Next Article in Journal
Synthesis of Sulfonic Acid-Functionalized Zirconium Poly(Styrene-Phenylvinyl-Phosphonate)-Phosphate for Heterogeneous Epoxidation of Soybean Oil
Previous Article in Journal
Immobilized rGO/TiO2 Photocatalyst for Decontamination of Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Bi3O5I2 Hollow Microsphere and Its Enhanced Photocatalytic Activity

1
Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, China
2
University of the Chinese Academy of Sciences, Beijing 100049, China
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(9), 709; https://doi.org/10.3390/catal9090709
Submission received: 29 July 2019 / Revised: 19 August 2019 / Accepted: 22 August 2019 / Published: 24 August 2019

Abstract

:
A new type of I-deficient bismuth oxyiodide Bi3O5I2 with a hollow morphology was prepared by the solvothermal process. The structure, composition, morphology, optical property and photoelectric property of the as prepared photocatalyst were investigated through some characterization methods. Those characterization results showed that Bi3O5I2 displayed a larger specific surface area, promising band structure and lower recombination of photoinduced carriers than pure BiOI. Bi3O5I2 had a higher photocatalytic activity than BiOI on the decomposition of methyl orange (MO) under simulated solar light irradiation. The superoxide (·O2) and hole (h+) were the dominating active species during the degradation of MO. Its stability and reusability performance showed its great promising application in the degradation of organic pollutant.

Graphical Abstract

1. Introduction

With increasing concerns on environmentally friendly processes, semiconductor photocatalysts have been given widely attention because of their promising applications. Among the multitudinous semiconductor photocatalysts, bismuth oxyiodide (BiOI) with unique layered crystal configuration [1,2,3] and excellent visible light absorption [4,5], is a promising photocatalytic material in the field of energy and environment [6,7,8,9,10]. Nevertheless, the photocatalytic activity of bismuth oxyiodide is far from potential industrial applications because of their relatively high recombination of photogenerated carriers and insufficient reduction and oxidation ability [11,12]. For the sake of increasing the photocatalytic performance of BiOI, measures should be taken to decrease the recombination of photogenerated carriers and improving its reduction and oxidation ability. So far, many measures have been taken to enhance the photocatalytic performance of BiOI, such as regulating morphology [13,14,15,16], exposing special crystal face [17,18,19], doping [20,21,22,23], compositing with other semiconductor [24,25,26,27] and synthesizing other bismuth oxyiodides such as Bi4O5I2 [12,28], Bi5O7I [29,30,31] and Bi7O9I3 [32,33].
Aside from BiOI, I-deficient bismuth oxyiodides have the positive oxidation and reduction ability [11,12], leading to its high photocatalytic activity. For example, Zhang et al. discovered that Bi4O5I2 and Bi5O7I displayed enhanced photocatalytic activity for degradation of tetracycline hydrochloride, bisphenol A and Rhodamine B [11]. Ye et al. synthesized Bi4O5I2 photocatalyst with excellent activities for photocatalytic H2 evolution [12]. Generally, I-deficient bismuth oxyiodides were prepared by four methods: (1) Hydrothermal progress using the pH modifier [34], (2) in situ calcination progress of BiOI [11], (3) hydrolytic process of the bismuth molecular precursor [12], (4) chemical etching method for bismuth oxides [35]. However, these methods were time consuming or introduce potential environmental and safety problems. Therefore, a simple method of synthesizing the I-deficient bismuth oxyiodide is needed.
It is generally accepted that the morphology of the catalysts has the close relation with the catalyst performance [36,37]. Photocatalysts with a hollow structure display excellent photocatalytic performance on account of their high specific surface area. In addition, both their high energy conversion efficiencies and their strong light-harvesting capacities contribute to their excellent photocatalytic performance [15,38,39]. For example, Li et al. found that BiOI with hollow structures presented a higher photocatalytic performance than those BiOI with low-dimensional [15]. Until now, the I-deficient bismuth oxyiodide with different morphologies are synthesized by several researchers, including sheet-like morphology [40], rod-like morphology [29] and flower-like morphology [33,41]. However, there are no reports about synthesizing the I-deficient bismuth oxyiodide with hollow structure. Therefore, a hollow structure of the I-deficient bismuth oxyiodide should be synthesized.
In this research, a new type of I-deficient bismuth oxyiodide with a hollow structure was prepared by the hydrothermal progress for the first time. Their structures as well as its photocatalytic ability for the demineralization of methyl orange (MO) were discussed under systematical characterization.

2. Results

2.1. Molecular Formula Characterization and Phase Transformation

The energy dispersive spectrometry (EDS) apparatus was employed to investigate the chemical composition on the surface of the as prepared I-deficient bismuth oxyiodide. As shown in Figure 1, the I-deficient bismuth oxyiodide contained Bi, O and I elements. Meanwhile, the associated EDS elemental maps confirmed that these elements were distributed homogeneously. As shown in Table S1, the average Bi:O:I ratio in the surface of the I-deficient bismuth oxyiodide was 3:4.7:2, in conformity with the molecular formula Bi3O5I2 (3:5:2). This result indicated that its chemical composition possibly was Bi3O5I2. Meanwhile, the average Bi:O:I ratio of the pure BiOI are 1:1.1:1 (Table S2), which is close to the ideal ratio for BiOI (1:1:1). This indicated that the result of the EDS was convincing.
The inductively coupled plasma (ICP) analysis showed that the percentage of the Bi element in the bulk bismuth oxyiodide was 64.95%, in conformity with the molecular formula Bi3O5I2 (65%). Consequently, based on the analysis of ICP and EDS, the molecular formula of the as prepared I-deficient BiOI can possibly be determined as Bi3O5I2. Therefore, the I-deficient bismuth oxyiodide was labelled as Bi3O5I2 in the following.
A thermogravimetric (TG) measurement was carried out to evaluate the phase transformation of Bi3O5I2 and BiOI and the result was shown in Figure 2. In Figure 2, the TG curve of the Bi3O5I2 sample was composed by three mass loss steps. The first mass loss step (250–360 °C) was attributed to the transformation from Bi3O5I2 to Bi4O5I2, whereas the second mass loss step (400–520 °C) was ascribed to the transformation from Bi4O5I2 to Bi5O7I, the third step (580–760 °C) was ascribed to the transformation from Bi5O7I to Bi2O3, which can be illustrated as Equation (1). The weight remaining of each phase transformation step in Equation (1) was calculated to be 91.3%, 80.2% and 73.8%, respectively, which is very similar to the experimental value (91.7%, 80.1% and 73.2%). This result shows that the molecular formula Bi3O5I2 is correct. According to the results of EDX, ICP and TG, the as prepared I-deficient bismuth oxyiodide can be determined as Bi3O5I2. The phase transformation of BiOI
Bi 3 O 5 I 2 91.3 % 3 4 B i 4 O 5 I 2 80.2 % 3 5 B i 5 O 7 I 73.8 % 3 2 B i 2 O 3

2.2. Structure Investigation and Morphology Observation

The crystal information of Bi3O5I2 and BiOI were characterized by the X-ray diffraction (XRD), which was shown in Figure 3. The main diffraction peaks of the as prepared Bi3O5I2 (Figure 3a) could be clearly seen at 2θ of 10.1, 20.5°, 29.2°, 31.7°, 45.7°,52.6°, 55.1° and 64.3°, respectively. According to the JCPDS file 73–2062, the diffraction peaks at 9.7°, 29.3°, 31.9°, 45.4°, 51.5° and 55.4°are corresponding to the (001), (012), (110), (020), (114) and (122) of BiOI, respectively, which is different with the main diffraction peaks of the as prepared Bi3O5I2.
The typical scanning electron microscope (SEM) and transmission electron microscopy (TEM) were shown in Figure 4. It was clearly seen that Bi3O5I2 had a hollow microsphere morphology with an average diameter of 1–3 μm and the entire sphere like structures were composed of numerous nanoplates (Figure 4a,b). In Figure 4c, there were clearly a contrast between the dark boundary and the relatively bright center, which further confirmed their hollow nature [42]. In Figure 4d, clear lattice fringes showed that the lattice spacing were 0.312 nm and 0.339 nm, which might be the (102) and (101) plane of Bi3O5I2.
The surface chemical compositions of the Bi3O5I2 and the pure BiOI, as well as the chemical states of them, were investigated by the X-ray photoelectron spectroscopy (XPS), which was shown in Figure 5. In Figure 5a, it could be seen that the surfaces of Bi3O5I2 and BiOI consisted of Bi, O, I and C elements (C may come from the reference sample), indicating a high purity of Bi3O5I2 and BiOI. In Figure 5b, the binding energy values of the Bi 4f spectrum of the BiOI was about 158.6 eV and 163.9 eV, which correspond to Bi 4f7/2 and Bi 4f5/2. It indicated the presence of Bi3+ in the sample [43]. As for the Bi3O5I2, the peak of Bi 4f7/2 (or 4f5/2) peak could be decomposed into two bimodal peaks of 158.5 eV and 159.2 eV (or 163.8 eV and 164.5 eV), which could be ascribed to Bi3+ and Bi5+, respectively [44]. The O 1s spectrum of BiOI (presented in Figure 5c) had two peaks. One was at 530.8 eV and the other was at 532.9 eV. The peak at 530.8 eV should be attributed to the Bi-O bonds the existed in BiOI. The peak at 532.9 eV should be attributed to the O-H bonds on the surface of BiOI [43]. The O 1s spectrum of Bi3O5I2 had three peaks at 530.7 eV, 531.4 eV and 532.9 eV. The peak at 530.7 eV belongs to the Bi3+-O bonds in the Bi3O5I2 lattice [45]. The peak at 531.4 eV belongs to the Bi5+-O bonds in the Bi3O5I2 lattice [43]. In addition, the peak at 532.9 eV belongs to the O-H bonds that existed on the Bi3O5I2 surface. As shown in Figure 5d, the two peaks can be observed at 630.1 eV and 618.6 eV. They were ascribed to I 3d5/2 and I 3d3/2, which could be assigned to the monovalent oxidation state of the I elements in BiOI [43]. It should be noted that the peaks for I 3d of the Bi3O5I2 samples (630.4 eV and 618.9 eV) presented a 0.3 eV shift to higher binding energy. This implied that the I in Bi3O5I2 possibly existed at a different ambient chemical environment [43]. The XPS character showed that the Bi in the Bi3O5I2 includes Bi3+ and Bi5+.
As shown in Figure 6, the specific surface area and the pore size distribution of Bi3O5I2 and BiOI were analyzed by the N2 absorption and desorption analysis. It was clear to see that all the isotherms were type IV, which exhibited the presence of mesoporous [43]. The hysteresis loop located at the range of 0.5–1.0 P/P0 were ascribed to type H3 in the IUPAC classification, indicating that the mesoporous could be considered as accumulated pores of the sample nanosheets [43]. Moreover, the hysteresis loop shifts approach P/P0 = 1, implying the existence of macroporous (>50 nm). The BET specific surface area of the Bi3O5I2 was 26.81 m2/g, while the surface area of BiOI was 13 m2/g, possibly due to its hollow microsphere structure. The PSD curves showed that the pore size range of Bi3O5I2 and BiOI was 2–140 nm. The main pore size of Bi3O5I2 was 21 nm, while the main pore size of BIOI was 39 nm. To our best knowledge, the larger specific surface area can lead to the increase of photocatalytic reaction sites [37]. Therefore, it could be concluded that the Bi3O5I2 might have a higher photocatalytic activity than BiOI.

2.3. Optical Property and Photoelectric Property of the Photocatalysts

The UV-vis diffuse reflectance spectroscopy of Bi3O5I2 and BiOI was shown in Figure 7a. In comparison with BiOI, Bi3O5I2 had a clearly blue shift in the absorption edge, indicating Bi3O5I2 had a higher energy gap (Eg). This agrees with the results of the I-deficient bismuth oxyiodide reported in the previous literature [46,47]. Meanwhile, the DRS displayed that the absorption band edge of Bi3O5I2 was 586 nm, which indicated that Bi3O5I2 had a good visible light absorption [29].
To estimate the Eg value of Bi3O5I2, a classical Tauc approach was introduced. According to the equation αhν = K(hν − Eg)n/2 (α, hν, K and Eg are the absorption coefficient, the energy of irradiation light, a constant and band gap energy, respectively), the Eg value can be calculated [43]. Considering BiOI and Bi3O5I2, were indirect transition semiconductor, the value of n is 4 [31,32,34]. The band gaps of Bi3O5I2 was calculated to be 2.02 eV and higher than that of BiOI (1.89 eV). In order to determine the relative positions of the valence band edge (VB), the VB-XPS spectra of Bi3O5I2 and BiOI were measured and shown in Figure 7b. The VB edge of Bi3O5I2 and BiOI was 1.50 eV and 1.42 eV. The conduction band (CB) edge of Bi3O5I2 and BiOI could be obtained using the equation of ECB = EVB − Eg. Therefore, the CB positions of Bi3O5I2 and BiOI were −0.72 eV and −0.47 eV. The suitable bang energy of Bi3O5I2 compared to BiOI implied that Bi3O5I2 has a stronger oxidation and reduction ability than BiOI [43]. As we all know, the electron in the less positive conduction band edge (CB) has higher reduction ability and the hole in the more positive valence band edge (VB) has higher oxidization ability. As we all know, the electron in the less positive conduction band edge (CB) has higher reduction ability and the hole in the more positive valence band edge (VB) has higher oxidization ability. Therefore, Bi3O5I2 with the less positive CB position and more positive VB position had the promising photacatalytic activity than BiOI. The energy band structure of Bi3O5I2 and BiOI were shown in Figure S2. The conduction band edge (CB) and the valence band edge (VB) of Bi3O5I2 were −0.72 eV and 1.50 eV (versus normal hydrogen electrode (NHE)). Since the CB value of Bi3O5I2 was less positive than E0(O2/•O2) (−0.046 eV vs. NHE) and the VB value of Bi3O5I2 was more positive than E0(•OH/OH) (2.38 eV vs. NHE) [48], the photogenerated electrons can reduce O2 to •O2 and the photogenerated holes cannot oxidize OH to •OH. Meanwhile, the VB value of Bi3O5I2 was more positive than the redox potential of MO (1.48 eV versus NHE) [49], the photogenerated holes can oxidize MO directly. The result of active species experiments shown in Figure 11 further confirm the above results. Based on the band structure analysis, Bi3O5I2 would exhibit effective photocatalytic performance, owing to its strong oxidation and reduction ability.
The transient photocurrent responses and electrochemical impedance spectroscopy (EIS) were further explored to estimate the generation and recombination rate of photogenerated electron-hole pairs, which was displayed in Figure 8. In Figure 8a, it was worth noting that the photocurrent density of the Bi3O5I2 electrode is much higher than BiOI for each switch-on and -off event. The enhanced photocurrent response of the Bi3O5I2 indicated a lower recombination of photogenerated electron-hole pairs than pure BiOI [50]. The photocurrent response mainly depends on the generation rate of photoinduced electron-hole pairs and the recombination rate of photogenerated electron-hole pairs. Meanwhile, the recombination rate of photogenerated electron-hole pairs were the rate controlling step of the photocurrent response. Just shown in Figure 7a, BiOI had a superior light harvesting than the I-deficient bismuth oxyiodides. It indicates that BiOI might have a higher generation rate of photoinduced electron-hole pairs than Bi3O5I2. However, it should be noted that the recombination rate of photogenerated electron-hole pairs of BiOI was much higher than Bi3O5I2. Therefore, the photocurrent response of Bi3O5I2 is stronger than that of BiOI. The result of the electrochemical impedance spectroscopy (EIS) of BiOI and Bi3O5I2 shown in Figure 8b can verify this result. Bi3O5I2 had a smaller arc radius of the EIS Nyquist plot than BiOI, suggesting an effective separation of photoinduced electron−hole pairs and fast interfacial charge transfer than BiOI [11]. In conclusion, Bi3O5I2 has a more efficient photocatalytic performance than BiOI.

2.4. Photocatalytic Performance of the Photocatalysts

The photocatalytic ability of Bi3O5I2 were measured on the decomposing of MO (20 mg/L) in water under the simulated visible light, which was shown in Figure 9. In Figure 9a, the maximum absorption of MO shifted from 464 nm to 420 nm in the presence of Bi3O5I2 after irradiation for 180 min [37]. Meanwhile, the color of MO changed from yellow to colorless during the photocatalytic reaction, which indicated that MO was degraded by Bi3O5I2 under the simulated solar light irradiation.
The variation of the MO content with the irradiation time was shown in Figure 9b. It can be seen that 81% of MO was photodegraded by Bi3O5I2 while only 11% of MO was degraded by BiOI. This confirmed that Bi3O5I2 present higher photoactivities than BiOI. Moreover, it was much higher than most of the activity data over other similar catalysts reported in previous works, as compared in Table S3.
The quantum efficiency of the photocatalytic activity is important for the photocatalysts. However, there is little information about the quantity of electrons needed to degrade a methyl orange molecule in literatures. Additionally it is also difficult to measure it in our Lab. In order to calculate the quantum efficiency (QE) of Bi3O5I2 and BiOI, the following equation about the apparent QE is put forward according to the literature and the calculated process is listed below in detail [12].
The photocatalytic reaction conditions: 100 mL, 20 mg/L methyl orange (MO 327.334 g/mol) solvent, the moles of MO is 6.11 × 10−6 mol. A light (λ ~ 420 nm) was used and the light intensity was 300 mW·cm−2. The energy can be obtained on the surface (1.3 cm−2) of the reaction mixture solution. We assumed that it takes m electrons to degrade a methyl orange molecule. n % of MO was degraded after the t hour. The apparent QE detected under the same photocatalytic reaction condition and calculated to the equation as below:
QE = number   of   reacted   eletrons number   of   incident   photons × 100 % = 6.11 × 10 6 m o l × m × n % E h ν   = 6.11 × 10 6 m o l × m × n % × h c E λ = 6.11 × 10 6 × m × n % × 6.02 × 10 23 × 6.63 × 10 34 × 3 × 10 8 300 × 1.3 × 10 3 × t × 3600 × 420 × 10 9 = 0.00124 m × n % t
When m = 1, it can be found that the QE of Bi3O5I2 and BiOI were 0.033% and 0.005% at 420 nm.
Based on the above analysis, the high photocatalytic activity of Bi3O5I2 should be ascribed to its hollow structure, suitable band structure and efficient separation of photogenerated electron–hole pairs. Firstly, with a hole in microspheres, Bi3O5I2 has a larger surface area (26.81 m2/g) than BiOI (13.00 m2/g). With a higher surface area, Bi3O5I2 had more active sites and absorb more MO on its surface. Moreover, the high surface area of hollow structures could promote the transfer of photoinduced carriers to the surface and enhance the decomposition of MO [37]. Secondly, Bi3O5I2 had a higher CB position and lower VB location than BiOI, which enable the photogenerated electron and hole of the Bi3O5I2 strong oxidation capability [9]. Thirdly, the photocurrent spectra showed that Bi3O5I2 had a lower recombination of photoinduced carriers under the simulated light irradiation. Therefore, Bi3O5I2 should have an enhanced photocatalytic performance since the photocatalytic activity of the photocatalyst mainly depend on the charge carrier generation and separation [9].

2.5. The Lifetime eValuation of the Photocatalysts

The reusability of the Bi3O5I2 material was investigated by recycling and reusing the catalyst (in Figure 10). Figure 10a shows no distinct loss in the photocatalytic activity when MO was degraded for the fourth time. Therefore, the Bi3O5I2 exhibited an excellent stability during the degradation of MO.
In order to evaluate the structural stability of Bi3O5I2, the fresh and used Bi3O5I2 were characterized by the XRD. The result was shown in Figure 10b. Obviously, there was no apparent change for the recycled Bi3O5I2 compared with the fresh Bi3O5I2, indicating the Bi3O5I2 are stable and reusable. Owing to its good photocatalytic activity as well as its stable and reusable character, Bi3O5I2 shows great potential for the photocatalytic degradation of organic pollutants in waste water.

2.6. The Possible Photacatalytic Mechanism

The probe mechanism Bi3O5I2 was further evaluated by the radicals and holes trapping experiments and the result was shown in Figure 11. Benzoquinone (BQ) was used as the quenching agent of superoxide radical (•O2). Potassium iodide (KI) was used as the scavenger of hole (h+) and hydroxyl radical (•OH). Tert-butyl alcohol (TBA) was used as the quenching agent of •OH. Their concentrations in the MO solution were 0.1 mM/L. Figure 11 showed that the degradation of MO was obviously decreased from 81% to 55% in the presence of KI, indicating that h+ or •OH played an important role for degrading MO [51]. While, the addition of TBA did not influence the MO degradation efficiency, indicating that •OH has no effect on the degrading MO. Therefore, it could be inferred that h+ was the major active species for the degradation of MO. When BQ was added, the MO degradation efficiency decreased significantly from 81% to 29%, suggesting that •O2 played the key role for degrading MO. In conclusion, •O2 was the primary active species and h+ was the main active species in the degradation of MO.
In conclusion of the analysis above, a possible photocatalytic program was proposed, which was displayed in Scheme 1. Firstly, the MO molecules were absorbed on the surface of the Bi3O5I2. Secondly, when the Bi3O5I2 was irradiated by simulated visible light, the photogengenerated electron would transfer from the valence band to the conduction band, leaving h+ in the valence band and electrons in the conduction band. Thirdly, h+ are capable of oxidizing MO directly and the electrons were also believed to be capable of benefiting the oxidation process through reduction of the absorbed O2 into •O2. As a result, the MO molecular was demineralized to CO2, H2O and other inorganic substances by the •O2 and h+.

3. Materials and Method

3.1. Materials

The bismuth nitrate pentahydrate (99.5 %, Sinopharm Chemical Reagent Co., Ltd. Shanghai, China), potassium iodide(99.5 %, Sinopharm Chemical Reagent Co., Ltd. Shanghai, China), ethyl alcohol and ethylene glycol (99.0 %, Sinopharm Chemical Reagent Co., Ltd. Shanghai, China) were used as received. The distilled water was synthesized by the water purification system.

3.2. Synthesis of the Bi3O5I2 and BiOI

In a typical synthesis route, 0.004 mol of bismuth nitrate pentahydrate and 0.004 mol of potassium iodide (KI) was dissolved in 40 mL of ethylene glycol under vigorously stirring for 0.5 h, respectively. After that, the KI solution was added dropwise into the solution which contained bismuth nitrate. Then, the mixture solution was stirred vigorously for 0.5 h at 25 °C and transferred into a 100 mL Teflon-lined stainless autoclave. The Teflon-lined stainless autoclave was heated to 160 °C and maintained for 12 h in a homogeneous reactor and air-cooled to room temperature. The synthesized precipitates were centrifuged and washed three times by ethanol and deionized water, and dried at 80 °C for 12 h in air. The sample was labelled as the I-deficient bismuth oxyiodide. Meanwhile, BiOI was prepared with the method above except using distilled water as the solvent of KI.

3.3. Photocatalyst Characterization

The samples are tested by the X-ray diffraction (XRD) with a D8 ADVANCE X-ray diffractometer (Cu Kα radiation, λ = 1.54178 Å Karlsruhe, German). A field emission scanning electron microscope (FE-SEM, JSM-7001F, Kyoto, Japan) was used to characterize the morphology of the samples. The FE-SEM was equipped with an energy dispersive analysis system of the X-ray spectroscope (EDS) as well as an elemental mapping system. The images of the transmission electron microscopy (TEM) for the samples were from a JEOL (JEM-2100F, Kyoto, Japan) operating at 200 kV. The N2 adsorption and desorption isotherms were investigated by a Tristar II 3020 sorptometer (Atlanta, GA, USA). The operating condition is at 77 K with the samples degassed at 100 °C in vacuo. The XPS spectrum was obtained with a KRATOS X-ray photoelectron spectroscopy(AXIS ULTRA DLD, Manchester, UK). The diffuse reflectance UV–vis absorbance spectroscopy were recorded by using a spectrophotometer (Shimadzu UV-2550, Kyoto, Japan) BaSO4 as the reference. The thermogravimetric analysis (TG) was carried out with a Netzsch STA 449C (Free State of Bavaria, German). In addition, the samples were heated with 10 °C/min from 25 °C to 900 °C in air atmosphere. The content of Bi was estimated by an Inductive Coupled Plasma emission spectrometer (Thermo iCAP6300, Waltham, MA, USA). The CHI 660E electrochemical system (Shanghai, China) was used for photoelectrochemical and electrochemical measurements and the sample preparation process was the same with the previous report [52].

3.4. Test of Photocatalytic Activity

To evaluate the photocatalytic activities of the as prepared samples, the photocatalytic degradation of MO was used under simulated visible light with a radiation source of the 300 W Xe lamp (PLS-SXE 300, Beijing, China). A 250 mL home-made reactor combined with a cooling water system was used. In each experiment, 100 mL of the MO solution (20 mg/L) was put in the reactor. A 25 mg catalyst was added into the MO solution. After that, the mixture of the MO solution and 25 mg catalyst placed in the dark was stirred for 60 min. Then, the mixture was exposed to simulated solar light with stirring. During the process, a 5 mL mixture was sampled at 30 min intervals. In addition, the concentration of MO in the mixture was measured by a UV–visible spectrophotometer. The spectrophotometer is the UV-2550 of Shimadzu, Japan, operating at 464 nm with water as a reference.

4. Conclusions

In summary, a new I-deficient bismuth oxyiodide Bi3O5I2 with a hollow structure was successfully prepared by an ethylene glycol-assisted solvothermal process. The average diameter of the Bi3O5I2 hollow microspheres was 1–3 μm, and the hollow nature contribute to the larger BET specific surface area than BiOI. The Bi3O5I2 presented a higher photocatalytic performance on the degradation of MO under simulated solar light irradiation than BiOI did. The enhanced photocatalytic activities of Bi3O5I2 could be contributed to its hollow structure, its appropriate band-gap and its low recombination rate of photogenerated carriers. The •O2 and h+ was the main active specie in the degradation of MO. The result of the stable and reusable performance indicated that Bi3O5I2 showed great potential for the photocatalytic degradation of organic pollutants in waste water.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/9/709/s1. Figure S1: SEM images and EDS analysis of as prepared samples in different locations: (a), (b) I-poor bismuth oxyiodide, (c), (d)BiOI, Figure S2: The energy band structure of Bi3O5I2 and BiOI, Table S1: The elemental composition of I-poor bismuth oxyiodide, Table S2: The elemental composition of BiOI, Table S3 Photocatalytic activity of different researches.

Author Contributions

B.C. designed and performed the experiments; B.C., Z.L., H.C. and J.W. analyzed the data; L.Z. provided financial support; B.C. wrote the paper; H.D. and X.L. reviewed and edited the paper.

Funding

This research was funded by the Science and Technology Major Project of Shanxi Province (No. 20181101018).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Peng, Y.; Yu, P.-P.; Zhou, H.-Y.; Xu, A.-W. Synthesis of BiOI/Bi4O5I2/Bi2O2CO3 p–n–p heterojunctions with superior photocatalytic activities. New J. Chem. 2015, 39, 8321–8328. [Google Scholar] [CrossRef]
  2. Dong, F.; Xiong, T.; Sun, Y.; Zhang, Y.X.; Zhou, Y. Controlling interfacial contact and exposed facets for enhancing photocatalysis via 2D–2D heterostructures. Chem. Commun. 2015, 51, 8249–8252. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, L.; Liu, X. Fast Degradation of Monochloroacetic Acid by BiOI-Enhanced UV/S(IV) Process: Efficiency and Mechanism. Catalysts 2019, 9, 460. [Google Scholar] [CrossRef]
  4. Liu, H.; Cao, W.-R.; Su, Y.; Chen, Z.; Wang, Y. Bismuth oxyiodide–graphene nanocomposites with high visible light photocatalytic activity. J. Colloid Interface Sci. 2013, 398, 161–167. [Google Scholar] [CrossRef] [PubMed]
  5. Cheng, H.; Huang, B.; Dai, Y. Engineering BiOX (X = Cl, Br, I) nanostructures for highly efficient photocatalytic applications. Nanoscale 2014, 6, 2009. [Google Scholar] [CrossRef] [PubMed]
  6. Zhu, G.; Hojamberdiev, M.; Zhang, S.; Din, S.T.U.; Yang, W. Enhancing visible-light-induced photocatalytic activity of BiOI microspheres for NO removal by synchronous coupling with Bi metal and graphene. Appl. Surf. Sci. 2019, 467, 968–978. [Google Scholar] [CrossRef]
  7. Ye, L.; Jin, X.; Ji, X.; Liu, C.; Su, Y.; Xie, H.; Liu, C. Facet-dependent photocatalytic reduction of CO2 on BiOI nanosheets. Chem. Eng. J. 2016, 291, 39–46. [Google Scholar] [CrossRef]
  8. Long, Y.; Wang, Y.; Zhang, D.; Ju, P.; Sun, Y. Facile synthesis of BiOI in hierarchical nanostructure preparation and its photocatalytic application to organic dye removal and biocidal effect of bacteria. J. Colloid Interface Sci. 2016, 481, 47–56. [Google Scholar] [CrossRef]
  9. Lee, G.-J.; Zheng, Y.-C.; Wu, J.J. Fabrication of hierarchical bismuth oxyhalides (BiOX, X = Cl, Br, I) materials and application of photocatalytic hydrogen production from water splitting. Catal. Today 2018, 307, 197–204. [Google Scholar] [CrossRef]
  10. Dai, W.W.; Zhao, Z.Y. Electronic structure and optical properties of BiOI as a photocatalyst driven by visible light. Catalysts 2016, 6, 133. [Google Scholar] [CrossRef]
  11. Huang, H.; Xiao, K.; Zhang, T.; Dong, F.; Zhang, Y. Rational design on 3D hierarchical bismuth oxyiodides via in situ self-template phase transformation and phase-junction construction for optimizing photocatalysis against diverse contaminants. App. Catal. B Environ. 2017, 203, 879–888. [Google Scholar] [CrossRef]
  12. Bai, Y.; Chen, T.; Wang, P.; Wang, L.; Ye, L. Bismuth-rich Bi4O5X2 (X = Br, and I) nanosheets with dominant {101} facets exposure for photocatalytic H2 eVolution. Chem. Eng. J. 2016, 304, 454–460. [Google Scholar] [CrossRef]
  13. Hu, J.; Weng, S.; Zheng, Z.; Pei, Z.; Huang, M.; Liu, P. Solvents mediated-synthesis of BiOI photocatalysts with tunable morphologies and their visible-light driven photocatalytic performances in removing of arsenic from water. J. Hazard. Mater. 2014, 264, 293–302. [Google Scholar] [CrossRef]
  14. Cao, J.; Zhou, C.; Lin, H.; Xu, B.; Chen, S. Direct hydrolysis preparation of plate-like BiOI and their visible light photocatalytic activity for contaminant removal. Mater. Lett. 2013, 109, 74–77. [Google Scholar] [CrossRef]
  15. Di, J.; Xia, J.; Ge, Y.; Xu, L.; Xu, H.; He, M.; Zhang, Q.; Li, H. Reactable ionic liquid-assisted rapid synthesis of BiOI hollow microspheres at room temperature with enhanced photocatalytic activity. J. Mater. Chem. A 2014, 2, 15864–15874. [Google Scholar] [CrossRef]
  16. Zhang, G.; Su, A.; Qu, J.; Xu, Y. Synthesis of BiOI flowerlike hierarchical structures toward photocatalytic reduction of CO2 to CH4. Mater. Res. Bull. 2014, 55, 43–47. [Google Scholar] [CrossRef]
  17. Han, J.; Zhu, G.; Hojamberdiev, M.; Peng, J.; Zhang, X.; Liu, Y.; Ge, B. Rapid adsorption and photocatalytic activity for Rhodamine B and Cr(vi) by ultrathin BiOI nanosheets with highly exposed {001} facets. New J. Chem. 2015, 39, 1874–1882. [Google Scholar] [CrossRef]
  18. Shan, L.-W.; He, L.-Q.; Suriyaprakash, J.; Yang, L.-X. Photoelectrochemical (PEC) water splitting of BiOI{001} nanosheets synthesized by a simple chemical transformation. J. Alloys Compd. 2016, 665, 158–164. [Google Scholar] [CrossRef]
  19. Huang, H.; Xiao, K.; He, Y.; Zhang, T.; Dong, F.; Du, X.; Zhang, Y. In situ assembly of BiOI@Bi12O17Cl2 p—n junction: Charge induced unique front-lateral surfaces coupling heterostructure with high exposure of BiOI {001} active facets for robust and nonselective photocatalysis. Appl. Catal. B Environ. 2016, 199, 75–86. [Google Scholar] [CrossRef]
  20. Zeng, L.; Zhe, F.; Wang, Y.; Zhang, Q.; Zhao, X.; Hu, X.; Wu, Y.; He, Y. Preparation of interstitial carbon doped BiOI for enhanced performance in photocatalytic nitrogen fixation and methyl orange degradation. J. Colloid Interf. Sci. 2019, 539, 563–574. [Google Scholar] [CrossRef]
  21. Dai, B.; Zhang, A.; Liu, Z.; Wang, T.; Li, C.; Zhang, C.; Li, H.; Liu, Z.; Zhang, X. Facile synthesis of metallic Bi deposited BiOI composites with the aid of EDTA-2Na for highly efficient Hg-0 removal. Catal. Commun. 2019, 121, 53–56. [Google Scholar] [CrossRef]
  22. Liu, H.; Cao, W.; Su, Y.; Wang, Y.; Wang, X. Synthesis, characterization and photocatalytic performance of novel visible-light-induced Ag/BiOI. Appl. Catal. B Environ. 2012, 111, 271–279. [Google Scholar] [CrossRef]
  23. Huang, H.; Liu, K.; Zhang, Y.; Chen, K.; Zhang, Y.; Tian, N. Tunable 3D hierarchical graphene–BiOI nanoarchitectures: Their in situ preparation, and highly improved photocatalytic performance and photoelectrochemical properties under visible light irradiation. RSC Adv. 2014, 4, 49386–49394. [Google Scholar] [CrossRef]
  24. Shi, X.; Wang, P.; Wang, L.; Bai, Y.; Xie, H.; Zhou, Y.; Ye, L. Change in photocatalytic NO removal mechanisms of ultrathin BiOBr/BiOI via NO3-adsorption. Appl. Catal. B Environ. 2019, 243, 322–329. [Google Scholar] [CrossRef]
  25. Al-Keisy, A.; Ren, L.; Xu, X.; Hao, W.; Dou, S.X.; Du, Y. Selective ferroelectric BiOI/Bi4Ti3O12 heterostructures for visible light-driven photocatalysis. J. Phys. Chem. C 2019, 123, 517–525. [Google Scholar] [CrossRef]
  26. Yan, P.; Jiang, D.; Li, H.; Cheng, M.; Xu, L.; Qian, J.; Bao, J.; Xia, J.; Li, H. Exploitation of a photoelectrochemical sensing platform for catechol quantitative determination using BiPO4 nanocrystals/BiOI heterojunction. Anal. Chim. Acta 2018, 1042, 11–19. [Google Scholar] [CrossRef]
  27. Gao, P.; Yan, T.; Liu, H.; Sun, M.; Wei, Q.; Xu, W.; Wang, X.; Du, B. Facile synthesized highly active BiOI/Zn2GeO4 composites for the elimination of endocrine disrupter BPA under visible light irradiation. New J. Chem. 2015, 39, 3964–3972. [Google Scholar]
  28. Jin, X.; Lv, C.; Zhou, X.; Zhang, C.; Meng, Q.; Liu, Y.; Chen, G. Molecular adsorption promotes carrier migration: Key step for molecular oxygen activation of defective Bi4O5I2. Appl. Catal. B Environ. 2018, 226, 53–60. [Google Scholar] [CrossRef]
  29. Cao, J.; Li, X.; Lin, H.; Xu, B.; Luo, B.; Chen, S. Low temperature synthesis of novel rodlike Bi5O7I with visible light photocatalytic performance. Mater. Lett. 2012, 76, 181–183. [Google Scholar] [CrossRef]
  30. Dai, B.J.; Zhang, A.C.; Zhang, D.; Liu, Z.C.; Li, H.X.; Wang, R.R.; Zhang, X.M. Effect of preparation method on the structure and photocatalytic performance of BiOI and Bi5O7I for Hg-0 removal. Atmos. Pollut. Res. 2019, 10, 355–362. [Google Scholar] [CrossRef]
  31. Sun, S.M.; Wang, W.Z.; Zhang, L.; Zhou, L.; Yin, W.Z.; Shang, M. Visible light-induced efficient contaminant removal by Bi5O7I. Environ. Sci. Technol. 2009, 43, 2005–2010. [Google Scholar] [CrossRef]
  32. Chuang, C.-W.; Siao, C.-W.; Lee, W.W.; Lu, C.-S.; Chen, Y.-J.; Fu, J.-Y. Synthesis of bismuth oxyiodides and their composites: Characterization, photocatalytic activity, and degradation mechanisms. RSC Adv. 2015, 5, 23450–23463. [Google Scholar]
  33. Xiao, X.; Zhang, W.D. Hierarchical Bi7O9I3 micro/nano-architecture: Facile synthesis, growth mechanism, and high visible light photocatalytic performance. Rsc. Adv. 2011, 1, 1099–1105. [Google Scholar] [CrossRef]
  34. Xiao, X.; Xing, C.; He, G.; Zuo, X.; Nan, J.; Wang, L. Solvothermal synthesis of novel hierarchical Bi4O5I2 nanoflakes with highly visible light photocatalytic performance for the degradation of 4-tert-butylphe. Appl. Catal. B Environ. 2014, 148, 154–163. [Google Scholar] [CrossRef]
  35. Zhang, L.; Gonçalves, A.A.S.; Jiang, B.; Jaroniec, M. Capture of Iodide by Bismuth Vanadate and Bismuth Oxide: An Insight into the Process and its Aftermath. ChemSusChem 2018, 11, 1486–1493. [Google Scholar] [CrossRef]
  36. Liu, Q.-C.; Ma, D.-K.; Hu, Y.-Y.; Zeng, Y.-W.; Huang, S.-M. Various Bismuth Oxyiodide Hierarchical Architectures: Alcohothermal-Controlled Synthesis, Photocatalytic Activities, and Adsorption Capabilities for Phosphate in Water. ACS Appl. Mater. Interfaces 2013, 5, 11927–11934. [Google Scholar] [CrossRef]
  37. Xia, J.; Yin, S.; Li, H.; Xu, H.; Yan, Y.; Zhang, Q. Self-Assembly and Enhanced Photocatalytic Properties of BiOI Hollow Microspheres via a Reactable Ionic Liquid. Langmuir 2011, 27, 1200–1206. [Google Scholar] [CrossRef]
  38. Ma, D.-K.; Zhou, S.-M.; Hu, X.; Jiang, Q.-R.; Huang, S.-M. Hierarchical BiOI and hollow Bi2WO6 microspheres: Topochemical conversion and photocatalytic activities. Mater. Chem. Phys. 2013, 140, 11–15. [Google Scholar] [CrossRef]
  39. Zhang, K.; Zhang, D.; Liu, J.; Ren, K.; Luo, H.; Peng, Y.; Li, G.; Yu, X. A novel nanoreactor framework of iodine-incorporated BiOCl core–shell structure: Enhanced light-harvesting system for photocatalysis. CrystEngComm 2012, 14, 700–707. [Google Scholar] [CrossRef]
  40. Ji, M.; Xia, J.; Di, J.; Wang, B.; Yin, S.; Xu, L.; Zhao, J.; Li, H. Ionic liquid-assisted bidirectional regulation strategy for carbon quantum dots (CQDs)/Bi4O5I2 nanomaterials and enhanced photocatalytic properties. J. Colloid Interface Sci. 2016, 478, 324–333. [Google Scholar] [CrossRef]
  41. Tu, S.; Zheng, C.; Zhong, H.; Lu, M.; Xiao, X.; Zuo, X.; Nan, J. Flower-like Bi4O5I2/Bi5O7 I nanocomposite: Facile hydrothermal synthesis and efficient photocatalytic degradation of propylparaben under visible-light irradiation. RSC Adv. 2016, 6, 44552–44560. [Google Scholar] [CrossRef]
  42. Ren, K.; Zhang, K.; Liu, J.; Luo, H.; Huang, Y.; Yu, X. Controllable synthesis of hollow/flower-like BiOI microspheres and highly efficient adsorption and photocatalytic activity. CrystEngComm 2012, 14, 4384. [Google Scholar] [CrossRef]
  43. Liu, C.; Wang, X.-J. Room temperature synthesis of Bi 4 O 5 I 2 and Bi 5 O 7 I ultrathin nanosheets with a high visible light photocatalytic performance. Dalton Trans. 2016, 45, 7720–7727. [Google Scholar] [CrossRef]
  44. Wang, W.; Chen, X.-Q.; Liu, G.; Shen, Z.; Xia, D.; Wong, P.K.; Yu, J.C. Monoclinic dibismuth tetraoxide: A new visible-light-driven photocatalyst for environmental remediation. Appl. Catal. B Environ. 2015, 176, 444–453. [Google Scholar] [CrossRef]
  45. Long, Y.; Li, L.; Wang, S.; Chen, Y.; Wang, L.; Zhang, S.; Luo, L.; Jiang, F. Photocatalytic Removal of 17α-Ethinyl Estradiol Using the Bi2O3/Bi2O4 Photocatalyst. Catal. Lett. 2018, 148, 3608–3617. [Google Scholar] [CrossRef]
  46. Xia, J.; Ji, M.; Di, J.; Wang, B.; Yin, S.; He, M.; Zhang, Q.; Li, H. Improved photocatalytic activity of few-layer Bi4O5I2 nanosheets induced by efficient charge separation and lower valence position. J. Alloys Compd. 2017, 695, 922–930. [Google Scholar] [CrossRef]
  47. Cui, S.; Shan, G.; Zhu, L. Solvothermal synthesis of I-deficient BiOI thin film with distinct photocatalytic activity and durability under simulated sunlight. Appl. Catal. B Environ. 2017, 219, 249–258. [Google Scholar] [CrossRef]
  48. Cheng, H.; Huang, B.; Dai, Y.; Qin, X.; Zhang, X. One-Step Synthesis of the Nanostructured AgI/BiOI Composites with Highly Enhanced Visible-Light Photocatalytic Performances. Langmuir 2010, 26, 6618–6624. [Google Scholar] [CrossRef]
  49. Wang, Y.; Deng, K.; Zhang, L. Visible Light Photocatalysis of BiOI and Its Photocatalytic Activity Enhancement by in Situ Ionic Liquid Modification. J. Phys. Chem. C 2011, 115, 14300–14308. [Google Scholar] [CrossRef]
  50. Ning, S.B.; Lin, H.X.; Tong, Y.C.; Zhang, X.Y.; Lin, Q.Y.; Zhang, Y.Q.; Long, J.L.; Wang, X.X. Dual couples Bi metal depositing and Ag@AgI islanding on BiOI 3D architectures for synergistic bactericidal mechanism of E-coli under visible light. Appl. Catal. B Environ. 2017, 204, 53–60. [Google Scholar] [CrossRef]
  51. Chang, C.; Zhu, L.; Fu, Y.; Chu, X. Highly active Bi/BiOI composite synthesized by one-step reaction and its capacity to degrade bisphenol A under simulated solar light irradiation. Chem. Eng. J. 2013, 233, 305–314. [Google Scholar] [CrossRef]
  52. Bao, C.; Wang, C.; Fan, D.; Ma, H.; Hu, L.; Fan, Y.; Wei, Q. A novel sandwich-type photoelectrochemical sensor for SCCA detection based on Ag2S-sensitized BiOI matrix and AucorePdshell nanoflower label for signal amplification. New J. Chem. 2018, 42, 15762–15769. [Google Scholar] [CrossRef]
Figure 1. The SEM image of the I-deficient bismuth oxyiodide (a) and Bi, O, I elemental maps of I-deficient bismuth oxyiodide (bd).
Figure 1. The SEM image of the I-deficient bismuth oxyiodide (a) and Bi, O, I elemental maps of I-deficient bismuth oxyiodide (bd).
Catalysts 09 00709 g001
Figure 2. Thermogravimetric (TG) profiles of Bi3O5I2 and bismuth oxyiodide (BiOI) samples in air flow.
Figure 2. Thermogravimetric (TG) profiles of Bi3O5I2 and bismuth oxyiodide (BiOI) samples in air flow.
Catalysts 09 00709 g002
Figure 3. The XRD patterns of: (a) Bi3O5I2; (b) BiOI.
Figure 3. The XRD patterns of: (a) Bi3O5I2; (b) BiOI.
Catalysts 09 00709 g003
Figure 4. (a,b) SEM image of Bi3O5I2; (c,d) TEM image of Bi3O5I2; (e,f) SEM image of BiOI.
Figure 4. (a,b) SEM image of Bi3O5I2; (c,d) TEM image of Bi3O5I2; (e,f) SEM image of BiOI.
Catalysts 09 00709 g004
Figure 5. XPS spectra of Bi3O5I2 and BiOI: (a) Survey, (b) Bi 4f, (c) O 1s, (d) I 3d.
Figure 5. XPS spectra of Bi3O5I2 and BiOI: (a) Survey, (b) Bi 4f, (c) O 1s, (d) I 3d.
Catalysts 09 00709 g005
Figure 6. Nitrogen absorption-desorption isotherms and the pore-size distribution (inset) of (a) Bi3O5I2 and (b) BiOI.
Figure 6. Nitrogen absorption-desorption isotherms and the pore-size distribution (inset) of (a) Bi3O5I2 and (b) BiOI.
Catalysts 09 00709 g006
Figure 7. (a) Diffuse reflectance spectra and band gaps of Bi3O5I2 and BiOI (inset), (b) valence band XPS spectra of Bi3O5I2 and BiOI.
Figure 7. (a) Diffuse reflectance spectra and band gaps of Bi3O5I2 and BiOI (inset), (b) valence band XPS spectra of Bi3O5I2 and BiOI.
Catalysts 09 00709 g007
Figure 8. (a) Photocurrent responses of Bi3O5I2 and BiOI in Na2SO4 solutions, (b) electrochemical impedance spectroscopy of Bi3O5I2 and BiOI.
Figure 8. (a) Photocurrent responses of Bi3O5I2 and BiOI in Na2SO4 solutions, (b) electrochemical impedance spectroscopy of Bi3O5I2 and BiOI.
Catalysts 09 00709 g008
Figure 9. (a) The UV-Vis absorption spectral changers during the degradation of MO (20 mg/L), (b) photocatalytic decomposition of MO in the presence of Bi3O5I2 and BiOI and photolysis of MO under simulated visible light irradiation.
Figure 9. (a) The UV-Vis absorption spectral changers during the degradation of MO (20 mg/L), (b) photocatalytic decomposition of MO in the presence of Bi3O5I2 and BiOI and photolysis of MO under simulated visible light irradiation.
Catalysts 09 00709 g009
Figure 10. (a) Recycling measurements for Bi3O5I2, (b) XRD patterns of the Bi3O5I2 before and after the cycling photocatalytic experiments.
Figure 10. (a) Recycling measurements for Bi3O5I2, (b) XRD patterns of the Bi3O5I2 before and after the cycling photocatalytic experiments.
Catalysts 09 00709 g010
Figure 11. Photocatalytic degradation of MO by the Bi3O5I2 under the presence of different scavenging species.
Figure 11. Photocatalytic degradation of MO by the Bi3O5I2 under the presence of different scavenging species.
Catalysts 09 00709 g011
Scheme 1. Hollow structure of Bi3O5I2 composite and probe processes of the photodegradation of MO over the Bi3O5I2 composite.
Scheme 1. Hollow structure of Bi3O5I2 composite and probe processes of the photodegradation of MO over the Bi3O5I2 composite.
Catalysts 09 00709 sch001

Share and Cite

MDPI and ACS Style

Cui, B.; Cui, H.; Li, Z.; Dong, H.; Li, X.; Zhao, L.; Wang, J. Novel Bi3O5I2 Hollow Microsphere and Its Enhanced Photocatalytic Activity. Catalysts 2019, 9, 709. https://doi.org/10.3390/catal9090709

AMA Style

Cui B, Cui H, Li Z, Dong H, Li X, Zhao L, Wang J. Novel Bi3O5I2 Hollow Microsphere and Its Enhanced Photocatalytic Activity. Catalysts. 2019; 9(9):709. https://doi.org/10.3390/catal9090709

Chicago/Turabian Style

Cui, Baoyin, Haitao Cui, Zhenrong Li, Hongyu Dong, Xin Li, Liangfu Zhao, and Junwei Wang. 2019. "Novel Bi3O5I2 Hollow Microsphere and Its Enhanced Photocatalytic Activity" Catalysts 9, no. 9: 709. https://doi.org/10.3390/catal9090709

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop