Efficient photocatalytic fuel cell via simultaneous visible

3 days ago - Photocatalytic fuel cell (PFC) has been proved effective for generating electricity and degrading pollutants aiming at resolving environm...
0 downloads 0 Views 1MB Size
Subscriber access provided by WEBSTER UNIV

Remediation and Control Technologies

Efficient photocatalytic fuel cell via simultaneous visible-photoelectrocatalytic degradation and electricity generation on a porous coral-like WO3/W photoelectrode Donglai Pan, Shuning Xiao, Xiaofeng Chen, Ruping Li, Yingnan Cao, Dieqing Zhang, Sisi Pu, Zhangcheng Li, Guisheng Li, and Hexing Li Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b05685 • Publication Date (Web): 28 Feb 2019 Downloaded from http://pubs.acs.org on March 2, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 32

Environmental Science & Technology

ACS Paragon Plus Environment

Environmental Science & Technology

1

Efficient photocatalytic fuel cell via simultaneous visible-

2

photoelectrocatalytic degradation and electricity generation on a

3

porous coral-like WO3/W photoelectrode

4

Donglai Pana, Shuning Xiaob, Xiaofeng Chena, Ruping Lia, Yingnan Caoa, Dieqing

5

Zhang a,*, Sisi Pua, Zhangcheng Lia, Guisheng Lia,*, Hexing Lia,c,*

6

a. Key Laboratory of Resource Chemistry of Ministry of Education, Shanghai Key

7

Laboratory of Rare Earth Functional Materials, College of Life and Environmental

8

Science, Shanghai Normal University, Shanghai, 200234, China. E-mail:

9

[email protected], [email protected]. [email protected]

10

b. International Collaborative Laboratory of 2D Materials for Optoelectronics Science

11

and Technology of Ministry of Education, College of Optoelectronic Engineering,

12

Shenzhen University, Shenzhen 518060, China.

13 14

c. Shanghai University of Electric Power, 2588 Changyang Rd., Shanghai 200090, China

15 16

ABSTRACT: Photocatalytic fuel cells (PFCs) have proven to be effective for

17

generating electricity and degrading pollutants with a goal to resolve environmental and

18

energy problems. However, the degradation of persistent organic pollutants (POPs),

19

such as perfluorooctanoic acid (PFOA), remains challenging. In the present work, a

20

porous coral-like WO3/W (PCW) photoelectrode with a well-designed energy band

21

structure was used for the photoelectrocatalytic degradation of POPs and the

22

simultaneous generation of electricity. The as-constructed bionic porous coral-like

23

nanostructure greatly improved the light-harvesting capacity of the PCW

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32

Environmental Science & Technology

24

photoelectrode. A maximum photocurrent density (0.31 mA/cm2) under visible light (λ

25

> 420 nm) irradiation and a high incident photon conversion efficiency (IPCE) value

26

(5.72 % at 420 nm) were achieved. Owing to the unique porous coral-like structure, the

27

suitable energy band position, and the strong oxidation ability, this PCW

28

photoelectrode-based PFC system exhibited a strong ability for simultaneous

29

photoelectrocatalytic degradation of PFOA and electricity generation under visible

30

light irradiation, with a power output of 0.0013 mV/cm2 using PFOA as the fuel. This

31

work provides a promising way to construct a reliable PFC using highly toxic POPs to

32

generate electricity.

33

Introduction

34

Persistent organic pollutants (POPs) are harmful to the natural environment and

35

human health.1-7 Perfluorooctanoic acid (PFOA) is recognized as the most stable and

36

toxic POP in nature due to its strong C-F chemical bond with an energy of 485 kJ/mol.8

37

This strong chemical bond means that the PFOA molecules are difficult to break down

38

by traditional methods, such as chemical oxidation, reverse osmosis and

39

nanofiltration.9-11 Recently, it was reported that PFOA could be degraded by

40

photocatalysis with the assistance of the persulfate ion (S2O82-) or the ferric ion as the

41

Fenton reagent.12-16 However, the applied photocatalytic process can only be driven by

42

ultraviolet light and suffers from low efficiency due to the high recombination rate of

43

photoelectron-hole pairs.17-21

44

Compared to photocatalysis, photoelectrocatalysis (PEC) has been used as an effective

45

way to remove organic pollutants and produce H2 due to the rapid separation of

46

photoelectrons from holes under the bias voltage.22-35 The photogenerated electrons can

47

be rapidly transferred to the cathode to form a photocurrent or drive reduction reactions

ACS Paragon Plus Environment

Environmental Science & Technology

48

such as H2 generation and CO2 reduction, 36-41 and the holes left on the anode are used

49

for oxidative degradation of POPs.42-45 Thus, a new photocatalytic fuel cell (PFC) based

50

on the entire process can be constructed for simultaneous electricity generation and

51

pollutant degradation.46-49

52

To construct a PFC that can decompose PFOA with a strong C-F bond, a highly

53

positive valence band (VB) position for the photoanodic semiconductor is needed for a

54

strong oxidizing capacity. A narrow band gap is also required for efficient utilization

55

of longer wavelength light in sunlight. Considering these two aspects, tungsten trioxide

56

(WO3) is selected as an ideal candidate as an n-type semiconductor.27, 50-57

57

Corals living under the sea have a unique dendrite morphology and a high sunlight

58

utilization efficiency. Inspired by this ability, porous coral-like structured

59

nanomaterials are desired to gain better light-harvesting properties. In this work, a

60

WO3/W photoelectrode with a porous coral-like nanostructure was prepared, and it

61

displayed high photoelectrochemical properties and an enhanced performance for PEC

62

degradation of POPs such as PFOA. The conduction band (CB) position of WO3 is

63

above -0.33 V, and the highly positive CB prevents the photogenerated electrons from

64

being trapped by O2, meaning that more electrons can be used to generate electricity.

65

Therefore, WO3 served as the photoanode in the PEC system to construct the PFC

66

(Scheme 1).

67 68

Scheme 1. Illustration of a photocatalytic fuel cell.

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32

Environmental Science & Technology

69

Experiment

70

Preparation of samples: Porous coral-like WO3/W photoelectrodes (PCW) were

71

prepared by constant potential electrolysis in a two-electrode system, in which a

72

polished tungsten foil (33 mm × 20 mm, 0.2 mm thick) and a platinum foil (20 mm ×

73

20 mm) served as the anode and cathode, respectively. A 50 mL aqueous solution

74

containing 0.2 mol/L NH4F and 0.1 mol/L H3PO4 was used as the electrolyte. The

75

process of anodic oxidation was performed under a controlled constant voltage from

76

DC 30 V to DC 80 V at 0 °C for 2.5 h. The as-obtained electrodes were then air-calcined

77

at 350, 450, 550 and 650°C for 3 h. The calcined electrodes were then named PCW

78

350, PCW 450, PCW 550 and PCW 650, respectively. Commercial WO3 (J&K Co. Ltd.)

79

loaded on W foil was also fabricated through the doctor-blade method reported

80

previously58 and then calcined at 550°C for 3 h. The resulting electrode was named

81

CW. The preparation of WNT, CdS NWs and BiVO4 photoelectrodes could be found

82

in the supporting information (SI Section S1-S3).

83

Characterization: X-ray diffraction (XRD) patterns were collected in parallel mode

84

(2θ from 20 to 40°) using a Rigaku Dmax-3C Advance X-ray diffractometer (Cu Kα

85

radiation, λ= 1.5406 Å). The morphologies were studied through field emission

86

scanning electron microscopy (FESEM, HITACHI S-4800) and transmission electron

87

microscopy (TEM, JEOL, 2010F). Diffuse reflectance spectra over a range of 200-800

88

nm were obtained through a UV-vis spectrophotometer (DRS, MC-2530) equipped

89

with an integrating sphere assembly using BaSO4 as a reference. The electron

90

paramagnetic resonance (EPR) measurements were conducted on a Bruker EMX-8/2.7

91

Electro-Spin Resonance Spectrometer (X-band).

ACS Paragon Plus Environment

Environmental Science & Technology

92

Photoelectrochemical measurement: All of the photoelectrochemical measurements

93

were performed on an electrochemical workstation (CHI660E). A standard three-

94

electrode setup was established using porous coral-like WO3/W (PCW, 4 cm2) as the

95

working electrode, a commercial platinum foil (99.99 %, 4 cm2) as the counter electrode,

96

and a Ag/AgCl electrode as the reference electrode. A 300 W Xenon lamp was used as

97

the visible light source with a cut off of λ < 420 nm. The photocurrent density was

98

detected in an aqueous solution containing 0.5 mol/L Na2SO4 by linear sweep

99

voltammetry (LSV). Electrochemical impedance spectroscopy (EIS) analyses were

100

recorded in the same solution at open-circuit voltage over the frequency range from 106

101

to 10-2 Hz with an AC voltage of 5 mV. The Motto-Schottky plots were obtained at a

102

fixed frequency of 1 kHz to determine the flat-band potential. The electrochemical

103

surface areas of different photoelectrodes were obtained by cyclic voltammetry (CV)

104

measurements in different scan rates to estimate electrochemical double layer (SI

105

Section S4). The incident photon conversion efficiency (IPCE) measurements were

106

carried out on the electrochemical workstation (ZAHNER-elektrik) at a potential of 0.5

107

V (vs. Ag/AgCl) under irradiation with λ ranging from 367 to 720 nm.

108

Photoelectrocatalytic activity test: A H-cell was used as the reactor for PEC

109

degradation of organic pollutants. In the anode, a 0.5 mol/L Na2SO4 aqueous solution

110

containing 5 ppm methyl orange (MO), rhodamine B (RhB), methylene blue (MB), 4-

111

chlorophenol (4-CP), bisphenol A (BPA) or PFOA was used as the electrolyte. An

112

aqueous solution containing 0.5 mol/L Na2SO4 was used in the cathode. The H-cell was

113

separated by a proton exchange membrane (Nafion 117). Before light irradiation, the

114

working electrode was immersed in the reaction solution in the dark for 20 min to reach

115

adsorption-desorption equilibrium. The cathode chamber electrolyte was purged with

116

nitrogen (99.999%) to remove dissolved O2. During the photoelectrocatalytic process,

ACS Paragon Plus Environment

Page 6 of 32

Page 7 of 32

Environmental Science & Technology

117

a constant potential (0.5 V vs. Ag/AgCl) was applied on the working electrode with a

118

constant magnetic stirring rate (600 rpm). A 300 W Xenon lamp was used as the visible

119

light source with a cut off of λ < 420 nm. All reactions lasted for 2 h. For comparison,

120

both the electrocatalytic (EC) degradation and the photocatalytic (PC) degradation of

121

organic pollutants were performed under the same conditions without light irradiation

122

or bias voltage, respectively. The concentration of MO, RhB, MB or 4-CP was

123

determined by measuring the UV absorbance at the characteristic wavelengths of 465,

124

564, 665 and 275 nm, respectively. The concentration of BPA was determined using an

125

HPLC system (UltiMate 3000 Thermo Fisher, phenyl column at 25 °C, a 1:1 v/v

126

water/methanol solution as the mobile phase, flow rate = 1.0 mL/min, detection

127

wavelength = 230 nm). For the degradation of PFOA, the resulting F- concentration

128

was determined using ion chromatography (DIONEX ICS-5000 Thermo Scientific),

129

and the degradation yield was calculated based on the following equation:16

130

PFOA defluorination ratio = CF-/(C0 × 15)

131

where CF- is the concentration of fluoride ion and C0 is the initial concentration of

132

PFOA. The factor of 15 is ascribed to the number of fluorine atoms in each PFOA

133

molecule. Similarly, the 4-CP degradation yield could also be calculated based on the

134

resulting F- concentration according to the following equation:

135

4-CP dechlorination ratio = CCl-/C0

136

In this equation, we used a factor of 1 instead of 15 since each 4-CP molecule contains

137

only one chlorine atom.

138

Photocatalytic fuel cell test: The PFC efficiency was tested in an electrochemical

139

single cell with a two-electrode system using an electrochemical workstation

ACS Paragon Plus Environment

Environmental Science & Technology

140

(CHI660E). The as-prepared PCW electrode was used as the working electrode, and

141

the Pt foil was used as the auxiliary and reference electrode. Both electrodes possessed

142

the same working area of 4 cm2. The distance between those two electrodes was fixed

143

at approximately 4 cm. A 300 W Xenon lamp was used as the visible light source by

144

cutting off all the lights with λ < 420 nm. Various organic pollutants (MO, RhB, MB,

145

4-CP, BPA, and PFOA) and methanol were used as fuel resources, respectively. An

146

aqueous solution containing 0.5 mol/L Na2SO4 served as the electrolyte. The current-

147

voltage (J-V) plots were recorded to characterize the PFC performance with a scan rate

148

of 1 mV/s.

149

Results and discussion

150

The oxidation voltage played a key role in fabricating the PCW photoelectrode

151

through the anodic oxidation method. Under a low oxidation voltage of DC 30 V, WO3

152

growing on the W foil displayed crystal nanoparticles (Figure 1a). When the anodic

153

voltage increased to DC 50 V, a porous coral-like nanostructured WO3 framework was

154

obtained (Figure 1b). Further increasing the anodic oxidation voltage resulted in the

155

collapse of the porous structure and the formation of large bumps with high thickness

156

(Figure 1c). High-resolution transmission electron microscopy (HRTEM) of PCW 550

157

showed two characteristic lattice spacings of 0.383 and 0.365 nm ascribed to the (002)

158

and the (200) lattice planes, respectively, of monoclinic WO3 (Figure 1d). The XRD

159

patterns (Figure S1a) also confirmed that the WO3 crystallization degree increased with

160

increasing anodic voltage. The UV-vis diffuse reflectance spectroscopy (DRS) results

161

(Figure S1b) demonstrated that the PCW photoelectrode prepared at 50 V DC exhibited

162

the strongest light absorption intensity from 200 to 800 nm. Fixing the DC voltage at

163

50 V, the porous coral-like WO3/W photoelectrode was easily obtained, and the

ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32

Environmental Science & Technology

164

thickness of the WO3 layers could be adjusted by increasing the anodic oxidation time

165

from 0.5 to 2.5 h, as determined by the SEM images (Figure S2) and XRD patterns

166

(Figure S3).

167 168

Figure 1. SEM images of WO3/W photoelectrodes calcined at 550 °C for 3 h and

169

treated with the anodic oxidation voltage of (a) 30 V, (b) 50 V and (c) 80 V; (d) HRTEM

170

image of PCW.

171

The SEM images of PCW 550 (Figure S4) revealed that the porous coral-like

172

morphology was preserved after being calcined for 3 h at temperatures below 550°C.

173

However, the porous coral-like nanostructure of WO3/W collapsed after being calcined

174

at 650°C, which occurred along with the serious aggregation of the WO3 pore wall. As

175

shown in Figure 2a, the PCW 550 calcined at 550°C for 3 h exhibited the strongest

176

visible light absorption compared with those obtained at other calcination temperatures

177

from 350 to 650 °C. The cross-section SEM image (Figure S5) confirmed that the

178

porous coral-like WO3 was directly grown on the metal tungsten substrate with a

179

thickness of 500 nm. These results demonstrated that the WO3 layer strongly interacted

180

with the tungsten substrate. This strong interaction could reduce the Schottky contact

181

between the WO3 and the W foil.59, 60 As shown in Figure 2b, no significant diffraction

182

peaks in the XRD patterns were found that indicated the presence of the crystal WO3

183

phase on the PCW photoelectrode before calcination. However, all the PCW

184

photoelectrodes calcined from 350 to 650 °C showed a stable monoclinic structure

ACS Paragon Plus Environment

Environmental Science & Technology

185

(PDF#43-1035) of WO361, 62 63, and the WO3 crystallization degree increased gradually

186

with increasing calcination temperature.64

187 188

Figure 2. (a) UV-vis DRS spectra and (b) XRD patterns of PCW calcined at different

189

temperatures.

190

Figure 3a shows the photocurrent density-potential curves of different PCW

191

photoelectrodes under visible light irradiation. PCW 550 exhibited the highest

192

photocurrent density of 0.31 mA/cm2 compared to PCW 350, PCW 450, PCW 650 and

193

CW, with corresponding photocurrent densities of approximately 0.20, 0.18, 0.12 and

194

0.08 mA/cm2 at 0.5 V (vs. Ag/AgCl). With increasing applied potential, the

195

photocurrent density first increased and then gradually stabilized with densities of 0.38,

196

0.57, 0.37 and 0.35 mA/cm2 for PCW 650, 550, 450 and 350 at 1.2 V (vs. Ag/AgCl),

197

respectively. PCW 550 also exhibited the highest IPCE under light irradiation with

198

wavelengths from 367 to 720 nm, and this result could be ascribed to the unique porous

199

structure favorable for light harvesting (Figure 3b). Moreover, the EIS spectra in Figure

200

3c revealed that PCW 550 displayed the smallest impedance arc diameter,

201

corresponding to the lowest charge-transfer impedance.65 In other words, PCW 550

ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32

Environmental Science & Technology

202

exhibited the fastest rate of electron transfer, which could efficiently reduce the

203

electron-hole recombination rate.

204 205

Figure 3. Photocurrent density-potential curves (a) IPCE measurements (b) and EIS

206

Nyquist plots (c) of all the PCW photoelectrodes calcined at different temperatures.

207

The PEC activity was evaluated by the degradation of MO in an aqueous solution

208

upon reaching an adsorption-desorption equilibrium in 20 min (Figure S6). As shown

209

in Figure 4a, PCW 550 exhibited the highest MO degradation rate of ~68.0 % under

210

visible light irradiation for 2 h, which is nearly 2 times higher than that of the CW

211

electrode. Compared to the PCW electrodes calcined below 550 °C, PCW 550

212

presented a higher crystallization degree of WO3, which can facilitate electron transfer

213

(see Figure 3c) and thus diminish electron-hole recombination. PCW 650, with the

ACS Paragon Plus Environment

Environmental Science & Technology

214

highest crystallization degree of WO3, exhibited a lower PEC activity than that of PCW

215

550, and this result can be ascribed to the collapse of the nanostructure and the reduced

216

light absorption (see Figure 3a). In addition, the aggregation of WO3 also increased the

217

impedance (see Figure 3c), which was unfavorable for electron transfer and enhanced

218

electron-hole recombination. For better illustrating the advantages of porous coral-like

219

structure, WO3/W photoelectrodes with other different morphologies were also

220

investigated in PEC degradation process (Figure S7). The results displayed that PCW

221

photoelectrode had a higher MO degradation efficiency than commercial WO3 (CW)

222

and nanotube-like WO3 (WNT) photoelectrodes, which could be contributed to the

223

higher adsorption capacity and surface area of PCW photoelectrode. The surface areas

224

of different samples were further measured based on the electrochemical surface area

225

(ESA) obtained via the electrochemical measurements and calculations. As shown in

226

Figure S8 and Table S1, PCW 550 photoelectrodes displayed a higher surface area than

227

both CW and WNT photoelectrodes. The decrease of electrochemical surface area from

228

PCW 350 to PCW 650 was attributed to the aggregation to the WO3 crystals during the

229

calcination treatment. Nevertheless, PCW 550 still possessed the highest activity owing

230

to its highly crystalline degree of WO3 and higher positive valence band. Therefore,

231

based on these results, it indicated that such unique porous coral-like nano-structure

232

could improve both the light and chemical absorption performances of WO3

233

photoelectrode. Meanwhile, PFOA was electronegative with a zeta potential of -4.83

234

mV (Figure S9), thus negatively-charged PFOA molecules could be easily trapped by

235

the PCW photoanode via electrostatic attraction. The recombination rate of photo-

236

induced charge carriers can be effectively suppressed in the PEC process under the

237

external circuit and bias potential, possibly accounting for the higher pollutant

238

degradation activity of PEC than that of the PC and EC processes (Figure 4b). With the

ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32

Environmental Science & Technology

239

assistance of the bias voltage, the photogenerated electron could be transferred to the

240

Pt cathode for the H2 evolution reaction, leading to enhanced activity. PCW 550

241

displayed the highest H2 evolution amount of 32.1 μmol within 2 h compared to that of

242

PCW 350, PCW 450, PCW 650 and CW, with H2 evolution amounts of 21.0, 29.5, 31.6

243

and 7.7 μmol, respectively (see Figure S10). The H2 evolution trend was also consistent

244

with the trend of the PEC degradation rates of MO, as shown in Figure 4a. The influence

245

of the initial MO concentration on the degradation efficiency and the H2 evolution rate

246

was also investigated during the PEC process (Figure S11, S12). The MO degradation

247

rate and the H2 evolution rate increased upon decreasing the concentration of MO,

248

which was in accordance with the Langmuir-Hinshelwood model.66-68 The total organic

249

carbon (TOC) value showed a similar trend to the trend of MO degradation yield in the

250

PEC process (Table S2). Furthermore, PCW 550 exhibited a high stability without an

251

obvious decrease in activity after 5 cycles (Figure S13). The long-time response of

252

current density were investigated to confirm the stability of PCW 550. As shown in

253

Figure S14, the current density presented no obvious decline even after 40 h, suggesting

254

an excellent stability of PCW 550.

ACS Paragon Plus Environment

Environmental Science & Technology

255 256

Figure 4. (a) The PEC degradation efficiency for a 5 ppm MO solution using different

257

photoelectrodes; (b) the MO degradation efficiency in PEC, PC and EC processes with

258

PCW 550; (c) the active species trapping experiments for the PEC degradation process;

259

and (d) the relationship of Eg and the VB value for the PCW photoelectrodes.

260

Active species trapping experiments were conducted to better understand the

261

degradation mechanism of the PEC process. As shown in Figure 4c, the degradation

262

efficiency of MO was greatly inhibited when 2.5 % methanol (ME) was added as the

263

photogenerated hole (h+) scavenger. Meanwhile, the degradation efficiency also

264

decreased abruptly after adding 2.5 % isopropyl alcohol (IPA) as the hydroxyl radical

265

(HO•) scavenger. We then concluded that the HO• radical acted as the major active

266

species during the PEC degradation process based on these results. The photogenerated

267

electrons were transferred to the counter electrode under the bias voltage, which could

268

inhibit the formation of •O2- for further oxidizing organic molecules on the photoanode.

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32

Environmental Science & Technology

269

Generally, the generation of HO• was dependent on the position of the VB of the

270

semiconductor. According to the UV-vis DRS results, the band gaps of PCW 350-650

271

were determined to be 2.5, 2.4, 2.4, and 2.3 eV (Figure S15 inset graph). Based on the

272

Motto-Schottky plots (Figure S15), the positions of the CBs of the PCW 350-650

273

samples were calculated to be approximately 0.50, 0.48, 0.40 and 0.32 V (vs. NHE).

274

Thus, the VB values of PCW 350-650 could be estimated as 3.0, 2.9, 2.8 and 2.6 V (vs.

275

NHE). The band gap (Eg) and the VB position were inversely related to the calcination

276

temperature, as described in Figure 4d. Therefore, a suitable balance of Eg and the VB

277

position could be pursued to realize both enhanced visible light harvesting and

278

oxidation ability. BiVO4 and CdS, as widely used visible-light photocatalysts, were also

279

used as references for treating 5 ppm MO in the PEC system as shown in Figure S7.

280

Both CdS and BiVO4 exhibited a lower MO removal efficiency than PCW 550, due to

281

the lower valence bandposition.69 It further illustrated the advantages of energy band

282

structure for PCW photoelectrode. Among the PCW photoelectrodes, PCW 550

283

exhibited an excellent visible light response, which generated more photoelectrons and

284

holes. The photoelectrons transfer to the counter electrode for H2 production via water

285

splitting (Figure S16). Due to the high oxidation potential, the photogenerated holes

286

left on the photoelectrodes could directly oxidize organic pollutants into CO2 and H2O.

287

In addition, partial holes could also oxidize H2O to ·OH radical (E0(H2O/·OH) = 2.4 V

288

vs. NHE), which could further accelerate the oxidative degradation of organic

289

pollutants.70, 71 The overall PEC reaction process is described as follows:

290

1) WO3 + hv → h+ + e-

291

2) H2O → H+ + OH-

292

3) h+ + OH- → HO•

ACS Paragon Plus Environment

Environmental Science & Technology

293

4) h+ + H2O → HO• + H+

294

5) h+ + organic pollutant → CO2 + H2O

295

6) HO• + organic pollutant → CO2 + H2O

296

The degradation efficiency of organic pollutants on PCW 550 was also evaluated using

297

5 ppm of RhB, MB, 4-CP, BPA and PFOA in the PEC, PC and EC processes (Figure

298

5). Meanwhile, the changes in the TOC value and H2 evolution amount within the 2 h

299

reaction were also detected (Table S3). For MB and RhB, the PEC process showed a

300

greatly improved performance in both the degradation efficiency and the H2 evolution

301

rate compared to the PC and EC processes (Figure 5a and b), and this result could be

302

attributed to the rapid separation of photoelectrons from holes in the presence of a bias

303

voltage. For POPs (4-CP, BPA and PFOA), the degradation performances were similar

304

to the dye-degradation process, as shown in Figure 5c-f and Table S3. The PEC

305

degradation rates of 4-CP and BPA can reach 19.2 % and 57.1 %, respectively. In

306

addition, the dechlorination ratio of 4-CP and the defluorination ratio of PFOA were

307

44.2 % and 3.4 %, respectively, after 2 h of reaction. It is suggested that the degradation

308

of PFOA in a PEC system using a PCW photoelectrode could be achieved under visible

309

light irradiation.72-74 The kinetic study of the pollutants-degradation process on PCW

310

550 (Figure S17) indicated that the degradation process was a first-order reaction. More

311

experiments including active species trapping, recyclability for treating PFOA solution

312

with various concentrations were studied to investigate the PEC degradation process of

313

PCW 550 photoelectrode. As shown in Figure S18a, the defluorination ratio declined

314

by using ME as the holes scavenger, indicating that the photogenerated holes were the

315

main activity species during the PFOA degradation process. Such slight decrease of

316

PFOA degradation rate in the presence of ME could be ascribed to strong electrostatic

ACS Paragon Plus Environment

Page 16 of 32

Page 17 of 32

Environmental Science & Technology

317

adsorption effect of PFOA molecules on the WO3 electrodes, thus, PFOA was still

318

easily trapped by the electrodes for further degradation. The EPR analysis also

319

presented four feature signals (hfsc αN =αH = 14.95 G) of DMPO-OH adducts in the

320

Na2SO4 electrolyte (Figure S18b), providing a direct evidence for the generation of ·OH

321

radical in the PEC degradation process. PFOA was considered one of the most difficult

322

pollutants to degrade, the high oxidation ability of photogenerated holes for PCW 550

323

could still effectively break the stable C-F bond on the surface of WO3 during the PEC

324

process. The repeated tests also verified that the as-prepared PCW 550 photoelectrode

325

exhibited an excellent stability (Figure S18c). Besides, the PEC performance for

326

degrading PFOA with various concentrations (Figure S18d) suggested that the as-

327

formed PEC system was effective to degrade PFOA with a Langmuir-Hinshelwood

328

model. The above degradation results indicated that the PCW photoelectrode exhibited

329

excellent universality for pollutant degradation and H2 generation in the PEC process

330

under visible light irradiation.

331

ACS Paragon Plus Environment

Environmental Science & Technology

332

Figure 5. The degradation rates for various organic pollutants in the PEC, PC and EC

333

processes: (a) RhB, (b) MB, (c) BPA and (d) 4-CP; dechlorination ratio of 4-CP (e) and

334

defluorination ratio of PFOA (f).

335

During the H2 generation process, the photogenerated electron could be transferred to

336

the cathode through an external circuit to construct a PFC system.75 The current-voltage

337

(J-V) plots and the current-power (J-JV) plots from the PCW electrodes are shown in

338

Figure 6. The open-circuit voltage (VOC), short-circuit current density (JSC) and fill

339

factor (ff) of the cell are listed in Table 1. The ff was calculated by the following

340

equation, which could directly reflect the performance of the PFC system:76

341

ff = J*Vmax/(Jsc*Voc)

342

where J*Vmax is the maximum power density of the cell obtained from the J-JV plots.

343

As shown in Figures 6a and 6b, the JSC was 0.0279 mA/cm2 in the presence of MO

344

using PCW 550 as the photoelectrode under visible light irradiation (λ > 420 nm).

345

Under the same conditions, the JSC values were approximately 0.0242, 0.0279, 0.0157,

346

and 0.0143 mA/cm2 for PCW 350, PCW 450, PCW 650 and CW photoelectrodes,

347

respectively. The ff value reached 0.213 for PCW 550 (Table 1), which was

348

significantly larger than those of the other photoelectrodes. PCW 550 also exhibited the

349

highest JVmax (0.003 mW/cm2) at JSC and VOC values of 0.0279 mA/cm2 and 0.505 V,

350

respectively. These excellent performances for PCW550 were ascribed to the suitable

351

Eg and VB position of PCW 550, which can enhance visible light absorption and

352

oxidation ability (see Figure 4d). In addition, the bionic porous structure of the PCW

353

photoelectrodes also improved the light-harvesting capacity over that of CW, thus

354

improving the PFC performance. However, the power output and the JSC of all the PCW

355

photoelectrodes decreased dramatically in the dark (Figure S19 and Table 2), which

ACS Paragon Plus Environment

Page 18 of 32

Page 19 of 32

Environmental Science & Technology

356

clearly confirmed the characteristics of the PFC. A series of organic molecules were

357

chosen as the fuel to obtain the J-V and J-JV plots to further investigate the PFC

358

performance of the PCW 550-Pt foil system. As shown in Figure 6c, surprisingly, the

359

Voc values in the presence of a POP (4-CP, BPA or PFOA) are higher than that in

360

methanol (0.180 V), which is commercially used in fuel cells. These higher Voc values

361

suggested that the as-formed PCW 550-Pt PFC system may possess a high power output

362

(JVmax) by using POPs as the fuel (Figure 6d). Methanol is much easier to oxidize

363

than POPs, and the Jsc value in a methanol solution should be much higher than those

364

obtained in a POP solution.75 Nevertheless, comparative Jsc values could still be

365

produced by choosing POPs as the fuel in the PCW 550-Pt PFC system (Table 1). Even

366

in a PFOA solution, a JSC of 0.0177 mA/cm2 was maintained, although the strong C-F

367

bond was very difficult to oxidize. These results suggested that the high oxidation

368

ability of PCW 550 could be used to directly oxidize the PFOA molecule to generate

369

electricity. The positive CB position of PCW 550 also prevented the reaction between

370

the photogenerated electrons and O2 to form •O2- (Fig S16); thus, more electrons could

371

be transferred to the external circuit to maintain the Jsc values. Nevertheless, the ff

372

values in the presence of POPs were slightly lower than those in the presence of

373

methanol. These results could be mainly ascribed to the slow oxidization process of

374

POPs under visible light irradiation. The decreased TOC values of different organic

375

pollutant solutions in the PFC process by using PCW 550 (Table S4), indicating that

376

organic pollutants could be used as the fuels to generate electricity during the PFC

377

process. These results confirmed that the as-formed PFC system could allow POPs to

378

be oxidized on the porous coral-like WO3/W photoanode; thus, more electrons could

379

be transferred to the Pt cathode to generate a higher photocurrent under visible light

380

irradiation. Moreover, the current-time plots under different output voltages in 5 ppm

ACS Paragon Plus Environment

Environmental Science & Technology

381

PFOA solution illustrated that the power output increased with the increasing of voltage

382

window in the PFC system (Figure S20a). And the stability of PCW photoelectrode

383

under acidic condition also could be verified by the long-time output tests (Figure S20b),

384

which showed that the PCW photoelectrode could maintain a stabilized power output

385

even in the acidic condition.

386 387

Figure 6. J-V (a) and J-JV (b) plots using different photoelectrodes with 5 ppm MO in

388

the PFC system. J-V (c) and J-JV (d) plots on PCW 550 photoelectrodes with 5 ppm of

389

different organic pollutants or 2.5 % of methanol. All tests were performed under

390

visible light irradiation in an aqueous solution containing 0.5 mol/L Na2SO4.

391

Table 1. Current-voltage characteristics from different photoelectrodes with 5 ppm of

392

different organic pollutants or 2.5 % methanol in a PFC system under visible light

393

irradiation or in the dark.

ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32

Environmental Science & Technology

Photoelectrode Jsc/mA*cm-2

Light source

Fuel

Voc/V

ff

Visible light

No fuel

PCW 550

0.0154

0.375

0.196

Visible light

4-CP

PCW 550

0.0198

0.451

0.156

Visible light

BPA

PCW 550

0.0187

0.441

0.154

Visible light

MB

PCW 550

0.0179

0.454

0.196

Visible light

Methanol

PCW 550

0.0186

0.180

0.252

Visible light

PFOA

PCW 550

0.0177

0.507

0.146

Visible light

MO

CW

0.0143

0.285

0.203

Visible light

MO

PCW 350

0.0242

0.518

0.183

Visible light

MO

PCW 450

0.0279

0.493

0.152

Visible light

MO

PCW 550

0.0279

0.505

0.213

Visible light

MO

PCW 650

0.0157

0.471

0.162

Dark

MO

CW

0.0124

0.282

0.208

Dark

MO

PCW 350

0.0184

0.496

0.192

Dark

MO

PCW 450

0.0252

0.491

0.162

ACS Paragon Plus Environment

Environmental Science & Technology

394

Page 22 of 32

Dark

MO

PCW 550

0.0184

0.504

0.237

Dark

MO

PCW 650

0.0074

0.432

0.165

All tests were performed in an aqueous solution containing 0.5 mol/L Na2SO4.

395 396

In summary, a novel WO3/W photoelectrode was developed to construct a PFC by

397

using various organic pollutants as the fuel under visible light irradiation. The prepared

398

photoelectrode displayed excellent performance for degrading POPs via visible light-

399

driven photoelectrocatalysis and showed a powerful ability to generate electricity due

400

to its unique porous coral-like nano-structure and well-designed energy band structure.

401

This work provides a new approach to fabricate functional photoelectrodes and may

402

offer more opportunities for designing PFC systems.

403

ASSOCIATED CONTENT

404

Supporting Information

405

The Supporting Information is available free of charge on the ACS Publication website at DOI:

406

10.1021/acs.est.

407

The preparation of nanotube-like WO3/W photoelectrode (WNT); The preparation of CdS

408

nanowires (CdS NWs) photoelectrode; The preparation of BiVO4 photoelectrode; The

409

measurement of electrochemical surface area (ESA); The standard potential conversion; Figure

410

with XRD patterns and UV-vis DRS spectra; SEM images of the as-obtained PCW

411

photoelectrodes; XRD patterns of PCW under different anodic oxidation times; SEM images

412

of the as-prepared PCW calcined at different temperatures; The cross-section SEM image of

413

PCW 550; The adsorption-desorption equilibrium for different photoelectrodes in 5 ppm

ACS Paragon Plus Environment

Page 23 of 32

Environmental Science & Technology

414

MO solution; The PEC degradation tests for different photoelectrodes at 0.5 V (vs.

415

Ag/AgCl) under the visible-light irradiation (300 W Xe lamp with a 420 nm cutoff

416

filter), the insect graphs of the morphologies for different photoelectrodes;

417

Electrochemical surface area measurements for different photoelectrodes; Zeta

418

potential of PFOA solution (5 ppm); The amount of H2 evolution during the process of

419

the PEC degradation of 5 ppm MO using different photoelectrodes; PEC degradation

420

efficiency of PCW 550 with different initial concentrations; The amount of H2

421

evolution in the PEC degradation process with different concentrations of MO by using

422

PCW 550; The cyclic measurements for the PCW 550 photoelectrode; The

423

photocurrent stability tests for PCW 550 in a 0.5 M Na2SO4 solution in the absence of

424

organic pollutant; Motto-Schottky plots and UV-vis DRS spectrum for the as-prepared

425

PCW photoelectrode; The proposed mechanism for the PEC pollutant degradation

426

process; The kinetics curves and constants for the PEC degradation process in different

427

pollutant solution using PCW 550 as the photoelectrode; PEC degradation of PFOA

428

solution: (a) The active species trapping experiments using different scavengers; (b)

429

EPR measurements for the DMPO-OH adducts; (c) The recyclability for the PCW 550

430

photoelectrode; (d) PEC degradation for different PFOA concentrations; J-V plots and

431

J-JV plots for both CW and PCW 550; The current-time plots at different output voltage

432

in a 5 ppm PFOA solution (a), The stability of PCW photoelectrode in acidic condition

433

(b); Table with Electrochemical surface areas of different photoelectrodes; The changes

434

of total organic carbon (TOC) for the MO degradation process after 2 h reaction with

435

the different photoelectrodes; The changes of total organic carbon (TOC) and the

436

amounts of H2 evolution for different organic degradation processes after 2 h reaction

437

time with the same photoelectrode; The TOC changes of different pollutants during the

438

PFC process by using PCW 550 as the photoelectrode for 4 h at 0.1 V.

ACS Paragon Plus Environment

Environmental Science & Technology

439

AUTHOR INFORMATION

440

Corresponding Author

441

* [email protected], [email protected], [email protected]

442

Notes

443

The authors declare no competing financial interest.

444

ACKNOWLEDGMENTS

445

This work was supported by the National Natural Science Foundation of China

446

(21207090, 21477079, 21876113, 21261140333, 21876112, 21803038), PCSIRT

447

(IRT1269), the China Postdoctoral Science Foundation (2018M630981) and a scheme

448

administrated by Shanghai Normal University (DXL122, and S30406).

449

REFERENCES

450

(1) Bennett, D. H.; Kastenberg, W. E.; McKone, T. E., General formulation of

451

characteristic time for persistent chemicals in a multimedia environment. Environ. Sci.

452

Technol. 1999, 33, (3), 503-509.

453

(2) Zhang, H.; Wang, Y.; Sun, C.; Yu, M.; Gao, Y.; Wang, T.; Liu, J.; Jiang, G., Levels

454

and distributions of hexachlorobutadiene and three chlorobenzenes in biosolids from

455

wastewater treatment plants and in soils within and surrounding a chemical plant in

456

China. Environ. Sci. Technol. 2014, 48, (3), 1525-1531.

457

(3) Chen, D.; Kannan, K.; Tan, H.; Zheng, Z.; Feng, Y.L.; Wu, Y.; Widelka, M.,

458

Bisphenol analogues other than BPA: environmental occurrence, human exposure, and

459

toxicity-a review. Environ. Sci. Technol. 2016, 50, (11), 5438-5453.

460

(4) Rochester, J. R., Bisphenol A and human health: A review of the literature. Reprod.

461

Toxicol. 2013, 42, 132-155.

462

(5) Lindstrom, A. B.; Strynar, M. J.; Libelo, E. L., Polyfluorinated compounds: Past,

463

Present, and Future. Environ. Sci. Technol. 2011, 45, (19), 7954-7961.

ACS Paragon Plus Environment

Page 24 of 32

Page 25 of 32

Environmental Science & Technology

464

(6) Johansson, N.; Fredriksson, A.; Eriksson, P., Neonatal exposure to perfluorooctane

465

sulfonate (PFOS) and perfluorooctanoic acid (PFOA) causes neurobehavioural defects

466

in adult mice. NeuroToxicology 2008, 29, (1), 160-169.

467

(7) Jones, K. C.; de Voogt, P., Persistent organic pollutants (POPs): state of the science.

468

Environ. Pollut. 1999, 100, (1), 209-221.

469

(8) Blondel, C.; Cacciani, P.; Delsart, C.; Trainham, R., High-resolution determination

470

of the electron affinity of fluorine and bromine using crossed ion and laser beams. 1989,

471

40, 3698-3701.

472

(9) Moriwaki, H.; Takagi, Y.; Tanaka, M.; Tsuruho, K.; Okitsu, K.; Maeda, Y.,

473

Sonochemical decomposition of perfluorooctane sulfonate and perfluorooctanoic acid.

474

Environ. Sci. Technol. 2005, 39, (9), 3388-3392.

475

(10) Mitchell, S. M.; Ahmad, M.; Teel, A. L.; Watts, R. J., Degradation of

476

Perfluorooctanoic acid by reactive species generated through catalyzed H2O2

477

propagation reactions. Environ. Sci. Technol. Lett. 2014, 1, (1), 117-121.

478

(11) Vecitis, C. D.; Park, H.; Cheng, J.; Mader, B. T.; Hoffmann, M. R., Treatment

479

technologies for aqueous perfluorooctanesulfonate (PFOS) and perfluorooctanoate

480

(PFOA). Front. Env. Sci. Eng. 2009, 3, (2), 129-151.

481

(12) Hori, H.; Yamamoto, A.; Hayakawa, E.; Taniyasu, S.; Yamashita, N.; Kutsuna, S.;

482

Kiatagawa, H.; Arakawa, R., Efficient decomposition of environmentally persistent

483

perfluorocarboxylic acids by use of persulfate as a photochemical oxidant. Environ.

484

Sci. Technol. 2005, 39, (7), 2383-2388.

485

(13) Hori, H.; Hayakawa, E.; Einaga, H.; Kutsuna, S.; Koike, K.; Ibusuki, T.;

486

Kiatagawa, H.; Arakawa, R., Decomposition of environmentally persistent

487

perfluorooctanoic acid in water by photochemical approaches. Environ. Sci. Technol.

488

2004, 38, (22), 6118-6124.

489

(14) Wang, Y.; Zhang, P., Photocatalytic decomposition of perfluorooctanoic acid

490

(PFOA) by TiO2 in the presence of oxalic acid. J. Hazard. Mater. 2011, 192, (3), 1869-

491

1875.

492

(15) Huang, D.; Yin, L.; Niu, J., Photoinduced hydrodefluorination mechanisms of

493

perfluorooctanoic acid by the SiC/graphene catalyst. Environ. Sci. Technol. 2016, 50,

494

(11), 5857-5863.

495

(16) Wang, Y.; Zhang, P.; Pan, G.; Chen, H., Ferric ion mediated photochemical

496

decomposition of perfluorooctanoic acid (PFOA) by 254 nm UV light. J. Hazard.

497

Mater. 2008, 160, (1), 181-186.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 32

498

(17) Chen, X.; Zhang, H.; Zhang, D.; Miao, Y.; Li, G., Controllable synthesis of

499

mesoporous multi-shelled ZnO microspheres as efficient photocatalysts for NO

500

oxidation, Appl. Surface Sci. 2018, 435, 468-475.

501

(18) Estrellan, C. R.; Salim, C.; Hinode, H., Photocatalytic decomposition of

502

perfluorooctanoic acid by iron and niobium co-doped titanium dioxide. J. Hazard.

503

Mater. 2010, 179, (1), 79-83.

504

(19) Wang, S.; Lin, S.; Zhang, D.; Li, G.; Leung, M. K. H., Controlling charge transfer

505

in quantum-size titania for photocatalytic applications. Appl. Catal. B Environ 2017,

506

215, 85-92.

507

(20) Wang, X.; Zhao, X.; Zhang, D.; Li, G.; Li, H., Microwave irradiation induced UIO-

508

66-NH2 anchored on graphene with high activity for photocatalytic reduction of CO2.

509

Appl. Catal. B Environ 2018, 228, 47-53.

510

(21) Pan, D.; Han, Z.; Miao, Y.; Zhang, D.; Li, G., Thermally stable TiO2 quantum dots

511

embedded in SiO2 foams: Characterization and photocatalytic H2 evolution activity.

512

Appl. Catal. B Environ 2018, 229, 130-138.

513

(22) Li, Z.; Wang, W.; Ding, C.; Wang, Z.; Liao, S.; Li, C., Biomimetic electron

514

transport via multiredox shuttles from photosystem II to a photoelectrochemical cell for

515

solar water splitting. Energy Environ. Sci. 2017, 10, (3), 765-771.

516

(23) Li, H.; Xiao, S.; Zhou, J.; Zhao, J.; Liu, F.; Li, G.; Zhang, D., A flexible CdS

517

nanorods-carbon

518

photoelectrocatalytic

519

DOI:10.1039/C9CC00050J.

520

(24) Zhang, H.; Ding, Q.; He, D.; Liu, H.; Liu, W.; Li, Z.; Yang, B.; Zhang, X.; Lei, L.;

521

Jin, S., A p-Si/NiCoSex core/shell nanopillar array photocathode for enhanced

522

photoelectrochemical hydrogen production. Energy Environ. Sci. 2016, 9, (10), 3113-

523

3119.

524

(25) Kubacka, A.; Fernández-García, M.; Colón, G., Advanced nanoarchitectures for

525

solar photocatalytic applications. Chem. Rev. 2012, 112, (3), 1555-1614.

526

(26) Ma, H.Y.; Zhao, L.; Wang, D.B.; Zhang, H.; Guo, L.H., Dynamic tracking of

527

highly toxic Intermediates in photocatalytic degradation of pentachlorophenol by

528

continuous flow chemiluminescence. Environ. Sci. Technol. 2018, 52, (5), 2870-2877.

529

(27) Li, L.; Zhao, X.; Pan, D.; Li, G., Nanotube array-like WO3/W photoanode

530

fabricated by electrochemical anodization for photoelectrocatalytic overall water

531

splitting. Chinese J. Catal. 2017, 38, (12), 2132-2140.

nanotubes/stainless hydrogen

steel

mesh

evolution,

photoanode

Chem.

ACS Paragon Plus Environment

for

boosted

Commun.

2019,

Page 27 of 32

Environmental Science & Technology

532

(28) Wang, W.; Li, F.; Zhang, D.; Leung, D. Y. C.; Li, G., Photoelectrocatalytic

533

hydrogen generation and simultaneous degradation of organic pollutant via CdSe/TiO2

534

nanotube arrays. Appl. Surf. Sci. 2016, 362, 490-497.

535

(29) Li, G.; Lian, Z.; Wang, W.; Zhang, D.; Li, H., Nanotube-confinement induced size-

536

controllable g-C3N4 quantum dots modified single-crystalline TiO2 nanotube arrays for

537

stable synergetic photoelectrocatalysis. Nano Energy 2016, 19, 446-454.

538

(30) Lian, Z.; Wang, W.; Xiao, S.; Li, X.; Cui, Y.; Zhang, D.; Li, G.; Li, H., Plasmonic

539

silver quantum dots coupled with hierarchical TiO2 nanotube arrays photoelectrodes

540

for efficient visible-light photoelectrocatalytic hydrogen evolution. Sci. Rep. 2015, 5,

541

10461.

542

(31) Jia, L.; Sun, X.; Jiang, Y.; Yu, S.; Wang, C., A Novel MoSe2–reduced graphene

543

oxide/polyimide

544

photoelectrocatalysis hydrogen evolution. Adv. Funct. Mater. 2014, 25, (12), 1814-

545

1820.

546

(32) Brereton, K. R.; Bonn, A. G.; Miller, A. J. M., Molecular photoelectrocatalysts for

547

light-driven hydrogen production. ACS Energy Lett. 2018, 3, (5), 1128-1136.

548

(33) Mert, M. E.; Mert, B. D.; Kardaş, G.; Yazici, B., The photoelectrocatalytic activity,

549

long term stability and corrosion performance of NiMo deposited titanium oxide nano-

550

tubes for hydrogen production in alkaline medium. Appl. Surf. Sci. 2017, 423, 704-715.

551

(34) Yang, L.; Wang, M.; Slattum, P. M.; Bunes, B. R.; Wang, Y.; Wang, C.; Zang, L.,

552

Donor–acceptor

553

photoelectrocatalysts for hydrogen production. ACS Appl. Mater. Interfaces 2018, 10,

554

(23), 19764-19772.

555

(35) Koo, M. S.; Cho, K.; Yoon, J.; Choi, W., Photoelectrochemical Degradation of

556

Organic Compounds Coupled with Molecular Hydrogen Generation Using

557

Electrochromic TiO2 Nanotube Arrays. Environ. Sci. Technol. 2017, 51, (11), 6590-

558

6598.

559

(36) Chen, F.; Zhu, Q.; Wang, Y.; Cui, W.; Su, X.; Li, Y., Efficient

560

photoelectrochemical hydrogen evolution on silicon photocathodes interfaced with

561

nanostructured NiP2 cocatalyst films. ACS Appl. Mater. Interfaces 2016, 8, (45), 31025-

562

31031.

563

(37) Kye, J.; Shin, M.; Lim, B.; Jang, J.-W.; Oh, I.; Hwang, S., Platinum Monolayer

564

electrocatalyst on gold nanostructures on silicon for photoelectrochemical hydrogen

565

evolution. ACS Nano 2013, 7, (7), 6017-6023.

composite

film

supramolecular

for

applications

organic

in

nanofibers

ACS Paragon Plus Environment

electrocatalysis

as

and

visible-light

Environmental Science & Technology

566

(38) Boettcher, S. W.; Warren, E. L.; Putnam, M. C.; Santori, E. A.; Turner-Evans, D.;

567

Kelzenberg, M. D.; Walter, M. G.; McKone, J. R.; Brunschwig, B. S.; Atwater, H. A.;

568

Lewis, N. S., Photoelectrochemical hydrogen evolution using Si microwire arrays. J.

569

Am. Chem. Soc. 2011, 133, (5), 1216-1219.

570

(39) Zhao, X. L; Pan, D. L.; Chen, X. F.; Li, R. P.; Jiang, T. G.; Wang, W. C.; Li, G.

571

S.; Leung, D. Y. C., g-C3N4 photoanode for photoelectrocatalytic synergistic pollutant

572

degradation and hydrogen evolution, Appl. Surf. Sci. 2019, 467, 658-665.

573

(40) Lianos, P., Review of recent trends in photoelectrocatalytic conversion of solar

574

energy to electricity and hydrogen. Appl. Catal. B Environ 2017, 210, 235-254.

575

(41) Huang, X.; Shen, Q.; Liu, J.; Yang, N.; Zhao, G., A CO2 adsorption-enhanced

576

semiconductor/metal-complex hybrid photoelectrocatalytic interface for efficient

577

formate production. Energy Environ. Sci. 2016, 9, (10), 3161-3171.

578

(42) Fan, J.; Shi, H.; Xiao, H.; Zhao, G., Double-layer 3D macro–mesoporous metal

579

oxide modified boron-doped diamond with enhanced photoelectrochemical

580

performance. ACS Appl. Mater. Interfaces 2016, 8, (42), 28306-28315.

581

(43) Wu, G.; Wang, J.; Thomas, D. F.; Chen, A., Synthesis of F-doped flower-like TiO2

582

nanostructures with high photoelectrochemical activity. Langmuir 2008, 24, (7), 3503-

583

3509.

584

(44) Sakthivel, S.; Janczarek, M.; Kisch, H., Visible light activity and

585

photoelectrochemical properties of nitrogen-doped TiO2. J. Phys. Chem. B 2004, 108,

586

(50), 19384-19387.

587

(45) Kim, C.; Kim, J. T.; Kim, K. S.; Jeong, S.; Kim, H. Y.; Han, Y. S., Immobilization

588

of TiO2 on an ITO substrate to facilitate the photoelectrochemical degradation of an

589

organic dye pollutant. Electrochimica Acta 2009, 54, (24), 5715-5720.

590

(46) Du, Y.; Zhou, X.; Qu, Y.; Liu, J.; Feng, Y.; Ren, N., Enhanced electricity

591

generation and pollutant degradation by hybrid photoelectrochemical and microbial

592

fuel cells. Energy Technol. 2016, 5, (3), 402-405.

593

(47) Wu, Z.; Zhao, G.; Zhang, Y.; Liu, J.; Zhang, Y.-n.; Shi, H., A solar-driven

594

photocatalytic fuel cell with dual photoelectrode for simultaneous wastewater treatment

595

and hydrogen production. J. Mater. Chem. 2015, 3, (7), 3416-3424.

596

(48) Zhou, Z.; Wu, Z.; Xu, Q.; Zhao, G., A solar-charged photoelectrochemical

597

wastewater fuel cell for efficient and sustainable hydrogen production. J. Mater. Chem.

598

2017, 5, (48), 25450-25459.

ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32

Environmental Science & Technology

599

(49) Zhang, L.; Niu, C.-G.; Zhao, X.-F.; Liang, C.; Guo, H.; Zeng, G.-M., Ultrathin

600

BiOCl Single-Crystalline Nanosheets with Large Reactive Facets Area and High

601

Electron Mobility Efficiency: A Superior Candidate for High-Performance Dye Self-

602

Photosensitization Photocatalytic Fuel Cell. ACS Appl. Mat. Interfaces 2018, 10, (46),

603

39723-39734.

604

(50) Esposito, D. V.; Forest, R. V.; Chang, Y.; Gaillard, N.; McCandless, B. E.; Hou,

605

S.; Lee, K. H.; Birkmire, R. W.; Chen, J. G., Photoelectrochemical reforming of glucose

606

for hydrogen production using a WO3-based tandem cell device. Energy Environ. Sci.

607

2012, 5, (10), 9091-9099.

608

(51) Kim, W.; Tachikawa, T.; Monllor-Satoca, D.; Kim, H. i.; Majima, T.; Choi, W.,

609

Promoting water photooxidation on transparent WO3 thin films using an alumina

610

overlayer. Energy Environ. Sci. 2013, 6, (12), 3732-3739.

611

(52) Shaner, M. R.; Fountaine, K. T.; Ardo, S.; Coridan, R. H.; Atwater, H. A.; Lewis,

612

N. S., Photoelectrochemistry of core–shell tandem junction n–p+-Si/n-WO3 microwire

613

array photoelectrodes. Energy Environ. Sci. 2014, 7, (2), 779-790.

614

(53) Kim, J.; Lee, C. W.; Choi, W., Platinized WO3 as an environmental Photocatalyst

615

that generates OH radicals under visible light. Environ. Sci. Technol. 2010, 44, (17),

616

6849-6854.

617

(54) Kwong, W. L.; Lee, C. C.; Messinger, J., Transparent nanoparticulate FeOOH

618

improves the performance of a WO3 photoanode in a tandem water-splitting device. J.

619

Phys. Chem. C 2016, 120, (20), 10941-10950.

620

(55) Feng, X.; Chen, Y.; Qin, Z.; Wang, M.; Guo, L., Facile fabrication of sandwich

621

structured WO3 nanoplate arrays for efficient photoelectrochemical water splitting.

622

ACS Appl. Mater. Interfaces 2016, 8, (28), 18089-18096.

623

(56) Chen, Z.; Peng, Y.; Liu, F.; Le, Z.; Zhu, J.; Shen, G.; Zhang, D.; Wen, M.; Xiao,

624

S.; Liu, C. P.; Lu, Y.; Li, H., Hierarchical nanostructured WO3 with biomimetic proton

625

channels and mixed ionic-electronic conductivity for electrochemical energy storage.

626

Nano Lett. 2015, 15, (10), 6802-6808.

627

(57) Zhang, D.; Wang, S.; Zhu, J.; Li, H.; Lu, Y., WO3 nanocrystals with tunable

628

percentage of (001)-facet exposure. Appl. Catal. B Environ 2012, 123-124, 398-404.

629

(58) Li, G.; Lian, Z.; Li, X.; Xu, Y.; Wang, W.; Zhang, D.; Tian, F.; Li, H., Ionothermal

630

synthesis of black Ti3+-doped single-crystal TiO2 as an active photocatalyst for

631

pollutant degradation and H2 generation. J. Mater. Chem. A 2015, 3, (7), 3748-3756.

ACS Paragon Plus Environment

Environmental Science & Technology

632

(59) Wan, F.; Kong, L.; Wang, C.; Li, Y.; Liu, Y.; Zhang, X., The W@WO3 ohmic

633

contact induces a high-efficiency photooxidation performance. Dalton Trans. 2017, 46,

634

(5), 1487-1494.

635

(60) Ab Kadir, R.; Zhang, W.; Wang, Y.; Ou, J. Z.; Wlodarski, W.; O'Mullane, A. P.;

636

Bryant, G.; Taylor, M.; Kalantar-zadeh, K., Anodized nanoporous WO3 Schottky

637

contact structures for hydrogen and ethanol sensing. J. Mater. Chem. A 2015, 3, (15),

638

7994-8001.

639

(61) Reyes-Gil, K. R.; Wiggenhorn, C.; Brunschwig, B. S.; Lewis, N. S., Comparison

640

between the quantum yields of compact and porous WO3 photoanodes. J. Phys. Chem.

641

C 2013, 117, (29), 14947-14957.

642

(62) Hong, S. J.; Jun, H.; Lee, J. S., Nanocrystalline WO3 film with high photo-

643

electrochemical activity prepared by polymer-assisted direct deposition. Scr. Mater.

644

2010, 63, (7), 757-760.

645

(63) Srinivasa Rao, K.; Bhogoju, R. K.; Srujana Devi, G.; Mukhopadhyay, P. K.,

646

Structural and optical properties of nanocrystalline WO3 thin films. 2011; Vol. 22.

647

(64) Yang, B.; Zhang, Y.; Drabarek, E.; Barnes, P. R. F.; Luca, V., Enhanced

648

photoelectrochemical activity of sol−gel tungsten trioxide films through textural

649

control. Chem. Mater. 2007, 19, (23), 5664-5672.

650

(65) Ye, M.; Gong, J.; Lai, Y.; Lin, C.; Lin, Z., High-efficiency photoelectrocatalytic

651

hydrogen generation enabled by palladium quantum dots-sensitized TiO2 nanotube

652

arrays. J. Am. Chem. Soc. 2012, 134, (38), 15720-15723.

653

(66) Wunder, S.; Polzer, F.; Lu, Y.; Mei, Y.; Ballauff, M., Kinetic analysis of catalytic

654

reduction of 4-Nitrophenol by metallic nanoparticles immobilized in spherical

655

polyelectrolyte brushes. J. Phys. Chem. C 2010, 114, (19), 8814-8820.

656

(67) Chakrabarti, S.; Dutta, B. K., Photocatalytic degradation of model textile dyes in

657

wastewater using ZnO as semiconductor catalyst. J. Hazard. Mater. 2004, 112, (3),

658

269-278.

659

(68) Kim, J.; Noh, M. C.; Doh, W. H.; Park, J. Y., Thermal evolution and instability of

660

CO-induced platinum clusters on the Pt(557) surface at ambient pressure. J. Am. Chem.

661

Soc. 2016, 138, (4), 1110-1113.

662

(69) Grigioni, I.; Abdellah, M.; Corti, A.; Dozzi, M. V.; Hammarström, L.; Selli, E.,

663

Photoinduced Charge-Transfer Dynamics in WO3/BiVO4 Photoanodes Probed through

664

Midinfrared Transient Absorption Spectroscopy. J. Am. Chem. Soc. 2018, 140, (43),

665

14042-14045.

ACS Paragon Plus Environment

Page 30 of 32

Page 31 of 32

Environmental Science & Technology

666

(70) Cong, Y.; Ge, Y.; Zhang, T.; Wang, Q.; Shao, M.; Zhang, Y., Fabrication of Z-

667

scheme

668

photoelectrocatalytic performance. Ind. Eng. Chem. Res. 2018, 57, (3), 881-890.

669

(71) Wang, J. C.; Zhang, L.; Fang, W. X.; Ren, J.; Li, Y. Y.; Yao, H. C.; Wang, J. S.;

670

Li, Z. J., Enhanced photoreduction CO2 activity over direct Z-scheme α-Fe2O3/Cu2O

671

heterostructures under visible light irradiation. ACS Appl. Mater. Interfaces 2015, 7,

672

(16), 8631-8639.

673

(72) Gomez-Ruiz, B.; Ribao, P.; Diban, N.; Rivero, M. J.; Ortiz, I.; Urtiaga, A.,

674

Photocatalytic degradation and mineralization of perfluorooctanoic acid (PFOA) using

675

a composite TiO2−rGO catalyst. J. Hazard. Mater. 2018, 344, 950-957.

676

(73) Peng, Y.P.; Chen, H.; Huang, C. P., The synergistic effect of photoelectrochemical

677

(PEC) reactions exemplified by concurrent perfluorooctanoic acid (PFOA) degradation

678

and hydrogen generation over carbon and nitrogen codoped TiO2 nanotube arrays (C-

679

N-TNTAs) photoelectrode. Appl. Catal. B Environ 2017, 209, 437-446.

680

(74) Wang, S.; Yang, Q.; Chen, F.; Sun, J.; Luo, K.; Yao, F.; Wang, X.; Wang, D.; Li,

681

X.; Zeng, G., Photocatalytic degradation of perfluorooctanoic acid and perfluorooctane

682

sulfonate in water: A critical review. Chem. Eng. J. 2017, 328, 927-942.

683

(75) Li, K.; Xu, Y.; He, Y.; Yang, C.; Wang, Y.; Jia, J., Photocatalytic fuel cell (PFC)

684

and dye self-photosensitization photocatalytic fuel cell (DSPFC) with BiOCl/Ti

685

photoanode under UV and visible light Irradiation. Environ. Sci. Technol. 2013, 47, (7),

686

3490-3497.

687

(76) O'Regan, B.; Gr; tzel, M., A Low-Cost, High-efficiency Solar Cell Based on Dye-

688

sensitized Colloidal TiO2 Films. 1991, 353, 737-740.

Fe2O3–MoS2–Cu2O

ternary

nanofilm

with

689 690 691 692 693 694 695 696 697 698

ACS Paragon Plus Environment

significantly

enhanced

Environmental Science & Technology

699

TOC Art

700

ACS Paragon Plus Environment

Page 32 of 32