Fabrication and photocatalytic performances of BiOCl nanosheets modified with ultrafine Bi2O3 nanocrystals

Xu Zhanga, Guangqing Xu*a, Jiajia Hua, Jun Lva, Jianmin Wanga and Yucheng Wu*ab
aSchool of Materials Science and Engineering, Hefei University of Technology, Hefei 230009, China. E-mail: gqxu1979@hfut.edu.cn; ycwu@hfut.edu.cn; Tel: +86 551 62901372
bAnhui Provincial Key Laboratory of Advanced Functional Materials and Devices, Hefei University of Technology, Hefei 230009, China

Received 17th April 2016 , Accepted 12th June 2016

First published on 13th June 2016


Abstract

BiOCl nanosheets modified with ultrafine Bi2O3 nanocrystals were synthesized by chemical deposition with the assistance of ascorbic acid and Pluronic F127. The structure, morphology, elemental composition and optical absorption performance were characterized using X-ray diffractometry, scanning electron microscopy, high-resolution transmission electron microscopy, X-ray photoelectron spectroscopy and UV-vis diffuse reflection spectroscopy, respectively. Square-like BiOCl nanosheets with ultrafine Bi2O3 nanocrystals evenly distributed on the surface were obtained. The photocatalytic activity of the composite was tested by degrading a 40 mg L−1 methyl orange solution under UV light illumination. A Bi2O3/BiOCl composite prepared with a Bi(NO3)3 concentration of 0.01 M achieved the highest photocatalytic rate of 100% in 6 min under UV light illumination. Repeated experiments on the degradation processes and recycling experiments on the photocatalysts were conducted under the same conditions. The results show that the as-synthesized Bi2O3/BiOCl composites possess excellent photocatalytic activity and outstanding recyclability. The photocatalytic mechanism of the composite is discussed.


1 Introduction

With the development of industry, environmental contamination has become more and more serious and it is extremely urgent to control pollution. As a rapidly growing green method for treating polluted water and air, photocatalytic degradation using a semiconductor as a catalyst has attracted more and more attention.1,2 TiO2 is one of the earliest and most widely studied semiconductors used as a photocatalyst owing to its excellent photocatalytic properties.3,4 Currently, more and more other photocatalysts have been studied as alternatives to TiO2, such as metallic oxides,5 metal sulfides6 and ternary system compounds.7 Among these semiconductors, bismuth-containing compounds have attracted a lot of attention. Numerous studies have concentrated on Bi2O3,8 BiOCl,9 Bi2S3,10 Bi4Ti3O12,11 Bi2MoO6 (ref. 12) and other bismuth compounds, revealing that bismuth-containing compounds are promising in the photocatalytic degradation of organic compounds and photocatalytic water splitting.

BiOCl is the most widely studied bismuth system photocatalyst owing to its high photocatalytic activity and high chemical stability.13 Double layers of Cl atoms and [Bi2O2] layers along the c-axis are arranged alternately and form the layered structure of BiOCl. The structure can also be expressed as [Cl–Bi–O–Bi–Cl],14 which possesses enough space to polarize the corresponding atoms and atomic orbitals. The induced dipole moment could separate photoinduced electrons and holes effectively, which can decrease the recombination rate of photoinduced electrons and holes, resulting in the high photocatalytic activity of BiOCl.

In order to improve the photocatalytic properties of BiOCl, numerous studies have concentrated on the modification of BiOCl including control of the crystal planes15 and modification with metal nanoparticles16 and semiconductors.17–19 Single-crystalline BiOCl nanosheets with exposed {001} and {010} facets were selectively synthesized by Zhang et al. via a facile hydrothermal route and were found to have different photocatalytic properties.15 Ao et al. prepared square-like BiOCl nanoplates with Ag@AgCl nanoparticles deposited via an in situ ultraviolet (UV) reduction method and their photoactivity was enhanced fivefold compared with that of pure BiOCl.16 Li et al. synthesized BiOCl–TiO2 heterostructures with TiO2 nanoparticles deposited on the surface of BiOCl nanoplates via a hydrothermal method. The obtained BiOCl–TiO2 heterostructures exhibited much higher photocatalytic efficiency than that of single-phase BiOCl and TiO2.17 Bi2O3/BiOCl heterojunctions were synthesized via the reaction of Bi2O3 and HCl and displayed excellent photocatalytic performance in the visible-light region.20

In our previous study, Bi and Bi2O3 nanocrystals were successfully deposited on BiOCl nanosheets by in situ reduction and oxidation, which can eventually enhance the photocatalytic performance.21,22 To further control the deposition process, Bi2O3 nanocrystals were obtained on BiOCl nanosheets by a facile and simple method with the assistance of ascorbic acid and Pluronic F127. Well-dispersed Bi2O3 nanocrystals can be controlled to be less than 10 nm in size, which is beneficial for promoting charge transfer between BiOCl and Bi2O3 and significantly enhancing the photocatalytic performance. Heterojunctions of Bi2O3/BiOCl restrict the recombination of photogenerated electron–hole pairs, resulting in an enhancement in the photocatalytic properties of BiOCl nanosheets.

2 Experimental

2.1 Synthesis

Bismuth nitrate pentahydrate (Bi(NO3)3·5H2O), sodium chloride (NaCl), ethylene glycol (EG), mannitol, ascorbic acid (AA) and terephthalic acid were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Pluronic F127 was purchased from Sigma-Aldrich Co. All reagents were of analytical grade and were used as received without further purification.

BiOCl nanosheets were synthesized by a solvothermal method in EG solution. In a typical process, 2 mmol Bi(NO3)3·5H2O was dissolved in 25 mL EG, and 2 mmol NaCl was dissolved in 25 mL deionized water. The aqueous solution of NaCl was added dropwise into the Bi(NO3)3 solution with continuous stirring. After being stirred for 30 min, the mixed solution was transferred to a Teflon-lined stainless-steel autoclave. The solvothermal synthesis was performed at 150 °C for 10 h in an electric oven. The resulting precipitate was centrifuged and washed with ethanol and deionized water several times to remove the chemical residues. Finally, white BiOCl powder was obtained by drying the precipitate at 40 °C in air.

Deposition of Bi2O3 was performed via the ordinary reduction and oxidation of Bi(NO3)3 in an aqueous solution of mannitol containing AA and F127, in which 0.0388 g to 0.776 g Bi(NO3)3 was added to a 20 mL mannitol solution with a concentration of 0.1 M. The solution was stirred continuously for full dissolution, and then mixed with 40 mL Pluronic F127 with a concentration of 1 mM.

Then, 0.2 g BiOCl nanosheets were put into the mixed solution and dispersed sufficiently. Next, 20 mL AA solution (0.1 M) was added to the solution and reacted for 60 min under ultrasonication. The resulting precipitate was washed with deionized water and ethyl alcohol several times to remove the chemical residues. Finally, a Bi2O3/BiOCl composite was obtained by drying the precipitate at 50 °C in an electric oven. Bi2O3/BiOCl composites synthesized with different Bi(NO3)3 concentrations of 1, 5, 10, 15 and 20 mM were labeled as Bi2O3/BiOCl-1, Bi2O3/BiOCl-5, Bi2O3/BiOCl-10, Bi2O3/BiOCl-15 and Bi2O3/BiOCl-20, respectively. As a comparison, yellow Bi2O3 nanoparticles were prepared by the same method without adding BiOCl nanosheets.

2.2 Characterization

The phase structure of the samples was measured by an X-ray diffractometer (D/MAX2500V) using Cu Kα radiation with 2θ ranging from 10° to 70°. The surface morphologies of the samples were observed with a SU8020 field emission scanning electron microscope (FESEM). A JEM-2100F high-resolution transmission electron microscope (HRTEM) was used to observe the morphologies and structures of the samples. X-ray photoelectron spectroscopy (XPS) analysis of the samples was performed using an ESCALAB 250 photoelectron spectrometer with a monochromatic Al Kα X-ray beam (1486.60 eV). The optical absorption performance of the samples was evaluated by a diffuse reflectance spectrometer (UV3600, Shimadzu) using BaSO4 as a reference.

2.3 Photocatalysis

The photocatalytic degradation of MO solution was used to evaluate the photocatalytic properties of the samples. The photocatalytic tests were conducted on an XPA-7 photochemical reactor (Nanjing Xujiang Machine-electronic Company, China) using a 300 W high-pressure mercury lamp as the UV light source (maximum emission wavelength of 365 nm). The distance between the solution and the lamp was kept at 10 cm, and the solution was continuously stirred during the photocatalytic degradation process.

In a typical procedure, 0.01 g photocatalyst was put into 10 mL MO solution with a MO concentration of 40 mg L−1. The suspension was vigorously stirred for 30 min in the dark to achieve adsorption/desorption equilibrium. After being illuminated every 2 min, a 5 mL solution was collected from the suspension and centrifuged. Then, the upper clear solution was tested by a UV1800 spectrometer (Shimadzu, Japan) to record the residual concentration of MO in the solution being photodegraded.

To further study the effect of the active species h+, ˙O2 and HO˙ on the photocatalytic performance of BiOCl nanosheets and Bi2O3/BiOCl composites, the three sacrificial agents NaI, BQ and IPA with concentrations of 10 mM were added, respectively, to the MO solution. The changes in the degradation rate reflect the effects of the corresponding active species on the photocatalytic properties of the photocatalysts. The concentration of HO˙ in the solution during the photocatalytic process was also detected by the terephthalic acid oxidation method.23 Terephthalic acid can react with HO˙ produced by photocatalysts illuminated with a 300 W high-pressure mercury lamp as the UV light source (maximum emission wavelength of 365 nm), and the product, dihydroxyterephthalic acid, exhibits excellent photoluminescence properties with an emission peak wavelength at around 420 nm. A F4500 fluorospectrophotometer (Hitachi) was used to determine the photoluminescence of terephthalic acid after being treated under UV light irradiation with the respective photocatalysts.

3 Results and discussion

3.1 Characterization

Fig. 1 shows the XRD patterns of products obtained in different conditions. Fig. 1(i) compares the phase structures of the products obtained by chemical deposition, solvothermal synthesis and a combination of the two methods. The chemical precipitation of Bi3+ ions in the presence of AA and Pluronic F127 gives rise to a broad diffraction peak at 28.2°, which fits the (111) lattice plane of the cubic phase of Bi2O3 (JCPDS file no. 27-0052), which indicates the existence of Bi2O3 with an ultrafine grain size in the precipitated products. The mean grain size of Bi2O3 can be calculated by the Scherrer equation to be 7.8 nm with a full width at half maximum of 1.25°.
image file: c6ra09919j-f1.tif
Fig. 1 XRD patterns of products obtained in different conditions: (i) comparison of Bi2O3 nanocrystals, BiOCl nanosheets and Bi2O3/BiOCl composites; (ii) comparison of Bi2O3/BiOCl composites prepared with different concentrations of Bi(NO3)3 ranging from 1 to 20 mM.

The solvothermal method is a commonly used method for the synthesis of BiOCl, and the product can be confirmed to be BiOCl (JCPDS no. 06-0249) by the XRD pattern of curve (b). The narrow peaks in curve (b) fit well with the standard diffraction peaks of BiOCl, which indicates the high crystallinity of the product. Curve (c) is the XRD pattern of the product (Bi2O3/BiOCl-10) obtained by a combination of the two methods. Only the diffraction peaks of BiOCl crystals can be observed in curve (c) owing to the small amount and low crystallinity of the deposited Bi2O3. The similarity of the high background noise between curve (c) and curve (a) indicates that the BiOCl nanosheets are covered with a heterogeneous composition.

Fig. 1(ii) shows the XRD patterns of Bi2O3/BiOCl composites prepared with Bi(NO3)3 concentrations ranging from 1 to 20 mM. Almost the same diffraction patterns with diffraction peaks of a single BiOCl phase can be observed, which indicates that no Bi2O3 can be detected in the composite products owing to the small amount and low crystallinity of Bi2O3.

SEM was used to investigate the morphologies of the as-synthesized samples. Fig. 2 shows the SEM morphologies of BiOCl, Bi2O3 nanocrystals and Bi2O3/BiOCl composites prepared with different Bi(NO3)3 concentrations of 5, 10, 15 and 20 mM, respectively. Fig. 2(i) shows the morphology of the solvothermal product without further modification, in which approximately square-like nanosheets with a width of 100–300 nm and a thickness of 30 nm can be observed, and the surface is quite smooth.


image file: c6ra09919j-f2.tif
Fig. 2 SEM morphologies of BiOCl (i), Bi2O3 (ii) and Bi2O3/BiOCl composites prepared with different Bi(NO3)3 concentrations of 5 (iii), 10 (iv), 15 (v) and 20 mM (vi), respectively.

The morphology of the chemical precipitation of Bi3+ ions in the presence of AA and Pluronic F127 is shown in Fig. 2(ii). A reticulation-like structure composed of nanowires can be observed, and some of the nanowires are intertwined to form spheres of a size of hundreds of nanometers.

Fig. 2(iii)–(vi) show the surface morphologies of the composites prepared with different Bi(NO3)3 concentrations. The results show that all the samples are mainly composed of approximately square-like nanosheets similar to the unmodified BiOCl nanosheets. However, the surfaces of these nanosheets are not smooth with ultrafine nanoparticles, which are too small to be characterized clearly by SEM. The amount of these nanoparticles increases with an increase in the Bi(NO3)3 concentration.

The morphologies and microstructures of the as-synthesized samples were further characterized by high-resolution transmission electron microscopy, of which the results are shown in Fig. 3. Fig. 3(i) shows the TEM morphology of the solvothermal product, in which smooth nanosheets with a width of about 100–200 nm can be observed. The smooth surface of the nanosheets indicates that there are no impurities on the nanosheets. The HRTEM morphology of a BiOCl nanosheet is shown in the inset of Fig. 3(i), demonstrating clear lattice fringes that were measured to be 0.278 nm, corresponding to the (110) lattice planes of the tetragonal phase of BiOCl.


image file: c6ra09919j-f3.tif
Fig. 3 TEM morphologies of BiOCl nanosheets (i), Bi2O3 nanocrystals (ii), and Bi2O3/BiOCl-10 (iii) and Bi2O3/BiOCl-20 (iv) composites. The insets show the HRTEM morphologies of the respective products.

Fig. 3(ii) shows the TEM morphology of the products of chemical precipitation with the assistance of AA and Pluronic F127. The synthesized products are composed of a large amount of nanowires, and some of the nanowires aggregated into a sphere with a size of about 300–400 nm. In the top-right inset of Fig. 3(ii) the TEM morphology of the nanowires with high magnification is presented, in which it can be observed clearly that the nanowires are composed of many nanoparticles with a size of a few nanometers. A HRTEM analysis of one of the nanoparticles is shown in the bottom-right inset of Fig. 3(ii), indicating clear lattice fringes of a single crystal with an interplanar spacing of 0.328 nm, which corresponds to the (111) lattice plane of cubic Bi2O3 crystals.

Fig. 3(iii) and (iv) show the morphologies of the composites prepared in Bi(NO3)3 solutions with different Bi3+ concentrations of 10 and 20 mM, respectively. The composite samples display similar nanosheets to the as-synthesized BiOCl nanosheets. However, the surfaces of these nanosheets are not smooth but have many nanoparticles on them. These nanoparticles are well distributed with uniform size. In comparison, the quantity of the nanoparticles in the Bi2O3/BiOCl-20 composite is greater than that in the Bi2O3/BiOCl-10 composite. The insets in both Fig. 3(iii) and (iv) show the HRTEM morphologies of the respective samples, in which a nanoparticle with a size of approximately 10 nm can be observed on the surface of a nanosheet. The clear lattice fringes show that the interplanar spacing is about 0.328 nm, which corresponds to the (111) lattice plane of the cubic Bi2O3 phase (JCPDS file no. 27-0052). Also, the clear and regular lattice fringes indicate the single-crystalline structure of the Bi2O3 nanocrystals. The interplanar spacing of the substrate is 0.277 nm, which corresponds to the (110) lattice plane of the tetragonal BiOCl phase (JCPDS file no. 06-0249). The results of TEM confirmed that Bi2O3 nanocrystals were successfully synthesized and well distributed on the surface of BiOCl nanosheets.

XPS spectra were recorded to reveal the elemental compositions of the samples. Fig. 4 shows the XPS spectra of as-synthesized BiOCl nanosheets, Bi2O3 nanocrystals and the Bi2O3/BiOCl-10 composite, and the survey spectra of all samples are shown in Fig. 4(i). The peak of C 1s at 284.7 eV is present for all samples and originated from the testing process. The main elemental components of the BiOCl nanosheets and Bi2O3/BiOCl-10 nanocomposite are Bi, O and Cl with binding energy peaks at 163 eV (Bi 4f), 442 eV (O 1s) and 198 eV (Cl 2p), respectively, whereas the Bi2O3 nanocrystals only contain Bi and O elements with no Cl element.


image file: c6ra09919j-f4.tif
Fig. 4 XPS spectra of BiOCl nanosheets, Bi2O3 nanocrystals and the Bi2O3/BiOCl-10 composite, (i) survey spectra; (ii) high-resolution spectra of Bi 4f electrons.

The high-resolution XPS spectra of Bi 4f electrons of the three samples are shown in Fig. 4(ii). Double peaks with binding energies of 158.9 and 164.2 eV can be observed in the spectrum of as-prepared BiOCl nanosheets, which correspond to the 4f7/2 and 4f5/2 electron levels of Bi3+ ions in BiOCl. Double peaks in the spectrum of Bi2O3 nanoparticles are observed at 159.3 eV and 164.6 eV, according to the binding energy of the 4f7/2 and 4f5/2 electron levels of Bi3+ ions in Bi2O3. However, the spectrum of the Bi2O3/BiOCl-10 nanocomposite is composed of two broadened peaks at about 159 eV and 164.3 eV, which can each be divided into two peaks by a Gaussian decomposition method. The four peaks agree with the 4f7/2 and 4f5/2 electron levels of Bi3+ in BiOCl and Bi2O3, which reveals that the sample is composed of BiOCl and Bi2O3, which can further confirm that Bi2O3 nanocrystals were successfully synthesized and deposited on BiOCl nanosheets via chemical deposition with the assistance of AA and Pluronic F127.

3.2 Growth mechanism discussion

Ascorbic acid is a commonly used reductant for synthesizing Ag nanomaterials.24 As a complexing agent, AA can react with Bi3+ ions to form bismuth ascorbate, and the Bi3+ ions can be reduced to Bi0 by AA with the assistance of ultrasonication according to the following reaction:
 
3C6H8O6 + 2Bi3+ → 2Bi0 + 3C6H6O6 + 6H+ (i)

The reduced Bi0 is not stable in aqueous solution and can be reoxidized by H2O according to the following reaction:

 
2Bi0 + 3H2O → Bi2O3 + 3H2 (ii)

Bi2O3 nanocrystals were obtained by the reduction and oxidation reactions. The presence of Pluronic F127 restrained the growth of the nanocrystals owing to its PEO-PPO-PEO triblock structure. The triblock copolymer formed micelles due to multimolecular aggregation spontaneously in aqueous solution. The component of PPO provided a local hydrophobic microenvironment in the aqueous phase. So the Bi3+ ions could not move easily, but assembled along long chains of Pluronic F127. That is why we observe nanowire structures of Bi2O3 in the SEM and TEM morphologies.

However, no nanowires or spheres composed of Bi2O3 nanocrystals can be observed in the Bi2O3/BiOCl composites, which indicates that Bi2O3 nanocrystals grew on the surface of BiOCl nanosheets but did not form nanowires or spheres by self-assembly. When BiOCl nanosheets were added to the solution, the nanocrystals tended to nucleate on the surface of the nanosheets. Thus, composites of BiOCl nanosheets deposited with Bi2O3 nanocrystals were obtained.

3.3 Photocatalytic activity

Fig. 5 shows the photocatalytic performance of as-prepared BiOCl nanosheets and Bi2O3/BiOCl composites synthesized in Bi(NO3)3 solutions with different concentrations. Before irradiation, adsorption/desorption equilibrium was achieved by stirring vigorously for 30 min in the dark. Fig. 5(i) shows the absorption rate in the dark and the photocatalytic degradation rate of MO solution irradiated with UV light using BiOCl nanosheets and Bi2O3/BiOCl-1, Bi2O3/BiOCl-5, Bi2O3/BiOCl-10, Bi2O3/BiOCl-15 and Bi2O3/BiOCl-20 composites as the photocatalyst, respectively. The as-prepared BiOCl nanosheets with smooth surfaces displayed relatively poor performance with an adsorption rate of less than 1% after 30 min. All the adsorption rates of the Bi2O3/BiOCl composites were less than 10% and a little higher than that of the BiOCl nanosheets.
image file: c6ra09919j-f5.tif
Fig. 5 Photocatalytic performance of BiOCl nanosheets and Bi2O3/BiOCl composites prepared with different concentrations of Bi3+ ions under UV light illumination: (i) degradation rates for BiOCl nanosheets, Bi2O3/BiOCl-1, Bi2O3/BiOCl-5, Bi2O3/BiOCl-10, Bi2O3/BiOCl-15 and Bi2O3/BiOCl-20 composites; (ii) recycling tests on the Bi2O3/BiOCl-10 composite. The MO concentration is 40 mg L−1.

When under illumination with UV light, the concentration of MO decreased rapidly, which indicates that all the samples are extremely sensitive to UV light and can be stimulated by UV light. After irradiation for 8 min, the degradation rate of MO using as-synthesized BiOCl nanosheets was 46%. The photocatalytic activity of Bi2O3/BiOCl-1 was similar to that of BiOCl owing to the low quantity of deposited Bi2O3 nanocrystals. However, when the concentration of the Bi(NO3)3 solution increased to 5 mM or above, the photocatalytic activity of the Bi2O3/BiOCl composites increased rapidly. All the Bi2O3/BiOCl composites, except for Bi2O3/BiOCl-1, can photodegrade 40 mg L−1 MO solution completely in 8 min under UV irradiation, which is a higher rate than that of the Bi2O3/BiOCl composites obtained by in situ reduction and oxidation in our previous work.22 Among these, the Bi2O3/BiOCl-10 nanocomposite achieved the highest photocatalytic performance and could degrade MO completely in 6 min under UV light.

As a comparison, the photocatalytic performance of TiO2 nanoparticles (P25) under the same conditions was tested, and the results are shown in Fig. 5(i). The photocatalytic performance of P25 was similar to that of unmodified BiOCl nanosheets, with a degradation rate of 48% after being illuminated with UV light for 8 min. Therefore, the optimized Bi2O3/BiOCl-10 nanocomposite displayed much higher photocatalytic activity than that of commercial TiO2 nanoparticles.

The kinetics of the photocatalytic degradation of MO can be described as follows:25

 
image file: c6ra09919j-t1.tif(iii)
here k is the reaction rate constant, and t is the irradiation time. The calculated rate constants k of the different samples are listed in Table 1.

Table 1 k values of BiOCl nanosheets and Bi2O3/BiOCl composites illuminated with UV light
Sample BiOCl Bi2O3/BiOCl-1 Bi2O3/BiOCl-5 Bi2O3/BiOCl-10 Bi2O3/BiOCl-15 Bi2O3/BiOCl-20
k/min−1 0.0748 0.0703 0.4980 0.5135 0.5085 0.4067


From Table 1, the kinetic constant of BiOCl nanosheets is 0.0748 min−1, and all the kinetic constants of the Bi2O3/BiOCl composites are much higher than that of BiOCl nanosheets, except for the Bi2O3/BiOCl-1 composite. The Bi2O3/BiOCl-10 composite possesses the highest kinetic constant with a value of 0.5135 min−1, which is about seven times that of BiOCl nanosheets, which indicates the excellent photocatalytic activity of the sample.

To evaluate the stability of the Bi2O3/BiOCl-10 nanocomposite in the photocatalytic process, the Bi2O3/BiOCl-10 nanocomposite was reclaimed and reused for 10 cycles under the same conditions. Fig. 5(ii) shows the curves for MO degradation when the Bi2O3/BiOCl-10 composite was reused 10 times. The first time, MO could be completely degraded by the Bi2O3/BiOCl-10 nanocomposite in 6 min and the next photocatalytic process needed 8 min to completely degrade MO. The photocatalytic activity of the sample displayed a slight decline. When the photocatalytic process was repeated for a further eight times, the photocatalyst retained the same degradation rate and degraded MO completely in 8 min, which indicates the excellent recyclability and outstanding photocatalytic performance of the Bi2O3/BiOCl-10 composite under UV light illumination.

3.4 Photocatalytic mechanism discussion

3.4.1 Optical absorption performance. UV-vis diffuse reflectance spectroscopy (DRS) is an effective method for characterizing the optical absorption performance of as-synthesized samples. Fig. 6 shows the UV-vis absorption spectra of the BiOCl nanosheets and Bi2O3/BiOCl-10 and Bi2O3/BiOCl-20 composites in the wavelength range from 200 to 800 nm. All samples exhibit strong absorption in the UV region. The absorption edge of the samples lies at approximately 330 to 360 nm. In addition, the BiOCl nanosheets are inert regarding optical absorption in the visible-light region, whereas the absorption intensities of the Bi2O3/BiOCl composites in the visible region are a little higher than that of the BiOCl nanosheets. However, the values of absorption are still very low in the entire visible-light region, which may be ascribed to the low amount of Bi2O3 nanocrystals deposited on BiOCl nanosheets.
image file: c6ra09919j-f6.tif
Fig. 6 UV-vis absorption spectra of BiOCl nanosheets and Bi2O3/BiOCl-10 and Bi2O3/BiOCl-20 composites.

Considering that the optical absorption of the Bi2O3/BiOCl composites displays no evident increase in the UV region, this indicates that optical absorption is not the major reason for the enhancement of the photocatalytic performance of BiOCl nanosheets in the UV region when deposited with Bi2O3 nanocrystals.

3.4.2 Surface area analysis. A key factor that affects adsorption performance is the surface area of samples. The surface areas of the samples are shown in Table 2. The surface area of unmodified BiOCl nanosheets was measured to be 18.1 m2 g−1. When modified with Bi2O3 nanocrystals, the surface area increased a little to 18.3, 19.2, 19.6, 19.4 and 19.4 m2 g−1 for the Bi2O3/BiOCl-1, Bi2O3/BiOCl-5, Bi2O3/BiOCl-10, Bi2O3/BiOCl-15 and Bi2O3/BiOCl-20 composites, respectively, resulting in a slight increase in the adsorption rate, as shown in Fig. 5(i). However, the surface area is not the main factor that enhances the photocatalytic performance.
Table 2 BET surface areas of BiOCl nanosheets and Bi2O3/BiOCl composites
Sample BiOCl Bi2O3/BiOCl-1 Bi2O3/BiOCl-5 Bi2O3/BiOCl-10 Bi2O3/BiOCl-15 Bi2O3/BiOCl-20
Surface area (m2 g−1) 18.1 18.3 19.2 19.6 19.4 19.4


3.4.3 Charge transfer. A schematic diagram of the photocatalytic process with Bi2O3/BiOCl composites is shown in Fig. 7. The photocatalytic processes of BiOCl nanosheets can be described as follows: when BiOCl nanosheets are excited by UV light with energy higher than the band gap, electrons are transferred from the valence band to the conduction band of BiOCl and form photogenerated electron–hole pairs. Photogenerated electrons and holes will be transferred to the surface of BiOCl, and some of them will recombine during the charge transfer process. Electrons on the surface of the photocatalyst can be captured by dissolved O2 in solution to produce the superoxide radical ˙O2−, which can degrade organic compounds into CO2, H2O and other inorganic substances. Also, electrons on the surface of BiOCl will enable hydrogen evolution when captured by H+ ions. Holes can react with OH ions to produce the strongly oxidizing hydroxyl radical HO˙, which is efficacious for degrading organic compounds. Also, holes can degrade organic compounds directly when they are captured by an organic compounds on the surface of BiOCl.
image file: c6ra09919j-f7.tif
Fig. 7 Schematic of the photocatalytic process with Bi2O3/BiOCl composites.

Well-combined Bi2O3/BiOCl composites lead to the transfer of photogenerated electrons and holes between Bi2O3 and BiOCl. The CB of BiOCl (−1.1 eV vs. NHE) is more negative than that of Bi2O3 (0.33 eV), which results in the transfer of electrons from the CB of BiOCl to that of Bi2O3. The photogenerated holes will be transferred from the VB of Bi2O3 to that of BiOCl due to the VB of Bi2O3 (3.13 eV) being more positive than that of BiOCl (2.3 eV). The different transfer directions of photogenerated electrons and holes result in charge separation and a low recombination rate of photogenerated electron–hole pairs. Therefore, the photocatalytic performance of the Bi2O3/BiOCl composites is enhanced.

3.4.4 Active species during the photocatalytic process. As analyzed above, there are three active species, hydroxyl radicals (HO˙), holes (h+) and superoxide radicals (˙O2), in the photocatalytic process of the degradation of MO. In order to reveal which is the key active species in the photocatalytic process, three sacrificial agents were added to the MO solution to consume the corresponding active species rapidly, namely, NaI for h+, BQ for ˙O2 and IPA for HO˙. Fig. 8 shows the photocatalytic performance of as-synthesized BiOCl nanosheets and the Bi2O3/BiOCl-10 composite for the degradation of MO with the addition of different sacrificial agents. All the degradation rates of MO decreased when the three sacrificial agents were added, respectively, which indicates that the three active species play partial roles in the photocatalytic process. For BiOCl nanosheets, when NaI and BQ radicals were added to the MO solution, the degradation rate decreased from 51.7% to 18.3% and 19.5%, respectively, which indicates that h+ and ˙O2 radicals would play key roles in the photocatalytic process. When IPA was added to the MO solution, the photocatalytic rate decreased from 51.7% to 40.9%. The decrease in the photocatalytic degradation rate was only about 10%, which indicates that HO˙ radicals would play a secondary role in the photocatalytic process. However, the changes in the degradation rate were very different for the Bi2O3/BiOCl composite. Decreases of only 17.7% and 8.7% in the degradation rate occurred when NaI and IPA were added to the MO solution, respectively, which indicates that h+ and HO˙ radicals would not play a key role in the photocatalytic process. When BQ was added to the MO solution as an inhibitor of ˙O2 radicals, the decrease in the degradation rate reached 82.4%, which indicates that ˙O2 radicals would play a major role in the photocatalytic process. The results show that the main photocatalytic process is the oxidation of ˙O2 produced by the Bi2O3/BiOCl composite, but not the degradation by h+ and HO˙ radicals. Considering that the photogenerated electrons and holes are equivalent in quantity, some of the holes may take part in the oxygen evolution reaction, and the dissolved O2 captures electrons to form ˙O2 radicals. This process was confirmed by the fact that the degradation rate changed little when N2 was pumped continuously to maintain a N2 atmosphere in the photocatalytic process.
image file: c6ra09919j-f8.tif
Fig. 8 Photocatalytic performance of BiOCl nanosheets and Bi2O3/BiOCl-10 composite for degradation of MO with the addition of different sacrificial agents: NaI for h+, BQ for ˙O2 and IPA for HO˙. The concentrations of all added sacrificial agents are 10 mM. As a comparison, N2 was pumped into MO solution for 1 h to remove the dissolved O2.

The concentrations of HO˙ during the photocatalytic process with BiOCl and Bi2O3/BiOCl nanosheets were also determined by the terephthalic acid oxidation method. Terephthalic acid would be transformed into dihydroxyterephthalic acid when reacted with HO˙. The emission intensities of dihydroxyterephthalic acid, which can indicate the HO˙ concentrations during the photocatalytic process, were measured. The photoluminescence patterns of a terephthalic acid solution after being illuminated with UV light for 15 min with BiOCl and Bi2O3/BiOCl nanosheets are shown in Fig. 9. The weak peak at 430 nm for BiOCl nanosheets reveals the low concentration of HO˙ when the photocatalyst was illuminated with UV light. The peak for Bi2O3/BiOCl is a little larger than that for BiOCl and reached about 300. However, the value for P25 (TiO2) is about 13[thin space (1/6-em)]700, which is about 45 times higher than that for the Bi2O3/BiOCl nanocomposites and about 140 times higher than that for BiOCl nanosheets, which reveals that the HO˙ concentration was much higher than that of BiOCl and the Bi2O3/BiOCl nanocomposites. It is well known that the photocatalytic properties of BiOCl are almost the same as or better than that of P25 (TiO2)26 under UV light illumination, so that the photocatalytic process with BiOCl and Bi2O3/BiOCl nanocomposites is very different to that with P25. HO˙ is not the major oxidizing active species, which is consistent with the analysis of Fig. 9.


image file: c6ra09919j-f9.tif
Fig. 9 Photoluminescence patterns of terephthalic acid after being illuminated with UV light with BiOCl, Bi2O3/BiOCl-10, and Bi2O3/BiOCl-20 nanosheets and TiO2 nanoparticles.

4 Conclusions

A facile method was used to deposit Bi2O3 nanocrystals on BiOCl nanosheets with the assistance of ascorbic acid and Pluronic F127. Bi2O3 nanocrystals were distributed evenly on the surface of BiOCl nanosheets with a size of approximately 10 nm. Bi2O3/BiOCl composites possess excellent photocatalytic activity with high stability under UV light illumination, which is much higher than that of unmodified BiOCl nanosheets. The optimized Bi2O3/BiOCl-10 composite achieved the highest photocatalytic degradation activity and could degrade 40 mg L−1 MO completely in 6 min under UV light illumination. The calculated reaction rate constant (0.5135 min−1) is seven times that of unmodified BiOCl nanosheets (0.0748 min−1). The enhancement mechanism of the modification of Bi2O3 nanocrystals has been discussed by analyzing the entire photocatalytic process, including optical absorption, surface area, charge transfer and surface reactions. The charge transfer between BiOCl and Bi2O3 restricts the recombination of photogenerated electrons and holes, which results in higher photocatalytic activity than that of unmodified BiOCl nanosheets. Photogenerated electrons on Bi2O3 nanoparticles play the dominant role in the photocatalytic process by forming ˙O2 radicals, and photogenerated holes take part in this reaction by providing dissolved O2.

Acknowledgements

This work was supported by the Natural Science Foundation of China (51102071, 51172059 and 51272063), Fundamental Research Funds for the Central Universities (2013HGQC0005), and the Natural Science Foundation of Anhui Province (1408085QE86).

References

  1. A. L. Linsebigler, G. Lu Jr and J. T. Yates, Photocatalysis on TiO2 surfaces: principles, mechanisms, and selected results, Chem. Rev., 1995, 95, 735–758 CrossRef CAS.
  2. N. Serpone and A. V. Emeline, Semiconductor Photocatalysis Past, Present, and Future Outlook, J. Phys. Chem. Lett., 2012, 3, 673–677 CrossRef CAS PubMed.
  3. Y. Kuwahara and H. Yamashita, Efficient photocatalytic degradation of organics diluted in water and air using TiO2 designed with zeolites and mesoporous silica materials, J. Mater. Chem., 2011, 21, 2407–2416 RSC.
  4. H. Irie, Y. Watanabe and K. Hashimoto, Carbon-doped anatase TiO2 powders as a visible-light sensitive photocatalyst, Chem. Lett., 2003, 32, 772–773 CrossRef CAS.
  5. S. Sakthivel, B. Neppolian and M. V. Shankar, et al., Solar photocatalytic degradation of azo dye: comparison of photocatalytic efficiency of ZnO and TiO2, Sol. Energy Mater. Sol. Cells, 2003, 77, 65–82 CrossRef CAS.
  6. Y. Bessekhouad, D. Robert and J. V. Weber, Bi2S3/TiO2 and CdS/TiO2 heterojunctions as an available configuration for photocatalytic degradation of organic pollutant, J. Photochem. Photobiol., A, 2004, 163, 569–580 CrossRef CAS.
  7. T. Cao, Y. Li and C. Wang, et al., Bi4Ti3O12 nanosheets/TiO2 submicron fibers heterostructures: in situ fabrication and high visible light photocatalytic activity, J. Mater. Chem., 2011, 21, 6922–6927 RSC.
  8. G. Cai, L. Xu, B. Wei, J. Che, H. Gao and W. Sun, Facile synthesis of β-Bi2O3/Bi2O2CO3 nanocomposite with high visible-light photocatalytic activity, Mater. Lett., 2014, 120, 1–4 CrossRef CAS.
  9. X. Zhang, Z. Ai and F. Jia, et al., 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.
  10. Y. Bessekhouad, D. Robert and J. V. Weber, Bi2S3/TiO2 and CdS/TiO2 heterojunctions as an available configuration for photocatalytic degradation of organic pollutant, J. Photochem. Photobiol., A, 2004, 163, 569–580 CrossRef CAS.
  11. D. Hou, W. Luo and Y. Huang, et al., Synthesis of porous Bi4Ti3O12 nanofibers by electrospinning and their enhanced visible-light-driven photocatalytic properties, Nanoscale, 2013, 5, 2028–2035 RSC.
  12. Z. Hao, L. Xu, B. Wei, L. Fan, Y. Liu, M. Zhang and H. Gao, Nanosize α-Bi2O3 decorated Bi2MoO6 via an alkali etching process for enhanced photocatalytic performance, RSC Adv., 2015, 5, 12346–12353 RSC.
  13. H. Peng, C. K. Chan and S. Meister, et al., Shape evolution of layer-structured bismuth oxychloride nanostructures via low-temperature chemical vapor transport, Chem. Mater., 2008, 21, 247–252 CrossRef.
  14. C. F. Guo, S. Cao and J. Zhang, et al., Topotactic transformations of superstructures: from thin films to two-dimensional networks to nested two-dimensional networks, J. Am. Chem. Soc., 2011, 133, 8211–8215 CrossRef CAS PubMed.
  15. J. Jiang, K. Zhao and X. Xiao, et al., Synthesis and facet-dependent photoreactivity of BiOCl single-crystalline nanosheets, J. Am. Chem. Soc., 2011, 134, 4473–4476 CrossRef PubMed.
  16. Y. Ao, H. Tang and P. Wang, et al., Deposition of Ag@AgCl onto two dimensional square-like BiOCl nanoplates for high visible-light photocatalytic activity, Mater. Lett., 2014, 131, 74–77 CrossRef CAS.
  17. D. Sun, J. Li and L. He, et al., Facile solvothermal synthesis of BiOCl–TiO2 heterostructures with enhanced photocatalytic activity, CrystEngComm, 2014, 16, 7564–7574 RSC.
  18. A. K. Chakraborty, S. B. Rawal, S. Y. Han, S. Y. Chai and W. I. Lee, Enhancement of visible-light photocatalytic efficiency of BiOCl/Bi2O3 by surface modification with WO3, Appl. Catal., A, 2011, 407, 217–223 CrossRef CAS.
  19. S. M. Zhou, D. K. Ma, P. Cai, W. Chen and S. M. Huang, TiO2/Bi2(BDC)3/BiOCl nanoparticles decorated ultrathin nanosheets with excellent photocatalytic reaction activity and selectivity, Mater. Res. Bull., 2014, 60, 64–71 CrossRef CAS.
  20. S. Y. Chaia, Y. J. Kima, M. H. Junga, A. K. Chakrabortya, D. Jungb and W. I. Lee, Heterojunctioned BiOCl/Bi2O3 a new visible light photocatalyst, J. Catal., 2009, 262(1), 144–149 CrossRef.
  21. J. Hu, G. Xu and J. Wang, et al., Photocatalytic properties of Bi/BiOCl heterojunctions synthesized using an in situ reduction method, New J. Chem., 2014, 38, 4913–4921 RSC.
  22. J. Hu, G. Xu and J. Wang, et al., Photocatalytic property of a Bi2O3 nanoparticle modified BiOCl composite with a nanolayered hierarchical structure synthesized by in situ reactions, Dalton Trans., 2015, 44, 5386–5395 RSC.
  23. K. I. Ishibashi, A. Fujishima and T. Watanabe, et al., Detection of active oxidative species in TiO2 photocatalysis using the fluorescence technique, Electrochem. Commun., 2000, 2, 207–210 CrossRef CAS.
  24. Y. Q. Qin, X. H. Ji, J. Jing, H. Liu, H. L. Wu and W. S. Yang, Size control over spherical silver nanoparticles by ascorbic acid reduction, Colloids Surf., A, 2010, 327, 172–176 CrossRef.
  25. H. Al-Ekabi and N. Serpone, Kinetics studies in heterogeneous photocatalysis. I. Photocatalytic degradation of chlorinated phenols in aerated aqueous solutions over titania supported on a glass matrix, J. Phys. Chem., 1988, 92, 5726–5731 CrossRef CAS.
  26. J. Xiong, G. Cheng and G. Li, et al., Well-crystallized square-like 2D BiOCl nanoplates: mannitol-assisted hydrothermal synthesis and improved visible-light-driven photocatalytic performance, RSC Adv., 2011, 1, 1542–1553 RSC.

This journal is © The Royal Society of Chemistry 2016