Field and Laboratory Dissipation of the Herbicide Fomesafen in the

Jun 7, 2016 - Food Chem. , 2016, 64 (25), pp 5156–5163 ..... with severe to extreme drought conditions across the Southern United States and Mexico...
0 downloads 0 Views 716KB Size
Subscriber access provided by UNIV OF NEBRASKA - LINCOLN

Article

Field and laboratory dissipation of the herbicide fomesafen in the southern Atlantic Coastal Plain (USA) Thomas L Potter, David D Bosch, and Timothy C Strickland J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b01649 • Publication Date (Web): 07 Jun 2016 Downloaded from http://pubs.acs.org on June 15, 2016

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.

Journal of Agricultural and Food Chemistry 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 34

Journal of Agricultural and Food Chemistry

Field and laboratory dissipation of the herbicide fomesafen in the southern Atlantic Coastal Plain (USA)

Thomas L. Potter,* David D. Bosch, Timothy C. Strickland

________________________________________________________________________ T.L. Potter*, D.D. Bosch, and T.C. Strickland, USDA-ARS, Southeast Watershed Research Laboratory, P.O. Box 748, Tifton, GA 31793. *Corresponding author ([email protected]).

Mention of trade names or commercial products in this article is solely for the purpose of providing specific information and does not imply recommendation or endorsement by the U.S. Department of Agriculture.

Keywords: pesticide transport, conservation tillage, strip tillage, conventional tillage, leaching, metabolites

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

ABSTRACT 1

To control weeds with evolved resistance to glyphosate, Southeastern (USA) cotton farmers have

2

increased fomesafen (5-(2-chloro-a,a,a-trifluoro-p-tolyloxy)-N-mesyl-2-nitrobenzamide) use. To

3

refine fomesafen risk assessments, data are needed that describe its dissipation following

4

application to farm fields. In our field studies relatively low runoff rates and transport by lateral

5

subsurface flow, 3 years after application. Findings suggest

10

low potential for fomesafen movement from treated fields however fate of fomesafen that

11

accumulated in subsoil and the identity of degradates are uncertain. Soil and water samples were

12

screened for degradates however none were detected.

13

2

ACS Paragon Plus Environment

Page 2 of 34

Page 3 of 34

14

Journal of Agricultural and Food Chemistry

INTRODUCTION

15

Use on crops with engineered resistance has made glyphosate the world’s most widely used

16

herbicide.1 This practice also appears to have contributed to evolved resistance in many weeds

17

and has created complex and costly problems for production of numerous crops.2,3 Highly

18

glyphosate resistant Palmer amaranth (Amaranthus palmeri) is particularly troublesome to cotton

19

growers in the Southeastern USA.4 Glyphosate continues to be used during cotton production in

20

the region but older more selective herbicides with alternate modes of action are often required

21

for Palmer amaranth control.5 This includes the protoporphyrinogen oxidase inhibitor, fomesafen

22

(5-(2-chloro-a,a,a-trifluoro-p-tolyloxy)-N-mesyl-2-nitrobenzamide).6 When applied

23

preemergence at recommended rates fomesafen did not negatively impact cotton yields and it

24

was highly effective in Palmer amaranth control.7,8

25 26

Fomesafen has been in use in the USA for nearly thirty years, primarily on soybean. It was

27

labeled, i.e. licensed for use, on cotton in 2006.9 This coincided with the first published report of

28

a glyphosate resistant Palmer amaranth biotype in central Georgia (USA).10 Since this time,

29

resistant biotypes have been identified throughout the region prompting a rapid increase in

30

fomesafen use.4,6 Between 2006 and 2012, there was an estimated 9-fold increase in Georgia and

31

a 4-fold increase nationwide.11,12 In Georgia, grower surveys indicated that >80% of all cotton

32

fields receive annual preemergence treatment with herbicides containing fomesafen.6 At the

33

recommended label rate this translates to application of 170 metric tons of fomesafen annually

34

on the approximately 0.5 million ha in cotton production.

35

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

36

Risk assessments that lead to fomesafen approval for cotton and more recently on several

37

vegetable crops concluded that human and ecological risks of exposure were within acceptable

38

margins.13-15 However, these assessments indicated that the compound’s environmental fate

39

properties make it mobile and persistent in terrestrial and aquatic environments suggesting that

40

the rapid increase in fomesafen use may have adverse impacts. Published water solubility, soil

41

Koc, and soil aerobic half-life (t1/2) are 50 mg L-1, 50 mL g-1, and 86 days, respectively.16

42 43

Human health concerns were linked to potential for contamination of water supplies. For

44

example, the New York Department of Environmental Conservation (NYDEC) determined that

45

risks of fomesafen application in an area with sandy soils underlain by a vulnerable sole-source

46

aquifer were unacceptable due to potential for the compound to leach to shallow groundwater.17

47

The decision concluded that 0.017 mg L-1 in drinking water was a screening level of concern.

48

The most sensitive surrogate species used in ecotoxicological assessments was freshwater green

49

algae. The no observable adverse effect concentration (NOAEC) was 0.010 mg L-1.18 High

50

sensitivity of green algae and fomesafen’s potential for persistence in aquatic environments were

51

confirmed in outdoor microcosm studies.19 Potential for drift during application and or runoff to

52

contact endangered plants at field margins have also contributed to proposals for relatively wide

53

buffer areas, 100 to 350 m, around treated fields.20

54 55

Several published investigations support conclusions regarding fomesafen’s relatively high

56

potential for runoff, leaching, and persistence. Rainfall simulations conducted on a silt-loam soil

57

in Georgia’s Atlantic Coastal Plain region showed that up to 5% of the herbicide applied may be

58

lost in surface runoff during a single storm event one day after fomesafen application at the label

4

ACS Paragon Plus Environment

Page 4 of 34

Page 5 of 34

Journal of Agricultural and Food Chemistry

59

rate.21 High frequency of detection in river water samples in a region in Central Canada where

60

fomesafen was used on soybean was another indication of runoff potential.22 Leaching was

61

indicated in a groundwater investigation in the North Carolina 23 and in the aforementioned

62

rainfall simulations. Before simulations 12.5 mm of irrigation was applied to selected plots

63

immediately after herbicide application. This reduced fomesafen runoff during subsequent

64

simulations by more than 2-fold. It was concluded that runoff availability was reduced by

65

movement of fomesafen into the soil with infiltrating irrigation water.21 Indications of

66

persistence were provided in field dissipation investigations. Days to 50% dissipation (DT50)

67

averaged 37, 47, and 50 days in agricultural fields in Brazil, Tennessee, and New York

68

respectively.24-26

69 70

While studies have demonstrated that fomesafen may persist in soil in farm fields and that runoff

71

and leaching may contribute to adverse water quality impacts, investigations were limited in

72

scale, scope, and region. Of particular interest to our research, is the lack of field scale

73

investigations in the Atlantic Coastal Plain region of Georgia. As noted, fomesafen use in the

74

area has increased rapidly in the past decade. The objective of the current study was to obtain

75

data from field studies needed to refine risk assessments and devise mitigation measures as

76

necessary.

77 78

MATERIALS AND METHODS

79

Site description and hydrologic monitoring. A topographic map that included a field plot

80

layout was recently published. 27,28 Plots were located in Tift County Georgia, USA

81

(N31o26’13”, W83o35’17”) within a gently sloping 1.2-ha field subdivided into two 0.6-ha fields

5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

82

running up and down the slope. Soil series identified in the fields, Tifton loamy sand (fine-

83

loamy, kaolinitic, thermic Plinthic Kandiudult), Fuquay loamy sand (loamy, kaolinitic, thermic

84

Arenic Plinthic Kandiudult), and Carnegie sandy loam (fine, kaolinitic, thermic Plinthic

85

Kandiudult) have closely related properties including permeable surface horizons with >85%

86

sand and dense argillic horizons at about 50 cm with plinthite at depth that impedes internal

87

drainage and promotes lateral subsurface flow.27 This flow was captured with 15-cm diameter

88

perforated tile drain pipes installed to a depth of 1.2 m below grade at the base of the slope and

89

along the northern boundaries of each field. Flow from each drain tile was conducted to 0.24-m

90

metal HS flumes through non-perforated 15-cm diameter drain pipe. Each 0.6-ha field was

91

further divided into three 0.2-ha plots running across the slope. Earthen berms that were

92

constructed around each plot directed surface runoff to 0.46-m metal H flumes located at NW

93

corners. Pressure transducers (Druck Inc., New Fairfield, CT, USA) linked to Campbell

94

Scientific data-loggers (Campbell Scientific, Inc., Logan, Utah, USA) measured depth of flow in

95

all flumes at 1 min intervals. Rainfall was measured with a tipping bucket rain gage (Texas

96

Electronics Inc., Dallas, TX) located 10-m from the NW field boundary.

97

Plot management. University of Georgia extension service recommendations guided

98

management. All crops planted in May and harvested in September-October (Table 1). After

99

harvest, rye (Secale cearale) was planted as a cover crop and maintained until termination by

100

glyphosate application. Prior to planting cotton in 2009, the block of plots on the NE side of the

101

field were conventionally tilled (CT) by inversion plowing to a depth of 20 cm, disking, and

102

ripping followed by seed bed formation. Cotton in the SW block was planted into 15-cm strips

103

tilled (ST) into the desiccated cover crop residue mulch. CT and ST practices had been

104

continuous use on these plots since 1999.27,28 The single fomesafen application in the study was

6

ACS Paragon Plus Environment

Page 6 of 34

Page 7 of 34

Journal of Agricultural and Food Chemistry

105

made to this cotton crop 1 h after planting at the label maximum, 0.42 kg ha-1, using the

106

commercial formulation, Reflex®(Syngenta, Greensboro, NC, USA). Application was with a

107

tractor-mounted boom sprayer. The herbicide was incorporated within 4 h with 12.5 mm

108

irrigation delivered by a solid set system. Other irrigation applied to meet crop water needs or

109

facilitate tillage ranged from 82 to 230 mm yr-1 (Table 2). All plots were converted to no-till for

110

the millet and sorghum crops (Table 1). The rye winter cover crops were terminated with

111

glyphosate, allowed to dry, and rolled prior to directly drilling seed using a Sukup no-till planter

112

(Sukup Manufacturing Co., Sheffield, Iowa, USA).

113

Water sample collection. For surface runoff, ISCO® (Lincoln, NE, USA) autosamplers were

114

programmed to collect 0.05 L into 9-L glass jars for every 1040 L that passed flumes during each

115

event.28 The minimum sample volume required for analysis (0.10 L) was equivalent to about 0.5-

116

mm runoff. No events exceeded the upper bound on the sample collection system, i.e. events

117

which would overfill the glass jars. Over the 4 years of runoff sample collection sampling

118

efficiency defined as the percent of total runoff for which samples were collected divided by

119

total flow was 88%. For lateral subsurface flow, refrigerated ISCO® automated samplers were

120

programmed to draw 0.05 L into 9-L glass collection jars at 0.5-h intervals during periods of

121

flow. Samples were retrieved at 2 to 4 day intervals. Sample collection efficiency, defined as the

122

percent of days when samples were collected compared to total days that flow occurred was

123

99%.

124

Soil sample collection. About 1 h after applying fomesafen and prior to irrigation incorporation,

125

composite soil samples to a depth of 2 cm were collected from each plot. Samples were

126

combined by tillage group, CT or ST. Soil samples was collected similarly in 2010 prior to

127

planting the millet and application of preemergence herbicides used for this crop. Four 7.5-cm

7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

128

diameter cores were collected from each of the plots to a depth of 1.2 m after sorghum harvest in

129

2012 using a tractor mounted Giddings hydraulic coring device (Giddens Machine Company,

130

Windsor, CO, USA). Polycarbonate sleeves were used. For sampling, plots were divided into

131

four quadrants with cores collected at quadrant centers. Cores were sectioned in 15 cm intervals

132

and combined by depth increment to yield a single set of 8 composite samples for each plot.

133

Soil incubations. After sieving though a 2-mm stainless steel screen, 24 field-moist 50-g

134

subsamples from the ST and CT treatments collected after fomesafen application, were placed in

135

250-mL French-square glass bottles. Soil water content in all bottles was adjusted to field

136

capacity (12% v/v) with deionized water. Fifty mL methanol was then added to three bottles

137

from each treatment groups. These bottles were sealed with Teflon®-lined caps and stored at -

138

20oC. All remaining bottles were similarly capped, shaken, and placed in a dark 25oC laboratory

139

incubator. Three, 7, 14, 28, 49, 77, and 100 days later all bottles were shaken and 50 mL

140

methanol was added to 3 bottles from each treatment group. After recapping they were also

141

stored at -20oC. The surface soil sample collected prior to planting millet in 2010 was handled

142

similarly with the exception that after placing soil subsamples in bottles, analytical grade

143

fomesafen coated on silica sand passing a 0.25-mm sieve was added to each. The target soil

144

fomesafen concentration was 1.4 µg g-1, was equivalent to the maximum label rate. Bottles were

145

then capped and shaken.

146

Soil and water sample handling and preparation. One L aliquots were used reserved for

147

analysis from all lateral subsurface flow samples. When runoff sample volume was 95% when analyzed by HPLC-MSMS

171

as described below. The surfactant and zinc powder were purchased from Sigma-Aldrich (St.

172

Louis, MO, USA).

9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

173

Instrumental analysis. Extracts were brought to room temperature, fortified with 0.2 mL 0.1%

174

formic acid, and analyzed by high-performance liquid chromatography (HPLC)-mass

175

spectrometry with a ThermoFinnigan TSQ® Quantum Mass Spectrometer system

176

(ThermoFisher, San Jose, CA, USA). The HPLC column was a 4.6 mm x 100 mm, 3.5 µm,

177

Zorbax Eclipse Plus C18 (Agilent Technologies, Santa Clara, CA, USA). Initial conditions of the

178

methanol (B) 0.1% formic acid (A) gradient elution were 90% A: 10% B. After injection, mobile

179

phase composition was changed linearly to 10% A: 90% B in 4.0 min and held isocratic for 7

180

min. Mobile phase was then increased 100% B in 0.5min and held isocratic for 2.5 min followed

181

by return to initial conditions in 1 min. The combined flow rate was 0.5 mL min-1. Data were

182

collected by selected reaction monitoring (SRM) using electrospray ionization (ESI). Negatively

183

charged parent and product ions were 437→286 (fomesafen) and 407→329 (fomesafen amine).

184

Quantification was based on the product ions. The method detection limit (MDL) for both

185

compounds in water samples was 0.01 to 0.03 µg L-1 (depending on volume of sample extracted

186

and in soil samples, 0.002 to 0.004 µg g-1 (depending on mass of soil extracted). To screen for

187

other degradation products, all extracts were also analyzed using the same HPLC conditions with

188

the mass spectrometer sequentially scanned from m/z = 100 to 500 in positive and negative ion

189

modes. Prior to these analyses the mass spectrometer response to m/z = 437 or 439 was

190

optimized while infusing a 10 µg mL-1 fomesafen solution in methanol.

191

Quality control. Target analytes were not detected in laboratory blanks prepared with HPLC

192

grade water. Calibration standards of the two target compounds were run with each sample set

193

(50 to 100 samples). Recovery of the analytes by SPE from water was evaluated by preparing

194

replicate (n=4) solutions of the compounds at 0.1 µg L-1 in HPLC grade water. The average

195

(standard error) of recoveries were fomesafen 96 (4.0) %, and fomesafen amine, 47 (6.0) %.

10

ACS Paragon Plus Environment

Page 10 of 34

Page 11 of 34

Journal of Agricultural and Food Chemistry

196

Spikes at 0.1 ug g-1 of Tifton surface soil collected in an untreated area yielded recoveries of 97

197

(3.9)% and 52 (3.8)% for fomesafen and fomesafen amine respectively.

198

Data analysis. When surface runoff occurred on consecutive days and or on weekends and

199

holidays, samples were composites of the runoff that occurred during these intervals. Measured

200

concentrations were assigned to each of the contributing days. No assignments were made for

201

days when runoff was recorded but samples were not collected due to sampler malfunction or

202

runoff