Comparative Analysis of Photocatalytic and ... - ACS Publications

Jul 8, 2019 - P.O. Box 217, 7500 AE Enschede, The Netherlands. ‡. Department .... reactor64 equipped with eight Philips UV lamps (TL-D 18W. BLB, λm...
0 downloads 0 Views 459KB Size
Full text access provided via ACS AuthorChoice

Environmental Processes

A comparative analysis of photocatalytic and electrochemical degradation of 4-Ethylphenol in saline conditions Robert Bruninghoff, Alyssa van Duijne, Lucas Braakhuis, Pradip Saha, Adriaan W Jeremiasse, Bastian Mei, and Guido Mul Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.9b01244 • Publication Date (Web): 08 Jul 2019 Downloaded from pubs.acs.org on July 21, 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 38

Environmental Science & Technology

1

A comparative analysis of photocatalytic and

2

electrochemical degradation of 4-Ethylphenol in

3

saline conditions

4

Robert Brüninghoff †, Alyssa K. van Duijne †, Lucas Braakhuis †, Pradip Saha ‡, Adriaan

5

W. Jeremiasse †,‡, Bastian Mei †, Guido Mul †,*

6

† PhotoCatalytic Synthesis Group, MESA+ Institute for Nanotechnology, Faculty of

7

Science and Technology, University of Twente, P.O. Box 217, 7500 AE Enschede, The

8

Netherlands.

9

‡ Department of Environmental Technology, Wageningen University and Research, P.O. Box

10

17, 6700 AA Wageningen, The Netherlands

11

†,‡ MAGNETO special anodes B.V. (an Evoqua brand), Calandstraat 109, 3125 BA Schiedam,

12

The Netherlands

13 14

* Corresponding author: Guido Mul

15

E-mail: [email protected]

16

Phone:+31534893890 1 ACS Paragon Plus Environment

Environmental Science & Technology

17

ABSTRACT

18

We evaluated electrochemical degradation (ECD) and photocatalytic degradation (PCD)

19

technologies for saline water purification, with focus on rate comparison and formation and

20

degradation of chlorinated aromatic intermediates using the same non-chlorinated parent

21

compound, 4-ethylphenol (4EP). At 15 mA.cm-2, and in the absence of chloride (0.6 mol.L-1

22

NaNO3 was used as supporting electrolyte), ECD resulted in an apparent zero-order rate of

23

30 μmol.L-1.h-1, whereas rates of ~300 μmol.L-1.h-1 and ~3750 μmol.L-1.h-1 were computed for

24

low (0.03 mol.L-1) and high (0.6 mol.L-1) NaCl concentration, respectively. For PCD, initial

25

rates of ~330 μmol.L-1.h-1 and 205 μmol.L-1.h-1 were found for low and high NaCl

26

concentrations, at a photocatalyst (TiO2) concentration of 0.5 g.L-1, and illumination at

27

λmax ≈ 375 nm, with an intensity ≈ 0.32 mW*cm-2. In the chlorine mediated ECD approach,

28

significant quantities of free chlorine (hypochlorite, Cl2) and chlorinated hydrocarbons were

29

formed in solution, while photocatalytic degradation did not show the formation of free

30

chlorine, nor chlorine-containing intermediates, and resulted in better removal of non-purgeable

31

hydrocarbons than ECD. The origin of the minimal formation of free chlorine and chlorinated

32

compounds in photocatalytic degradation is discussed based on photoelectrochemical results

33

and existing literature, and explained by a chloride-mediated surface-charge recombination

34

mechanism.

35 36

2 ACS Paragon Plus Environment

Page 2 of 38

Page 3 of 38

Environmental Science & Technology

37

INTRODUCTION

38

Water pollution is one of the greatest challenges of modern society.1,2 The treatment of polluted

39

water to enable its re-use is essential,3 while efficient removal of (recalcitrant) organic

40

pollutants, present in low concentrations, requires development of innovative technology.4–7

41

Particularly difficult is treatment of industrial effluents containing high amounts of sodium

42

chloride, which impedes biological treatment due to negative effects of the NaCl on the

43

microbial flora such as plasmolysis.8,9

44

Advanced oxidation processes (AOP) such as photocatalytic degradation (PCD) or

45

electrochemical degradation (ECD) are promising methodologies for the removal of

46

(recalcitrant) organics in saline conditions. Highly reactive oxygen species (ROS), e.g.

47

hydroxyl radicals (●OH) are generated oxidatively (anodically) in both methodologies.10

48

Sodium chloride, ubiquitous in many industrial wastewaters,8 is known to further enhance

49

degradation rates in ECD, due to (anodic) chloride oxidation and the consecutive formation of

50

reactive chlorine species (RCS: e.g. Cl●, Cl2, or HOCl).11–15 However, the major disadvantage

51

of ECD in saline media is the formation of chlorinated by-products.12,16–20 These by-products

52

typically induce high water toxicity, and thus deteriorate the environment.21–23

53

For TiO2-based photocatalytic degradation, chloride is usually considered to be an inhibitor.

54

Several mechanisms have been proposed to explain the inhibition: scavenging of holes or OH-

55

radicals by chloride ions;24–27 blocking of active surfaces sites by chloride ions;25,28–32,

56

formation of an inorganic salt layer;33 or chloride acting as surface-charge-recombination-

57

center for photogenerated charge carriers.34 Often, more than one mechanism is used to explain

58

the inhibiting effect.35–40 In addition, aggregation of TiO2 particles has been reported to reduce

59

the number of absorbed photons influencing degradation rates.41 While the inhibiting effect of

60

chloride, using TiO2 as photocatalyst, has been discussed frequently in the literature, detailed

61

studies regarding the formation of reactive chlorine species (RCS), such as hypochlorite, and 3 ACS Paragon Plus Environment

Environmental Science & Technology

62

chlorinated intermediates in aqueous solution (starting from a non-chlorinated parent molecule)

63

are rare, and obtained results seem to be inconsistent. For example, negligible amounts of

64

organochloride compounds have been reported during phenol degradation in high saline media

65

[50 g.L-1 NaCl],38,42,43 which appears in disagreement with detection of several chlorinated

66

intermediates arising from the photocatalytic degradation of the azo dye Acid Orange 7 at

67

moderate salinity [5.8-11.8 g.L-1].39

68

The role of chloride ions in TiO2-based PCD of organic pollutants is thus not fully

69

understood44,45 and requires further studies.46 In particular, the formation of toxic by-products

70

needs attention, to reveal the environmental impact of AOPs.4,20,47–52

71

Although some comparative studies between different AOPs, or a combination of technologies

72

have been reported,53–62 to the best of our knowledge a comparative study specifically

73

addressing the concentration dependent formation of chlorinated intermediates during ECD,

74

PCD and photoelectrochemical degradation (PECD), starting from a non-chlorinated parent

75

compound in saline solutions, has not yet been performed.

76

In this study the influence of NaCl on the degradation rate and mechanism of the model

77

compound 4-ethylphenol (4EP) is investigated. Two salt concentrations, corresponding to

78

brackish (1.75 g.L-1) and sea water salt concentrations (35 g.L-1), covering a relevant

79

concentration range of industrial effluents,8 have been studied. TiO2 particle suspensions or

80

TiO2 thin films illuminated by UV light are used for PCD and PECD, respectively. A platinized

81

Ti (Ti/Pt) anode was utilized for ECD, and the performance compared to a Boron-Doped-

82

Diamond (BDD) electrode.

83

The obtained results clearly demonstrate that the rate of degradation of 4EP in ECD & PECD

84

is mainly the result of chemical chlorination, since free chlorine and various chlorinated

85

hydrocarbons were detected. In contrast, the formation of chlorinated compounds was 4 ACS Paragon Plus Environment

Page 4 of 38

Page 5 of 38

Environmental Science & Technology

86

significantly smaller in PCD, and the overall removal effective, despite an inhibiting effect of

87

chloride.

88

MATERIALS AND METHODS

89

Chemical and Reagents

90

The following chemicals were used in this study: titanium dioxide (Hombikat UV 100,

91

Sachtleben (Venator)), 4-Ethylphenol (Sigma Aldrich 99%), sodium chloride (Sigma Aldrich

92

≥ 99%), sodium nitrate (Sigma Aldrich ≥ 99.0%), demineralised water (Merck MilliQ system,

93

resistivity > 18MΩ*cm), water (LC-MS grade, Biosolve BV), acetonitrile (LC-MS grade,

94

Biosolve BV), formic acid (98-100%, LC-MS grade, Merck), 1-(4-Hydroxylphenyl)ethanol

95

(synthesis

96

4-Hydroxybenzaldehyde (Sigma Aldrich 98%), 4-Hydroxyacetophenone (Sigma Aldrich

97

99%), 4-Ethylresorcinol (Alfa Aesar 98%), 4-Ethylcatechol (Sigma Aldrich 95%), and

98

2-Chloro-4-ethylphenol (AKos GmbH, > 95%), 2,6-Dichloro-4-ethylphenol (AKos GmbH,

99

> 90%).

see

SI),

2-(4-Hydroxylphenyl)ethanol

(Sigma

Aldrich

98%),

100

Degradation experiments

101

If not otherwise stated all experiments were performed in solutions containing NaCl or NaNO3

102

at various concentrations (0.03 mol.L-1 or 0.6 mol.L-1). Sodium nitrate (NaNO3) was used for

103

comparison, since NaNO3 is supposed to be an “inert” water additive in oxidative processes.

104

Photocatalytic degradation

105

Pre-treatment of TiO2 photocatalyst, via annealing at 600 °C for 4 h (named in the following

106

H600), was performed to achieve an improved photocatalytic performance (optimum between

107

adsorbed surface OH groups and available holes at the surface), as reported in earlier work by

108

our group.63 The presence of the anatase phase was confirmed by XRD analysis (see XRD

109

patterns in Figure S1). 5 ACS Paragon Plus Environment

Environmental Science & Technology

110

4EP belongs to the group of alkylphenols, which have been found in process-water from the

111

petrochemical industry65, or as degradation product from alkylphenol ethoxylates (used in the

112

chemical industry).66 These compounds are suspected to induce endocrine disruptive effects,

113

and thus pose a health risk.67 In this study 4EP was used as model compound offering several

114

experimental advantages, for example a moderate toxicity and an adequate water solubility.

115

Moreover, the structure consists of three functional groups (benzene ring, alcohol group, alkyl

116

chain) representing important functional groups of more complex compounds found in

117

wastewater5,6 and is suitable to study the selectivity in reactions with chlorine.68,69

118

PCD experiments were performed as described in the following: 4EP stock solution (approx.

119

50 mg.L-1) was pre-saturated with air for 20 min. The catalyst H600 [25(+/- 1) mg], the required

120

amount of salt and 50 mL stock solution were mixed in a quartz glass beaker, covered with a

121

quartz glass lid and stirred (350 rpm) in the dark for 30 min before the first sample was taken

122

(0 h measurement). UV-irradiation of the suspension was achieved using a custom-made closed

123

box reactor64 equipped with eight Philips UV lamps (TL-D 18W BLB, λmax ≈ 375 nm, intensity

124

≈ 0.32 mW.cm-2). Any influence of UV light absorption by the salt solutions on the PCD rate

125

was excluded (Figure S2)33. Liquid samples were taken after 0.5 h, 1 h, 2 h, 3 h, 4 h and 5 h of

126

continuous illumination. Each time approx. 1.5 mL solution was taken. Afterwards, the samples

127

were filtered using a Phenex RC membrane filter (0.2 µm; Phenomenex), and analysed by

128

HPLC and LC-MS.

129

Electrochemical degradation

130

ECD of 4EP was performed in a custom-made single-compartment cell (see scheme in

131

Figure S3). A platinized Ti plate electrode or BDD electrode (Magneto special anodes B.V;

132

geometric surface area 3.14cm2) was used to study the electrochemical degradation of 4EP. A

133

Pt mesh and a Ag/AgCl electrode (3 M NaCl, BASi) were used as a counter electrode and as a

134

reference electrode, respectively. The reactor was filled with 70 mL 4EP stock solution (pre6 ACS Paragon Plus Environment

Page 6 of 38

Page 7 of 38

Environmental Science & Technology

135

saturated with air). Degradation experiments were carried out under galvanostatic conditions at

136

15 mA.cm-2 (VersaSTAT4 potentiostat, PAR or biologic SP-300/VMP3). The electrolyte was

137

constantly stirred during the experiments. Sampling was performed as described for PCD. For

138

high NaCl concentration, samples were also taken after 1 min, 2.5 min, 5 min, 10 min, 15 min,

139

20 min and 30 min. The cell was not purged with air, and the cathodic reaction is thus likely

140

the formation of hydrogen. The surface area of the Pt Gauze was significantly larger than the

141

applied anode, and therefore the cathodic reaction was not limiting the degradation of

142

hydrocarbons (See Figure S3).

143

To evaluate reactivity of 4EP with reactive chlorine species, 9.5 mL of 4EP stock solution

144

[50mg/L] was mixed with sodium hypochlorite solution (0.5 mL, equal to approx. 20 mg.L-1

145

free chlorine) and treated for 15 min. Filtration was followed by LC MS analysis.

146

Photoelectrochemical degradation

147

PECD experiments were carried out in a custom-made Teflon based PEC-cell. The reactor was

148

equipped with a thin film of TiO2 on a Ti-electrode (for the preparation see SI; geometric surface

149

area 3.14 cm2 [TiO2/Ti]) or H600 coated on FTO ([H600/FTO], 2.54 cm2]) and a Pt-mesh

150

counter electrode. The reactor was filled with the 4EP stock solution containing 0.03 mol.L-1

151

NaCl, and pretreated by purging air. Experiments were carried out under potentiostatic

152

conditions at 0.1 V vs. Ag/AgCl reference electrode (3 M NaCl, BASi) using a potentiostat

153

(VersaStat3, PAR or biologic SP-300). The electrolyte was constantly stirred and illuminated

154

with a 365 nm LED [for TiO2/Ti experiments] or 375 nm LED [for H600/FTO experiments]

155

(Roithner Lasertechnik GmbH, APG2C1-365/375-E, 200 mW output power) placed in front of

156

the glass window of the PEC-cell.

157

Analytical methods

158

HPLC analysis was performed by a ThermoFisher Scientific Dionex Ultimate 3000 HPLC

159

system equipped with a UV detector and a reversed phase Luna Omega polar C18 column 7 ACS Paragon Plus Environment

Environmental Science & Technology

160

(3 µm, 150 x 2.1 mm, protected by a polar C18 security guard column (both Phenomenex)).

161

The sample injection volume was 5 µL. The oven temperature was set to 30 °C. A water ([A];

162

0.1% formic acid) - acetonitrile (0.1% formic acid) gradient system was used at a flow rate of

163

0.200 mL/min and a total run time of 60 min. (0 min 100% A, 1 min 100% A, 30 min 50% A,

164

35 min 50% A, 45 min 5% A, 50 min 5% A, 55 min 100% A, 60 min 100% A). For detection,

165

UV absorption was measured at 276 nm. For the quantification, an external standard calibration

166

was applied (for more information see SI Figure S4 & Table S1-3).

167

For the identification of intermediates, a Bruker amaZon SL ion trap using electrospray

168

ionization (conditions: capillary 4500 V, end plate offset 500 V, 15 psi nebulizer gas pressure

169

(N2), 8 L/min dry gas flow (N2), 200 °C dry gas temperature, scan between 50 to 1000 m/z) was

170

coupled to the HPLC system. A divert-valve was used for salt separation. Ionization was

171

performed both in positive and negative mode under alternating conditions using automatic

172

fragmentation for data-dependent acquisition of both MS and LC MS data.

173

Free chlorine ((FC); total concentration of Cl2 + HOCl + OCl-) measurements were performed

174

in adaption of the EPA-DPD method 330.5, using a Hanna Instruments free chlorine test kit

175

(HI93701) and photometer (HI83099) for a concentration range of 0.00 to 2.50 mg.L-1

176

[accuracy +/- 0.03 mg.L-1]. For higher free chlorine concentrations a Hanna Instruments test kit

177

(HI95771) and photometer (HI96771) were used [accuracy +/- 2 mg.L-1].

178

Non-purgeable organic carbon (NPOC) content was determined with a SHIMADZU total

179

organic carbon analyser, TOC-L CPH/CPN (see SI for more information)

180

8 ACS Paragon Plus Environment

Page 8 of 38

Page 9 of 38

Environmental Science & Technology

181

RESULTS AND DISCUSSION

182

Electrochemical degradation (ECD) of 4EP

183

Degradation curves of 4EP obtained for ECD at various salt conditions (NaNO3 or NaCl

184

containing electrolytes), at constant current density of j = 15 mA.cm-2, are shown in Figure 1.

185

A linear degradation behavior was obtained for both NaNO3 concentrations (Figure 1a), and

186

zero-order kinetic constants can be estimated to be ~20 μmol.L-1.h-1 and 30 μmol.L-1.h-1 for the

187

low or high salt concentrations, respectively (voltage-time curves are shown in Figure S5 and

188

quantified numbers are presented in Table S4).

189

In contrast to NaNO3, which cannot be oxidized to reactive radicals, the degradation of 4EP in

190

chloride-containing media proceeded much faster (Figure 1b). Again assuming linear

191

regression (zero-order), constants of ~300 μmol.L-1.h-1 and ~3750 μmol.L-1.h-1 can be

192

computed for low and high NaCl concentration, respectively. The fast 4EP degradation in the

193

NaCl experiments can be explained by chlorine mediated electrochemical oxidation, where free

194

chlorine (i.e. hypochlorite) and chlorine radicals support the degradation of organic

195

pollutants.11–15 In agreement with this hypothesis, the measured concentrations of FC amounted

196

to 0.95 mg.L-1 (after 5 h at 0.03 mol.L-1 NaCl) and 111 mg.L-1 (after 30 min at 0.6 mol.L-1).

4EP [0.03 mol*L-1 NaCl]

400

concentration [mol*L-1]

concentration [mol*L-1]

400

300

200

100

4EP [0.03 mol*L-1 NaNO3]

4EP [0.6 mol*L-1 NaCl]

300

200

100

4EP [0.6 mol*L-1 NaNO3] 0

0 0

1

2

3

4

5

0

1

2

3 time (h)

4

197

a)

198

Figure 1. Electrochemical degradation of 4EP at various salt conditions: a) 0.03 mol.L-1

199

NaNO3 (light blue), 0.6 mol.L-1 NaNO3 (dark blue); b) 0.03 mol.L-1 NaCl (light green),

time (h)

b)

9 ACS Paragon Plus Environment

5

Environmental Science & Technology

200

0.6 mol.L-1 NaCl (dark green). ECD was performed at a Ti/Pt working electrode at constant

201

current density of 15 mA.cm-2.

202

Theoretically an electric charge of 110 Coulomb is required for full mineralization of 4EP,

203

corresponding to approx. 40 min. of continuous treatment at the applied current density of

204

j = 15 mA.cm-2.

205

The presence of chlorine radicals was found to induce the formation of undesired chlorinated

206

organic intermediates and by-products (RCl), in agreement with literature.12,16–20 LC-MS

207

analysis resulted in the identification of 2-chloro-4-ethylphenol (2C4EP) and 2,6-dichloro-4-

208

ethylphenol (26DC4EP), see Figure 2. The same intermediates were also formed during

209

chemical chlorination of 4EP with sodium hypochlorite solution (20 mg.L-1 FC), showing that

210

conversion of 4EP in saline conditions resembles chemical chlorination. The chlorination of the

211

aromatic ring at the ortho position is directed by the adjacent OH-group and is a typical reaction

212

of phenolic compounds with electrophilic oxidants.70,71 Furthermore, the identified

213

intermediates are in good agreement with other literature reports on phenol and bisphenol

214

conversion, using degradation at BDD electrodes at low salt concentrations16,18,19 and the

215

chemical chlorination of methylphenols.72

216

217 218

Figure 2. Reaction path of 4EP with RCS (mainly hypochlorite) during ECD at 15 mA.cm-2

219

at a Ti/Pt electrode or BDD electrode in NaCl media (0.03 mol.L-1 NaCl & 0.6 mol.L-1 NaCl). 10 ACS Paragon Plus Environment

Page 10 of 38

Page 11 of 38

Environmental Science & Technology

220

2-chloro-4-ethylphenol (2C4EP) and 2,6-dichloro-4-ethylphenol (26DC4EP) were identified

221

as main aromatic intermediates.

222

In order to further understand the subsequent degradation process of the two intermediates the

223

concentration vs time profiles were determined (Figure 3). Obviously, 4EP is converted into

224

2C4EP almost instantly after the start of the experiment. For the low NaCl concentration

225

(Figure 3a) a maximum of 2C4EP formation is obtained after 1 - 1.5 h of treatment time, with

226

2C4EP already exceeding the concentration of 4EP. The measured concentration of

227

149 μmol.L-1 [23.3 mg.L-1] of 2C4EP corresponds to a molar fraction of 34% relative to the

228

initial molar concentration of 4EP (for more information see also Figure S6).

11 ACS Paragon Plus Environment

Environmental Science & Technology

4EP [0.03 mol*L-1 NaCl]_Pt 2C4EP [0.03 mol*L-1 NaCl]_Pt 26DC4EP [0.03 mol*L-1 NaCl]_Pt

300 34 mol% [23.3 mg*L-1]

200

100

4EP [0.6 mol*L-1 NaCl]_Pt 2C4EP [0.6 mol*L-1 NaCl]_Pt 26DC4EP [0.6 mol*L-1 NaCl]_Pt

400 concentration (mol*L-1)

concentration (mol*L-1)

400

10 mol% [8.6 mg*L-1]

300

200 21 mol% [14.1 mg*L-1] 100

7 mol% [5.7 mg*L-1]

0

0 0

1

2

3

4

0

5

time (h)

5

-1

4EP [0.03 mol*L NaCl]_BDD 2C4EP [0.03 mol*L-1 NaCl]_BDD 26DC4EP [0.03 mol*L-1 NaCl]_BDD

300 35 mol% [24.0 mg*L-1]

200

13 mol% [10.5 mg*L-1]

100

15 20 time (min)

25

30

4EP [0.6 mol*L-1 NaCl]_BDD 2C4EP [0.6 mol*L-1 NaCl]_BDD 26DC4EP [0.6 mol*L-1 NaCl]_BDD

400 concentration (mol*L-1)

concentration (mol*L-1)

400

10

b)

a)

300

200

24 mol% [16.4 mg*L1] 8 mol% [7.1 mg*L1]

100

0

0 0

c)

Page 12 of 38

1

2

3

4

0

5

time (h)

5

10

15 20 time (min)

25

30

d)

229

Figure 3. Formation and degradation of 2C4EP (red) and 26DC4EP (blue) formed during

230

ECD of 4EP (black) in 0.03 mol.L-1 NaCl (a & c) and 0.6 mol.L-1 NaCl solution (b & d)

231

using Ti/Pt (a & b) or BDD electrode (c & d).

232

26DC4EP is already formed during accumulation of 2C4EP indicating a continuous

233

consecutive transformation of 4EP into 2C4EP and 26DC4EP (Figure 2). Again these results

234

are in agreement with the chemical chlorination of phenol70 and recent studies on phenol

235

degradation at a BDD electrode in low saline solution.16 A concentration maximum of

236

45 μmol.L-1 [8.6 mg.L-1] of 26DC4EP corresponding to a molar fraction of 10% relative to the

237

initial molar concentration of 4EP was obtained after 2 h. It took as long as 4 h of ECD at low

238

NaCl concentration, to remove both intermediates completely from solution (below the limit of

239

detection (LOD), see Table S3 for more information).

12 ACS Paragon Plus Environment

Page 13 of 38

Environmental Science & Technology

240

Similar concentration profiles were obtained for the high salt conditions, however within

241

shorter reaction times (Figure 3b). Already after 5 min., 2C4EP reached its maximum

242

concentration (90 μmol.L-1 [14.1 mg.L-1], corresponding to 21 mol% of the initial 4EP

243

concentration). 26DC4EP reached its maximum concentration after 10 min (30 μmol.L-1

244

[5.7 mg.L-1], corresponding to 7 mol% of the initial 4EP concentration). After 20 min of

245

treatment both intermediates were already removed from solution. For comparison, the

246

performance of a BDD electrode is shown in Figure 3 c & d) for the low and high NaCl

247

concentration. While small changes in the kinetic curves of the degradation of 4EP and

248

formation of chlorinated intermediates are present, generally both electrodes perform similarly

249

under the studied conditions.

250

The results suggest that significant amounts of the same mono- and di-chlorinated intermediates

251

were formed and built-up in the solution for both NaCl concentrations and both electrode types.

252

Although the apparent removal of the starting compound 4EP is fast (compared to NaNO3) a

253

large fraction of 4EP is directly converted into chlorinated compounds in a consecutive reaction

254

path (Figure 2). Interestingly for the higher salt concentration less 2C4EP and 26DC4EP were

255

measured, however the higher FC concentration led to a faster removal rate which most likely

256

suppresses more accumulation of these intermediates.

257

In addition to the two identified intermediates various other chlorinated compounds were

258

detected by MS. While identification of their structure also exceeds the scope of this work, it is

259

interesting to note that the degradation of these intermediates is significantly delayed

260

(Figure S7). The degradation pattern indicates that these compounds are obtained by additional

261

reactions of the two main intermediates, 2C4EP and 26DC4EP. Hydroxylation72,73 and

262

subsequent ring opening73,74 might cause the formation of several other intermediates. In

263

addition, dimerization has been proposed for chemical chlorination of methylphenols leading

264

to polychlorinated methylphenoxymethylphenols72 and the formation of polychlorinated phenol 13 ACS Paragon Plus Environment

Environmental Science & Technology

265

oligomers was observed during ECD of phenol and 2-chlorophenol.74,75 Therefore, it is likely

266

that the detected intermediates are congeners of the reported polychlorinated phenol oligomers.

267

The formation of small chlorinated end products such as chloromethanes17,76 and chlorinated

268

acetic acids17 has been reported for ECD in saline water, which processes likely occur 4-5 hours

269

after start of the experiment. In order to analyze the residual organic content after oxidation of

270

4EP, non purgeable organic carbon (NPOC) measurements were performed. The NPOC values

271

(Figure 4) clearly indicate that large amounts, between 64% to 86% of the initial organic

272

carbon, were still present. Even after extended treatment times of 5 h (high salinity) only 23%

273

of organic carbon was effectively removed. These non-volatile compounds are most probably

274

low molecular weight organic compounds such as organic acids, which have been detected after

275

aromatic ring cleavage during ECD in NaCl solution74 and might also consist of chlorinated

276

derivates such as chlorinated organic acids.17

100

NPOC c/c(0) (%)

80

ECD (0.5 h or 5 h) with Ti/Pt or BDD 86% (5h)

83% (0.5h)

84% (0.5h)

77% (5h)

64% (5h)

60

40

20

0

0.03

0.6 0.03 0.6 salt concentration (mol*L-1)

0.6

277 278

Figure 4. Non purgeable organic carbon (NPOC) measurements of treated waters after ECD

279

of 4EP at constant current conditions (15 mA.cm-2) at a Ti/Pt or BDD electrode in NaCl

14 ACS Paragon Plus Environment

Page 14 of 38

Page 15 of 38

Environmental Science & Technology

280

media. NPOC values were measured after 5 h ECD in 0.03 mol.L-1 NaCl media and after

281

30 min and 5 h for 0.6 mol.L-1 NaCl.

282

The experimental observations highlight a major bottleneck of electrochemical AOP

283

techniques. Though ECD in saline conditions appeared to be efficient, especially compared to

284

inert salt conditions, undesirable (poly)chlorinated aromatic intermediates were formed for both

285

electrodes that require long treatment times. Moreover, the degradation of 4EP and

286

(chlorinated) intermediates on a Ti/Pt electrode led only to a breakdown of the aromatic

287

compounds and the total removal rate of organic carbon was quite poor. Only for BDD at low

288

salt concentration slightly higher NPOC removal was achieved, most likely due to an increased

289

contribution of generated OH radicals. Still, 64 % of the NPOC remained after 5 h treatment

290

and significant amounts of RCl were formed (Figure 3), leading to the conclusion that for both

291

electrodes chlorination is the main degradation pathway under the studied conditions, which is

292

in agreement with a recent report of high faradaic efficiency for chloride oxidation on BDD

293

electrodes.77

294 295

Photocatalytic degradation (PCD) of 4EP

296

The photocatalytic degradation of 4EP at different salt conditions is shown in Figure 5. Using

297

a pseudo first order approximation, implying surface mediated conversion is dominating, rate

298

constants can estimated to be 1.13 h-1 in the absence of salt, 0.8 h-1 in the presence of NaNO3

299

(high and low concentration), and 0.8 h-1 and 0.5 h-1 for low and high concentration of NaCl.

300

To compare the rates to ECD, initial rates can be calculated (using the initial concentration of

301

410 μmol.L-1 [50 mg.L-1]) to be ~330 μmol.L-1.h-1 or 205 μmol.L-1.h-1 for the low or high NaCl

302

concentration respectively. At low NaCl concentration, the rates are comparable, but in contrast

303

to ECD, addition of salts, and in particular NaCl at high concentration, had a negative effect on 15 ACS Paragon Plus Environment

Environmental Science & Technology

304

the degradation rates of the parent compound 4EP, leading to an order of magnitude difference

305

under these conditions (~3750 μmol.L-1.h-1 vs 205 μmol.L-1.h-1). The inhibiting effect of NaCl

306

is in agreement with observations reported for PCD of phenol.38 Since the degradation at

307

0.6 mol.L-1 NaCl is clearly more effected than for 0.6 mol.L-1 NaNO3, the inhibition is mainly

308

attributed to the presence of Cl-.24

309

4EP [no salt] 4EP [0.03 mol*L-1 NaCl] 4EP [0.6 mol*L-1 NaCl] 4EP [0.03 mol*L-1 NaNO3]

concentration (mol*L-1)

400

300

4EP [0.6 mol*L-1 NaNO3]

200

100

0

310

0

1

2

3 time (h)

4

5

311

Figure 5. Photocatalytic degradation of 4EP at various salt conditions: no salt (black curve),

312

0.03 mol.L-1 NaNO3 (light blue), 0.6 mol.L-1 NaNO3 (dark blue), 0.03 mol.L-1 NaCl (light

313

green), 0.6 mol.L-1 NaCl (dark green). PCD was performed under 375 nm illumination in a

314

TiO2 (H600) slurry [0.5 g.L-1].

315 316

While the exact reason for the negative effect of anions is still under debate, blocking of active

317

surface sites25,28–32 appears unlikely as the main explanation. Considering a point of zero charge

318

(PZC) for TiO2 between 5 to 678, the TiO2 particle surface charge in this study is neutral as

319

measurements were performed at circumneutral conditions. Moreover, under illumination the

320

build-up of a negative surface charge has been observed for TiO2 particles (pH > isoelectric

321

point)79,80 hindering a competitive adsorption or blocking of active surface sites by chloride 16 ACS Paragon Plus Environment

Page 16 of 38

Page 17 of 38

Environmental Science & Technology

322

anions. In the case of chloride absorption on the TiO2 particles (pH < PZC, for example possible

323

due to a pH drop during PCD because of dissolved CO2 produced during degradation; see also

324

Figure S8) the oxidation of chloride ions, by either photogenerated valence band holes and/or

325

hydroxyl radicals becomes likely and is proposed frequently in literature.24–27,35–37,39,40,57 As a

326

consequence RCS formation would occur supporting the 4EP removal as observed in ECD.

327

However, our results indicate that the degradation pathways of 4EP during PCD and ECD are

328

different and RCS likely does not contribute to PCD. Thus, just blocking of active sites by Cl-

329

cannot fully explain the observations. In order to verify this conclusion we analyzed the

330

intermediates formed during (PCD).

331

Unlike in the case of ECD, chlorinated intermediates were identified only in extremely low

332

quantities, and LC-MS showed the following aromatic intermediates in significantly larger

333

quantities:

334

Hydroxyacetophenone, (4) 4-Ethylresorcinol and (5) 4-Ethylcatechol (Figure 6a; more

335

information SI Table S1). Thus, the model compound 4EP is oxidized at least at three carbon

336

atoms during PCD, which is typical for non-selective radical reactions (i.e. by •OH). The

337

formation of the isomer 2-(4-Hydroxyphenyl)ethanol could not be revealed due to overlapping

338

retention times with (1) (SI Table S1). However, the increased stability of secondary alkyl

339

radicals favors the formation of intermediate (1) and its subsequent oxidation product (3).

340

Therefore, the formation of 2-(4-Hydroxyphenyl)ethanol is of minor contribution during PCD

341

of 4EP. The same intermediates were detected for all PCD conditions. The non-selective radical

342

reaction pathway is also in good agreement with other reports, e.g. for PCD of ethylbenzene,81

343

phenol37,42,82 or p-cresol,83 confirming that RCS are barely involved in the degradation path.

(1)

1-(4-Hydroxyphenyl)ethanol,

(2)

4-Hydroxybenzaldehyde,

344

17 ACS Paragon Plus Environment

(3)

4-

Environmental Science & Technology

concentration (mol*L-1)

12

Page 18 of 38

4HAP [no salt] 4HAP [0.03 mol*L-1 NaCl] 4HAP [0.6 mol*L-1 NaCl] 4HAP [0.03 mol*L-1 NaNO3]

10

4HAP [0.6 mol*L-1 NaNO3]

8

1.7 mol% [0.90 mg*L-1]

6 4 2 0 0

1

2

3

4

5

345

a)

346

Figure 6. a) Chemical structures of the identified aromatic intermediates during PCD of 4EP

347

observed for all studied salt conditions (no salt, 0.03 mol.L-1 NaNO3, 0.6 mol.L-1 NaNO3,

348

0.03 mol.L-1 NaCl, 0.6 mol.L-1 NaCl: (1) 1-(4-Hydroxyphenyl)ethanol, (2) 4-

349

Hydroxybenzaldehyde, (3) 4-Hydroxyacetophenone, (4) 4-Ethylresorcinol, (5) 4-

350

Ethylcatechol. b) Formation and degradation of intermediate (3) 4-Hydroxyacetophenone

351

during PCD of 4EP at various salt conditions: no salt (black curve), 0.03 mol.L-1 NaNO3

352

(light blue), 0.6 mol.L-1 NaNO3 (dark blue), 0.03 mol.L-1 NaCl (light green), 0.6 mol.L-1 NaCl

353

(dark green).

b)

time (h)

354 355

For the most abundant intermediate (3), its formation and degradation at different salt

356

conditions is shown in Figure 6b. At high NaCl concentration a pronounced inhibition of

357

formation and subsequent degradation was observed, whereas the influence of NaNO3 or low

358

NaCl concentration is less. The observations are related to the 4EP removal, where the strongly

359

inhibited 4EP degradation explains the delayed formation of (3) at high NaCl concentration.

360

This is in agreement with reports for PCD of phenol.38

18 ACS Paragon Plus Environment

Page 19 of 38

Environmental Science & Technology

361

As 2C4EP and 26DC4EP were detected as main chlorinated intermediates in ECD the obtained

362

chromatograms were thoroughly analyzed in PCD. Only for high salt concentration small

363

amounts of 2C4EP could be measured and quantified (Figure 7). Similarly to ECD

364

measurements, formation of 2C4EP occurs immediately, slowly accumulates in the reactor, and

365

its degradation is strongly retarded in agreement with intermediate (3) monitored during PCD

366

at high saline conditions (see Figure 6b). If at all formed, the concentration of 26DC4EP

367

remained below the LOD.

368

2C4EP [0.6 mol*L-1 NaCl]

1.4

concentration (mol*L-1)

1.2 1.0 0.3 mol%

0.8

[0.20 mg*L-1]

0.6 0.4 0.2 0.0

369

0

1

2

3 time (h)

4

5

370

Figure 7. Formation of the monochlorinated intermediate 2C4EP during PCD of 4EP in

371

0.6 mol.L-1 NaCl solution.

372 373

Again FC concentrations were determined. For low NaCl concentrations the measured values

374

are below the limit of accuracy; for high NaCl concentrations only small amounts of FC could

375

be detected [0.15 mg.L-1]. This shows PCD is hardly mediated by FC species, in agreement

376

with the low quantities of chlorinated intermediates.

19 ACS Paragon Plus Environment

Environmental Science & Technology

377

In general the oxidation of chloride by scavenging OH radicals is not favored at the conditions

378

of this study.45,52,84 Zhang et al. concluded chloride being a poor hydroxyl radical scavenger at

379

circumneutral to alkaline conditions, because the equilibrium of the reaction is favored for the

380

reverse reaction to Cl- and •OH (> 99.98%).52 The absence of FC for smaller Cl- concentrations

381

is also in agreement with findings reported by Krivec et al.,32 who did not detect Cl• radicals.

382

For higher NaCl concentrations the results match the observations by Azevedo et al.38 who

383

reported small amounts of 4-Chlorophenol (not quantified) formed during phenol PCD at high

384

salt concentrations. Anyway, the results indicate that chloride is barely oxidized during PCD,

385

and formation of RCS is negligible.

386

As for the ECD experiments, the overall removal of organic carbon during the PCD treatment

387

was estimated by determination of the NPOC content. Results obtained after 5 h of PCD in

388

saline water are shown in Figure 8. In the absence of any salt NPOC was below the LOD. This

389

is in agreement with the non-affected 4EP and intermediate removal (compared to salt

390

containing solutions) and indicates a complete removal of the organic carbon within the

391

duration of the PCD experiments. For the low and high NaCl concentration, residual NPOC

392

remained in the solution in agreement with the delay in 4EP degradation. For example, although

393

4EP and the main intermediate (3) were removed after 5 h treatment in 0.03 mol.L-1 NaCl

394

solutions, NPOC of 13% was measured. Since the UV chromatogram did not show any

395

significant peaks after 5 h of PCD, the residual organic carbon most likely exists in compounds

396

arising from ring-opening reactions, such as short chain aliphatic compounds and carboxylic

397

acids.82,85,86 At high NaCl concentration, a residual NPOC of 46% remained. Calculated

398

photonic efficiencies of PCD experiments for the various salt conditions are shown in Table S5

399

and an estimation of the energy consumption compared to ECD experiments is shown in

400

Table S6. For the applied conditions in this study the energy consumption of the PCD treatment

401

was approx. between one and two orders of magnitude higher compared to the ECD; however, 20 ACS Paragon Plus Environment

Page 20 of 38

Page 21 of 38

Environmental Science & Technology

402

it is important to note that the applied reactors were not optimized for energy consumption, and

403

the applied light sources have low efficiency for illumination of the photocatalytic slurry (less

404

than 1%). Although ECD thus seems to appear beneficial in terms of consumed energy per kg

405

removed non purgeable organic carbon (NPOC), degradation of 4EP by PCD leads to

406

significantly smaller NPOC than ECD at comparable rate of 4EP conversion (low NaCl

407

concentration). Moreover, the concentration of chlorinated intermediates formed during PCD

408

is significantly lower (for low & high NaCl concentration) compared to ECD, thus

409

photocatalysis appears the preferred method for application, in particular if illumination

410

efficiency and reactor design are optimized.

100

PCD in NaCl (5h)

NPOC c/c(0) (%)

80

60

46% 40

20

0

13% 0.03

0.6

salt concentration (mol*L-1)

411 412

Figure 8. Non purgeable organic carbon (NPOC) measurements of treated waters after 5 h

413

PCD of 4EP in 0.03 mol.L-1 NaCl and 0.6 mol.L-1 NaCl.

414 415

Finally, photoelectrochemical degradation (PECD), a hybrid technology, was studied. The

416

experiments were performed with a TiO2/Ti electrode at an applied anodic bias of 0.1 V vs

417

Ag/AgCl irradiated with UVA light (see Figure S9 for more information). In recent reports

21 ACS Paragon Plus Environment

Environmental Science & Technology

418

Zanoni et al. showed already that oxidation of chloride to FC at TiO2 photoelectrodes occurs87

419

and the generation of RCS has been proposed to assist the PECD of organic compounds.46,57,61,88

420

Still, reports discussing the formation of intermediates are scarce.50,89 After 15 min of

421

irradiation at low NaCl concentration, 92% of 4EP was removed and an intensive formation of

422

RCS was observed (as detected by FC [27 mg.L-1]). The HPLC analysis (Figure 9) revealed

423

significant amounts of the same intermediates, as observed during ECD (57 μmol.L-1 [9 mg.L-

424

1]

425

(resembling PCD conditions) intermediate (3) was observed as main intermediate and the

426

formation of chlorinated compounds was excluded (Figure S10). Thus, the 4EP reaction

427

pathway can be easily modified by the applied bias. The blank experiment (PECD in dark at

428

0.1 V vs. Ag/AgCl) did not lead to any photocurrent, RCS formation, or 4EP degradation

429

(Figure S11), because the applied potential was below any relevant standard redox potential

430

(e.g 1.36 V vs. SHE for chlorine evolution or 2.8 V vs. SHE for ●OH generation). To verify the

431

observations and exclude a particular effect of the applied TiO2, PECD experiments were

432

conducted with the H600 photocatalyst coated on FTO glass, showing similar trends (see

433

Table S7).

2C4EP and 89 μmol.L-1 [17 mg.L-1] 26DC4EP). Interestingly, without additional bias

22 ACS Paragon Plus Environment

Page 22 of 38

Page 23 of 38

Environmental Science & Technology

100 4EP

dark UVA

Intensity (a.u.)

80

60

40 2C4EP 20

26DC4EP

RCl

0 28

30

32

34

36

38

40

time (min)

434 435

Figure 9. HPLC chromatogram obtained after 15 min PECD of 4EP in 0.03 mol.L-1 NaCl

436

solution using a TiO2/Ti electrode with 0.1 V[vs. Ag/AgCl] bias and 365 nm UV light. The

437

majority of 4EP was removed and transformed into significant amounts of the

438

monochlorinated 2C4EP and dichlorinated 26DC4EP, similar as observed during ECD

439

experiments.

440

The effect of an applied anodic bias can be explained by the following: In a photocatalytic

441

approach charge carrier (electrons and holes) are usually generated close to the particle surface

442

and physical separation of reduction and oxidation sites is hardly achieved on the nanoscale.

443

Thus, oxidation and reduction occur in close proximity and the recombination of oxidized

444

chloride with photogenerated CB electrons is likely. This surface-charge recombination

445

mechanism (quenching of photogenerated charge carriers by chloride ions)34 is effectively

446

suppressed when an anodic bias is applied, e.g. in PECD experiments. Upon biasing the

447

electrode photogenerated conduction band (CB) electrons from the TiO2 surface are pulled

448

towards the back contact and in contrast to PCD, reduction of Cl-radicals by available CB

449

electrons is unfeasible. Therefore, pronounced RCS generation could be observed during

450

PECD, leading to a distinctive formation of RCl. In addition, the coulombic repulsion between

451

the TiO2 surface and chloride ions90 is lowered upon polarization of the TiO2 surface.50 This 23 ACS Paragon Plus Environment

Environmental Science & Technology

452

facilitates the adsorption of Cl- on the positively charged TiO2 surface allowing for chloride

453

oxidation (RCS formation) and subsequent formation of RCl.

454

Combining all information, it is evident that although not all of the organic carbon could be

455

removed during PCD, the removal of organic carbon was more effective in the purely light-

456

driven approach compared to ECD (under the conditions applied in this study). Especially, the

457

significantly smaller formation of RCS and RCl intermediates during PCD is of importance.

458

During ECD in saline media 4EP was mainly removed by conversion into (polychlorinated)

459

compounds and only small amounts of 4EP could be completely mineralized. In contrast, during

460

PCD only for high NaCl concentrations minor formation of monochlorinated 2C4EP could be

461

detected. The measured 2C4EP concentration during ECD at 0.6 mol.L-1 was nearly 2 orders

462

of magnitude higher than that detected after 5 h of PCD. The difference in the AOP techniques

463

is likely governed by a surface-charge recombination mechanism, which suppresses FC and

464

RCl formation. Overall, the formation of chlorinated intermediates in both AOP technologies

465

is important for consideration of practical wastewater treatment applications.

466

24 ACS Paragon Plus Environment

Page 24 of 38

Page 25 of 38

467

Environmental Science & Technology

ASSOCIATED CONTENT

468 469

Supporting Information. Synthesis of 1-(4-Hydroxyphenyl)ethanol; NPOC measurements;

470

synthesis TiO2 photoelectrodes; 4EP and identified intermediates; coefficient of determination

471

(calibration); limit of detection; iR drop corrected potentials; XRD of pre-treated photocatalyst;

472

UV transmission of salt solutions; scheme of electrochemical cell; calibration curves; potential-

473

time curves; ECD intermediates (molar fraction); ECD additional intermediate; PCD pH

474

development; PCD photonic efficiencies; energy consumption; PECD linear sweep

475

voltammograms; PECD without bias (chromatogram); PECD in dark (photocurrent); PECD

476

intermediates with H600/FTO electrode

477

478

AUTHOR INFORMATION

479

Corresponding Author

480

* E-mail: [email protected]

481

Phone: +31534893890

482

ACKNOWLEDGEMENTS

483 484

This research is financed by the Netherlands Organisation for Scientific Research (NWO),

485

which is partly funded by the Ministry of Economic Affairs and Climate Policy, and co-

486

financed by the Netherlands Ministry of Infrastructure and Water Management and partners of

487

the Dutch Water Nexus consortium (project number 14301). Dyllian Huijgens is acknowledged

488

for PCD experiments. Erna Fräntzel-Luiten and Robert Meijer are acknowledged for technical

489

support. 25 ACS Paragon Plus Environment

Environmental Science & Technology

490

REFERENCES

491

(1)

492 493

Development Report 2015: Water for a Sustainalbe World; UNESCO: Paris, 2015. (2)

494 495

WWAP (United Nations World Water Assessment Programme). The United Nations World

Smalley, R. E. Future Global Energy Prosperity: The Terawatt Challenge. MRS Bull. 2005, 30 (6), 412–417. https://doi.org/10.1557/mrs2005.124.

(3)

WWAP (United Nations World Water Assessment Programme) The United Nations World

496

Water Development Report 2017: Wastewater The Untapped Rescource; UNESCO: Paris,

497

2017.

498

(4)

Garcia-Segura, S.; Ocon, J. D.; Chong, M. N. Electrochemical Oxidation Remediation of Real

499

Wastewater Effluents — A Review. Process Saf. Environ. Prot. 2018, 113, 48–67.

500

https://doi.org/10.1016/j.psep.2017.09.014.

501

(5)

Margot, J.; Rossi, L.; Barry, D. A.; Holliger, C. A Review of the Fate of Micropollutants in

502

Wastewater Treatment Plants. Wiley Interdiscip. Rev. Water 2015, 2 (5), 457–487.

503

https://doi.org/10.1002/wat2.1090.

504

(6)

Luo, Y.; Guo, W.; Ngo, H. H.; Nghiem, L. D.; Hai, F. I.; Zhang, J.; Liang, S.; Wang, X. C. A Review

505

on the Occurrence of Micropollutants in the Aquatic Environment and Their Fate and Removal

506

during Wastewater Treatment. Sci. Total Environ. 2014, 473–474, 619–641.

507

https://doi.org/http://dx.doi.org/10.1016/j.scitotenv.2013.12.065.

508

(7)

Gupta, A.; Thakur, I. S. Treatment of Organic Recalcitrant Contaminants in Wastewater. In

509

Biological Wastewater Treatment and Resource Recovery; Farooq, R., Ahmad, Z., Eds.; InTech,

510

2017. https://doi.org/10.5772/66346.

511

(8)

Lefebvre, O.; Moletta, R. Treatment of Organic Pollution in Industrial Saline Wastewater: A

512

Literature Review. Water Res. 2006, 40 (20), 3671–3682.

513

https://doi.org/http://dx.doi.org/10.1016/j.watres.2006.08.027. 26 ACS Paragon Plus Environment

Page 26 of 38

Page 27 of 38

514

Environmental Science & Technology

(9)

Kargi, F.; Dincer, A. R. Effect of Salt Concentration on Biological Treatment of Saline

515

Wastewater by Fed-Batch Operation. Enzyme Microb. Technol. 1996, 19 (7), 529–537.

516

https://doi.org/10.1016/S0141-0229(96)00070-1.

517

(10)

Dewil, R.; Mantzavinos, D.; Poulios, I.; Rodrigo, M. A. New Perspectives for Advanced

518

Oxidation Processes. J. Environ. Manage. 2017, 195, 93–99.

519

https://doi.org/https://doi.org/10.1016/j.jenvman.2017.04.010.

520

(11)

521 522

Comninellis, C.; Nerini, A. Anodic Oxidation of Phenol in the Presence of NaCl for Wastewater Treatment. J. Appl. Electrochem. 1995, 25 (1), 23–28. https://doi.org/10.1007/bf00251260.

(12)

Panizza, M.; Cerisola, G. Electrochemical Oxidation of 2-Naphthol with in Situ

523

Electrogenerated Active Chlorine. Electrochim. Acta 2003, 48 (11), 1515–1519.

524

https://doi.org/https://doi.org/10.1016/S0013-4686(03)00028-8.

525

(13)

526 527

Panizza, M.; Cerisola, G. Direct And Mediated Anodic Oxidation of Organic Pollutants. Chem. Rev. 2009, 109 (12), 6541–6569. https://doi.org/10.1021/cr9001319.

(14)

Martinez-Huitle, C. A.; Ferro, S. Electrochemical Oxidation of Organic Pollutants for the

528

Wastewater Treatment: Direct and Indirect Processes. Chem. Soc. Rev. 2006, 35 (12), 1324–

529

1340. https://doi.org/10.1039/B517632H.

530

(15)

Sirés, I.; Brillas, E.; Oturan, M. A.; Rodrigo, M. A.; Panizza, M. Electrochemical Advanced

531

Oxidation Processes: Today and Tomorrow. A Review. Environ. Sci. Pollut. Res. 2014, 21 (14),

532

8336–8367. https://doi.org/10.1007/s11356-014-2783-1.

533

(16)

Saylor, G. L.; Chen, L.; Kupferle, M. J. Using Toxicity Testing to Evaluate Electrochemical

534

Reactor Operations. Environ. Toxicol. Chem. 2012, 31 (3), 494–500.

535

https://doi.org/10.1002/etc.1719.

536 537

(17)

Jasper, J. T.; Yang, Y.; Hoffmann, M. R. Toxic Byproduct Formation during Electrochemical Treatment of Latrine Wastewater. Environ. Sci. Technol. 2017, 51 (12), 7111–7119. 27 ACS Paragon Plus Environment

Environmental Science & Technology

538 539

https://doi.org/10.1021/acs.est.7b01002. (18)

Li, H.; Long, Y.; Zhu, X.; Tian, Y.; Ye, J. Influencing Factors and Chlorinated Byproducts in

540

Electrochemical Oxidation of Bisphenol A with Boron-Doped Diamond Anodes. Electrochim.

541

Acta 2017, 246, 1121–1130. https://doi.org/10.1016/j.electacta.2017.06.163.

542

(19)

Burgos-Castillo, R. C.; Sirés, I.; Sillanpää, M.; Brillas, E. Application of Electrochemical

543

Advanced Oxidation to Bisphenol A Degradation in Water. Effect of Sulfate and Chloride Ions.

544

Chemosphere 2018, 194, 812–820. https://doi.org/10.1016/j.chemosphere.2017.12.014.

545

(20)

Chaplin, B. P. Critical Review of Electrochemical Advanced Oxidation Processes for Water

546

Treatment Applications. Environ. Sci. Process. Impacts 2014, 16 (6), 1182–1203.

547

https://doi.org/10.1039/C3EM00679D.

548

(21)

549 550

Appl. Chem. 1996, 68 (9), 1791–1799. https://doi.org/10.1351/pac199668091791. (22)

551 552

Hanberg, A. Toxicology of Environmentally Persistent Chlorinated Organic Compounds. Pure

Schoeny, R. Disinfection By-Products: A Question of Balance. Environ. Health Perspect. 2010, 118 (11), A466–A467. https://doi.org/10.1289/ehp.1003053.

(23)

Villanueva, C. M.; Cordier, S.; Font-Ribera, L.; Salas, L. A.; Levallois, P. Overview of Disinfection

553

By-Products and Associated Health Effects. Curr. Environ. Heal. reports 2015, 2 (1), 107–115.

554

https://doi.org/10.1007/s40572-014-0032-x.

555

(24)

Lindner, M.; Bahnemann, D. W.; Hirthe, B.; Griebler, W. D. Solar Water Detoxification: Novel

556

TiO2 Powders as Highly Active Photocatalysts. J. Sol. Energy Eng. 1997, 119 (2), 120–125.

557

https://doi.org/10.1115/1.2887890.

558

(25)

Burns, R. A.; Crittenden, J. C.; Hand, D. W.; Selzer, V. H.; Sutter, L. L.; Salman, S. R. Effect of

559

Inorganic Ions in Heterogeneous Photocatalysis of TCE. J. Environ. Eng. 1999, 125 (1), 77–85.

560

https://doi.org/doi:10.1061/(ASCE)0733-9372(1999)125:1(77).

28 ACS Paragon Plus Environment

Page 28 of 38

Page 29 of 38

561

Environmental Science & Technology

(26)

Calza, P.; Pelizzetti, E. Photocatalytic Transformation of Organic Compounds in the Presence

562

of Inorganic Ions. Pure Appl. Chem. 2001, 73, 1839.

563

https://doi.org/10.1351/pac200173121839.

564

(27)

565 566

Yang, S.; Chen, Y.; Lou, L.; Wu, X. Involvement of Chloride Anion in Photocatalytic Process. J. Environ. Sci. 2005, 17 (5), 761–765.

(28)

Chen, H. Y.; Zahraa, O.; Bouchy, M. Inhibition of the Adsorption and Photocatalytic

567

Degradation of an Organic Contaminant in an Aqueous Suspension of TiO2 by Inorganic Ions.

568

J. Photochem. Photobiol. A Chem. 1997, 108 (1), 37–44. https://doi.org/10.1016/S1010-

569

6030(96)04411-5.

570

(29)

Zhang, W.; An, T.; Cui, M.; Sheng, G.; Fu, J. Effects of Anions on the Photocatalytic and

571

Photoelectrocatalytic Degradation of Reactive Dye in a Packed-Bed Reactor. J. Chem. Technol.

572

Biotechnol. 2005, 80 (2), 223–229. https://doi.org/10.1002/jctb.1185.

573

(30)

Chong, M. N.; Jin, B.; Chow, C. W. K.; Saint, C. Recent Developments in Photocatalytic Water

574

Treatment Technology: A Review. Water Res. 2010, 44 (10), 2997–3027.

575

https://doi.org/https://doi.org/10.1016/j.watres.2010.02.039.

576

(31)

Umar, M.; Abdul Aziz, H. Photocatalytic Degradation of Organic Pollutants in Water. In

577

Organic Pollutants - Monitoring, Risk and Treatment; Nageeb Rashed, M., Ed.; InTech, 2013.

578

https://doi.org/10.5772/53699.

579

(32)

Krivec, M.; Dillert, R.; Bahnemann, D. W.; Mehle, A.; Strancar, J.; Drazic, G. The Nature of

580

Chlorine-Inhibition of Photocatalytic Degradation of Dichloroacetic Acid in a TiO2-Based

581

Microreactor. Phys. Chem. Chem. Phys. 2014, 16 (28), 14867–14873.

582

https://doi.org/10.1039/C4CP01043D.

583 584

(33)

Guillard, C.; Puzenat, E.; Lachheb, H.; Houas, A.; Herrmann, J.-M. Why Inorganic Salts Decrease the TiO2 Photocatalytic Efficiency. Int. J. Photoenergy 2005, 7 (1), 1–9. 29 ACS Paragon Plus Environment

Environmental Science & Technology

585 586

https://doi.org/10.1155/s1110662x05000012. (34)

Sunada, F.; Heller, A. Effects of Water, Salt Water, and Silicone Overcoating of the TiO2

587

Photocatalyst on the Rates and Products of Photocatalytic Oxidation of Liquid 3-Octanol and

588

3-Octanone. Environ. Sci. Technol. 1998, 32 (2), 282–286. https://doi.org/10.1021/es970523f.

589

(35)

Abdullah, M.; Low, G. K. C.; Matthews, R. W. Effects of Common Inorganic Anions on Rates of

590

Photocatalytic Oxidation of Organic Carbon over Illuminated Titanium Dioxide. J. Phys. Chem.

591

1990, 94 (17), 6820–6825. https://doi.org/10.1021/j100380a051.

592

(36)

Piscopo, A.; Robert, D.; Weber, J. V. Influence of PH and Chloride Anion on the Photocatalytic

593

Degradation of Organic Compounds. Appl. Catal. B Environ. 2001, 35 (2), 117–124.

594

https://doi.org/http://dx.doi.org/10.1016/S0926-3373(01)00244-2.

595

(37)

Sökmen, M.; Özkan, A. Decolourising Textile Wastewater with Modified Titania: The Effects of

596

Inorganic Anions on the Photocatalysis. J. Photochem. Photobiol. A Chem. 2002, 147 (1), 77–

597

81. https://doi.org/10.1016/S1010-6030(01)00627-X.

598

(38)

Azevedo, E. B.; Neto, F. R. de A.; Dezotti, M. TiO2-Photocatalyzed Degradation of Phenol in

599

Saline Media: Lumped Kinetics, Intermediates, and Acute Toxicity. Appl. Catal. B Environ.

600

2004, 54 (3), 165–173. https://doi.org/http://dx.doi.org/10.1016/j.apcatb.2004.06.014.

601

(39)

Yuan, R.; Ramjaun, S. N.; Wang, Z.; Liu, J. Photocatalytic Degradation and Chlorination of Azo

602

Dye in Saline Wastewater: Kinetics and AOX Formation. Chem. Eng. J. 2012, 192 (Supplement

603

C), 171–178. https://doi.org/https://doi.org/10.1016/j.cej.2012.03.080.

604

(40)

Adishkumar, S.; Kanmani, S.; Rajesh Banu, J. Solar Photocatalytic Treatment of Phenolic

605

Wastewaters: Influence of Chlorides, Sulphates, Aeration, Liquid Volume and Solar Light

606

Intensity. Desalin. Water Treat. 2014, 52 (40–42), 7957–7963.

607

https://doi.org/10.1080/19443994.2013.834522.

608

(41)

Pellegrino, F.; Pellutiè, L.; Sordello, F.; Minero, C.; Ortel, E.; Hodoroaba, V. D.; Maurino, V. 30 ACS Paragon Plus Environment

Page 30 of 38

Page 31 of 38

Environmental Science & Technology

609

Influence of Agglomeration and Aggregation on the Photocatalytic Activity of TiO 2

610

Nanoparticles. Appl. Catal. B Environ. 2017, 216, 80–87.

611

https://doi.org/10.1016/j.apcatb.2017.05.046.

612

(42)

L’Amour, R. J. A.; Azevedo, E. B.; Leite, S. G. F.; Dezotti, M. Removal of Phenol in High Salinity

613

Media by a Hybrid Process (Activated Sludge+photocatalysis). Sep. Purif. Technol. 2008, 60 (2),

614

142–146. https://doi.org/https://doi.org/10.1016/j.seppur.2007.08.008.

615

(43)

Azevedo, E. B.; Tôrres, A. R.; Aquino Neto, F. R.; Dezotti, M. TiO2-Photocatalyzed Degradation

616

of Phenol in Saline Media in an Annular Reactor: Hydrodynamics, Lumped Kinetics,

617

Intermediates, and Acute Toxicity. Brazilian J. Chem. Eng. 2009, 26, 75–87.

618

(44)

619 620

66 (6), 185–297. https://doi.org/https://doi.org/10.1016/j.surfrep.2011.01.001. (45)

621 622

Henderson, M. A. A Surface Science Perspective on TiO2 Photocatalysis. Surf. Sci. Rep. 2011,

Yang, Y.; Pignatello, J. Participation of the Halogens in Photochemical Reactions in Natural and Treated Waters. Molecules 2017, 22 (10), 1684.

(46)

Su, Y.; Wang, G.-B.; Kuo, D. T. F.; Chang, M.; Shih, Y. Photoelectrocatalytic Degradation of the

623

Antibiotic Sulfamethoxazole Using TiO2/Ti Photoanode. Appl. Catal. B Environ. 2016, 186

624

(Supplement C), 184–192. https://doi.org/https://doi.org/10.1016/j.apcatb.2016.01.003.

625

(47)

Ghernaout, D.; Naceur, M. W.; Aouabed, A. On the Dependence of Chlorine By-Products

626

Generated Species Formation of the Electrode Material and Applied Charge during

627

Electrochemical Water Treatment. Desalination 2011, 270 (1–3), 9–22.

628

https://doi.org/10.1016/j.desal.2011.01.010.

629

(48)

Radjenovic, J.; Sedlak, D. L. Challenges and Opportunities for Electrochemical Processes as

630

Next-Generation Technologies for the Treatment of Contaminated Water. Environ. Sci.

631

Technol. 2015, 49 (19), 11292–11302. https://doi.org/10.1021/acs.est.5b02414.

632

(49)

Bessegato, G. G.; Guaraldo, T. T.; de Brito, J. F.; Brugnera, M. F.; Zanoni, M. V. B. 31 ACS Paragon Plus Environment

Environmental Science & Technology

633

Achievements and Trends in Photoelectrocatalysis: From Environmental to Energy

634

Applications. Electrocatalysis 2015, 6 (5), 415–441. https://doi.org/10.1007/s12678-015-0259-

635

9.

636

(50)

Garcia-Segura, S.; Brillas, E. Applied Photoelectrocatalysis on the Degradation of Organic

637

Pollutants in Wastewaters. J. Photochem. Photobiol. C Photochem. Rev. 2017, 31, 1–35.

638

https://doi.org/10.1016/J.JPHOTOCHEMREV.2017.01.005.

639

(51)

Hao, Z.; Yin, Y.; Wang, J.; Cao, D.; Liu, J. Formation of Organobromine and Organoiodine

640

Compounds by Engineered TiO2nanoparticle-Induced Photohalogenation of Dissolved Organic

641

Matter in Environmental Waters. Sci. Total Environ. 2018, 631–632, 158–168.

642

https://doi.org/10.1016/j.scitotenv.2018.03.027.

643

(52)

644 645

Zhang, K.; Parker, K. M. Halogen Radical Oxidants in Natural and Engineered Aquatic Systems. Environ. Sci. Technol. 2018, acs.est.8b02219. https://doi.org/10.1021/acs.est.8b02219.

(53)

Selcuk, H.; Bekbolet, M. Photocatalytic and Photoelectrocatalytic Humic Acid Removal and

646

Selectivity of TiO2coated Photoanode. Chemosphere 2008, 73 (5), 854–858.

647

https://doi.org/10.1016/j.chemosphere.2008.05.069.

648

(54)

Zhang, F.; Li, M.; Li, W.; Feng, C.; Jin, Y.; Guo, X.; Cui, J. Degradation of Phenol by a Combined

649

Independent Photocatalytic and Electrochemical Process. Chem. Eng. J. 2011, 175 (1), 349–

650

355. https://doi.org/10.1016/j.cej.2011.09.122.

651

(55)

Osiewała, L.; Socha, A.; Perek, A.; Socha, M.; Rynkowski, J. Electrochemical, Photochemical,

652

and Photoelectrochemical Treatment of Sodium p-Cumenesulfonate. Water. Air. Soil Pollut.

653

2013, 224 (9). https://doi.org/10.1007/s11270-013-1657-3.

654

(56)

Hurwitz, G.; Pornwongthong, P.; Mahendra, S.; Hoek, E. M. V. Degradation of Phenol by

655

Synergistic Chlorine-Enhanced Photo-Assisted Electrochemical Oxidation. Chem. Eng. J. 2014,

656

240 (Supplement C), 235–243. https://doi.org/https://doi.org/10.1016/j.cej.2013.11.087. 32 ACS Paragon Plus Environment

Page 32 of 38

Page 33 of 38

657

Environmental Science & Technology

(57)

Tantis, I.; Bousiakou, L.; Karikas, G.-A.; Lianos, P. Photocatalytic and Photoelectrocatalytic

658

Degradation of the Antibacterial Agent Ciprofloxacin. Photochem. Photobiol. Sci. 2015, 14 (3),

659

603–607. https://doi.org/10.1039/C4PP00377B.

660

(58)

Suhadolnik, L.; Pohar, A.; Likozar, B.; Čeh, M. Mechanism and Kinetics of Phenol

661

Photocatalytic, Electrocatalytic and Photoelectrocatalytic Degradation in a TiO2-Nanotube

662

Fixed-Bed Microreactor. Chem. Eng. J. 2016, 303, 292–301.

663

https://doi.org/10.1016/j.cej.2016.06.027.

664

(59)

Serna-Galvis, E. A.; Silva-Agredo, J.; Giraldo, A. L.; Flórez-Acosta, O. A.; Torres-Palma, R. A.

665

Comparative Study of the Effect of Pharmaceutical Additives on the Elimination of Antibiotic

666

Activity during the Treatment of Oxacillin in Water by the Photo-Fenton, TiO2-Photocatalysis

667

and Electrochemical Processes. Sci. Total Environ. 2016, 541, 1431–1438.

668

https://doi.org/10.1016/j.scitotenv.2015.10.029.

669

(60)

Granda-Ramírez, C. F.; Hincapié-Mejía, G. M.; Serna-Galvis, E. A.; Torres-Palma, R. A.

670

Degradation of Recalcitrant Safranin T Through an Electrochemical Process and Three

671

Photochemical Advanced Oxidation Technologies. Water. Air. Soil Pollut. 2017, 228 (11).

672

https://doi.org/10.1007/s11270-017-3611-2.

673

(61)

Cao, D.; Wang, Y.; Zhao, X. Combination of Photocatalytic and Electrochemical Degradation of

674

Organic Pollutants from Water. Curr. Opin. Green Sustain. Chem. 2017, 6, 78–84.

675

https://doi.org/10.1016/j.cogsc.2017.05.007.

676

(62)

Escudero, C. J.; Iglesias, O.; Dominguez, S.; Rivero, M. J.; Ortiz, I. Performance of

677

Electrochemical Oxidation and Photocatalysis in Terms of Kinetics and Energy Consumption.

678

New Insights into the p-Cresol Degradation. J. Environ. Manage. 2017, 195, 117–124.

679

https://doi.org/https://doi.org/10.1016/j.jenvman.2016.04.049.

680

(63)

Carneiro, J. T.; Savenije, T. J.; Moulijn, J. A.; Mul, G. Toward a Physically Sound

33 ACS Paragon Plus Environment

Environmental Science & Technology

681

Structure−Activity Relationship of TiO2-Based Photocatalysts. J. Phys. Chem. C 2010, 114 (1),

682

327–332. https://doi.org/10.1021/jp906395w.

683

(64)

Romão, J. S.; Hamdy, M. S.; Mul, G.; Baltrusaitis, J. Photocatalytic Decomposition of Cortisone

684

Acetate in Aqueous Solution. J. Hazard. Mater. 2015, 282 (Supplement C), 208–215.

685

https://doi.org/https://doi.org/10.1016/j.jhazmat.2014.05.087.

686

(65)

Boitsov, S.; Mjøs, S. A.; Meier, S. Identification of Estrogen-like Alkylphenols in Produced

687

Water from Offshore Oil Installations. Mar. Environ. Res. 2007, 64 (5), 651–665.

688

https://doi.org/http://dx.doi.org/10.1016/j.marenvres.2007.07.001.

689

(66)

Priac, A.; Morin-Crini, N.; Druart, C.; Gavoille, S.; Bradu, C.; Lagarrigue, C.; Torri, G.; Winterton,

690

P.; Crini, G. Alkylphenol and Alkylphenol Polyethoxylates in Water and Wastewater: A Review

691

of Options for Their Elimination. Arab. J. Chem. 2017, 10, S3749–S3773.

692

https://doi.org/10.1016/j.arabjc.2014.05.011.

693

(67)

Kochukov, M. Y.; Jeng, Y. J.; Watson, C. S. Alkylphenol Xenoestrogens with Varying Carbon

694

Chain Lengths Differentially and Potently Activate Signaling and Functional Responses in

695

GH3/B6/F10 Somatomammotropes. Environ. Health Perspect. 2009, 117 (5), 723–730.

696

https://doi.org/10.1289/ehp.0800182.

697

(68)

Reckhow, D. A.; Singer, P. C.; Malcolm, R. L. Chlorination of Humic Materials: Byproduct

698

Formation and Chemical Interpretations. Environ. Sci. Technol. 1990, 24 (11), 1655–1664.

699

https://doi.org/10.1021/es00081a005.

700

(69)

Bond, T.; Henriet, O.; Goslan, E. H.; Parsons, S. A.; Jefferson, B. Disinfection Byproduct

701

Formation and Fractionation Behavior of Natural Organic Matter Surrogates. Environ. Sci.

702

Technol. 2009, 43 (15), 5982–5989. https://doi.org/10.1021/es900686p.

703 704

(70)

Lee, G. F.; Morris, J. C. Kinetics of Chlorination of Phenol-Chlorophenolic Tastes and Odors. Int.J.Air Wat.Poll. 1962, 6 (567), 419–431. 34 ACS Paragon Plus Environment

Page 34 of 38

Page 35 of 38

705

Environmental Science & Technology

(71)

Rule, K. L.; Ebbett, V. R.; Vikesland, P. J. Formation of Chloroform and Chlorinated Organics by

706

Free-Chlorine-Mediated Oxidation of Triclosan. Environ. Sci. Technol. 2005, 39 (9), 3176–3185.

707

https://doi.org/10.1021/es048943+.

708

(72)

Onodera, S.; Yamada, K.; Yamaji, Y.; Ishikura, S.; Suzuki, S. Chemical Changes of Organic

709

Compounds in Chlorinated Water: X. Formation of Polychlorinated

710

Methylphenoxymethylphenols (Predioxins) during Chlorination of Methylphenols in Dilute

711

Aqueous Solution. J. Chromatogr. 1986, 354, 293–303.

712

(73)

Onodera, S.; Tabata, M.; Suzuki, S.; Ishikura, S. Gas Chromatographic Identification and

713

Determination of Chlorinated Quinones Formed during Chlorination of Dihydric Phenols with

714

Hypochlorite in Dilute Aqueous Solution. J. Chromatogr. 1980, 200, 137–144.

715

(74)

Vallejo, M.; San Román, M. F.; Ortiz, I. Quantitative Assessment of the Formation of

716

Polychlorinated Derivatives, PCDD/Fs, in the Electrochemical Oxidation of 2-Chlorophenol As

717

Function of the Electrolyte Type. Environ. Sci. Technol. 2013, 47 (21), 12400–12408.

718

https://doi.org/10.1021/es403246g.

719

(75)

Chen, L.; Campo, P.; Kupferle, M. J. Identification of Chlorinated Oligomers Formed during

720

Anodic Oxidation of Phenol in the Presence of Chloride. J. Hazard. Mater. 2015, 283, 574–581.

721

https://doi.org/10.1016/j.jhazmat.2014.10.001.

722

(76)

Zöllig, H.; Remmele, A.; Fritzsche, C.; Morgenroth, E.; Udert, K. M. Formation of Chlorination

723

Byproducts and Their Emission Pathways in Chlorine Mediated Electro-Oxidation of Urine on

724

Active and Nonactive Type Anodes. Environ. Sci. Technol. 2015, 49 (18), 11062–11069.

725

https://doi.org/10.1021/acs.est.5b01675.

726

(77)

Mostafa, E.; Reinsberg, P.; Garcia-Segura, S.; Baltruschat, H. Chlorine Species Evolution during

727

Electrochlorination on Boron-Doped Diamond Anodes: In-Situ Electrogeneration of Cl2, Cl2O

728

and ClO2. Electrochim. Acta 2018, 281, 831–840.

35 ACS Paragon Plus Environment

Environmental Science & Technology

729 730

https://doi.org/10.1016/j.electacta.2018.05.099. (78)

Kosmulski, M. PH-Dependent Surface Charging and Points of Zero Charge. IV. Update and New

731

Approach. J. Colloid Interface Sci. 2009, 337 (2), 439–448.

732

https://doi.org/10.1016/j.jcis.2009.04.072.

733

(79)

Dunn, W. W.; Aikawa, Y.; Bard, A. J. Characterization of Particulate Titanium Dioxide

734

Photocatalysts by Photoelectrophoretic and Electrochemical Measurements. J. Am. Chem.

735

Soc. 1981, 103 (12), 3456–3459. https://doi.org/10.1021/ja00402a033.

736

(80)

737 738

Boxall, C.; Kelsall, G. H. Photoelectrophoresis of Colloidal Semiconductors, Part 1.- The Technique and Its Applications. J.Chem.Soc.Faraday Trans. 1991, 87 (21), 3537–3545.

(81)

Vidal, A.; Herrero, J.; Romero, M.; Sanchez, B.; Sanchez, M. Heterogeneous Photocatalysis:

739

Degradation of Ethylbenzene in TiO2 Aqueous Suspensions. J. Photochem. Photobiol. A Chem.

740

1994, 79 (3), 213–219. https://doi.org/http://dx.doi.org/10.1016/1010-6030(93)03763-7.

741

(82)

Dang, T. T. T.; Le, S. T. T.; Channei, D.; Khanitchaidecha, W.; Nakaruk, A. Photodegradation

742

Mechanisms of Phenol in the Photocatalytic Process. Res. Chem. Intermed. 2016, 42 (6),

743

5961–5974. https://doi.org/10.1007/s11164-015-2417-3.

744

(83)

Khunphonoi, R.; Grisdanurak, N. Mechanism Pathway and Kinetics of P-Cresol Photocatalytic

745

Degradation over Titania Nanorods under UV-Visible Irradiation. Chem. Eng. J. 2016, 296,

746

420–427. https://doi.org/10.1016/j.cej.2016.03.117.

747

(84)

748 749

Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental Implications of Hydroxyl Radicals (•OH). Chem. Rev. 2015, 115 (24), 13051–13092. https://doi.org/10.1021/cr500310b.

(85)

Devi, L. G.; Krishnamurthy, G. TiO2- and BaTiO3-Assisted Photocatalytic Degradation of

750

Selected Chloroorganic Compounds in Aqueous Medium: Correlation of

751

Reactivity/Orientation Effects of Substituent Groups of the Pollutant Molecule on the

752

Degradation Rate. J. Phys. Chem. A 2011, 115 (4), 460–469. 36 ACS Paragon Plus Environment

Page 36 of 38

Page 37 of 38

Environmental Science & Technology

753 754

https://doi.org/10.1021/jp103301z. (86)

Pang, X.; Chen, C.; Ji, H.; Che, Y.; Ma, W.; Zhao, J. Unraveling the Photocatalytic Mechanisms

755

on TiO2surfaces Using the Oxygen-18 Isotopic Label Technique. Molecules 2014, 19 (10),

756

16291–16311. https://doi.org/10.3390/molecules191016291.

757

(87)

Zanoni, M. V. B.; Sene, J. J.; Selcuk, H.; Anderson, M. A. Photoelectrocatalytic Production of

758

Active Chlorine on Nanocrystalline Titanium Dioxide Thin-Film Electrodes. Environ. Sci.

759

Technol. 2004, 38 (11), 3203–3208. https://doi.org/10.1021/es0347080.

760

(88)

An, T.; Zhang, W.; Xiao, X.; Sheng, G.; Fu, J.; Zhu, X. Photoelectrocatalytic Degradation of

761

Quinoline with a Novel Three-Dimensional Electrode-Packed Bed Photocatalytic Reactor.

762

2004, 161, 233–242. https://doi.org/10.1016/j.nainr.2003.08.004.

763

(89)

Selcuk, H. Disinfection and Formation of Disinfection By-Products in a Photoelectrocatalytic

764

System. Water Res. 2010, 44 (13), 3966–3972.

765

https://doi.org/10.1016/J.WATRES.2010.04.034.

766

(90)

Boxall, C. The Electrophoresis of Semiconductor Particles. Chem. Soc. Rev. 1994, 23, 137–145.

767

37 ACS Paragon Plus Environment

Environmental Science & Technology

768

TOC/ABSTRACT GRAPHIC

769

770

38 ACS Paragon Plus Environment

Page 38 of 38