Next Article in Journal
Methane to Methanol through Heterogeneous Catalysis and Plasma Catalysis
Next Article in Special Issue
Cost Effective Solvothermal Method to Synthesize Zn-Doped TiO2 Nanomaterials for Photovoltaic and Photocatalytic Degradation Applications
Previous Article in Journal
Palladium-Catalyzed Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

SnS2/TiO2 Nanocomposites for Hydrogen Production and Photodegradation under Extended Solar Irradiation

by
Sivagowri Shanmugaratnam
1,2,
Balaranjan Selvaratnam
3,
Aravind Baride
3,
Ranjit Koodali
4,*,
Punniamoorthy Ravirajan
1,
Dhayalan Velauthapillai
2 and
Yohi Shivatharsiny
5,*
1
Clean Energy Research Laboratory (CERL), Department of Physics, University of Jaffna, Jaffna 40000, Sri Lanka
2
Faculty of Engineering and Science, Western Norway University of Applied Sciences, P.O. Box 7030, 5020 Bergen, Norway
3
Department of Chemistry, University of South Dakota, Vermillion, SD 57069, USA
4
Western Kentucky University, 1906 College Heights Blvd, Bowling Green, KY 42101, USA
5
Department of Chemistry, University of Jaffna, Jaffna 40000, Sri Lanka
*
Authors to whom correspondence should be addressed.
Catalysts 2021, 11(5), 589; https://doi.org/10.3390/catal11050589
Submission received: 2 April 2021 / Revised: 24 April 2021 / Accepted: 27 April 2021 / Published: 30 April 2021

Abstract

:
Earth–abundant transition metal chalcogenide materials are of great research interest for energy production and environmental remediation, as they exhibit better photocatalytic activity due to their suitable electronic and optical properties. This study focuses on the photocatalytic activity of flower-like SnS2 nanoparticles (composed of nanosheet subunits) embedded in TiO2 synthesized by a facile hydrothermal method. The materials were characterized using different techniques, and their photocatalytic activity was assessed for hydrogen evolution reaction and the degradation of methylene blue. Among the catalysts studied, 10 wt. % of SnS2 loaded TiO2 nanocomposite shows an optimum hydrogen evolution rate of 195.55 µmolg−1, whereas 15 wt. % loading of SnS2 on TiO2 exhibits better performance against the degradation of methylene blue (MB) with the rate constant of 4.415 × 10−4 s−1 under solar simulated irradiation. The improved performance of these materials can be attributed to the effective photo-induced charge transfer and reduced recombination, which make these nanocomposite materials promising candidates for the development of high-performance next-generation photocatalyst materials. Further, scavenging experiments were carried out to confirm the reactive oxygen species (ROS) involved in the photocatalytic degradation. It can be observed that there was a 78% reduction in the rate of degradation when IPA was used as the scavenger, whereas around 95% reduction was attained while N2 was used as the scavenger. Notably, very low degradation (<5%) was attained when the dye alone was directly under solar irradiation. These results further validate that the •OH radical and the superoxide radicals can be acknowledged for the degradation mechanism of MB, and the enhancement of degradation efficiency may be due to the combined effect of in situ dye sensitization during the catalysis and the impregnation of low bandgap materials on TiO2.

Graphical Abstract

1. Introduction

Emission-less, energy efficient, and low-carbon power production is vital for a green economy. Therefore, renewable energy production from hydro, wind, and solar power paves the way toward overcoming the current environmental issues of fossil fuels [1]. Photocatalytic water splitting involved in hydrogen and oxygen evolution reactions can produce chemical energy in H2 and O2 by utilizing solar energy [2]. The practical application of overall water splitting remains limited due to the lack of effective and stable catalysts to reduce reaction energy barriers. Therefore, the development of efficient and affordable photocatalysts is of great significance.
Various materials, such as TiO2 [3,4,5,6], ZnO [7], mixed oxide materials [8], dye [9,10], and metal-doped oxide materials [11,12], have been intensively focused on for photocatalytic applications. Among these, TiO2 is considered a gold standard due to its significant characteristics, such as strong optical absorption, favorable band edge position, nontoxicity, and abundant availability [13,14]. However, the wide bandgap energy (~3.2 eV) of TiO2 necessitates UV irradiation, but the composites of TiO2 with co-catalysts enable the catalysts to absorb visible light abundant in solar radiation [15].
Recently, transition metal chalcogenides (TMCs) have gained more attention due to their electrocatalytic properties that includes indirect bandgaps, optoelectronic behavior, and their stability [16,17]. Moreover, the stronger edge effects and the quantum confinement effects make these nanodots (quantum dots) or nanostructures of metal chalcogenides possible to utilize considerable amounts of solar irradiation [17,18,19]. These are playing an increasingly important role in different applications, such as photo degradation [20], capacitors [21], and hydrogen evolution [22] due to their suitable electronic and optical properties.
The generation of an exciton pair and the effective charge separation in these semiconductor photocatalysts lead to an efficient photocatalytic process. The efficiency of charge transport mainly depends on the crystallinity, particle, and crystallite size of the materials. Although the small bandgap energy of TMCs necessitates the use of visible irradiation, composites of TMCs with co-catalysts, for example, graphene [23], graphene oxides [24], carbon nitride [25], metal oxides [26,27], and metals [28], were found to increase the conductivity of electrons, provide active sites, and effectively separate the electron and hole pairs generated by the semiconductor photocatalysts. Among the transition metal chalcogenides, SnS2 receives significant attention due to its interesting optical and electronic properties [29]. It has a CdI2-type layered n-type semiconductor and a bandgap of about 2.18 eV [30,31]. Until now, different SnS2 nanostructures with a range of morphologies have been prepared in various conditions, exhibiting intriguing physical and chemical properties; for example, well uniform nanoflake SnS2 was synthesized by the hydrothermal method at 145 °C in 48 h by Feng et al. [32], whereas in a different study, Huijuan Geng et al. used Titon-X 100 as a surfactant to get hexagonal nanoflake SnS2 [33]. However, the photocatalytic performance of SnS2 is still considered unsatisfactory, and only few works have been reported on SnS2 nanoparticles for photocatalytic water splitting and photodegradation. The objective of this paper is to investigate whether SnS2/TiO2 nanocomposite will enhance the photocatalytic hydrogen production and rate of degradation of organics. Herein, we report SnS2-embedded TiO2 nanocomposites synthesized by the facile hydrothermal method for photocatalytic hydrogen production and photocatalytic degradation of methylene blue under simulated solar irradiation. We also investigate the optimal loading weight percentage of SnS2 in the nanocomposite for these applications. These results will be of enormous interest to the scientific community involved in material chemistry research, especially with metal chalcogenide materials for various photocatalysis applications.

2. Results

2.1. Characterization of Materials

The powder XRD pattern of pure SnS2, TiO2, and different weight percentages of SnS2/TiO2 (5, 10, 15, and 20) nanocomposites are illustrated in Figure 1. The peaks observed at the 2 theta value of 25.29, 37.88, 48.25, 53.74, 55.14, 62.71, 68.86, 70.30, and 75.06 for the TiO2 material are due to (101), (004), (200), (105), (211), (204), (116), (220), and (215) diffraction planes of anatase phase of TiO2 (JCPDS:00-004-0477), and the peaks at 2 theta values of 14.76, 27.97, 32.00, 41.79, 49.92, 52.41, 54.79, and 60.55 for pristine SnS2 are due to (001), (100), (101), (102), (110), (111), (103), and (201) diffraction planes, which confirms the formation of SnS2 (JCPDS:00-023-0677). A combination of SnS2 and TiO2 peaks observed with the 5, 10, 15, and 20 wt. % SnS2/TiO2 (ST-5, ST-10, ST-15, and ST-20, respectively) nanocomposite confirms good impregnation of SnS2 on TiO2.
Scanning electron microscopic images of pristine TiO2 and SnS2 are illustrated in Figure 2a,b and Figure 2c,d, respectively. It can be clearly seen from Figure 2a,b that the pure TiO2 nanoparticle showed an irregular 3D block-like structure covered with sponge-like particles. The micrograph obtained at higher magnification (Figure 2b) clearly illustrates that the aggregated particles have spongy-like structures, supported by the data published elsewhere [26]. A flower-like structure composed with nanosheets was attained for pristine SnS2 and is shown in Figure 2c,d.
SEM images of different wt. % of SnS2. embedded TiO2 nanocomposites, ST-5 (e), ST-10 (f), ST-15 (g), and ST-20 (h), are shown in Figure 3. These SnS2/TiO2 nanocomposites in different wt. % possess nanosheets with a morphology decorated with sponge-like material and, therefore, exhibit aggregation.
The atomic ratios between Sn and Ti were calculated by using energy dispersive spectroscopy (EDS), with the values found to be 0.029, 0.039, 0.052, and 0.101 for ST-5, ST-10, ST-15, and ST-20, respectively. The calculated values closely matched with the expected values, which are tabulated in Table S1.
The bandgap energies of the pristine SnS2, and different wt. % of SnS2/TiO2 nanocomposite materials were estimated by using the Tauc plot (Figure 3), which transformed the absorption spectra (Figure 4a) via the Kubelka–Munk function [26,34], [F(R∞)E]n vs. E, when n = 0.5, for a direct allowed transition (K = F(R∞)). Figure 4c shows the Tauc plot with the absorption spectra of pristine TiO2. Bandgaps estimated from the Tauc plots, by extrapolating the steep portion of the plot in Figure 4b,c to the x-axis, suggest the bandgaps of pristine SnS2 to be 1.890 eV, which is closer to some literature studies [35], and the bandgap values of SnS2-embedded TiO2 nanocomposite to be (ST-5—2.250 eV, ST-10—2.015 eV, ST-15—2.005 eV, and ST-20—1.980 eV), whereas the bandgaps for pristine TiO2 (3.100 eV) materials lie in the range between 1.89 and 3.10 eV. The pure SnS2 sample shows strong absorption in the visible region with an absorption edge of ~555 nm, corresponding to a bandgap of 1.89 eV. However, the loading of TiO2 with SnS2 elevates the absorption level/ bandgap gradually, which is consistent with the visual color of the samples gradually changing from yellow to white.

2.2. Photocatalytic Hydrogen Evolution

Transition metal chalcogenides (TMCs) have been in the spotlight due to their optoelectronic properties; in addition, they exhibit better performance against photocatalytic applications. To study SnS2/TiO2’s activity for photocatalytic application, the synthesized nanocomposites for hydrogen production and photocatalytic dye degradation were used. The amount of H2 evolution was measured under solar simulated illumination, and the results show an optimum hydrogen evolution of 195.55 µmolg−1 with 10 wt. % of SnS2/TiO2 nanocomposite, which is more than double the amount of hydrogen attained with pure TiO2 (89.20 µmolg−1), as presented in Table 2. The rate of hydrogen evolution was found to increase with increasing wt. % of SnS2 on TiO2 up to 10 wt. % of loading, and a further increase resulted with a decrease in hydrogen evolution. This result can be correlated with the bandgap energies of the materials, which may effectively induce the formation of electron–hole pairs. In the case of pure SnS2 (Figure 5), there was no hydrogen evolution obtained; this can be related to the faster recombination of exciton pairs due to its relatively small bandgap energy (1.89 eV). These results may suggest that the SnS2 nanoparticles act as a co-catalyst when they are in the SnS2/TiO2 nanocomposites and, thus, enhance excited electrons in the reactive site of titanium dioxide (Figure 7).
It is noteworthy to compare the result obtained in this study with the literature; although the experimental conditions are varied and the results are attained in different units, the comparison may provide insight into the TMC materials for hydrogen evolution. Table 1 compares the amount of hydrogen evolution from various transition metal chalcogenides studied under different experimental conditions. It can be clearly seen that different levels of metal doping enhance the hydrogen production, where the carbonaceous substrates improve the performance by providing better surface sites. Further, dye sensitization on these materials and their influence on hydrogen production were also studied. In addition, variable morphologies, such as nanosheets, films, and nanotubes of TMC materials were studied and are illustrated in the table below. Although the comparison of different photocatalysts is challenging, it can be clearly seen that the catalyst prepared in this work exhibits a reasonably good amount of hydrogen in addition to better degradation ability.

2.3. Photocatalytic Degradation

Photocatalytic degradation of methylene blue was also employed to study the synergic effect of the prepared nanocomposite materials. The reaction suspension was allowed to establish adsorption equilibrium by stirring the suspension in the dark for 30 min. After direct sunlight illumination, the solution withdrawn from the reactor in a time interval was tested by studying the absorbance of the remnant dye solution. The absorption spectra for different samples, such as (a) pure TiO2 (ST-0), (b) pure SnS2 (ST-100), SnS2/TiO2 nanocomposites ((c) ST-5, (d) ST-10, (e) ST-15, and (f) ST-20 are illustrated in Figure 6, which clearly explains the better photocatalytic activity of these nanocomposites. The rate of the reaction was determined by the linear plot of ln (A/A0) vs. time A0-absorbance of initial solution (t = 0), where A is absorption of the solution at time t. The highest rate of degradation was obtained for the sample ST-15 (15 wt. % of SnS2/TiO2) with the rate of 4.415 × 10−4 s−1 compared to the pure TiO2 (2.745 × 10−4 s−1) and SnS2 (1.955 × 10−4 s−1). These results may suggest the essential role of SnS2 in the nanocomposite photocatalyst, which enhances the separation of photo-generated electron–hole pairs and thus maximizes the utilization of photon under direct sunlight. It is well known that there are two possible paths for the degradation of the dye: (i) the visible light can be absorbed by the dye molecules, resulting in excited electrons and injection of these electrons into the conduction band of the catalyst, and (ii) the catalyst itself can absorb photons, and the electrons ejected from the valence band to the conduction band will then migrate to its active site. Further, in this study, the scavenging experiments were conducted in the presence of an N2 environment (absence of O2) and in the presence of IPA (to scavenge the •OH radical). It can be clearly seen that there was a 78% reduction in the rate when IPA was used as the scavenger (Figure S3, Supplementary Materials), whereas around 95% reduction was attained while N2 was used as the scavenger (Figure S4, Supplementary Materials). Notably, very low degradation (<5%) was attained when the dye alone was placed under direct light (Figure S5, Supplementary Materials). These results further validate that the •OH radical and the superoxide radicals ( O 2 ) can be acknowledged in the degradation mechanism of MB (Figure 7). The effective degradation efficiency can be attributed to the combined effect of in situ dye sensitization and the impregnation of chalcogenide materials on TiO2.
A comparison between the rate of degradation and the hydrogen production is listed in Table 2. As indicted in the table, the bandgap energy of the materials can be correlated with the better performance of degradation and hydrogen production. The materials ST-10 and ST-15 show very similar bandgaps, and that may be attributed to the higher performance of those materials. The replication studies also showed a similar trend in the photocatalytic degradation and hydrogen production.

3. Materials and Methods

3.1. Materials

The following chemicals were used without further purification. Titanium tetra isopropoxide, 98+% (Sigma–Aldrich Norway AS, Oslo, Norway) was employed as the precursor for TiO2 preparation. Concentrated HNO3 (Sigma–Aldrich Norway AS) PHARMCO-AAPER ethyl alcohol (200 proof; absolute, anhydrous, Sigma–Aldrich Norway AS) was used as catalyst and solvent, respectively. Tin (II) chloride, pentahydrate, ACS reagent, ≥98% (Sigma–Aldrich Norway AS) was used as the tin precursor; thiourea, ACS reagent, 99.0% (Sigma–Aldrich Norway AS) was utilized as sulfur source; and conc. HCl (used as hydrolyzer) and deionized water (resistivity > 18 M cm, Velp/AREC, VELP Scientifica Srl, Usmate, MB, Italy) were used to prepare the solution mixtures.

3.2. Methods

3.2.1. Synthesis

Titanium Dioxide

Nanocrystalline titanium dioxide was prepared using the method we reported previously [26] under hydrothermal conditions using the sol–gel technique. In a typical synthesis, 32.5 mL of ethanol is acidified with 0.3 mL concentrated HNO3 in a Teflon liner and stirred at a constant speed (300 rpm); 6.60 mL of titanium (iv) isopropoxide is then added drop wise into it with continuous stirring; finally, 3.0 mL of water is added to the solution for the formation of a gel. It is then transferred into an autoclave (AUTOCLAVE-PTFE-0100, TECINSTRO, Maharashtra, India) and kept at 180 °C for 6 h. The final product is then washed with ethanol, and the dried product calcined at 500 °C for 6 h.

SnS2-Embedded TiO2 Nanocomposite

The appropriate amount of SnCl2.2H2O was taken into a 50 mL beaker containing 0.8 mL of concentrated hydrochloric acid (37%) in 10 mL of deionized water. It was then stirred at 300 rpm for 30 min, followed by the addition of the appropriate amount of TiO2 (hydrothermally synthesized) before being completely dispersed by constant stirring for 10 min. Then, an appropriate amount of thiourea was added into the solution made up to 40 mL volume by adding deionized water. Finally, the resulting solution was transferred in an autoclave and kept at 145 °C for 12 h. The final product was collected by centrifugation, and it was washed with the mixture of deionized water and ethanol (v/v = 1:1) several times. It was then air-dried to prepare X wt. % (X = 0 (ST-0), 5 (ST-5), 10 (ST-10), 15 (ST-15), 20 (ST-20), and 100 (ST-100)) of the SnS2-embedded TiO2 material. Similar conditions were followed in the preparation of pristine SnS2 nanoparticles without adding titanium dioxide. We refer to this as ST-0—SnS2 is denoted by S; T stands for TiO2; and the numbers 0, 5, 10, 15, 20, and 100 are the wt. % of SnS2 in TiO2.

3.2.2. Characterization

Synthesized nanocomposite materials were characterized using different techniques, such as powder X-ray diffraction (patterns of the catalysts were recorded using D8 ADVANCE ECO in a 1 kW copper X-ray tube diffractometer with the scan range from 3 to 90 degrees (2theta)), diffuse reflectance spectra (DRS Cary 100 Bio UV-Visible spectrophotometer, Santa Clara, CA, USA) by measuring the intensity in the wavelength range of 800–200 nm, and scanning electron microscopy (SEM, Oxford instrument, Nano Analysis, Concord, MA, USA).

3.2.3. Photocatalytic Hydrogen Evolution

The photocatalytic experiments were carried out for pristine TiO2, SnS2, and ST-X (X = 05, 10, 15, and 20). Firstly, the synthesized nanocomposite was suspended in a solution containing 1.5 mL of deionized water and 0.5 mL of methanol (hole scavenger). The suspension was degassed for 30 min with high-purity argon prior to irradiation. The final suspensions were continuously stirred throughout the experiment. A 300 W Xenon lamp (Oriel light source, Xenon arc lamp, Newport 1000W, Irvine, CA, USA) with an AM 1.5 G filter was used as the source of radiation. Finally, the amount of H2 produced was measured by using gas chromatography (SRI 8610 C, SRI Instruments, Torrance, CA, USA) equipped with a molecular sieve column and a TCD (thermal conductivity detector), and the amount of hydrogen produced was quantified by using a calibration curve prepared previously.

3.2.4. Photocatalytic Degradation

Photocatalytic degradation of methylene blue (MB) solution was performed using a ST-X nanocomposite photocatalyst. In a typical experiment, 25.0 mg of the photocatalyst is suspended with 50 mL of MB solution (initial concentration of 10 ppm) under direct sunlight. Prior to irradiation, the suspensions are sonicated in the dark for 30 min to ensure the establishment of adsorption or desorption equilibrium. Periodically, 3 mL of suspension is withdrawn, and absorption spectra of the solution are obtained using a UV/Vis spectrometer (JENWAY 6800 UV/Vis spectrophotometer (OSA, UK)). Scavenging experiments were also carried out under the IPA and N2 environment.

4. Conclusions

The pure SnS2, TiO2, and different wt. % of SnS2/TiO2 nanocomposite samples were successfully synthesized by the facile hydrothermal process, affording an effective photocatalyst that can be utilized for hydrogen production and degradation of methylene blue. The SnS2/TiO2 nanocomposite with 10 wt. % shows an optimum rate of hydrogen production of 195.55 µmolg−1, whereas SnS2/TiO2 nanocomposites with 15 wt. % exhibit a high rate of degradation of methylene blue (4.415 × 10−4 s−1). The pure SnS2 was found to be inactive due to a lower bandgap, and pure TiO2 shows less photoactivity compared with the nanocomposites. This study concludes that SnS2 acts as co-catalyst and thus effectively induces the formation of electron–hole pairs that leads to the higher photocatalytic activity in both hydrogen production and photodegradation applications. The scavenging experiments showed evidence of a 78% reduction in the rate of degradation when IPA was used as the scavenger, whereas around 95% reduction was attained in the presence of N2. Nearly no (<5%) degradation was attained when the dye alone was under direct solar irradiation. Furthermore, our analysis clearly showed that enhancement in degradation MB could be attributed to hydroxyl radicals and superoxide radicals.

Supplementary Materials

The following information is available online at https://www.mdpi.com/article/10.3390/catal11050589/s1, Figure S1: Elemental dispersive spectra for ST-0, ST-5, ST-10, ST-15, ST-20, and ST-100; Figure S2: First-order kinetics of degradation of methylene blue by using samples ST-0, ST-5, ST-10, ST-15, ST-20, and ST-100; Table S1: Theoretical and experimental value of ratio between Sn and Ti from EDS analysis.

Author Contributions

Conceptualization, S.S. and Y.S.; methodology, S.S. and Y.S.; software, S.S.; validation, S.S., Y.S., D.V., P.R., B.S., R.K., and A.B.; formal analysis, S.S. and B.S.; investigation, S.S. and Y.S.; resources, Y.S., P.R., D.V., R.K., and A.B.; data curation, S.S. and Y.S.; writing—original draft preparation, S.S.; writing—review and editing, Y.S., B.S., D.V., P.R., and R.K.; visualization, Y.S.; supervision, Y.S., P.R., and D.V.; project administration, P.R. and D.V.; funding acquisition, P.R. and D.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Capacity Building and Establishment of a Research Consortium (CBERC) project, grant number LKA-3182-HRNCET and Higher education and Research collaboration on Nanomaterials for Clean Energy Technologies (HRNCET) project, grant number NORPART/2016/10237.

Data Availability Statement

The data in this paper is available in Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ran, J.; Zhang, J.; Yu, J.; Jaroniec, M.; Qiao, S.Z. Earth-abundant cocatalysts for semiconductor-based photocatalytic water splitting. Chem. Soc. Rev. 2014, 43, 7787–7812. [Google Scholar] [CrossRef]
  2. Nguyen, P.D.; Duong, T.M.; Tran, P.D. Current progress and challenges in engineering viable artificial leaf for solar water splitting. J. Sci. Adv. Mater. Devices 2017, 2, 399–417. [Google Scholar] [CrossRef]
  3. Leung, D.Y.C.; Fu, X.; Wang, C.; Ni, M.; Leung, M.K.H.; Wang, X.; Fu, X. Hydrogen production over titania-based photocatalysts. ChemSusChem 2010, 3, 681–694. [Google Scholar] [CrossRef]
  4. Yin, Y.; Jin, Z.; Hou, F. Enhanced solar water-splitting efficiency using core/sheath heterostructure CdS/TiO2 nanotube arrays. Nanotechnology 2007. [Google Scholar] [CrossRef] [PubMed]
  5. Kim, S.B.; Hong, S.C. Kinetic study for photocatalytic degradation of volatile organic compounds in air using thin film TiO2 photocatalyst. Appl. Catal. B Environ. 2002, 35, 305–315. [Google Scholar] [CrossRef]
  6. Rasalingam, S.; Peng, R.; Koodali, R.T. Removal of Hazardous Pollutants from Wastewaters: Applications of TiO2-SiO2 Mixed Oxide Materials Shivatharsiny. J. Nanomater. 2014, 2014, 617405. [Google Scholar] [CrossRef] [Green Version]
  7. Chakrabarti, S.; Dutta, B.K. Photocatalytic degradation of model textile dyes in wastewater using ZnO as semiconductor catalyst. J. Hazard. Mater. 2004, 112, 269–278. [Google Scholar] [CrossRef]
  8. Mugunthan, E.; Saidutta, M.B.; Jagadeeshbabu, P.E. Photocatalytic degradation of diclofenac using TiO2–SnO2 mixed oxide catalysts. Environ. Technol. 2019, 40, 929–941. [Google Scholar] [CrossRef]
  9. Kong, C.; Min, S.; Lu, G. Dye-sensitized NiSx catalyst decorated on graphene for highly efficient reduction of water to hydrogen under visible light irradiation. ACS Catal. 2014, 4, 2763–2769. [Google Scholar] [CrossRef]
  10. Pirashanthan, A.; Murugathas, T.; Mariappan, K.; Ravirajan, P.; Velauthapillai, D.; Yohi, S. A multifunctional ruthenium based dye for hybrid nanocrystalline titanium dioxide/poly(3-hexylthiophene) solar cells. Mater. Lett. 2020, 274, 127997. [Google Scholar] [CrossRef]
  11. Zhang, W.; Xiao, X.; Zeng, X.; Li, Y.; Zheng, L.; Wan, C. Enhanced photocatalytic activity of TiO2 nanoparticles using SnS2/RGO hybrid as co-catalyst: DFT study and photocatalytic mechanism. J. Alloys Compd. 2016, 685, 774–783. [Google Scholar] [CrossRef]
  12. Jafari, T.; Moharreri, E.; Amin, A.S.; Miao, R.; Song, W.; Suib, S.L. Photocatalytic water splitting—The untamed dream: A review of recent advances. Molecules 2016, 21, 900. [Google Scholar] [CrossRef] [PubMed]
  13. Pirashanthan, A.; Murugathas, T.; Robertson, N.; Ravirajan, P.; Velauthapillai, D. A Quarterthiophene-Based Dye as an Efficient. Polymers 2019, 11, 1752. [Google Scholar] [CrossRef] [Green Version]
  14. Eidsvåg, H.; Bentouba, S.; Vajeeston, P.; Yohi, S.; Velauthapillai, D. TiO2 as a Photocatalyst for Water Splitting—An Experimental and Theoretical Review. Molecules 2021, 26, 1687. [Google Scholar] [CrossRef] [PubMed]
  15. Rajaramanan, T.; Natarajan, M.; Ravirajan, P.; Senthilnanthanan, M.; Velauthapillai, D. Ruthenium (Ru) Doped Titanium Dioxide (P25) electrode for dye sensitized solar cells. Energies 2020, 13, 1532. [Google Scholar] [CrossRef] [Green Version]
  16. Li, Y.; Zhou, X.; Xing, Y. In situ thermal-assisted loading of monodispersed Pt nanoclusters on CdS nanoflowers for efficient photocatalytic hydrogen evolution. Appl. Surf. Sci. 2020, 506, 144933. [Google Scholar] [CrossRef]
  17. Shanmugaratnam, S.; Rasalingam, S. Transition Metal Chalcogenide (TMC) Nanocomposites for Environmental Remediation Application over Extended Solar Irradiation. Intech 2019. [Google Scholar] [CrossRef] [Green Version]
  18. Zheng, L.; Zhang, W.; Xiao, X. Preparation of titanium dioxide/tungsten disulfide composite photocatalysts with enhanced photocatalytic activity under visible light. Korean J. Chem. Eng. 2016, 33, 107–113. [Google Scholar] [CrossRef]
  19. Wang, Q.; Huang, J.; Sun, H.; Ng, Y.H.; Zhang, K.Q.; Lai, Y. MoS2 Quantum Dots@TiO2 Nanotube Arrays: An Extended-Spectrum-Driven Photocatalyst for Solar Hydrogen Evolution. ChemSusChem 2018, 11, 1708–1721. [Google Scholar] [CrossRef]
  20. Meng, Z.; Oh, W. Photodegradation of Organic Dye by CoS2 and Carbon(C60, Graphene, CNT)/TiO2 Composite Sensitizer. Chin. J. Catal. 2012, 33, 1495–1501. [Google Scholar] [CrossRef]
  21. Kajana, T.; Velauthapillai, D.; Shivatharsiny, Y.; Ravirajan, P.; Yuvapragasam, A.; Senthilnanthanan, M. Structural and photoelectrochemical characterization of heterostructured carbon sheet/Ag2MoO4-SnS/Pt photocapacitor. J. Photochem. Photobiol. A Chem. 2020, 401, 112784. [Google Scholar] [CrossRef]
  22. Chen, T.Y.; Chang, Y.H.; Hsu, C.L.; Wei, K.H.; Chiang, C.Y.; Li, L.J. Comparative study on MoS2 and WS2 for electrocatalytic water splitting. Int. J. Hydrog. Energy 2013, 38, 12302–12309. [Google Scholar] [CrossRef]
  23. Ali, A.; Oh, W.C. Photocatalytic performance of CoS2-graphene-TiO2 ternary composites for reactive black B (RBB) degradation. J. Korean Ceram. Soc. 2017, 54, 308–313. [Google Scholar] [CrossRef] [Green Version]
  24. Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. MoS2 nanoparticles grown on graphene: An advanced catalyst for the hydrogen evolution reaction. J. Am. Chem. Soc. 2011, 133, 7296–7299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. He, K.; Xie, J.; Li, M.; Li, X. In situ one-pot fabrication of g-C3N4 nanosheets/NiS cocatalyst heterojunction with intimate interfaces for efficient visible light photocatalytic H2 generation. Appl. Surf. Sci. 2018, 430, 208–217. [Google Scholar] [CrossRef]
  26. Shanmugaratnam, S.; Velauthapillai, D.; Ravirajan, P.; Christy, A.A.; Shivatharsiny, Y. CoS2/TiO2 in nanocomposites for hydrogen production under UV irradiation. Materials 2019, 12, 3882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Khan, Z.; Chetia, T.R.; Vardhaman, A.K.; Barpuzary, D.; Sastri, C.V.; Qureshi, M. Visible light assisted photocatalytic hydrogen generation and organic dye degradation by CdS-metal oxide hybrids in presence of graphene oxide. RSC Adv. 2012, 2, 12122–12128. [Google Scholar] [CrossRef]
  28. Sun, B.; Liu, A.; Li, J.; Wang, J.; Wang, S. Development of novel highly stable synergistic quaternary photocatalyst for the efficient hydrogen evolution reaction. Appl. Surf. Sci. 2020, 510, 145498. [Google Scholar] [CrossRef]
  29. Zhu, W.; Yang, Y.; Ma, D.; Wang, H.; Zhang, Y.; Hu, H. Controlled growth of flower-like SnS2 hierarchical structures with superior performance for lithium-ion battery applications. Ionics 2014, 21, 19–26. [Google Scholar] [CrossRef]
  30. Mondal, C.; Ganguly, M.; Pal, J.; Roy, A.; Jana, J.; Pal, T. Morphology controlled synthesis of SnS2 nanomaterial for promoting photocatalytic reduction of aqueous Cr(VI) under visible light. Langmuir 2014, 30, 4157–4164. [Google Scholar] [CrossRef]
  31. Gaur, R.; Jeevanandam, P. Synthesis of SnS2 nanoparticles and their application as photocatalysts for the reduction of Cr(VI). J. Nanosci. Nanotechnol. 2018, 18, 165–177. [Google Scholar] [CrossRef]
  32. Feng, J.; Chen, J.; Geng, B.; Feng, H.; Li, H.; Yan, D.; Zhuo, R.; Cheng, S.; Wu, Z.; Yan, P. Two-dimensional hexagonal SnS2 nanoflakes: Fabrication, characterization, and growth mechanism. Appl. Phys. A Mater. Sci. Process. 2011, 103, 413–419. [Google Scholar] [CrossRef]
  33. Geng, H.; Su, Y.; Wei, H.; Xu, M.; Wei, L.; Yang, Z.; Zhang, Y. Controllable synthesis and photoelectric property of hexagonal SnS2 nanoflakes by Triton X-100 assisted hydrothermal method. Mater. Lett. 2013, 111, 204–207. [Google Scholar] [CrossRef]
  34. Rasalingam, S.; Wu, C.M.; Koodali, R.T. Modulation of pore sizes of titanium dioxide photocatalysts by a facile template free hydrothermal synthesis method: Implications for photocatalytic degradation of rhodamine B. ACS Appl. Mater. Interfaces 2015, 7, 4368–4380. [Google Scholar] [CrossRef]
  35. Rangappa, A.P.; Kumar, D.P.; Gopannagari, M.; Reddy, D.A.; Hong, Y.; Kim, Y.; Kim, T.K. Highly efficient hydrogen generation in water using 1D CdS nanorods integrated with 2D SnS2 nanosheets under solar light irradiation. Appl. Surf. Sci. 2020, 508, 144803. [Google Scholar] [CrossRef]
  36. Zong, X.; Han, J.; Ma, G.; Yan, H.; Wu, G.; Li, C. Photocatalytic H2 evolution on CdS loaded with WS2 as cocatalyst under visible light irradiation. J. Phys. Chem. C 2011, 115, 12202–12208. [Google Scholar] [CrossRef]
  37. Li, Y.; Wang, H.; Peng, S. Tunable photodeposition of MoS2 onto a composite of reduced graphene oxide and CdS for synergic photocatalytic hydrogen generation. J. Phys. Chem. C 2014, 118, 19842–19848. [Google Scholar] [CrossRef]
  38. Chang, K.; Mei, Z.; Wang, T.; Kang, Q.; Ouyang, S.; Ye, J. MoS2/graphene cocatalyst for efficient photocatalytic H2 evolution under visible light irradiation. ACS Nano 2014, 8, 7078–7087. [Google Scholar] [CrossRef]
  39. Yuan, Y.; Lu, H.; Ji, Z.; Zhong, J.; Ding, M.; Chen, D.; Li, Y.; Tu, W.; Cao, D.; Yu, Z.; et al. Enhanced visible-light-induced hydrogen evolution from water in a noble-metal-free system catalyzed by ZnTCPP-MoS2/TiO2 assembly. Chem. Eng. J. 2015, 275, 8–16. [Google Scholar] [CrossRef]
  40. Liu, E.; Chen, J.; Ma, Y.; Feng, J.; Jia, J.; Fan, J.; Hu, X. Fabrication of 2D SnS2 /g-C3N4 heterojunction with enhanced H2 evolution during photocatalytic water splitting. J. Colloid Interface Sci. 2018, 524, 313–324. [Google Scholar] [CrossRef] [PubMed]
  41. Yan, C.; Xue, X.; Zhang, W.; Li, X.; Liu, J.; Yang, S.; Hu, Y.; Chen, R.; Yan, Y.; Zhu, G.; et al. Well-designed Te/SnS2/Ag artificial nanoleaves for enabling and enhancing visible-light driven overall splitting of pure water. Nano Energy 2017, 39, 539–545. [Google Scholar] [CrossRef]
  42. Yu, J.; Xu, C.Y.; Ma, F.X.; Hu, S.P.; Zhang, Y.W.; Zhen, L. Monodisperse SnS2 nanosheets for high-performance photocatalytic hydrogen generation. ACS Appl. Mater. Interfaces 2014, 6, 22370–22377. [Google Scholar] [CrossRef] [PubMed]
  43. Barba-Nieto, I.; Christoforidis, K.C.; Fernández-García, M.; Kubacka, A. Promoting H2 photoproduction of TiO2-based materials by surface decoration with Pt nanoparticles and SnS2 nanoplatelets. Appl. Catal. B Environ. 2020, 277, 119246. [Google Scholar] [CrossRef]
Figure 1. XRD for pristine TiO2 (ST-0), SnS2 (ST-100), and X wt. % of SnS2/TiO2 (ST-5, ST-10, ST-15, and ST-20) nanocomposites.
Figure 1. XRD for pristine TiO2 (ST-0), SnS2 (ST-100), and X wt. % of SnS2/TiO2 (ST-5, ST-10, ST-15, and ST-20) nanocomposites.
Catalysts 11 00589 g001
Figure 2. SEM images for (a,b): pristine TiO2 (ST-0), and (c,d): SnS2 (ST-100) nanomaterials.
Figure 2. SEM images for (a,b): pristine TiO2 (ST-0), and (c,d): SnS2 (ST-100) nanomaterials.
Catalysts 11 00589 g002
Figure 3. SEM images for X wt. % of SnS2/TiO2: (ae) ST-5, (bf) ST-10, (cg) ST-15, and (dh) ST-20 nanocomposites.
Figure 3. SEM images for X wt. % of SnS2/TiO2: (ae) ST-5, (bf) ST-10, (cg) ST-15, and (dh) ST-20 nanocomposites.
Catalysts 11 00589 g003
Figure 4. (a) Absorption spectra for samples ST-5, ST-10, ST-15, ST-20, and ST-100; (b) Tauc plot for S-5, ST-10, ST-15, ST-20, and ST-100 and; (c) Tauc plot of pure TiO2 sample (ST-0) (inset shows the respective absorption spectra).
Figure 4. (a) Absorption spectra for samples ST-5, ST-10, ST-15, ST-20, and ST-100; (b) Tauc plot for S-5, ST-10, ST-15, ST-20, and ST-100 and; (c) Tauc plot of pure TiO2 sample (ST-0) (inset shows the respective absorption spectra).
Catalysts 11 00589 g004
Figure 5. The amount of hydrogen evolved vs. time in the presence of pristine TiO2, SnS2, and SnS2/TiO2 nanocomposite with different weight percentage (ST-5, ST-10, ST-15, and ST-20) in different time intervals.
Figure 5. The amount of hydrogen evolved vs. time in the presence of pristine TiO2, SnS2, and SnS2/TiO2 nanocomposite with different weight percentage (ST-5, ST-10, ST-15, and ST-20) in different time intervals.
Catalysts 11 00589 g005
Figure 6. Absorbance vs. wavelength graph for degradation of methylene blue by using pristine (a) TiO2 (ST-0), (b) SnS2 (ST-100) and SnS2/TiO2 nanocomposites (c) ST-5, (d) ST-10, (e) ST-15 and (f) ST-20) as photocatalysts at different time intervals under direct sunlight.
Figure 6. Absorbance vs. wavelength graph for degradation of methylene blue by using pristine (a) TiO2 (ST-0), (b) SnS2 (ST-100) and SnS2/TiO2 nanocomposites (c) ST-5, (d) ST-10, (e) ST-15 and (f) ST-20) as photocatalysts at different time intervals under direct sunlight.
Catalysts 11 00589 g006
Figure 7. Photocatalytic mechanism.
Figure 7. Photocatalytic mechanism.
Catalysts 11 00589 g007
Table 1. A comparison of amount of hydrogen evolved with different transition metal chalcogenides.
Table 1. A comparison of amount of hydrogen evolved with different transition metal chalcogenides.
MaterialPreparation MethodAmount of Hydrogen ProducedLigtht SourceSacrificial Agent
CdS on WS2Impregnation–sulfidation0.198 mmolh−1VisibleLatic acid[36]
Dye-sensitized NiSx on grapheneIn situ chemical deposition0.34 mmolh−1--[9]
MoS2 on RGO and CdSPhotoreduction 0.099 mmolh−1VisibleLatic acid[37]
MoS2 on grapheneHydrothermal1.80 mmolh−1VisibleNa2S-Na2S2O3[38]
MoS2 QDs on TiO2 NTAElectrodeposition0.065 mmolcm−2h−1
0.053 mmolcm−2h−1
0.016 mmolcm−2h−1
UV
Visible
NIR
-[19]
ZnTCPP-MoS2 on TiO2Hydrothermal0.010 mmolh−1-triethanolamine (TEOA)[39]
10 wt. % CoS2 on TiO2Hydrothermal2.55 mmolg−1UVMethanol[26]
2D SnS2 on g-C3N4Hydrothermal0.972 mmolh−1g−1VisibleTEOA and H2Pt2Cl6.6H2O[40]
Te/SnS2/AgHydrothermal0.332 mmolh−1UV–visible-[41]
SnS2 nanosheetsSolvothermal1.06 mmolh−1g−1UV–visibleNa2S
Na2S2O3
[42]
CdS on SnS2Hydrothermal20.2 mmolh−1g−1UV–visibleLatic acid[35]
Pt nanoparticles on oxide,Hydrothermal10.0 mmolh−1g−1UV [43]
Pt nanoparticles on SnS2 nanoplatelets, and 9.0 mmolh−1g−1
Pt nanoparticles on SnS2 and oxide3.0 mmolh−1g−1
10 wt. % SnS2 on TiO2 nanocompositeHydrothermal 0.195 mmolg−1UV–visibleMethanol
Table 2. Comparison of the amount of hydrogen evolved and rate constant of the degradation reaction of MB by using pure SnS2, TiO2, and SnS2-embedded TiO2 nanocomposite over simulated irradiation with the bandgap of these materials.
Table 2. Comparison of the amount of hydrogen evolved and rate constant of the degradation reaction of MB by using pure SnS2, TiO2, and SnS2-embedded TiO2 nanocomposite over simulated irradiation with the bandgap of these materials.
SampleAmount of Hydrogen Evolved (µmolg−1)Rate Constant of Photodegradation Reaction (×10−4 s−1)Calculated BandGap Value (eV)
ST-1000.001.955 ± 0.1851.890
ST-20142.352.768 ± 0.1811.980
ST-15171.304.415 ± 0.2582.005
ST-10195.553.948 ± 0.1102.015
ST-528.252.661 ± 0.3882.250
ST-089.202.745 ± 0.5133.100
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shanmugaratnam, S.; Selvaratnam, B.; Baride, A.; Koodali, R.; Ravirajan, P.; Velauthapillai, D.; Shivatharsiny, Y. SnS2/TiO2 Nanocomposites for Hydrogen Production and Photodegradation under Extended Solar Irradiation. Catalysts 2021, 11, 589. https://doi.org/10.3390/catal11050589

AMA Style

Shanmugaratnam S, Selvaratnam B, Baride A, Koodali R, Ravirajan P, Velauthapillai D, Shivatharsiny Y. SnS2/TiO2 Nanocomposites for Hydrogen Production and Photodegradation under Extended Solar Irradiation. Catalysts. 2021; 11(5):589. https://doi.org/10.3390/catal11050589

Chicago/Turabian Style

Shanmugaratnam, Sivagowri, Balaranjan Selvaratnam, Aravind Baride, Ranjit Koodali, Punniamoorthy Ravirajan, Dhayalan Velauthapillai, and Yohi Shivatharsiny. 2021. "SnS2/TiO2 Nanocomposites for Hydrogen Production and Photodegradation under Extended Solar Irradiation" Catalysts 11, no. 5: 589. https://doi.org/10.3390/catal11050589

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