Direct conversion mechanism from BiOCl nanosheets to BiOF, Bi7F11O5 and BiF3 in the presence of a fluorine resource

Yandong Kan, Fei Teng*, Yang Yang, Juan Xu and Liming Yang
Jiangsu Engineering and Technology Research Center of Environmental Cleaning Materials (ECM), Jiangsu Key Laboratory of Atmospheric Environment Monitoring and Pollution Control (AEMPC), Jiangsu Joint Laboratory of Atmospheric Pollution Control (APC), Collaborative Innovation Center of Atmospheric Environment and Equipment Technology (AEET), School of Environmental Science and Engineering, Nanjing University of Information Science & Technology, 219 Ningliu Road, Nanjing 210044, China. E-mail: tfwd@163.com; Fax: +86-25-58731090; Tel: +86-25-58731090

Received 7th May 2016 , Accepted 27th June 2016

First published on 28th June 2016


Abstract

In this work, we report the facile conversion from BiOCl nanosheets to BiOF, Bi7F11O5 and BiF3 by a novel ion exchange approach, in which the effects of fluorine source, F/Bi molar ratio and reaction medium (ethanol/water) on the products are mainly investigated. A plausible conversion mechanism is proposed to illustrate the formation of BiOF, Bi7F11O5 and BiF3. Furthermore, the photocatalytic activities of the samples are also investigated. It is amazing that under ultraviolet light irradiation (λ ≤ 420 nm), the activity of the as-formed Bi7F11O5/BiOCl sample is 3.28 times higher than that of BiOCl for the degradation of methyl orange (MO). It is demonstrated that the improved activity is mainly attributed to the formation of Bi7F11O5/BiOCl heterojunction, which has significantly improved the separation and transfer efficiency of charges. It is important that this study provides an initiative post-synthesis strategy to develop new, efficient photocatalysts.


1. Introduction

Solar energy-driven photocatalysis, as a green advanced oxidation technology, has attracted much attention. However, its practical applications is still limited by low quantum efficiency, light utilization efficiency and photoactivity.1–3 Thus, it remains a big challenge to develop inexpensive, highly efficient photocatalysts for practical applications. To date, various photocatalysts have been developed.4–8 Recently, bismuth-based compounds have been intensively investigated due to the outstanding optical and photocatalytic properties, including Bi2O(OH)2SO4,9,10 Bi(C2O4)OH,11 [Bi6O6(OH)3](NO3)3·1.5H2O,12 BiOCl,13 BiPO4,14 Bi2O2CO3,15 BiOF,16–18 BiF3,19 Bi7F11O5,20 BiOI,21 BiVO4[thin space (1/6-em)]22,23 and so on. However, their photocatalytic activities are still not high enough to satisfy the practical needs. Fluorination of semiconductors has been demonstrated to be an effective method to improve the photocatalytic activities.24–26 For example, Chen et al.27 have reported that the photocatalytic activity of the fluorinated ZnWO4 is obviously enhanced for the decomposition of rhodamine B (RhB) under ultraviolet light irradiation. On one hand, the surface fluorination can promote the dye adsorption and the interfacial charge transfer dynamics; on the other hand, the doping of fluorine in lattice can introduce special localized electronic and defect states.28 These results have also been observed in fluorinated TiO2,29 Bi2O3,30 SrTiO3,31 BiPO4,32 etc.

Inspired by the results above, a simple ion exchange approach is developed to convert BiOCl to BiOF, Bi7F11O5 and BiF3, in which fluoride is mainly employed as the fluorination chemical. A plausible conversion mechanism is proposed and discussed; meanwhile, we have investigated their photocatalytic activities for the degradation of MO. The results show that Bi7F11O5/BiOCl heterojunction shows the significantly improved photocatalytic activity. To the best of our knowledge, this ion exchange approach has not been explored for the conversion of bismuth-based semiconductors so far. This work provides an initiative post-synthesis strategy to prepare the other photocatalysts.

2. Experimental

2.1. Sample preparation

All reagents were of analytical grade, purchased from Beijing Chemical Reagents Industrial Company of China, and were used without further purification.
2.1.1. Preparation of BiOCl. BiOCl sample was prepared according to previously reported method.33 Typically, 1.5 mmol of Bi(NO3)3·5H2O was dissolved in 30 mL of ethylene glycol (EG) under stirring at room temperature. After the Bi(NO3)3·5H2O was completely dissolved, 10 mmol of NaCl was added to the above solution. After being stirring for 30 min, the solution was transferred into a 50 mL Teflon-lined stainless steel autoclave, and then maintained at 170 °C for 6 h. After reaction, the autoclave was cooled to room temperature naturally. The resulting sample was centrifuged and washed with ethanol and distilled water three times, and then dried at 60 °C for 3 h.
2.1.2. Ion exchange. Typically, the as-obtained BiOCl (0.2605 g) above was dispersed into a 20 mL of ethanol/water mixture containing the desired amount of NH4F. Then, the dispersion was transferred to Teflon-lined stainless steel autoclave, and maintained at 170 °C for 6 h. The resulting precipitate was centrifuged, washed with ethanol and distilled water three times and dried at 60 °C for 3 h.

To investigate the effect of NH4F amount added, F/Bi molar ratios (RF) were varied from 0 to 0.5, 1, 2, 4, 6 and 8, while keeping the other conditions constant.

To investigate the effect of fluorine source, NH4F was substituted by HF or NaF at RF = 2, while keeping the other conditions constant.

To investigate the effect of reaction medium, VEth/Vw was varied from 0/20 to 2/18, 10/10 and 20/0 at RF = 2, while keeping the other conditions constant.

2.2. Characterization

The crystal structures of the samples were determined by X-ray powder polycrystalline diffractometer (XRD, Rigaku D/max-2550VB), using graphite monochromatized Cu Kα radiation (λ = 0.154 nm), operating at 40 kV and 50 mA. The XRD patterns were obtained in the range of 10–80° (2θ) at a scanning rate of 7° min−1. The samples were characterized on a scanning electron microscope (SEM, Hitachi SU-1510) with an acceleration voltage of 15 keV. The samples were coated with 5 nm-thick gold layer before observations. The fine surface structures of the samples were determined by high-resolution transmission electron microscopy (HRTEM, JEOL JEM-2100F) equipped with an electron diffraction (ED) attachment with an acceleration voltage of 200 kV. UV-vis diffused reflectance spectra of the samples were obtained using a UV-vis spectrophotometer (UV-2550, Shimadzu, Japan). BaSO4 was used as a reflectance standard in a UV-vis diffuse reflectance experiment. Surface area was calculated by the Brunauer–Emmett–Teller (BET) method. Electrochemical impedance spectroscopy (EIS) was performed from 0.01 Hz to 100 kHz at an open circuit potential of 0.3 V and alternating current (AC) voltage amplitude of 5 mV. The photo electrodes were prepared by a dip-coating method and 0.1 M Na2SO4 was used as the electrolyte solution. The data were analyzed by ZSimWin software.

2.3. Photocatalytic degradation reactions

The photodegradation of MO dye was carried out to evaluate the catalytic activities under ultraviolet light irradiation (λ ≤ 420 nm), using a 500 W Xe arc lamp (CEL-HXF 300) equipped with a cutoff filter (λ > 420 nm) as a light source. Typically, 0.1 g of powders was suspended in 200 mL aqueous solution of MO (12.5 mg L−1) under continuous magnetic stirring. Before light-on, the suspension was stirred for 30 min to reach an adsorption–desorption equilibrium of dye molecules on the surface of the photocatalyst. The adsorption curves of the samples are shown in Fig. S1 (ESI), which show that an adsorption–desorption equilibrium of dye molecules on the surface of the photocatalyst can be reached in 30 min. During reaction, 3 mL of suspension was collected at a given interval time and centrifuged to remove the powders. The concentration of dye was determined by using UV-vis spectrophotometer. Herein, the photodegradation of MO can be simplified as a pseudo-first-order kinetic reaction, the apparent reaction rate constants (ka) are calculated by the formula (1) as follows.
 
ln(C0/C) = kat (1)
where C is the concentration of MO remaining in solution after irradiation, and C0 is the initial concentration of MO before irradiation.

2.4. Theoretical calculations

The simulations of energy band structures, total and a part of densities of states (T- and P-DOS) were calculated by density functional theory (DFT) as implemented in the CASTEP. The calculations were carried out using the generalized gradient approximation (GGA) level and Perdew–Burke–Ernzerhof (PBE) formalism for combination of exchange and correlation function. The cut-off energy is chosen as 380 eV, and a density of (3 × 2 × 5) Monkhorst–Pack K-point was adopted to sample the Brillouin zone.

3. Results and discussion

3.1. Effect of F/Bi molar ratio (RF) on the samples

3.1.1. Crystal phases, morphologies and textures of the samples. Fig. 1 shows the XRD patterns of the untreated BiOCl and treated samples prepared at different NH4F amounts added. All the diffraction peaks of the untreated sample are in good agreement with the standard BiOCl (JCPDS no. 06-0249), indicating that the formation of phase-pure BiOCl. At RF = 0, the crystalline phase of the sample does not change, however, the crystallinity becomes stronger than the untreated one. At RF = 0.5, BiOF (no. 73-1595) has formed besides BiOCl, indicating that a part of BiOCl has converted into BiOF. At RF = 1, Bi7F11O5 (no. 50-0003) has formed, besides both BiOF and BiOCl; meanwhile, the diffraction peaks of BiOF become weak. At RF = 2, both Bi7F11O5 and BiOCl were observed, indicating the Bi7F11O5/BiOCl heterojunction is formed. At RF = 4, the phase-pure Bi7F11O5 has formed. Herein, it should be noted that the standard no. 50-0003 only ranges in 10–60°, thus only the weak diffraction peaks in the range of 10–60° are checked. At RF = 6, the sample consists of BiF3 (no. 73-1988) as the main phase, besides a small amount of Bi7F11O5. At RF = 8, the phase-pure BiF3 has formed. It is obvious that BiOCl can be converted into phase-pure Bi7F11O5 and BiF3 in the ethanol/water mixture (VEth/Vw = 2/18). For Bi7F11O5/BiOCl sample (synthesized at RF = 2), we have also investigated the influence of reaction time on the sample (Fig. S2, ESI). By calculating the intensity ratios of the strongest peaks of Bi7F11O5 and BiOCl phases (1–12 h), the reaction time nearly have no obvious influence on the composition in the Bi7F11O5/BiOCl mixture.
image file: c6ra11877a-f1.tif
Fig. 1 X-ray diffraction (XRD) patterns of the untreated BiOCl and treated samples by NH4F at different molar ratios of F to Bi (RF): (a) untreated BiOCl; (b) RF = 0 (BiOCl); (c) RF = 0.5 (BiOF/BiOCl); (d) RF = 1 (BiOF/Bi7F11O5/BiOCl); (e) RF = 2 (Bi7F11O5/BiOCl); (f) RF = 4 (Bi7F11O5); (g) RF = 6 (Bi7F11O5/BiF3); (h) RF = 8 (BiF3). Preparation conditions: VEth/Vw = 18/2 (volumetric ratio of ethanol to water).

Additionally, the EDS of Bi7F11O5/BiOCl (synthesized at RF = 2) is shown in Fig. S3 (ESI), which confirm that Bi, O, Cl and F elements were existed in the sample, moreover, the detail contents of Bi, O, Cl and F elements are listed in Table S1, which demonstrated that the molar ratio of Bi7F11O5 and BiOCl is about 35% in Bi7F11O5/BiOCl sample.

The SEM images of the untreated BiOCl samples are shown in Fig. S4a (ESI). The untreated BiOCl sample is composed of 4–7 μm microspheres assembled by the nanosheets. After hydrothermal treatment at RF = 0, a part of microspheres have broken into nanosheets (Fig. S4b, ESI). After treatment with NH4F, the microspheres have apparently broken into nanosheets and eventually evolved into the other shapes (Fig. S4c–h, ESI). It is obvious that NH4F has a significant impact on the phase composition and morphology of the sample. It is interesting that BiOCl can be converted to the other new materials merely by a simple ion exchange approach.

The surface areas of the samples are shown in Table 1. It is obvious that the sample treated at RF = 0 has a smaller surface area (2.4 m2 g−1) than the untreated BiOCl sample (16.5 m2 g−1). Fig. S4a and b (ESI) shows that the untreated sample is composed of uniform, dense microspheres assembled by numerous nanosheets, leading to the formation of more pores;33 for the treated samples, the microspheres have been broken into the grains, the increased grains may result in the reduced surface area.

Table 1 BET areas of the untreated BiOCl and treated samples by NH4F at different RF at VEth/Vw = 18/2 at 170 °C for 6 ha
Sample BET areab (m2 g−1) Sample BET area (m2 g−1)
a RF: the molar ratios of F to Bi; VEth/Vw: volumetric ratio of ethanol to water.b Calculated by the Brunauer–Emmett–Teller (BET) method.
Untreated BiOCl 16.5 RF = 2 (Bi7F11O5/BiOCl) 5.0
RF = 0 (BiOCl) 2.4 RF = 4 (Bi7F11O5) 3.9
RF = 0.5 (BiOF/BiOCl) 8.1 RF = 6 (Bi7F11O5/BiF3) 5.8
RF = 1 (BiOF/Bi7F11O5/BiOCl) 4.4 RF = 8 (BiF3) 7.2


Typically, the Bi7F11O5/BiOCl sample obtained at RF = 2 is characterized by the high-resolution transmission electron microscopy (HRTEM) and selected-area electron diffraction (SAED) patterns. It is clear that Bi7F11O5 nanosheets distribute on BiOCl nanosheets (Fig. 2a). In Fig. 2b, the clear diffraction spots correspond to the (200) and (110) Bragg reflections of BiOCl, and the diffraction rings correspond to (112) and (004) Bragg reflections of Bi7F11O5, respectively. The SAED patterns reveal the single crystalline nature of BiOCl and the polycrystalline nature of Bi7F11O5. Moreover, Fig. 2c shows the lattice fringe images of the Bi7F11O5/BiOCl sample. The lattice spacing of 0.278 nm is well indexed to (110) crystal plane of BiOCl, and the lattice spacings of 0.230 nm and 0.335 nm match (004) and (112) planes of Bi7F11O5, respectively. The EDS analysis of Bi7F11O5/BiOCl sample is shown in Fig. S3 and Table S1 (ESI), which is also given as follows. The Bi, O, Cl and F are detected, which can also confirm the phase evolution and the formation of Bi7F11O5/BiOCl heterojunction. The results demonstrate that a heterojunction has formed between BiOCl and Bi7F11O5 due to their matched energy bands, which will be further discussed in the latter sections.


image file: c6ra11877a-f2.tif
Fig. 2 High-resolution transmission electron microscopy (HRTEM) images and selected-area electron diffraction (SAED) patterns of Bi7F11O5/BiOCl sample prepared at RF = 2 at VEth/Vw = 18/2: (a) TEM; (b) SAED; (c) lattice fringe images.
3.1.2. Light absorptions and electrical conductivities of the samples. Fig. 3a displays the UV-visible diffuse reflectance spectra (UV-DRS) of the samples. All the samples exhibit the light absorption in UV light ranges. Compared with the untreated BiOCl sample, all the treated samples show the obvious blue-shifts, indicating the band gap increase, which is mainly due to their phase compositions. It can be determined that the absorption edges of BiOCl, Bi7F11O5 and BiF3 are 374, 317 and 300 nm, respectively.
image file: c6ra11877a-f3.tif
Fig. 3 (a) UV-visible diffuse reflectance spectra (UV-DRS) and (b) electrochemical impedance spectra (EIS) of the untreated BiOCl and treated samples by NH4F at different RF: VEth/Vw = 18/2.

Besides, the electrochemical impedance spectroscopy (EIS) spectra of the samples are shown in Fig. 3b. It is well known that the smaller arc radius indicates a higher electrical conductivity that will favor for an efficient electron transfer.34,35 The arc radius of the samples follow the order as follows: RF = 0 < BiOCl < RF = 0.5 < RF = 1 < RF = 2 < RF = 4 < RF = 6 < RF = 8. It is obvious that with the increase of RF, the conductivities of the treated samples decrease, which is relative to the phase composition. But, the treated BiOCl sample at RF = 0 has a higher conductivity than the untreated one, which can be attributed to the improved crystallinity of the former one.

3.1.3. Photocatalytic activities of the samples. Fig. 4 shows the degradation curves and reaction kinetic curves of methyl orange (MO) dye under UV light irradiation (λ ≤ 420 nm). The degradation activities of the samples follow the order as follows: RF = 2 > RF = 1 > RF = 0 > RF = 0.5 > untreated BiOCl > RF = 4 > RF = 6 > RF = 8. It is obvious that the Bi7F11O5/BiOCl sample (RF = 2) shows the highest activity amongst.
image file: c6ra11877a-f4.tif
Fig. 4 Degradation curves (a) and reaction kinetic curves (b) of MO over the untreated BiOCl and treated samples by NH4F at different RF under ultraviolet light irradiation (λ ≤ 420 nm): VEth/Vw = 18/2.

First, after irradiation 30 min, 80% of MO has been degraded by Bi7F11O5/BiOCl (RF = 2), but only 17% and 21% of MO are degraded by Bi7F11O5 (RF = 4) and untreated BiOCl, respectively. The degradation rate constants (ka) are given in Table 2. The degradation rate constant (ka = 0.0642 min−1) of Bi7F11O5/BiOCl sample is 6.42 and 3.28 times higher than those of Bi7F11O5 (ka = 0.0100 min−1) and the untreated BiOCl (ka = 0.0196 min−1), respectively. From Table 1 and Fig. 3, the untreated BiOCl has a higher BET area (16.5 m2 g−1), a stronger light absorption and a higher conductivity than Bi7F11O5/BiOCl sample, but its activity is low. This suggests that the BET area, light absorbance and conductivity may not be the main affecting factors for the photocatalytic activity. It is well known that photocatalytic activity is strongly dependent on the separation and transfer of photogenerated charges.36,37 We could assume that the high activity of Bi7F11O5/BiOCl may be relative to the formation of a heterojunction between BiOCl and Bi7F11O5. The conduct band (CB) and valence band (VB) are 0.49 and 3.81 eV for BiOCl, whereas 1.12 and 5.1 eV for Bi7F11O5, respectively. Their energy bands are well matched, the efficient Bi7F11O5/BiOCl heterojunctions can form, at which an interfacial electrical field can form. Under UV light irradiation, the photogenerated electrons at the CB of BiOCl would transfer to the CB of Bi7F11O5; meanwhile the generated holes at the VB of Bi7F11O5 would migrate to the VB of BiOCl. Driven by the interfacial electrical field, the photogenerated electron holes can be effectively separated.

Table 2 The apparent rate constants (ka) over the untreated BiOCl and treated samples by NH4F at different RF for the degradation of MOa
Sample ka (min−1) Sample ka (min−1)
a VEth/Vw = 18/2; fluorine source: NH4Cl; 170 °C/6 h.
Untreated BiOCl 0.0196 RF = 2 0.0642
RF = 0 0.0415 RF = 4 0.0100
RF = 0.5 0.0234 RF = 6 0.0079
RF = 1 0.0532 RF = 8 0.0072


Second, for the untreated and treated BiOCl (RF = 0) samples, the degradation rate (ka = 0.0415 min−1) of the treated BiOCl sample is obviously higher than that (ka = 0.0196 min−1) of the former sample, although the BET area (2.4 m2 g−1) of the latter sample is greatly smaller than that (16.5 m2 g−1) of the former sample (Table 1). This suggests the improved degradation efficiency cannot be ascribed to the BET area. Since the BiOCl sample treated at RF = 0 has a higher conductivity (Fig. 3b) and a higher crystallinity (Fig. 1) than the untreated BiOCl, so the high degradation efficiency of the former sample may be attributed to its high charge transportation rate and high crystallinity.

Third, the degradation efficiencies of the phase-pure BiF3, Bi7F11O5 and untreated BiOCl samples follow the order as follows: BiF3 < Bi7F11O5 < untreated BiOCl; but their BET areas follow the order as follows: Bi7F11O5 (3.9 m2 g−1) < BiF3 (7.2 m2 g−1) < the untreated BiOCl (16.5 m2 g−1). Therefore, the BET area may not be the main affecting factor for the photocatalytic activity. Compared with BiOCl, the blue shifts of Bi7F11O5 and BiF3 at absorption thresholds are obvious (Fig. 3a). Thus, their different photocatalytic activities can be attributed to their different light absorptions.

In addition, the degradation curve of MO over P25 also have been tested shown in Fig. S5a (ESI). The results show that about 90% of MO can be degraded by P25 after irradiated for 25 min, which was little higher than Bi7F11O5/BiOCl sample. What is more, in order to avoiding dye photosensitization, we have also tested the photodegradation activity of 2-nitrophenol by the Bi7F11O5/BiOCl sample. Fig. S5b (ESI) shows the UV-vis absorption spectra of 2-nitrophenol (200 mL, 10 mg L−1) over Bi7F11O5/BiOCl sample at different reaction time under UV light irradiation (λ ≤ 420 nm). It is clear that the absorbance (at about 277 nm) are decreased gradually with the increasing of irradiation time. After been irradiated for 40 min, nearly 90% of 2-nitrophenol can be degraded by Bi7F11O5/BiOCl sample.

3.2. Effect of fluorine source on the samples

To investigate the effect of fluorine source on the sample, the samples are prepared by using different fluorine sources (NH4F, NaF and HF) at RF = 2 and VEth/Vw = 18/2, while keeping the other condition same. Fig. 5a shows the XRD patterns of the samples. With NH4F, the sample is mainly composed of Bi7F11O5 and a small amount of BiOCl; with NaF, nevertheless, the crystal phase has not changed (i.e. BiOCl); with HF, the sample mainly consists of BiOCl and a small amount of Bi7F11O5. Further, Fig. S6 (ESI) shows the SEM images of the samples synthesized with different fluorine sources. When using NH4F, the formed sample consists of nanosheets (Fig. S6a, ESI). When NaF is added, however, the sample remains the original microspheres (Fig. S6b, ESI). While HF is used, the sample is composed of loose microspheres and a few of nanosheets (Fig. S6c, ESI). It is obvious that the fluorine source has a significant impact on the phase composition and morphology of the sample. The alkaline condition (NaF) has little impact on the sample, while the neutral (NH4F) and acid (HF) conditions have a significantly influences on the conversion process. As shown in Table 3, the BET areas of the samples with different fluorine sources follow the order as follows: NaF (14.8 m2 g−1) > HF (6.8 m2 g−1) > NH4F (5.0 m2 g−1). Compared with the untreated BiOCl sample (16.5 m2 g−1), the BET area of the sample treated by NaF almost does not change, which may be attributed to the almost unchanged microspheres morphology. However, the BET areas of the samples treated by NH4F and HF decrease obviously, which can be mainly ascribed to the morphology change from porous microspheres to nanosheets.
image file: c6ra11877a-f5.tif
Fig. 5 XRD patterns (a) and UV-DRS (b) of the samples treated at VEth/Vw = 18/2 using different fluorine sources (RF = 2): NH4F: Bi7F11O5/BiOCl; HF: Bi7F11O5/BiOCl; NaF: BiOCl.
Table 3 BET areas and the apparent rate constants (ka) for the degradation of MO of the treated samples with different fluorine sourcesa
Fluorine source NH4F NaF HF
a RF = 2; VEth/Vw = 18/2; 170 °C/6 h.
Crystal phase Bi7F11O5/BiOCl BiOCl Bi7F11O5/BiOCl
BET area (m2 g−1) 5.0 14.8 6.8
ka (min−1) 0.0642 0.0264 0.0672


Fig. 5b displays UV-DRS of the samples treated with different fluorine sources. The three samples exhibit almost the same absorption. Compared with the NaF-treated sample (BiOCl), the samples with NH4F and HF show a slight blue shift, which is attributed to the formation of Bi7F11O5. In the sample with HF, a small amount of Bi7F11O5 is contained; but a large amount of Bi7F11O5 is contained in the sample with NH4F. As determined in Section 3.1.2, Bi7F11O5 has a wider band gap than BiOCl.

Fig. 6 shows the degradation curves and reaction kinetic curves of the samples, and the calculated rate constants (ka) are summarized in Table 3. Their degradation rates follow the order as follows: HF (ka = 0.0672 min−1) ≈ NH4F (ka = 0.0642 min−1) > NaF (ka = 0.0264 min−1). The photocatalytic activities of the samples treated with NH4F (Bi7F11O5/BiOCl) and HF (Bi7F11O5/BiOCl) are higher than that of NaF (BiOCl), although the sample treated by NaF has the larger BET area (Table 3). Herein, we hold that the improved activities of the former samples are mainly due to the formation of heterojunction between BiOCl and Bi7F11O5, which has efficiently improve the charge separation.


image file: c6ra11877a-f6.tif
Fig. 6 Degradation curves (a) and reaction kinetic curves (b) of the samples at VEth/Vw = 18/2 with different fluorine sources (RF = 2) under ultraviolet light irradiation (λ ≤ 420 nm).

3.3. Effect of reaction medium composition on the samples

In order to investigate the effect of reaction medium composition, ethanol/water volumetric ratio (VEth/Vw) is varied from 0/20 to 2/18, 10/10, 18/2 and 20/0 while keeping the other preparations parameters constant (NH4F, RF = 2). The XRD patterns of the samples are shown in Fig. 7a. At VEth/Vw = 0/20, 2/18 and 10/10, all products are phase-pure BiOF. However, Bi7F11O5/BiOCl heterojunctions have formed at VEth/Vw = 18/2, and BiOCl has converted into BiF3/BiOCl at VEth/Vw = 20/0. In brief, at lower VEth/Vw ratios, BiOCl can completely convert to BiOF; on the contrary, Bi7F11O5/BiOCl and BiF3/BiOCl can form at higher VEth/Vw ratios (18/2 and 20/0). Fig. S7 (ESI) shows their SEM images. At VEth/Vw = 0/20, 2/18 and 10/10, the part of microspheres have broken and the microspheres become loose (Fig. S7a–c, ESI). At VEth/Vw = 18/2, all the microspheres have broken into nanosheets (Fig. S7d, ESI). At VEth/Vw = 20/0, however, some of irregular particles form, besides microspheres (Fig. S7e, ESI). It is clearly observed that the VEth/Vw ratio has an important influence on the phase composition and morphology of the sample.
image file: c6ra11877a-f7.tif
Fig. 7 XRD patterns (a) and UV-DRS (b) of the samples prepared with NH4F (RF = 2) at different VEth/Vw ratios: BiOF (0/20, 2/18, 10/10); Bi7F11O5/BiOCl (18/2); BiF3/BiOCl (20/0).

The BET areas of the samples are shown in Table 4. At VEth/Vw = 0/20–10/10, the BET areas of the phase-pure BiOF samples decrease gradually, which may be attributed to the broken microspheres that reduces the porosity. The Bi7F11O5/BiOCl sample obtained at VEth/Vw = 18[thin space (1/6-em)]:[thin space (1/6-em)]2 has a smaller BET area (5.0 m2 g−1) than that (8.2 m2 g−1) of the BiF3/BiOCl sample at VEth/Vw = 20/0, which is due to their different morphology. According to the SEM results, the former sample is composed of nanosheets resultant from completely broken porous microspheres, but the latter sample consists of the porous microsphere and a small number of irregular particles (Fig. S7e, ESI). Fig. 7b shows the UV-DRS of the samples. With increasing VEth/Vw ratio from 0/20 to 2/18 and 10/10, the light absorption intensities of the BiOF samples decrease gradually, which can be attributed to their decreased BET areas (Table 4). Compared with BiOF, however, the Bi7F11O5/BiOCl (VEth/Vw = 18/2) and BiF3/BiOCl (VEth/Vw = 20/0) samples show a slight red shift at absorption edges, which can be attributed to the different phase compositions from BiOF. As shown in Table 5, the band gaps are 3.32, 3.98, 4.01 and 4.15 eV for BiOCl, Bi7F11O5, BiOF and BiF3, respectively.

Table 4 BET areas and the apparent rate constants (ka) for the degradation of MO of the treated samples at different VEth/Vw ratiosa
VEth/Vw 0/20 2/18 10/10 18/2 20/0
a RF = 2; fluorine source: NH4F; 170 °C/6 h.
Crystal phase BiOF BiOF BiOF Bi7F11O5/BiOCl BiF3/BiOCl
BET area (m2 g−1) 4.0 3.8 2.2 5.0 8.2
ka (min−1) 0.0075 0.0064 0.0066 0.0642 0.0789


Table 5 The band gap (Eg), VBM (EVB) and CBM (ECB) of the samplesa
Chemicals χ (eV) Eg (eV) EVB (eV) ECB (eV)
a χ, the geometric mean of Mulliken's electronegativities; VBM, valence band maximum; CBM, conduction band minimum.
BiOCl 6.65 3.32 3.81 0.49
BiOF 7.17 4.01 4.68 0.67
Bi7F11O5 7.61 3.98 5.1 1.12
BiF3 8.53 4.15 6.11 1.96


Fig. 8 shows their degradation curves and reaction kinetic curves for degradation of MO under UV irradiation (λ ≤ 420 nm), and the calculated rate constants (ka) are given in Table 4. Their degradation rates follow the order as follows: VEth/Vw = 20/0 (BiF3/BiOCl, ka = 0.0789 min−1) > VEth/Vw = 18/2 (Bi7F11O5/BiOCl, ka = 0.0642 min−1) ≫ VEth/Vw = 0/20 (BiOF, ka = 0.0075 min−1) > VEth/Vw = 10/10 (BiOF, ka = 0.0066 min−1) ≅ VEth/Vw = 2/18 (BiOF, ka = 0.0064 min−1). Among them, the BiF3/BiOCl sample prepared at VEth/Vw = 20/0 shows the highest photocatalytic activity. Fig. 7b and Table 4 demonstrate that the orders of light absorption and BET area are consistent with the photocatalytic activity order. Moreover, the BiF3/BiOCl and Bi7F11O5/BiOCl heterojunctions also contribute to their obviously improved activities, compared with BiOF.


image file: c6ra11877a-f8.tif
Fig. 8 Degradation curves (a) and reaction kinetic curves (b) of MO over the samples prepared with NH4F (RF = 2) at different VEth/Vw ratios under ultraviolet light irradiation (λ ≤ 420 nm).

3.4. Theory calculations

Further, we have mainly compared UV-DRS spectra of the phase-pure BiOCl, BiOF, Bi7F11O5 and BiF3 samples (Fig. 9). For BiOCl, BiOF, Bi7F11O5 and BiF3, with the gradual increase of fluorine atoms, the samples show an obvious blue-shift in the absorption thresholds (Fig. 9a). The band gap energy (Eg) of a semiconductor can be calculated by eqn (2) as follows.38
 
αhv = A(hvEg)n/2 (2)
where α, h, v, Eg and A represent absorption coefficient, Plank constant, light frequency, band gap energy and a constant, respectively. Among them, n is determined by the optical transition type of a semiconductor (n = 1 for a direct optical transition and n = 4 for an indirect optical transition).39 The values of n are 4, 1, 1 and 1 for BiOCl, BiOF, Bi7F11O5 and BiF3, respectively.19,20,40 Fig. 9b further shows their Tauc plots. The bandgap energies are determined to be 3.32, 4.01, 3.98 and 4.15 eV for BiOCl, BiOF, Bi7F11O5 and BiF3, respectively (Fig. 9b, Table 5).

image file: c6ra11877a-f9.tif
Fig. 9 UV-DRS spectra (a) Tauc plots, (b) energy bands and (c) energy level diagram of the samples.

Further, the conduct band and valence band positions of a semiconductor can be calculated by the following empirical formula:41

 
EVB = χEe + 0.5Eg (3)
 
ECB = EVBEg (4)
where χ is the absolute electronegativity of a semiconductor, which is defined as the geometric mean of the absolute electronegativity of constituent atoms; Ee is the energy of free electrons on the hydrogen scale (ca. 4.5 eV); ECB and EVB are the conduct and valence band edge potentials, and Eg is band gap. Herein, the calculated ECB and EVB values of the samples are summarized in Table 5. It is obvious that with increase of F, their CB and VB down shift and their band gap become wider (Fig. 9c).

In our work, the energy band and electronic structures of BiOCl, BiOF and BiF3 are calculated by Density Functional Theory (DFT) with the generalized-gradient approximation (GGA) and the exchange–correlation functional of Perdew–Burke–Ernzerhof of (PBE) by using the Materials Studios software. The adopted CASTEP energy cutoff is 340 eV (Fig. 10). The calculated band gaps of BiOCl, BiOF and BiF3 are 2.61, 3.34 and 3.70 eV, respectively. The band gap of Bi7F11O5 has been calculated to be 3.81 eV by Hu et al.20 It can be observed that the calculated results by DFT are slightly smaller than the experimental results, which is a common feature of DFT calculations.42 The DFT calculations also demonstrate that with the increase of fluorine atoms, the band gaps of the BiOCl, BiOF, Bi7F11O5 and BiF3 samples increases. From Fig. 10d–f, we can see that the VB top of BiOCl is mainly composed of the O 2p and Cl 3p orbits, and its CB bottom is mainly composed of Bi 6p and a part of O 2p orbits (Fig. 10d). The VB top of BiOF is found to be mainly composed of both O 2p and F 2p orbits, while the CB bottom is dominantly mainly composed of Bi 6p and a little of O 2p orbits (Fig. 10e). The results show that O 2p orbit contributes to both CB and VB of both BiOCl and BiOF. Hu et al. have reported that F 2p and O 2p orbits mainly contribute to the VB of Bi7F11O5, and Bi 6p orbit mainly contributes to the CB.20 Meanwhile, the VB top of BiF3 is mainly attributed by the F 2p orbital, and the CB bottom is mainly attributed by the Bi 6p, as well as a little of F 2p orbits (Fig. 10f). Comparing BiF3 with BiOF, it is interesting that F 2p orbit contributes to both CB and VB of BiF3, which is different from BiOF. This further confirms that with the increase of fluorine, the band gaps of BiOCl, BiOF, Bi7F11O5 and BiF3 increase, similar to the experimental result above.


image file: c6ra11877a-f10.tif
Fig. 10 Energy band structures (a–c) and total and a part of densities of states (TDOS and PDOS) (d–f) of the samples: (a and d) BiOCl; (b and e) BiOF and (c and f) BiF3.

3.5. Conversion mechanism

It is well known that BiOCl, BiOF have tetragonal structures and Bi7F11O5, BiF3 have monoclinic, cubic structures; respectively. Due to the layer structure of BiOCl, Cl, instead of O, is easily substituted by F. Thus, BiOCl can be converted to BiOF, Bi7F11O5 and BiF3. In the presence of NH4F at VEth/Vw = 18/2, BiOCl can be facilely converted to BiOF, Bi7F11O5 and BiF3. At VEth/Vw = 18/2, however, we cannot achieve the complete conversion from BiOCl to phase-pure BiOF while only changing the amount of NH4F added. Nevertheless, BiOCl can completely convert to phase-pure BiOF at low VEth/Vw ratios (0/20, 2/18, 10/10) at RF = 2. According to the experiment results, the schematic conversion process is proposed and shown in Fig. 11, and two main conclusions are made as follows.
image file: c6ra11877a-f11.tif
Fig. 11 Schematic conversion process.

First, generally, a higher dielectric constant results in a higher solubility of metal oxide,43 which favors for the nucleation and growth processes. With increasing VEth/Vw ratio, the dielectric constant of reaction medium decreases (εEth = 24.5, εw = 78.5). Thus we assume that the medium composition may mainly affect the crystal nucleation and growth in crystal conversion.

Second, NH4+ as a Lewis acid can destroyed the [Bi2O2] layers obviously. At RF = 4 (NH4F), the [Bi2O2] slabs may have been destroyed partially, where the partial O of [Bi2O2] slabs will be replaced by F. Thus, tetragonal BiOCl has completely converted to monoclinic Bi7F11O5. To further increase the amount of NH4F to RF = 8, the original [Bi2O2] slabs will be destroyed completely, where the O atoms of [Bi2O2] slabs may be eventually replaced by F. As a result, tetragonal BiOCl has completely converted into cubic BiF3. When HF is used as the fluorine source, on the other hand, BiOCl convert to hybrid phase Bi7F11O5/BiOCl, similar to the result by using NH4F as the fluorine source. However, BiOCl cannot be changed while using NaF as the fluorine source, suggesting that the alkaline condition has no effect on [Bi2O2] slabs of BiOCl. Herein, we could assume that the crystal conversion is mainly controlled through altering the [Bi2O2] slabs. However, the real conversion mechanism needs further studying.

4. Conclusions

A facile conversion from BiOCl nanosheets to BiOF, Bi7F11O5 and BiF3 can be achieved through a simple ion exchange method. The added amounts of NH4F, fluorine source and medium have significant influences on the composition and morphology of the samples. The photocatalytic activity of Bi7F11O5/BiOCl heterojunction (RF = 2) is 3.28 times higher than that of BiOCl for the degradation of methyl orange (MO). The as-formed Bi7F11O5/BiOCl heterojunction benefits the separation and transfer of charges, leading to an improved photocatalytic activity. This work provides an initiative post-synthesis strategy to prepare the other photocatalysts.

Acknowledgements

This work is financially supported by National Science Foundation of China (21377060), Scientific Research Foundation for the Returned Overseas Chinese Scholars of State Education Ministry (20121707), Six Talent Climax Foundation of Jiangsu (20100292), the Key Project of Environmental Protection Program of Jiangsu (2013005), “333” Outstanding Youth Scientist Foundation of Jiangsu (20112015), the Project Funded by the Science and Technology Infrastructure Program of Jiangsu (BM201380277), and A Project Funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD) sponsored by SRF for ROCS, SEM (2013S002).

References

  1. K. L. Zhang, C. M. Liu, F. Q. Huang, C. Zheng and W. D. Wang, Appl. Catal., B, 2006, 68, 125–129 CrossRef CAS.
  2. H. Gnayem and Y. Sasson, ACS Catal., 2013, 3, 186–191 CrossRef CAS.
  3. J. Cao, B. Y. Xu, H. L. Lin, B. D. Luo and S. F. Chen, Catal. Commun., 2012, 26, 204–208 CrossRef CAS.
  4. M. L. Guan, C. Xiao, J. Zhang, S. J. Fan, R. An, Q. M. Cheng, J. F. Xie, M. Zhou, B. J. Ye and Y. Xie, J. Am. Chem. Soc., 2013, 135, 10411–10417 CrossRef CAS PubMed.
  5. K. Zhao, L. Z. Zhang, J. J. Wang, Q. X. Li, W. W. He and J. J. Yin, J. Am. Chem. Soc., 2013, 135, 15750–15753 CrossRef CAS PubMed.
  6. C. J. Huang, J. L. Hu, S. Cong, Z. G. Zhao and X. Q. Qiu, Appl. Catal., B, 2015, 174, 105–112 CrossRef.
  7. C. L. Yu, F. F. Cao, G. Li, R. F. Wei, J. C. Yu, R. C. Jin, Q. Z. Fan and C. Y. Wang, Sep. Purif. Technol., 2013, 120, 110–122 CrossRef CAS.
  8. H. F. Cheng, B. B. Huang, X. Y. Qin, X. Y. Zhang and Y. Dai, Chem. Commun., 2012, 48, 97–99 RSC.
  9. J. Xu, F. Teng, Y. X. Zhao, Y. D. Kan, L. M. Yang, Y. Yang, W. Q. Yao and Y. F. Zhu, Dalton Trans., 2016, 45, 6866–6877 RSC.
  10. H. Tian, F. Teng, J. Xu, S. Q. Lou, N. Li, Y. X. Zhao and M. D. Chen, Sci. Rep., 2015, 5, 7770–7778 CrossRef CAS PubMed.
  11. J. Xu, F. Teng, W. Q. Yao and Y. F. Zhu, RSC Adv., 2016, 6, 23537–23549 RSC.
  12. S. R. Wang, F. Teng and Y. X. Zhao, RSC Adv., 2015, 5, 76588–76598 RSC.
  13. N. Li, X. Hua, K. Wang, Y. J. Jin, J. Xu, M. D. Chen and F. Teng, J. Mol. Catal. A: Chem., 2014, 395, 428–433 CrossRef CAS.
  14. Q. Y. Zhang, H. Tian, N. Li, M. D. Chen and F. Teng, CrystEngComm, 2014, 16, 8334–8339 RSC.
  15. P. Sun, Y. J. Jin, Y. X. Zhao, J. Xu, M. D. Chen, W. Q. Yao, Y. F. Zhu and F. Teng, Nano, 2014, 9, 1450094–1450104 CrossRef.
  16. H. B. Kang, S. T. Myung, K. Amine, S. M. Lee and Y. K. Sun, J. Power Sources, 2010, 195, 2023–2028 CrossRef CAS.
  17. W. Y. Su, J. Wang, Y. X. Huang, W. J. Wang, L. Wu, X. X. Wang and P. Liu, Scr. Mater., 2010, 62, 345–348 CrossRef CAS.
  18. S. J. Zou, F. Teng, C. Chang, Z. L. Liu and S. R. Wang, RSC Adv., 2015, 5, 88936–88942 RSC.
  19. C. K. Feng, F. Teng, Z. L. Liu, C. Chang, Y. X. Zhao, S. R. Wang, M. D. Chen, W. Q. Yao and Y. F. Zhu, J. Mol. Catal. A: Chem., 2015, 401, 35–40 CrossRef CAS.
  20. W. Hu, P. Shan, T. Q. Sun, H. D. Liu, J. M. Zhang, X. W. Liu, Y. F. Kong and J. J. Xu, J. Alloys Compd., 2016, 658, 788–794 CrossRef CAS.
  21. Q. C. Liu, D. K. Ma, Y. Y. Hu, Y. W. Zeng and S. M. Huang, ACS Appl. Mater. Interfaces, 2013, 5, 11927–11934 CAS.
  22. D. K. Ma, M. L. Guan, S. S. Liu, Y. Q. Zhang, C. W. Zhang, Y. X. He and S. M. Huang, Dalton Trans., 2012, 41, 5581–5586 RSC.
  23. P. Cai, S. M. Zhou, D. K. Ma, S. N. Liu, W. Chen and S. M. Huang, Nano-Micro Lett., 2015, 7, 183–193 CrossRef.
  24. J. C. Yu, J. G. Yu, W. K. Ho, Z. T. Jiang and L. Z. Zhang, Chem. Mater., 2002, 14, 3808–3816 CrossRef CAS.
  25. S. W. Liu, K. Yin, W. S. Ren, B. Cheng and J. G. Yu, J. Mater. Chem., 2012, 22, 17759–17767 RSC.
  26. H. M. Wu, J. Z. Ma, Y. B. Li, C. B. Zhang and H. He, Appl. Catal., B, 2014, 152, 82–87 CrossRef.
  27. S. H. Chen, S. X. Sun, H. G. Sun, W. L. Fan, X. Zhao and X. Sun, J. Phys. Chem. C, 2010, 114, 7680–7688 CAS.
  28. S. W. Liu, J. G. Yu, B. Cheng and M. Jaroniec, Adv. Colloid Interface Sci., 2012, 173, 35–53 CrossRef CAS PubMed.
  29. J. G. Yu, W. G. Wang, B. Cheng and B. L. Su, J. Phys. Chem. C, 2009, 113, 6743–6750 CAS.
  30. H. Y. Jiang, J. J. Liu, K. Cheng, W. B. Sun and J. Lin, J. Phys. Chem. C, 2013, 117, 20029–20036 CAS.
  31. J. S. Wang, S. Yin, Q. W. Zhang, F. Saito and T. Sato, J. Mater. Sci., 2004, 39, 715–717 CrossRef CAS.
  32. Y. F. Liu, Y. H. Lv, Y. Y. Zhu, D. Liu, R. L. Zong and Y. F. Zhu, Appl. Catal., B, 2014, 147, 851–857 CrossRef CAS.
  33. N. Li, X. Hua, K. Wang, Y. J. Jin, J. J. Xu, M. D. Chen and F. Teng, Dalton Trans., 2014, 43, 13742–13750 RSC.
  34. Y. Y. Zhu, Q. Ling, Y. F. Liu, H. Wang and Y. F. Zhu, Phys. Chem. Chem. Phys., 2015, 17, 933–940 RSC.
  35. Y. D. Hou, A. B. Laursen, J. S. Zhang, G. G. Zhang, Y. S. Zhu, X. C. Wang, S. Dahi and I. Chorkendorff, Angew. Chem., Int. Ed., 2013, 52, 3621–3625 CrossRef CAS PubMed.
  36. C. S. Pan, J. Xu, Y. J. Wang, D. Li and Y. F. Zhu, Adv. Funct. Mater., 2012, 22, 1518–1524 CrossRef CAS.
  37. M. Guerrero, A. Altube, E. Garcia-Lecina, E. Rossinyol, M. D. Baro, E. Pellicer and J. Sort, ACS Appl. Mater. Interfaces, 2014, 6, 13994–14000 CAS.
  38. J. Zeng, H. Wang, Y. C. Zhang, M. K. Zhu and H. Yan, J. Phys. Chem. C, 2007, 111, 11879–11887 CAS.
  39. S. M. Sun, W. Z. Wang, L. Zhang, L. Zhou, W. Z. Yin and M. Shang, Environ. Sci. Technol., 2009, 43, 2005–2010 CrossRef CAS PubMed.
  40. Y. Z. Zhao and W. W. Dai, Inorg. Chem., 2014, 53, 13001–13011 CrossRef PubMed.
  41. N. Zhang, S. Q. Liu, X. Z. Fu and Y. J. Xu, J. Phys. Chem. C, 2011, 115, 9136–9145 CAS.
  42. A. Kudo, H. Kato and S. Nakagawa, J. Phys. Chem. B, 2000, 104, 571–575 CrossRef CAS.
  43. N. Garti and H. Zour, J. Cryst. Growth, 1997, 172, 486–498 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: This material, SEM images of the samples with NH4F at different RF and fluorine sources and VEth/Vw. See DOI: 10.1039/c6ra11877a

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