DOI:
10.1039/D5NA00533G
(Paper)
Nanoscale Adv., 2025,
7, 6032-6048
Fabrication of TiO2@SnS2 core–shell nanocomposites via a thermal decomposition approach for sunlight-driven photodegradation of crystal violet
Received
31st May 2025
, Accepted 1st August 2025
First published on 5th August 2025
Abstract
In the current study, TiO2@SnS2 core–shell nanocomposites were prepared via a facile thermal decomposition method. The synthesis consists of thermal decomposition of tin chloride pentahydrate and thiourea in the presence of NaOH-modified TiO2 microspheres in diphenyl ether (DPE) at ∼200 °C in air. The synthesized TiO2@SnS2 core–shell nanocomposites were characterized by XRD, FT-IR, TGA, FESEM, TEM, EDXA, UV-DRS, PL and XPS. The XRD results indicate the presence of both TiO2 and SnS2 phases in the core–shell nanocomposites. FESEM, TEM and EDX analyses results confirm uniform coating of SnS2 nanoparticles on the surface modified TiO2 microspheres. XPS analysis results confirm the presence of Sn4+, Ti4+, O2− and S2− in the TiO2@SnS2 nanocomposites. After their characterization, the TiO2@SnS2 core–shell nanocomposites were explored for their catalytic activity towards the photodegradation of a toxic dye (crystal violet (CV)) in an aqueous solution under sunlight. The photodegradation efficiency of TiO2@SnS2 core–shell nanocomposites is better than that of TiO2 microspheres, SnS2 nanoparticles and other metal sulfide nanoparticles/nanocomposites reported in the literature.
1 Introduction
Water pollution from dyes threatens ecosystems across the globe. As human population and economies expand, water pollution has become one of the major issues. The unavailability of sufficient freshwater not only affects human life but it also affects biodiversity and other ecosystems.1 Textile industries produce organic contaminants such as synthetic dyes in high amounts. The dyes do not bind tightly to the fabric and about 20% of them are discharged without any prior treatment into water systems and cause pollution.2 The toxic dyes are often stable, carcinogenic and poisonous.3 In the past few decades, scientists have been working on different methods for the treatment of contaminated water. Photocatalysis is widely explored for the degradation of toxic dyes. It has advantages such as use of sunlight, low cost and being an environment friendly process.4
Photocatalysts based on metal oxide nanoparticles such as TiO2,5 MoO3−x,6 Cu2O,7 WO3,8 and NiCo2O4,9 and metal sulfide nanoparticles such as NiS,10 MoS2,11 CuS,12 and SnS2
13 have been explored as catalysts for the degradation of toxic dyes. Several core–shell nanocomposites such as CuS/ZnO,14 ZnO@ZnS,15 CoFe2O4@CoWO4,16 MoO3@MoS2,17 CdS@ZnO,18 TiO2@CoFe3O4,19 WO3@TiO2,20 ZnIn2S4@CeO2,21 ZnO@SnS2,22 and α-Fe2O3@SnO2/Ti3C2
23 have been explored for the photodegradation of dyes such as methylene blue, congo red, rhodamine B, methyl orange, malachite green, etc. and they show better activity than pure metal sulfide and metal oxide nanoparticles.
Different SnS2 based nanocomposites are reported in the literature such as SnS2/chitosan, synthesized via the precipitation method, for photodegradation of crystal violet,24 SnS2/biochar nanocomposite, synthesized via the hydrothermal route, for photodegradation of amoxycillin and congo red,25 SnS2/BiVO4 nanocomposite, synthesized via the hydrothermal route, for photocatalytic degradation of ciprofloxacin,26 SnS2/CdO nanocomposite, synthesized via the precipitation method, for photodegradation of RhB and congo red,27 SnS2/Ag3VO4 nanocomposite, synthesized via the hydrothermal method, for photodegradation of methylene blue,28 SnS2/NiCo-LDH nanocomposite, synthesized via the co-precipitation method, for photodegradation of thiamethoxam,29 polyaniline/SnS2 nanocomposite, synthesized by the hydrothermal method, for hydrogen evolution,30 PVDF/SnS2 nanocomposite, synthesized via the hydrothermal method, for piezoelectric energy harvesting,31 CeO2/SnS2/polyaniline nanocomposite, synthesized via the hydrothermal method, for photocatalytic reduction of Cr(VI),32 PbS
:
SnS2 nanocomposite, synthesized via the precipitation method, for ethanol sensing,33 Pd/SnS2/SnO2 nanocomposite, synthesized via the hydrothermal method, for detection of H2,34 Au/SnS2 nanocomposite, synthesized via chemical reduction method, as a NO2 gas sensor,35 MoS2/SnS2 nanocomposite, synthesized via the hydrothermal method, as an anode in sodium ion batteries,36 SnS2/CuO nanocomposite, synthesized by the precipitation method, in broadband photodetectors,37 BiOI@SnS2 core–shell nanocomposite, synthesized via the co-precipitation method, for photodegradation of RhB,38 and MnFe2O4@SnS2 core–shell nanocomposite, synthesized via the hydrothermal method, for photodegradation of methylene blue.39
TiO2 is a semiconductor with a wide band gap (Eg (bulk) = 3.2 eV).40 It is a well-known semiconductor for environmental remediation and it exhibits high recombination of photogenerated electrons and holes leading to its low photocatalytic activity. SnS2 is economical and non-toxic with good thermal stability.41,42 It has a narrow band gap of about 1.9 eV
43 and is known to exhibit good photocatalytic activity.42,44 To improve the photocatalytic activity of TiO2 further, formation of a nanocomposite of TiO2 with SnS2 has been proposed. Both the valence band (VB) and conduction band (CB) of SnS2 lie below those of TiO2. Due to this, during light irradiation, the photogenerated electrons from the CB of SnS2 are easily transferred to that of TiO2. This enhances the separation of electrons and holes and improves the photocatalytic efficiency.45 TiO2–SnS2 nanocomposites are useful for various applications such as photodegradation of dyes such as methyl orange, methylene blue and RhB,46–48 photodegradation of antibiotics such as tetracycline,49 NO2 gas sensors,50,51 humidity sensors,52,53 protection of stainless steel,54 photodetectors,55 H2 gas production,56,57 biosensors,58 photoreduction of Cr(VI),45 photocatalytic H2 evolution,59 anode material in lithium-ion batteries,60 photodegradation of diclofenac,61 self-decontaminating textiles,62 photoelectrochemical water splitting,63 photodegradation of diethyl sulfide,64 electrocatalytic oxygen evolution,65 and photocatalytic reduction of CO2.66
In the current study, TiO2@SnS2 core–shell nanocomposites have been synthesized. The novelty of this work is as follows. In the literature, TiO2–SnS2 nanocomposites have been reported (e.g., yolk–shell SnS2–TiO2 microsphere composite67 and SnS2@TiO2 double shell nanocomposite47) but the core–shell structure of the nanocomposite (TiO2@SnS2) has not been previously reported. The thermal decomposition method has been used for the synthesis of different metal sulfide nanoparticles and their nanocomposites. However, the synthesis of a TiO2@SnS2 nanocomposite has not been previously reported using this approach. The reported methods used for the synthesis of TiO2–SnS2 nanocomposites in the literature (solvothermal,50 hydrothermal,52 chemical bath deposition,55 and sol–gel method61) have disadvantages such as use of high pressure (50 bar to 2 kbar) and prolonged reaction time (12 h to 16 h). For example, the sol–gel method requires a reaction time of up to 7 days, and the chemical bath deposition method is a two-step process and often requires an inert environment. The current study reports the synthesis of TiO2@SnS2 core–shell nanocomposites using a thermal decomposition method with a shorter synthesis time (60 min) without the need of an inert environment. The TiO2@SnS2 core–shell nanocomposites were first characterized using various techniques and then they were explored for their photocatalytic activity towards the photodegradation of CV.
2 Experimental
2.1 Materials
Titanium isopropoxide (TIP) (97%, Sigma-Aldrich), SnCl4·5H2O (98%, Sigma-Aldrich), thiourea (98%, SRL Chemicals), methanol (MeOH) (99%, Rankem), acetonitrile (99.5%, Rankem), dodecylamine (DDA) (99%, Rankem), sodium hydroxide (99%, Rankem), diphenyl ether (DPE, 99%, Sigma-Aldrich) and crystal violet (CV, 97%, Sigma-Aldrich) were used.
2.2 Preparation of materials
2.2.1 Synthesis of TiO2 microspheres.
TiO2 microspheres were prepared using a reported wet chemical method.68 Typically, 0.18 mL of H2O and 0.345 mL of dodecyl amine (DDA) were taken in a mixture of solvents (100 mL of methanol and 50 mL of acetonitrile) and the contents were stirred for 10 min. After this, 1 mL of titanium isopropoxide was added and the reaction mixture was further stirred for about 6 h at room temperature. The contents were centrifuged, and the collected product was washed with MeOH and dried in an oven at 60 °C for 12 h to obtain as-prepared TiO2 microspheres. Dodecylamine acts as a surfactant during the synthesis of TiO2 microspheres. Its hydrophilic amine group interacts with polar solvent molecules (methanol), while the hydrophobic tail interacts with non-polar titanium isopropoxide. During the hydrolysis and condensation of TIP, it forms TiO2 nuclei. The hydrophobic tail of DDA binds to TiO2 particles via van der Waals forces preventing agglomeration. This stabilizes the dispersion, lowers surface energy, and leads to controlled nucleation and growth, resulting in uniform TiO2 microspheres.69,70 The as-prepared TiO2 microspheres were calcined at 500 °C for 3 h (heating rate = 2° min−1) to obtain calcined TiO2 microspheres. The calcined TiO2 microspheres were surface-modified using NaOH.71 For this, 100 mg of calcined TiO2 microspheres were added to 10 mL of 5 M NaOH aqueous solution and left at room temperature for about 24 h. The contents were centrifuged, washed with de-ionised water and with 0.1 N HCl solution. The sample was dried in an oven at 60 °C for 12 h to obtain NaOH-modified TiO2 microspheres (TiO2–NaOH).
Zeta potential measurements were done for calcined TiO2 and NaOH-modified TiO2. The surface charge as measured from the zeta potential analyzer was +10.1 mV and −11.8 mV for TiO2-calc. and TiO2–NaOH, respectively. The change in surface charge from positive to negative indicates successful surface modification of TiO2 microspheres with NaOH.
2.2.2 Synthesis of TiO2@SnS2 core–shell nanocomposites.
To synthesize TiO2@SnS2 core–shell nanocomposites, 100 mg of TiO2–NaOH microspheres were taken in 10 mL of DPE in a 50 mL RB flask. The contents were sonicated for one min followed by addition of x mmol of SnCl4·5H2O and 2x mmol of thiourea (Table 1). After refluxing the reaction mixture for about 60 min in air at about 200 °C, the contents were allowed to cool to RT, followed by the addition of about 20 mL of MeOH. The obtained brown precipitate was centrifuged and washed (2–3 times) with methanol. The precipitate was dried in a vacuum desiccator overnight to obtain the TiO2@SnS2 core–shell nanocomposite. Pure SnS2 nanoparticles were also synthesized using a similar synthetic approach in the absence of TiO2–NaOH microspheres.
Table 1 Synthetic details of TiO2@SnS2 core–shell nanocomposites and their sample codes
| Sample code |
TiO2–NaOH (mg) |
SnCl4·5H2O (mmol) |
Thiourea (mmol) |
Solvent |
Reaction temperature (°C) |
Reaction time (min) |
| SnS2 |
— |
0.50 |
1.0 |
Diphenyl ether |
200 ± 2 |
60 ± 5 |
| TiO2–SnS2-0.05 |
100 |
0.05 |
0.1 |
Diphenyl ether |
200 ± 2 |
60 ± 5 |
| TiO2–SnS2-0.1 |
100 |
0.10 |
0.2 |
Diphenyl ether |
200 ± 2 |
60 ± 5 |
| TiO2–SnS2-0.15 |
100 |
0.15 |
0.3 |
Diphenyl ether |
200 ± 2 |
60 ± 5 |
2.3 Photocatalytic studies
Crystal violet is toxic and carcinogenic with a complex structure and is widely used in textiles, microbiological staining, printing, etc., and leads to water pollution. CV is a cationic triphenylmethane dye which is highly resistant to natural degradation.
All the photodegradation experiments were done under sunlight irradiation between 1 PM and 2 PM at Indian Institute of Technology Roorkee, Roorkee, India between March 2024 and May 2024. The solar irradiance data were obtained using the average solar data sourced from ProfileSolar.72 The average intensity of the sun during spring (March–May) is about 6.62 kWh per m2 per day, i.e., 276 W m−2 h−1. The photocatalytic experiments were done by taking 5 mg each of TiO2@SnS2 core–shell nanocomposites in 5 mL of CV aqueous solution (2 × 10−5 M) in a test tube followed by sonication for one minute. The adsorption–desorption equilibrium was achieved by keeping the mixture in the dark for about 30 min. After that, the contents were exposed to sunlight for 60 min. The contents were centrifuged to remove the photocatalyst and the supernatant solutions were analysed using UV-Vis spectroscopy. A blank experiment (i.e., no catalyst) was also carried out to understand the self-degradation of crystal violet under sunlight. The % degradation of crystal violet was estimated using the following formula:
| Degradation% = (1 − Ct/C0) × 100 |
where
C0 and
Ct denote the concentration of crystal violet at adsorption equilibrium and at any irradiation time ‘
t’, respectively.
2.4 Characterization
Powder X-ray diffraction (XRD) measurements were carried out for phase identification using a Rigaku Miniflex 600 powder X-ray diffractometer (Cu-Kα, λ = 1.5406 Å). A Thermo Nicolet FT-IR spectrophotometer was used for recording FT-IR spectra of the samples in the range of 4000–400 cm−1 using KBr pellets. For thermogravimetric analysis (TGA), a PerkinElmer Pyris Diamond instrument was employed in the temperature range of 30–1000 °C (heating rate = 10 °C min−1) in air. The morphological and EDX analyses of the samples were done using a Thermofisher Apreo field emission SEM (operating voltage = 20 kV). TEM measurements were done using a Thermofisher TALOS F200X transmission electron microscope (operating voltage = 200 kV). For the TEM analysis, 1 mg of each sample was taken in a test tube to which 5 mL of ethanol was added. The mixture was sonicated for about 45 min and then drop-cast over a copper grid followed by drying overnight at room temperature. HRTEM images and SAED patterns of the nanocomposites were recorded using the same instrument. EDS line scan and elemental mapping of the nanocomposites were recorded using a JEOL JEM-3200FS transmission electron microscope (operating voltage = 300 kV; energy resolution = 127 eV). The measurements were performed by carefully optimizing the acquisition parameters, such as beam current and dwell time, to minimize beam effects and no significant beam-induced sample damage was observed during the EDS measurements. The oxidation states of the elements present in the TiO2@SnS2 core–shell nanocomposites were determined using a Physical Electronics PHI 5000 Versa Probe III XPS spectrometer using Al-Kα radiation (λ = 1486.6 eV). The optical studies on the core–shell nanocomposites were done using a Cary 5000 UV-Vis-NIR spectrophotometer in the reflectance mode in the wavelength range 300 nm to 800 nm. A diffuse reflectance accessory was used for this purpose. About 10 mg of each sample powder was used to measure the reflectance spectra. The specific surface area of the samples was measured via N2 physisorption at 77 K on a Nova 2200e Quantachrome instrument using the BET method. A Shimadzu UV-2600 UV-Vis spectrophotometer was used for recording absorption spectra during the photodegradation studies. Photoluminescence (PL) spectral measurements were made using a Fluoromax-4 spectrofluorometer at room temperature. 1 mg of sample powder was dispersed in about 3 mL of deionized water by sonication and PL spectra were recorded for the dispersion with an excitation wavelength of 450 nm.
3 Results and discussion
3.1 X-ray diffraction
The powder XRD patterns of calcined TiO2 (before NaOH modification) and TiO2 (after modification with NaOH) are displayed in Fig. S1. The XRD patterns of both the samples show peaks at similar 2θ values and match with that of anatase TiO2 (JCPDS file no. 84-1286) indicating that there is no change in the TiO2 phase after modification with NaOH. The XRD patterns of TiO2–NaOH, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites are displayed in Fig. 1. The XRD pattern of TiO2–NaOH microspheres shows peaks at 2θ = 25.39°, 36.96°, 37.93°, 38.76°, 48.02°, 53.94°, 55.19°, 62.74°, 68.93°, 70.39°, 75.14° and 76.2° which correspond to (101), (103), (004), (112), (200), (105), (211), (204), (116), (220), (215) and (301) planes, respectively, of the anatase phase of TiO2. The XRD pattern of SnS2 nanoparticles shows peaks at 2θ = 28.32°, 32.16°, 41.83°, 45.94°, 50.04°, 52.51°, 58.48°, 60.76°, 63.05°, 67.30° and 70.48° indexed to (100), (011), (012), (003), (110), (111), (200), (201), (004), (202) and (113) planes, respectively, of the Berndtite-2T phase of hexagonal SnS2 (JCPDS file no. 83-1705). The powder XRD patterns of TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15) show peaks due to the anatase phase of TiO2 and also show peaks at 2θ = 28.59° and 50.35° due to (100) and (110) planes of the Berndtite-2T phase of hexagonal SnS2. It is known that the crystallite coherence and stacking faults of nanoparticles may influence the activity.73 The crystallite size of SnS2 nanoparticles was estimated from the XRD patterns of SnS2, TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15 and the values are 11.6 nm, 10.2 nm, 8.8 nm and 8.4 nm, respectively. The crystallite size of TiO2 was estimated from the XRD patterns of TiO2–NaOH, TiO2–SnS2-0.05, TiO2–SnS2-0.1, and TiO2–SnS2-0.15 and the values are 16.2 nm, 15.9 nm, 15.1 nm, and 14.4 nm, respectively.
 |
| | Fig. 1 XRD patterns of TiO2–NaOH, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites. | |
The dislocation density and crystallite strain for TiO2 and SnS2 were estimated for the TiO2@SnS2 core–shell nanocomposites. Dislocation density (ρ) is defined as the number of dislocation lines per unit volume and is estimated using eqn. (1).74
where
D is the crystallite size.
The dislocation density values for SnS2 in SnS2, TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15 are 0.0074 nm−2, 0.0096 nm−2, 0.0129 nm−2, and 0.014 nm−2, respectively. The dislocation density values for TiO2 in TiO2–NaOH, TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15 are 0.0038 nm−2, 0.0039 nm−2, 0.0043 nm−2, and 0.0048 nm−2, respectively. The dislocation density of both TiO2 and SnS2 lattices is small. This suggests a high degree of crystallinity of TiO2–NaOH, SnS2, and TiO2@SnS2 core–shell nanocomposites. Also, the dislocation density increases as the thickness of the shell in the core–shell nanocomposites increases. Crystallite strain is defined as the small distortion in the crystal lattice due to defects or dislocations, causing a slight variation in atomic positions. Due to this, broadening of XRD peaks is observed without change in peak positions. The crystallite strain was estimated using eqn. (2).75
where
ε is the crystallite strain (or microstrain) and is determined from the slope of linear fit of the plot between
β![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
cos
θ and 4
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
sin
θ, commonly known as the Williamson–Hall plot. Fig. S2 provides the Williamson–Hall plots for TiO
2–NaOH and TiO
2@SnS
2 core–shell nanocomposites. The crystallite strain (
ε) values for TiO
2–NaOH, TiO
2–SnS
2-0.05, TiO
2–SnS
2-0.1 and TiO
2–SnS
2-0.15 are 0.00378, 0.00395, 0.00432, and 0.00468, respectively. There is an increase in crystallite strain as shell (SnS
2) thickness increases. The presence of more number of dislocation lines/defects (dislocation density) leads to higher microstrain in the lattice.
76,77 The crystallite strain for SnS
2 could not be estimated as there are only two peaks due to SnS
2 in the XRD patterns of TiO
2@SnS
2 core–shell nanocomposites.
The TiO2 microspheres (as prepared) were calcined at 750 °C and XRD analysis of the sample was performed to confirm the phase of TiO2. The XRD pattern of TiO2 calcined at 500 °C matches with the anatase phase of TiO2 (JCPDS file no. 84-1286) whereas the XRD pattern of TiO2 calcined at 750 °C matches with the rutile phase of TiO2 (JCPDS file no. 04-0551) (Fig. S3). This indicates that phase transformation of TiO2 from anatase into rutile takes place between 650 °C and 800 °C.
3.2 Functional group analysis
The FT-IR spectra of unmodified calcined TiO2 microspheres and TiO2–NaOH are given in Fig. S4. The intensities of IR bands due to OH stretching and bending are more intense in the case of TiO2–NaOH microspheres compared to unmodified TiO2 indicating the successful modification of calcined TiO2 microspheres with NaOH. The FT-IR spectra of TiO2–NaOH, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites are displayed in Fig. S5. The IR band positions and their assignments for all the samples are given in Table S1. The IR spectrum of TiO2–NaOH microspheres shows bands at 3428 cm−1 and 1637 cm−1 due to the stretching and bending vibration of physisorbed water molecules.78 The characteristic band due to Ti–O vibration is observed at about 624 cm−1.79 The IR spectra of SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites show vibrational bands at about 3397 cm−1 and 1638 cm−1 assigned to OH stretching and bending of physisorbed water molecules.80 The IR bands observed at about 2925 cm−1, 2852 cm−1 and 1384 cm−1 are attributed to asymmetric stretching, symmetric stretching and bending vibration of C–H bond, respectively.81 The IR band at about 1178 cm−1 corresponds to C–N stretching.82 The IR spectrum of SnS2 nanoparticles shows an additional band at 862 cm−1 assigned to C–H bending.83 The band at about 611 cm−1 is assigned to Sn–S stretching.84
3.3 Thermogravimetric analysis
The TGA curves of TiO2–NaOH, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites are given in Fig. S6. The DTG plot for TiO2–NaOH is given in Fig. S7. The TiO2–NaOH microspheres show a two-step weight loss of about 3% in the temperature range of 30–1000 °C. The first weight loss of ∼1.3% below 350 °C is due to the loss of physisorbed water molecules.85 The second weight loss of about 1.7% between 650 °C and 800 °C is due to the loss of chemically bonded OH groups from the surface of TiO2 (dehydroxylation) and conversion of anatase TiO2 to rutile phase.68 The TGA curve of SnS2 nanoparticles shows a three-step weight loss pattern. The first weight loss of about 3% below 230 °C is due to the loss of surface-adsorbed water molecules. The second weight loss of about 48.5% between 230 °C and 440 °C is attributed to the removal of organic moieties. The third weight loss of about 15.5% is due to oxidation of SnS2 into SnO2.86 The TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15) show two-step weight loss patterns. The first weight loss of ∼3.0%, 3.1% and 3.8%, respectively, observed below 440 °C is due to the loss of physisorbed water molecules and removal of organic moieties. The second weight loss of about 1.9%, 3.4%, and 3.6%, respectively, between 440 °C and 530 °C is attributed to oxidation of SnS2 into SnO2.87 The TGA studies indicate a total weight loss of about 4.9%, 6.5% and 7.4% in TiO2–SnS2-0.05, TiO2–SnS2-0.1, and TiO2–SnS2-0.15. In the core–shell nanocomposites, the % weight loss due to TiO2 is negligible and it is primarily due to the presence of SnS2 in the TiO2@SnS2 core–shell nanocomposites. In the TiO2@SnS2 core–shell nanocomposites, an increase in weight loss is observed with increase in the concentration of SnS2 nanoparticles (shell) on the TiO2 core.
3.4 FESEM and EDX analyses
The coating of SnS2 nanoparticles over the TiO2 microspheres was first done using unmodified as-prepared TiO2 (i.e., before calcination) and unmodified calcined TiO2 microspheres. Fig. S8 shows the FESEM images of TiO2@SnS2 nanocomposites synthesized using unmodified as-prepared TiO2 and unmodified calcined TiO2 microspheres. The EDX results (Table S2) give distribution of different elements (Ti, O, Sn and S) in the thus prepared TiO2@SnS2 nanocomposites. Both FESEM and EDX results of TiO2@SnS2 nanocomposites, synthesized using as-prepared TiO2 and unmodified calcined TiO2 microspheres, indicate non-uniform coating of SnS2 nanoparticles on the TiO2 microspheres. Hence, the synthesis of TiO2@SnS2 core–shell nanocomposites was attempted using NaOH-modified TiO2 microspheres. The SEM images of TiO2–NaOH, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites, synthesized using TiO2–NaOH microspheres as the core, are given in Fig. S9. The FESEM image of TiO2–NaOH microspheres displays spherical particles with a diameter of about 471 ± 17 nm. The FESEM image of SnS2 nanoparticles shows flake-like morphology. The FESEM images of TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15) show that SnS2 nanoparticles are uniformly coated over the TiO2–NaOH microspheres. The EDX results (Table S3) indicate uniform distribution of different elements (Ti, O, Sn, and S) in all the core–shell nanocomposites. In the TiO2@SnS2 core–shell nanocomposites, an increase in weight% of Sn and S is observed with an increase in [Sn4+]
:
[S2−] ratio used during their synthesis.
3.5 TEM analysis
The TEM micrographs of TiO2–NaOH, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites are displayed in Fig. 2. The TiO2–NaOH microspheres show spherical morphology with a diameter of about 488 nm. The SnS2 nanoparticles show flake-like morphology. The flake thickness was estimated using the TEM image. The particle size histogram is shown in Fig. S10, and the mean thickness is about 52 ± 12 nm. The TEM micrographs of TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15) show uniform coating of SnS2 nanoparticles on the TiO2 microspheres. Table 2 summarizes the particle diameter, core size and shell thickness observed from the TEM micrographs of TiO2@SnS2 core–shell nanocomposites. The overall diameter of TiO2@SnS2 core–shell nanocomposites varies from 457 nm to 503 nm. The core diameter in the TiO2@SnS2 core–shell nanocomposites varies from 434 nm to 489 nm. The thickness of the shell (SnS2 nanoparticles) over the TiO2 microspheres varies from 14 nm to 26 nm. This variation of thickness of the shell is due to the synthetic conditions. Use of different amounts of SnCl4·5H2O and thiourea during the synthesis of TiO2@SnS2 core–shell nanocomposites leads to variation in shell thickness. The EDS spectrum and elemental analysis data (from TEM measurements) for TiO2–SnS2-0.15 are given in Fig. S11 and Table S4, respectively. The TEM EDS analysis was performed at three different spots and the results confirm the uniform presence of Ti, Sn, S, and O in the core–shell nanocomposites (TiO2–SnS2-0.15). The EDS line scanning data and elemental mapping images (from SEM) of TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.15) are shown in Fig. 3. The EDS line scan (Fig. 3(a)) indicates that the concentration of titanium and oxygen at the edge is less and that of tin and sulfur is high. On moving from the edge to the centre of the particle, the concentration of Ti and O increases and that of Sn and S decreases. The elemental mapping results (Fig. 3(b)) also indicate that Sn and S are predominantly located at the edges of TiO2 microspheres. On the other hand, Ti and O are present more at the centre. This indicates the successful formation of TiO2@SnS2 core–shell nanocomposites.
 |
| | Fig. 2 TEM images of TiO2–NaOH, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites. | |
Table 2 Summary of results from TEM studies of TiO2@SnS2 core–shell nanocomposites
| Sample code |
Particle diameter (nm) |
Flake thickness (nm) |
Core size (nm) |
Shell thickness (nm) |
| TiO2–NaOH |
488 |
— |
— |
— |
| SnS2 |
— |
52 ± 12 |
— |
— |
| TiO2–SnS2-0.05 |
457 |
— |
434 |
14 |
| TiO2–SnS2-0.1 |
478 |
— |
456 |
21 |
| TiO2–SnS2-0.15 |
503 |
— |
489 |
26 |
 |
| | Fig. 3 (a) EDS line scan and (b) elemental mapping data of TiO2–SnS2-0.15 core–shell nanocomposite. | |
The SAED patterns of TiO2–NaOH microspheres, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites are displayed in Fig. S12. The SAED pattern of TiO2–NaOH microspheres shows rings that correspond to the planes of anatase TiO2 (JCPDS file no. 84-1286). The rings, observed in the SAED pattern of SnS2 nanoparticles, correspond to the planes of SnS2 (JCPDS file no. 83-1705). The SAED patterns of TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15) show rings attributed to reflections due to both TiO2 and SnS2.
The HRTEM images of TiO2–NaOH microspheres, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites are given in Fig. S13. The observed lattice fringes in the HRTEM image of TiO2–NaOH with a d spacing of 3.49 Å are attributed to the (101) plane of TiO2 (JCPDS file no. 84-1286). The lattice fringes in the HRTEM image of SnS2 nanoparticles with a d spacing of 3.15 Å is attributed to the (100) plane of SnS2 (JCPDS file no. 83-1705). The d spacing values (estimated from the lattice fringes) of TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15) are 3.49 Å and 3.15 Å attributed to the (101) plane of TiO2 and (100) plane of SnS2, respectively. The specific surface area of TiO2–NaOH microspheres, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15), as measured by the BET method, are 9 m2 g−1, 17 m2 g−1, 10 m2 g−1, 13 m2 g−1, and 19 m2 g−1, respectively.
3.6 Optical properties
The optical properties of TiO2–NaOH microspheres, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites were studied using UV-Vis DRS spectroscopy (Fig. 4) and the Tauc plots (Fig. S14) were used to estimate the band gap. The UV-Vis DRS measurements were done three times for each sample. Table S5 includes error bars for the band gap.88 The band gap of TiO2–NaOH microspheres and SnS2 nanoparticles are 3.1 eV and 2.03 eV, respectively. The TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15) have band gaps of 2.23 eV, 2.18 eV and 2.15 eV, respectively. The TiO2@SnS2 core–shell nanocomposites exhibit band gap only due to SnS2 suggesting that the TiO2 microspheres are coated with a thick shell of SnS2 nanoparticles. The band gap values were also estimated using K–M plots89 (Fig. S15) and the values are summarized in Table S5. Although the band gap of a semiconductor can be estimated using both Tauc plots and K–M plots, the band gap values estimated using K–M plots can often be misleading. This is because the K–M function is proportional to the ratio of the absorption and scattering coefficients (ɑ) and is considered as an approximate. Furthermore, the K–M function does not consider the nature of the electronic transition (direct or indirect in a semiconductor).90
 |
| | Fig. 4 UV-Vis DRS spectra of TiO2–NaOH, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites. | |
3.7 PL spectroscopy
The PL spectra of TiO2@SnS2 core–shell nanocomposites (λexc = 450 nm) are shown in Fig. 5. The PL spectra of TiO2–NaOH microspheres, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites exhibit an emission peak at ∼590 nm. This emission peak is due to the recombination of electrons with holes that are trapped above the valence state.91–93 The emission peak at about 590 nm in SnS2 nanoparticles is due to the near band edge emission of the SnS2 phase.94 The order of PL intensity in the samples is: TiO2–NaOH > SnS2 > TiO2–SnS2-0.05 > TiO2–SnS2-0.1 > TiO2–SnS2-0.15. The PL intensity is the least for TiO2–SnS2-0.15 and it is expected to show the highest photocatalytic activity among all the samples.
 |
| | Fig. 5 PL spectra of TiO2–NaOH, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites. | |
3.8 XPS analysis
The oxidation states of different elements present in TiO2–NaOH microspheres, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposite (TiO2–SnS2-0.15) were determined by XPS analysis, and the results are given in Fig. 6. Table S6 summarizes binding energies of all the elements (Ti, O, Sn and S) present in the samples. As shown in the Ti 2p spectrum of TiO2–NaOH (Fig. 6(a)), the peaks at 456.2 eV and 462.3 eV correspond to Ti 2p3/2 and Ti 2p1/2 of Ti4+, respectively. The O 1s spectrum of TiO2–NaOH shows peaks at 530.9 eV and 532.1 eV due to lattice oxygen (O2−) and surface hydroxyl (OH) oxygen, respectively.95,96 The Sn 3d spectrum of SnS2 nanoparticles (Fig. 6(b)) shows peaks at 486.0 eV and 494.4 eV due to Sn 3d5/2 and Sn 3d3/2 of Sn4+, respectively. The S 2p spectrum of SnS2 nanoparticles shows peaks at 165.0 eV and 166.2 eV corresponding to S 2p3/2 and S 2p1/2, respectively.97 The XPS spectrum of the TiO2@SnS2 core–shell nanocomposite (TiO2–SnS2-0.15) is shown in Fig. 6(c). The Ti 2p spectrum of TiO2–SnS2-0.15 shows peaks at 458.0 eV and 463.7 eV attributed to Ti 2p3/2 and Ti 2p1/2 of Ti4+, respectively. The O 1s spectrum of TiO2–SnS2-0.15 shows peaks at 531.0 eV and 532.3 eV assigned to lattice oxygen and surface hydroxyl oxygen, respectively.95,96 The Sn 3d spectrum of TiO2–SnS2-0.15 shows peaks at 485.9 eV and 494.3 eV due to Sn 3d5/2 and Sn 3d3/2 of Sn4+, respectively. The S 2p spectrum of TiO2–SnS2-0.15 shows peaks at 164.9 eV and 166.2 eV due to S 2p3/2 and S 2p1/2, respectively.97
 |
| | Fig. 6 XPS spectra of (a) TiO2–NaOH microspheres, (b) SnS2 nanoparticles and (c) TiO2@SnS2 core–shell nanocomposite (TiO2–SnS2-0.15). | |
There is a minor decrease in Sn 3d and S 2p binding energy values on going from pure SnS2 nanoparticles to the TiO2@SnS2 core–shell nanocomposites. Also, a minor increase in Ti 2p and O 1s binding energies is observed on going from pure TiO2 microspheres (TiO2–NaOH) to the core–shell nanocomposites. The shifts in binding energies in the XPS spectrum of the TiO2@SnS2 core–shell nanocomposite could be due to electron transfer from the core (TiO2) to the shell (SnS2) during the formation of the core–shell nanocomposite. The electron transfer leads to a partial positive charge on TiO2, leading to an increase in binding energies of Ti 2p and O 1s and a partial negative charge on SnS2, leading to a decrease in binding energies of Sn 3d and S 2p.98–100
To investigate the charge transfer and band (VB and CB) positions in TiO2@SnS2 core–shell nanocomposites, XPS-VB (UPS spectra) measurements were performed. Fig. S16 provides the XPS-VB spectra of TiO2–NaOH microspheres, SnS2 nanoparticles and TiO2–SnS2-0.15. The valence band positions estimated from the XPS-VB spectra for TiO2–NaOH microspheres, SnS2 nanoparticles and TiO2–SnS2-0.15 are 2.83 eV, 1.35 eV and 1.78 eV, respectively. In the TiO2@SnS2 core–shell nanocomposites, electron transfer occurs from the core (TiO2) to the shell (SnS2). The intermediate valence band position of TiO2–SnS2-0.15 suggests successful heterostructure formation and interfacial charge transfer between TiO2 and SnS2.101,102 From the optical studies (DRS spectra), the band gaps for TiO2–NaOH and SnS2 are 3.1 eV and 2.02 eV, respectively. On applying the equation, ECB = EVB − Eg, the conduction band positions of TiO2–NaOH and SnS2 nanoparticles are −0.27 eV and −0.67 eV, respectively. To further confirm the charge transfer in TiO2@SnS2 core–shell nanocomposites, EIS measurements were also performed (see Section 3.15 for more details).
3.9 Mechanism of formation of TiO2@SnS2 core–shell nanocomposites
Scheme 1 illustrates the proposed mechanism of formation of TiO2@SnS2 core–shell nanocomposites. At first, on surface modification with NaOH, hydroxyl groups are attached on the surface of TiO2 microspheres. This can be evidenced from the IR spectral studies (Fig. S4). Due to the presence of hydroxyl groups on the surface of TiO2 microspheres, Sn4+ ions are attached on the surface of TiO2via electrostatic interaction. This is followed by the thermal decomposition of thiourea to form H2S and the formation of TiO2@SnS2 core–shell nanocomposites. The reactions involved in the formation of SnS2 nanoparticles over TiO2–NaOH microspheres are given in eqn (3) and (4).103–105| |  | (3) |
 |
| | Scheme 1 Mechanism of formation of TiO2@SnS2 core–shell nanocomposites. | |
SnS2 nanoflakes are formed by the thermal decomposition of SnCl4·5H2O and thiourea in DPE. The formation of nanoflakes is a two-step process, i.e., nucleation followed by growth.106 At first, tin chloride reacts with thiourea leading to the formation of a small cluster of SnS2 (nuclei). The presence of chloride ions leads to oriented growth of particles. The chloride ions adsorbed on the transverse planes of SnS2 particles inhibit growth in those directions. As a result, the growth occurs in the longitudinal axis leading to the formation of SnS2 nanoflakes.82
3.10 Photocatalytic activity
The photocatalytic activity of TiO2@SnS2 core–shell nanocomposites was studied using crystal violet as the dye pollutant (see Section 2.3 for more details). As indicated by the UV-Vis spectral results (Fig. 7) on the photodegradation experiments, TiO2–SnS2-0.15 achieved about 99% degradation of crystal violet within 60 minutes under sunlight. TiO2–NaOH microspheres, SnS2 nanoparticles, TiO2–SnS2-0.05, and TiO2–SnS2-0.1 degrade about 53.1%, 60.2%, 61.5%, and 78.2% of crystal violet, respectively. This indicates that TiO2–SnS2-0.15 is a better photocatalyst for the degradation of crystal violet compared to TiO2–NaOH microspheres, SnS2 nanoparticles and other core–shell nanocomposites. The self-degradation of crystal violet under sunlight, tested in the absence of any catalyst, was about 1.9%.
 |
| | Fig. 7 UV-Vis spectral results indicating the photodegradation of CV using TiO2@SnS2 core–shell nanocomposites. | |
Furthermore, dark experiments were carried out in the absence of sunlight to understand adsorption of CV on the TiO2@SnS2 core–shell nanocomposites. In this experiment, 5 mg of each catalyst was taken in a test tube to which 5 mL of CV aqueous solution (2 × 10−5 M) was added. The test tubes were sonicated for one min followed by keeping them in the dark for 30 min/90 min. The UV-Vis absorbance spectra recorded for the supernatant solutions obtained after centrifugation of each mixture are shown in Fig. S17. The % adsorption by TiO2–NaOH microspheres, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15) are 15%, 28.1%, 31.3%, 37.9% and 49%, respectively in 30 min and 26%, 47.9%, 34.8%, 39.9% and 53.8%, respectively in 90 min. The adsorption and photodegradation studies indicate that the TiO2@SnS2 core–shell nanocomposites act as a good photocatalyst for the degradation of CV and not as an adsorbent.
Furthermore, kinetic experiments were carried out by taking 6 test tubes each containing 5 mg of catalyst (TiO2–NaOH microspheres, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15)) and 5 mL of CV aqueous solution (2 × 10−5 M). The mixtures were kept in the dark for 30 min which was followed by irradiation with sunlight for different time periods (10–60 min). UV-Vis spectra were recorded for the supernatant solutions obtained after centrifuging the mixture in each test tube.
The kinetics results are shown in Fig. 8(a) and TiO2–SnS2-0.15 shows maximum degradation in 60 min. The kinetics of photodegradation of CV by TiO2–SnS2-0.15 follows pseudo 1st order kinetics as indicated by −ln(Ct/C0) against irradiation time plots shown in Fig. 8(b). The rate constant values were determined using the kinetic curves. Table 3 summarizes the % adsorption (30 min/90 min), % total degradation (90 min), % photodegradation only, rate constant and R2 values for TiO2–NaOH microspheres, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15). The rate constant values are in the order: TiO2–SnS2-0.15 > TiO2–SnS2-0.1 > TiO2–SnS2-0.05 > SnS2 > TiO2–NaOH indicating that TiO2–SnS2-0.15 is the best photocatalyst among all the samples. The better photocatalytic activity of TiO2–SnS2-0.15 is attributed to a lower rate of recombination of electron and hole pairs in TiO2–SnS2-0.15 as indicated by the PL spectral studies (Fig. 5). The photocatalytic degradation results were compared with those reported in the literature (Table S7).24,107–128 In the literature, more time (70–390 min) is required for the photodegradation of CV. Wherever lesser time is reported (30–45 min),116,117 either the amount of catalyst used is high (250 mg/100 mL) or the % photodegradation is less (83–93%). The TiO2@SnS2 core–shell nanocomposites (e.g., TiO2–SnS2-0.15), reported in the current work, achieved 99% degradation of CV within 60 minutes under sunlight.
 |
| | Fig. 8 (a) (Ct/C0) plots and (b) −ln(Ct/C0) vs. time plots for TiO2–NaOH microspheres, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites as the catalyst for the photodegradation of CV. | |
Table 3 Photocatalytic degradation results of crystal violet using TiO2@SnS2 core–shell nanocomposites
| Sample code |
% Adsorption (30 min) |
% Adsorption (90 min) |
Total degradation% (30 min dark + 60 min sunlight) |
Photodegradation % only (60 min) |
Rate constant for photodegradation (min−1) |
R
2
|
| TiO2–NaOH |
15.0 |
26.0 |
53.1 |
38.1 |
9.1 × 10−3 |
0.98 |
| SnS2 |
28.1 |
47.9 |
60.2 |
32.1 |
9.5 × 10−3 |
0.98 |
| TiO2–SnS2-0.05 |
31.3 |
34.8 |
61.5 |
30.2 |
1.0 × 10−2 |
0.99 |
| TiO2–SnS2-0.1 |
37.9 |
39.9 |
78.2 |
38.3 |
1.7 × 10−2 |
0.99 |
| TiO2–SnS2-0.15 |
49.0 |
53.8 |
99.0 |
50.0 |
6.4 × 10−2 |
0.99 |
3.11 Effect of pH
The photocatalytic degradation efficiency of a catalyst is affected by the pH of the solution. The effect of pH on the photodegradation efficiency of TiO2–SnS2-0.15 was studied by varying the pH of the solution between 2 and 10. The pH of the solutions was varied using 0.1 M HCl and 0.1 M NaOH. The experiment was performed by taking 5 mg of catalyst (TiO2–SnS2-0.15) in 5 mL of CV aqueous solution (2 × 10−5 M) at different pH values (2–10) in different test tubes. The contents were kept in the dark (for 30 min) followed by irradiation with sunlight for 60 min and UV-Vis absorbance spectra were measured for the supernatant solutions obtained after centrifugation. The variation of the photodegradation efficiency of TiO2–SnS2-0.15 with pH is shown in Fig. S18(a). The results indicate that maximum photocatalytic efficiency is observed at pH = 7.
At lower pH, there are more H+ ions in the solution that make the surface of the catalyst positively charged (zeta potential = + 19.0 mV at pH = 4). Crystal violet is a cationic dye, and it will experience electrostatic repulsion towards the surface of the catalyst in an acidic medium leading to a decrease in the photocatalytic efficiency. On the other hand, in basic medium, the presence of more OH− ions in the solution will lead to a negatively charged catalyst surface (zeta potential = −21.4 mV at pH = 10). This results in an electrostatic attraction towards crystal violet leading to higher photodegradation efficiency by the catalyst.129
3.12 Effect of catalyst dosage
To determine the optimum amount of catalyst required to degrade CV using TiO2@SnS2 core–shell nanocomposites, the effect of catalyst dosage on the photodegradation efficiency was studied. The catalyst (TiO2–SnS2-0.15) was taken in different amounts (1–8 mg) in 5 mL of aqueous CV solution (2 × 10−5 M) in test tubes. The mixtures were kept in the dark (for 30 min) followed by sunlight irradiation for 60 min and the UV-Vis absorbance spectra were measured for the supernatant solutions obtained after centrifugation. The UV-Vis spectral results shown in Fig. S18(b) indicate that 5 mg of TiO2–SnS2-0.15 in 5 mL of CV is the optimum catalyst dosage for the photodegradation of CV.130
3.13 Scavenger tests
To understand the mechanism of photodegradation of CV by TiO2–SnS2-0.15, scavenger tests were performed. Different scavengers (isopropyl alcohol (IPA), ammonium oxalate (AO) and p-benzoquinone (PBQ)) were used to trap OH˙, O2˙− and h+, respectively.127,131 In a typical experiment, 10 mM of scavenger was added to 5 mL aqueous solution of CV (2 × 10−5 M) followed by the addition of 5 mg of catalyst (TiO2–SnS2-0.15). The contents were kept in the dark (for 30 min) followed by irradiation with sunlight for 60 min and UV-Vis absorbance spectra were measured for the supernatant solutions obtained after centrifugation. The scavenger test results (Fig. S19) show a decrease of about 15.2%, 39.4% and 53.4% in the presence of IPA, PBQ, and AO, respectively. This indicates that O2˙− and h+ are the major species responsible for the photodegradation of CV using TiO2@SnS2 core–shell nanocomposites as the catalyst.
3.14 Mechanism of photodegradation of crystal violet
Crystal violet degrades via two different pathways, i.e., N-de-methylation and oxidative degradation.106,132 In the first case, reactive oxygen species (ROS) attack the N,N-dimethyl groups of the CV molecules leading to the formation of N-de-methylated CV species. In the second case, the ROS species attack the central carbon of CV facilitated by its conjugated aminotriphenylmethane structure leading to disruption in the conjugation and decolourisation of CV. This step leads to the formation of intermediates such as 4-(N,N-dimethylamino)-4′-(N′,N′-dimethylamino)benzophenone and 4-(N,N-dimethylamino)phenol. Both these steps can occur simultaneously or one-by-one.132 The formed intermediates undergo further oxidation leading to ring opening and complete mineralisation of CV leading to formation of CO2 and H2O.133
Scheme 2 gives the proposed mechanism of photodegradation of crystal violet in the presence of TiO2@SnS2 core–shell nanocomposites as the catalyst. The valence band (VB) and conduction band (CB) positions of TiO2 are located at 2.83 eV and −0.27 eV, respectively.134 The VB and CB of SnS2 are located at 1.35 eV and −0.67 eV, respectively,46 indicating a more negative band potential than that of TiO2 (see Section 3.8, XPS-VB results).
 |
| | Scheme 2 Mechanism of photodegradation of crystal violet using TiO2@SnS2 core–shell nanocomposites as the catalyst. | |
Upon irradiation with sunlight, electrons in the VB of SnS2 are excited to its CB, leaving behind photogenerated holes. Due to the favourable band alignment, the excited electrons in the CB of SnS2 migrate to the CB of TiO2, while the holes in the VB of TiO2 migrate to that of SnS2. This charge transfer effectively suppresses electron–hole recombination, enhancing the photocatalytic activity. The electrons in the CB of TiO2 interact with dissolved O2 molecules to generate superoxide radicals (O2˙−), while the holes in the VB of SnS2 oxidize water molecules to produce OH˙. These highly reactive species then attack and degrade crystal violet molecules into harmless end products (CO2 and H2O).129,135,136
The standard potential for O2/O2˙− is −0.33 V and that for OH−/OH˙ is 2.38 V. The VB of TiO2 (2.83 V) is more positive than that of OH−/OH˙ (2.38 V) indicating that OH˙ could be generated. Similarly, the conduction band of SnS2 (–0.67 V) is more negative than that of O2/O2˙− (−0.33 V) indicating that O2˙− could also be generated. The generation of both OH˙ and O2˙− by TiO2@SnS2 core–shell nanocomposites on solar irradiation leads to the degradation of crystal violet into CO2 and H2O as shown in Scheme 2.
3.15 EIS studies
Electrochemical Impedance Spectroscopy (EIS) measurements were performed for TiO2–NaOH, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15) to understand the charge separation. The samples were deposited onto FTO (fluorine-doped tin oxide) glass substrates (dimensions: 1 cm × 2 cm). For each sample, 30 mg of the sample (TiO2–NaOH, SnS2 nanoparticles and TiO2@SnS2 core–shell nanocomposites) was dispersed in 3 mL of ethanol to form a homogeneous slurry. The mixture was subjected to sonication for 45 minutes to ensure uniform dispersion. Subsequently, the slurry was drop-cast onto the FTO substrates and dried in an oven (at ∼50 °C).
The electrochemical measurements were conducted using a conventional three-electrode system. The modified FTO glass slides (coated with TiO2–NaOH, SnS2 and TiO2@SnS2 core–shell nanocomposites) served as the working electrode. Ag/AgCl electrode was employed as the reference electrode. A platinum wire was employed as the counter electrode. The EIS measurements were performed in 0.1 M KOH aqueous solution as the electrolyte, over a frequency range of 0.1 Hz to 100 kHz (Multi Autolab/M204 electrochemical workstation (Metrohm Autolab B·V., Netherlands)).
Fig. S20 provides the Nyquist plots for all the samples. The order of arc radius is as follows: TiO2–NaOH > SnS2 > TiO2–SnS2-0.05 > TiO2–SnS2-0.1 > TiO2–SnS2-0.15. The arc radius in the Nyquist plot corresponds to charge transfer resistance (Rct) at the interface of the electrode and electrolyte.137 A smaller arc radius indicates lower charge-transfer resistance (Rct), which suggests faster interfacial charge transfer. The smaller arc radius observed in TiO2–SnS2-0.15 among all the samples suggests faster interfacial charge transfer and thus its better photocatalytic efficiency towards degradation of crystal violet compared to the other samples.
3.16 Time resolved PL decay experiments
PL decay experiments were performed using an FLS-1000-xs-t (Edinburgh Instruments) system for TiO2–NaOH microspheres, SnS2 nanoparticles, and TiO2@SnS2 core–shell nanocomposites (TiO2–SnS2-0.05, TiO2–SnS2-0.1 and TiO2–SnS2-0.15) and the data are presented in Fig. S21. The excitation and emission wavelengths were 450 nm and 590 nm, respectively. The PL decay curves were well fitted according to the biexponential function in the form of R(t) = A + B1
exp(−t/τ1) + B2
exp(−t/τ2).
Where, A is a constant background offset, B1 and B2 are amplitude coefficients of the two decay components, τ1 is the lifetime of the faster decay component and τ2 is the lifetime of the slower decay component.
The average lifetime τavg was calculated using the formula τavg= (B1τ12 + B2τ22)/(B1τ1 + B2τ2).138 Table S8 summarizes the fluorescence decay components for all the samples. The average lifetime (τavg) values estimated for TiO2–NaOH microspheres, SnS2 nanoparticles, TiO2–SnS2-0.05, TiO2–SnS2-0.1, and TiO2–SnS2-0.15 are 8.37 ns, 10.82 ns, 10.91 ns, 12.44 ns, and 14.39 ns, respectively. The observed trend in lifetime follows the order: TiO2–SnS2-0.15 > TiO2–SnS2-0.1 > TiO2–SnS2-0.05 > SnS2 >TiO2. TiO2–SnS2-0.15 exhibits a higher average PL decay lifetime indicating a slower recombination rate of photogenerated excitons, indicating more effective charge separation and better photocatalytic efficiency towards the degradation of crystal violet.
3.17 Detection of hydroxyl radicals
To prove the formation of hydroxyl radicals during the photodegradation of CV by TiO2–SnS2-0.15, PL spectral studies were performed using terephthalic acid. Terephthalic acid is non-fluorescent but on reaction with hydroxyl radicals, it produces 2-hydroxyl terephthalic acid, which is fluorescent.139 About 40 mg of catalyst (TiO2–SnS2-0.15) was taken in a beaker containing an aqueous solution of terephthalic acid (5 × 10−4 M) and NaOH (2 × 10−3 M). After sunlight irradiation, PL spectra were recorded at periodic time intervals for the supernatant solutions obtained after centrifugation (λexc = 315 nm). An emission band at 425 nm (Fig. S22) is observed with an increase in PL intensity as a function of time indicating the generation of hydroxyl radicals with time.
3.18 Recyclability and scalability
Photodegradation experiments were performed repeatedly to understand the stability and reusability of TiO2@SnS2 core–shell nanocomposites. In this experiment, an aqueous solution containing 70 mg of catalyst (TiO2–SnS2-0.15) and 70 mL of CV aqueous solution (2 × 10−5 M) was kept in the dark (for 30 min) followed by irradiation with sunlight for 60 min. The UV-Vis absorbance spectrum was measured for the supernatant solution obtained after centrifugation. The catalyst was recovered by centrifugation and washed with methanol several times and finally dried in an oven at 60 °C for 12 h. Five cycles of photocatalytic degradation experiments were carried out using the catalyst recovered after each cycle. The recyclability test results (Fig. S23) show a decrease of about 12.5% in photodegradation efficiency of TiO2–SnS2-0.15 after 5 cycles indicating that TiO2–SnS2-0.15 is a photocatalyst with good recyclability. Furthermore, to determine the stability of the catalyst (TiO2–SnS2-0.15), characterization of the recovered catalyst was performed. Fig. S24 provides the XRD pattern, FT-IR spectrum and FESEM image of the catalyst recovered after the recyclability test. The post-recyclability EDX analysis results are provided in Table S9. The results indicate no change in the phase or morphology of the catalyst after repeated use indicating its high stability.
Nanomaterials have been explored as catalysts for photodegradation of toxic dyes by easy design using different synthetic approaches. Their size, chemical composition and morphology can be easily controlled using optimized synthetic conditions but due to various environmental concerns and economic viability, their large-scale use for environmental remediation is still limited. In the current work, TiO2@SnS2 core–shell nanocomposites demonstrate promising photocatalytic activity under laboratory conditions. The synthetic method and photodegradation results in the current study are reproducible, but large-scale production and long-term stability are unexplored.140
3.19 Potential application of TiO2@SnS2 core–shell nanocomposites as C-2 catalysts
Chemical reactions in which two carbon containing products (e.g., ethanol, ethylene or acetic acid) are formed through C–C coupling reactions are called C-2 reactions. This is particularly relevant for electrocatalytic and photocatalytic CO2 reduction and syngas conversion. For a material to be used as a catalyst in a C-2 reaction, the following properties are important: (a) the presence of active sites that can stabilise key intermediates such as *CO, *CH2 and *CHO, (b) efficient separation of charge carriers (in photocatalysis), and (c) a favourable and stable surface that can promote C–C coupling under the reaction conditions.
Metal sulfide nanoparticles and their nanocomposites have been employed as catalysts for C-2 reactions. For example, ZnIn2S4 nanosheets activated by nitrogen-doped carbon have been used for electrocatalytic CO2 reduction to ethanol.141 SnO2/SnS2/Cu2SnS3 heterojunction has been used for photocatalytic reduction of CO2 to ethanol.142 The TiO2@SnS2 core–shell nanocomposites are potential materials as catalysts for C-2 reactions. They can catalyze C–C coupling reactions in the following ways. The formation of TiO2@SnS2 core–shell nanocomposites builds a type-II heterojunction between TiO2 and SnS2. On solar irradiation, SnS2 absorbs photons and generates an electron–hole pair (see Section 3.14), reducing the recombination of photogenerated excitons and promoting long-lived charge carriers for surface reactions. In the second step, the photogenerated electrons in the CB of TiO2 can be used for the formation of intermediate species such as *CO, *CHO or *CH3 from CO2/CO. SnS2 is well known to stabilize intermediates such as *CO. Furthermore, the C–C coupling can occur in the third step via reductive dimerization of surface-bound *CO or *CH3 on the catalyst's surface. This leads to the formation of C-2 species such as acetaldehyde. In the final step, the photogenerated electrons and protons (typically from water oxidation) can hydrogenate acetaldehyde to produce ethanol.
From XPS results (Section 3.8), it was found that charge transfer takes place from core TiO2 to shell SnS2, indicating interfacial charge redistribution.46 An increase in binding energy is observed for Ti 2p and O 1s, whereas a decrease in binding energy is observed for Sn 3d and S 2p in the TiO2@SnS2 core–shell nanocomposites. This suggests that there is more electron density on the surface of SnS2 after the formation of TiO2@SnS2 core–shell nanocomposites. The electron-rich SnS2 surface is beneficial in stabilizing electrophilic species (C-1) such as CHO and provides them sufficient residence time for C–C coupling reactions and can enhance catalytic activity compared to the individual components.
4 Conclusions
In the present study, TiO2@SnS2 core–shell nanocomposites were synthesized using the thermal decomposition method using different [Sn4+]
:
[S2−] ratios. The core–shell nanocomposites were characterized using several analytical techniques. The XRD results confirmed the presence of both TiO2 and SnS2 phases in the core–shell nanocomposites. FESEM and TEM analyses confirmed uniform coating of SnS2 nanoparticles on the TiO2 microspheres indicating their core–shell structure. The TiO2@SnS2 core–shell nanocomposites demonstrate good catalytic activity towards the photodegradation of CV in aqueous solutions under sunlight. The photocatalytic efficiency of TiO2@SnS2 core–shell nanocomposites is better than that of the individual constituents and other metal-sulfide based nanocomposites that are reported in the literature. The TiO2@SnS2 core–shell nanocomposites, reported in the present study, have other potential applications, e.g., sodium-ion batteries, gas sensors, water splitting, etc.
Author contributions
Nainy Khera: Methodology, data collection and interpretation, writing – original draft. Pethaiyan Jeevanandam: Conceptualization, project administrat
Conflicts of interest
There are no conflicts to declare.
Data availability
All the data supporting this study have been included in the main manuscript and the SI.
Some additional data supporting this article have been included as part of SI. See DOI: https://doi.org/10.1039/d5na00533g.
Acknowledgements
Nainy Khera acknowledges Council of Scientific and Industrial Research (CSIR), Govt of India for the award of a fellowship (JRF/SRF). The authors express gratitude to Institute Instrumentation Centre (IIT Roorkee) for helping with necessary instrumental facilities. Thanks are due to the Department of Metallurgical and Materials Engineering, IIT Roorkee for providing the HRTEM facility.
References
- Z. Kılıç, Int. J. Hydrol., 2020, 4, 239–241 CrossRef.
- A. Rafiq, M. Ikram, S. Ali, F. Niaz, M. Khan, Q. Khan and M. Maqbool, J. Ind. Eng. Chem., 2021, 97, 111–128 CrossRef CAS.
- R. Al-Tohamy, S. S. Ali, F. Li, K. M. Okasha, Y. A. G. Mahmoud, T. Elsamahy, H. Jiao, Y. Fu and J. Sun, Ecotoxicol. Environ. Saf., 2022, 231, 113160 CrossRef CAS.
- Z. Zheng, J. He, Z. Zhang, A. Kumar, M. Khan, C. W. Lung and I. M. C. Lo, Environ. Sci. Nano, 2024, 11, 1784–1816 RSC.
- C. C. Chen, C. S. Lu, Y. C. Chung and J. L. Jan, J. Hazard. Mater., 2007, 141, 520–528 CrossRef CAS.
- Y. Zhang, X. Yu, H. Liu, X. Lian, B. Shang, Y. Zhan, T. Fan, Z. Chen and X. Yi, Environ. Sci. Nano, 2021, 8, 2049–2058 RSC.
- Y. Zhang, Z. Zhang, Y. Zhang, Y. Li and Y. Yuan, J. Colloid Interface Sci., 2023, 651, 117–127 CrossRef CAS.
- X. Yin, L. Liu and F. Ai, Front. Chem., 2021, 9, 1–9 Search PubMed.
- Y. Wan, J. Chen, J. Zhan and Y. Ma, J. Environ. Chem. Eng., 2018, 6, 6079–6087 CrossRef CAS.
- W. Xu, S. Zhu, Y. Liang, Z. Li, Z. Cui, X. Yang and A. Inoue, Sci. Rep., 2015, 5, 18125 CrossRef CAS PubMed.
- B. Sheng, J. Liu, Z. Li, M. Wang, K. Zhu, J. Qiu and J. Wang, Mater. Lett., 2015, 144, 153–156 CrossRef CAS.
- M. Baláž, E. Dutková, Z. Bujňáková, E. Tóthová, N. G. Kostova, Y. Karakirova, J. Briančin and M. Kaňuchová, J. Alloys Compd., 2018, 746, 576–582 CrossRef.
- R. R. Srivastava, P. Kumar Vishwakarma, U. Yadav, S. Rai, S. Umrao, R. Giri, P. S. Saxena and A. Srivastava, Front. Nanotechnol., 2021, 3, 1–13 Search PubMed.
- D. Hong, W. Zang, X. Guo, Y. Fu, H. He, J. Sun, L. Xing, B. Liu and X. Xue, ACS Appl. Mater. Interfaces, 2016, 8, 21302–21314 CrossRef CAS.
- A. Sadollahkhani, O. Nur, M. Willander, I. Kazeminezhad, V. Khranovskyy, M. O. Eriksson, R. Yakimova and P. O. Holtz, Ceram. Int., 2015, 41, 7174–7184 CrossRef CAS.
- H. D. Abdul kader, I. Sh. Mohammed and S. H. Ammar, Environ. Nanotechnol. Monit. Manag., 2022, 17, 100664 Search PubMed.
- J. Chen, Y. Liao, X. Wan, S. Tie, B. Zhang, S. Lan and X. Gao, J. Solid State Chem., 2020, 291, 121652 CrossRef.
- X. Dong, J. Deng, M. Ge, S. Lu and Q. Zhu, Chem. Phys. Lett., 2023, 833, 140950 CrossRef.
- M. F. Mubarak, H. Selim and R. Elshypany, J. Environ. Health Sci. Eng., 2022, 20, 265–280 CrossRef.
- A. Habtamu and M. Ujihara, RSC Adv., 2023, 13, 12926–12940 RSC.
- Q. Gao, J. Li, B. Liu and C. Liu, J. Alloys Compd., 2023, 931, 167430 CrossRef.
- T. T. Salunkhe, V. Kumar, A. N. Kadam, M. Mali and M. Misra, Ceram. Int., 2024, 50, 1826–1835 CrossRef.
- X. Zhang, X. Jia, R. Xu, X. Lu, H. Liu and Y. Niu, J. Alloys Compd., 2022, 923, 166315 CrossRef.
- V. Gadore, S. R. Mishra and M. Ahmaruzzaman, J. Hazard. Mater., 2023, 444, 130301 CrossRef.
- V. Gadore, S. R. Mishra and M. Ahmaruzzaman, J. Environ. Manage., 2023, 334, 117496 CrossRef.
- G. Kumar, J. Inorg. Organomet. Polym. Mater., 2023, 33, 2710–2720 CrossRef.
- J. Srivind, V. S. Nagarethinam, M. Suganya, S. Balamurugan, D. Prabha and A. R. Balu, Mater. Sci. Eng. B, 2020, 255, 114530 CrossRef.
- Q. Li, S. He, L. Wang, J. Song, J. Wang, C. Shao, Z. Tian and Y. Liu, CrystEngComm, 2023, 25, 2882–2891 Search PubMed.
- V. Gadore, S. R. Mishra and M. Ahmaruzzaman, Chemosphere, 2024, 359, 142343 CrossRef CAS.
- Y. Zhang, L. Hu, H. Zhou, H. Wang and Y. Zhang, ACS Appl. Nano Mater., 2022, 5, 391–400 CrossRef CAS.
- G. Gautam, M. Kumar and B. Singh, Mater. Today Proc., 2022, 62, 3239–3243 CrossRef CAS.
- H. Wei, Y. Zhang, Y. Zhang and Y. Zhang, Colloids Interface Sci. Commun., 2021, 45, 100550 CrossRef CAS.
- P. Salimi Kuchi, H. Roshan and M. H. Sheikhi, J. Alloys Compd., 2020, 816, 152666 CrossRef CAS.
- X. Meng, M. Bi, Q. Xiao and W. Gao, Sens. Actuators, B, 2022, 359, 131612 CrossRef CAS.
- D. Gu, W. Liu, J. Wang, J. Yu, J. Zhang, B. Huang, M. N. Rumyantseva and X. Li, Chemosensors, 2022, 10, 165 Search PubMed.
- J. Yan, Q. Li, Y. Hao, C. Dai and Y. Chen, Funct. Mater. Lett., 2020, 13, 1–4 CrossRef.
- A. K. Dwivedi and S. Tripathi, IEEE Trans. Electron Devices, 2023, 70, 2378–2383 Search PubMed.
- L. Li, Z. Wei, S. Huang, C. Li, Q. Lu and J. Ma, J. Mater. Sci. Mater. Electron., 2022, 33, 18884–18896 CrossRef CAS.
- W. Zhao, Z. Wei, X. Zhang, M. Ding and S. Huang, Mater. Res. Bull., 2020, 124, 110749 CrossRef CAS.
- T. Luttrell, S. Halpegamage, J. Tao, A. Kramer, E. Sutter and M. Batzill, Sci. Rep., 2015, 4, 4043 CrossRef PubMed.
- C. S. Diko, M. Abitonze, Y. Liu, Y. Zhu and Y. Yang, Nanomaterials, 2022, 12, 4497 CrossRef CAS PubMed.
- X. Guo, F. Zhang, Y. Zhang and J. Hu, J. Mater. Chem. A, 2023, 11, 7331–7343 RSC.
- L. A. Burton, T. J. Whittles, D. Hesp, W. M. Linhart, J. M. Skelton, B. Hou, R. F. Webster, G. O'Dowd, C. Reece, D. Cherns, D. J. Fermin, T. D. Veal, V. R. Dhanak and A. Walsh, J. Mater. Chem. A, 2016, 4, 1312–1318 RSC.
- S. Mondal, S. Das and U. K. Gautam, J. Colloid Interface Sci., 2021, 603, 110–119 Search PubMed.
- L. Deng, H. Liu, X. Gao, X. Su and Z. Zhu, Ceram. Int., 2016, 42, 3808–3815 CrossRef CAS.
- N. Carl, M. Fiaz, H. S. Oh and Y. K. Kim, Catalysts, 2024, 14, 442 CrossRef CAS.
- J. Zheng, H. Yang, F. Zheng, F. Liu, L. Lin and S. Li, J. Phys. Chem. C, 2022, 126, 11629–11635 CrossRef.
- J. Gao, X. Sun, L. Zheng, G. He, Y. Wang, Y. Li, Y. Liu, J. Deng, M. Liu and J. Hu, New J. Chem., 2021, 45, 16131–16142 RSC.
- S. Ding, W. Gan, J. Guo, R. Chen, R. Liu, Z. Zhao, J. Li, M. Zhang and Z. Sun, J. Mater. Chem. C, 2024, 12, 7079–7094 RSC.
- P. Bharathi, S. Harish, M. Shimomura, M. K. Mohan, J. Archana and M. Navaneethan, Chemosphere, 2024, 346, 140486 CrossRef.
- Q. Sun, Y. Li, J. Hao, S. Zheng, T. Zhang, T. Wang, R. Wu, H. Fang and Y. Wang, ACS Appl. Mater. Interfaces, 2021, 13, 54152–54161 CrossRef.
- W. Yu, D. Chen, J. Li and Z. Zhang, Crystals, 2023, 13, 1–11 Search PubMed.
- J. Zhang, H. Shen, Y. Li, X. Li, J. Sun, D. Yan, X. Jin and X. Wu, Surf. Interfaces, 2024, 54, 105136 CrossRef.
- J. Shao, X.-T. Wang, H. Xu, X.-D. Zhao, J.-M. Niu, Z.-D. Zhang, Y.-L. Huang and J.-Z. Duan, J. Electrochem. Soc., 2021, 168, 016511 CrossRef.
- H. Yao and L. Liu, Vacuum, 2024, 222, 113030 CrossRef.
- S. Zhu, G. Wu, Z. Liu, S. Zhao, D. Cao, C. Li and G. Liu, Ceram. Int., 2023, 49, 5893–5904 CrossRef.
- S. Shanmugaratnam, B. Selvaratnam, A. Baride, R. Koodali, P. Ravirajan, D. Velauthapillai and Y. Shivatharsiny, Catalysts, 2021, 11, 589 CrossRef.
- P. Yao, S. Yu, H. Shen, J. Yang, L. Min, Z. Yang and X. Zhu, New J. Chem., 2019, 43, 16748–16752 RSC.
- L. Sun, Z. Zhao, S. Li, Y. Su, L. Huang, N. Shao, F. Liu, Y. Bu, H. Zhang and Z. Zhang, ACS Appl. Nano Mater., 2019, 2, 2144–2151 CrossRef.
- Y. Wu, G. Lin, X. Zhou, J. Chen, J. Zhuang, Q. Chen, Y. Luo, D. Lu, V. Ganesh and R. Zeng, J. Electroanal. Chem., 2020, 857, 113740 CrossRef.
- M. Kovacic, H. Kusic, M. Fanetti, U. L. Stangar, M. Valant, D. D. Dionysiou and A. L. Bozic, Environ. Sci. Pollut. Res. Int., 2017, 24, 19965–19979 CrossRef PubMed.
- L. Truong-Phuoc, K. C. Christoforidis, F. Vigneron, V. Papaefthimiou, G. Decher, N. Keller and V. Keller, ACS Appl. Mater. Interfaces, 2016, 8, 34438–34445 CrossRef.
- J. Lin, Y. Liu, Y. Liu, C. Huang, W. Liu, X. Mi, D. Fan, F. Fan, H. Lu and X. Chen, ChemSusChem, 2019, 12, 961–967 CrossRef PubMed.
- K. C. Christoforidis, A. Sengele, V. Keller and N. Keller, ACS Appl. Mater. Interfaces, 2015, 7, 19324–19334 CrossRef.
- M. Li, H. Liu, T. Lv and M. Ding, J. Mater. Chem. A, 2018, 6, 3488–3499 RSC.
- H. She, H. Zhou, L. Li, Z. Zhao, M. Jiang, J. Huang, L. Wang and Q. Wang, ACS Sustain. Chem. Eng., 2019, 7, 650–659 CrossRef.
- J. Wang, X. Li, J. Zhu, H. Li and X. Le, Nanoscale, 2013, 5, 1876–1881 RSC.
- S. Tanaka, D. Nogami, N. Tsuda and Y. Miyake, J. Colloid Interface Sci., 2009, 334, 188–194 CrossRef.
- C. Rosiles-Perez, M. Ocampo Gaspar, O. J. Padilla González, L. F. Román Flores and A. E. Jiménez- González, Emergent Mater, 2024, 7, 1445–1462 CrossRef.
-
P. Somasundaran, B. Markovic, X. Yu and S. Krishnakumar, Handb. Surf. Colloid Chem. Second Ed., CRC Press, Taylor and Francis Group, London, New York, 2002, pp. 387–436 Search PubMed.
- P. Rana and P. Jeevanandam, ChemNanoMat, 2024, 10, 1–11 CrossRef.
-
Profile Solar, Solar PV analysis of Roorkee, India, https://profilesolar.com/locations/India/Roorkee/accessed July 11, 2025.
- A. H. Ragab, N. F. Gumaah, A. A. El Aziz Elfiky and M. F. Mubarak, BMC Chem., 2024, 18, 121 CrossRef PubMed.
- L. B. Chandrasekar, R. Chandramohan, R. Vijayalakshmi and S. Chandrasekaran, Int. Nano Lett., 2015, 5, 71–75 CrossRef.
- J. Madhavi, SN Appl. Sci., 2019, 1, 1–12 Search PubMed.
- A. Dastider, H. Saha, M. J. F. Anik, M. Jamal and M. M. Billah, RSC Adv., 2024, 14, 11677–11693 RSC.
- M. S. Jamal, S. A. Shahahmadi, P. Chelvanathan, H. F. Alharbi, M. R. Karim, M. Ahmad Dar, M. Luqman, N. H. Alharthi, Y. S. Al-Harthi, M. Aminuzzaman, N. Asim, K. Sopian, S. K. Tiong, N. Amin and M. Akhtaruzzaman, Results Phys., 2019, 14, 102360 CrossRef.
- S. Wadhai and P. Thakur, Environ. Sci. Pollut. Res. Int., 2024, 31, 60836–60851 CrossRef.
- P. Praveen, G. Viruthagiri, S. Mugundan and N. Shanmugam, Spectrochim. Acta A Mol. Biomol. Spectrosc., 2014, 117, 622–629 CrossRef.
- H. Liu, C. Du, H. Bai, Y. Su, D. Wei, Y. Wang, G. Liu and L. Yang, J. Mater. Sci., 2018, 53, 10743–10757 CrossRef.
- Z. Wu, Y. Xue, Y. Zhang, J. Li and T. Chen, RSC Adv., 2015, 5, 24640–24648 RSC.
- R. Gaur and P. Jeevanandam, J. Nanosci. Nanotechnol., 2017, 18, 165–177 CrossRef.
- A. K. El-Sawaf, A. A. Nassar, A. A. El Aziz Elfiky and M. F. Mubarak, Polym. Bull., 2024, 81, 12451–12476 CrossRef.
- S. Balu, K. Uma, G. T. Pan, T. C. K. Yang and S. K. Ramaraj, Materials, 2018, 11, 1030 CrossRef PubMed.
- Y. L. Pang, S. Lim and R. K. L. Lee, Environ. Sci. Pollut. Res. Int., 2020, 27, 34638–34652 CrossRef PubMed.
- M. Sathish, S. Mitani, T. Tomai, A. Unemoto and I. Honma, J. Solid State Electrochem., 2012, 16, 1767–1774 CrossRef.
- B. Liu, W. Wang, H. Li, Z. Jian, Y. Xing and S. Zhang, Int. J. Electrochem. Sci., 2018, 13, 6870–6879 CrossRef.
- M. F. Mubarak, H. Selim, H. B. Hawash and M. Hemdan, Environ. Sci. Pollut. Res. Int., 2024, 31, 2297–2313 CrossRef.
- S. Landi, I. R. Segundo, E. Freitas, M. Vasilevskiy, J. Carneiro and C. J. Tavares, Solid State Commun., 2022, 341, 1–7 CrossRef.
- P. R. Jubu, E. Danladi, U. I. Ndeze, O. Adedokun, S. Landi, A. J. Haider, A. T. Adepoju, Y. Yusof, O. S. Obaseki and F. K. Yam, Results Opt., 2024, 14, 100606 Search PubMed.
- A. Saha, A. Moya, A. Kahnt, D. Iglesias, S. Marchesan, R. Wannemacher, M. Prato, J. J. Vilatela and D. M. Guldi, Nanoscale, 2017, 9, 7911–7921 RSC.
- C. C. Mercado, F. J. Knorr, J. L. McHale, S. M. Usmani, A. S. Ichimura and L. V. Saraf, J. Phys. Chem. C, 2012, 116, 10796–10804 Search PubMed.
- F. J. Knorr, D. Zhang and J. L. McHale, Langmuir, 2007, 23, 8686–8690 CrossRef PubMed.
- D. Prabha, S. Ilangovan, J. Srivind, M. Suganya, S. Anitha, S. Balamurugan and A. R. Balu, J. Mater. Sci. Mater. Electron., 2017, 28, 15556–15564 CrossRef.
- I. Iatsunskyi, M. Kempiński, G. Nowaczyk, M. Jancelewicz, M. Pavlenko, K. Załeski and S. Jurga, Appl. Surf. Sci., 2015, 347, 777–783 CrossRef.
- A. K. Mohamedkhair, Q. A. Drmosh and Z. H. Yamani, Front. Mater., 2019, 6, 1–10 CrossRef.
- S. Yan, K. Li, Z. Lin, H. Song, T. Jiang, J. Wu and Y. Shi, RSC Adv., 2016, 6, 32414–32421 RSC.
-
P. van der Heide, X-Ray Photoelectron Spectroscopy: an Introduction to Principles and Practices, John Wiley & Sons, Inc., Hoboken, New Jersey, 2011 Search PubMed.
- S. Jain, N. Kumar, S. Sharma, D. Parmar, R. K. Sharma, M. Tahir, K. Kumari and G. Rani, Mater. Today Sustain., 2023, 24, 100539 Search PubMed.
- M. Andrulevičius, S. Tamulevičius, Y. Gnatyuk, N. Vityuk, N. Smirnova and A. Eremenko, Mater. Sci., 2008, 14, 8–14 Search PubMed.
- L. Shi, C. Xu, X. Sun, H. Zhang, Z. Liu, X. Qu and F. Du, J. Mater. Sci., 2018, 53, 11329–11342 Search PubMed.
- D. M. Mwangangi, T. A. Makhetha, J. C. Ngila and L. N. Dlamini, FlatChem, 2025, 51, 100865 CrossRef.
- S. Ahmed, M. Ahmad, M. H. Yousaf, S. Haider, Z. Imran, S. S. Batool, I. Ahmad, M. I. Shahzad and M. Azeem, Front. Chem., 2022, 10, 1–9 Search PubMed.
- A. Balakrishnan, J. D. Groeneveld, S. Pokhrel and L. Mädler, Chem.–Eur. J., 2021, 27, 6390–6406 CrossRef PubMed.
- J. Gajendiran and V. Rajendran, Adv. Nat. Sci. Nanosci. Nanotechnol., 2011, 2, 015001 CrossRef.
- H. Zhang, T. Van Pelt, A. N. Mehta, H. Bender, I. Radu, M. Caymax, W. Vandervorst and A. Delabie, 2D Mater., 2018, 5, 035006 CrossRef.
- S. Samira and A. Raja P, J. Thermodyn. Catal., 2012, 3, 1–6 Search PubMed.
- R. K. Upadhyay, M. Sharma, D. K. Singh, S. S. Amritphale and N. Chandra, Sep. Purif. Technol., 2012, 88, 39–45 CrossRef CAS.
- G. Xing, L. Zhang, Y. Zhao, S. Li, T. Li, T. Lv, C. Yu and C. Zhao, J. Mol. Struct., 2023, 1274, 134409 CrossRef CAS.
- F. Karimi, H. R. Rajabi and L. Kavoshi, Ultrason. Sonochem., 2019, 57, 139–146 CrossRef CAS.
- S. D. Oladipo and B. Omondi, J. Coord. Chem., 2022, 75, 2170–2188 CrossRef CAS.
- M. Sharma, T. Jain, S. Singh and O. P. Pandey, Sol. Energy, 2012, 86, 626–633 Search PubMed.
- Z. A. Piranshahi, M. Behbahani and F. Zeraatpisheh, Appl. Organomet. Chem., 2018, 32, 1–11 CrossRef.
- A. Dasari and V. Guttena, Mater. Today Commun., 2019, 19, 157–169 CrossRef CAS.
- V. Manikandan, R. Elancheran, P. Revathi, P. Suganya and K. Krishnasamy, Bull. Mater. Sci., 2020, 43, 265 CrossRef CAS.
- A. K. Garg, B. Singh, S. Naskar, R. K. Prajapati, C. Dalal and S. K. Sonkar, Langmuir, 2023, 39, 11036–11047 CrossRef CAS PubMed.
- A. Nezamzadeh-Ejhieh and Z. Banan, Desalination, 2012, 284, 157–166 CrossRef CAS.
- V. Balchander, D. Ayodhya and R. S. Sunder, Inorg. Chem. Commun., 2024, 162, 112181 CrossRef CAS.
- R. Mohammed and M. E. M. Ali, Environ. Nanotechnol., Monit. Manag., 2023, 20, 100885 CAS.
- G. Murugadoss, S. Salla, M. R. Kumar, N. Kandhasamy, H. Al Garalleh, M. Garaleh, K. Brindhadevi and A. Pugazhendhi, Environ. Res., 2023, 220, 115171 CrossRef CAS PubMed.
- H. Xie, T. Zeng, S. Jin, Y. Li, X. Wang, X. Sui and X. Zhao, J. Nanosci. Nanotechnol., 2013, 13, 1461–1466 CrossRef CAS PubMed.
- S. V. P. Vattikuti, I. L. Ngo and C. Byon, Solid State Sci., 2016, 61, 121–130 CrossRef CAS.
- B. S. Goud, G. Koyyada, J. H. Jung, G. R. Reddy, J. Shim, N. D. Nam and S. V. P. Vattikuti, Int. J. Hydrogen Energy, 2020, 45, 18961–18975 CrossRef CAS.
- H. Fakhri, A. R. Mahjoub and A. H. C. Khavar, Appl. Surf. Sci., 2014, 318, 65–73 CrossRef CAS.
- C. Yu, K. Wang, P. Yang, S. Yang, C. Lu, Y. Song, S. Dong, J. Sun and J. Sun, Appl. Surf. Sci., 2017, 420, 233–242 CrossRef CAS.
- G. Murugadoss, S. Muruganandam, M. R. Kumar and A. Pugazhendhi, Chemosphere, 2023, 336, 139100 CrossRef CAS.
- Y. Dan, J. Xu, J. Jian, L. Meng, P. Deng, J. Yan, Z. Yuan, Y. Zhang and H. Zhou, Molecules, 2023, 28, 6882 CrossRef CAS.
- Q. Wei, Q. Dong, D. W. Sun and H. Pu, Spectrochim. Acta - Part A Mol. Biomol. Spectrosc., 2023, 285, 121895 CrossRef CAS.
- V. N. Nguyen, D. T. Tran, M. T. Nguyen, T. T. T. Le, M. N. Ha, M. V. Nguyen and T. D. Pham, Res. Chem. Intermed., 2018, 44, 3081–3095 CrossRef CAS.
- N. P. de Moraes, F. N. Silva, M. L. C. P. da Silva, T. M. B. Campos, G. P. Thim and L. A. Rodrigues, Mater. Chem. Phys., 2018, 214, 95–106 CrossRef.
- P. Ghasemipour, M. Fattahi, B. Rasekh and F. Yazdian, Sci. Rep., 2020, 10, 4414 CrossRef CAS.
- H. J. Fan, S. T. Huang, W. H. Chung, J. L. Jan, W. Y. Lin and C. C. Chen, J. Hazard. Mater., 2009, 171, 1032–1044 CrossRef CAS PubMed.
- A. Farajollahi, A. Poursattar Marjani, N. Noroozi Pesyan and H. Alamgholiloo, Appl. Surf. Sci., 2023, 622, 156903 CrossRef CAS.
- M. Bledowski, L. Wang, A. Ramakrishnan, O. V. Khavryuchenko, V. D. Khavryuchenko, P. C. Ricci, J. Strunk, T. Cremer, C. Kolbeck and R. Beranek, Phys. Chem. Chem. Phys., 2011, 13, 21511–21519 RSC.
- O. P. Kumar, M. N. Ashiq, M. Ahmad, S. Anjum and A. ur Rehman, J. Mater. Sci. Mater. Electron., 2020, 31, 21082–21096 CrossRef CAS.
- Q. U. Ain, U. Rasheed, M. Yaseen, H. Zhang and Z. Tong, J. Hazard. Mater., 2020, 397, 122758 CrossRef CAS.
- Z. Cao, Y. Yin, P. Fu, D. Li, Y. Zhou, Y. Deng, Y. Peng, W. Wang, W. Zhou and D. Tang, Nanoscale Res. Lett., 2019, 14, 1–9 CrossRef CAS.
- H. Wang, L. Xu, R. Zhang, Z. Ge, W. Zhang, J. Xu, Z. Ma and K. Chen, Nanoscale Res. Lett., 2015, 10, 128 CrossRef.
- G. Žerjav, A. Albreht, I. Vovk and A. Pintar, Appl. Catal. A Gen., 2020, 598, 117566 CrossRef.
- K. R. M. Oraby, A. Villalonga, F. S. M. Hassan, M. A. Zayed, M. F. Mubarak, I. Ojeda, A. Sánchez and R. Villalonga, Process Biochem., 2025, 148, 10–16 CrossRef CAS.
- F. Cai, X. Hu, F. Gou, Y. Chen, Y. Xu, C. Qi and D. K. Ma, Appl. Surf. Sci., 2023, 611, 155696 CrossRef CAS.
- F. Wang, S. Zhang, W. Jing, H. Qiu, Y. Liu and L. Guo, J. Mater. Sci. Technol., 2024, 189, 146–154 CrossRef CAS.
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.