Identifying the Nonradical Mechanism in the Peroxymonosulfate

May 23, 2018 - ... Magnetic Phenomena · Nuclear Phenomena · Nuclear Technology · Optical, .... Persulfate activation that does not rely on the oxidizi...
0 downloads 0 Views 665KB Size
Subscriber access provided by University of Winnipeg Library

Remediation and Control Technologies

Identifying the Nonradical Mechanism in the Peroxymonosulfate Activation Process: Singlet Oxygenation Versus Mediated Electron Transfer Eun-Tae Yun, Jeong Hoon Lee, Jaesung Kim, Hee-Deung Park, and Jaesang Lee Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b00959 • Publication Date (Web): 23 May 2018 Downloaded from http://pubs.acs.org on May 23, 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 36

Environmental Science & Technology

1

Identifying the Nonradical Mechanism in the Peroxymonosulfate Activation

2

Process: Singlet Oxygenation Versus Mediated Electron Transfer

3

Eun-Tae Yun1†, Jeong Hoon Lee1†, Jaesung Kim1, Hee-Deung Park1, and Jaesang Lee1*

4 5 6

1

7

Abstract. Select persulfate activation processes were demonstrated to initiate oxidation not

8

reliant on sulfate radicals, though the underlying mechanism has yet to be identified. This study

9

explored singlet oxygenation and mediated electron transfer as plausible nonradical mechanisms

10

for organic degradation by carbon nanotube (CNT)-activated peroxymonosulfate (PMS). The

11

degradation of furfuryl alcohol (FFA) as a singlet oxygen (1O2) indicator and the kinetic

12

retardation of FFA oxidation in the presence of L-histidine and azide as 1O2 quenchers apparently

13

supported a role of 1O2 in the CNT/PMS system. However, the 1O2 scavenging effect was

14

ascribed to a rapid PMS depletion by L-histidine and azide. A comparison of CNT/PMS and

15

photoexcited Rose Bengal (RB) excluded the possibility of singlet oxygenation during

16

heterogeneous persulfate activation. In contrast to the case of excited RB, solvent exchange (H2O

17

to D2O) did not enhance FFA degradation by CNT/PMS and the pH- and substrate-dependent

18

reactivity of CNT/PMS did not reflect the selective nature of 1O2. Alternatively, concomitant

19

PMS reduction and trichlorophenol oxidation were achieved when PMS and trichlorophenol

20

were physically separated in two chambers using a conductive vertically aligned CNT membrane.

21

This result suggested that CNT-mediated electron transfer from organics to persulfate was

22

primarily responsible for the nonradical degradative route.

23

Keywords: persulfate activation, nonradical mechanism, singlet oxygenation, mediated electron

24

transfer, carbon nanotubes

School of Civil, Environmental, and Architectural Engineering, Korea University, Seoul 136-701, Korea

*Corresponding author: E-mail: [email protected]; phone: +82-2-3290-4864; fax: +82-2-928-7656 †

These authors contributed equally to this work.

25 26 27 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 36

29

INTRODUCTION

30

The electron donating potentials of transition metals (e.g., cobalt and iron) can initiate the one-

31

electron reduction of peroxymonosulfate (PMS) and the associated production of sulfate radicals

32

(SO4•−), thus achieving radical-induced oxidation of organics.1-3 This process is in accordance

33

with various physicochemical activation strategies (e.g., photocatalysis,4,

34

radiolysis7) in which the generated (conduction band) electrons reductively cleave the peroxide

35

bond of PMS to form SO4•−. In contrast, recent studies8-15 have raised the likelihood of persulfate

36

activation via a degradative reaction pathway involving no radical attack based on the following

37

universal experimental results: no quenching by methanol and chloride as SO4•− scavengers and

38

the absence of electron paramagnetic resonance (EPR) spectra assigned to SO4•− adducts.

39

Persulfate activation that does not rely on the oxidizing power of SO4•− has been demonstrated to

40

proceed with metal- and carbon-based activators, and two hypotheses, i.e., singlet oxygenation

41

and mediated electron transfer, have been postulated for the underlying mechanism.8-10, 12-18

5

electrolysis,6 and

42

Singlet oxygen (1O2) has been hypothesized to have a role in persulfate activation

43

processes based on the following results: oxidation of organics by activated PMS is decelerated

44

in the presence of excess azide and L-histidine as 1O2 quenchers, and PMS activation in the

45

presence of 2,2,6,6-tetramethylpiperidine (TEMP) as a spin-trapping agent produces the EPR

46

spectrum assigned to the corresponding 1O2 adduct.12,

47

alkaline aqueous PMS solutions when phenols are selected as target substrates or benzoquinone

48

is added.22, 23 As ketones (characterized by the presence of a carbonyl moiety (C=O)) have been

49

found to accelerate 1O2 production associated with PMS decay under alkaline conditions (i.e.,

50

HSO5− + SO52− → HSO4− + SO42− + 1O2),24 quinones formed as intermediates during phenol

51

oxidation or supplied externally could provide carbonyl groups to activate PMS and produce 1O2

13, 16-22

ACS Paragon Plus Environment

Singlet oxygenation occurs in

Page 3 of 36

Environmental Science & Technology

52

at basic pH.22, 23 The addition of benzoquinone, which serves as a O2•− scavenger, reduces the

53

efficiency of PMS activation by base, suggesting an alternative 1O2 formation route that involves

54

the superoxide radical anion (O2•−) as an intermediate (i.e., 2O2•− + 2H+ → H2O2 + 1O2).21

55

Carbonaceous nanomaterials, such as N-doped graphene and carbon nanotubes (CNTs), could

56

catalyze the self-decomposition of persulfate, even at acidic or neutral pH (note that alkaline

57

conditions favor 1O2 generation through homogeneous persulfate activation21-23), mediating the

58

nonphotochemical production of 1O2.13, 16, 17, 20 Surface modification with glutaraldehyde, which

59

remarkably increases the surface density of carbonyl groups on CNTs, enhances 1O2 formation

60

from peroxydisulfate (PDS),13 suggesting that nonradical persulfate activation occurs via

61

carbonyl moieties intrinsically present on nanocarbon surfaces.13, 20 Perovskite oxide- and metal-

62

derived activators also enable singlet oxygenation upon PMS addition, though the relevant

63

mechanism has not been empirically verified.12, 18

64

Electron transfer from organic compounds to persulfate, facilitated by select activators,

65

could be responsible for the nonradical mechanism.8-10, 15, 25 PMS activation systems utilizing

66

CNTs and graphitized nanodiamonds have been clearly distinguished from Co2+/PMS as a

67

reference system2 based on the negligible hydroxylation characteristics of SO4•−-mediated

68

oxidation and substrate-dependent degradation efficacies that contradicted the reactivity of

69

SO4•−.8, 9, 25 PMS decay and current generation (in a three-electrode cell) is markedly pronounced

70

when the organic substrate, PMS, and activator co-exist, revealing that the electron-transfer

71

mediating action of the activator allows PMS to abstract electrons from organics.10, 25 Duan et

72

al26 suggested that surface-complexed PMS on N-doped CNTs could trigger the mediated

73

electron transfer from phenol to PMS involved in the complexation reaction. Our recent work25

74

also supports a degradative route based on mediated electron transfer, as any oxyanion that could

ACS Paragon Plus Environment

Environmental Science & Technology

75

serve as an electron acceptor (e.g., periodate, peracetate, PMS, and PDS) was found to facilitate

76

the oxidative degradation of organics in the presence of CNTs. The similarities in substrate-

77

specificity and product distribution confirmed that the reaction pathway induced on CNTs was

78

not unique to the type of oxyanion.25 In contrast, zerovalent nanosized iron (nFe0) as an activator

79

causes the reductive conversion of oxyanions into the corresponding radicals; thus, the

80

degradative reaction pathway is dependent on the nature of the oxyanion-derived radical.25 As

81

aforementioned, two hypothetical mechanisms have been established for persulfate activation not

82

involving SO4•−, but there still remains a fundamental gap in the understanding of the nonradical

83

mechanism.

84

To improve our understanding of the PMS activation mechanism involving no radical

85

attack, in this study, we examined the possibility of oxidative degradation of organics via singlet

86

oxygenation and mediated electron transfer during PMS activation by CNTs (CNT-activated

87

persulfate was demonstrated to cause the nonradical degradative pathway9, 13, 14). In an effort to

88

identify a role of 1O2 in PMS activation, we compared CNT-activated PMS (i.e., CNT/PMS)

89

with photoexcited Rose Bengal (RB), which is well-known to photosensitize singlet

90

oxygenation.27 For these systems, the effects of chemical reagents able to scavenge 1O2 (L-

91

histidine and azide) and extend the lifetime of 1O2 (D2O),27 the dependence of oxidizing capacity

92

on substrate type and pH, and EPR spectral features were compared. Furthermore, to explore the

93

electron-transfer mediating action of CNTs during PMS activation, we tested a vertically aligned

94

CNT membrane (VA-CNT membrane) that physically separates the reaction system into two

95

zones but allows electron transport through the CNT arrays for interzone electron delivery from

96

organics to PMS across the membrane.

97

MATERIALS AND METHODS

ACS Paragon Plus Environment

Page 4 of 36

Page 5 of 36

Environmental Science & Technology

98

Chemicals and Materials. Single-walled CNT (> 95%) was purchased from NanoLab Inc.

99

Other chemicals were of reagent grade (see Supporting Information), and used without further

100

purification or treatment. Ultrapure deionized water (>18 MΩ•cm), produced with a Millipore

101

system, was used to prepare all experimental suspensions and solutions

102

Preparation and Characterization of VA-CNT Membrane. VA-CNTs were synthesized by a

103

water-vapor-assisted chemical vapor deposition (CVD) technique according to the previously

104

reported procedure.28 Briefly, VA-CNTs were grown on a SiO2/Si wafer coated with aluminum

105

and iron as a catalyst layer. The coated SiO2/Si wafer (width: 1 cm, length: 1 cm) was placed in a

106

CVD reactor. The growth of millimeter-scale VA-CNT arrays (thickness: ca. 900 µm) was

107

carried out at 750 °C for 2 h under a constant flow of ethylene as a carbon source and hydrogen

108

and argon as carrier gases at rates of 100, 200, and 300 sccm (standard cubic centimeters per

109

minute), respectively. Water vapor, a known agent for enhancing and preserving the performance

110

of metal catalysts, was supplied at a feed rate of 30 sccm. Scanning electron microscopy (SEM;

111

Quanta 250 FEG, Thermo Scientific) revealed a densely packed forest of aligned CNTs on the

112

silicon wafer (Figure S1). The Raman spectrum of the CNTs (500–3000 cm−1; LabRam

113

ARAMIS, Horiba Jobin-Yvon; argon ion laser excitation (514.5 nm)) included a G-band

114

corresponding to graphite structures and a D-band sensitive to defects at 1582 and 1350 cm−1,

115

respectively29 (Figure S2). Scheme S1 illustrates the procedure for VA-CNT membrane

116

fabrication. An epoxy resin (Epon Resin 828, Miller-Stephenson Inc.) was mixed with a curing

117

agent (Jeffamine D-230, Huntsman Corporation) at a 3:1 weight ratio.30 The resin mixture was

118

infiltrated into the interstitial spaces of the as-grown CNTs. Then, the VA-CNT/epoxy composite

119

was incubated under vacuum for 3 h on a mold. The resultant composite was allowed to cure at

120

room temperature for 24 h. Finally, the residual resin and catalysts were removed, the CNT-

ACS Paragon Plus Environment

Environmental Science & Technology

121

based membrane was detached from the silicon substrate, and the VA-CNTs were uncapped by

122

cutting the top surface and bottom of the composite using a microtome (HM 340 E, MICROM

123

Lab.) to produce the VA-CNT membrane (with open CNT tips). X-ray photoelectron

124

spectroscopy (XPS; PHI X-tool, ULVAC-PHI) confirmed that no detectable amounts of metal

125

species (e.g., iron) that may activate persulfate remained on the surface of the VA-CNT

126

membrane (Figure S3).

127

To explore the possibility of interchamber electron transport across the VA-CNT

128

membrane, we monitored concurrent oxidation of organics and reduction of PMS in a crossflow

129

filtration system in which the feed water migrated tangentially across the VA-CNT membrane

130

surface. The VA-CNT membrane (effective surface area: 4 cm2, thickness: 0.9 cm) was

131

vertically mounted in a module to partition the reaction system into two chambers. Water flow in

132

the two physically separated compartments was circulated by a gear pump (REGLO-Z,

133

ISMATEC) at a feed rate of 500 mL min−1. To guarantee the complete removal of metallic

134

species that may activate persulfate, the pristine VA-CNT membrane was subjected to UV-C

135

treatment for 5 h (UV-C irradiation could photochemically cleave some organic linkers that

136

might attach metals to the membrane, which allowed for the release of loosely bound metallic

137

species) and washed thrice with deionized water prior to its application in the filtration process.

138

Experimental Procedure and Analytical Methods. Oxidative degradation of organic substrates

139

by activated PMS was performed in a magnetically stirred 40 mL reactor under air-equilibrated

140

conditions. A typical experimental suspension contained 0.1 g L−1 CNTs, 1 mM PMS, and 0.05

141

mM target pollutant. Photosensitized singlet oxygenation of organics proceeded in a 40 mL

142

cylindrical quartz reactor with six fluorescent lamps (output power: 4 W; Philips Co.). The

143

significant overlap between the emission spectrum of the light source and the absorption

ACS Paragon Plus Environment

Page 6 of 36

Page 7 of 36

Environmental Science & Technology

144

spectrum of RB (1O2 photosensitizer; applied at an initial concentration of 0.05 mM) indicated

145

that the photoexcitation of RB readily occurred under fluorescent lamp irradiation (Figure S4).

146

The incident light intensity, measured using a pyranometer (Apogee, PYR-P), was determined to

147

be 1.105 mW cm−2. The suspensions (or solutions) were initially adjusted to pH 7 and buffered

148

using 1 mM phosphate buffer in most cases. No significant pH change was observed over the

149

course of PMS activation. To investigate pH effects, the initial pH of the aqueous suspensions

150

(or solutions) was adjusted to the desired value using concentrated HClO4 and NaOH, and 1 mM

151

phosphate and carbonate buffers were used for maintaining neutral and alkaline pH, respectively.

152

Sample aliquots were withdrawn from the reactor at fixed time intervals using a 1 mL syringe,

153

filtered through a 0.45 µm PTFE filter (Millipore), and injected into a 2 mL amber glass vial. In

154

PMS activation experiments, excess methanol (0.5 M) was added to quench any remaining

155

radicals. The residual concentrations of organic pollutants were measured using an HPLC

156

(Agilent Infinity 1260) system equipped with a C-18 column (ZORBAX Eclipse XDB-C18) and

157

a UV/vis detector (G1314F 1260VWD). The mobile phase comprised 0.1% (v/v) aqueous

158

phosphoric acid solution and acetonitrile at a volume ratio of 45:55. According to the method

159

proposed by Liang et al., PMS was colorimetrically determined based on the amount of iodine

160

(λmax = 352 nm) formed via the oxidation of iodide by PMS.31 For EPR analysis, 5,5-dimethyl-

161

pyrroline N-oxide (DMPO) and TEMP were used as spin-trapping agents for SO4•−, azidyl

162

radical (N3•), and 1O2, respectively. EPR spectra of the aqueous CNT/PMS suspensions (aqueous

163

binary mixture of azide and PMS or fluorescent-light-irradiated RB solutions) were recorded

164

using a JES-TE 300 spectrometer (JEOL) under the following conditions: microwave power = 1

165

mW, microwave frequency = 9.4136 GHz, center field = 3350 G, modulation width = 0.1 mT,

166

and modulation frequency = 100 kHz. Raman spectra of fresh and used CNTs were acquired on a

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 36

167

LabRam ARAMIS Raman spectrometer (Horiba Jobin-Yvon) using an argon ion laser

168

(excitation at 514.5 nm). Sulfur-containing chemical moieties on CNTs (after exposure to PMS

169

solution) were identified by Fourier transform infrared spectroscopy (FT-IR, Thermo Scientific

170

Nicolet 6700) performed in ATR (attenuated total reflectance) mode. The zeta potential of the

171

aqueous-suspended CNTs was recorded as a function of pH using a Zetasizer Nano ZS

172

(Malvern).

173

RESULTS AND DISCUSSION

174

Oxidation of furfuryl alcohol (FFA) by Activated PMS. To explore whether 1O2 was produced

175

during PMS activation, oxidative degradation of FFA, as a 1O2 indicator,27 by the CNT/PMS

176

system was examined (Figure 1a). Sorption on CNTs caused no noticeable FFA decay, but PMS

177

alone directly oxidized FFA to a certain extent. However, CNTs in the presence of PMS

178

decomposed FFA more rapidly, with k(FFA) = 0.0054 ± 0.0014 min−1 for PMS alone versus

179

k(FFA) = 0.3750 ± 0.0207 min−1 for CNT/PMS. To further confirm that 1O2 had a role in CNT-

180

induced PMS activation, the effects of excess azide and L-histidine as 1O2 scavengers were

181

investigated (Figure 1a). FFA oxidation was completely quenched upon addition of L-histidine,

182

and kinetic retardation of FFA decay was significant upon addition of azide. The results

183

appeared to corroborate the previous findings12, 16-18 that singlet oxygenation was responsible for

184

the nonradical reaction pathway in PMS activation processes. However, solvent exchange (H2O

185

to D2O) did not accelerate FFA degradation at all, which contradicted the usual behavior of 1O2

186

in D2O, as the lifetime of 1O2 is extended up to 10 times when H2O is replaced with D2O.32 In

187

particular, the solvent exchange did not kinetically affect the PMS decay in the aqueous CNT

188

suspension (Figure S5). Since the highly accelerated self-decomposition of PMS was presumed

189

to result in 1O2 production during PMS activation by carbocatalysts,13,

ACS Paragon Plus Environment

16, 17, 20

the results

Page 9 of 36

Environmental Science & Technology

190

indicated that the alternative use of D2O had no influence on the kinetics of the reaction that

191

might allow the CNTs to transform PMS into 1O2.

192

Figure 1b shows the change in FFA degradation efficiency with quencher addition and

193

solvent exchange for photoexcited RB as a benchmark 1O2 producer.27 As observed in the

194

CNT/PMS system, FFA decomposition in the fluorescent-light-irradiated solution of RB was

195

drastically decelerated in the presence of azide and L-histidine. However, the use of D2O as a

196

solvent kinetically enhanced FFA degradation by RB under photoillumination. This behavior

197

was in marked contrast to the effect of D2O on the FFA degradation efficiency of CNT/PMS but

198

reflected the known properties of 1O2. Thus, this result suggested that singlet oxygenation may

199

not occur during persulfate activation. We also monitored H2O2 production as an alternative

200

indicator of singlet oxygenation during the oxidation of ascorbate by CNT/PMS and

201

photoexcited RB. The singlet oxygenation of ascorbate is accompanied by significant H2O2

202

formation.33 Though we found no difference in FFA oxidation efficiency between CNT/PMS and

203

photoexcited RB (Figures 1a and 1b), a much higher H2O2 formation yield was achieved with the

204

fluorescent-light-irradiated RB solution (Figure S6). The result also revealed that singlet

205

oxygenation likely contributed insignificantly to the oxidizing capacity of CNT/PMS.

206

The observed PMS concentrations in the PMS/quencher systems (Figure 2) revealed that and azide as 1O2 scavengers may cause misinterpretation of the

207

the choice of

208

experimental results on PMS activation processes, even though these compounds have been

209

widely employed to evidence the role of 1O2 as an oxidant in environmental processes.27, 34 The

210

inhibition of FFA degradation by CNT/PMS in the presence of excess L-histidine (Figure 1a)

211

appeared to be consistent with the retarding effect of L-histidine reported in the literature.12, 19

212

However, this drastic reduction in FFA oxidation efficiency resulted not from the scavenging

L-histidine

ACS Paragon Plus Environment

Environmental Science & Technology

213

activity of L-histidine toward 1O2 but from rapid PMS depletion by excess L-histidine (Figure 2a).

214

Note that direct reaction with L-histidine at an initial concentration of 100 mM reduced the PMS

215

concentration to the undetectable level within 5 min, and the reaction accelerated as L-histidine

216

concentration increased (Figure 2a). The quenching effect of azide (Figure 1a) also seemed to

217

support previous studies16-18 that suggest singlet oxygenation as the nonradical mechanism, but

218

we found that a binary mixture of PMS (100 mM) and azide exhibited significant FFA

219

degradation efficiency, which was not affected by the addition of CNTs (Figure 2b, inset). In fact,

220

FFA decay by CNT/PMS in the presence of excess azide, which was initially believed to be

221

kinetically hindered by the 1O2 scavenging activity of azide (Figure 1a), could instead be mainly

222

ascribed to the oxidizing capacity of azide/PMS (direct PMS reduction by L-histidine was not

223

accompanied by FFA oxidation (Figure S7)). In this case, the majority of PMS initially added

224

was rapidly consumed through direct reaction with azide as a reducing anion35 (Figure 2b),

225

rendering PMS unavailable for further reaction with residual azide (or CNTs).

226

To explore the reactivity of azide/PMS, we examined the binary mixture of azide and

227

PMS for the oxidative degradation of diverse organic substrates including 4-chlorophenol (4-CP),

228

2,4,6-trichlorophenol (TCP), 4-nitrophenol, and carbamazepine (Figure S8). Unlike FFA, which

229

was rapidly decomposed by azide/PMS, the other organic compounds were barely degraded in

230

aqueous azide/PMS solutions, which implied that direct azide oxidation by PMS led to the

231

production of a highly selective oxidant. The previous finding that N3• exhibited a very low

232

reactivity toward aromatic compounds substituted with electron-withdrawing groups36 likely

233

reveals the possibility that PMS could oxidatively convert azide into N3•. Further, the EPR

234

spectrum obtained for the aqueous azide/PMS mixture showed features that are assignable to the

235

formation of N3• (Figure S9).37 Although the kinetic rate of PMS degradation increased

ACS Paragon Plus Environment

Page 10 of 36

Page 11 of 36

Environmental Science & Technology

236

proportionally with the initial azide concentration (Figure 2b), the FFA removal efficiency

237

decreased with increasing azide concentration (inset of Figure 2b). This concentration-dependent

238

efficiency may be because excess azide likely favored the reaction routes that can deactivate N3•

239

(e.g., 2N3• → 3N2 (k = 4.5 × 109 M−1 s−1)38; N3• + N3− → N6•− (k = 1.0 × 106 M−1 s−1)38), but a

240

further study is required for an in-depth investigation into the mechanism underlying the PMS-

241

mediated production of N3• from PMS. Overall, the apparent inhibitory effects of 1O2 quenchers

242

that have been considered as evidence for 1O2 formation in persulfate activation systems are

243

attributable to the reactivity of L-histidine and azide toward PMS, which was consistent with the

244

previous report39 on the effective PMS consumption by L-histidine and azide.

245

FFA degradation was insignificant in aqueous suspensions of CNTs when PDS was

246

applied instead of PMS (Figure S10), which is consistent with the previous finding14 that

247

CNT/PMS was more effective for decomposing FFA than CNT/PDS. In contrast, a comparison

248

of the 4-CP degradation efficiencies for PMS and PDS activated with CNTs indicated that the

249

oxidizing powers of these two systems were similar (Figure S11). To examine the possibility of

250

1

251

4-CP oxidation efficiency was investigated (Figure S11). These reagents, as a 1O2 scavenger and

252

a singlet oxygenation enhancer, respectively, had no effect on the kinetic rate of 4-CP oxidation;

253

L-histidine

254

alternative solvent did not accelerate the 4-CP decay (Figure S11). The results eliminated the

255

possible contribution of 1O2 to the decomposition of organics by activated PDS. In contrast to

256

CNT/PDS, 4-CP was barely oxidized when excess

257

suspension of CNT/PMS (Figure S11). Although PMS was reductively decomposed by L-

258

histidine (Figure 2a), no loss of PDS was observed in the presence of 100 mM L-histidine

O2 formation during CNT-induced PDS activation, the effect of L-histidine and D2O addition on

negligibly quenched the degradation of 4-CP by CNT/PDS, and the use of D2O as an

L-histidine

ACS Paragon Plus Environment

was added to an aqueous

Environmental Science & Technology

259

(Figure S12). This result confirmed that the significantly inhibited oxidation of organics by

260

CNT/PMS (Figures 1a and S11) was due to rapid PMS degradation by excess L-histidine rather

261

than 1O2 scavenging. No acceleration in 4-CP degradation occurred in aqueous CNT/PMS

262

suspension when H2O was replaced with D2O (applied as an enhancer for singlet oxygenation)

263

(Figure S11).

264

While CNT/PMS caused much more rapid FFA oxidation than CNT/PDS (Figures 1a and

265

S10), similar 4-CP degradation efficiencies were observed, irrespective of whether PMS or PDS

266

was used (Figure S11). This reveals that the production of the reactive oxygen species (e.g., 1O2

267

and SO4•−) through persulfate activation may not be primarily responsible for the oxidizing

268

capacities of CNT/PMS and CNT/PDS. If the CNT-induced activation of persulfate involved 1O2

269

formation, the experiments using FFA as a 1O2 probe (Figures 1a and S10) would suggest that

270

CNT/PMS was much superior to CNT/PDS with respect to singlet oxygenation, which

271

contradicted with the lack of difference observed in the 4-CP treatment efficiency between

272

CNT/PMS and CNT/PDS (Figure S11). On the other hand, the nonradical mechanism in which

273

the CNTs effectively facilitated the transfer of electrons from organics to persulfates may

274

provide a plausible explanation for the substrate-specific reactivity of CNT/persulfate. The

275

mediated electron transfer should occur depending on how the electron flow from the organic

276

substrate to persulfate is favored. The observation that PMS achieved more rapid direct FFA

277

oxidation than PDS (Figure S13) clearly indicated that electrons were more preferentially

278

transferred from FFA to PMS than to PDS. This likely led to a much higher efficiency of

279

CNT/PMS for FFA degradation. In contrast, since 4-CP is susceptible to oxidative degradation

280

via direct electron transfer40, the electron delivery from 4-CP to either PMS or PDS appears to be

ACS Paragon Plus Environment

Page 12 of 36

Page 13 of 36

Environmental Science & Technology

281

thermodynamically plausible, which could render CNT/PMS and CNT/PDS with comparable 4-

282

CP degradation efficiencies.

283

Dependence of Reactivity on pH and Substrate Type. The substrate specificity of CNT/PMS

284

(in the dark) versus RB (under visible light irradiation) was compared (Figure 3). The systems

285

that produce identical reactive species are expected to exhibit similar substrate specificity during

286

oxidative degradation. The reactivity of the photoexcited RB varied considerably depending on

287

the type of organic compound (Figure 3b), which is consistent with the selective nature of 1O2.34,

288

41

289

degradation, whereas the other compounds, namely, benzoic acid, bisphenol A, and

290

carbamazepine, were highly persistent (Figure 3b). The oxidizing capacity of CNT/PMS was

291

also dependent on the type of substrate, but CNT/PMS caused significant degradation of all the

292

organics tested in this study, except benzoic acid (Figure 3a). In particular, organic compounds

293

that exhibited negligible or slow decomposition by photoexcited RB (e.g., bisphenol A,

294

carbamazepine, and propranolol) were effectively removed by aqueous CNT/PMS suspensions.

295

The substrate-specific reactivities of the CNT/PMS and photoexcited RB systems were clearly

296

distinguishable, suggesting that singlet oxygenation was minor (if present at all) in the

297

CNT/PMS system.

For example, cimetidine rapidly decomposed and propranolol underwent relatively slow

298

Three model substrates (4-CP, TCP, and pentachlorophenol (PCP)) were selected to

299

explore the effect of pH on the kinetic rates of chlorophenol oxidation by the CNT/PMS and

300

photoexcited RB systems (Figure 4). In general, alkaline conditions favor oxidative degradation

301

of phenolic compounds by 1O2 because phenolates, which are more electron-rich forms of

302

phenols that become predominant as the pH increases, are more susceptible to singlet

303

oxygenation than neutral phenols.42 The rate of phenol oxidation by 1O2 increases by two orders

ACS Paragon Plus Environment

Environmental Science & Technology

304

of magnitude at pH values above the pKa, with k(phenolate + 1O2) = 1.8 × 108 M−1 s−1 and

305

k(neutral phenol + 1O2) = 3.0 × 106 M−1 s−1.43 Consistent with the intrinsic reactivity of 1O2,34, 42,

306

43

307

solutions accelerated significantly with increasing pH up to or above the pKa (pKa(4-CP) =

308

9.41;44 pKa(TCP) = 6.23;44 pKa(PCP) = 4.7045). Fast 4-CP degradation proceeded in the visible-

309

light-irradiated RB solution at basic pH (pH = 11), whereas no removal of 4-CP was observed at

310

pH 7 or 4.5 (Figures 4b and S14b). However, photoexcited RB still mediated the rapid oxidation

311

of TCP at neutral pH, and the PCP decomposition efficiency was relatively constant, regardless

312

of the pH, i.e., k(PCP) = 0.0469 ± 0.0015 min−1 at pH 4.5, k(PCP) = 0.0671 ± 0.0056 min−1 at pH

313

7, and k(PCP) = 0.0711 ± 0.0025 min−1 at pH 11 (Figures 4b and S14b). In contrast, the pH

314

dependence of chlorophenol oxidation by activated PMS (Figures 4a and S14a) was different

315

from that of photosensitized chlorophenol oxidation by RB. All the tested chlorophenols were

316

significantly degraded in aqueous CNT/PMS suspensions under acidic and neutral pH conditions

317

(Figures 4a and S14a) (note that the scale of the y-axis in Figure 4a is 10 times greater than that

318

of Figure 4b). The rate of chlorophenol degradation by CNT/PMS at acidic pH was comparable

319

to or higher than the maximal rate observed in photoirradiated RB solution (k(4-CP at pH 4.5) =

320

0.159 ± 0.009 min−1 for CNT/PMS versus k(4-CP at pH 11) = 0.165 ± 0.006 min−1 for

321

photoexcited RB; k(TCP at pH 4.5) = 0.380 ± 0.017 min−1 for CNT/PMS versus k(TCP at pH 7)

322

= 0.111 ± 0.009 min−1 for photoexcited RB). Further, alkaline conditions in which phenolate

323

anions preferentially exist inhibited oxidative degradation of TCP and PCP by activated PMS.

324

These results also contradicted the possibility that heterogeneous PMS activation was

325

accompanied by singlet oxygenation.

Figure 4b demonstrates that the oxidation of chlorophenols in visible-light-irradiated RB

ACS Paragon Plus Environment

Page 14 of 36

Page 15 of 36

Environmental Science & Technology

326

Considering the possibility that the phenolate anion may have a stronger tendency to lose

327

electrons than neutral phenol, the oxidative degradation of chlorophenols could also be

328

accelerated in alkaline suspensions of CNT/PMS. However, the pH of the heterogeneous

329

persulfate activation system affects not only the interconversion between the phenolate anion and

330

the un-ionized phenol but also the surface charge of the carbocatalysts. The zeta potential

331

measured as a function of pH demonstrated that the CNT surface was negatively charged when

332

pH increased above ca. 6.2 (Figure S15), which implied that the alkaline conditions in which the

333

phenolate anion dominates the speciation caused electrostatic repulsions between the CNTs and

334

chlorophenol or PMS, kinetically hindering the electron transfer process in the ternary system.

335

EPR Study. PMS activation by CNTs in the presence of DMPO as a spin trapping agent led to

336

the formation of 5,5-dimethylpyrrolidone-2-(oxy)-(1) (DMPOX), a known product of direct

337

DMPO oxidation,46 which confirmed that SO4•− and hydroxyl radicals (•OH) were not involved

338

in the degradative mechanism (Figure S16). No EPR signals corresponding to DMPO adducts of

339

free radicals were observed for the photoilluminated RB solution (Figure S16). Signals

340

corresponding to 2,2,6,6-tetramethylpiperidine-N-oxyl (TEMPO), which is assignable to the

341

formation of a TEMP-1O2 adduct, have been demonstrated in the EPR spectra of persulfates in

342

the presence of activators (e.g., N-doped graphene,16,

343

provided a basis for suggesting singlet oxygenation as an alternative degradative route. We also

344

observed EPR spectral features corresponding to TEMPO generation in aqueous CNT/PMS

345

suspensions and photoirradiated RB solutions (Figure 5). However, D2O as a singlet oxygenation

346

enhancer affected the EPR spectral patterns in a different way. The peaks assigned to TEMPO

347

for CNT/PMS slightly decreased in the presence of D2O (Figure 5a), but the solvent exchange

348

caused a ca. 50% increase in the intensity of the EPR signal for photoexcited RB (Figure 5b).

17

CNTs,13 and Pd/g-C3N412), which

ACS Paragon Plus Environment

Environmental Science & Technology

349

This result implies that the TEMPO signal observed during PMS activation in the presence of

350

TEMP as a spin trap may not indicate the formation of 1O2. Nardi et al.47 suggested that EPR

351

detection of TEMPO may not be associated with 1O2 production; abstraction of one electron

352

from TEMP by the excited sensitizer results in the formation of the TEMP radical cation

353

(TEMP•+), which undergoes deprotonation and combination with dissolved oxygen to form

354

TEMPO. Therefore, TEMPO signals considered as evidence for singlet oxygenation during

355

persulfate activation could instead correspond to an electron-transfer mechanism, in which the

356

CNTs mediate electron transfer from TEMP to persulfate, leading to TEMP•+ generation.

357

Intermediate Distribution. In order to further clarify the difference in the degradative

358

mechanism between CNT/PMS (mediated electron transfer) and photoexcited RB (singlet

359

oxygenation), we compared the intermediate distribution from TCP oxidation by CNT/PMS and

360

fluorescent-light-irradiated RB. LC/MS analysis demonstrated a clear distinction in the

361

intermediate distribution between the two systems (Table S1). For instance, the main products

362

formed during the photosensitized singlet oxygenation of TCP included 1,2,3-trihydroxybenzene

363

2,6-dichloro-3-hydroxy-1,4-benzoquinone, which were not detectable over the course of TCP

364

decomposition by CNT/PMS. On the other hand, TCP oxidation in the aqueous CNT/PMS

365

suspension led to the formation of dihydroxydiphenyl ether and trichlorobenzene, which barely

366

formed when applying the photoirradiated RB for photochemical TCP degradation. The result

367

further confirmed that the degradative mechanism induced by CNT/PMS could be distinguished

368

from singlet oxygenation.

369

PMS Reduction and TCP Oxidation in Two Chambers Separated by a CNT Membrane. As

370

presented above, the empirical results supporting a role of 1O2 in persulfate activation, including

371

kinetic retardation in the presence of 1O2 quenchers and EPR spectral features characteristic of

ACS Paragon Plus Environment

Page 16 of 36

Page 17 of 36

Environmental Science & Technology

372

1

373

pathway. Further, the lack of a D2O enhancing effect and the incompatibility of the reactivity of

374

CNT-activated PMS with the pH-dependent and substrate-specific oxidizing capacity of 1O2

375

collectively implied that oxidative degradation during the heterogeneous activation of persulfate

376

was unlikely to involve 1O2. Thus, as an alternative nonradical degradative route, we examined

377

CNT-mediated electron transfer from organic substrates to PMS using a VA-CNT membrane

378

that physically partitioned the reaction system into two chambers containing PMS and TCP

379

(Figure 6a). Pilgrim et al.48 demonstrated electron exchange between photoexcited CdSe

380

quantum dots and methyl viologen located on opposite sides of a VA-CNT membrane. In this

381

system, photogenerated electrons were transported over hundreds of micrometers across the

382

CNT-based membrane. In contrast, as an extreme pressure (>120 bar) is required to allow water

383

entry into the inner pores of superhydrophobic virgin CNT membranes (with unmodified

384

entrances and exits),49 we found that VA-CNT membranes blocked the passage of water

385

molecules under ambient pressure (even with an external pressure of 5 bar, water did not pass

386

through the nanotube channels). As a VA-CNT membrane with high electric conductivity48, 50

387

and water impermeability is likely to reject organic/inorganic impurities and reactive oxygen

388

species (e.g., SO4•− and 1O2), concomitant TCP oxidation and PMS reduction in the physically

389

separated chambers (i.e., chambers A and B in Figure 6a) would be strong proof for an electron-

390

transfer mechanism, in which CNTs mediate electron transfer from organics to persulfates.

O2 formation, were not convincing evidence for singlet oxygenation as the nonradical reaction

391

When an aqueous chloride solution (1 M) and pure water were placed in the separated

392

compartments, no movement of Cl− from one side of the membrane (chamber A) to the other

393

side (chamber B) occurred as the conductivity of the pure water side did not change at all (Figure

394

S17). This result indicated that hydrophilic ions are unable to pass through the nonwettable pores

ACS Paragon Plus Environment

Environmental Science & Technology

395

of the VA-CNT membrane. Further, the PMS concentration decreased negligibly over 8 h when

396

Co2+ and PMS were placed on opposite sides of membrane, which confirmed that there was no

397

simultaneous transport of the ionic species across the CNT membrane (Figure S18). PMS

398

transport across the VA-CNT membrane would lead to a substantial increase in the PMS

399

concentration in chamber A, and the delivery of Co2+ to the other side would also cause

400

significant PMS reduction in chamber B. In contrast, when PMS and TCP were placed in the

401

separated chambers, noticeable decomposition was observed (Figure 6b). This result suggested

402

that electrons released from TCP on one side of the membrane (leading to TCP oxidation in

403

chamber A) crossed the conductive VA-CNT membrane (serving as an electron-transfer

404

mediator), eventually being accepted by PMS on the other side of the membrane (leading to PMS

405

reduction in chamber B). PMS was overconsumed for the observed TCP degradation efficiency

406

(~60%). When we repeated the experiment without the phosphate buffer, the performance with

407

respect to TCP treatment was almost unchanged, though PMS consumption was drastically

408

reduced (Figure S19). Considering ca. 10% of PMS was removable via sorption (Figure S18),

409

the observed PMS decay was insignificant. Therefore, the exposure to the phosphate buffer over

410

a relatively long reaction time (8 h) likely resulted in PMS being consumed in excessive

411

quantities (anions such as bicarbonate and phosphate added in high concentrations are capable of

412

direct PMS reduction51). On the other hand, PDS decay was minor when the TCP oxidation

413

associated with PDS activation (presented later) was performed even in the case of the buffered

414

solution (Figure S20), since PDS was unreactive toward anions present in excess

415

concentrations.51 Neither PMS nor TCP was detected on the opposite side of the membrane

416

(Figure 6b), which implied that interchamber transport did not contribute to the significant loss

417

of PMS or TCP in each chamber. TCP removal by sorption on the VA-CNT membrane was

ACS Paragon Plus Environment

Page 18 of 36

Page 19 of 36

Environmental Science & Technology

418

marginal (Figure S21). TCP degradation would not occur with PMS-derived oxidants (e.g.,

419

SO4•−, 1O2) (if any formed) owing to the impermeability of the VA-CNT membrane to water-

420

soluble species.

421

The distribution of the intermediates resulting from TCP oxidation in one chamber

422

physically separated from the other chamber in which PMS reduction concurrently occurred was

423

fairly similar to that observed when TCP was subjected to oxidation in the aqueous CNT/PMS

424

suspension (Table S1). The result confirmed that the organic oxidation associated with PMS

425

activation was achieved in the same manner irrespective of the type of CNTs used (i.e., an

426

aqueous CNT suspension versus VA-CNT membrane). Minor differences in the intermediate

427

distribution may rule out the possible role of the radical-induced reaction pathway in the CNT-

428

induced PMS activation process. If the surface-bound or free SO4•− derived from the PMS

429

molecules (if any) contributed to oxidative TCP decomposition, there would be a significant

430

distinction in the intermediate distribution between the two PMS activation systems: aqueous

431

CNT/PMS/TCP mixture and CNT and PMS physically separated via the VA-CNT membrane.

432

Note that no other (transient) chemical species (except electrons and protons) is transferable to

433

the other chamber across the VA-CNT membrane. TCP decomposition was noticeable, though it

434

was kinetically retarded when we used PDS alternatively (Figure S20), which supports our

435

hypothetic nonradical reaction pathway based on mediated electron transfer; CNT-induced

436

activation involving no radical formation is not unique to PMS and is achievable using proper

437

chemicals that can serve as electron acceptors.

438

Spectroscopic analysis suggested the possibility that PMS could form a complex on the

439

surface of graphitized nanodiamond10, and the resultant complex could effectively facilitate the

440

transfer of electrons from organic substrates to the PMS molecules involved in the surface

ACS Paragon Plus Environment

Environmental Science & Technology

441

complexation. More electronegative nitrogen atoms doped to CNTs could induce positive

442

charges on the adjacent carbons, which allowed the anionic PMS to form a reactive complex

443

with the N-doped CNTs.52 The surface complex was also suggested to initiate organic oxidation

444

by abstracting electrons from the organics through the conductive doped CNTs.52 The nonradical

445

mechanism is not far from our hypothetic mechanism, in which the CNTs likely mediate the

446

delivery of electrons from the organic compound to PMS, apart from the assumption that surface

447

complexation involving PMS would be a prerequisite for the electron exchange between the

448

organics and PMS. However, the formation of the reactive complex of PMS with CNTs is

449

unlikely to be indispensable to the electron transfer mechanism on account of the following

450

reasons. First, our previous work25 demonstrated that not only persulfate (PMS and PDS) but

451

also any oxyanions that serve as effective electron acceptors (e.g., periodate and percarboxylate)

452

could initiate oxidative organic degradation not reliant on the reactive radicals in the aqueous

453

CNT suspensions. Second, surface characterization of CNTs (after exposure to the PMS solution

454

for 1 h) using ATR-FTIR showed no occurrence of the spectral features assignable to the sulfur-

455

containing chemical moieties (Figure S22a); the infrared absorption peak at 1160 cm−1 is

456

attributed to the asymmetric stretching vibrations of C-S-C and S=O.53 Moreover, the

457

comparison of Raman spectra (Figure S22b) showed that the intensity of G-band, indicative of

458

graphitic carbon (1582 cm−1), did not change significantly after the use of CNTs in PMS

459

activation, implying no variation in defect density. Overall, the surface complexation of PMS can

460

promote the nonradical reaction pathway on carbonaceous activators, but will not contribute as

461

an essential step to PMS activation not relying on SO4•−.

462

Environmental Applications. In this study, we tested singlet oxygenation and mediated electron

463

transfer as hypothetical nonradical mechanisms underlying heterogeneous PMS activation. A

ACS Paragon Plus Environment

Page 20 of 36

Page 21 of 36

Environmental Science & Technology

464

comparison of CNT/PMS and photoirradiated RB showed that the reactivity of CNT/PMS

465

contradicted the intrinsic properties of 1O2 in terms of the effects of chemical reagents that

466

scavenge or enhance singlet oxygenation, substrate specificity, the pH dependence of

467

chlorophenol oxidation efficiency, and the effects of solvent exchange on EPR signal intensity.

468

However, simultaneous PMS reduction and TCP oxidation were demonstrated to occur when

469

these compounds were placed in two chambers physically separated by a VA-CNT membrane

470

that prevented interchamber transport of chemical species (e.g., water, hydrophilic anions, and

471

reactive oxygen species) but allowed electron conduction. Collectively, these results implied that

472

electron transfer from organics to persulfate, effectively facilitated by nanocarbon activators, was

473

primarily responsible for persulfate activation not involving reactive radicals. Here, we

474

demonstrated that PMS was superior to PDS in terms of the organic oxidation associated with

475

persulfate activation (i.e., FFA oxidation in aqueous CNT suspensions; TCP decomposition in

476

the two-chamber system). The observation was consistent with the previous finding that CNTs

477

achieved a more rapid destruction of a variety of organics in the presence of PMS rather than

478

PDS.25 Together with the general recognition that metal-induced activation results in more

479

effective SO4•− production from PMS than PDS,8, 54 the results justified the use of PMS in the

480

persulfate activation processes, even though the choice of PDS is more economically feasible.

481

The persulfate activation process utilizing SO4•− as the main oxidant (e.g., Co2+/PMS)

482

enables effective the oxidative degradation and mineralization of a wide range of organic

483

pollutants, whereas nonradical persulfate activation leads to substrate-specific oxidation. On the

484

other hand, the selective nature of the nonradical degradation pathway (typically induced by

485

carbocatalysts) has a competitive advantage over the radical-induced pathway on account of the

486

following reasons. First, a substrate-dependent oxidizing capacity allows the persulfate activation

ACS Paragon Plus Environment

Environmental Science & Technology

487

system to better target the priority pollutants present at trace levels in the complicated water

488

matrix; the majority of non-selective radicals (e.g., SO4•−, •OH) would be fruitlessly consumed

489

through the reactions with background organic and inorganic substrates. Second, since halide

490

ions are highly susceptible to one-electron oxidation by SO4•−, it is probable that the oxidative

491

treatment by SO4•− in the presence of bromide ions is inevitably accompanied by the high-yield

492

production of bromate.55 In contrast, persulfate activation based on the mediated electron transfer

493

mechanism would not result in bromate production from bromide.56 Finally, the nonradical

494

persulfate activation takes place in the co-presence of the organic pollutant (electron donor),

495

persulfate (electron acceptor), and activator (electron-transfer mediator), which likely minimizes

496

persulfate consumption; persulfate barely decomposes once the pollutant concentration is

497

significantly reduced. On the other hand, since the SO4•− in the heterogeneous persulfate

498

activation process is generated through the one-electron reduction of persulfate by the activators,

499

persulfate continues to be degraded until the residual PMS concentration reduces to virtually

500

zero.

501

The oxidative degradation of organics by CNT/PMS not reliant on SO4•− was suggested

502

to result from the CNT-induced electron exchange between the organic pollutant and PMS.

503

Carbon-based activators can kinetically enhance the transfer of electrons from organic substrates

504

to persulfates only when the electron flow is thermodynamically favored, which implies that the

505

thermodynamics of the electron transfer process is likely responsible for the selective nature of

506

the nonradical degradation pathway induced by carbocatalysts. As a result, a comparison of the

507

redox potentials of the target contaminant and persulfate will allow us to explore the treatability

508

of organic pollutants in the persulfate activation processes based on a mechanism that is not

509

radical-induced. Considering the key role of activators in the nonradical mechanism, the

ACS Paragon Plus Environment

Page 22 of 36

Page 23 of 36

Environmental Science & Technology

510

properties of the carbonaceous activators (e.g., electric conductivity, surface charge, and surface

511

affinity) that affect their capability of electron transfer mediation should be considered while

512

evaluating the performance of nanocarbon materials in heterogeneous persulfate activation. The

513

persulfate activation capacity of carbonaceous materials that can initiate oxidative degradation

514

without oxidizing radicals could possibly be improved in two ways: i) improving the electric

515

conductivity and ii) increasing the surface affinity toward organic contaminants (i.e., electron

516

donor) and persulfate (i.e., electron acceptor). Accordingly, possible strategies for developing

517

high-performance nanocarbon activators include doping with heteroatoms, combining with

518

metals/metal oxides, and surface modification with chemical moieties that interact

519

electrostatically with persulfate or organics.

520

Acknowledgements

521

This study was supported by a National Research Foundation of Korea grant funded by the

522

Korean Government (No. 2017R1A2B4002235) and a grant from the National Research

523

Foundation of Korea funded by the Ministry of Science, ICT, and Future Planning (No.

524

2016M3A7B4909318).

525

Supporting Information Available

526

Description of chemicals used in this study (Text S1), VA-CNT membrane fabrication (Scheme

527

S1), SEM images and Raman spectrum of aligned CNTs (Figures S1 and S2), XPS survey

528

spectrum of VA-CNT membrane (Figure S3), emission spectrum of fluorescent lamp and

529

absorption spectrum of RB (Figure S4), PMS decay by CNTs in H2O and D2O (Figure S5), H2O2

530

production during ascorbate oxidation by CNT/PMS and photoexcited RB (Figure S6), FFA

531

oxidation by PMS with excess L-histidine (Figure S7), reactivity of azide/PMS towards organics

532

(Figure S8), EPR spectra of azide only, PMS only, and azide/PMS with DMPO as a spin trap

ACS Paragon Plus Environment

Environmental Science & Technology

533

(Figure S9), FFA degradation by CNT/PMS and CNT/PDS (Figure S10), effects of L-histidine

534

and D2O on 4-CP oxidation by CNT/PMS and CNT/PDS (Figure S11), direct reduction of PMS

535

and PDS by L-histidine (Figure S12), direct FFA oxidation by PMS and PDS (Figure S13), effect

536

of pH on chlorophenol degradation by CNT/PMS and photoexcited RB (Figure S14), zeta

537

potential of CNTs (Figure S15), EPR spectra of CNT/PMS and photoirradiated RB with DMPO

538

as a spin trap (Figure S16), chloride permeability of VA-CNT membrane (Figure S17), time-

539

dependent changes in PMS concentration with Co2+ and PMS in chambers A and B (Figure S18),

540

concurrent TCP oxidation (chamber A) and PMS reduction (chamber B) without a phosphate

541

buffer (Figure S19), concurrent TCP oxidation (chamber A) and PDS reduction (chamber B)

542

(Figure S20), time-dependent changes in TCP concentration with aqueous TCP and pure water in

543

chambers A and B (Figure S21), and ATR-FTIR and Raman spectra of fresh and used CNTs

544

(Figure S22). This information is available free of charge via the Internet at http://pubs.acs.org/.

545 546

Literature Cited

547

1. Waclawek, S.; Lutze, H. V.; Grubel, K.; Padil, V. V. T.; Cernik, M.; Dionysiou, D. D.,

548

Chemistry of persulfates in water and wastewater treatment: A review. Chem. Eng. J. 2017, 330,

549

44-62.

550

2. Anipsitakis, G. P.; Dionysiou, D. D., Degradation of organic contaminants in water with

551

sulfate radicals generated by the conjunction of peroxymonosulfate with cobalt. Environ. Sci.

552

Technol. 2003, 37, (20), 4790-4797.

553

3. Oh, W. D.; Dong, Z. L.; Lim, T. T., Generation of sulfate radical through heterogeneous

554

catalysis for organic contaminants removal: Current development, challenges and prospects.

555

Appl. Catal. B: Environ. 2016, 194, 169-201.

556

4. Chen, X. Y.; Wang, W. P.; Xiao, H.; Hong, C. L.; Zhu, F. X.; Yao, Y. L.; Xue, Z. Y.,

557

Accelerated TiO2 photocatalytic degradation of Acid Orange 7 under visible light mediated by

558

peroxymonosulfate. Chem. Eng. J. 2012, 193, 290-295.

ACS Paragon Plus Environment

Page 24 of 36

Page 25 of 36

Environmental Science & Technology

559

5. Kim, H.; Yoo, H. Y.; Hong, S.; Lee, S.; Lee, S.; Park, B. S.; Park, H.; Lee, C.; Lee, J., Effects

560

of inorganic oxidants on kinetics and mechanisms of WO3-mediated photocatalytic degradation.

561

Appl. Catal. B: Environ. 2015, 162, 515-523.

562

6. Chen, W. S.; Huang, C. P., Mineralization of aniline in aqueous solution by electrochemical

563

activation of persulfate. Chemosphere 2015, 125, 175-181.

564

7. Roshani, B.; Leitner, N. K. V., Effect of persulfate on the oxidation of benzotriazole and

565

humic acid by e-beam irradiation. J. Hazard. Mater. 2011, 190, (1-3), 403-408.

566

8. Ahn, Y. Y.; Yun, E. T.; Seo, J. W.; Lee, C.; Kim, S. H.; Kim, J. H.; Lee, J., Activation of

567

peroxymonosulfate by surface-loaded noble metal nanoparticles for oxidative degradation of

568

organic compounds. Environ. Sci. Technol. 2016, 50, (18), 10187-10197.

569

9. Lee, H.; Lee, H. J.; Jeong, J.; Lee, J.; Park, N. B.; Lee, C., Activation of persulfates by carbon

570

nanotubes: Oxidation of organic compounds by nonradical mechanism. Chem. Eng. J. 2015, 266,

571

28-33.

572

10. Lee, H.; Kim, H. I.; Weon, S.; Choi, W.; Hwang, Y. S.; Seo, J.; Lee, C.; Kim, J. H.,

573

Activation of persulfates by graphitized nanodiamonds for removal of organic compounds.

574

Environ. Sci. Technol. 2016, 50, (18), 10134-10142.

575

11. Zhang, T.; Chen, Y.; Wang, Y. R.; Le Roux, J.; Yang, Y.; Croue, J. P., Efficient

576

peroxydisulfate activation process not relying on sulfate radical generation for water pollutant

577

degradation. Environ. Sci. Technol. 2014, 48, (10), 5868-5875.

578

12. Wang, Y. B.; Cao, D.; Liu, M.; Zhao, X., Insights into heterogeneous catalytic activation of

579

peroxymonosulfate by Pd/g-C3N4: The role of superoxide radical and singlet oxygen. Catal.

580

Commun. 2017, 102, 85-88.

581

13. Cheng, X.; Guo, H. G.; Zhang, Y. L.; Wu, X.; Liu, Y., Non-photochemical production of

582

singlet oxygen via activation of persulfate by carbon nanotubes. Water Res. 2017, 113, 80-88.

583

14. Guan, C. T.; Jiang, J.; Pang, S. Y.; Luo, C. W.; Ma, J.; Zhou, Y.; Yang, Y., Oxidation

584

kinetics of bromophenols by nonradical activation of peroxydisulfate in the presence of carbon

585

nanotube and formation of brominated polymeric products. Environ. Sci. Technol. 2017, 51,

586

(18), 10718-10728.

587

15. Cheng, X.; Guo, H. G.; Zhang, Y. L.; Liu, Y.; Liu, H. W.; Yang, Y., Oxidation of 2,4-

588

dichlorophenol by non-radical mechanism using persulfate activated by Fe/S modified carbon

589

nanotubes. J. Colloid Interface Sci. 2016, 469, 277-286.

ACS Paragon Plus Environment

Environmental Science & Technology

590

16. Liang, P.; Zhang, C.; Duan, X. G.; Sun, H. Q.; Liu, S. M.; Tade, M. O.; Wang, S. B., An

591

insight into metal organic framework derived N-doped graphene for the oxidative degradation of

592

persistent contaminants: Formation mechanism and generation of singlet oxygen from

593

peroxymonosulfate. Environ. Sci. Nano 2017, 4, (2), 315-324.

594

17. Liang, P.; Zhang, C.; Duan, X. G.; Sun, H. Q.; Liu, S. M.; Tade, M. O.; Wang, S. B., N-

595

doped graphene from metal-organic frameworks for catalytic oxidation of p-hydroxylbenzoic

596

acid: N-functionality and mechanism. ACS Sustainable Chem. Eng. 2017, 5, (3), 2693-2701.

597

18. Tian, X. K.; Gao, P. P.; Nie, Y. L.; Yang, C.; Zhou, Z. X.; Li, Y.; Wang, Y. X., A novel

598

singlet oxygen involved peroxymonosulfate activation mechanism for degradation of ofloxacin

599

and phenol in water. Chem. Commun. 2017, 53, (49), 6589-6592.

600

19. Dai, D. J.; Yang, Z. Y.; Yao, Y. Y.; Chen, L. K.; Jia, G. S.; Luo, L. S., Highly efficient

601

removal of organic contaminants based on peroxymonosulfate activation by iron phthalocyanine:

602

Mechanism and the bicarbonate ion enhancement effect. Catal. Sci. Technol. 2017, 7, (4), 934-

603

942.

604

20. Li, D. G.; Duan, X. G.; Sun, H. Q.; Kang, J.; Zhang, H. Y.; Tade, M. O.; Wang, S. B., Facile

605

synthesis of nitrogen-doped graphene via low-temperature pyrolysis: The effects of precursors

606

and annealing ambience on metal-free catalytic oxidation. Carbon 2017, 115, 649-658.

607

21. Qi, C. D.; Liu, X. T.; Ma, J.; Lin, C. Y.; Li, X. W.; Zhang, H. J., Activation of

608

peroxymonosulfate by base: Implications for the degradation of organic pollutants. Chemosphere

609

2016, 151, 280-288.

610

22. Zhou, Y.; Jiang, J.; Gao, Y.; Ma, J.; Pang, S. Y.; Li, J.; Lu, X. T.; Yuan, L. P., Activation of

611

peroxymonosulfate by benzoquinone: A novel nonradical oxidation process. Environ. Sci.

612

Technol. 2015, 49, (21), 12941-12950.

613

23. Zhou, Y.; Jiang, J.; Gao, Y.; Pang, S. Y.; Yang, Y.; Ma, J.; Gu, J.; Li, J.; Wang, Z.; Wang, L.

614

H.; Yuan, L. P.; Yang, Y., Activation of peroxymonosulfate by phenols: Important role of

615

quinone intermediates and involvement of singlet oxygen. Water Res. 2017, 125, 209-218.

616

24. Lange, A.; Brauer, H. D., On the formation of dioxiranes and of singlet oxygen by the

617

ketone-catalysed decomposition of Caro's acid. J. Chem. Soc. Perkin Trans. 2 1996, (5), 805-

618

811.

ACS Paragon Plus Environment

Page 26 of 36

Page 27 of 36

Environmental Science & Technology

619

25. Yun, E. T.; Yoo, H. Y.; Bae, H.; Kim, H. I.; Lee, J., Exploring the role of persulfate in the

620

activation process: Radical precursor versus electron acceptor. Environ. Sci. Technol. 2017, 51,

621

(17), 10090-10099.

622

26. Duan, X. G.; Sun, H. Q.; Wang, Y. X.; Kang, J.; Wang, S. B., N-doping-induced nonradical

623

reaction on single-walled carbon nanotubes for catalytic phenol oxidation. ACS Catal. 2015, 5,

624

(2), 553-559.

625

27. Haag, W. R.; Hoigne, J., Singlet oxygen in surface waters .3. Photochemical formation and

626

steady-state concentrations in various types of waters. Environ. Sci. Technol. 1986, 20, (4), 341-

627

348.

628

28. Lee, K. J.; Park, H. D., The most densified vertically-aligned carbon nanotube membranes

629

and their normalized water permeability and high pressure durability. J. Membrane Sci. 2016,

630

501, 144-151.

631

29. Dresselhaus, M. S.; Jorio, A.; Hofmann, M.; Dresselhaus, G.; Saito, R., Perspectives on

632

carbon nanotubes and graphene Raman spectroscopy. Nano Lett. 2010, 10, (3), 751-758.

633

30. Du, F.; Qu, L. T.; Xia, Z. H.; Feng, L. F.; Dai, L. M., Membranes of vertically aligned

634

superlong carbon nanotubes. Langmuir 2011, 27, (13), 8437-8443.

635

31. Liang, C. J.; Huang, C. F.; Mohanty, N.; Kurakalva, R. M., A rapid spectrophotometric

636

determination of persulfate anion in ISCO. Chemosphere 2008, 73, (9), 1540-1543.

637

32. Gorman, A. A.; Rodgers, M. A. J., Singlet molecular oxygen. Chem. Soc. Rev. 1981, 10, (2),

638

205-231.

639

33. Kramarenko, G. G.; Hummel, S. G.; Martin, S. M.; Buettner, G. R., Ascorbate reacts with

640

singlet oxygen to produce hydrogen peroxide. Photochem. Photobiol. 2006, 82, (6), 1634-1637.

641

34. Lee, J.; Hong, S.; Mackeyev, Y.; Lee, C.; Chung, E.; Wilson, L. J.; Kim, J. H.; Alvarez, P. J.

642

J., Photosensitized oxidation of emerging organic pollutants by tetrakis C60 aminofullerene-

643

derivatized silica under visible light irradiation. Environ. Sci. Technol. 2011, 45, (24), 10598-

644

10604.

645

35. Thompson, R. C.; Wieland, P.; Appelman, E. H., Oxidation of azide and

646

azidopentaaminechromium(III) by peroxymonosulfate in aqueous solution. Inorg. Chem. 1979,

647

18, (7), 1974-1977.

648

36. Alfassi, Z. B.; Schuler, R. H., Reaction of azide radicals with aromatic compounds. Azide as

649

a selective oxidant. J. Phys. Chem. 1985, 89, (15), 3359-3363.

ACS Paragon Plus Environment

Environmental Science & Technology

650

37. Kalyanaraman, B.; Janzen, E. G.; Mason, R. P., Spin trapping of the azidyl radical in

651

azide/catalase/H2O2 and various azide/peroxidase/H2O2 peroxidizing systems. J. Biol. Chem.

652

1985, 260, (7), 4003-4006.

653

38. Ram, M. S.; Stanbury, D. M., Electron-transfer reactions involving the azidyl radical. J.

654

Phys. Chem. 1986, 90, (16), 3691-3696.

655

39. Yang, Y.; Banerjee, G.; Brudvig, G. W.; Kim, J.-H.; Pignatello, J. J., Oxidation of organic

656

compounds in water by unactivated peroxymonosulfate. Environ. Sci. Technol. 2018, 52, 5911-

657

5919.

658

40. Kim, W.; Park, J.; Jo, H. J.; Kim, H. J.; Choi, W., Visible light photocatalysts based on

659

homogeneous and heterogenized tin porphyrins. J. Phys. Chem. C 2008, 112, (2), 491-499.

660

41. Kim, H.; Kim, W.; Mackeyev, Y.; Lee, G. S.; Kim, H. J.; Tachikawa, T.; Hong, S.; Lee, S.;

661

Kim, J.; Wilson, L. J.; Majima, T.; Alvarez, P. J. J.; Choi, W.; Lee, J., Selective oxidative

662

degradation of organic pollutants by singlet oxygen-mediated photosensitization: Tin porphyrin

663

versus C60 aminofullerene systems. Environ. Sci. Technol. 2012, 46, (17), 9606-9613.

664

42. Scully, F. E.; Hoigne, J., Rate constants for reactions of singlet oxygen with phenols and

665

other compounds in water. Chemosphere 1987, 16, (4), 681-694.

666

43. Wilkinson, F.; Helman, W. P.; Ross, A. B., Rate constants for the decay and reactions of the

667

lowest electronically excited singlet-state of molecular oxygen in solution - An expanded and

668

revised compilation. J. Phys. Chem. Ref. Data 1995, 24, (2), 663-1021.

669

44. Serjeant, E. P.; Dempsey, B., Ionisation Constants of Organic Acids in Aqueous Solution.

670

Pergamon: Oxford, U.K., 1979.

671

45. Cessna, A. J.; Grover, R., Spectrophotometric determination of dissociation constants of

672

selected acidic herbicides. J. Agric. Food Chem. 1978, 26, 289-292.

673

46. Wang, Y. X.; Sun, H. Q.; Ang, H. M.; Tade, M. O.; Wang, S. B., 3D-hierarchically

674

structured MnO2 for catalytic oxidation of phenol solutions by activation of peroxymonosulfate:

675

Structure dependence and mechanism. Appl. Catal. B: Environ. 2015, 164, 159-167.

676

47. Nardi, G.; Manet, I.; Monti, S.; Miranda, M. A.; Lhiaubet-Vallet, V., Scope and limitations

677

of the TEMPO/EPR method for singlet oxygen detection: The misleading role of electron

678

transfer. Free Radical Biol. Med. 2014, 77, 64-70.

ACS Paragon Plus Environment

Page 28 of 36

Page 29 of 36

Environmental Science & Technology

679

48. Pilgrim, G. A.; Amori, A. R.; Hou, Z. T.; Qiu, F.; Lampa-Pastirk, S.; Krauss, T. D., Carbon

680

nanotube-based membrane for light-driven, simultaneous proton and electron transport. ACS

681

Energy Lett. 2017, 2, (1), 129-133.

682

49. Walther, J. H.; Ritos, K.; Cruz-Chu, E. R.; Megaridis, C. M.; Koumoutsakos, P., Barriers to

683

superfast water transport in carbon nanotube membranes. Nano Lett. 2013, 13, (5), 1910-1914.

684

50. Pilgrim, G. A.; Leadbetter, J. W.; Qiu, F.; Siitonen, A. J.; Pilgrim, S. M.; Krauss, T. D.,

685

Electron conductive and proton permeable vertically aligned carbon nanotube membranes. Nano

686

Lett. 2014, 14, (4), 1728-1733.

687

51. Yang, S. Y.; Wang, P.; Yang, X.; Shan, L.; Zhang, W. Y.; Shao, X. T.; Niu, R., Degradation

688

efficiencies of azo dye Acid Orange 7 by the interaction of heat, UV and anions with common

689

oxidants: Persulfate, peroxymonosulfate and hydrogen peroxide. J. Hazard. Mater. 2010, 179,

690

(1-3), 552-558.

691

52. Duan, X. G.; Sun, H. Q.; Shao, Z. P.; Wang, S. B., Nonradical reactions in environmental

692

remediation processes: Uncertainty and challenges. Appl. Catal. B: Environ. 2018, 224, 973-982.

693

53. Chen, L. S.; Cui, X. Z.; Wang, Y. X.; Wang, M.; Qiu, R. H.; Shu, Z.; Zhang, L. X.; Hua, Z.

694

L.; Cui, F. M.; Weia, C. Y.; Shi, J. L., One-step synthesis of sulfur doped graphene foam for

695

oxygen reduction reactions. Dalton Transact. 2014, 43, (9), 3420-3423.

696

54. Anipsitakis, G. P.; Dionysiou, D. D., Radical generation by the interaction of transition

697

metals with common oxidants. Environ. Sci. Technol. 2004, 38, (13), 3705-3712.

698

55. Fang, J. Y.; Shang, C., Bromate formation from bromide oxidation by the UV/persulfate

699

process. Environ. Sci. Technol. 2012, 46, (16), 8976-8983.

700

56. Yun, E.-T.; Moon, G.-H.; Lee, H.; Jeon, T. H.; Lee, C.; Choi, W.; Lee, J., Oxidation of

701

organic pollutants by peroxymonosulfate activated with low-temperature-modified

702

nanodiamonds: Understanding the reaction kinetics and mechanism. Appl. Catal. B: Environ. In

703

press.

704

ACS Paragon Plus Environment

Environmental Science & Technology

(a)

CNT/PMS CNT/PMS w / L-histidine CNT/PMS w / azide

1.0

CNT/PMS w / D2O

0.8 FFA Conc. (C/C0 )

Page 30 of 36

CNT only PMS only 0.6

0.4

0.2

0.0 0

10

20

30

40

50

60

Reaction Time (min) (b)

RB RB w / L-histidine RB w / azide RB w / D2 O

1.0

FFA Conc. (C/C0)

0.8

0.6

0.4

0.2

0.0 0

705

10

20

30

40

50

60

Fluorescent Light Irradiation Time (min)

706

FIGURE 1. Degradation of furfuryl alcohol (FFA) by (a) CNT/PMS (in the dark) and (b) RB

707

(under photoirradiation) in the absence and presence of L-histidine, azide, and D2O ([CNT]0 =

708

0.1 g L−1; [RB]0 = 0.05 mM; [PMS]0 = 1 mM; [FFA]0 = 0.05 mM; [L-histidine]0 = [azide]0 = 100

709

mM; [phosphate buffer]0 = 1 mM; pHi = 7.0).

ACS Paragon Plus Environment

Page 31 of 36

Environmental Science & Technology

(a)

1.0

5 mM L-histidine 10 mM L-histidine 25 mM L-histidine 100 mM L-histidine

PMS Conc. (C/C0 )

0.8

0.6

0.4

0.2

0.0 0

10

20

30

40

50

60

(b) 1.0 FFA Conc. (C/C0)

1.0

PMS Conc. (C/C0 )

0.8

0.6

Azide only PMS/5 mM azide PMS/25 mM azide PMS/100 mM azide CNT /PMS/100 mM azide

0.8

0.6

5 mM azide 10 mM azide 25 mM azide 100 mM azide

0.4

0.2

0.0

0.4

0

10

20

30

40

50

60

Reaction Time (min)

0.2

0.0 0

10

20

30

40

50

60

Reaction Time (min)

710 711

FIGURE 2. Effects of initial concentrations of (a)

712

decomposition ([PMS]0 = 1 mM; [furfuryl alcohol (FFA)]0 = 0.05 mM; [phosphate buffer]0 = 1

713

mM; pHi = 7.0). Inset: FFA degradation during PMS activation with increasing concentrations of

714

azide.

L-histidine

715

ACS Paragon Plus Environment

and (b) azide on PMS

Environmental Science & Technology

(a)

Benzoic acid

1.0 Organic Compound Conc. (C/C0 )

Page 32 of 36

Bisphenol A Carbamazepine Cimetidine

0.8

Propranolol 0.6

0.4

0.2

0.0 0

10

20

30

40

50

60

Reaction Time (min) (b)

Benzoic acid Bisphenol A

Organic Compound Conc. (C/C0)

1.0

Carbamazepine Cimetidine

0.8

Propranolol

0.6

0.4

0.2

0.0 0

716

10

20

30

40

50

60

Fluorescent Light Irradiation Time (min)

717

FIGURE 3. Degradation of various organic compounds by (a) CNT/PMS (in the dark) and (b)

718

RB (under photoirradiation) ([CNT]0 = 0.1 g L−1; [RB]0 = 0.05 mM; [PMS]0 = 1 mM; [benzoic

719

acid]0 = [bisphenol A]0 = [cimetidine]0 = [propranolol]0 = [carbamazepine]0 = 0.05 mM;

720

[phosphate buffer]0 = 1 mM; pHi = 7.0).

ACS Paragon Plus Environment

Page 33 of 36

Environmental Science & Technology

Pseudo-First-Order Rate Constant (min-1 )

2.0

(a)

pH 4.5 pH 7 pH 11

1.5

1.0

0.5

0.0 4-CP

TCP

PCP

0.20 Pseudo-First-Order Rate Constant (min-1 )

(b)

pH 4.5 pH 7 pH 11

0.15

0.10

0.05

0.00 4-CP

TCP

PCP

721 722

FIGURE 4. Pseudo-first-order rate constants for chlorophenol degradation by (a) CNT/PMS (in

723

the dark) and (b) RB (under photoirradiation) under various pH conditions ([CNT]0 = 0.1 g L−1;

724

[RB]0 = 0.05 mM; [PMS]0 = 1 mM; [4-chlorophenol (4-CP)]0 = [trichlorophenol (TCP)]0 = 0.05

725

mM; [pentachlorophenol (PCP)]0 = 0.04 mM; [phosphate buffer]0 = [carbonate buffer]0 = 1 mM).

ACS Paragon Plus Environment

Environmental Science & Technology

(a)

Page 34 of 36

CNT only PMS only CNT/PMS (in H2O)

Intensity (Arb. Unit)

CNT/PMS (in D2O)

330.4

330.6

330.8

331.0

331.2

(b)

RB (in H2O)

Intensity (Arb. Unit)

RB (in D2O)

330.6

330.7

330.8

330.9

331.0

331.1

Magnetic Field (mT)

726 727

FIGURE 5. EPR spectra recorded in H2O- and D2O-based (a) CNT/PMS suspensions (in the

728

dark) and (b) RB solutions (under photoirradiation) ([CNT]0 = 0.1 g L−1; [RB]0 = 0.05 mM;

729

[PMS]0 = 1 mM; [TEMP]0 = 1 mM; [phosphate buffer]0 = 1 mM; pHi = 7.0). TEMP was used as

730

a spin trap.

731 732

ACS Paragon Plus Environment

Page 35 of 36

Environmental Science & Technology

733 (a)

734 735 736 737 738 739 740 741 1.5

(b)

1.0

TCP (chamber A) TCP (chamber B) PMS (chamber A) PMS (chamber B)

1.2

0.9

0.6

0.6

0.4

0.3

0.2

0.0 0

742

PMS Conc. (mM)

TCP Conc. (C/C0 )

0.8

100

200

300

400

0.0 500

Reaction Time (min)

743

FIGURE 6. (a) Experimental set-up for PMS activation in the reaction system partitioned into

744

two chambers by a vertically aligned CNT membrane and (b) simultaneous trichlorophenol (TCP)

745

oxidation (chamber A) and PMS reduction (chamber B) ([PMS]0 = 1.5 mM; [TCP]0 = 0.01 mM;

746

[phosphate buffer]0 = 1 mM; pHi = 7.0).

ACS Paragon Plus Environment

Environmental Science & Technology

747

Table of Contents Figure:

748 749 750 751 752 753

ACS Paragon Plus Environment

Page 36 of 36