparticles, as well as factors like pollutant concentration, solution pH, and the spectrum and intensity of the light source. Comparing its performance directly with that of a wide bandgap semiconductor like TiO2, even under very similar experimental conditions, may provide limited insights due to the significant differences in their photocatalytic mechanisms and properties.
Zhang et al. synthesized crystalline SnS nanoparticles using an anaerobic reaction method and found limited photoactivity toward RhB dye under visible light [25]. Mittal et al. fabricated TiO2 and SnS nanoparticles with the precipitation technique. They observed that, in the presence of either TiO2 or SnS suspended nanoparticles, RhB concentration follows pseudo-first-order kinetics and that TiO2 exhibited higher photocatalytic activity under Xe UV-vis light compared to SnS [17]. In Figure 7, we observe similar trends in the degradation behavior of RhB upon Xe light illumination. We also notice differences in the reaction kinetic order compared to Mittal et al.’s study.
As indicated in the potential diagram of Figure 8, the VB edge of TiO2 lies deeper than SnS. Therefore, the oxidizing power of a valence hole for RhB is stronger in TiO2 than in SnS. Likewise, a VB hole in TiO2 can react with an adsorbed water molecule, producing a hydroxyl radical, but this reaction route is not allowed in SnS. It is worthwhile mentioning that in our case, photocatalysis occurs only for the dye molecules adsorbed on the top few layers of the thin film. The reaction is active adsorption site limited, following pseudo-zero-order kinetics, and requires a longer irradiation time to reduce the concentration of RhB.
TiO2 and SnS have CB edge potentials quite close to each other. The formation of nano-heterojunction with intimate contacts facilitates electron transfer across the interface from one semiconductor to another, effectively reducing the electron–hole recombination rate. However, hole transfer from the VB of TiO2 to that of SnS has dual effects: it enhances charge survival but reduces oxidizing potential. The balance of these two competing effects depends on the details of the band alignment.
We attribute the enhancement of the photocatalytic activity of a small amount of SnS-decorated TiO2 mainly to reduced electron–hole recombination rates in both semiconductors. With increasing χ, more photocatalysis occurs on the SnS surface, which has a lower efficiency than that on TiO2. The reduction in SnS bandgap energy increases the generation of electron–hole pairs, but the upward shift of the VBM decreases the oxidative power of holes to some extent.
Studies conducted by Jiang et al. on SnS@TiO2 nanobelts and by Yang et al. on amorphous SnS@TiO2 NPs also showed improvements in the photocatalytic efficiency of various dyes compared to pure TiO2 under different light illuminations [15,24]. Mittal et al. reported similar behavior of RhB bleaching rates with various χ as that observed in Figure 7 [17].
These findings collectively describe the complex interplay of semiconductor properties, coating ratios, and experimental conditions in determining photocatalytic efficacy for pollutant degradation.

5. Conclusions

We have coated porous TiO2 nanoparticle thin films with SnSx and characterized their properties with SEM, XRD, XPS, and UV-vis absorption techniques. We found that the effectiveness of SnSx@TiO2 nanoparticle thin films for photocatalytic applications hinges on finding a balance in the SnSx content. Decoration with a 2% molar concentration of SnSx enhances the photocatalytic activity towards rhodamine B, whereas excessive SnSx reduces efficiency.
The observed photocatalytic behavior suggests a delicate balance of several competing factors. The ILAR growth technique ensures intimate contact at interfaces. Optimal alignment of CBMs can facilitate electron transfer, reducing electron–hole recombination rates and promoting the utilization of more photogenerated charges. However, when a hole transfers from the VB of TiO2 to that of SnS, it decreases the oxidative potential, which is crucial for the pollutant degradation process.

Author Contributions

Conceptualization, F.X. and C.V.; methodology, F.X., C.V. and N.S.; validation, F.X. and C.V.; formal analysis, F.X. and C.V.; investigation, F.X., C.V. and N.S.; resources, F.X. and C.V.; data curation, F.X.; writing—original draft preparation, F.X.; writing—review and editing, F.X. and C.V.; funding acquisition, C.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors sincerely thank M. Davoli, I. D. Perrotta, G. Niceforo, and G. Desideri for their technical assistance and support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ren, G.; Han, H.; Wang, Y.; Liu, S.; Zhao, J.; Meng, X.; Li, Z. Recent Advances of Photocatalytic Application in Water Treatment: A Review. Nanomaterials 2021, 11, 1804. [Google Scholar] [CrossRef] [PubMed]
  2. Ge, J.; Zhang, Y.; Heo, Y.J.; Park, S.J. Advanced Design and Synthesis of Composite Photocatalysts for the Remediation of Wastewater: A Review. Catalysts 2019, 9, 122. [Google Scholar] [CrossRef]
  3. Kumari, H.; Suman, S.; Ranga, R.; Chahal, S.; Devi, S.; Sharma, S.; Kumar, S.; Kumar, P.; Kumar, S.; Kumar, A.; et al. A Review on Photocatalysis Used for Wastewater Treatment: Dye Degradation. Water Air Soil Pollut. 2023, 234, 349. [Google Scholar] [CrossRef]
  4. Mishra, S.; Sundaram, B. A Review of the Photocatalysis Process Used for Wastewater Treatment. Mater. Today Proc. 2023, in press. [CrossRef]
  5. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO 2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef]
  6. Jain, P.; Arun, P. Influence of Grain Size on the Band-Gap of Annealed SnS Thin Films. Thin Solid Films 2013, 548, 241–246. [Google Scholar] [CrossRef]
  7. Henry, J.; Mohanraj, K.; Kannan, S.; Barathan, S.; Sivakumar, G. Structural and Optical Properties of SnS Nanoparticles and Electron-Beam-Evaporated SnS Thin Films. J. Exp. Nanosci. 2015, 10, 78–85. [Google Scholar] [CrossRef]
  8. Peng, H.; Jiang, L.; Huang, J.; Li, G. Synthesis of Morphologically Controlled Tin Sulfide Nanostructures. J. Nanoparticle Res. 2007, 9, 1163–1166. [Google Scholar] [CrossRef]
  9. Tarkas, H.S.; Marathe, D.M.; Mahajan, M.S.; Muntaser, F.; Patil, M.B.; Tak, S.R.; Sali, J.V. Synthesis of Tin Monosulfide (SnS) Nanoparticles Using Surfactant Free Microemulsion (SFME) with the Single Microemulsion Scheme. Mater. Res. Express 2017, 4, 025018N. [Google Scholar] [CrossRef]
  10. Aparna, N.; Philip, R.S.; Mathew, M. Photocatalytic Activities of SnS Thin Films Deposited at Room Temperature. Appl. Surf. Sci. Adv. 2023, 13, 100374. [Google Scholar] [CrossRef]
  11. Hegde, S.S.; Surendra, B.S.; Talapatadur, V.; Murahan, P.; Pamesh, K. Visible Light Photocatalytic Properties of Cubic and Orthorhombic SnS Nanoparticles. Chem. Phys. Lett. 2020, 754, 137665. [Google Scholar] [CrossRef]
  12. Chaki, S.H.; Chaudhary, M.D.; Deshpande, M.P. SnS Thin Films Deposited by Chemical Bath Deposition, Dip Coating and SILAR Techniques. J. Semicond. 2016, 37, 053001. [Google Scholar] [CrossRef]
  13. Nengzi, L.C.; Yang, H.; Hu, J.Z.; Zhang, W.M.; Jiang, D.A. Fabrication of SnS/TiO2 NRs/NSs Photoelectrode as Photoactivator of Peroxymonosulfate for Organic Pollutants Elimination. Sep. Purif. Technol. 2020, 249, 117172. [Google Scholar] [CrossRef]
  14. Sharma, S.; Mittal, A.; Singh Chauhan, N.; Makgwane, P.R.; Kumari, K.; Maken, S.; Kumar, N. Developments in Visible-Light Active TiO2/SnX (X = S and Se) and Their Environmental Photocatalytic Applications—A Mini-Review. Inorg. Chem. Commun. 2021, 133, 108874. [Google Scholar] [CrossRef]
  15. Jiang, Y.; Yang, Z.; Zhang, P.; Jin, H.; Ding, Y. Natural Assembly of a Ternary Ag-SnS-TiO2 Photocatalyst and Its Photocatalytic Performance under Simulated Sunlight. RSC Adv. 2018, 8, 13408–13416. [Google Scholar] [CrossRef]
  16. Barberio, M.; Imbrogno, A.; Grosso, D.R.; Bonanno, A.; Xu, F. Low-Cost Carbon-Based Counter Electrodes for Dye-Sensitized Solar Cells. Mater. Res. Express 2015, 2, 075502. [Google Scholar] [CrossRef]
  17. Mittal, A.; Sharma, S.; Kumar, T.; Chauhan, N.S.; Kumari, K.; Maken, S.; Kumar, N. Surfactant-Assisted Hydrothermally Synthesized Novel TiO2/SnS@Pd Nano-Composite: Structural, Morphological and Photocatalytic Activity. J. Mater. Sci. Mater. Electron. 2020, 31, 2010–2021. [Google Scholar] [CrossRef]
  18. Whittles, T.J.; Burton, L.A.; Skelton, J.M.; Walsh, A.; Veal, T.D.; Dhanak, V.R. Band Alignments, Valence Bands, and Core Levels in the Tin Sulfides SnS, SnS2, and Sn2S3: Experiment and Theory. Chem. Mater. 2016, 28, 3718–3726. [Google Scholar] [CrossRef]
  19. Miyauchi, M. Tailoring of SnS Quantum Dots in Mesoporous Media for Efficient Photoelectrochemical Device. Chem. Phys. Lett. 2011, 514, 151–155. [Google Scholar] [CrossRef]
  20. Watanabe, T.; Takizawa, T.; Honda, K. Photocatalysis through Excitation of Adsorbates. 1. Highly Efficient N-Deethylation of Rhodamine B Adsorbed to CdS. J. Phys. Chem. 1977, 81, 1845–1851. [Google Scholar] [CrossRef]
  21. Rauf, A.; Arif Sher Shah, M.S.; Lee, J.Y.; Chung, C.H.; Bae, J.W.; Yoo, P.J. Non-Stoichiometric SnS Microspheres with Highly Enhanced Photoreduction Efficiency for Cr(VI) Ions. RSC Adv. 2017, 7, 30533–30541. [Google Scholar] [CrossRef]
  22. Maheu, C.; Cardenas, L.; Puzenat, E.; Afanasiev, P.; Geantet, C. UPS and UV Spectroscopies Combined to Position the Energy Levels of TiO2 Anatase and Rutile Nanopowders. Phys. Chem. Chem. Phys. 2018, 20, 25629–25637. [Google Scholar] [CrossRef] [PubMed]
  23. Takizawa, T.; Watanabe, T.; Honda, K. Photocatalysis through Excitation of Adsorbates. 2. A Comparative Study of Rhodamine B and Methylene Blue on Cadmium Sulfide. J. Phys. Chem. 1978, 82, 1391–1396. [Google Scholar] [CrossRef]
  24. Yu, K.; Yang, S.; He, H.; Sun, C.; Gu, C.; Ju, Y. Visible Light-Driven Photocatalytic Degradation of Rhodamine B over NaBiO3: Pathways and Mechanism. J. Phys. Chem. A 2009, 113, 10024–10032. [Google Scholar] [CrossRef]
  25. Zhang, G.; Fu, Z.; Wang, Y.; Hao, W. Facile Synthesis of Hierarchical SnS Nanostructures and Their Visible Light Photocatalytic Properties. Adv. Powder Technol. 2015, 26, 1183–1190. [Google Scholar] [CrossRef]
Figure 1. SEM images of a pure TiO2 thin film (A,C) and an SnSx (Sn100S600)-coated TiO2 film (B,D). The beam energy was 5 keV for (A,B) and 10 keV for (C,D), while the magnification was 500 k for all images.
Figure 1. SEM images of a pure TiO2 thin film (A,C) and an SnSx (Sn100S600)-coated TiO2 film (B,D). The beam energy was 5 keV for (A,B) and 10 keV for (C,D), while the magnification was 500 k for all images.
Coatings 14 01185 g001
Figure 2. Cross-sectional SEM image showing a TiO2 thin film about 11 μm thick.
Figure 2. Cross-sectional SEM image showing a TiO2 thin film about 11 μm thick.
Coatings 14 01185 g002
Figure 3. EDX spectra for a pure TiO2 thin film (A) and two SnSx-coated TiO2 thin films (B,C). Also indicated are the molar ratios between various elements.
Figure 3. EDX spectra for a pure TiO2 thin film (A) and two SnSx-coated TiO2 thin films (B,C). Also indicated are the molar ratios between various elements.
Coatings 14 01185 g003
Figure 4. XRD pattern for a pure TiO2 nanoparticle (P25) thin film deposited on an FTO-grazed glass substrate (a); decorated with SnSx Sn020S120 (b); Sn100S600 (c); and Sn100S600 (5 cycles) (d).
Figure 4. XRD pattern for a pure TiO2 nanoparticle (P25) thin film deposited on an FTO-grazed glass substrate (a); decorated with SnSx Sn020S120 (b); Sn100S600 (c); and Sn100S600 (5 cycles) (d).
Coatings 14 01185 g004
Figure 5. XPS spectra of Ti 2p, O 1s, Sn 3d, and S 2p levels for films of pure TiO2 (Sn000S000) and SnSx-coated (Sn100S600) TiO2 films.
Figure 5. XPS spectra of Ti 2p, O 1s, Sn 3d, and S 2p levels for films of pure TiO2 (Sn000S000) and SnSx-coated (Sn100S600) TiO2 films.
Coatings 14 01185 g005
Figure 6. Upper panel: UV-vis optical absorption spectra of pure TiO2 thin film and decorated TiO2 films with SnSx NPs of different amounts; Lower panel: bandgap energy determination made using the Tauc plots.
Figure 6. Upper panel: UV-vis optical absorption spectra of pure TiO2 thin film and decorated TiO2 films with SnSx NPs of different amounts; Lower panel: bandgap energy determination made using the Tauc plots.
Coatings 14 01185 g006
Figure 7. Upper panel: Time evolution of the RhB absorption spectra for the sample Sn005S030. Middle panel: Relative RhB concentration change C(t)/C0 as a function of Xe lamp irradiation time t for various SnSx@TiO2 films. The lines are the linear fittings. Lower panel: The zero-order degradation rate constant vs. Sn-to-Ti molar ratio.
Figure 7. Upper panel: Time evolution of the RhB absorption spectra for the sample Sn005S030. Middle panel: Relative RhB concentration change C(t)/C0 as a function of Xe lamp irradiation time t for various SnSx@TiO2 films. The lines are the linear fittings. Lower panel: The zero-order degradation rate constant vs. Sn-to-Ti molar ratio.
Coatings 14 01185 g007
Figure 8. Energy level potential diagram of SnS, TiO2, RhB, and relevant redox reactions.
Figure 8. Energy level potential diagram of SnS, TiO2, RhB, and relevant redox reactions.
Coatings 14 01185 g008
Table 1. SnS nanoparticles formed with a constant [Sn2+] = 10 mM and various [S2−].
Table 1. SnS nanoparticles formed with a constant [Sn2+] = 10 mM and various [S2−].
Sn-to-Ti Ratio χ ± 0.01S-to-Sn Ratio x ± 0.05Direct Band Gap Eg(SnSx) (eV) ± 0.05 eV
Sn010S0100.030.202.10
Sn010S0200.030.402.05
Sn010S0400.030.602.00
Sn010S0600.030.852.00
Sn010S1000.030.852.00
Table 2. NPs formed with a constant [S2−] to [Sn2+] precursor concentration ratio of α = 6 (S-to-Sn ratio x = 0.85 ± 0.05).
Table 2. NPs formed with a constant [S2−] to [Sn2+] precursor concentration ratio of α = 6 (S-to-Sn ratio x = 0.85 ± 0.05).
Sn-to-Ti Ratio χ ± 0.01Direct Band Gap Eg(SnSx) (eV) ± 0.05 eV
Sn000S0000.00-
Sn001S0060.01-
Sn005S0300.022.10
Sn010S0600.032.00
Sn020S1200.041.90
Sn100S6000.081.40
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

Xu, F.; Scaramuzza, N.; Versace, C. Decorating TiO2 Nanoparticle Thin Film with SnSx (x < 1): Preparation, Characterization, and Photocatalytic Activity. Coatings 2024, 14, 1185. https://doi.org/10.3390/coatings14091185

AMA Style

Xu F, Scaramuzza N, Versace C. Decorating TiO2 Nanoparticle Thin Film with SnSx (x < 1): Preparation, Characterization, and Photocatalytic Activity. Coatings. 2024; 14(9):1185. https://doi.org/10.3390/coatings14091185

Chicago/Turabian Style

Xu, Fang, Nicola Scaramuzza, and Carlo Versace. 2024. "Decorating TiO2 Nanoparticle Thin Film with SnSx (x < 1): Preparation, Characterization, and Photocatalytic Activity" Coatings 14, no. 9: 1185. https://doi.org/10.3390/coatings14091185

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