Role of Glutamine Synthetase Isogenes and Herbicide Metabolism in

May 8, 2019 - (4) The obligate-outcrossing, self-incompatible mating system of this .... species with copy number variation of an herbicide target sit...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF LOUISIANA

Agricultural and Environmental Chemistry

Role of glutamine synthetase isogenes and herbicide metabolism in the mechanism of resistance to glufosinate in Lolium perenne L. spp. multiflorum biotypes from Oregon Caio Augusto Brunharo, Hudson K Takano, Carol A. Mallory-Smith, Franck E. Dayan, and B. D. Hanson J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.9b01392 • Publication Date (Web): 08 May 2019 Downloaded from http://pubs.acs.org on May 9, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

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

Page 1 of 38

Journal of Agricultural and Food Chemistry

1 1

Role of glutamine synthetase isogenes and herbicide metabolism in the mechanism of

2

resistance to glufosinate in Lolium perenne L. spp. multiflorum biotypes from Oregon

3 4

Caio A. C. G. Brunharo1*

5

1Department

6

ORCID: 0000-0001-9735-1648

7

Hudson K. Takano2

8

2Department

9

Colorado State University, Fort Collins, CO 80523, USA

of Crop and Soil Science, Oregon State University, Corvallis, OR 97331, USA

of Bioagricultural Sciences and Pest Management, 1777 Campus Delivery,

10

ORCID: 0000-0002-8018-3868

11

Carol A. Mallory-Smith1

12

1Department

13

Franck E. Dayan2

14

2Department

15

Colorado State University, Fort Collins, CO 80523, USA

16

ORCID: 0000-0001-6964-2499

17

Bradley D. Hanson3

18

3One

19

95616, USA

20

ORCID: 0000-0003-4462-5339

21

*Corresponding author: 3050 SW Campus Way, Crop Sciences Building 331B, Department of

22

Crop and Soil Science, Oregon State University, Corvallis, OR 97331, USA

23

[email protected]

of Crop and Soil Science, Oregon State University, Corvallis, OR 97331, USA

of Bioagricultural Sciences and Pest Management, 1777 Campus Delivery,

Shields Avenue, Department of Plant Science, MS-4, University of California, Davis, CA

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 38

2 24

ABSTRACT

25

Glufosinate-resistant Lolium perenne L. spp. multiflorum biotypes from Oregon exhibited

26

resistance levels up to 2.8-fold the field rate. One resistant biotype (MG) had an amino acid

27

substitution in the glutamine synthetase 2 (GS2), whereas the other (OR) exhibited the wild-type

28

genotype. We hypothesized that the amino acid substitution in GS2 is involved in the resistance

29

mechanism in MG, and that non-target site resistance mechanisms are present in OR. OR

30

metabolized glufosinate faster than the other two biotypes, with >75% of the herbicide

31

metabolized, compared to 50% in MG and the susceptible biotype. A mutation in the GS2 co-

32

segregating with resistance in MG did not reduce the enzyme activity, results further supported

33

by our enzyme homology models. This research supports the conclusion that a metabolism

34

mechanism of glufosinate resistance is present in OR, and that glufosinate resistance in MG is

35

not due to an altered target-site.

36

Keywords: Enzyme modelling, gene expression, herbicide absorption, herbicide degradation,

37

herbicide translocation, non-target-site resistance, phosphinothricin,

38 39 40 41 42 43 44 45 46

ACS Paragon Plus Environment

Page 3 of 38

Journal of Agricultural and Food Chemistry

3 47

INTRODUCTION

48

Lolium perenne L. spp. multiflorum (L. multiflorum) is a winter annual grass species native to

49

Europe and initially introduced to the United States as a forage crop.1 Lolium multiflorum is also

50

a major weed in agricultural areas in many parts of the world, primarily in regions of temperate

51

climate. In Oregon, L. multiflorum interferes with a wide range of crops.2 Modern agriculture

52

relies heavily on herbicides because of their efficacy and economic returns, whereas other

53

management practices may be more costly or less selective.3 Overreliance on herbicides for weed

54

management has, however, selected for herbicide-resistant L. multiflorum populations in at least

55

12 countries.4 The obligate-outcrossing, self-incompatible mating system of this diploid (2n = 2

56

= 14) species facilitates the dispersal of herbicide resistance genes within and among

57

populations.5-6

58

Glufosinate (2-amino-4-(hydroxymethylphosphinyl) butyric acid) is a commercially

59

available nonselective herbicide composed of a racemic mixture (L/D-phosphinothricin) used for

60

postemergence control of weeds in many agricultural and non-agricultural systems. The L-

61

phosphinothricin form has herbicidal activity, irreversibly inhibiting glutamine synthetases (EC

62

6.3.1.2; GS), a family of enzymes responsible for the ATP-dependent assimilation of ammonia

63

as an amide moiety of glutamine, central to the assimilation of nitrogen in plants.7 A number of

64

GS isoforms have been identified in Arabidopsis thaliana, Zea mays and L. multiflorum, and are

65

commonly classified as cytosolic (GS1) and plastidic (GS2), depending on their cellular

66

localization and specific role in plants.8-9 The GS1a isoform in Triticum aestivum, for example, is

67

involved in nitrogen remobilization during leaf senescence,10 while GS1b is associated with

68

nitrogen assimilation in roots of Oryza sativa.11 Inhibition of GS causes a dramatic decrease in

69

glutamate and glutamine, leading to the accumulation of downstream metabolites in

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 38

4 70

photorespiration (glyoxylate, phosphoglycolate and glycolate).12-13 The accumulation of high

71

levels of glyoxylate leads to the inhibition of ribulose-1,5-bisphosphate carboxylase/oxygenase

72

activase, a key enzyme in the Calvin-Benson cycle, resulting in the accumulation of reactive

73

oxygen species, lipid peroxidation of membranes and cell death.14

74

Although glufosinate has been used for over 40 years, the first glufosinate-resistant weed

75

species were confirmed in 2009 in a vegetable farm in Malaysia15 and in a Corylus avellana L.

76

orchard in Oregon.2 Despite recent efforts to understand the biological processes involved in

77

glufosinate resistance, little information has been obtained about the underlying mechanisms

78

involved in these phenotypes.16 Among the glufosinate-resistant L. multiflorum biotypes

79

identified in Oregon by Avila-Garcia et al.,2, 17 one biotype (MG) exhibited an amino acid

80

substitution in the target site, whereas the other biotype (OR) presented the wild-type genotype.

81

The single nucleotide polymorphism (SNP) found in the GS2 gene encoded for a GS isozyme

82

with an amino acid substitution at position 173 (D173N) (equivalent to position 171 on the T.

83

aestivum GS2 sequence). If GS2 is altered only in MG (compared to the GS2 sequence of the

84

susceptible L. multiflorum biotype; SFS), then it is possible that a conformational alteration in

85

GS2 leads to an altered binding affinity between the target enzyme and glufosinate. For instance,

86

a single amino acid substitution in the 5-enolpyruvylshikimate-3-phosphate synthase (EPSPS;

87

EC. 2.5.1.19) at position 102 and/or 106, located in the enzyme active site, reduced the binding

88

affinity with the inhibitor glyphosate, conferring resistance to this herbicide in several weed

89

species.18-21 Conversely, because alterations in the glufosinate target site were not found in OR, it

90

is possible that non-target site resistance (NTSR) mechanisms were selected in this population.

91

Glufosinate metabolism, for example, has been found in Daucus carota L. and Nicotiana

92

tabacum L.,22 as well as in a number of weed species.23 Although enhanced metabolism has not

ACS Paragon Plus Environment

Page 5 of 38

Journal of Agricultural and Food Chemistry

5 93

been established as a glufosinate resistance mechanism, up-regulation of certain groups of

94

xenobiotic-detoxifying enzymes may confer resistance to other herbicides.24

95

Elucidating the mechanisms of resistance to herbicides in weeds is important for

96

understanding the underlying basic biological processes in plants, and for allowing practitioners

97

to develop better weed management decisions. In this context, the objective of this research was

98

to characterize the putative target-site and NTSR mechanisms of glufosinate resistance in L.

99

multiflorum biotypes from Oregon.

100 101

MATERIALS AND METHODS

102

Plant material

103

Previously characterized2, 17 L. multiflorum biotypes were used in this research: (i) the

104

glufosinate-resistant biotype MG that has an amino acid substitution in the plastidic GS2

105

(D173N); (ii) the glufosinate-resistant biotype OR that does not exhibit mutations in GS2; and

106

(iii) the susceptible biotype SFS. The reported GR50 values (herbicide rate required to reduce

107

plant growth by 50%) for MG, OR and SFS were 0.45, 0.49 and 0.15 kg ai ha-1, respectively.2, 17

108 109

In vivo glutamine synthetase activity - dose-response assay

110

A dose-response experiment was conducted to compare the response of MG, OR and SFS

111

to glufosinate. Plants were treated at the BBCH-23 growth stage25 in a spray cabinet calibrated to

112

deliver 187 L ha-1, and kept at 30/20 C and 13/11 h (day/night) in a greenhouse throughout the

113

experiment. Eight glufosinate rates (Liberty 280 SL, Bayer CropScience, Research Triangle PK,

114

NC) ranging from 0 to 2,240 g ai ha-1 were applied to single-plant experimental units. Enzyme

115

activity, as percentage of the control, was evaluated using a modified procedure outlined in

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 38

6 116

Dayan et al.,26 that relies on the formation of γ-glutamyl hydroxamate and further reaction with

117

acidic ferric chloride. Twenty-four hours after herbicide treatment (HAT), crude leaf extracts

118

were obtained from 200 mg of fully expanded leaves in the presence of 333 µL extraction buffer

119

[50 mM tris(hydroxymethyl)methylamine (99%, Bio-Rad, Hercules, CA), 1 mM

120

ethylenediaminetetraacetic acid (99%, Bio-Rad, Hercules, CA), 2 mM dithiothreitol (99%,

121

Thermo Scientific, Waltham, MA), 10 mM MgCl2 (99%, Fisher Scientific, Waltham, MA), 5%

122

polyvinylpyrrolidone (w/v) (99%, Sigma Aldrich, St Louis, MO)], filtered with

123

Miracloth®(EMD Millipore, San Diego, CA), centrifuged at 12,000  g for 10 min at 4 C and

124

kept on ice throughout the procedure. Twenty microliters of crude leaf extract were added to 180

125

µL of assay buffer [25 mM imidazole-HCL (pH 7.5) (99%, Fisher Scientific), 50 mM L-

126

glutamine (99%, Sigma Aldrich), 4 mM MnCl2 (99%, Fisher Scientific), 5 mM adenosine

127

diphosphate (99%, Sigma Aldrich), 40 mM sodium arsenate (99%, Fisher Scientific), 25 mM

128

hydroxylamine (99%, VWR, Radnor, PA)] and incubated for 30 min at 30 C. After the

129

incubation period, 100 µL of ferric chloride reagent [250 mM trichloroacetic acid (100%, Sigma

130

Aldrich), 200 mM anhydrous ferric chloride (99%, Fisher Scientific), 0.5 M HCl (100%, Sigma

131

Aldrich)] were added to each sample to stop the reaction before spectrophotometric measurement

132

at 540 nm. Each treatment was replicated three times and the experiment was repeated.

133 134 135

In vitro glutamine synthetase activity – dose-response assay To evaluate the role of D173N in the mechanism of resistance to glufosinate in MG, a GS

136

enzyme assay comparing MG, OR and SFS was performed in vitro to isolate other potential

137

mechanisms that each biotype might exhibit to reduce glufosinate damage. Because glufosinate

138

inhibits the GS isoforms with similar inhibition constant (Ki) values,27 an alteration in the

ACS Paragon Plus Environment

Page 7 of 38

Journal of Agricultural and Food Chemistry

7 139

binding kinetics between glufosinate and GS2 should be quantifiable with this assay. To test this

140

hypothesis, GS inhibition by glufosinate was measured in vitro from plant leaf extracts.26 Briefly,

141

200 mg of leaf tissue were ground with extraction buffer, filtered with Miracloth® and

142

centrifuged. Twenty microliters of crude leaf extract were added to 180 µL of assay buffer in the

143

absence of L-glutamine, along with 5 µL of glufosinate at different doses (to yield final

144

concentrations of 0, 1, 3, 10, 30, 100 or 300 µM), and incubated for 30 min at 30 C to allow

145

glufosinate binding to the GS isoforms. After the incubation period with the inhibitor, 20 µL of

146

500 mM L-glutamine was added and incubated for another 30 min at 30 C. Absorbance was read

147

at 540 nm after the addition of 100 µL of ferric chloride reagent. Three replications of each

148

glufosinate treatment were performed and the experiment was conducted twice.

149 150 151

In vitro time-dependent GS inhibition assay GS isoforms exhibit time-dependent inhibition dependent on the Ki and on the interaction

152

between the enzyme active site and the inhibitor. If an amino acid substitution was involved in

153

the mechanism of resistance to glufosinate in MG, then a differential time-dependent loss of

154

enzyme activity would be observed when MG was compared to SFS. To test this hypothesis, GS

155

activity was quantified over time in MG, OR and SFS. A 34 factorial design (three biotypes and

156

four incubation periods) with three replications. Plant extracts were obtained by grinding plant

157

tissue in the presence of the previously described GS extraction buffer, filtered with Miracloth®

158

and centrifuged. Twenty microliters of crude leaf extract were added to 180 µL of assay buffer

159

along with 5 µL of glufosinate (to yield a final concentration of 100 µM of glufosinate). Samples

160

were incubated at 30 C for 0, 5, 15 and 30 min after addition of glufosinate and the reaction was

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 38

8 161

stopped by adding 100 µL of ferric chloride reagent. Twenty microliters of 500 mM L-glutamine

162

were added and incubated for another 30 min at 30 C. Absorbance was read at 540 nm.

163 164 165

Effects of glufosinate on glutamate and glutamine levels Because glufosinate inhibits GS isoforms that are crucial to the photorespiratory nitrogen

166

cycle, we expected that glutamate and glutamine concentrations in glufosinate-affected tissues

167

should be altered. Furthermore, a differential alteration of these amino acids between

168

glufosinate-resistant and -susceptible biotypes may help to elucidate the mechanisms of

169

resistance. Therefore, glutamate and glutamine levels were quantified in MG, OR and SFS 24

170

HAT, using a method described previously.14 Plants were treated with 560 g ai ha-1 glufosinate

171

as described in the in vivo glutamine synthetase activity - dose-response assay section, in a 32

172

factorial (three biotypes, two herbicide treatments) for each amino acid with three replications.

173

Plants were frozen in liquid nitrogen, ground to a fine powder with mortar and pestle, and 200

174

mg of plant tissue was weighed and transferred to 2-mL Eppendorf microtubes. Five milliliters

175

of a 75:25 (v/v) methanol:water solution were added to each sample, sonicated for 30 min at

176

(400 kHz), and centrifuged at 16000  g for 10 min. Supernatant was filtered through a 0.2 µm

177

nylon filter and transferred to a 2-mL injection vial before mass spectrometry analysis.

178

A triple quadrupole liquid chromatography mass spectrometer (LC-MS/MS; LCMS-

179

8040, Shimadzu, Columbia, MD, USA) equipped with an electro spray ionization source was

180

used to analyze glutamate, glutamine and the glufosinate parent compound in the plant samples.

181

The LC–MS/MS system consisted of a Nexera X2 UPLC with 2 LC-30 AD pumps, an SIL-30

182

AC MP autosampler, a DGU-20A5 Prominence degasser, a CTO-30A column oven, and SPD-

183

M30A diode array detector coupled to an 8040-quadrupole mass-spectrometer. Separation

ACS Paragon Plus Environment

Page 9 of 38

Journal of Agricultural and Food Chemistry

9 184

occurred in a hydrophilic interaction column (iHilic-Fusion 100 × 2.1 mm, 3.5 µm; Hilicon,

185

Umeå, Sweden) at a flow rate of 0.2 mL min-1 using a linear gradient of acetonitrile (B) and 25

186

mM ammonium acetate (A): 2 min, 80% B; 8 min, 30% B; 12 min, 30% B; 12.1 min, 80% B.

187

The MRMs were optimized to 181.95 → 136.05, 147.95 → 130.10, 147.10 → 130.00 for

188

glufosinate, glutamate, and glutamine, respectively.

189 190 191

GS1a, GS1b and GS2 expression Glufosinate, as previously noted, inhibits not only the plastid form of GS (GS2), but also

192

its cytosolic counterparts. Therefore, an enhanced expression in one of the GS isogenes would

193

indicate that more enzyme is present in the cellular pool and, consequently, more glufosinate

194

must be absorbed to be effective, as is observed in other plant species with copy number

195

variation of an herbicide target-site gene.28 First, primers were designed to amplify fragments of

196

GS1a, GS1b and GS2 followed by a primer efficiency experiment. Sequence information was

197

obtained from the International Weed Genomics Consortium29-30 of L. multiflorum GS annotated

198

sequences, as well as the gene encoding the enzyme acetolactate synthase (ALS) which was used

199

as internal control gene31 (for primer information, see Table S1). Approximately 50 mg of plant

200

tissue from the youngest fully expanded leaves were collected, immediately frozen in liquid

201

nitrogen for RNA extraction (RNeasy Plant Mini Kit, Qiagen, Germantown, MD) and cDNA

202

synthesis (iSCRIPT, BIO-RAD, Hercules, CA) following manufacturer recommendations. PCR

203

reactions contained 1.5 mM MgCl2, 0.2 mM dNTP mix, 0.2 mM forward primer, 0.2 mM reverse

204

primer, 10 ng DNA template, 2 U Platinum™ Taq DNA polymerase (Platinum™ Taq DNA

205

polymerase kit, Life Technologies, Carlsbad, CA). The following PCR program was used: 10

206

min at 94 C, 35 cycles of 30 s at 94 C, 30 s at 55 C and three min at 72 C, and a final extension

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 38

10 207

step of 10 min at 72 C. PCR products sequenced resulted in mixed sequences, suggesting that

208

primers were annealing in conserved regions of multiple GS isogenes. Therefore, a cloning step

209

was performed to obtain individual sequences from GS1a, GS1b and GS2. Amplicons were

210

ligated into a vector, transformed into Escherichia coli, and further selected with blue-white

211

screen with X-Gal following the manufacturer’s protocol (TOPO TA Cloning Kit One Shot

212

TOP10, Life Technologies, Carlsbad, CA) before sequencing (BigDye Terminator version 3.1

213

Cycle Sequencing Kit, Applied Biosystems, California, USA) with the universal primer M13F(-

214

21). The sequencing data were used to design q-RT-PCR primers that amplified GS isogene

215

fragments (Table S2). The relevant sequences have been deposited to NCBI (accession number

216

MK572811-813).

217

To evaluate GS isogene expression, the plants were treated with 35 g ai ha-1 of

218

commercial formulation of glufosinate as previously described. Tissue from three individual

219

plants from each biotype was collected from nontreated and treated plants 24 HAT and the

220

experiment was conducted twice. The q-RT-PCR reactions were performed in a 7500/7500 Fast

221

Real-Time PCR System (Applied Biosystems, Foster City, California, USA) with 10 µL

222

SsoAdvanced Universal SYBR® Green Supermix (Bio-Rad, Hercules, California, USA), 1 µL

223

of forward primer, 1 µL of reverse primer (10 µM), and 10 ng of DNA in a total reaction volume

224

of 20 µL. Primer efficiency experiments were performed and all primer combinations were

225

within 100-110% efficiency (Table S2). Specificity was checked with a melt curve step in the q-

226

RT-PCR program and no sign of primer dimers was observed.

227 228

Absorption and translocation of 14C-glufosinate

ACS Paragon Plus Environment

Page 11 of 38

Journal of Agricultural and Food Chemistry

11 229

MG, OR and SFS seeds were germinated and seedlings grown in a growth chamber

230

throughout the duration of the experiment with an 11/13 h and 16/10 C (day/night) regime, and

231

light intensity of 800 µmol m-2 s-1. At the 13-BBCH stage,12 seedlings were transferred from soil

232

to a hydroponic system comprised of 40 mL vials with PTFE/silicon septa containing a dilute

233

nutrient solution and returned to the growth chamber.32 A completely randomized 35 factorial

234

(biotype  time point) design with four replications was used and the experiment was repeated. Approximately 1.4 kBq of 14C-glufosinate (specific activity 6.35 MBq mg-1) were applied

235 236

to the youngest fully expanded leaf in a 1-µL droplet 2 cm from the ligule towards the leaf apex.

237

At 0, 6, 24, 48 and 72 HAT, plants were sectioned into treated leaf, aboveground tissue, and

238

roots. The treated leaf was further sectioned into the upper portion of the treated leaf, the treated

239

middle section, and the lower portion of the treated leaf. Treated leaves were individually

240

washed with 10 mL of deionized water (diH2O) to remove non-absorbed herbicide and the rinsate

241

samples were concentrated to approximately 1 mL in an evaporator (RapidVap, Labconco,

242

Kansas City, MO) and 10 mL of scintillation cocktail was added for 14C quantification using

243

liquid scintillation spectrometry (LSS) techniques. Glufosinate exudation was quantified by

244

collecting a 10-mL aliquot from the growth solution, evaporating this aliquot to dryness and

245

adding 10 mL of scintillation cocktail for LSS analysis. Plant parts were oven-dried and oxidized

246

in a sample oxidizer (307 Sample Oxidizer, Perkin Elmer, Waltham, MA) for quantification of

247

14C

248

found in all plant parts compared to the amount applied).

using LSS techniques. The overall mass balance of 14C was 95.3 ± 2.0% (proportion of 14C

249 250

Metabolism of glufosinate

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 38

12 251

Plants were grown as described in the 14C-glufosinate absorption and translocation

252

section. Approximately 16.5 kBq of radiolabeled glufosinate in a total of 10 µL of mixture were

253

applied to the youngest fully expanded leaves. A completely randomized 33 factorial (biotype 

254

time point) design with four replications was used and the experiment was repeated. At 24, 48

255

and 72 HAT, treated leaves were sectioned and washed with 10 mL of diH2O to quantify

256

unabsorbed glufosinate and whole-plant tissue was frozen and kept in a -80 C freezer until

257

further analysis. Plant tissue was ground in liquid nitrogen using a mortar and a pestle, and

258

ground samples were transferred to 50-mL Falcon tubes. Samples received 4 mL of 2%

259

ammonium hydroxide (v/v, 28-30% NH3 basis, Sigma-Aldrich, St. Louis, MO) extraction

260

solution, and were sonicated for 45 min at 65 C. After sonication, 10 mL of dichloromethane

261

(>99.8%, Sigma-Aldrich, St. Louis, MO) were added to each sample for a liquid-liquid

262

extraction and centrifuged for 1 h at 3,800  g. Three milliliters of the upper aqueous layer

263

containing the majority (>99.9%) of the extracted 14C-glufosinate were subjected to a solid-phase

264

extraction (SPE) step in a vacuum manifold (Visiprep SPE vacuum manifold, Supelco,

265

Bellefonte, PA) set at -20 kPa. An additional 0.5 mL aliquot was collected from the aqueous

266

layer after centrifugation for mass balance quantification. SPE cartridges (Bond Elut Plexa PAX

267

500 mg, Agilent Technologies, Folsom, CA) were conditioned with 3 mL of methanol (>99.9%,

268

Sigma-Aldrich, St. Louis, MO) and equilibrated with 3 mL of diH2O before sample loading.

269

Wash steps were performed with 2 mL of diH2O and 2 mL of methanol. Finally, 14C-glufosinate

270

elution was performed with four 1-mL 10% formic acid (>95%, Sigma-Aldrich, St. Louis, MO)

271

washes in methanol. The four elution steps were combined into a 4-mL sample and were

272

subjected to a solvent exchange step, where eluted samples were dried and resuspended with

273

0.48 mL of 0.025 M sodium tetraborate (99%, Sigma-Aldrich, St. Louis, MO) adjusted to pH 10

ACS Paragon Plus Environment

Page 13 of 38

Journal of Agricultural and Food Chemistry

13 274

with sodium hydroxide (>97%, Fisher Scientific, Fair Lawn, NJ). A derivatization step (Figure

275

S1) was performed by adding 0.5 mL of 20 mM 9-fluorenylmethoxycarbonyl chloride in excess

276

(FMOC, >99%, Sigma-Aldrich, St. Louis, MO) in acetonitrile (>99.9%, Sigma-Aldrich, St.

277

Louis, MO) to each sample and incubated on a reciprocating shaker for 2 h. The pre-column

278

derivatization step with FMOC-Cl reduces the polar characteristics of glufosinate, facilitating

279

chromatographic retention and reducing the lower limit of detection. Because FMOC-Cl reacts

280

with amine groups in alkaline medium, the derivatization reaction was restricted to glufosinate,

281

as its known metabolites do not have the amine groups present on the parent compound.

282

Furthermore, by manipulating the 14C-glufosinate-FMOC retention time, it was possible to

283

reduce co-elution of peaks related to the matrix and metabolites which increased method

284

precision and accuracy. After the 2-h incubation period, reactions were stopped by the addition

285

of 20 µL of formic acid and samples injected in a HPLC system (1200 Infinity LC, Agilent,

286

Santa Clara, CA, USA) with an inline radioactivity detector (FlowStar LB 513, Berthold

287

Technologies, Bad Wildbad, Germany).

288

Analyte separation occurred in a Zorbax Eclipse Plus (C18, 250 mm  4.6 mm  5 µm)

289

with a gradient elution composed of mobile phase A (MP A) as 10% acetonitrile in 1% formic

290

acid (v/v) and mobile phase B (MP B) as 1% formic acid in acetonitrile (v/v), as follows: 100%

291

MP B from 0 to 3 min, 100% to 0% in 1 min and held for 6 min, with a total run time of 10 min

292

at 1.5 mL min-1. 14C-glufosinate-FMOC calibration curves were produced in the sample matrix

293

and subjected to the sample extraction procedure to reduce enhancement/suppression effects and

294

assess 14C-glufosinate degradation.

295 296

Data analysis

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 14 of 38

14 297

In vivo GS activity dose-response assay and in vitro GS activity dose-response assay data were

298

fit to three-parameter log-logistic models and outliers were identified and removed based on the

299

internally studentized residuals method with α = 0.05.33 In vivo dose-response data did not meet

300

the nonlinear regression assumptions (i.e. normality of residuals and homoscedasticity of

301

variance), and were transformed based on a Box-Cox transformation with λ = 0.747.34 The

302

resistance indexes (RI) were calculated for MG and OR by comparing the amount of glufosinate

303

required to inhibit the GS activity by 50% compared to SFS (I50R/I50S) using a t-test. Data from

304

the in vitro time-dependent GS inhibition assay were analyzed as a factorial 34 (three biotypes

305

by four time points) using ANOVA and comparisons were made with Tukey’s HSD test.

306

Glutamate, glutamine, and glufosinate concentrations after glufosinate treatment were analyzed

307

as a 32 factorial for each compound, where glutamate and glufosinate data required a Box-Cox

308

transformation with λ = 0.1 and 0.2, respectively, to meet the ANOVA assumptions. Gene

309

expression data were analyzed as a 32 factorial (three biotypes and two glufosinate treatments)

310

for each gene, followed by Tukey’s HSD test.

311

Absorption data were subjected to nonlinear regression comparison and regression

312

parameters including Amax (the maximum absorption of glufosinate) and t90 (time required to

313

reach 90% of maximum absorption) were used to compare the biotypes.35 Translocation data

314

were analyzed as a 35 factorial with eight replications. An ANOVA was conducted and means

315

were separated using Tukey’s HSD test. Metabolism data were subjected to ANOVA and means

316

were separated using Tukey’s HSD test. All multiple comparisons performed had α = 0.05

317

adjusted with the Bonferroni correction to control for family-wise type I errors. Data from

318

different runs were pooled in accordance with the Levene’s homogeneity of variance test with α

319

= 0.05.

ACS Paragon Plus Environment

Page 15 of 38

Journal of Agricultural and Food Chemistry

15 320 321

Homology modeling of GS2 from SFS and MG

322

Sequences of GS2 from the SFS and MG L. multiflorum biotypes were aligned to the

323

sequence of Zea mays GS1 crystal structure (2d3c.pdb)36 using the Clustal Omega multiple

324

alignment tool.37 Lolium multiflorum sequences were longer than the GS1 sequence 2d3c for which

325

a crystal structure was available, so residue 58-414 from SFS and MG L. multiflorum sequences

326

were selected for homology modeling. Pairwise alignment of these segments was obtained using

327

EMBOSS Needle37 and calculated a 74% sequence identity and 87% sequence homology between

328

these sequences (Supplemental Information 1).

329

The 3D structure models of a GS2 from SFS and MG L. multiflorum proteins were

330

developed by aligning the primary structure and building a preliminary tertiary structure of the

331

proteins using MODELLER (version 9.21)38-39 based on the crystal structure of Z. mays GS1

332

(2d3c.pdb) serving as a template.36 Models for SFS and MG L. multiflorum with the lowest

333

DOPE scores were selected for further refinement using GROMACS (version 2018.3)40-41 (see

334

Supplemental Information 2 for codes) on Dell Precision T7500 workstation equipped with

335

100GB of DDR3 ECC RAM, 24-Core Intel® Xeon® 5600 series processors and a GeForce GTX

336

1060 6GB GDDR5 DirectX 12 graphics card. Briefly, the .pdb output from Modeller was

337

converted to .gro format with accompanied topology, restraint file, and post-processed structure,

338

and submitted to molecular dynamic simulation. A virtual box extending 1 nm around the protein

339

was built, solvated using the spc216 water model, and the charges on the protein were

340

neutralized by including a 0.15 M solution of NaCl.

341 342

The structure was relaxed to ensure that no steric clashes or inappropriate geometry were present prior to molecular dynamics simulation. The system was then subjected to two two-step

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 38

16 343

equilibrations to optimize temperature and pressure. Finally, the protein was subjected to full

344

molecular dynamics for 1-ns without restraints. GS2 MG and SFS homology models were

345

evaluated using MolProbity42-43 to identify the conformation of residues with potential to reduce

346

glufosinate binding. Proteins and ligand interactions were visualized using PyMOL (The PyMOL

347

Molecular Graphics System, Version 2.0 Schrödinger, LLC).44

348 349

RESULTS

350

Glutamine synthetase activity

351

MG and OR exhibited RI’s of 2.3 and 3.3, respectively (Table 1). Little variation was

352

observed within each treatment or between experimental runs, as observed by the 95%

353

confidence intervals around the regressions (Figure S2) and the residual standard deviation of

354

1.93. GS activity reached values close to zero with the highest glufosinate dose.

355

On the other hand, there was no difference in GS inhibition by glufosinate in MG, OR

356

and SFS when tested in vitro, with little variation within treatments as observed by the 95%

357

confidence intervals around the regressions (Figure S3) (Table 2), with residual standard error of

358

9.4. No differences were observed among the biotypes when the time-dependency of GS

359

inhibition was evaluated with 100 µM of glufosinate (Figure 1).

360 361 362

Effects of glufosinate on glutamate and glutamine levels Overall, glutamate levels were greater than glutamine levels in nontreated samples.

363

Levels of free glutamate and glutamine decreased 24 h after glufosinate application in MG, OR

364

and SFS (Figure 2). Glutamate concentrations in SFS plants treated with glufosinate were four

365

times less than in OR, but similar to the concentrations in MG. There were no differences among

ACS Paragon Plus Environment

Page 17 of 38

Journal of Agricultural and Food Chemistry

17 366

the biotypes for glutamine, with or without glufosinate treatment. Glufosinate parent compound

367

concentration in OR was two less than in SFS and similar between MG and SFS (Figure 2).

368 369

GS1a, GS1b and GS2 expression

370

The overall expression of GS1a and GS1b was less than that of GS2 in all L. multiflorum

371

biotypes, with mean expression of GS2 of 0.44, GS1a of 0.14, and GS1b of 0.07 compared to the

372

normalization gene. GS1a and GS1b expression were similar among the biotypes with or without

373

glufosinate treatments (Figure 3). GS2 expression was greater in MG compared to OR and SFS

374

before glufosinate application, but was similar to the other biotypes 24 HAT. Upon glufosinate

375

treatment, GS2 expression in OR increased.

376 377

Absorption and translocation of 14C-glufosinate

378

Herbicide absorption dynamics were similar among MG, OR and SFS (Figure S4). The

379

regression estimates Amax and t90 from MG and OR were not different from those of SFS (Table

380

3). Most of the 14C-glufosinate was absorbed within 24 HAT and the maximum absorption of

381

glufosinate observed was greater than 85% for all biotypes within 72 HAT.

382

Herbicide translocation out of the treated leaf was similar among the biotypes up to 24

383

HAT (Figure 4). At 48 HAT, the resistant biotypes MG and OR had less translocation and by 72

384

HAT biotype OR translocated the least amount of 14C of the three biotypes (16.5% for OR,

385

19.2% for MG, and 20.7% for SFS). The majority of the 14C-glufosinate retained in the treated

386

leaf was translocated toward the leaf apex in all biotypes (data not shown). The maximum

387

differences in translocation were less than 5% when OR and SFS were compared at 48 and 72

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 18 of 38

18 388

HAT, with up to 20% of the radioactivity moving out of the treated leaf. Glufosinate exudation

389

from roots was negligible (data not shown).

390 391 392

Metabolism of glufosinate A high coefficient of determination was observed when 14C-glufosinate-FMOC was

393

analyzed using HPLC in line with a radioactivity detector (>99%), with a wide range of linearity

394

(43 to 813 dpm µL-1) in the L. multiflorum matrix (data not shown). Biotypes MG, OR and SFS

395

had 14C-glufosinate metabolites at all sampling times (Figure 5). At 24 HAT, no differences in

396

the amount of metabolite were observed among biotypes, whereas at 48 and 72 HAT biotype OR

397

had greater metabolite amounts compared to MG and SFS, in some instances up to 30% more.

398

The method used in this research resulted in a mass balance >90% which was a significant

399

improvement compared to previously reported analytical methods that analyzed glufosinate

400

metabolism in weed species.16

401 402 403

Homology modeling of GS2 from glufosinate SFS and MG L. multiflorum Sequences of GS2 from the SFS and MG L. multiflorum biotypes had 74% sequence

404

identity and 87% sequence homology to the Z. mays GS1 sequence, making it possible to build

405

homology models that had 1.23 root mean square deviation of the α-C atoms from the crystal

406

structure (those having the most influence on the secondary and tertiary conformation) (Figure

407

S5). After GROMACS, 98.9% (350/354) of all GS2 residues from the SFS had torsion and bond

408

angles within allowable ranges. The four outliers were residues 2 Arg, 18 Lys, 171 Ile, and 242

409

Asp (Supplemental Information 3). Similarly, 98.69% of all GS2 residues from MG had torsion

ACS Paragon Plus Environment

Page 19 of 38

Journal of Agricultural and Food Chemistry

19 410

and bond angles within allowable ranges. The 5 outliers were residues 64 Pro, 138 Asp, 142 Pro,

411

204 Pro, and 336 Ser. In both models, all of the Asn, Gln, and His were oriented correctly.

412 413 414

DISCUSSION Weed resistance to herbicides threatens food security and the sustainability of cropping

415

systems, as weed competition reduces the yield potential and aesthetic value of crops. Non-

416

target-site resistance (NTSR) is of major concern to practitioners, as it is a more general adaptive

417

trait and may confer resistance to different classes of herbicides and herbicides that have not yet

418

been discovered. Here, we report a glufosinate-resistant L. multiflorum biotype with enhanced

419

herbicide metabolism. Furthermore, our initial hypothesis that the D173N amino acid

420

substitution in GS2 was involved in the glufosinate resistance phenotype in MG was not

421

supported by our data.

422

In vivo GS activity measured in whole-plants sprayed with different rates of glufosinate

423

was consistent with resistance indices from Avila-Garcia et al. (Table 1), even though these

424

authors adopted an ammonia accumulation bioassay to quantify the resistance levels.2, 17

425

Conversely, the resistance indices in our research were considerably different from Avila-Garcia

426

et al.17 when the GS was quantified, likely induced by the large standard errors obtained by the

427

authors. No differences in GS activity from the in vitro experiment were observed among the

428

biotypes, suggesting the the D173N amino acid substitution is not involved in the mechanism of

429

glufosinate resistance in MG. Alterations in target-site enzymes are known to confer resistance to

430

other herbicides such as the protoporphyrinogen oxidase (E.C. 1.3.3.4) and EPSPS inhibitors,45-46

431

primarily due to conformational changes in the enzyme active site. As demonstrated in Z. mays,

432

GS2 active sites involve interactions of amino acid residues at positions 131, 192, 249, 291, 311

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 38

20 433

and 332 with glufosinate, whereas positions 127, 187, 251, 253, 328 and 316 interact with an

434

ADP molecule.36 Because the position 173 in L. multiflorum is equivalent to amino acid residue

435

167 in Z. mays, this amino acid substitution does not seem to interact directly with the catalitic

436

site of GS2. No crystal structure comparison has been conducted to evaluate the effect of D173N

437

in the L. multiflorum GS2, nor have studies been conducted to evaluate the activity of GS2 in

438

isolation of the other GS isozymes in MG. Our data suggest, however, that the overall combined

439

catalytic efficiency of the GS isozymes was not significantly reduced in biotype MG with

440

D173N.

441

Sequences of GS2 from the SFS and MG L. multiflorum biotypes have strong sequence

442

identity and homology to the Z. mays GS1 sequence, resulting in the production of highly

443

reliable homology models (Figure S5) suitable for examining the potential role of the D173N

444

mutation on resistance to glufosinate.17 The homology models of the GS2 from L. multiflorum

445

had binding domains that could hold both ADP and glufosinate based on the coordinate from the

446

crystal structure of Z. mays GS1 (2d3c.pdb), illustrating the highly conserved residues involved

447

in the activity of all GS (Figure 6A).36 A closer examination of the catalytic domain confirmed

448

that the interactions between ADP, glufosinate, and the three Mn atoms were maintained and that

449

the complex was stabilized via interactions to several amino acid side chains (Figure 6B).

450

While a D173N mutation in the MG biotype has been associated with resistance to

451

glufosinate,17 the biochemical characterization of GS activity from SFS and MG did not support

452

that this mutation imparted resistance to the enzyme in vitro (Table 2 and Figure S3). Analysis

453

of the tertiary structure of L. multiflorum GS2 revealed that this mutation was positioned

454

between 20.5 and 32.5 Å from either of the ligands and separated by a β-sheet and an α-helix

455

(Figure 7), which makes it very unlikely to affect GS activity or the binding of glufosinate. This

ACS Paragon Plus Environment

Page 21 of 38

Journal of Agricultural and Food Chemistry

21 456

observation is consistent with our biochemical characterization of the GS activity indicating no

457

difference in sensitivity to glufosinate between the SFS and MG biotypes. Consequently, more

458

work needs to be conducted to elucidate the potential function of the D173N mutation (if any) in

459

the MG biotype.

460

Differential expression of GS isogenes was not associated with resistance, as no

461

differences among L. multiflorum biotypes were observed after glufosinate treatment. These data

462

agree with the overall results from the enzyme activity assays, and further suggest that there is a

463

close relationship between GS transcripts and enzyme translated. Plastid GS is likely encoded by

464

a single gene in L. multiflorum,9 whereas cytoplasmic GS is encoded by a family of genes

465

depending on the plant tissue, enzyme function and developmental stage.9, 10, 47 Nord-Larsen et

466

al.9 observed that nitrogen starvation (a symptom similar to glufosinate action in plants) in L.

467

perenne increased the expression of GS1a and a decreased GS1b expression, likely because

468

GS1a is associated with nitrogen remobilization in tissues.48 In our experiments, however, GS1a

469

expression did not change after glufosinate treatment.

470

As expected, concentrations of glutamine and glutamate decreased after glufosinate

471

application. Because OR exhibited the highest resistance level, differences in the glutamate

472

concentration compared to SFS were more evident after glufosinate treatment (Figure 2). The

473

inhibition of GS activity by glufosinate inherently reduces glutamine and glutamate

474

concentration in plants,14 and a differential reduction among the biotypes would help elucidate

475

the mechanism of resistance to glufosinate. Furthermore, our experiments suggest that GS

476

enzyme activity assays are more reliable biomarkers for glufosinate resistance diagnostics than

477

glutamate and glutamine quantification.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 22 of 38

22 478

Although the physicochemical properties of glufosinate would suggest it is a phloem-

479

mobile herbicide (i.e. low pKa, low Kow),49 its ability to be translocated to other parts of the plant

480

is in fact very limited,50 explaining the observations from this research that the majority of the

481

applied herbicide was found in the treated leaf. The maximum differences in translocation were

482

less than 5% when OR and SFS were compared at 48 and 72 HAT, and these differences seem to

483

only partially explain the glufosinate resistance phenotype observed at the whole-plant level.

484

Enhanced glufosinate metabolism was observed in OR compared to SFS, and is the major

485

mechanism of resistance in OR. Furthermore, mass spectrometry data of the glufosinate parent ion

486

corroborate the metabolism experiment with radiolabeled glufosinate, as the lowest concentration

487

of the glufosinate parent compound was found in OR and the greatest concentration in SFS. L-

488

Phosphinothricin metabolites are not toxic and D-phosphinothricin seems to be stable in plants,

489

supporting the hypothesis that glufosinate metabolism could confer resistance.22, 51 Considering

490

that L-phosphinothricin has herbicidal activity and is subjected to biological degradation while D-

491

phosphinothricin is not metabolized in plants,22 a comparative analysis between biotypes seems

492

reasonable as D-phosphinothricin had similar behavior in all L. multiflorum biotypes studied and

493

the differences observed likely originated from L-phosphinothricin metabolism (Figure 5).

494

In summary, the D173N amino acid substitution does not account for the glufosinate

495

resistance in MG and the mechanism(s) of resistance in this biotype remains unclear. Our results

496

strongly suggest that the differences in glufosinate metabolism in OR are involved in the

497

mechanism of resistance to glufosinate. These results support our initial hypothesis that NTSR

498

mechanisms are involved in glufosinate resistance in OR. Ongoing research will elucidate the

499

mechanisms of resistance to glufosinate in MG, as well as the inheritance mechanisms, and the

500

genetic mechanisms contributing to the observed phenotypes in MG and OR.

ACS Paragon Plus Environment

Page 23 of 38

Journal of Agricultural and Food Chemistry

23 501 502

SUPPORTING AND SUPPLEMENTAL INFORMATION

503

Supporting information contains primer sequences for amplification of both initial GS

504

isogenes (Table S1) and q-RT-PCR (Table S2), the molecular structure of glufosinate, FMOC

505

and the derivatization product (Figure S1), in vivo and in vitro dose-response curves (Figure S2

506

and S3), glufosinate absorption (Figure S4), evaluation of the homology models using

507

MolProbity (Figure S5). Supplemental information contains the sequence analysis of SFS, MG

508

and Z. mays GS1 (Supplemental Information 1), GROMACS code used to refine the structures of

509

GS2 from SFS and MG (Supplemental information 2), and overlay of homology model of L.

510

multiflorum GS2 on crystal structure of Z. mays GS1 (Supplemental information 3).

511 512 513 514 515 516 517 518 519 520 521 522 523

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 38

24 524

REFERENCES

525 526

1. DiTomaso, J. M.; Healy, E. A. Weeds of California and Other Western States; University of California Division of Agriculture and Natural Resources: Oakland, CA, 2007. 1760p.

527 528

2. Avila-Garcia, W. V.; Mallory-Smith, C. Glyphosate-resistant Italian ryegrass (Lolium perenne) populations also exhibit resistance to glufosinate. Weed Sci. 2011, 59, 305-309.

529

3.

530 531

4. Heap, I. The International Survey of Herbicide Resistant Weeds. www.weedscience.org (accessed 2/1/2019).

532 533

5. Loureiro, I.; Escorial, M. C.; Chueca, M. C. Pollen-mediated movement of herbicide resistance genes in Lolium rigidum. PLoS One 2016, 11, e0157892.

534 535 536 537

6. Brunharo, C. A. C. G.; Hanson, B. D. Multiple herbicide–resistant Italian ryegrass [Lolium perenne L. spp. multiflorum (Lam.) Husnot] in California perennial crops: characterization, mechanism of resistance, and chemical management. Weed Sci. 2018, 66, 696701.

538 539

7. Forde, B. G.; Lea, P. J. Glutamate in plants: metabolism, regulation, and signalling. J. Exp. Bot. 2007, 58, 2339-2358.

540 541 542

8. Thomsen, H. C.; Eriksson, D.; Møller, I. S.; Schjoerring, J. K. Cytosolic glutamine synthetase: a target for improvement of crop nitrogen use efficiency? Trends Plant Sci. 2014, 19, 656-663.

543 544 545 546

9. Nord-Larsen, P. H.; Kichey, T.; Jahn, T. P.; Jensen, C. S.; Nielsen, K. K.; Hegelund, J. N.; Schjoerring, J. K. Cloning, characterization and expression analysis of tonoplast intrinsic proteins and glutamine synthetase in ryegrass (Lolium perenne L.). Plant Cell Rep. 2009, 28, 1549-1562.

547 548 549 550 551

10. Bernard, S. M.; Moller, A. L.; Dionisio, G.; Kichey, T.; Jahn, T. P.; Dubois, F.; Baudo, M.; Lopes, M. S.; Terce-Laforgue, T.; Foyer, C. H.; Parry, M. A.; Forde, B. G.; Araus, J. L.; Hirel, B.; Schjoerring, J. K.; Habash, D. Z. Gene expression, cellular localisation and function of glutamine synthetase isozymes in wheat (Triticum aestivum L.). Plant Mol. Biol. 2008, 67, 89105.

552 553 554

11. Funayama, K.; Kojima, S.; Tabuchi-Kobayashi, M.; Sawa, Y.; Nakayama, Y.; Hayakawa, T.; Yamaya, T. Cytosolic glutamine synthetase1;2 is responsible for the primary assimilation of ammonium in rice roots. Plant Cell Physiol. 2013, 54, 934-943.

555

12.

556 557 558 559

13. Masclaux-Daubresse, C.; Reisdorf-Cren, M.; Pageau, K.; Lelandais, M.; Grandjean, O.; Kronenberger, J.; Valadier, M.-H.; Feraud, M.; Jouglet, T.; Suzuki, A. Glutamine synthetaseglutamate synthase pathway and glutamate dehydrogenase play distinct roles in the sink-source nitrogen cycle in tobacco. Plant Physiol. 2006, 140, 444-456.

560 561 562

14. Takano, H. K.; Beffa, R.; Preston, C.; Westra, P.; Dayan, F. E. Reactive oxygen species trigger the fast action of glufosinate. Planta 2019, in press. https://doi.org/10.1007/s00425-01903124-3.

Zimdahl, R. L. Fundamentals of Weed Science; Elsevier/AP: Amsterdam; Boston, 2018.

Hess, F. D. Light-dependent herbicides: an overview. Weed Sci. 2000, 48, 160-170.

ACS Paragon Plus Environment

Page 25 of 38

Journal of Agricultural and Food Chemistry

25 563 564 565

15. Jalaludin, A.; Ngim, J.; Bakar, B. H. J.; Alias, Z. Preliminary findings of potentially resistant goosegrass (Eleusine indica) to glufosinate-ammonium in Malaysia. Weed Biol. Manag. 2010, 10, 256-260.

566 567

16. Jalaludin, A.; Yu, Q.; Zoellner, P.; Beffa, R.; Powles, S. B. Characterisation of glufosinate resistance mechanisms in Eleusine indica. Pest Manag. Sci. 2017, 73, 1091-1100.

568 569 570

17. Avila-Garcia, W. V.; Sanchez-Olguin, E.; Hulting, A. G.; Mallory-Smith, C. Target-site mutation associated with glufosinate resistance in Italian ryegrass (Lolium perenne L. ssp. multiflorum). Pest Manag. Sci. 2012, 68, 1248-1254.

571 572

18. Vila-Aiub, M. M.; Yu, Q.; Powles, S. B. Do plants pay fitness cost to be resistant to glyphosate. New Phytol [Online], 2018. (doi: doi.org/10.1111/nph.15733).

573 574 575

19. Cross, R. B.; McCarty, L. B.; Tharayil, N.; McElroy, J. S.; Chen, S.; McCullough, P. E.; Powell, B. A.; Bridges, W. C. A Pro(106) to Ala substitution is associated with resistance to glyphosate in annual bluegrass (Poa annua). Weed Sci. 2015, 63, 613-622.

576 577 578

20. Karn, E.; Jasieniuk, M. Nucleotide diversity at dite 106 of EPSPS in Lolium perenne L. ssp. multiflorum from California indicates multiple evolutionary origins of herbicide resistance. Front. Plant Sci. 2017, 8, e777.

579 580 581

21. Kaundun, S. S.; Zelaya, I. A.; Dale, R. P.; Lycett, A. J.; Carter, P.; Sharples, K. R.; McIndoe, E. Importance of the P106S target-site mutation in conferring resistance to glyphosate in a goosegrass (Eleusine indica) population from the Philippines. Weed Sci. 2008, 56, 637-646.

582 583 584

22. Dröge, W.; Broer, I.; Puhler, A. Transgenic plants containing the phosphinothricin-Nacetyltransferase gene metabolize the herbicide L-phosphinothricin (glufosinate) differently from untransformed plants. Planta 1992, 187, 142-151.

585 586

23. Jansen, C.; Schuphan, I.; Schmidt, B. Glufosinate metabolism in excised shoots and leaves of twenty plant species. Weed Sci. 2000, 48, 319-326.

587 588 589

34. Mersey, B. G.; Hall, J. C.; Anderson, D. M.; Swanton, C. J. Factors affecting the herbicidal activity of glufosinate-ammonium: absorption, translocation, and metabolism in barley and green foxtail. Pestic. Biochem. Physiol. 1990, 37, 90-98.

590 591 592 593

24. Iwakami, S.; Endo, M.; Saika, H.; Okuno, J.; Nakamura, N.; Yokoyama, M.; Watanabe, H.; Toki, S.; Uchino, A.; Inamura, T. Cytochrome P450 CYP81A12 and CYP81A21 are associated with resistance to two acetolactate synthase inhibitors in Echinochloa phyllopogon. Plant Physiol. 2014, 165, 618-629.

594 595 596

25. Hess, M.; Barralis, G.; Bleiholder, H.; Buhr, L.; Eggers, T.; Hack, H.; Stauss, R. Use of the extended BBCH scale - general for the descriptions of the growth stages of mono- and dicotyledonous weed species. Weed Res. 1997, 37, 433-441.

597 598 599

26. Dayan, F. E.; Owens, D. K.; Corniani, N.; Silva, F. M. L.; Watson, S. B.; Howell, J.; Shaner, D. L. Biochemical markers and enzyme assays for herbicide mode of action and resistance studies. Weed Sci. 2015, 63, 23-63.

600 601

27. Logusch, E. W.; Walker, D. M.; McDonald, J. F.; Franz, J. E. Inhibition of plant glutamine synthetases by substituted phosphinothricins. Plant Physiol. 1991, 95, 1057-1062.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 38

26 602 603 604

28. Brunharo, C. A. C. G.; Morran, S.; Martin, K.; Moretti, M. L.; Hanson, B. D. EPSPS duplication and mutation involved in glyphosate resistance in the allotetraploid weed species Poa annua L. Pest Manag. Sci. [Online], 2018. (doi.org/10.1002/ps.5284).

605 606 607 608 609

29. Ravet, K.; Patterson, E. L.; Krahmer, H.; Hamouzova, K.; Fan, L.; Jasieniuk, M.; Lawton-Rauh, A.; Malone, J. M.; McElroy, J. S.; Merotto, A., Jr.; Westra, P.; Preston, C.; VilaAiub, M. M.; Busi, R.; Tranel, P. J.; Reinhardt, C.; Saski, C.; Beffa, R.; Neve, P.; Gaines, T. A. The power and potential of genomics in weed biology and management. Pest Manag. Sci. 2018, 74, 2216-2225.

610 611

30. McElroy, S.; Sivaraj, S.; Wilkhu, S.; Chen, S.; Barnhardt, T. International Weed Genome Consortium. http://www.weedgenomics.com (accessed 03/01/2018).

612 613

31. Schmittgen, T. D.; Livak, K. J. Analyzing real-time PCR data by the comparative CT method. Nat. Protoc. 2008, 3, 1101.

614 615

32. Brunharo, C.; Hanson, B. D. Vacuolar sequestration of paraquat is involved in the resistance mechanism in Lolium perenne L. spp. multiflorum. Front. Plant Sci. 2017, 8, 9.

616

33.

617 618

34. Box, G. E. P.; Cox, D. R. An analysis of transformations. J. R. Stat. Soc. Ser. B-Stat. Methodol. 1964, 26, 211-252.

619 620

35. Kniss, A. R.; Vassios, J. D.; Nissen, S. J.; Ritz, C. Nonlinear regression analysis of herbicide absorption studies. Weed Sci. 2011, 59, 601-610.

621 622 623

36. Unno, H.; Uchida, T.; Sugawara, H.; Kurisu, G.; Sugiyama, T.; Yamaya, T.; Sakakibara, H.; Hase, T.; Kusunoki, M. Atomic structure of plant glutamine synthetase - A key enzyme for plant productivity. J. Biol. Chem. 2006, 281, 29287-29296.

624 625 626

37. Li, W.; Cowley, A.; McWilliam, H.; Uludag, M.; Buso, N.; Squizzato, S.; Gur, T.; Li, W.; Park, Y. M.; Lopez, R. The EMBL-EBI bioinformatics web and programmatic tools framework. Nucleic Acids Res. 2015, 43, 580-584.

627 628

38. Sali, A.; Potterton, L.; Yuan, F.; Van Vlijmen, H.; Karplus, M. Evaluation of comparative protein modeling by MODELLER. Proteins 1995, 23, 318-326.

629 630

39. Eswar, N.; Webb, B.; Sali, A. Comparative protein structure modeling using MODELLER. Curr Protoc Bioinformatics 2014, 47, 5.6.1-5.6.32.

631 632 633

40. Abraham, M. J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J. C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1-2, 19-25.

634 635 636

41. Hess, B.; Kutzner, C.; van der Spoel, D.; Lindahl, E. GROMACS 4: Algorithms for highly efficient, load-balanced, and scalable molecular simulation. J. Chem. Theory Comput. 2008, 4, 435-447.

637 638 639

42. Chen, V. B.; Arendall, W. B., III; Headd, J. J.; Keedy, D. A.; Immormino, R. M.; Kapral, G. J.; Murray, L. W.; Richardson, J. S.; Richardson, D. C. MolProbity: all-atom structure validation for macromolecular crystallography. Act. Crystall. Sec. D 2010, 66, 12-21.

640 641

43. Davis, I. W.; Leaver-Fay, A.; Chen, V. B.; Block, J. N.; Kapral, G. J.; Wang, X.; Murray, L. W.; Arendall, I. I. I. W. B.; Snoeyink, J.; Richardson, J. S.; Richardson, D. C. MolProbity: all-

Kutner, M. H. Applied linear statistical models; McGraw-Hill Irwin: Boston, 2005.

ACS Paragon Plus Environment

Page 27 of 38

Journal of Agricultural and Food Chemistry

27 642 643

atom contacts and structure validation for proteins and nucleic acids. Nucleic Acids Res. 2007, 35, 375-383.

644 645

44. Delano, W. L. The PyMOL molecular graphics system, DeLano Scientific: Palo Alto, CA, 2002.

646 647 648

45. Dayan, F. E.; Daga, P. R.; Duke, S. O.; Lee, R. M.; Tranel, P. J.; Doerksen, R. J. Biochemical and structural consequences of a glycine deletion in the α-8 helix of protoporphyrinogen oxidase. Biochim. Biophys. Acta 2010, 1804, 1548-1556.

649 650 651

46. Healy-Fried, M. L.; Funke, T.; Priestman, M. A.; Han, H.; Schonbrunn, E. Structural basis of glyphosate tolerance resulting from mutations of Pro(101) in Escherichia coli 5enolpyruvylshikimate-3-phosphate synthase. J. Biol. Chem. 2007, 282, 32949-32955.

652 653 654

47. Sakurai, N.; Hayakawa, T.; Nakamura, T.; Yamaya, T. Changes in the cellular localization of cytosolic glutamine synthetase protein in vascular bundles of rice leaves at various stages of development. Planta 1996, 200, 306-311.

655 656 657 658 659 660

48. Martin, A.; Lee, J.; Kichey, T.; Gerentes, D.; Zivy, M.; Tatout, C.; Dubois, F.; Balliau, T.; Valot, B.; Davanture, M.; Terce-Laforgue, T.; Quillere, I.; Coque, M.; Gallais, A.; GonzalezMoro, M. B.; Bethencourt, L.; Habash, D. Z.; Lea, P. J.; Charcosset, A.; Perez, P.; Murigneux, A.; Sakakibara, H.; Edwards, K. J.; Hirel, B. Two cytosolic glutamine synthetase isoforms of maize are specifically involved in the control of grain production. Plant Cell. 2006, 18, 32523274.

661 662

49. Bromilow, R. H.; Chamberlain, K.; Evans, A. A. Physicochemical aspects of phloem translocation of herbicides. Weed Sci. 1990, 38, 305-314.

663 664 665

50. Beriault, J. N.; Horsman, G. P.; Devine, M. D. Phloem transport of D, L-glufosinate and acetyl-L-glufosinate in glufosinate-resistant and -susceptible Brassica napus. Plant Physiol. 1999, 121, 619-627.

666 667 668

51. Dröge-Laser, W.; Siemeling, U.; Puhler, A.; Broer, I. The metabolites of the herbicide L phosphinothricin (glufosinate) - Identification, stability, and mobility in transgenic, herbicideresistant, and untransformed plants. Plant Physiol. 1994, 105, 159-166.

669 670 671 672 673 674 675

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 38

28 676

Table 1. In vivo glutamine synthetase dose-response assay with MG, OR and SFS L. multiflorum

677

biotypes. Log-logistic regression estimates/a Biotype/b

b

d

I50

RI/c

MG

1.21 ± 0.06

95.01 ± 2.09

51.17 ± 3.79

2.27 ± 0.19***

OR

1.73 ± 0.11

91.74 ± 1.76

72.71 ± 4.19

3.30 ± 0.26***

SFS

1.03 ± 0.04

101.01 ± 2.16

21.07 ± 1.44

-

678

/aEquation:

679

the upper limit, and e is the I50 (amount of glufosinate required to inhibit glutamine synthetase

680

activity by 50% in g a.i. ha-1). /bValues are means ± SE of the detransformed data. /cRI =

681

Resistance Index [e(MG or OR)/e(SFS)]. /ns not statistically significant (α = 0.05). /*** P-value