uv/nitrilotriacetic acid process as a novel strategy for efficient … · 2018. 2. 5. · 1 1...

42
Subscriber access provided by Caltech Library Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties. Article UV/nitrilotriacetic acid process as a novel strategy for efficient photoreductive degradation of perfluorooctane sulfonate Zhuyu Sun, Chaojie Zhang, Lu Xing, Qi Zhou, Wenbo Dong, and Michael R Hoffmann Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b05912 • Publication Date (Web): 03 Feb 2018 Downloaded from http://pubs.acs.org on February 5, 2018 Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Upload: others

Post on 01-Feb-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • Subscriber access provided by Caltech Library

    Environmental Science & Technology is published by the American Chemical Society.1155 Sixteenth Street N.W., Washington, DC 20036Published by American Chemical Society. Copyright © American Chemical Society.However, no copyright claim is made to original U.S. Government works, or worksproduced by employees of any Commonwealth realm Crown government in the courseof their duties.

    Article

    UV/nitrilotriacetic acid process as a novel strategy for efficientphotoreductive degradation of perfluorooctane sulfonate

    Zhuyu Sun, Chaojie Zhang, Lu Xing, Qi Zhou, Wenbo Dong, and Michael R HoffmannEnviron. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b05912 • Publication Date (Web): 03 Feb 2018

    Downloaded from http://pubs.acs.org on February 5, 2018

    Just Accepted

    “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are postedonline prior to technical editing, formatting for publication and author proofing. The American ChemicalSociety provides “Just Accepted” as a service to the research community to expedite the disseminationof scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear infull in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fullypeer reviewed, but should not be considered the official version of record. They are citable by theDigital 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. Aftera manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Website and published as an ASAP article. Note that technical editing may introduce minor changesto the manuscript text and/or graphics which could affect content, and all legal disclaimers andethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors orconsequences arising from the use of information contained in these “Just Accepted” manuscripts.

  • 1

    UV/nitrilotriacetic acid process as a novel strategy for 1

    efficient photoreductive degradation of perfluorooctane 2

    sulfonate 3

    4

    Zhuyu Sun,†‡

    Chaojie Zhang,*†‡

    Lu Xing,†‡

    Qi Zhou,†‡

    Wenbo Dong,&

    5

    Michael R. Hoffmann§ 6

    †State Key Laboratory of Pollution Control and Resources Reuse, College 7

    of Environmental Science and Engineering, Tongji University, Shanghai 8

    200092, China 9

    ‡ Shanghai Institute of Pollution Control and Ecological Security, 10

    Shanghai 200092, China 11

    &Shanghai Key Laboratory of Atmospheric Particle Pollution and 12

    Prevention, Department of Environmental Science & Engineering, Fudan 13

    University, Shanghai 200433, China 14

    §Linde-Robinson Laboratories, California Institute of Technology, 15

    Pasadena, California 91125, United States 16

    *Corresponding author. Tel: +86 21 65981831; fax: +86 21 65983869; 17

    E-mail address: [email protected] 18

    19

    Page 1 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 2

    Abstract: Perfluorooctane sulfonate (PFOS) is a toxic, bioaccumulative and highly 20

    persistent anthropogenic chemical. Hydrated electrons (eaq–) are potent nucleophiles 21

    that can effectively decompose PFOS. In previous studies, eaq– are mainly produced 22

    by photoionization of aqueous anions or aromatic compounds. In this study, we 23

    proposed a new photolytic strategy to generate eaq– and in turn decompose PFOS, 24

    which utilizes nitrilotriacetic acid (NTA) as a photosensitizer to induce water 25

    photodissociation and photoionization, and subsequently as a scavenger of hydroxyl 26

    radical (·OH) to minimize the geminate recombination between ·OH and eaq–. The net 27

    effect is to increase the amount of eaq– available for PFOS degradation. The UV/NTA 28

    process achieved a high PFOS degradation ratio of 85.4% and a defluorination ratio of 29

    46.8% within 10 h. A pseudo first-order rate constant (k) of 0.27 h-1

    was obtained. The 30

    laser flash photolysis study indicates that eaq– is the dominant reactive species 31

    responsible for PFOS decomposition. The generation of eaq– is greatly enhanced and 32

    its half-life is significantly prolonged in the presence of NTA. The electron spin 33

    resonance (ESR) measurement verified the photodissociation of water by detecting 34

    ·OH. The model compound study indicates that the acetate and amine groups are the 35

    primary reactive sites. 36

    Page 2 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 3

    1. Introduction 37

    Perfluorooctane sulfonate (PFOS, C8F17SO3−) is a useful anthropogenic chemical 38

    that has been extensively utilized in industrial and commercial applications for 39

    decades.1,2

    However, due to its high toxicity, environmental persistence, 40

    bioaccumulation and global distribution, PFOS is considered hazardous to 41

    environmental and human health.3,4

    Moreover, because of the high thermal and 42

    chemical stability of C–F bond (~116 kcal/mol, almost the strongest in nature2), PFOS 43

    is extremely resistant to traditional chemical and biological degradations. 44

    Recently, hydrated electrons (eaq–) mediated photoreductive approaches have 45

    garnered special attention for PFOS decomposition owing to their extraordinarily high 46

    efficiency and relatively mild conditions. However, in order to produce sufficient 47

    amount of eaq– to defluorinate PFOS, chemicals such as iodide

    5, sulfite

    6, chloride

    7 and 48

    indole derivatives8 are essential. Lyu et al. developed a catalyst-free PFOS 49

    photodecomposition method, but it relied on using strong alkaline conditions (pH = 50

    11.8) and high temperatures (100 °C).9 Furthermore, due to the rapid reaction between 51

    eaq– and oxygen (Eq. 1, rate constant = 1.9×10

    10 M

    -1·s

    -1)10

    , the reduction efficiency of 52

    eaq– is significantly affected by dissolved oxygen (O2). Thereby, previous eaq

    –53

    -mediated photoreductive approaches normally require strict anoxic conditions.9, 11

    In 54

    view of the drawbacks of current photoreductive technologies, we were motivated to 55

    propose a new strategy for PFOS removal, which can be conducted under more 56

    environmentally relevant conditions and with minimization of undesirable 57

    byproducts. 58

    Page 3 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 4

    In addition to electron photodetachment from anions and aromatic compounds12,

    59

    13, UV photolysis of liquid water generates eaq

    – as well (Eq. 2).

    14, 15 However, direct 60

    photoionization of liquid water under 254 nm irradiation is difficult due to a 61

    negligible absorption coefficient at λ > 200 nm.16

    In addition, the quantum yield of 62

    eaq– from water photoionization is low because the radical species undergo significant 63

    geminate recombination. In pure water photolysis, energy deposition takes place in 64

    well-separated local volumes called spurs, where the primary products including eaq–, 65

    hydroxyl radicals (·OH), hydrogen atoms (·H) and H3O+ are formed in close vicinity 66

    with high initial local concentrations.17

    However, as the spurs expand through 67

    diffusion, a large proportion of the primary products undergo very efficient back 68

    reactions (Eq. 3-5). Thus, only a small fraction of eaq– actually escape into the bulk 69

    solution to initiate reductive chemical processes.10

    Therefore, there is scarcely any net 70

    generation of eaq– in pure water photolysis upon low energy excitation. Based on these 71

    facts, we now propose a new strategy to enhance the generation of eaq–, i.e., by 72

    minimizing the geminate recombination of eaq– with oxidative species after inducing 73

    water photoionization. 74

    ���– + O� → O�

    ∙ (1)

    H�O��� ���

    – +∙ OH +∙ H + H�O� (2)

    ���– +∙ OH → OH (3)

    ���– +∙ H → H� + OH

    (4)

    ���– + H�O

    � →∙ H + H�O (5)

    Aminopolycarboxylic acids (APCAs) are compounds that contain several 75

    Page 4 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 5

    carboxylate groups bound to one or more nitrogen atoms. Due to their excellent 76

    chelating properties18

    , interests in previous studies were mainly focused in their 77

    chelation with metal ions.19-21

    Furthermore, APCAs contain several donor atoms 78

    whose free electron pairs are easily attacked by ·OH. Leitner et al.22

    reported that the 79

    rate constants for the reactions of APCAs with ·OH can exceed 1010

    M-1

    ·s-1

    . These 80

    near diffusion controlled values are 2 to 3 orders of magnitude higher than reactions 81

    of APCAs with eaq–. Therefore, APCAs are considered to be excellent candidates to 82

    scavenge ·OH, and thus increase the apparent quantum yields of eaq– and in turn 83

    promote the photoreductive degradation of PFOS. 84

    In this study, we chose to use nitrilotriacetic acid (NTA) as a representative 85

    APCA to explore its impact on the photoreductive degradation of PFOS. NTA is the 86

    first chemical synthetic APCA23

    and is more biodegradable than other APCAs24

    . In 87

    order to unravel the underlying reaction mechanisms, laser flash photolysis, ESR 88

    measurement and a model compound study were carried out. Furthermore, the effects 89

    of NTA concentration, pH and air were investigated. To our best knowledge, this is 90

    the first study that utilizes APCAs as ·OH scavengers to promote the photoreductive 91

    degradation of refractory organic pollutants. The findings in this study can provide a 92

    new strategy for PFOS remediation and a novel application field for APCAs. 93

    Page 5 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 6

    2. Materials and methods 94

    2.1. Chemicals 95

    Perfluorooctanesulfonic acid (PFOS, ~40% in water), HPLC grade methanol 96

    (≥99.9%), 5,5-Dimethyl-1-pyrroline N-oxide (DMPO, ≥98.0%) and model 97

    compounds including trisodium citrate (≥99.0%), ethylenediamine-N,N’-disuccinic 98

    acid trisodium salt solution (EDDS, ~35% in water) and methylglycine diacetic acid 99

    (MGDA, ≥99.0%) were purchased from Sigma-Aldrich Chemical Co. (St. Louis, Mo, 100

    USA). Nitrilotriacetic acid (NTA, ≥98.5%), ammonium hydroxide solution (25%), 101

    ammonium chloride (≥99.8%), ethylenediaminetetraacetic acid (EDTA, ≥99.5%), 102

    iminodiacetic acid (IDA, ≥98.0%), glycine (≥99.5%), oxamic acid (≥98.0%) and 103

    sodium oxalate (≥99.8%) were obtained from Sinopharm Chemical Reagent Co. 104

    (Shanghai, China). HPLC grade ammonium acetate (97.0%) was purchased from 105

    TEDIA (Fairfield, OH, USA). Sodium perfluoro-1-[1,2,3,4-13

    C4]octanesulfonate 106

    (MPFOS, ≥99%, 13

    C4) acquired from Wellington Laboratories Inc. (Guelph, ON, 107

    Canada) was used as the internal standard for the quantification of PFOS. Milli-Q 108

    water and deionized water were used throughout the whole experiment. 109

    2.2. Reductive defluorination 110

    The photoreductive degradation of PFOS was conducted under anoxic conditions 111

    in a stainless steel cylindrical reactor (Fig. S1, SI). The outer and inner diameters of 112

    the reactor were 100 mm and 60 mm, respectively. A low-pressure mercury lamp (14 113

    W, Heraeus, Germany) with quartz tube protection was placed in the center of the 114

    reactor, emitting 254 nm UV light. 720 mL solution was added to the reactor. Before 115

    Page 6 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 7

    the reaction, the mixture was bubbled with highly purified nitrogen for 20 mins to 116

    remove oxygen. The solution pH was adjusted with ammonium hydroxide-ammonium 117

    chloride (NH3·H2O-NH4Cl) buffer. The reaction temperature was held constant at 30

    118

    oC by the circulating cooling system. After various time intervals, samples of the 119

    liquid were analyzed after filtration through 0.22 µm nylon filter (ANPEL Laboratory 120

    Technologies, Shanghai, China). 121

    2.3. Analytical methods 122

    The concentrations of PFOS and possible aqueous-phase intermediate analytes 123

    were determined by high-performance liquid chromatography/tandem mass 124

    spectrometry (HPLC−MS/MS, TSQTM

    Quantum AccessTM

    , Thermo Finnigan, San 125

    Jose, CA, USA). Ion Chromatography (Dionex, ICS-3000, Thermo Fisher Scientific, 126

    USA) equipped with a conductivity detector was used for the analysis of fluoride, 127

    nitrate and short chain organic acids. NTA and its degradation products including IDA, 128

    oxamate and oxalate were detected by UPLC−MS/MS (ACQUITY UPLC&SCIEX 129

    SelexION Triple Quad 5500 System, Waters, MA, USA). More detailed information 130

    is available in the SI. 131

    2.4. Laser flash photolysis experiment and electron spin resonance (ESR) 132

    measurement 133

    Nanosecond laser flash photolysis for the detection of eaq– was performed using a 134

    Quanta Ray LAB-150-10 Nd:YAG laser at an excitation wavelength of 266 nm. The 135

    components of the laser flash photolysis apparatus were as described by Ouyang et 136

    al..25

    Detailed information is available in the SI. 137

    Page 7 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 8

    The ESR spectra were obtained on a Bruker EMXplus-10/12 ESR spectrometer 138

    at room temperature. The instrumental parameters for ESR analysis were as follows: 139

    microwave frequency, 9.852 GHz; microwave power, 20 mW; modulation amplitude 140

    1 G; modulation frequency, 100 kHz; center field, 3500 G; sweep width, 100 G. The 141

    simulations of ESR spectra were obtained with the use of Spinfit in Xenon software. 142

    Details are provided in the SI. 143

    144

    Page 8 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 9

    3. Results and Discussion 145

    3.1. NTA-assisted photoreductive degradation and defluorination of PFOS 146

    147

    Figure 1. The time profiles of PFOS degradation (a) and defluorination (b) under different 148

    Page 9 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 10

    conditions. The treatment experiment: PFOS (0.01 mM), NTA (2 mM), UV irradiation, N2 149

    saturated, pH (10.0); the 1st control experiment (direct photolysis): PFOS (0.01 mM), UV 150

    irradiation, N2 saturated; the 2nd

    control experiment (UV/buffer): PFOS (0.01 mM), UV irradiation, 151

    N2 saturated, pH (10.0); the 3rd

    control experiment (without UV irradiation): PFOS (0.01 mM), 152

    NTA (2mM), N2 saturated, pH (10.0); the 4th

    control experiment (UV/sulfite): PFOS (0.01 mM), 153

    SO32-

    (2mM), UV irradiation, N2 saturated, pH (10.0). Error bars represent standard deviations of 154

    triplicate assays. 155

    The photochemical decomposition of PFOS was conducted under 254 nm light 156

    irradiation in the presence of NTA under anoxic condition. In order to determine the 157

    effect of each parameter, four control experiments were conducted. The results of 158

    experiments under different conditions are shown in Fig. 1. Since PFOS has a weak 159

    absorption at 254 nm and the quantum yield of eaq– by pure water photolysis is 160

    negligible26

    , the direct photolysis of PFOS was poor. Only 12.9% of the initial PFOS 161

    was decomposed after 10 h of irradiation, with a low defluorination ratio of 3.7%. 162

    Adjusting solution pH to 10.0 with NH3·H2O-NH4Cl buffer somewhat enhanced the 163

    decomposition of PFOS, with the 10-h PFOS degradation and defluorination ratios 164

    increased to 20.9% and 9.9%, respectively. Compared with the UV/buffer process, 165

    addition of 2 mM NTA significantly accelerated the degradation and defluorination of 166

    PFOS. After 10 h of irradiation, the degradation and defluorination ratios of PFOS in 167

    the presence of NTA were 85.4% and 46.8%, respectively. In particular, PFOS 168

    degraded rapidly during the initial 0.5 h. The 0.5-h PFOS photodegradation and 169

    defluorination ratios by UV/NTA process achieved 45.1% and 18.6%, respectively, 170

    Page 10 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 11

    which are 5.2-fold and 8.5-fold, respectively higher than by UV/buffer process. The 171

    results of the 3rd

    control experiment indicate that PFOS does not degrade without UV 172

    irradiation, even in the presence of NTA. Therefore, both UV and NTA are necessary 173

    for the efficient degradation and defluorination of PFOS. 174

    Gu et al. (2016) demonstrated that UV/sulfite system exhibited a high efficiency 175

    in decomposing PFOS.6 In order to better evaluate the efficiency of UV/NTA process 176

    for PFOS degradation, the newly developed approach was compared with UV/sulfite 177

    process (the 4th

    control). The 10-h degradation and defluorination ratios of PFOS by 178

    UV/sulfite process were 60.1% and 29.6%, respectively, which were 0.7-fold and 179

    0.63-fold, respectively lower than by UV/NTA process. PFOS degradation by the 180

    UV/NTA process follows pseudo first-order kinetics (Fig. S4, SI), with an apparent 181

    reaction rate constant (k) of 0.27 h-1

    , and a half-life (t1/2) of 2.6 h. The k value for the 182

    UV/NTA process was higher than other photochemical approaches including UV/KI 183

    process5, UV/sulfite process, UV/Fe

    3+ process

    27, UV/alkaline 2-propanol process

    26 184

    and UV/K2S2O8 process28

    (Table S1, SI). Therefore, UV/NTA process exhibited a 185

    high efficiency in the degradation and defluorination of PFOS. 186

    As PFOS decomposed, short-chain-length perfluorinated intermediates including 187

    PFBS, PFBA, PFHxS, PFHxA, PFHpA and PFOA were detected (Fig. S5, SI). Based 188

    on the concentrations of PFOS, intermediates and F−, the mass balance of F was 189

    calculated (Fig. S6, SI). The loss of F recovery during the reaction implies the 190

    formation of partially fluorinated intermediates. Therefore, based on the product 191

    distribution and F balance, three possible degradation pathways of PFOS are likely to 192

    Page 11 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 12

    occur in UV/NTA process: a) PFOS undergoes direct defluorination to form 193

    polyfluorinated intermediates; b) PFOS firstly undergoes desulfonation to form PFOA, 194

    and then degrades to shorter-chain-length perfluoroalkyl carboxylic acids (PFCAs); c) 195

    C-C bond in PFOS molecules cleaves, in which case short-chain-length 196

    perfluoroalkane sulfonates (PFSAs), such as PFHxS and PFBS, were produced. The 197

    proposed PFOS degradation pathways in UV/NTA process are similar with other 198

    photochemical processes5, 9, 26

    , and accord well with the mechanisms suggested by 199

    molecular orbitals and thermodynamic analyses6. 200

    As mentioned above, most photochemical technologies for PFOS degradation are 201

    faced with challenges of secondary pollution due to by-product formation. In order to 202

    evaluate the potential post-treatment impact of the UV/NTA process, the degradation 203

    products of NTA were quantified. As is shown in Fig. S7, SI, NTA degraded rapidly 204

    during the first 1 h of PFOS degradation. 97.7% of the initial NTA decomposed within 205

    1 h, along with the concomitant formation of three intermediates including IDA, 206

    oxamate and oxalate. The maximum concentrations of IDA, oxamate and oxalate 207

    were 0.75 mM, 0.26 mM and 0.68 mM, respectively, observed at 0.5 h, 2 h and 6 h, 208

    respectively. IDA and oxamate underwent nearly complete decomposition after 10 h, 209

    with their concentrations falling below 0.03 mM. Oxalate was completely 210

    decomposed after 14 h (data not shown). Ammonium (NH4+) and nitrate (NO3

    −) were 211

    the major N-containing end products. Based on the product distribution and the NTA 212

    photooxidation mechanism in literatures29, 30

    , a possible NTA decomposition pathway 213

    is illustrated in Fig. S8, SI. The end products of NTA degradation are NH4+, NO3

    − and 214

    Page 12 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 13

    carbon dioxide (CO2), which are innocuous and can be easily removed by biological 215

    transformation. Furthermore, NTA is reported to have relatively low environmental 216

    risks to sewage treatment or aquatic life.18, 31

    It is readily biodegradable by natural 217

    microbial processes, and the biodegradation is complete without accumulation of 218

    unwanted intermediates.24

    In contrast, the secondary pollutions of other approaches 219

    appear to be severer. For example, the iodide by-products include residual iodide, 220

    iodine, polyiodide and iodate, which may cause detrimental effects to aquatic 221

    organisms32

    and human health33

    . The various iodide species are also potential sources 222

    for iodinated disinfection byproducts (iodo-DBPs). Therefore, UV/NTA process is 223

    expected to provide a relatively green alternative for efficient photoreductive 224

    degradation of PFOS. 225

    As is shown in Fig. 1, the PFOS decomposition rate attenuated after 1 h. This is 226

    primarily due to the nearly complete degradation of NTA within 1 h. The subsequent 227

    PFOS degradation and defluorination after 1 h is attributed to the effect of NTA 228

    degradation products such as IDA. The effects of NTA degradation products are 229

    discussed in section 3.3. 230

    Page 13 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 14

    3.2. Mechanism of NTA-assisted UV photoreductive degradation of PFOS 231

    Photodegradation of organic compounds can take place through direct and/or 232

    indirect photolysis. Since direct PFOS photolysis under UV irradiation is very slow, 233

    PFOS degradation in UV/NTA process is believed to occur through indirect pathway. 234

    In order to ascertain the dominant reactive species that is responsible for the 235

    degradation of PFOS, experiments with nitrous oxide (N2O) were conducted. As is 236

    shown in Fig. S9 in the SI, PFOS degradation and defluorination were substantially 237

    suppressed in the presence of N2O. N2O is a well-known scavenger of eaq–, which 238

    quenches eaq– rapidly to form ·OH (Eq. 6, rate constant = 9.1×10

    9 M

    -1·s

    -1)

    10, whereas 239

    ·OH has a poor reactivity towards perfluorinated acids that it cannot decompose PFOS 240

    effectively.34, 35

    Therefore, these results imply the crucial role of eaq– in the PFOS 241

    degradation in UV/NTA process. Besides eaq–, N2O can also react with ·H.

    10 However, 242

    under alkaline conditions, the yield of ·H is much lower than that of eaq–, and the rate 243

    constant for the reaction of N2O and ·H at alkaline pH is 2.1×106 M

    -1·s

    -1, which is 244

    over 3 orders of magnitude lower than that of N2O and eaq–. 10

    Therefore, the effect of 245

    ·H can be excluded. 246

    ���– +N�O → OH

    +∙ OH + N� (6)

    247

    Page 14 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 15

    248

    Figure 2. Transient absorption spectra following the laser flash photolysis of 10 mM NTA 249

    solution at pH 10.0. Insertion in Figure 2 is the decay of eaq– detected at 630 nm in the presence of 250

    PFOS with different concentrations. The transient absorption curves are fitted. 251

    The N2O scavenging experiment provides an indirect proof for the potential role 252

    of eaq–. However, a more direct proof is needed to reveal the formation and decay of 253

    eaq– in UV/NTA process. Therefore, a laser flash photolysis study was conducted. The 254

    absorption spectrum of intermediates was obtained at 10 nm intervals between 260 255

    nm and 700 nm upon the photolysis of 10 mM NTA in N2 saturated water (Fig. 2). 256

    The broad optical absorption band with a peak at around 630 nm is attributed to eaq– 257

    since it is identical to the eaq– absorption spectrum acquired both theoretically

    36 and 258

    experimentally37

    . The absorption peak of eaq– decayed continuously with monitoring 259

    time, indicating eaq– gradually reacts back with other primary species. The decays of 260

    Page 15 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 16

    eaq– in the presence of PFOS with different concentrations are shown in the inserted 261

    figure in Fig. 2. In the absence of NTA (i.e., the laser photolysis of pure water), eaq– 262

    was not produced. By adding 10 mM NTA, the absorbance at 630 nm increased 263

    significantly, confirming the enhanced production of eaq– in the presence of NTA. The 264

    addition of PFOS accelerated the decay of eaq–, and its decay rate increased at higher 265

    PFOS concentration level. These results further verify the role of eaq– in PFOS 266

    degradation. 267

    Thus, the generation of eaq– in UV/NTA process has been clearly confirmed by 268

    the laser flash photolysis study. However, compared with common eaq– source 269

    chemicals like sulfite38

    , ferrocyanide39

    , iodide40

    , and aromatic compounds like 270

    pyrenetetrasulfonate41

    , the transient yield of eaq– by UV/NTA process is much lower. 271

    The relatively low transient yield of eaq– cannot explain the high efficiency of 272

    UV/NTA process in PFOS degradation and defluorination. For instance, the 10-h 273

    degradation and defluorination ratios of PFOS by UV/NTA process were 1.42-fold 274

    and 1.58-fold, respectively higher than by UV/sulfite process. Therefore, we speculate 275

    that besides direct photoejection of eaq– from NTA, there may be another dominant 276

    mechanism for the efficient generation of eaq–. 277

    Grossweiner et al. once proposed an alternative mechanism for the generation 278

    of eaq– based on their observations, that is the photoionizaton of water itself sensitized 279

    by the light-absorbing substances.12

    In this case, a hydroxyl radical is produced by 280

    dissociation of the photoionized water.12

    According to the UV-Vis spectra (Fig. S10, 281

    SI), direct photoionization of water under 254 nm light irradiation can be excluded 282

    Page 16 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 17

    since water is transparent in the near UV spectral domain (λ > 200 nm)16

    . However, 283

    NTA has an absorbance at 254 nm, which means NTA is able to absorb photons and 284

    possibly sensitize water ionization. Furthermore, APCAs can form hydration 285

    complexes with several water molecules through hydrogen bonding. For instance, 286

    eight water molecules are required to fully hydrate the first hydration shell of 287

    deprotonated glycine.42

    The solvation process may make it easier for the interactions 288

    between NTA and water. Therefore based on foregoing facts, we speculate that water 289

    photoionization sensitized by NTA is another major source of eaq– in UV/NTA 290

    process. 291

    In addition, an important reason for the low quantum yield of eaq– from pure 292

    water photolysis is the rapid recombination of eaq– with ·OH, ·H and H3O

    + (Eq. 3-5), 293

    especially the reaction with ·OH, which accounts for more than 82% ± 3% of the 294

    recombination of eaq–.43

    The geminate recombination between eaq– and ·OH greatly 295

    decreases the survival probability of eaq–. However in UV/NTA process, NTA is 296

    highly reactive towards ·OH,30, 44

    with a high reaction rate constant of 4.2×109 M

    -1·s

    -1 297

    at pH of 10.0.22, 45

    The strong scavenging capacity of NTA towards ·OH can 298

    effectively protect eaq– from being quenched by ·OH, which increases the steady-state 299

    concentration of eaq– and in turn facilitates the decomposition of PFOS. The eaq

    –300

    protection mechanism can be demonstrated by the long lifetime of eaq– in UV/NTA 301

    process (Fig. 2). For instance, the half-life of eaq– in UV/NTA process is about 11.6 µs, 302

    whereas it is only 1 ns in UV/sulfite process38

    and lower than 2 ns in UV/KI process40

    . 303

    Therefore, the presence of NTA can dramatically prolong the survival time of eaq– by 304

    Page 17 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 18

    scavenging ·OH, thus promoting the degradation of PFOS. This explains why 305

    UV/NTA process has a low transient quantum yield of eaq–, whereas has a high 306

    efficiency in PFOS degradation. 307

    Overall, the generation of eaq– during the UV/NTA process appears to involve 308

    three steps. In aqueous solution, NTA is fully hydrated with both primary and 309

    secondary hydration spheres. The fully hydrated NTA induces water photodissociation 310

    and photoionization as a photosensitizer with the generation of eaq– and ·OH. Finally, 311

    the NTA core scavenges ·OH to decrease the geminate recombination between ·OH 312

    and eaq–, thus increasing the amount of eaq

    – available for PFOS degradation. 313

    314

    Figure 3. (A) DMPO spin-trapping ESR spectra recorded in the UV/NTA and UV/buffer 315

    (NH3·H2O-NH4Cl) processes. (B) Simulated data of various radicals trapped by DMPO in 316

    UV/NTA process after 60 s irradiation: (c) total ESR signal; (e) simulation of DMPO-·OH; (f) 317

    simulation of DMPO-·CH2COOH; (g) simulation of NO·. 318

    In order to verify our speculation, DMPO was used as an ESR spin-trap to 319

    identify possible short-lived radicals produced in the UV/NTA process. As shown in 320

    Fig. 3, after 30 s irradiation, several peaks were clearly observed in the ESR spectra, 321

    Page 18 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 19

    indicating formation of radicals and their involvement in the reaction. As the 322

    irradiation time increased to 60 s, the ESR signal intensity also increased. No 323

    well-defined ESR signal was observed in the UV/buffer process, indicating that the 324

    radicals were not produced in the absence of NTA. In order to further confirm the 325

    radical species, the ESR data of UV/NTA-60s reaction were simulated. The ESR 326

    signal of 1:2:2:1 quartets is thus assigned to DMPO-·OH (curve e). The spectrum of 327

    curve f is speculated to be the signal of DMPO-·CH2COOH. The three-line spectrum 328

    (curve g) indicates the detection of free nitroxide (NO·).46, 47

    The occurrence of ·OH 329

    suggests that water indeed undergoes photodissociation in the UV/NTA process. NO· 330

    and ·CH2COOH are presumably the reaction products of NTA with ·OH, which is 331

    consistent with the reactive sites in NTA molecule (discussed in detail in section 3.3) 332

    and also consistent with the known steady-state products of NTA degradation (Fig. S7, 333

    SI). The ESR data of UV/NTA-30s has a similar fitting result with UV/NTA-60s 334

    (shown in Fig. S11, SI). Therefore, the ESR measurement detected the occurrence of 335

    ·OH, ·CH2COOH and NO·, further confirming the mechanism that NTA induces water 336

    photodissociation and photoionization, and scavenges ·OH to increase the quantum 337

    yield of eaq–.338

    Page 19 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 20

    3.3. Model compound study 339

    340

    Page 20 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 21

    Figure 4. The time profiles of PFOS degradation and defluorination in the presence of model 341

    compounds (NTA, IDA, EDTA, EDDA, citrate, glycine, oxamate, EDDS and MGDA): PFOS 342

    (0.01 mM), model compounds (2 mM), UV irradiation, N2 saturated, pH (10.0). Error bars 343

    represent standard deviations of triplicate assays. 344

    In order to further validate the proposed mechanism and to test the effects of 345

    other chelating agents, IDA, EDTA, EDDA, citric acid, glycine, oxamic acid, oxalic 346

    acid, EDDS and MGDA were chosen as model compounds to investigate the 347

    structure-activity relationship of APCAs in terms of impacting the photodegradation 348

    of PFOS. The structural formulas of the model compounds are shown in Table S2, SI. 349

    The results of the model compound study are shown in Fig. 4. Except oxamate and 350

    oxalate, all of the tested model compounds accelerated the PFOS degradation and 351

    defluorination compared with the control experiment, although their enhancement 352

    effects varied a lot. For example, NTA was clearly better than IDA in accelerating 353

    PFOS degradation. The 10-h PFOS degradation and defluorination ratios in presence 354

    of NTA were 85.4% and 46.8%, respectively, while in the case of IDA they were only 355

    54.8% and 25.6%, respectively. NTA and IDA are APCAs with only one nitrogen 356

    atom. However, the nitrogen atom in NTA molecule is attached one more acetate 357

    group (-CH2COOH) than IDA. Similarly, in the comparison of EDDA to EDTA, the 358

    PFOS degradation and defluorination ratios were found to be higher in the presence of 359

    EDTA (10-h values of 78.5% and 43.7%, respectively for EDTA and 51.1% and 360

    24.2%, respectively for EDDA). This is because each nitrogen atom in EDTA 361

    molecule is connected to one more acetate group than EDDA. Therefore, the acetate 362

    Page 21 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 22

    group density appears to be a factor that determines the efficiency of the specific 363

    APCA in promoting PFOS photodegradation. 364

    Citric acid is an important organic tricarboxylic acid. According to the results 365

    shown in Fig. 4, citrate also promotes PFOS degradation and defluorination as 366

    compared with the control experiment in its absence. However, the enhancement 367

    effect of citrate is lower than NTA. The 10-h PFOS degradation ratios in the presence 368

    of NTA and citrate were 85.4% and 50.7%, respectively, with the corresponding 10-h 369

    defluorination ratios of 46.8% and 24.1%, respectively. Citric acid has a similar 370

    chemical structure with NTA, both containing acetate groups, whereas citric acid has 371

    no amine group. Therefore, the greater acceleration in the presence of NTA than 372

    citrate can be probably attributed to the electron-rich center of amine group, whose 373

    lone pair on the nitrogen atom favors the electrophilic attack of ·OH. 374

    Glycine and oxamic acid both contain an amine group and a carboxyl group, but 375

    their relative impacts on PFOS degradation are totally different. Glycine clearly 376

    promoted PFOS degradation compared with the control experiment, whereas oxamate 377

    has a significant inhibitory effect. The 10-h PFOS degradation and defluorination 378

    ratios in the presence of glycine were 61.5% and 29.4%, respectively, whereas they 379

    were only 1.8% and 0.4%, respectively in the presence of oxamate. These results are 380

    mainly due to the different moieties present in glycine and oxamic acid that connect 381

    amine group and carboxyl group, i.e., methylene group (CH2) for glycine and 382

    carbonyl group (C=O) for oxamic acid. Therefore, it is presumably the methylene 383

    group in acetate group that offers a reactive site for ·OH attack. The significant 384

    Page 22 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 23

    inhibition of oxamate is resulted from the electron-withdrawing property of 385

    C=O-COOH group, which inductively withdraws electron density from the 386

    neighboring N atom, thus reducing its reactivity with ·OH. Oxamic acid is reported to 387

    be highly persistent during ·OH oxidation process.48, 49

    In contrast, oxamic acid has a 388

    high reactivity towards eaq–. The rate constant for the reaction of oxamate with eaq

    – is 389

    reported to be 5.7×109 M

    -1·s

    -1 (pH = 9.2). The second-order rate constant for the 390

    reactions between oxamate and eaq– is more than 3 orders of magnitude higher than 391

    the corresponding rate constant for eaq– with glycine (1.7×10

    6 M

    -1·s

    -1, pH = 11.8). 392

    Therefore, oxamic acid competes with PFOS for eaq–, and thus inhibits PFOS 393

    degradation. Similar to oxamic acid, the presence of oxalate significantly inhibited 394

    PFOS degradation and defluorination (Fig. S12, SI). This result coincides with the 395

    low reaction rate constant for oxalate with ·OH (7.7×106 M

    -1·s

    -1, pH = 6.0)

    10 and 396

    further indicates that carboxyl group alone was unable to accelerate PFOS 397

    degradation. In other words, the methylene group in acetate moieties is essential for 398

    the effective attack by ·OH via H-atom abstraction. 399

    In comparing EDTA to EDDS, we find that EDTA is more efficient with respect 400

    to acceleration of PFOS degradation than EDDS. Although EDTA and EDDS both 401

    have four acetate groups, two acetate groups in EDDS molecule are attached to the 402

    α-C instead of the N atom. Therefore, EDTA is a tertiary amine while EDDS is a 403

    secondary amine. The hydrogen bonding to the N atom in EDDS decreases the 404

    availability of N-electrons and hinders the N-electrons transfer. Furthermore, the 405

    presence of hydrogen on carbon next to the amine (α-hydrogen) was reported to be a 406

    Page 23 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 24

    key factor for electron-transfer interactions.50

    Therefore, an extra α-hydrogen in 407

    EDTA may result in a higher electron-transfer capability. Compared with NTA, the 408

    enhancement is more favored for MGDA. This is because MGDA has an additional 409

    CH3 group on the carbon α of amine, which increases the electron density of N due to 410

    the electron-donating inductive effect, and adds a reactive site for ·OH attack. 411

    Therefore, the electronic distribution of N atom, especially the availability of the lone 412

    pair on N atom also makes a difference for the efficiencies of APCAs. 413

    In summary, the acetate group and amine group in APCAs play important roles 414

    in accelerating the photodegradation of PFOS. This conclusion is consistent with the 415

    reactive sites for the reactions of APCAs with ·OH as pointed out in literatures.22, 29, 45,

    416

    50-53 Therefore, the results of our model compound study confirm that the scavenging 417

    effect on ·OH by APCAs is the primary mechanism for the enhanced 418

    photodegradation of PFOS. 419

    The effects of NTA degradation products including IDA, glycine, oxamate and 420

    oxalate were summarized in Fig. S12, SI. Their efficiencies for the degradation and 421

    defluorination of PFOS followed the order of NTA > glycine ≈ IDA > control > 422

    oxamate ≈ oxalate. Compared with the control experiment, NTA, glycine and IDA 423

    obviously accelerated the photodegradation of PFOS, whereas oxamate and oxalate 424

    significantly inhibited. Since NTA degradation products have either lower efficiencies 425

    than NTA or inhibiting effect, the PFOS degradation and defluorination rates 426

    decreased gradually as NTA degraded, especially after 1 h of reaction. 427

    Page 24 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 25

    3.4. Effects of pH and air 428

    429

    Figure 5. The effect of pH on the degradation (a) and defluorination (b) of PFOS: PFOS (0.01 430

    mM), NTA (2 mM), UV irradiation, N2 saturated. Error bars represent standard deviations of 431

    Page 25 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 26

    triplicate assays. 432

    As is shown in Fig. 5, the degradation and defluorination of PFOS by UV/NTA 433

    process is strongly dependent on pH. The PFOS degradation ratios under the pH 434

    conditions of 6.0, 7.0, 8.0, 9.0, 10.0 and 11.0 after 10 h were 17.4%, 29.9%, 47.7%, 435

    56.6%, 85.4% and 99.5%, respectively. The 10-h defluorination ratios of PFOS under 436

    the pH conditions of 6.0, 7.0, 8.0, 9.0, 10.0 and 11.0 were 8.5%, 8.0%, 15.1%, 34.9%, 437

    46.8% and 72.3%, respectively. After 2 h irradiation, the degradation and 438

    defluorination ratios of PFOS at the pH value of 11.0 had reached 82.2% and 48.8%, 439

    respectively, whereas they were only 2.1% and 0.1%, respectively at the pH value of 440

    6.0. The strong dependence of PFOS degradation on pH is mainly attributed to the 441

    following two reasons. First, at low pH values, eaq– is quickly quenched by H

    + with a 442

    high reaction rate constant of 2.3×1010

    M−1

    ·s−1

    (Eq. 5). 10

    The excessive consumption 443

    of eaq– by H

    + inhibits the degradation of PFOS and results in a low defluorination 444

    efficiency under acidic condition. Second, pH affects the reactivity of NTA towards 445

    ·OH. The pKa values for successive deprotonation of NTA are 0.8, 1.9, 2.48 and 446

    9.65.45, 54

    The principal forms of NTA at pH 2.0 are HN+(CH2COOH)2(CH2COO

    −) 447

    and HN+(CH2COOH)(CH2COO

    −)2; its major form at pH 6.0 is HN

    +(CH2COO

    −)3; 448

    while at pH 10.0, NTA is near fully deprotonated in the form of :N+(CH2COO

    −)3. As 449

    concluded in the model compound study, the lone pair on the nitrogen atom is one of 450

    the primary reactive sites for the electrophilic attack of ·OH on APCAs molecules. At 451

    lower pH values, the addition of a single proton to :N(CH2COO−)3 or to 452

    HN+(CH2COO

    −)3 can decrease the rate constant for ·OH reaction by about one order 453

    Page 26 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 27

    of magnitude.45

    Therefore, the deprotonated form of NTA is the more reactive species 454

    towards ·OH than its protonated or partially protonated counterparts. Leitner et al. 455

    drew a similar conclusion that the reactivity of ·OH with NTA is related to the amount 456

    of deprotonated nitrogen.22

    The rate constant for the reaction of ·OH with NTA was 457

    reported to be 6.1×107

    M−1

    ·s−1

    , 5.5×108

    M−1

    ·s−1

    and 4.2×109

    M−1

    ·s−1

    , respectively, at 458

    the pH values of 2.0, 6.0 and 10.0.45

    Furthermore, since the pKa4 value of NTA is 459

    9.65,45

    the PFOS degradation rate increased significantly as pH increased from 9.0 to 460

    10.0. Therefore, alkaline conditions are most favorable for the reaction of ·OH with 461

    NTA, thus making more eaq– available for the degradation of PFOS. 462

    Page 27 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 28

    463

    Figure 6. The effect of air on the degradation (a) and defluorination (b) of PFOS: PFOS (0.01 464

    mM), NTA (2 mM), UV irradiation, pH (10.0). Error bars represent standard deviations of 465

    triplicate assays. 466

    Page 28 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 29

    Besides pH, air is another important parameter that affects the photoreductive 467

    degradation efficiency of PFOS. In general, the eaq–-mediated degradation of PFOS 468

    requires strict anoxic conditions to prevent eaq– from being quenched by molecular 469

    oxygen (Eq. 1).9, 11

    For instance, Park et al. reported that PFOS was degraded rapidly 470

    by UV/Iodide process in the presence of Ar with a rate constant of 6.5×10-3

    min-1

    , 471

    whereas it was not degraded in the presence of air.5 Therefore in this study, PFOS 472

    degradation was conducted under both N2 and air saturated conditions to evaluate the 473

    effect of air. Under N2 saturation, the solution was pre-bubbled with N2 for 20 mins 474

    and kept bubbling during the whole reaction period. Under air saturation, the solution 475

    was not pre-bubbled and the reactor was kept open with the solution exposed to air 476

    during the whole reaction period. The results are shown in Fig. 6. The degradation 477

    ratios of PFOS under N2 and air saturated conditions after 10 h irradiation were 85.4% 478

    and 75.2%, respectively. The 10-h defluorination ratios of PFOS under N2 and air 479

    saturation were 46.8% and 35.8%, respectively. The PFOS degradation and 480

    defluorination efficiencies under N2 saturation were just slightly higher than under air 481

    condition. No significantly negative impact of air was observed. This is because the 482

    excited NTA under UV irradiation (NTA*) can react with O2.30

    Meanwhile, ·OH can 483

    abstract the alpha hydrogen from the methylene group in NTA molecule.45

    The 484

    hydrogen abstracted radical intermediate can also react with O2.45

    As a reductant, 485

    NTA can scavenge a variety of oxidizing species produced in the presence of O2 such 486

    as superoxide radicals (O2·−) and HO2 radicals (HO2·), which are also active 487

    quenchers of eaq–.55

    The sequestration of these oxidants by NTA protects eaq–, and thus 488

    Page 29 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 30

    enhances the degradation of PFOS. Therefore, the photodegradation of PFOS by 489

    UV/NTA process is less sensitive to air than previously reported photoreductive 490

    approaches like UV/KI11

    and UV/sulfite56

    processes. This result bodes well for future 491

    applications of in situ or ex situ PFOS photo-remediation since strict anoxic 492

    conditions are often difficult to achieve in practical engineering applications. 493

    The effect of NTA concentration was also evaluated in this study. A higher NTA 494

    concentration over the range of 0 to 4.0 mM resulted in higher PFOS degradation and 495

    defluorination ratios (Fig. S13, SI). A detailed discussion is available in the SI. 496

    4. Environmental Implications 497

    This study presents a new strategy for efficient photoreductive degradation and 498

    defluorination of PFOS. Compared with previous photochemical approaches, the 499

    newly developed UV/NTA process has three remarkable advantages. First, PFOS 500

    undergoes rapid photodegradation in the presence of NTA, with a high defluorination 501

    rate. Second, no significant detrimental impact of air was observed, which means 502

    UV/NTA process has a certain tolerance to oxygen. Third, the end products of NTA 503

    degradation are CO2, NH4+ and NO3

    −, which are easily biodegradable with minimized 504

    secondary pollution risks. Therefore, UV/NTA may provide a novel, efficient and 505

    relatively green technology for future in situ or ex situ PFOS remediation. 506

    In previous studies, the most common way for the generation of eaq– to 507

    decompose PFOS was by adding eaq– source chemicals such as iodide, sulfite and 508

    indole derivatives. Meanwhile, attentions were more often paid on the improvement 509

    of the transient generation of eaq–, such as increasing the reaction temperature or 510

    Page 30 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 31

    applying high photon flux. Actually, the apparent quantum yield of eaq– or its ultimate 511

    efficiency is not just determined by its transient quantum yield. It is also greatly 512

    influenced by the twinborn species such as ·OH and oxidative byproducts. For 513

    instance, the unsatisfactory activity of UV/iodide process towards PFOS degradation 514

    is likely a result of triiodide scavenging of eaq–.57

    In this study, we proposed a new 515

    way to generate eaq–, which utilizes NTA to sensitize water photoionization, and 516

    subsequently to scavenge ·OH, in order to minimize the geminate recombination 517

    between ·OH and eaq–. The net effect is to increase the steady-state level of eaq

    – that is 518

    available for PFOS degradation. The laser flash photolysis results indicate that 519

    although UV/NTA process has a low transient yield of eaq–, it has a long half-life of 520

    eaq–, which is considered to be the main reason for the high efficiency of UV/NTA 521

    process. 522

    APCAs is a class of compounds acting as chelating agents. Interests in previous 523

    studies were mainly focused in their chelating effect with metal ions. Other APCAs’ 524

    properties receive little attention. For example, they are excellent electron donors and 525

    they have a high reactivity with ·OH. In this study, we revealed that the 526

    photoreductive degradation of PFOS can be enhanced by various APCAs such as 527

    NTA, IDA, EDTA, EDDA, EDDS and MGDA. The acetate and amine groups are the 528

    primary reactive sites in APCAs molecules. Therefore, APCAs-mediated 529

    photoreductive process may be a novel application field for APCAs, which is also 530

    expected to be a promising replacement for the current photoredutive technologies for 531

    PFOS decomposition. 532

    Page 31 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 32

    Associated content 533

    Supporting Information 534

    Schematic diagram of the photochemical reactor; Technical data of the low-pressure 535

    mercury lamp; additional details on analytical methods; kinetics for PFOS 536

    degradation; abbreviation and chemical structure of model compounds. Figures 537

    showing time profiles of PFOS degradation products; mass balance of F; NTA and 538

    NTA degradation products; NTA degradation pathway; effect of N2O; UV-Vis 539

    absorption spectra; ESR spectra; effect of NTA degradation products and effect of 540

    NTA concentration. Discussion on the effect of NTA concentration. This information 541

    is available free of charge via the Internet at http://pubs.acs.org. 542

    Acknowledgements 543

    The authors greatly thank Dr. Jiahui Yang from Bruker (Beijing) Scientific 544

    Technology Co., Ltd for her kind assistance on the ESR result analysis. The authors 545

    also gratefully acknowledge Prof. Side Yao and Dr. Huijie Shi for their valuable 546

    comments on the discussion. This study has been supported by the National Natural 547

    Science Foundation of China (Project No. 21677109), and the State Key Laboratory 548

    of Pollution Control and Resource Reuse Foundation (No. PCRRT16001). 549

    550

    References 551

    1. Renner, R. Growing concern over perfluorinated chemicals. Environ. Sci. 552

    Technol. 2001, 35 (7), 154A-160A. 553

    2. Giesy, J. P.; Kannan, K. Peer reviewed: perfluorochemical surfactants in the 554

    Page 32 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 33

    environment. Environ. Sci. Technol. 2002, 36 (7), 146A-152A. 555

    3. Lau, C.; Butenhoff, J. L.; Rogers, J. M. The developmental toxicity of 556

    perfluoroalkyl acids and their derivatives. Toxicology and Applied Pharmacology 557

    2004, 198 (2), 231-241. 558

    4. Lau, C.; Anitole, K.; Hodes, C.; Lai, D.; Pfahles-Hutchens, A.; Seed, J. 559

    Perfluoroalkyl acids: A review of monitoring and toxicological findings. 560

    Toxicological Sciences 2007, 99 (2), 366-394. 561

    5. Park, H.; Vecitis, C. D.; Cheng, J.; Choi, W.; Mader, B. T.; Hoffmann, M. R. 562

    Reductive Defluorination of Aqueous Perfluorinated Alkyl Surfactants: Effects 563

    of Ionic Headgroup and Chain Length. J. Phys. Chem. A 2009, 113 (4), 690-696. 564

    6. Gu, Y. R.; Dong, W. Y.; Luo, C.; Liu, T. Z. Efficient Reductive Decomposition of 565

    Perfluorooctanesulfonate in a High Photon Flux UV/Sulfite System. Environ. Sci. 566

    Technol. 2016, 50 (19), 10554-10561. 567

    7. Guo, R.; Zhang, C. J.; Zhang, G.; Zhou, Q. Degradation of Perfluorooctanoic 568

    Acid by UV/Chloride Process. Chem. J. Chin. Univ.-Chin. 2016, 37 (8), 569

    1499-1508. 570

    8. Tian, H. T.; Gao, J.; Li, H.; Boyd, S. A.; Gu, C. Complete Defluorination of 571

    Perfluorinated Compounds by Hydrated Electrons Generated from 572

    3-Indole-acetic-acid in Organomodified Montmorillonite. Scientific Reports 573

    2016, 6, 9. 574

    9. Lyu, X. J.; Li, W. W.; Lam, P. K. S.; Yu, H. Q. Insights into perfluorooctane 575

    sulfonate photodegradation in a catalyst-free aqueous solution. Scientific Reports 576

    Page 33 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 34

    2015, 5, 1-6. 577

    10. Buxton, G. V.; Greenstock, C. L.; Helman, W. P.; Ross, A. B. Critical review of 578

    rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl 579

    radicals (⋅OH/⋅O−) in aqueous solution. Journal of physical and chemical 580

    reference data 1988, 17 (2), 513-886. 581

    11. Qu, Y.; Zhang, C. J.; Li, F.; Chen, J.; Zhou, Q. Photo-reductive defluorination of 582

    perfluorooctanoic acid in water. Water Research 2010, 44 (9), 2939-2947. 583

    12. Grossweiner, L. I.; Swenson, G. W.; Zwicker, E. F. Photochemical Generation of 584

    the Hydrated Electron. Science (New York, N.Y.) 1963, 141 (3583), 805-6. 585

    13. Sauer, M. C.; Crowell, R. A.; Shkrob, I. A. Electron photodetachment from 586

    aqueous anions. 1. Quantum yields for generation of hydrated electron by 193 587

    and 248 nm laser photoexcitation of miscellaneous inorganic anions. J. Phys. 588

    Chem. A 2004, 108 (25), 5490-5502. 589

    14. Elles, C. G.; Jailaubekov, A. E.; Crowell, R. A.; Bradforth, S. E. 590

    Excitation-energy dependence of the mechanism for two-photon ionization of 591

    liquid H2O and D2O from 8.3 to 12.4 eV. J. Chem. Phys. 2006, 125 (4), 12. 592

    15. Elles, C. G.; Shkrob, I. A.; Crowell, R. A.; Bradforth, S. E. Excited state 593

    dynamics of liquid water: Insight from the dissociation reaction following 594

    two-photon excitation. J. Chem. Phys. 2007, 126 (16), 8. 595

    16. Nikogosyan, D. N.; Angelov, D. A. Formation of free-radicals in water under 596

    high-power laser UV irradiation. Chem. Phys. Lett. 1981, 77 (1), 208-210. 597

    17. Gobert, F.; Pommeret, S.; Vigneron, G.; Buguet, S.; Haidar, R.; Mialocq, J. C.; 598

    Page 34 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 35

    Lampre, I.; Mostafavi, M. Nanosecond kinetics of hydrated electrons upon water 599

    photolysis by high intensity femtosecond UV pulses. Res. Chem. Intermed. 2001, 600

    27 (7-8), 901-910. 601

    18. Schmidt, C. K.; Brauch, H. J. Impact of aminopolycarboxylates on aquatic 602

    organisms and eutrophication: Overview of available data. Environ. Toxicol. 603

    2004, 19 (6), 620-637. 604

    19. Repo, E.; Warchol, J. K.; Bhatnagar, A.; Mudhoo, A.; Sillanpaa, M. 605

    Aminopolycarboxylic acid functionalized adsorbents for heavy metals removal 606

    from water. Water Research 2013, 47 (14), 4812-4832. 607

    20. Saifullah; Meers, E.; Qadir, M.; de Caritat, P.; Tack, F. M. G.; Du Laing, G.; Zia, 608

    M. H. EDTA-assisted Pb phytoextraction. Chemosphere 2009, 74 (10), 609

    1279-1291. 610

    21. De Luca, A.; Dantas, R. F.; Esplugas, S. Study of Fe(III)-NTA chelates stability 611

    for applicability in photo-Fenton at neutral pH. Appl. Catal. B-Environ. 2015, 612

    179, 372-379. 613

    22. Leitner, N. K. V.; Guilbault, I.; Legube, B. Reactivity of OH center dot and 614

    e(aq)(-) from electron beam irradiation of aqueous solutions of EDTA and 615

    aminopolycarboxylic acids. Radiat. Phys. Chem. 2003, 67 (1), 41-49. 616

    23. Heintz, W. Ueber dem Ammoniaktypus angehörige organische Säuren. Justus 617

    Liebigs Annalen der Chemie 1862, 122 (3), 257-294. 618

    24. Bucheli-Witschel, M.; Egli, T. Environmental fate and microbial degradation of 619

    aminopolycarboxylic acids. Fems Microbiol. Rev. 2001, 25 (1), 69-106. 620

    Page 35 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 36

    25. Ouyang, B.; Dong, W. B.; Hou, H. Q. A laser flash photolysis study of nitrous 621

    acid in the aqueous phase. Chem. Phys. Lett. 2005, 402 (4-6), 306-311. 622

    26. Yamamoto, T.; Noma, Y.; Sakai, S. I.; Shibata, Y. Photodegradation of 623

    perfluorooctane sulfonate by UV irradiation in water and alkaline 2-propanol. 624

    Environ. Sci. Technol. 2007, 41 (16), 5660-5665. 625

    27. Jin, L.; Zhang, P. Y.; Shao, T.; Zhao, S. L. Ferric ion mediated 626

    photodecomposition of aqueous perfluorooctane sulfonate (PFOS) under UV 627

    irradiation and its mechanism. J. Hazard. Mater. 2014, 271, 9-15. 628

    28. Yang, S. W.; Cheng, J. H.; Sun, J.; Hu, Y. Y.; Liang, X. Y. Defluorination of 629

    Aqueous Perfluorooctanesulfonate by Activated Persulfate Oxidation. PLoS One 630

    2013, 8 (10), 10. 631

    29. Chen, D.; Martell, A. E.; McManus, D. Studies on the mechanism of chelate 632

    degradation in iron-based, liquid redox H2S removal processes. Can. J. 633

    Chem.-Rev. Can. Chim. 1995, 73 (2), 264-274. 634

    30. Sorensen, M.; Frimmel, F. H. Photodegradation of EDTA and NTA in the 635

    UV/H2O2 process. Z.Naturforsch.(B) 1995, 50 (12), 1845-1853. 636

    31. Anderson, R. L.; Bishop, W. E.; Campbell, R. L. A review of the environmental 637

    and mammalian toxicology of the nitrilotriacetic acid. Crc Critical Reviews in 638

    Toxicology 1985, 15 (1), 1-102. 639

    32. Laverock, M. J.; Stephenson, M.; Macdonald, C. R. Toxicity of iodine, iodide, 640

    and iodate to daphnia-magna and rainbow-trout (oncorhynchus-mykiss). 641

    Archives of Environmental Contamination and Toxicology 1995, 29 (3), 344-350. 642

    Page 36 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 37

    33. Burgi, H.; Schaffner, T.; Seiler, J. P. The toxicology of iodate: A review of the 643

    literature. Thyroid 2001, 11 (5), 449-456. 644

    34. Hori, H.; Hayakawa, E.; Einaga, H.; Kutsuna, S.; Koike, K.; Ibusuki, T.; 645

    Kiatagawa, H.; Arakawa, R. Decomposition of environmentally persistent 646

    perfluorooctanoic acid in water by photochemical approaches. Environ. Sci. 647

    Technol. 2004, 38 (22), 6118-6124. 648

    35. Zhang, Z.; Chen, J. J.; Lyu, X. J.; Yin, H.; Sheng, G. P. Complete mineralization 649

    of perfluorooctanoic acid (PFOA) by gamma-irradiation in aqueous solution. 650

    Scientific Reports 2014, 4, 1-6. 651

    36. Abramczyk, H.; Kroh, J. Absorption spectrum of the solvated electron. 2. 652

    Numerical calculations of the profiles of the electron in water and methanol at 653

    300 K. The Journal of Physical Chemistry 1991, 95 (16), 6155-6159. 654

    37. Hart, E. J.; Boag, J. Absorption spectrum of the hydrated electron in water and in 655

    aqueous solutions. Journal of the American Chemical Society 1962, 84 (21), 656

    4090-4095. 657

    38. Lian, R.; Oulianov, D. A.; Crowell, R. A.; Shkrob, I. A.; Chen, X.; Bradforth, S. 658

    E. Electron photodetachment from aqueous anions. 3. Dynamics of geminate 659

    pairs derived from photoexcitation of mono-vs polyatomic anions. The Journal 660

    of Physical Chemistry A 2006, 110 (29), 9071-9078. 661

    39. Huang, L.; Dong, W. B.; Hou, H. Q. Investigation of the reactivity of hydrated 662

    electron toward perfluorinated carboxylates by laser flash photolysis. Chem. 663

    Phys. Lett. 2007, 436 (1-3), 124-128. 664

    Page 37 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 38

    40. Qu, Y.; Zhang, C. J.; Chen, P.; Zhou, Q.; Zhang, W. X. Effect of initial solution 665

    pH on photo-induced reductive decomposition of perfluorooctanoic acid. 666

    Chemosphere 2014, 107, 218-223. 667

    41. Yuan, H. X.; Pan, H. X.; Wu, Y. L.; Zhao, J. F.; Dong, W. B. Laser Flash 668

    Photolysis Mechanism of Pyrenetetrasulfonate in Aqueous Solution. Acta 669

    Phys.-Chim. Sin. 2012, 28 (4), 957-962. 670

    42. Yao, Y. H.; Chen, D.; Zhang, S.; Li, Y. L.; Tu, P. H.; Liu, B.; Dong, M. D. 671

    Building the First Hydration Shell of Deprotonated Glycine by the MCMM and 672

    ab Initio Methods. J. Phys. Chem. B 2011, 115 (19), 6213-6221. 673

    43. Thomsen, C. L.; Madsen, D.; Keiding, S. R.; Thogersen, J.; Christiansen, O. 674

    Two-photon dissociation and ionization of liquid water studied by femtosecond 675

    transient absorption spectroscopy. J. Chem. Phys. 1999, 110 (7), 3453-3462. 676

    44. Sillanpaa, M. E. T.; Kurniawan, T. A.; Lo, W. H. Degradation of chelating agents 677

    in aqueous solution using advanced oxidation process (AOP). Chemosphere 2011, 678

    83 (11), 1443-1460. 679

    45. Sahul, K.; Sharma, B. K. Gamma-radiolysis of nitrilotriacetic aicd (NTA) in 680

    aqueous solutions. J. Radioanal. Nucl. Chem.-Artic. 1987, 109 (2), 321-327. 681

    46. Rehorek, D.; Janzen, E. G. On the formation of arseno aminoxyls (nitroxides) by 682

    spin trapping of arseno radicals. Polyhedron 1984, 3 (5), 631-634. 683

    47. Buettner, G. R. Spin trapping - electron-spin resonance parameters of spin 684

    adducts. Free Radic. Biol. Med. 1987, 3 (4), 259-303. 685

    48. Ganiyu, S. O.; Oturan, N.; Raffy, S.; Esposito, G.; van Hullebusch, E. D.; Cretin, 686

    Page 38 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 39

    M.; Oturan, M. A. Use of Sub-stoichiometric Titanium Oxide as a Ceramic 687

    Electrode in Anodic Oxidation and Electro-Fenton Degradation of the 688

    Beta-blocker Propranolol: Degradation Kinetics and Mineralization Pathway. 689

    Electrochim. Acta 2017, 242, 344-354. 690

    49. Antonin, V. S.; Garcia-Segura, S.; Santos, M. C.; Brillas, E. Degradation of 691

    Evans Blue diazo dye by electrochemical processes based on Fenton's reaction 692

    chemistry. J. Electroanal. Chem. 2015, 747, 1-11. 693

    50. Chen, Y.; Hu, C.; Hu, X. X.; Qu, J. H. Indirect Photodegradation of Amine Drugs 694

    in Aqueous Solution under Simulated Sunlight. Environ. Sci. Technol. 2009, 43 695

    (8), 2760-2765. 696

    51. Furlong, D. N.; Wells, D.; Sasse, W. H. F. Photooxidation at platinum colloidal 697

    TiO2 aqueous-solution interfaces 1. Ethylenediaminetetraacetic acid and 698

    related-compounds. Aust. J. Chem. 1986, 39 (5), 757-769. 699

    52. Anipsitakis, G. P.; Dionysiou, D. D. Radical generation by the interaction of 700

    transition metals with common oxidants. Environ. Sci. Technol. 2004, 38 (13), 701

    3705-3712. 702

    53. Morooka, S.; Ikemizu, K.; Kamano, H.; Kato, Y. Ozonation rate of water-soluble 703

    chelates and related-compounds. J. Chem. Eng. Jpn. 1986, 19 (4), 294-299. 704

    54. Martell, A. E.; Smith, R. M. Critical Stability Constants Vol. Amino Acids. 705

    Plenum Press: New York, 1974. 706

    55. Larson, R. A.; Stabler, P. P. Sensitized photooxidation of nitrilotriacetic and 707

    iminodiacetic acids. Journal of Environmental Science & Health Part A 1978, 13 708

    Page 39 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • 40

    (8), 545-552. 709

    56. Song, Z.; Tang, H.; Wang, N.; Zhu, L. Reductive defluorination of 710

    perfluorooctanoic acid by hydrated electrons in a sulfite-mediated UV 711

    photochemical system. J. Hazard. Mater. 2013, 262 (22), 332-338. 712

    57. Park, H.; Vecitis, C. D.; Cheng, J.; Dalleska, N. F.; Mader, B. T.; Hoffmann, M. 713

    R. Reductive degradation of perfluoroalkyl compounds with aquated electrons 714

    generated from iodide photolysis at 254 nm. Photochemical & Photobiological 715

    Sciences 2011, 10 (12), 1945-1953. 716

    Page 40 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology

  • water photoionization

    O

    Geminate

    recombination

    hv

    H2O

    eaq−

    NTA

    scavenges

    ∙OH

    CO2, NH4+, NO3

    ∙OH

    NTA sensitized

    H C N O F S

    CC

    CC

    C

    C

    N

    O

    O

    O

    OO

    HH

    H

    H

    HH

    O

    Degradation & DefluorinationPFOS

    Page 41 of 41

    ACS Paragon Plus Environment

    Environmental Science & Technology