p-BiOI/n-SnS2 heterojunction flowerlike structure with enhanced visible-light photocatalytic performance

Tingting Wanga, Hui Meng*a, Xiang Yub, Yizhu Liua, Hongbin Chenb, Yi Zhua, Jinpeng Tangb, Yexiang Tongc and Yuanming Zhang*b
aCollege of Science and Engineering, Jinan University, Guangzhou 510632, P. R. China. E-mail: tmh@jnu.edu.cn
bDepartment of Chemistry, Jinan University, Guangzhou 510632, P. R. China. E-mail: tzhangym@jnu.edu.cn
cSchool of Chemistry & Chemical Engineering, Sun Yat Sen University, Guangzhou 510275, P. R. China

Received 4th December 2014 , Accepted 26th January 2015

First published on 26th January 2015


Abstract

A BiOI/SnS2 p–n junction is prepared by depositing BiOI particles on flowerlike SnS2 flakes. The heterojunction formed between p-type BiOI and n-type SnS2 brings greatly improved photocatalytic activity compared with pure SnS2 or BiOI in the degradation of Rhodamine B dye under visible light and sunlight irradiation. The improved photocatalytic performance is explained by the synergetic effects of the formation of a p–n junction between BiOI and SnS2. The migration of photogenerated carriers can be promoted by the electric field established at the heterojunction interfaces, leading to effective separation of the photogenerated electron–hole pairs and reduced electron–hole recombination. The BiOI/SnS2 heterojunction will find promising applications in photocatalysis and solar energy conversion.


1. Introduction

Photocatalysis is a hot topic in solar energy conversion and photo-oxidation of organic pollutants in environmental management.1–6 Most photocatalysts are semiconductors of oxides and sulfides such as TiO2,7–9 ZnO,10,11 SnO2,12,13 Bi2O3 (ref. 14 and 15) and CdS.16,17 However the low energy conversion efficiency limits the application of these materials.18,19 Doping with a foreign semiconductor is an effective way to improve the performance of photocatalysts, as the difference between the conduction band and valence band of different semiconductors will benefit the separation of photon-generated electron–hole pairs, reduce the recombination and improve the quantum efficiency.20–23

SnS2 is n-type semiconductor with intriguing optical and electrical properties which makes it an attractive candidate in photocatalysis.24–26 However, the band gap (2.18–2.44 eV) of SnS2 (wider compared with the narrow band gap semiconductor such as BiOI) weakens its photocatalytic activity in the visible light region. Various ways have been tried to improve the visible light absorption of SnS2.26–28 The construction of p–n junction between n-type SnS2 and p-type narrow band gap semiconductors is an effective way to solve this problem, as the external electric field of p–n junction benefits the separation of electron–hole and enhances the photocatalytic capability dramatically.29–31 In our work BiOI, a p-type semiconductor with a narrow band gap of 1.70–1.83 eV, is chosen as a coupling semiconductor to SnS2 to achieve high visible light response. The valence band and conduction band potentials of BiOI are more negative than those of SnS2, which thermodynamically allows the electrons on the conduction band of BiOI transfers to that of SnS2 while the holes remain in BiOI.30–34 The photogenerated electron–hole pairs will be separated effectively by the p–n junction formed at the p-BiOI/n-SnS2 interface,29,34 resulting in a reduced electron–hole recombination. Thus, the p-BiOI/n-SnS2 heterojunction is expected to possess higher visible-light-driven photocatalytic activity than individual SnS2 and BiOI. BiOI exhibits the best photocatalytic activity for the degradation of organic pollutants under visible light irradiation among the BiOX (X = Cl, Br, I) family.35 To reduce the high recombination probability of electron–hole pairs and improve the slow rate of charge transfer, a lot of composites have been synthesized to improve the photocatalytic activity of BiOI, such as BiOI/TiO2,36 BiOBr/BiOI,37 AgI/BiOI (ref. 38) and ZnO/BiOI.39 A latest work reported a BiPO4/BiOI composite with excellent enhancement of visible light utilization ability and photocatalytic activity than pure BiOI.40 But there is no report on the BiOI/SnS2 composite till now.

In this work, for the first time BiOI was coated on the surface of the flowerlike SnS2 to form p-BiOI/n-SnS2 heterojunction. The as-prepared p–n BiOI/SnS2 junction shows high visible-light photocatalytic activity toward the degradation of RhB in water with high stability. To the best of our knowledge, this is the first report on the preparation of p-BiOI and n-SnS2 heterojunction with enhanced visible-light photocatalytic performance.

2. Experimental

All the reagents were purchased from Aladdin reagent (China) Co., Ltd and were used as received without further purification. All water was deionized water prepared in lab.

2.1. Synthesis

Synthesis of flowerlike SnS2. In a typical preparation process, 1 mmol SnCl4·5H2O, 8.0 mmol thioacetamide and 0.3 g polyvinyl pyrrolidone (PVP) were dissolved in 40 ml deionized water. The mixture was vigorously stirred for 30 min to ensure homogeneous dispersion. The solution was transferred to a 50 ml Teflon-lined stainless steel autoclave and maintained at 160 °C for 12 h. After cooling to room temperature the product was collected by centrifugation, followed by washing thoroughly with deionized water and ethanol. The final product was dried at 60 °C in vacuum for 8 h.
Synthesis of BiOI/SnS2. BiOI was coated onto the flowerlike SnS2 by an impregnating hydroxylation method using BiI3 as the precursor. In a typical synthesis, 0.94 g of BiI3 was added into 20 ml ethanol containing 2 drops of 35% HCl. The flowerlike SnS2 were first immersed in BiI3 solution for 30 min followed by a slight rinse with ethanol and then dried in an oven at 80 °C for 1 h to ensure that BiI3 was adhered to the surface of SnS2.

The hydroxylation of BiI3 is proposed by the following reactions:

Bi3+ + 2H2O → BiO(OH) + 3H+

BiO(OH) + H+ + I → BiOI + H2O

2.2. Physical characterizations

The morphologies of the samples were characterized by field emission scanning electron microscopy (FESEM) (JEOL JSM-T300, operated at 10 kV) and field emission transmission electronic micrograph (FETEM) (JEOL SM-6330F, operated at 200 kV). Energy dispersive X-ray spectroscopy (EDX) (JEOL JSM-6330F equipped with an X-ray energy dispersive spectrometer) was used to analyze the element composition. The crystal structure of the samples were determined on an X-ray diffractometer (XRD) (MSALXD-2 with Cu-Kα radiation), D/Max-IIIA, Rigaku Co., Japan, CuK1, λ = 1.54056 Å radiation and selected area electron diffraction (SAED). X-ray photoelectron spectroscopy (XPS) spectra were acquired using an ESCA Lab 250 (Thermo VG) with 200 W Al Kα radiation in twin anode and the distance between X-ray gun and sample is about 1 cm. The analysis chamber pressure is about 2 × 10−7 Pa and the pass energy is 20 eV for high resolution scans. The Brunauer–Emmett–Teller (BET) surface areas of the samples were investigated on a TriSTRA 3000 instrument at liquid nitrogen temperature (77.3 K). The light absorption ability was analyzed by UV-vis diffuse reflectance spectra (DRS) (Cary 5000) using BaSO4 as a reference.

2.3. Photocatalytic performance characterizations

Photocatalytic activities of the samples were evaluated by photo-bleaching of Rhodamine B (RhB) under UV-vis (385–740 nm) and visible light (λ > 420 nm) irradiations. Experiments were carried out in a 250 ml cylindrical-shaped-glass reactor at room temperature in the air and at neutral pH conditions. A suspension of 100 mg photocatalyst and 100 ml aqueous solution of RhB (10 mg l−1) was magnetically stirred for at least 2 h before irradiation to establish an adsorption–desorption equilibrium. The optical system used for the photocatalytic reaction consisted of a xenon lamp (XQ350W) and a 420 nm cutoff filter, which was placed under the reaction cell to completely remove all incoming light with wavelengths shorter than 420 nm to ensure irradiation with visible light only. At specific time intervals, 4 ml suspension were sampled and centrifugated at 5000 rpm for 10 minutes to remove the photocatalyst powder. The concentration of remnant RhB in the solution after irradiation was analyzed by UV-vis spectrophotometer by recording the variations of the absorbance band maximum at λ = 554 nm (RhB).

In the cycling test, the most active catalyst was regenerated by centrifugation after the degradation and washed with water for 3–4 times for the complete removal of dye from the catalysts. Then the catalyst was dried at 60 °C for 8 h and reused for another cycle keeping the concentration of dye and photocatalyst unchanged.

2.4. Photoelectrochemical measurements

In the fabrication of the photo-anode, 20 mg as-prepared photocatalyst powder and 2 ml ethanol solution were mixed homogeneously. The mixture was sprayed on an FTO (fluorine doped tin oxide) glass and allowed to dry under ambient conditions. Photocurrent was measured by an electrochemical workstation (SP-150 from Bio-Logic Science Instruments) in a standard three-electrode system with the as-prepared samples as the working electrodes with an active area of ca. 6 cm2, a Pt wire as the counter electrode, and a saturated calomel electrode (SCE) as reference electrode. A 350 W Xe lamp equipped with an ultraviolet cutoff filter (λ > 420 nm) was utilized as the visible-light source. A 0.1 M Na2SO4 aqueous solution was used as the electrolyte.

3. Results and discussion

3.1. Physical characterization of the BiOI/SnS2 composite

FESEM image (Fig. 1a) shows the flowerlike morphology of the obtained SnS2 with the average overall diameter of 1–1.5 μm. A high-magnified picture in Fig. 1b shows that the SnS2 flower is comprised by several self-assembled irregular plates in the similar way like the petals connected with each other to form a Rieger begonia (inset of Fig. 1b). Fig. 1c and d are SEM images of the flowerlike SnS2 with BiOI deposited on the surface. SnS2 keeps its structure after the deposition, the flakes of the SnS2 flower are covered by BiOI particles with the diameter of 5–40 nm.
image file: c4ra15770b-f1.tif
Fig. 1 FESEM images of self-organized SnS2 (a and b) and BiOI/SnS2 (c and d). The inset picture in (b) is the picture of a Rieger begonia for illustrating.

It is proposed that the surfactant (PVP) present in the solution plays a key role in the assembly of SnS2 flakes into the flower structure. In the crystal growth process the surface of the crystal is covered by PVP molecules, the van der Waals force and hydrogen bond play crucial role in the formation of the flowerlike structure. The formation of flowerlike morphology is based on heterogeneous nucleation. SnS2 nanoparticles are formed during the initial stage of the hydrothermal process, PVP molecules stabilize the particles and prevent the aggregation of the particles. After a heterogeneous nucleation process the particles grow along the specific energetically favorable sites that are likely to be promoted by the surfactant molecules into the flowerlike structure.41

The microstructure of SnS2 was further characterized by means of TEM and HRTEM. As shown in Fig. 2a and b, the SnS2 flowers were decomposed into flakes after the ultrasonic treatment in the sample preparation for TEM observation. The morphology was the same with that observed in FESEM images (Fig. 1a and b). The lattice fringe was observed by HRTEM of a selected area of a single flake as shown in Fig. 2c. The distance between the lattice fringes is 0.278 nm, corresponding to the d-spacing of (011) lattice planes of hexagonal phase SnS2. The SAED in Fig. 2d shows hexagonal array of spots, which revealed the single crystal structure of the flake. The d value of the (011) facet from the SAED is in accordance with that from HRTEM. From SAED it is concluded that the [001] orientation is the preponderant growth direction of the SnS2 flake, which is parallel to the (011) facet.


image file: c4ra15770b-f2.tif
Fig. 2 FETEM images (a–c) and SEAD pattern (d) of as prepared SnS2.

TEM verifies the formation of BiOI/SnS2 p–n junction. The size and morphology of both BiOI and SnS2 in the TEM images of Fig. 3a and b are in consistent with the SEM images in Fig. 1c. Fig. 3b is an enlarged picture of the BiOI/SnS2 heterojunction, from which it is observed that the deposited particles disperse uniformly throughout the surface of the SnS2 flake and the particle size is 5–20 nm. In TEM the big particles of 40 nm in SEM are not found, this may be caused by the ultrasonic treatment in the TEM sample preparation, only leaving small particles on the surface. The crystal property of the BiOI/SnS2 heterojunction is studied by HRTEM as shown in Fig. 3c. Two d values of lattice spacing are indexed to be 0.278 and 0.301 nm, corresponding to the spacing of the (011) plane of SnS2 and (012) plane of BiOI, respectively. The (011) peak of SnS2 and (012) peak of BiOI are the strongest peak presented in the XRD of Fig. 4a and b, which indicates the presence of the mixed phase of SnS2 and BiOI in the flowerlike composite. The SAED pattern (Fig. 3d) indexed out the diffraction patterns of (012), (110) and (122) planes of BiOI. The (012) plane is also observed in the HRTEM in Fig. 3c and the (110) plane is parallel to the growth direction of BiOI. The (011), (102) and (022) facets of SnS2 are also indexed.


image file: c4ra15770b-f3.tif
Fig. 3 FETEM images (a–c) and SEAD pattern (d) of as prepared BiOI/SnS2 flowerlike heterojunction.

image file: c4ra15770b-f4.tif
Fig. 4 EDS mapping of BiOI/SnS2. The elements presented are (a), Sn (b), S (c), Bi (d), O (e), I (f).

Energy dispersive X-ray spectroscopy (EDS) mapping were conducted to study the detailed information about the elemental dispersion of Sn, S, Bi, O and I elements in the BiOI/SnS2 composite (Fig. 4). Fig. 4b and c proved the main composition of the flake is SnS2 since Sn and S are predominant elements throughout the flake. It should be noted that Bi, O and I elements are randomly distributed throughout the SnS2 flake and the density of these three elements is not so high as Sn and S, this is because BiOI exist in particles on the surface of SnS2 flakes, which is in accordance with the morphology information from the SEM and TEM results.

XRD is used to study the crystal composition of the BiOI/SnS2 composite. Fig. 5 shows the XRD patterns of the as-synthesized pure SnS2, BiOI and BiOI/SnS2 composites. From Fig. 5a and b hexagonal SnS2 phase and tetragonal BiOI phase are identified from standard card JCPDS 89-2028 and JCPDS 73-2062, respectively. BiOI/SnS2 composites are composed by tetragonal BiOI phase and hexagonal SnS2 phase, which is proved by the coexistence of all characteristic peaks of both SnS2 and BiOI phases (Fig. 5c). It is found the full width at half maximum (FHWM) of the BiOI peaks in Fig. 5c is much wider than that of Fig. 5b, this is a sign of smaller crystal size of BiOI in the BiOI/SnS2 composites than pure BiOI.


image file: c4ra15770b-f5.tif
Fig. 5 XRD patterns of (a) pure SnS2, (b) pure BiOI, (c) BiOI/SnS2.

The chemical state and valence band states of the elements on the surface of the BiOI/SnS2 composite are studied with the X-ray photoelectron spectroscopy (XPS). In this work, XPS results can also provide useful information on the structure of the material and help to explain the enhanced photocatalytic performance. XPS spectrum in Fig. 6a disclosed the presence of Bi, O, I, Sn and S elements. The result is in accordance with the XRD and EDS results. Typical high-resolution XPS spectra of Sn 3d and Bi 4f are shown in Fig. 6b and d. It is clear that the binding energies of Sn 3d5/2 (486.4 eV) and Sn 3d3/2 (494.9 eV) in BiOI/SnS2 composites are higher than those of pure SnS2 (487.1 eV and 493.4 eV), which is characteristic of Sn4+.34,42,43 At the same time the binding energies of Bi 4f7/2 and Bi 4f5/2 peaks of SnS2/BiOI are located at 159.0 eV and 164.3 eV which are lower than those of pure BiOI (159.4 eV and 164.3 eV), which is characteristic of Bi3+.30–32 This is explained by the fact that the conduction band edge of BiOI is lower than that of SnS2, and the Fermi level of BiOI is lower than that of SnS2, so that the electrons on Sn atom can transfer to the Bi atom in BiOI dispersed on the surface of SnS2, which changes the outer electron cloud density of Sn and Bi and makes the binding energies of Sn 3d5/2 and Sn 3d3/2 increase and that of Bi 4f7/2 and Bi 4f5/2 decrease.33,34 The interaction of the Fermi level electron and electron cloud may change the property of individual SnS2 and BiOI, leading to new properties. Two peaks at 161.4 eV and 162.5 eV shown in Fig. 6c are attributed to S 2p3/2 and S 2p1/2, which are characteristic of S2− in SnS2.34,42,43 Fig. 6e shows the high-resolution XPS spectrum of O 1s, which can be fitted into two peaks. The main peak at 530.9 eV belongs to the Bi–O bonds in (BiO)22+ an BiOI, and the peak at 532.2 eV is assigned to the hydroxyl groups on the surface.30–32 As to the high-resolution spectra of the I 3d (Fig. 6f), two peaks at 619.1 and 630.3 eV attributed to I 3d5/2 and I 3d3/2 respectively, can be ascribed to I in BiOI and BiOI/SnS2 composites.30–32 In conclusion, according to the XPS results, the coexistence of SnS2 and BiOI in the composites is confirmed and the formation of a p–n junction caused the interaction of the Fermi level electron and electron cloud of SnS2 and BiOI.


image file: c4ra15770b-f6.tif
Fig. 6 XPS spectra of SnS2, BiOI, and BiOI/SnS2, survey spectrum (a), Sn 3d (b), S 2p (c), Bi 4f (d), O 1s (e), I 3d (f).

The specific BET surface areas (SBET) and pore structure of the prepared samples were investigated using nitrogen adsorption–desorption measurements, the results are in the ESI. Fig. S1 shows the nitrogen adsorption–desorption isotherms of the samples. According to the Brunauer–Deming–Deming–Teller (BDDT) classification, the isotherms and the shapes of hysteresis loops in Fig. S1 are types IV and H3, indicating the pore-size distributions in the mesoporous regions, which is associated with mesopores formed due to deposition of BiOI particles. As shown in Table 1, the SBET of the BiOI/SnS2 composite is 24.75 m2 g−1, which is sixfold larger in comparison to pure SnS2 (4.09 m2 g−1). Considering the fact that the morphology of the SnS2 remains almost unchanged after the formation of BiOI/SnS2 composite (SEM in Fig. 1), the increased SBET of the BiOI/SnS2 composite compared with pure SnS2 can only be ascribed to the BiOI particles on the surface of the SnS2 flakes. Moreover, the SnS2/BiOI composites also demonstrate nearly 6 times greater pore volume of 0.1475 cm3 g−1, compared to pure SnS2 (0.0255 cm3 g−1). Insert in Fig. S1 is the pore-size distribution (PSD). The PSD curve of pure SnS2 is unimodal with mesopore value of 7.4 nm. The PSD curve of SnS2/BiOI composite is tri-modal with the value of 2.2 nm, 4.6 nm and 29 nm. These results prove that BiOI coatings on SnS2 result in a greater surface area and pore volume, which can benefit the availability of more surface active sites and facilitate mass transfer of the reactants. The larger SBET and pore volume is an important factor leading to an enhanced photocatalytic activity.

Table 1 Specific BET surface areas, pore parameters for the as-synthesized samples
Samples SBET (m2 g−1) Total volume (cm3 g−1) Peak pore diameter (nm)
SnS2 4.09 0.0255 2.2/4.6/29
SnS2/BiOI 24.75 0.1475 7.4


3.2. Photocatalytic characterization of the BiOI/SnS2 composite

The optical absorption properties of the unmodified SnS2, BiOI and BiOI/SnS2 are characterized by UV-vis diffuse reflectance spectroscopy. Fig. 7a shows that pure SnS2 primarily absorbs the ultraviolet light with wavelength below 400 nm, which is determined by the intrinsic band gap of SnS2. In contrast, BiOI and BiOI/SnS2 exhibit an obvious red shift and show strong absorptions in the wavelength ranging from 400 to 650 nm. This indicates that the incorporation of BiOI can effectively extend the light absorption edge of SnS2 from ultraviolet into visible region. Based on the theory of optical absorption for direct band gap semiconductors, the band gap energy (Eg) of the samples are calculated according to the following equation:17,44
αhv = A(hvEg)1/2
where α, h, v, Eg, A are the absorption coefficient, Plank constant, light frequency, band gap energy, and a constant, respectively.44,45 By extrapolating the linear portion of (αhv)2 vs. (hv) curves (Fig. 7b), the Eg of SnS2, BiOI and BiOI/SnS2 are estimated to be about 2.45, 1.82, 2.02 eV, respectively. The Eg value of BiOI/SnS2 is between those of pure SnS2 and BiOI, this proves the formation of the BiOI/SnS2 p–n junction.

image file: c4ra15770b-f7.tif
Fig. 7 UV-vis diffuse reflectance spectra (A) and (B) the plot of (αhv)1/2 vs. band gap energy (hv) of pure SnS2, pure BiOI, BiOI/SnS2.

The conduction band (CB) of SnS2 and valence band (VB) of BiOI can be determined by the photoelectrochemical method. In heterojunction the flow direction of electrons depends on their relative band edge positions. The photocurrent spectra of SnS2 and BiOI are shown in Fig. S2. In the photocurrent spectra, the flat band potential of SnS2 is observed at −0.96 V versus SCE. For n-type semiconductor, the flat band potential is closely related to the bottom of the CB. According to the measured Eg value of SnS2 (2.45 eV), the VB is calculated to be 1.50 eV.26,33 The photocurrent onset potential of BiOI gives the value VB edge of +0.53 V versus SCE. The band gap energy of BiOI is 1.82 eV, the CB minimum is calculated to be −1.29 eV.33,45 Scheme 1 shows the energy band edge positions form above calculations. Before contact the conduction band edge and Fermi level of BiOI are lower than those of SnS2. After contact when a p–n junction is formed, the charge carriers transfer from BiOI to SnS2 to form an external electric field at the heterojunction interface.30–33 The Fermi levels of n-SnS2 and p-BiOI are aligned under thermal equilibrium conditions, and the external electric field is directed from n-SnS2 to p-BiOI, thus preventing the charge migrating from n-SnS2 into p-BiOI. Meanwhile, the energy band positions of SnS2 in BiOI/SnS2 heterojunction are shifted toward the downward direction and that p-BiOI in BiOI/SnS2 heterojunction toward the upward direction along with the Fermi level.32,33,46 Therefore, the band positions of BiOI and SnS2 in the heterojunction have a type-II band structure, which means the CB and VB of BiOI lie above those of SnS2.29,33 BiOI absorbs photons of energy greater than the band gap energy under visible-light irradiation, exciting the electrons in the VB to the CB and leaving holes in BiOI. The electrons in the conduction band of the p-type BiOI then transfer to the n-type SnS2, and the holes remain in the valence band of BiOI. The migration of photogenerated carries can be promoted by the inner electric field established at the heterojunction interfaces.30–33,46 Thus, the photogenerated electron–hole pairs will be effectively separated due to the formation of a junction between the p-BiOI and n-SnS2 interface, resulting in a reduced electron–hole recombination.


image file: c4ra15770b-s1.tif
Scheme 1 Schematic diagrams for (A) energy bands of p-BiOI and n-SnS2 before contact and (B) the formation a p–n junction and its energy band diagram at equilibrium and transfer of photoinduced electrons form p-BiOI and n-SnS2 under visible-light irradiation.

The photocatalytic activity of BiOI/SnS2 heterojunction was evaluated by the degradation of RhB, a typical organic contaminant. Fig. 8 presents the changement of RhB concentration (C/C0) with irradiation time at wavelengths in the range of 385–740 nm and visible light irradiations at λ > 420 nm, respectively. For comparison, the photocatalytic activity of Degussa's P25 and the degradation of RhB without photocatalysts were tested under the same conditions. The blank test without photocatalysts shows that RhB is very stable and does not decompose even after long-time illumination with UV-vis or visible light irradiations. A rapid decrease in the concentration of RhB with time is realized in the presence of the as-prepared BiOI/SnS2 heterojunction photocatalyst. For comparison, the photocatalytic property of pure SnS2 and BiOI were also tested. All the samples exhibited much higher photocatalytic activity under UV-vis and visible light irradiation compared to the standard (Degussa's P25). Among them BiOI/SnS2 heterojunction showed the highest RhB removal rate with total decomposition within only 30 min under UV-vis light irradiation and within only 50 min under visible light irradiation, which is 6.6 times and 5.3 times higher than that of the standard (Degussa's P25). After 50 minutes of visible irradiation the degradation efficiency of RhB on pure SnS2 is less than 38%, which is significantly less than that of BiOI/SnS2 heterojunction.


image file: c4ra15770b-f8.tif
Fig. 8 Comparison of visible-light photocatalytic activities, (a) BiOI/SnS2, (b) BiOI, (c) SnS2, (d) P25, (e) adsorption in dark, (f) degradation of 10 mg l−1 RhB without photocatalyst under UV-vis (385–740 nm) (A) and visible light (λ > 420 nm) irradiations (B).

The degradation of RhB by the BiOI/SnS2 heterojunction may involve the following steps: under illumination, CB electrons (e) will accumulate on the surface of SnS2 and holes (h+) on BiOI. Super oxide radicals (O2˙) are then formed by the combination of electrons (e) and oxygen on the surface of the catalyst, which again react with protons and photogenerated electrons to form hydroxyl radical species (OH˙). Finally some highly active species, i.e., superoxide radicals, hydroxyl radicals, hydroxyl radicals are formed, together with holes on BiOI, these species are responsible for the degradation of organic pollutants. The details are presented in below reactions:

BiOI/SnS2 + hv → SnS2(e) + BiOI(h+)

SnS2(e) + O2 → O2˙

O2˙ + H+ → HO2˙

HO2˙ + H+ + SnS2(e) → H2O2

H2O2 + SnS2(e) → OH˙ + OH

RhB + OH˙ → degradation products

RhB + BiOI(h+) → degradation products

The remarkable improvement of the photocatalytic performance of BiOI/SnS2 heterojunction is explained by the following two reasons: (i) the photo absorption range of SnS2 is extended to the visible region by the composition with BiOI due to the sensitization of BiOI. (ii) The p–n junction formed between the p-type BiOI and n-type SnS2 helps in separating electron–hole pairs at the interface. The separated electrons and holes are then free to initiate the degradation reaction of RhB dye adsorbed on the surfaces of photocatalyst with enhanced photocatalytic activity.

The regeneration and stability of the photocatalyst are the two major factors to be considered in practical applications. The results for the recycling performance of the BiOI/SnS2 heterojunction are shown in Fig. 9. It is obvious that photocatalytic efficiency does not exhibit significant loss after ten recycles, indicating that the BiOI/SnS2 heterojunction displays high stability and do not suffer from photocorrosion during the photocatalytic degradation of RhB.


image file: c4ra15770b-f9.tif
Fig. 9 Cycling degradation curves of BiOI/SnS2 on the degradation of 10 mg l−1 RhB under UV-visible light (385–740 nm) irradiations.

To further investigate the photocatalytic mechanism, the transient photocurrent responses of the samples were recorded via several irradiation on–off cycles. The representative plots are shown in Fig. 10. When the light was turned on the photocurrent reached a high point and followed by a fast decline, forming a spike, then decreased gradually until a constant value. The photocurrent was instantaneously close to zero as long as the light was switched off. The initial anodic photocurrent spike is caused by the separation of electron–hole pairs at the BiOI/SnS2 p–n junction interface.32,47 Holes move to the BiOI surface, where they are trapped or captured by reduced species in the electrolyte. At the same time the electrons are transported to the conduction band of SnS2. The decay of the photocurrent is caused by the recombination of photogenerated electron–hole pairs. Instead of capturing electrons from the electrolyte, a portion of the holes at the surface of BiOI recombine with electrons from the conduction band of SnS2 or just accumulate at the surface.32,39,47 The electrons in the conduction band of SnS2 take part in the reduction of the photogenerated oxidized species in the electrolyte, which results in the decay of the photocurrent. When the seperation and recombination of the electron–hole pairs reach an equilibration, a constant current is reached. When the light is turned off, the holes accumulated at the BiOI surface and the electrons in SnS2 conduction band are recombined simultaneously, leading to a fast drop of the photocurrent to zero. It can be seen from Fig. 10 that BiOI/SnS2 p–n junction exhibits a prompt generation of photocurrent with good reproducibility. This is because the p–n junction facilitates the separation of electron–hole pairs by the external electrostatic field at the heterojunction interface, which is the origin of the synergetic effect resulting in the high photocatalytic performance of the BiOI/SnS2 p–n junction.


image file: c4ra15770b-f10.tif
Fig. 10 Comparison of transient photocurrent response of (a) pure SnS2, (b) pure BiOI, (c) BiOI/SnS2 in 0.1 M Na2SO4 aqueous solution under visible-light irradiation vs. SCE.

4. Conclusions

In summary, a novel p-BiOI/n-SnS2 heterojunction was synthesized by coating BiOI on the flowerlike SnS2. The formation of BiOI/SnS2 heterojunctions was confirmed by XRD, TEM and XPS. The wide band gap of SnS2 was effectively sensitized in the visible region, which is confirmed by UV-vis DRS. The p–n junction greatly improves the separation of the photogenerated charge carriers, which is proved by the transient photocurrent response experiments. A dramatic enhancement in the visible light photocatalytic performance of BiOI/SnS2 heterojunction is observed, which is explained by the synergetic effects of several factors including the strong visible absorption, the formation of p–n junction, and the formation of an external electrostatic field. The easy synthesis and high degradation efficiency under UV and visible light makes the novel photocatalyst a very promising candidate for application in the degradation of organic pollutants.

Acknowledgements

This work was supported by National Natural Science Foundation of China (21276104, 21106190, 21476096), Program of the Pearl River Young Talents of Science and Technology in Guangzhou (2013J2200040).

References

  1. K. Maeda, K. Teramura, D. Lu, T. Takata, N. Saito, Y. Inoue and K. Domen, Photocatalyst Releasing Hydrogen from Water, Nature, 2006, 440, 295 CrossRef CAS PubMed .
  2. S. U. M. Khan, M. Al-Shahry and W. B. Ingler, Efficient Photochemical Water Splitting by a Chemically Modified n-TiO2, Science, 2002, 297, 2243–2245 CrossRef CAS PubMed .
  3. J. S. Zhang, J. H. Sun, K. Maeda, K. Domen, P. Liu, M. Antonietti, X. Z. Fu and X. C. Wang, Sulfur-Mediated Synthesis of Carbon Nitride: Band-gap Engineering and Improved Functions for Photocatalysis, Energy Environ. Sci., 2011, 4, 675–678 CAS .
  4. K. Maeda, A. Xiong, T. Yoshinaga, T. Ikeda, N. Sakamoto, T. Hisatomi, M. Takashima, D. Lu, M. Kanehara and T. Setoyama, Photocatalytic Overall Water Splitting Promoted by Two Different Cocatalysts for Hydrogen and Oxygen Evolution under Visible Light, Angew. Chem., 2010, 122, 4190–4193 CrossRef .
  5. C. Pan, J. Xu, Y. Wang, D. Li and Y. Zhu, Dramatic Activity of C3N4/BiPO4 Photocatalyst with Core–Shell Structure Formed by Self-Assembly, Adv. Funct. Mater., 2012, 22, 1518–1524 CrossRef CAS .
  6. W. Zhou, Z. Yin, Y. Du, X. Huang, Z. Zeng, Z. Fan, H. Liu, J. Wang and H. Zhang, Synthesis of Few-Layer MoS2 Nanosheet-Coated TiO2 Nanobelt Heterostructures for Enhanced Photocatalytic Activities, Small, 2013, 9, 140–147 CrossRef CAS PubMed .
  7. S. Chuangchote, J. Jitputti, T. Sagawa and S. Yoshikawa, Photocatalytic Activity for Hydrogen Evolution of Electrospun TiO2 Nanofibers, ACS Appl. Mater. Interfaces, 2009, 1, 1140–1143 CAS .
  8. S. Liu, J. Yu and M. Jaroniec, Tunable Photocatalytic Selectivity of Hollow TiO2 Microspheres Composed of Anatase Polyhedra with Exposed {001} Facets, J. Am. Chem. Soc., 2010, 132, 11914–11916 CrossRef CAS PubMed .
  9. L. K. Tan, M. K. Kumar, W. W. An and H. Gao, Transparent, Well-Aligned TiO2 Nanotube Arrays with Controllable Dimensions on Glass Substrates for Photocatalytic Applications, ACS Appl. Mater. Interfaces, 2010, 2, 498–503 CAS .
  10. J. Wang, Z. Wang, B. Huang, Y. Ma, Y. Liu, X. Qin, X. Zhang and Y. Dai, Oxygen Vacancy Induced Band-Gap Narrowing and Enhanced Visible Light Photocatalytic Activity of ZnO, ACS Appl. Mater. Interfaces, 2012, 4, 4024–4030 CAS .
  11. J. Mu, C. Shao, Z. Guo, Z. Zhang, M. Zhang, P. Zhang, B. Chen and Y. Liu, High Photocatalytic Activity of ZnO–Carbon Nanofiber Heteroarchitectures, ACS Appl. Mater. Interfaces, 2011, 3, 590–596 CAS .
  12. Y. C. Zhang, Z. N. Du, K. W. Li, M. Zhang and D. D. Dionysiou, High-performance Visible-Light-Driven SnS2/SnO2 Nanocomposite Photocatalyst Prepared via In Situ Hydrothermal Oxidation of SnS2 Nanoparticles, ACS Appl. Mater. Interfaces, 2011, 3, 1528–1537 CAS .
  13. A. K. Sinha, M. Pradhan, S. Sarkar and T. Pal, Large-Scale Solid-State Synthesis of Sn–SnO2 Nanoparticles from Layered SnO by Sunlight: a Material for Dye Degradation in Water by Photocatalytic Reaction, Environ. Sci. Technol., 2013, 47, 2339–2345 CrossRef CAS PubMed .
  14. H. Y. Jiang, J. Liu, K. Cheng, W. Sun and J. Lin, Enhanced Visible Light Photocatalysis of Bi2O3 upon Fluorination, J. Phys. Chem. C, 2013, 117, 20029–20036 CAS .
  15. H. Fan, H. Li, B. Liu, Y. Lu and T. Xie, Photoinduced Charge Transfer Properties and Photocatalytic Activity in Bi2O3/BaTiO3 Composite Photocatalyst, ACS Appl. Mater. Interfaces, 2012, 4, 4853–4857 CAS .
  16. Y. Shi, H. Li, L. Wang, W. Shen and H. Chen, Novel α-Fe2O3/CdS cornlike nanorods with enhanced photocatalytic performance, ACS Appl. Mater. Interfaces, 2012, 4, 4800–4806 CAS .
  17. M. Luo, Y. Liu, J. Hu, H. Liu and J. Li, One-Pot Synthesis of CdS and Ni-Doped CdS Hollow Spheres with Enhanced Photocatalytic Activity and Durability, ACS Appl. Mater. Interfaces, 2012, 4, 1813–1821 CAS .
  18. V. Etacheri, M. K. Seery, S. J. Hinder and S. C. Pillai, Oxygen Rich Titania: a Dopant Free, High Temperature Stable, and Visible Light Active Anatase Photocatalyst, Adv. Funct. Mater., 2011, 21, 3744–3752 CrossRef CAS .
  19. N. Roy, Y. Sohn and D. Pradhan, Synergy of Low-energy {101} and High-Energy {001} TiO2 Crystal Facets for Enhanced Photocatalysis, ACS Nano, 2013, 7, 2532–2540 CrossRef CAS PubMed .
  20. R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki and Y. Taga, Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides, Science, 2001, 293, 269–271 CrossRef CAS PubMed .
  21. M. Wang, J. Iocozzia, L. Sun, C. Lin and Z. Lin, Semiconductor TiO2 Nanotube Arrays for Photocatalysis, Energy Environ. Sci., 2014, 7, 2182–2202 CAS .
  22. Q. Mu, Q. Zhang, H. Wang and Y. Li, Facile Growth of Vertically Aligned BiOCl Nanosheet Arrays on Conductive Glass Substrate with High Photocatalytic Properties, J. Mater. Chem., 2012, 22, 16851–16857 RSC .
  23. C. Wang, Y. Guo, Y. Yang, S. Chu, C. Zhou, Y. Wang and Z. Zou, Sulfur-Doped Polyimide Photocatalyst with Enhanced Photocatalytic Activity under Visible Light Irradiation, ACS Appl. Mater. Interfaces, 2014, 6, 4321–4328 CAS .
  24. R. Wei, J. Hu, T. Zhou, X. Zhou, J. Liu and J. Li, Ultrathin SnS2 Nanosheets with Exposed {001} Facets and Enhanced Photocatalytic Properties, Acta Mater., 2014, 66, 163–171 CrossRef CAS PubMed .
  25. Y. C. Zhang, J. Li, M. Zhang and D. D. Dionysiou, Size-Tunable Hydrothermal Synthesis of SnS2 Nanocrystals with High Performance in Visible Light-Driven Photocatalytic Reduction of Aqueous Cr(VI), Environ. Sci. Technol., 2011, 45, 9324–9331 CrossRef CAS PubMed .
  26. X. Hu, G. Song, W. Li, Y. Peng, L. Jiang, Y. F. Xue, Q. Liu, Z. G. Chen and J. Q. Hu, Phase-Controlled Synthesis and Photocatalytic Properties of SnS, SnS2 and SnS/SnS2 Heterostructure Nanocrystals, Mater. Res. Bull., 2013, 48, 2325–2332 CrossRef CAS PubMed .
  27. 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, Hierarchical Assembly of Ultrathin Hexagonal SnS2 Nanosheets onto Electrospun TiO2 Nanofibers: Enhanced Photocatalytic Activity Based on Photoinduced Interfacial Charge Transfer, Nanoscale, 2013, 5, 606–618 RSC .
  28. X. An, C. Y. Jimmy and J. Tang, Biomolecule-Assisted Fabrication of Copper Doped SnS2 Nanosheet–Reduced Graphene Oxide Junctions with Enhanced Visible-Light Photocatalytic Activity, J. Mater. Chem. A, 2014, 2, 1000–1005 CAS .
  29. Y. J. Wang, Q. S. Wang, X. Y. Zhan, F. M. Wang, M. Safdar and J. He, Visible Light Driven Type II Heterostructures and Their Enhanced Photocatalysis Properties: a Review, Nanoscale, 2013, 5, 8326–8339 RSC .
  30. D. L. Jiang, L. L. Chen, J. J. Zhu, M. Chen, W. D. Shi and J. M. Xie, Novel P–N Heterojunction Photocatalyst Constructed by Porous Graphite-Like C3N4 and Nanostructured BiOI: Facile Synthesis and Enhanced Photocatalytic Activity, Dalton Trans., 2013, 15726–15734 RSC .
  31. D. Hou, X. Hu, P. Hu, W. Zhang, M. Zhang and Y. Huang, Bi4Ti3O12 Nanofibers–BiOI Nanosheets P–N Junction: Facile Synthesis and Enhanced Visible-Light Photocatalytic Activity, Nanoscale, 2013, 5, 9764–9772 RSC .
  32. G. Dai, J. Yu and G. Liu, Synthesis and Enhanced Visible-Light Photoelectrocatalytic Activity of P–N Junction BiOI/TiO2 Nanotube Arrays, J. Phys. Chem. C, 2011, 115, 7339–7346 CAS .
  33. K. H. Reddy, S. Martha and K. M. Parida, Fabrication of Novel P-BiOI/N-ZnTiO3 Heterojunction for Degradation of Rhodamine 6G under Visible Light Irradiation, Inorg. Chem., 2013, 52, 6390–6401 CrossRef CAS PubMed .
  34. F. Yang, G. Han, D. Fu, Y. Chang and H. Wang, Improved Photodegradation Activity of TiO2 via Decoration with SnS2 Nanoparticles, Mater. Chem. Phys., 2013, 140, 398–404 CrossRef CAS PubMed .
  35. X. Z. Zhang, H. Ai, F. L. Jia and L. Z. Zhang, Generalized one-pot synthesis, characterization, and photocatalytic activity of hierarchical BiOX (X = Cl, Br, I) nanoplate microspheres, J. Phys. Chem. C, 2008, 112, 747–753 CAS .
  36. X. Zhang, L. Z. Zhang, T. F. Xie and D. J. Wang, Low-Temperature Synthesis and High Visible-Light-Induced Photocatalytic Activity of BiOI/TiO2 Heterostructures, J. Phys. Chem. C, 2009, 113, 7371–7378 CAS .
  37. J. Cao, B. Y. Xu, H. L. Lin, B. D. Luo and S. F. Chen, Chemical etching preparation of BiOI/BiOBr heterostructures with enhanced photocatalytic properties for organic dye removal, Chem. Eng. J., 2012, 91–99 CrossRef CAS PubMed .
  38. H. F. Cheng, W. J. Wang, B. B. Huang, Z. Y. Wang, J. Zhan, X. Y. Qin, X. Y. Zhang and Y. Dai, Tailoring AgI nanoparticles for the assembly of AgI/BiOI hierarchical hybrids with size-dependent photocatalytic activities, J. Mater. Chem. A, 2013, 1, 7131–7136 CAS .
  39. J. Jiang, X. Zhang, P. B. Sun and L. Z. Zhang, ZnO/BiOI Heterostructures: Photoinduced Charge-Transfer Property and Enhanced Visible-Light Photocatalytic Activity, J. Phys. Chem. C, 2011, 115, 20555–20564 CAS .
  40. Y. F. Liu, W. Q. Yao, D. Li, R. L. Zong, M. Zhang, X. G. Ma and Y. F. Zhu, Enhancement of visible light mineralization ability and photocatalytic activity of BiPO4/BiOI, Appl. Catal., B, 2015, 163, 547–553 CrossRef CAS PubMed .
  41. J. Ota, P. Roy, S. K. Srivastava, B. B. Nayak and A. K. Saxena, Morphology Evolution of Sb2S3 under Hydrothermal Conditions: Flowerlike Structure to Nanorods, Cryst. Growth Des., 2008, 8, 2019–2023 CAS .
  42. J. Wang, X. Li, X. Li, J. Zhu and H. Li, Mesoporous Yolk–Shell SnS2–TiO2 Visible Photocatalysts with Enhanced Activity and Durability in Cr(VI) Reduction, Nanoscale, 2013, 5, 1876–1881 RSC .
  43. Y. Lei, S. Song, W. Fan, Y. Xing and H. Zhang, Facile Synthesis and Assemblies of Flowerlike SnS2 and In3+-Doped SnS2: Hierarchical Structures and Their Enhanced Photocatalytic Property, J. Phys. Chem. C, 2009, 113, 1280–1285 CAS .
  44. M. Shang, W. Wang and L. Zhang, Preparation of BiOBr Lamellar Structure with High Photocatalytic Activity by CTAB as Br Source and Template, J. Hazard. Mater., 2009, 167, 803–809 CrossRef CAS PubMed .
  45. X. Xiao and W. D. Zhang, Facile Synthesis of Nanostructured BiOI Microspheres with High Visible Light-Induced Photocatalytic Activity, J. Mater. Chem., 2010, 20, 5866–5870 RSC .
  46. L. Chen, S. F. Yin, S. L. Luo, R. Huang, Q. Zhang, T. Hong and C. T. Peter, Bi2O2CO3/BiOI Photocatalysts with Heterojunctions Highly Efficient for Visible-Light Treatment of Dye-Containing Wastewater, Ind. Eng. Chem. Res., 2012, 51, 6760–6768 CrossRef CAS .
  47. J. Yu, G. Dai and B. Huang, Fabrication and Characterization of Visible-Light-Driven Plasmonic Photocatalyst Ag/AgCl/TiO2 Nanotube Arrays, J. Phys. Chem. C, 2009, 113, 16394–16401 CAS .

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra15770b

This journal is © The Royal Society of Chemistry 2015