Liquid exfoliation of layered metal sulphide for enhanced photocatalytic activity of TiO2 nanoclusters and DFT study

Weiping Zhang, Xinyan Xiao*, Yang Li, Xingye Zeng, Lili Zheng and Caixia Wan
School of Chemistry and Chemical Engineering, South China University of Technology, Guangzhou 510640, China. E-mail: cexyxiao@scut.edu.cn

Received 7th February 2016 , Accepted 21st March 2016

First published on 24th March 2016


Abstract

Layered metal sulphides (LMSs) such as MoS2, WS2 and SnS2 have attracted much attention in the field of photocatalysis due to their excellent properties. Herein, a facile and effective liquid exfoliation solvothermal method for fabricating TiO2/LMS (LMS = MoS2, WS2 or SnS2) photocatalysts has been developed. The optimum molar ratio of Ti–Mo, Ti–W and Ti–Sn was determined to be 50[thin space (1/6-em)]:[thin space (1/6-em)]0.8, 50[thin space (1/6-em)]:[thin space (1/6-em)]0.1 and 50[thin space (1/6-em)]:[thin space (1/6-em)]0.1, respectively. The optical properties of TiO2/LMS with a matching solar spectrum contribute to converting the solar energy to chemical energy by photon-driven photocatalytic reactions. The combined effect of liquid exfoliation and solvothermal reforming has been demonstrated as an effective method to obtain high efficiency photocatalysts using bulk metal sulphides as sensitizers. The binding site of TiO2 and the LMS at the interface of a composite photocatalyst was investigated by the density functional theory (DFT) method at a molecular cluster level, and the calculation results showed that firm structures were formed at the interfaces of TiO2 nanoparticles and the LMS. The photocatalytic activity evaluation of TiO2/LMS showed that the LMS played the crucial role in separation of photogenerated e/h+ pairs and utilization of photons for enhancing the photocatalytic activity of TiO2. The study of the electron transfer mechanism indicated that the synergetic effect of superoxide radicals (·O2) and hydroxyl radicals (·OH) plays the leading role in the dye degradation process.


Introduction

Semiconductor photocatalysis, as a green technology, has been widely applied in environmental purification (non-selective process),1 photocatalytic synthesis,2 and clean energy.3,4 Recently, titanium oxide (TiO2) nanomaterials are generally believed to be the most reliable materials for the decomposition of toxic and hazardous organic pollutants and heavy metals due to their non-toxicity, physical and chemical stability, availability, and unique optical properties.5 However, TiO2 in a pure state can solely absorb UV light (≤380 nm) and has lower capacity for the utilization of solar energy due to its large band gap (3.2 eV).6–10 Hence, it is highly desirable to develop photocatalysts with high catalytic activities under solar light.11–15 In recent years, metal sulphide has been extensively studied due to its excellent performance and potential application for solar energy conversion. Certain metal sulphides, especially single- and few-layer 2D transition metal disulphides such as MoS2,16 WS2,17 and SnS2,18 have attracted much attention in photocatalysis due to their excellent properties for utilization of solar light. Thus, using these metal disulphides as photosensitizers will be a highly efficient and low cost way for the modification of TiO2.19–21 Numerous research has reported that LMS (MoS2, WS2 and SnS2) nanosheets can be synthesized by a chemical method using the relevant metal salt as raw material. For instance, Liu et al.22 reported layered nano-MoS2 decorated on TiO2 nanotube arrays by photocatalytic reduction of (NH4)2MoS4. Nano-MoS2 can obviously enhance the electrocatalytic activity of TiO2 nanotube array electrodes in hydrogen evolution reactions. Shao et al.23 reported ultrathin hexagonal SnS2 nanosheets assembled onto TiO2 nanofibers by a hydrothermal process using SnCl4·5H2O and thioacetamide as the raw materials. The results showed that the enhancement of photocatalytic activity was mainly due to the photosynergistic effect of TiO2 and SnS2, leading to the better separation efficiency of photoinduced electron–hole pairs. Hou et al.24 reported a WS2/g-C3N4 photocatalyst prepared via an impregnation sulfidation approach using (NH4)2WS4 as raw material. The layered WS2 was loaded on the surface of g-C3N4 and significantly enhanced the H2 production performance. These reviews concentrate on LMS nanosheets synthesized by chemical methods, and surely, a stable heterojunction system could be obtained. However, there are also numerous problems emerging in these methods, such as rigorous processing, lower purity and lower quality, which will limit their industrial applications.

In recent years, the physical method, mechanical exfoliation, has attracted much attention in the preparation of two-dimensional nanosheet materials.25,26 A simple liquid exfoliation method would allow the formation of novel hybrid and composite materials. The layered compounds, such as MoS2, WS2 and SnS2, can be efficiently dispersed in common solvents and can be deposited as individual flakes or formed into films.25–29 By blending these materials with the suspensions of other nanomaterials, the hybrid composites could be obtained.28,30,31 However, the simple blending method may not make the hybrid composites stable and also some effective strategies are needed, such as thermal treatment and coating treatment, to obtain a stable heterojunction system.

Here, we demonstrated exfoliation of bulk LMS crystals (MoS2, WS2 and SnS2) in common solvent (EtOH) to give mono- and few-layer nanosheets. This method is insensitive to air and water and can potentially be scaled up to give large quantities of exfoliated material. Subsequently, using these LMS nanosheets as sensitizers, TiO2/LMS composites were synthesized by the combined method of ultrasonic exfoliation and solvothermal reforming. In a typical procedure, LMS nanosheets were embedded into the TiO2 cross-linked structure by the hydrolytic polymerization of TiCl4, which preferably protected the LMS from the influence of the external system. This special coated structure would be very favourable to the electrons transferred to the surface of the composite, leading to effective separation of photoinduced electron/hole pairs, which significantly enhanced the photocatalytic degradation of organic dye. Moreover, the possible structures were predicted by the density functional theory (DFT) method. The photocatalytic mechanism of the TiO2/LMS composites was also discussed.

Experimental details

Materials

All of the chemicals used for the synthesis of the catalysts were of analytical grade and were used as-received without any further purification. Tungsten disulphide (WS2, 99%) was obtained from Aladdin Reagent Co. Tin disulphide (SnS2, ≥99.5%) was purchased from Changsha Huajing Powder Material Technological Co., China. Titanium tetrachloride (TiCl4, ≥99%), methyl blue and molybdenum disulphide (MoS2, ≥98.5%) were purchased from Tianjin Fuchen Chemical Reagent Factory, China. Ethanol was purchased from Nanjing Chemical Reagent Co., China, and both glycerol (C3H8O3, ≥99%) and sodium hexametaphosphate ((NaPO3)6) were obtained from Tianjin Kemiou Chemical Reagent Co., China.

Preparation of TiO2/LMS composite photocatalysts

The preparation procedure of the TiO2/LMS composite photocatalysts contains two steps: (1) the exfoliation process of the LMS; and (2) the solvothermal process of the TiO2/LMS precursor. The detailed steps are shown as follows.

0.05 g of LMS particles, 30 mL of ethanol and 1 mL of glycerol were added into the beaker and exfoliated by the ultrasonic method for 2–3 h. The upper dispersion was collected and centrifuged at 9000 rpm for 5 min and washed with ethanol 3 times. The exfoliated LMS nanosheets were obtained after being dried at 60 °C for 15 h.

4 mL of titanium tetrachloride ethanol solution (TiCl4, 2 mol L−1) and certain exfoliated LMS nanosheets were added into a beaker with 12.5 mL of EtOH at room temperature (30 °C), and the homogeneous dispersion solution was obtained by the ultrasound method for 120 min. Then, 3 mL of glycerol (2%) aqueous solution containing a certain dispersant (SHMP, 0.4%) as the hydrolytic agent was slowly added into the dispersion, and the TiO2/LMS vitreosol was obtained after ultrasonic treatment for 10 min. Subsequently, the obtained solution vitreosol was transferred into a 25 mL Teflon hydrothermal reactor and kept at 140 °C for 3 h. The as-prepared composite photocatalysts were isolated by centrifugation, washed with absolute ethanol three times and dried at 80 °C for 5 h in an air dry oven.

Characterization

The crystal forms of the composite photocatalyst were analyzed by an X-ray diffractometer (D8 ADVANCE, Bruker) with a Cu tube for generating Cu Kα radiation (k = 1.5418 Å). The morphology of the photocatalysts was measured using scanning electron microscopy (SEM, Hitachi, S-3700N). The morphologies and internal structure were examined by transmission electron microscopy (TEM, JEOL, JEM-2100F, Japan). X-ray photoelectron spectroscopy (XPS) measurements were performed on a Phi X-tool instrument. Chemical compositions and molecular structure of the as-prepared materials were analyzed by a Raman spectrometer (LabRAM Aramis, France). UV-vis diffuse reflectance spectra were recorded on a Hitachi UV-3010 spectrophotometer (BaSO4 as a reflectance standard).

Evaluation of photocatalytic activities of the TiO2/LMS composites

The photocatalytic activities of the samples for mineralizing methyl blue (20 mg L−1) were measured as follows: a certain amount of the as-prepared (50 mg) powders was combined with 250 mL of organic substance solution in a water-cooled reactor and stirred for 30 min to reach the absorption–desorption equilibrium in a dark box. Subsequently, the dispersed suspension was irradiated by a XG500 xenon long-arc lamp. At the given irradiation time intervals, 4 mL samples were collected and then filtered with a centrifuge to remove the photocatalyst. The supernatant was analysed by a SHIMADZU UV-2450 spectrophotometer to record the maximum absorbance (599 nm).

Capture of active species (h+, e, ·O2 and ·OH)

Benzoquinone, tert-butyl alcohol, silver nitrate and EDTA-2Na were employed as the scavengers of superoxide radicals (·O2), hydroxyl radicals (·OH), electrons (e) and holes (h+).32,33 50 mg of the as-prepared sample was dispersed in the MB aqueous solution (250 mL), and then 0.3 mmol scavenger was added into the aqueous solution. The photocatalytic experiment process was similar to the evaluation process of photocatalytic activities.

Results and discussion

The bulk LMS was exfoliated by the liquid exfoliation method in a liquid system of ethanol and glycerol solution for fabricating TiO2/LMS (LMS = MoS2, WS2 or SnS2). The SEM images before and after exfoliation of the LMS are shown in Fig. 1. A microsheet structure was the main type in the LMS compounds before exfoliation. After exfoliation, the LMS compounds were mainly in the form of nanosheets being mono- or few-layer. This indicated that the ultrasonic exfoliation was a better way to obtain nanoscale LMS sheets, which would provide high quality sensitizers for fabricating TiO2-based hybrid composites.
image file: c6ra03534e-f1.tif
Fig. 1 SEM images of LMS (MoS2, WS2 and SnS2). LMS before exfoliation (a1, b1 and c1), and LMS after exfoliation (a2, b2 and c2).

Using these LMS nanosheets as sensitizers, the TiO2/LMS composites were synthesized by a solvothermal method. From Fig. 2, TEM images were given to investigate the inter-structures of the TiO2/LMS composites. In Fig. 2a, TiO2 exists in the form of a cross-linked structure. It is noted that MoS2 nanosheets are embedded in the TiO2 cross-linked structure and nanopores were distributed evenly on the composite, which could improve the adsorbing capacity of hazardous substances (Fig. 2b).


image file: c6ra03534e-f2.tif
Fig. 2 TEM images of pure TiO2 (a), and the TiO2/MoS2 (b), TiO2/WS2 (c) and TiO2/SnS2 (d) composite photocatalysts.

For the TiO2/WS2 composite (Fig. 2c), a clear boundary structure was observed between TiO2 and WS2, including 9 nm of a TiO2 shell and nano-WS2 core, implying that a stable contact interface was formed between TiO2 and WS2. The TiO2/SnS2 composite existed in the formation of aggregates and the nano-SnS2 was coated by TiO2 nanoparticles (Fig. 2d). These formations of the composites can be explained as follows: the TiCl4 ethanol fluid system has a lower surface tension, which could help Ti4+ permeate the interlayer of the LMS. With the hydrolysis of Ti4+, the Ti–OH sol was uniformly dispersed around the LMS nanosheets, and subsequently, Ti–OH was polymerized to form a TiO2 cross-linked structure and coated on the surface of the LMS nanosheets (Fig. 3). In this typical process, TiO2/LMS heterojunctions could be formed at the interfaces of the TiO2 nanostructure and LMS. These special structures are beneficial for the electron transfer and separation of photogenerated e/h+ pairs at the contact interface of the composite photocatalysts.


image file: c6ra03534e-f3.tif
Fig. 3 Schematic illustration for the fabrication of the TiO2/LMS (LMS = MoS2, WS2 or SnS2) composite photocatalyst.

The XRD patterns of the as-prepared samples are shown in Fig. 4. The crystal form of TiO2 in the photocatalyst composites was anatase, since the characteristic diffraction peaks at 25.3°, 37.8°, 48.0°, 53.9°, 55.1°, 62.7°, 68.8°, 70.3° and 75.0° are attributed to the (101), (004), (200), (105), (211), (204), (116), (220) and (215) crystal faces of anatase TiO2 (PDF#21-1272). For the TiO2/MoS2 composite, the modest peaks at 33.0°, 48.1° and 52.1° (PDF#17-0744) corresponded to the (101), (107) and (018) crystal faces of molybdenite-3R-MoS2. A characteristic peak located at 26.6° corresponded to the (111) crystal faces of MoO3, which was attributed to part of the MoS2 being oxidized in the fabricating procedure. For the TiO2/WS2 composite, an apparent peak observed at the position of 14.4° was the diffraction peak of tungstenite-3R (PDF#35-0651). For the TiO2/SnS2 composite, a diffraction peak at the position of 15.0° is observed, which corresponds to the (001) crystal face of berndtite-2T (PDF#23-0677).


image file: c6ra03534e-f4.tif
Fig. 4 XRD patterns of pure TiO2, and the TiO2/MoS2, TiO2/WS2 and TiO2/SnS2 composite photocatalysts.

The crystal sizes of TiO2 in the composites were calculated by applying Scherrer’s equation to the anatase (101) diffraction peak which represented the highest intensity peak for each pure phase.34

 
image file: c6ra03534e-t1.tif(1)
where D is the average crystallite size, k = 0.89, λ is the wavelength of Cu Kα radiation (0.15406 nm), β is the full width at half maximum (FWHM), and θ is the Bragg diffraction angle. The average crystallite size of anatase TiO2 in the different composites based on eqn (1) are given in Table 1. The effectiveness of the hydrothermal process in the presence of metal sulphide to deform the TiO2 was evident from the changes to the crystallite size. The results showed that the presence of the LMS in TiO2/LMS could increase the grain size of TiO2. This indicated that the LMS could enhance the crystallization degree of TiO2, leading to the surface energy of the composite being decreased, which could make the composite photocatalysts maintain stability in aqueous solutions.

Table 1 The average anatase TiO2 crystallite size of TiO2/LMS
Sample TiO2 TiO2/MoS2 TiO2/WS2 TiO2/SnS2
TiO2 (d, nm) 9.1 11.4 11.3 11.1


X-ray photoelectron spectroscopy (XPS) was used to investigate the chemical states of the metal and S in the TiO2/LMS heterostructures (Fig. 5). The high resolution XPS spectra show that the binding energies of the S 2p1/2 and S 2p3/2 peaks in TiO2/LMS are located at 162.3 and 161.1 eV (TiO2/MoS2),35 163.5 and 161.5 eV (TiO2/WS2),31 and 163 and 161.5 eV (TiO2/SnS2),23,36 respectively, implying that S2− existed in the TiO2/LMS composites. Furthermore, the binding energies of Mo 3d3/2 and Mo 3d5/2 in TiO2/MoS2, W 4f5/2 and W 4f7/2, and Sn 3d3/2 and Sn 3d5/2 are located at 233.4 and 230.2 eV,35,37 34.5 and 32.4 eV,31 and 495.6 and 486.8 eV,23,36 respectively, suggesting that Mo4+, W4+ and Sn4+ existed in the TiO2/MoS2, TiO2/WS2 and TiO2/SnS2 composites, respectively. Meanwhile, Mo6+ and W6+ were confirmed to exist in the TiO2/MoS2 and TiO2/WS2 composites, respectively, which was consistent with the XRD results. The high energy component at 166.1 eV can be assigned to S4+ species in sulphate groups (SO32−), and these groups could locate at the edges of the SnS2 layers, which was attributed to the oxidation of S2− at the surface of SnS2.


image file: c6ra03534e-f5.tif
Fig. 5 XPS spectra of the TiO2/LMS composites. S 2p of TiO2/LMS (a), Mo 3d of TiO2/MoS2 (b), W 4f of TiO2/WS2 (c), and Sn 3d of TiO2/SnS2 (d).

The compositions of the TiO2/LMS samples were investigated by Raman spectroscopy. The shifting of the Raman band has a close relationship with the particle size and phase structure of the composites. Fig. 6 shows that the TiO2 of the anatase structure had six modes including the Raman bands of 153.1 cm−1 (Eg), 211.5 cm−1 (Eg), 400.5 cm−1 (B1g), 515.2 cm−1 (A1g + B1g), and 639.2 cm−1 (Eg).38 For the TiO2/MoS2 composite, the 2Eg, 2B1g and A1g Raman modes of anatase TiO2 in TiO2/MoS2 have a blue shift, and the Raman bands were located at 151.4 cm−1 (Eg), 388.4 cm−1 (B1g), 506.8 cm−1 (A1g + B1g), and 629.7 cm−1 (Eg). This indicated that the vibrational energy was elongated due to the contact interface of the TiO2/MoS2 composite formed in the fabrication process.39 The Raman mode near 210 cm−1 was weak and almost disappeared, and instead a new peak at 201.1 cm−1 was observed, which was speculated to be the characteristic peak of MoO3. For the TiO2/WS2 sample, the Raman band at 400.5 cm−1 was divided into three new peak positions of 351.8 cm−1, 399.6 cm−1 and 418.8 cm−1. Among them, the positions of 351.8 cm−1 and 418.8 cm−1 were speculated to be for the A1g and E12g vibrational modes of WS2, respectively, which is consistent with previous results.40 From the Raman spectrum of the TiO2/SnS2 composite, a new peak was observed at 313.4 cm−1, corresponding to the A1g mode of the SnS2 hexagonal phase.31,41 These results displayed the shifting or splitting of the characteristic peak, indicating that a heterostructure was formed at the interface of the TiO2/LMS composites.


image file: c6ra03534e-f6.tif
Fig. 6 Raman spectra of pure TiO2 and the TiO2/LMS composites.

The absorption spectra of the as-prepared samples were measured using UV-vis diffuse reflectance spectroscopy. As shown in Fig. 7a, the basal absorption edge of TiO2 occurs at a wavelength shorter than 400 nm, whereas TiO2/MoS2, TiO2/WS2 and TiO2/SnS2 have notable absorption in the visible light region between 400 and 800 nm, which is attributed to the excellent visible light response of LMSs. And the absorption edge of the TiO2/SnS2 composite shifts to the visible light area (420 nm) as compared to that of P25. For the TiO2/WS2 composite, one obvious absorption band was observed around 626 nm due to the excellent optical property of WS2 of improving visible light absorption and utilizing photons.42 Similarly, for TiO2 coupled with MoS2, the optical absorption threshold of the TiO2/MoS2 band gap transition has been radically decreased due to the fine matching energy level with solar energy.


image file: c6ra03534e-f7.tif
Fig. 7 UV-vis absorption spectra (a) and the corresponding Kubelka–Munk function (b) of pure TiO2 and the TiO2/LMS composites.

The band gap (Eg) of TiO2/MoS2, TiO2/WS2, TiO2/SnS2 and pure TiO2 was estimated using the Kubelka–Munk function:43,44

 
image file: c6ra03534e-t2.tif(2)
where F(R) is the Kubelka–Munk function, and R is reflectivity. The curve graph of F(R) ∼ λ can be exported during the test, which can be transformed into a curve graph corresponding to image file: c6ra03534e-t3.tif, where E is the energy (E = 1240/λ). Thus, as shown in Fig. 7b, the band gap (Eg) of TiO2/MoS2, TiO2/WS2, TiO2/SnS2 and pure TiO2 is estimated to be 3.01 eV, 3.06 eV, 3.07 eV and 3.21 eV, respectively. The results showed that the conduction band edge of TiO2/LMS was about 0.14–0.20 eV more positive than that of pure TiO2. The results suggested that the existence of LMSs in composites can enlarge the light abstraction width of TiO2, implying that a stable TiO2/LMS heterostructure was formed at the interface of TiO2 and the LMS. Meanwhile, the red shift also indicated that the surface energy of TiO2 in the TiO2/LMS composites decreased, which was consistent with the XRD results.

A series of experiments were carried out to further investigate the photocatalytic activities of the TiO2/LMS composites and the photocatalytic degradation of methyl blue (MB) in the liquid phase, which were conducted under simulated sunlight and the results are shown in Fig. 8. The optimum mole ratios of TiO2/LMS were determined to be 50[thin space (1/6-em)]:[thin space (1/6-em)]0.8 (TiO2/MoS2), 50[thin space (1/6-em)]:[thin space (1/6-em)]0.1 (TiO2/WS2) and 50[thin space (1/6-em)]:[thin space (1/6-em)]0.1 (TiO2/SnS2). As compared to the TiO2/MoS2 composite, the lower molar ratios of SnS2 and WS2 were more beneficial for enhancing the photocatalytic activity of TiO2. The reason for the results was mainly related to the exfoliated-type of LMS nanosheets. The exfoliated-type of the LMSs was mainly MoS2 nanosheets in the TiO2/MoS2 composite, and was mainly particles or few-layered nanosheets in the TiO2/WS2 and TiO2/SnS2 composites. These typical exfoliated structures of the LMSs made the form of the composites different: an embedded structure of TiO2/MoS2, a core–shell structure of TiO2/WS2, and a coating structure of TiO2/SnS2. Hence, the composite structure of TiO2/LMS and the exfoliated-type of LMS jointly determined the photocatalytic activity of the composites. However, the coating structure was not more stable than the embedded structure and the core–shell structure, so the photocatalytic activity of TiO2/SnS2 was lower than that of TiO2/WS2 and TiO2/MoS2.


image file: c6ra03534e-f8.tif
Fig. 8 Photocatalytic degradation of MB using TiO2/MoS2 (a), TiO2/WS2 (b), TiO2/SnS2 (c), and P25, TiO2 and a physical mixture of TiO2 and LMS (d) as the photocatalyst.

In addition, the photocatalytic activities for the physical mixture of TiO2 and the LMS (MoS2, WS2 and SnS2) were also investigated by MB degradation (Fig. 8d). The results showed that the degradation rates of a physical mixture of TiO2 and MoS2, a mixture of TiO2 and WS2, and a mixture of TiO2 and SnS2 were obviously lower than those of the TiO2/MoS2, TiO2/WS2 and TiO2/SnS2 composites obtained by chemical reaction, which suggested that the TiO2 and LMS in the TiO2/LMS composites (TiO2/MoS2, TiO2/WS2 and TiO2/SnS2) was not combined by van der Waals forces but by chemical bonds. Also, the photocatalytic activities of pure TiO2 and P25 were lower than those of the TiO2/LMS composites, suggesting that the presence of the LMS in the composites could effectively enhance the photocatalytic activity of TiO2, and the properties and structures of the LMS would be beneficial for electron transfer and reduction of e/h+ pair recombination, leading to promotion of the photocatalytic degradation of MB.45

To investigate the photostability of the photocatalysts, cycling tests were conducted of MB degradation under simulated sunlight irradiation. For each cycle, the photocatalysts were not recovered and a fresh solution of MB was added into the photoreactor to undergo degradation under identical conditions (the MB initial concentration was 20 mg L−1). As shown in Fig. 9a, the degradation rates of MB (TiO2/MoS2, 91.4%; TiO2/WS2, 92.0%; TiO2/SnS2, 90.0%) slightly decrease after four cycles. The samples of TiO2/LMS after cycling tests were characterized by XRD and the results are shown in Fig. 9b. From the XRD results, it can be seen that the crystal structure of TiO2/LMS had not changed significantly. These results demonstrate the better stability and promising application of TiO2/LMS in the treatment of dye-containing wastewater. This was consistent with the capture test of hydroxyl radicals because the intensity of hydroxyl radicals over time was linear (Fig. S1).


image file: c6ra03534e-f9.tif
Fig. 9 Photostability test of the TiO2/LMS composite photocatalysts for MB degradation (MB initial concentration, 20 mg L−1; irradiation time, 60 min) (a). XRD analysis of the TiO2/LMS composite photocatalysts after photocatalytic reaction (b).

To deeply understand the formation mechanism of TiO2/LMS, the DFT method was used to study the possible active site of TiO2/LMS combination. As shown in Fig. 10, these hexagonal platelets exhibit two kinds of edges, which are named respectively as the metallic and sulphur edge. Generally, the initial form of the Ti–OH sol will couple with the LMS from the two edges. The metallic edges should theoretically be the centres of electron deficiency, which need external electrons to form a structure; meanwhile, the surface of the Ti–OH sol contains vast numbers of negative charges. Under proper conditions, the formation of a chemical bond between the LMS and TiO2 is possible.


image file: c6ra03534e-f10.tif
Fig. 10 Representation of the LMS (MoS2, WS2 or SnS2) crystallographic plan (Mo atoms in green, W atoms in blue, Sn atoms in gray and S atoms in yellow).

In Fig. 11, the structures of the global minima of TiO2 coupled with the LMS are presented. A key feature shown is that the two oxygen atoms of Ti–OH prefer to coordinate with the metallic edges of MoS2 and WS2. For SnS2, instead the oxygen atoms generally will not coordinate with the support, suggesting that the composite formation is closely related to its trigonal crystal system. And yet, MoS2 and WS2 have a similar structure of the hexagonal crystal system, leading to the similar results. It is noted that the distances between the preferred sites for Mo and O, W and O, and Sn and O are 2.08 Å, 2.00 Å and 2.36 Å, respectively. Meanwhile, the absorption energies were also calculated from the optimum results, and were −478.89 kJ mol−1 (TiO2/MoS2), −640.91 kJ mol−1 (TiO2/WS2), and −245.42 kJ mol−1 (TiO2/SnS2), respectively. The results indicated that the solid–solid contact interface between TiO2 and the LMS was connected by chemical bonds, and not van der Waals interaction. From the experimental results, the photocatalytic activities of the TiO2/MoS2, TiO2/WS2 and TiO2/SnS2 composite photocatalysts were higher than that of the TiO2/LMS physical mixture. Meanwhile, the photostabilities and photocatalytic activities of the TiO2/MoS2 and TiO2/WS2 composites were higher than that of the TiO2/SnS2 composite due to the absorption energies of the TiO2/MoS2 and TiO2/WS2 composites, which suggested that the experimental and the optimization results are highly consistent with the these theoretical calculation results.


image file: c6ra03534e-f11.tif
Fig. 11 Global minima of TiO2 coupled with the LMS (MoS2, WS2 or SnS2) (Mo atoms in cyan, W atoms in blue, Sn atoms in gray, S atoms in yellow, and H, O and Ti atoms of Ti–OH in silver-gray, red and silver, respectively).

The degradation of dye molecules commonly corresponds to the active species such as h+, e, ·O2 and ·OH.33,46,47 Herein, benzoquinone (BQ), tert-butyl alcohol (TBA), silver nitrate (AgNO3) and EDTA-2Na were used as scavengers of superoxide radicals (·O2), hydroxyl radicals (·OH), electrons (e) and holes (h+) to investigate the possible mechanism for the degradation of MB. Different scavengers were employed individually to remove the corresponding active species so that the function of different active species in the degradation process was understood. As shown in Fig. 12, the degradation rates are 93.4% (TiO2/MoS2), 93.8% (TiO2/WS2), and 92.1% (TiO2/SnS2), respectively, without a scavenger. When BQ was added into the reaction system, the degradation rates decrease to 34.9% (TiO2/MoS2), 33.3% (TiO2/WS2) and 29.7% (TiO2/SnS2), respectively. When TBA was added into the reaction system, the degradation rates decrease to 71.3% (TiO2/MoS2), 70.6% (TiO2/WS2) and 68.7% (TiO2/SnS2), respectively. The results suggested that ·O2 significantly influences the photocatalytic activity in the degradation of MB, while ·OH has implications for the degradation of MB. From the capture of the free electrons and h+, the influence of free electrons and h+ seems to be negligible in the reaction. From Fig. S1, the formation rate of commercial TiO2 (P25) is higher than that of TiO2/LMS, while the photocatalytic activity is lower than that of TiO2/LMS, which suggests that ·OH was not the main active species for the degradation of dye in the TiO2/LMS photocatalytic system.


image file: c6ra03534e-f12.tif
Fig. 12 Effects of scavengers on the degradation of MB (irradiation time = 60 min, scavenger dosage = 0.3 mmol L−1).

To further understand the mechanism of electron transfer, the electron transfer of the photocatalytic reaction in a heterojunction-type photocatalytic system was elaborated at the molecular level. Based on quantum confinement effects,48,49 the band gap of the bulk LMS can be significantly increased after being transformed into LMS nanosheets, leading to a more negative potential of the conduction band and a more positive potential of the valence band of the LMS. When the LMS nanosheets couple with TiO2, the higher energy level of the conduction band of the LMS than that of TiO2 would promote photoinduced electron transfer from the nanoscale LMS to TiO2. As shown in Fig. 13, the photogenerated electrons in the CB of the LMS migrate to the CB of TiO2, while the photogenerated holes in the VB of TiO2 move to the VB of the LMS. Thus, the photogenerated electrons and holes are spatially separated, which greatly reduces the recombination of charge carriers. Due to the negative redox potential in the CB of TiO2, the photogenerated electrons can reduce O2 to ·O2. Moreover, ·O2 can also oxidize H2O to produce ·OH, and the formation of ·OH depends on the oxidation of ·O2, which can explain why ·O2 was the main active species. Nevertheless, the synergetic effect of ·O2 and ·OH plays the leading role in the degradation of dye.


image file: c6ra03534e-f13.tif
Fig. 13 Schematic diagram of the photogenerated charge transfer process on the TiO2/LMS photocatalysts.

Conclusions

LMS nanosheets were prepared from bulk LMS by a sonication method in a liquid system. The TiO2/LMS (LMS = MoS2, WS2 or SnS2) photocatalysts were synthesized by a combined method of liquid exfoliation and the solvothermal method. The obtained TiO2/LMS composite photocatalysts showed high photocatalytic activity under solar irradiation. The calculation results of DFT analysis and the experimental results indicated that the solid–solid contact interface between TiO2 and the LMS was connected by chemical bonds, and not van der Waals interaction. At present, this approach of using bulk metal sulphide as a direct sensitizer to synthesize TiO2-based composite photocatalysts has still rarely been previously reported. LMS in the composites acted as an electron donor for the interfacial electron transfer, and thus enhanced the separation of the photogenerated electron–hole pairs. The application of TiO2/LMS for photocatalytic decontamination of MB demonstrated a higher photocatalytic activity for the TiO2/metal sulphide composites, which have great potential in the detoxification of harmful pollutants in wastewater.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No. 21376099, 21546002).

Notes and references

  1. M. Zeng, Y. Z. Li, M. Y. Mao, J. L. Bai, L. Ren and X. J. Zhao, ACS Catal., 2015, 5, 3278 CrossRef CAS.
  2. D. Sannino, V. Vaiano and P. Ciambelli, Catal. Today, 2013, 205, 159 CrossRef CAS.
  3. S. Ida, N. Kim, E. Ertekin, S. Takenaka and T. Ishihara, J. Am. Chem. Soc., 2015, 137, 239 CrossRef CAS PubMed.
  4. D. J. Martin, P. J. T. Reardon, S. J. A. Moniz and J. W. Tang, J. Am. Chem. Soc., 2014, 136, 12568 CrossRef CAS PubMed.
  5. H. Tong, S. X. Ouyang, Y. P. Bi, N. Umezawa, M. Oshikiri and J. H. Ye, Adv. Mater., 2012, 24, 229 CrossRef CAS PubMed.
  6. R. Long, N. J. English and O. V. Prezhdo, J. Phys. Chem. Lett., 2014, 5, 2941 CrossRef CAS PubMed.
  7. J. J. Murcia, M. C. Hidalgo, J. A. Navío, J. Araña and J. M. Dona-Rodríguez, Appl. Catal., B, 2015, 179, 305 CrossRef CAS.
  8. M. Planells, A. Abate, H. J. Snaith and N. Robertson, ACS Appl. Mater. Interfaces, 2014, 6, 17226 CAS.
  9. D. Dolat, S. Mozia, B. Ohtani and A. W. Morawski, Chem. Eng. J., 2013, 225, 358 CrossRef CAS.
  10. E. Evgenidou, K. Fytianos and I. Poulios, Appl. Catal., B, 2005, 59, 81 CrossRef CAS.
  11. M. Jin, X. T. Zhang, S. Nishimoto, Z. Y. Liu, D. A. Tryk, A. V. Emeline, T. Murakami and A. Fujishima, J. Phys. Chem. C, 2007, 111, 658 CAS.
  12. B. J. Ma, J. S. Kim, C. H. Choi and S. I. Woo, Int. J. Hydrogen Energy, 2013, 38, 3582 CrossRef CAS.
  13. M. Shalom, S. Rühle, I. Hod, S. Yahav and A. Zaban, J. Am. Chem. Soc., 2009, 131, 9876 CrossRef CAS PubMed.
  14. X. J. Wang, W. Y. Yang, F. J. Li, Y. B. Xue, R. H. Liu and Y. J. Hao, Ind. Eng. Chem. Res., 2013, 52, 17140 CrossRef CAS.
  15. H. H. Yang, S. V. Kershaw, Y. Wang, X. Z. Gong, S. Kalytchuk, A. L. Rogach and W. Y. Teoh, J. Phys. Chem. C, 2013, 117, 20406 CAS.
  16. Q. J. Xiang, J. G. Yu and M. Jaroniec, J. Am. Chem. Soc., 2012, 134, 6575 CrossRef CAS PubMed.
  17. H. Y. Guo, N. Lu, J. Dai, X. J. Wu and X. C. Zeng, J. Phys. Chem. C, 2014, 118, 14051 CAS.
  18. X. Z. Yuan, H. Wang, Y. Wu, X. H. Chen, G. M. Zeng, L. J. Leng and C. Zhang, Catal. Commun., 2015, 61, 62 CrossRef CAS.
  19. D. Merki and X. Hu, Energy Environ. Sci., 2011, 4, 3878 CAS.
  20. J. G. Wang, X. R. Li, X. Li, J. Zhu and H. X. Li, Nanoscale, 2013, 5, 1876 RSC.
  21. H. L. Li, K. Yu, C. Li, B. J. Guo, X. Lei, H. Fu and Z. Q. Zhu, J. Mater. Chem. A, 2015, 3, 20225 CAS.
  22. C. H. Meng, Z. Y. Liu, T. R. Zhang and J. Zhai, Green Chem., 2015, 17, 2764 RSC.
  23. Z. Y. Zhang, C. L. Shao, X. H. Li, Y. Y. Sun, M. Y. Zhang, J. B. Mu, P. Zhang, Z. C. Guo and Y. C. Liu, Nanoscale, 2013, 5, 606 RSC.
  24. Y. D. Hou, Y. S. Zhu, Y. Xu and X. C. Wang, Appl. Catal., B, 2014, 156–157, 122 CrossRef CAS.
  25. V. Nicolosi, M. Chhowalla, M. G. Kanatzidis, M. S. Strano and J. N. Coleman, Science, 2013, 340, 1421 CrossRef.
  26. J. N. Coleman, M. Lotya, A. O’Neill, S. D. Bergin, P. J. King, U. Khan, K. Young, A. Gaucher, S. De, R. J. Smith, I. V. Shvets, S. K. Arora, G. Stanton, H. Y. Kim, K. Lee, G. T. Kim, G. S. Duesberg, T. Hallam, J. J. Boland, J. J. Wang, J. F. Donegan, J. C. Grunlan, G. Moriarty, A. Shmeliov, R. J. Nicholls, J. M. Perkins, E. M. Grieveson, K. Theuwissen, D. W. McComb, P. D. Nellist and V. Nicolosi, Science, 2011, 331, 568 CrossRef CAS PubMed.
  27. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman and M. S. Strano, Nat. Nanotechnol., 2012, 7, 699 CrossRef CAS PubMed.
  28. Y. C. Liu, H. Y. Kang, L. F. Jiao, C. C. Chen, K. Z. Cao, Y. J. Wang and H. T. Yuan, Nanoscale, 2015, 7, 1325 RSC.
  29. M. Acerce, D. Voiry and M. Chhowalla, Nat. Nanotechnol., 2015, 10, 313 CrossRef CAS PubMed.
  30. S. J. Xu, Z. Y. Lei and P. Wu, J. Mater. Chem. A, 2015, 3, 16337 CAS.
  31. J. P. Zou, J. Ma, J. M. Luo, J. Yu, J. k. He, Y. T. Meng, Z. Luo, S. K. Bao, H. L. Liu, S. L. Luo, X. B. Luo, T. C. Chen and S. L. Suib, Appl. Catal., B, 2015, 179, 220 CrossRef CAS.
  32. J. F. Zhang, Y. F. Hu, X. L. Jiang, S. F. Chen, S. G. Meng and X. L. Fu, J. Hazard. Mater., 2014, 280, 713 CrossRef CAS PubMed.
  33. N. Tian, H. W. Huang, Y. X. Guo, Y. He and Y. H. Zhang, Appl. Surf. Sci., 2014, 322, 249 CrossRef CAS.
  34. J. Fang, F. Wang, K. Qian, H. Z. Bao, Z. Q. Jiang and W. X. Huang, J. Phys. Chem. C, 2008, 112, 18150 CAS.
  35. L. H. Yuwen, F. Xu, B. Xue, Z. M. Luo, Q. Zhang, B. Q. Bao, S. Su, L. X. Weng, W. Huang and L. H. Wang, Nanoscale, 2014, 6, 5762 RSC.
  36. H. X. Zhong, G. Z. Yang, H. W. Song, Q. Y. Liao, H. Cui, P. K. Shen and C. X. Wang, J. Phys. Chem. C, 2012, 116, 9319 CAS.
  37. G. P. Chen, D. M. Li, F. Li, Y. Z. Fan, H. F. Zhao, Y. H. Luo, R. C. Yu and Q. B. Meng, Appl. Catal., A, 2012, 443–444, 138 CrossRef CAS.
  38. S. Anandan, T. Lana-Villarreal and J. J. Wu, Ind. Eng. Chem. Res., 2015, 54, 2983 CrossRef CAS.
  39. M. C. Mathpal, A. K. Tripathi, M. K. Singh, S. P. Gairola, S. N. Pandey and A. Agarwal, Chem. Phys. Lett., 2013, 555, 182 CrossRef CAS.
  40. N. Goswami, A. Giri and S. K. Pal, Langmuir, 2013, 29, 11471 CrossRef CAS PubMed.
  41. Q. F. Wang, Y. Huang, J. Miao, Y. Zhao and Y. Wang, Electrochim. Acta, 2013, 93, 120 CrossRef CAS.
  42. P. Zhang, Z. L. Mo, L. J. Han, Y. W. Wang, G. P. Zhao, C. Zhang and Z. Li, J. Mol. Catal. A: Chem., 2015, 402, 17 CrossRef CAS.
  43. X. M. Yu, B. Kim and Y. K. Kim, ACS Catal., 2013, 3, 2479 CrossRef CAS.
  44. M. A. Mohamed, W. N. W. Salleh, J. Jaafar, A. F. Ismail, M. A. Mutalib, N. A. A. Sani, S. E. A. M. Asri and C. S. Ong, Chem. Eng. J., 2016, 284, 202 CrossRef CAS.
  45. M. Niu, D. J. Cheng and D. P. Cao, J. Phys. Chem. C, 2014, 118, 5954 CAS.
  46. J. G. Radich and P. V. Kamat, ACS Nano, 2013, 7, 5546 CrossRef CAS PubMed.
  47. K. Li, S. M. Gao, Q. Y. Wang, H. Xu, Z. Y. Wang, B. B. Huang, Y. Dai and J. Lu, ACS Appl. Mater. Interfaces, 2015, 7, 9023 CAS.
  48. J. P. Wilcoxon, J. Phys. Chem. B, 2000, 104, 7334 CrossRef CAS.
  49. H. Wang, L. L. Yu, Y. H. Lee, S. Cramm, W. Eberhardt, A. Mrzel and D. Mihailovic, Nano Lett., 2012, 12, 4674 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: DFT computational method, and capture tests of hydroxyl radical (·OH). See DOI: 10.1039/c6ra03534e

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.