Biodegradation and Uptake of the Pesticide Sulfluramid in a Soil

Feb 8, 2018 - We also evaluated potential losses from freeze-drying by analyzing ... by comparing MS/MS product ions and retention times to that of PF...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIVERSITY OF TOLEDO LIBRARIES

Article

BIODEGRADATION AND UPTAKE OF THE PESTICIDE SULFLURAMID IN A SOIL/CARROT MESOCOSM Itsaso Zabaleta, Ekhine Bizkarguenaga, Deborah B. O. Nunoo, Lara Schultes, Juliana Leonel, Ailette Prieto, Olatz Zuloaga, and Jonathan P. Benskin Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b03876 • Publication Date (Web): 08 Feb 2018 Downloaded from http://pubs.acs.org on February 9, 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.

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 26

Environmental Science & Technology

1 2

BIODEGRADATION AND UPTAKE OF THE PESTICIDE SULFLURAMID IN A SOIL/CARROT MESOCOSM

3 Itsaso Zabaleta1,2*, Ekhine Bizkarguenaga1,2*, Deborah B. O. Nunoo2, Lara Schultes2, Juliana Leonel3, Ailette Prieto1,4, Olatz Zuloaga1,4, Jonathan P. Benskin2

4 5 6 7 8

1

Department of Analytical Chemistry, Faculty of Science and Technology, University of the Basque Country (UPV/EHU), Bilbao, Spain. 2

9 10

Department of Environmental Science and Analytical Chemistry (ACES), Stockholm University, Stockholm, Sweden.

11 12

3

Department of Geoscience, Universidade Federal de Santa Catarina, Florianópolis, Brazil.

13 14

4

Research Centre for Experimental Marine Biology and Biotechnology (PIE), University of the Basque Country (UPV/EHU), Plentzia, Spain.

15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31

* Corresponding authors: [email protected] Tel.: 0034 946015551 Fax: 0034 946013500 [email protected] Tel.: 0034 946015551 Fax: 0034 946013500

32 1

ACS Paragon Plus Environment

Environmental Science & Technology

33

Page 2 of 26

TOC Art

34

2

ACS Paragon Plus Environment

Page 3 of 26

Environmental Science & Technology

35

ABSTRACT

36

N-ethyl perfluorooctane sulfonamide (EtFOSA) is the active ingredient of Sulfluramid, a

37

pesticide which is used extensively in South America for control of leaf-cutting ants. Despite

38

being a known precursor to perfluorooctane sulfonate (PFOS), the importance of EtFOSA as a

39

source of environmental PFOS remains unclear. In the present work, uptake, leaching, and

40

biodegradation of EtFOSA and its transformation products were assessed over 81 days in

41

soil/carrot (Daucus carota ssp sativus) mesocosms for the first time. Experiments performed in

42

the presence of carrot produced PFOS yields of up to 34 % using a technical EtFOSA standard,

43

and up to 277 % using Grão Forte, a commercial Sulfluramid bait formulation containing

44

0.0024 %

45

sulfonamide (FOSA), and perfluorooctanoic acid (PFOA) also formed over the course of the

46

experiments. Leachate contained low levels of transformation products and a high FOSA:PFOS

47

ratio, consistent with recent observations in Brazilian surface water. In carrot, the more

48

hydrophilic transformation products (e.g. PFOS) occurred primarily in the leaves, while the

49

more hydrophobic products (e.g. FOSA, FOSAA and EtFOSA) occurred in peel and core.

50

Remarkably, isomer-specific analysis revealed that the linear EtFOSA isomer biodegraded

51

significantly faster than branched isomers. These data collectively show that the application of

52

Sulfluramid baits can lead to the occurrence of PFOS in crops and in the surrounding

53

environment, in considerably higher yields than previously thought.

EtFOSA.

Perfluorooctane

sulfonamido

acetate

(FOSAA),

perfluorooctane

54 55

KEYWORDS: EtFOSA, Sulfluramid, soil, biodegradation, carrot, PFAS isomers

56 57

Supporting Information: Materials and reagents, sample extraction, and supporting figures

58

and tables.

3

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 26

59

INTRODUCTION

60

Perfluorooctane sulfonate (PFOS; C8F17SO3-) has attracted considerable international regulatory

61

and scientific attention due to its widespread occurrence and links to adverse health effects in

62

humans and wildlife.1 On account of these risks, PFOS and its precursors were added to

63

Annex B of the United Nations Stockholm Convention on Persistent Organic Pollutants in

64

2009,2 and to the list of priority hazardous substances in the EU water policy Directive

65

2013/39/EU in 2013.3 Presently, manufacturing of PFOS and PFOS-precursors continues in

66

some countries under Stockholm Convention production and use exemptions. These

67

contemporary sources of PFOS are poorly characterized and may pose a considerable ongoing

68

risk to humans and wildlife.4–6

69

Brazil is currently among the main global producers of the PFOS-precursor N-ethyl

70

perfluorooctane sulfonamide (EtFOSA; C8F17SO2NHC2H5), which is the active ingredient in

71

Sulfluramid, a commercial pesticide. EtFOSA is produced from the starting material

72

perfluorooctane sulfonyl fluoride (POSF; C8F17SO2F), which is imported into Brazil from

73

China. Brazil holds an exemption under Annex B to manufacture and use Sulfluramid to manage

74

leaf-cutting ants from the genus Atta ssp. and Acromyrmex spp., which jeopardize agricultural

75

activities in parts of Latin America.7 Alternatives to Sulfluramid are not currently available; and

76

while the country is phasing out production and use of baits for domestic use, commercial

77

applications in agriculture are expected to continue into the foreseeable future.8

78

The manufacture and use of Sulfluramid in Brazil from 2004 to 2015 is expected to produce

79

between 167 and 603 tonnes of PFOS.9,10 However, there are considerable uncertainties

80

surrounding these estimates, owing to an absence of manufacturing data but also a lack of

81

information surrounding PFOS yields in the environment. For example, the only study to

82

investigate soil biodegradation of EtFOSA reported very low (4 mol%) yields of PFOS

83

following incubation of a pure standard of EtFOSA over 182 days.11 Studies involving other

84

perfluorooctane sulfonamides have demonstrated considerably higher PFOS yields (and in some

85

cases formation of perfluoroalkyl carboxylic acids) under biological12–15 and abiotic16,17

86

conditions (reviewed elsewhere18). Among these studies, a soil-vegetable mesocosm study

87

involving perfluorooctane sulfonamide (FOSA), the N-dealkylation product of EtFOSA,

88

demonstrated that FOSA was totally degraded to PFOS in presence of carrot while no

89

degradation was observed in absence of vegetable.12 Collectively these data suggest that in the

90

natural environment (and in particular in the presence of a vegetable crop), yields of PFOS from

91

EtFOSA may be considerably higher than 4%. However, to date there are no soil-vegetable

92

mesocosm studies involving EtFOSA or commercial sulfluramid formulations. 4

ACS Paragon Plus Environment

Page 5 of 26

Environmental Science & Technology

93

Data on the environmental occurrence of EtFOSA in South and Central America are also

94

scarce.9 One study reported low but detectable levels of EtFOSA in air samples from Costa

95

Rica19 while others have observed elevated concentrations of potential EtFOSA transformation

96

products in both South American surface waters9,20,21 and biota.22 The unusually high ratio of

97

FOSA:PFOS in Brazilian surface water is hypothesized to be a marker of Sulfluramid use, but

98

this requires further investigation. To date, there are no studies which have examined the

99

occurrence of EtFOSA or its transformation products around agricultural regions where

100

Sulfluramid is deployed. Such data, together with improved estimates of EtFOSA production

101

and PFOS degradation yields, are clearly needed in order to determine the importance of

102

Sulfluramid as a source of environmental PFOS.

103

Despite some recent work involving leaching and plant uptake12,23–25 of PFAAs, only two

104

studies have investigated the fate of PFOS-precursors in a soil-vegetable mesocosm.12,26 There

105

are no peer-reviewed studies investigating the fate and behavior of commercial Sulfluramid

106

baits in soil-vegetable mesocosm. Considering the use pattern of Sulfluramid, this information

107

is urgently needed in order to characterize the likelihood of environmental contamination

108

arising from the use of this commercial pesticide. The purpose of this study was to investigate

109

biodegradation, leaching, plant uptake, and distribution of EtFOSA and its transformation

110

products in soil-carrot mesocosms. Experiments were performed with both technical standards

111

and a commercially available, characterized Sulfluramid bait, providing new estimates for

112

EtFOSA-derived PFOS formation under environmentally-relevant conditions. Furthermore,

113

since commercial EtFOSA is manufactured as an isomeric mixture, we studied the fate and

114

behavior of individual isomers using isomer-specific analytical methodologies. To our

115

knowledge this is the first isomer-specific study of any PFAS in a soil and/or soil-vegetable

116

mesocosm. Collectively, these data provide valuable new insight on the importance of EtFOSA

117

as a contemporary source of PFOS.

118

EXPERIMENTAL METHODS

119

Standards and Reagents

120

A full list of chemicals and reagents is provided in the Supporting Information (SI). Technical

121

EtFOSA (95%) originated from Lancaster Synthesis (Wyndham, NH).27 Isomeric purity could

122

not be determined due to a lack of purified branched isomer standard. Grão Forte, a commercial

123

Sulfluramid formulation (determined to contain 0.0024 % EtFOSA (∑branched+linear

124

isomers)) was obtained from Insetimax Industrial Chemicals (Brazil). L-EtFOSA, L-FOSA,

125

perfluorooctane sulfonamido acetate (L-FOSAA), perfluorodecanoate (PFDA), characterized

126

isomeric mixtures of PFOS and perfluorooctanoate (PFOA), and the isotopically labeled 5

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 26

127

standards of EtFOSA, FOSA, PFOS, PFOA and PFDA were purchased from Wellington

128

Laboratories (Guelph, ON, Canada) (Table S1).

129

Experimental Design and Soil Fortification

130

A total of six, 81-day mesocosm experiments were carried out concurrently (Table 1).

131

Experiments 1 and 2 were conducted in duplicate and involved incubation of technical EtFOSA

132

(3.8 mg kg-1; ∑branched+linear isomers) in microbially-active soil (referred to herein as `active´

133

soil), with and without carrot (Daucus carota ssp sativus), respectively. Experiments 3, 4, and 5,

134

were designed as control incubations: Experiment 3 was conducted in duplicate and involved

135

fortifying soil autoclaved at 112 °C under vacuum for 4 hours (referred to herein as `inactive´

136

soil) with technical EtFOSA (3.8 mg kg-1, ∑branched+linear isomers) to monitor leaching and

137

abiotic losses. Experiments 4 and 5 were single blank experiments which contained unfortified

138

active soils without and with carrot, respectively, to monitor contamination introduced from

139

water and air. Finally, Experiment 6 was carried out in duplicate and involved fortifying active

140

soil containing carrot with the commercial Sulfluramid formulation Grão Forte (0.0024 %

141

EtFOSA, ∑branched+linear isomers). All experiments contained PFDA (100 ng g-1) which

142

functioned as an internal negative control, as previously described.15

143

An acidic sandy loam soil (pH = 5.7 ± 0.2), which is common to regions of Brazil28 was used in

144

the present work. Soil chemistry parameters are provided in Table S2. In Experiments 1-3

145

(Table 1), soils were weighed, covered with acetone and fortified with technical EtFOSA in

146

order to achieve a 3.8 mg kg-1 nominal concentration. After stirring for 24 hours, the soil-

147

acetone mixture was placed under a fume hood in order to let the solvent evaporate. Soil was

148

then aged for one week. For Experiment 6, 10 g of Grão Forte was added to the surface of each

149

pot containing 2 kg of soil.

150

Plant Cultivation and Sampling

151

All experiments were performed in a climate-controlled greenhouse with interior conditions set

152

to 25 °C / 50 % humidity during the day (14 h) and 18 °C / 60 % humidity at night (10 h). Prior

153

to germination, seeds were soaked in Milli-Q water. The washed seeds were distributed

154

randomly on dampened filter paper in a Petri dish and covered with moistened filter paper.

155

Upon germination (12 - 14 days), 4 seedlings were transplanted to each pot containing 2 kg of

156

soil (fortified or non-fortified). Each pot represented a single time point, and a total of 5 time

157

points were sampled over the course of the experiment. Pots were arranged randomly and

158

regularly watered with distilled water and Hoagland nutritive solution (Hoagland reagents and

159

preparation procedure are included in SI). Leachate from each pot was collected at the same 6

ACS Paragon Plus Environment

Page 7 of 26

Environmental Science & Technology

160

time as soil sampling on days 14, 28, 56 and 81, resulting in 14-day, 28-day, 56-day, and 81-day

161

composite leachates samples. Exact volumes collected for each pot are provided in Table S3

162

(SI). Blanks, consisting of Milli-Q water stored in the same PE bottles, were also analyzed in

163

parallel to assess background contamination. On the last two time points, carrots were collected

164

and divided into peel, core and leaf compartments. Soil was air-dried and carrots were freeze-

165

dried. All samples were stored at -80 °C prior to extraction and analysis.

166

Extraction and Cleanup

167

The extraction procedure for determination of EtFOSA and its transformation products in Grão

168

Forte bait is described in the SI. Validation of this method is described elsewhere.29 Soil and

169

carrot extractions were performed according to Avendaño and Liu,11 with slight modifications.

170

Briefly, 0.5 g of dried sample was fortified with 2 ng of isotopically-labeled standards and 8 mL

171

acetonitrile (ACN), and then sonicated for 20 min. After sonication, the mixture was placed in

172

an angular shaker for 40 min, centrifuged at 2900 rpm for 20 min and the ACN was transferred

173

into a clean 15 mL polypropylene test tube. The procedure was repeated using ACN with

174

25 mM sodium hydroxide and the supernatants were combined. The extracts were placed in a

175

Turvobap LV evaporator and reduced to dryness (soil extracts) or to approx. 1 mL (carrot

176

extracts). Soil extracts were reconstituted in 400 µL methanol (MeOH):Milli-Q water (1:1, v/v)

177

with 20 mM formic acid and 20 mM ammonium formate, while a portion (200 µL) of the carrot

178

extract was mixed with 200 µL of Milli-Q water containing 20 mM formic acid and 20 mM

179

ammonium formate. Extracts were transferred to microvials prior to instrumental analysis.

180

Leachate was filtered through borosilicate glass fiber filters prior to extraction using the

181

procedure described by Gilljam et al.9 (see SI for details). The filters were extracted separately

182

to assess potential sorption of target analytes, as previously described for solid matrices.

183

Instrumental Analysis

184

Quantitative analysis of EtFOSA and its transformation products was carried out by ultra

185

performance liquid chromatography-tandem mass spectrometry (UPLC-MS/MS) using a Waters

186

Acquity UPLC coupled to a Xevo TQ-S triple quadrupole mass spectrometer (Waters). The

187

method, which has been previously described,30 facilitates chromatographic separation and

188

quantification of individual PFAS isomers (see example chromatograms in Figures S1-S4, SI).

189

Extracts (10 µL) were injected onto an Ascentis Express F5 guard column (2.7 µm, 2.1 mm ×

190

0.5 cm) coupled to an Ascentis Express F5 (2.7 µm, 2.1 mm × 10 cm) analytical column

191

maintained at 30 °C. The mobile phase consisted of 20 mM formic acid and 20 mM ammonium

192

formate in Milli-Q water (mobile phase A) and 100 % MeOH (mobile phase B). The flow rate

193

was maintained at 0.25 mL min-1. The gradient profile started at 90 % A (hold time 1 min), 7

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 26

194

followed by a linear decrease to 40 % A by 3 min, then to 12 % A by 14 min and finally 0 % by

195

14.5 min (hold time 1 min). The mobile phase composition was returned to initial conditions by

196

16.5 min and then equilibrated by 21.5 min. The mass spectrometer was operated under selected

197

reaction monitoring (SRM) mode, with 2 to 5 transitions per analyte (see Table S1, SI).

198

Quality Control

199

Prior to analysis of samples, spike/recovery experiments were performed in soil (n = 4), carrot

200

(n = 4) and glass filters (n = 4) at a fortification level of 30 ng g-1, and in water (n = 4) at a

201

fortification level of 10 ng mL-1. Limits of detection (LODs) and quantification (LOQs) were

202

estimated as the concentration producing a signal-to-noise ratio of 3 and 10, respectively (Table

203

S4). We also evaluated potential losses from freeze-drying by analyzing soil fortified with target

204

analytes with and without a freeze-drying step. Following method validation, ongoing

205

assessment of method performance was carried out through the inclusion of blanks and spiked

206

samples in every batch.

207

Isomer Nomenclature and Identification

208

In all cases, isomers were denoted by either ‘L-’ (linear isomer), ‘Br-’ (∑branched isomers), or

209

a number denoting the location of the perfluoromethyl branching point (1-, 2-, 3-, etc.).

210

Individual PFOS and PFOA isomers could be identified in chromatograms (Figure S1-S2) by

211

matching their relative retention times and MS/MS product ions to those reported

212

previously.30,31 In the case of EtFOSA (Figure S3), tentative structural assignments were made

213

by comparing MS/MS product ions and retention times to that of PFOS. For example, 6-

214

EtFOSA and 6-PFOS both produced m/z 169 product ion and eluted closest to their respective

215

linear isomers, while 1-EtFOSA and 1-PFOS both eluted between 5- and 6-isomers and

216

produced a unique m/z 419 product ion.

217

Data Handling and Statistical Analysis

218

Concentrations for a single time point in each experimental replicate were based on analysis of

219

n = 3 soil or carrot samples or n = 1 sample of composite leachate. Quantification of target

220

analytes was performed using an isotope dilution approach, with the exception of FOSAA,

221

where matrix-matched calibration approach was performed due to the lack of a homologous

222

isotopically labeled standard. Calibration curves (1/x weighting) were prepared from around the

223

limit of quantification (LOQ) to 250 ng mL-1 and determination coefficients, R2, were always in

224

the range of 0.994-0.998. Individual PFOS and PFOA isomers were determined using isomer-

225

specific calibration curves prepared from characterized technical standards (see standards and

226

reagents section). For targets where characterized isomeric mixtures were unavailable (i.e. 8

ACS Paragon Plus Environment

Page 9 of 26

Environmental Science & Technology

227

EtFOSA, FOSA and FOSAA), Σ branched isomers were quantified separately from the linear

228

isomer using a linear isomer calibration curve. In this case, the concentration of branched

229

isomers should be considered semi-quantitative, owing to differences in response factors

230

between branched and linear isomers.

231

EtFOSA rate constants were determined by fitting soil concentrations (Csoil) to the equation

232

ln(Csoil)= a - kdt, where kd is the apparent depletion rate constant, t is time, and a is a constant

233

(See Figure S5). The apparent half-life (t1/2) was calculated by dividing ln(2) by the kd. Since the

234

concentrations in replicate pots for a given experiment were not significantly different (p >

235

0.05; Analysis of Variance (ANOVA) test), t1/2 was determined for each of the replicates, and

236

these were used to calculate an average half-life and pooled standard deviation. Losses not

237

accounted for by the internal negative control (e.g. from volatilization or irreversible sorption)

238

as well as the potential for novel product formation were monitored by calculating the total

239

number of moles in the system at each time point and comparing this to the total number of

240

moles at t = 0. Finally, bioaccumulation factors (BAFs) were determined in carrot peel, core and

241

leaf as a ratio between the concentration determined in each of the carrot compartment (d.w.)

242

and the concentration detected in soil (d.w.).

243

RESULTS AND DISCUSSION

244

Quality Control

245

Spike/recovery experiments involving L-FOSAA, L-FOSA, and isomeric mixtures of EtFOSA,

246

PFOS, and PFOA resulted in internal standard-corrected percent recoveries ranging from 76-

247

109 % for soil, 49-120 % for carrot, 76-130 % for leaching water and 65-130 % for filters (see

248

Table S4), indicating good accuracy of the method. No significant differences (p > 0.05) were

249

observed between freeze-dried and non-freeze dried soils, indicating that losses during the

250

freeze-drying step were negligible (Figure S6). The internal negative control PFDA, which was

251

incubated in all the experiments to monitor losses in situ, was recovered quantitatively from all

252

pots and displayed no significant change in concentration over the course of the experiments

253

(Figure 1). Monitoring of unfortified soil (Experiment 5) and soil-carrot (Experiment 4)

254

mesocosms revealed the occurrence of PFOS PFOA, and FOSA in both soils and leachate. For

255

PFOS and FOSA, soil and leachate concentrations in unfortified experiments were always