0% found this document useful (0 votes)
58 views54 pages

Destruction of Per-And Polyfluoroalkyl Substances (Pfas) With Advanced Reduction Processes (Arps) : A Critical Review

Uploaded by

Aissa Dehane
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
58 views54 pages

Destruction of Per-And Polyfluoroalkyl Substances (Pfas) With Advanced Reduction Processes (Arps) : A Critical Review

Uploaded by

Aissa Dehane
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd

Subscriber access provided by UNIV VAN AMSTERDAM

Critical Review
Destruction of Per- and Polyfluoroalkyl Substances (PFAS)
with Advanced Reduction Processes (ARPs): A Critical Review
Junkui Cui, Panpan Gao, and Yang Deng
Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.9b05565 • Publication Date (Web): 12 Mar 2020
Downloaded from pubs.acs.org on March 13, 2020

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 53 Environmental Science & Technology

ACS Paragon Plus Environment


Environmental Science & Technology Page 2 of 53

6 Destruction of Per- and Polyfluoroalkyl Substances

7 (PFAS) with Advanced Reduction Processes

8 (ARPs): A Critical Review

9 Junkui Cui†, Panpan Gao†,‡, Yang Deng†,*

10 † Department of Earth and Environmental Studies, Montclair State University, Montclair,

11 New Jersey 07043, United States

12 ‡ School of Environmental Studies, China University of Geosciences, Wuhan, Hubei

13 Province 430074, P. R. China

14

15 ABSTRACT.

16 Advanced reduction processes (ARPs) have emerged as a promising method for

17 destruction of persistent per-and poly-fluoroalkyl substances (PFAS) in water due to the

18 generation of short-lived and highly reductive hydrated electrons (eaq-). This study provides a

19 critical review on the mechanisms and performance of reductive destruction of PFAS with eaq-.
2

ACS Paragon Plus Environment


Page 3 of 53 Environmental Science & Technology

20 Unique properties of eaq- and its generation in different ARP systems, particularly UV/sulfite and

21 UV/iodide, are overviewed. Different degradation mechanisms of PFAS chemicals, such as

22 perfluorooctanoic acid (PFOA), perfluorooctane sulfonate (PFOS), and others (e.g. short chain

23 perfluorocarboxylic acids (PFCAs) and perfluorosulfonic acids (PFSAs), per- and polyfluoro

24 dicarboxylic acids, and fluorotelomer carboxylic acids), are reviewed, discussed, and compared.

25 The degradation pathways of these PFAS chemicals rely heavily upon their head groups. For

26 specific PFAS types, fluoroalkyl chain lengths may also affect their reductive degradation

27 patterns. Degradation and defluorination efficiencies of PFAS are considerably influenced by

28 solution chemistry parameters and operating factors, such as pH, dose of chemical solute (i.e.

29 sulfite or iodide) for eaq- photo-production, dissolved oxygen, humic acid, nitrate, and

30 temperature. Furthermore, implications of the state-of-the-art knowledge on practical PFAS

31 control actions in water industries are discussed and the priority research needs are identified.

32

33 INTRODUCTION

34 Per-and poly-fluoroalkyl substances (PFAS) have been globally incorporated into various

35 industrial and consumer products since the 1940s.1 The anthropogenic chemicals, comprising

36 perfluoroalkyl substances and polyfluoroalkyl substances, encompass a family of fluorinated

37 aliphatic substances containing one or more carbon (C) atoms on which all the hydrogen (H)

38 substituents present in the nonfluorinated analogues, from which they are notionally derived, are

39 replaced with fluoride (F) atoms.2 Therefore, PFAS possess at least one perfluoroalkyl moiety

40 (CnF2n+1-). For polyfluoroalkyl substances, a more general concept is also proposed to embrace

ACS Paragon Plus Environment


Environmental Science & Technology Page 4 of 53

41 the aliphatic substances on which at least one, but not all, C atoms are bonded to fluorine, so that

42 more partially fluorinated compounds are comprised,2, 3 such as those having “scattered” or

43 “grouped” multiple F atoms. Because of their unique resistance to heat, water, oil, and stains, in

44 addition to the reduction of friction,1 PFAS have been historically and presently used in a broad

45 range of products (e.g. food packaging and cookware, paper, textiles, leathers, carpet, mattress,

46 and firefighting materials) and in many industries (e.g. metal plating and etching, building and

47 constriction, photolithography, and semiconductor).1, 4

48 However, concerns on PFAS have gradually been growing due to their prevalence,

49 mobility, persistence, bioaccumulation, and adverse health effects.2, 5-9 Many PFAS have been

50 demonstrated to bioaccumulate, can bind to blood proteins, and have long half-lives in humans.

51 Human PFAS exposure is linked to cancer, obesity, elevated cholesterol, immune suppression,

52 and endocrine disruption.10-12 Various PFAS, particularly perfluorooctanoic acid (PFOA) and

53 perfluorooctane sulfonate (PFOS), have been frequently identified in surface freshwater,13-15

54 groundwater,7, 16-19 drinking water,20-22 and landfill leachate.23-27 The emerging anthropogenic

55 chemicals are challenging water and wastewater treatment, water reclamation, site remediation,

56 and landfill leachate disposal, because of a lack of effective, efficient, and practical treatment

57 technologies. Meanwhile, a tremendous pressure potentially derives from the regulatory

58 determination. In 2016, US Environmental Protection Agency (EPA) issued a Lifetime Health

59 Advisory (LHA) for PFOA and PFOS at 70 ng/L.21 Presently, US EPA is moving forward with

60 the maximum contaminant level (MCL) process for PFOA, PFOS, and probably more PFAS

61 chemicals in drinking water.28 The pursuit of public health and potential regulatory pressure

62 require the water industry to stay current and proactive for advancing innovative PFAS treatment

ACS Paragon Plus Environment


Page 5 of 53 Environmental Science & Technology

63 technologies. Among very few technically effective PFAS treatment methods, advanced

64 reduction processes (ARPs) have recently emerged as a promising option.

65 The term of ARPs was early used in the late 1990s to explain unintended degradation of

66 highly oxidized orgnanic compounds (e.g. carbon tetrachloride) by reducing radical species (e.g.

67 hydrated electrons (eaq-) and H atoms 7@ produced during electron beam irridation of polluted

68 groundwater.29 As a purposeful water treatment option, ARPs were proposed for addressing

69 water contaminants more than one decade later.30, 31 ARPs represent a chemical degradation

70 process producing sufficient and highly reductive radicals for destruction of contaminants in

71 water. The reducing radicals involved typically include eaq- and 7@2 in addition to others such as

72 sulfite radical anions (SO3@ -) and sulfur dioxide radical anions (SO2@ -), depending on the used

73 activation methods and chemical solutes.31 Although the term ARPs for water and wastewater

74 treatment has been very recently adopted, the investivigations on reactions of these reudcing

75 radicals with different chemicals in water began several decades ago.32-36 For environmental

76 applications, ARPs have proven very effective for the removal of various oxidized contaminants,

77 such as vinyl chloride,37, 38 perchlorate,39 bromate40, 41, nitrate42, chromium (VI)43, 44, and 2, 4, 6-

78 trichlorophenol.45

79 Special attention has been recently directed to ARPs due to their capability of effectively

80 destructing different PFAS chemicals in water,31, 46-59 advantageous over hydroxyl radical @57

81 or sulfate radical (SO4@ -) – based advanced oxidation processes (AOPs).47, 50, 58 These

82 encouraging findings demonstrate that ARPs are a promising approach to mitigating the PFAS

83 impacts in an aqueous environment. This article aimed to provide a state-of-the-art review on

84 chemical mechanisms of reductive destruction of PFAS in an advanced reduction process and on

ACS Paragon Plus Environment


Environmental Science & Technology Page 6 of 53

85 the evaluation of the effects of key factors affecting the PFAS degradation and defluorination

86 efficiencies. Furthermore, environmental implications of the knowledge for water industries are

87 discussed and future research needs are identified.

88 HYDRATED ELECTRONS AND THEIR GENERATION FOR DESTRUCTION OF

89 PFAS

90 Although different activation methods (e.g. photolysis, radiolysis, and sonolysis) with or

91 without the assistance of chemical solutes can produce different free reducing radicals in water,

92 photo-irradiation of sulfite (SO32-), iodide (I-), dithionite (S2O62-), ferrocyanide (Fe(CN)64-), or

93 certain organic compounds (e.g. aminopolycarboxylic acids (APCAs) and indole derivatives)

94 was reported for effective decomposition of PFAS chemicals in water in literature.31, 56, 60-64

95 Among these solutes, SO32- and I- are the most frequently used. The effective destruction of

96 PFAS molecules during the ARP treatment is principally ascribed to the generation of eaq-, rather

97 than other reducing agents produced, as evidenced by results of scavenging experiments.56

98 Excess electrons can be generated in electrically neutral water directly by pulse radiolysis or

99 using ionization or detachment of a specific solute that enables the existence of electrons in

100 water rather than association with their parent cations.65, 66 Negative charges of the secondary

101 electrons polarize neighboring neutral water molecules, so that the electrons are bound to these

102 water molecules to create a metastable localized species called hydrated electrons.65, 67

103 Experimental evidence for detection of eaq- was first found in the late 1950s.67 Afterwards,

104 substantial debates have remained over the different models proposed to describe molecular

105 structure of eaq-.68 In a traditionally accepted cavity model, a hydrated electron excludes water to

ACS Paragon Plus Environment


Page 7 of 53 Environmental Science & Technology

106 create and occupy a small quasi-spherical region, which is surrounded by water molecules with

107 O-H bonds pointing toward the cavity.69 However, the existence of a cavity structure has been

108 occasionally questioned.65, 70

109 Hydrated electrons are a potent reducing agent with an extremely negative standard

110 reduction potential of -2.9 V.71, 72 The short-lived radicals can have a half-life time of more than

111 300 microseconds.67 It is conventionally believed that they tend to react with chemical species

112 via a one-electron transfer mechanism.73 Whereas rate constants of these reactions broadly vary

113 from ~10 M-1@ -1 to a diffusion controlled mode, the activation energies range narrowly within 6 -

114 30 E@. -1, suggesting that the kinetics is limited by the availability of a vacant orbital on the

115 target reactant species.60

116 UV/sulfite. Occurrence of a chain reaction during photoionization of sulfite in water was

117 early reported in the 1920s.74 Efforts were later made to elucidate the underlying reaction

118 mechanisms.75-79 Because the UV/sulfite system for reductive destruction of PFAS is mostly

119 operated at an alkaline condition (the effect of pH would be discussed later) and bisulfite (HSO3-

120 ) has a pKa of 7.21,80 the major sulfite species in the PFAS treatment system is SO32- rather than

121 HSO3-. The photoionization of SO32- produces eaq- together with equimolar SO3@ - as shown in

122 Eq.(1).

123 SO32- + hv H SO3@ - + eaq- (1)

124 When vacuum UV (VUV) irradiation (185 nm) is adopted in a UV/sulfite system, additional eaq-

125 can be directly created from photoionization of water, but the phenomena cannot be observed

126 using other spectral regions (e.g. 254 nm).55, 81

ACS Paragon Plus Environment


Environmental Science & Technology Page 8 of 53

127 At an anoxic condition, SO3@ - is subsequently recombined through two parallel pathways

128 to generate S2O62- (Eq. (2)) and sulfate (SO42-) (Eq. (3)), respectively.

129 SO3@ - + SO3@ - H S2O62- 2k2 = 1.1 × 109 M-1@ -1, 76, 82 (2)

130 SO3@ - + SO3@ - + H2O H SO42- + H+ + HSO3- k3/k2 = 0.37, 79 (3)

131 Meanwhile, eaq- may be consumed through three mechanisms in the absence of target

132 chemical species. Firstly, eaq- can react with H+ to produce the conjugate acid, 7@ (Eq. (4)).

133 Subsequently, the produced 7@ can further promptly react with eaq- to form hydrogen gas (H2)

134 (Eq. (5)).

135 eaq- + H+ H 7@ k4 = 2.3 × 1010 M-1@ -1, 60 (4)

136 eaq- + 7@ + H2O H H2 + OH- k5 = 3.0 × 1010 M-1@ -1, 60 (5)

137 Secondly, eaq- can be recombined to produce H2 and OH- (Eq.(6)).

138 eaq- + eaq- + 2H2O H H2 + 2OH- 2k6 = 1.1 × 1010 M-1@ -1, 60 (6)

139 Thirdly, eaq- may react with S2O62- generated from Eq. (2) (Eq. (7)).

140 eaq- + S2O62- H SO32- + SO3@ - k7 K 2 × 105 M-1@ -1, 79 (7)

141 Of note, SO32- can be partially recycled from Eq. (2) and (7) for generation of additional eaq-. The

142 final photo-decomposition products of SO32- are sulfate and dithionite with a reported

143 approximate molar ratio of SO42- to S2O62- of 2:1.78, 79

144 When dissolved oxygen (DO) is present, eaq- generated can be substantially consumed by

145 O2 to produce superoxide (O2@ -) (Eq.(8)), which may further react with eaq- as Eq. (9).60
8

ACS Paragon Plus Environment


Page 9 of 53 Environmental Science & Technology

146 eaq- + O2 H O2@ - k8 = 1.9 × 1010 M-1@ -1 (8)

147 eaq- + O2@ - H O2- k9 = 1.3 × 1010 M-1@ -1 (9)

148 UV/iodide. Photochemistry of iodide in water was also early investigated.83-85

149 Photoexcitation of I- in water first occurs to produce an excited iodide (I@H2O*)

150 (Eq. (10)).

151 I- + H2O + hv H I@H2O -* (10)

152 I@H2O* can decay to its ground state (I-) though Eq. (11). It can also be thermally degraded to an

153 intermediate state, i.e. a caged complex <@2 e-), in which electrons are less confined and undergo

154 a quasi-random walk diffusion process in the proximity of I atom <@ over water molecules (Eq.

155 (12)).84 Finally, eaq- can be generated through dissociation of the caged complex, accompanied

156 with the production of <@ (Eq. (13)).85

157 I@H2O -* H I- + H2O (11)

158 I@H2O -* H <@2 e-) + H2O (12)

159 <@2 e-) H <@ + eaq- (13)

160 Here ( ) represents the solvent cage in which <@ and eaq- are formed in pairs.

161 More detailed mechanistic information regarding the eaq- generation from photo-

162 irradiation of I- in water was further explored. Initial photoexcitation involves a state above the

163 charge-transfer-to-solvent (CTTS) absorption, which is an absorption band for detachment of the

164 electron from I-. The photo-detached electron is an electron (etrapped-) trapped in a polarization

ACS Paragon Plus Environment


Environmental Science & Technology Page 10 of 53

165 well created by water molecules oriented around I-.86, 87 88 The trapped electron would

166 subsequently experience electron solvation, which is viewed as an excited state relaxation

167 process between different electronic energy levels of the solvated electrons.89-91 Consequently,

168 two other electron species are sequentially produced primarily through a nonadiabatic electron

169 transfer process,86 including wet electrons (ewet-) and eaq-. ewet- is a pre-solvated electron with the

170 lowest excited state, while eaq- is the equilibrium solvated electron with the ground state

171 (Eq.(14)).90 Transitions among the three electron species are extremely rapid with a time scale in

172 the order of femtoseconds.86, 89, 92

173 etrapped- H ewet- H eaq- (14)

174 <@ generated from Eq. (13) can react with the added I- to generate various I species, which

175 react with another one in the water as follows.

176 <@ + I- M I2@ - k15f = 8.8 × 109 M-1@ -1, K = 1.28 × 105 M-1, 93 (15)

177 <@ + <@ H I2 2k16 = 3.0 × 1010 M-1@ -1, 93 (16)

178 <@ + I2@ - H I3- k17 = 4.6 × 109 M-1@ -1, 93 (17)

179 I2@ - + I2@ - H I3- + I- 2k18 = 4.6 × 109 M-1@ -1, 93 (18)

180 I- + I2 M I3- k19f = 5.6 × 109 M-1@ -1, K = 747 M-1, 93 (19)

181 Here, the subscript of f in the above rate constants means that the rate constant is for the forward

182 reaction. In the absence of target pollutants, eaq- can be consumed by several I species (Eq.(20)

183 and (21)),47 besides reactions with H+ (Eq. (4) and (5)) and self-recombination (Eq.(6)).

10

ACS Paragon Plus Environment


Page 11 of 53 Environmental Science & Technology

184 eaq- + I3- H I2@ - + I- (20)

185 eaq- + I2@ - H 2I- (21)

186 UV/Others. Other solutes were also used under the UV irradiation to generate eaq- for

187 destruction of PFAS in water, such as Fe(CN)64-, which can be photolyzed to generate eaq- (Eq.

188 (22)) for degradation of PFAS.62

189 Fe(CN)64- + hv H eaq- + Fe(CN)63- (22)

190 It should be noted that a fraction of eaq- generated from the UV/sulfite or UV/iodide

191 system can be scavenged by oxidizing species, leading to a low utilization efficiency of and a

192 short lifetime of eaq-. To address this issue, organic compounds were proposed for the generation

193 of eaq-. One such example is nitrilotriacetic acid (NTA), which plays a dual role in water.56 It is

194 firstly fully hydrated with hydration spheres and then serves as a photosensitizer to facilitate

195 water photo-dissociation and photo-ionization for generation of eaq- and @57,94

196 2H2O + 2hv H eaq- + @57 + H3O+ (23)

197 Afterwards, NTA with a high reactivity toward @57 (rate constant = 4.2 × 109 M-1@ -1 at pH 10.0)

198 minimizes the germinate recombination between eaq- and @57295 promoting the eaq- utilization for

199 PFAS. Therefore, under the identical conditions (0.01 mM PFOS; 2 mM NTA or SO32-; pH

200 10.0), UV/NTA was reported to achieve better degradation and defluorination of PFOS than

201 UV/sulfite.56

11

ACS Paragon Plus Environment


Environmental Science & Technology Page 12 of 53

202 The other organic solutes used are indole (Eq. (24)) and its derivatives (e.g. 3-indole-

203 acetic-acid (IAA)) to generate organic radical cations and eaq- for destruction of PFAS in water.96,

204 97

205 Indole + hv H eaq- + < @+ (24)

206 The reductive degradation can be dramatically enhanced in the presence of organomodified

207 montmorillonite (a naturally occurring clay).63, 64 The organic modification materials are cationic

208 surfactants, including hexadecyltrimethylammonium (HDTMA) bromide,

209 dioctodecyldimethylammonium (DODMA) bromide, hexadecylpyridinium (HDPY) bromide,

210 and trimethylphenylammonium (TMPA) bromide. The modified clay possesses an excellent

211 adsorption capability for both PFAS and indole solutes, creating a constrained interlayer space

212 for ensuing reactions. When eaq- and organic radical cations are simultaneously generated there

213 upon photolysis, the cationic radicals can be rapidly stabilized by the negatively charged planar

214 clay structure to minimize their recombination reactions with eaq-. Consequently, eaq- can more

215 efficiently attack the co-sorbed PFAS on the clay. Degradation performance of this approach is

216 only slightly inhibited in the presence of DO or with a pH decreae.63

217 MECHANISMS OF PFAS DESTRUCTION

218 Fundamental information regarding ARP degradation of PFAS in water has been

219 explored using different approaches, including identification of intermediate and final

220 degradation products, closing mass balance of F and C over the reaction time, measurements of

221 PFAS degradation and defluorination rates, and calculation with density functional theory

222 (DFT). Of note, the extent of defluorination has been quantified using three parameters with

12

ACS Paragon Plus Environment


Page 13 of 53 Environmental Science & Technology

223 different meanings, including the F index (Eq.(25)),46 overall defluorination ratio (overall deF%)

224 (Eq.(26)),59 and molecular defluorination ratio (molecular deF%) (Eq.(27)). 59

[F ]released
225 F index = [PFAS]degraded (25)

[F ]released
226 Overall deF% = [PFAS]0 × NC F
× 100% (26)

[F ]released Overall deF%


227 Molecular deF% = [PFAS]degraded × NC F
× 100% = DP (27)

228 Here, [F-]released is the molar concentration of F- released due to degradation; [PFAS]0 is the initial

229 molar concentration of a specific parent PFAS chemical; [PFAS]degraded is the molar concertation

230 of the PFAS that has been degraded; NC-F is the number of C-F bonds in the parent PFAS
[PFAS]degraded
231 molecule; and DP is the degradation portion of the parent PFAS (i.e. [PFAS]0 ). F index

232 indicates the average number of F- produced from each decomposed parent PFAS molecule at a

233 specific reaction time. Overall and molecular deF% are the ratios of F- released at a specific

234 reaction time to F present in the total (i.e. degraded + residual) and degraded parent PFAS,

235 respectively. Overall deF% indicates an overall conversion efficiency from organic F in the

236 original PFAS to inorganic F anions, while molecular deF% shows the average conversion

237 efficiency of organic F to inorganic F on each decomposed parent PFAS molecule. Among the

238 three defluorination parameters, overall deF% has been the most commonly used. In the

239 following discussion, the defluorination efficiency used is an overall defluorination ratio.

240 PFOA. Reductive degradation of PFOA with eaq- follows a second order kinetic reaction

241 pattern with a rate constant of (1.7 ± 0.5) ×107 M-1@ -1 (0.01 M NaClO4; pH 10.0).62 Rate of

242 PFOA defluorination is lower than its degradation rate,48, 59 typically at the initial reaction phase,

13

ACS Paragon Plus Environment


Environmental Science & Technology Page 14 of 53

243 implying that part of PFOA degradation reactions, such as the cleavage of C-C bonds on PFOA

244 for production of F-containing intermediate products, cannot synchronously translate into the

245 release of F-. PFOA degradation efficiencies can reach up to 100% in an ARP system,47, 48, 51, 56,

246 63, 64, 98 while the maximum defluorination efficiencies are observed to broadly vary between

247 ~55% and 96% in the most literature,47, 48, 51, 56, 59, 63, 64, 98, except for two studies that reported

248 nearly complete defluorination.47, 63 Disparity of the maximum defluorination efficiencies is

249 likely ascribed to the different experimental conditions used in these studies. However, a

250 complete defluorination is argued to be hardly achieved from a perspective of the reductive

251 degradation mechanisms of PFOA, as discussed latter.59

252 Two major intermediate product types are identified from the PFOA reactions with eaq-,

253 including less fluorinated carboxylic acids and shorter-chain PFCAs.47-49, 59 The presence of the

254 former products indicates the cleavage of C-F bonds as well as ensuing H/F exchange, while

255 formation of the latter ones implies the scission of C-C bonds.47 Concentrations of the

256 aforementioned degradation intermediate products are typically increased with time as a result of

257 accumulation due to the degradation of parent compounds at the initial reaction phase, followed

258 by a gradual decrease because the intermediates are increasingly degraded due to their own

259 reactivity toward eaq- and with the decrease of the parent compound concentrations.47, 48, 59

260 Two major parallel reaction pathways have been proposed for the PFOA degradation

261 with eaq- (Figure 1) including H/F exchange and chain shortening. Among all the – CF2 – in a

262 PFOA molecule, the T - one adjacent to the carboxyl group has exhibited a high activity

263 likely due to the inductive effect of the head group, thereby providing the preferential reaction

264 center.47, 59 Furthermore, DFT calculations reveal that for the PFCA radical anion produced after

14

ACS Paragon Plus Environment


Page 15 of 53 Environmental Science & Technology

265 the attachment of an extra electron, T - C-F bonds are spontaneously stretched,

266 facilitating the bond cleavage.59 The ensuing cleavage of C-F bonds leads to the formation of less

267 fluorinated organic radical anions and the elimination of F- (Eq.(28)) followed by H addition

268 (Eq.(29)) to accomplish H/F exchange (for PFOA, n = 7).47

269 CnF2n+1-COO- + eaq- H @ CnF2n-COO- + F- (28)

270 @ CnF2n-COO- + H2O H CnF2nH-COO- + @ OH (29)

271 The most probable structure after the two sequential H/F exchanges on T - C-F bonds is

272 Cn-1F2n-1 –CH2-COOH.59 Specifically, for PFOA (n = 7), the intermediate is C6F13 –CH2-

273 COOH.59 Of note, the presence of –CH2- is believed to increase chemical recalcitrance of the

274 degradation product.59 Meanwhile, other hydrofluorinated degradation products generated from

275 more H/F exchanges are identified (e.g. C7F12H3-COO-, C7F11H4-COO-, and C7F9H6-COO-),

276 suggesting that the C-F bonds on the middle positions of carbon chain can also be cleaved.59

277 Furthermore, the H/F exchange can occur in a similar fashion on shorter-chain PFCAs that are

278 produced via chain shortening.

279 Chain shortening is acknowledged as the other primary PFOA degradation pathway with

280 the formation of shorter-chain PFCAs as a consequence of the scission of C-C bonds. However,

281 controversial pathways have been proposed to explain the mechanism. In an early study,47

282 photo-reductive destruction of PFOA in a UV/iodide system was investigated in 18O-water with

283 the approach used previously to explore the mechanisms governing direct UV irradiation of

284 PFOA.99 Experimental evidence showed the formation of [C6F13C(16O)(16O)]-,

285 [C6F13C(16O)(18O)]-, and [C6F13C(18O)(18O)]-. Therefore, two C-C scission routes were

286 presumed to simultaneously occur during reductive degradation of PFOA with eaq-,47-49 which
15

ACS Paragon Plus Environment


Environmental Science & Technology Page 16 of 53

287 lead to the production of [C6F13C(16O)(16O)]- and [C6F13C(16O)(18O)]- /[C6F13C(18O)(18O)]- ,

288 respectively.

289 In the first one, UV photolysis is believed to excite the PFCA degradation products after

290 two H/F exchanges (i.e. CnF2n-1H2-COO-) to form @ Cn-1F2n-1, @'55-, and CH carbenes (i.e. :CH2)

291 (Eq.(30)). The produced @ Cn-1F2n-1 and @'55- can be recombined to form Cn-1F2n-1-COO- with

292 one fewer CF2 unit than its parent compound (Eq.(31)).47-49, 56

293 CnF2n-1H2-COO- + hv H @ Cn-1F2n-1 + @'55- + :CH2 (30)

294 @ Cn-1F2n-1 + @'55- H Cn-1F2n-1-COO- (31)

295 The stepwise –CH2– removal through direct UV irradiation and aforementioned hydrolysis

296 explains further C-C scission of PFOA. However, the pathway has recently been questioned due

297 to two reasons.59 Firstly, the degradation of Cn-1F2n-1 –CH2–COO- is challenging due to the

298 presence of -CH2-, which is supported by the sluggish degradation of CF3-CH2-COO- with eaq-. If

299 the above pathway was the case, all the PFOA molecules would finally be shortened to

300 trifluoroacetic acid (TFA). Bentel et al.59 found that TFA could be completely defluorinated in an

301 advanced reduction system, but a complete defluorination of PFOA was barely achieved.59

302 Secondly, even though Eq.(30) truly occurs, the possibility of recombination of @ Cn-1F2n-1 and

303 @'55- (Eq.(31)) is low due to their very low concentrations.59

304 The second route is involved with four sequential steps, including decarboxylation

305 initiated by direct UV irradiation (Eq.(32)), hydroxylation (Eq.(33)), F elimination (Eq.(34)), and

306 hydrolysis (Eq.(35)).47, 59 The pathway is called decarboxylation-hydroxylation-elimination-

307 hydrolysis (DHEH) mechanism.

16

ACS Paragon Plus Environment


Page 17 of 53 Environmental Science & Technology

308 CnF2n+1-COO- + hv H @ CnF2n+1 + @'55- (32)

309 @ CnF2n+1 + H2O H CnF2n+1OH + 7@ (33)

310 CnF2n+1OH H Cn-1F2n-1COF + H+ + F- (34)

311 Cn-1F2n-1COF + H2O H Cn-1F2n-1COO- + 2H+ + F- (35)

312 Among the reactions, speculative Eq.(32) and (33) were proposed in a prior study on the direct

313 UV irradiation of PFOA.99 HF elimination from CnF2n+1OH in Eq.(34) was previously proposed

314 for chlorine atom – initiated photo-oxidation of methyl perfluoroalkyl ethers.100 Along the

315 DHEH pathway, the produced Cn-1F2n-1COOH can be further degraded repeatedly in a similar

316 manner, so that multiple CF2 units can be stepwise lost. However, the direct photolysis

317 hypothesis for scission of C-C bonds on PFOA is not supported by the reported evidence that a

318 very limited defluorination efficiency is achieved during direct UV photolysis of PFOA in

319 water.47, 48 In another study supporting the DHEH for chain shortening,59 how the

320 decarboxylation is initiated is not clearly stated. Particularly, whether or how hydrated electrons

321 are involved in the C-C cleavage of PFOA via the DHEH mechanism was not clarified in the

322 available literature.

323 Of note, low energy electrons are capable of inducing decarboxylation of formic acid

324 (HCOOH) and TFA in a condensed phase.101, 102 Therefore, it is plausible that photo-induced

325 hydrated electrons can initiate decarboxylation of PFOA in water in a similar manner, which

326 may serve as the first step of the DHEH mechanism. The hypothesis was proposed to explain the

327 scission of C-C bonds on PFOA during TiO2 photocatalytic decomposition of PFOA in the

17

ACS Paragon Plus Environment


Environmental Science & Technology Page 18 of 53

328 presence of oxalic acid.103 The hydrated electron-driven decarboxylation of PFCAs in water is as

329 follows.

330 CnF2n+1-COO- + e-aq H @ CnF2n+1 + COO- (36)

331 Following the decarboxylation involved with e-aq, the PFOA degradation product would be

332 further degraded via Eq.(33)-(35) to accomplish the cleavage of C-C bonds via a new probable

333 DHEH pathway. Between the two aforementioned chain shortening mechanisms (i.e. Eq.(30)-

334 (31) and Eq.(32)-(36)), the DHEH route seems more plausible. In-depth mechanistic studies are

335 required to elucidate the detailed C-C scission mechanisms.

336 Quantitative information regarding relative contributions of the H/F exchange and chain

337 shortening pathways to PFOA degradation is not reported. However, the measurement of overall

338 defluorination ratio may be informative to determine relative importance of the two competitive

339 mechanisms. Theoretically, a complete defluorination is barely accomplished via the H/F

340 exchange due to chemical persistence of the degradation products, while 100% defluorination

341 can be achieved by chain shortening via the DHEH pathway with the final products of CO2, H2O,

342 and F- as shown in Figure 1.59

343 Besides the two aforementioned reaction pathways, other side reactions may

344 simultaneously occur, as evidenced by identification of other intermediate products, depending

345 on the activation modes of eaq-. In the UV/iodide system, various gaseous intermediates produced

346 are identified, such as CFHI2, C2F4HI, C5F6HI, C6F13I, and C6F10HI. Their formation may be a

347 result of the incorporation of <@ (Eq. (13)) into fluoridated intermediates (RF@ as follows.46

348 RF@ + <@ H RFI (37)

18

ACS Paragon Plus Environment


Page 19 of 53 Environmental Science & Technology

349 Similarly, in the UV/sulfite system, shorter chain-length fluorinated alkyl sulfonates are also

350 observed, such as C7F15SO3-, C6F13SO3-, C5F11SO3-, C4F9SO3-, and C3F7SO3-,48 likely due to the

351 incorporation of SO3-@ generated from Eq.(1).

352 PFOS. PFOS degradation with eaq- occurs through mechanisms different from the

353 decomposition of PFOA owing to the presence of a different head group (i.e. sulfonic group).46,

354 52, 55-57, 59 Two types of intermediate PFOS degradation products are identified, including 1)

355 short-chain fully and partially fluorinated PFSAs; and 2) PFOA and short-chain fully and

356 partially fluorinated PFCAs (e.g. perfluoroheptanoic acid (PFHpA), perfluorohexanoic acid

357 (PFHxA), perfluoropentanoic acid (PFPeA), and perfluorobutanoic acid (PFBA)). 52, 56, 57 Other

358 degradation products are also observed due to the incorporation of <@ or SO3-@, Particularly in a

359 UV/iodide system, more diverse iodinated gaseous products are produced from the degradation

360 of PFOS than the decomposition of PFOA.46 Similar to the intermediate products of PFOA

361 degradation, these daughter chemicals gradually build up and subsequently diminish in an ARP

362 system as the reactions proceed.52, 55, 59 Formation of the different intermediate degradation

363 products implies that the PFOS decomposition with eaq- occurs through multiple competitive

364 reaction mechanisms.

365 Three reaction pathways have been proposed to explain the reductive destruction of

366 PFOS with e-aq, including desulfonation, H/F exchange, and chain shortening via direct C-C

367 cleavage (Figure 2). 52, 55, 57, 59 During desulfonation, the attachment of eaq- breaks down the C-S

368 bond between the head group and perfluoroalkyl chain. The C-S scission relatively readily

369 occurs (Eq. (38)-(40)) because C-S (272 kJ/mol) in PFOS has a lower bond energy than C-C

370 (346 kJ/mol). Moreover, the distance between C and S atoms (4.463 Å) is longer than that of C-

19

ACS Paragon Plus Environment


Environmental Science & Technology Page 20 of 53

371 C (1.529-1.627 Å) or S-O (1.651Å) on PFOS.104 DFT calculations also indicate the C-S bond

372 stretching after the formation of @ CnF2n+1SO32- (n=8) in Eq. (38), suggesting dissociation of the

373 sulfonate group and the perfluoroalkyl chain.59

374 CnF2n+1SO3- + eaq- H @ CnF2n+1SO32- (38)

375 However, two slightly different dissociation routes have been proposed to describe the following

376 C-S scission. In the first one, the desulfonation leads to the formation of CnF2n+1- (Eq.(39))

377 followed by the production of C8F17OH (Eq. (40)), which would be subsequently transformed to

378 PFOA (n=8) via Eq. (34) and (35).52

379 @ CnF2n+1SO32- H CnF2n+1- + SO3@- (39)

380 CnF2n+1- + H2O+ H CnF2n+1OH + @7 (40)

381 In the other pathway, the dissociation produces @ CnF2n+1 (Eq.(41)), which can be transformed to

382 PFOA (n=8) through Eq. (33)-(35).59

383 @ CnF2n+1SO32- H @'nF2n+1 + SO32- (41)

384 The second desulfonation pathway may be more plausible because the alkaline condition for the

385 ARP treatment disfavors the occurrence of Eq.(40). Although desulfonation does not directly

386 eliminate F or shorten the C-C chain, H/F exchange and chain shortening can take place during

387 the subsequent reductive destruction of PFOA, as discussed above.

388 H-F exchange represents another PFOS degradation pathway, which can eliminate F but

389 cannot break down C-C bonds. Besides the occurrence of the aforementioned H/F exchange on

390 the PFCAs (the PFOS degradation products), H/F exchange may happen on relatively weak C-F

391 bonds of PFOS after the attachment of eaq-.46, 55, 59 The eaq- attachment preferentially attacks one

20

ACS Paragon Plus Environment


Page 21 of 53 Environmental Science & Technology

392 of the centermost C-F bonds, producing PFOS raidcals.46, 52, 105 Recent DFT calculations show

393 that the two center ones on PFOS have the lowest bond dissociation energy (BDE) (446.31

394 E@. -1) among all the C-F bonds, implying that the C-F cleavages preferentially take place in

395 the middle –CF2– chain.59 One H/F exchange plausibly proceeds successively through Eq. (42)

396 and (43),55 in a manner similar to H/F exchange for PFOA (Eq. ((28) and (29)) except for the

397 different C-F locations.

398 CnF2n+1SO3- + eaq- H @ CnF2n+1SO32-H @ CnF2nSO3- + F- (42)

399 @ CnF2nSO3- + H2O H CnF2nHSO3- + @ OH (43)

400 Cyclic H/F exchanges can lead to the formation of multiple H/F exchange products, depending

401 on the C-F BDE.59

402 The third degradation mechanism is direct cleavage of C-C bonds on PFOS. Lowest

403 unoccupied molecular orbitals (LUMO) of PFOS are situated on the moiety of C4 - C8 atoms in

404 the perfluoroalkyl chain, which exhibits a sigma antibonding nature.52 Following the attachment

405 of eaq- on the middle – CF2 –, the binding between the C and C atoms would be attenuated,

406 facilitating the breakdown of these C-C bonds. Particularly, from a thermodynamic point of

407 view, the formation of C3F7-, C4F9-, and C5F11- is favorable because these dissociation fragments

408 of PFOS have the lowest relative energies W= with respect to 5 @,52 For example, two

409 pathways are proposed to explain plausible occurrence of reductive C-C scission on C5. The first

410 C-C scission mechanism begins with the attachment of eaq- on the middle – CF2 – to produce

411 CF3(CF2)2CF2 X –(CF2)4SO3-, which then complexes with a proton followed by C-C scission due

412 to an intramolecular e transfer (Eq (44)).46

413 F F F F
21

ACS Paragon Plus Environment


Environmental Science & Technology Page 22 of 53

414 Y Y Y Y
415 CF3(CF2)2C @ Z C (CF2)3SO3 + H H CF3(CF2)2'Z7 + @ C(CF2)3SO3-
– - + (44)
416 Y Y Y Y
417 F F F F
418
419 However, the contribution of Eq. (44) to the overall chain shortening is insignificant because of

420 limited availability of protons at an alkaline condition typically used for ARP treatment of PFAS.

421 The other reaction pathway firstly proceeds through F elimination (Eq.(42)) to produce

422 @'nF2nSO3- , which subsequently transforms to a carbanion (C8F16SO32-) after eaq- attachment

423 (Eq. 45).52

424 @ CnF2nSO3- + eaq- H CnF2nSO32- (45)

425 Thereafter, the carbanion is dissociated due to reductive cleavage of C-C to generate a shorter

426 chain carbanion CF3CF2CF2- (i.e. C3F7- fragment) and an olefin (Eq.(46)).46, 52

427

428

429 F F F F
430 Y Y Y Y
431 CF3CF2'Z'Z' -Z CF2(CF2)2SO3- H CF3CF2C - + C [ C(CF2)3SO3- (46)
432 Y Y Y Y Y Y
433 F F F F F F
434

435 Carbanion C3F7- would recombine @'557 from the PFOA degradation to generate PFBA. The

436 reaction pathway is indirectly evidenced from the abundance of PFBA observed among the

437 PFOS degradation prodcuts.52 Or the carbanion reacts with H2O+ (Eq.(40)), followed by F

438 elimination (Eq.(34)) and hydrolysis (Eq.(35)) to produce pentafluoropropionic acid (PFPrA).

22

ACS Paragon Plus Environment


Page 23 of 53 Environmental Science & Technology

ACS Paragon Plus Environment


Environmental Science & Technology Page 24 of 53

ACS Paragon Plus Environment


Page 25 of 53 Environmental Science & Technology

446 Other PFAS. ARP destruction of other PFAS chemicals in water have also been

447 investigated, including shorter chain PFCAs and PFSAs, per- and polyfluoro dicarboxylic acids

448 (PFdiCAs, HOOC-CnF2n-COOH), and fluorotelomer carboxylic acids (FTCAs, CnF2n+1-CH2CH2-

449 COOH), in addition to PFOA and PFOS alternatives (e.g. GenX and F-53B).46, 58, 59, 106-108

450 Generally, degradation behaviors of the PFAS chemicals rely heavily upon their head groups.

451 For specific PFAS types, the degradation patterns may be also influenced by lengths of

452 fluoroalkyl chain.46, 59 Below are an in-depth discussion on hydrated electron reductive

453 degradation of other PFCAs and PFSAs, PFdiCAs, and FTCAs in water.

454 Degradation of other PFCAs (n ^ 2) in an ARP system proceeds through the two

455 aforementioned competitive PFOA decomposition mechanisms (i.e. H/F exchange and chain

456 shortening via DHEH). Their decomposition rates are typically greater than the rates of F-

457 release.59 Moreover, both degradation and defluorination rates of these PFCAs (n ^ 2) appear

458 independent of their chain lengths. Park et al.46 found that three PFCAs (n = 4, 6, and 8) had

459 similar pseudo-first-order rate constants (~ 1.3×10-3 min-1) in a UV/iodide system and their F

460 indexes slightly varied between 1 and 2, suggesting a chain length-independent degradation

461 mechanism. And Bentel et al.59 reported that the overall defluorination ratios of PFCAs (n = 2-

462 10) ranged narrowly within 49.1-58.2% in a UV/sulfite system.

463 In contrast, different mechanisms for reductive degradation of TFA (n = 1 PFCA) with

464 eaq- have been proposed.47, 59 One viewpoint ascribes the decomposition of TFA to two parallel

465 pathways:47 1) sequential H/F exchange (Eq. (28) and (29)) that successively produces

466 difluoroacetic acid (DFA, CF2HCOOH) and monofluoroacetic acid (MFA, CFH2COOH), finally

467 generating acetic acid (CH3COOH); and 2) the direct cleavage of C-C leads to the formation of

25

ACS Paragon Plus Environment


Environmental Science & Technology Page 26 of 53

468 fluoroform (CHF3) and CO2, and/or generates two fragments (i.e. @' 3 and @'557 2 which react

469 with water to form fluoroform and formic acid, respectively, or @' 3 of which recombines with

470 itself to generate hexafluoroethane (C2F6). Acetic acid, formic acid, fluoroform, and

471 hexafluoroethane are identified among the intermediate products of eaq- -initiated degradation of

472 PFOA, which can be stepwise degraded to TFA.47, 49

473 However, the aforementioned TFA degradation pathways are not supported by

474 experimental evidence from direct UV/sulfite degradation of TFA, DFA, and MFA.59 ARP is

475 observed to achieve complete defluorination for TFA, but not for DFA or MFA, thereby ruling

476 out the principle role of the sequential H/F exchange pathway. Meanwhile, the rates of TFA

477 degradation and defluorination are nearly synchronized, indicating that F- is rapidly released

478 accompanied with TFA degradation. The observation does not support the aforementioned direct

479 C-C cleavage route, because the alteration of TFA structure cannot cause an immediate release

480 of F-. Instead, the two unique properties of the TFA degradation, i.e. 100% defluorination and

481 the synchronized degradation and defluorination patterns, suggest that the DHEH route is a

482 plausibly dominant mechanism, which finally leads to the production of inorganic F-, H2O, and

483 CO2 (Figure 1).59

484 However, degradation and defluorination rates of different PFSAs are noticeably reduced

485 with the decreasing chain length.46, 59 For example, in a UV/iodide system, the decay rate of

486 perfluorobutane sulfonate (PFBS) was reported to be only 13% of the PFOS decomposition rate

487 under identical experimental conditions.46 At n = 1, the decay and F- release of CF3SO3- are

488 almost marginal.59 The chain-length reliance is ascribed to the increasing strengths of primary

489 and secondary C-F bonds with the decreasing chain length. During the initial defluorination

26

ACS Paragon Plus Environment


Page 27 of 53 Environmental Science & Technology

490 phase, F- release preferentially occurs from the middle – CF2 – where extra electrons tend to

491 attach.

492 Reductive destruction of three other PFAS including PFdiCAs, FTCAs, and GenX in

493 water has also been studied.58, 59 PFdiCAs have two – COO- on the two chain ends, different

494 from corresponding PFCAs (with the same perfluoroalkyl chain length) having a – COO- on one

495 end and a recalcitrant – CF3 on the other. One more relatively weak link between the

496 perfluoroalkyl chain and – COO- on PFdiCA favors its degradation via H/F exchange and/or

497 DHEH mechanisms. Bentel et al.59 found that PFdiCAs were more rapidly decomposed than

498 corresponding PFCAs (n=3-10) in a UV/sulfite system. Moreover, they observed very similar

499 overall deF% (~67%) of these compounds, indicating that defluorination of PFdiCAs is

500 independent of their chain lengths.

501 Reductive degradation of FTCAs was investigated in detail.59 The presence of – CH2CH2

502 – considerably increases chemical persistence to eaq- attack, so that FTCAs are more recalcitrant

503 to reductive degradation or defluorination than the PFCAs with the same perfluoroalkyl chain

504 length.59 As their perfluoroalkyl chain lengths decrease, reactivity of FTCAs is lowered.

505 Consequently, significant degradation can only be observed for n ^ 6 FTCAs. Similar to the

506 degradation of PFSAs, the decay of FTCAs is believed to firstly occur on the middle C-F bonds

507 with relatively low BDEs via H/F exchange. The assumption is supported by identification of

508 H/F exchange degradation products of FTCAs. In the degradation product analysis, shorter-chain

509 PFCAs are also found in low abundance, implying that dissociation of the head group

510 simultaneously plays a role in the FTCA degradation.59

511

27

ACS Paragon Plus Environment


Environmental Science & Technology Page 28 of 53

512 FACTORS AFFECTING ARP DESTRUCTION OF PFAS

513 Hydrated electron–driven reductive degradation has proven highly effective for chemical

514 destruction of PFAS in water. Nearly complete degradation of various PFAS can be

515 accomplished at appropriate operating conditions.47-49, 52, 53, 56, 57, 63, 64 Key factors affecting the

516 PFAS decomposition behaviors are discussed as follows.

517 Solution pH. Initial pH can significantly influence PFAS reductive degradation and

518 defluorination with eaq-. Meanwhile, pH may evolve as the reaction proceeds, depending on the

519 degradation extent and solution buffer capacity. The pH range studied in literature broadly varies

520 within 2.4 - 11.48, 50, 53, 55, 56 Experimental evidence unanimously indicates that an increasing pH

521 favors the reductive destruction, and thus the optimal pH for reductive degradation of PFAS falls

522 within an alkaline range.

523 The enhancement effect of an alkaline pH is principally because less H+ scavenges eaq-

524 (Eq. (4)-(5)) with the increasing pH, so that eaq- can be more efficiently utilized for the

525 degradation of PFAS.60 Based on Eq.(4), a pH increase by 1.0 means that the reaction rate of H+

526 scavenging by eaq- is decreased by one order of magnitude. Therefore, a slight pH change may

527 noticeably alter the PFAS degradation rate.

528 In a UV/NTA system, the presence of NTA can greatly eliminate @57 due to its high

529 reactivity toward @572 thereby promoting the utilization efficiency of eaq- for destruction of

530 PFAS.56 Distributions of successively deprotonated NTA species with the pKa values of 0.8, 1.9,

531 2.48, and 9.65 depend on solution pH. Because a fully deprotonated NTA is more reactive

532 toward @57 than its fully or partially protonated forms, an alkaline solution would benefit the

533 reactions of NTA and @57 to reduce the consumption of @57 for eaq-. 56
28

ACS Paragon Plus Environment


Page 29 of 53 Environmental Science & Technology

534 However, in an ARP system using photo-irradiation of indole derivatives with

535 organomodified clay, the inhibiting effect of pH decrease (3.0-11.0) on the PFAS degradation

536 was reported to be insignificant.64 The pH-independence is ascribed to a unique micro-

537 environment on the modified clay, where protons in the clay interlayers are appreciably replaced

538 by large modification organic cations and eaq- is generated in the vicinity of sorbed PFAS.

539 On the other hand, solution pH may change as the PFAS degradation proceeds in an ARP

540 system with a limited solution pH buffering capacity.46, 55, 57 Various pH evolution patterns have

541 been reported in literature, likely because the pH variation is simultaneously controlled by

542 multiple factors. A major reason for the pH increase is the H+ consumption by eaq- (Eq.(4)).

543 Moreover, when direct photolysis (e.g. VUV irradiation) generates @57 in the presence of sulfite,

544 @57 can react with sulfite to generate OH- and increase pH (Eq.(47)).79

545 @57 + SO32- H OH- + SO3@ - k47 = 4.5 × 109 M-1@ -1 (47)

546 For example, in a UV/sulfite system (2.4 g/L sulfite; UV irradiation: 500 W and 365 nm; and

547 400 mL solution), pH at an initial level of 7.00 was reported to consistently increase by 0.16 and

548 0.22 during degradation of PFOA (20 mg/L) and PFOS (20 mg/L), respectively.57

549 In contrast, other reasons can contribute to a pH decrease. For example, a rapid F- release

550 from PFAS defluorination can lead to form a weak acid HF (pKa = 3.45) in water, thus reducing

551 solution pH.46, 55 In a UV/iodide solution (10 mM iodide; UV irradiation: 8 W and 254 nm; and

552 30 mL solution) for the destruction of PFOS (24 µM), pH was noticed to initially increase and

553 then decrease, exhibiting a unique bell-shaped pH curve.46

554 Besides the aforementioned reasons, others such as concentrations and species of

555 intermediate degradation products generated over the reactions likely influence the activity of H+
29

ACS Paragon Plus Environment


Environmental Science & Technology Page 30 of 53

556 in water. Consequently, the final pH change patterns may be very complex, depending on

557 specific operational conditions. Gu et al.55 treated PFOS with VUV/sulfite (37.2 µM PFOS; pH

558 10; UV irradiation: 10 W and 185 nm; and 800 mL solution) using a broad range of sulfite dose

559 (1-20 mM). A consistent pH decrease was observed at 1 mM sulfite, whereas the pH exhibited an

560 initial decrease followed by an increase for other doses, at which the pH increase extent was

561 increased with the increasing sulfite dose.

562 Solute dose. Dose of the chemical solute for photo-production of eaq- is essential to the

563 reductive degradation patterns of PFAS because it is directly related to the yield of eaq-. When

564 SO32- or I- is used, the PFAS degradation and defluorination efficiencies are typically increased

565 with an increasing solute dose until a critical level, beyond which the PFAS destruction

566 efficiencies would decrease as the dose further increases.47, 48, 57 For example, Sun et al.57

567 reported that the degradation efficiencies of five PFAS all increased with the SO32- dose, peaked

568 almost at 100% at 2.4 g/L SO32-, and decreased, to different degrees, as the SO32- concentration

569 increased to 4.0 g/L (20 mg/L PFDA, PFNA, PFOA, PFBA, or PFOS; pH 7.0; UV irradiation:

570 500 W and 365 nm; and 400 mL solution). And Qu et al.47 found that the initial PFOA

571 decomposition rate increased with an increase in the I- concentration within [I-]:[PFOA] = 0 – 12,

572 but declined as the I- concentration increased at [I-]:[PFOA] = 12-28 (25 µM PFOA; 0.0-0.7 mM

573 I-; pH 9.0; UV irradiation: 15 W and 254 nm; and 740 mL solution).

574 The positive dependence of solute dose observed within a low solute dose range is

575 ascribed to the fact that more solutes for photo-excitation can translate into more eaq- through Eq.

576 (1) or Eq. (10)-(13), provided that sufficient photons are available. For example, the relative

577 quasi-stationary concentration (RQSC) of eaq- in a UV/SO32-/N2 system was concluded to reveal

30

ACS Paragon Plus Environment


Page 31 of 53 Environmental Science & Technology

578 that RQSC is approximately linearly increased with the increasing sulfite dose at pH> 9.48

579 However, the potential scavenging effects for eaq- can also be gradually enhanced with the

580 increase in the solute dose. In a UV/sulfite system, eaq- can be increasingly consumed with the

581 increasing sulfite dose through the enhanced self-recombination of eaq- (Eq.(6)) and/or reaction

582 with intermediate S2O62- (Eq.(7)). And for a UV/iodide system, an increasing I- dose can enhance

583 the reactions of eaq- with different I species (Eq.(20) and (21)), besides its self-recombination.

584 Once the solute dose reaches a specific level beyond which the scavenging begins to overweigh

585 the eaq- generation, the negative dependence of the solute dose would become dominant.

586 DO. The presence of DO can substantially decrease rates and efficiencies of PFAS

587 degradation and defluorination in the most ARP systems.46-48, 50, 55 The negative impact is

588 principally attributed to the scavenging effects of O2 (Eq.(8) and (9)) .60 Moreover, DO is

589 capable of adsorbing part of UV photons to reduce the yield of eaq- and thus compromise the

590 PFAS decomposition.50, 109 Of note, the inhibiting extent of PFAS decomposition due to DO is

591 typically substantial. For example, in a UV/sulfite system for PFOA degradation (20 µM PFOA;

592 10 mM sulfite; UV irradiation: 10 W and 254 nm; 200 mL solution; and pH 10.3), an overall

593 defluorination ratio was reported to considerably decline from 88.5% in a N2 purged solution to

594 6.4% in an air-aerated solution within 24 hours.48 These findings suggest that effective

595 destruction of PFAS mostly occurs in a strictly oxygen deficient condition.

596 Efforts have been made to mitigate the negative DO impact. A strategy is the use of a

597 high photon flux UV irradiation. Gu et al.52 found that PFOS remained a rapid decomposition at

598 an initial 5 mg/L DO in an open UV/sulfite system (9.93 × 10-8 einstein/cm2@ using a high

599 pressure mercury UV lamp. The UV lamp provided a higher emission density and a broader

31

ACS Paragon Plus Environment


Environmental Science & Technology Page 32 of 53

600 emission spectrum than low or medium-pressure UV lamps, thus promoting the eaq- quantum

601 yield. Another approach is adoption of appropriate chemical solutes for photo-ionization to

602 minimize the scavenging effect of DO for eaq-. Sun et al.56 noticed a slight difference of PFOS

603 degradation and defluorination at oxic and anoxic conditions in a UV/NTA treatment, because

604 NTA and UV-excited NTA (NTA*) could rapidly react with O2 and different reactive oxygen

605 species (ROS) (e.g. O2@ -) produced in the presence of O2. Consequently, the quenching of eaq- by

606 O2 or ROS was lessened.95, 110, 111

607 Water matrix constituents. Water matrix constituents co-existing with PFAS may

608 significantly affect reductive destruction of PFAS in water through different mechanisms. Two

609 constituents extensively studied are humic acid (HA) and nitrate (NO3-).

610 Humic acid. Dissolved organic matter (DOM) is ubiquitously present in different PFAS-

611 polluted water matrixes, such as natural organic matter (NOM) in surface freshwater,112-114

612 effluent organic matter (EfOM) in biologically treated municipal wastewater,115, 116 soil organic

613 matter (SOM) in groundwater, and leachate organic matter (LOM) in landfill leachate.117-119

614 Because DOM is typically much more abundant than PFAS in polluted water, its impact on ARP

615 destruction of PFAS is of interest. Although various organic types are comprised in aquatic

616 DOM, only the effect of HA, a primary hydrophobic DOM fraction, was investigated in

617 literature,53, 55, 98 among which in-depth studies for evaluation of the role of HA were

618 implemented in a UV/iodide system, rather than for a UV/sulfite ARP.

619 For a UV/iodide ARP treatment, enhanced and inhibiting effects of HA on the PFAS

620 decomposition are observed at different HA concentrations,53, 98 suggesting that multiple

32

ACS Paragon Plus Environment


Page 33 of 53 Environmental Science & Technology

621 competitive mechanisms govern the influence of HA. Experimental evidence reveals that HA

622 can accelerate the PFAS degradation with the increasing HA concentration, but the

623 decomposition is slowed down once HA is beyond a specific concentration. At a too high HA

624 concentration, the PFAS degradation kinetic rate is even below that in the absence of HA.

625 Therefore, an optimal HA appears to exist to maximize the rate of PFAS decomposition in a

626 UV/iodide system.53 Of note, though HA may considerably alter the kinetic patterns of PFAS

627 degradation, it may not accordingly change the final PFAS degradation efficiency. On the other

628 hand, HA may exhibit different impacts on the defluorination behaviors. In a UV/iodide system

629 (30 µM PFOA; 0.3 mM iodide; UV irradiation: 14 W and 254 nm; and pH 10.0), rapid PFOA

630 degradation and defluorination were noticed in the presence of HA (1.0 mg/L) only within the

631 initial 3 hours. Thereafter, the PFOA decomposition and defluorination efficiencies with time

632 were independent of the initial HA concentration.98 In another set of PFOS degradation

633 experiments with the nearly identical operational conditions,53 HA (0.0-30.0 mg/L) played a

634 more complicated role. HA at 1.0 and 30.0 mg/L significantly promoted and inhibited the initial

635 PFOS decomposition rate, respectively. However, the greatest defluorination efficiency (almost

636 100%) was finally achieved at 30.0 mg/L HA.53

637 Four mechanisms governing HA-enhanced PFAS degradation are proposed for the

638 UV/iodide system. The principal enhancement pathway is that HA reacts with oxidizing I-

639 containing intermediates (e.g. I2, HOI, IO3- and I3-) to again produce I-, which is subsequently

640 photo-activated to generate additional eaq-.49, 53 In the second mechanism, specific functional

641 groups on HA (e.g. quinone moieties) may serve as an electron shuttle,120, 121 thereby mediating

642 electron transfer between I- and PFAS. The third one is associated with the confinement effect of

33

ACS Paragon Plus Environment


Environmental Science & Technology Page 34 of 53

643 HA, in which HA bridges I- and PFAS to form I- -HA-PFAS adducts.53 The adducts can enable

644 the occurrence of ensuing eaq- formation and PFAS decomposition in a local region, so that the

645 quenching of eaq- by others becomes minimal and the diffusion distance for eaq- to PFAS is

646 shortened. And the fourth mechanism is the direct formation of extra eaq- from photo-ionization

647 of HA in water.122 However, contribution of this pathway is minor due to its low quantum yield

648 of eaq-.53

649 Meanwhile, three plausible mechanisms can inhibit the reductive degradation of PFAS in

650 the presence of HA. Firstly, an inhibition can be potentially caused by the UV quenching

651 property of HA. UV blocking effects of HA fractions in NOM, EfOM, and LOM have been

652 reported elsewhere.119, 123, 124 Competitive absorption of HA with SO32- or I- for photons can

653 directly alleviate the yield of eaq-. Secondly, the HA absorption of UV can generate various

654 reactive species (e.g. @572 1O2, H2O2, and excited triplet state DOM (DOM*)), which

655 ineffectually decompose PFAS, but can oxidize the generated eaq-.50 Thirdly, the presence or

656 absence of specific moieties may inhibit the reductive decomposition of PFAS. Electron

657 withdrawing groups (EWGs) (e.g. carboxylic acids on aromatic rings) on HA preferentially react

658 with the generated eaq-.125 And the scarcity of quinone moieties would lessen electron transfer

659 mediation. Both of them disfavor the reductive destruction of PFAS.

660 A net effect of HA on the PFAS degradation in the UV/iodide system relies upon

661 competition among the aforementioned enhancement and inhibiting mechanisms. For the eaq-

662 reductive decomposition of PFAS, the high molecular weight (MW) HA faction with more

663 electron donating groups (EDGs) and electron transfer mediators exhibits an overall

34

ACS Paragon Plus Environment


Page 35 of 53 Environmental Science & Technology

664 enhancement effect, whereas the low MW molecules characterized with abundant EWGs and

665 saturated aliphatic moieties marginally promote or even inhibit the PFAS degradation.53

666 Nitrate. NO3- is a common surface water and groundwater solute. NO3--N concentrations

667 in the U.S. groundwater are mostly below 3 mg/L,126 while its level in surface freshwater without

668 nutrient pollution is typically less than 1 mg/L.127 Gu et al. 55 observed that PFAS degradation

669 was slowed down with the increasing NO3- level (0.0-0.5 mM) in a VUV/sulfite system (37.2

670 µM PFOS; pH 10; UV irradiation: 10 W and 185 nm; and 800 mL solution). However, a

671 significant difference among the final PFOS degradation efficiencies was not observed at 0.0-0.5

672 mM NO3-, unless the NO3- was increased to 1.0 mM. The NO3- suppression of PFAS

673 decomposition is due to its rapid reactions with eaq- as follows.128

674 NO3- + eaq- H (NO3 @ 2- k48 = 1.0 × 1010 M-1@ -1 (48)

675 (NO3 @ 2- + H2O M (NO3H)@ - + OH- pK49 K 7.5 (49)

676 (NO3H)@ - H NO2@ + OH- (50)

677 Temperature. Hydrated electron destruction of PFAS is also influenced by temperature.

678 The reductive degradation of PFAS is improved with an increasing temperature,50, 51 likely due to

679 the growing collision frequency of molecules in water as the temperature increases. Pseudo-

680 first-order rate constants of PFAS degradation or defluorination are reported to fit the Arrhenius

681 equation at different temperature ranges (20-40oC for PFOA;51 and 35-100oC for PFOS50).

682 Moreover, the same intermediate degradation products are identified in the UV/iodide

683 degradation of PFOA at different temperatures, suggesting that the temperature change cannot

684 alter the pathways of PFAS degradation.51 However, the time required for a specific intermediate

35

ACS Paragon Plus Environment


Environmental Science & Technology Page 36 of 53

685 product to reach its maximum concentration declines with the increasing temperature, because

686 intermediate products become more reactive at a higher temperature and thus tend to more

687 promptly react with eaq- rather than accumulate in the system. 51

688 IMPLICATIONS AND THE WAY FORWARD

689 Implication. Destruction of aqueous of PFAS through eaq- -driven chemical reduction

690 provides a technically viable treatment approach to mitigation of the PFAS pollution for water

691 and wastewater industries. This approach is capable of effectively degrading a broad range of

692 PFAS molecules as compared to many other chemical oxidative or reductive methods, offering a

693 promising ultimate solution to the PFAS impacts. Unique nature of the advanced chemical

694 reductive process has profound implications for its application in realistic treatment.

695 Firstly, effluent total dissolved solids (TDS) can influence how ARPs are applied to the

696 treatment of different PFAS-polluted water matrixes. TDS would be inevitably increased after

697 the ARP treatment due to the formation of inert salt residuals, such as sulfate and iodide. In the

698 literature, a typical dose of sulfite or iodide for the generation of eaq- for PFAS treatment ranges

699 from several to a few tens of mM, which leads to a TDS increase by several hundreds to

700 thousands of mg/L. Further TDS increase is expected when acid and/or alkaline is used for pH

701 adjustment. Consequently, the direct application mode is infeasible for drinking water treatment

702 or municipal wastewater reuse, which typically has a strict limit on the effluent salinity. For

703 example, TDS and sulfate in U.S. potable water should be a 500 and 250 mg/L, respectively,

704 according to the U.S. secondary drinking water standards. Instead, under the circumstances,

705 indirect application of ARPs to treatment of the water or wastewater may be considered. One

36

ACS Paragon Plus Environment


Page 37 of 53 Environmental Science & Technology

706 such example is that reverse osmosis (RO) concentrates PFAS into membrane brine, which is

707 later treated with ARPs to remove PFAS accumulated in the brine. For remediating subsurface

708 PFAS-polluted groundwater, TDS is not a limiting factor in most cases. However, in-situ

709 remediation with the ARPs is impossible because direct UV irradiation is barely applied to

710 subsurface. Rather, PFAS-polluted groundwater, after pumped out, can be readily ex-situ treated

711 with ARPs.

712 Secondly, complexity of the system design, operation, and maintenance can be

713 significantly increased due to the pH adjustment and the requirement for an anoxic environment.

714 Solution pH in the most water, particularly drinking water sources and municipal wastewater, is

715 nearly neutral. However, a moderately to highly alkaline pH is preferred to ensure an effective

716 reductive destruction of PFAS, thus requiring a pH adjustment. However, addition of acid and/or

717 base increases the system complexity, effluent TDS, and treatment costs. Another factor

718 inhibiting the PFAS decomposition efficiency is DO, which is ubiquitous in various PFAS-

719 polluted water. To remove dissolved oxygen from water and subsequently maintain a low or no

720 oxygen condition in realistic water and wastewater treatment is challenging, particularly for

721 high-capacity treatment facilities. To control DO in water and wastewater typically requires

722 additional treatment devices and costs. However, DO may not be an issue for treatment of a low-

723 DO water or wastewater, such as landfill leachate and groundwater.

724 Thirdly, the formation of undesirable degradation byproducts is a concern. Various

725 fluoride-containing degradation products may be produced, of which some remain toxic or are

726 greenhouse gases. Furthermore, when I- is adopted for the eaq- generation, different iodinated

727 byproducts have been identified as a result of iodide incorporation. Formation and toxicity of

37

ACS Paragon Plus Environment


Environmental Science & Technology Page 38 of 53

728 iodide-containing byproducts produced from drinking water disinfection have been well

729 documented. 129-133 These compounds generally exhibit more significant mammalian cell

730 cytotoxicity and genotoxicity than their brominated and chlorinated analogues.132

731 Future research needs.

732 Although current studies well demonstrate encouraging results on eaq- destruction of PFAS in

733 water, further investigations are needed, including:

734 1) The dilemma of ARPs for realistic treatment is that eaq- generated is highly reactive

735 toward both target PFAS compounds and many water matrix constituents, which are

736 typically much more abundant than PFAS at trace concentrations. Consequently, the

737 fraction of eaq- allocated for the PFAS degradation is limited, leading to a low treatment

738 efficiency. Therefore, novel approaches to minimizing the scavenging effects of different

739 co-existing species (e.g. H+, DO, and NOM) for eaq- are highly demanded. Particularly,

740 alleviation of the H+ inhibition can reduce the reliance of ARPs on an alkaline pH and

741 thus avoid costly pH adjustment. Another strategy is to improve the yield of eaq- using

742 new and more efficient eaq- generation methods.

743 2) Production and toxicity of PFAS transformation products after the ARP treatment need to

744 be focused. And effects of solution chemistry parameters and operating conditions on

745 formation of these unwanted byproducts deserve in-depth evaluation. Attention should be

746 paid to production of the transformation products produced from the reactions of eaq- with

747 both PFAS and water matrix constituents (e.g. NOM) and to formation of the byproducts

748 due to the incorporation of dosed solutes (e.g. iodide).

38

ACS Paragon Plus Environment


Page 39 of 53 Environmental Science & Technology

749 3) Can ARPs be combined with other physical/chemical processes and even biological

750 methods to build more efficient multiple barriers for PFAS in water? For example, the

751 PFAS reduction products generated from the ARP treatment may be efficiently degraded

752 by certain chemical oxidation processes if abundant reducing functional groups remain

753 on them, or further removed by biological degradation when ARPs improve

754 biodegradability of these compounds.

755 4) Development of cost and energy-efficient technologies with minimal environmental

756 impacts is essential to address the PFAS pollutions. However, it is inappropriate now to

757 carry out techno-economic assessment (TEA) and life-cycle analysis (LCA) for ARPs

758 and compare the technologies with exiting PFAS treatment options, considering that the

759 emerging treatment process is still at the early stage. With the advances in the ARP

760 chemistry as well as the technology development and optimization, the TEA and LCA

761 will be needed at a future and more-developed phase. During these assessments, residual

762 disposal and management will also need to be considered.

763
764
765 AUTHOR INFORMATION

766 Corresponding Author

767 Yang Deng, CELS 220, Department of Earth and Environmental Studies, Montclair State

768 University, Montclair, New Jersey 07043, USA

769 Tel: 01-973-655-6678; Email: [email protected]

770 Notes

39

ACS Paragon Plus Environment


Environmental Science & Technology Page 40 of 53

771 The authors declare no competing financial interest.

772

773 ACKNOWLEDGMENT

774 This project was supported through the Career Development Grant of Montclair State University

775 (MSU). Junkui Cui was sponsored under the Doctor Assistantship of PhD Program in

776 Environmental Science and Management at MSU. Panpan Gao worked at MSU with the support

777 from the MoE 111 Project (No. B18049). Special thanks to MSU Office of International

778 Engagement. We greatly appreciate valuable comments from three anonymous reviewers.

779

780 REFERENCES

781 1. CDC An Overview of Perfluoroalkyl and Polyfluoroalkyl Substances and Interim


782 Guidance for Clinicians Responding to Patient Exposure Concerns; 2017.
783 2. Buck, R. C.; Franklin, J.; Berger, U.; Conder, J. M.; Cousins, I. T.; De Voogt, P.;
784 Jensen, A. A.; Kannan, K.; Mabury, S. A.; van Leeuwen, S. P., Perfluoroalkyl and
785 polyfluoroalkyl substances in the environment: terminology, classification, and origins.
786 Integrated environmental assessment and management 2011, 7, (4), 513-541.
787 3. ATSDR The family tree of per- and polyfluoroalkyl substances (PFAS) for
788 environmental health professionals: Names and abbreviations; 2017.
789 4. ITRC History and Use of Per- and Polyfluoroalkyl Substances (PFAS); 2017.
790 5. Higgins, C. P.; Field, J. A.; Criddle, C. S.; Luthy, R. G., Quantitative
791 determination of perfluorochemicals in sediments and domestic sludge. Environmental
792 science & technology 2005, 39, (11), 3946-3956.
793 6. Higgins, C. P.; Luthy, R. G., Sorption of perfluorinated surfactants on sediments.
794 Environmental Science & Technology 2006, 40, (23), 7251-7256.
795 7. Moody, C. A.; Hebert, G. N.; Strauss, S. H.; Field, J. A., Occurrence and
796 persistence of perfluorooctanesulfonate and other perfluorinated surfactants in

40

ACS Paragon Plus Environment


Page 41 of 53 Environmental Science & Technology

797 groundwater at a fire-training area at Wurtsmith Air Force Base, Michigan, USA. Journal
798 of Environmental Monitoring 2003, 5, (2), 341-345.
799 8. Naile, J. E.; Khim, J. S.; Wang, T.; Chen, C.; Luo, W.; Kwon, B.-O.; Park, J.; Koh,
800 C.-H.; Jones, P. D.; Lu, Y., Perfluorinated compounds in water, sediment, soil and biota
801 from estuarine and coastal areas of Korea. Environmental Pollution 2010, 158, (5),
802 1237-1244.
803 9. Moody, C. A.; Martin, J. W.; Kwan, W. C.; Muir, D. C.; Mabury, S. A., Monitoring
804 perfluorinated surfactants in biota and surface water samples following an accidental
805 release of fire-fighting foam into Etobicoke Creek. Environmental science & technology
806 2002, 36, (4), 545-551.
807 10. USEPA Basic Information on PFAS. https://www.epa.gov/pfas/basic-information-
808 pfas
809 11. Gounaris, V.; Anderson, P. R.; Holsen, T. M., Characteristics and environmental
810 significance of colloids in landfill leachate. Environmental science & technology 1993,
811 27, (7), 1381-1387.
812 12. Braun, J. M.; Chen, A.; Romano, M. E.; Calafat, A. M.; Webster, G. M.; Yolton,
813 K.; Lanphear, B. P., Prenatal perfluoroalkyl substance exposure and child adiposity at 8
814 years of age: The HOME study. Obesity 2016, 24, (1), 231-237.
815 13. Taniyasu, S.; Kannan, K.; Horii, Y.; Hanari, N.; Yamashita, N., A survey of
816 perfluorooctane sulfonate and related perfluorinated organic compounds in water, fish,
817 birds, and humans from Japan. Environmental Science & Technology 2003, 37, (12),
818 2634-2639.
819 14. So, M.; Miyake, Y.; Yeung, W.; Ho, Y.; Taniyasu, S.; Rostkowski, P.; Yamashita,
820 N.; Zhou, B.; Shi, X.; Wang, J., Perfluorinated compounds in the Pearl river and
821 Yangtze river of China. Chemosphere 2007, 68, (11), 2085-2095.
822 15. Nakayama, S. F.; Strynar, M. J.; Reiner, J. L.; Delinsky, A. D.; Lindstrom, A. B.,
823 Determination of perfluorinated compounds in the Upper Mississippi River Basin.
824 Environmental science & technology 2010, 44, (11), 4103-4109.

41

ACS Paragon Plus Environment


Environmental Science & Technology Page 42 of 53

825 16. Murakami, M.; Kuroda, K.; Sato, N.; Fukushi, T.; Takizawa, S.; Takada, H.,
826 Groundwater pollution by perfluorinated surfactants in Tokyo. Environmental science &
827 technology 2009, 43, (10), 3480-3486.
828 17. Houtz, E. F.; Higgins, C. P.; Field, J. A.; Sedlak, D. L., Persistence of
829 perfluoroalkyl acid precursors in AFFF-impacted groundwater and soil. Environmental
830 science & technology 2013, 47, (15), 8187-8195.
831 18. Xiao, F.; Simcik, M. F.; Halbach, T. R.; Gulliver, J. S., Perfluorooctane sulfonate
832 (PFOS) and perfluorooctanoate (PFOA) in soils and groundwater of a US metropolitan
833 area: Migration and implications for human exposure. Water research 2015, 72, 64-74.
834 19. Barzen-Hanson, K. A.; Roberts, S. C.; Choyke, S.; Oetjen, K.; McAlees, A.;
835 Riddell, N.; McCrindle, R.; Ferguson, P. L.; Higgins, C. P.; Field, J. A., Discovery of 40
836 classes of per-and polyfluoroalkyl substances in historical aqueous film-forming foams
837 (AFFFs) and AFFF-impacted groundwater. Environmental science & technology 2017,
838 51, (4), 2047-2057.
839 20. Mak, Y. L.; Taniyasu, S.; Yeung, L. W.; Lu, G.; Jin, L.; Yang, Y.; Lam, P. K.;
840 Kannan, K.; Yamashita, N., Perfluorinated compounds in tap water from China and
841 several other countries. Environmental science & technology 2009, 43, (13), 4824-4829.
842 21. Hu, X. C.; Andrews, D. Q.; Lindstrom, A. B.; Bruton, T. A.; Schaider, L. A.;
843 Grandjean, P.; Lohmann, R.; Carignan, C. C.; Blum, A.; Balan, S. A., Detection of poly-
844 and perfluoroalkyl substances (PFASs) in US drinking water linked to industrial sites,
845 military fire training areas, and wastewater treatment plants. Environmental science &
846 technology letters 2016, 3, (10), 344-350.
847 22. Guelfo, J. L.; Adamson, D. T., Evaluation of a national data set for insights into
848 sources, composition, and concentrations of per-and polyfluoroalkyl substances
849 (PFASs) in US drinking water. Environmental pollution 2018, 236, 505-513.
850 23. Busch, J.; Ahrens, L.; Sturm, R.; Ebinghaus, R., Polyfluoroalkyl compounds in
851 landfill leachates. Environmental Pollution 2010, 158, (5), 1467-1471.

42

ACS Paragon Plus Environment


Page 43 of 53 Environmental Science & Technology

852 24. Huset, C. A.; Barlaz, M. A.; Barofsky, D. F.; Field, J. A., Quantitative
853 determination of fluorochemicals in municipal landfill leachates. Chemosphere 2011, 82,
854 (10), 1380-1386.
855 25. Benskin, J. P.; Li, B.; Ikonomou, M. G.; Grace, J. R.; Li, L. Y., Per-and
856 polyfluoroalkyl substances in landfill leachate: patterns, time trends, and sources.
857 Environmental science & technology 2012, 46, (21), 11532-11540.
858 26. Yan, H.; Cousins, I. T.; Zhang, C.; Zhou, Q., Perfluoroalkyl acids in municipal
859 landfill leachates from China: Occurrence, fate during leachate treatment and potential
860 impact on groundwater. Science of the Total Environment 2015, 524, 23-31.
861 27. Lang, J. R.; Allred, B. M.; Field, J. A.; Levis, J. W.; Barlaz, M. A., National
862 estimate of per-and polyfluoroalkyl substance (PFAS) release to US municipal landfill
863 leachate. Environmental science & technology 2017, 51, (4), 2197-2205.
864 28. USEPA EPA’s Per- and Polyfluoroalkyl Substances (PFAS) Action Plan (EPA
865 823R18004); 2019.
866 29. Gehringer, P.; Eschweiler, H., Ozone/electron beam process for water treatment:
867 design, limitations and economic considerations. Ozone: science & engineering 1999,
868 21, (5), 523-538.
869 30. Yoon, S. H.; Abdel-Wahab, A.; Batchelor, B. In Advanced Reduction Processes
870 for Hazardous Waste Treatment, Qatar Foundation Annual Research Forum
871 Proceedings, 2011; Bloomsbury Qatar Foundation Journals: 2011; p EVP16.
872 31. Vellanki, B. P.; Batchelor, B.; Abdel-Wahab, A., Advanced reduction processes: a
873 new class of treatment processes. Environmental engineering science 2013, 30, (5),
874 264-271.
875 32. Gordon, S.; Hart, E.; Matheson, M.; Rabani, J.; Thomas, J., Reactions of the
876 hydrated electron. Discussions of the Faraday Society 1963, 36, 193-205.
877 33. Gordon, S.; Hart, E. J.; Matheson, M. S.; Rabani, J.; Thomas, J., Reaction
878 constants of the hydrated electron. Journal of the American Chemical Society 1963, 85,
879 (10), 1375-1377.

43

ACS Paragon Plus Environment


Environmental Science & Technology Page 44 of 53

880 34. Hart, E. J.; Gordon, S.; Thomas, J., Rate Constants of Hydrated Electron
881 Reactions with Organic Compounds1. The Journal of Physical Chemistry 1964, 68, (6),
882 1271-1274.
883 35. Keene, J., The absorption spectrum and some reaction constants of the hydrated
884 electron. Radiation Research 1964, 22, (1), 1-13.
885 36. OZAWA, T.; SETAKA, M.; YAMAMOTO, H.; KWAN, T., On the reaction of the
886 sulfite radical anions with thioureas. Chemical and Pharmaceutical Bulletin 1974, 22,
887 (4), 962-964.
888 37. Liu, X.; Yoon, S.; Batchelor, B.; Abdel-Wahab, A., Degradation of vinyl chloride
889 (VC) by the sulfite/UV advanced reduction process (ARP): Effects of process variables
890 and a kinetic model. Science of the Total Environment 2013, 454, 578-583.
891 38. Liu, X.; Yoon, S.; Batchelor, B.; Abdel-Wahab, A., Photochemical degradation of
892 vinyl chloride with an Advanced Reduction Process (ARP) - Effects of reagents and pH.
893 Chemical Engineering Journal 2013, 215, 868-875.
894 39. Vellanki, B. P.; Batchelor, B., Perchlorate reduction by the sulfite/ultraviolet light
895 advanced reduction process. Journal of Hazardous Materials 2013, 262, 348-356.
896 40. Jung, B.; Nicola, R.; Batchelor, B.; Abdel-Wahab, A., Effect of low- and medium-
897 pressure Hg UV irradiation on bromate removal in advanced reduction process.
898 Chemosphere 2014, 117, 663-672.
899 41. Botlaguduru, V. S. V.; Batchelor, B.; Abdel-Wahab, A., Application of UV-sulfite
900 advanced reduction process to bromate removal. Journal of Water Process Engineering
901 2015, 5, 76-82.
902 42. Bensalah, N.; Nicola, R.; Abdel-Wahab, A., Nitrate removal from water using UV-
903 M/S 2 O 4 M advanced reduction process. International Journal of Environmental
904 Science and Technology 2014, 11, (6), 1733-1742.
905 43. Moussavi, G.; Jiani, F.; Shekoohiyan, S., Advanced reduction of Cr(VI) in real
906 chrome-plating wastewater using a VUV photoreactor: Batch and continuous-flow
907 experiments. Separation and Purification Technology 2015, 151, 218-224.

44

ACS Paragon Plus Environment


Page 45 of 53 Environmental Science & Technology

908 44. Xie, B.; Shan, C.; Xu, Z.; Li, X.; Zhang, X.; Chen, J.; Pan, B., One-step removal
909 of Cr (VI) at alkaline pH by UV/sulfite process: reduction to Cr (III) and in situ Cr (III)
910 precipitation. Chemical Engineering Journal 2017, 308, 791-797.
911 45. Yazdanbakhsh, A.; Eslami, A.; Moussavi, G.; Rafiee, M.; Sheikhmohammadi, A.,
912 Photo-assisted degradation of 2, 4, 6-trichlorophenol by an advanced reduction process
913 based on sulfite anion radical: Degradation, dechlorination and mineralization.
914 Chemosphere 2018, 191, 156-165.
915 46. Park, H.; Vecitis, C. D.; Cheng, J.; Choi, W.; Mader, B. T.; Hoffmann, M. R.,
916 Reductive defluorination of aqueous perfluorinated alkyl surfactants: effects of ionic
917 headgroup and chain length. The Journal of Physical Chemistry A 2009, 113, (4), 690-
918 696.
919 47. Qu, Y.; Zhang, C.; Li, F.; Chen, J.; Zhou, Q., Photo-reductive defluorination of
920 perfluorooctanoic acid in water. Water research 2010, 44, (9), 2939-2947.
921 48. Song, Z.; Tang, H.; Wang, N.; Zhu, L., Reductive defluorination of
922 perfluorooctanoic acid by hydrated electrons in a sulfite-mediated UV photochemical
923 system. Journal of hazardous materials 2013, 262, 332-338.
924 49. Qu, Y.; Zhang, C.-J.; Chen, P.; Zhou, Q.; Zhang, W.-X., Effect of initial solution
925 pH on photo-induced reductive decomposition of perfluorooctanoic acid. Chemosphere
926 2014, 107, 218-223.
927 50. Lyu, X.-J.; Li, W.-W.; Lam, P. K.; Yu, H.-Q., Insights into perfluorooctane
928 sulfonate photodegradation in a catalyst-free aqueous solution. Scientific reports 2015,
929 5, 9353.
930 51. Zhang, C.; Qu, Y.; Zhao, X.; Zhou, Q., Photoinduced reductive decomposition of
931 Perflurooctanoic acid in water: effect of temperature and ionic strength. CLEAN–Soil,
932 Air, Water 2015, 43, (2), 223-228.
933 52. Gu, Y.; Dong, W.; Luo, C.; Liu, T., Efficient reductive decomposition of
934 perfluorooctanesulfonate in a high photon flux UV/sulfite system. Environmental science
935 & technology 2016, 50, (19), 10554-10561.

45

ACS Paragon Plus Environment


Environmental Science & Technology Page 46 of 53

936 53. Sun, Z.; Zhang, C.; Chen, P.; Zhou, Q.; Hoffmann, M. R., Impact of humic acid
937 on the photoreductive degradation of perfluorooctane sulfonate (PFOS) by UV/Iodide
938 process. Water research 2017, 127, 50-58.
939 54. Gu, J.; Ma, J.; Jiang, J.; Yang, L.; Yang, J. X.; Zhang, J. Q.; Chi, H. Z.; Song, Y.;
940 Sun, S. F.; Tian, W. Q., Hydrated electron (e(aq)(-)) generation from phenol/UV:
941 Efficiency, influencing factors, and mechanism. Applied Catalysis B-Environmental
942 2017, 200, 585-593.
943 55. Gu, Y.; Liu, T.; Wang, H.; Han, H.; Dong, W., Hydrated electron based
944 decomposition of perfluorooctane sulfonate (PFOS) in the VUV/sulfite system. Science
945 of the Total Environment 2017, 607, 541-548.
946 56. Sun, Z.; Zhang, C.; Xing, L.; Zhou, Q.; Dong, W.; Hoffmann, M. R.,
947 UV/nitrilotriacetic acid process as a novel strategy for efficient photoreductive
948 degradation of perfluorooctanesulfonate. Environmental science & technology 2018, 52,
949 (5), 2953-2962.
950 57. Sun, M.; Zhou, H.; Xu, B.; Bao, J., Distribution of perfluorinated compounds in
951 drinking water treatment plant and reductive degradation by UV/SO 3 M process.
952 Environmental Science and Pollution Research 2018, 25, (8), 7443-7453.
953 58. Bao, Y.; Deng, S.; Jiang, X.; Qu, Y.; He, Y.; Liu, L.; Chai, Q.; Mumtaz, M.; Huang,
954 J.; Cagnetta, G., Degradation of PFOA Substitute: GenX (HFPO–DA Ammonium Salt):
955 Oxidation with UV/Persulfate or Reduction with UV/Sulfite? Environmental science &
956 technology 2018, 52, (20), 11728-11734.
957 59. Bentel, M. J.; Yu, Y.; Xu, L.; Li, Z.; Wong, B. M.; Men, Y.; Liu, J., Defluorination of
958 Per-and Polyfluoroalkyl Substances (PFASs) with Hydrated Electrons: Structural
959 Dependence and Implications to PFAS Remediation and Management. Environmental
960 science & technology 2019.
961 60. Buxton, G. V.; Greenstock, C. L.; Helman, W. P.; Ross, A. B., Critical review of
962 rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl
963 radicals (! OH/! M in aqueous solution. Journal of physical and chemical reference data
964 1988, 17, (2), 513-886.

46

ACS Paragon Plus Environment


Page 47 of 53 Environmental Science & Technology

965 61. Trojanowicz, M.; Bojanowska-Czajka, A.; Bartosiewicz, I.; Kulisa, K., Advanced
966 oxidation/reduction processes treatment for aqueous perfluorooctanoate (PFOA) and
967 perfluorooctanesulfonate (PFOS)–a review of recent advances. Chemical Engineering
968 Journal 2018, 336, 170-199.
969 62. Huang, L.; Dong, W.; Hou, H., Investigation of the reactivity of hydrated electron
970 toward perfluorinated carboxylates by laser flash photolysis. Chemical Physics Letters
971 2007, 436, (1-3), 124-128.
972 63. Tian, H.; Gao, J.; Li, H.; Boyd, S. A.; Gu, C., Complete defluorination of
973 perfluorinated compounds by hydrated electrons generated from 3-indole-acetic-acid in
974 organomodified montmorillonite. Scientific reports 2016, 6, 32949.
975 64. Tian, H.; Gu, C., Effects of different factors on photodefluorination of
976 perfluorinated compounds by hydrated electrons in organo-montmorillonite system.
977 Chemosphere 2018, 191, 280-287.
978 65. Larsen, R. E.; Glover, W. J.; Schwartz, B. J., Does the hydrated electron occupy
979 a cavity? Science 2010, 329, (5987), 65-69.
980 66. Devonshire, R.; Weiss, J. J., Nature of the transient species in the
981 photochemistry of negative ions in aqueous solution. The Journal of physical chemistry
982 1968, 72, (11), 3815-3820.
983 67. Hart, E. J., The hydrated electron. Science 1964, 146, (3652), 1664-1664.
984 68. Herbert, J. M.; Coons, M. P., The hydrated electron. Annual review of physical
985 chemistry 2017, 68, 447-472.
986 69. Kevan, L., Solvated electron structure in glassy matrixes. Accounts of Chemical
987 Research 1981, 14, (5), 138-145.
988 70. Casey, J. R.; Kahros, A.; Schwartz, B. J., To be or not to be in a cavity: the
989 hydrated electron dilemma. The Journal of Physical Chemistry B 2013, 117, (46),
990 14173-14182.
991 71. Swallow, A. J., Radiation chemistry. An introduction. 1973.
992 72. Schwarz, H. A., Free radicals generated by radiolysis of aqueous solutions.
993 Journal of Chemical Education 1981, 58, (2), 101-105.

47

ACS Paragon Plus Environment


Environmental Science & Technology Page 48 of 53

994 73. Hart, E. J.; Anbar, M., Hydrated electron. 1970.


995 74. Bäckström, H. L., The chain-reaction theory of negative catalysis1. Journal of the
996 American Chemical Society 1927, 49, (6), 1460-1472.
997 75. Dogliotti, L.; Hayon, E., Flash photolysis study of sulfite, thiocyanate, and
998 thiosulfate ions in solution. The Journal of Physical Chemistry 1968, 72, (5), 1800-1807.
999 76. Hayon, E.; Treinin, A.; Wilf, J., Electronic spectra, photochemistry, and
1000 autoxidation mechanism of the sulfite-bisulfite-pyrosulfite systems. SO2-, SO3-, SO4-,
1001 and SO5-radicals. Journal of the American Chemical Society 1972, 94, (1), 47-57.
1002 77. Chawla, O. P.; Arthur, N.; Fessenden, R. W., Electron spin resonance study of
1003 the photolysis of aqueous sulfite solutions. The Journal of Physical Chemistry 1973, 77,
1004 (6), 772-776.
1005 78. Deister, U.; Warneck, P., Photooxidation of sulfite (SO32-) in aqueous solution.
1006 Journal of Physical Chemistry 1990, 94, (5), 2191-2198.
1007 79. Fischer, M.; Warneck, P., Photodecomposition and photooxidation of hydrogen
1008 sulfite in aqueous solution. The Journal of Physical Chemistry 1996, 100, (37), 15111-
1009 15117.
1010 80. Haynes, W. M., CRC handbook of chemistry and physics. CRC press: 2014.
1011 81. Zoschke, K.; Börnick, H.; Worch, E., Vacuum-UV radiation at 185 nm in water
1012 treatment–a review. Water research 2014, 52, 131-145.
1013 82. Eriksen, T. E., pH Effects on the pulse radiolysis of deoxygenated aqueous
1014 solutions of sulphur dioxide. Journal of the Chemical Society, Faraday Transactions 1:
1015 Physical Chemistry in Condensed Phases 1974, 70, 208-215.
1016 83. Franck, J.; Scheibe, G., Über absorptionsspektren negativer halogenionen in
1017 lösung. Zeitschrift für Physikalische Chemie 1928, 139, (1), 22-31.
1018 84. Jortner, J.; Levine, R.; Ottolenghi, M.; Stein, G., The photochemistry of the iodide
1019 ion in aqueous solution. The Journal of Physical Chemistry 1961, 65, (7), 1232-1238.
1020 85. Jortner, J.; Ottolenghi, M.; Stein, G., On the photochemistry of aqueous solutions
1021 of chloride, bromide, and iodide ions. The Journal of Physical Chemistry 1964, 68, (2),
1022 247-255.

48

ACS Paragon Plus Environment


Page 49 of 53 Environmental Science & Technology

1023 86. Long, F. H.; Lu, H.; Shi, X.; Eisenthal, K. B., Femtosecond studies of electron
1024 photodetachment from an iodide ion in solution: The trapped electron. Chemical Physics
1025 Letters 1990, 169, (3), 165-171.
1026 87. Long, F. H.; Shi, X.; Lu, H.; Eisenthal, K. B., Electron photodetachment from
1027 halide ions in solution: Excited-state dynamics in the polarization well. The Journal of
1028 Physical Chemistry 1994, 98, (30), 7252-7255.
1029 88. Sheu, W.-S.; Rossky, P. J., The electronic dynamics of photoexcited aqueous
1030 iodide. Chemical physics letters 1993, 202, (3-4), 186-190.
1031 89. Long, F. H.; Lu, H.; Eisenthal, K. B., Femtosecond studies of the presolvated
1032 electron: An excited state of the solvated electron? Physical review letters 1990, 64,
1033 (12), 1469.
1034 90. Shi, X.; Long, F. H.; Lu, H.; Eisenthal, K. B., Femtosecond electron solvation
1035 kinetics in water. The Journal of Physical Chemistry 1996, 100, (29), 11903-11906.
1036 91. Sheu, W.-S.; Rossky, P. J., Electronic and solvent relaxation dynamics of a
1037 photoexcited aqueous halide. The Journal of Physical Chemistry 1996, 100, (4), 1295-
1038 1302.
1039 92. Tang, Y.; Shen, H.; Sekiguchi, K.; Kurahashi, N.; Mizuno, T.; Suzuki, Y.-I.;
1040 Suzuki, T., Direct measurement of vertical binding energy of a hydrated electron.
1041 Physical Chemistry Chemical Physics 2010, 12, (15), 3653-3655.
1042 93. Elliot, A. J., A pulse radiolysis study of the reaction of OH with I2 and the decay
1043 of 8 M2 Canadian Journal of Chemistry 1992, 70, (6), 1658-1661.
1044 94. Lian, R.; Oulianov, D. A.; Shkrob, I. A.; Crowell, R. A., Geminate recombination
1045 of electrons generated by above-the-gap (12.4 eV) photoionization of liquid water.
1046 Chemical physics letters 2004, 398, (1-3), 102-106.
1047 95. Sahul, K.; Sharma, B., Gamma radiolysis of nitrilotriacetic acid (NTA) in aqueous
1048 solutions. Journal of radioanalytical and nuclear chemistry 1987, 109, (2), 321-327.
1049 96. Mialocq, J.; Amouyal, E.; Bernas, A.; Grand, D., Picosecond laser photolysis of
1050 aqueous indole and tryptophan. The Journal of Physical Chemistry 1982, 86, (16),
1051 3173-3177.

49

ACS Paragon Plus Environment


Environmental Science & Technology Page 50 of 53

1052 97. Tian, H.; Guo, Y.; Pan, B.; Gu, C.; Li, H.; Boyd, S. A., Enhanced photoreduction
1053 of nitro-aromatic compounds by hydrated electrons derived from indole on natural
1054 montmorillonite. Environmental science & technology 2015, 49, (13), 7784-7792.
1055 98. Guo, C.; Zhang, C.; Sun, Z.; Zhao, X.; Zhou, Q.; Hoffmann, M. R., Synergistic
1056 impact of humic acid on the photo-reductive decomposition of perfluorooctanoic acid.
1057 Chemical Engineering Journal 2019, 360, 1101-1110.
1058 99. Hori, H.; Hayakawa, E.; Einaga, H.; Kutsuna, S.; Koike, K.; Ibusuki, T.;
1059 Kiatagawa, H.; Arakawa, R., Decomposition of environmentally persistent
1060 perfluorooctanoic acid in water by photochemical approaches. Environmental science &
1061 technology 2004, 38, (22), 6118-6124.
1062 100. Nohara, K.; Toma, M.; Kutsuna, S.; Takeuchi, K.; Ibusuki, T., Cl atom-initiated
1063 oxidation of three homologous methyl perfluoroalkyl ethers. Environmental science &
1064 technology 2001, 35, (1), 114-120.
1065 101. Bertin, M.; Martin, I.; Duvernay, F.; Theulé, P.; Bossa, J.-B.; Borget, F.;
1066 Illenberger, E.; Lafosse, A.; Chiavassa, T.; Azria, R., Chemistry induced by low-energy
1067 electrons in condensed multilayers of ammonia and carbon dioxide. Physical Chemistry
1068 Chemical Physics 2009, 11, (11), 1838-1845.
1069 102. Lafosse, A.; Bertin, M.; Azria, R., Electron driven processes in ices: Surface
1070 functionalization and synthesis reactions. Progress in Surface Science 2009, 84, (5-6),
1071 177-198.
1072 103. Wang, Y.; Zhang, P., Photocatalytic decomposition of perfluorooctanoic acid
1073 (PFOA) by TiO2 in the presence of oxalic acid. Journal of hazardous materials 2011,
1074 192, (3), 1869-1875.
1075 104. Erkoç, V2: Erkoç, F., Structural and electronic properties of PFOS and LiPFOS.
1076 Journal of Molecular Structure: THEOCHEM 2001, 549, (3), 289-293.
1077 105. Paul, A.; Wannere, C. S.; Schaefer, H. F., Do linear-chain perfluoroalkanes bind
1078 an electron? The Journal of Physical Chemistry A 2004, 108, (43), 9428-9434.

50

ACS Paragon Plus Environment


Page 51 of 53 Environmental Science & Technology

1079 106. Bao, Y.; Cagnetta, G.; Huang, J.; Yu, G., Degradation of hexafluoropropylene
1080 oxide oligomer acids as PFOA alternatives in simulated nanofiltration concentrate:
1081 Effect of molecular structure. Chemical Engineering Journal 2020, 382, 122866.
1082 107. Bao, Y.; Huang, J.; Cagnetta, G.; Yu, G., Removal of F–53B as PFOS alternative
1083 in chrome plating wastewater by UV/Sulfite reduction. Water research 2019, 163,
1084 114907.
1085 108. Bentel, M. J.; Yu, Y.; Xu, L.; Kwon, H.; Li, Z.; Wong, B. M.; Men, Y.; Liu, J.,
1086 Degradation of Perfluoroalkyl Ether Carboxylic Acids with Hydrated Electrons:
1087 Structure–Reactivity Relationships and Environmental Implications. Environmental
1088 Science & Technology 2020.
1089 109. Hasson, V.; Nicholls, R., Absolute spectral absorption measurements on
1090 molecular oxygen from 2640-1920 AA. II. Continuum measurements 2430-1920 AA.
1091 Journal of Physics B: Atomic and Molecular Physics 1971, 4, (12), 1789.
1092 110. Sörensen, M.; Frimmel, F. H., Photodegradation of EDTA and NTA in the
1093 UV/H2O2 process. Zeitschrift für Naturforschung B 1995, 50, (12), 1845-1853.
1094 111. Larson, R. A.; Stabler, P. P., Sensitized photooxidation of nitrilotriacetic and
1095 iminodiacetic acids. Journal of Environmental Science & Health Part A 1978, 13, (8),
1096 545-552.
1097 112. Ma, H.; Allen, H. E.; Yin, Y., Characterization of isolated fractions of dissolved
1098 organic matter from natural waters and a wastewater effluent. Water research 2001, 35,
1099 (4), 985-996.
1100 113. Croue, J.-P.; Korshin, G. V.; Benjamin, M. M., Characterization of natural organic
1101 matter in drinking water. American Water Works Association: 2000.
1102 114. Frimmel, F., Characterization of natural organic matter as major constituents in
1103 aquatic systems. Journal of Contaminant Hydrology 1998, 35, (1-3), 201-216.
1104 115. Shon, H.; Vigneswaran, S.; Snyder, S. A., Effluent organic matter (EfOM) in
1105 wastewater: constituents, effects, and treatment. Critical reviews in environmental
1106 science and technology 2006, 36, (4), 327-374.

51

ACS Paragon Plus Environment


Environmental Science & Technology Page 52 of 53

1107 116. Jin, P.; Jin, X.; Bjerkelund, V. A.; Østerhus, S. W.; Wang, X. C.; Yang, L., A study
1108 on the reactivity characteristics of dissolved effluent organic matter (EfOM) from
1109 municipal wastewater treatment plant during ozonation. Water research 2016, 88, 643-
1110 652.
1111 117. Andreottola, G.; Cannas, P., 2.4 Chemical and Biological Characteristics of
1112 Landfill Leachate. Landfilling of waste: leachate 1992, 1, 65.
1113 118. Shouliang, H.; Beidou, X.; Haichan, Y.; Liansheng, H.; Shilei, F.; Hongliang, L.,
1114 Characteristics of dissolved organic matter (DOM) in leachate with different landfill
1115 ages. Journal of Environmental Sciences 2008, 20, (4), 492-498.
1116 119. Zhao, R.; Jung, C.; Trzopek, A.; Torrens, K.; Deng, Y., Characterization of
1117 ultraviolet-quenching dissolved organic matter (DOM) in mature and young leachates
1118 before and after biological pre-treatment. Environmental Science: Water Research &
1119 Technology 2018, 4, (5), 731-738.
1120 120. Kang, S.-H.; Choi, W., Oxidative degradation of organic compounds using zero-
1121 valent iron in the presence of natural organic matter serving as an electron shuttle.
1122 Environmental Science & Technology 2008, 43, (3), 878-883.
1123 121. Scott, D. T.; McKnight, D. M.; Blunt-Harris, E. L.; Kolesar, S. E.; Lovley, D. R.,
1124 Quinone moieties act as electron acceptors in the reduction of humic substances by
1125 humics-reducing microorganisms. Environmental science & technology 1998, 32, (19),
1126 2984-2989.
1127 122. Wang, W.; Zafiriou, O. C.; Chan, I.-Y.; Zepp, R. G.; Blough, N. V., Production of
1128 hydrated electrons from photoionization of dissolved organic matter in natural waters.
1129 Environmental science & technology 2007, 41, (5), 1601-1607.
1130 123. Lester, Y.; Sharpless, C. M.; Mamane, H.; Linden, K. G., Production of photo-
1131 oxidants by dissolved organic matter during UV water treatment. Environmental science
1132 & technology 2013, 47, (20), 11726-11733.
1133 124. Zhang, D.; Yan, S.; Song, W., Photochemically induced formation of reactive
1134 oxygen species (ROS) from effluent organic matter. Environmental science &
1135 technology 2014, 48, (21), 12645-12653.

52

ACS Paragon Plus Environment


Page 53 of 53 Environmental Science & Technology

1136 125. Li, C.; Zheng, S. S.; Li, T. T.; Chen, J. W.; Zhou, J. H.; Su, L. M.; Zhang, Y. N.;
1137 Crittenden, J. C.; Zhu, S. Y.; Zhao, Y. H., Quantitative structure-activity relationship
1138 models for predicting reaction rate constants of organic contaminants with hydrated
1139 electrons and their mechanistic pathways. Water Research 2019, 151, 468-477.
1140 126. Spalding, R. F.; Exner, M. E., Occurrence of nitrate in groundwater—a review.
1141 Journal of environmental quality 1993, 22, (3), 392-402.
1142 127. USEPA Monitoring and Assessing Water Quality - Volunteer Monitoring.
1143 http://water.epa.gov/type/rsl/monitoring/vms510.cfm
1144 128. Gonzalez, M. G.; Oliveros, E.; Wörner, M.; Braun, A. M., Vacuum-ultraviolet
1145 photolysis of aqueous reaction systems. Journal of Photochemistry and Photobiology C:
1146 Photochemistry Reviews 2004, 5, (3), 225-246.
1147 129. Bichsel, Y.; Von Gunten, U., Formation of iodo-trihalomethanes during
1148 disinfection and oxidation of iodide-containing waters. Environmental Science &
1149 Technology 2000, 34, (13), 2784-2791.
1150 130. Hua, G.; Reckhow, D. A.; Kim, J., Effect of bromide and iodide ions on the
1151 formation and speciation of disinfection byproducts during chlorination. Environmental
1152 Science & Technology 2006, 40, (9), 3050-3056.
1153 131. Plewa, M. J.; Wagner, E. D.; Richardson, S. D.; Thruston, A. D.; Woo, Y.-T.;
1154 McKague, A. B., Chemical and biological characterization of newly discovered iodoacid
1155 drinking water disinfection byproducts. Environmental science & technology 2004, 38,
1156 (18), 4713-4722.
1157 132. Richardson, S. D.; Fasano, F.; Ellington, J. J.; Crumley, F. G.; Buettner, K. M.;
1158 Evans, J. J.; Blount, B. C.; Silva, L. K.; Waite, T. J.; Luther, G. W., Occurrence and
1159 mammalian cell toxicity of iodinated disinfection byproducts in drinking water.
1160 Environmental Science & Technology 2008, 42, (22), 8330-8338.
1161 133. Pan, Y.; Li, W.; An, H.; Cui, H.; Wang, Y., Formation and occurrence of new polar
1162 iodinated disinfection byproducts in drinking water. Chemosphere 2016, 144, 2312-
1163 2320.

1164

53

ACS Paragon Plus Environment

You might also like