Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination under solar light irradiation

Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination under solar light irradiation

JES-01691; No of Pages 10 JOUR N AL OF EN VI R ONM E NT AL S CI E N CE S XX ( XXX X) X XX Available online at www.sciencedirect.com ScienceDirect ww...

2MB Sizes 0 Downloads 51 Views

JES-01691; No of Pages 10 JOUR N AL OF EN VI R ONM E NT AL S CI E N CE S XX ( XXX X) X XX

Available online at www.sciencedirect.com

ScienceDirect www.elsevier.com/locate/jes

Q34Q2 5 6

F

3

O

2

Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination under solar light irradiation

Yonghua Zhang⁎⁎, Guoqiang Shan⁎⁎, Fengfeng Dong, Chenshuai Wang, Lingyan Zhu⁎

R O

1Q1

Ministry of Education, Tianjin Key Laboratory of Environmental Remediation and Pollution Control, Key Laboratory of Pollution Processes and Environmental Criteria, College of Environmental Science and Engineering, Nankai University, Tianjin 300350, China

P

7

AR TIC LE I NFO

ABSTR ACT

11 16

Article history:

BiOI powder has been proved to be an efficient photocatalyst, but the difficulty in removing it

12 17

Received 28 October 2018

from water after reaction limits its application in real water treatment. To solve this problem,

13 18

Revised 5 January 2019

a thin-film fixed-bed reactor (TFFBR) was set-up by developing a BiOI thin film on glass fiber

14 19

Accepted 7 January 2019

cloth (GFC). The composition and structure of the as-prepared films were characterized with

15 20

Available online xxxx

X-ray diffraction, X-ray photoelectron spectroscopy, field emission microscopy, transmis-

21 35 22

Keywords:

was made by painting a silica sol containing BiOI on GFC, which could be tailored to desired

36 23 37 24

BiOI

sizes to accommodate the TFFBR. The mass of BiOI on the GFC increased with the number of

Sol–gel

iterations of the painting process. SiO2 sol glued the BiOI particles tightly onto the GFC,

38 25 39 26

Photocatalytic film

making the thin film strong enough to resist fluid flushing in the TFFBR. The photocatalytic

TFFBR

activity of the BiOI thin film was investigated by degrading bisphenol A (BPA) under

40 27 41 28

BPA

simulated sunlight. Ninety eight percent of BPA (20 mg/L in 2 L) was degraded by the BiOI thin

T

E

D

1 90

R

E

C

sion electron microscopy and Fourier transform infrared spectroscopy. The BiOI thin film

film sample of seven layers (GFC-7) on the TFFBR within 8 hr irradiation. The GFC-7 displayed

29

O R

good photocatalytic ability toward artificial sewage containing BPA in a wide pH range (5–9),

30

and also demonstrated excellent durability and reusability. The working conditions were

31

optimized and it was found that the thickness of the fluid film and residence time over the

32

thin film were key factors affecting the photocatalytic efficiency.

33

N C

© 2019 The Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences.

34 44 43 42 45

Introduction

48

Environmental pollution is becoming one of the major challenges in the twenty-first century, especially for the aquatic environment, which is contaminated by a variety of emerging organic contaminants (Rodriguez-Narvaez et al., 2017). Among the available biological, physical and chemical treatment techniques, photocatalytic remediation driven by

49 50 51 52 53

U

47 46

Published by Elsevier B.V.

solar light has received intensive interest since it has the advantages of being energy-saving and generating little or no secondary contaminants (Malato et al., 2009, 2016). As members of the class of visible-light-driven photocatalysts, bismuth oxyhalides (BiOX, X= Cl, Br, I) display superior photocatalytic activity (Cheng et al., 2014; Zhang et al., 2018b). The excellent visible photo-response properties of BiOX are attributed to their outstanding optical and electrical properties (Ye et al.,

⁎ Corresponding author. E-mail address: [email protected]. (L. Zhu). ⁎⁎ The authors contributed equally to this work

https://doi.org/10.1016/j.jes.2019.01.004 1001-0742 © 2019 The Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences. Published by Elsevier B.V.

Please cite this article as: Y. Zhang, G. Shan, F. Dong, et al., Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination unde..., Journal of Environmental Sciences, https://doi.org/10.1016/j.jes.2019.01.004

54 55 56 57 58 59 60 61

2

76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121

F

75

O

74

R O

73

P

72

D

71

122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144

1. Experimental

146 145

1.1. Materials and reagents

147

E

70

many excellent photocatalysts have been fabricated, there is a large gap between laboratory studies and application of these photocatalysts in real-water treatment systems to eliminate pollutants. At present, the photocatalytic reactors mainly include the parabolic trough solar reactor (Hong et al., 2012), double skin sheet reactor (Braham and Harris, 2009) and thinfilm fixed-bed reactor (TFFBR) (Braham and Harris, 2009; Zayani et al., 2009). The TFFBR has the advantages of simple structure, convenient operation and low cost. In addition, the TFFBR has a high utilization rate of light, and favors transfer of oxygen to the liquid film efficiently (Braham and Harris, 2009; Khan et al., 2012; Zayani et al., 2009). In this study, BiOI thin films on GFC were fabricated by a sol–gel method. The obtained films were characterized and applied in a TFFBR for water treatment under simulated solar light irradiation. Bisphenol A (BPA) is regarded as an endocrine disrupting compound and can cause negative effects on immune, reproductive and nerve systems. Thus, BPA was chosen as model pollutant for evaluating the decontamination efficiency of the reactor (Huang et al., 2012; Krishnan et al., 1993). To evaluate the application in real-water treatment, artificial sewage containing BPA was prepared, and the durability and reusability of the films were evaluated.

T

69

C

68

E

67

R R

66

O

65

C

64

2014; Kong et al., 2018). The internal electric field between [Bi2O2]2+ and X− layers in the materials enables easy separation of the photogenerated electron hole pairs, leading to excellent photocatalytic performance. With increasing atomic number, the BiOX displays descending band gaps from BiOCl (3.2 eV) to BiOI (1.8 eV). Thus, BiOI displays superior visible light adsorption ability among the BiOX. Currently, BiOI is generally employed in powdered form, which may significantly limit its practical application. The photocatalytic performance could decrease due to agglomeration of the BiOI nanoparticles (Lopez-Salinas et al., 2010). Moreover, it is very difficult to separate powdered photocatalysts from the reaction system, which inevitably increases the cost of the water purification process and might lead to secondary contamination. One efficient technique is to immobilize catalysts on a solid substrate (Lopez-Salinas et al., 2010; Zhang et al., 2009; Sudrajat and Babel, 2016). In previous studies, many efforts have been made to prepare TiO2 films supported on solid substrates, which were used in photocatalytic reactors driven by UV light irradiation (Abdel-Maksoud et al., 2016; Jafarikojour et al., 2015; McCullagh et al., 2011; Tugaoen et al., 2018). However, there are very few reports on preparation of BiOX thin films, which could be used for building a solar-light-driven photocatalytic reactor for real water decontamination. To build up TiO2 thin-film-based UV-light driven reactors, various substrates such as glass plates (Cui et al., 2017; Mehrjouei et al., 2013; Vella et al., 2010; Ye et al., 2013), quartz fibers (Jia et al., 2015; Kim et al., 2012), glass beads (Chiou et al., 2006; Shen et al., 2012), stainless steel (Momeni et al., 2017; Wang et al., 2017; Yu et al., 2003) and glass fibers (Erjavec et al., 2016; Zazhigalov et al., 2017) have been used as supporting substrates. Among them, glass fiber cloth (GFC) is particularly promising due to its special characteristics, such as low cost, good light transmittance, flexibility, corrosion resistance, large surface area and adaptability (being easily cut into desired sizes) (Erjavec et al., 2016; Jafarikojour et al., 2015). Various methods including chemical vapor transport (Starr et al., 2016; Ye et al., 2013; Ye and Chen, 2016), electrodeposition methods (Cruz-Ortiz et al., 2016; Hachisu et al., 2016) and sol– gel methods (Mohamed et al., 2017; Zhou et al., 2017), have been developed to prepare substrate-supported photocatalytic thin films. Among these methods, the sol–gel method has found wide application since it does not require expensive equipment or complicated operations (Yuan et al., 2016). In recent years, some efforts have been made to immobilize BiOX on the surface of supporting substrates such as glass and cement using direct coating or sol–gel silica coating methods (Cui et al., 2017; Jia et al., 2015; Shen et al., 2015; Wang et al., 2018). Direct coating followed by calcination is not suitable for preparation of a film in a large-scale reactor. A sol–gel coating with silica as an adhesion agent might be a noteworthy choice for BiOX thin-film preparation (Shen et al., 2015; Wang et al., 2018; Yuan et al., 2016). Previous reports on composite catalysts containing SiO2, such as TiO2@SiO2 and BiOBr@SiO2, indicated that the catalytic activity of TiO2 or BiOBr was not apparently affected by SiO2, or was even higher than the original activity (Wang et al., 2018; Yuan et al., 2016). Developing photocatalytic reactors is the key to large-scale application of photocatalysts in water treatment. Although

N

63

U

62

JOUR N AL OF EN VI RON M EN T AL S CI E NC ES XX ( XXX X) X XX

Bisphenol A (BPA, 98%) was purchased from Aladdin Industrial Corporation (Shanghai, China). Bismuth nitrate pentahydrate (Bi(NO3)3•5H2O), potassium iodide (KI), tetraethylorthosilicate (TEOS) and other conventional reagents were all analytical grade and obtained from native chemical reagent companies. Glass fiber cloth (GFC, alkalifree, 200 g/m2) was purchased from Japan East Electrician Ltd. (Suzhou, China). Deionized water was used throughout all experiments.

148

1.2. Synthesis of BiOI thin films on GFC

157

The silica coating of BiOI was carried out using a sol–gel method by mixing the as-prepared BiOI powder with silica sol. Five mL of TEOS was dissolved in 2.6 mL of ethanol, and then vigorously stirred for 0.5 hr. Afterwards, 1.7 mL of 0.7 mol/L HCl was added (Jia et al., 2015), followed by refluxing at 78°C for 13 min, which resulted in SiO2 sol. Subsequently, 2.0 g of BiOI powder, which was prepared using a solvothermal method described in a previous report (Chang et al., 2013), was thoroughly dispersed in the mixture of freshly prepared SiO2 sol (1.5 mL) and ethanol (30 mL). The mixture was stirred magnetically for 1 hr to obtain a uniform suspension. The GFC was pretreated by a sintering method (Sangkhun et al., 2012) to remove impurities. The prepared suspension was evenly painted onto the sintered GFC using a soft brush. The coated cloth was dried in an oven at 100°C for 1 hr, and the obtained thin film sample was named as GFC-1. This process was repeated several times to get GFC-3, -5 and -7, where the number represents the layers of BiOI on the cloth. The immobilized BiOI of GFC-7 was scraped off, which was

158

Please cite this article as: Y. Zhang, G. Shan, F. Dong, et al., Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination unde..., Journal of Environmental Sciences, https://doi.org/10.1016/j.jes.2019.01.004

149 150 151 152 153 154 155 156

159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176

3

JOUR N AL OF EN VI R ONM E NT AL S CI E N CE S XX ( XXX X) X XX

1.4. Photocatalytic test and BPA analysis

214

181

1.3. Characterization of the BiOI powder and as-prepared films

215

182

The X-ray diffraction patterns (XRD) were obtained on a D/ MAX 2500V diffractometer (Rigaku, Japan) with Cu Kα radiation to obtain the crystal structure of the samples. The morphologies of the samples were observed using a field emission scanning electron microscope (FESEM, LEO, Germany) and a transmission electron microscope (TEM, JEM2100FEF, JEOL, Japan). X-ray photoelectron spectroscopy (XPS, Axis Ultra DLD, Kratos, Japan) was performed to confirm the elemental compositions and chemical states of the materials. The Fourier transform infrared (FT-IR) spectra were obtained on a MAGNA-560 Infrared Spectrometer (Nicolet, USA). Photocurrent measurement was performed on a CHI 600D electrochemical system (Chenhua Instruments Co., Shanghai, China) (Cui et al., 2017). The photoluminescence (PL) spectra for solid samples were obtained on a FP-8500 fluorescence spectrophotometer (JASCO, Japan) with an excitation wavelength of 320 nm. The surface areas were obtained using the Brunauer–Emmett–Teller (BET) method, and the average pore diameter was obtained from the sorption/desorption isotherms of N2 using the Barrett–Joyner–Halenda (BJH) equation on data obtained by a surface area and porosimetry analyzer (TriStar 3000, Micromeritics, USA). A UV–Vis spectrometer (U3010, Hitachi, Japan) in the range of 200–800 nm was used to measure UV–Vis diffuse reflectance spectra (DRS) of the samples. The band gap of the as-prepared catalysts was calculated using the following equation (Ma et al., 2012): αhv = A(hv-Eg)n/2. α, h, v and Eg present the absorption coefficient, Planck constant, photonic frequency and semiconductor band gap energy, respectively, and A is a constant. The parameter n depends on the type of transition, with n = 1

The photocatalytic activity of the as-prepared BiOI thin films and the scraped BiOI-7 powder was evaluated on an XPA-7 photochemical reactor (Xujiang Electromechanical Plant, Nanjing, China). An 800 W Xe lamp was used as light source. The prepared BiOI thin films (3 × 15 cm) were vertically placed in quartz tubes containing 40 mL of BPA (20 mg/L). For comparison, 20 mg of pure BiOI powder was added in one of the tubes containing 40 mL of BPA (20 mg/L). Before irradiation, the suspensions were stirred in the dark for 1 hr to establish adsorption–desorption equilibrium. At certain time intervals, 1 mL of solution was taken out and centrifuged at 10,000 r/min. The supernatant was sampled and analyzed on a High-Performance Liquid Chromatograph (HPLC, Agilent 1260 infinity, Agilent Corporation, USA) with a Fluorescence Detector. More detailed instrumental conditions are provided in Appendix A. The stability of the as-prepared thin-film GFC7 was tested through a recycling test.

1.5. Photocatalytic activity evaluation of the TFFBR

232

A thin-film fixed-bed reactor (TFFBR) was set-up, consisting of a peristaltic pump, two reservoirs, a light source and a sloping plate for supporting the film and collecting light (Fig. 1), according to a previous design with minor modifications (Khan et al., 2012; Sudrajat and Babel, 2016). The BiOI thin-film with a working area of 25 × 30 cm was attached on the sloping plate. At the top of the sloping plate, there was a water buffer zone with a splitter, which guaranteed that the reaction solution flowed over the immobilized film smoothly and evenly. The BPA (20 mg/L) solution in a reservoir of 2.5 L was pumped by a peristaltic pump to the top buffer zone through the splitter, and flowed along the sloping plate as a fluid film. The slope of the plate and the flow rate were adjustable. The position of an 800 W Xe lamp was adjusted to make sure the films were perpendicular to the light source at a 10 cm distance. After a given irradiation time, an aliquot of the solution was sampled and analyzed by HPLC. In order to make the experiment more realistic, 2 L of artificial sewage (Chen et al., 1995) (the ingredients of the artificial sewage are provided in Appendix A Table S1) containing BPA (20 mg/L) was used throughout the reactor tests. The pH of the solution was adjusted to the desired level using diluted NaOH and HCl. The reusability of the TFFBR with BiOI-7 thin film was evaluated by running it for 12 cycles. Thereafter, the GFC-7 was dried and kept in the dark at room temperature for 6 months. Then, it was used for a degradation experiment to evaluate the stability of the GFC-7.

233

2. Results and discussion

260 259

2.1. Characterization of the BiOI thin film

261

192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211

O

R O

191

P

190

D

189

E

188

T

187

C

186

E

185

R

184

O R

183

N C

179

F

for a direct transition and n = 4 for an indirect transition. For 212 BiOI, n was 4 (Zhang et al., 2018a). 213

180

named as BiOI-7, and was subjected to further characterization. The coated GFCs were cut into pieces of 3 × 15 cm and 25 × 30 cm to investigate their photocatalytic activity in the photochemical reactor.

178

U

177

Fig. 1 – Set-up of the thin-film fixed-bed reactor (TFFBR).

216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231

234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258

As shown in Fig. 2, the XRD patterns of pure BiOI and BiOI-7 262 were similar to each other, and were well indexed to the pure 263 phase of tetragonal BiOI (JCPDS file No. 10-0445) with good 264

Please cite this article as: Y. Zhang, G. Shan, F. Dong, et al., Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination unde..., Journal of Environmental Sciences, https://doi.org/10.1016/j.jes.2019.01.004

4

JOUR N AL OF EN VI RON M EN T AL S CI E NC ES XX ( XXX X) X XX

P

D

T C E R R

269

O

268

C

267

N

266

crystallinity (Liu et al., 2012; Zhang et al., 2018a). This suggested that mixing with SiO2 sol during the thin-film preparation did not evidently affect the crystal phase of BiOI. The broad diffraction peak of SiO2 at ca. 23o was not detected due to its low weight ratio (< 3.0%) (Yuan et al., 2016).

U

265

E

Fig. 2 – X-ray diffraction (XRD) patterns of BiOI, BiOI-7 and GFC-7.

R O

O

F

The full scan XPS spectrum demonstrated that there were no elements other than Bi, O, I and C in the pure BiOI powder (Fig. 3a). The weak signal of C 1s was derived from the adventitious hydrocarbon in the XPS instrument itself, reflecting that impurities could be considered negligible in the samples. In the XPS spectra of pure BiOI and BiOI-7, which was the powder scraped from GFC-7, the two peaks of Bi 4f at 164.0 and 159.0 eV (Fig. 3b) were assigned to Bi 4f5/2 and Bi 4f7/2 of Bi(III), respectively (Chang et al., 2013; Gao et al., 2017; Hu et al., 2018) and the peaks with binding energies of I 3d3/2 and I 3d3/5 were at 630.8 and 619.1 eV, respectively (Fig. 3c), confirming the existence of I− (Gao et al., 2017). The two peaks of O1s at 530.2 and 531.7 eV corresponding to the lattice oxygen of Bi-O and hydroxyl groups or O2 adsorbed on the surface of the catalysts, respectively (Fig. 3d) (Wang et al., 2011). The silica sol method introduced Si on the surface of the BiOI powder. Besides the signals attributed by BiOI, a weak signal was detected in the region of Si 2p in BiOI-7 with a binding energy (BE) of 103.4 eV, which corresponded to that of silica sol (Fig. 3a), suggesting the existence of Si–O–Si in BiOI-7 (Yuan et al., 2016). For BiOI-7, the third peak of O1s appearing at 533.3 eV could also be ascribed to Si–O–Si bonds (Gao et al., 2017), which was confirmed by the Si–O signal of silica sol (Fig. 3a). In addition, the EDX spectra of BiOI-7 (Appendix A Fig. S1) also indicated that there was Si in BiOI-7. Fig. 3b illustrates that the BEs of Bi in BiOI-7 shifted to lower binding energies by ca. 0.4–0.7 eV compared to the

Fig. 3 – X-ray photoelectron spectroscopy (XPS) results of BiOI and BiOI-7.

Please cite this article as: Y. Zhang, G. Shan, F. Dong, et al., Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination unde..., Journal of Environmental Sciences, https://doi.org/10.1016/j.jes.2019.01.004

270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296

5

O R

R

E

C

T

E

D

P

R O

O

F

JOUR N AL OF EN VI R ONM E NT AL S CI E N CE S XX ( XXX X) X XX

298 299 300 301 302 303 304 305 306 307 308 309 310 311 312

pure BiOI, implying that a chemical interaction might be formed between the Bi atoms and SiO2. The smaller electronegativity of Si (1.8) than Bi (2.0) might lead to an increase of electron density around Bi atoms. Thus, the negative charge on Bi increased, resulting in a decrease in the BEs of Bi (Yuan et al., 2016). In the FT-IR spectra of both pure BiOI and BiOI-7 (Appendix A Fig. S2), the absorption peak at 3440 cm− 1 was ascribed to the stretching vibration absorption of O–H (Chang et al., 2013). The absorption peaks at 881 and 772 cm− 1 corresponded to the stretching vibrations of Bi–O bonds (Chang et al., 2013). There was an additional peak near 957 cm− 1 in BiOI-7, which was not observed in BiOI. This peak might be attributed to Si–OH groups, further supporting the existence of silica in the thin films (Davis and Liu, 1997; Neumann and Levin-Elad, 1997; Rasalingam et al., 2013).

U

297

N C

Fig. 4 – Field emission scanning electron microscope (FESEM) images of GFC-1 (a), -3 (b), -5 (c) and -7 (d) films.

Agreeing with a previous study (Chang et al., 2013), both pure BiOI and BiOI-7 presented as 3D hierarchical microspheres assembled from nanosheets (Appendix A Fig. S3). The surface morphology of the microspheres immobilized on the GFC was not apparently affected by the presence of silica. The silica present in the sol–gel served as an adhesion agent which glued the BiOI microspheres to the fiber surface or the interstices (Fig. 4a). With the increase in the iterations of the painting/drying process, more and more BiOI catalyst grew on the previously formed layers, forming a dense and bulky film of catalyst on the GFC (Fig. 4b–d). The BiOI catalysts gradually filled the interstices of the glass fiber cloth and the surface became smoother. The loading amounts of BiOI on the GFC were estimated by the weight difference of the GFC after and before loading BiOI, which were 5.6, 21.5 and 42.2 g/m2, respectively, on GFC-3, -5 and -7. Efforts were made to further

Please cite this article as: Y. Zhang, G. Shan, F. Dong, et al., Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination unde..., Journal of Environmental Sciences, https://doi.org/10.1016/j.jes.2019.01.004

313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328

6

JOUR N AL OF EN VI RON M EN T AL S CI E NC ES XX ( XXX X) X XX

b

a

1.0

1.0

0.8

0.4

0.6 0.4

0.2

0.2

0.0

0.0

-60

-40

-20 0 Time (min)

20

-60 -40 -20

40

c

0

20 40 60 Time (min)

d 1.0

0.6

0.6

0.2 0.0

pH 3 pH 5 pH 7 pH 9 pH 11

3rd

4th

5th

0.2

0

20 40 60 Time (min)

80 100 120

0.0

Runs

T

-60 -40 -20

0.4

2nd

D

0.4

Ct /C0

0.8

80 100 120

P

1st 0.8

E

Ct /C0

1.0

Light on

R O

Dark

F

0.6

Blank GFC-3 GFC-5 GFC-7

O

Blank Pure BiOI BiOI-7

Ct /C0

Ct /C0

0.8

333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352

O

332

C

331

increase the thickness of the BiOI thin film. However, if the number of coatings was > 7, an obvious flaking of BiOI powder from the GFC was observed. High resolution transmission electron microscope (HRTEM) (Appendix A Fig. S4) was performed to further illustrate the crystal structure of the as-prepared catalysts. The HRTEM image of BiOI-7 scraped from GFC-7 shows the lattice spacing of 0.3 and 0.289 nm, corresponding to that of (102) and (110) diffraction peaks of BiOI, respectively. This coincides with the XRD result, supporting the supposition that the as-prepared catalysts were BiOI, and that mixing with SiO2 sol did not change its crystal structure. The UV–vis diffuse reflectance spectra (DRS) were obtained to investigate the light absorption of the as-prepared catalysts, and the results are depicted in Appendix A Fig. S5a. Their absorption edges were at around 650 nm, while the intensity of BiOI-7 was slightly greater than that of BiOI in the visible region. As shown in Appendix A Fig. S5b, the band gaps of BiOI and BiOI-7 were estimated to be 1.76 and 1.57 eV, respectively. This reveals that the light absorption ability of BiOI-7 in the visible region was slightly higher than that of pure BiOI, which might be because the small amount of silica present enhanced the transmittance in the visible region (Shen et al., 2015).

N

330

U

329

R R

E

C

Fig. 5 – Photocatalytic degradation of bisphenol A (BPA) in pure water was evaluated on a photochemical reactor with quartz tubes: (a) Degradation of BPA using BiOI and BiOI-7; (b) Degradation of BPA using GFC-3, -5 and -7; (c) Photocatalytic efficiency of GFC-7 under different initial pH conditions. (d) Recycling test for photocatalytic degradation of BPA using GFC-7.

2.2. Photocatalytic properties of the BiOI thin films

353

The photocatalytic activity of the BiOI thin films was evaluated by BPA degradation using the scraped BiOI-7 power from GFC-7 in quartz tubes. Fig. 5a shows that BPA was very stable under solar light irradiation without any photocatalyst. Compared to the pure BiOI, the BiOI-7 sample displayed lower degradation efficiency. The photoluminescence spectra and transient photocurrent results (Appendix A Fig. S6) indicated that the presence of SiO2 depressed the transfer of photogenerated electrons, which might explain the relatively lower degradation efficiency of BiOI7 than pure BiOI. A similar result was reported for the BiOBr@SiO2 catalyst (Wang et al., 2018). Another reason explaining the lower adsorption and weaker photocatalytic ability of BiOI-7 than BiOI was that BiOI-7 had smaller pores than BiOI (Wang et al., 2018). The N2 adsorption–desorption isotherms and pore size distributions of BiOI and BiOI-7 are illustrated in Appendix A Fig. S7. The BET surface area of BiOI-7 (24.8 m2/g) was higher than that of pure BiOI (16.9 m2/g). However, the average pore size of BiOI-7 (10.5 nm) was much smaller than that of pure BiOI (31.8 nm), which could be due to blocking of the pores by SiO2 sol. Although BiOI-7 had relatively lower degradation efficiency than BiOI, it was still able to achieve satisfactory

354

Please cite this article as: Y. Zhang, G. Shan, F. Dong, et al., Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination unde..., Journal of Environmental Sciences, https://doi.org/10.1016/j.jes.2019.01.004

355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375

7

JOUR N AL OF EN VI R ONM E NT AL S CI E N CE S XX ( XXX X) X XX

a

b 0.7

250 mL/min 200 mL/min 180 mL/min 150 mL/min

1.0 0.8 Ct /C0

Ct /C0

0.6

0.1

0.6 0.4 0.2

15° Inclination

0

30°

0.8 0.6 0.4

1.0 0.8 0.6 0.4

D

0.2

400

500

pH 3.8 pH 5.2 pH 7.5 pH 9.5 pH 11.0

P

1.0

Ct /C0

pH 3.8 pH 5.2 pH 7.5 pH 9.5 pH 11.0

200 300 Time (min)

R O

d

c

Ct /C0

100

O



F

0.0 0.0

0.2

0

100

200

300

400

500

0

100

200 300 Time (min)

400

500

T

Time (min)

E

0.0

376

R

E

C

Fig. 6 – Photocatalytic degradation of bisphenol A (BPA) in pure water by the TFFBR applied with GFC-7: (a) at different inclination angles (Q = 150 mL/min, pH 7.5), (b) at different flow rates Qs (β = 15o, pH 7.5) and (c) at different initial pHs (Q = 150 mL/min, 15o); (d) Photocatalytic degradation of artificial sewage containing BPA at different initial pHs (Q = 150 mL/min, β = 15o).

390

2.3. Degradation performance of the TFFBR applied with GFC-7

391

Fig. 6 shows the photocatalytic degradation of BPA in the TFFBR under different working conditions, including the mass flow-rate (Q) and the inclination angle of the sloping plate (β). The fluid film thickness (d), residence time over the thin-film (Rt) and the recycle index of the continuous flow (RI) are important hydrodynamic parameters affecting degradation rates in a continuous flow TFFBR (Bousselmi et al., 2004). The calculation of these parameters is provided in Appendix A.

380 381 382 383 384 385 386 387 388

392 393 394 395 396 397 398

N C

379

U

378

O R

389

performance in the degradation of BPA (Fig. 5b). For the BiOI thin films, the photocatalytic activity increased with increasing numbers of catalyst layers, which could be ascribed to more available BiOI photocatalyst being present on the GFC. In all cases, the removal rate of BPA by the thin films was greater than 90% in 2 hr, and BPA was almost completely degraded by GFC-7. The effect of pH on the photocatalytic performance of GFC7 was investigated (Fig. 5c). BPA could be well degraded with a degradation rate of 99.8% in 2 hr at pH 5–9. Reusability experiments were conducted to investigate the films' lifetime. Fig. 5d illustrates that there was no significant decrease of photocatalytic efficiency after 5 cycles under the same reaction conditions, indicating that the as-prepared BiOI thin films were very stable for reuse.

377

The β of the plate was found to be a vital parameter to affect the degradation efficiency (Fig. 6a). At pH of 7.5 and Q of 150 mL/L, the degradation rate (98%) at the β of 15o was significantly higher than those at 5o (23%) and 30o (40%). At β of 30o, the fluid film became thinner, with the d value decreasing from 0.22 to 0.18 mm, which should be beneficial for degradation since a thinner fluid film favors the mass transfer of BPA (Appendix A Table S2). However, the Rt decreased by 20% at the β of 30o compared to the value at 15o. The shorter Rt meant that BPA in the solution did not have enough time to react with the produced reactive oxygen species (ROS), leading to lower degradation efficiency. The longer Rt (by 43%) at the β of 5o compared to that at 15o should favor the degradation. Thus, the reduced degradation efficiency could be attributed to the increase in d (from 0.22 to 0.32 mm) (Appendix A Table S2). In a thicker fluid, the mass transfer of BPA in the solid–liquid interface became more difficult, making BPA unable to efficiently contact the produced ROS, then reducing the degradation efficiency. Thus, the β of 15o was selected as the optimum inclination angle for subsequent experiments. Fig. 6b indicates the degradation rates at pH of 7.5, β of 15o and Qs of 150–250 mL/min. Under this condition, the Qs did not affect the degradation rate distinctly, which was in

Please cite this article as: Y. Zhang, G. Shan, F. Dong, et al., Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination unde..., Journal of Environmental Sciences, https://doi.org/10.1016/j.jes.2019.01.004

399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423

8

438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457

462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480

Appendix A. Supplementary data

490 489

F

437

C

436

E

435

R R

434

O

433

C

432

N

431

U

430

484 485 486 487 488 Q4

Supplementary data to this article can be found online at 491 https://doi.org/10.1016/j.jes.2019.01.004. 492

O

A novel photocatalytic thin film, consisting of highly active BiOI nanoparticles immobilized on GFC using silica sol, was conveniently prepared and applied on a TFFBR for degradation of BPA. The photocatalytic degradation efficiency of the films increased with the number of coatings, and the sample with seven layers exhibited the highest photocatalytic performance. The as-prepared BiOI thin films were very durable and maintained high performance in a wide range of pH (pH 5–9) after running for many degradation cycles and after long-term storage. The GFC-7 applied on the TFFBR could degrade 98% of 2 L of BPA at 20 mg/L within 8 hr irradiation. The degradation efficiency was distinctly affected by the slope of the plate in the TFFBR. Although the degradation rate was relatively lower for artificial sewage, an 83% removal rate could be achieved in 8 hr irradiation. At the end of reaction, no separation process is necessary to remove the photocatalyst from the solution. Thus, the TFFBR based on glass fiber-supported BiOI films shows good prospects for practical wastewater treatment.

429

483

R O

461

428

This work was supported by the National Science Foundation of China (Nos. 21737003, 21677081), the Tianjin Municipal Science and Technology Commission (Nos. 16PTSYJC00020, 17JCYBJC23200), the Ministry of Science and Technology (No. 2017ZX07301002), and the Yangtze River scholar program, and 111 program, Ministry of Education, China (No. T2017002).

REFERENCES

493

Abdel-Maksoud, Y., Imam, E., Ramadan, A., 2016. TiO2 solar photocatalytic reactor systems: selection of reactor design for scale-up and commercialization-analytical review. Catalysts 6, 138–164. Bousselmi, L., Geissen, S.U., Schroeder, H., 2004. Textile wastewater treatment and reuse by solar catalysis: results from a pilot plant in Tunisia. Water Sci. Technol. 49, 331–337. Braham, R.J., Harris, A.T., 2009. Review of major design and scaleup considerations for solar photocatalytic reactors. Ind. Eng. Chem. Res. 48, 8890–8905. Chang, C., Zhu, L., Fu, Y., Chu, X., 2013. 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. 233, 305–314. Chen, G., Miao, S., Tam, N.Y., Wong, Y.S., Li, S., Lan, C., 1995. Effect of synthethic waster on young kandelia candel plants growing under greenhouse conditions. Hydrobiologia 295, 263–273. Cheng, H., Huang, B., Dai, Y., 2014. Engineering BiOX (X = Cl, Br, I) nanostructures for highly efficient photocatalytic applications. Nanoscale 6, 2009–2026. Chiou, C.S., Shie, J.L., Chang, C., Liu, C., Chang, C., 2006. Degradation of di-n-butyl phthalate using photoreactor packed with TiO2 immobilized on glass beads. J. Hazard. Mater. 137, 1123–1129. Cruz-Ortiz, B.R., Garcia-Lobato, M.A., Larios-Duran, E.R., MuzquizRamos, E.M., Ballesteros-Pacheco, J.C., 2016. Potentiostatic electrodeposition of nanostructured NiO thin films for their application as electrocatalyst. J. Electroanal. Chem. 772, 38–45. Cui, S., Shan, G., Zhu, L., 2017. Solvothermal synthesis of Ideficient BiOI thin film with distinct photocatalytic activity and durability under simulated sunlight. Appl. Catal. B Environ. 219, 249–258. Davis, R.J., Liu, Z., 1997. Titania-silica: a model binary oxide catalyst system. Chem. Mater. 9, 2311–2324. Erjavec, B., Hudoklin, P., Perc, K., Tisler, T., Dolenc, M.S., Pintar, A., 2016. Glass fiber-supported TiO2 photocatalyst: efficient mineralization and removal of toxicity/estrogenicity of bisphenol A and its analogs. Appl. Catal. B Environ. 183, 149–158. Gao, S., Guo, C., Lv, J., Wang, Q., Zhang, Y., Hou, S., et al., 2017. A novel 3D hollow magnetic Fe3O4/BiOI heterojunction with enhanced photocatalytic performance for bisphenol A degradation. Chem. Eng. J. 307, 1055–1065. Hachisu, T., Shi, K., Yokoshima, T., Sugiyama, A., Kuroiwa, S., Osaka, T., et al., 2016. Preparation of anatase phase titanium dioxide film by non-aqueous electrodeposition. Electrochem. Commun. 65, 5–8.

494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540

P

3. Conclusions

427

482 481

D

460 459

426

Acknowledgments

T

458

the range of 95%–98% within 8 hr irradiation. This might be due to the integrated effects of the three hydrodynamic parameters: d, Rt and RI. As the Q increased, d and RI would increase while Rt decreased (Bousselmi et al., 2004), as shown in Appendix A Table S2. Greater RI should favor the reaction, whereas shorter Rt and greater d should be against it. Due to the combined results, the degradation efficiency within a fixed time did not change much. At both 150 and 250 mL/min, the degradation rate was ca. 98%. Considering that higher flow rate would consume more electric energy, 150 mL/min and β of 15o were chosen for the following experiments. In agreement with the tests in quartz tubes using BiOI thin films, the degradation efficiency of the TFFBR reactor with GFC-7 was the highest at pH 7.5 (Fig. 6c). The reusability test for the TFFBR showed that after 12 runs, the removal rate of BPA only slightly decreased to 96%. After a 6-month storage, the film in the TFFBR still displayed satisfactory degradation efficiency (>95%). To further assess the applicability of the TFFBR, artificial sewage containing BPA was prepared and the degradation of BPA was tested. About 83% of BPA in the artificial sewage was degraded at pH of 7.5 within 8 hr (Fig. 6d), which was lower than the value in pure water. This may be due to the presence of inorganic ions, which were adsorbed on the thin films and depressed the photocatalytic degradation (Sudrajat and Babel, 2016). Some anions such as NO−3, which has strong light absorption at a wavelength of < 300 nm, may compete with the catalyst for the irradiation, thus reducing the photocatalytic degradation as well (Zhang et al., 2005). Previous studies reported that cations such as Cu2+ and Ni2+ inhibited the photocatalytic rate by trapping the electron-hole pairs (Hu et al., 2003). Further study is warranted to understand the impacts of the components in artificial sewage on the degradation efficiency.

425

E

424

JOUR N AL OF EN VI RON M EN T AL S CI E NC ES XX ( XXX X) X XX

Please cite this article as: Y. Zhang, G. Shan, F. Dong, et al., Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination unde..., Journal of Environmental Sciences, https://doi.org/10.1016/j.jes.2019.01.004

9

JOUR N AL OF EN VI R ONM E NT AL S CI E N CE S XX ( XXX X) X XX

N C

O R

R

E

C

D

P

R O

O

F

template assisted sol–gel synthesis and photocatalytic activity. Appl. Surf. Sci. 393, 46–59. Momeni, M., Saghafian, H., Golestani-Fard, F., Barati, N., Khanahmadi, A., 2017. Effect of SiO2 addition on photocatalytic activity, water contact angle and mechanical stability of visible light activated TiO2 thin films applied on stainless steel by a sol gel method. Appl. Surf. Sci. 392, 80–87. Neumann, R., Levin-Elad, M., 1997. Metal Oxide (TiO2, MoO3, WO3) substituted silicate xerogels as catalysts for the oxidation of hydrocarbons with hydrogen peroxide. J. Catal. 166, 206–217. Rasalingam, S., Kibombo, H.S., Wu, C., Budhi, S., Peng, R., Baltrusaitis, J., et al., 2013. Influence of Ti-O-Si hetero-linkages in the photocatalytic degradation of Rhodamine B. Catal. Commun. 31, 66–70. Rodriguez-Narvaez, O.M., Peralta-Hernandez, J.M., Goonetilleke, A., Bandala, E.R., 2017. Treatment technologies for emerging contaminants in water: a review. Chem. Eng. J. 323, 361–380. Sangkhun, W., Laokiat, L., Tanboonchuy, V., Khamdahsag, P., Grisdanurak, N., 2012. Photocatalytic degradation of BTEX using W-doped TiO2 immobilized on fiberglass cloth under visible light. Superlattices Microst. 52, 632–642. Shen, C., Wang, Y., Xu, J., Luo, G., 2012. Facile synthesis and photocatalytic properties of TiO2 nanoparticles supported on porous glass beads. Chem. Eng. J. 209, 478–485. Shen, F., Zhou, L., Shi, J., Xing, M., Zhang, J., 2015. Preparation and characterization of SiO2/BiOX (X = Cl, Br, I) films with high visible-light activity. RSC Adv. 5, 4918–4925. Starr, B., Tarabara, V., Herrera-Robledo, M., Zhou, M., Roualdes, S., Ayral, A., 2016. Corrigendum to “Coating porous membranes with a photocatalyst: Comparison of LbL self-assembly and plasma-enhanced CVD techniques”[J. Membr. Sci. 514 (2016) 340–349]. J. Membr. Sci. 520, 812–813. Sudrajat, H., Babel, S., 2016. An innovative solar photoactive system N-WO3@polyester fabric for degradation of amaranth in a thin-film fixed-bed reactor. Sol. Energy Mater. Sol. Cells 149, 294–303. Tugaoen, H.O., Garcia-Segura, S., Hristovski, K., Westerhoff, P., 2018. Compact light-emitting diode optical fiber immobilized TiO2 reactor for photocatalytic water treatment. Sci. Total Environ. 613, 1331–1338. Vella, G., Imoberdorf, G.E., Sclafani, A., Cassano, A.E., Alfano, O.M., Rizzuti, L., 2010. Modeling of a TiO2-coated quartz wool packed bed photocatalytic reactor. Appl. Catal. B Environ. 96, 399–407. Wang, Y., Deng, K., Zhang, L., 2011. Visible light photocatalysis of BiOI and its photocatalytic activity enhancement by in situ ionic liquid modification. J. Phys. Chem. C 115, 14300–14308. Wang, Y., Long, Y., Zhang, D., 2017. Facile in situ growth of high strong BiOI network films on metal wire meshes with photocatalytic activity. ACS Sustain. Chem. Eng. 5, 2454–2462. Wang, D., Hou, P., Yang, P., Cheng, X., 2018. BiOBr@SiO2 flower-like nanospheres chemically-bondedon cement-based materials for photocatalysis. Appl. Surf. Sci. 430, 539–548. Ye, L., Chen, S., 2016. Fabrication and high visible-light-driven photocurrent response of g-C3N4 film: the role of thiourea. Appl. Surf. Sci. 389, 1076–1083. Ye, L., Chen, J., Tian, L., Liu, J., Peng, T., Deng, K., et al., 2013. BiOI thin film via chemical vapor transport: photocatalytic activity, durability, selectivity and mechanism. Appl. Catal. B Environ. 130, 1–7. Ye, L., Su, Y., Jin, X., Xie, H., Zhang, C., 2014. Recent advances in BiOX (X = Cl, Br and I) photocatalysts: synthesis, modification, facet effects and mechanisms. Environ. Sci. Nano 1, 90–112. Yu, J., Ho, W., Lin, J., Yip, K.Y., Wong, P., 2003. Photocatalytic activity, antibacterial effect, and photoinduced hydrophilicity of TiO2 films coated on a stainless steel substrate. Environ. Sci. Technol. 37, 2296–2301. Yuan, L., Han, C., Pagliaro, M., Xu, Y., 2016. Origin of enhancing the photocatalytic performance of TiO2 for artificial

E

T

Hong, H., Liu, Q., Jin, H., 2012. Operational performance of the development of a 15 kW parabolic trough mid-temperature solar receiver/reactor for hydrogen production. Appl. Energy 90, 137–141. Hu, C., Yu, J., Hao, Z., Wong, P., 2003. Effects of acidity and inorganic ions on the photocatalytic degradation of different azo dyes. Appl. Catal. B Environ. 46, 35–47. Hu, T., Yang, Y., Dai, K., Zhang, J., Liang, C., 2018. A novel Zscheme Bi2MoO6/BiOBr photocatalyst for enhanced photocatalytic activity under visible light irradiation. Appl. Surf. Sci. 456, 473–481. Huang, Y., Wong, C., Zheng, J., Bouwman, H., Barra, R., Wahlstrom, B., et al., 2012. Bisphenol A (BPA) in China: a review of sources, environmental levels, and potential human health impacts. Environ. Int. 42, 91–99. Jafarikojour, M., Mohammadi, M.M., Sohrabi, M., Royaee, S.J., 2015. Evaluation and modeling of a newly designed impinging stream photoreactor equipped with a TiO2 coated fiberglass cloth. RSC Adv. 5, 9019–9027. Jia, Y., Yang, Y., Guo, Y., Guo, W., Qin, Q., Yang, X., et al., 2015. Simulated sunlight photocatalytic degradation of aqueous pnitrophenol and bisphenol A in a Pt/BiOBr film-coated quartz fiber photoreactor. Dalton Trans. 44, 9439–9449. Khan, S.J., Reed, R.H., Rasul, M.G., 2012. Thin-film fixed-bed reactor (TFFBR) for solar photocatalytic inactivation of aquaculture pathogen Aeromonas hydrophila. BMC Microbiol. 12, 80–87. Kim, D.H., Koo, H.J., Jur, J.S., Woodroof, M., Kalanyan, B., Lee, K., et al., 2012. Stable anatase TiO2 coating on quartz fibers by atomic layer deposition for photoactive light-scattering in dyesensitized solar cells. Nanoscale 4, 4731–4738. Kong, T., Wei, X., Zhu, G., Huang, Y., 2018. Electronic structure and optical properties of BiOI {001} monolayer under biaxial strain. J. Mater. Sci. 53, 708–715. Krishnan, A.V., Stathis, P., Permuth, S.F., Tokes, L., Feldman, D., 1993. Bisphenol a: an estrogenic substance is released from polycarbonate flasks during autoclaving. Endocrinology 132, 2279–2286. Liu, H., Cao, W., Su, Y., Wang, Y., Wang, X., 2012. Synthesis, characterization and photocatalytic performance of novel visible-light-induced Ag/BiOI. Appl. Catal. B Environ. 111, 271–279. Lopez-Salinas, F.I., Martinez-Castanon, G.A., Martinez-Mendoza, J. R., Ruiz, F., 2010. Synthesis and characterization of nanostructured powders of Bi2O3, BiOCl and Bi. Mater. Lett. 64, 1555–1558. Ma, H., Shen, J., Shi, M., Lu, X., Li, Z., Long, Y., et al., 2012. Significant enhanced performance for Rhodamine B, phenol and Cr(VI) removal by Bi2WO6 nancomposites via reduced graphene oxide modification. Appl. Catal. B Environ. 121, 198–205. Malato, S., Fernandez-Ibanez, P., Maldonado, M.I., Blanco, J., Gernjak, W., 2009. Decontamination and disinfection of water by solar photocatalysis: recent overview and trends. Catal. Today 147, 1–59. Malato, S., Maldonado, M.I., Fernandez-Ibanez, P., Oller, I., Polo, I., Sanchez-Moreno, R., 2016. Decontamination and disinfection of water by solar photocatalysis: the pilot plants of the plataforma solar de Almeria. Mat. Sci. Semicon. Proc. 42, 15–23. McCullagh, C., Skillen, N., Adams, M., Robertson, P.J., 2011. Photocatalytic reactors for environmental remediation: a review. J. Chem. Technol. Biotechnol. 86, 1002–1017. Mehrjouei, M., Mueller, S., Moeller, D., 2013. Design and characterization of a multi-phase annular falling-film reactor for water treatment using advanced oxidation processes. J. Environ. Manag. 120, 68–74. Mohamed, M.A., Salleh, W.W., Jaafar, J., Rosmi, M.S., Hir, Z.M., Abd Mutalib, M., Tanemura, M., et al., 2017. Carbon as amorphous shell and interstitial dopant in mesoporous rutile TiO2: bio-

U

541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609

Please cite this article as: Y. Zhang, G. Shan, F. Dong, et al., Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination unde..., Journal of Environmental Sciences, https://doi.org/10.1016/j.jes.2019.01.004

610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678

10

photoreduction of CO2 through a SiO2 coating strategy. J. Phys. Chem. C 120, 265–273. Zayani, G., Bousselmi, L., Mhenni, F., Ghrabi, A., 2009. Solar photocatalytic degradation of commercial textile azo dyes: performance of pilot plant scale thin film fixed-bed reactor. Desalination 246, 344–352. Zazhigalov, S., Elyshev, A., Lopatin, S., Larina, T., Cherepanova, S., Mikenin, P., et al., 2017. Copper-chromite glass fiber catalyst and its performance in the test reaction of deep oxidation of toluene in air. React. Kinet. Mech. Catal. 120, 247–260. Zhang, W., An, T., Cui, M., Sheng, G., Fu, J., 2005. Effects of anions on the photocatalytic and photoelectrocatalytic degradation of reactive dye in a packed-bed reactor. J. Chem. Technol. Biotechnol. 80, 223–229.

Zhang, L., Wang, W., Zhou, L., Shang, M., Sun, S., 2009. Fe3O4 coupled BiOCl: a highly efficient magnetic photocatalyst. Appl. Catal. B Environ. 90, 458–462. Zhang, J., Fu, J., Wang, Z., Cheng, B., Dai, K., Ho, W., 2018a. Direct Zscheme porous g-C3N4/BiOI heterojunction for enhanced visible-light photocatalytic activity. J. Alloy. Compd. 766, 841–850. Zhang, J., Lv, J., Dai, K., Liang, C., Liu, Q., 2018b. One-step growth of nanosheet-assembled BiOCl/BiOBr microspheres for highly efficient visible photocatalytic performance. Appl. Surf. Sci. 430, 639–646. Zhou, P., Le, Z., Xie, Y., Fang, J., Xu, J., 2017. Studies on facile synthesis and properties of mesoporous CdS/TiO2 composite for photocatalysis applications. J. Alloy. Compd. 692, 170–177.

F

679 680 681 682 683 684 685 686 687 688 689 690 691 692

JOUR N AL OF EN VI RON M EN T AL S CI E NC ES XX ( XXX X) X XX

U

N

C

O

R R

E

C

T

E

D

P

R O

O

708

Please cite this article as: Y. Zhang, G. Shan, F. Dong, et al., Glass fiber supported BiOI thin-film fixed-bed photocatalytic reactor for water decontamination unde..., Journal of Environmental Sciences, https://doi.org/10.1016/j.jes.2019.01.004

693 694 695 696 697 698 699 700 701 702 703 704 705 706 707