Enhanced generation of reactive oxygen species under visible light

Resources Processing and Environment, State Key Laboratory. 8 of Silicate Materials for Architectures, Wuhan University of Technology, 122 Luoshi ...
0 downloads 0 Views 905KB Size
Subscriber access provided by Bibliothèque de l'Université Paris-Sud

Energy and the Environment

Enhanced generation of reactive oxygen species under visible light irradiation by adjusting the exposed facet of FeWO4 nanosheets to activate oxalic acid for organic pollutant removal and Cr(VI) reduction Jun Li, Chun Xiao, Kai Wang, Yuan Li, and Gaoke Zhang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.9b00641 • Publication Date (Web): 20 Aug 2019 Downloaded from pubs.acs.org on August 20, 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 28

Environmental Science & Technology

1

Enhanced generation of reactive oxygen species

2

under visible light irradiation by adjusting the

3

exposed facet of FeWO4 nanosheets to activate

4

oxalic acid for organic pollutant removal and Cr(VI)

5

reduction

6

Jun Li‡, Chun Xiao‡, Kai Wang, Yuan Li and Gaoke Zhang*

7

Hubei Key Laboratory of Mineral Resources Processing and Environment, State Key Laboratory

8

of Silicate Materials for Architectures, Wuhan University of Technology, 122 Luoshi Road,

9

Wuhan 430070, China

10

*Corresponding author. E-mail: [email protected]; Phone/Fax: 86-27-87887445.

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 28

11

ABSTRACT

12

In this work, taking FeWO4 nanosheets as an example, the activation of oxalic acid (OA) based

13

on facet engineering for the enhanced generation of active radical species is reported, revealing

14

unprecedented surface Fenton activity for pollutant degradation. Density functional theory (DFT)

15

calculations confirmed the more efficient generation of reactive oxygen species over FeWO4

16

nanosheets with the {001} facet exposed (FWO-001) under visible light irradiation compared to

17

the efficiency of FeWO4 nanosheets with the {010} facet exposed (FWO-010), which could be

18

attributed to a higher density of iron and the efficient activation of OA on the {001} facet. The

19

H2O2-derived •OH tended to diffuse away from the active sites of FWO-001 into solution to

20

favor the continuous activation of OA into the active radicals for pollutant redox reactions, but

21

preferred to remain on FWO-010 to hinder the further activation of OA on the {010} facet.

22

Additionally, the generation of •CO2- endowed FeWO4 with a strong reduction ability. Compared

23

with FWO-010, FWO-001 exhibited enhanced redox activity for the catalytic degradation of

24

organic pollutants and Cr( Ⅳ ) in the optimized conditions. These findings can help in

25

understanding the facet dependent surface Fenton chemistry of catalytic redox reactions and in

26

designing efficient catalysts for environmental decontamination.

ACS Paragon Plus Environment

2

Page 3 of 28

28

Environmental Science & Technology

INTRODUCTION

29

The presence of persistent organic pollutants, such as industrial dyes, antibiotics, and

30

hexavalent chromium Cr(VI), in the wastewater has caused great concern because of their

31

incomplete elimination and persistent pollution of the environment.1-6 Due to the large volume of

32

industrial and municipal wastewater, advanced water treatment techniques are essential for

33

maintaining healthy water circulation. Advanced water treatment techniques for organic

34

pollutant degradation rely on the Fenton reaction (Fe2+/H2O2) to generate •OH.7-10 Nevertheless,

35

in a traditional Fenton reaction system, the free ferrous ions can be quickly expended, and this

36

process requires the addition of Fe2+. The excess addition of Fe2+ can result in the accumulation

37

of Fe2+. To avoid the disadvantages of traditional Fenton reactions, heterogeneous Fenton

38

systems, such as Fe0, WS2, CuFeO2, α-FeOOH and maghemite/montmorillonite (MMT)

39

composites, have been developed for environmental purification.11-15 Although numerous

40

Fenton-like catalysts have been extensively investigated for contaminant removal, developing

41

highly efficient, stable and low-cost catalysts is still challenging and a promising research field.

42

Different from traditional Fenton systems, the Fe(III)-oxalate system, which can produce

43

strong oxidized radicals to degrade organic pollutants under visible light irradiation without the

44

addition of H2O2, has received considerable attention.16-18 The photochemical Fe(III)-oxalate

45

system has been demonstrated to be more efficient for catalytic organic pollutant degradation

46

than the Fenton reaction (Fe(II/III)/H2O2) due to the rapid cycling of iron and the generation of •

47

OH.19-21 To date, numerous studies have been reported regarding the superior photo-Fenton

48

catalytic activity and mechanism of the Fe(II/III)-oxalate system. Mazellier and Sulzberger22

49

reported an α-FeOOH/oxalate system and systematically explored the mechanism of Fe(II)

50

formation and the role of oxalate. Wei et al.23 studied a zero-valent iron (Fe0)/oxalate system for

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 28

51

the simultaneous and rapid redox removal of chromium (Cr(VI)) and orange II and its relative

52

mechanism. Recent studies of Fe/oxalate systems have focused on designing composite

53

structures, different Fe-based compounds, etc. to improve the stability and activity of the

54

system.24-27 Zhang et al.28 reported the hematite with exposed different facets for U(VI) removal.

55

The U(VI) adsorption site densities on the {012} and {110} facets of hematite were higher than

56

that on the {001} facet, suggesting that the U(VI) adsorption activity by hematite was strongly

57

dependent on the coordination type of U(VI) on the hematite facets. Like this, due to the

58

difference of the surface structure of FeWO4 nanosheets with different exposed facets, the OA

59

adsorption activity by FeWO4 nanosheets was determined by the coordination environment of

60

OA on FWO-001 and FWO-010, which would affect the catalytic performance in

61

FeWO4/OA/vis system. Typically, FeWO4 as a significant functional material, has been widely

62

used in catalysis, dyes, pigments, sensors, preservatives and magnetic materials. However, a

63

detailed study about the Fenton-like reaction by controlling exposed facets for pollutant removal

64

has not been reported up to now.

65

Herein, inspired by facet engineering, FeWO4 nanosheets with different exposed facets were

66

successfully synthesized via a hydrothermal method, representing a promising approach for

67

adding OA to a catalyst/aqueous organic pollutant system under visible light irradiation to

68

produce active radicals and realizing the removal of pollutants. This system for the catalysts-

69

assisted activation of oxalic acid Fenton catalysis was validated in FeWO4/acid red G (ARG),

70

FeWO4/methyl orange (MO), FeWO4/4-nitrophenol (4-NP) and FeWO4/Cr( Ⅵ ) systems under

71

visible light irradiation. Compared with FWO-010, FWO-001 exhibited improved H2O2 and •OH

72

generation, resulting in a greatly enhanced redox activity for the degradation of ARG and Cr(VI)

73

under optimal conditions. A possible mechanism for FeWO4 nanosheets with different exposed

ACS Paragon Plus Environment

4

Page 5 of 28

Environmental Science & Technology

74

facets activating OA to generate active species was proposed. Our work provides atomic scale

75

insights for understanding the influence of different facets on the adsorption and activation of

76

oxalate, H2O2 dissociation and •OH generation behaviors.

77

MATERIALS AND METHODS

78

The Synthesis of FWO-001. Firstly, 0.014 mol of Na2WO4·2H2O was added to 100 mL of

79

water with magnetic stirring until a homogeneous solution was formed. The pH of the Na2WO4

80

solution was adjusted to 1.2 by using a 2 M HCl solution. Second, 0.035 mol C2H2O4·2H2O was

81

added to the above solution, which was then diluted to 250 mL with continuous magnetic stirring

82

for 30 min at room temperature. Finally, 30 ml of the above solution was measured and

83

transferred into a 50 ml stainless steel polyphenylene (PPL)-lined autoclave, and simultaneously,

84

0.012 mol of FeSO4·7H2O was added. The autoclave was sealed and heated to 220 °C for 24 h,

85

and then naturally cooled to room temperature. The final products were collected by

86

centrifugation, washed with distilled water, dried at 60 °C for 6 h in a vacuum drying chamber,

87

and finally labelled as FWO-001.

88

The Synthesis of FWO-010. First, 0.99 g of Na2WO4·2H2O and 1.176 g of

89

Fe(NH4)2(SO4)2·6H2O were dissolved in 30 mL of deionized water with continuously magnetic

90

stirring until a homogeneous solution was formed. Afterwards, 6 mL of a NaOH solution (1

91

mol/L) was added to the above solution until a dark green solution formed. Finally, the

92

suspension was transferred into a stainless steel PPL-lined autoclave, sealed and heated to 180

93

°C for 12 h. The final products were collected by centrifugation, washed with distilled water,

94

dried at 60 °C for 6 h in a vacuum drying chamber, and finally labelled as FWO-010.

95

Characterization. The phase and crystallinity of the catalysts were determined by powder X-

96

ray diffraction (XRD, Japan) over the diffraction angle range from 5° to 70° at a step size of

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 28

97

0.02°. The X-ray source was Cu Kα radiation. The morphologies and microstructures of the

98

products were observed by scanning electron microscopy (SEM), transmission electron

99

microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) using a

100

JEM 2100 F electron microscope at an accelerating voltage of 200 kV. X-ray photoelectron

101

spectroscopy (XPS) analysis was performed using an ESCALAB 250Xi system (Thermo

102

Scientific, USA), all binding energies were calibrated in reference to the binding energy of the C

103

1s peak. Fourier transform infrared (FT-IR) spectroscopy (Thermo Nicolet, USA) was used to

104

characterize the chemical bonds of the products.

105

Pollutant Degradation. The degradation of acid red G (ARG) in water was performed to

106

evaluate the photocatalytic activities of the samples. In a typical experiment, 35 mg of the

107

catalyst was dispersed in 50 mL of an aqueous ARG solution (50 mg/L). To eliminate the effect

108

of adsorption, the mixed suspensions were constantly stirred for 30 min in the dark to reach the

109

adsorption equilibrium. Next, 6.3 mg of oxalic acid dihydrate was added to the above suspension

110

and the suspension was irradiated with visible light. In this work, 100 W LED lamp with the

111

wavelengths 420 nm was selected as visible light source and its light intensity of visible light

112

sources was measured to be 55 mW•cm-2. Then, 7 mL of the suspension was collected after a

113

certain time interval (15 min) and then centrifuged (5000 rpm, 5 min). Finally, the concentration

114

of ARG was measured by using a UV-Vis spectrophotometer (Orion AquaMate 8000, China) at

115

the maximum absorption wavelength of ARG (554 nm).

116

Density Functional Theory Calculations. DFT calculations were performed using the Vienna

117

Ab initio Simulation Package with projector-augmented wave (PAW) pseudopotentials.29 The

118

generalized gradient approximation (GGA) in the Perdew, Burke and Ernzerhof (PBE)

119

parametrization was selected as the exchange-correlation functional.30 A plane-wave basis set

ACS Paragon Plus Environment

6

Page 7 of 28

Environmental Science & Technology

120

with a cutoff energy of 420 eV was set. The lattice parameters of FWO-001 were a = 4.75 Å, b =

121

5.72 Å, c= 22.04 Å, and that of FWO-010 were a = 4.97 Å, b = 4.75 Å, and c= 22.18 Å. In

122

addition, the k-point meshes of FWO-001 and FWO-010 were both selected to be 5 × 5 × 1. The

123

residual force and iterative energy difference of all atoms were allowed to converge to 1×10-4 eV

124

and 0.02 eV‧Å-1.

125

RESULTS AND DISCUSSION

126

To investigate the structures of the samples, a series of characterization methods, including

127

XRD, FT-IR and XPS measurements, were performed. Figure 1a shows the XRD patterns of the

128

samples. All the characteristic peaks agree well with those of the standard card for FeWO4

129

(JCPDS: 46-1446), and no other peaks are found, indicating the high crystallinity and purity of

130

the samples. Figure 1b shows the FTIR spectra of FWO-001 and FWO-010. The peak at 854 cm-

131

1

132

band at 650 cm-1 can be assigned to the stretching vibration of W-O in the WO6 octahedral, and

133

the asymmetric deformation vibration of Fe-O in FeO6 is observed at 508 cm-1.31-33 XPS

134

measurements were then carried out to elucidate the chemical states of the samples. The two

135

peaks at 709.7 and 723.9 eV correspond to Fe 2p3/2 and Fe 2p1/2, respectively, indicating the

136

presence of the Fe2+ species (Figure S1a).34 Figure S1b shows two obvious peaks located at 35.3

137

and 37.4 eV, which can be attributed to W 4f7/2 and W 4f5/2, respectively, confirming the

138

presence of W6+ in the FeWO4 structure.35 The O 1s region can be deconvoluted into two peaks

139

corresponding to the lattice oxygen of the W-O bond (530.3 eV) and -OH (532.1 eV) (Figure

140

S1c).36

corresponds to the symmetric vibration peak of the oxygen atom in Fe-W-O, the characteristic

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 28

141 142

Figure 1. (a) XRD patterns, (b) FTIR spectra, (c, f) TEM images, (d, g) HRTEM images and (e,

143

h) FFT images of FWO-001 and FWO-010.

144

TEM images of pure FWO-001 and FWO-010 are shown in Figure 1c-h. It can be seen that

145

FWO-001 and FWO-010 both exhibit nanosheet morphology. HRTEM images provide further

146

insight into the microstructures of the as-prepared samples. Lattice fringe spacings of 0.37 nm

147

for the FWO-001 nanosheet and 0.247 nm for the FWO-010 nanosheets correspond to the (011)

148

and (021) lattice planes, respectively, and their corresponding fast Fourier transform (FFT)

149

images reveal their single-crystalline nature. By analysis of the FFT images we can conclude that

150

the FWO-001 nanosheets and the FWO-010 nanosheets were grown with the [001] and [010]

ACS Paragon Plus Environment

8

Page 9 of 28

Environmental Science & Technology

151

orientations, respectively, as the main exposed facets, suggesting the main exposed facet are the

152

{001} and {010} facets, respectively.

153

The photo-Fenton activity of the FeWO4 nanosheets with different exposed facets in the

154

presence of OA for visible-light-driven photo-Fenton catalysis was evaluated by the degradation

155

of ARG and the results are shown in Figures 2a, b and S2. In the dark, with or without the

156

addition of H2O2, FWO-001 showed weak adsorption performance and negligible activity toward

157

ARG degradation. Under visible light irradiation, FWO-001/H2O2 system still exhibited

158

ignorable performance toward ARG degradation. In the presence of both OA and visible light,

159

the degradation ratio of ARG by FWO-001 is 98%, confirming that FWO-001 can significantly

160

accelerate the degradation rate of ARG in the presence of OA under visible light irradiation. The

161

reaction kinetic constant (k) for FWO-001 is more than 4 times greater than that of FWO-010 for

162

the photo-Fenton catalytic oxidative degradation of ARG. Additionally, the photo-Fenton

163

catalytic reductive degradation of Cr(VI) in an aqueous solution was also measured (Figure 2c

164

and d). Only a 10% removal rate is observed for a Cr(VI) solution mixed with FWO-001 and

165

C2H2O4 in the dark, showing the weak adsorption of Cr(VI) onto FWO-001. After adding oxalic

166

acid or FWO-001, the removal rate of Cr(VI) is only 11% and 19%, respectively. A 100%

167

removal ratio of Cr(VI) is achieved by FWO-001 in the presence of oxalic acid under visible

168

light irradiation. However, a 38% removal rate is observed after adding FWO-010 to the aqueous

169

solution under the same conditions. It can be seen that the reaction kinetic constant (k) of FWO-

170

001 is approximately 10 times greater than that of FWO-010 for the photo-Fenton catalytic

171

reductive degradation of Cr(VI). XPS measurement was further employed to analyze the surface

172

species after Cr(VI) reduction (Figure S3). The peaks located at 579.3 and 588.4 eV could be

173

assigned to the binding energies of Cr(VI) 2p3/2 and Cr(VI) 2p1/2, respectively. The peaks at

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 28

174

577.8 and 586.9 eV could be ascribed to the binding energies of Cr(III) 2p3/2 and Cr(III) 2p1/2,

175

respectively, suggesting the Cr(VI) was actually reduced into Cr(III) in the presence of the

176

FWO-001 and OA under visible light irradiation. In comparison with previously reported

177

catalysts/OA/vis systems (Table S1), the FWO-001/OA/vis system revealed superior or

178

comparable catalytic performance for pollutants removal. The above results confirm that FWO-

179

001 can more favorably activate OA for pollutant removal.

180

The influences of the concentrations of the catalyst and oxalic acid and the solution pH on

181

ARG degradation were examined. When the dosage of catalysts was increased from 0.3 to 1.5 g•

182

L-1, more than 90 % degradation ratio of ARG was maintained within 60 min (Figure S4a, Table

183

S2). The effect of OA concentration on the catalytic degradation of ARG was observed from 0.5

184

to 2.0 mM. The results showed when the concentration of OA was 1.0 mM, FWO-001

185

nanosheets exhibited the optimized catalytic activity (Figure S4b). The excess oxalate inhibited

186

the degradation of ARG initially. The reason may be that oxalate can also act as a scavenger of

187

·OH produced in the Fenton reaction. Figure S4c displayed the influence of initial solution pH on

188

the degradation of ARG by FWO-001 nanosheets in a wide pH range of 2.2-9.6. As shown in

189

Figure 3c, the catalytic activity of FWO-001 nanosheets increased with increasing pH between

190

2.2 and 5.8. The pH-dependence of ARG degradation in the photo/ferrioxalate system can be

191

explained by the dependence of Fe(II) speciation on pH value, which affects the rate of the

192

Fenton reaction. At higher solution pH, the solubility of Fe(III) and Fe(II) strongly decreased and

193

the main species were Fe(III)-OH and Fe(II)-OH which could form precipitations and lost

194

photoactive. In order to observe the intermediate products of ARG degradation, the Liquid

195

chromatography-mass spectrometer (LC-MS) was performed. According to the analysis of mass

196

spectrometry, the main transformation products can be identified and products with m/z of 291,

ACS Paragon Plus Environment

10

Page 11 of 28

Environmental Science & Technology

197

274, 231, 160 and 156 could be attributed to the degraded products of ARG in aqueous solution.

198

The corresponding mass spectra and chemical structures of the possible intermediates are

199

presented in Table S3. In addition, MO, RhB, 4-NP and tetracycline (TC) were also selected as

200

target pollutants to generalize the effect of OA activation by the FeWO4 nanosheets (Figure S5),

201

showing that the addition of OA is a promising approach that can significantly enhance visible-

202

light-driven photo-Fenton catalysis for pollutant removal. The stability of FWO-001 was

203

monitored over multiple runs (Figures 2e and S6). After six rums, the FWO-001 nanosheets

204

retain their high photo-Fenton catalytic activity for ARG removal and Cr(VI) reduction. Figure

205

2f shows the identical XRD spectra of the fresh and used FWO-001 nanosheets. It can be seen

206

that their XRD spectra did not show significant changes, further indicating the structural stability

207

of the FWO-001 nanosheets.

208 209

Figure 2. Degradation activity curves for (a) ARG and (c) Cr( Ⅵ ) in aqueous solutions under

210

different conditions. The corresponding kinetic curves for (b) ARG and (d) Cr(Ⅵ) degradation.

211

(e) The cyclical performance of the FWO-001 nanosheet catalyst for the degradation of ARG in

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 28

212

an aqueous solution. (f) XRD patterns of the catalyst before and after utilization. Initial

213

conditions: CARG = 10.0 mg/L, Wcatalyst = 35 mg, COA = 1.0 mM.

214

To identify the radical species involved in the pollutant degradation, trapping experiments of

215

the active species were carried out.37-38 As shown in Figure 3a, the catalytic performance for the

216

degradation of ARG can be inhibited in the presence of isopropanol and p-benzoquinone

217

suggesting that •O2- play crucial roles in the degradation process. To identify the source of •O2-, a

218

comparison experiment was employed by excluding dissolved O2 and conducting the experiment

219

under a N2 atmosphere. The results reveal that the catalytic activity did not significantly

220

decrease, confirming that ·O 2- was originated from the decomposition of ·C2O- 4 rather than a

221

reaction between dissolved O2 and the electrons of the catalyst.

222 223

Figure 3. (a) Photo-Fenton activities of FWO-001 for the degradation of ARG with different

224

radical scavengers. Quantitative determination of the amounts of (b) H2O2 and (c) free •OH

225

generated by the as-prepared samples in the presence of H2C2O4. ESR signals of (d) DMPO-•O2-

226

and (e) DMPO-•OH for FWO-001 and FWO-010 after 2 min in the dark and under visible light

ACS Paragon Plus Environment

12

Page 13 of 28

Environmental Science & Technology

227

irradiation, respectively. (f) Changes in the leached iron concentration at different irradiation

228

times for FWO-001. Initial conditions: CARG = 10.0 mg/L, Wcatalyst = 35 mg, COA = 1.0 mM.

229

The generation and decomposition of H2O2 in photo-Fenton systems determine the reaction

230

rate. The generation of H2O2 over FWO-010/OA and FWO-001/OA systems was evaluated as

231

shown in Figure 3b. As the irradiation time increases, the concentration of H2O2 gradually

232

increases in the presence of FeWO4 and OA. Compared with that of FWO-010, FWO-001

233

exhibits a relative higher H2O2 generation efficiency of 160 μmol/L. Similar phenomena are also

234

observed for the generation of •OH (Figure 3c). The high rate of H2O2 production is positively

235

correlated to the surface iron ion concentration, which is in agreement with the following DFT

236

study. The generation of H2O2 was further demonstrated by the color variation of KI paper, as

237

shown in Figure S7. DMPO-ESR spectra were further used to detect the active species. The

238

characteristic signals of DMPO-·O 2- and DMPO-·OH can be obviously detected for the FWO-

239

001 and FWO-010 samples under visible light irradiation (Figure 3d-e). The ·O 2- and ·OH

240

signals of FWO-001 are stronger than those of FWO-010, indicating the amount of ·O2- and ·OH

241

generated by the FWO-001 activation of OA significantly exceeds that produced by the FWO-

242

010 activation of OA. Figure 3f shows the changes in the Fe2+ and Fe3+ concentrations during the

243

catalytic process. Fe2+ and Fe3+ have the same variation tendencies and their maximum

244

concentrations are less than 2.5 mg/L, suggesting low Fe leaching rates and high stabilities for

245

the samples. A comparison of the homogeneous and heterogeneous Fenton catalytic performance

246

for the same Fe content was also performed, indicating that Fe leaching has little influence on the

247

activity of the heterogeneous Fenton catalytic degradation of pollutants in this work (Figure

248

S4d).

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 28

249

In-situ IR experiments were performed to observe the relative information on both adsorption

250

modes and the relative affinities of OA/H2O2 to the two facets. It can be seen in Figure S8 that as

251

the oxalic acid concentration increased from 0 to 2.0 mM, the wavenumbers at 677 and 1948 cm-

252

1

253

FWO-010 surface. The possible coordination types of Fe-OA on FWO-001 surface were given in

254

Figure S9. In this work, FWO-001 revealed better catalytic activity for removal of Cr(Ⅵ) and

255

pollutants in the presence of OA under visible light irradiation. According to previous reports,39-

256

40

257

removal of Cr(Ⅵ) and pollutants. Therefore, the second coordination type in Figure S9b was

258

regarded as the optimized Fe-OA coordination type.

appeared, suggesting the different Fe-OA formed different coordination type on FWO-001 and

the second coordination type (Figure S9b) of Fe-OA on FWO-001 surface was helpful to the

259

To gain further insight into the effect of the exposed facet and iron concentration on the

260

catalytic activity of the FeWO4 nanosheets, density functional theory (DFT) calculations were

261

performed to compare the {001} and {010} facets. The representative relaxed atomic geometries

262

of each slab are shown in Figure 4a and c. The calculated surface energies of the {001} and

263

{010} facets are summarized in Table S4. The surface energies are 1.86 and 2.13 J/m2 for the

264

{001} and {010} facets, respectively, indicating that the {010} facet of FeWO4 nanosheets is the

265

more active facet. In general, crystals prefer to expose thermodynamically stable facets during

266

the crystal growth process, resulting in minimization of the surface energy. Therefore, the

267

exposed percentage of the {001} facet is 97%, while the exposed percentage of the {010} facets

268

is only 52%. Considering the effect of specific surface area of the samples on their catalytic

269

activity, the BET measurements were carried out (Figure S10). Baaed on the IUPAC

270

classification, the N2 adsorption-desorption isotherm of FWO-001 and FWO-010 are both of

271

type IV with obvious hysteresis hoops of type H3. The specific surface areas of FWO-001 and

ACS Paragon Plus Environment

14

Page 15 of 28

Environmental Science & Technology

272

FWO-010 are 10.879 and 17.641 m2/g, respectively (shown in Table S5). The results showed

273

that FWO-001 with smaller specific surface area, but had the better catalytic activity as

274

compared to FWO-010. Therefore, it can be concluded that the Fe density on different facets of

275

FeWO4 may dominate the catalytic activity. Therefore, it is necessary to investigate the Fe

276

charge density in the FeWO4 nanosheets with different exposed facets. Figure 4b and d reveal the

277

calculated charge densities of the Fe atoms in the {001} and {010} facets, respectively. As such,

278

our DFT results confirm that an overwhelming majority of the Fe is distributed in the {001}

279

facet, which allows for a higher carrier density as well as more efficient carrier transport. The

280

higher concentration of Fe in the {001} facet provides a greater number of active sites in the

281

photo-Fenton reaction system, favoring a highly efficient catalytic reaction.

282 283

Figure 4. (a, c) Surface structures of the relaxed stoichiometric {001} and {010} facets and (b,

284

d) their corresponding charge density distributions. (e) Calculated adsorption energy profiles for

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 28

285

H2O2 dissociation over FWO-001 and FWO-010. (f) Schematic illustration of the •OH

286

generation process over FWO-001.

287

The above theoretical calculation results indicate that the higher concentration of Fe in FWO-

288

001 provides a greater number of active sites in the photo-Fenton reaction system, which

289

contributes to its highly efficient catalytic reactivity. In a catalyst/OA/visible light system, the

290

catalytic performance is closely related to the adsorption and activation of OA, the adsorption

291

and dissociation of H2O2, and the diffusion of •OH. To further understand these processes on the

292

different facets of the FeWO4 nanosheets, the adsorption energies of OA and H2O2 on the

293

different facets of the FeWO4 nanosheets were calculated. Tables S6 and S7 show the calculated

294

adsorption energies for the different facets with OA (or H2O2) molecules. In a

295

catalyst/OA/visible light system, the OA molecules are first adsorbed onto the FeWO4

296

nanosheets, forming an FeII-OA complex and generating H2O2. Once the OA has been activated,

297

OA and H2O2 competitively adsorb onto the exposed iron atoms. Due to the higher adsorption

298

energy of H2O2 on the {010} facet of the FeWO4 nanosheets than that of OA on the {010} facet

299

of the FeWO4 nanosheets, the generated H2O2 molecules will preferentially adsorb onto the

300

{010} facet of the FeWO4 nanosheets. The preferential adsorption of H2O2 molecules onto the

301

{010} facet of the FeWO4 nanosheets will occupy the abundant active sites, further hindering the

302

adsorption and activation of OA on the {010} facet of the FeWO4 nanosheets and limiting the

303

catalytic reaction. Furthermore, the slow diffusion of •OH from the {010} facet of the FeWO4

304

nanosheets affects its reactivity with pollutants (Figure 4e). The process for OA activation by

305

FWO-001 is presented in Figure 4f. For FWO-001, the adsorption energy of OA on the {001}

306

facet of the FeWO4 nanosheets is more negative than that of the adsorption energy of H2O2 on

307

the {001} facet of the FeWO4 nanosheets, indicating that OA molecules are preferentially

ACS Paragon Plus Environment

16

Page 17 of 28

Environmental Science & Technology

308

adsorbed onto the {001} facet of the FeWO4 nanosheets compared to H2O2, which favors the

309

activation of OA and the diffusion of H2O2. In addition, the decomposition products of H2O2 can

310

easily diffuse into solution to oxidize the pollutants. Subsequently, the freed active sites can

311

participate in the next reaction step. Therefore, the FWO-001 nanosheets can exhibit a better

312

catalytic performance for pollutant degradation in the activated OA system than can the exposed

313

{010} facet of the FeWO4 nanosheets. Obviously, controlling the activation of OA, adsorption

314

and dissociation of H2O2, and diffusion of •OH by facet engineering is a promising strategy to

315

control their catalytic reactivity for pollutant removal in catalyst/OA systems.

316

According to the reported work and the experimental results, a possible mechanism for the

317

catalytic degradation of organic pollutants and Cr(VI) in the FeWO4/OA system under visible

318

light irradiation is proposed (Figure 5). In the reaction system, FeII and OA can form an FeII-OA

319

complex and then transform into an [FeIII(C2O4)3]3- complex, which is a reciprocal reaction

320

(reaction 1) that has a high photochemical reactivity. The excitation of [FeIII(C2O4)3]3- under

321

visible light irradiation involves an intramolecular electron transfer from oxalate to Fe(III),

322

forming ·C2O4- (reaction 2), and then ·C2O4- can be decomposed into ·CO2- and ·O2- radicals

323

(reaction 3-4), which can react with Fe(II) to generate H2O2 (reactions 5-6). In this process,

324

visible light irradiation promotes the generation of the intermediate free radicals, contributing to

325

the catalytic performance, and then, [FeII(C2O4)2]2- reacts with H2O2 to produce ·OH (reaction 7).

326

H2O2 can also be trapped by Fe3+ on the surface of the catalyst, forming Fe2+ and completing the

327

cycle of Fe2+/Fe3+. Finally, the produced active species can attack the pollutants causing the

328

degradation of the organic pollutants (reactions 8) and the reduction of Cr(VI) into Cr(III)

329

(reactions 9-10). The possible reaction pathway is shown in reactions 1-1017, 19, 41-42:

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 28

330 331

Figure 5. Schematic illustration for the activation of OA in the FeWO4/OA system under visible

332

light irradiation for the degradation of organic pollutants and Cr(VI).

333

[FeII(C2O4)2]2-↔ [FeIII(C2O4)3]3-

(1)

334

[FeIII(C2O4)3]3- + hν → ·C2O- 4 + [FeII(C2O4)2]2-

(2)

335

C2O- 4·→ ·O2- + CO2

(3)

336 337

C2O- 4·→ ·CO2- + CO2 (4)

338

H+ + ·O2-→ ·HO2

339 340

FeII + ·O- 2/HO2· + H+→ FeIII + H2O2 (6)

341

[FeII(C2O4)2]2- + H2O2 →[FeIII(C2O4)2]+ + ·OH + OH-

(7)

342

OH/·O2- + organic pollutants → CO2 + H2O

(8)

343

[FeII(C2O4)2]2- + Cr(VI)→ [FeIII(C2O4)2]+ + Cr(III)

(9)

344

·CO- 2 + Cr(VI)→ CO2 + Cr(III)

(10)

(5)

345

Environmental Implications. In this work, the effect of the exposed facet of FeWO4

346

nanosheets on oxalic acid activation and the catalytic performance for pollutant degradation was

347

investigated for the first time. In FeWO4/OA/vis system, once the OA has been activated, OA

348

and H2O2 competitively adsorb on the surface of catalysts. Compared to FWO-001, the

349

preferential adsorption of H2O2 molecules on FWO-010 will occupy the abundant active sites,

350

further hindering the continuous adsorption and activation of OA to generation H2O2 and

ACS Paragon Plus Environment

18

Page 19 of 28

Environmental Science & Technology

351

limiting the generation of •OH and other active radicals. The easier generation of active radicals

352

on FWO-001 resulted in the efficiently catalytic performance toward the degradation of organic

353

pollutants, especially for Cr(VI) reduction. Therefore, regulating the exposed facet of a catalyst

354

allows for controlling the adsorption and activation of oxalic acid and the generation and

355

decomposition of H2O2. Our findings provide comprehensive insights into facet-dependent

356

surface Fenton chemistry for environmental decontamination.

357

ASSOCIATED CONTENT

358

Supporting Information: Details for the time dependent UV-visible absorption spectra of ARG

359

in an aqueous solution; COD removal; effect of catalyst concentration, C2H2O4 concentration

360

and initial pH on the catalytic performance of FWO-001; the catalytic degradation of MO, RhB,

361

4-NP and TC in the FeWO4/OA system under visible light irradiation; and the adsorption

362

energies of the OA and H2O2 molecules on the FeWO4 nanosheets with different exposed facets.

363

AUTHOR INFORMATION

364

*Corresponding

365

Phone/fax: +86-27-87887445. E-mail: [email protected]

366

Author Contributions

367

‡These authors contributed equally.

368

Notes

369

The authors declare no competing financial interest.

370

ACKNOWLEDGMENTS

Author

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 28

371

This work was supported by NSFC (No. 51472194), National Program on Key Basic Research

372

Project of China (973 Program) 2013CB632402 and the NSF of Hubei Province (2016CFA078).

ACS Paragon Plus Environment

20

Page 21 of 28

374 375

Environmental Science & Technology

REFERENCES (1)

Zhang, D.; Lee, C.; Javed, H.; Yu, P.; Kim, J. H.; Alvarez, P. J. J. Easily recoverable,

376

micrometer-sized TiO2 hierarchical spheres decorated with cyclodextrin for enhanced

377

photocatalytic degradation of organic micropollutants. Environ. Sci. Technol. 2018, 52 (21),

378

12402-12411; DOI 10.1021/acs.est.8b04301.

379

(2)

Jo, Y.; Kim, C.; Moon, G. H.; Lee, J.; An, T.; Choi, W. Activation of peroxymonosulfate

380

on visible light irradiated TiO2 via a charge transfer complex path. Chem. Eng. J. 2018, 346,

381

249-257; DOI 10.1016/j.cej.2018.03.150.

382

(3)

Ma, J.; Ma, W.; Chen, C.; Ji, H.; Zhao, J. An efficient anthraquinone-resin hybrid co-

383

catalyst for Fenton-like reactions: acceleration of the Iron cycle using a quinone cycle under

384

visible-light irradiation. Chem. - Asian J. 2011, 6 (9), 2264-2268; DOI 10.1002/asia.201100347.

385

(4)

Tai, C.; Zhang, S.; Yin, Y.; Dai, Z.; Li, Y.; Jiang, G.; Cai, Y.; Huang, C.; Shi, J. Facile

386

photoinduced generation of hydroxyl radical on a nitrocellulose membrane surface and its

387

application in the degradation of organic pollutants. ChemSusChem 2018, 11 (5), 843-847; DOI

388

10.1002/cssc.201800047.

389

(5)

Choi, Y.; Koo, M. S.; Bokare, A. D.; Kim, D. H.; Bahnemann, D. W.; Choi, W.

390

Sequential process combination of photocatalytic oxidation and dark reduction for the removal of

391

organic pollutants and Cr(VI) using Ag/TiO2. Environ. Sci. Technol. 2017, 51 (7), 3973-3981;

392

DOI 10.1021/acs.est.6b06303.

ACS Paragon Plus Environment

21

Environmental Science & Technology

393

(6)

Page 22 of 28

Sun, J.; Mao, J. D.; Gong, H.; Lan, Y. Fe(III) photocatalytic reduction of Cr(VI) by low-

394

molecular-weight organic acids with α-OH. J. Hazard. Mater. 2009, 168 (2-3), 1569-1574; DOI

395

10.1016/j.jhazmat.2009.03.049.

396

(7)

Guo, S.; Yang, Z.; Wen, Z.; Fida, H.; Zhang, G.; Chen, J. Reutilization of iron sludge as

397

heterogeneous Fenton catalyst for the degradation of rhodamine B: role of sulfur and

398

mesoporous

399

10.1016/j.jcis.2018.08.005.

400

(8)

structure.

J.

Colloid

Interface

Sci.

2018,

532,

441-448;

DOI

Xing, M.; Xu, W.; Dong, C.; Bai, Y.; Zeng, J.; Zhou, Y.; Zhang, J.; Yin, Y. Metal

401

sulfides as excellent co-catalysts for H2O2 decomposition in advanced oxidation processes. Chem

402

2018, 4 (6), 1359-1372; DOI 10.1016/j.chempr.2018.03.002.

403

(9)

Huang, G. X.; Wang, C. Y.; Yang, C. W.; Guo, P. C.; Yu, H. Q. Degradation of

404

Bisphenol A by peroxymonosulfate catalytically activated with Mn1.8Fe1.2O4 nanospheres:

405

synergism between Mn and Fe. Environ. Sci. Technol. 2017, 51 (21), 12611-12618; DOI

406

10.1021/acs.est.7b03007.

407

(10) Mu, Y.; Ai, Z.; Zhang, L. Phosphate shifted oxygen reduction pathway on Fe@Fe2O3

408

core–shell nanowires for enhanced reactive oxygen species generation and aerobic 4-

409

chlorophenol degradation. Environ. Sci. Technol. 2018, 51 (14), 8101-8109; DOI

410

10.1021/acs.est.7b01896.

411

(11) Zou, H.; Hu, E.; Yang, S.; Gong, L.; He, F. Chromium(VI) removal by

412

mechanochemically sulfidated zero valent iron and its effect on dechlorination of trichloroethene

413

as

414

10.1016/j.scitotenv.2018.09.003.

a

co-contaminant.

Sci.

Total

Environ.

2019,

650,

419-426;

DOI

ACS Paragon Plus Environment

22

Page 23 of 28

Environmental Science & Technology

415

(12) Jin, M.; Long, M.; Su, H.; Pan, Y.; Zhang, Q.; Wang, J.; Zhou, B.; Zhang, Y.

416

Magnetically separable maghemite/montmorillonite composite as an efficient heterogeneous

417

Fenton-like catalyst for phenol degradation. Environ. Sci. Pollut. Res. 2017, 24 (2), 1926-1937;

418

DOI 10.1007/s11356-016-7866-8.

419

(13) Dai, C.; Tian, X.; Nie, Y.; Lin, H. M.; Yang, C.; Han, B.; Wang, Y. Surface facet of

420

CuFeO2 nanocatalyst: a key parameter for H2O2 activation in fenton-like reaction and organic

421

pollutant

422

10.1021/acs.est.8b01448.

degradation.

Environ.

Sci.

Technol.

2018,

52

(11),

6518-6525;

DOI

423

(14) Qian, X.; Ren, M.; Zhu, Y.; Yue, D.; Han, Y.; Jia, J.; Zhao, Y. Visible light assisted

424

heterogeneous Fenton-like degradation of organic pollutant via α-FeOOH/mesoporous carbon

425

composites. Environ. Sci. Technol. 2017, 51 (7), 3993-4000; DOI 10.1021/acs.est.6b06429.

426

(15) Dong, C.; Ji, J.; Shen, B.; Xing, M.; Zhang, J. Enhancement of H2O2 decomposition by

427

the co-catalytic effect of WS2 on the Fenton reaction for the synchronous reduction of Cr(VI)

428

and remediation of phenol. Environ. Sci. Technol. 2018, 52 (19), 11297-11308; DOI

429

10.1021/acs.est.8b02403.

430

(16) Zhou, T.; Wu, X.; Mao, J.; Zhang, Y.; Lim, T. T. Rapid degradation of sulfonamides in a

431

novel

heterogeneous

sonophotochemical

432

(US/UV/Fe3O4/oxalate)

433

10.1016/j.apcatb.2014.05.036.

system.

Appl.

Catal.,

magnetite-catalyzed B

2014,

160-161,

Fenton-like 325-334;

DOI

434

(17) Liu, S. Q.; Feng, L. R.; Xu, N.; Chen, Z. G.; Wang, X. M. Magnetic nickel ferrite as a

435

heterogeneous photo-Fenton catalyst for the degradation of rhodamine B in the presence of

436

oxalic acid. Chem. Eng. J. 2012, 203, 432-439; DOI 10.1016/j.cej.2012.07.071.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 28

437

(18) Liu, Z. J.; Liu, W.; Wang, Y.; Guo, M. L. Preparation of β-ferrous oxalate dihydrate

438

layered nanosheets by mechanochemical method and its visible-light-driven photocatalytic

439

performance. Mater. Lett. 2016, 178, 83-86; DOI 10.1016/j.matlet.2016.04.201.

440

(19) Kwan, C. Y.; Chu, W. Effect of ferrioxalate-exchanged resin on the removal of 2,4-D by

441

a photocatalytic process. J. Mol. Catal. A: Chem. 2006, 255 (1-2), 236-242; DOI

442

10.1016/j.molcata.2006.03.036.

443

(20) Lei, J.; Liu, C.; Li, F.; Li, X.; Zhou, S.; Liu, T.; Gu, M.; Wu, Q. Photodegradation of

444

orange I in the heterogeneous iron oxide–oxalate complex system under UVA irradiation. J.

445

Hazard. Mater. 2006, 137 (2), 1016-1024; DOI 10.1016/j.jhazmat.2006.03.028.

446

(21) Lan, Q.; Li, F.; Liu, C.; Li, X. Z. Heterogeneous photodegradation of pentachlorophenol

447

with maghemite and oxalate under UV illumination. Environ. Sci. Technol. 2008, 42 (21), 7918-

448

7923; DOI 10.1021/es801220n.

449

(22) Mazellier, P.; Sulzberger, B. Diuron degradation in irradiated, heterogeneous iron/oxalate

450

systems: the rate-determining step. Environ. Sci. Technol. 2001, 35 (16), 3314-3320; DOI

451

10.1021/es001324q.

452

(23) Wei, S.; Ren, H.; Li, J.; Shi, J.; Shao, Z. Decolorization of organic dyes by zero-valent

453

iron in the presence of oxalic acid and influence of photoirradiation and hexavalent chromium. J.

454

Mol. Catal. A: Chem. 2013, 379, 309-314; DOI 10.1016/j.molcata.2013.09.002.

455

(24) Li, X.; Chen, W.; Li, L. Catalytic ozonation of oxalic acid in the presence of Fe2O3-

456

loaded

activated

carbon.

Ozone:

457

10.1080/01919512.2018.1462142.

Sci.

Eng.

2018,

40

(6),

448-456;

DOI

ACS Paragon Plus Environment

24

Page 25 of 28

Environmental Science & Technology

458

(25) Lan, Q.; Li, F. B.; Sun, C. X.; Liu, C. S.; Li, X. Z. Heterogeneous photodegradation of

459

pentachlorophenol and iron cycling with goethite, hematite and oxalate under UVA illumination.

460

J. Hazard. Mater. 2010, 174 (1-3), 64-70; DOI 10.1016/j.jhazmat.2009.09.017.

461

(26) Chen, N.; Shang, H.; Tao, S.; Wang, X.; Zhan, G.; Li, H.; Ai, Z.; Yang, J.; Zhang, L.

462

Visible light driven organic pollutants degradation with hydrothermally carbonized sewage

463

sludge and oxalate via molecular oxygen activation. Environ. Sci. Technol. 2018, 52 (21), 12656-

464

12666; DOI 10.1021/acs.est.8b03882.

465

(27) Pang, H.; Zhang, Q.; Wang, H.; Cai, D.; Ma, Y.; Li, L.; Li, K.; Lu, X.; Chen, H.; Yang,

466

X.; Chen, J. Photochemical aging of guaiacol by Fe(III)–oxalate complexes in atmospheric

467

aqueous phase. Environ. Sci. Technol. 2019, 53 (1), 127-136; DOI 10.1021/acs.est.8b04507.

468

(28) Huang, X. P.; Hou, X. J.; Wang, F.; Guo, B. H.; Song, F. H.; Ling, L.; Zhao, J. C.; Zhang,

469

L. Z. Molecular-scale structures of uranyl surface complexes on hematite facets. Environ. Sci.:

470

Nano, 2019, 6, 892-903; DOI 10.1039/c8en00899j.

471 472

(29) Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59 (3), 1758-1775; DOI 10.1103/physrevb.59.1758.

473

(30) Kresse, G.; Furthmüller, J. Efficient iterative schemes forab initiototal-energy

474

calculations using a plane-wave basis set. Phys. Rev. B 1996, 54 (16), 11169-11186; DOI

475

10.1103/physrevb.54.11169.

476

(31) Gao, Y.; Zhao, J.; Zhu, Y.; Ma, S.; Su, X.; Wang, Z. Wet chemical process of rod-like

477

tungsten nanopowders with iron (II) as reductive agent. Mater. Lett. 2006, 60 (29-30), 3903-

478

3905; DOI 10.1016/j.matlet.2006.03.137.

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 28

479

(32) Zhang, J.; Zhang, Y.; Yan, J. Y.; Li, S. K.; Wang, H. S.; Huang, F. Z.; Shen, Y. H.; Xie,

480

A. J. A novel synthesis of star-like FeWO4 nanocrystals via a biomolecule-assisted route. J.

481

Nanopart. Res. 2012, 14 (4), 1-10; DOI 10.1007/s11051-012-0796-6.

482

(33) Buvaneswari, K.; Karthiga, R.; Kavitha, B.; Rajarajan, M.; Suganthi, A. Effect of FeWO4

483

doping on the photocatalytic activity of ZnO under visible light irradiation. Appl. Surf. Sci. 2015,

484

356, 333-340; DOI 10.1016/j.apsusc.2015.08.060.

485

(34) He, G. L.; Chen, M. J.; Liu, Y. Q.; Li, X.; Liu, Y. J.; Xu, Y. H. Hydrothermal synthesis of

486

FeWO4-graphene composites and their photocatalytic activities under visible light. Appl. Surf.

487

Sci. 2015, 351, 474-479; DOI 10.1016/j.apsusc.2015.05.159.

488

(35) Sadiq, M. M. J.; Shenoy, U. S.; Bhat, D. K. Enhanced photocatalytic performance of N-

489

doped RGO-FeWO4 /Fe3O4 ternary nanocomposite in environmental applications. Materials

490

Today Chemistry 2017, 4, 133-141; DOI 10.1016/j.mtchem.2017.04.003.

491

(36) Rajagopal, S.; Nataraj, D.; Khyzhun, O. Y.; Djaoued, Y.; Robichaud, J.; Mangalaraj, D.

492

Hydrothermal synthesis and electronic properties of FeWO4 and CoWO4 nanostructures. J.

493

Alloys Compd. 2010, 493 (1-2), 340-345; DOI 10.1016/j.jallcom.2009.12.099.

494

(37) Li, J.; Wu, X.; Wan, Z.; Chen, H.; Zhang, G. Full spectrum light driven photocatalytic in-

495

situ epitaxy of one-unit-cell Bi2O2CO3 layers on Bi2O4 nanocrystals for highly efficient

496

photocatalysis and mechanism unveiling. Appl. Catal., B 2019, 243, 667-677; DOI

497

10.1016/j.apcatb.2018.10.067.

ACS Paragon Plus Environment

26

Page 27 of 28

Environmental Science & Technology

498

(38) Li, J.; Wang, J.; Zhang, G.; Li, Y.; Wang, K. Enhanced molecular oxygen activation of

499

Ni2+-doped BiO2-x nanosheets under UV, visible and near-infrared irradiation: mechanism and

500

DFT study. Appl. Catal., B 2018, 234, 167-177; DOI 10.1016/j.apcatb.2018.04.016.

501

(39) Huang, X. P.; Hou, X. J.; Wang, F.; Guo, B. H.; Song, F. H.; Ling, L.; Zhao, J. C.; Zhang,

502

L. Z. Molecular-scale structures of uranyl surface complexes on hematite facets. Environ. Sci.:

503

Nano, 2019, 6, 892-903; DOI 10.1039/c8en00899j.

504

(40) Xu, T. Y.; Zhu, R. L.; Shang, H.; Xia, Y. B.; Liu, X.; Zhang, L. Z. Photochemical

505

behavior of ferrihydrite-oxalate system: Interfacial reaction mechanism and charge transfer

506

process. Water Res. 2019, 159, 10-19.10.1016/j.watres.2019.04.055.

507

(41) Jeong, J.; Yoon, J. Dual roles of CO2- for degrading synthetic organic chemicals in the

508

photo/ferrioxalate

system.

509

10.1016/j.watres.2004.05.016.

Water

Research

2004,

38

(16),

3531-3540;

DOI

510

(42) Zhou, T.; Wu, X.; Zhang, Y.; Li, J.; Lim, T. T. Synergistic catalytic degradation of

511

antibiotic sulfamethazine in a heterogeneous sonophotolytic goethite/oxalate Fenton-like system.

512

Appl. Catal., B 2013, 136-137, 294-301; DOI 10.1016/j.apcatb.2013.02.004.

ACS Paragon Plus Environment

27

Environmental Science & Technology

514

Page 28 of 28

TOC GRAPHIC

515

ACS Paragon Plus Environment

28