Multiphase Porous Electrochemical Catalysts Derived from Iron-Based

6 days ago - ... mineralization of herbicides in aqueous solution at circum-neutral pH. ... sufficient contact of electrolyte and substrate with the a...
0 downloads 0 Views 818KB Size
Subscriber access provided by University of Victoria Libraries

Remediation and Control Technologies

Multi-Phase Porous Electrochemical Catalysts Derived from Iron-Based Metal-Organic Framework Compounds Kai Liu, Menglin Yu, Haiying Wang, Juan Wang, Weiping Liu, and Michael R Hoffmann Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.9b01143 • Publication Date (Web): 10 May 2019 Downloaded from http://pubs.acs.org on May 12, 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 30

Environmental Science & Technology

1

Multi-Phase Porous Electrochemical Catalysts Derived from Iron-Based

2

Metal-Organic Framework Compounds

3 4

Kai Liu,†, ‡ Menglin Yu,† Haiying Wang,† Juan Wang,† Weiping Liu,† Michael R. Hoffmann‡,*

5

† College

6 7 8 9 10 11

of Environmental and Resource Science

Zhejiang University, Hangzhou 310058, China ‡

Department of Environmental Science and Engineering

California Institute of Technology, Pasadena, California 91126, United States # Author Contributions K. Liu and M. Yu contributed equally to the work. *Corresponding

Author: Michael R. Hoffmann

12

Tel: +1-626-395-4391. Fax: +1-626-395-4391. E-mail: [email protected]

13

Table of Contents Art

E-Fenton

14

O2 Reduction

15 16 1

ACS Paragon Plus Environment

Environmental Science & Technology

17

Abstract

18

Herbicide use has attracted attention recently due to potential damage to human health and

19

lethality to the honey bees and other pollinators. Fenton reagent treatment processes can be applied

20

for the degradation of herbicidal contaminants from water. However, the need to carry out the

21

normal Fenton reactions under acidic conditions often hinders their practical application for

22

pollution control. Herein, we report on the synthesis and application of multi-phasic porous

23

electro-Fenton catalysts prepared from calcinated metal-organic framework compounds,

24

CMOF@PCM, and their application for the mineralization of herbicides in aqueous solution at

25

circum-neutral pH. CMOF nanoparticles (NPs) are anchored on porous carbon monolithic (PCM)

26

substrates, which allow for binder-free application. H2O2 is electrochemically generated on the

27

PCM substrate which serves as a cathode, while ·OH is generated by the CMOF NPs at low applied

28

potentials (-0.14 V). Results show that the structure and reactivity of the CMOF@PCM electro-

29

Fenton catalysts is dependent on the specific MOF precursor used during synthesis. For example,

30

CMIL-88-NH2, which is prepared from MIL-88(Fe)-NH2, is a porous core-shell structured NP

31

comprised of a cementite (Fe3C) intermediate layer that is sandwiched between a graphitic shell

32

and a magnetite (Fe3O4) core. The electro-Fenton production of hydroxyl radical on the

33

CMOF@PCM composite material is shown to effectively degrades an array of herbicides.

2

ACS Paragon Plus Environment

Page 2 of 30

Page 3 of 30

35

Environmental Science & Technology

Introduction

36

Due to extensive application of herbicides, they are now considered to be pseudo-persistent

37

pollutants.1 Among the widely used herbicides, glyphosate is considered to be carcinogenic2, 3.

38

Furthermore, several herbicides are reported to impact pollinators resulting in a dramatic reduction

39

of bee populations worldwide.4-7 Herbicides are also suspected of causing harm to crustaceans,

40

which are an important component of freshwater and oceanic food webs.8 In spite of these growing

41

concerns, there is little research on their removal from the aquatic environment compared to other

42

contaminants of emerging concern (CECs).

43

Advanced oxidation processes (AOP) including electrochemical oxidation are suitable for

44

herbicide degradation and removal from contaminated water.9-11 However, electrochemical

45

degradation of organic compounds is often energy intensive with high capital expenditure

46

(CAPEX) costs. This is due to use of expensive material such as platinum group metals (PGM) or

47

boron-doped diamonds as anodes or in the case of platinum as cathodes. In contrast, alternative

48

oxidative approaches employing Fenton-reagent reactions (H2O2 + Fe(II) → ∙OH + Fe(III) + OH-

49

) utilize iron, an earth abundant metal.12, 13 Fenton-like systems can achieve complete degradation

50

of some of the most persistent contaminants such as perfluorinated surfactants and

51

perfluorotelomer alcohols.14, 15 However, a major drawback of using the Fenton process is that

52

above pH 5, Fe(II) is readily oxidized to Fe(III) in the form of Fe(OH)3 by dissolved oxygen, thus

53

limiting the pH range to carry out the overall reactions. Electro-Fenton reaction, on the other hand,

54

overcomes this problem by reducing Fe(III) to Fe(II) at cathode.16,

55

concentrations as high as 0.01 M need to be applied for the effective removal of contaminants.15

17

Unfortunately, Fe(II)

56

We have shown previously that a composite Fe-NP and nitrogen co-doped graphite

57

material can be used as an electrode for heterogeneous electro-Fenton oxidation.18 This approach

3

ACS Paragon Plus Environment

Environmental Science & Technology

58

only required a loading level of 0.36 wt% Fe in order to maximize ·OH production. However, the

59

percentage of Fe that was accessible on the surface was limited.18 In order to explore the full

60

potential of a heterogeneous electro-Fenton process, a porous electrode where all potential Fe sites

61

are accessible is desired. Given this objective, inorganic porous nanomaterials derived from MOFs,

62

or carbonized MOFs (CMOFs), appear to provide a solution that would achieve this. For example,

63

metal NPs porous catalyst derived from MOFs have been shown to be useful for organic

64

synthesis.19

65

In order to be useful in practical applications, the electro-Fenton catalysts must also meet

66

the following criteria: 1) the metal nano-catalyst supported on a substrate must be well dispersed

67

in order to allow for sufficient contact of electrolyte and substrate with the active catalyst; 2) the

68

metal catalyst must form a transient bond with the substrate; and 3) both the substrate and

69

nanoparticulate metal catalysts must be attached to a highly porous material. The first criterion

70

allows the utilization of the electrochemically-active nano-sized catalyst. The third criterion allows

71

for the construction of 3D electrodes that have form-factor advantages, which can maximize the

72

electrode surface to electrolyte volume ratio during application.

73

Herein, we use binder-free multi-phase porous electro-Fenton catalysts made from

74

pyrolyzed MOFs for the electrochemical degradation of an array of herbicides. The resulting

75

catalysts are anchored onto porous carbon monoliths (PCM), which are denoted as CMOF@PCMs.

76

In contrast to a bulk-phase catalyst, the narrow distribution of the pores within the CMOF@PCMs

77

allows for facile access to the active Fe(II)/Fe(III) catalytic centers. In addition, the PCM substrate

78

allows in-situ generation of hydrogen peroxide (H2O2) that is needed for the electro-Fenton

79

reaction. The CMOF@PCMs are shown to be efficient at removing herbicides at neutral pH while

80

Fe leaching is negligible. Furthermore, we discovered the resulting structure of CMOFs are

4

ACS Paragon Plus Environment

Page 4 of 30

Page 5 of 30

Environmental Science & Technology

81

independent of the topology of parent MOF. But the resulting structure appears to depend on the

82

specific functional group and on the identity of the organic linker. A core-shell structured

83

CMOF@PCM is obtained from a MIL-88(Fe)-NH2 precursor. The results provided herein provide

84

insights into the development of stable and porous materials for use as sustainable electro-Fenton

85

catalysts.

86

Materials and Methods

87

Materials

88

All chemicals were obtained from Sinopharm Chemical Reagent Co. Ltd. and were used

89

without further purification. Solvents were HPLC grade unless otherwise stated. Pesticides with

90

chemical purity ≥ 97.0% (Sinopharm Chemical Reagent Co. Ltd.) used in this study are listed in

91

Table S1.

92

Preparation of Porous Carbon Substrate

93

The PCM substrate was synthesized based on a previously reported procedure.20 Briefly,

94

resorcinol (3.0 g, 27.3 mmol) and Pluronic F127 (1.25 g) were dissolved in a solvent mixture of

95

ethanol (11.4 mL) and deionized water (9 mL). 1,6-diaminohexane (0.078 g, 0.67 mmol) was

96

added to the above solution and stirred for 30 min. Subsequently, formalin (4.42 g, 54.5 mmol)

97

was added. The reaction mixture was stirred for 40 min before sealed and heated at 90°C for 4 h.

98

The polymer monolith, poly(benzoxazine-co-resol), was dried at 90°C for 48 h. Pyrolysis was

99

performed at 800°C for 2 h under a nitrogen protection. Crack-free PCM was obtained after

100

cooling down to room temperature.

101

Preparation of MOFs

102

A microwave-assisted synthesis of Fe-MOFs was based on a similar method using

103

microwave synthesizer (CEM, Discover SP).21 In brief, FeCl3 (0.88 g, 0.54 mmol) and terephthalic

5

ACS Paragon Plus Environment

Environmental Science & Technology

104

acid (0.9 g, 0.54 mmol) were dissolved in N, N-dimethylformamide (DMF) (15 mL) inside a 30

105

mL microwave tube. The reaction was carried at 150°C for predetermined time before cooled to

106

room temperature. The nano-crystals of MIL-88(Fe) were separated by centrifugation at 11,000

107

rpm, washed with DMF and methanol before drying in vacuo at 150°C. MIL-100(Fe), MIL-

108

101(Fe), MIL-88(Fe)-NH2, and MIL-101(Fe)-NH2 were synthesized using the same method.

109

Preparation of CMOF@PCM

110

PCM was mixed with the Fe-MOF powder at a given mass ratio. The powdered MIL-

111

88(Fe)@PCM was carbonized at 600 ºC for 0.5 h under a N2 atmosphere to produce CMIL-

112

88@PCM. The same procedure was applied for the preparation of the other CMOF@PCMs using

113

the other MOFs precursors. The CMOF@PCMs used in the current study were designated as

114

CMIL-88@PCM, CMIL-100@PCM, CMIL-101@PCM, CMIL-88-NH2@PCM, and CMIL-101-

115

NH2@PCM for the hybrid materials prepared from MIL-88(Fe), MIL-100(Fe), MIL-101(Fe),

116

MIL-88(Fe)-NH2, and MIL-101(Fe)-NH2, respectively. Fe-MOF powder was replaced with Fe3O4

117

nanoparticle to produce Fe3O4@PCM using similar method.

118

Preparation of Electrodes

119

Evaluation of electro-catalytic property of CMOF@PCMs required identical geometrical

120

surface areas for the working electrode. As the result, rigid carbon paper was coated with these

121

electro-catalysts and used as the working electrode in order to ensure identical electro-active areas.

122

The ground CMOF@PCM substrates (50 mg) were suspended in 4.47 ml of mili-Q water, 1.5 ml

123

of isopropanol, and 100 μl of a Nafion solution. The resulting ink was spray-coated onto carbon

124

paper (HCP030P Hesen) at a catalyst loading of 0.6 mg/cm2. Both sides of carbon paper were

125

uniformly coated (with total effective area of 8 cm2) to avoid exposing the supporting carbon paper

126

substrate to the electrolyte.

6

ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30

127

Environmental Science & Technology

Characterization of Catalysts

128

The surface morphology of the PCM substrate was examined using a Hitachi SU8010 field-

129

emission scanning electron microscope (FE SEM), while the surface morphologies and elemental

130

compositions of the CMOF@PCMs were examined using a ZEISS Merlin FE SEM equipped with

131

an Oxford INCA X-MAX50 energy dispersive spectrometer (EDS). TEM micrographs and TEM

132

EDS measurements were obtained using a FEI Tecnai G2 F20 (200 kV) transmission electron

133

microscope (TEM). X-ray powder diffraction patterns were collected using a Bruker D8 A25

134

Advance diffractometer with Cu-Kα radiation (λ=1.5418 Å). The element composition of catalysts

135

was characterized by X-ray photoelectron spectroscopy (XPS) using Surface Science Instruments

136

Thermo Scientific ESCALAB 250Xi surface spectrometer. Monochromatic Al-Kα radiation

137

(1486.6 eV) was used.

138

Electrochemical Experiments

139

Electrochemical experiments were performed using a CH CHI660E C17171 potentiostat

140

with a three-electrode beaker cell at room temperature. A platinum (Pt) wire and saturated calomel

141

electrode (SCE) were used as the counter electrode and the reference electrode, respectively (Fig.

142

S1). Recorded potentials were converted to the reversible hydrogen electrode (RHE) scale (RHE

143

= SCE + 0.059 × pH + 0.241 V). Cyclic voltammetry (CV) was conducted in Ar- or O2-saturated

144

electrolyte (0.1 M Na2SO4) at a scan rate of 10 mV/s. Constant potentials were applied to determine

145

H2O2 yields obtained with the PCM substrate and electro-Fenton efficiencies of herbicide

146

degradation by CMOF@PCMs. 0.1 M Na2SO4 was used as electrolyte, and solution pH was

147

adjusted to 4-10 using H2SO4 and NaOH. H2O2 yield measurements were performed in triplicate.

148

Electro-Fenton degradation experiments were performed in duplicate with the average value being

149

reported.

7

ACS Paragon Plus Environment

Environmental Science & Technology

150

Analytical Methods

151

To quantify the H2O2 produced, samples collected at certain time intervals were analyzed

152

by titanium oxalate spectrophotometric method.22 Briefly, reaction sample was added into 3 M

153

sulfuric acid solution before reacted with potassium titanium oxalate solution (0.5 wt% in water)

154

to produce a yellow pertitanic acid (TiO(H2O2)2+) complex. The absorbance of the solution was

155

measured using a JASCO V-750 spectrophotometer at 400 nm. H2O2 solutions with known

156

concentration were used to construct a calibration curve.

157

During the electro-Fenton process, the concentrations of herbicides were analyzed using a

158

high-performance liquid chromatography (Waters Acquity UPLC) system coupled to e2998

159

photodiode array detector and an e2695 separation module. Sunfire C18 column (4.6x50 mm; 5

160

µm particles) was used. (A) 0.1% Phosphoric acid and (B) methanol were used as mobile phase

161

with flow rate of 0.8 mL min-1.

162

RESULTS AND DISCUSSION

163

Synthesis of E-Fenton Catalysts from Fe-MOFs

164

The iron-containing MOFs MIL-88, MIL-100, and MIL-101 were used as templates in the

165

preparation of electro-Fenton catalysts. MIL-88(Fe) and MIL-101(Fe) were synthesized by using

166

the same organic ligand, terephthalic acid (BDC), with FeCl3. MIL-88(Fe) consists of an infinite

167

framework with three-dimensional channels formed by [Fe3(μ3-O)(COO)6(H2O)2Cl]6 secondary

168

building units (SBUs) and BDC linker (Fig. 1). MIL-101(Fe), however, consists of quasi-

169

spherical cages formed by Fe3O-carboxylate trimers SBU and BDC linkers (Fig. 1). The

170

mesoporous cages with dimensions of 2.9 by 3.4 nm are accessible via microporous windows or

171

openings (1.2 and 1.6 nm). Comparison between MIL-88(Fe) and MIL-101(Fe) allows us to study

8

ACS Paragon Plus Environment

Page 8 of 30

Page 9 of 30

Environmental Science & Technology

172

the impact of MOFs topology on the subsequent electrocatalytic properties of the CMOFs. In

173

addition, MIL-100(Fe) was synthesized by replacing the BDC ligand with 1,3,5-

174

benzenetricarboxylic acid (BTC) in order to investigate the effect of the ligand on the

175

electrocatalytic properties of the CMOFs. MIL-101(Fe) and MIL-100(Fe) have mesoporous cages

176

formed by the combination of iron octahedra trimers of [(Fe3(m3-O))(OH)(H2O)2]6 SBUs and BTC

177

linkers (Fig. 1). The MIL-100(Fe) cages (2.5 and 2.9 nm) have microporous apertures of 0.55 and

178

0.86 nm, which are significantly smaller than that of MIL-101(Fe)’s. In order to investigate the

179

functional group effect, MIL-88(Fe)-NH2 and MIL-101(Fe)-NH2 were synthesized by using an

180

aminated ligand, BDC-NH2. Since the MOF crystal size is directly proportional to the overall

181

reaction time, the microwave reaction conditions needed to be optimized in order to produce nano-

182

scaled MOF crystals. Thus, MIL-88(Fe) was synthesized as a function of microwave reaction time.

183

The SEM images of the various MIL-88(Fe) products are summarized in Figure S2. Initially, the

184

size of the MIL-88(Fe) crystals increased as the microwave irradiation time was lengthened. After

185

15 mins, the average diameter of the MIL-88(Fe) crystals were stabilized at 500 nm which is most

186

likely due to the depletion of starting materials. Since irradiation for 40 min produced the highest

187

yield, we then used a 40 min microwave irradiation time for the synthesis of the other MOFs. The

188

PXRD patterns of synthesized MOFs were consistent with the theoretically predicted PXRD

189

patterns (Fig. S3); thus, the identities of MOFs used in this study were confirmed.

190

The electro-Fenton catalysts were prepared by direct pyrolysis of the MOFs at 873 K for

191

30 mins under a N2 atmosphere. The synthesized catalysts are denoted as CMIL-88, CMIL-100,

192

CMIL-101, CMIL-88-NH2, and CMIL-101-NH2, for MIL-88(Fe), MIL-100(Fe), MIL-101(Fe),

193

MIL-88(Fe)-NH2, and MIL-101(Fe)-NH2 respectively. SEM micrographs show that CMOFs

194

retained crystal morphologies similar to that of parent MOFs (Fig. S4). PXRD pattern showed that

9

ACS Paragon Plus Environment

Environmental Science & Technology

195

the main iron containing species in the CMOFs is Fe3O4 (Fig. 2), except for CMIL-88-NH2 for

196

which cementite (Fe3C) was also identified as an iron component. The compositions of the CMOFs

197

were confirmed by energy dispersive X-ray spectroscopy (EDS). The results confirmed the

198

relatively even distribution of Fe within the CMOFs (Fig. S5).

199

The MOFs precursor and CMOFs were also analyzed by the FTIR; the results are shown

200

in Figures S6 and S7. For MIL-88(Fe)-NH2, peaks at 3414 and 3370 are assigned to the symmetric

201

and asymmetric stretching of the primary amine, while peaks at 1620 and 1392 cm−1 are assigned

202

to the asymmetric and symmetric vibrations of the carboxyl groups, respectively. The absence of

203

peak at 1700 cm-1 indicates that no uncoordinated BDC is present in MIL-88(Fe). In addition,

204

doublet at 572, 487 cm-1 is most likely due to the Fe-O vibration in MIL-88(Fe). For CMIL-88, a broad

205

peak at 3397 cm-1 is assigned to the O-H stretching vibrations of C-OH group. A newly formed

206

peak at 1644 cm-1 is assigned to the vibration of C=C bonds in the graphitic carbon skeleton. A

207

strong peak at 522 cm-1 is associated with the stretching vibration of Fe-O from the Fe3O4-NPs. These

208

peaks are assigned based on previously reported assignments.23, 24 MIL-101(Fe)-NH2 and CMIL-

209

101-NH2 gave similar FTIR spectra (Fig. S7). This result implies that the formation of graphitic carbon

210

and Fe3O4 during MOFs pyrolysis is independent of the specific MOF topology.

211

To prepare the PCM supported bi-functional electrochemical catalysts (CMOF@PCM),

212

resorcinol-formaldehyde (RF) polymers were synthesized as a carbon precursor. The RF polymer

213

was pyrolyzed at 1,073 K for 2 h under N2 atmosphere to produce PCM. After formation, the

214

product PCM was impregnated with MOFs before carbonization at 873 K for 30 mins under N2.

215

The MOFs loading amount was adjusted by changing the amount of the specific MOF mixed with

216

PCM during impregnation. The MOFs loading of MOF@PCM catalyst precursor was fixed at 25

217

wt%. Pyrolysis of MOF@PCM at 873 K for 30 mins under a N2 atmosphere produced mesoporous

10

ACS Paragon Plus Environment

Page 10 of 30

Page 11 of 30

Environmental Science & Technology

218

material designated as CMOF@PCM. SEM analysis indicates that the PCM substrate has an open

219

porous network that is fully accessible to both electrolytes, substrates, and dissolved gas molecules

220

(Fig. 3). In addition, SEM micrograph shows that the CMOF s were deposited inside the

221

macropores of PCM material (Fig. S8). Based on these analyses, the loaded CMOFs have mean

222

particle sizes ranging between 150 to 200 nm, 100 to 200 nm, 100 to 150 nm for CMIL-88@PCM,

223

CMIL-100@PCM, and CMIL-101@PCM, respectively.

224

Electro-generation of H2O2 by Porous Carbon Substrates

225

The electrochemical activities of the PCM support for the oxygen reduction reaction (ORR)

226

were investigated. Cyclic voltammetry (CV) was used to examine the oxygen reduction potential.

227

As shown in Figure S9, the working electrode prepared from the PCM substrate was featureless

228

for an Ar-saturated electrolyte solution, but showed a pronounced reduction peak in an O2-

229

saturated electrolyte solution. The current density measured at the ORR peak was 2.3 mA/cm2. A

230

similar value was found at pH 4, 7, and 10. The onset potentials at pH 4, 7, and 10 were 0.49,

231

0.75, and 0.94 V vs. RHE, respectively. These results suggest the PCM substrate can

232

electrochemically catalyze the ORR on site over a wide range of solution pH. Such a result bodes

233

well for promoting the coupled Fenton reactions occurring under these conditions.

234

Oxygen reduction taking place on graphitic materials are selective towards a two-electron

235

reduction to produce H2O2 with a selectivity over 90%.25, 26 As a baseline, the amount of H2O2

236

produced via O2 reduction on the PCM support was determined (Fig. 4). These results show that

237

PCM substrate can produce H2O2 at 3 pH values which are consistent with the cyclic voltammetry

238

results at each pH (Fig. S9). In addition, the rate of formation of H2O2 increases with increasing

239

applied potential. However, once the applied potential exceeds to that of ORR peak potential, there

240

is a declining increase in H2O2 productivity for all solution pH tested. This is due to a diffusion

11

ACS Paragon Plus Environment

Environmental Science & Technology

241

controlled limit for H2O2 release from the porous structure of PCM substrate. Thus, with the porous

242

structure there is an elevated concentration of H2O2 that works against a further increase in H2O2

243

production rate as measured in solution. The rates of H2O2 production on 6 mg of the PCM

244

substrate at pH 4, 7, and 10 were 9.06, 9.41, and 9.48 mmol/L/h, respectively, which suggests

245

solution pH has limited effect on its H2O2 productivity. In contrast, nonporous carbon material

246

produced H2O2 at a rate of 4.41 mmol/L/h at pH 7 under similar reaction conditions.27 Since the

247

oxygen reduction performance of carbon based materials is known to be unaffected by buffer,15

248

this suggests solution pH can be easily maintained during the oxygen reduction reaction using

249

PCM material.

250

Electro-Fenton Degradation of Napropamide

251

In order to demonstrate the feasibility of catalytically enhancing heterogeneous electro-

252

Fenton reactions on CMOF@PCM, napropamide was selected as a model compound for testing

253

the electro-Fenton process using CMIL-88@PCM. The initial concentration of napropamide was

254

10 ppm in 50 mL 0.1 M Na2SO4; the effective surface area of the CMIL-88@PCM electrode was

255

6 cm2. For comparison, we investigated the removal efficiency of napropamide via the E-Fenton

256

process under multiple sets of conditions; these results are summarized in Figure 5. H2O2 alone

257

was unable to degrade napropamide. The kinetics of napropamide removal using the CMIL-

258

88(Fe)@PCM electro-Fenton process, electrosorption on CMIL-88(Fe)@PCM, electrosorption

259

on PCM in an Ar-saturated 0.1 M Na2SO4 electrolyte, and electrosorption on PCM in an O2-

260

saturated 0.1 M Na2SO4 electrolyte resulted in loss percentages of 75.5%, 35.4%, 14.6%, and

261

16.0%, respectively. The electrosorption of napropamide by the PCM support material rapidly

262

reaches equilibrium within 30 min due to the limited surface area of the PCM material. On the

263

other hand, CMIL-88(Fe)@PCM removed substantially more napropamide by electrosorption due

12

ACS Paragon Plus Environment

Page 12 of 30

Page 13 of 30

Environmental Science & Technology

264

to the additional CMOFs embedded in the PCM support. The removal efficiency of napropamide

265

on CMIL-88(Fe)@PCM in Ar saturated electrolyte is similar to that of the O2-saturated electrolyte

266

at the presence of isopropanol (IPA) as a ·OH trap. This result suggests that ·OH is the primary

267

oxidant generated during the electro-Fenton process.

268

CMOF@PCM is a unique, multifunctional electrochemical catalyst in which the electro-

269

Fenton reaction is catalyzed by CMOF-NPs while H2O2 is generated on site by the PCM support

270

substrate. Because the electrochemical performance of CMOFs may be controlled by the MOFs

271

precursors, various of MOFs with different topology including MIL-88(Fe), MIL-100(Fe) and

272

MIL-101(Fe) were synthesized to investigate the MOFs topology effects on CMOF’s electro-

273

Fenton performance. Among the MOFs, MIL-88(Fe) and MIL-101(Fe) were synthesized using an

274

identical metal salt and a bidentate linker, BDC. In contrast, a tridentate linker, BTC, was used to

275

synthesize MIL-100(Fe). The relative napropamide removal efficiency decreased in the following

276

order: CMIL-100@PCM > CMIL-88@PCM > CMIL-101@PCM resulting in removal efficiencies

277

of 82.3%, 64.1% and 60.56%, respectively, after 60 min of electrolysis (Fig. 6). The pseudo first-

278

order kinetic plots versus time for napropamide degradation are shown in Figure S10. The

279

corresponding pseudo first-order constants for napropamide removal were determined to be 1.70,

280

0.99, and 0.87 h-1 for CMIL-100@PCM, CMIL-88@PCM, and CMIL-101@PCM, respectively

281

(Table S2).

282

In order to investigate the effects of MOF functional groups on CMOF@PCM electro-

283

Fenton kinetics, amine functionalized MOFs using an aminated linker were synthesized. CMOFs

284

were prepared from BDC and BDC-NH2. However, due to the steric hinderance of the tridentate

285

trismic acid, CMIL-100-NH2@PCM was not prepared in the same fashion. A comparison of

286

Napropamide electro-Fenton removal rates using CMOF@PCM and CMOF-NH2@PCM is given

13

ACS Paragon Plus Environment

Environmental Science & Technology

287

in Figure 7. The removal percentages after 60 min were 73.42%, 64.1%, 60.56%, and 53.63% for

288

CMIL-88-NH2@PCM, CMIL-88@PCM, CMIL-101@PCM, and CMIL-101-NH2@PCM,

289

respectively. Again a pseudo first-order kinetic decay of napropamide was observed (Fig. S11)

290

with first-order constants of 1.26, 0.99, 0.87, 0.74 h-1 for CMIL-88-NH2@PCM, CMIL-88@PCM,

291

CMIL-101@PCM, and CMIL-101-NH2@PCM, respectively (Table S2).

292

The effect loading levels of CMIL-100 on the observed electro-Fenton kinetics of

293

napropamide degradation was studied for samples that were prepared from mixing 10, 25, 50, 75

294

wt% of CMIL-100 into the PCM support. These sample were designated as CMIL-100@PCM10,

295

CMIL-100@PCM25, CMIL-100@PCM50, CMIL-100@PCM75, respectively. Napropamide

296

degradation was carried out with CMIL-100 alone without PCM substrate for comparison. As

297

illustrated in Figure S12, the napropamide degradation percentage after 60 min is 11.8, 89.2, 97.0,

298

94.7, 92.6, and 91.3% for PCM, CMIL-100@PCM10, CMIL-100@PCM25, CMIL-100@PCM50,

299

CMIL-100@PCM75, and CMIL-100, respectively. Napropamide removal also follows apparent

300

pseudo first-order kinetics (Fig. S13). The first-order kinetic constants are summarized in Table

301

S2. The rate of napropamide degradation increased with the loading amount of CMIL-100 from

302

10 to 25 wt% but declined at higher doping concentrations. The maximum degradation efficiency

303

was obtained using CMIL-100@PCM25 (i.e., 97.0% after 60 min of electrolysis). In traditional

304

applications of the electro-Fenton process, contaminant removal efficiencies have been found to

305

increase with increasing amount of heterogeneous electro-Fenton catalyst until a maximum

306

efficiency is reached at higher loading levels.28 In these aforementioned systems, nonporous bulk-

307

phase Fe3O4 was used but there are limits in that non-surface accessible Fe(II) does not participate

308

in the Fenton reaction. Thus, a relatively low utilization efficiency of electro-Fenton catalyst

309

normally requires large amounts of Fe3O4 to be used in order to fully consume the added hydrogen

14

ACS Paragon Plus Environment

Page 14 of 30

Page 15 of 30

Environmental Science & Technology

310

peroxide. In contrast, for CMOF@PCM, the main electro-Fenton active species is porous CMOFs.

311

The active form in our case is Fe3O4-NP that is embedded in graphitic carbon derived from the

312

specific organic linker. The highly uniform crystalline structure of the MOF precursors allows

313

even distribution of the Fe3O4-NPs throughout the CMOF-NPs after carbonization (Fig. S5).

314

Therefore, the embedded Fe3O4-NPs in the graphitic molecular support have similar local

315

coordination environments, which appear to be a prerequisite for maximizing the number of

316

electrocatalytically active sites.29 Since both the PCM support substrate and CMIL-100 NPs are

317

porous materials, they are able to maximize the interactions of H2O2 with the Fe3O4-NPs. However,

318

at a low loading level of CMOFs (e.g., CMIL-100@PCM5), the electro-Fenton reaction rate is

319

limited by the amount of Fe3O4 that was embedded in the multifunctional electrocatalysts,

320

CMOF@PCM. Low loading of Fe3O4 leads to an insufficient amount of Fe(II) available for the

321

reaction with H2O2 to produce a sufficient amount of ·OH radical. However, the electrochemical

322

reduction of O2 as catalyzed by the PCM substrate is the rate limiting step at higher loading levels

323

of CMOF in the form of CMIL-100@PCM75.

324

The impact of pH on napropamide degradation using CMIL-100@PCM was minimal as

325

shown in Figures S14. and S15. The corresponding pseudo first-order rate constants as given

326

Table S2 suggest the overall rate of degradation is somewhat higher at circum-neutral pH than at

327

low or high pH. The applied potential was set at -0.14 V, which is a potential that produces the

328

highest concentration of H2O2 (Fig. 4). Any pH dependence may be due to impact on the sorption

329

of napropamide onto the CMIL-100@PCM electrode.

330

Because actual wastewater has a wide range of conductivity, we have evaluated

331

napropamide removal efficiency at different electrolyte concentration. Figures S16 shows the

332

napropamide removal efficiency was slightly reduced at reduced electrolyte concentration. In

15

ACS Paragon Plus Environment

Environmental Science & Technology

333

addition, Figure S16 suggests napropamide can be successfully removed by CMIL100@PCM in

334

simulated river water prepared according to Table S5.

335 336

Figure 8 shows that about 90% mineralization of napropamide can be achieved within 2 h of electrolysis via the electro-Fenton reaction as catalyzed by CMIL-100@PCM at pH 7.

337

The CMIL-100@PCM electro-Fenton was applied for the degradation of other herbicides

338

beyond (Fig. 9). The degradation of metolachlor, glyphosate, atrazine, acetochlor, and dichloprop

339

also followed apparent first-order kinetics (Fig. S17) with degradation percentages of 83.0, 78.5,

340

70.9, 91.1, and 76.2% after 60 min of electrolysis and kobs values of 1.50, 1.34, 0.99, 1.90, and

341

1.27 h-1, respectively. On the basis of reported electro-Fenton and Fenton degradation

342

mechanism,16 we propose a generalized mechanistic pathway for ·OH mediated degradation of

343

pesticides using CMOF@PCM electrodes (Fig. S18).

344

Reusability and Stability

345

The recyclability of the CMIL-100@PCM electro-Fenton catalyst has been assessed for

346

three consecutive cycles. No significant change to the napropamide degradation efficiency has

347

been observed (Fig. S19), demonstrating the excellent stability of CMOF@PCM. Over time

348

leaching of reactive Fe from CMIL-100@PCM may impact the durability of this system. However,

349

Figure S20 shows that only approximately 6.5 ppb of the Fe was detected in the solution, and the

350

Fe concentration remain unchanged throughout the reaction using CMIL-100@PCM. The initial

351

Fe concentration maybe due to the dissociation of CMIL-100@PCM that was loosely bonded to

352

the carbon paper substrate by nafion binding agent. In comparison, significant Fe leaching has

353

been observed for Fe3O4@PCM, which consists Fe3O4-NP embedded on PCM substrate. The

354

concentration of leached Fe in the electrolyte had increased linearly and reached about 200 ppb

355

after 2 h reaction.

16

ACS Paragon Plus Environment

Page 16 of 30

Page 17 of 30

356

Environmental Science & Technology

The Primary Basis for High E-Fenton Reactivity

357

In order to increase the electrochemical catalytic efficiency, it is important to be able to

358

access the largest number of surface active sites that are capable of participating in an

359

electrocatalytic reaction. However, the reactivity of the electrochemical active sites is affected by

360

both the surface structure of the catalyst and its interactions with molecules or molecular moieties

361

close to the site of catalysis. For example, heterogeneous catalytic reactions are dependent on the

362

catalytically active surface area. In our system, it is difficult to accurately determine the actual

363

electro-chemically active surface area of the Fe3O4-NPs embedded within the porous carbon

364

matrix due to the dielectric behavior of magnetite. Thus, Brunauer-Emmett-Teller (BET) surface

365

area measurements are used as a surrogate. BET measurements of the CMOF surfaces were

366

determined at 77 K by N2 adsorption with the results summarized in Table S3. For non-

367

functionalized CMOFs, the BET surface areas ranged from 287.9, 340.9, and 361.6 m2/g for

368

CMIL-88, CMIL-100, and CMIL-101, respectively. Hysteresis (type H3) between adsorption and

369

desorption branches at medium pressure (P/P0 = 0.4) were observed for both CMIL-88 and CMIL-

370

101 (Fig. S21). This behavior is consistent with the presence of wide distribution of mesopores.

371

In contrast, CMIL-100 gaves a type H4 isotherm, which suggests the presence of narrow slit-like

372

pores that are dominated by micropores with a limited number of mesopores. In fact, the pore size

373

distribution (Fig. S22) showed that the average pore diameters were 13.12, 4.85, 13.65 nm for

374

CMIL-88, CMIL-100, and CMIL-101, respectively. This result indicates that both the BET surface

375

areas and the pore structures of CMOFs are dependent on the linker and core metal ion rather than

376

the topology of parent MOFs. The BET surface areas of the CMOFs synthesized from the aminated

377

MOFs are given in Table S3. These results show that amine functionalization resulted in slightly

378

lower BET surface areas for CMIL-88-NH2 and CMIL-101-NH2 of 217.88 and 211.85 m2/g,

17

ACS Paragon Plus Environment

Environmental Science & Technology

379

respectively. Both CMIL-88-NH2 and CMIL-101-NH2 gave type H4 BET isotherms (Fig. S23)

380

that indicated that micropores control the pore structure. The average pore diameter for CMIL-88-

381

NH2 and CMIL-101-NH2 were determined to be 8.50 and 11.10 nm, respectively (Fig. S24). These

382

results clearly suggest that amine functional groups in the linker lead to a reduction in pore

383

diameters of the CMOFs. These results are also consistent with the location of amine functional

384

group in the MOFs precursor, which is at the edge of the aperture (Fig. S25). These observations

385

highlight the importance of BET surface area in evaluating the reactivity of electro-Fenton

386

catalysts. Furthermore, the results presented in Table S3 also show that the CMOF@PCMs have

387

greater BET surface area than CMOFs without PCM substrate. This is consistent with the higher

388

BET surface area of PCM. In light of these results, we have normalized the pseudo first- order rate

389

constants to the BET surface areas (kSA) to give surface area normalized heterogeneous rate

390

constants. However, since the BET surface area of the PCM substrate is the same for all of the

391

CMOF@PCMs, the BET surface areas of the CMOFs were used instead of CMOF@PCMs for

392

normalization. These results are summarized in Table S4, which show that the normalized rate

393

constants, kSA, follow the order of CMIL-88-NH2@PCM > CMIL-100@PCM > CMIL-101-

394

NH2@PCM = CMIL-88@PCM > CMIL-101@PCM.

395

The normalized rate constants for the electro-Fenton reactions show that the catalytic

396

activity of the CMIL-88-NH2@PCM catalyst is almost twice that of the CMIL-88@PCM in spite

397

of a lower BET surface area (Fig. S26). The PXRD spectra indicate that the CMIL-88 NPs are

398

present primarily as Fe3O4 without an additional phase (Fig. S27). This result is consistent with

399

the high-resolution TEM images that clearly identify nano-particulate Fe3O4 in Figure S28. The

400

observed lattice lines have a lattice spacing of 4.90 Å, which is consistent with the (110) plane of

401

Fe3O4. In contrast, the CMIL-88-NH2 NPs TEM images show the presence of two iron phases,

18

ACS Paragon Plus Environment

Page 18 of 30

Page 19 of 30

Environmental Science & Technology

402

Fe3O4 and Fe3C. The high-resolution TEM images shown in Figure 10 show that the CMIL-88-

403

NH2 NPs are uniformly dispersed within the graphitic PCM substrate. CMIL-88-NH2 has a unique

404

core-shell structure. The graphic carbon shell thickness is ∼2 nm, while a lattice spacing of the

405

intermediate layer of 2.38 Å, which corresponds to the (210) plane of Fe3C. The core is Fe3O4 with

406

a lattice spacing of 2.53 Å. These results are also consistent with the PXRD spectrum (Fig. S27).

407

The difference in electro-Fenton reaction performance between CMIL-88-NH2@PCM and

408

CMIL-88@PCM is most likely due to the difference in their structures. In the case of CMIL-

409

88@PCM, Fe3O4 NPs are imbedded in the carbon substrate matrix. The unit cell for Fe3O4 has

410

cubic inverse spinel structure, each containing 8 Fe atoms occupying Wyckoff position 8a and 16

411

Fe atoms occupying Wyckoff position 18d in space group Fd3m (Fig. S29a). The Fe atoms are

412

directly accessible by the H2O2 for electro-Fenton reaction. In terms of CMIL-88-NH2@PCM,

413

core-shell structured NPs are imbedded in the graphitic carbon substrate. Within the core-shell

414

nanoparticulate structures, the Fe3C interfacial layer is sandwiched between the graphitic carbon

415

shell and Fe3O4 core. Fe3C has orthorhombic structure in a Pnma space group. Each unit cell

416

contains 12 Fe atoms and 4 C atoms. 8 Fe atoms occupy Wyckoff position 4c and the remaining

417

Fe atoms occupy Wyckoff position 8d (Fig. S29b). Compared to the base elemental metal, the

418

formation of a metal carbide (Fe3C) leads to a lengthening of the metal-metal bond distance due

419

to presence of carbon. As the result, a metal d-band contraction occurs, generating more density

420

of states (DOS) near the Fermi level.30-32 Therefore, metal carbides are quasi electron conductors.

421

Incorporation Fe3C as an intermediate layer within CMIL-88-NH2@PCM composite material

422

improves the electro-Fenton reactivity by: (1) facilitating the electron transfer between the

19

ACS Paragon Plus Environment

Environmental Science & Technology

423

graphitic carbon shell and Fe3O4 core, (2) increasing the catalytic reactivity by reducing the lattice

424

mismatch between the graphitic carbon shell and Fe3O4 core,33 and (3) by fixing in three-

425

dimensions Fe(II) species within the Fe3O4 core as well as locking Fe atoms within Fe3C. The

426

latter effect essentially prevents Fe(II) leaching into solution and provides for long-term stability

427

of the electrode. This is confirmed by Fe leaching measurement from the composite catalyst (Fig.

428

S20).

429

In conclusion, the degradation of a set of herbicides has been achieved using an electro-

430

Fenton process based on multifunctional CMOF@PCM catalysts. In this version of an electro-

431

Fenton reaction system, porous carbonized Fe-based metal organic framework compounds

432

(CMOFs) optimize the availability of Fe(II) sites to initiate the initial oxidation of Fe(II) by H2O2

433

to Fe(III) and ∙OH. The reverse reaction, reduction of Fe(III) is achieved by the negative applied

434

potential on the composite carbon substrate electrode (Fig. S18). A low applied potential (-0.14

435

V) combined with a broad operational pH range may make this particle electro-Fenton reaction

436

system cost-effective for the degradation and mineralization of a wide range of herbicides or other

437

oxidizable organic chemical water contaminants.

20

ACS Paragon Plus Environment

Page 20 of 30

Page 21 of 30

Environmental Science & Technology

438

Associated Contents

439

Supporting Information

440

Includes SEM, TEM images, XPS spectrum, and kinetic study.

441 442

Author Information

443

Corresponding Author

444

*(Michael

445

[email protected]

R.

Hoffmann)

Tel:

+1-626-395-4391.

Fax:

+1-626-395-4391.

E-mail:

446 447

Notes

448

The authors declare no competing financial interest.

449 450

Acknowledgments

451

This work was supported by the Bill and Melinda Gates Foundation (Grant OPP1149755),

452

National Natural Science Foundations of China (Grant No. 21777137), International Cooperation

453

Grant of China (Grant No. 21320102007), and the Caltech Rosenburg Innovation Initiative.

21

ACS Paragon Plus Environment

Environmental Science & Technology

455

References

456

1.

457

Glyphosate and AMPA, "pseudo-persistent" pollutants under real world agricultural

458

management practices in the Mesopotamic Pampas agroecosystem, Argentina.

459

Environmental Pollution 2017, 229, 771-779.

460

2.

461

Guha, N.; Scoccianti, C.; Mattock, H.; Straif, K.; Int Agcy Res Canc Monog, W.,

462

Carcinogenicity of tetrachlorvinphos, parathion, malathion, diazinon, and glyphosate.

463

Lancet Oncology 2015, 16, (5), 490-491.

464

3.

465

claim that an herbicide causes cancer.

466

https://www.eenews.net/eedaily/2018/02/07/stories/1060073069 (July 05, 2018),

467

4.

468

1975, 115, (5), 178-179.

469

5.

470

vanEngelsdorp, D.; Pettis, J. S., High levels of miticides and agrochemicals in North

471

American apiaries: implications for honey bee health. Plos One 2010, 5, (3).

472

6.

473

Effects of the herbicide dicamba on nontarget plants and pollinator visitation.

474

Environmental Toxicology and Chemistry 2016, 35, (1), 144-151.

475

7.

476

Effects of sublethal doses of glyphosate on honeybee navigation. Journal of

477

Experimental Biology 2015, 218, (17), 2799-2805.

478

8.

479

crustaceans. Water Pollution Control Federation 1970, 42, (8), 1544-1550.

480

9.

481

chlorophenoxyacetic acid by advanced electrochemical oxidation methods.

482

Environmental Science & Technology 2002, 36, (13), 3030-3035.

Primost, J. E.; Marino, D. J. G.; Aparicio, V. C.; Costa, J. L.; Carriquiriborde, P.,

Guyton, K. Z.; Loomis, D.; Grosse, Y.; El Ghissassi, F.; Benbrahim-Tallaa, L.;

Hiar, C. Under fire by U.S. politicians, World Health Organization defends its

Moffett, J.; Morton, H., How herbicides affect honey bees. American Bee Journal Mullin, C. A.; Frazier, M.; Frazier, J. L.; Ashcraft, S.; Simonds, R.;

Bohnenblust, E. W.; Vaudo, A. D.; Egan, J. F.; Mortensen, D. A.; Tooker, J. F.,

Balbuena, M. S.; Tison, L.; Hahn, M. L.; Greggers, U.; Menzel, R.; Farina, W. M.,

Sanders, h., Toxicities of some herbicides to six species of freshwater Boye, B.; Dieng, M. M.; Brillas, E., Degradation of herbicide 4-

22

ACS Paragon Plus Environment

Page 22 of 30

Page 23 of 30

Environmental Science & Technology

483

10.

Medel, A.; Lugo, F.; Meas, Y., Application of electrochemical processes for

484

treating effluents from hydrocarbon industries. In Electrochemical water and wastewater

485

treatment, Martínez-Huitle, C. A., Ed. Elsevier: 2018; pp 365-392.

486

11.

487

oxidation by using ozone, hydrogen peroxide, and ferrate. In Electrochemical water and

488

wastewater treatment, Martínez-Huitle, C. A., Ed. Elsevier: 2018; pp 165-192.

489

12.

490

ingredients and commercial formulations in water by the photo-assisted Fenton

491

reaction. Water Research 1999, 33, (5), 1238-1246.

492

13.

493

P. L.; Comninellis, C., Electrochemical destruction of chlorophenoxy herbicides by

494

anodic oxidation and electro-Fenton using a boron-doped diamond electrode.

495

Electrochimica Acta 2004, 49, (25), 4487-4496.

496

14.

497

membrane stack for flow-through sequential regenerative electro-Fenton. Environmental

498

Science & Technology 2015, 49, (4), 2375-2383.

499

15.

500

mineralization of perfluorooctanoate by electro-Fenton with H2O2 electro-generated on

501

hierarchically porous carbon. Environmental Science & Technology 2015, 49, (22),

502

13528-13533.

503

16.

504

electrochemical technologies based on Fenton's reaction chemistry. Chemical Reviews

505

2009, 109, (12), 6570-6631.

506

17.

507

and applications. In Electrochemical Water and Wastewater Treatment, Martínez-Huitle,

508

C. A., Ed. Elsevier: 2018; pp 193-221.

509

18.

510

Mineralization of Carbamazepine Using an Electro-Fenton Reaction Catalyzed by

511

Magnetite Nanoparticles Fixed on an Electrocatalytic Carbon Fiber Textile Cathode.

512

Environmental Science & Technology 2018, 52, (21), 12667-12674.

Sáez, C.; Rodrigo, M.; Fajardo, A.; Martínez-Huitle, C., Indirect electrochemical

Huston, P. L.; Pignatello, J. J., Degradation of selected pesticide active

Brillas, E.; Boye, B.; Sires, I.; Garrido, J. A.; Rodriguez, R. M.; Arias, C.; Cabot,

Gao, G. D.; Zhang, Q. Y.; Hao, Z. W.; Vecitis, C. D., Carbon nanotube

Liu, Y. M.; Chen, S.; Quan, X.; Yu, H. T.; Zhao, H. M.; Zhang, Y. B., Efficient

Brillas, E.; Sires, I.; Oturan, M. A., Electro-Fenton process and related

Oturan, N.; Oturan, M., Electro-Fenton process: background, new developments,

Liu, K.; Yu, J. C. C.; Dong, H.; Wu, J. C. S.; Hoffmann, M. R., Degradation and

23

ACS Paragon Plus Environment

Environmental Science & Technology

513

19.

Jagadeesh, R. V.; Murugesan, K.; Alshammari, A. S.; Neumann, H.; Pohl, M. M.;

514

Radnik, J.; Beller, M., MOF-derived cobalt nanoparticles catalyze a general synthesis of

515

amines. Science 2017, 358, (6361), 326-+.

516

20.

517

formaldehyde. Journal of Materials Science 1989, 24, (9), 3221-3227.

518

21.

519

microporous lanthanide organic framework. Crystal Growth & Design 2010, 10, (5),

520

2025-2028.

521

22.

522

potassium titanium(IV) oxalate. Analyst 1980, 105, (1255), 950-954.

523

23.

524

adsorption and degradation of acid orange 7 in aqueous solution via persulfate

525

activation. Applied Surface Science 2016, 369, 130-136.

526

24.

527

Ponpandian, N., Solvent-free mechanochemical synthesis of graphene oxide and

528

Fe3O4-reduced graphene oxide nanocomposites for sensitive detection of nitrite.

529

Journal of Materials Chemistry A 2015, 3, (30), 15529-15539.

530

25.

531

doped carbon for the electrocatalytic synthesis of hydrogen peroxide. Journal of the

532

American Chemical Society 2012, 134, (9), 4072-4075.

533

26.

534

nitrogen-doped carbon. Applied Surface Science 2009, 256, (1), 335-341.

535

27.

536

Lanza, M. R. V., Electrogeneration of hydrogen peroxide in gas diffusion electrodes

537

modified with tert-butyl-anthraquinone on carbon black support. Carbon 2013, 61, 236-

538

244.

539

28.

540

Fe@Fe2O3/ACF composite cathode for wastewater treatment. Journal of Hazardous

541

Materials 2009, 164, (1), 18-25.

542

29.

543

single-atom electrocatalysts. Nature Catalysis 2018, 1, (1).

Pekala, R. W., Organic aerogels from the polycondensation of resorcinol with Silva, P.; Valente, A. A.; Rocha, J.; Paz, F. A. A., Fast microwave synthesis of a

Sellers, R. M., Spectrophotometric determination of hydrogen-peroxide using Li, X. H.; Guo, W. L.; Liu, Z. H.; Wang, R. Q.; Liu, H., Fe-based MOFs for efficient

Bharath, G.; Madhu, R.; Chen, S. M.; Veeramani, V.; Mangalaraj, D.;

Fellinger, T. P.; Hasche, F.; Strasser, P.; Antonietti, M., Mesoporous nitrogen-

Okamoto, Y., First-principles molecular dynamics simulation of O2 reduction on Valim, R. B.; Reis, R. M.; Castro, P. S.; Lima, A. S.; Rocha, R. S.; Bertotti, M.;

Li, J. P.; Ai, Z. H.; Zhang, L. Z., Design of a neutral electro-Fenton system with

Xu, H.; Cheng, D.; Cao, D.; Zeng, X., A universal principle for a rational design of

24

ACS Paragon Plus Environment

Page 24 of 30

Page 25 of 30

Environmental Science & Technology

Heine, V., s-d interaction in transition metals. Physical Review 1967, 153, (3),

544

30.

545

673-&.

546

31.

547

Materials for Low Temperature Fuel Cells. Energies 2009, 2, (4), 873-899.

548

32.

549

catalysis. Science 1973, 181, (4099), 547-549.

550

33.

551

Nanoparticles for Optimizing Catalytic Activity. Nano Letters 2015, 15, (6), 4089-4095.

Ham, D. J.; Lee, J. S., Transition Metal Carbides and Nitrides as Electrode Levy, R. B.; Boudart, M., Platinum-like behavior of tungsten carbide in surface Moseley, P.; Curtin, W. A., Computational Design of Strain in Core-Shell

552

25

ACS Paragon Plus Environment

Environmental Science & Technology

554

Page 26 of 30

Figure 1. Structure of (a) MIL-88(Fe), (b) MIL-100(Fe), and (c) MIL-101(Fe).

555 (a)

(b)

(c)

556 557 558

Figure 2. PXRD spectrum of CMOFs used in the current study.

CMIL-101-NH2 CMIL-101 CMIL-100 CMIL-88-NH2 CMIL-88 Simulated Fe3O4 Simulated FeC3 Simulated Fe0 20

30

40

50

60

70

80

90



2  

559 560

Figure 3. SEM micrographs of (a) CMIL-88@PCM, (b) CMIL-100@PCM, and (c) CMIL-

561

101@PCM. (a)

(b)

(c)

26

ACS Paragon Plus Environment

Page 27 of 30

562

Environmental Science & Technology

400 nm

563

Figure 4. The concentration of H2O2 produced by 6 mg of PCM substrate at (a) pH 4, (b) pH 7,

564

and (c) pH 10.

565

(b) 0.08V -0.12V -0.32V

9 6 3 0

0

30 60 Time (min)

90

(c) 0.26V 0.06V -0.14V

12 9 6 3 0

0

30 60 Time (min)

90

H2O2 Concentration (mmol/L)

12

H2O2 Concentration (mmol/L)

H2O2 Concentration (mmol/L)

(a)

0.42V 0.24V 0.04V

12 9 6 3 0

0

30 60 Time (min)

90

566 567 568 569 570

Figure 5. Napropamide (10 ppm) removal by H2O2 oxidation (400 ppm), PCM substrate (-0.14

571

V), and CMIL-88@PCM (-0.14 V). All experiments were performed in 0.1 M Na2SO4

572

electrolyte at pH 7.

573

27

ACS Paragon Plus Environment

Environmental Science & Technology

1.00

C/C0

0.75 H2O2 PCM (O2) PCM (Ar2) CMIL-88@PCM (O2) CMIL-88@PCM (Ar2) CMIL-88@PCM (O2)/IPA

0.50 0.25 0

30 60 Time (min)

574

90

575 576

Figure 6. Napropamide (10 ppm) removal by electro-Fenton using CMOF@PCMs prepared

577

from MIL-88(Fe), MIL-100(Fe), and MIL-101(Fe). All experiments were performed in 0.1 M

578

Na2SO4 electrolyte at pH7 (-0.14 V).

579 1.0

CMIL-101@PCM CMIL-100@PCM CMIL-88@PCM

C/C0

0.8 0.6 0.4 0.2 0

580

15

30 45 Time (min)

60

581 582 583 584

Figure 7. Napropamide (10 ppm) removal by electro-Fenton using CMOF@PCMs and CMOFs-

585

NH2@PCMs (0.1 M Na2SO4, pH7, -0.14 V). 28

ACS Paragon Plus Environment

Page 28 of 30

Page 29 of 30

Environmental Science & Technology

586 CMIL-101@PCM CMIL-101-NH2@PCM CMIL-88@PCM CMIL-88-NH2@PCM

1.0

C/C0

0.8 0.6 0.4 0.2

0

15

587

30 45 Time (min)

60

588

Figure 8. TOC removal efficiency of electro-Fenton removal of 10 ppm napropamide by CMIL-

589

100@PCM (0.1 M Na2SO4, pH 7, -0.14 V).

590

1.0

C/C0

0.8 0.6 0.4 0.2 0.0

591

0

30

60 90 120 150 180 Time (min)

592 593 594 595 596

Figure 9. Electro-Fenton degradation of metolachlor, glyphosate, atrazine, acetochlor, and

597

dichlorprop (10 ppm) by CMIL-100@PCM (0.1 M Na2SO4, -0.14 V, pH 7). 29

ACS Paragon Plus Environment

Environmental Science & Technology

598

Metolachlor Glyphosate Atrazine Acetochlor Dichlorprop

1.00

C/C0

0.75 0.50 0.25 0.00

599

30 45 60 Time (min)

600

Figure 10. (a) TEM images of CMIL-88-NH2@PCM, (b) high-resolution TEM image of CMIL-

601

88-NH2@PCM core-shell structure. (a)

0

15

75

(b)

602 603

30

ACS Paragon Plus Environment

90

Page 30 of 30