Enhancement of H2O2 Decomposition by the Cocatalytic Effect of

Sep 4, 2018 - The greatest problem in the Fe(II)/H2O2 Fenton reaction is the low production of •OH owing to the inefficient Fe(III)/Fe(II) cycle and...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIV OF DURHAM

Remediation and Control Technologies

Enhancement of H2O2 Decomposition by the Cocatalytic Effect of WS2 on the Fenton Reaction for the Synchronous Reduction of Cr(VI) and Remediation of Phenol Chencheng Dong, Jiahui Ji, Bin Shen, Mingyang Xing, and Jinlong Zhang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b02403 • Publication Date (Web): 04 Sep 2018 Downloaded from http://pubs.acs.org on September 5, 2018

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 40

Environmental Science & Technology

1

Enhancement of H2O2 Decomposition by the Cocatalytic Effect of WS2 on the

2

Fenton Reaction for the Synchronous Reduction of Cr(VI) and Remediation of

3

Phenol

4

Chencheng Dong, Jiahui Ji, Bin Shen, Mingyang Xing* and Jinlong Zhang*

5

Key Laboratory for Advanced Materials and Institute of Fine Chemicals, School of

6

Chemistry & Molecular Engineering, East China University of Science and

7

Technology, 130 Meilong Road, Shanghai 200237, P.R. China.

8 9

ABSTRACT

10

The greatest problem in the Fe(II)/H2O2 Fenton reaction is the low production of •OH

11

owing to the inefficient Fe(III)/Fe(II) cycle and the low decomposition efficiency of

12

H2O2 ( 420 nm, 500 W).

94

Simultaneous Oxidation of Phenol and Reduction of Cr(VI). The Fenton

95

oxidation of phenol and the synergistic reduction of Cr(VI) were tested under visible

96

light irradiation by using a tungsten lamp (> 420 nm, 500 W). In a typical experiment,

97

certain amounts of ferrous sulfate, hydrogen peroxide, and WS2 were suspended in 100

98

mL 10 mg/L phenol solution with the addition of 1.0 mL K2Cr2O7 (40 mg/L) aqueous

99

solution. In addition, the pH of the suspension was adjusted by using 0.1 M H2SO4

100

solution and 0.1 M NaOH solution. During visible light irradiation, a certain amount of

101

suspension was collected from the reaction cell at given time intervals and then

102

centrifuged to remove the catalysts. The concentration of Cr(VI) was determined by the

103

diphenylcarbazide (DPC) method.54 The absorbance of sample solutions was measured

104

by a UV-vis spectrometer at 540 nm after full color development. The concentration of

105

phenol was measured by a High Performance Liquid Chromatography (HPLC) system.

106

Detection of Generated Hydroxyl Radicals. The number of •OH molecules

107

generated during the WS2 cocatalytic AOP was detected by measurement of the

108

photoluminescence (PL) signal of hydroxybenzoic acid resulting from the capture of

109

•OH by benzoic acid. The details are as follows: 400 mg WS2, 4 mg FeSO4·7H2O and

ACS Paragon Plus Environment

Page 4 of 40

Page 5 of 40

Environmental Science & Technology

110

a certain amount of benzoic acid (0.2 mmol/L) were mixed into 100 mL H2O. Then, 4.0

111

μL H2O2 (30% wt) was added into the solution by pipette. After vortexing for 10

112

seconds, the mixed solution was placed under visible light irradiation for 30 min. The

113

solution was filtered, and the filtrate was measured by PL emission spectroscopy to

114

indirectly measure the amount of •OH (excitation wavelength: 330 nm).

115

Detection of Fe(II) Ions. In this work, to detect the ferrous ions, we employed

116

deionized water instead of phenol solution. Certain amounts of ferrous sulfate,

117

hydrogen peroxide, and WS2 were added into the water. In addition, the pH of the

118

suspension was adjusted by using 0.1 M H2SO4 solution and 0.1 M NaOH solution.

119

During visible light irradiation, approximately 1.0 mL of the suspension was collected

120

from the reaction cell at given time intervals. The 1.0 mL supernatant was mixed with

121

1.0 mL 1 mg/mL 1,10-phenanthroline monohydrate. After that, the absorbance was

122

investigated by UV-vis spectroscopy to evaluate the corresponding amounts of Fe2+

123

ions.

124

Characterization. The concentration of the pollutant was measured using a UV-vis

125

spectrophotometer (Shimadzu, UV-2450). Raman measurements were performed at

126

room temperature using a Via+ Reflex Raman spectrometer with an excitation

127

wavelength of 514 nm. Transmission electron microscopy (TEM) was conducted on a

128

JEOL JEM-2100EX electron microscope, operated at an accelerating voltage of 200 kV.

129

The intensity of hydroxyl radicals was also measured using luminescence spectrometry

130

(Cary Eclipse) at room temperature under an excitation wavelength of 330 nm. X-ray

131

diffraction (XRD) measurements were performed with a Rigaku Ultima IV (Cu Ka

132

radiation, λ = 1.5406 Å) in the range of 10-80°(2θ). The instrument employed for XPS

133

measurements was a Perkin-Elmer PHI 5000C ESCA system with Al Kα radiation

134

operated at 250 W. The total organic carbon (TOC) concentration of the degradation

135

agent was recorded using a SHIMADZU TOC-L CPN analyzer.

136

3. RESULTS AND DISCUSSION

ACS Paragon Plus Environment

Environmental Science & Technology

137

WS2 Cocatalytic AOP Performance for the Remediation of Phenol. Phenol is

138

commonly known as a general reagent in chemical analysis, especially for the

139

manufacture of medical and industrial organic compounds.55, 56, 57 In addition, phenol

140

is a common component of oil refinery wastes, which may enter the environment via

141

discharges from oil refineries, coal conversion plants, and municipal waste treatment

142

plants.58 Due to the stability of its aromatic ring and the hydrophilicity of its hydroxyl

143

group, phenol has served as a common model pollutant for the development of AOPs.59-

144

61

145

performance. Thus, we firstly carried out systematic studies to reveal the substantial

146

impact of WS2 on cocatalytic performance for H2O2 decomposition. To be more specific,

147

we used the degradation of phenol as a model reaction and investigated a number of

148

factors that may influence the performance of AOPs, such as the pH value and the

149

dosages of WS2, Fe(II) and H2O2, as shown in Figure 1.

AOPs are always considered complicated systems, and many factors influence their

150 151

Figure 1. Different influencing factors of the WS2-cocatalytic Fenton process under

152

visible light irradiation (λ> 420 nm) for the degradation of phenol (10 mg/L). (a) pH

153

value influence (100 mL solution including 0.04 g/L Fe(SO4)·7H2O, 4.0 g/L WS2, and

154

0.1 mmol/L H2O2). (b) Various concentrations of Fe(SO4)·7H2O (100 mL solution

155

including 4.0 g/L WS2 and 0.1 mmol/L H2O2, pH = 3.8). (c) Various concentrations of

ACS Paragon Plus Environment

Page 6 of 40

Page 7 of 40

Environmental Science & Technology

156

WS2 (100 mL solution including 0.02 g/L Fe(SO4)·7H2O and 0.1 mmol/L H2O2, pH =

157

3.8). (d) Various concentrations of H2O2 (100 mL solution including 0.02 g/L

158

Fe(SO4)·7H2O and 4.0 g/L WS2, pH = 3.8).

159

Some reports have demonstrated that photoassisted AOPs are a very efficient way to

160

further improve the mineralization of organic pollutants.62-64 Hence, in our case, we

161

carried out all the Fenton reaction experiments under visible light irradiation (λ > 420

162

nm). First, we explored the impact of pH value on the photo-Fenton reaction for the

163

degradation of phenol. As seen from Figure 1a, when the pH value is fixed at 3.8, WS2

164

exhibits the optimal cocatalytic activity for the photo-Fenton reaction. L. Clarizia et

165

al.46 have reported that too-low or too-high pH values decrease the concentration of

166

reactive species of FeOH2+. Indeed, at pH values higher than 4.0, dissolved iron

167

precipitates as ferric hydroxide, which leads to catalyst poisoning. Afterward, we tested

168

the activity of WS2 cocatalytic AOPs with various amounts of Fe(II) ions. The results

169

demonstrate that when ferrous sulfate is fixed at 0.04 g/L, it shows the highest

170

efficiency (Figure 1b). Then, we investigated the effects of the amounts of added WS2

171

and H2O2 on the degradation of phenol and found that the optimal amounts of WS2 and

172

H2O2 were fixed at 4.0 g/L and 0.4 mmol/L, respectively (Figure 1c, d). Excessive Fe(II)

173

ions or H2O2 would act as •OH scavengers, which are harmful to the enhancement effect

174

of the Fenton reaction.65, 66 Meanwhile, too much WS2 powders would weaken the light

175

absorption.

ACS Paragon Plus Environment

Environmental Science & Technology

176 177

Figure 2. (a) Activity comparison of various Fenton reactions for the degradation of

178

phenol (10 mg/L). (100 mL solution including 0.04 g/L Fe(SO4)·7H2O, 4.0 g/L WS2,

179

and 0.4 mmol/L H2O2, pH= 3.8; Vis: under visible light illumination (λ> 420 nm), Dark:

180

under dark conditions). (b) EPR spectra for the detection of •OH in the presence of 50

181

µL 5,5-dimethyl-1-pyrrolidine-N-oxide (DMPO, 0.14 M) at room temperature. The

182

EPR signals are marked as follows: green rhombus, hydroxyl free radicals.67 (c) PL

ACS Paragon Plus Environment

Page 8 of 40

Page 9 of 40

Environmental Science & Technology

183

spectra of hydroxybenzoic acid generated under different conditions. The

184

decomposition efficiency of H2O2 can be obtained by the division of PL intensities.

185

Ultimately, in our case, the optimal visible-light-driven (λ > 420 nm) Fenton reaction

186

conditions were as follows: 100 mL of phenol solution at pH 3.8, containing 4.0 mg

187

FeSO4·7H2O, 400 mg WS2, and 4.0 µL H2O2. As shown in Figure 2a, under the optimal

188

conditions, the WS2 cocatalytic AOPs using 0.04 g/L Fe(SO4)·7H2O, 4.0 g/L WS2 and

189

0.4 mmol/L H2O2 display a high phenol degradation rate of 81% under visible light

190

irradiation for 1 min only, which is much better than the degradation rate of the Fenton

191

reaction in the absence of WS2 (43%). The low efficiency of pure WS2 in the dark and

192

under light irradiation could negate the adsorption capacity and photocatalytic activity

193

of WS2. In addition, the poor activity of WS2+H2O2 indicates that WS2 cannot directly

194

decompose H2O2. It is worth mentioning that the activity of WS2 cocatalytic AOPs

195

under visible light irradiation is better than that in the dark because light is beneficial

196

to the decomposition of H2O2 and the conversion of Fe3+ to Fe2+.68 After introducing

197

visible light, the efficiency of conventional Fenton reaction has a little increase, owing

198

to the low recycling efficiency of iron ions in the absence of WS2. Although a

199

significant enhancement can already be obtained in the dark, performing the reaction

200

under visible light irradiation may further improve the oxidation power of AOPs. On

201

the other side, the commercial WS2’s BET surface area (6 m2·g-1) is too small to adsorb

202

phenol molecules. Nevertheless, we have done a comparison of WS2 co-catalytic

203

Fenton system with the adsorbent systems for the removal of phenol, as shown in Table

204

S1. As a result, the WS2 co-catalytic Fenton system exhibits a high phenol removal rate

205

of 240 mg·g-1·h-1, which is much better than the reported adsorbent systems (0.44~

206

44.82 mg·g-1·h-1). 5,5-Dimethyl-1-pyrrolidine-N-oxide (DMPO) was used as the

207

electron paramagnetic resonance (EPR) probe to detect the •OH in the Fenton reaction.

208

Compared with the very weak EPR signal for WS2+H2O2, the FeSO4+H2O2 and

209

FeSO4+WS2+H2O2 systems displayed the characteristic quartet signals of •OH (Figure

210

2b).67 Furthermore, the FeSO4+WS2+H2O2 system showed the strongest EPR signals,

211

supporting the observation of the efficient cocatalytic activity of WS2 for the

ACS Paragon Plus Environment

Environmental Science & Technology

212

decomposition of H2O2. In addition, the production of •OH in AOPs was measured by

213

observing the increase in the PL signal of hydroxybenzoic acid resulting from the

214

capture of •OH by benzoic acid.69 Compared with the data for FeSO4+H2O2, the

215

FeSO4+WS2+H2O2 solution displays a significantly enhanced PL signal (Figure 2c), an

216

observation that further confirms the cocatalytic effect of WS2 on the decomposition of

217

H2O2 into •OH. The tert-butyl alcohol (TBA) are always employed as the quenching

218

agents of •OH radicals.70 Seen from Figure S1, after adding TBA (100 mg/L) into the

219

phenol (10 mg/L) solution, the phenol removal rate of cocatalytic Fenton system has an

220

obvious decrease from 81% to 22%. The result no doubt confirms that •OH radicals

221

play a key role in cocatalytic Fenton reaction for the degradation of organic pollutants.

222

In addition, we also employed AgNO3, 1,4-benzoquinone (PBA) and CH3OH as

223

scavengers for electrons, superoxide radicals and holes, respectively. Seen from Figure

224

S2, the phenol removal rate was obviously impeded by the adding of AgNO3 or CH3OH.

225

It is because that the AgNO3 and CH3OH would react with the Fe2+ (or W4+) and •OH

226

radicals respectively, which is harmful to the Fenton reaction for the oxidation of phenol.

227

On the other hand, the adding of AgNO3 would greatly hinder the reduction of Cr6+,

228

indicating the Fe2+ plays the key role for the Cr6+ reduction in the cocatalytic Fenton

229

system. The AgNO3 would react with Fe2+ and W4+ to greatly decrease the

230

concentration of Fe2+ in the Fenton process. The decomposition efficiency of H2O2

231

represents an important index to measure in order to evaluate the performance potential

232

of AOPs. The H2O2 (0.4 mmol/L) was heated at 90 °C for 120 min in order to

233

completely decompose the H2O2 into •OH, as shown in Figure S3. After careful

234

calculation by the division of PL intensities in Figure 2c, the conversion efficiency of

235

H2O2 into •OH was as high as 60.1% by the promotion of WS2, which is much higher

236

than the yield of 22.9% for the conventional Fenton reaction in the absence of WS2.

ACS Paragon Plus Environment

Page 10 of 40

Page 11 of 40

Environmental Science & Technology

237 238

Figure 3. Schematic diagram of the photo-Fenton degradation mechanism of phenol in

239

the presence or absence of WS2. Insets are the mass spectra of intermediates observed

240

from high-performance liquid chromatography (HPLC) that appeared in the

241

degradation of phenol after 30 min.

242

In addition, we investigated the degradation mechanism of phenol in the WS2

243

cocatalytic Fenton system, as shown in Figures 3 and S4. In the original phenol solution,

244

there was an obvious mass spectra peak at m/z=78, which was assigned to the benzene

245

ring (Figure S4). After the Fenton reaction, the benzene ring is disintegrated into a

246

variety of intermediate products. Among them, various organic acids and aldehydes are

247

the main intermediate products. Interestingly, compared with the traditional Fenton

248

process, there are significantly fewer intermediates from the WS2 cocatalytic Fenton

249

reaction in the degradation of phenol (Figure 3), indicating that the cocatalytic system

250

has a faster reaction rate for the remediation of organic pollutants. Another interesting

251

finding is that after 30 min of reaction, the content of small-molecule fatty acids in the

252

WS2 cocatalytic system is much lower than that of the traditional Fenton reaction (insets

253

of Figure 3), suggesting the better mineralization of the cocatalytic system for the

254

degradation of phenol.

255

Apparent Kinetic Modeling. In the classic Fenton process, zero-,71 first-,72 and

256

second-order73 reaction kinetics have been used to study the degradation of phenolic

257

pollutants. In our case, the reaction rate equation can be described as the following Eq.

258

4 or expressed in terms of logarithms (Eq. 5) owing to the presence of the cocatalyst

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 40

259

WS2. In Eqs. 4 and 5, a, b, and c represent the reaction order, and K represents the total

260

reaction rate constant. Here, we specifically analyzed three parameters: the

261

concentrations of ferrous sulfate, hydrogen peroxide and tungsten disulfide. For this

262

analysis, we assumed that the initial concentration of tungsten disulfide in the Fenton

263

system is [WS2]0, the concentration of ferrous ion is [Fe2+]0, the concentration of

264

hydrogen peroxide is [H2O2]0, and the rate at t = 0 can be expressed as in Eq. 6 or in the

265

logarithmic form of Eq. 7.

266

V  (dc / dt )  K [Fe 2+ ]a [H 2O2 ]b [WS2 ]c

(4)

-lg(dc / dt )  lg(K )  a  lg[Fe 2+ ]+b  lg[H 2O2 ]+c  lg[WS2 ]

(5)

V  (dc / dt )  K [Fe 2+ ]0a [H 2O2 ]0b [WS2 ]0c

(6)

-lg(dc / dt )  lg(K )  a  lg[Fe 2+ ]0 +b  lg[H 2O2 ]0 +c  lg[WS2 ]0

(7)

267

We employed a series of ferrous solutions with different concentrations in the phenol

268

solution, but the same concentrations of tungsten disulfide and hydrogen peroxide were

269

used. Then, we recorded the curve of phenol concentration versus time. We can obtain

270

the equations of these curves, which use t as the independent variable and c as the

271

dependent variable, and the derivative of the equation that is equal to -lg(dc/dt) at

272

different [Fe2+]0 values when t = 0. Meanwhile, we can obtain a group of related data,

273

and -lg(dc/dt) is inversely related to lg[Fe2+]0. Since the amount of tungsten disulfide

274

and hydrogen peroxide is fixed, lgK + b[H2O2]0 + c[WS2]0 is constant. Thus, -lg(dc/dt)

275

is linear with respect to lg[Fe2+]0. We can obtain a straight line by using lg[Fe2+]0 as the

276

abscissa and -lg(dc/dt) as the y-axis. The slope is a, and the intercept is lgK + b[H2O2]0

277

+ c[WS2]0. Similarly, we can obtain the data for b, lgK + alg[Fe2+]0 + c[WS2]0, c and

278

lgK + alg[Fe2+]0 + b[H2O2]0; thereby, we can obtain the value of K. Accordingly, the

279

apparent kinetic equation of the WS2 cocatalytic Fenton reaction for the degradation of

280

phenol can be obtained. The details of these calculations and results are shown in Figure

281

4 and Tables S2~S4.

ACS Paragon Plus Environment

Page 13 of 40

Environmental Science & Technology

282 283

Figure 4. Apparent kinetic calculations. (a) The fitting line between -lg(dc/dt) and

284

lg[Fe2+]0 under visible light irradiation. (b) The fitting line between -lg(dc/dt) and

285

lg[H2O2]0 under visible light irradiation. (c) The fitting line between -lg(dc/dt) and

286

lg[WS2]0 under visible light irradiation. We chose the equation model ExpDec1: y =

287

y0+A1e-x/t to fit our degradation data. As illustrated in Tables S2~S4, we acquired a

288

series of fitting equations for the degradation rate under visible light. For t = 0, the

289

derivatives of the equations are listed in Tables S2~S4.

290

We employed a different initial concentration of FeSO4 to explore its kinetic model.

291

By utilizing the above method, we obtained fitting equations for the degradation

ACS Paragon Plus Environment

Environmental Science & Technology

292

experiments (as listed in Table S2). The derivative of the equation -lg(dc/dt)0 is listed

293

in Table S2 as well. Under visible conditions, -lg(dc1/dt) = 0.68999, -lg(dc2/dt) =

294

0.73225, -lg(dc3/dt) = 0.31878, and -lg(dc4/dt) = 0.20616. As mentioned above, -

295

lg(dc/dt) is inversely proportional to lg[Fe2+]0. Thus, we can obtain a fitting line (Figure

296

4a) and Eq. 8. This indicates that the reaction order a is equal to -0.85691 and that lgK

297

+ blg[H2O2]0 + clg[WS2]0 = -0.67339.

298

-lg  dc/dt  = -0.67339-0.85691 lg[FeSO4 ]0

(8)

299

Similarly, we chose a series of different initial concentrations of H2O2 to represent

300

its kinetic model. By employing the above calculation method, we obtained fitting

301

equations for the degradation experiments, as listed in Table S3. The derivative of the

302

equation -lg(dc/dt)0 is listed in Table S3 as well. As a result, under visible conditions,

303

we can obtain a fitting line (Figure 4b) and the equation of Eq. 9. Regarding the different

304

initial concentrations of the WS2 cocatalyst, the fitting equations supported by Origin

305

software are listed in Table S4. Unsurprisingly, we can obtain the fitting line shown in

306

Figure 4c and Eq. 10.

307

308

[ 0.523 -lg(dc/dt)=0.5896+( )e 0.514 π/2

-2(lg(H 2 O2 )0 +0.377)2

[ 0.0817 -lg(dc/dt)=0.1694+( )e 0.1231 π/2

0.5142

]

-2(lg(WS2 )0 -0.608)2 0.12312

(9)

]

(10)

309

To further explore the total reaction rate constant K, we combined three equations

310

(Eqs. 8~10) to obtain the apparent kinetic equation of Eq. 11: 3lgK + 2alg[FeSO4]0 +

311

2blg[H2O2]0 + 2clg[WS2]0= 0.08561, where [FeSO4]0 = 0.02 g/L, [H2O2]0 = 0.1 mmol/L,

312

[WS2]0 = 4 g/L, a =-0.85691, b = 0.81186, and c =0.52955. Thus, K equals 0.24359,

313

and the total reaction order of a + b + c equals 0.4845. Careful kinetic measurements

314

revealed that the reaction order of [WS2] is between zero- and first-order kinetics, which

315

indicates that increased WS2 is beneficial for the degradation of phenol. In addition,

316

compared with the very slow rate-limiting step of Eq. 2 for the conventional Fenton

ACS Paragon Plus Environment

Page 14 of 40

Page 15 of 40

Environmental Science & Technology

317

reaction, the addition of the cocatalyst WS2 can greatly improve the total reaction rate.

318

V=-lg  dc/dt  = 0.24359[FeSO4 ]-0.85691[H2O2 ]0.81186 [WS2 ]0.52955

(11)

319

Mechanism Investigation. To investigate the mechanism of the WS2 cocatalytic

320

effect in the Fe(II)/H2O2 Fenton system, we detected the soluble Fe2+ ions in solution

321

by adding 1,10-phenanthroline monohydrate (Figure 5a).74 Unsurprisingly, in the

322

traditional Fenton reaction, there is a negligible peak ascribed to Fe2+ due to the rapid

323

oxidation of Fe2+ to form Fe3+ (Eq. 1). Interestingly, after adding the cocatalyst WS2,

324

an obvious peak at 511 nm assigned to Fe2+ can be observed owing to the acceleration

325

of the Fe(III)/Fe(II) cycle. In addition, the inset picture shows a color comparison

326

between the conventional Fenton and WS2 cocatalytic Fenton process. That is, the color

327

of the solution in the presence of WS2 is much darker than that of the solution without

328

WS2, indicating the higher concentration of Fe2+ in the WS2 cocatalytic system. Hence,

329

we can conclude that WS2 can indeed promote the reduction of Fe(III) to Fe(II) in the

330

Fe(II)/H2O2 Fenton system, which plays the key role in H2O2 decomposition.

331

ACS Paragon Plus Environment

Environmental Science & Technology

332 333

Figure 5. (a) Fe(II) concentration detection: UV-vis spectra of the conventional Fenton

334

process (Fe2++H2O2) and WS2 cocatalytic Fenton process with the addition of 1,10-

335

phenanthroline monohydrate (inset picture shows a photograph of this extracted

336

solution). (b) Low-temperature EPR spectra of WS2 after being washed by water (g =

337

2.0). (c) UV spectra of AgNO3, WS2 supernatant, and their mixture. (d) W4f XPS

338

spectra of WS2 before and after mixing with commercial Fe3O4 particles.

339

To investigate the reduction of Fe(III) by the cocatalytic effect of WS2 in the Fenton

340

process, we examined the surface chemical properties of WS2. We found an interesting

341

phenomenon in that WS2 is always acidic (pH = 2.5) in aqueous solution. Therefore,

342

we can infer that WS2 could release a large number of protons in solution. To determine

343

why, we performed EPR analysis at room temperature. Interestingly, WS2 after water

344

washing treatment shows an obvious EPR signal at g = 2.0 (Figure 5b), which can be

345

ascribed to the sulfur vacancies generated on the surface of WS2. It is generally accepted

346

that the unsaturated S atoms capture protons during H2 evolution.75-77 However, in our

347

case, the unsaturated S atoms could be removed from the WS2 surface by binding with

ACS Paragon Plus Environment

Page 16 of 40

Page 17 of 40

Environmental Science & Technology

348

protons to form H2S,76 leading to the low pH value of the WS2 aqueous solution. To

349

further demonstrate the release of H2S, we compared the extinction spectra of an

350

aqueous solution of pure AgNO3, the supernatant of a WS2 aqueous dispersion, and a

351

mixture of AgNO3+WS2, as shown in Figure 5c. A broad absorption curve at 267 nm in

352

the UV range corresponding to Ag2S was observed,78 indicating the loss of S atoms

353

(H2S) from WS2. The loss of S atoms exposes more W4+ active sites, which are easily

354

oxidized to W6+ by the adsorbed Fe3+. In addition, we calculated the released content

355

of H2S during the Fenton reaction, which is approximately 0.05 mg/L according to

356

inductively coupled plasma optical emission spectrometry (ICP-OES), indicating

357

negligible secondary pollution.

358

To further investigate the reduction capacity of WS2 in the Fenton reaction, we added

359

WS2 to the heterogeneous Fenton reaction of the Fe3O4/H2O2 system. The W3d XPS

360

spectra, shown in Figure 5d, can be used to detect the variation of states in WS2+Fe3O4.

361

The peaks at 31.9 and 34.1 eV were assigned to W4+, and a distinct characteristic peak

362

of W6+ was visible at 35.2 and 37.5 eV after WS2 was mixed with Fe3O4, indicating the

363

oxidation of W4+ to W6+ by forming a W6+-O-Fe bond. Meanwhile, there is an obvious

364

“blueshift” for the S2p characteristic XPS peak after the WS2 mixed with the Fenton

365

reagent (Figure S5) owing to the formation of sulfur vacancies on the surface of WS2.

366

Although the Raman peak positions of WS2 were almost unchanged after mixing with

367

Fe3O4 (Figure S6), the intensity of the peaks changed significantly: the characteristic

368

peak of the A1g mode decreased obviously after the Fenton reaction, which indicates

369

the generation of sulfur vacancies on the surface of WS2.79, 80 The above results indicate

370

that the release of H2S and the generation of S vacancies are responsible for the

371

production of W6+-O-Fe bonds between WS2 and Fe3O4, which results from the

372

reduction of Fe3+ to Fe2+ by the exposed W4+. Interestingly, bond formation can also be

373

evidenced by the fact that a simple mixture of WS2 and Fe3O4 powders in aqueous

374

solution can be completely separated by applying an external magnetic field (Figure

375

S7). To understand the source of oxygen in the W6+-O-Fe bonds, we mixed WS2 and

376

Fe3O4 powders in the anaerobic solvent of cyclohexane. As shown in Figure S8, the

ACS Paragon Plus Environment

Environmental Science & Technology

377

resulting mixture can also be completely separated by magnet, indicating that the

378

oxygen in the W6+-O-Fe bonds originated from Fe3O4 rather than water molecules or

379

dissolved oxygen. To further eliminate the effect of dissolved oxygen on homogeneous

380

Fenton performance, we performed the WS2 cocatalytic Fe2+/H2O2 Fenton experiment

381

in the absence of oxygen by N2-bubbling treatment, as shown in Figure S9. In the

382

absence of dissolved oxygen, WS2 still shows a very high cocatalytic Fenton activity

383

for the degradation of phenol. To demonstrate the reduction of W6+ by H2O2 during the

384

Fenton reaction, we added H2O2 (4 mmol/L) to the mixed WS2 and Fe3O4 solution. The

385

WS2 powders could not be separated by magnet (Figure S10) mainly due to the

386

disruption of W6+-O-Fe bonds by the reduction of W6+ to W4+.

387

Figure S11 illustrates the proposed mechanism of the WS2 cocatalytic effect in the

388

Fe(II)/H2O2 Fenton reaction. The first step is the capture of protons by unsaturated S

389

atoms on the surface of WS2 to form H2S molecules. As H2S is formed, many W4+

390

active sites are exposed on the surface of WS2 and can be easily oxidized by Fe3+ ions

391

to generate W6+. In addition, the oxidation reaction is accompanied by the reduction of

392

Fe3+ to Fe2+, thus greatly improving the reaction rate of the originally rate-limiting step

393

Eq. 2 in the conventional Fe(II)/H2O2 Fenton system. Afterward, the further reduction

394

of W6+ back to W4+ with the help of H2O2 based on the Fenton reaction ensures the

395

cocatalytic cycling of WS2. WS2 greatly enhances the conversion efficiency from Fe3+

396

to Fe2+, which is very helpful to the decomposition of H2O2 and the suppression of iron

397

sludge in the Fenton reaction. As seen from the above cocatalytic mechanism, the

398

addition of WS2 can not only promote the Fe3+/Fe2+ cycle reaction but also cause the

399

Fenton reaction to exhibit oxidation (·OH) and reduction (Fe2+) activity at the same

400

time, which is different from conventional AOPs.81 Therefore, we can boldly predict

401

that the WS2 cocatalytic Fenton reaction could achieve the synchronous reduction of

402

heavy metal ions and remediation of organic pollutants.

403

Simultaneous Oxidation of Phenol and Reduction of Cr(VI). Toxic inorganic

404

pollutants, especially those such as hexavalent chromium (Cr(VI)) that result from the

405

manufacture of paints, ceramics and corrosion inhibitors, are usually associated with

ACS Paragon Plus Environment

Page 18 of 40

Page 19 of 40

Environmental Science & Technology

406

toxicity and bioaccumulation, even at a low concentration.82 As a result, Cr(VI) has

407

been listed as one of the priority pollutants regulated by the United States

408

Environmental Protection Agency (USEPA) due to its potential threat to public health.83,

409

84

410

water treatment, even though Cr(VI) ions can be directly reduced by H2O2 to form

411

Cr(III), which also participates in a Fenton-like reaction.81, 85 However, the problem is

412

that Cr(III) is easily precipitated as insoluble chromium hydroxide [Cr(OH)3] in neutral

413

and alkaline conditions (pH > 5), which terminates the decomposition of H2O2,81 and

414

under acidic conditions, free Cr(III) in the form of [Cr(H2O)6]3+ is completely

415

unreactive toward H2O2.86 To maintain the oxidation activity of the Cr(VI)/H2O2

416

Fenton-like reaction, we must keep a high concentration of Cr(VI) in the system, which

417

means the reaction system is still very toxic. There is thereby an urgent need to develop

418

advanced AOPs for the synchronous reduction of Cr6+ and remediation of organic

419

pollutants, but this is still a significant challenge. Here, to highlight the potential of WS2

420

cocatalytic AOPs for industrial environmental applications, we attempted to apply the

421

WS2 cocatalytic Fenton system to the simultaneous oxidation of phenol and reduction

422

of Cr(VI), according to the excellent reduction performance of exposed W4+ in WS2.

The treatment of Cr(VI) has been an ongoing challenge in wastewater and drinking

423 424

Figure 6. Simultaneous (a) reduction of Cr(VI) and (b) degradation of phenol via

425

different Fenton reactions (100 mL solution including 0.04 g/L Fe(SO4)·7H2O, 4.0 g/L

426

WS2, 0.4 mmol/L H2O2, 10 mg/L phenol, and 40 mg/L Cr(VI), pH= 3.8, Vis: under

ACS Paragon Plus Environment

Environmental Science & Technology

427

visible light illumination (λ> 420 nm)). (c) Cycle test of the visible-light-driven WS2

428

cocatalytic Fenton reaction for the simultaneous degradation of phenol and reduction

429

of Cr(VI). (d) Reduction of CrO42- ions in the dark (in the absence of phenol and H2O2,

430

100 mL solution including 40 mg/L Cr(VI) and 4.0 g/L WS2, 0.04 g/L Fe2(SO4)3 or 0.04

431

g/L Fe(SO4)·7H2O, pH= 3.8). (e) Reduction of CrO42- and Fe3+ with WS2 in the dark

432

(100 mL solution including 4.0 g/L WS2, 0.04 g/L Fe2(SO4)3 and 40 mg/L Cr(VI), pH=

433

3.8; KSCN and o-phenanthroline were employed as the color-developing agents for the

434

detection of Fe3+ and Fe2+ by UV-visible spectrophotometry, respectively). (f) Fe2+

435

concentration detection: statistics for the absorbance of the color-developing agent (o-

436

phenanthroline) after the complexing of Fe2+.

437

As seen from Figure 6a, when phenol and Cr(VI) ions coexist, the WS2 cocatalytic

438

system exhibits a significant enhancement to the reduction rate of Cr(VI), compared

439

with the conventional Fenton reaction. On the other hand, the removal rate of phenol

440

via the WS2 cocatalytic system also shows an increasing trend compared with that of

441

the Fenton reaction (Figure 6b). We have explored the individual effect of H2O2 towards

442

chromium reduction and phenol oxidation, as shown in Figure S12. The result shows

443

the pure H2O2 can only remove about 12% phenol, however, the Cr6+ ions can be

444

reduced about 85% with 30 minutes, owing to the reduction of H2O2.87 Nevertheless,

445

the WS2+H2O2+Fe2+ Fenton system can achieve 80% reduction rate of Cr6+ just within

446

5 minutes (Figure 6a), which is much higher than that of the individual H2O2 system

447

(30% within 5 minutes). In conclusion, the individual H2O2 cannot possess the ability

448

of simultaneous oxidation and reduction. Most importantly, the WS2 cocatalytic system

449

shows an excellent cycling stability for the remediation of phenol and reduction of

450

Cr(VI), as shown in Figure 6c. In addition, the TOC degradation rate of phenol can

451

reach up to 54% after the 5th cycle test. The TEM and XRD results of WS2 before and

452

after the Fenton reaction indicate the microstructure stability of WS2 during the cycle

453

test (Figure S13). After 5 cycle test, the microstructure of WS2 is almost unchanged

454

(Figure S13a), and the HRTEM images indicate that the WS2 after cycle test maintains

455

the lattice spacing of 0.279 nm (Figure S13b), which corresponds to the (100) plane of

ACS Paragon Plus Environment

Page 20 of 40

Page 21 of 40

Environmental Science & Technology

456

WS2.88 The corresponding plane in XRD patterns is shown in Figure S13c. The

457

diffraction peaks located at 14.2o, 28.8o, 33o, 33.7o,39.4o, 44.1o, 49.7o, 58.4o and 60.6o

458

are correspond to (002), (004), (100), (101), (103), (006), (105), (110) and (112) planes,

459

respectively. The data agree well with the standard values of WS2 (JCPDS 08-0237).

460

After 5 cycles test, the planes of WS2 are almost unchanged, indicating the stability of

461

WS2 cocatalytic Fenton system. In the reaction system of coexistent phenol and Cr(VI),

462

the exposed W4+ in WS2 not only can act as the cocatalyst for the Fenton reaction but

463

also plays a key role in the reduction of Cr(VI). We have taken various concentration

464

of phenol and chromium for simultaneous oxidation and reduction. Seen from Figure

465

S14, when the concentration of Cr6+ was confined, the phenol removal rate could be

466

limited slowly with the increasing concentration of phenol. And when phenol was fixed

467

at 5 mg/L, the phenol removal rate can achieve above 99%. When we fixed the

468

concentration of phenol, it was found that the reduction rate of Cr6+ showed a downward

469

trend with the increase concentration of Cr6+ from 40 mg/L to 80 mg/L. In summary,

470

the WS2 co-catalytic Fenton system can indeed realize the synchronous reduction of Cr

471

(VI) and remediation of phenol. To understand the remediation mechanism of Cr(VI)

472

and phenol, the reduction of CrO42- ions in the absence of phenol and H2O2 with

473

different catalysts is shown in Figure 6d. As a result, Fe3+ cannot reduce Cr(VI), but

474

Fe2+ can efficiently reduce Cr(VI). Interestingly, after adding WS2, the Fe3++WS2

475

system shows a significant enhancement in the Cr(VI) reduction rate (98.3%), which is

476

much higher than the sum of the reduction rates of WS2 (53.5%) and Fe3+ (0.4%), owing

477

to the generation of Fe2+. The pure WS2 shows a relatively low activity for the reduction

478

of Cr(VI) because CrO42- is not readily adsorbed on the surface of WS2. In Fig. 6d, it

479

was observed that the WS2+Cr6+ exhibited ~40% efficiency. This is mainly ascribed

480

to the reduction ability of exposed W4+. Compared with CrO42- ions, WS2 can much

481

more easily adsorb Fe3+ ions in an acidic solution (pH ≈ 4.0) owing to the negatively

482

charged surface of WS2.89 We have tested the Zeta potentials of the WS2 aqueous

483

solution under different pH values (pH=2, 3, 5, 7, 9, 11), as shown in Figure S15. As a

484

result, the isoelectric point of WS2 is about 2.29. Hence, we can conclude that the WS2

485

surface is negatively charged at a pH value of 4.0. The adsorbed Fe3+ ions are

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 40

486

continuously reduced by the exposed W4+ to form Fe2+. The regenerated Fe2+ ions are

487

redistributed into aqueous solution for H2O2 decomposition due to the excellent water

488

solubility of Fe2+.90 Therefore, the only possibility is that Fe2+ ions or H2O2 molecules

489

directly reduce the Cr(VI) in the WS2 cocatalytic Fenton system. Actually, Fe2+ or H2O2

490

can individually directly reduce Cr(VI), as shown in Figure S16. However, the mixture

491

of Fe2+ and H2O2 (Fenton system) cannot further improve the Cr(VI) reduction rate

492

owing to the occurrence of the reaction between Fe2+ and H2O2. As a result, the excess

493

Fe3+ produced not only inhibits H2O2 decomposition but also limits Cr(VI) reduction.

494

Figure 6e shows the reduction of CrO42- and Fe3+ with WS2 in the dark. As expected,

495

with decreasing Fe3+ and Cr(VI), there is an increasing trend of Fe2+, confirming that

496

the generation of Fe2+ is responsible for the reduction of Cr(VI) in the WS2 cocatalytic

497

Fenton system.

498

In our case, we demonstrated that the addition of WS2 to the Fenton system can

499

greatly improve the reduction of Fe3+ to Fe2+. Hence, we can infer that in the system

500

designed for the synchronous reduction of Cr(VI) and remediation of phenol, Fe2+ is

501

responsible for the reduction of Cr(VI) and the decomposition of H2O2. We measured

502

the Fe2+ concentration in the synchronous system, as shown in Figure 6f. Interestingly,

503

the Fe2++Cr6++phenol and Fe2++H2O2+Cr6++phenol systems display a decreasing trend

504

of

505

Fe2++H2O2+Cr6++WS2+phenol system shows a significant increasing trend in Fe2+

506

concentration, further confirming that the concentration of Fe2+ is the reason for the

507

synchronous reduction of Cr(VI) and remediation of phenol. Additionally, the activity

508

of the reduction of Cr(VI) and remediation of phenol is directly proportional to the mass

509

of WS2 added, as shown in Figure S17. No obvious poisoning effect was observed,

510

which is consistent with the nature of the cocatalytic effect of WS2 in the Fenton

511

reaction. When the amount of WS2 was increased from 400 mg to 500 mg, the

512

degradation rate of phenol was increased from 57.8% to 80.9%. Meanwhile, the

513

reduction rate of Cr(VI) was improved from 74.0% to 90.9% within merely 1 min of

514

irradiation; both of these rates are much higher than that obtained for the conventional

Fe2+

concentration

with

increasing

reaction

ACS Paragon Plus Environment

time.

Conversely,

the

Page 23 of 40

Environmental Science & Technology

515

Fenton reaction in the absence of WS2. Furthermore, we performed a careful

516

comparison between past reports and our research on the Fenton reaction for the

517

remediation of pollutants, as shown in Table S5. In all the reported literature, it often

518

took a long time for the complete degradation of phenol (> 5 min), but in our case, the

519

WS2 cocatalytic Fenton system required only 1 min. Importantly, the Fenton systems

520

reported in the literature require a very high level of H2O2 (136~6800 mg/L) and Fe2+

521

(10~500 mg/L), but the WS2 cocatalytic Fenton system involved here only needs 13.6

522

mg/L H2O2 and 1.47 mg/L Fe2+. Moreover, the WS2 cocatalytic Fenton system exhibits

523

remarkable stability for the synchronous reduction of Cr(VI) and remediation of phenol,

524

as shown in Figure 6c. Therefore, compared with the reported Fenton reaction, the WS2

525

cocatalytic Fenton system has an efficient and steady capacity for the remediation of

526

organic/inorganic pollutants.

527

4. ENVIRONMENTAL IMPLICATIONS.

528

Unlike traditional organic chelators such as the Fe(III)-salen complex, PCA, and

529

cysteine, WS2 is an inorganic compound that shows very stable physicochemical

530

properties and is widely found in many tungsten ores. The addition of WS2 could

531

effectively promote H2O2 decomposition in the Fenton reaction and enhance the

532

degradation of phenol because it can expose active W4+ to realize the effective

533

Fe(III)/Fe(II) cycle and prevent the precipitation of iron sludge. More importantly, WS2

534

cocatalytic AOPs can achieve the simultaneous mineralization of phenol molecules and

535

reduction of toxic Cr(VI), which can reduce the environmental risk caused by

536

multicomponent wastewater. In addition to the Fe(II)/H2O2 Fenton system, the WS2

537

cocatalytic effect in the remediation of phenol is also suitable for other Fenton-like

538

systems, such as Fe(III)/H2O2 and Ni(II)/H2O2 (Figure S18). Thus, we expect that this

539

work could lead to the development of an attractive and reliable platform for promoting

540

the conversion and mineralization of organic-inorganic contaminants in natural aquatic

541

environments via the WS2 cocatalytic Fenton reaction.

542

ASSOCIATED CONTENT

ACS Paragon Plus Environment

Environmental Science & Technology

543

Supporting Information. The Supporting Information is available free of charge on the ACS

544

Publications website at DOI: XXXXX.

545

Detailed reaction kinetics data; Mass spectra of intermediates; H2O2 total

546

decomposition by heating treatment; EPR, TEM and XRD spectra of WS2; XPS spectra

547

of S2p; Digital image of magnetic separation; Simultaneous degradation of phenol and

548

reduction of Cr(VI) with different amounts of WS2; WS2 cocatalytic Fenton-like

549

reactions.

550

AUTHOR INFORMATION

551

Corresponding Author

552

*Email: [email protected]; [email protected]

553

Author Contributions

554

The manuscript was written with contributions from all authors. All authors have

555

approved the final version of the manuscript.

556

Notes

557

The authors declare no competing financial interest.

558

ACKNOWLEDGMENT

559

This work was supported by the State Key Research Development Program of China

560

(2016YFA0204200), the National Natural Science Foundation of China (21822603,

561

21773062, 21577036, 21377038, 5171101651), the Shanghai Education Development

562

Foundation and Shanghai Municipal Education Commission (16JC1401400,

563

17520711500), the Shanghai Pujiang Program (17PJD011), and the Fundamental

564

Research Funds for the Central Universities (22A201514021, 22221818014).

565

REFERENCES

ACS Paragon Plus Environment

Page 24 of 40

Page 25 of 40

566 567

Environmental Science & Technology

1. Ahmaruzzaman, M.; Gupta, V. K., Rice husk and its ash as low-cost adsorbents in water and wastewater treatment. Ind. Eng. Chem. Res. 2011, 50, (24), 13589-13613.

568

2. Li, J.; Luo, C.; Song, M.; Dai, Q.; Jiang, L.; Zhang, D.; Zhang, G., Biodegradation

569

of phenanthrene in polycyclic aromatic hydrocarbon-contaminated wastewater

570

revealed by coupling cultivation-dependent and -independent approaches. Environ. Sci.

571

Technol. 2017, 51, (6), 3391-3401.

572

3. Fan, J.; Grande, C. D.; Rodrigues, D. F., Biodegradation of graphene oxide-

573

polymer nanocomposite films in wastewater. Environ. Sci: Nano 2017, 4, (9), 1808-

574

1816.

575

4. Mir-Tutusaus, J. A.; Parladé, E.; Llorca, M.; Villagrasa, M.; Barceló, D.;

576

Rodriguez-Mozaz, S.; Martinez-Alonso, M.; Gaju, N.; Caminal, G.; Sarrà, M.,

577

Pharmaceuticals removal and microbial community assessment in a continuous fungal

578

treatment of non-sterile real hospital wastewater after a coagulation-flocculation

579

pretreatment. Water Res. 2017, 116, (Supplement C), 65-75.

580

5. Zhuang, J.; Xu, J.; Wang, X.; Li, Z.; Han, W.; Wang, Z., Improved microfiltration

581

of prehydrolysis liquor of wood from dissolving pulp mill by flocculation treatments

582

for hemicellulose recovery. Sep. Purif. Technol. 2017, 176, (Supplement C), 159-163.

583

6. Langsa, M.; Heitz, A.; Joll, C. A.; von Gunten, U.; Allard, S., Mechanistic Aspects

584

of the Formation of Adsorbable Organic Bromine during Chlorination of Bromide-

585

containing Synthetic Waters. Environ. Sci. Technol. 2017, 51, (9), 5146-5155.

ACS Paragon Plus Environment

Environmental Science & Technology

586

7. Du, Y.; Wu, Q.-Y.; Lu, Y.; Hu, H.-Y.; Yang, Y.; Liu, R.; Liu, F., Increase of

587

cytotoxicity during wastewater chlorination: Impact factors and surrogates. J. Hazard.

588

Mater. 2017, 324, (Part B), 681-690.

589

8. Dong, S.; Masalha, N.; Plewa, M. J.; Nguyen, T. H., Toxicity of wastewater with

590

elevated bromide and iodide after chlorination, chloramination, or ozonation

591

disinfection. Environ. Sci. Technol. 2017, 51, (16), 9297-9304.

592

9. Soltermann, F.; Abegglen, C.; Tschui, M.; Stahel, S.; von Gunten, U., Options and

593

limitations for bromate control during ozonation of wastewater. Water Res. 2017, 116,

594

(Supplement C), 76-85.

595

10.

Saleh, T. A.; Gupta, V. K., Synthesis and characterization of alumina nano-

596

particles polyamide membrane with enhanced flux rejection performance. Sep. Purif.

597

Technol. 2012, 89, 245-251.

598

11.

Gupta, V. K.; Nayak, A.; Agarwal, S.; Tyagi, I., Potential of activated carbon

599

from waste rubber tire for the adsorption of phenolics: Effect of pre-treatment

600

conditions. J. Colloid Interf. Sci. 2014, 417, 420-430.

601

12.

Mittal, A.; Mittal, J.; Malviya, A.; Gupta, V. K., Removal and recovery of

602

Chrysoidine Y from aqueous solutions by waste materials. J. Colloid Interf. Sci. 2010,

603

344, (2), 497-507.

ACS Paragon Plus Environment

Page 26 of 40

Page 27 of 40

604

Environmental Science & Technology

13.

Robati, D.; Mirza, B.; Rajabi, M.; Moradi, O.; Tyagi, I.; Agarwal, S.; Gupta,

605

V. K., Removal of hazardous dyes-BR 12 and methyl orange using graphene oxide as

606

an adsorbent from aqueous phase. Chem. Eng. J. 2016, 284, 687-697.

607

14.

Gupta, V. K.; Saleh, T. A., Sorption of pollutants by porous carbon, carbon

608

nanotubes and fullerene- An overview. Environ. Sci. Pollut. Res. 2013, 20, (5), 2828-

609

2843.

610

15.

Asfaram, A.; Ghaedi, M.; Agarwal, S.; Tyagi, I.; Kumar Gupta, V., Removal

611

of basic dye Auramine-O by ZnS:Cu nanoparticles loaded on activated carbon:

612

optimization of parameters using response surface methodology with central composite

613

design. RSC Adv. 2015, 5, (24), 18438-18450.

614

16.

Ghaedi, M.; Hajjati, S.; Mahmudi, Z.; Tyagi, I.; Agarwal, S.; Maity, A.; Gupta,

615

V. K., Modeling of competitive ultrasonic assisted removal of the dyes - Methylene

616

blue and Safranin-O using Fe3O4 nanoparticles. Chem. Eng. J. 2015, 268, 28-37.

617

17.

International occupational, s.; health information, c.; Gupta, V. K.; Nayak, A.;

618

Agarwal, S., Bioadsorbents for remediation of heavy metals: Current status and their

619

future prospects. Environ. Eng. Res. 2015, 20, (1), 1-18.

620

18.

Gupta, V. K.; Jain, R.; Nayak, A.; Agarwal, S.; Shrivastava, M., Removal of

621

the hazardous dye-Tartrazine by photodegradation on titanium dioxide surface. Mater.

622

Sci. Eng. C 2011, 31, (5), 1062-1067.

ACS Paragon Plus Environment

Environmental Science & Technology

623

19.

Gupta, V. K.; Atar, N.; Yola, M. L.; Üstündağ, Z.; Uzun, L., A novel magnetic

624

Fe@Au core–shell nanoparticles anchored graphene oxide recyclable nanocatalyst for

625

the reduction of nitrophenol compounds. Water Res. 2014, 48, 210-217.

626

20.

Behera, A.; Kandi, D.; Majhi, S. M.; Martha, S.; Parida, K., Facile synthesis

627

of ZnFe2O4 photocatalysts for decolourization of organic dyes under solar irradiation.

628

Beilstein J. Nanotech. 2018, 9, 436-446.

629

21.

Pradhan, A. C.; Parida, K. M.; Nanda, B., Enhanced photocatalytic and

630

adsorptive degradation of organic dyes by mesoporous Cu/Al2O3–MCM-41: intra-

631

particle mesoporosity, electron transfer and OH radical generation under visible light.

632

Dalton Trans. 2011, 40, (28), 7348-7356.

633

22.

Devaraj, M.; Saravanan, R.; Deivasigamani, R.; Gupta, V. K.; Gracia, F.;

634

Jayadevan, S., Fabrication of novel shape Cu and Cu/Cu2O nanoparticles modified

635

electrode for the determination of dopamine and paracetamol. J. Mol. Liq. 2016, 221,

636

930-941.

637

23.

Saravanan, R.; Thirumal, E.; Gupta, V. K.; Narayanan, V.; Stephen, A., The

638

photocatalytic activity of ZnO prepared by simple thermal decomposition method at

639

various temperatures. J. Mol. Liq. 2013, 177, 394-401.

640

24.

Saravanan, R.; Sacari, E.; Gracia, F.; Khan, M. M.; Mosquera, E.; Gupta, V.

641

K., Conducting PANI stimulated ZnO system for visible light photocatalytic

642

degradation of coloured dyes. J. Mol. Liq. 2016, 221, 1029-1033.

ACS Paragon Plus Environment

Page 28 of 40

Page 29 of 40

643

Environmental Science & Technology

25.

Saravanan, R.; Karthikeyan, N.; Gupta, V. K.; Thirumal, E.; Thangadurai, P.;

644

Narayanan, V.; Stephen, A., ZnO/Ag nanocomposite: An efficient catalyst for

645

degradation studies of textile effluents under visible light. Mater. Sci. Eng. C 2013, 33,

646

(4), 2235-2244.

647

26.

Saravanan, R.; Gupta, V. K.; Prakash, T.; Narayanan, V.; Stephen, A.,

648

Synthesis, characterization and photocatalytic activity of novel Hg doped ZnO

649

nanorods prepared by thermal decomposition method. J. Mol. Liq. 2013, 178, 88-93.

650

27.

Saravanan, R.; Karthikeyan, S.; Gupta, V. K.; Sekaran, G.; Narayanan, V.;

651

Stephen, A., Enhanced photocatalytic activity of ZnO/CuO nanocomposite for the

652

degradation of textile dye on visible light illumination. Mater. Sci. Eng. C 2013, 33,

653

(1), 91-98.

654

28.

Saravanan, R.; Khan, M. M.; Gupta, V. K.; Mosquera, E.; Gracia, F.;

655

Narayanan, V.; Stephen, A., ZnO/Ag/Mn2O3 nanocomposite for visible light-induced

656

industrial textile effluent degradation, uric acid and ascorbic acid sensing and

657

antimicrobial activity. RSC Adv. 2015, 5, (44), 34645-34651.

658

29.

Saravanan, R.; Mansoob Khan, M.; Gupta, V. K.; Mosquera, E.; Gracia, F.;

659

Narayanan, V.; Stephen, A., ZnO/Ag/CdO nanocomposite for visible light-induced

660

photocatalytic degradation of industrial textile effluents. J. Colloid Interf. Sci. 2015,

661

452, 126-133.

ACS Paragon Plus Environment

Environmental Science & Technology

662

30.

Rajendran, S.; Khan, M. M.; Gracia, F.; Qin, J.; Gupta, V. K.; Arumainathan,

663

S., Ce3+-ion-induced visible-light photocatalytic degradation and electrochemical

664

activity of ZnO/CeO2 nanocomposite. Sci. Rep. 2016, 6, 31641.

665

31.

Saravanan, R.; Joicy, S.; Gupta, V. K.; Narayanan, V.; Stephen, A., Visible

666

light induced degradation of methylene blue using CeO2/V2O5 and CeO2/CuO catalysts.

667

Mater. Sci. Eng. C 2013, 33, (8), 4725-4731.

668 669

670

32.

Qi, H.; Liu, H.; Gao, Y., Removal of Sr(II) from aqueous solutions using

polyacrylamide modified graphene oxide composites. J. Mol. Liq. 2015, 208, 394-401. 33.

Mohammadi, N.; Khani, H.; Gupta, V. K.; Amereh, E.; Agarwal, S.,

671

Adsorption process of methyl orange dye onto mesoporous carbon material–kinetic and

672

thermodynamic studies. J. Colloid Interf. Sci. 2011, 362, (2), 457-462.

673

34.

Saleh, T. A.; Gupta, V. K., Functionalization of tungsten oxide into MWCNT

674

and its application for sunlight-induced degradation of rhodamine B. J. Colloid Interf.

675

Sci. 2011, 362, (2), 337-344.

676

35.

Gupta, V. K.; Kumar, R.; Nayak, A.; Saleh, T. A.; Barakat, M. A., Adsorptive

677

removal of dyes from aqueous solution onto carbon nanotubes: A review. Adv. Colloid

678

Interf. Sci. 2013, 193-194, 24-34.

679 680

36.

Khani, H.; Rofouei, M. K.; Arab, P.; Gupta, V. K.; Vafaei, Z., Multi-walled

carbon nanotubes-ionic liquid-carbon paste electrode as a super selectivity sensor:

ACS Paragon Plus Environment

Page 30 of 40

Page 31 of 40

Environmental Science & Technology

681

Application to potentiometric monitoring of mercury ion(II). J. Hazard. Mater. 2010,

682

183, (1), 402-409.

683

37.

Saleh, T. A.; Gupta, V. K., Photo-catalyzed degradation of hazardous dye

684

methyl orange by use of a composite catalyst consisting of multi-walled carbon

685

nanotubes and titanium dioxide. J. Colloid Interf. Sci. 2012, 371, (1), 101-106.

686

38.

Mazioti, A. A.; Stasinakis, A. S.; Psoma, A. K.; Thomaidis, N. S.; Andersen,

687

H. R., Hybrid Moving Bed Biofilm Reactor for the biodegradation of benzotriazoles

688

and hydroxy-benzothiazole in wastewater. J. Hazard. Mater. 2017, 323, (Part A), 299-

689

310.

690

39.

Castronovo, S.; Wick, A.; Scheurer, M.; Nödler, K.; Schulz, M.; Ternes, T. A.,

691

Biodegradation of the artificial sweetener acesulfame in biological wastewater

692

treatment and sandfilters. Water Res. 2017, 110, (Supplement C), 342-353.

693

40.

Carucci, A.; Cappai, G.; Piredda, M., Biodegradability and toxicity of

694

pharmaceuticals in biological wastewater treatment plants. J. Environ. Sci. Heal. A

695

2006, 41, (9), 1831-1842.

696

41.

Alves, M. M.; Mota Vieira, J. A.; Álvares Pereira, R. M.; Pereira, M. A.; Mota,

697

M., Effects of lipids and oleic acid on biomass development in anaerobic fixed-bed

698

reactors. Part II: Oleic acid toxicity and biodegradability. Water Res. 2001, 35, (1), 264-

699

270.

ACS Paragon Plus Environment

Environmental Science & Technology

700

42.

Tyagi, V. K.; Lo, S.-L., Application of physico-chemical pretreatment

701

methods to enhance the sludge disintegration and subsequent anaerobic digestion: an

702

up to date review. Rev. Environ. Sci. Bio/Technol. 2011, 10, (3), 215.

703

43.

Ma, J.; Sun, Y.; Zhang, M.; Yang, M.; Gong, X.; Yu, F.; Zheng, J.,

704

Comparative study of graphene hydrogels and aerogels reveals the important role of

705

buried water in pollutant adsorption. Environ. Sci. Technol. 2017, 51, (21), 12283-

706

12292.

707

44.

Luo, W.; Phan, H. V.; Xie, M.; Hai, F. I.; Price, W. E.; Elimelech, M.; Nghiem,

708

L. D., Osmotic versus conventional membrane bioreactors integrated with reverse

709

osmosis for water reuse: Biological stability, membrane fouling, and contaminant

710

removal. Water Res. 2017, 109, (Supplement C), 122-134.

711

45.

Mu, Y.; Ai, Z.; Zhang, L., Phosphate shifted oxygen reduction pathway on

712

Fe@Fe2O3 core-shell nanowires for enhanced reactive oxygen species generation and

713

aerobic 4-chlorophenol degradation. Environ. Sci. Technol. 2017, 51, (14), 8101-8109.

714

46.

Clarizia, L.; Russo, D.; Di Somma, I.; Marotta, R.; Andreozzi, R.,

715

Homogeneous photo-Fenton processes at near neutral pH: A review. Appl. Catal. B

716

2017, 209, 358-371.

717

47.

Karthikeyan, S.; Gupta, V. K.; Boopathy, R.; Titus, A.; Sekaran, G., A new

718

approach for the degradation of high concentration of aromatic amine by heterocatalytic

719

Fenton oxidation: Kinetic and spectroscopic studies. J. Mol. Liq. 2012, 173, 153-163.

ACS Paragon Plus Environment

Page 32 of 40

Page 33 of 40

720

Environmental Science & Technology

48.

Pham, A. L.-T.; Doyle, F. M.; Sedlak, D. L., Kinetics and efficiency of H2O2

721

activation by iron-containing minerals and aquifer materials. Water Res. 2012, 46, (19),

722

6454-6462.

723

49.

Qin, Y.; Song, F.; Ai, Z.; Zhang, P.; Zhang, L., Protocatechuic acid promoted

724

alachlor degradation in Fe(III)/H2O2 Fenton system. Environ. Sci. Technol. 2015, 49,

725

(13), 7948-7956.

726

50.

Ma, J.; Ma, W.; Song, W.; Chen, C.; Tang, Y.; Zhao, J.; Huang, Y.; Xu, Y.;

727

Zang, L., Fenton degradation of organic pollutants in the presence of low-molecular-

728

weight organic acids:  Cooperative effect of quinone and visible Light. Environ. Sci.

729

Technol. 2006, 40, (2), 618-624.

730

51.

Gazi, S.; Rajakumar, A.; Singh, N. D. P., Photodegradation of organic dyes in

731

the presence of [Fe(III)-salen]Cl complex and H2O2 under visible light irradiation. J.

732

Hazard. Mater. 2010, 183, (1), 894-901.

733

52.

Li, T.; Zhao, Z.; Wang, Q.; Xie, P.; Ma, J., Strongly enhanced Fenton

734

degradation of organic pollutants by cysteine: An aliphatic amino acid accelerator

735

outweighs hydroquinone analogues. Water Res. 2016, 105, 479-486.

736

53.

Jiang, C.; Garg, S.; Waite, T. D., Hydroquinone-mediated redox cycling of

737

iron and concomitant oxidation of hydroquinone in oxic waters under acidic conditions:

738

Comparison with iron-natural organic matter interactions. Environ. Sci. Technol. 2015,

739

49, (24), 14076-14084.

ACS Paragon Plus Environment

Environmental Science & Technology

740

54.

Liu, F.; Yu, J.; Tu, G.; Qu, L.; Xiao, J.; Liu, Y.; Wang, L.; Lei, J.; Zhang, J.,

741

Carbon nitride coupled Ti-SBA-15 catalyst for visible-light-driven photocatalytic

742

reduction of Cr(VI) and the synergistic oxidation of phenol. Appl. Catal. B 2017, 201,

743

(Supplement C), 1-11.

744

55.

Han, C.; Jing, M.; Shen, X.; Qiao, G., Electrospinning fabrication of

745

mesoporous nano Fe2O3-TiO2@activated carbon fiber membrane for hybrid removal of

746

phenol from waste water. Russ. J. Appl. Chem. 2016, 89, (12), 2008-2015.

747

56.

Pradhan, A. C.; Nanda, B.; Parida, K. M.; Das, M., Quick photo-Fenton

748

degradation of phenolic compounds by Cu/Al2O3-MCM-41 under visible light

749

irradiation: small particle size, stabilization of copper, easy reducibility of Cu and

750

visible light active material. Dalton Trans. 2013, 42, (2), 558-566.

751

57.

Parida, K. M.; Pradhan, A. C., Fe/meso-Al2O3: An efficient Photo-Fenton

752

catalyst for the adsorptive degradation of phenol. Ind. Eng. Chem. Res. 2010, 49, (18),

753

8310-8318.

754

58.

Wu, Z.-L.; Ondruschka, B.; Cravotto, G., Degradation of phenol under

755

combined irradiation of microwaves and ultrasound. Environ. Sci. Technol. 2008, 42,

756

(21), 8083-8087.

757

59.

Kavitha, V.; Palanivelu, K., Degradation of phenol and trichlorophenol by

758

heterogeneous photo-Fenton process using Granular Ferric Hydroxide®: comparison

759

with homogeneous system. Int. J. Environ. Sci. Technol. 2016, 13, (3), 927-936.

ACS Paragon Plus Environment

Page 34 of 40

Page 35 of 40

760

Environmental Science & Technology

60.

Wang, J.; Yao, Z.; Wang, Y.; Xia, Q.; Chu, H.; Jiang, Z., Preparation of

761

immobilized coating Fenton-like catalyst for high efficient degradation of phenol.

762

Environ. Pollut. 2017, 224, (Supplement C), 552-558.

763

61.

Gao, H.-Y.; Mao, L.; Li, F.; Xie, L.-N.; Huang, C.-H.; Shao, J.; Shao, B.;

764

Kalyanaraman, B.; Zhu, B.-Z., Mechanism of intrinsic chemiluminescence production

765

from the degradation of persistent chlorinated phenols by the Fenton system: A

766

structure-activity relationship study and the critical role of quinoid and semiquinone

767

radical intermediates. Environ. Sci. Technol. 2017, 51, (5), 2934-2943.

768

62.

Guo, S.; Zhu, Y.; Yan, Y.; Min, Y.; Fan, J.; Xu, Q., Holey structured graphitic

769

carbon nitride thin sheets with edge oxygen doping via photo-Fenton reaction with

770

enhanced photocatalytic activity. Appl. Catal. B 2016, 185, 315-321.

771

63.

Li, M.; Qiang, Z.; Pulgarin, C.; Kiwi, J., Accelerated methylene blue (MB)

772

degradation by Fenton reagent exposed to UV or VUV/UV light in an innovative micro

773

photo-reactor. Appl. Catal. B 2016, 187, (Supplement C), 83-89.

774

64.

Hems, R. F.; Hsieh, J. S.; Slodki, M. A.; Zhou, S.; Abbatt, J. P. D., Suppression

775

of OH generation from the Photo-Fenton reaction in the presence of α-pinene secondary

776

organic aerosol material. Environ. Sci. Technol. Lett. 2017. 4, (10), 439–443.

777 778

65.

Walling, C., Fenton's reagent revisited. Accounts Chem. Res. 1975, 8, (4), 125-

131.

ACS Paragon Plus Environment

Environmental Science & Technology

779 780

781

66.

Tang, W. Z.; Huang, C. P., 2,4-Dichlorophenol oxidation kinetics by Fenton's

reagent. Environ. Technol. 1996, 17, (12), 1371-1378. 67.

Feng, G.; Cheng, P.; Yan, W.; Boronat, M.; Li, X.; Su, J.-H.; Wang, J.; Li, Y.;

782

Corma, A.; Xu, R.; Yu, J., Accelerated crystallization of zeolites via hydroxyl free

783

radicals. Science 2016, 351, (6278), 1188-1191.

784 785

786

68.

Wu, K.; Xie, Y.; Zhao, J.; Hidaka, H., Photo-Fenton degradation of a dye under

visible light irradiation. J. Mol. Catal. A 1999, 144, (1), 77-84. 69.

Xing, M.; Zhang, J.; Qiu, B.; Tian, B.; Anpo, M.; Che, M., A brown

787

mesoporous TiO2-x/MCF composite with an extremely high quantum yield of solar

788

energy photocatalysis for H2 evolution. Small 2015, 11, 1920–1929.

789

70.

Wang, Y.; Sun, H.; Ang, H. M.; Tadé, M. O.; Wang, S., 3D-hierarchically

790

structured MnO2 for catalytic oxidation of phenol solutions by activation of

791

peroxymonosulfate: Structure dependence and mechanism. Appl. Catal. B 2015, 164,

792

159-167.

793

71.

Xu, L.; Wang, J., A heterogeneous Fenton-like system with nanoparticulate

794

zero-valent iron for removal of 4-chloro-3-methyl phenol. J. Hazard. Mater. 2011, 186,

795

(1), 256-264.

796

72.

Lin, S. H.; Lin, c. M.; Leu, H. G., Operating characteristics and kinetic studies

797

of surfactant wastewater treatment by Fenton oxidation. Water Res. 1999, 33, (7), 1735-

798

1741.

ACS Paragon Plus Environment

Page 36 of 40

Page 37 of 40

799

Environmental Science & Technology

73.

Karthikeyan, S.; Gupta, V. K.; Boopathy, R.; Titus, A.; Sekaran, G., A new

800

approach for the degradation of high concentration of aromatic amine by heterocatalytic

801

Fenton oxidation: Kinetic and spectroscopic studies. J. Mol. Liq. 2012, 173,

802

(Supplement C), 153-163.

803

74.

Tamura, H.; Goto, K.; Yotsuyanagi, T.; Nagayama, M., Spectrophotometric

804

determination of iron(II) with 1,10-phenanthroline in the presence of large amounts of

805

iron(III). Talanta 1974, 21, (4), 314-318.

806

75.

Ding, Q.; Song, B.; Xu, P.; Jin, S., Efficient electrocatalytic and

807

photoelectrochemical hydrogen generation using MoS2 and related compounds. Chem

808

2016, 1, (5), 699-726.

809 810

811

76.

Chianelli, R. R.; Berhault, G.; Torres, B., Unsupported transition metal sulfide

catalysts: 100 years of science and application. Catal. Today 2009, 147, (3), 275-286. 77.

Voiry, D.; Fullon, R.; Yang, J.; de Carvalho Castro e Silva, C.; Kappera, R.;

812

Bozkurt, I.; Kaplan, D.; Lagos, M. J.; Batson, P. E.; Gupta, G.; Mohite, A. D.; Dong,

813

L.; Er, D.; Shenoy, V. B.; Asefa, T.; Chhowalla, M., The role of electronic coupling

814

between substrate and 2D MoS2 nanosheets in electrocatalytic production of hydrogen.

815

Nat. Mater. 2016, 15, 1003-1009.

816 817

78.

Jarosz, A. P.; Yep, T.; Mutus, B., Microplate-based colorimetric detection of

free hydrogen sulfide. Anal. Chem. 2013, 85, (7), 3638-3643.

ACS Paragon Plus Environment

Environmental Science & Technology

818

79.

Yu, S.; Kim, J.; Yoon, K. R.; Jung, J.-W.; Oh, J.; Kim, I.-D., Rational design

819

of efficient electrocatalysts for hydrogen evolution reaction: Single layers of WS2

820

nanoplates anchored to hollow nitrogen-doped carbon nanofibers. ACS Appl. Mater.

821

Interf. 2015, 7, (51), 28116-28121.

822

80.

Dai, Y.; Wu, X.; Sha, D.; Chen, M.; Zou, H.; Ren, J.; Wang, J.; Yan, X., Facile

823

self-assembly of Fe3O4 nanoparticles@WS2 nanosheets: A promising candidate for

824

supercapacitor electrode. Electron. Mater. Lett. 2016, 12, (6), 789-794.

825

81.

Bokare, A. D.; Choi, W., Review of iron-free Fenton-like systems for

826

activating H2O2 in advanced oxidation processes. J. Hazard. Mater. 2014, 275, 121-

827

135.

828

82.

Barrera-Díaz, C. E.; Lugo-Lugo, V.; Bilyeu, B., A review of chemical,

829

electrochemical and biological methods for aqueous Cr(VI) reduction. J. Hazard.

830

Mater. 2012, 223-224, (Supplement C), 1-12.

831 832

833

83.

Keith, L.; Telliard, W., ES&T Special Report: Priority pollutants: I-a

perspective view. Environ. Sci. Technol. 1979, 13, (4), 416-423. 84.

Patnaik, S.; Das, K. K.; Mohanty, A.; Parida, K., Enhanced photo catalytic

834

reduction of Cr(VI) over polymer-sensitized g-C3N4/ZnFe2O4 and its synergism with

835

phenol oxidation under visible light irradiation. Catal. Today 2018, 315, 52-56.

ACS Paragon Plus Environment

Page 38 of 40

Page 39 of 40

836

Environmental Science & Technology

85.

Bokare, A. D.; Choi, W., Chromate-induced activation of hydrogen peroxide

837

for oxidative degradation of aqueous organic pollutants. Environ. Sci. Technol. 2010,

838

44, (19), 7232-7237.

839 840

841

86.

Bokare, A. D.; Choi, W., Advanced oxidation process based on the

Cr(III)/Cr(VI) redox cycle. Environ. Sci. Technol. 2011, 45, (21), 9332-9338. 87.

Dehghani, M. H.; Heibati, B.; Asadi, A.; Tyagi, I.; Agarwal, S.; Gupta, V. K.,

842

Reduction of noxious Cr(VI) ion to Cr(III) ion in aqueous solutions using H2O2 and

843

UV/H2O2 systems. J. Ind. Eng. Chem. 2016, 33, 197-200.

844

88.

Cao, S.; Liu, T.; Hussain, S.; Zeng, W.; Peng, X.; Pan, F., Hydrothermal

845

synthesis, characterization and optical absorption property of nanoscale WS2/TiO2

846

composites. Physica E: Low-dimensional Systems and Nanostructures 2015, 68, 171-

847

175.

848

89.

Xu, B.-H.; Lin, B.-Z.; Chen, Z.-J.; Li, X.-L.; Wang, Q.-Q., Preparation and

849

electrical conductivity of polypyrrole/WS2 layered nanocomposites. J. Colloid Interf.

850

Sci. 2009, 330, (1), 220-226.

851

90.

Ishimaru, Y.; Suzuki, M.; Tsukamoto, T.; Suzuki, K.; Nakazono, M.;

852

Kobayashi, T.; Wada, Y.; Watanabe, S.; Matsuhashi, S.; Takahashi, M.; Nakanishi, H.;

853

Mori, S.; Nishizawa, N. K., Rice plants take up iron as an Fe3+-phytosiderophore and

854

as Fe2+. Plant J. 2006, 45, (3), 335-346.

855

ACS Paragon Plus Environment

Environmental Science & Technology

856

SYNOPSIS (Word Style “SN_Synopsis_TOC”).

857

ACS Paragon Plus Environment

Page 40 of 40