New Phytotoxic Cassane-Like Diterpenoids from Eragrostis plana

Jan 28, 2019 - New Phytotoxic Cassane-Like Diterpenoids from Eragrostis plana. Adriana Favaretto , Charles Lowell Cantrell , Frank R Fronczek , Stephe...
0 downloads 0 Views 487KB Size
Subscriber access provided by EDINBURGH UNIVERSITY LIBRARY | @ http://www.lib.ed.ac.uk

Bioactive Constituents, Metabolites, and Functions

New Phytotoxic Cassane-Like Diterpenoids from Eragrostis plana Adriana Favaretto, Charles Lowell Cantrell, Frank R Fronczek, Stephen O. Duke, David E. Wedge, Abbas Ali, and Simone Meredith Scheffer-Basso J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b06832 • Publication Date (Web): 28 Jan 2019 Downloaded from http://pubs.acs.org on January 29, 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

1

Journal of Agricultural and Food Chemistry

New Phytotoxic Cassane-Like Diterpenoids from Eragrostis plana

2 3

Adriana Favaretto†*, Charles L. Cantrell‡*, Frank R. Fronczek§, Stephen O.

4

Duke‡, David E. Wedge‡, Abbas Ali# and Simone M. Scheffer-Basso†

5 6

†University

7

Brazil.

8

‡United

9

Unit, Mississippi, USA.

of Passo Fundo, Agronomy Graduate Program, Passo Fundo, Rio Grande do Sul,

States Department of Agriculture, USDA-ARS, Natural Products Utilization Research

10

§Department

11

‡United

12

Unit, Mississippi, USA.

13

#National

14

38677, USA.

of Chemistry, Louisiana State University, Baton Rouge, Louisiana 70803, USA.

States Department of Agriculture, USDA-ARS, Natural Products Utilization Research

Center for Natural Products Research, The University of Mississippi, University, MS

15 16 17 18 19 20 1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

22

ABSTRACT

23

Eragrostis plana (Nees) is an allelopathic plant with invasive potential in South American

24

pastures. To isolate and identify phytotoxic compounds from leaves and roots of E. plana, a

25

bioassay-directed isolation of the bioactive constituents was performed. This is the first report on

26

a new diterpene carbon skeleton, the neocassanes, and of three new neocassane diterpenes,

27

neocassa-1,12(13),15-triene-3,14-dione, 1, 19-norneocassa-1,12(13),15-triene-3,14-dione, 2,

28

and 14-hydroxyneocassa-1,12(17),15-triene-3-one, 3, identified from the roots. Compounds 1, 2,

29

and 3 inhibited the growth of duckweed 50% at concentrations of 109 ± 28, 200 ± 37, and 59 ±

30

15 µM, respectively. Compound 2 was fungicidal to Colletotrichum fragariae, C. acutatum and

31

C. gloeosporioides. The compounds identified here could explain the allelopathy of E. plana.

32

The description of the newly discovered compounds, besides contributing to the chemical

33

characterization of the species, may be the first step in the study of the potential of these

34

compounds as bioherbicides.

Page 2 of 38

35 36

KEYWORDS: allelopathy, diterpenes, Eragrostis plana, phytotoxicity

2 ACS Paragon Plus Environment

Page 3 of 38

Journal of Agricultural and Food Chemistry

37

INTRODUCTION

38

Eragrostis plana (Nees), known as tough lovegrass and South African lovegrass (or capim-annoni

39

in Brazil), is an African grass initially introduced in Argentina, later in Brazil (1950-1960) and

40

dispersed throughout Uruguay. In 1971, it was considered as a good forage species and for that

41

purpose was propagated in several Brazilian states. Unfortunately, the grass has low forage quality

42

compared with native species and has several undesirable traits. It is considered the main invasive

43

plant of pastures of the Pampa Biome, with competitive characteristics that stand out over local

44

species. Beyond South America, E. plana is currently present in several regions of Asia, India and

45

the USA.1,2

46

The allelopathic effect of this species é is responsible for the ability to colonize extensive

47

areas.3-5 The use of allelopathy in weed management has been considered promising6-8 and one of

48

the major expected applications is discovery and development of bioherbicides.9 The first step in

49

the search for bioherbicides from plants is the knowledge about the chemical composition of plants

50

and the phytotoxic activity of isolated compounds. Bioassay-guided fractionation is the preferred

51

method for determining which compounds are related to phytotoxicity .10

52

Most reported allelochemicals are phenolic or terpenoid compounds.11 Terpenoids are the

53

largest and most diverse class of the plant biochemicals, with over 25,000 known structures.12 In

54

E. plana, some phenolic compounds were identified previously: caffeic, ferulic, vanillic and p-

55

coumaric acids, catechin, epicatechin, and coumarin,13 as well as rutin, quercetin, and chlorogenic,

56

ellagic and gallic acids. These compounds are present in many plant species (both allelopathic and

57

non allelopathic), and the role of most of them in allelopathy has been questioned.15 Even though

58

the allelopathic activity of E. plana and the phytotoxicity of some of its chemical constituents have

59

been described previously, the compounds responsible for the activity of the crude extracts were

3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 38

60

not isolated and identified. Therefore, the objective of this study was to isolate and identify the

61

phytotoxic compounds from E. plana by bioassay-directed isolation of the most potent bioactive

62

constituents.

63 64 65 66

MATERIALS AND METHODS Chemicals. All the solvents used in this work were HPLC grade and purchased from SigmaAldrich (St. Louis, MO).

67

Plant Material and Extraction Procedures. Leaves and roots were collected from E. plana

68

plants in vegetative stage in the city of Passo Fundo, Rio Grande do Sul state, Brazil (28°13'28.8"S

69

52°23'00.3"W), in December 2016. Identification of E. plana was according to the voucher

70

specimen deposited at the herbarium of Zoobotanical Museum Augusto Ruschi of the University

71

of Passo Fundo, under the code RSPF 11832. The material was dried in an oven at 40 °C with air

72

circulation; and then crushed in a Wiley mill. Ground roots (100.07 g) and leaves (100.31 g) were

73

consecutively extracted by soaking with solvents of increasing polarity: hexane (HEX),

74

dichloromethane/methanol (dichloromethane/MeOH) (1:1) and water (H2O), for 24 h each using

75

800 mL for each solvent. After the extraction, the extracts were filtered through 0.22 µm paper in

76

a Büchner funnel coupled to a vacuum pump and the solvents were evaporated in a rotary

77

evaporator, except for the aqueous extract, which was freeze-dried. This process yielded the

78

following extracts: LeavesHEX (2.91 g), Leavesdichloromethane/MeOH (2.89 g), LeavesH2O (5.39 g),

79

RootsHEX (1.85 g), Roots dichloromethane/MeOH (3.12 g), RootsH2O (1.57 g), which were stored at -20 oC.

80

Phytotoxicity Bioassays with Lactuca sativa and Agrostis stolonifera. The phytotoxic

81

activity of the extracts from the initial fractionation, column chromatography fractions and pure

82

compounds were evaluated using lettuce (Lactuca sativa) and creeping bentgrass (Agrostis

4 ACS Paragon Plus Environment

Page 5 of 38

Journal of Agricultural and Food Chemistry

83

stolonifera) as test plants, according described Dayan et al.16 In each well of a 24-well plate were

84

placed a filter paper (Whatman No. 1) and five lettuce seeds (L. sativa L. cv. iceberg A from

85

Burpee Seeds, Warminster, PA) or 10 mg of creeping bentgrass (A. stolonifera var. penncross from

86

Turf-Seed Inc., Gervais, OR). The test extracts or fractions were dissolved in acetone for

87

preparation of the stock solutions (10 mg/mL). Plus the 20 µL of the stock solution or acetone as

88

solvent control, 180 µL of distilled water was added to each well. The final concentration per well

89

was 1 mg/mL for extracts or fractions and 10% v/v for acetone. Plates were sealed with Parafilm,

90

and incubated at 26 °C in a growth chamber set at 173 µmol/m2/s continuous photosynthetically

91

active radiation. The visual comparison in each well with solvent control at 7 d determined the

92

phytotoxic activity. The qualitative estimation of phytotoxicity was evaluated by using a rating

93

scale of 0 to 5, where 0 = no effect and 5 = no growth or germination of the seeds. Each experiment

94

was repeated in duplicate.

95

Phytotoxicity-Guided Fractionation. Guided by the lettuce and creeping bentgrass bioassays,

96

RootsHEX and Leavesdichloromethane/MeOH extracts of E. plana were subjected to column

97

chromatography using an Isolera One system (Biotage) (Uppsala, Sweden), equipped with a UV

98

detector (254 and 280 nm) and an automatic fraction collector. Separation was performed by

99

normal-phase chromatography. The column used was a SNAP Cartridge KP-Sil, 37 mm x 157

100

mm, 50 μm irregular silica, 100 g (Biotage) and a pre-packaged SNAP Samplet Cartridge KP-Sil,

101

37 mm x 17 mm (Biotage). The separation for both extracts was performed using a gradient of

102

hexane (solvent A) and ethyl acetate (solvent B), beginning with 0-10% B, over 2310 mL, followed

103

by 20-50% B over 800 mL, then 50-100% B over 400 mL, and finishing with a MeOH wash (300

104

mL). Flow rate was 40 mL min. Portions of 22 mL each were collected in 16x150 mm test tubes.

105

According their TLC and chromatogram profiles fractions were recombined, giving seventeen

5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 38

106

fractions from RootsHex, named A to Q, and thirteen fraction from Leavesdichloromethane/MeOH, named

107

A to M. These fractions were evaluated for their phytotoxicity activity.

108

Fraction L from RootsHex was submitted to normal-phase chromatography as above, using a

109

linear gradient from 100% CHCl3 to 80/20 (CHCl3/Et2O) over 2400 mL. Flow rate was 40 mL/min.

110

By TLC similarity test tubes were grouped into three new subfractions: La, Lb and Lc. Subfraction

111

Lb was purified by HPLC on a model 1200 system (Agilent Technologies, Santa Clara, CA)

112

equipped with a quaternary pump, autosampler, diode-array detector, and vacuum degasser, using

113

a Zorbax RX-SIL (Agilent) 5 µm 9.4 x 250 nm HPLC column. Running at 4.5 mL/min. A linear

114

gradient of 100% hexane to 25% hexane/75% Et2O over 40 min at a flow rate of 4.5 mL/min and

115

with the diode array detector set at 254 and 280 nm was used to afford compound 1 (14.1 mg). 254

116

and 280 nm were used by the diode array detector for collection.

117

Fraction M from the RootsHex was also directly purified by HPLC using the same equipment

118

and HPLC column as described above for purification of fraction Lb. The linear gradient consisted

119

of 100% hexane (solvent A) and hexane/isopropyl alcohol (97:3, v/v) (solvent B) changed from

120

20-80%B over 35 min at a flow rate of 4.5 mL/min, with diode array detection at 254 and 280 nm,

121

providing compounds 2 (11.4 mg) and 3 (13.2 mg). 254 and 280 nm were used by the diode array

122

detector for collection.

123

Fraction J from Leavesdichloromethane/MeOH was identified as compound 4 (20.3 mg).

124

Chemical Analysis and Compound Identification. The isolated compounds were analyzed

125

by GC/MSD on a 7890A GC system coupled to a 5975C Inert XL MSD (Agilent). The GC was

126

equipped with a DB-5 fused silica capillary column (30 m × 0.25 mm, film thickness of 0.25 μm)

127

operated using the following conditions: injector temperature, 240 °C; column temperature, 60–

128

240 °C at 3 °C/min then held at 240 °C for 5 min; carrier gas, He; injection volume, 1 μL (splitless).

6 ACS Paragon Plus Environment

Page 7 of 38

Journal of Agricultural and Food Chemistry

129

The MS range was m/z 40-650, with a filament delay of 3 min, target TIC of 20,000, a prescan

130

ionization of 100 μs, an ion-trap temperature of 150 °C, a manifold temperature of 60 °C, and a

131

transfer line temperature of 170 °C. 1H- and 13C-NMR spectra were recorded in CDCl3 on a Bruker

132

400 MHz spectrometer (Bruker, Billerica, MA). High-resolution mass (ESI-MS) spectra of

133

isolated compounds in MeOH were acquired by direct injection of 20 μl of sample (approximately

134

0.1 mg/mL) on an JMS-T100LC AccuTOF liquid chromatograph (JEOL, Peabody, MA).

135

Neocassa-1,12(13),15-triene-3,14-dione, 1: High-resolution ESI-MS m/z 299.19964 [M+H]+

136

calculated for C20H27O2 299.20110, mass difference (mmu) -1.46; 1H NMR (400 MHz, CDCl3)

137

and 13C NMR (101 MHz, CDCl3) (Table 1).

138

Neocassa-1,12(13),15-triene-3,14-dione X-ray structure. The crystal structure of 1 was

139

determined using X-ray data collected at 180K with CuKα radiation (1.54184 Å), on a Kappa

140

Apex-II DUO diffractometer (Bruker, Billerica, MA) equipped with a microfocus Cu source.

141

C20H26O2, orthorhombic space group P212121, a=7.2346(2), b=9.1010(3), c=25.1969(8) Å, Z=4,

142

Dcalcd=1.195 g/cm3. A total of 20,857 data was collected to =68.3˚, R=0.032 for 2961 data with

143

I>2(I) of 3018 unique data (Rint=0.033) and 203 refined parameters. H atoms were visible in

144

difference maps, but were placed in idealized positions for refinement. Maximum and minimum

145

residual densities were 0.22 and -0.17 e/Å3. The absolute configuration was confirmed from the

146

Flack parameter, x=0.00(7), based on 1203 Friedel pair quotients. The CIF has been deposited at

147

the Cambridge Crystallographic Data Centre, CCDC 1821493.

148

19-norneocassa-1,12(13),15-triene-3,14-dione, 2: High-resolution ESI-MS m/z 285.18744

149

[M+H]+ calculated for C19H25O2 285.18545, mass difference (mmu) 1.99; 1H NMR (400 MHz,

150

CDCl3) and 13C NMR (101 MHz, CDCl3) (Table 1).

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 38

151

14-hydroxyneocassa-1,12(17),15-triene-3-one, 3: High-resolution ESI-MS m/z 301.21826

152

[M+H]+ calculated for C20H29O2 301.21675, mass difference (mmu) 1.51; 1H NMR (400 MHz,

153

CDCl3) and 13C NMR (101 MHz, CDCl3) (Table 1).

154

9,12,15 Octadecatrienoic acid, 4: High-resolution ESI-MS m/z 277.22184 [M+H]+ calculated

155

for C18H29O2 277.21675, mass difference (mmu) 5.08; 1H-NMR and 13C-NMR data in agreement

156

with that previously reported.17-18

157

Phytotoxicity Bioassay with Duckweed. Phytotoxicity of isolated compounds was evaluated

158

on duckweed (Lemna paucicostata (L.) Hegelm.).19 Briefly, modified Hoagland media was used

159

to grown duckweed stocks from a single colony constituted by a mother and two daughter fronds.

160

The medium was adjusted to pH 5.5 with 1 M NaOH and filtered through a 0.2 µm filter. Each

161

well of CoStar 3506 non-pyrogenic polystyrene sterile 6-well plates (Corning Inc., Corning, NY)

162

was filled with 4,850 µL of Hoagland media and, 150 µL of acetone in solvent control or 150 µL

163

of acetone containing the appropriate concentration of test compound. The final concentration of

164

acetone was 1 % (v/v). The pure compounds were tested at the concentrations 0.1; 0.3; 1; 3; 10;

165

33; 100 and 333 (µM), with three replicates for each concentration. A positive control of atrazine

166

(Sigma-Aldrich, St. Louis, MO) at the same concentrations was used. Two, three-frond colonies

167

from 4 to 5 days-old stock cultures were placed in each well. Plates were placed in an incubator

168

with white light (94.2 µE/m2/s). Total frond area per well was recorded by Scanalyzer image

169

analysis system (LemnaTec, Würselen, Germany) from days 0-7. Percentage of increase between

170

days 1 and 7 was determined relative to baseline area at day zero.

171

Electrolyte Leakage Assay. The pure compounds isolated from E. plana were tested on

172

membrane stability using cucumber cotyledon disks. Experimental units were Petri plates (60mm

173

x 15 mm, sterile, polystyrene) containing 4,850 µL of 1µM MES buffer, pH 6.5, with 2% sucrose

8 ACS Paragon Plus Environment

Page 9 of 38

Journal of Agricultural and Food Chemistry

174

plus 150 µL acetone as control, or 150 µL acetone containing the appropriate concentration of test

175

compound. The pure compounds were tested at the concentrations 10, 33, 100, 333, and 1000 µM,

176

with three replicates for each concentration and compared with the solvent control (acetone) and

177

the herbicide acifluorfen (50 μM) (Chem Service. West Chester, PA). Acifluorfen causes rapid

178

plasma membrane leakage in the light. Four-mm leaf disks were cut from cucumber cotyledons

179

with a cork borer under dim green light and 50 disks placed in each plate. Fifty disks were also

180

deposited in 5 mL of buffer in test tubes. The tubes were placed in boiling water for 15 min, and

181

allowed to cool to room temperature. A model 3082 electrical conductivity meter with an 865

182

multi-cell (Pt) probe (Amber Science, Eugene, OR) was used to take electrical conductivity

183

measurements. Conductivity measurements were carried out at the beginning of the dark

184

incubation period, after 18 h, when the samples were placed under high light intensity, and

185

subsequently, evaluated at 19, 20, 22 and 24 h from the beginning of the experiment. The data

186

were subtracted from the initial reading values, and a graph with conductivity versus time was

187

plotted. The maximal possible conductivity was determined by measuring the conductivity of the

188

solution in which the cotyledon disks were boiled.

189

Fungicide Bioassay with Colletotrichum spp. Isolates of Colletotrichum acutatum Simmonds,

190

C. fragarie Brooks, and C. gloeosporioides (Penz.) Penz & Sacc. were used by bioautography TLC

191

technique to detect antifungal activity of the compounds isolated from E. plana according to

192

published methods.20 Isolated substances were tested at 10 and 100 µg/spot, in duplicate. Each

193

plate was subsequently sprayed with a spore suspension (106 spores/mL) of the fungus and

194

incubated in a moisture chamber for 4 d at 26 oC with a 12 h photoperiod. Clear zones of fungal

195

growth inhibition on the TLC plate indicated the presence of antifungal constituents21 and their

196

diameter measured.

9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 38

197

Statistical Analysis. Data from dose–response experiments were analyzed using the dose–

198

response curve module22 of R version 2.2.1.23 This software calculates IC50 values. Data from

199

phytotoxicity bioassays were analyzed by ANOVA using the Assistat software version 7.7,24 with

200

comparison of means by Tukey test at a 5% probability.

201 202

RESULTS AND DISCUSSION

203

Phytotoxicity Bioassay-guided Isolation. Initially, the leaves and roots of E. plana were

204

extracted with hexane, dichloromethane/MeOH and water, sequentially, giving three extracts for

205

leaves and three for roots, which were which were evaluated for their phytotoxicity. Evaluation

206

against lettuce and creeping bentgrass at 1 mg/mL indicated that RootsHEX and Leaves

207

dichloromethane/MeOH were

the most phytotoxic (Table 2).

208

Based on these results, the RootsHEX was subjected to normal-phase flash column

209

chromatography, giving 17 fractions (A-Q), which were, in the same way, evaluated for their

210

phytotoxicity on lettuce and creeping bentgrass at 1 mg/mL. Fractions L and M were the most

211

active. Fraction L was submitted again to normal-phase flash column chromatography followed

212

by normal phase HPLC purification providing compound 1 (Figures 1 and 2). Fraction M was

213

purified by normal phase HPLC providing compounds 2 and 3 (Figure 1).

214

Identification of phytotoxic compounds. Compound 1 was identified as neocassa-

215

1,12(13),15-triene-3,14-dione based on its spectroscopic data, both one- and two-dimensional

216

NMR spectroscopy and X-ray diffraction analysis. Its molecular formula was C20H26O2, as

217

revealed by the pseudomolecular ion [M+H]+ acquired in positive ion mode. The 1H NMR

218

spectrum of 1 showed the presence of four methyl singlets (δ 2.08, 1.15, 1.14, and 1.09), one of

219

which was an olefinic methyl, and olefinic protons (δ 6.40, 5.46, and 5.41) of a monosubstituted

10 ACS Paragon Plus Environment

Page 11 of 38

Journal of Agricultural and Food Chemistry

220

double bond, in addition to olefinic protons of an ,-unsaturated double bond ( 7.00 and 5.94).

221

The COSY, 13C NMR and DEPT spectra confirmed the presence of the ,unsaturated double

222

bond as well as the presence of a 1,3-diene (δ 154.5, 132.6, 130.1 and 120.3) containing the

223

monosubstituted double bond. The above together with the presence of three methyl singlets (δ

224

27.1, 22.1, and 16.2) and one olefinic singlet at δ 22.5 suggested a cassane-type skeleton. Upon

225

inspection of the X-ray crystallographic data it was realized that the structure was not a cassane

226

skeleton but rather a new carbon skeleton containing the C-17 olefinic methyl at C-12 rather than

227

at C-14 (Figures 2 and 3). This was further supported by the HMBC correlations between H-5 and

228

C-4, C-6, C-7, and C-10 as well as those between H-1 and C-3 and C-5; and H-20 and C-1, and C-

229

10 (Figure 4). Critical to the establishment of the diene moiety at C-11 were the HMBC

230

correlations between H-17 and C-11, C-12, and C-13 as well as those between H-16 and C-13.

231

COSY and HSQC correlations helped establish all assignment data in Table 1. X-ray

232

crystallographic data established both the relative and absolution configuration of 1.

233

Literature analysis revealed the presence mainly of the isopimarane, labdane, and cassane

234

diterpene skeletal types from the genus Eragrostis.25 They also isolated two new diterpenes from

235

the roots of E. ferruginea (Thunb.): isopimara-9(11),15-dien-19-ol-3-one and cassa-13(14),15-

236

diene-3,12-dione. They have also isolated the known diterpene diol, isopimara-9(11),15-diene-3β,

237

19-diol. Four labdanes with a 8a,15-epoxy ring (8a,15-epoxylabdan-16b-oic acid; 8a,15-epoxy-

238

16-norlabdan-13-one; 8a,15-epoxy-16-norlabdane; and 16-acetoxy-8a,15-epoxylabdane) and the

239

known compound ambreinolide were isolated from the hexane extract of the aerial parts of the E.

240

viscosa (Retz.).26 It appears that our diterpene carbon skeletons for the compound 1 does not fit

241

any of these isopimaranes, labdanes, and cassanes structural types. However, they are very similar

242

to that of the cassanes (Figure 3) aside from the position of the C-17 methyl group. In the cassanes,

11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 38

243

the C-17-methyl group is positioned at the C-14 carbon; however, the C-17 methyl group in

244

compound 1 is attached to the C-12 carbon giving rise to an entirely new diterpene carbon skeleton.

245

Due to the similarity to the cassanes, we have chosen to name this new carbon skeleton

246

neocassanes. Figure 3 shows an example of the typical dienes found with a cassane carbon

247

skeleton, compared with the typical corresponding diene found in the neocassanes which

248

corresponds to those we isolated.

249

Compound 2 from fraction M was identified as 19-norneocassa-1,12(13),15-triene-3,14-dione

250

according of its one- and two-dimensional NMR spectroscopic data. Its molecular formula was

251

C19H24O2 as revealed by the pseudomolecular ion at m/z 285.18744 [M+H]+ acquired in positive

252

ion mode, which suggested one fewer carbons than compound 1. The 1H NMR spectrum of 2

253

looked nearly identical to that observed for 1 apart from the absence of a C-19 methyl singlet and

254

presence of a methyl doublet at δ 1.17 (C-18). The 13C NMR and 90° and 135° DEPT specta of 2

255

looked nearly identical to that observed for 1 except for the presence of only 19 carbons with the

256

clear absence of a C-19 methyl and a singlet carbon signal. Instead of the singlet carbon, we

257

observed a doublet carbon signal corresponding to C-4 at δ 42.1. The above suggested a 19-nor

258

neocassane and led us to assign the structure of 2 as that drawn in Figure 1. COSY and HMBC

259

correlations were in agreement with those observed for compound 1 in the C ring and the diene

260

with the addition of a correlation between H-16 and C-14. HMBC correlations between H-4 and

261

C-3; H-18 and C-3 and C-5 confirmed the absence of the dimethyl group at C-4 as was found in

262

compound 1. NOESY correlations between H-5 and H-9 and those between H-5 and H-18

263

established the H-18 methyl at C-4 with an α-orientation. The remaining relative configurations

264

were all in agreement with that observed for compound 1 as shown by NOESY correlations

265

between H-4 and H-20 and those between H-20 and H-8.

12 ACS Paragon Plus Environment

Page 13 of 38

Journal of Agricultural and Food Chemistry

266

Compound 3 from fraction M was identified as 14-hydroxyneocassa-1,12(17),15-triene-3-one

267

on the basis of its one- and two-dimensional NMR spectroscopic data. The pseudomolecular ion

268

at m/z 301.21826 [M+H]+acquired in positive ion mode, indicated the molecular formula as

269

C20H28O2, suggesting the same number of carbons as compound 1 but with one less site of

270

unsaturation. The

271

ketone (δ 204.9), three olefins (δ 155.6, 145.5, 136.8, 126.2, 116.8, 114.3), and one hydroxylated

272

carbon (δ 75.5). Two of the olefinic carbon signals are triplets (δ 116.8, 114.3) suggesting the

273

presence of two exocyclic methylenes. Based on reasoning described above for compound 1, it

274

was clear that compound 3 was a neocassane-type with both A and B rings similar to 1. HMBC

275

data correlations observed are listed in Figure 4 and clearly agree with the A and B ring correlations

276

observed for compound 1. The C ring of 3 was missing the C-14 ketone found in 1 which was

277

replaced by a hydroxylated carbon. 1H-NMR, HSQC, and HMBC spectral data confirmed this and

278

the presence of two exocyclic methylenes with protons at C-16 (δ 5.18, 5.15) and C-17 (δ 5.04,

279

4.94). COSY correlations clearly established connections between H-16 to H-15 and H-15 to H-

280

13. H-14 HMBC correlations were instrumental in the establishment of the location of the hydroxyl

281

group. H-14 correlated to C-7, C-13, and C-9. The above suggested a neocassane reduced isomer

282

of 1 and led us to assign the structure of 3 as that show in Figure 1.

13C

NMR and 90° and 135° DEPT spectra of 3 revealed the presence of one

283

By normal-phase flash column chromatography were obtained 13 fractions (A-M) from

284

Leavesdichloromethane/MeOH, which were evaluated for their phytotoxicity on lettuce and creeping

285

bentgrass at 1 mg/mL. The fractions I, K and L had good activity, but contained a mixture of

286

different compounds. Fraction J was analyzed by GC-MS, LC-MS and 1H and 13C NMR data, and

287

by comparison with the literature it was identified as 9,12,15 octadecatrienoic acid, 4.17-18,27

13 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 14 of 38

288

Phytotoxic Effect of the Pure Compounds on Lettuce, Creeping Bentgrass and Duckweed.

289

All the compounds isolated from E. plana leaves and roots were tested in a dose-response

290

bioassays on lettuce and creeping bentgrass. This bioassay provides a determination of the

291

phytotoxicity of a compound on both a monocotyledonous (creeping bentgrass) and a

292

dicotyledonous (lettuce) plant species using small amounts of compounds. The concentration of

293

333 µM was inhibitory for 3 and 4 (Table 3). At 1000 µM, all the compounds were inhibitory,

294

especially on monocot creeping bentgrass. Compound 4 (-linolenic acid) is a ubiquitous

295

compound in plants, and therefore was not tested further. After 7 d the fresh weight of the plants

296

was measured. We found a statistical difference only for compound 3, where at 333 µM the lowest

297

fresh weight within creeping bentgrass plants was measured.

298

During dose-response bioassays on duckweed, the pure compounds 1, 2 and 3 inhibited growth.

299

Compound 3 was the most phytotoxic, but all of them significantly reduced the growth at higher

300

concentrations as in Figure 5. Duckweed (Lemna spp.) has several advantages as test organism,

301

among other things its simple structure and small size, allowing small volumes of sample toxicants

302

to be used.28 Effects on growth can be monitored over time without harvesting or making

303

destructive measurements.19, 29 Furthermore, duckweed has a rapid rate of growth, ease of culture

304

and handling, and a good homogeneity with in a clone.30 It is sensitive to most phytotoxins and is

305

uniquely suitable for testing herbicides.19, 29

306

Based on growth inhibition of duckweed, IC50 values were 109 ± 28, 200 ± 37, and 59 ± 15 µM

307

for 1, 2 and 3, respectively. In a repeated study, the IC50 value for 3 was 68 ± 23 µM and that for

308

atrazine, a commercial herbicide was 1.95 ± 46 µM. Statistically, the IC50 value for 3 was the same

309

as in Figure 5, and that for atrazine was close to that reported before (0.93 µM) by Michel et

310

al.19 using the same bioassay. Studying the phytotoxicity of 26 herbicides with as many as 19

14 ACS Paragon Plus Environment

Page 15 of 38

Journal of Agricultural and Food Chemistry

311

different modes of action, Michel et al.19 found the IC50 values ranged from 0.003 μM for

312

sulcotrione and 0.005 μM for chlorsulfuron to 388 μM for glyphosate isopropylamine salt and 407

313

μM for asulam. Such growth differences are due physicochemical properties, uptake, metabolic

314

degradation, and molecular target sites of the herbicides. Therefore, the IC50 values of the isolated

315

compounds in this work are within the range of values found for commercial herbicides.

316

Duckweed can exhibit many symptoms when it is under stress. These include chlorosis,

317

necrosis, colony breakup, root destruction, loss of buoyancy, and gibbosity (humpback or

318

swelling).29 In this work, the image analysis by the LemnaTec software allowed the determination

319

of the percentage of the healthy, chlorotic, and necrotic tissues at duckweed plants. There was a

320

reduction on the percentage of healthy tissues at the 333 µM, and there was an increase on the

321

percentage of necrotic tissues of 12, 47 and 26% for compounds 1, 2, and 3, respectively (Figure

322

6).

323

Compounds 1, 2 and 3 are diterpenes. Many phytotoxic compounds are derived from the

324

mevalonic acid pathway,31 but few of these phytotoxins have a mode of action completely

325

understood. 32 Modes of action of various plant-derived terpenoids have been related to inhibition

326

of ATP formation, alkylation of nucleophiles, disruption of hormonal activity, complexation with

327

protein, binding with free sterols, inhibition of respiration, and increasing relative electron

328

partitioning to the alternative oxidase pathway.33 Despite many studies with terpenoids, most of

329

them are related with diterpenes from fungi while few phytotoxins are known from plants.32

330

Because of their phytotoxicity when compared to commercial herbicides, these compounds could

331

be used as herbicide models34. An example of active natural diterpenes is the quassinoids

332

chaparrinune and glaucaxubulone, which inhibit growth and germination of dicots and monocots

333

at concentrations below 10 μM.31 Moreover, several diterpenoids act as phytoalexins

15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 38

334

(antimicrobial compounds) in monocots35 such as momilactones A and B,36 oryzalexins A–F37,

335

and oryzalexin S38 from rice (Oryza sativa L.). Momilactone B is a well established allelochemical

336

found in the most allelopathic rice varieties, however its mode of action is unknown.39

337

Electrolyte Leakage Assay. None of the compounds isolated in this work caused the kind of

338

changes in the electrolyte leakage (Figure 7) expected with compounds that directly modify the

339

integrity of the plasma membrane, which is a good biomarker to help identify modes of action of

340

herbicides and their dependence on light.40 The fact that the isolated compounds from E. plana did

341

not influence the membrane integrity indicates that the mechanism of action of these compounds

342

is not directly related to cell membrane damage but to other, unknown molecular targets.

343

Bioassay against Colletotrichum spp. A bioautographic technique can be used to identify

344

fungitoxic substances. Visible inhibition zones observed after incubation indicate the presence of

345

fungitoxic compounds.41

346

In this work, a direct antifungal bioautography was performed against the strawberry pathogens

347

C. acutatum, C. fragariae and C. gloeosporioides. These pathogens work can occur singly or in

348

combination in strawberry crops, causing a disease loosely referred to as anthracnose.42 Compound

349

2 was the most active, especially at 100 µg/spot, although at 10 µg/spot it was possible to observe

350

activity (Table 4). Compounds 1 and 3 it produced a diffuse inhibition zone (Table 4). Previous

351

work reported antifungal activity of crude extracts from E. plana leaves on the plant pathogen

352

Drechslera tritici-repentis.43 Another species from the same genus, E. cynosuroides (Retz.), has

353

antifungal activity against Aspergillus niger.44

354

In the present study, three new diterpenes were extracted, isolated and identified for the first

355

time as potential allelochemicals produced by E. plana. Previously, only phenolic compounds were

356

identified as potential allelochemicals in this species.13 Many phenolics and terpenoids have been

16 ACS Paragon Plus Environment

Page 17 of 38

Journal of Agricultural and Food Chemistry

357

reported as allelopathic compounds.45-47 However, simply finding a phytotoxic compound in a

358

plant extract does not prove that it is an allelochemical.48 Bioassay-guided isolation of active

359

compounds is a first step in proof of an allelochemical. A potential complicating factor is the

360

interaction of potential allelochemicals. For example, little attention has been paid to interactions

361

of phenolics and terpenoids from the same plant.49 Compounds identified in this work should be

362

studied in combination. The description of the newly discovered compounds 1, 2 and 3, besides

363

contributing to the chemical characterization of the species, may be the first step in the study of

364

potential of these compounds as bioherbicides. Allelochemicals such as diterpenes or

365

sesquiterpene lactones could be used as models in the development of herbicides of natural origin.

366 367 368 369

ACKNOWLEDGMENTS

370

We are grateful to Amber Reichley, Solomon Green, Robert Johnson, Jesse Linda Robertson and

371

Ramona Pace for their excellent technical assistance.

372 373

FUNDING SOURCES

374

We acknowledge the CAPES – Coordination for the Improvement of Higher Education Personnel

375

(Brazil) (Project PDSE-No 19/2016).

376 377

SUPPORTING INFORMATION

378

The Supporting Information is available free of charge on the ACS Publications website at DOI:

17 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 18 of 38

379

X-ray crystallography information for compound 1 including crystal data, data collection,

380

refinement, fractional atomic coordinates and isotropic or equivalent isotropic displacement

381

parameters (Å2), atomic displacement parameters (Å2), and geometric parameters (Å, º). 1H, 13C,

382

1H-1H

383

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

COSY, 1H-13C HSQC, and 1H-13C HMBC NMR spectra for compounds 1, 2, and 3. This

384 385

CORRESPONDING AUTHOR

386

*Corresponding author. Tel.: 55 54 991702687. Email: [email protected].

387

*Co-corresponding author. Tel.: +1-662-915-5898. Email: [email protected].

388 389

18 ACS Paragon Plus Environment

Page 19 of 38

390

Journal of Agricultural and Food Chemistry

REFERENCES

391 392 393 394 395 396 397

1. Boechat, S. C.; Longhi-Wagner, H. M. Padrões de distribuição geográfica dos táxons brasileiros de Eragrostis (Poaceae, Chloridoideae). Braz. J. Bot., 2000, 23, 177–194. 2. USDA, (United States Department of Agriculture). Plants Database (Version 3.5), http://plants.usda.gov/ (accessed Oct 24, 2017). 3. Coelho, R. W. Substâncias fitotóxicas presentes no capim-annoni-2. Pesqui. Agropecu. Bras., 1986, 21, 255-63.

398

4. Ferreira, N. R.; Medeiros, R. B.; Soares, G. L. Potencial alelopático de capim-annoni

399

(Eragrostis plana Nees) na germinação de sementes de gramíneas estivais. Rev. Bras. Sementes,

400

2008, 30, 43-50.

401

5. Favaretto, A.; Scheffer-Basso, S. M.; Felini, V.; Zoch, A. N.; Carneiro, C. M. Growth of white

402

clover seedlings treated with aqueous extracts of leaf and root of tough lovegrass. Rev. Bras.

403

Zootec., 2011, 40, 68-72.

404 405 406 407 408 409

6. Belz, R. G. Allelopathy in crop/weed interactions–an update. Pest Manage. Sci., 2007, 63, 308– 326. 7. Bhowmik, P. C.; Inderjit. Challenges and opportunities in implementing allelopathy for natural weed management. Crop Prot., 2003, 22, 661-671. 8. Macías, F. A.; Molinillo, J. M. G.; Varela, R. M.; Galindo, J. C. G. Allelopathy- a natural alternative for weed control. Pest Manage. Sci., 2007, 63, 327–348.

410

9. Soltys, D.; Krasuska, U.; Bogatek, R.; Gniadowska, A. Allelochemicals as bioherbicides –

411

Present and perspectives. In Herbicides current research and case studies in use, Price, A. J.,

412

Kelton, J., Eds.; InTech: Warsaw, Poland, 2013; pp. 517-542.

19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 38

413

10. Queiroz, S. C. N.; Cantrell, C. L.; Duke, S. O.; Wedge, D. E.; Nandula, V. K.; Moraes, R. M.;

414

Cerdeira, A. L. Bioassay-directed isolation and identification of phytotoxic and fungitoxic

415

acetylenes from Conyza canadensis. J. Agric. Food Chem., 2012, 60, 5893-5898.

416

11. Inderjit; Cheng, H. H.; Nishimura, H. Plant phenolics and terpenoids: transformation,

417

degradation, and potential for allelopathic interactions. In: Principles and practices in plant

418

ecology: allelochemical interactions; Inderjit, Dakshini, K. M. M., Foy, C. L., Eds.; CRC: Boca

419

Raton, Florida, 1999; pp. 255–266.

420 421

12. Gershenzon, J.; Dudareva, N. The function of terpene natural products in the natural world. Nat. Chem. Biol., 2007, 3, 408–414.

422

13. Favaretto, A.; Chini, S. O.; Scheffer-Basso, S. M.; Sobottka, A. M.; Bertol, C. D.; Perez, N. B.

423

Pattern of allelochemical distribution in leaves and roots of tough lovegrass (Eragrostis plana

424

Nees). Aust. J. Crop Sci., 2015, 9, 1119-1125.

425

14. Fiorenza, M.; Dotto, D. B.; Boligon, A. A.; Boligon, A. A.; Athayde, M. L.; Vestena, S. Análise

426

fitoquímica e atividade alelopática de extratos de Eragrostis plana Nees (capim-annoni).

427

Iheringia, 2016, 71, 193-200.

428

15. Dayan, F. E.; Duke, S. O. Biological activity of allelochemicals. In: Plant-Derived Natural

429

Products - Synthesis, Function and Application; Osbourn, A., Lanzotti, V., Eds.; Springer:

430

Dordrecht, 2009; pp. 361-384.

431 432 433 434

16. Dayan, F. E.; Romagni, J. G.; Duke, S. O. Investigating the mode of action of natural phytotoxins. J. Chem. Ecol., 2000, 26, 2079-2094. 17. Lee, S. O.; Choi, S. Z.; Choi, S. U.; Ryu, S. Y.; Lee, K. R. Phytochemical constituents of the aerial parts from Aster hispidus. Nat. Prod. Sci., 2004, 10, 335-340.

20 ACS Paragon Plus Environment

Page 21 of 38

Journal of Agricultural and Food Chemistry

435

18. Lee, W. B.; Kwon, H. C.; Cho, O. R.; Lee, K. C.; Choi, S. U.; Baek, N. I.; Lee, K. R.

436

Phytochemical constituens of Cirsium setidens Nakai and their cytotoxicity against human

437

cancer cell lines. Arch. Pharm. Res., 2002, 25, 628-635.

438

19. Michel, A.; Johnson, R. D.; Duke, S. O.; Scheffler, B. E. Dose-response relationships between

439

herbicides with different modes of action and growth of Lemna paucicostata: an improved

440

ecotoxicological method. Environ. Toxicol. Chem., 2004, 23, 1074-1079.

441 442

20. Wedge, D. E.; Nagle, D. G. A. New 2D-TLC bioautography method for the discovery of novel antifungal agents to control plant pathogens. J. Nat. Prod., 2000, 63, 1050-1054.

443

21. Tellez, M. R.; Dayan, F. E.; Schrader, K. K.; Wedge, D. E.; Duke, S. O. Composition and some

444

biological activities of the essential oil of Calliarpa americana (L.). J. Agric. Food Chem.,

445

2000, 48, 3008-3012.

446

22. Ritz, C.; Streibig, J. C. Bioassay analysis using R. J. Stat. Software, 2005, 12, 1-22.

447

23. R-Development Core Team. R: A Language and Environment for Statistical Computing; R

448 449 450 451 452

Foundation for Statistical Computing: Viena, Austria, 2009; p. 409. 24. Silva, F. A. S.; Azevedo, C. A. V. Versão do programa computacional Assistat para o sistema operacional Windows. Rev. Bras. Prod. Agr., 2002, 4, 71-78. 25. Nishiya, K.; Kimura, T.; Takeya, K.; Tokawa, H.; Lee, S. R. Diterpenoids from Eragrostis ferruginea. Phytochemistry. 1991, 30, 2410-2411.

453

26. Sebastião, N. N.; Cordeiro, I. J. S.; dos Santos, A. F.; Gaspar, J. F.; Martins, C.; Rueff, J.;

454

Diakanamwa, C.; Sant’Ana, A. E. G.; de Mendonça, D. I. M. 8,15-Epoxylabdane and

455

norlabdane diterpenoids from Eragrostis viscosa. Phytochemistry. 2010, 71, 798-803.

21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 22 of 38

456 27. Yang, Q.; Cao, W.; Zhou, X.; Cao, W.; Xie, Y.; Wang, S. Anti-thrombotic effects of α-linolenic 457

acid isolated from Zanthoxylum bungeanum Maxim seeds. BMC Complementary Altern. Med.,

458

2014, 14, 348-356.

459 28. Kumar, K. S.; Han, T. Physiological response of Lemna species to herbicides and its probable 460

use in toxicity testing. Toxicol. Environ. Health Sci., 2010, 2, 39-49.

461 29. Wang, W. Literature review on duckweed toxicity testing. Environ. Res., 1990, 52, 7-22. 462 30. Lahive, E.; Halloran, J. O.; Jansen, M. A. K. Differential sensitivity of four Lemnaceae species 463

to zinc sulphate. Environ. Exp. Bot., 2011, 71, 25-33.

464 31. Macías, F. A.; Molinillo, J. M. G.; Galindo, J. C. G.; Varela, R. M.; Simonet, A. M.; Castellano, 465

D. The use of allelopathic studies in the search for natural herbicides. J. Crop. Prod., 2001, 4,

466

237-255.

467 32. Duke, S. O.; Oliva, A. Mode of action of phytotoxic terpenoids. In Allelopathy. Chemistry and 468

mode of action of allelochemicals; Macías, F. A., Galindo, J. C. G., Molinillo, J. M. G., Cutler,

469

H. G., Eds.; CRC Press: Boca Raton, Florida, 2004, pp. 201-216.

470 33. Peñuelas, J.; Ribas-Carbo, M.; Giles, L. Allelochemical effects of plant respiration and on 471

oxygen discrimination by alternative oxidase. J. Chem. Ecol., 1995, 22, 801-805.

472 34. Rial, C.; Varela, R. M.; Molinillo, J. M. G.; Bautista, E.; Hernández, A. O.; Macías, F. A. 473

Phytotoxicity evaluation of sesquiterpene lactones and diterpenes from species of the

474

Decachaeta, Salvia and Podachaenium genera. Phytochem. Lett., 2016, 18, 68-76.

475 35. Schmelz, E. A.; Huffaker, A.; Sims, J. W.; Shawn, A.; Lu, X. C., Okada, K.; Peters, R. J. 476

Biosynthesis, elicitation and roles of monocot terpenoid phytoalexins. Plant J., 2014, 79, 659-

477

678.

22 ACS Paragon Plus Environment

Page 23 of 38

Journal of Agricultural and Food Chemistry

478 36. Hwang, B. K.; Sung, N. K. Effect of metalaxyl on capsidiol production in stems of pepper plants 479

infected with Phytophthora capsici. Plant Dis., 1989, 73, 748–751.

480 37. Peters, R. J. Uncovering the complex metabolic network underlying diterpenoid phytoalexin 481

biosynthesis in rice and other cereal crops. Phytochemistry. 2006, 67, 2307–2317.

482 38. Tamongani, S.; Mitani, M. Oryzalexin S structure: a new stemarane-type rice plant phytoalexin 483

and its biosynthesis. Tetrahedron, 1993, 49, 2025–2032.

484 39. Xu, M., Galhano, R., Wiemann, P.; Bueno, E., Tiernan, M., Chung, I. M., Gershenzon, J., 485

Tudzynski, B., Sesma, A.; Peters, R. J. Genetic evidence for natural product-mediated plant-

486

plant alleleopathy in rice (Oryza sativa). New Phytol., 2012, 193, 570-575.

487 40. Dayan, F. E.; Watson, S. B. Plant cell membrane as a marker for light-dependent and light488

independent herbicide mechanisms of action. Pestic. Biochem. Physiol., 2011, 101, 182-190.

489 41. Dewanjee, S.; Gangopadhyay, M.; Bhattacharya, N.; Khanra, R.; Dua, T. K. Bioautography and 490

its scope in the field of natural product chemistry. J. Pharm. Anal., 2015, 5, 75-84.

491 42. Freeman, S.; Rodriguez, R. J. Differentiation of Colletotrichum species responsible for 492

anthracnose of strawberry by arbitrarily primed PCR. Mycol. Res., 1995, 99, 501-504.

493 43. Favaretto, A.; Tonial, F.; Bertol, C. D.; Scheffer-Basso, S. M. Antimicrobial activity of leaf and 494

root extracts of tough lovegrass. Comun. Sci., 2016, 7, 420-427.

495 44. Barath, M.; Aravind, J.; Sivasamy, R. Investigation of antimicrobial activity and chemical 496

constituents of Eragrostis cynosuroides by GC-MS. Res. J. Pharm. Technol., 2016, 9, 267-271.

497 45. Putnam, A. R.; Tang, C. S. The science of allelopathy. John Wiley: New York, 1986. 498 46. Macias, F. A. Allelopathy in the search for natural herbicide models. In Allelopathy: organisms, 499

processes, and applications; Inderjit, Dakshini, K. M. M., Einhellig, F. A., Eds.; American

500

Chemical Society: Washington, D.C., 1995; pp. 310-329.

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 38

501 47. Waller, G. R. Allelochemicals: role in agriculture and forestry. Amer. Chem. Soc. Symp. Ser. 502

#330. Washington, D. C., 1987. 606 pp. 48. Duke, S. O. Proving allelopathy in crop-weed

503

interactions. Weed Sci., 2015, 63, 121-132.

504 49. Inderjit; Muramatsu, M.; Nishimura, H. On the allelopathic potential of certain terpenoids, 505

phenolics, and their mixtures, and their recovery from soil. Can. J. Bot., 1997, 75, 888-891.

24 ACS Paragon Plus Environment

Page 25 of 38

506

Journal of Agricultural and Food Chemistry

FIGURE CAPTIONS

507 508

Figure 1. Structures of compounds isolated from E. plana roots: neocassa-1,12(13),15-triene-

509

3,14-dione, 1, 19-norneocassa-1,12(13),15-triene-3,14-dione, 2, and 14-hydroxyneocassa-

510

1,12(17),15-triene-3-one, 3.

511 512

Figure 2. Molecular structure of neocassa-1,12(13),15-triene-3,14-dione, 1, from X-ray analysis,

513

with 50% ellipsoids.

514 515

Figure 3. Carbon skeleton examples for cassa-13(14),15-diene and neocassa-12(13),15-diene

516

analogs.

517 518

Figure 4. HMBC correlation data for neocassa-1,12(13),15-triene-3,14-dione, 1, 19-norneocassa-

519

1,12(13),15-triene-3,14-dione, 2, and 14-hydroxyneocassa-1,12(17),15-triene-3-one, 3.

520 521

Figure 5. Growth-response curve of the effects of neocassa-1,12(13),15-triene-3,14-dione, 1, 19-

522

norneocassa-1,12(13),15-triene-3,14-dione, 2, 14-hydroxyneocassa-1,12(17),15-triene-3-one, 3,

523

on L. pauciscostata growth 7 d after treatment. The solvent control value is 0 µM.

524 525

Figure 6. Proportion of healthy, chlorotic and necrotic tissues in L. pauciscostata plants, when

526

submitted to different concentrations of the compounds 1: neocassa-1,12(13),15-triene-3,14-

527

dione; 2: 19-norneocassa-1,12(13),15-triene-3,14-dione; 3: 14-hydroxyneocassa-1,12(17),15-

528

triene-3-one.

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 38

529

Figure 7. Electrolyte leakage induced by the compounds neocassa-1,12(13),15-triene-3,14-

530

dione,

531

1,12(17),15-triene-3-one, 3, submitted to 18 h dark + 6 h light (arrows indicate the start of the

532

light exposure). Data represent means of three replications ± SD.

1,

19-norneocassa-1,12(13),15-triene-3,14-dione,

2,

and

14-hydroxyneocassa-

533 534

26 ACS Paragon Plus Environment

Page 27 of 38

Journal of Agricultural and Food Chemistry

Table 1. 13C and 1H NMR Assignmenta Data for Compounds 1, 2, and 3 in CDCl3. position

δC

1 2 3 4 5 6

154.2 126.9 204.7 44.6 51.6 21.0

7

26.9

8

45.2b

9 10 11

46.2b 38.9 32.8

12 13 14 15

154.5 132.6 199.5 130.1

16

120.3

17 18 19 20

1 δH (multiplicity, J in Hz)

2c δH (multiplicity, J in Hz)

δC

7.00 (d, 10.2) 5.94 (d, 10.2) 1.66 (dd, 12.0, 2.0) 1.53 (ddd, 12.0, 12.8, 3.6) 2.41 (m) 1.31 (m) 2.29 (ddd, 13.4, 11.5, 4.8) 1.77 (m) 2.52 (dd, 18.1, 4.6) 2.43 (m) 6.40 (dd, 17.9, 11.6)

153.6 127.5 201.3 42.1 49.2 23.4

120.2

22.5

5.46 (dd, 17.9, 2.1) 5.41 (dd, 11.6, 2.1) 2.08 (s)

22.1 27.1 16.2

1.09 (s) 1.15 (s) 1.14 (s)

δC

7.03 (d, 10.2) 5.95 (d, 10.2) 2.35 (m) 1.58 (dd, 12.2, 2.8) 1.88 (m) 1.33 (m) 2.38 (m) 1.31 (m) 2.33 (m)b

155.6 126.2 204.9 44.5 51.7 21.4

44.7 38.9 29.5

22.3

1.81 (m)b 2.60 (dd, 18.1, 4.4) 2.46 (m) 6.43 (ddd, 17.9, 11.6, 11.6) 5.50 (dd, 17.9, 2.2) 5.44 (dd, 11.6, 2.2) 2.11 (s)

12.1 13.6 -

1.17 (d, 6.7) 1.11 (s) -

22.0 27.1 15.7

26.0 44.8 44.6 38.9 32.9 154.3 132.5 199.4 129.9

29.9 36.1

145.5 55.1 75.5 136.8 116.8 114.3

a

Assignments of NMR data are based on 1H, 13C, DEPT, 1H−1H COSY, HSQC, and HMBC NMR experiments.

b

Signals in same columns interchangeable.

3c δH (multiplicity, J in Hz) 7.07 (d, 10.3) 5.93 (d, 10.3) 1.62 (m) 1.56 (m) 1.72 (m) 1.65 (m) 1.30 (m) 1.76 (m) 1.49 (m) 2.35 (dd, 12.9, 3.8) 2.14 (t, 12.9) 3.13 (m) 3.73 (dd, 2.4) 5.87 (ddd, 16.5, 10.5, 5.4) 5.18 (dd, 10.5, 1.9) 5.15 (dd, 16.5, 1.9) 5.04 (t, 2.0) 4.94 (t, 2.0) 1.10 (s) 1.16 (s) 1.08 (s)

27 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 38

Table 2. Phytotoxic Activity of Extracts Prepared with Leaves and Roots of Eragrostis plana Using Different Solvents phytotoxicity at 7 da

extracts (1mg/mL)

lettuce

creeping bentgrass

rootsHEX

2

4

rootsdichloromethane/MeOH

0

2

rootsH2O

0

0

leavesHEX

0

1

leavesdichloromethane/MeOH

1

3

leavesH2O

0

0

Bioassay rating based on scale of 0 to 5: 0 = no effect and 5 = no growth or germination. a

28 ACS Paragon Plus Environment

Page 29 of 38

Journal of Agricultural and Food Chemistry

Table 3. Phytotoxic Activity of Novel Pure Compounds Isolated From Roots of Eragrostis plana compounds

1

2

3

tested concentration (µM) 0* 0** 10 33 100 333 1000 0* 0** 10 33 100 333 1000 0* 0** 10 33 100 333 1000

phytotoxicity at 7 da lettuce creeping bentgrass 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 1 4 0 0 0 0 0 0 0 0 0 0 2 3 1 4

a Bioassay * Water

rating based on scale of 0 to 5: 0 = no effect and 5 = no growth or germination. control. ** Solvent control.

Table 4. Antifungal Activity of Pure Compounds Isolated from Leaves and Roots of Eragrostis plana 29 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

compounds

1 2 3 Benomyl Captan Cyprodinil Azoxystrobin

tested concentration (µg) 10 100 10 100 10 100

Page 30 of 38

inhibitory zone diameter (mm) Colletotrichum Colletotrichum Colletotrichum acutatum fragariae gloeosporioides *5.0±0.0 0.0±0.0 4.0±0.0 *7.5±0.8 *11.5±0.7 *10.0±0.0 *4.5±0.6 3.5±0.7 5.5±0.7 9.0±0.0 12.7±1.0 7.6±0.5 0.0±0.0 0.0±0.0 0.0±0.0 *8.5±0.7 0.0±0.0 0.0±0.0 0±0 0±0 0±0 15 14 17 *20 *21 0 *33 *26 *23

* diffuse.

30 ACS Paragon Plus Environment

Page 31 of 38

Journal of Agricultural and Food Chemistry

FIGURES

17 15 20 9 1 3

5

19

18

O

H

H

13

H

16

O

H

7

O

H

H

1

O

H 2

H H H O

H

OH

H 3

Figure 1.

31 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 32 of 38

Figure 2.

32 ACS Paragon Plus Environment

Page 33 of 38

Journal of Agricultural and Food Chemistry 15 13

20 9

H

1 3

5

19

18

H

12 16

15

H

17

H

7

H

H

cassa-13(14),15-diene

neocassa-12(13),15-diene

Figure 3.

33 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 34 of 38

H O

H O

1

H

H O

H O

H H CH3

2

H OH

H O

H

3

Figure 4.

34 ACS Paragon Plus Environment

Page 35 of 38

Lemna growth (%)

600

Journal of Agricultural and Food Chemistry

1

500 400 300 200 100 0 0 0.1 0.3 1

3

10 33 100 333

Concentration (µM)

Lemna growth (%)

2 400 300 200 100 0 0 0.1 0.3 1

3

10 33 100 333

Concentration (µM)

3

Lemna growth (%)

500 400 300 200 100 0 0 0.1 0.3 1

3

10 33 100 333

Concentration (µM) Figure 5.

35 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

healthy

chlorotic

Page 36 of 38

1

necrotic

100 tissue (%)

80 60 40 20 0 C

CS 0.1 0.3

1

3

10

33 100 333

Concentration (µM) 2

tissue (%)

100 80 60 Figure 6.

40 20

3 33

0 10

33

10

3

1

0. 3

0. 1

S C

C

0 3

Concentration (µM)

80 60 40

33

10

Figure 6. 3

1

0. 3

0. 1

C

S

0

10 0 33 3

20 C

tissue (%)

100

Concentration (µM)

36 ACS Paragon Plus Environment

Page 37 of 38

Journal of Agricultural and Food Chemistry

Figure 7. 37 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 38 of 38

Table of Contents Graphic

Eragrostis plana

New neocassanes

38 ACS Paragon Plus Environment