and Base-Catalyzed Hydrolysis of ... - ACS Publications

Jan 15, 2018 - include rainfall/snow melting, laundry wastewater, stormwater runoff, and desorption from soil,3−6 .... (detailed procedure for prepa...
5 downloads 11 Views 2MB Size
Subscriber access provided by Universitaetsbibliothek | Johann Christian Senckenberg

Article

Mineral- and base-catalyzed hydrolysis of organophosphate flame retardants – A potential major fate-controlling sink in soil and aquatic environments Yida Fang, Erin Kim, and Timothy J. Strathmann Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b05911 • Publication Date (Web): 15 Jan 2018 Downloaded from http://pubs.acs.org on January 15, 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 free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

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

Page 1 of 33

Environmental Science & Technology

1

2

Mineral- and base-catalyzed hydrolysis of

3

organophosphate flame retardants – A potential

4

major fate-controlling sink in soil and aquatic

5

environments

6

Yida Fang, Erin Kim, Timothy J. Strathmann*

7

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

8

Colorado 80401, United States

9

Table of Contents (TOC)/Abstract Art

ACS Paragon Plus Environment

Environmental Science & Technology

10

Abstract: The ubiquitous occurrence of organophosphate flame retardants (OPFRs) in aquatic and

11

soil environments poses significant risks to human health and ecosystems. Here, we report on the

12

hydrolysis of six OPFRs and three structural analogues in the absence and presence of metal

13

(hydr)oxide minerals. Eight of the target compounds showed marked degradation in alkaline

14

solutions (pH 9-12) with half-lives ranging from 0.02 – 170 d. Kinetics follow a second-order rate

15

law with apparent rate constants for base-catalyzed hydrolysis (kB) ranging from 0.69 – 42,000 M-

16

1

17

2 y, except for tris(2,2,2-trichloroethy) phosphate), rapid degradation is observed in the presence

18

of metal (hydr)oxide minerals, with half-lives reduced to

27

28 29 30

ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33

31

Environmental Science & Technology

Introduction

32

Organophosphate flame retardants (OPFRs; Table 1) are widely used as additives in residential

33

and commercial products, including building materials, furniture, plastics, and electronic

34

equipment.1 Their production over the past decade has increased exponentially due to the phaseout

35

of brominated flame retardants. Because OPFRs are not chemically bonded to the materials they

36

are added to, they readily leach over time.2 Major pathways for entry into soil and aquatic

37

environments include rainfall/snow melting, laundry wastewater, stormwater runoff, and

38

desorption from soil,3-6 and OPFRs have been detected in a number of aquatic systems, including

39

wastewater treatment plants, surface water, sediments and groundwater.7-12 For example, the

40

estimated discharge of tris(1-chloro-2-propyl) phosphate (TCPP) to U.S. rivers is 400 metric tons

41

per year.4

42

Despite benefits to public safety and property losses,7 release of OPFRs into the environment

43

poses significant risks to human health and ecosystems. OPFRs have been detected in human milk,

44

urine, serum, and adipose tissues,13-16 and potential health effects include cancer, neurotoxicity,

45

endocrine disruption, and reproductive impacts.17-22 They have also been detected in various

46

aquatic organisms, raising additional concerns about their effects on impacted ecosystems.23-25

47

Risks associated with OPFR release is compounded by their recalcitrance to many biotic and

48

abiotic degradation processes.26,27 To date, only limited information is available regarding their

49

natural attenuation.12 Reports of biodegradation have been limited to studies of aromatic OPFRs

50

in activated sludge28 and tris(2-chloroethyl) phosphate (TCEP) in enriched cultures.29

51

Investigations of photochemical degradation of OPFRs in lake and river water revealed variable

52

persistence, with chloroalkyl phosphates being the most recalcitrant group.30,31 Their

53

organophosphate ester core structure suggests that hydrolysis may be an important natural

ACS Paragon Plus Environment

Environmental Science & Technology

54

attenuation mechanism.32 However, a recent report indicates that aqueous hydrolysis of OPFRs is

55

extremely slow at pH conditions of natural waters, with appreciable hydrolysis only being

56

observed under highly alkaline conditions (e.g., pH 13).33 Similarly, reports of OPFR reactions

57

with other relevant nucleophiles (e.g., reduced sulfur species) have been documented, but

58

experiments were limited to elevated temperatures and nucleophile concentrations (e.g., t1/2 = 6.3

59

h at 50C and 10 mM thiophenolate, pH 9).34

60

Past reports show that dissolved metal ions and mineral surfaces can catalyze the hydrolysis of

61

some organophosphate pesticides (e.g., Mevinphos) and warfare agents (e.g., Soman).35,36

62

Elevated concentration of selected transition metal ions, including FeII, CoII, NiII, were found to

63

catalyze the hydrolysis of some organophosphate pesticides.37 Moreover, several mineral phases

64

were found to catalyze the hydrolysis of organophosphates (e.g., p-nitrophenyl phosphate and

65

paraoxon).38,39 While this suggests a potential sink for OPFRs, the nature of the side chains in

66

OPFR structures differ somewhat from those found in other organophosphates. Thus, further

67

studies are needed to evaluate the potential role of common soil minerals in catalyzing hydrolysis

68

of OPFRs and assess the relative importance of these pathways to natural attenuation.

69

This study examines for the first time the effects of common metal (hydr)oxide soil minerals on

70

the hydrolytic transformation of OPFRs. Specifically, the aqueous degradation of nine OP triesters

71

(six OPFRs and three structural analogs, Table 1) were simultaneously monitored in the absence

72

and presence of six naturally occurring minerals [iron, aluminum, titanium and silicon

73

(hydr)oxides] over a range of pH conditions. Application of sensitive liquid chromatography-

74

tandem mass spectrometry (LC-MS/MS) methods enabled simultaneous measurement of the

75

hydrolysis of all nine OPFRs in solution at concentrations low enough to prevent competition for

76

surface catalytic sites. High resolution mass spectrometry (quadrupole time-of-flight; LC-qToF-

ACS Paragon Plus Environment

Page 4 of 33

Page 5 of 33

Environmental Science & Technology

77

MS) was also used to identify and compare OPFR transformation products in both homogenous

78

and heterogeneous reactions. Results from kinetics experiments and product identification studies

79

were used to confirm the controlling reaction pathways, reveal structure-reactivity trends, and

80

identify mineral characteristics that contribute to the catalysis of OPFR degradation.

81 Table 1. OPFRs and Structural Analogues General Structure

Structures used in this study Cl O R

O

Tris(1,3-dichloro-2-propyl) phosphate (TDCPP) a

Tris(2,3-dibromopropyl) phosphate (TBPP)

Tris(2-chloroethyl) phosphate (TCEP)

Tris(1-chloro-2-propyl) phosphate (TCPP)

Tris(2-butoxyethyl) phosphate (TBEP)

Triethyl phosphate (TEP)

Tris(2,2,2-trichloroethyl) phosphate (TTCEP) b

Tris(2,3-dichloropropyl) phosphate (2,3-TDCPP)

Triisopropyl phosphate (TIPP)

a

Structures with bolded titles are organophosphate compounds used as flame retardants (OPFRs) or precursors for plasticizers and OPFRs. b Structures with italicized labels are structural analogues of OPFRs used here for structure-reactivity analyses.

ACS Paragon Plus Environment

Environmental Science & Technology

82

Materials and Methods

83

Reagents and Minerals. A full listing of chemical reagents and mineral solids is provided in

84

Supporting Information (SI). The target compounds studied here are collectively referred to as

85

OPFRs, but three structures not used as flame retardants or precursors were included to examine

86

structure-reactivity trends.

87

Batch Reaction Kinetics. A series of abiotic batch experiments were conducted to characterize

88

the degradation of OPFRs. Detailed reaction conditions are provided in Table S1 of SI. Reactors

89

were covered with aluminum foil and conducted under dark and anoxic conditions inside a

90

thermostatted anaerobic glovebox (25C, 95% N2, 5% H2; Pd catalyst: Coy Laboratory; control

91

experiments confirmed similar rates of OPFR degradation are observed in air-saturated and anoxic

92

conditions). Individual reactors contained 30 mL water amended with 5 mM pH buffer and 10

93

mM NaCl. For homogeneous reactions (pH 2-12), degradation of OPFRs was monitored for 64 d.

94

For heterogeneous reactions (pH 6), sufficient quantity of each mineral was added to individual

95

reactors to yield suspensions with 15 m2/L of mineral surface area, and degradation of the OPFRs

96

was monitored for 16 d. Additional reactions were conducted with goethite over a wider pH range

97

(6-9), and reactions were monitored for 64 d. Mineral suspensions were continuously mixed during

98

reactions by magnetic stirrers to maximize the exposure of mineral surface to the surrounding

99

solution. After equilibrating individual batch reactors overnight, reactions were initiated by

100

introducing a mixture of the nine OPFRs. The initial concentration of each OPFR was 200 μg/L,

101

corresponding to a total initial OPFR concentration (∑[OPFR]0) of ~4 μM (detailed procedure for

102

preparing the OPFR mixture provided in the SI-1). This was much lower than the estimated

103

mineral surface site concentration of all selected minerals. For example, α-FeOOH, which has the

104

lowest surface site density (1.68 nm-2) among the six minerals, has an estimated surface site

ACS Paragon Plus Environment

Page 6 of 33

Page 7 of 33

Environmental Science & Technology

105

concentration of ~42 μM under conditions examined.40-42 Furthermore, tests conducted with

106

individual OPFRs showed no difference in reaction rates when experiments were conducted using

107

the OPFR mixture. Aliquots of reaction solution were periodically collected for LC-MS/MS

108

analysis. Procedures for quenching reactions in the collected aliquots are described in SI.

109

Separate batch reactions were prepared with individual OPFRs to identify transformation

110

products. The conditions of these reactions (Table S2 in SI) were designed to allow for collection

111

of sample aliquots after the parent OPFR degraded through ~1, 2, and 3 half-lives. At these times,

112

aliquots were collected and processed as described above before analyzing by LC-qToF-MS.

113

Analysis and Modeling. Details of LC-MS/MS and LC-qToF-MS methods are provided in SI.

114

Acid dissociation constants (pKa) of alcohol structures corresponding to the OPFR leaving groups

115

were predicted using SPARC (Archem LLC). Pseudo first-order rate constants for individual batch

116

reactions were determined by linear regression methods; second-order rate constants for base-

117

catalyzed and neutral hydrolysis of individual OPFRs were determined by least-squares regression

118

analysis using Scientist for Windows (Micromath).

119 120

Result and Discussion

121

Homogenous Aqueous Hydrolysis. Degradation of OPFRs was observed in both homogeneous

122

and heterogeneous systems, with degradation kinetics following a pseudo-first-order rate law

123

(Figures S3-S4 in SI). Rate constants (kobs, d-1) are summarized in Table S1 in the SI. In the absence

124

of mineral surfaces, OPFR degradation was limited to alkaline pH conditions. Eight of the nine

125

target compounds (TTCEP, TDCPP, TBPP, 2,3-TDCPP, TCEP, TCPP, TBEP, and TEP) showed

126

significant degradation (defined as 20%) within the 64-d monitoring period; no apparent

127

degradation of TIPP was observed within this period regardless of the pH conditions. Figure 1A

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 33

128

shows that measured kobs values increased dramatically with increasing pH, exhibiting a linear

129

relationship with changing OH- concentration. Apart from TTCEP, no significant degradation was

130

observed at neutral or acidic pH conditions. It follows that OPFR degradation can be described by

131

a second-order rate law, where kobs is proportional to [OH-]: 

132

d [OPFR ]  k obs [OPFR] = k B [OH - ][OPFR] dt

(1)

133

where kB is the apparent second-order rate constant for reaction with OH- (M-1 d-1). Such a rate law

134

is consistent with a base-catalyzed ester hydrolysis mechanism, wherein nucleophilic attack on the

135

electrophilic phosphorus atom by OH- is followed by ester bond cleavage and release an alcohol

136

leaving group.43 The reaction produces the corresponding phosphate diester and an alcohol as

137

transformation products:

(2) 138 139

Solid lines in Figure 1A show the fit of Eq 1 to measured data, and the resulting kB values are

140

provided in Table 2. It follows that hydrolysis of OPFRs is exceedingly slow at neutral and acidic

141

conditions, and the reaction is only an important contributor to natural attenuation in highly

142

alkaline waters. For the OPFR exhibiting the highest reactivity, TTCEP, appreciable rates of

143

degradation are also observed at neutral and acidic pH conditions. Below pH 7, rates are

144

independent of pH conditions, consistent with H2O reacting as the nucleophile. Thus, Eq 1 can be

145

modified to consider both nucleophiles reacting in parallel:

ACS Paragon Plus Environment

Page 9 of 33

Environmental Science & Technology



146

d [OPFR]  kobs [OPFR] = (kH2O [H 2 O]  k B [OH - ])[OPFR] dt

(3)

147

Similar rate laws have been documented for carboxylic esters and carbamates.44,45 Comparing fit-

148

derived kB and kH2O values (Table 2), it is apparent that H2O is a much weaker nucleophile for

149

TTCEP than OH-. This is consistent with the relative reactivity of different nucleophiles reported

150

previously,43,46 as commonly described by the Swain-Scott equation:

k  log  Nu  =s  nNu,CH3Br  kH 2O 

151

(4)

152

where [kNu /kH2O] is the ratio of second-order rate constants for different nucleophiles in comparison

153

to kH2O (representing reactivity of the background solvent), nNu,CH Br is the nucleophilicity of the

154

attacking nucleophile determined empirically from reactions with CH3Br, and s is the nucleophilic

155

sensitivity of the reaction of interest compared to the reference CH3Br reaction. The value of

156

nNu,CH Br for OH- is 4.2, indicating that OH- is 104.2-fold more reactive with CH3Br than H2O (s =

157

1 for reactions with CH3Br). Application of Eq 4 together with the reference nucleophilicity values

158

for OH- (4.2) and H2O (1.0) provides an estimate of s = 1.92 for reaction with TTCEP. This

159

suggests that nucleophile strength plays a greater role in reactions with TTCEP than the reference

160

reaction with CH3Br.47 Previous studies on hydrolysis of organophosphate esters trimethyl

161

phosphate (TMP) and TEP also reported greater sensitivity to nucleophile strength than CH3Br,

162

exhibiting s values of 1.35 and 1.20, respectively.32,48 Since TMP and TEP have comparable s×

163

nNu, CH Br values, we might assume other OPFRs that have similar structures with TTCEP will

164

exhibit similar sensitivity towards OH- and H2O. Using this assumption, the kH2O value of TDCPP

165

(the third most reactive target OPFR) can be estimated to be 110-6 M-1d-1, corresponding to a half-

3

3

3

ACS Paragon Plus Environment

Environmental Science & Technology

166

life of ~30 y at pH < 10, consistent with the fact that no TDCPP degradation was observed within

167

the 64-d monitoring period under these conditions.

168

To our knowledge, this is the first confirmation that degradation of commonly detected OPFRs

169

follow a second-order rate law for base-catalyzed hydrolysis. Although a recent report by Su and

170

coworkers33 included measurements of reaction half-lives of TDCPP and TCEP at pH 13, they

171

were unable to measure rate constants at lower pH conditions (pH 7-11) to confirm a first-order

172

dependence on [OH-]33. This is surprising in light of the present findings where the observed half-

173

lives at pH 11 (~5.5 d for TDCPP, ~15 d for TCEP) are much shorter than the time period that the

174

authors monitored. The source of the discrepancy is unclear, but findings from the current study

175

are consistent with the expected pH-dependence for base-catalyzed hydrolysis reactions (i.e., 10-

176

fold change in kobs value with each integer change in pH conditions). A full comparison between

177

the half-lives of other OPFRs are provided in Figure S5 in SI. In the absence of minerals, the

178

observed half-lives of most OPFRs ranged from 0.02 – 170 d between pH 9-12. However, under

179

neutral pH conditions, most OPFRs appear to be rather inert. Table S3 shows that, apart from

180

TTCEP, the estimated half-lives of the target OPFRs range from 1 – 5,800 years for pH 6 - 8. Thus,

181

in the absence of a suitable catalyst, OPFR hydrolysis in natural aquatic environment is not

182

expected to be an important natural attenuation mechanism.

ACS Paragon Plus Environment

Page 10 of 33

Page 11 of 33

Environmental Science & Technology

183 Table 2. Second-order rate constants for base-catalyzed and uncatalyzed hydrolysis of OPFRs measured in homogeneous solutions. kB (M-1d-1) kH2O (M-1d-1) TTCEP 3.48 (±0.86) x 10-4 4.20 (±0.90 a)  104 TDCPP

1.42 (±0.27)  102

TBPP 2,3-TDCPP TCEP TCPP TBEP TEP

1.38 (±0.18)  103 1.27 (±0.19)  102 7.65 (±3.20)  101 2.46 (±0.60)  100 4.53 (±0.21)  100

TIPP

6.95 (±1.46)  10-1 c < 3.47  10-1

d

< 1.25  10-4 b < 1.25  10-4 < 1.25  10-4 < 1.25  10-4 < 1.25  10-4 < 1.25  10-4 < 1.25  10-4 < 1.25  10-4

a

Uncertainties represent one standard deviation of the least squares fit-derived values determined using Scientist for Windows (Micromath). b

Estimated by dividing the minimum observable kobs threshold value by H2O concentration (55.5 M). c Estimated based on kobs value measured at pH 12. d Estimated by dividing the minimum observable kobs threshold value by [OH-] at the highest pH examined (pH 12).

ACS Paragon Plus Environment

Environmental Science & Technology

A

B

184

Figure 1. (A) Influence of pH conditions on the kinetics of homogeneous aqueous hydrolysis of OPFRs. (B) Time courses for degradation of TDCPP in the presence of metal (hydr)oxide soil minerals at pH 6. Reaction conditions: [OPFR]0 = 200 g/L, 25C, 15 m2/L mineral solid surface area, 0.01 M NaCl. Solid lines in panel A represent model fits of Eqs 1 and 3 (resulting second-order rate constants listed in Table 2). The criterion threshold for the minimum value of kobs that could be accurately measured corresponds to a half-life of 200 d. Lines in panel B represent pseudo-first order model fits (resulting rate constants listed in Table S1 in SI). Error bars represent the triplicate-averaged standard error. Downward arrow in panel A indicates that the rate constant is an upper-bound estimate. ACS Paragon Plus Environment

Page 12 of 33

Page 13 of 33

Environmental Science & Technology

185

Mineral-Catalyzed Hydrolysis. Previous reports have shown that selected soil minerals can be

186

potent catalysts of hydrolysis reactions,35,36,39 so a series of metal (hydr)oxide phases were

187

screened for their effects on OPFR degradation. Figure 1B shows a representative group of time

188

courses for TDCPP degradation observed at pH 6 in the absence and presence of metal (hydr)oxide

189

minerals (other timecourses provided in Figure S5 in SI). Whereas the predicted half-lives (Table

190

3) of most OPFRs in homogeneous solution at pH 6 are estimated to exceed 10 y, half-lives are

191

reduced to 50 d). d Uncertainties represent triplicate-averaged standard errors.

192 193

Among the minerals screened, OPFRs degraded most rapidly in the presence of Fe(III)

194

(hydr)oxides (α-FeOOH, γ-FeOOH, and α-Fe2O3), whereas no apparent catalysis was observed in

195

the presence of 15 m2 L-1 of the aluminum and silicon oxides (γ-Al2O3 and SiO2). TiO2 catalyzed

196

degradation of TBPP and TBEP, but had no apparent effect on the other OPFRs. Within the three

197

Fe(III) (hydr)oxides, γ-FeOOH displayed greatest catalytic effect. The variation in reaction rate

198

between Fe(III) (hydr)oxides is possibly due to the different position of Fe(III) centers within each

199

oxide complexes,49 which can generate different catalytic effects to selected compounds.38 The

200

percentage of wetting surface on each Fe(III) (hydr)oxides is another reason that might affect their

ACS Paragon Plus Environment

Environmental Science & Technology

201

ability to catalyzing hydrolytic reactions.50 As with homogeneous solutions, no TIPP degradation

202

was observed with any of the mineral surfaces, providing further indication that this particular

203

structure is very resistant to hydrolysis. More surprising was that no degradation of TCEP or TEP

204

was observed in mineral suspensions, despite being sensitive to base-catalyzed hydrolysis. In

205

general, much less variation in rates of degradation are observed among the target OPFR structures

206

in suspensions of individual minerals (i.e., range of measured kobs values varies ~10-fold, Figure

207

2B) than in the homogeneous base-catalyzed reactions (range of measured kobs values varies

208

~100,000-fold, Figure 1A). The compressed range of reactivities is attributed to decreased

209

importance of leaving group acidity in the rate-controlling step of the surface-catalyzed hydrolysis

210

reactions. This is consistent with OPFR-surface coordination promoting ester bond cleavage,

211

possibly by stabilizing formation of the resulting alkoxide (-OR) group. The finding that mineral-

212

catalyzed hydrolysis is observed for TCPP but not TCEP, despite the latter exhibiting higher base-

213

catalyzed reactivity, suggests that surface-TCPP interactions have a more pronounced effect on

214

ester cleavage than surface-TCEP interactions. Hydrolysis of other classes of organophosphate

215

contaminants (e.g., pesticides, nerve agents) has also been reported to be associated with metal

216

oxide surfaces. For example, the presence of Al and Ti oxides increased the hydrolysis rate of

217

chlorpyrifos-methyl oxonate by ~10-fold at pH 7 while the presence of goethite caused a ~3-fold

218

increase.39 On the other hand, the presence of Fe(III) (hydr)oxide minerals reduced the half-life of

219

p-nitrophenyl phosphate, a phosphate monoester, by more than 30-fold, and Al and Ti oxides

220

caused similar catalytic effects as Fe(III) (hydr)oxides.38 In contrast, no mineral catalysis was

221

observed for degradation of the pesticide demeton-S.51 These studies further support the

222

conclusion that hydrolysis-promoting interactions between OPFRs and mineral oxides are

223

compound and phase selective.

ACS Paragon Plus Environment

Page 14 of 33

Page 15 of 33

Environmental Science & Technology

224

Goethite (α-FeOOH) was further used as a representative Fe(III) (hydr)oxide mineral to examine

225

the influence of solution conditions and heterogeneous reaction mechanism. Figure 2A shows the

226

concentration change of TDCPP over the 64-d monitoring period in goethite suspensions

227

(comparable kinetic profiles for other OPFRs are provided in Figure S6). Most of the target OPFRs

228

showed accelerated degradation in the presence of goethite over the studied pH range (pH 6-8),

229

but the pH-dependent trends (Figure 2B) are muted in comparison to the effects of pH on

230

homogeneous solution. Furthermore, kobs values increase slightly when pH drops from pH 8 to 6,

231

opposite to the direction of the trend observed in alkaline homogenous solutions. The lack of strong

232

pH dependence is similar to earlier reports organophosphate hydrolysis catalyzed by metal

233

(hydr)oxides surfaces,38,39 and consistent with the pH-independent adsorption of nonionic

234

solutes.52,53 Like other nonionic organics, adsorption of the target OPFRs in the aqueous metal

235

(hydr)oxide suspensions is generally weak (i.e., extent of adsorption was too low to measure, being

236

less than the variability in the analytical measurements),52 but surface interactions are still critical

237

to promoting hydrolysis reactions. The lack of strong pH dependence also suggests that H2O

238

(rather than OH-) is the attacking nucleophile. The small variation in degradation rates observed

239

with changing pH may result from changes in weak hydrogen bonding interactions that occur upon

240

protonation/deprotonation of surface functional groups.

241

The range of OPFR half-lives were between 0.56 and 40 d, with TTCEP at pH 6 being the

242

shortest and TCPP at pH 8 being the longest in the goethite suspensions (Figure 2B). Their

243

corresponding kobs values were calculated and compared with the kobs values estimated in

244

homogeneous solutions at the same conditions (Figure 2C). The significant difference (i.e. 10 to

245

10,000,000-fold) between those kobs values shows goethite-catalyzed hydrolysis is much faster at

246

environmentally relevant conditions.

ACS Paragon Plus Environment

Environmental Science & Technology

A

B

C

247

Figure 2. Influence of pH conditions on goethite-catalyzed hydrolysis of OPFRs. Detailed reaction conditions listed in Table S1; [OPFR]0 = 200 μg/L. (A) Observed time courses for dissipation of TDCPP, (B) Effect of pH on kobs values, (C) Differences between log(kobs) values measured in goethite suspensions versus homogeneous solutions at the same pH conditions. Error ACS Paragon Plus Environment bars represent triplicate-averaged standard errors.

Page 16 of 33

Page 17 of 33

Environmental Science & Technology

248

Transformation Products. Alkaline homogeneous batch reactors and goethite suspension (pH

249

6) batch reactors were prepared and amended with individual target OPFRs, and samples were

250

collected after ~1, 2, and 3 reaction half-lives to screen for transformation products (TPs) using

251

high resolution LC-qToF-MS. Potential TPs were analyzed from each qToF-MS signal and

252

compared with the unreacted samples as controls to avoid biased interpretation of the raw data,

253

despite hydrolytic TPs (i.e., OP diester, OP monoester) being the most plausible in this study.

254

Table 4 summarizes the TPs detected for each target OPFR. This analysis shows that all OPFR

255

transformation products detected were consistent with hydrolysis reactions, verifying that OH- and

256

goethite are acting principally to catalyze such transformations. Figure 3A-B show qToF-MS

257

spectra of two TPs identified during TDCPP reactions in goethite suspension (corresponding

258

MS/MS spectral fragmentation patterns are provided in SI). Figure 3C shows how MS spectral

259

peak areas associated with these two TPs grow over three consecutive half-lives of TDCPP

260

degradation, providing further confirmation that they are transformation products resulting from

261

degradation of the parent OPFR. qToF-MS analysis identified the corresponding OP diesters as

262

the major TP formed over the first three half-lives for most OPFRs in both homogeneous and

263

heterogeneous reactions, similar to the degradation products observed for other OP compounds.54

264

This indicates that the diesters are less reactive than their parent triesters. For TCPP, the

265

corresponding monoester was also detected in alkaline homogeneous solution, consistent with

266

further hydrolysis of the initial diester TP. Moreover, phosphate, the terminal hydrolysis

267

endproduct, was also detected in selected homogeneous alkaline reactions (TDCPP, 2,3 TDCPP,

268

and TCEP) and goethite-catalyzed reactions (TDCPP, 2,3 TDCPP, and TCPP), despite the strong

269

affinity between Fe(III) based minerals and phosphate.55,56 This is due to the very high sensitivity

270

of the analytical method used (LC-qToF-MS) in this study, which can measure many aqueous

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 33

271

constituents to sub-ppt levels. Tests confirmed that sufficient residual aqueous phosphate remains

272

in solution after equilibration to be detected by LC-qToF-MS (e.g., 0.11 µM aqueous phosphate is

273

detected after equilibrating 0.46 µM phosphate with 1 g/L goethite for 2 d). The fact that the

274

corresponding monoesters were not detected under any of these reaction conditions indicates that

275

either the monoesters are unstable and rapidly hydrolyze to the phosphate terminal product or that

276

the phosphate ion results from an alternative reaction pathway for the parent OPFR or diester. OP

277

monoester and phosphate have been detected previously as hydrolysis products for other

278

organophosphate contaminants, but only at more extreme solution conditions (e.g., 75 0C, pH 4).57

279

To our knowledge, this is the first time OP monoester and phosphate have been confirmed as TPs

280

of organophosphate hydrolysis at environmentally relevant conditions.

281 Table 4. OPFR transformation products identified by LC-qToF-MS in alkaline homogeneous solutions and in goethite suspensions at neutral pH conditions.a Target OPFR TTCEP TDCPP TBPP 2,3-TDCPP TCEP TCPP TBEP TEP TIP

Diester TP Alkaline -FeOOH solution suspension             

Monoester TP Alkaline -FeOOH solution suspension



a

PO43- TP Alkaline -FeOOH solution suspension 



 

 

Detailed conditions for alkaline solution and goethite suspension reactions for each target OPFR are listed in Table S2 in SI.

ACS Paragon Plus Environment

Page 19 of 33

Environmental Science & Technology

A

B

C

282

Figure 3. LC-ESI(+)-qToF-MS spectra for TPs identified during TDCPP reaction in goethite suspensions. Reaction conditions are provided in Table S2 in SI. (A) MS spectrum of the TDCPP diester observed at t = 3 day, (B) MS spectrum of phosphate observed at t = 6 day, (C) Timecourse for TDCPP degradation and the corresponding evolution of peak areas associated with the two TPs. Confirmation MS/MS spectra are provided in the SI. Error bars represent the triplicateaveraged standard errors.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 33

283

Influence of OPFR Structure. Results presented in Figures 1-2 demonstrate that both base-

284

and mineral-catalyzed degradation reaction rates are dependent upon OPFR structure. This is

285

especially the case for base-catalyzed reactions where kB values determined for the target OPFRs

286

range from 12), and the strong dependence on the estimated pKa values for base-catalyzed hydrolysis

ACS Paragon Plus Environment

Page 24 of 33

Page 25 of 33

Environmental Science & Technology

368

reactions support mechanism 2. Clearly, formation of the unstable alkoxide intermediate during

369

ester bond cleavage is a major rate-determining factor, so direct surface interactions with this group

370

may be critical to promoting cleavage under non-alkaline conditions. Still, mechanism 1 cannot be

371

excluded, and additional experimental and theoretical investigations are needed to further

372

discriminate between the two pathways.

373 374

Environmental Significance

375

Growing detection of OPFRs in natural environments has increased concerns related to

376

potentially adverse human health and ecosystem effects, which are exacerbated by the limited

377

information on their attenuation mechanisms in natural systems. Given their ester-based structure,

378

hydrolysis was a logical candidate for controlling degradation of OPFRs in aquatic systems.

379

However, while hydrolysis of OPFRs was previously documented in highly alkaline solutions,

380

results presented here confirm that aqueous-phase hydrolysis is not expected to be a major sink for

381

these contaminants in aquatic systems. At the same time, results presented for the first time here

382

document the potential importance of common Fe(III) hydr(oxide) minerals as catalysts of OPFR

383

degradation via hydrolytic pathways. These phases are ubiquitous in many soils, so these reactions

384

may be a critical sink for OPFRs released into such systems (e.g., by stormwater percolation,

385

irrigation with treated wastewater). The fact that three different Fe(III) hydr(oxide) minerals all

386

exert significant catalytic effects suggests that such reactions will be widespread in Fe-rich soil

387

systems. Moreover, high surface area Si and Al oxide minerals failed to exert a catalytic effect,

388

indicating that OPFRs will be more persistent in soil and aquifer systems dominated by these

389

phases. Such findings need to be confirmed with whole soil samples of varying mineralogical

390

composition.

ACS Paragon Plus Environment

Environmental Science & Technology

391

Findings from this study also indicate that OPFR diesters are likely to be a stable product of

392

OPFR hydrolysis processes, both in alkaline waters and Fe-rich soil systems. It follows that

393

concentrations of the diesters may exceed the parent OPFRs in many environments, yet very

394

limited information is available on the fate or toxicity of OP diesters in comparison to the parent

395

triesters, and their occurrence has been detected in human bodies, promoting environmental health

396

concerns.64,65 Based on these findings, it is recommended that OPFR occurrence studies add the

397

analysis of diesters to their protocols to provide a more comprehensive assessment of mass fluxes

398

in important environmental systems. This also leads to a recommendation for development of

399

authentic reference standards for diester TPs.

400

The observed structure reactivity trends, especially in alkaline solutions, can be potentially

401

applied to select less persistent OPFRs or design new OPFRs that will have lower persistence when

402

released into natural environments. The non-halogenated structures (TEP and TIPP), which

403

possess the least acidic leaving groups, are resistant to both alkaline and mineral-catalyzed

404

hydrolysis reactions. Thus, use of TEP might be discouraged based upon this finding. At the same

405

time, halogenated structures are often found to be less biodegradable than non-halogenated

406

analogues, but further study is needed to address the relative importance of biodegradation

407

processes and structure-reactivity trends before a definitive recommendation can be made.

408

ACS Paragon Plus Environment

Page 26 of 33

Page 27 of 33

Environmental Science & Technology

409

Associated Content

410

Supporting Information. Experimental conditions and results of kinetics experiments, sample

411

quenching and storage protocols, description of analytical methods and characteristics of

412

transformation products and their MS/MS spectrums, figures illustrating degradation of OPFRs in

413

homogeneous and heterogeneous reactions, prediction and comparison of OPFRs half-life in the

414

absence of minerals, and a figure illustrating the relationship between kobs and leaving group

415

acidity in mineral suspensions are provided. This material is available free of charge via the

416

Internet at http://pubs.acs.org.

417 418

Author Information

419

Corresponding Author

420

*E-mail: [email protected]; phone: (303)-384-2226

421 422

Acknowledgments

423

Financial support was provided by the United States Department of Agriculture (Agriculture and

424

Food Research Initiative, Number 2014-67019-24024). Xin Xiao (Zhejiang University), Karl

425

Oetjen, and Christopher Higgins (Colorado School of Mines) provided assistance with mass

426

spectrometry analysis.

427 428 429 430 431

ACS Paragon Plus Environment

Environmental Science & Technology

432

References

433 434

1. van der Veen, I.; de Boer, J., Phosphorus flame retardants: Properties, production, environmental occurrence, toxicity and analysis. Chemosphere 2012, 88, (10), 1119-1153.

435 436 437

2. Rodríguez, I.; Calvo, F.; Quintana, J. B.; Rubí, E.; Rodil, R.; Cela, R., Suitability of solidphase microextraction for the determination of organophosphate flame retardants and plasticizers in water samples. Journal of Chromatography A 2006, 1108, (2), 158-165.

438 439

3. Mihajlovic, I.; Fries, E., Atmospheric deposition of chlorinated organophosphate flame retardants (OFR) onto soils. Atmos Environ 2012, 56, 177-183.

440 441 442

4. Schreder, E. D.; La Guardia, M. J., Flame Retardant Transfers from U.S. Households (Dust and Laundry Wastewater) to the Aquatic Environment. Environ. Sci. Technol. 2014, 48, (19), 11575-11583.

443 444 445

5. Regnery, J.; Püttmann, W., Occurrence and fate of organophosphorus flame retardants and plasticizers in urban and remote surface waters in Germany. Water Research 2010, 44, (14), 40974104.

446 447 448

6. Cristale, J.; Álvarez-Martín, A.; Rodríguez-Cruz, S.; Sánchez-Martín, M. J.; Lacorte, S., Sorption and desorption of organophosphate esters with different hydrophobicity by soils. Environmental Science and Pollution Research 2017, 24, (36), 27870-27878.

449 450 451

7. Cristale, J.; García Vázquez, A.; Barata, C.; Lacorte, S., Priority and emerging flame retardants in rivers: Occurrence in water and sediment, Daphnia magna toxicity and risk assessment. Environment International 2013, 59, 232-243.

452 453 454

8. Rodil, R.; Quintana, J. B.; Concha-Graña, E.; López-Mahía, P.; Muniategui-Lorenzo, S.; Prada-Rodríguez, D., Emerging pollutants in sewage, surface and drinking water in Galicia (NW Spain). Chemosphere 2012, 86, (10), 1040-1049.

455 456

9. Salamova, A.; Ma, Y. N.; Venier, M.; Hites, R. A., High Levels of Organophosphate Flame Retardants in the Great Lakes Atmosphere. Environ. Sci. Technol. Lett. 2014, 1, (1), 8-14.

457 458 459

10. Marklund, A.; Andersson, B.; Haglund, P., Organophosphorus Flame Retardants and Plasticizers in Swedish Sewage Treatment Plants. Environ. Sci. Technol. 2005, 39, (19), 74237429.

460 461 462

11. Kim, U. J.; Oh, J. K.; Kannan, K., Occurrence, Removal, and Environmental Emission of Organophosphate Flame Retardants/Plasticizers in a Wastewater Treatment Plant in New York State. Environ. Sci. Technol. 2017, 51, (14), 7872-7880.

463 464 465 466

12. Stepien, D. K.; Regnery, J.; Merz, C.; Püttmann, W., Behavior of organophosphates and hydrophilic ethers during bank filtration and their potential application as organic tracers. A field study from the Oderbruch, Germany. Science of The Total Environment 2013, 458-460, (Supplement C), 150-159.

ACS Paragon Plus Environment

Page 28 of 33

Page 29 of 33

Environmental Science & Technology

467 468 469

13. Sundkvist, A. M.; Olofsson, U.; Haglund, P., Organophosphorus flame retardants and plasticizers in marine and fresh water biota and in human milk. J. Environ. Monit. 2010, 12, (4), 943-951.

470 471 472

14. Meeker, J. D.; Cooper, E. M.; Stapleton, H. M.; Hauser, R., Urinary Metabolites of Organophosphate Flame Retardants: Temporal Variability and Correlations with House Dust Concentrations. Environ Health Perspect 2013, 121, (5), 580-585.

473 474 475 476

15. Ma, Y. L.; Jin, J.; Li, P.; Xu, M.; Sun, Y. M.; Wang, Y.; Yuan, H. D., Organophosphate ester flame retardant concentrations and distributions in serum from inhabitants of Shandong, China, and changes between 2011 and 2015. Environmental Toxicology and Chemistry 2017, 36, (2), 414-421.

477 478 479 480

16. Papachlimitzou, A.; Barber, J. L.; Losada, S.; Bersuder, P.; Deaville, R.; Brownlow, A.; Penrose, R.; Jepson, P. D.; Law, R. J., Organophosphorus flame retardants (PFRs) and plasticisers in harbour porpoises (Phocoena phocoena) stranded or bycaught in the UK during 2012. Marine Pollution Bulletin 2015, 98, (1-2), 328-334.

481 482

17. Dishaw, L. V.; Macaulay, L. J.; Roberts, S. C.; Stapleton, H. M., Exposures, mechanisms, and impacts of endocrine-active flame retardants. Curr Opin Pharmacol 2014, 19, 125-133.

483 484 485

18. Ta, N.; Li, C. N.; Fang, Y. J.; Liu, H. L.; Lin, B. C.; Jin, H.; Tian, L.; Zhang, H. S.; Zhang, W.; Xi, Z. G., Toxicity of TDCPP and TCEP on PC12 cell: Changes in CAMKII, GAP43, tubulin and NF-H gene and protein levels. Toxicol Lett 2014, 227, (3), 164-171.

486 487 488

19. Kojima, H.; Takeuchi, S.; Itoh, T.; Iida, M.; Kobayashi, S.; Yoshida, T., In vitro endocrine disruption potential of organophosphate flame retardants via human nuclear receptors. Toxicology 2013, 314, (1), 76-83.

489 490 491

20. Zhang, Q.; Lu, M. Y.; Dong, X. W.; Wang, C.; Zhang, C. L.; Liu, W. P.; Zhao, M. R., Potential Estrogenic Effects of Phosphorus-Containing Flame Retardants. Environ. Sci. Technol. 2014, 48, (12), 6995-7001.

492 493 494 495

21. Dishaw, L. V.; Powers, C. M.; Ryde, I. T.; Roberts, S. C.; Seidler, F. J.; Slotkin, T. A.; Stapleton, H. M., Is the PentaBDE replacement, tris (1,3-dichloro-2-propyl) phosphate (TDCPP), a developmental neurotoxicant? Studies in PC12 cells. Toxicology and Applied Pharmacology 2011, 256, (3), 281-289.

496 497 498 499

22. Van den Eede, N.; Heffernan, A. L.; Aylward, L. L.; Hobson, P.; Neels, H.; Mueller, J. F.; Covaci, A., Age as a determinant of phosphate flame retardant exposure of the Australian population and identification of novel urinary PFR metabolites. Environment International 2015, 74, 1-8.

500 501 502

23. Ross, P. S., Fireproof killer whales (Orcinus orca): flame-retardant chemicals and the conservation imperative in the charismatic icon of British Columbia, Canada. Canadian Journal of Fisheries and Aquatic Sciences 2006, 63, (1), 224-234.

503 504

24. Fu, J.; Han, J.; Zhou, B. S.; Gong, Z. Y.; Santos, E. M.; Huo, X. J.; Zheng, W. L.; Liu, H. L.; Yu, H. X.; Liu, C. S., Toxicogenomic Responses of Zebrafish Embryos/Larvae to Tris(1,3-

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 33

505 506

dichloro-2-propyl) Phosphate (TDCPP) Reveal Possible Molecular Mechanisms Developmental Toxicity. Environ. Sci. Technol. 2013, 47, (18), 10574-10582.

507 508 509 510

25. Sun, L. W.; Tan, H. N.; Peng, T.; Wang, S. S.; Xu, W. B.; Qian, H. F.; Jin, Y. X.; Fu, Z. W., Developmental Neurotoxicity of Organophosphate Flame Retardants in Early Life Stages of Japanese Medaka (Oryzias Latipes). Environmental Toxicology and Chemistry 2016, 35, (12), 2931-2940.

511 512

26. Fries, E.; Puttmann, W., Occurrence of organophosphate esters in surface water and ground water in Germany. J. Environ. Monit. 2001, 3, (6), 621-626.

513 514 515

27. Kawagoshi, Y.; Fukunaga, I.; Itoh, H., Distribution of organophosphoric acid triesters between water and sediment at a sea-based solid waste disposal site. Journal of Material Cycles and Waste Management 1999, 1, (1), 53-61.

516 517 518 519

28. Jurgens, S. S.; Helmus, R.; Waaijers, S. L.; Uittenbogaard, D.; Dunnebier, D.; Vleugel, M.; Kraak, M. H. S.; de Voogt, P.; Parsons, J. R., Mineralisation and primary biodegradation of aromatic organophosphorus flame retardants in activated sludge. Chemosphere 2014, 111, 238242.

520 521 522

29. Takahashi, S.; Miura, K.; Abe, K.; Kera, Y., Complete detoxification of tris(2-chloroethyl) phosphate by two bacterial strains: Sphingobium sp strain TCM1 and Xanthobacter autotrophicus strain GJ10. J Biosci Bioeng 2012, 114, (3), 306-311.

523 524 525

30. Regnery, J.; Puttmann, W., Occurrence and fate of organophosphorus flame retardants and plasticizers in urban and remote surface waters in Germany. Water Research 2010, 44, (14), 40974104.

526 527 528

31. Cristale, J.; Dantas, R. F.; De Luca, A.; Sans, C.; Esplugas, S.; Lacorte, S., Role of oxygen and DOM in sunlight induced photodegradation of organophosphorous flame retardants in river water. Journal of Hazardous Materials 2017, 323, (Part A), 242-249.

529 530 531

32. Mabey, W.; Mill, T., Critical review of hydrolysis of organic compounds in water under environmental conditions. Journal of Physical and Chemical Reference Data 1978, 7, (2), 383415.

532 533 534

33. Su, G.; Letcher, R. J.; Yu, H., Organophosphate Flame Retardants and Plasticizers in Aqueous Solution: pH-Dependent Hydrolysis, Kinetics, and Pathways. Environ. Sci. Technol. 2016, 50, (15), 8103-8111.

535 536

34. Saint-Hilaire, D.; Jans, U., Reactions of three halogenated organophosphorus flame retardants with reduced sulfur species. Chemosphere 2013, 93, (9), 2033-2039.

537 538 539

35. Qian, C.; Sanders, P. F.; Selber, J. N., Accelerated degradation of organophosphorus pesticides with sodium perborate. Bulletin of Environmental Contamination and Toxicology 1985, 35, (1), 682-688.

540 541

36. Kim, K.; Tsay, O. G.; Atwood, D. A.; Churchill, D. G., Destruction and Detection of Chemical Warfare Agents. Chemical Reviews 2011, 111, (9), 5345-5403.

ACS Paragon Plus Environment

of

Page 31 of 33

Environmental Science & Technology

542 543 544

37. Smolen, J. M.; Stone, A. T., Divalent metal ion-catalyzed hydrolysis of phosphorothionate ester pesticides and their corresponding oxonates. Environmental Science and Technology 1997, 31, (6), 1664-1673.

545 546

38. Baldwin, D. S.; Beattie, J. K.; Coleman, L. M.; Jones, D. R., Phosphate Ester Hydrolysis Facilitated by Mineral Phases. Environ. Sci. Technol. 1995, 29, (6), 1706-1709.

547 548 549

39. Smolen, J. M.; Stone, A. T., Metal (hydr)oxide surface-catalyzed hydrolysis of chlorpyrifos-methyl, chlorpyrifos-methyl oxon, and paraoxon. Soil Science Society of America Journal 1998, 62, (3), 636-643.

550 551

40. Liger, E.; Charlet, L.; Van Cappellen, P., Surface catalysis of uranium(VI) reduction by iron(II). Geochim Cosmochim Ac 1999, 63, (19), 2939-2955.

552 553 554 555

41. Peacock, C. L.; Sherman, D. M., Copper(II) sorption onto goethite, hematite and lepidocrocite: a surface complexation model based on ab initio molecular geometries and EXAFS spectroscopy1 1Associate editor: D. L. Sparks. Geochim Cosmochim Ac 2004, 68, (12), 26232637.

556 557 558

42. Vasudevan, D.; Stone, A. T., Adsorption of 4-Nitrocatechol, 4-Nitro-2-Aminophenol, and 4-Nitro-1,2-Phenylenediamine at the Metal (Hydr)Oxide/Water Interface: Effect of Metal (Hydr)Oxide Properties. Journal of Colloid and Interface Science 1998, 202, (1), 1-19.

559

43. Hine, J. S., Physical Organic Chemistry. McGraw-Hill: 1962.

560 561 562

44. Klausen, J.; Meier, M. A.; Schwarzenbach, R. P., Assessing the Fate of Organic Contaminants in Aquatic Environments: Mechanism and Kinetics of Hydrolysis of a Carboxylic Ester. Journal of Chemical Education 1997, 74, (12), 1440.

563 564

45. Dittert, L. W.; Higuchi, T., Rates of hydrolysis of carbamate and carbonate esters in alkaline solution. Journal of Pharmaceutical Sciences 1963, 52, (9), 852-857.

565 566

46. Haag, W. R.; Mill, T., Some reactions of naturally occurring nucleophiles with haloalkanes in water. Environmental Toxicology and Chemistry 1988, 7, (11), 917-924.

567 568 569

47. Swain, C. G.; Scott, C. B., Quantitative Correlation of Relative Rates. Comparison of Hydroxide Ion with Other Nucleophilic Reagents toward Alkyl Halides, Esters, Epoxides and Acyl Halides1. Journal of the American Chemical Society 1953, 75, (1), 141-147.

570 571

48. Faust, S. D.; Gomaa, H. M., Chemical hydrolysis of some organic phosphorus and carbamate pesticides in aquatic environments. Environmental Letters 1972, 3, (3), 171-201.

572 573

49. Zinder, B.; Furrer, G.; Stumm, W., The coordination chemistry of weathering: II. Dissolution of Fe(III) oxides. Geochim Cosmochim Ac 1986, 50, (9), 1861-1869.

574 575

50. Torrents, A.; Stone, A. T., Hydrolysis of phenyl picolinate at the mineral/water interface. Environ. Sci. Technol. 1991, 25, (1), 143-149.

ACS Paragon Plus Environment

Environmental Science & Technology

576 577 578

51. Dannenberg, A.; Pehkonen, S. O., Investigation of the heterogeneously catalyzed hydrolysis of organophosphorus pesticides. Journal of Agricultural and Food Chemistry 1998, 46, (1), 325-334.

579 580 581

52. Van Emmerik, T. J.; Angove, M. J.; Johnson, B. B.; Wells, J. D., Sorption of Chlorpyrifos to Selected Minerals and the Effect of Humic Acid. Journal of Agricultural and Food Chemistry 2007, 55, (18), 7527-7533.

582 583

53. Kovaios, I. D.; Paraskeva, C. A.; Koutsoukos, P. G.; Payatakes, A. C., Adsorption of atrazine on soils: Model study. Journal of Colloid and Interface Science 2006, 299, (1), 88-94.

584 585 586

54. Kirby, A. J.; Medeiros, M.; Mora, J. R.; Oliveira, P. S. M.; Amer, A.; Williams, N. H.; Nome, F., Intramolecular General Base Catalysis in the Hydrolysis of a Phosphate Diester. Calculational Guidance to a Choice of Mechanism. J Org Chem 2013, 78, (4), 1343-1353.

587 588

55. Zeng, L.; Li, X.; Liu, J., Adsorptive removal of phosphate from aqueous solutions using iron oxide tailings. Water Research 2004, 38, (5), 1318-1326.

589 590 591

56. Kang, S. K.; Choo, K. H.; Lim, K. H., Use of Iron Oxide Particles as Adsorbents to Enhance Phosphorus Removal from Secondary Wastewater Effluent. Separation Science and Technology 2003, 38, (15), 3853-3874.

592 593 594

57. Grzyska, P. K.; Czyryca, P. G.; Purcell, J.; Hengge, A. C., Transition state differences in hydrolysis reactions of alkyl versus aryl phosphate monoester monoanions. Journal of the American Chemical Society 2003, 125, (43), 13106-13111.

595 596

58. Lee Wolfe, N.; Zepp, R. G.; Paris, D. F., Use of structure-reactivity relationships to estimate hydrolytic persistence of carbamate pesticides. Water Research 1978, 12, (8), 561-563.

597 598 599

59. Donarski, W. J.; Dumas, D. P.; Heitmeyer, D. P.; Lewis, V. E.; Raushel, F. M., Structureactivity relationships in the hydrolysis of substrates by the phosphotriesterase from Pseudomonas diminuta. Biochemistry 1989, 28, (11), 4650-4655.

600 601 602

60. Hilal, S. H.; Karickhoff, S. W.; Carreira, L. A., A Rigorous Test for SPARC's Chemical Reactivity Models: Estimation of More Than 4300 Ionization pKas. Quantitative StructureActivity Relationships 1995, 14, (4), 348-355.

603 604 605

61. Mäkie, P.; Persson, P.; Österlund, L., Adsorption of trimethyl phosphate and triethyl phosphate on dry and water pre-covered hematite, maghemite, and goethite nanoparticles. Journal of Colloid and Interface Science 2013, 392, (Supplement C), 349-358.

606 607 608 609

62. Evers, A.; Hancock, R. D.; Martell, A. E.; Motekaitis, R. J., Metal ion recognition in ligands with negatively charged oxygen donor groups. Complexation of iron(III), gallium(III), indium(III), aluminum(III), and other highly charged metal ions. Inorganic Chemistry 1989, 28, (11), 2189-2195.

610 611 612

63. Suh, J.; Chun, K. H., Metal-ion catalysis by blocking inhibitory reverse paths in the hydrolysis of 3-carboxyaspirin. Journal of the American Chemical Society 1986, 108, (11), 30573063.

ACS Paragon Plus Environment

Page 32 of 33

Page 33 of 33

Environmental Science & Technology

613 614 615

64. Su, G.; Letcher, R. J.; Yu, H., Determination of organophosphate diesters in urine samples by a high-sensitivity method based on ultra high pressure liquid chromatography-triple quadrupole-mass spectrometry. Journal of Chromatography A 2015, 1426, 154-160.

616 617 618

65. Van den Eede, N.; Neels, H.; Jorens, P. G.; Covaci, A., Analysis of organophosphate flame retardant diester metabolites in human urine by liquid chromatography electrospray ionisation tandem mass spectrometry. Journal of Chromatography A 2013, 1303, 48-53.

619

ACS Paragon Plus Environment