Plasma-Based Water Treatment - ACS Publications - American

Jan 12, 2017 - Southern Nevada Water Authority, Water Quality Research and Development Division, Henderson, Nevada 89015, United States...
0 downloads 0 Views 801KB Size
Subscriber access provided by University of Newcastle, Australia

Article

Plasma-based water treatment: Efficient transformation of perfluoroalkyl substances (PFASs) in prepared solutions and contaminated groundwater Gunnar R Stratton, Fei Dai, Christopher L Bellona, Thomas M. Holsen, Eric Reyvell Velazquez Dickenson, and Selma Mededovic Thagard Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b04215 • Publication Date (Web): 12 Jan 2017 Downloaded from http://pubs.acs.org on January 13, 2017

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 free 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 accessible to all readers and 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.

Environmental Science & Technology 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 19

Environmental Science & Technology

1

Plasma-based water treatment: Efficient transformation of perfluoroalkyl substances

2

(PFASs) in prepared solutions and contaminated groundwater

3 4

Gunnar R. Stratton1, Fei Dai2, Christopher L. Bellona3, Thomas M. Holsen2, Eric R. V.

5

Dickenson4, Selma Mededovic Thagard1,*

6

1

7

Engineering, Potsdam, NY 13699 USA

8

2

9

USA

Clarkson University, Plasma Research Laboratory, Department of Chemical and Biomolecular

Clarkson University, Department of Civil and Environmental Engineering, Potsdam, NY 13699

10

3

11

80401 USA

12

4

13

Henderson, NV 89015 USA

Colorado School of Mines, Department of Civil and Environmental Engineering, Golden, CO

Southern Nevada Water Authority, Water Quality Research and Development Division,

14 15

* Corresponding author. Tel.: +1 315 2684423; fax: +1 315 2686654.

16

E-mail address: [email protected] (S. Mededovic Thagard)

17 18 19

Abstract: A process based on electrical discharge plasma was tested for the transformation of

20

perfluorooctanoic acid (PFOA). The plasma-based process was adapted for two cases, high

21

removal rate and high removal efficiency. During a 30 minute treatment, the PFOA

22

concentration in 1.4 L aqueous solutions was reduced by 90% with the high rate process (76.5 W

23

input power) and 25% with the high efficiency process (4.1 W input power). Both achieved

24

remarkably high PFOA removal and defluorination efficiencies compared to leading alternative

25

technologies. The high efficiency process was also used to treat groundwater containing PFOA

26

and several co-contaminants including perfluorooctane sulfonate (PFOS), demonstrating that the

27

process was not significantly affected by co-contaminants and that the process was capable of

28

rapidly degrading PFOS. Preliminary investigation into the byproducts showed that only about

29

10% of PFOA and PFOS is converted into shorter-chain perfluoroalkyl acids (PFAAs).

30

Investigation into the types of reactive species involved in primary reactions with PFOA showed

31

that hydroxyl and superoxide radicals, which are typically the primary plasma-derived reactive 1 ACS Paragon Plus Environment

Environmental Science & Technology

32

species, play no significant role. Instead, scavenger experiments indicated that aqueous electrons

33

account for a sizable fraction of the transformation, with free electrons and/or argon ions

34

proposed to account for the remainder.

2 ACS Paragon Plus Environment

Page 2 of 19

Page 3 of 19

35

Environmental Science & Technology

Introduction

36

There is considerable and growing concern over perfluoroalkyl substances (PFASs) due to

37

their ubiquitous presence and recalcitrance in the environment, and toxicity in humans and

38

wildlife.1-5 Manufacture, disposal and use of formulations and products containing PFASs or

39

PFAS-precursors (e.g., aqueous film-forming foams) has resulted in PFAS contamination of

40

groundwater and drinking water supplies.3,

41

acid (PFOA) and perfluorooctane sulfonate (PFOS) is problematic due to their particularly high

42

prevalence, toxicity and resistance to transformation, which has prompted the U. S. EPA to issue

43

health advisories for both compounds.8-9

6-7

In particular, the presence of perfluorooctanoic

44

Conventional water treatment processes are not effective for the removal of perfluoroalkyl

45

acids (PFAAs).10-12 Past research has also demonstrated that commonly used advanced oxidation

46

processes (AOPs) such as ultraviolet light (UV) or ozone (O3) with hydrogen peroxide (H2O2)

47

are not effective for PFAA transformation, due to the stability of the carbon-fluorine bond.12-13

48

While some success has been found for PFAA transformation using alternative processes, such

49

as sonolysis, activated persulfate and electrolysis, these typically involve significant chemical

50

and/or energy additions for decomposition reactions to proceed.13-17 As a result, researchers and

51

practitioners have focused on the use of adsorbents such as activated carbon, and to a lesser

52

extent, reverse osmosis for the treatment of PFAS-contaminated water.12,

53

relatively short breakthrough times have been reported for activated carbon for shorter-chain

54

PFAAs, and both processes produce a residual requiring disposal or further treatment.12, 18

18-20

However,

55

For this work, plasma-based water treatment (PWT) was evaluated for degrading PFOA,

56

which served as the model PFAA. Similar to other AOPs, PWT makes use of the highly

57

oxidative radicals to oxidize chemical contaminants. Unlike other AOPs however, PWT involves

58

the generation of radicals in situ and does not require significant chemical inputs. Additionally,

59

plasma is capable of producing a broad range of reactive species (OH, O, H, O3, H2O2, eaq),

60

including strong oxidants and reductants.21,22 Previous attempts have been made to utilize plasma

61

to degrade PFASs; however, these involved the use of inefficient reactor types and DC

62

discharges, which are less efficient than the pulsed discharges used in this study.23-27

63

The primary objective of this study was to evaluate the efficacy of PWT for PFOA

64

transformation and defluorination compared to the leading alternative technologies by operating

65

the PWT process under two sets of parameters. The first was designed to target high removal rate 3 ACS Paragon Plus Environment

Environmental Science & Technology

66

by using high input power (76.5 W); the second was designed to target high removal efficiency

67

by using far lower input power (4.1 W). Both cases involved the “laminar jet with bubbling”

68

(LJB) reactor, which was found to be the most effective of several reactor types investigated in

69

previous work.28 Additional objectives were to 1) determine whether co-contaminants affect the

70

performance of PWT by treating samples of PFAS-contaminated groundwater, 2) quantify

71

shorter-chain PFAAs that are formed as byproducts from the transformation of PFOA and PFOS,

72

and 3) to determine the types of reactive species that play significant roles in the transformation

73

of PFOA by PWT.

74

4 ACS Paragon Plus Environment

Page 4 of 19

Page 5 of 19

75

Environmental Science & Technology

Experimental

76

Equipment and procedures. A custom-built high voltage (HV) pulsed power supply was

77

used to generate the plasma. The electrical and operating parameters (discharge voltage,

78

discharge frequency and load capacitance) were varied between experiments (Table 1). The

79

voltage and current in the plasma reactor were measured using a Tektronix P6015A high voltage

80

probe and a Tektronix P6021 current probe connected to a Tektronix TDS 3032C oscilloscope.

81

Voltage and current waveforms are provided in the Supporting Information for the cases of

82

laminar jet with bubbling with high rate and high efficiency (Figure S1). The general circuit

83

diagram for the HV pulsed power supply can be found in a previous publication.28

84

Table 1. Electrical and operational parameters for each set of experiments. Reactor type

Figure

Laminar jet with bubbling (high rate) Laminar jet with bubbling (high efficiency) Laminar jet with bubbling (byproduct trials) Liquid discharge Gas discharge with bubbling

1(a) 1(a) 1(a) 1(b) 1(c)

Discharge Capacitance Discharge frequency (Hz) (nF) voltage (kV) 120 2 +25.0 20 1 +16.0 43 0.94 +20.0 60 1 +18.8 and -16.5 60 1 +18.8 and -16.5

Discharge energy (J) 0.63 0.13 0.19 0.18 and 0.14 0.18 and 0.14

85 86

Three different reactor types were used in this study (Figure 1) and consisted of a 17.3 cm

87

diameter glass vessel (total volume = 3.8 L) fitted with an airtight polymer cap, which was

88

adapted to allow for sample extraction, solution recirculation, and integration of the electrodes.

89

The reactors were operated in semi-batch mode, with liquid recirculating at 1.4 L/min. The liquid

90

recirculation loop ensured thorough mixing and included a heat exchanger to keep the solution at

91

15°C. The headspace was purged with argon at 3.9 L/min either directly or through a submerged

92

diffuser. It must be noted that although the argon was sourced directly from a pressurized

93

cylinder, the power required to achieve the same flow rates with a gas pump was included in the

94

input power calculations used in the following section (discussed further in the PFAS byproducts

95

section). The general characteristics (size, shape, location, etc.) of the plasma discharges in the

96

LJB and GDB reactors are shown in Figure S2. Detailed descriptions of each reactor are

97

provided in a previous publication.28

5 ACS Paragon Plus Environment

Environmental Science & Technology

98 99 100

Page 6 of 19

Figure 1. Reactor diagrams: (a) laminar jet with bubbling (LJB), (b) liquid discharge (LD) and (c) gas discharge with bubbling (GDB).

101

The high concentration (20 µM) PFOA solutions were prepared by dissolving the PFOA

102

(96% purity, Sigma-Aldrich, St. Louis, MO) in deionized water and adjusting the solution

103

electrical conductivity to the desired value. For the aqueous electron ( eaq ) scavenger

104

experiments, the PFOA solution contained 10 mM sodium nitrate (NaNO3), which yielded a

105

conductivity of 1360 µS/cm. For all other experiments with high concentrations of PFOA, NaCl

106

was used to adjust the conductivity to 1360 µS/cm. The groundwater experiments were

107

conducted using unmodified samples from the effluent of an air stripper within the Former

108

NAWC Warminster Groundwater Treatment Plant in Warminster, Pennsylvania. The conductivity

109

of the groundwater was 1150 µS/cm. For the complimentary experiments with prepared solutions

110

containing low concentrations of PFOA (3.1 nM) and PFOS (0.2 nM), the solutions were

111

prepared

112

heptadecafluorooctanesulfonic acid potassium salt, Sigma-Aldrich, St. Louis, MO) in deionized

113

water and adding NaCl to adjust the conductivity to 1150 µS/cm. The solutions for the byproduct

114

experiments were prepared in the same manner as the previous low-concentration solutions, but

115

with a conductivity of 300 µS/cm.



by

dissolving

PFOA

(same

as

above)

and

PFOS

(98%

purity

116

Analysis. Analysis of PFASs was carried out using a Waters Acquity UPLC coupled with a

117

Xevo G2 QToF mass spectrometer and equipped with an Acquity HSS T3 (2.1 mm x 100 mm,

118

1.8 µm) column and a 100 µL injection loop. The method employed for PFAS analysis has been

119

described previously in full detail.29

6 ACS Paragon Plus Environment

Page 7 of 19

Environmental Science & Technology

120

For the byproduct quantification experiments, the samples were concentrated using solid

121

phase extraction and analyzed using isotope dilution liquid chromatography with tandem mass

122

spectrometry (LC/MS-MS); further details of the method have been described previously.12

123

Fluoride was analyzed by the EPA method 9214,30 using a Fisher Scientific accumet Excel

124

XL60 meter kit with an accumet combination electrode and total ionic strength adjustment buffer

125

(TISAB) obtained from VWR Chemicals.

7 ACS Paragon Plus Environment

Environmental Science & Technology

126

Page 8 of 19

Results and discussion

127

PWT efficacy for prepared solutions. Figure 2 compares the performance of the high

128

efficiency and high rate embodiments of the LJB reactor in terms of reduction in PFOA

129

concentration and percentage of fluorine recovered as F-.

130

131 132

Figure 2. Normalized PFOA concentration and defluorination profiles for the LJB reactor

133

configured for high treatment efficiency and high treatment rate.

134 135

The performances of these PWT processes were compared to those of some leading (sonolysis,31

alternative

137

treatment17) and DC plasma in O2 bubbles23 (Table 2), in terms of the observed first-order

138

removal rate constant (kobs) divided by the power density (PD = input power/treated volume),

139

which is a measure of the PFOA removal efficiency. Compared to the other processes, PWT

140

performed well, particularly the high efficiency PWT, which was about eight times more

141

efficient than activated persulfate, about four times more efficient than electrochemical

142

treatment, and over 57 times more efficient than sonolysis.

treatment

technologies,

activated

persulfate,32

136

electrochemical

143

Other performance indicators and corresponding experimental parameters for these seven

144

processes are shown in Table 2. Because non-mineralized transformation byproducts may be

145

more mobile than the parent compound and may still be harmful to the environment, the rate and

146

efficiency of mineralization (transformation of PFOA to F- and CO2) is important. Though CO2

147

must be measured to directly confirm that the PFOA is actually mineralized, defluorination has

148

been used previously to represent mineralization13 and is used here for the same purpose. 8 ACS Paragon Plus Environment

Page 9 of 19

Environmental Science & Technology

149

However, it must be noted that there may be discrepancies between rates and extents of

150

defluorination and mineralization, and so the defluorination results and comparisons presented

151

here should not be taken as directly representing mineralization.

152

To make fair comparisons between defluorination capabilities, the differences in

153

transformation rates must be accounted for. To determine this F50 was defined as the percentage

154

of fluorine recovered in the form of fluoride normalized by the time corresponding to 50%

155

reduction in PFOA concentration (t50 which is equal to −ln(1/2)/kobs). Therefore, F50/t50

156

effectively represents the defluorination rate, while (F50/t50)/PD represents the defluorination

157

efficiency. It should be noted that although t50 was not reached during the high efficiency

158

treatments, PFOA transformation closely follows pseudo-first-order kinetics and defluorination

159

rates are fairly constant, which allowed for t50 and F50 to be obtained by extrapolation.

160

The high efficiency PWT process performed well compared with the alternative treatment

161

methods, achieving a defluorination efficiency about 30 times greater than that of activated

162

persulfate, 10 times greater than that of sonolysis and 15% greater than electrochemical

163

treatment. Although the high rate PWT process does not compare as well in terms of efficiency,

164

it provides insights into the operational versatility of the PWT and how process performance can

165

be varied by changing operational parameters.

166 167

Table 2. Performance indicators and corresponding experimental parameters for high rate PWT,

168

high efficiency PWT, sonolysis, UV-activated persulfate, electrochemical treatment and DC

169

plasma in O2 bubbles. Parts of this table have been adapted and used in [33] (expected publication

170

date: 2017). Treatment High rate PWT

171 172

[PFOA]0 (µM) 20

PD (W/L) 54.6 a a

kobs (min-1)

k obs  −4 min −1   10 ⋅  PD  W L

F50  %    t 50  min 

0.074

14

2.1

3.8

this work

0.012

41

0.31

11

this work

F50 t 50  −2 % min  10 ⋅  PD  W L 

Ref.

High efficiency PWT

20

2.90

Sonolysis

20

250

0.018

0.72

2.5

0.99

[31]

UV/persulfate

50 b

23

0.012

5.2

0.09

0.38

[32]

Electrochemical

0.031 b

5.0

0.0057

11

0.47

9.5

[17] c

DC plasma in O2

100 b

1550

0.030

0.20

4.4

0.28

[23]

a

Input power includes power requirements for gas pump and plasma generation. b Performance may be sensitive to initial PFOA concentration, thus comparisons are approximate. c For a current density of 10 mA/cm2.

9 ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 19

173

PWT efficacy for contaminated groundwater. To determine whether the performance of

174

PWT would transfer well to practical applications, where co-contaminants may interfere,

175

samples of contaminated groundwater were treated using the high efficiency LJB reactor (Figure

176

3). In addition to PFOA (~2.4 nM), the groundwater samples contained measurable

177

concentrations of PFOS (~0.5 nM) and perfluorohexane sulfonate (PFHxS; ~1.0 nM) as well as

178

non-fluorinated co-contaminants such as trichloroethene (3.6 µg/L), tetrachloroethene (0.33

179

µg/L), and had a TOC concentration of 0.67 mg/L. For comparison, the high efficiency LJB

180

reactor was also used to treat a prepared solution, containing similar concentrations of PFOA

181

(3.1 nM) and PFOS (0.2 nM), but without any other co-contaminants.

182

PFOA was degraded at about the same rate in both cases (within 2.5%), indicating that the

183

non-PFAS co-contaminants present in the groundwater had no significant effect on PWT

184

efficacy. Potential reasons for this are discussed in the key reactants section. Compared to the

185

previous case of high efficiency LJB reactor with 20 µM PFOA, treatment of lower

186

concentrations of PFOA yielded a value for kobs/PD about six times as large (250 min-1/(W/L)).

187

While this highlights the magnitude of the influence of initial concentration, the mechanism

188

underlying this influence requires further investigation.

189 190

These results also demonstrate that PWT is capable of degrading other PFASs, most notably PFOS, whose transformation rate constant was more than twice that of PFOA.

191 192 193

Figure 3. Normalized concentration profiles for PFASs in the contaminated groundwater and prepared solution.

194 10 ACS Paragon Plus Environment

Page 11 of 19

Environmental Science & Technology

195

PFAS byproducts. Previous studies on the transformation of PFOA and PFOS have cited

196

shorter-chain PFAAs as a major class of byproducts. A preliminary investigation was conducted

197

to quantify shorter-chain PFAAs produced during treatment of a mixture of PFOS (detection

198

limit: 1 ng/L) and PFOA (detection limit: 5 ng/L) in the LJB reactor. The anticipated byproducts

199

that were tested for were perfluoroheptanoic acid (PFHpA), perfluorohexanoic acid (PFHxA)

200

and perfluoropentanoic acid (PFPnA), which had detection limits of 0.5, 1 and 2 ng/L,

201

respectively. Though shorter-chain PFAAs are clearly being produced (Figure 4), the difference

202

between the sum of the concentrations of PFOA and PFOS and the sum of the concentrations of

203

all PFASs is never greater than 0.1 nmol/L. This indicates that shorter-chain PFAAs account for

204

only about 10% of the degraded PFOA and PFOS, which is much lower than for oxidation-based

205

processes, where shorter-chain PFAAs account for most of the degraded PFOA (85-95% for

206

activated persulfate32). PFAAs smaller than PFPnA may also be formed, however, they would be

207

formed via the much slower transformation of PFPnA, thus they are not expected to be produced

208

in significant quantities. Further investigation is required to fully understand the byproducts

209

produced, particularly those in the gas phase.

210

While identification of gaseous byproducts is of interest due to the insights they offer into the

211

reaction pathways, they are expected to be of significantly less importance in relation to process

212

viability because the process gas is argon, which is neutral and therefore largely unaffected by

213

the plasma. This will allow the argon to be recycled (power requirements for gas pump were

214

included in the input power calculations for the high rate and high efficiency cases), which will

215

allow any gaseous byproducts to be re-treated and further degraded. The quantities of gaseous

216

byproducts formed and the extent to which argon recycling will reduce their emission is a subject

217

of ongoing investigation.

218

11 ACS Paragon Plus Environment

Environmental Science & Technology

219 220

Figure 4. Concentration profiles showing the reduction in concentrations of PFOA and PFOS,

221

and a corresponding increase in concentrations of PFHpA, PFHxA and PFPnA.

Page 12 of 19

222

− Key Reactants. Past studies on PFOA transformation have provided evidence supporting eaq

223

as the primary reactant for many reduction-based technologies.34,35 In previous plasma studies it

224

− has been observed that negative polarity discharges in water generate substantial amounts of eaq ,

225

− while positive polarity discharges do not.36,37 To investigate the potential role of eaq in the

226

transformation of PFOA, experiments were conducted using the LD reactor with both positive

227

and negative polarity. The rate constant for positive polarity is very small compared to that for

228

− negative polarity (Figure 5a), which supports the notion that eaq are important in PFOA

229

transformation. To further test this finding, experiments were performed with negative polarity

230

− liquid discharges in the presence of 10 mM NaNO3, which is an effective eaq scavenger (Figure

231

5a).34 The NaNO3 suppressed the transformation of PFOA almost entirely, providing additional

232

− evidence that the eaq produced by the plasma are primarily responsible for degrading PFOA, and

233

likely account for a significant fraction of the overall removal rate achieved by the PWT

234

processes.

235

The discrepancy between the rate constants for positive and negative polarity discharges in

236

liquid also indicates that hydrogen radicals, hydroxyl radicals and other oxidants generated by

237

the plasma play an insignificant role in initiating primary reactions. These radicals are

238

observably present for both polarities (positive polarity produces greater quantities than

12 ACS Paragon Plus Environment

Page 13 of 19

Environmental Science & Technology

239

negative),38-39 and thus their presence alone is not enough to initiate PFOA transformation at an

240

appreciable rate.

241

242 243

Figure 5. (a) Observed PFOA removal rate constants for (+) NaCl: positive polarity LD, (-)

244

− NaCl: negative polarity LD and (-) NaNO3: negative polarity LD with eaq scavenger. (b)

245

Normalized observed PFOA removal rate constants for (+) NaCl: positive polarity GDB, (-)

246

− NaNO3: negative polarity GDB with eaq scavenger, (-) NaCl: negative polarity GDB and (-)

247

− NaNO3: negative polarity GDB with eaq scavenger. For each polarity, kobs was normalized with

248

− respect to the case without eaq scavenger.

249 250

Conclusions drawn from liquid phase discharge experiments do not necessarily apply to gas

251

phase discharge reactors (LJB and GDB). Therefore, a similar set of experiments was conducted

252

using the GDB reactor. For positive and negative polarity, the presence of NaNO3 caused a

253

significant (at the 90% CI) reduction in the rate of PFOA removal (Figure 5b); however, the

254

effect of NaNO3 is not as extreme as in the case of negative polarity discharges in liquid. This

255

− suggests that, in addition to eaq , there is at least one other reactive species responsible for

256

initiating a significant fraction of the primary reactions with PFOA in the gas phase discharge

257

reactors.

258

PFOA adsorbs to the gas-water interface (the region in which most primary reactions occur

259

in plasma reactors)28 such that much of its hydrophobic tail (5-6 carbons) protrudes into the gas 13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 19

260

phase,14 thus it is likely that the other reactive species is present in the plasma interior.40 Within

261

the plasma interior, there are two species that are both abundant and likely capable of initiating

262

reactions with the fluorocarbon tail of PFOA: high-energy free electrons and argon ions

263

(evidence of excited-state argon ions is provided in the Supporting Information (Figure S3)).

264

High-energy free electrons may initiate a reaction through excitation or ionization and argon ions

265

may initiate a reaction through charge transfer. The ionization potential of argon (15.7 eV) is

266

much greater than the ionization potential of PFOA (~11 eV, based on calculations carried out by

267

our group using the Gaussian 09 program), which, in the event of charge transfer, will provide

268

the PFOA molecule with a large excess of energy and cause its rapid fragmentation.41 It is also

269

possible that PFOA is thermally decomposed (which occurs at 300-350ºC)13, due to the high

270

temperature in the plasma interior, as this has been confirmed as an important transformation

271

mechanism for PFOS in a different plasma system.42 However, due to the steep temperature

272

gradients near the plasma-liquid interface and the difficulty of obtaining reliable spatially-

273

resolved estimates for the temperature of the plasma interior,43-45 it is uncertain whether the

274

fluorocarbon tail is exposed to temperatures high enough for thermal decomposition to occur in

275

this system. It may be possible to confirm the presence or absence of these proposed mechanisms

276

via our continued investigation of the transformation byproducts.

277

Because most reactions are taking place at or above the gas-liquid interface, PWT should be

278

far less sensitive than most other treatment processes to the presence of co-contaminants, such as

279

NOM or other organic compounds, which is consistent with the results for the groundwater

280

treatment. This lack of sensitivity to co-contaminants coupled with the high PFOA removal and

281

defluorination efficiencies makes PWT a promising technology for remediation of PFAS-

282

contaminated water.

283 284 285

Acknowledgements

286

The authors thank the U.S. EPA for its financial, technical, and administrative assistance in

287

funding and managing this project (Agreement Number 83533201). The comments and views

288

detailed herein may not necessarily reflect the views of the U.S. EPA. The authors thank Bernard

289

Crimmins and Adam Point for their assistance with the UPLC-ToF-MS analysis, Xiangru Fan for

290

calculating the ionization potential of PFOA and Timothy Appleman for helping to arrange our 14 ACS Paragon Plus Environment

Page 15 of 19

Environmental Science & Technology

291

procurement of the groundwater samples. The authors also thank the following personnel at the

292

Southern Nevada Water Authority for analytical support: Brett Vanderford, Oscar Quiñones, and

293

Janie Zeigler-Holady.

294 295

Supporting Information Available

296

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

15 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 19

297

References

298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342

1. Schultz, M. M.; Barofsky, D. F.; Field, J. A., Fluorinated Alkyl Surfactants. Environ. Eng. Sci. 2003, 20 (5), 487-501. 2. Houtz, E. F.; Higgins, C. P.; Field, J. A.; Sedlak, D. L., Persistence of Perfluoroalkyl Acid Precursors in AFFF-Impacted Groundwater and Soil. Environ. Sci. Technol. 2013, 47 (15), 8187-8195. 3. Post, G. B.; Cohn, P. D.; Cooper, K. R., Perfluorooctanoic acid (PFOA), an emerging drinking water contaminant: A critical review of recent literature. Environ. Res. 2012, 116, 93117. 4. Stahl, T.; Mattern, D.; Brunn, H., Toxicology of perfluorinated compounds. Env. Sci. Eur. 2011, 23 (1), 1-52. 5. Lau, C.; Butenhoff, J. L.; Rogers, J. M., The developmental toxicity of perfluoroalkyl acids and their derivatives. Toxicol. Appl. Pharmacol. 2004, 198 (2), 231-241. 6. Guelfo, J. L.; Higgins, C. P., Subsurface Transport Potential of Perfluoroalkyl Acids at Aqueous Film-Forming Foam (AFFF)-Impacted Sites. Environ. Sci. Technol. 2013, 47 (9), 41644171. 7. Hu, X. C.; Andrews, D. Q.; Lindstrom, A. B.; Bruton, T. A.; Schaider, L. A.; Grandjean, P.; Lohmann, R.; Carignan, C. C.; Blum, A.; Balan, S. A., Detection of Poly-and Perfluoroalkyl Substances (PFASs) in US Drinking Water Linked to Industrial Sites, Military Fire Training Areas, and Wastewater Treatment Plants. Environ. Sci. Technol. Lett. 2016, 3 (10), 344-350. 8. Drinking Water Health Advisory for Perfluorooctane Sulfonate (PFOS); EPA 822-R-16004; U.S. Environmental Protection Agency; Washington, DC, 2016. 9. Drinking Water Health Advisory for Perfluorooctanoic Acid (PFOA); EPA 822-R-16005; U. S. Environmental Protection Agency; Washington, DC, 2016. 10. Quiñones, O.; Snyder, S. A., Occurrence of Perfluoroalkyl Carboxylates and Sulfonates in Drinking Water Utilities and Related Waters from the United States. Environ. Sci. Technol. 2009, 43 (24), 9089-9095. 11. Xiao, F.; Simcik, M. F.; Gulliver, J. S., Mechanisms for removal of perfluorooctane sulfonate (PFOS) and perfluorooctanoate (PFOA) from drinking water by conventional and enhanced coagulation. Water Res. 2013, 47 (1), 49-56. 12. Appleman, T. D.; Higgins, C. P.; Quiñones, O.; Vanderford, B. J.; Kolstad, C.; ZeiglerHolady, J. C.; Dickenson, E. R. V., Treatment of poly- and perfluoroalkyl substances in U.S. full-scale water treatment systems. Water Res. 2014, 51, 246-255. 13. Vecitis, C. D.; Park, H.; Cheng, J.; Mader, B. T.; Hoffmann, M. R., Treatment technologies for aqueous perfluorooctanesulfonate (PFOS) and perfluorooctanoate (PFOA). Front. Environ. Sci. Eng. China 2009, 3 (2), 129-151. 14. Campbell, T. Y.; Vecitis, C. D.; Mader, B. T.; Hoffmann, M. R., Perfluorinated surfactant chain-length effects on sonochemical kinetics. J. Phys. Chem. A 2009, 113 (36), 9834-9842. 15. Liu, C. S.; Higgins, C. P.; Wang, F.; Shih, K., Effect of temperature on oxidative transformation of perfluorooctanoic acid (PFOA) by persulfate activation in water. Sep. Purif. Technol. 2012, 91 (0), 46-51. 16. Mitchell, S. M.; Ahmad, M.; Teel, A. L.; Watts, R. J., Degradation of Perfluorooctanoic Acid by Reactive Species Generated through Catalyzed H2O2 Propagation Reactions. Environ. Sci. Technol. Lett. 2014, 1 (1), 117-121. 17. Schaefer, C. E.; Andaya, C.; Urtiaga, A.; McKenzie, E. R.; Higgins, C. P., Electrochemical treatment of perfluorooctanoic acid (PFOA) and perfluorooctane sulfonic acid 16 ACS Paragon Plus Environment

Page 17 of 19

343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387

Environmental Science & Technology

(PFOS) in groundwater impacted by aqueous film forming foams (AFFFs). J. Hazard. Mater. 2015, 295, 170-175. 18. Appleman, T. D.; Dickenson, E. R. V.; Bellona, C.; Higgins, C. P., Nanofiltration and granular activated carbon treatment of perfluoroalkyl acids. J. Hazard. Mater. 2013, 260, 740746. 19. Yan, H.; Cousins, I. T.; Zhang, C.; Zhou, Q., Perfluoroalkyl acids in municipal landfill leachates from China: Occurrence, fate during leachate treatment and potential impact on groundwater. Sci. Total Environ. 2015, 524–525, 23-31. 20. Watanabe, N.; Takata, M.; Takemine, S.; Yamamoto, K., Thermal mineralization behavior of PFOA, PFHxA, and PFOS during reactivation of granular activated carbon (GAC) in nitrogen atmosphere. Environ. Sci. Pollut. Res. 2015, 1-6. 21. Joshi, R. P.; Thagard, S. M., Streamer-Like Electrical Discharges in Water: Part II. Environmental Applications. Plasma Chem. Plasma Process. 2013, 33 (1), 17-49. 22. Locke, B. R.; Sato, M.; Sunka, P.; Hoffmann, M. R.; Chang, J. S., Electrohydraulic Discharge and Nonthermal Plasma for Water Treatment. Ind. Eng. Chem. Res. 2005, 45 (3), 882905. 23. Hayashi, R.; Obo, H.; Takeuchi, N.; Yasuoka, K., Decomposition of perfluorinated compounds in water by DC plasma within oxygen bubbles. Electr. Eng. Japan. 2015, 190 (3), 916. 24. Matsuya, Y.; Takeuchi, N.; Yasuoka, K., Relationship Between Reaction Rate of Perfluorocarboxylic Acid Decomposition at a Plasma–Liquid Interface and Adsorbed Amount. Electr. Eng. Japan. 2014, 188 (2), 1-8. 25. Takeuchi, N.; Kitagawa, Y.; Kosugi, A.; Tachibana, K.; Obo, H.; Yasuoka, K., Plasma– liquid interfacial reaction in decomposition of perfluoro surfactants. Journal of Physics D: Applied Physics 2013, 47 (4), 045203. 26. Yasuoka, K.; Sasaki, K.; Hayashi, R.; Kosugi, A.; Takeuchi, N., Degradation of perfluoro compounds and F-recovery in water using discharge plasmas generated within gas bubbles. Int. J. Plasma Environ. Sci. Technol 2010, 4 (2), 113-117. 27. Yasuoka, K.; Sasaki, K.; Hayashi, R., An energy-efficient process for decomposing perfluorooctanoic and perfluorooctane sulfonic acids using dc plasmas generated within gas bubbles. Plasma Sources Sci. Technol. 2011, 20 (3), 034009. 28. Stratton, G. R.; Bellona, C. L.; Dai, F.; Holsen, T. M.; Thagard, S. M., Plasma-based water treatment: Conception and application of a new general principle for reactor design. Chem. Eng. J. 2015, 273, 543-550. 29. Crimmins, B. S.; Xia, X.; Hopke, P. K.; Holsen, T. M., A targeted/non-targeted screening method for perfluoroalkyl carboxylic acids and sulfonates in whole fish using quadrupole timeof-flight mass spectrometry and MSe. Anal. Bioanal. Chem. 2014, 406 (5), 1471-1480. 30. Method 9214 - Potentiometric determination of fluoride in aqueous samples with ionselective electrode; Environmental Protection Agency; 1996. 31. Vecitis, C.; Park, H.; Cheng, J.; Mader, B.; Hoffmann, M., Enhancement of perfluorooctanoate and perfluorooctanesulfonate activity at acoustic cavitation bubble interfaces. The Journal of Physical Chemistry C 2008, 112 (43), 16850-16857. 32. Chen, J.; Zhang, P., Photodegradation of perfluorooctanoic acid in water under irradiation of 254 nm and 185 nm light by use of persulfate. Water science and technology 2006, 54 (1112), 317-325.

17 ACS Paragon Plus Environment

Environmental Science & Technology

388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426

Page 18 of 19

33. Advanced Oxidation Processes for Water Treatment: Fundamentals and Applications, 2017. http://www.iwapublishing.com/books/9781780407180/advanced-oxidation-processeswater-treatment-fundamentals-and-applications. 34. Song, Z.; Tang, H.; Wang, N.; Zhu, L., Reductive defluorination of perfluorooctanoic acid by hydrated electrons in a sulfite-mediated UV photochemical system. J. Hazard. Mater. 2013, 262, 332-338. 35. Park, H.; Vecitis, C. D.; Cheng, J.; Choi, W.; Mader, B. T.; Hoffmann, M. R., Reductive defluorination of aqueous perfluorinated alkyl surfactants: effects of ionic headgroup and chain length. J. Phys. Chem. A 2009, 113 (4), 690-696. 36. Thagard, S. M.; Takashima, K.; Mizuno, A., Chemistry of the positive and negative electrical discharges formed in liquid water and above a gas–liquid surface. Plasma Chem. Plasma Process. 2009, 29 (6), 455-473. 37. Rumbach, P.; Bartels, D. M.; Sankaran, R. M.; Go, D. B., The solvation of electrons by an atmospheric-pressure plasma. Nat. Commun. 2015, 6. 38. Sun, B.; Sato, M.; Clements, J. S., Optical study of active species produced by a pulsed streamer corona discharge in water. J. Electrostat. 1997, 39 (3), 189-202. 39. Miyahara, T.; Oizumi, M.; Nakatani, T.; Sato, T., Effect of voltage polarity on oxidationreduction potential by plasma in water. AIP Advances 2014, 4 (4), 047115. 40. Thagard, S. M.; Stratton, G. R.; Dai, F.; Bellona, C. L.; Holsen, T. M.; Bohl, D. G.; Paek, E.; Dickenson, E. R., Plasma-based water treatment: development of a general mechanistic model to estimate the treatability of different types of contaminants. Journal of Physics D: Applied Physics 2016, 50 (1), 014003. 41. Johnstone, R. A. W., Mass spectrometry for organic chemists. CUP Archive: 1972; p. 166. 42. Tachibana, K.; Takeuchi, N.; Yasuoka, K., Reaction Process of Perfluorooctanesulfonic Acid (PFOS) Decomposed by DC Plasma Generated in Argon Gas Bubbles. Ieee T Plasma Sci 2014, 42 (3), 786-793. 43. Bruggeman, P.; Kushner, M. J.; Locke, B. R.; Gardeniers, J.; Graham, W.; Graves, D. B.; Hofman-Caris, R.; Maric, D.; Reid, J. P.; Ceriani, E., Plasma–liquid interactions: a review and roadmap. Plasma Sources Sci. Technol. 2016, 25 (5), 053002. 44. Bruggeman, P.; Sadeghi, N.; Schram, D.; Linss, V., Gas temperature determination from rotational lines in non-equilibrium plasmas: a review. Plasma Sources Sci. Technol. 2014, 23 (2), 023001. 45. Zhang, S.; van Gaens, W.; van Gessel, B.; Hofmann, S.; van Veldhuizen, E.; Bogaerts, A.; Bruggeman, P., Spatially resolved ozone densities and gas temperatures in a time modulated RF driven atmospheric pressure plasma jet: an analysis of the production and destruction mechanisms. Journal of Physics D: Applied Physics 2013, 46 (20), 205202.

18 ACS Paragon Plus Environment

Page 19 of 19

427 428

Environmental Science & Technology

For Table of Contents Only

19 ACS Paragon Plus Environment