mechanisms of herbicide resistance in wild oats (avena spp.)

206
_*"Íot[il*uu **: {P" -."_ Mechanisms of He in Witd Oats Chanya Maneechote Thesis Submitted for the Degree of Doctor of Philosophy in Faculty of Agricultural and Natural Resource Sciences The University of Adelaide Department of Crop Protection Waite Agricultural Research Institute July 1995

Upload: khangminh22

Post on 02-May-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

_*"Íot[il*uu **:{P" -."_

Mechanisms of He

in Witd Oats

Chanya Maneechote

Thesis Submitted for the Degree of

Doctor of Philosophy

in Faculty of Agricultural and Natural Resource Sciences

The University of Adelaide

Department of Crop Protection

Waite Agricultural Research Institute

July 1995

u

(;'ì,'i i "Table of Contents

Title page

Table of Contents

AbsEact

Declaration

Acknowledgements

Abbreviations

Chapter I Introduction

I .1 History of Herbicide Resistance

1.2 Herbicide Resistance in Ausualia

1.3 Mechanisms of Herbicide Resistance

1.3.1 Reduced Herbicide Uptake and Translocation

I.3.2 Enhanced Herbicide Metabolism

1.3.3 Modified Site of Action

I.3.4 Sequesfration of Herbicide within Plant Cells

1.4 Inheritance of Herbicide Resistance

1.5 Aryloxyphenoxypropionate and Cyclohexanedione Herbicides

1.5.1 Antagonism of APP and CHD Herbicides

L.5.2 Uptake and Translocation

1.5.3 Metabolism

I .6 Mode of Action of APP and CHD Herbicides

I .6.1 Inhibition of Fatty Acid Biosynthesis

1.6.2 Characteristics of ACCase Enzyme

1.6.3 Kinetics of Inhibition by APP and CHD Herbicides

I.6.4 Inheritance of Herbicide Resistant ACCase in Formerly

Susceptible Weed Species

I

ü

vtl

x

xi

1

I

2

4

4

5

6

8

8

10

L2

12

t4

20

2t

23

25

25

ru

1.7 Disruption of Membrane Function by APP and CHD Herbicides

t.7 .l Diclofop as a ProtonoPhore

1.7 .2 Recovery from Diclofop Acid Induced Depolarization of the

PlasmaMembrane Electrogenic Potential in Some Resistant

V/eeds

1.8 Resistance to Acetyl-CoA Carboxylase-krhibiting Herbicides

1.9 V/ild Oat(Avena spp.)

1.9.1 Seed Dormancy and Germination

I.9.2 Seed Bank of lVild Oat

1.9.3 ControlofWildOat

1.10 Herbicide Resistance in Wild Oat

l.l 1 Objectives of this Project

Chapter 2 Materials and Methods

2.1 PlantMaærial

2.2 PlantCulture

2.2.I Germination

2.2.2 Pot culnue

2.2.3 Hydroponic Culnue

2.3 Dose Response Experiments

2.4 ACCase Extraction and Assay

2.5 Uptake and Translocation of ¡l4qDiclofop-Methyl

2.6 Metabolism of Herbicides in Plants

2'6.1 ¡l4clDiclofoP-MethYl

2.6.2 ¡4-t+ClSethoxydim

2.7 Measr¡rement of Plasma Membrane Potentials

26

27

27

28

30

3l

32

33

33

34

35

35

37

37

37

37

39

40

43

42

42

43

44

lv

Chapter 3 Spectrum of Resistance to Arylox¡phenoxy¡lropionate and

Cyclohexanedione Herbicides in lVild Oat Biot¡pes:

\ilhole Plant and Target Site Inhibition Studies

3.1 Introduction

3.2 Whole Plant Assessment of Resistance

3.3 Inhibition of ACCase Activity by Herbicides

3.4 Discussion

Chapter 4 Herbicide Resistance Endowed by an APP Insensitive

ACCase in BiotYPe SAS 1

4.t Introduction

4.2 lau'flaærials andMethods

4.2.1 Variations in ACCase During Development

4.3 Results

4.3.1 ACCase Activity During Growth Development

4.3.2 Uptake and Translocation of [14C]D'clofop-Methyl

4.3.3 Meøbolism of l14c]D'clofop-Methyl

4.3 .4 Effect of Diclofop Acid on Plasma Membrane Potential

4.3.5 Inhibition of ACCase Activity InVivo

4.4 Discussion

Chapter 5 Herbicide Resistance Endowed by an APP- and CHD-

Insensitive ACCase in Biotype NAF 6

5.1 Intoduction

5.2 Results

5.2.1 Uptake, Translocation and Metabolism of ¡llgtD'tlofop-Methyl

5.2.2 Uptake and Metabolism of ¡4-l4clsethoxydim

46

46

47

58

69

74

74

74

74

75

75

77

80

83

83

86

88

88

89

89

93

v

5.2.3 Effect of Diclofop Acid on Plasma Membrane Potential

5.3 Discussion

Chapter 6 Resistance to APP Herbicides Endowed by Two

Mechanisms in BiotYPe NAS 4

6.1 Introduction

6.2 Maænals andmethods

6.2.1 Effects of the Cytochrome P450Inhibitor Tetcyclasis on Diclofop

Metabolism

6.2.2 Effects of the Cytochrome P450 Inhibitor Tetcyclasis on Plant

Growth

6.2.3 Inhibition of ACCase by Diclofop in Single Lines of NAS 4

6.3 Results

6.3.1 Uptake and Translocation of ¡l4C¡Diclofop-Methyl

6.3.2 Metabolism of l14c]D'ctofop-Methyl

6.3.3 Effect of Tetcyclasis on Diclofop Metabolism

6.3.4 Effects of Tetcyclasis on Plant Growth

6.3.5 Inhibition of ACCase Activity by Diclofop Acid in Single Lines

of Biotype NAS 4

6.3.6 Effect of Diclofop Acid on Plasma Membrane Potential

6.4 Discussion

Chapter 7 Non-Target Site Mechanism Conferring Resistance to

ACCase-Inhibiting Herbicides in Biotype NAS 14

7.1 Innoduction

7 .2 Matenals and Methods

7 .2.1 Uptake, Translocation and Metabolism of ¡l+grD'.lofop-Methyl

7 .2.2 Absorption of ll4c]D'clofop-Methyl to Leaf Cuticle

97

97

100

100

t0l

l0l

101

t02

t02

t02

105

108

It2tt7

119

il9

t23

r23

t24

t24

t24

7.3 Results

7.3.t

7.3.2

7.3.3

7.3.4

7.3.5

Uptake and Metabolism of I l4C]Diclofop-Methyl

Translocation of [14c]Diclofop-Methyl

Absorption of Herbicide to I-eaf Cuticle

ACCase Activity inVivo

Effect of Diclofop on Plasma Membrane Potential

vl

t24

124

t27

t27

t29

t32

132

135

135

136

136

136

t37

137

138

142

t45

152

185

7.4 Discussion

Chapter 8 Investigations of Alterations of Membrane Responses to

Diclofop as Mechanism of Resistance in Wild Oat

8.1 InEoduction

8.2 Materials and Methods

8.2.I Effect of pH on Membrane Response to Diclofop Acid

8.2.2 Effect of Diclofop Acid on Acidification of External Solution

8.3 Results

8 .3 . 1 Effect of Diclofop Acid on Plasma Membrane Potential

8.3.2 Effects of pH on PlasmaMembrane Repolarization

8 .3 . 3 Acidification of the External Solution by Roots of Intact Plants

8.4 Discussion

Chapter 9 General Conclusions

Literature Cited

Appendix

159

vll

Abstract

Herbicide resistance, specifically to the acetyl coenzyme A carboxylase (ACCase)-inhibiting

herbicides aryloxyphenoxypropionate (APP) and cyclohexanedione (CIID) herbicides, in

wild oat has become a serious problem in cropping regions of Australia. This study was

conducted to elucidate the mechanisms endowing resistance to herbicides in wild oat. Ten

biotypes of wild oat (Avena fatua and A. sterilis ssp. ludovicíana) were examined for

resistance to ACCase-inhibiting herbicides. At the whole plant level, all biotypes were

resistant to at least one APP (diclofop-methyl, fluazifop-butyl, and haloxyfop-ethoxyethyl),

or CHD (sethoxydim and tralkoxydim) herbicide. The response of the target site, ACCase,

to the APP and CHD herbicides was examined in vitro. All resistant biotypes excePt one

contained a target site resistant to at least one APP and CHD herbicide. The response of the

target site of different biotypes to APP and CHD herbicides varied widely suggesting that

several different mutations within ACCase can endow resistance to these herbicides. Four

resistant biotypes (SAS 1, NAF 6, NAS 4 and NAS 14) of wild oat with contrasting

patterns of resistance at the whole plant and ACCase levels were studied further to fully

elucidate mechanisms endowing resistance.

Biotype SAS I is highly-resistant to APP herbicides but is not resistant to CHD herbicides.

The ACCase from this biotype displays high-level resistance to APP herbicides and low-

level resistance to CHD herbicides. Absorption, translocation and metabolism of

¡l4Cldiclofop-methyl were simila¡ in both resistant biotype SAS I and susceptible biotype

SAS 2. Hence, the target-site based mechanism is the only mechanism responsible for

resistance in biotype SAS 1.

Biotype NAF 6 is moderately resistant to both APP and CHD herbicides at the whole plant

level and contains a target enzyme ACCase with moderate resistance to both groups of

herbicides. There were no differences in uptake, translocation and metabolism of

llaC]diclofop-methyl between susceptible SAS 2 and resistant NAF 6 wild oat biotypes. In

vlu

addition, absorption and meøbolism of ¡+laClsethoxydim in both biotypes were simila¡. A

modified target site is solely responsible for resistance to both APP and CHD herbicides in

biotype NAF 6.

Biotype NAS 4 is highly resistant to diclofop-methyl in the field yet contains an ACCase

which displays only moderate resistance to this herbicide. No differences in uptake and

translocation of diclofop-methyl were observed in this biotype compared to the susceptible.

However, this biotype showed a 1.S-fold enhanced rate of metabolism of diclofop-methyl.

In the prcsence of the cytochrome P450 inhibitor tetcyclasis, the rate of diclofop metabolism

in this biotype was decreased to the same level as that of the susceptible biotype SAF 19

suggesting the involvement of cytochrome P450 monooxygenases in resistance. In

addition, tetcyclasis increased the phytotoxicity of diclofop-methyl against the resistant

biotype. Studies with ten individual families of biotype NAS 4 indicate that two

mechanisms of resistance, an altered target site and enhanced metabolism, are present in each

individual of the NAS 4 population.

Biotype NAS 14 shows low-level resistance to APP and CHD herbicides but contains á

sensitive ACCase. Reduced uptake or enhanced metabolism of diclofop-metþl do not

account for resistance in this biotype. A small difference between susceptible and resistant

biotypes was observed in translocation of diclofop to the meristematic region following

application of ll4c]d'clofop-methyl to the leaf axil of twoleaf stage plants. This reduced

translocation of diclofop was correlated with a reduction in the inhibition of ACCase in vivo

following application of diclofop-methyl to intact plants. This non-target site-based

mechanism may be responsible for the low level of resistance to diclofop-methyl in this

biotype.

The relationship between resistance to graminicides and the ability to repolarize the plasma

membrane potential following diclofop acid-induced depolarization was studied in all

biotypes of wild oat. No biotype demonstrated recovery of the plasma membrane potential

at pH 5.8 following the removal of diclofop acid. pH was found to be an important

determinant of the ability of plasma membranes to repolarize following removal of diclofop

D(

acid. An initial dictofop acid-induced depolarization of the plasma membrane potential was

reversible in herbicide-free solution buffered at pH 6.5, but not when buffered at pH 5.8. It

is clear from this study that resistance to graminicides is not correlated with differences in the

ability of wild oat biotypes to recover from diclofop acid-induced depolarization of the

plasma membrane potential.

In conclusion, at least three mechanisms of resistance to ACCase-inhibiting herbicides, a

modified target-site, enhanced metabolism and reduced translocation are found in Avena

populations. The target-site based mechanism, which is responsible for the moderate and

high resistance to herbicides at the whole plant level, is the most cornmon mechanism. The

non-target site based mechanisms, enhanced herbicide metabolism and reduced translocation

of herbicide to the target site, were observed in one resistant biotype each. The variety of

resistance mechanisms which have appea¡ed in wild oat biotypes may lead to cross-

resistance to other herbicide classes and make control of resistant wild oat with alternative

herbicides difficult.

x

Declaration

This work contains no material which has been accepted for the awa¡d of any other degree in

any university or other tertiary institution and to the best of my knowledge and belief

contains no material previously published or written by another person, except where due

reference has been made in the text.

I give consent to this copy of my thesis, when deposited in the University Library, being

available for loan and photocopying.

Chanya Maneechote

July 1995

xr

Acknowledgements

I wish to thank my supervisors, Dr. S.B. Powles and Dr. C. Preston for their valuable

supervision, advice and constructive criticism during the course of this study. I express my

sincere thanks to Dr. J.A.M. Holtum for his expert advice and also special thanks to Dr.

F.J. Tardif and Dr. J. Karotam for critical reading of this manuscript.

Sincere thanks is extended to Dr. S.D. Tyerman from School of Biological Sciences,

Flinders University, Bedford Parh South Australia for providing facilities for the membrane

potential work.

I would like to thank Dr. A.M. Mansooji for providing the dose response data of wild oat

biotypes SAS I and SAS 2.

I would also like to thank the members of herbicide resistance group at the V/aite

Agricultural Research Institute: Dr. J.C. Christopher, Dr. E. Purba, J.M. Matthews, P.

Boutsalis, S. Hole, B. Nietschke, H. Alizadeh, N. Farida and I. Hidayat for their help,

encouragement and companionship.

Special thanks a¡e extended to my dear friends, S. Jamjod, C. Koshawatana, S. Agung and

K. Beaton for their help and freindship.

I thank the Australian Agency for International Development (AusAID) for the financial

support and also thank Thai Goverment for study leave permission.

Lastly, my deepest gratitude to my parents and family for their support and understanding

throughout the study.

xlt

Abbreviations

ai

APP

ACCase acetyl co€nzyme A carboxylase (EC 6.4.1.2)

active ingredient

Aryloxyphenoxypropionate

CHD Cyclohexanedione

Clethodim (E,E)-(!)-2-[ I -[[(3-chloro-2-propenyl) oxy] iminol propyll-S-

[(ethylthio) propyl] -3-hydroxy-2-cyclohexene- I -one

Cycloxydim (t)-2-[ 1 -(ethoxyimino) butyl] -3-hydroxy-5-thian-3-ylcyclohex-

2-enone

Diclofopmethyl methyl(t)-2-Í4-(2,4-dtchlorophenoxy) phenoxyl propanoate

EDTA

dithiothreitol

ethylenediamineteEaacetic acid

ethyl- I -(2,4-dicilorphenyl)-5-trichloromethyl- lH-t,2,4-

tiazole-3-carboxylic acid

ethyl(t)2-[4-[(6-chloro -2-betuoxazolyl) oxy] phenoxyl

propanoate

buty(t)-2-[a-[[5-(trifluoromethyl)-2-pyridinyl] phenoxyl

propanoate

fresh weight

fenchloraz ole-ethyl

Fenoxaprop-ethyl

Fluazifop-butyl

DTT

FW

ruu

Haloxyfoperhoxyethyl ethoxyerhyt(t)-2-[4-[[3-chloro-5-(trifluoromethyl)-2-pyndinyu

oxyl phenoxYlProPanoate

HPLC High Pressure Liquid Chromatography

herbicide concenEation required to reduce invitro ACCase

activity 507o withrespect to control assay without herbicide

Iso

LDso

LSS

MES

PMSF

PPO

POPOP

PVP

Quizalofopethyl

Sethoxydim

Tralkoxydim

Tricine

dose of herbicide required to kill 507o of plants

Liquid Scintillation SpecEometrY

2-(N-morpholino)ethanesulfonic acid

phenylmethylsulfonyl fluoride

2,5-diphenyloxazole

1,4-bis (5 -phenyloxazoyl)benzene

polyvinylpyrrolidone

ethyl(+)-2- [4- [(6-chloro-2-quinoxdinyl) oxy]phenoxyl

propanoate

2- [ I -ethoxyimino)butyl] -5 - l2-(ethylthio)propyl] -3-hydroxy-2-

cyclohexen-1-one

2- [ 1 (ethoxyimino)propyl] -3-hydroxy -5-(2,4,6'

trimethylphenyl)cyclohex-2-enone

N-[2-hydroxy- 1 - 1 -bis (hydroxymethyl) ethyl] glycine

Tris (hydroxymethyl) aminomethaneTris

xivErrata

Page 6,

Page 13,

Page 2I,

Page 24,

Page 25,

Page 26,

Page 27,

Page 29,

Page 34,

Page 40,

Page 44,

Page 72,

Page 77 ,

Page 80,

line 4. Change to: Reduced herbicide sensitivity ...

line 14. Change to: The predominant translocated form is diclofop acid ...

line 7-8. Change to: Initially, the ATP-dependent carboxylation of biotin is

catalysed in the biotin carboxylase domain.

line 26. Change "containg" to containing.

lirc 21. Change "diclots" to dicots.

lines 5-6. Remove last sentence from paragraph.

lines 19-20. Remove reference to Stoltenberg et al. 1989.

line 10. Change to: A second mode of action of aryloxyphenoxypropionic acids

and cyclohexanediones is hypothesised to ...

line 11. Change to: ...proton-translocating plasmalemma ATPase.

Table 1.9. Remove line containing Festuca rubra.

line 2I. Change to: Ten putative resistant biotypes with varying levels of resistance

were examined.

line 11. Change to: The supernatant was slowly...

line 22. Add to the end of the sentence , giving a final concentration of | 7o

acetone in the assay.

line 4. Add to the end of paragraph. Recovery of radiolabel was in the

range of 90 to L00 7o of. that applied in all cases.

Iine 24. Change to: To test the assumption...

line 12. Change "as" to \ilas.

line2l. Change to: Metabolism occurred in both the meristematic regions (Figure

4.4) and in the leaves (data not shown).

figure legend. Change to: ...extracted from the meristematic region, 0 to 180 h

after treatment.

Page 81,

Page 82, figure legend. Change to: HPLC elution profiles of extract, frorr, 1I"

meristematic region of susceptible ...

Page 92, figure legend. Change to: Percentage of radioactivity extracted from the

meristematic region as diclofop-methyl ...

Page 109, table title. Change to: The percentage of radioactivity extracted from the

meristematic region as ¡14çrd'"lofop acid ...

Page 1 lT,line 22. Change to: ...possessing an ACCase that is 3- to 8-fold less resistant.

Page l22,line 7. Change to: In L. rigidum, biotype SLR 31 has at least two mechanisms of

resistance, enhanced metabolism to detoxify diclofop-metþl and resistant ACCase,

as well as an alteration of the plasma membrane that has been correlated with

resistance (Haüsler et al. 1991, Holtum et a1.,1991, Tardif and Powles 1994).

Page t24,line 5. Change "absorption" to adsorption.

Page 127,line 21. Change "absorption" to adsorption.

Page l29,line 17. Change "diclofop-methyl" to diclofop.

Page 130, Figure legend. Change "[14C]diclofop" to radiolabel.

Page 132,line22. Change "diclofop-methyl" to diclofop.

line 25. Change "diclofop-methyl" to diclofop.

Page 135, line 7 . Change "resistance" to action.

Page [49,line23. Change "have" to has.

Page 153, line 16. Change to: ...where three mutant alleles of the ACCase structural

gene...

Page 156, line 25. Change to: The recovery of membrane polarity is dependent on the pH...

Page 157 ,line l. Change "documented as a mechanism of resistance" to correlated with

resistance.

Iine 2. Change "appeared" to apparent.

line 27. Change "has" to have.

Page 158,line 3. Change "could" to can.

Page 159, line 2. Change "Physiological" to physiological.

Page 160,line 15. Change "herbicide" to Herbicide.

Page l7O,line 7. Change "ishikawa" to Ishikawa.

..""' - ,,-*"*.*"'*l

' Lr+.qliË {},¿,i,,{Pli$ L!ffi{p,nY :

';'iiã tl|,åTF:[1¡ii]]l tf: li,'[L/,i¡i ,Chapter 1

Introduction

1.1 History of Herbicide Resistance

The exposure of a weed population to persistent herbicide application can lead to the

occrurence of herbicide resistance in that population. The first report of herbicide resistance

was in a biotype of. Senecio vulgaris L. from Washington State, USA, in 1968 which was

no longer connolled by the photosystemtr-inhibiting herbicide simazine (Ryan 1970). This

biotype appeared in a conifer plantation where simazine had been sprayed repeatedly for

several years. By t914,triazine-resistant weeds were documented as a widespread problem

in maize fields of north-west'Washington and throughout the Pacific North West, from

California to British Columbia. Later, triazine resistant weeds became cortmon in eastern

Europe, Israel and Australia (Ritter 1989). Now more than 57 weed species (40 dicots and

17 monocots) have been identified as triazine resistant (Holt et at. 1993).

Resistance to some other early and widely-used herbicide groups has developed more

slowly. For example, resistance to the bipyridiliums (paraquat and diquat), released in

1960, has been identified in 18 weed species (Preston L994). Purba (1993) noted that

resistance to paraquat and diquat has evolved slowly compared to the other groups of

herbicides. Resistance to dinitroanilines has been identified in 4 weed species following

annual applications of herbicides for at least l0 years (Smeda and Vaughn 1994).

Resistance to auxin analog herbicides, released in 1940s, has been documented in 8 species

following repeated herbicide application (Coupland 1994).

Recently, resistance to herbicides inhibiting acetyl coenzyme A carboxylase (ACCase) and

acetolactate synthase (ALS) has become widespread. Resistance to ACCase-inhibiting

herbicides (aryloxyphenoxypropionate and cyclohexanedione) has been reported in nine

grass weeds (Devine and Shimabukuro L994). Resistance to AlS-inhibiting herbicides

2

(sulfonylureas, imidazolinones, triazolopyrimidines and pyrimidinyl thiobenzoates) has

evolved in 13 weed species in North America and Australia (Saari et al. 1994). These

herbicides have only been in use for less than two decades yet resistance to these herbicides

has been evolved more rapidly than to other herbicide groups. Thus, increased use of these

herbicides will probably increase the number of resistant weeds developing in the future.

1.2 Herbicide Resistance in Australia

The fust report of herbicide resistance in Australia was a diclofop-methyl resistant biotype of

Lolium rigidum in 1982 following the use of this herbicide in four consecutive years (Heap

and Knight 1982). Subsequently, hundreds of populations of L. rigidum in Australia have

developed resistance to diclofop-methyl and some are cross-resistant to different classes of

herbicides which have a different mode of action (Heap and Knight 1990). To date, eleven

weed species (seven monocots and four dicots) have developed resistance to ten classes of

herbicides with six different modes of action (Table 1.1).

3

Table 1.1 Herbicide-resistant weed species in Australia

Species Class of herbicide (site of action) Reference

Arcøtlvca calenùtla

Avenafatrn

Avena sterilß

Brassica touneþrtü

Cyperus diformis

Hordeum gbucum

Hordeum leporinum

Lolium rigidum

Sisymbrium orientale

Sonchus oleraceus

Bipyridilium (photosystem I)

Aryloxyphenoxypropionates (ACCase)

Aryloxyphenoxypropionates (ACCase)

Sulfonylurea (ALS)

Sulfonylurea (ALS)

Bipyridilium (photosystem I)

Bipyridilium (photosystem I)

Aryloxyphenoxypropionate (ACCase)

Cyclohexanedione (ACCase)

Dinitroaniline (microtubule forrration)

Phenylurea (photosystem II)

Sulfonylurea (ALS)

Triazine (photosystem tr)

Triazinone (photosystem tr)

Triazole (pigment biosynthesis)

Imidazoünone (ALS)

Sulfonylurea (ALS)

Imidazolinone (ALS)

Sulfonylurea t.qlSl

Triazolopyrimidine (ALS)

Bipyridilium (photosystem I)

Powles et al. 1989

Mansooji e¡ al. 1992

Mansooji et al. 1992

Boutsalis and Powles 1995a

Gratram et al. unpublised

Powles 1986

Tucker and Powles 1991

Heap and Knight 1982

Heap and lftright 1990

McAlister et al. 1995

Burnetetal. l99l

Christopher etal. t992

Burnetetal. l99l

Burnet et al. l99l

Burnet et al. l99l

Boutsalis and Powles 1995a

Boutsalis and Powles 1995a

Vulpiabromoides Purba et al. 1993b

4

1.3 Mechanisms of Herbicide Resistance

To be effective, a herbicide must enter the plant, reach the target site at a sufficient

concentration and be maintained there for a sufficient period to kill the plant. Changes in

morphological, physiological and/or biochemicat processes in plants that exclude herbicide

from the tatget site can lead to herbicide resistance. Possible mechanisms of resistance to

herbicides are discussed below.

1.3.1 Reduced Herbicide Uptake and Translocation

Herbicide can be prevented from reaching the target tissue by reduced uptake and

translocation. To date, there have been no reports of resistance due to reduced absorption of

herbicide into resistant weed biotypes. Examples of reduced translocation of herbicide

conferring resistance are given in Table 1.2.

Table 1.2 Reduced franslocation as a mechanism of resistance in weeds

Species Herbicide Reference

Arctothcca calendula diquat

Conyza bonariensis paraquat

Erigeron canadensis pa¡aquat

Erigeron philadelphicus pa¡aquat

Hordeum glaucum paraquat

Preston etal.L994

Fuerstet al. 1985

Tanaka et al. 1986

Tanaka et al. 1986

Bishop et al. 1987

Preston etal.t992

Preston etaI. t992Hordeum leporinum paraquat

5

1.3.2 Enhanced Herbicide Metabolism

After entering the plant, the herbicide may be metabolized within the cell to an inactive

product. Changes that enhance metabolism of the herbicide within the plant, may result in a

reduction in the amount of the active form of herbicide. Several weed species have evolved

resistance to herbicides due to an increased capacity to detoxify the herbicide into less toxic

compounds as shown in Table 1.3.

Table 1.3 Enhanced metabolism of herbicide as a mechanism of resistance in weeds

Species Herbicide Reference

Abutilon theophrasti atazine Gronwald et al. 1989

Alopecurus myosuroides chlorotoluron Moss 1990

Alopecurus myosuroides diclofop, fenoxaprop Hall et al. 1993

Lolium rigid.um diclofop-methyl Holtum et al. 1991

Loliurn rigi.dum PS tr inhibitors Burnet et al. 1993a and 1993b

Inlium rigidum chlorsulfuron

Loliutn rigidum

Stellariamedia

Ealkoxydim

mecoprop

Christopher et al. 1991

Cotterman and Saari 1992

Preston et al. unpublished

Coupland et al. 1990

6

1.3.3 Modified Site of Action

When the herbicide reaches the target site in a sufficient concentration to inhibit the

biochemical target site, a modification of the target site may reduce the herbicidal effect and

provide resistance. Reduced sensitivity of photosystem II, ACCase and ALS are

documented as target site mechanisms of resistance. Most of the 57 species with resistance

to PS tr-inhibiting herbicides have a target site change which renders them resistant to the

triazine herbicides. A single amino acid change (ser 264 to gly) in the Qs-binding site

within the D1 protein of PS II is responsible for resistance in these plants (Gronwald 1994).

Examples of resistant species exhibiting reduced sensitivity of herbicides at the target sites

ACCase and ALS are shown in Tables 1.4 and 1.5, respectively.

Table 1.4 Reduced sensitivity of ACCase enzyme as a mechanism of resistance in plants

Species Herbicide Reference

Festucarubra APP, CHD Stoltenberg et al. 1989

Festtrca ovina

Festuca amethystina

APP, CHD Catanzaro et al. 1993

Lolium multiforum diclofop-methyl Gronwald et aL. 1992

I-oliutn rigidum APP, CHD Tardif et al. 1993

Tardif and Powles 1994

Setarinviridis APP, CHD Ma¡les et al. 1993

Tzamays haloxyfop,sethoxydim Parkeretal. 1990a

7

T¡ble 15 Insensitive ALS enzyme as a mechanism of resistance in weeds

Species Refercnce

Brassicatourneþrtü BoutsalisandPowlesunpublished

Kochia scoparia Saarietal.1990

Loliwn perenne Saa¡i et al.1992

Lolium rigidum Christopher etal.1992

Salsola iberica Saari et aL.1992

Sisynbrium orientale Boutsalis and Powles unpublished

Sonchus oleraceus Boutsalis and Powles 1995b

Stellariamedia Devine et al. 1991

Xanthium strumarium Schmiøer et al. 1993

In addition, some resistant weed species containing a sensitive target enzyme ACCase

exhibit altered membrane function. This is manifest as a recovery of the plasma membrane

electrogenic potential following diclofop acid-induced depolarization (Holtum et al. 1991,

DiTomaso 1993, Devine et al. 1993). Membrane depolarization has been proposed as a

second site of action of ACCase-inhibiting herbicides (more detail in section 1.7).

Therefore, repolarization of the plasma membrane potential following removal of herbicide

has been proposed to be a mechanism of resistance to these herbicides in some weed species

(Table 1.6).

8

Table 1.6 Alæred plasma membrane response as a mechanism of resistance in weeds

Species Herbicide Reference

Avenafatua

I¡Iiutn rigidum

Loliumrigidum

APP Devine et al. 1993

APP, CHD Häusler et al. 1991

Holtum et aI. 1991

DiTomaso 1993

APP Shimabuhuo and Hoffer 1992

1.3.4 Sequestration of Herbicide within Plant Cells

Sequestration is another possible mechanism of herbicide resistance. Holtum et al. (1991)

proposed that resistant L. rigidurø plants which were able to grow in the presence of

substantial pools of diclofop acid in the tissue could do so because the toxic form is

sequestered in the vacuole or in the apoplastic space. Under this scenario, the concentration

of diclofop acid at the active site is too low to substantially inhibit ACCase activity or

depolarize the plasma membrane potential. However, this phenomenon has yet to be

proved.

1.4 Inheritance of Herbicide Resistance

In most cases of resistance, inheritance of resistance is conferred by a single gene. In the

case of triazine resistance, inheritance is usually maternal, whereas in other cases of

resistance it is often a partially-dominant, nuclear gene as shown in Table 1.7. Resistance to

dinitroaniline herbicides in Setaria viridis is an exception and is due to a single recessive

9

gene (Jasieniuk et aI. 1994). However, resistance to ALS inhibitors in some biotypes of

Loliutn rigidwt in Australia is controlled by at least ttto genes (Christopher 1993).

Table 1.7 Inheritance of herbicide resistance in weed biotypes

Herbicide Species Number of gene Reference

a¡razme Abutilon theophrasti I partially-dominant Anderson and Gronwald 1987

cblorsulfuron Sonchus oleraceus I partially-dominant Boutsalis and Powles 1995b

chlortoluron Alopecurus myosuroides 2 addiúve Chauvel l99l

diclofop Lolium multiþrum l partially-dominant Betts et al.1992

fenoxaprop Avena sterilis 1 partially-dominant Barr et al. 1992

fluazifop Avena sterilis 1 partially-dominant Barr et al.1992

haloxyfop Lolium rigidum 1 partially-dominant Ta¡dif et al. submined

metsulfuron l-actuca seniala 1 partially-dominant Mallory-Snith et al. 1990

paraquat Erigeroncanadensß l dominant Itoh and Miyahara 1984

paraquaf Conyzt bonariensis l dominant Shaaltiel et al. 1988

pa¡aquat Hordeum glaucum I partially-dominant Islam and Powles 1988

praquat/diquat Arcthotheca calendulo I partially-dominant Purba et al. 1993a

paraqur Hordeum leporinum I partially-dominant Pu¡ba et al. 1993a

siduron Hordeum jubatum 3 complementary Shooler et al. 1972

trifluralin Setariaviridis I recessive Jasieniuk er al. 1994

l0

1.5 Aryloxyphenoxypropionate and Cyclohexanedione Herbicides

ACCase-inhibiting herbicides are represented by two classes of herbicides: the

aryloxyphenoxypropionate (APP) and substituted l,3-cyclohexanedione (CID) herbicides.

These are stn¡cturally different but exhibit simila¡ ph¡otoxicity and symPtomology. The

chemical structure and common name of these herbicides are given in Table 1.8. APP and

CHD herbicides are sometimes called grarninicides because they selectively control grass

weeds in dicot crops (Gronwald 1991). Diclofop-methyl, fenoxaprop-ethyl and

tralkoxydim are used for selective post-emergent wild oat control in wheat, barley, rye and

dicot crops. Fluazifop-butyl, haloxyfop-ethoxyethyl and sethoxydim are widely used as

selective post-emergent herbicides in dicot crops. Better confrol of grass weeds is obtained

when these herbicides a¡e applied at the two to th¡ee leaf stage (Ashton and Monaco t99I,

Harker and Blackshaw 1991). The site of action of APP and CHD herbicides is mainly in

the meristematic regions of the susceptible plants (Hoerauf and Shimabukuro 1979, Hosaka

et al. 1984, Hosaka et at. 1987). Injury symptoms produced by these graminicides on

susceptible plants are necrosis in the meristematic tissue, necrosis or chlorosis in developing

leaf tissue and appearance of purple leaf colour due to anthocyanin accumulation (Hoerauf

and Shimabukuro Ig7g,Ishigawa et al. 1985, Köcher et aI. 1982, Swisher and Corbin

1982, Asare-Boamah and Fletcher 1983). General usage, mode of action, uptake,

translocation and metabolism of APP (diclofop, fluazifop and haloxyfop) and CHD

(sethoxydim and tralkoxydim) herbicides will be discussed in the following sections.

The APP herbicides are chiral due to an asymetrical carbon atom in the propionic side chain.

These herbicides are usually a racemic mixture of two isomers, R(+) and S(-) (Duke and

Kenyon 1988). The R-enantiomer is the herbicidally-active form (Dicks et al. 1985,

Uchiyama et al. 1986, Harwood et al. 1987, Gerwick et al. 1988, Secor et al. 1989). The

APP herbicides a¡e formulated as esters of the parent acid to facilitate penetration into the

leaf cuticle. After entering the plant, the ester form is hydrolysed rapidly to the phytotoxic

1t

Table 1.8 Structure of graminicides

(a) Aryloxlphenoxlpropionate (APP) herbicides

F'o&R1-

R1 Rz Common tame

CH¡ diclofop-methyl

\)- CzHs fenoxaprop-ethyl

c¿Hs fluazifopbutyl

cH¡ haloxyfop-methyl

czHs quizalofopethyl

(b) Cyclohexanedione (CtD) herbicides

o-R¿R1

R1 R2 R3 R4 Commonname

l"' CzHs CH2CIFCH clethodimC2H5S-C-CH2- H2

H2

H2

H2

H

CtHt CzHs cycloxydim

CtHt CzHs sethoxydim

cHr

c¡I

S-r(}

1",qH5rc-cH2-H

cHs

CH¡

CzHs CzHs Ealkoxydim

12

carboxylic acid (Shimabukuro et al. 1979, Stonebridge 1981, Duke and Kenyon 1988,

Harwood et al. 1989) which is then translocaæd to meristematic regions (Buhler et al. 1985,

Hendley et al. 1985, Chandrasena and Sagar 1987 and 1989).

1.5.1 Antagonism of APP and CIID Herbicides

The general lack of activity of APP and CHD herbicides on dicot weeds makes mixtures

with other herbicides important in order to broaden the spectrum of weed control.

However, auxinic compounds and some other herbicides can antagonise the efficacy of APP

and CHD herbicides when applied as a tank-mixture. For example, addition of 2,4-D

dimethylamine reduced the efficiency of diclofop-methyl on wild oat (Todd and Stobbe

1980), fenoxaprop and haloxyfop on Sorghum halepense (Mueller et al. 1989 and 1990),

and tralkoxydim on wild oat (Jensen and Caseley 1990). In most cases, 2,4-D reduced

uptake and/or translocation of those herbicide into plants. Dicamba can also antagonise the

effect of haloxyfop-methyl onFestuca arund.inaceø (Aguero-Alva¡ado et al. 1991).

Bentazon is another compound that can antagonize the effects of APP and CHD herbicides.

Penetration of haloxyfop into Setaria lutescens leaves was reduced3TTo in the presence of

bentazon due to crystallization of bentazon on the leaf surface, thus forming a physical

barrier that impaired the penetration of other xenobiotics through the cuticle (Gerwick 1988).

Aguero-Alvarado et al. (1991) found that bentazon alone or mixed with haloxyfop

significantly inhibited ACCase activity only at high concentrations. For example, l0 mM

bentazon was required to reduce ACCase activity of a susceptible grass (F. arundinøcea)by

40Vo. They suggested that antagonism may be due to mechanisms that reduce the amount of

haloxyfop reaching the target site. In addition, bentazon reduced the uptake and

translocation of tralkoxydim resulting in a reduction of wild oat control (Jensen and Caseley

1990).

1.5.2 Uptake and Translocation

The ester forms of APP herbicides are readily absorbed by leaves and roots. Tritter et al.

(1987) found that uptake of diclofop-methyl was ten times greater than diclofop acid into oat

13

protoplasts over the same period. They suggested that this might be due to the acid group of

diclofop being dissociated at physiological pH giving a charged molecule which would be

less penneable to plant membranes and result in reduced uptake compared to diclofop-

methyl which is less polar. Uptake of [lag¡Ooazifop-butyl was more rapid in young leaves

compared to older leaves (Ca¡r 1985). Absorption of fluazifop was greater at 30'C to 35'C

than at 18'C to 20'C (Ca¡r 1985). Cycloxydim absorption by suspension-cultured cells of

Abutilon theophrasti was significantly greater at2O'C compared with 4 and 12"C, but no

difference was observed between the two lower temperatures @gyneberg et al. 1994).

Translocation of APP and CHD herbicides varies between individual herbicides.

Translocation of diclofop-methyl in plants is limited. A number of studies have established

that less than 5Vo of absorbed [14C]diclofop-methyl is translocated out of the treated zone in

a range of species (Brezeanu et al. 1976, Boldt and Putman 1980, Hall et al. t982,

Jacobson and Shimabukuro t982, Dahroug and Muller 1990, Baker and Chamel 1990,

Devine etal.1992). The predominant tanslocated forms are diclofop-methyl and diclofop

acid (Jacobson and Shimabukuro 1982). Diclofop was found to move mainly in an

acropetal direction when applied to different parts of wild oat plants (Friesen et al. 1976).

Beckett et al. (1992) found that translocation of ¡l+grn rizalofop out of the treated area was

limited and translocation occurred both acropetally and basipetatly in similar arnounts.

Fluazifop acid or its glucose conjugated products a¡e the mobile forms of fluazifop (Hendley

et al. 1985, Carr et al. 1985). Fluazifop is readily translocated from treated leaves to

rhizomes and stolon of perennial grasses and accumulates in meristematic tissue

(Stonebridge 1981, Ca¡r et al. 1985, Chandrasena and Sagar 1987). Fluazifop moves in the

phloem with the assimilate stream along 'source-sink' relationships of the plants (Kell et al.

1984, Hendley et al. 1985, Chandrasena and Sagar 1987). As a result, applications of

[l4C]fluazifop to mature leaves resulted in large quantities of radioactivity accumulating in

lower stem regions, roots and rhizomes of Agropyron repens, compared with treatments to

younger leaves (Chandrasena and Sagar l9S7). In addition, translocation of [l4C]fluazifop

t4

was greater from lamina base applications than from treatments to the lamina apex and

middle (Chandrasena and Sagar 1987).

Little translocation of l4C out of the treated leaf occurred following [l4C]traloxyfopmethyl

application to the youngest leaf of susceptible and tolerant species @uhler et al. 1985).

These authors also found that highly susceptible Sorghum bicolor translocated more 14C

than did the susceptible Setaria glauca. Following llaC]haloxyfop-methyl application to

leaves of a stolon node of Cynodon dactylon less than 37o of the radiolabel translocated to

the apex and base (Marorder et al. 1987).

CHD herbicides a¡e absorbed rapidly into plants and translocated to the target site. Rapid

uptake of sethoxydim occurred in the first 3 h and by 6 h more than half of heöicide applied

had entered wild oat leaves (Smith and Vanden Born 1992). Absorption of

[laC]cycloxydim by suspension-cultured velvetleaf (Abutilon theophrasti) cells was

dependent on concentration, temperature and pH (Rgyneberg et al. 1994). Absorption of

cycloxydim was linea¡ly related to the extracellular herbicide concentration. Cycloxydim

absorption was significantly greater at2O'C compared to at 4 or l2'C. Cycloxydim was

absorbed >5.5 fold at pH 3.7 compared to pH 5.7 (Røyneberg et al. L994). Ammonium

sulfate can increase the efficiency of sethoxydim by increased absorption and translocation

in Digitaria sanguinalis, Hordeum vulgare and,Avenafatua (Jordanet al. 1989, Smith and

Vanden Born 1992).

1.5.3 Metabolism

APP herbicides are absorbed into plants as an ester form and then are hydrolyzed to the

ph¡otoxic acid form. Grass crops such as wheat rapidly metabolize some APP herbicides

to inactive products (Shimabukuro et al. 1979 and 1987). However, wheat is susceptible to

fluazifop, haloxyfop and some other APP herbicides because it is unable to detoxify these

herbicides. The metabolic pathway for diclofop-methyl in resistant (e.g. wheat) and

susceptible plants (e.9. wild oat, muze and other susceptible grasses) is shown in Figure

1.1. In susceptible species, the phytotoxic diclofop acid is mostly converted to the neutral

15

glycoester which can be hydrolyzed back to the toxic form (Shimabuku¡o 1990). In wheat

diclofop acid undergoes aryl-hydroxylation by a cytochrome P450 monooxygenase

(Zimmerlin and Durst 1992). Subsequently, most of the hydroxylated diclofop is

conjugated rapidly to glucose to form the aryl-O-glucoside metabolite (Shimabukuro et al.

1987).

Metabolism of fenoxaprop-ethyl was studied in soybean (Wink et al. 1984) and grass

species (Lefsrud and Hall 1989, Tal et al. 1993). In soybean, fenoxaprop-ethyl is rapidly

hydrolyzed to the ph¡otoxic form, fenoxaprop acid, which is subsequently metabolized to

polar conjugates $fink et al. 1984). Tal et al. (1993) proposed that fenoxaprop-ethyl is

detoxified by glutathione, cysteine and glucose conjugations in some grass species. In

wheat (Triticum aestivum), barley (Hordeum vulgare), crabgrass (Digitaria ischaemum) and

oat (Avena sativa), fenoxaprop-ethyl is rapidly hydrolyzed to fenoxaprop acid (Figure 1.2).

After 8 h of incubation, in susceptible crabgrass and oat, most of the radioactivity was

recovered as the ph¡otoxic form of fenoxaprop acid and only small amounts were convefed

to water-soluble metabolites. In contrast, in wheat and barley, fenoxaprop acid underwent

rapid displacement of the phenyl group by glutathione (GSH) and/or cysteine, resulting in

production of S-(6-chlorobenzoxazole-2-yl)-glutathione (GSH conjugate), 5-(6-

chlorobenzoxazole-2-yl)-cysteine (cysteine conjugate), and 4-hydroxyphenoxy-propanoic

acid. The GSH conjugate may be catabolized to form the cysteine conjugate which was

subsequently metabolized to an unidentified metabolite. The 4-hydroxyphenoxypropionic

acid was further metabolized to form a glucoside conjugate.

l6

o

o

F.,IH

diclofopmethyl

diclofop acid

H

susceptible

COO-Glucose

H

neutral glucose ester of diclofop

CH¡t-

oresistant

ct

o cooHIH

OH

aryl-hydroxylated diclofop

oCH.t-IH

cooH

acidic aryl-O-glucoside of diclofop

Figure 1.1 Metabolism of diclofop-methyl in plants (redrawn from Shimabukuro 1990)

t7

1",O- CH- COOC2H5

CH¡l-

o- cH- cooH

1",o- cH- cooH

4-hydroxylated-propanoic acidS.

cl o co-NH-cH2-cooH

S-(6-chlorobenzoxazole-2-yl)-glutathione

S-

N

cl o

fenoxaproprthyl

o

fenoxaprop acid

,l

cooH

l"'o- cH- cooH

N

cl

NHnl'

N NH-CO-CHr-CHI-CH-COOHl-

CHz-1

H

pO-glucoside conjugate

NHr

cHz-lN

ocl

S-(6-ctrlorobenzoxazole-2-yl)-cysteine N-glucoside

unidentified metabolite

Figure 1.2 Metabolism of fenoxaprop-etþl in wheat, barley, crabgrass and oat (redrawn

from Tal et al. 1993)

l8

Less information is available on the metabolism of CHD herbicides in plants.

Cyclohexanediones are unstable compounds in both biotic and abiotic systems (Swisher and

Corbin 1982, Røyneberg et al. 1994). Decomposition of sethoxydim is enhanced by

alkaline conditions (Shoaf and Carlson 1986) and ultraviolet light (Shoaf and Ca¡lson

1992). Sethoxydim decomposes in water, especially at basic pH, leading to generation of at

least 14 possible degradation products (Shoaf and Carlson 1986). The metabolism of

sethoxydim in plants is rapid and most of the parent herbicide is degraded within 6 to 72 h

after application (Campbell and Penner 1985, Hosaka et al. 1987, Ishihara et al. 1988,

Tardif et al. 1993). Campbell and Penner (1985) found six major breakdown products of

sethoxydim and two of them were speculated to be toxic forms. However, it is unclear

whether sethoxydim or its breakdown products are the actual herbicidal agent. The

metabolism of tralkoxydim was studied in wheat gfolvn under field condition and sprayed at

a rate of 345 g ai ha-l (Hadfield et al. 1994). In wheat, the metabolism of tralkoxydim

involved hydroxylation of the p-methyl function group of the a¡omatic ring (Figure 1.3).

This oxidation may occur in conjunction with either transformation of the

ethoxy(imino)propyl function or with oxidative cleavage of the cyclohexanedione ring. All

the initial products were then conjugated to a range of endogeneous sugars. Although there

is no conclusive evidence, it is likely that the selectivity of tralkoxydim between susceptible

and resistant grass species is based on different rates of herbicide metabolism.

/CH

cH2 - co2H

CHO -COOHcll3

3-(2,4,Gtimethylphenyl)pentanedioic acid

l9

QHs

NOC2H5

CzHs

cH¡ o

2-( I -(ethoxyimino)propyl)-3-hydroxy-S-(zl-hydroxymethyl-2,Gdimethylphenyl)cyclohex-2-enone

cH3 o

Tralkoxydim

CuHs

N

cH3

2- ethyl-G (4-hydroxymethyl-2, 6-dimethylphenyl)-4,5,6,7-tetrahydrozoxazol-4-one

HO HorcH2 cH2

OH

NI{

czHs

cH¡ o

3-hydroxy-2-( 1 -iminopropyl)-5-(a-hydroxymethyl-2,6-dimethylphenyl)cyclohex-2-enone

cH, - co,H/

CH

CH¡cH2 - co2H

3-(zt-hydroxymethyl-2,Gdinethylphenyl)pentanedioic acid

Figure 1.3 Metabolic pathway of tralkoxydim in wheat foliage and straw (redrawn from

Hadfield et al. 1994)

"ol

Hol

cH.

cHz

20

1.6 Mode of Action of APP and CHD Herbicides

Two basic mechanisms of action have proposed for both APP and CHD herbicides. One is

a biochemically-based mechanism involving the inhibition of acetyl-CoA carboxylase

(ACCase) and the second is a biophysically-based mechanism involving a perturbation of

the transmembrane proton gradient across plasma membrane (Figure 1.4).

PlasmalemmaInside

ATP

Outside

IT+.ATP

ADP+Pi

ÂpH

PlastidsH+ channel (?)

¿tH'1

¿H.

I

tH'1

diclofop-methyl

diclofop

and/or diclofop-methyl

Em

G120mV) (0mV)

Figure 1.4 A diagram of biophysical and biochemical mechanisms of action of APP

herbicides (redrawn from Devine and Shimabukuro 1994)

diclofopI +

+

+

+

+

+

IAA,2,4-D

TTITITTTTTTTI

acetyl-CoA

malonyl-CoA

()ØCÚ

UU

+

2l

1.6.1 Inhibition of Fatty Acid Biosynthesis

Fatty acid biosynthesis in plants occrus exclusively in plastids (Harwood 1988). ACCase

(acetyl-coerizyme A:bicarbonate ligase [ATP], E.C. 6.4.1.2), a biotin-containing ewzymø

catalyzes the first step in fatty acid biosynthesis via ATP-dependent carboxylation of acetyl-

CoA to fonn malonyl-CoA (Harwood 1989). Plant ACCase is a high molecular weight,

multifunctional enzyme comprised of three catalytic domains, biotin carboxylase, biotin

carboxyl carrier protein and carboxyltansferase domains (Figure 1.5). Initially, a carboxyl

group is donated from a bica¡bonate anion, and ATP hydrolysis is used to a ca¡boxybiotin

intermediate by biotin carboxylase. Carboxybiotin is attached to an e-amino group of a

lysine residue on the biotin carboxyl ca¡rier protein (BCCP). Carboxybiotin then functions

as a CO2 donor in malonyl-CoA formation (Harwood 1989). Malonyl-CoA is an

intermediate in a number of metabolic pathways, such as the biosynthesis of cuticular waxes

and flavonoids (Harwood 1989). The condensation of acetyl-CoA and malonyl-ACP

catalyzedby 3-ketoacyl-ACP synthase is the fust step in a series of reactions in the fatty acid

synthase complex that results in 16:0-ACP and l8:I-ACP as major products of fatty acid

synthesis in plastids (Browse and Sommerville 1991). Phospholipids are synthesized at the

endoplasmic reticulum through initial acylation of glycerol phosphate by acyl-CoA

transferase.

ACCase is the target site of the aryloxyphenoxypropionic acid and the cyclohexanedione

herbicides in susceptible grass species (Burton et al. 1987 and 1989, Rendina et al. 1988

and 1989, Secor and Cséke 1988, Secor et al. 1989, Walker et al. 1989, Howard and

Ridley 1990). Generally, ACCase from monocots is sensitive to both classes of herbicides

and that of dicots is not (Cho et al. 1986, Burton et al. 1987, Litchtenthaler and Kobek

1987, Kobek et al. 1988, Rendina and Felts 1988, Rendina et al. 1988, Secor and Cséke

1988, Walker et al. 1989, Boldt and Barett 1991). However, some monocot crops such as

wheat are tolerant to diclofop and fenoxaprop despite containg a herbicide-sensitive

ACCase, as a result of rapid metabolism of these herbicides to nontoxic products

(Shimabukuro et al. 1979, Lefsrud and Hall 1989).

22

acetyl-CoA carboxylase

biotin biotin carboxyl BCCP:acetyl-CoAca¡rier protein fianscarboxylase

(BCCP)carboxylase

aryloxyphenoxypropionic acids

cyclohexanediones

MSATP

biotin \ carboxylbiotinbiotin-CO,

malonyl-CoA

"4 biotin

/pyruvate acetate

malonyl-ACPoCË

ÊøEo6!

àd

J3-ketoacyl-ACP

+l6:GACP

+18:O-ACP

Plastid vl8:1-ACP

vacyl-CoA

phospholipidsacyl-CoA transferase

Figure 1.5 Schematic of fatty acid biosynthesis in plants

glycerol-3-

reticulum

23

1.6.2 Characteristics of ACCase Enzyme

ACCase is found in plastids and is active over a wide pH range with an optimum at pH 8.0

in purifred plastids of developing endosperm of castor seeds (Finlayson and Dennis 1983)

and with optimum at pH 8.2 in soybean seeds(G/ycine max L. Men.) (Charles et al. 1986).

A divalent cation, Vtn2+ or Mg2+, is necessary for its activity @nlayson and Dennis 1983),

but Mn2+ could not replace Mg2+ as an essential activator of ACCase in soybean (Charles et

al. 1986). Based on experiments on the utilization of acetate by chloroplasts in the light and

dark, the biosynthesis of fatty acids from acetate in chloroplasts was induced by light

(Nakamura and Yamada 1975) with little or no accumulation of fatty acids in leaves in the

dark (Browse et al. 1981). Nakamura and Yamada (1979) proposed that the carbon flow

for fatty acid synthesis is limited at the step of ACCase in a light-dependent reaction.

Moreover, the kinetic properties of this enzyme in maize chloroplasts demonstrated that its

activity may be modulated by light-dependent changes of metabolite levels, particularly by

alterations in the stromal concentration of ATP, ADP, Mg2+ and by pH (Nikolau and

Hawke 1984). During seed maturation, the level of ACCase markedly increased just before

the onset of major lipid accumulation (Turnham and Northcote 1983).

There are two types of ACCase: prokaryotic ACCase, in which the three functional

domains-biotin carboxylase (BC), biotin carboxyl carrier protein (BCCP) and

carboxyltransferase (CT) a¡e located in separable proteins (as found in Escherichia coli and

probably pea chloroplast) (Li and Cronan l992aa¡d 1992b, Alban et al. 1994, Konishi and

Sasaki 1994), and the eukaryotic ACCase, in which all the domains a¡e located on one large

polypeptide (as found in rat, chicken, yeast, diatom and wheat) (I-ópez-Casillas et al. 1988,

Takai et al. 1988, Al-Feel et al. 1992). In plants, both eukaryotic and prokaryotic forms

have been found (Kannangara and Stumpf 1972, Nikolau and Hawke 1984, Egli et al.

1993, Sasaki et al. 1993, Alban et al. 1994, Konishi and Sasaki 1994).

Plants contain two isoforms of ACCase that appear to have different subcellular localization

and metabolic roles. A major isoform, which is responsible for the bulk of the cell ACCase

activity, is located in the chloroplasts and other plastids, whereas a minor isoform is

24

presumably found in only the cytosol @gli et al. 1993, Alban et al. 1994, Ashton et al.

1994, Konishi and Sasaki 1994). The major isoform plays a key role in fatty acid

metabolism and the minor isoform is involved in plant defence against pathogens and

elongation of very long chain fatty acids (Shorrosh et al. 1994).

Recent evidence indicates that the types of ACCase present in dicotyledonous species are

fundamentally different to the types present in monocot species. Pea contains a eukaryotic

ACCase in the cytoplasm but a prokaryotic multi-subunit ACCase in the chloroplasts (Alban

et aL. t994, Konishi and Sasaki 1994, Sasaki et al. 1994). In contrast, grasses contain

eukaryotic ACCase in both the plastids and cytoplasm (Egli et al. 1993, Ashton et al. 1994).

In young pea leaves, a minor eukaryotic ACCase (-20Vo of the total ACCase activity in the

whole leaf), located in the epidermis, \r,as strongly inhibited by diclofop (Alban et al. 1994).

In contrast, a major prokaryotic ACCase, present in both epidermis and mesophyll, was

highly resistant to herbicide (Alban et al. 1994). Konishi and Sasaki (1994), working with

pea, wheat and rice, found that prokaryotic ACCase \ryas located in the plastids and the

eukaryotic ACCase was presumably in the cytosol. The prokaryotic ACCase was

insensitive to ACCase-inhibitng herbicides whereas the eukaryotic ACCase was sensitive to

herbicide inhibition. Pea is resistant to ACCase-inhibiting herbicides since it contains the

prokaryotic, herbicide-insensitive form of ACCase. In contrast, wheat and rice are

susceptible to ACCase-inhibiting herbicides because they lack the prokaryotic ACCase

(Konishi and Sasaki 1994). This probably explains why grasses a¡e sensitive to APP and

CHD herbicides, whereas diclots are susceptible.

In grass species like maize, a herbicide-insensitive form of ACCase firstly repofed by

Howa¡d and Ridley (1990) has been observed (Egli et al. 1993, Ashton etal.1994} Egli et

al. (1993) found that a major isoform of ACCase (ACCase I), located in most tissues, was

strongly inhibited by haloxyfop and sethoxydim whereas a minor isoform (ACCase II), not

found in mesophyll plastids, was insensitive to those herbicides. Ashton et aI. (1994) also

found that a minor form of ACCase from maize leaves was insensitive to fluazifop and

talkoxydim while the major ACCase form was strongly inhibited.

25

ACCase from susceptible and APP herbicide resistant biotypes of Lolium multiþrumwas

purified and characteizedby Evenson et al. (L994). They found that both susceptible and

resistant forms of ACCase were simila¡ in subunit molecular mass, pH optimum and

substrate kinetics. However, the resistant form of ACCase exhibited altered kinetics of

diclofop binding and a different isoelectric point. They speculated that the ACCase fraction

from the resistant biotype contains both a diclofop-susceptible and -resistant forms.

1.6.3 Kinetics of ACCase Inhibition by APP and CHD Herbicides

The kinetics of inhibition of ACCase by APP (diclofop, fluazifop, haloxyfop) and CHD

(clethodim and sethoxydim) herbicides are linear, reversible and non-competitive with

respect to all three substrates, acetyl-CoA, MgATP and bicarbonate (Rendina and Felts

1988, Rendina et al. 1989, Howa¡d and Ridley 1990, Rendina et al. 1990, Burton et al.

1989 and l99l). Kinetic analyses suggested that APP and CHD herbicides are mutually

exclusive inhibitors of ACCase erzyme and both classes of herbicides compete for the same

binding site (Rendina et al. 1989, Rendina et al. 1990, Burton et al. 1991). In addition,

APP and CHD herbicides inhibit the transcarboxylation rather than the biotin carboxylation

reaction (Rendina et al. 1990, Burton et at. 1991).

1.6.4 Inheritance of Herbicide Resistant ACCase in Formerly Susceptible

Weed Species

Recently, a resistant form of ACCase has been reported in several grass species previously

controlled by ACCase-inhibiting herbicides (Stoltenberg et al. 1989, Betts et aI. t992,

Holtum and Powles 1991, Mansooji et al. 1992. Tardif and Powles 1993) and in maize

(Parker et al. 1990a). In each species, the degree of resistance at the enzyme level is

correlated with resistance at the whole plant level. Inheritance of resistance in wild oat (r4,.

sterilis ssp. ludoviciana) and Italian ryegrass (L. multiþrum) is due to a single, partially-

dominant, nuclear gene (Barr et al. 1992, Betts et al. 1992). Inheritance of resistance to

ACCase-inhibiting herbicides n L. rigidun biotype rWLR 96 is also controlled by a single,

partially-dominant, nuclear gene (Tardif and Powles 1993). Inheritance studies conducted

26

with regenerated plants from graminicide-resistant maize lines selected in tissue culture,

showed that resistance is due to a single nuclea¡ gene er.hibiting partial dominance (Parker

et al. 1990a, Marshall et al. 1992). Among those graminicide-resistant maize lines, th¡ee

mutant ACCase alleles (Accl-S,AccI-HI,Accl-H2) were identified (Ma¡shall et al.

1992). The Accl-S and Accl-H2 alleles conferred target-site cross-resistance to both APP

and CHD herbicides, whereas the AccI-Hl allele conferred selective resistance to the APP

herbicides.

1.7 Disruption of Membrane Function by APP and CHD

Herbicides

The second mode of action of aryloxyphenoxypropionic acids and cyclohexanediones is to

depolarize the plasma membrane potential of susceptible species. Plant cells establish and

maintain a transmembrane proton gradient which is vital to their growth and development.

Any chemical such as diclofop-methyl that disrupts proton electrochemical gradients may

have significant effects on many aspects of the physiology of the plant (Shimabukuro

1ee0).

The free acids of APP herbicides (such as diclofop) depolarize membrane potentials in

Chara spp. (Lucas et al. 1984), parenchyma cells of oat and wheat (Wright and

Shimabukuro 1987), coleoptiles and root tips of A. fatua and Z. rigidum (Häusler et al.

L99t, Holtum et al. 1991, Shimabukuro and Hoffer 1992, Devine et al. 1993). Similar

effects have been found in isolated vesicles of plasma membrane and tonoplast of oat root

(Ratterman and Balke 1988, 1989). Similarly, CHD herbicides such as sethoxydim

depolarize membrane potentials of leaf cells from sensitive grasses (Weber and Lünge 1988,

Weber et al. 1988), and coleoptiles of susceptible and resistant L. rigidum biotypes (Häusler

et al. 1991). The mechanism by which diclofop affects the electrogenic properties of the cell

membrane is poorly understood.

t1

1.7.1 Diclofop as a ProtonoPhore

Diclofop was proposed to act as a specific protonophore that moves protons across the

plasmalemma of plant cells resulting in an influx of protons into the protoplast (Lucas et al.

1984,V/right and Shimabukuro 1987). According to this model, the methyl ester form of

diclofop crosses the plasmalemma and is rapidly hydrolyzed to the free acid form. This

undissociated acid then releases a proton on the c¡oplasmic side of the membrane (Figure

1.4). The deprotonated anionic species then moves back across the plasmalemma to the

apoplasmic side in response to the root cell membrane potential (inside negative) and is

again protonated. This cycle is repeated until the proton gradient is dissipated and

eventually leads to a reduction in cellular ATP levels by short circuiting the proton-

translocating plasmalerlma. Later, Shimabukuro and Hoffer (1992) found that diclofop did

not act like a classic proton ionophore such as z-chlorophenylhydrüzone (CCCP) and

membrane depolarization by diclofop could be prevented by p-

chloromercuribenzenesulphonic acid (PCMBS), a non-permeant molecule that binds -SH

groups on proteins. This suggests that diclofop may interact with a specific plasma

membrane-associated protein to increase proton influx into sensitive cells.

rWhether the membrane depolarization caused by this herbicide leads to phytotoxicity in

plants is the subject of some debate. DiTomaso et al. (1991) found that diclofop acid was

capable of depolarizing the plasma membrane potential of susceptible maize and oat roots, as

well as resistant pea roots. The similar membrane response of resistant pea and susceptible

muze to diclofop acid suggested that the membrane depolarization is not phytotoxic. In

addition, after removal of diclofop acid from the treatment solution, complete repolarization

was observed in both maize and pea root cells (DiTomaso et al. 1991).

1.7.2 Recovery from Diclofop Acid Induced Depolarization of the Plasma

Membrane Electrogenic Potential in Some Resistant Weeds

Evidence that differences in the ability of plasma membranes to recover from depolarization

may be correlated with the resistance responses of some resistant weeds has been found in

28

diclofop-resistant L. rigidum and.á. fatua biotypes (Holn¡m et al. 1991, Häusler et al.

1991, Shimabukuro and Hoffer 1992, Devine et al. 1993). Etiolated coleoptiles and root

tips of foru diclofop-methyl-resistant annual ryegrass biotlpes recovered membra¡e polarity

following exposure to diclofop acid, whereas the susceptible plants did not (Häusler et al.

199L, Holtum et al. 1991, Shimabukuro and Hoffer t992). Although there is a good

correlation between the recovery and resistance, the mechanism producing the difference in

membrane response between the susceptible and resistant biotypes is still unclear.

DiTomaso (1993) suggested that the differential membrane response between susceptible

and resistant L. rigidurn biotypes may be the result of differences in cell wall pH. He also

demonsEated that the repolarization of the membrane potential was pH dependent and use of

a buffered solution could eliminate the differential membrane responses. In solution

buffered at pH 5 or pH 6 (2 mM MES-Tris), both susceptible and resistant L. rigidum

biotypes behaved similarly following diclofop acid-induced depolarization (DiTomaso

1993).

1.8 Resistance to Acetyl-CoA Carboxylase-Inhibiting Herbicides

The first reported case of resistance to ACCase-inhibiting herbicides was a diclofop-methyl

resistant biotype of L. rigi.dum in Ausfralia reported by Heap and Knight (1982). To date, at

least ten species of weeds have become resistant to ACCase-inhibiting herbicides (Table

1.9). Three of these resistant species, L. rigidum,Avenafatua andA. sterilis, a¡e found in

Australia.

29

Table 1.9 Developed resistance to ACCase-inhibiting herbicides in grass species

Species Herbicide Origin Reference

Avenalattn APP, CHD Canada Heap et al. 1993

Avenafatrn APP Austalia Mansooji et al. 1992

Avena sterilis APP Australia Mansooji et al. 1992

Eleusine itúica fluazifop Malaysia Leach et al.t993

Fesucarubm APP, CHD USA Stoltenberg et al. 1989

I-olium multþrum diclofop USA Stanger and Appleby 1989

I-olium rigidum APP, CHD Australia Heap and Knight 1982 and 1986

Alopecurus myosuroides APP, CHD TJK Hall et al.1993

Setariafaberi APP, CHD USA Stoltenberg and V/iederholt 1993

Senriaviridis APP, CHD Anada Ma¡les et al. 1993

Sorghum lulepense APP, CHD USA Barrentine etal. 1992

Smeda et al. 1993

30

1.9 \üild Oat (Avena spp.)

lVild oat is an important weed of cereal and a¡able crops throughout the world, and is of

major importance in Australia. It is estirnaæd that wild oat infestations cost Austalian wheat

growers at least M2 million per year (Medd and Pandey 1990) due to the reduction of yields

through wild oat competition and the cost of herbicides applied to control wild oat. There

are three species of wild oat, Avena latuaL., A. sterilis ssp. ludoviciana Dur. G&M and A.

barbata Pott. ex Link, occur in Australia. Only the first two species are documented as a

major weed of cereal crops while /,. barbata is found mainly in vacant, undisturbed land and

in pastures (V/halley and Burfit t972, Paterson t976, Ma¡tin et al. 1988). The detailed

descriptions for those two species of wild oat are given in Table 1 .10.

Table 1.10 Diagnostic cha¡acters of the two main wild oat weeds (Thomas and Jones 1976)

Character A. fatuaL. A. sterilis ssp. ludoviciana Dw. G&M

Chromosomenumber Hexaploid(2n=6x42) Hexaploid(2n=6x=42)

Season of germination

Awn on third seed

Abscission sca¡

Shedding of ripe seed

Spring

Present

Present at the base of

every seed

All seeds of spikelet fall

separately

V/inter

Absent

At the base of first seed only (second

and third seeds end in stalk)

Seeds of spikelet fall together as a

unit and force necessary to separate

the grains benveen a spikelet

31

1.9.1 Seed Dormancy and Germination

Dormancy in wild oat seeds is a major problem in controlling this weed throughout the

world. Seeds of. A. fatua and A. sterilis are dormant for at least two months after harvest

(Thurston 1959). About 957o of A. faru¿ seeds are dormant at the time of shedding whereas

orty 50Vo of A. sterilis seeds are dormant at this time (Chancellor 1976). ln A. steríI¡s, the

fust seed in the spikelet is usually non-dormant and germinates within a few weeks of being

shed (Thurston 1951, Gwynne and Murray 1985). Dorrrancy of wild oats is due to the

impermeability of seed coat to oxygen (Naylor and Simpson 1961) and is more prolonged

under cool, moist conditions (Gwynne and Murray 1985).

There are several treatments for breaking dormancy in wild oat seeds:

(a) Scarification. The hull (lemma and palea) of wild oat seed retards water uptake and

contains germination inhibitors. Thus, the germination of dormant seeds can be brought

about by removing their hulls @lack 1959). It was reported that over 9O7o germination of

wild oat seeds can be obtained by pricking the seeds (Whaltey and Burfitt 1972).

(b) Fertilizers. Watkins (1971) found that the number of germinating wild oats increased as

the rate of the applied nitrogen increased. Adkins et at. (1984a and 1984b) demonstrated

that sodium nitrate and nitrite (50-100 mM) induced germination in a partly after-rþened

dormanthneof A.fatuaby promotingoxygenuptakeasmuch asTOVo priortothefustsigns

of germination. Nitrogen fertilizers significantly enhanced seedling emergence of partially-

dormant seeds of wild oat (Agenbag and DeVillier 1989).

(c) Chemical treatments. Ethanol at 50-200 mM overcame dormancy in freshly han¡ested

and partially after-ripened caryopses of A. fatua after five days of incubation (Adkins et al.

1984c). Cairns and DeVilliers (1986) broke A.fatua seed dormancy by treating the air-dry

seeds with ammonia for 48 h. Ammonia caused an increase in leakage of electrolytes from

treated seeds, indicating an increase in permeability of their seed coats. Giberellic acid (GA)

32

at > 0.1 ppm was sufficient to induce germination of both newly matured and old embryos

of A. fatua (Naylor and Simpson 1961).

Wild oat seed must pass through a wa¡m imbibed condition in the soil before chilling can

stimulate its germination. Quail and Carter (1968) proposed that the optimum temperature

for germination of .4. sterilis seeds was between 0-10'C whereas A. fatua required 20-30'C.

However, Peters (1986) found that germination of dormant A. fatw seeds occurred when

removed from the soil and placed at 15'C. Paterson et al. (1976) found that maximum

germination of A. fatua occurred in 11-23'C range and temperatures above 23'C reduced

germination.

Wild oat seeds do not germinate under water-logged conditions. No germination was

observed for seeds following submersion for 6, 8 or 42 months in fresh water (Brun 1965).

In addition, seeds being in water-logged soil for only t to 2 months could not survive

(Lewis 196l).

1.9.2 Seed Bank of Wild Oat

The seed bank consists of seeds of different ages, some of which a¡e dormant with some

being exposed to favourable and some to unfavourable conditions for germination (Wilson

1988). Seed bank life of wild oat varies depending on soil type, soil depth and cultivation.

For example, under grassland or in undisturbed a¡able soil, the survival period of A. fatua

seeds is up to 8-10 years and about half of that for A. sterilis (Gwynne and Murray 19S5).

Longevity of buried wild oat seeds can range from 2 to 6 years (Thurston 1966) or 7 to 9

years (Chancellor 1976, Miller and Nalewaja 1990) depending on the local environmental

and cultural practices. Injection of wild oat seed into the seed bank can be reduced by straw

burning (Wilson and Cussans 1975) or lost by natural deterioration or by predation

(Marshall and Jain 1970, V/ilson 1972).

33

1.9.3 Control of Wild Oat

Wild oats are highly adaptable and produce large number of seeds which ænd to persist from

year to year. The control of wild oat not only reduces the yield loss in the current year but

also affects future infestation by reducing the numbers of seeds in the soil (seed bank)

carried over to the next year. The earliest herbicides used for wild oat control were non-

selective graminicides such as dalapon, propham, chloropham (Hutson and Roberts 1987).

In the late 1950s, diallate and ba¡ban were inEoduced and these herbicides were selective for

the control of wild oat in cereals. However, all varieties of oats were susceptible whereas

wheat and barley were relatively resistant. Triallate proved to be better for wheat and barley

than diallate (both species being more tolerant to triallate) and the use of diallate for wild oat

control was eventually discontinued (Hutson and Roberts 1987). In the late 1970s,

diclofop-methyl was fustly used as a post-emergence herbicide for the control of wild oat in

wheat and barley (Friesen et al. 1976). Since then, many more APP and CHD herbicides

have become available for the control of wild oat and other grass weeds. Of these, diclofop-

methyl, fenoxaprop-ethyl and tralkoxydim give selective control in grass crops such as

wheat and barley. The other APPs (such as fluazifop-butyl, haloxyfop-ethoxyethyl,

quizalofop-ethyl) and CHDs (such as sethoxydim, cycloxydim and alloxydim) are widely

used to control wild oat in dicot crops (Worthing and Hance 1991). In addition,

fenoxaprop-etþI, flamaprop-methyl or tralkoxydim can be used during stem elongation and

booting stage of wild oat to reduce the seed production (Medd et al. 1992).

1.10 Herbicide Resistance in Wild Oat

The occurrence of resistance to ACCase-inhibiting herbicides in wild oat is of considerable

importance as wild oat is one of the most serious weeds of cereal cropping in many

countries. Also there are few altemative selective chemistries for wild oat control. Resistant

wild oat has now appeared in N. America, Europe and Australia. In western Canada, there

are 15 populations of Avena fatua that have developed high-level resistance to triallate

(O'Donavan et al. 1994). Most triallate-resistant populations are also resistant to

34

difenzoquat. However, the mechanisms conferring resistance to triallaæ and difenzoquat in

those populations have not yet been elucidated. In addition, wild oat resistant to APP and

CHD herbicides in N. America developed following 6 to l0 herbicide applications (Heap et

al. 1993). Among these populations, A. fatua biotype UM-l has been examined for

possible mechanisms conferring resistance to APP herbicides. This study showed that this

biotype was able to repolarize the plasma membrane potential following depolarization by

diclofop acid and fenoxaprop acid (Devine et al. 1993) and has a resistant ACCase (Marles

and Devine, unpublished).

In Australia, resistance to diclofop-methyl tnAvenafatwtwas first evident in 1990 (T. Pþr

unpublished). This biotype was subsequently shown low-level resistance to diclofop-

methyl but was not resistant to other APP and CHD herbicides (Mansooji et al. 1992). The

mechanism of resistance in this biotype has not been elucidated but this biotype does contain

a sensitive target ACCase (Boutsalis et al. 1990). At present, resistance to APP and CHD

herbicides has been confirmed in populations of A. fatua and A. sterilis from New South

Wales, South Australia, Victoria and S/estern Australia (Mansooji et aI. 1992). At least 20

populations of Avena spp. have developed resistance to APP and/or CHD herbicides

(Mansooji 1994).

1.11 Objectives of this Project

The aim of this project is to elucidate the mechanism or mechanisms by which biotypes of

wild oat (Avena fatua and Avena sterilis) in Australia have become resistant to ACCase-

inhibiting herbicides. Ten putative resistant biotypes were examined with varying levels of

resistance. Four biotypes which showed different patterns of resistance were examined in

more detail.

35

Chapter 2

Materials and Methods

In this chapter, general methods will be described. Some modifications to these methods

and specific techniques are detailed in the relevant chapters.

2.1 Plant Material

Wild oat (4. fatw L. andÁ. sterilis L.) biotypes were collected from the following sites:

SAS I Collected in 1989 from a barley field near Bordertown, South Australia.

NAS 4 Collected in 1990 from a wheat field near Greenthorpe, New South V/ales.

NAF 6 Collected in 1991 from a wheat field near Orange, New South Wales.

NAS 14 Collected in 1989 from a wheat field nea¡ Henty, New South V/ales.

NAF 8 Collected in 1991 from a wheat field near Junee, New South rWales.

NAS 2 Collected in 1990 from a wheat field nea¡ Greenthorpe, New South rùfales.

NAS 6 Collected in 1991 from a wheat trial site nea¡ Tamworttr, New South Wales.

SAF 14 Collected in 1990 from a wheat field near Bordertown, South Australia.

SAF 34 Collected in 1991 from a wheat field near Mintaro, South Australia.

V/AF 12 Collected in 1991 from a wheat field nea¡ Coomberdale,'Western Austalia.

SAS 2 Collected from unsprayed area adjacent to biotype SAS 1.

SAF 19 Collected from unsprayed pasture area near Bordertown, South AusEalia

36

A coding system was used to classify the wild oat seed samples. The fust letter is taken

from the state of Australia in which the biotlpe was collected. The following letters are the

genus and species of wild oat. The number is the order of collection from each state. For

exarrple, bioqpe SAS I reprcsents wild oat from South Australia, Avena sterílis, collection

No. 1. Biotype \MAF 12 represents wild oat from'Western Australia, Avena fatua,

collection No. 12. The full details of histories of herbicide applications of all resistant

biotypes are shown in Appendix l. The number of herbicide applications in four

representative resistant biotypes of wild oat are shown in Table 2.1. Biotypes SAS 2 and

SAF 19 a¡e known to be susceptible to all ACCase-inhibiting herbicides.

Other species used:-

Oat (Avena sativa L.) cultivar Echidna

Wheat (Triticum aestivum L.) cultivar Halberd

Table 2.1 Number of herbicide applications prior to the development of herbicide

resistance in four resistant biotypes of wild oats

Biotype APP herbicides CHD herbicides Other herbicides

SAS 1 3 diclofop, 2fluaztfop

NAF 6 2 diclofop

NAS 4 5 diclofop, 1 fluazifop, 2 tralkoxydim

I haloxyfop, 1 fenoxaprop

0

0

0

0

0

3 trifluralin

NAS 14 4 diclofop, 2f\razifop 2 simazine

37

2.2 Plant Culture

2.2.1 Germination

Seeds of wild oats were dehusked and gerurinated in containers containing O.6Vo agar and

kept at 5"C for 7 d then transferred to a seed germinator at 15'C 12 h light/dark cycle.

Seedlings were transplanted at one-leaf stage approximately 5 d after conrmencement of

germination. Plants of even size and vigour were selected. V/heat and oat seeds were

incubated for 48 h at 5'C prior to transfer to l5'C. These seedlings were transplanted one

day later than wild oat seedlings so that plants reached the two-leaf stage at the same time.

2.2.2 Pot Culture

Germinated seedlings were transplanted into sterile potting mix based on sand and peat (1:1)

in 15-cm pots. For dose response experiments, plants were grown outdoors during autumn

and winter. For all other experiments, plants were grown in the growth chamber with 20"C,

14 h, 330 p mol photons m-2 s-1 light period/16'C, 10 h dark period.

2.2.3 Hydroponic Culture

Seedlings germinated as in 2.2.1 were transferred to I L rectangular plastic containers

containing 500 mL of nutrient solution (Table 2.2). Containers were $rrapped with

aluminium foil to prevent light reaching the roots. Seedlings were supported by black poly

propylene beads (400 ml/container) then transferred to the growth room with the same

conditions as in section2.2.2. The containers were topped up with 100 mL water every day

to replace evapo-transpirational losses in the first week. Afterwards, half strength nutrient

solution was added until the end of the experiments.

38

Table2.2 Hydroponic nutrient solution (Hoagland and Arnon 1938)

Source of nutrient Final concenEation (¡tM)

Ca(NOg).4HzO

CaSO¿.2HzO

KHzPO¿

K2SOa

KNO¡

MgSO+.7HzO

FeSOa.TH2O

EDTA Na2.2H2O

H3B03

CuSOa.5H2O

NazMoO¿.2HzO 0.25

MnCl2.4H2O 4.60

7-nSO+.7HzO 0.38

1670

800

500

400

t670

1000

72

&

23

0.16

39

2.3 l)ose Response Experiments

Plants were sprayed at the two to three-leaf stage, the most susceptible stage under

agricultural conditions. Herbicides assigned as treatments (Table 2.3) werc applied as

commercial fonnulations with O.2Vo (vlv) non-ionic surfactant (Agral600@) in a laboratory

spray cabinet and delivered via two 110' flat fan hydraulic nozzles suspended 40 cm above

the plant axils. At a pressure of 25O kPa and a boom speed of 1 m s-1, the sprayer output

was equivalent to 113 L ha-l.

Six plants of each biotype were grown in each pot. rü/ithin each herbicide treatment, pots

\ilere aranged as a completely randomized block design with six replicates. Plants were

harvested 4 weeks after spraying which allowed surviving plants to make noticeable growth.

Plants which were green, had live meristems and had increased in size beyond the two leaf-

stage were classified as live plants and were harvested at the ground level. Plant shoots

were dried at 80"C for 48 h prior to dry weight measurement.

Table 2.3 Commercial formulations of aryloxyphenoxypropionates and cyclohexanedione

herbicides registered for use in Australia were used as treatments for dose response

experiments.

Common name Trade name Recommendedrate Manufacturer

(g ai ha-l)

diclofopmethyl Hoegrass@

fluazifopbutyl Fusilade@

haloxyfop-ethoxyethyl Verdict@

cycloxydim Focus@

sethoxydim Sertin@

Hoechst AustraliaLtd.

104 ICI AustraliaLtd.

52 Dow Elancoltd.

40 BASF

100 Schering Australia Ltd.

563

Grasp@talkoxydim 150 Hoechst Australia Ltd.

40

2.4 ACCase Extraction and Assay

The shoot meristematic region of wo to th¡ee-leaf plants was harvested and 3 g of this tissue

was ground in a chilled mortar with 5 rnl- of Buffer A (Table 2.4). All subsequent

operations were performed at 4'C. The brei was centrifuged for 10 min at 27,0009. The

supernatant (2.7 mL) was centrifuged for a few minutes at 10,0009 to remove non-protein

solid material and 2.5 rnL was loaded onto a column containing Sephadex G-25 (PD-10

column, Phannacia, Uppsala, Sweden) previously equilibrated with Buffer B (Table 2.4).

The void volume was eluted with 3.5 mT of Buffer B. The eluate was stored on ice and

assayed immediately for ACCase activity.

To obtain partially purified ACCase enzyme, plant tissue was ground in Buffer A and

centrifuged as mentioned above. The supernatant slowly brought to 1.0Vo (NH4)2SO4

saturation, kept on ice for l0 min and then centrifuged for 20 min at 27,OC[lg at 4'C. The

precipitate was discarded, the supernatant was broughtto 4O7o saturation with (NHa)zSO¿

and centrifuged as above. The supernatant was disca¡ded and the pellet resuspended in

2.7 nL Buffer B and 2.5 nL applied to a column of Sephadex G-25 previously equilibrated

with Buffer B. The eluate (3.5 mL) was stored on ice and immediately assayed for ACCase

activity.

ACCase activity was assayed by following the incorporation of l4C from NaHl4CO, into

acid-stable products at 30'C. Assays, final volume 200ltL, contained Buffer C (Table 2.4),

30 pL protein extract and herbicides or HrO as required. Herbicide stock solutions (1 mM

diclofop acid, fluazifop acid, haloxyfop acid, sethoxydim and tralkoxydim) were dissolved

in 50Vo acetone and made up as serial dilutions in tÙVo acetone. Assay duration was 3 min

and the reaction was terminated by the addition of 30 pL glacial acetic acid. A piece of

glass-micro fibre filter (GF-A; Whatman, Maidstone, Kent, UK) was added to each reaction

tube and the liquid evaporated for 5 h under a stream of air at 70-80'C. Scintillation fluid

(3 mL of toluene:Triton X-100 f2:L, vlvf containing O.4Vo wlv PPO, and O.OlVo (w/v)

POPOP was added. Radioactivity was determined by LSS.

4l

Table 2.4 ExEaction and assay buffers in ACCase assay

BufferA BufferB Buffer C

100 mI\d Tris (pH 8.0)

I nIVIEDTA

2 mM isoascorbic acid

I mMPMSF

20mMDTT

0.57o (wlv) PVP-40

0.5Vo (wlv) insoluble PVP

lOVo (vlv) glycerol

50 mN,f Tricine (pH 8.0)

2.5 mM MgClr.6HrO

50 mM KCI

I mMDTT

50 mM Tricine (pH 8.0)

2.5 mM MgClr.6HrO

50 mM KCI

I TnIvIDTT

I mMATP

0.1 mM acetyl-CoA

10 mM NaH14CO3

containing 0.5 pCi of 14C

2.5 Uptake and Translocation of llaC]Diclofop-Methyl

Resistant (NAF 6, NAS 4 and NAS 14) and susceptible (SAF 19 and SAS 2) plants were

grown hydroponically as in section 2.2.3. At the two leaf-stage, plants were treated with I

pL of 5 mM [l4C] ¿ictofop-methyl (4.5 mM llaC]diclofop-methyl with specific activity of

592 Mbq ml--l dissolved in 0.5 mM commercial product, Hoegrass) onto the leaf æril. Ten

plants of NAS 14 (one plant/replicate) were harvested L,3, 6, 12,24,48 and 72 h after

treatment. For NAF 6 and NAS 4 biotypes, five plants were used in each replicate.

Unabsorbed radioactivity was removed by washing the plant in 5 mL of ZOVo (v/v) methanol

plus 0.17o (v/v) Triton X-100 (contained in 20 ml-scintillation vials). Radioactivity in the

washes was measured by LSS. The plant was sectioned into four fractions:- shoot

42

meristematic region (l cm above the initiation root zone), stem (from meristematic zone up to

2 cm above the leaf axil), leaves and roots (Figure 2.1) except for biotype NAF 6 where

plants were sectioned into the shoot meristematic region and the rest.

second leaffirst leaf

stem

shoot meristematic region

root

Figure 2.1 Plant sections in translocation experiments when harvested

Plant parts were combusted in a Biological Sample Oxidizer OX-600 (R.J. Harvey

Corporation, Milldale, N.Y.) and the l4COzreleased was trapped in 14 mL of scintillation

fluid lCarbosorb@:Permafluor@E+, l:1 (v/v), Canberra-Packard, Tampa, FL, USA] and

laC quantified by LSS.

2.6 Metabolism of Herbicides in Plants

2.6.1 [14C]nictofop-Methyl

Plants \ilere grown in the soil and treated as for the uptake and translocation experiments.

Treated plants were rinsed, blotted dry and frozen in liquid nitrogen. Frozen tissue was

pulverised in liquid nitrogen using a mortar and pestle. Tissue powder was extacted with 5

43

mI . of 807o methanol at -20'C. The brei was centrifrrged at 15,0009 for 20 min at 2'C. The

pellet was washed twice in 2 mI of 807o methanol and recentriñ¡ged at 15,0009 for 10 min

at 4'C. The supenratants were pooled and stored at -20'C for 6 h, then centrifuged for 3

min to remove any precipitate before evaporation. The pooled supernatant was reduced in

volume to near dryness under vacuum and resuspended in enough 507o methanol to provide

radioactivity levels suitable for the radioactivity detector. Samples were filtered through a

O.22 ¡tmNylon filter prior to injection onto the HPLC column.

Radiolabelled metabolites were separated by HPLC equipped with a reverse-phase C18

column (250 mm long, 4.6 mm i.d.; Brownlee Labs ODS-5 Spherisorb; Applied

Biosystems, Lincoln City, California, USA). Radioactivity was detected with an inline

radioactivity detector (Flo-one 1 Beta Model A-140, Radiomatic Instmments and Chemical,

Canberra-Packard, Tampa, FL, USA). Solvents used were 89:10:1 (v/v)

H2O:acetonitile:acetic acid (solvent A) and 90:9:l (v/v) acetoninile:H2O:acetic acid (solvent

B). Elution conditions were a lO-rnin linear gradient from 30 to 35Vo solvent B, followed

by a 12-min linear gradient from 35 to 50Vo solvent B, followed by a 3-min linea¡ gradient

from5OVo to lOÙVo solvent B. The column was then eluted with 1007o solvent B for l0 min

and re-equilibrated with307o solvent B for l0 min prior to injection of next sample. At all

times, the combined flow rate of solvents A and B was 1.5 mL min-l.

2.6.2 ¡4-l4ClSethoxydim

A single droplet of I pL of I mM ¡4-laClsethoxydim (a33 Mbq mmol-l) dissolved in

water:methanol:Triton X-l0O:toluene (96:16.4:L.7:I, by vol.) was placed on the leaf axil of

the resistant NAF 6 and susceptible SAS 2 plants at the two-leaf stage. Plants were

harvested after varying periods of exposure (1,3, 6, t2, 24 and 48 h). Tissue was rinsed

and exfiacted with the same methods as for diclofop metabolism. Six plants were pooled at

each time point. Extracted [laC]sethoxydim and metabolites were separated by HPLC.

Solvents A and B as described in section 2.6.1were used. The elution conditions were a 5-

min linear gradient from 19 to 607o solvent B, followed by a l0-min linear gradient from 60

to 85Vo solvent B, followed by a l0-min linear gradient from 85 to LÙOVo solvent B. Then

4

the column was eluted with IOOEI solvent B for 10 min and re-equilibrated with l9%o

solvent B for l0 min before injection of the next sample. Flow rate was 1.2 mL min-I.

Recovery of radiolabel was deterurined by adding the radioactivity in the wash to the

radioactivity in the extract and dividing by the radioactivity applied.

2.7 Measurement of Plasma Membrane Potentials

Etiolated coleoptiles were collected 3 to 5 d after gerrrination on0.67o agar. The top 3 mm

of the coleoptiles were decapitated and the leaves within the coleoptiles were removed. The

epidermis was peeled using fine forceps under Higinbotham's solution (Table 2.5). A

peeled coleoptile was placed in a plexiglass chamber attached to the stage of a microscope.

During the impalement of a cell, the coleoptile was continuously inigated with

Higinbotharn's solution, pH adjusted to 5.8, at a flow rate of 2.5 mL min-I.

Table 2.5 Higinbotham's solution (Higinbotham et al. 1970).

Source of nutrient Concentation (¡tM)

KCI

NaII2POa

Na2HPOa

Ca(NO¡)z.4HzO

MgSOa.TH2O

Micropipettes \ilere pulled from borosilicate glass capillaries, 0.86 mm i.d. with an intcrnal

filament (GCl50F-10, Clark Electrochemical Instn¡ments, UK) to a tip diameter of about I

pm. The electrodes, with resistances of between 2 to L5 Mll, were mounted on a head-

1000

9U

48

1000

250

45

stage and holder with a Ag-AgCl electrode. The reference electrode, filled with lM KCI in

2Vo (wlv) agar, was placed close to the coleoptile in the bathing solution, but downstreart

from the measuring electrode. Electrode offsets were nulled before each impalement. The

membrane potential was measured by inserting a glass microelectrode into a single cell by a

micromanipulator. Membrane potentials were monitored using a Neuroprobe Amplifier

model 1600 (A-M Systems, USA) connected to a recorder. The initial membrane potential

was allowed to stabilize for 5 min before herbicide application. The basal solution was

replaced with 50 pM diclofop acid in Higinbotham's solution with the final concentration of

l%o (vlv) acetone. After 15 min exposure, the Eeafinent solution was replaced with the basal

solution. Ten coleoptiles from resistant and susceptible seedlings were used and the data

presented a typical run of each biotype.

46

Chapter 3

Spectrum of Resistance to Aryloxyphenoxypropionate and

Cyclohexanedione Herbicides in WiId Oat Biotypes:

Whole Ptant and Target Site Inhibition Studies

3.1 Introduction

The aryloxyphenoxypropionate (APP) and cyclohexanedione (CHD) herbicides are post-

emergent herbicides used to control annual and perennial grass weeds in dicot and some

monocot crops. These are structurally different chemical groups with the same mode of

action, inhibition of ACCase. Weeds can develop resistance to these herbicides after

exposure to these compounds for as few as three consecutive years (Tardif et al. 1993). In

the past 15 years diclofop-methyl and subsequent herbicides which target the plastidic

ACCase enzyme have become widely used for wild oat control in world agriculture.

V/idespread and persistent use of ACCase-inhibiting herbicides has resulted in the recent

appearance of resistant populations of wild oats in AusEalia and North America (Mansooji et

al. L992, Devine et al. 1992, Heap et al. 1993). In Australia, there a¡e increasing numbers

of biotypes of two species of wild oats (Avena fatua L. and A. sterilis L.) exhibiting

resistance to ACCase-inhibiting herbicides. In addition, biotypes of several other grass

weeds are resistant to APP and CHD herbicides such as Alopecurus rnyosuroides (Moss

1990), Lolium multiþrum (Gronwald e¡ aL. 1992), L. rigidum (Heap and Knight 1982,

Powles and Howat 1990) and Setaria viridis (Marles et al. 1993).

Putative resistant biotypes of wild oat were investigated for their responses to ACCase-

inhibiting herbicides. It is suspected that there would be some va¡iations in response to

ACCase-inhibiting herbicides as a result of different herbicide treaünents in the field. The

response of ten putative biotypes of wild oat were tested with the two chemical groups, APP

and CHD herbicides. Herbicides examined were diclofop-methyl, fluazifop-butyl,

47

haloxyfop-ethoxyethyl, sethoxydim and tralkoxydim. The ability of these herbicides to

inhibit ACCase activity invitro was also determined.

3.2 Whole Plant Assessment of Resistance

(a) Response to APP Herbicides

Ten putative resistant wild oat biotypes with varying histories of exposure to APP and CHD

herbicides (Appendix l) were exaurined for resistance to APP and CHD herbicides. These

biotypes were found to express very different patterns of resistance to APP and CHD

herbicides. Detailed dose response experiments were conducted to the ACCase-inhibiting

herbicides as shown in Figure 3.1. At the recoûrmended rate for wild oat control in

Australia, 563 g diclofop-methyl ha-1, the susceptible biotypes exhibited no survival

whereas three resistant biotypes, NAS 14, NAF 6 and NAS 4 exhibited 60, 100 and l00%o

survival, respectively (Figure 3.lA). Biotype NAS 14 and the susceptible biotype were

easily controlled by the recommended rate of fluazifop-butyl (104 g ha:l) whereas biotypes

NAS 4 and NAF 6 were resistant at this rate exhibiting90%o andTOVo survival, respectively

(Figure 3.lB). The three resistant biotypes were also resistant to haloxyfop-ethoxyethyl

with3OVo,TOVo and 907o survival at the recommended rate of 52 gha-l for biotypes

NAS 14, NAF 6 and NAS 4, respectively whereas the susceptible biotype was controlled

by this rate (Figure 3.lC).

48

1finO 2finO 3finO

100

00 200 400 600 800

100

Herbicide rate (g ai h"-1)

Figure 3.14.C Survival of susceptible SAF 19 (O) and putative resistant NAS 14 (A),

NAF 6 (O) and NAS 4 (I) wild oat biotypes 30 d after exposure to diclofop-methyl (A),

fluazifop-butyl (B) and haloxyfop-ethoxyethyl (C). A¡rows indicate rates recommended for

wild oat control. Each point is the mean + standard error of six replicates.

100

s80ã60a-bã40u)

20

s80ã60.-340

ct)20

s80ã60'E

40(n

20

00

04003002001000

Adiclofop-methyl

Bfluazifop-butyl

ü

chaloxyfop-ethoxyethyl

49

The susceptible wild oat biotypes SAF 19 and SAS 2 werc easily controlled by the

reconrmended rates of all APP and CHD herbicides tested. The rates of APP herbicides

required to kill 50Vo of. the individuals of the wild oat populations are given in Tables 3.1 to

3.3. The level of resistance to the herbicides among the resistant populations was classified

into three $oups by using the resistanUsusceptible ratio of LD5g values as follows:

Level of resistance APP herbicides CHD herbicides

>50

Moderate >10

Low

Sensitive <2

Diclofop-methyl is widely used to selectively control grass weeds in wheat. rrl/heat is

resistant to diclofop-methyl due to the ability to rapidly detoxify the herbicide (Shimabukr¡ro

et al. 1979). However, resistant biotypes SAS I and NAS 4 were less sensitive to diclofop-

methyl than wheat cv. Halberd. At the rate of 16 kg diclofop-methyl ha-l they had l00Vo

survival while in this wheat cultivar some mortality was observed. The biotypes SAS l,

NAS 4, NAF 8, SAF 14 and NAS 2 were highly resistant to diclofop-methyl with LD59

ratios of. >2L2, >192, 190, 159, and 113, respectively (Table 3.1). Biotype NAF 6 was

moderately resistant to diclofop-methyl with an LDso ratio of 45. The other four resistant

biotypes, NAS 14, ÌWAF 12, NAS 6 and SAF 34, expressed either a low level resistance

to diclofop-methyl; 3.2 and 2.3 times, respectively for biotypes NAS L4 and WAF 12, or

showed no resistance <2 fold (Table 3.1).

>20

>4

>2

<2

>2

High

50

Table 3.1 Amounts of diclofop-methyl in g ai ha-I, required to kill 50Vo ofpopulations

of twelve biotypes of wild oat and wheat cv. Halberd. The recommended rate of

diclofo¡methyl for wild oat control in Austalia is 563 g ai hrl.

Biotype LDso G ai ha:t) Resistant/Susceptible

wheat cv. Halberd

Resistant biotvoes

SAS I*

NAS 4

NAF 8

SAF 14

NAS 2

NAF 6

NAS 14

rwAF 12

NAS 6

SAF 34

Susceotible biowoes

SAS 2*

SAF 19

>16,000

>30,000

>32,000

31,660

26,550

18,890

7,490

392

227

t4t

167

>2t2

>192

190

159

113

45

534

286

3.2

2.3

t.7

1.4

* Data from Mansooji et al. 1992

51

Growth response of biotypes having moderate and high levels of resistance to APP

herbicides were not the same for all the herbicide rates. When exposed to the rccommended

rate, plants looked unaffected. At very high rates, from 8,000 to 32,0O0 g ha-l diclofop-

methyl, growth was obviously inhibited and transient yellowing or leaf burn appeared on the

older leaves. However, emerging new leaves appeared normal. In contrast, biotypes with a

low level of resistance like biotype NAS 14 were badly affected by application rates as low

as 2,000 g ha-l diclofop-methyl. Newly emerging leaves of these biotypes showed severe

chlorosis.

The biotypes SAS l, NAS 2, SAF 14 and NAF 8 were highly resistant to fluazifop-butyl

and had LDso ratios of >1,000, >67,>67 and 53 times that of the susceptible biotypes,

respectively (Table 3.2). Moderate resistance to fluazifop-butyl was observed in the

biotypes NAS 4, NAF 6 and NAS 6 with LDso ratios ranging from 7.7 to 37. The

remaining biotypes 1WAF 12, SAF 34 and NAS 14, showed low level resistance to

fluazifop-butyl with LD59 ratios of 4.3,3.3 and 2.7 times that of the susceptible,

respectively (Table 3.2). These last three biotypes would be controlled by the recommended

rate of fluazifop-butyl.

Biotype SAS 1 was the only biotype to show high level resistance to haloxyfop-ethoxyethyl

(Table 3.3). Biotypes NAS 2, NAF 6 and NAS 4 were moderately resistant to haloxyfop-

ethoxyethyl with LDso ratios of 25,10 and 10 times that of the susceptible, respectively.

The remaining biotypes exhibited low-level resistance to haloxyfop-ethoxyethyl with LDso,

ranging from 8.5 to 4 times that of the susceptible.

52

Table 3.2 Amount of fluazifopbutyl in g ai ha-I, requi¡ed to kill SOVo of populations

of twelve biotypes of wild oat. The recoûlmended rate of fluazifop-butyl for wild oat

conEol in Australia is 106 g ai ha-I.

Biotype LDso G ai ha-t) Resistant/Susceptible

Resistant biotvoes

SAS I*

NAS 2

SAF 14

NAF 8

NAS 4

NAF 6

NAS 6

IüVAF 12

SAF 34

NAS 14

Susceotible biowoes

SAS 2*

SAF 19

>16960

>800

>800

uo

48

t28

92

52

40

25

>1000

>66.7

>66.7

53.3

37.3

t0.7

7.7

4.3

3.3

2.t

<16

t2

* Data fromMansooji et al.1992

53

Table 3.3 Amount of haloxyfop-ethoxyethyl in g ai ha-I, required to kilt SOVo of

populations of twelve biotypes of wild oat. The recommended rate of haloxyfop-

ethoxyethyl for wild oat control is 52 g ai ha-I.

Biotype LD59 G ai tra-t¡ ResistanUsusceptible

Resistant biotvoes

SAS 1* 177

NAS 2

NAF 6 10

NAS 4 t0

NAF 8 8.5

SAF 14

NAS ó 5.5

NAS 14

WAF 12 4.8

SAF 34

SAS 2* l0

SAF 19

* Data from Mansooji et al. 1992

5

1768

200

80

80

68

56

4

40

38

32

25

7

4

8

54

O) Response to CHD Herbicides

None of the putative resistant biotypes tested were found to be highly resistant to

tralkoxyrlim (Table 3.4). Biotype NAF 6 exhibited 507o survival at the recommended rate of

100 g ha-l tralkoxydim whereas biotypes NAS 4 and NAS 14 and the susceptible biotype

SAF 19 were controlled by this rate of herbicide (Figure 3.2). Biotypes NAF 6 and

V/AF 12 were moderately resistant to tralkoxydim with LDso. 5 and 3.6 times that of the

susceptible, respectively (Table 3.4). Biotypes NAS 14, NAF 8, NAS 2 and NAS 6 were

slightly resistant to tralkoxydim but were completely contolled by the recommended rate of

100 g ¡¿-l (Table 3.4). The other biotypes were not resistant with LD5g ratios of less

than 2.

Only a single biotype of wild oat, NAF 6, showed substantial survival at the recommended

rate of 100 g ha-l tralkoxydim. This biotype was also resistant to the other CHD herbicides,

cycloxydim and sethoxydim with 90 and lÙ07o survival, respectively, at the recommended

rates (Figure 3.34-B). Rates of CHD herbicides, tralkoxydim, cycloxydim, and

sethoxydim required to kill 5OVo of biotype NAF 6 were 5 to 7 times those required to kill

SOVI of the susceptible population (Table 3.4 and Figure 3.34-8).

55

Table 3.4 Amount of tralkoxydim, in g ai ha:I, required to kill SOVo of populations of

twelve biotypes of wild oat. The recommended rate of tralkoxydim for wild oat contol in

Australia is 100 g ai ha-l.

Biotype LD5s (g ai ha-t) ResistanUsusceptible

Resistant biotvoes

NAF 6

rùrAF 12

NAS 14

NAF 8

NAS 2

NAS 6

NAS 4

SAS 1*

SAF 14

SAF 34

Susceotible biowoes

SAS 2*

SAF 19

5100

72

52

50

50

50

36

34

28

28

3.6

2.6

2.5

2.5

2.5

1.8

2.6

1.4

1.4

l3

20

* Data from Mansooji et aL.1992

56

100

80

èa

ã60.-¡i

ã40

20

00 100 300 400

Tralkoxydim (g ai ha'1)

Figure 3.2 Survival of susceptible SAF 19 (O) and resistant NAS 14 (A), NAF 6 (O)

and NAS 4 (f) wild oat biotypes 30 d after exposure to tralkoxydim, Arrow indicates the

rate recommended for wild oat control. Point symbols a¡e the mean t standa¡d error of the

six replicate experiments.

240

57

100 A I Bsethoxydim

20

0 40 80 t20 160 0 100 2t0 300 400

Herbicide rate (g ai h"-1)

Figure 3.34-B Survival of susceptible SAS 2 (O) and resistant NAF 6 (O) wild oat

biotypes following treatment with cycloxydim (A) and sethoxydim (B). Point symbols are

the mean * standard error of six replicate experiments. Arrows indicate the recommended

rate of herbicides for use in AusEalia.

EO

èa

:60G

.l

i40ct)

0

58

3.3 Inhibition of ACCase Activity by Herbicides

Partially purified ACCase enzyme extracted from resistant and susceptible plants was

assayed (as described in Chapter 2) to determine the effect of ACCase-inhiþi¡i¡g heùicides.

The level of resistance to the herbicides among the resistant populations w¿ts classified as in

the whole plant level (section 3.2)by using the resistant and susceptible ratio of I5g values.

Inhibition of ACCase by diclofop acid and tralkoxydim a¡e shown in Figures 3.4 to 3.7 for

four representative resistant biotypes compared to susceptible biotypes. These figures show

that ACCase extracted from the susceptible biotypes SAS 2 and SAF 19 was sensitive to

inhibition by diclofop acid with an Iso of less than 1.1 UM. Similarly, the enzyme from the

susceptible biotypes was sensitive to tralkoxydim with an I5g of about I to 2 ¡tM. ACCase

from four resistant biotypes displayed varying responses to these two herbicides. ACCase

from biotype SAS I was highly resistant to diclofop acid with an Iso >10 pM (Figure 3.4A)

but was only moderately resistant to tralkoxydim (Figure 3.48). ACCase from biotype

NAS 4 showed moderate resistance to diclofop acid, I5g -6 ttM (Figure 3.54), and only

slight resistance to tralkoxydim (Figure 3.58). ACCase from biotype NAF 6 showed

moderate resistance to both diclofop acid and tralkoxydim with 15¡ of -4 ¡rM for diclofop

acid and 4 lrM for tralkoxydim (Figure 3.6A-8). ACCase from biotype NAS 14 showed

identical inhibition kinetics to the susceptible biotype for both diclofop acid and tralkoxydim

(Figure 3.74-B).

ACCase was extracted from all biotypes and herbicide inhibition curves were generated for

diclofop acid, fluazifop acid, haloxyfop acid, sethoxydim and tralkoxydim. The I5gs and

R/S ratios for these herbicides for each biotype are shown in Tables 3.5 to 3.9. The

ACCase from biotype SAS I was the most resistant to diclofop acid. ACCase from biotypes

NAS 4, NAF 8 and SAF 14 was also strongly resistant to this herbicide. ACCase from

biotypes NAF 6 and NAS 2 was moderately resistant, whereas that from biotypes NAS 6,

NAS 14, WAF 12 and SAF 34 was not resistant (Table 3.5).

59

The ACCase from biotype NAS 2 was most resistant to fluazifop acid (Table 3.6). ACCase

from biotypes SAF 14, NAF 8, NAS 4, NAS 6 and SAS I was also strongly resistant,

whereas the ACCase from biotypes NAF 6, WAF 12 and SAF 34 showed low-level

resistance to fluazifop acid. ACCase from biorype NAS 14 was susceptible to fluazifop

acid.

None of the biotypes examined had ACCase with high-level resistance to haloxyfop acid

(Table 3.7), such as was seen for diclofop acid and fluazifop acid. ACCase from biotypes

SAS I and NAF 8 were moderately resistant to this herbicide, whereas that from biotypes

SAF 14, NAS 4, NAS 2, NAF 6, NAS 6, WAF 12 and SAF 34 exhibited low-level

resistance to this herbicide. ACCase from biotype NAS 14 was susceptible to haloxyfop

acid.

ACCase from biotypes SAS 1, NAF 6, WAF 12 and NAS 2 was moderately resistant to

sethoxydim (Table 3.8). The ACCase from biotypes NAF 8, NAS 4, NAS 6, SAF 34 and

SAF 14 showed low-level resistance to sethoxydim whereas that from biotype NAS 14 was

as sensitive as the susceptible enzyme.

For tralkoxydim inhibition, ACCase from biotypes SAS l, NAF 6 and IWAF 12 was

moderately resistant to this herbicide (Table 3.9). The enzyme from biotypes NAS 14 and

SAF 34 was sensitive to tralkoxydim whereas that from biotypes NAF 8, NAS 2, NAS 6,

NAS 4 and SAF 14 showed lowJevel resistance to tralkoxydim.

60

-oLۃoI

CHoñ

-a-

a-ìIcÉ(ÐU2GI(JU

100

80

60

40

20

100

80

60

40

20

0

00.1 1 10 100

Concentration (pM)

Figure 3.44-8 Inhibition of ACCase activity by diclofop acid (A) and tralkoxydim (B)

in susceptible SAS 2 (O) and resistant SAS 1 (O) wild oat biotypes. Data are the means of

seven replicate experiments. Vertical bars represent the standard errors.

A

B

a

61

0.1 I 10 100

Concentration (ttM)

Figure 3.54-8 Inhibition of ACCase activity by diclofop acid (A) and tralkoxydim (B)

in susceptible SAF 19 (O) and resistant NAS 4 (O) wild oat biotypes. Data are the means of

four replicate experiments. Vertical bars represent the standard enors.

100

80

60

40

20

0

100

80

60

40

20

-oLrÅIocJ

CHo

Ba

È¡.-.a€I

c)at2c!QL)

0

e

B

62

0.1 1 10

Concentration (pM)

Figure 3.64-B Inhibition of ACCase activity by diclofop acid (A) and tralkoxydim (B)

in susceptible SAF 19 (O) and resistant NAF 6 (O) wild oat biotypes. Data are the means of

four replicate experiments. Vertical bars represent the standard errors.

-oLÈIoI

CHoBa

.aJ

-IcÉ(¡)u)6lU(J

100

80

60

40

20

00

80

60

40

20

I0

0100

A

B

0

63

100

80

60

-oËã40c.)c:o

Ð20*¡'F0.lË 1oocË

é)(t)

880t)

40

20

A

B

0.1 I 100

100

Concentration (pM)

Figure 3.74-B Inhibition of ACCase activity by diclofop acid (A) and tralkoxydim (B)

in susceptible SAF 19 (O) resistant NAS 14 (O) wild oat biotypes. Data are the means of

four replicate experiments. Venical bars represent the standa¡d enors.

&

Table 3.5 Concentrations of diclofop acid giving 507o inhibition of ACCase from

susceptible and resistant wild oat biotypes. Ratios are the value for the resistant divided

by the value for the susceptible biotype. ACCase activity \ilas measured in partially

purifred extracts of the shoot meristematic regions. Values are the mean È standard error

of four replicate experiments.

Biotype Iso (tM) ResisanUSusceptible

Resistant biotvoes

SAS 1*

NAS 4

NAF 8

SAF 14

NAF 6

NAS 2

NAS 6

NAS 14

SAF 34

WAF 12

Susceptible biotypes

SAS 2*

SAF 19

11.0 r 3.0

6.30 + 0.8

6.95 t 0.8

5.85 + 0.6

3.80 + 0.7

2.90 + 0.8

o.93 to.2

0.93 + 0.4

1.51 + 0.3

r.22 t 0.3

0.2 r 0.1

1.1 + 0.9

52

6

6

5

3

3

1

I

I

I

* Values a¡e the mean t standard error of seven replicate experiments

65

Table 3.6 Concentrations of fluazifop acid giving 507o inhibition of ACCase from

susceptible and resistant wild oat biotypes. Ratios are the value for the resistant divided

by the value for the susceptible biotype. ACCase activity was measured in partially

purified extracts of the shoot meristematic region. Values are the mean t standa¡d error

of foru replicate experiments.

Biot¡pe Iso (tM) Resistant/Susceptible

Resistant biowoes

NAS 2

SAF 14

NAF 8

NAS 4

SAS 1*

NAF 6

NAS 6

I9VAF 12

SAF 34

NAS 14

Susceptible biotypes

sAs 2*

SAF 19

157 + t5

49+tt

97 t7.0

78 + 6.0

23 + 6.0

14 + 6.0

3l + 3.0

r0 t 2.5

16 t 4.0

6t2.0

2.5 + 0.8

3.2 + r.0

50

15

30

25

9

4

10

3

5

1.9

* Values are the mean + standard error of seven replicate experiments

66

Table 3.7 Concentrations of haloxyfop acid giving 507o inhibition of ACCase from

susceptible and resistant wild oat biotypes. Ratios are the value for the resistant divided

by the value for the susceptible biotype. ACCase activity was measured in partially

purifred extracts of the shoot meristematic region. Values are the mean * standard error

of four replicate experiments.

Biotype Iso (tM) ResistanlSusceptible

Resistant biotvoes

SAS I*

NAS 2

NAF 6

NAS 4

NAF 8

SAF 14

NAS 6

NAS 14

WAF 12

SAF 34

Susceotible biotvoes

SAS 2*

SAF 19

17 r 1.0

1l + 1.0

5r1.6

16 r 3.0

24 t3.5

18 r 0.4

5.6 r 0.1

2.0 + 0.6

4.7 + 1.2

3.4 r 0.3

0.7 r 0.1

1.8 + 0.6

25

6

3

9

13

10

3

1

3

2

* Values a¡e the mean t standa¡d error of seven replicate experiments

67

Table 3.8 Concentrations of sethoxydim giving 507o inhibition of ACCase from

susceptible and resistant wild oat biotypes. Ratios a¡e the value for the resistant divided

by the value for the susceptible biotype. ACCase activity was measured in partially

purified extracts of the shoot meristematic region. Values a¡e the mean t standard error

of four replicate experiments.

Biotype Iso (pM) ResistanlSusceptible

Resistant biotvoes

NAF 6

V/AF 12

NAS 2

sAs 1*

NAF 8

NAS 4

NAS 6

SAF 34

SAF 14

NAS 14

Susceotible biowoes

SAS 2*

SAF 19

38 r 6.8

28 t2.8

25 + 5.t

22 t 5.0

22t 4.3

t4+ 2.5

t5 + 2.6

16 + 5.0

18 r 3.6

410.8

2.7 X0.7

5.4 + 0.9

7

5

5

8

4

3

3

3

3

I

* Values a¡e the mean t standa¡d error of seven replicate experiments

68

Table 3.9 Concentrations of tralkoxydim giving 507o nhtbition of ACCase from

susceptible and resistant wild oat biotypes. Ratios are the value for the resistant divided

by the value for the susceptible biotype. ACCase activity $,as measured in partially

purified extracts of the shoot meristematic region. Values a¡e the mean t standard error

of four replicate experiments.

Biotype Iso (pM) Resistant/Susceptible

Resistant biotvoes

SAS I*

NAF 6

WAF 12

NAS 14

NAF 8

NAS 2

NAS 6

NAS 4

SAF 14

SAF 34

Susceotible biowoes

sAs 2*

SAF 19

6.s + 5.0

3.7 + 0.9

5.5 + 1.4

0.9 r 0.1

2.t + o.3

1.8 t 0.4

t.9 t 0.2

1.7 r 0.3

2.6 + t.r

1.0 + 0.3

2.1t 0.6

0.8 r 0.r

3

4

6

I

2

2

2

2

3

I

* Values are the mean t standard error of seven replicate experiments

69

3.4 l)iscussion

Of the ten putative resistant populations of wild oat tested for resistance, all proved to be

resistant to at least one APP or CHD herbicide (Tables 3.1 to 3.4). Not all biotypes were

resistant to diclofop-methyl (Table 3.1), and only one biotype NAF 6 displayed

substantial resistance to the CHD herbicides (Table 3.4). The pattern of resistance at the

whole plant level could be generally classified into four groups (Table 3.10). A group

with high resistance to diclofop-methyl but sensitive to tralkoxydim consisted of biotypes

SAS l, NAS 4 and SAF 14. High resistance to diclofop-methyl and low resistance to

tralkoxydim was found in biotypes NAF 8 and NAS 2. Moderate resistance to both

diclofop-methyl and ualkoxydim was found in biotype NAF 6. Low resistance to both

herbicides was observed in biotypes NAS 14 and }VAF 12. Apart from those groups,

biotypes NAS 6 and SAF 34 were rather sensitive to both herbicides.

Most of these populations also contained an ACCase that was resistant to one or more of

the APP herbicides. In parallel with whole plant experiments, there was much variation

in the level of resistance exhibited at the enzyme level. Differences in the spectra of

resistance to APP and CHD herbicides at the whole plant and enzyme level in Table 3.10

for the populations of A. fatua and A. sterilis suggest that some of the va¡iation in

patterns of resistance in wild oat may be due to the presence of different physiological and

biochemical mechanisms of resistance.

Resistance to ACCase-inhibiting herbicides has been documented in several weed species

such as Lolium rigidum (Heap and Knight L982), L. tnultiþrum (Gronwald et al.1992),

Avena spp. (Joseph et al. 1990, Heap et al. 1993, Mansooji et aL L992), Setaria viridis

(Marles et al. 1993), Eleusine indica, Digitaria sanguinalis, Sorghum halepense and,

S. faberii (reviewed in Devine and Shimabukuro 1994). In most cases, resistance has

become evident after repeated applications of either, or both, APP and CHD herbicides for

several years. The patterns of resistance to these herbicides in resistant biotypes vary.

There may be resistance to diclofop only, to all APP but not CHD herbicides, or to both

APP and CHD herbicides.

70

Table 3.10 Degree of resistance at the whole plant level and at the target enzyme,

ACCase, of ten resistant biotypes of wild oat.

Biotype Whole plant response to: Enzyme rcsponse to:

Diclofop-methyl Tralkoxydim Diclofop acid Tralkoxydim

SAS I

NAS 4

NAF 8

NAS 2

SAF 14

NAF 6

NAS 14

V/AF 12

NAS 6

SAF 34 sensitive

sensitive

sensitive

moderate

loilmoderate

sensitive

high

high

high

high

high

low

low

low

low

low

high

low

low

sensitive

moderate

moderate

moderate

moderate

moderate

sensitive

sensitive

sensitive

sensitive

sensitive

moderate

sensitive

moderate moderate

lowlowsensitive

lowlow

low

7t

As for the whole plant response, the pattern of resistance of ACCase to herbicides also

varies. Most ACCase from resistant weed biotypes demonstrated a greater level of resistance

to APP herbicides than to CHD herbicides (Table 3.10) e.g. biotypes SAS I, NAS 4,

NAF 8, NAS 2 and SAF 14. Similar results were obtained from some resistant biotypes of

L. rigidum which were 30 to 85 fold resistant to diclofop but I to 8 fold resistant to

sethoxydim (Tardif et al. 1993; Tardif, Preston, Holtum and Powles unpublished; Preston,

Tardif, Christopher and Powles unpublished). ACCase from biotype NAF 6 was

moderately resistant to both APP and CHD herbicides. Likewise, two biotypes of .á..

myosuroider were moderately resistant to both diclofop and sethoxydim but they were more

resistant to fenoxaprop (Hall, Moss and Powles unpublished). In contrast to the biotypes

above, a biotype of S. viridir was highty resistant to both diclofop and sethoxydim (Marles

et al. 1993) and ACCase from a l27o subset of L. rigidum biotype SLR 3l showed low level

resistance to diclofop but was highly resistant to sethoxydim (Tardif and Powles 1994).

ACCase from biotypes NAS 14, ÌWAF 12, NAS 6 and SAF 34 was sensitive to diclofop and

only ACCase from biotypes NAS 6 and ÌWAF 12 showed a moderate level of resistance to

tralkoxydim (Table 3. 10).

Target site cross-resistance across APP and CHD herbicides in L. rigidum can result from

selection either with an APP herbicide (Holtum and Powles 1991) or a CHD herbicide

(Tardif et al. 1993). Target site cross-resistance has occurred in biotype NAF 6 which has

been selected with two applications of diclofop-methyl and has a moderate cross-resistance to

CHD herbicides. In contrast, wild oat biotypes SAS I and NAS 4, which had been exposed

to APP herbicides for at least five years, are highly resistant to APP herbicides but remain

sensitive to a CHD herbicide, tralkoxydim. A similar result was observed in a biotype of Z.

multiþrum which did not exhibit cross-resistance to CHD herbicides following selection

with APP herbicides (Gronwald et ^1.

1992). These results suggest that, although the

herbicide is a strong selection pressure for weeds carrying a herbicide-resistant ACCase,

cross-resistance across the APP and CHD classes cannot be predicted by herbicide exposure.

It is likely that there are a number of possible mutations of ACCase which can confer

72

resistance to the APP and CHD herbicides and those different mutations of ACCase affect the

action of different herbicides leading to resistance either within or across herbicide classes.

The data presented in this chapter establishes that there are clear differences in the spectra of

herbicide resistance among resistant wild oat biotypes. Thus, in the following chapters, four

resistant biotypes SAS 1, NAS 4, NAF 6 and NAS 14 were used to further investigate the

mechanisms of herbicide resistance. These biotypes are representative of the different

responses of the wild oat biotypes to both application of herbicides to plants and inhibition of

ACCase by herbicides (Table 3.10). Resistant biotypes SAS I and NAS 4 both have a high

level of resistance to APP herbicides, however, there are differences in the responses of the

ACCase from these two biotypes to the APP and CHD herbicides. ACCase from biotype

SAS I is highly resistant to diclofop and slightly resistant to tralkoxydim whereas ACCase

from biotype NAS 4 is moderately resistant to diclofop and slightly resistant to tralkoxydim.

Resistant biotype NAF 6 has moderate resistance to both APP and CHD herbicides at the

whole plant and ACCase levels, whereas biotype NAS 14 has a low level of resistance to

APP and CHD herbicides but contains a sensitive form of the ACCase erizyme.

ACCase from the susceptible biotype SAS 2 was strongly inhibited by a range of herbicides

whereas that of the resistant biotype SAS 1 was much less inhibited (Figure 3.4 and Tables

3.5 to 3.9). ACCase from the resistant biotype was markedly less sensitive to the APP

herbicides (diclofop acid, fluazifop acid and haloxyfop acid) than was ACCase from the

susceptible biotype by 52,25 and 9 fold, respectively. Likewise, the enzyme extracted from

resistant plants was less sensitive to CHD herbicides (sethoxydim and tralkoxydim) by 8 and

6 fold, respectively. Hence, a mechanism for resistance to ACCase inhibiting herbicides in

SAS 1 biotype is the possession of a less sensitive, mutant, target enzyme which has a good

correlation to the resistance at the whole plant level (Mansooji et al. 1992). To support the

assumption that an APP insensitive ACCase is a single mechanism in this biotype, other

possible mechanisms conferring resistance were investigated in Chapter 4.

Biotype NAF 6 shows moderate resistance to both diclofop-methyl and tralkoxydim and

contains an ACCase less sensitive to the APP herbicides diclofop acid, fluazifop acid and

73

haloxyfop acid by 3,4 and 3 times, respectively (Tables 3.5 to 3.7). This ACCase is also

less sensitive to the CHD herbicides sethoxydim and tralkoxydim by 7 and 4 fold,

respectively (Tables 3.8 to 3.9). The effects of APP and CHD herbicides on the target site

correlates well with the results of response to the herbicides at the whole plant level.

A modified target site, ACCase, which is moderately resistant to both APP and CHD

herbicides is probably the mechanism of resistance in this resistant biotype. Other possible

¡¡scþanism of resistance in this biotlpe will be examined in Chapter 5.

Resistant biotype NAS 4 exhibited a greater level of resistance in the field than biotype

SAS 1, whereas the ACCase enzyme from biotype NAS 4 is less sensitive to diclofop acid,

fluazifop acid and haloxyfop acid than ACCase from susceptible SAF 19 by 6,25 and 9 fold,

respectively (Figure 3.54 and Tables 3.5 to 3.7). Similarly, ACCase from resistant NAS 4

plants is slightly less sensitive to CHD herbicides sethoxydim and tralkoxydim by 3 and?

fold, respectively (Tables 3.8 to 3.9). Despite being more resistant to diclofop-methyl than

biotype SAS l, the ACCase of this biotype is considerably more sensitive to diclofop acid

than the ACCase of biotype SAS l. This suggests that other mechanisms may contribute to

resistance in this biotype, which will be examined in Chapter 6.

Resistant biotype NAS 14 is three times less sensitive than the susceptible biotype to

diclofop-methyl and tralkoxydim in the field (Tables 3.1 and 3.4), however, this biotype

contains an ACCase enzyme that is as sensitive to diclofop acid and tralkoxydim as that of the

susceptible (Figure 3.74-8). In addition, the ACCase from biotype NAS 14 exhibits the

same sensitivity to inhibition by haloxyfop acid and sethoxydim as that of the susceptible

plants (Tables 3.5 to 3.9). These results suggest that a modified target enzyme ACCase is

not the mechanism of resistance in this biotype. Possible mechanisms of resistance in this

biotype will be investigated in Chapter 7.

74

Chapter 4

Herbicide Resistance Endowed by an APP Insensitive ACCase

in Biotype SAS 1

4.1 Introduction

The response of wild oat biotype SAS I at the whole plant and enzyme level suggested that

this biotype contains a target site ACCase that is insensitive to these herbicides (Chapter 3).

Reduced sensitivity of the target site is the most commonly encountered mechanism of

herbicide resistance (Holt et al. 1993). This is also true of ACCase-inhibiting herbicides

where, for example, herbicide-resistant ACCase has been found in resistant biotypes of

Lolium multiþrum @etts et al. 1992, Gronwald et al. 1992); L. rigidum (Tardif et al. 1993,

Tardif and Powles 1994) and Setaria viridis (Marles et al. 1993). In wild oat, however, no

cases of resistance due to ACCase insensitivity to APP and CHD herbicides have been

reported (Boutsalis et al. 1990, Devine et al. 1992)

Herbicide resistance may occur as a result of one or more mechanisms. Therefore,

investigations of herbicide uptake, translocation, metabolism and herbicide-induced

membrane depolarization were conducted to establish if any other mechanisms were

contributing to herbicide resistance in this biotype.

4.2 Materials and Methods

Most materials and methods are as described in Chapter 2. Additional methods are described

below.

4.2.1 Variations in ACCase During Development

To verify whether there were differences between the two biotypes, susceptible SAS 2 and

resistant SAS 1, in the amount of ACCase enzyme during development, meristematic tissue

was harvested 5, 10, 20, 30, 40 and 45 days after planting. Crude extracts from both

75

biotypes were assayed with 0.2 ml\d acetyl-CoA. To measure ACCase activity at day 0,

seeds were imbibed with water fot 24 h and I g of seed was extracted and assayed as

described in Chapter 2.

4.3 Results

4.3.1 ACCase Activity During Growth Development

There were no large differences in the level of ACCase activity in crude extracts from

susceptible and resistant plants in any of the plant fractions examined (Table 4.1). The

highest relative ACCase activity was found in the meristematic regions in both biotypes,

while the leaf had the lowest relative activity. As relative activity was highest in meristematic

regions this tissue was used in all subsequent assays.

Table 4.1 ACCase activity (nmoles-rlCOz fixed g-1 FW min-l) in crude extract from

different parts of susceptible (SAS 2) and resistant (SAS 1) A. sterilis plants at the two leaf

stage. Values are the mean + standard error of six replicates.

Plant part Susceptible Resistant

Meristematic region 205.6 + 7.0 224.5 + 6.3

Stem 174.2 + 7.6 180.1 + 6.3

I-eaf lamina 160.7 + 3.2 t&.3 i.3.7

In addition, the amount of ACCase activity in the shoot meristematic regions of both

resistant and susceptible biotypes was found to be similar during the course of plant

development @gure 4.1). ACCase activity in the shoot meristem of both biotypes was low

76

in the imbibed seed at planting. Activity increased in a linea¡ fashion reaching a peak in bottt

biotypes 30 d after planting. ACCase activity in the meristematic regions subsequently

rapidly decreased such that 45 d after planting when plants \ilere at the booting stage

ACCase activity was very low. ACCase activity in the susceptible shoot meristem was

forurd to be slightly higher than in the resistant shoot meristem at most time points.

100

0 10 20 30 40 50

Days after planting

Figure 4.1 ACCase activity in the shoot meristem of the susceptible SAS 2 (O) and

resistant SAS 1 (a) A. sterilis from zero to 45 d after planting. ACCase activity was

measured in crude extracts and each point is the average of three replicates. Vertical bars

represent standard errors and are sometimes obscured by point symbols. The arrow

indicates that soil emergence occurred 5 d after planting the seeds.

î80-I

.l--'tB60.-i

Ë l¡i

Ë ä40z\ (9xãv-

920

0

77

4.9.2 Uptake and Translocation of [lag¡Oictofop-Methyl

To quantify the rates of uptake and Eanslocation of diclofopmethyl, radiolabelled herbicide

was applied to the leaf axil of seedlings at the two-leaf stage. The amount of [tag¡6¡clofop-

methyl absorbed into treated plants was mærimal after ?.4hin both biotypes (Figure 4.2).

The pattems of [l4C1¿¡clofop-methyl uptake in the resistant and susceptibte biotypes rvere

similar with SOVo of uptake evident 6 h following treatment. The amount of radioactivity

that could be washed from the leaves declined rapidly and reached 0 at 180 h after

application. Not all of the label appeared in the extract with significant losses occurring due

either to volatilization, or binding of the label to insoluble material.

For the fi¡st 12 h after treatment, more than SOVo of the radioactivity absorbed was found in

the stem (Figure 4.34). After this time, increasing amounts of label were observed in the

second leaf until about 50Vo of the radioactivity as in the second leaf from 48 h (Figure

4.38). Little radiolabel was detected in the first or third leaves at any time @gure 4.3D-E).

After 6 h exposure, l%o and 3Vo of the radiolabel absorbed was located in the shoot

meristematic regions of the susceptible and resistant biotypes respectively and then rapidly

declined (Figure 4.3C). Although the radiolabel in the shoot meristematic region of the

susceptible biotype was higher than that in the resistant, they were not significantly different.

At no time were there clear differences between the resistant and susceptible biotypes.

Although the total amount of radiolabel recovered in the susceptible and resistant biotypes

during the period of the experiment did not differ, the susceptible plants showed necrosis on

the first, second and third leaves whereas the resistant plants showed some chlorosis at 24 h

at the point of application but newly emerged leaves were unaffected. Very little

radioactivity lvas found in the roots of either biotype (data not shown).

78

100

80

60

40

20

0

E(¡).--ÈÊG-ctto€tHo

Ba

J.l.lIIGo.-EÉË

ú

061218z-4896144192Time after application (h)

Figure 4.2 The absorbed (O, O) and unabsorbed (1, tr) radioactivity in the seedlings of

susceptible SAS 2 (open symbols) and resistant SAS I (closed symbols) A. sterilis. Each

point is the mean of two experiments with two replicates.

79

100

50

100

50

0

20

10

0

10

0

0q)¡cl-È=-cl€o.aJ

tHo

èa

tal

.¡tIcúo.tE6É

ú

10

006121824 48 96

Time after application (h)

Figure 4.34-E Distribution of radioactivity following application of [l4ç¡*.lofop to the

leaf ædls of the susceptible SAS 2(O) and resistant SAS I (O) A. sterilis; stem (A), second

leaf (B), shoot meristematic region (C), fust leaf (D) and third leaf (E). Each point is the

mean of two experiments with two replicates. The thi¡d leaf of resistant plants emerged after

48 h but the thi¡d leaf of susceptible plants did not emerge until 140 h. Vertical bars represent

standa¡d enors and are sometimes obscured by symbols.

192

A

B

c

D

E

80

4.3.3 Metabolism of [14C]n¡ctofop-Methyl

[laC]Diclofop-methyl was converted to diclofop acid at a similar rate in both susceptible and

resistant biotypes. Conversion of diclofop-methyl to diclofop acid was rapid with less than

lÙVo of the applied radioactivity recovered as diclofop-methyl in both biotypes 12 h after

treatment @igure 4.4A). The amount of radioactivity recovered as diclofop acid, the

phytotoxic form of the herbicide, reached a peak 6 h after treatment. The level of

[l4C]diclofop acid rapidly decreased to 20Vo after 48 h. Little label was recovered as

ll4Cldictofop acid in either biotype 180 h after treatment (Figure 4.48). HPLC elution

profiles suggest that diclofop acid was metabolized to similar polar compounds in both

biotypes (Figure 4.5). At 180 h after treatment, more than 95Vo of the radioactivity

recovered from both biotypes was in the form of metabolites of diclofop acid with no

differences between the two biotypes (Figure 4.4C). Therefore, in both resistant and

susceptible biotypes, diclofop-methyl was rapidly converted to diclofop acid, which was in

turn converted to more polar metabolites. Although most of the herbicide in both biotypes

was metabolized, only the susceptible plants were killed by herbicide application.

The HPLC elution profiles of the parent herbicide a¡rd its metabolites at24h after treatment

were similar for both susceptible and resistant biotypes (Figure 4.54-8). Although the

elution profile of diclofop acid and metabolites was simila¡ between susceptible and resistant

wild oat (Figure 4.54-8), cultivated oat (4. sativa cv. Echidna) showed a different pattern

with more polar metabolites eluting between 2 and l0 min than either wild oat biotype

(Figure 4.5C). Metabolism occurred in both the shoot meristematic regions and in the

leaves. In translocation experiments, about 25Vo of radiolabel in the leaves was diclofop

acid in both biotypes 48 h after application and the HPLC elution profiles were similar (data

not shown). Therefore, differences in the amount of unmetabolized herbicide at the

meristem cannot account for resistance.

81

A

0

0612182A-4896144192

Time after application (h)

Figure 4.4L-C Percentage of [laqdt.lofop-methyl (A), diclofop acid (B) and turther

metabolites (C) in the susceptible SAS 2 (O) and resistant SAS I (O),A. stertHs extracted

from 0 to 180 h after treatment. Each point is the mean of two experiments with two

replicates. Venical bars represent standard enors and are sometimes obscr¡red by symbols.

100

t0

60vt-Ë40€Xo20-I

äos 100CH

:BoB\

ã60.-

Ë40GIoË206l¡rts0Ë 100o.i

=80Êa-L.Ë60ä

40

20

B

c

82

:rA

susceptible wild oat

:l€B

resistant wild oat

{(

c

oat cv. Echidna

Ð.--taLGIL-.l¡¡ictl

Ì.l

.l

-9cËo.-E6lú

0 5 10 15 20

Elution time (min)25 30

Figure 4.54-C HPLC elution profiles of extracts from susceptible SAS 2 (A), resistant

SAS 1 (B) ,4. sterilis, and cultivated oat .4. sativa cv. Echidna (C) 24 h after trearnent with

l14C]diclofop-methyl. ll4C]piclofop acid eluted at the retention time of 26 min (*).

83

4.3.4 Effect of Diclofop Acid on Plasma Membrane Potential

In the presence of 50 ¡rM diclofop acid, the plasma membrane potential of etiolaæd coleoptile

cells from both susceptible and resistant wild oat depolarized within 15 min @igure 4.6).

After removal of diclofop acid, the plasma membrane potential remained depolarized for

more than 20 min for both biotypes (Figure 4.64-8). Therefore in this biotype of wild oat,

resistance to diclofop does not correlate with recovery of the plasma membrane electrogenic

potential following removal of diclofop acid.

4.3.5 Inhibition of ACCase Activity Iz Vivo

Translocation experiments revealed that the herbicide concentration reached a mærimum in

the meristem of both resistant and susceptible biotypes 6 h after treatment (Figure 4.3C).

Hence, following herbicide application to intact seedlings, invivo ACCase activity from the

susceptible biotype should become inhibited due to the susceptibility of this enzyme to

ACCase-inhibiting herbicides. However, as the ACCase in the resistant biotype is much less

sensitive to these herbicides (Table 3.5), in vivo ACCase activity should be less inhibited.

To measure this directly, plants were Eeated with herbicide and then the activity of ACCase

from crude extract of the meristems was measured at l, 3, 6,48, 140 and 180 h after

herbicide application. These experiments revealed that extractable ACCase activity was

reduced by SOVo for the susceptible biotype 6 h after treatment whereas ACCase activity

extracted from the treated resistant biotype was reduced by only 3OVo (Figure 4.7). The

extractable ACCase activity in the susceptible biotype remained inhibited up to 180 h after

treatment. However, in the resistant biotype, extractable ACCase activity was only reduced

in the ftrst?4 h after treatment and completely recovered within 180 h. This in vivo result

confirmed the in vitro measurements demonstrating strong inhibition of ACCase activity

from the susceptible but not resistant plants (Table 3.5) and allows the conclusion that

resistance to ACCase-inhibiting herbicides in the resistant biotype is endowed by a mutant

form of ACCase which is markedly less sensitive to inhibition by herbicides.

84

20 30

mtn

Figure 4.64-B The effect of 50 pM diclofop acid on plasma membrane potential in

coleoptile cells of the susceptible SAS 2 (A) and resistant SAS I (B) A. sterilis. Arrows,

addition (+) and removal (-) of diclofop acid (DA). The data shown is from a single

coleoptile and is representative of ten coleoptiles each of the susceptible and resistant

biotypes.

-120

-90

-60

-30

0

-tI

-CE.-Ð-:()*aoÈc)IGL¡-tI()Hà

-120

-90

-60

-30

0100

+ A+DA

.DA

¡

- +DA B

DA

I

85

100

50

061218z-140 180

Time after herbicide application (h)

Figure 4.7 Extractable ACCase activity in the meristem of the susceptible SAS 2 (O) and

resistant SAS I (O) A. sterilis after treatment with a cornmercial formulation of 5 mM

diclofop-methyl on the leaf axil. ACCase activity was measured in crude extracts and values

are expressed as a percentage of those for untreated plants. Vertical bars represent standard

errors of four replicates.

-oLr-toICIoèa

Ða-

a-IIcË(¡)trhG()L)

0

86

4.4 Discussion

There were no differences in diclofop absorption, translocation or metabolism between the

resistant A. sterìlis biotype SAS 1 and susceptible SAS 2 (Figures 4.2, 4.3A-E and 4.4).

Therefore, these potential resistance mechanisms do not account for diclofop-methyl

resistance in this biotype. To date, there have been no reports of APP or CHD resistance

due to altered absorption or translocation of these herbicides in the related species A. fatua

(Devine et al. 1992). Similarly, absorption and translocation of diclofop-methyl a¡e simila¡

for susceptible and resistant biotypes of L. rigidum (Holtum et al. 1991), L. multiþrum

(Gronwald et ^1.

L992) and S. viridis (Marles et al. 1993).

Häusler et al. (1991) reported that in a number of biotypes of L. rigidurn resistance to APP

and CHD herbicides was correlated with the ability of the plasma membrane electrogenic

potential to recover from diclofop acid-induced depolarization following removal of the

herbicide. A similar situation has been observed for a resistant biotype of A. fatua (Devine

et al. 1992). In this study, neither the resistant nor susceptible A. sterilis biotypes exhibited

recovery of the plasma membrane potential following removal of diclofop acid (Figure 4.6).

Therefore, resistance to APP herbicides in biotype SAS I differs from resistance in the

resistant biotype of A. fatw (Devine et al. 1992) in that it does not correlate with recovery of

electrogenic membrane potential. This phenomenon will be considered at greater length in

Chapter 8.

In certain tolerant maize cell lines selected in vitro, resistance to APP and CHD herbicides is

due to the overproduction of the normal, sensitive ACCase (Parker et al. 1990b). The 58-

and 9O-fold resistance to sethoxydim and haloxyfop, respectively, in the selected line B50S

was conferred by a 2.6-fold increase in herbicide-sensitive ACCase activity. However, in

biotype SAS 1, differences in the activity of the target enzyme, ACCase, cannot explain

resistance as the level of activity in the susceptible and resistant biotypes were similar at all

stages of development (Figure 4.1). The distribution of ACCase activity in shoot meristem,

stem and leaf lamina in the resistant and susceptible biotypes was also simila¡ (Table 4.1).

87

However, ACCase activity in panially-purifred exEacts from the resistant biotype was much

less susceptible to inhibition by APP herbicides than that of the susceptible (Tables 3.5 to

3.e).

At the whole plant level, the resistant biotype exhibits a high level of resistance to APP

herbicides and slight resistance to CHD herbicides (Mansooji et al. 1992). This difference

between herbicide groups is reflected at the ACCase level with the level of resistance of the

enzyme being greater for APP herbicides (9- to 52-fold) than CHD herbicides (6- to 8-fold)

(Tables 3.5 to 3.9). Furthennore, when ACCase was extracted from tissue of susceptible

and resistant plants at various periods after herbicide treatment, the enzyme extracted from

susceptible plants was found to be inhibited to a gteater extent and for a longer period than

that of the resistant (Figure 4.6). Therefore, there is a good correlation between the

inhibition of ACCase activity in vivo and in vitro (Table 3.5 and Figure 4.7). Clearly, the

effects of herbicides on ACCase activity observed both in vivo and in vitro provide direct

evidence that the resistance mechanism in this resistant biotype involves a modified form of

ACCase. Preliminary genetic studies with this resistant A. sterilis biotype suggest that

resistance is endowed by a single, major nuclear gene (Barr etal.1992} It is likely that this

single gene is that encoding for the resistant ACCase enzyme in the biotype SAS l.

88

Chapter 5

Herbicide Resistance Endowed by an APP- and CHD-Inseruitive

ACCase in Biotype NAF 6

5.1 Introduction

Many of the wild oat biotypes examined in Chapter 3 showed resistance to APP herbicides

due to an APP-insensitive ACCase. A few of these biotypes show slight resistance to the

CHD herbicides, however, biotype NAF 6 is moderately resistant to both APP and CHD

herbicides. The mechanism of resistance to herbicides in this resistant biotype NAF 6

appears due to a modified target enzyme, ACCase, which is moderately resistant to both

APP herbicides and CHD herbicides (Chapter 3). This is the only wild oat biotype to show

substantial resistance to CHD herbicides.

The mechanisms providing resistance to CHD herbicides have been elucidated in a number

of studies. Non-regenerable muze cell lines selected with sethoxydim were resistant due to

an overproduction of the target enzyme, ACCase (Parker et al. 1990b). Resistance to

sethoxydim n Inlium rigidum biotype SLR 3 is due to a modified form of ACCase (Tardif

et al. 1993), as is resistance ín Setaria viridis (Marles et al. 1993). A L2Vo subset of L.

rigidum biotype SLR 3l is highly resistant to sethoxydim and other ACCase-inhibiting

herbicides and contains a resistant ACCase target enzyme (Tardif and Powles 1994),

however, the remainder of the population is susceptible to this herbicide. It seems therefore

that most cases of resistance to CHD herbicides, and especially sethoxydim, is conferred by

an insensitive ACCase. Although target site modification appeÍìrs to confer resistance to

CHD herbicides in some species, there may be more than one mechanism of resistance to

CHD herbicides in biotype NAF 6.

To verify whether other mechanisms were contributing to resistance in biotype NAF 6,

uptake, translocation and metabolism of diclofop were examined in both susceptible SAS 2

89

and resistant NAF 6 biotypes. Uptake and metabolism of sethoxydim in this biotype were

also investigated.

5.2 Results

5.2.1 Uptake, Translocation and Metabolism of [14C]O¡ctofop-Methyl

Foliar absorption of diclofop-methyl occurred readily in both susceptible and resistant wild

oat biotypes (Figure 5.1). At I and 6 h after treatment, 25 and427o of the radioactivity was

absorbed in the resistant biotype NAF 6 compared with 27 and 45Vo for the susceptible

biotype SAS 2. Absorption of radioactivity reached a mærimum at 80Vo by 24 h after

application in both biotypes.

The distribution of radioactivity between the meristematic region and the rest of the plant was

found to be identical in both susceptible and resistant biotypes. Translocation of

[lac]diclofop was limited with approximately lOVo reaching the meristematic tissue in both

biotypes by 6 h after application (Figure 5.2A). Most of the radioactivrty absorbed (-907o)

was found in the rest of the foliage (Figure 5.28).

The rate of metabolism of llaC]¿ictofop-methyl was identical in both susceptible and

resistant biotypes (Figure 5.3). Metabolism of diclofop-methyl by susceptible and resistant

plants was nearly complete within 48 h (Figure 5.34). Diclofop-methyl was rapidly

converted to the phytotoxic form, diclofop acid in the fust 6 h following treatment (Figure

5.3B). At72 h after treatment, more than SOVo of the label recovered from both biotypes

was in the form of metabolites of diclofop acid with no differences between two biotypes

(Figure 5.3C). The major pathways for metabolism of diclofop-methyl appear to be similar

in the susceptible SAS 2 and resistant NAF 6 biotypes.

90

Time after application (h)

Figure 5.1 The absorption of radioactivity following application of [lag¡diclofop-methyl

to the axil of two-leaf stage seedlings of susceptible SAS 2 (O) and resistant NAF 6 (O)

wild oat biotypes. Points are the mean t standard error of three replicate experiments.

100

80

60

40

20

0

Eq).I

-ÈÈ6l

-6tto¡fHo

èa

-.-.--IGIo.-EcÉ

ú

A0 7248

9l

20

10

0

100

80

60

40

20

(¡),¡lcÉÐÈ!a-

-acË

-ÌCHoèa

-.ã-CJ6úoEc!ú

0

u18t260

Time after application (h)

Figure 5.24-B Radioactivity found in the meristematic region (A) and the rest of the

foliage (B) of susceptible SAS 2 (O) and resistant NAF 6 (O) wild oat biotypes following

application of [l4c]diclofop-methyl to the axil of nvo-leaf seedlings. Point symbols are the

mean of five replicate experiments. Vertical bars represent the standard errors of the mean

and a¡e sometimes obscured by point symbols.

A

I I

B

I

92

100

80

60

40

20

0100

80

60

40

20

0100

80

60

40

20

0u 48 72

Time after application (h)

Figure 5.34-C Percentage of extracted radioactivity as diclofop-methyl (A), diclofop

acid (B) and further metabolites (C) following application of ¡l4cldiclofop-methyl to the

susceptible SAS 2 (O) and resistant NAF 6 (O) biotypes of wild oat up to 72 h after

treatnent. Each point is the mean of th¡ee replicates. Vertical bars represent standard errors

and are sometimes obscured by symbols.

et)

-IG'LÈX(¡)

-L.l

-ÉË

-o-CHo

èa

5.I

.--IeÉo.lËc!¡rr:o-I

.tÐI-¡

alL-(t2â

0

A

B

c

93

5.2.2 Upteke and Metabolism of [4-149]Sethoxydim

Absorption of [4-laC¡sethoxydim was identical in both susceptible SAS 2 and resistant

NAF 6 biotypes. Absorption of [4-1aC¡sethoxydim, when applied onto the leaf axil of two-

leaf stage seedlings, increased rapidly from I to 6 h after treatment and reached a maximum

of about 9OVo by 72 h after herbicide application (Figure 5.4). The susceptible biotype

SAS 2 absorbed slightly more radioactivity than did the resistant biotype NAF 6, however,

this difference was not significant.

The parent herbicide, ¡4-laClsethoxydim, was metabolized by both biotypes at the same rate

(Figure 5.54). By t h after treatment, only 30Vo of the radioactivity was recovered as

sethoxydim and by 24 h very little sethoxydim was detected. Sethoxydim was rapidly

converted to a metabolite (Ml) which was more polar than sethoxydim (Figure 5.6). This

metabolite reached a maximum concentration 6 h following application (Figure 5.5B) and

was subsequently converted to other products. Radioactivity in other peaks was summed as

polar metabolites (Figure 5.5C). These metabolites increased very rapidly from I to 36 h

after treatment and represented 85Vo of the radioactivity in the extract at the end of the

sampling period @igure 5.5C). There was no difference between susceptible SAS 2 and

resistant NAF 6 in the amount of either the major metabolite or the other polar metabolites

produced. HPLC chromatograms of sethoxydim and metabolites were identical in the

susceptible and resistant biotypes as shown in Figure 5.64-8.

94

00

80

60

40

20

0

1Eq)al

-ÈÈeÉ

-ÉÉ€oICEoñÐ.-.l

-Ic!o

.lEc!ú

72

Time after application (h)

Figure 5.4 Absorption of [4-14c]sethoxydim by susceptible SAS 2 (O) and resistant

NAF 6 (O) wild oat biotypes following application to the axil of two-leaf stage seedlings.

Point symbols are the mean * standard error of three replicate experiments.

5436180

95

100

80

60

40

20

0100

80

60

40

20

0100

80

60

40

20

0

E)-IqB¡r-Xq)

-I.--6!-otÞ

cHo

Ba

>¡-.-a--c.)clo.l€ÉtlL

CHo-Io.----I

li€ct2.-â

54

Time after application (h)

Figure 5.54-C Metabolism of [4-l4c¡sethoxydim in the susceptible SAS 2 (O) and

resistant NAF 6 (O) wild oat biotypes when applied to the leaf axil at the twoleaf stage.

Percentages of [4-l4g]sethoxydim (A), metabolite Ml (B) and other metabolites (C) in the

extracts. Points are the mean t standard efrors of the experiment with four replicates. Error

ba¡s a¡e sometimes obscured by point symbols.

720 18 36

A

B

c

96

*

*

^^^^

A

susceptible wild oat

:lc

*

B

resistant wild oat

Ð.l-tãliÉËL-.Fl¡L(!I.-.-ùI6!o.-EcË

ú

051015 30

Elution time (min)

Figure 5.64-8 HPLC elution profiles of the extracts from the susceptible SAS 2 (A) and

resistant NAF 6 (B) t h after treatment with ¡4-laClsethoxydim. Metabolite (Ml) and

sethoxydim eluted at a retention time of 13 (tt) and 20 min (*), respectively.

97

5.2.3 Effect of Diclofop Acid on Plasma Membrane Potential

To determine whether an altered plasma membrane response to diclofop acid is a mechanism

conferring resistance in this biotype, the plasma membrane potential of etiolated coleoptile

cells of both susceptible SAS 2 and resistant NAF 6 were measured. In the presence of

50 pM diclofop acid, the plasma membrane potential from both susceptible and resistant

biotypes was depolarized within 15 min (Figure 6.6,{-8). After removal of diclofop acid,

the plasma membrane potential from both biotypes remained depolarized for more than

20 min. Hence, in this biotype resistance is not correlated with a recovery of the plasma

membrane electrogenic potential following removal of diclofop acid.

5.3 l)iscussion

The mechanism of resistance to APP herbicides in biotype NAF 6 does not involve changes

in uptake, translocation or metabolism of diclofop-methyl (Figures 5.1 to 5.3) or an

alteration of the plasma membrane response to diclofop acid (Figure 5.6). In addition,

resistance to CHD herbicides in this biotype is not due to reduced uptake or enhanced

metabolism of sethoxydim (Figures 5.4 and 5.5). A modified target site, ACCase enzyme,

appe¿us to be the only mechanism responsible for resistance for both APP and CHD

herbicides in this biotype (Chapter 3). The ACCase enzyme from this resistant biotype is 3,

4 and 3 times more resistant to diclofop acid, fluazifop acid and haloxyfop acid, respectively

(Tables 3.5 to 3.7) and is 7 and 4 times more resistant to sethoxydim and tralkoxydim

(Tables 3.8 and 3.9) than that of the susceptible biotype.

The level of resistance to diclofop-methyl may be correlated with the history of herbicide

application. Herbicide resistance to ACCase-inhibiting herbicides has been shown to occur

after three years or more of herbicide application (Tardif et al. t993, Devine and

Shimabukuro 1994). Exposure to APP herbicides for five years in biotype SAS I and nine

years in biotype NAS 4 resulted in a high level of resistance of both biotypes to diclofop-

methyl (Table 2.l,Table 3.10). In contrast, biotype NAF 6 had been exposed to only two

applications of APP herbicide before developing a moderate level of resistance to diclofop-

98

-LI-Grt-I()i¡cÈé)I6lL¡-Iq)

FTÀ

-100

-80

-60

-120

-120

0

-100

-80

-60

00 10 20

Time (min)

30

Figure 5.64-B The effect of 50 pM diclofop acid on plasma membrane potentials in

coleoptile cells of the susceptible SAS 2 (A) and resistant NAF 6 (B) biotypes. Arrows,

addition (+) and removal (-) of diclofop acid (DA). The data shown is from a single

coleoptile and is representative of ten coleoptiles each of the susceptible and resistant

biotypes.

+DA

ü A

.DA

,

+DA

ü B

.DA

+

,I I

99

methyl. It is difficult to correlate resistance to CHD herbicides with herbicide use histories.

Biotype NAF 6 had not been exposed to CHD herbicides yet it exhibited a moderate

resistance to CHD herbicides (cycloxydim, sethoxydim and tralkoxydim). In contrast,

biotype NAS 4 had been exposed to two applications of CHD herbicides but this biotype is

still sensitive to tralkoxydim (Table 3.10). In the case of biotype NAF 6, it is clea¡ that

selection by diclofop-methyl has concomitantly provided target site cross-resistance to CHD

herbicides. This does not always occur and many wild oat biotypes only have resistance to

APP herbicides.

Some resistant biotypes possessing resistant ACCase are highly resistant to both APP and

CHD herbicides. For example, a high level of resistance to APP and CHD herbicides has

been described in Setaria viridis (Marles et ai. 1993) and Z. rigidum (Tardif et al. 1993,

Tardif and Powles 1994) based on a modified form of ACCase. Yet, wild oat biotypes

SAS I and NAS 4 that possess APP-resistant forms of ACCase are highly resistant to APP

herbicides only. A similar pattern is also evident in a resistant biotype of. L. multiþrum

(Gronwald et al. L992). Biotype NAF 6 also contains an APP-resistant ACCase but it is

moderately resistant to both groups of herbicides.

The ACCase from biotype NAF 6 appears to be different to that of biotypes SAS I and NAS

4 since it exhibits moderate resistance to both APP and CHD herbicides. Probably, the

mutations of the herbicide binding sites within ACCase are not the same among the three

resistant wild oat biotypes. Modified forms of ACCase from mutantmuze cell cultures,

selected by sethoxydim or haloxyfop are encoded by different allelic mutations at the same

locus (Marshall et al.1992). These data indicate that there a¡e several potential modifications

of ACCase which can confer resistance to APP and CHD herbicides. The differences in

sensitivity of resistant forms of ACCase to CHD herbicides indicates that the CHD and APP

herbicides probably bind to overlapping but non-identical sites on the enzyme. In the ten

resistant biotypes examined in this thesis, modifications of ACCase which confer resistance

to APP herbicides only are more common than those which confer target site cross-

resistance to CHD herbicides (Chapter 3).

100

Chapter 6

Resistance to APP Herbicides Endowed by Two Mechanisms in

Biotype NAS 4

6.1 Introduction

At the whole plant level, wild oat biotypes NAS 4 and SAS I could not be controlled by the

very high rate of 30,000 g diclofop-methyl ¡¿-l (Table 3.1). However, at the enzyme level,

biotype NAS 4 exhibited an ACCase that was six-fold less sensitive to diclofop (Table 3.6),

whereas the ACCase from biotype SAS I showed 52-fold less sensitivity to diclofop than

that of susceptible plants (Table 3.5). Thus, the modified target site, ACCase enzyme, in

biotype NAS 4 may not be sufficient to confer the high level of resistance observed in the

whole plants. Therefore, studies were undertaken with biotype NAS 4 to establish whether

herbicide resistance in this biotype may involve more than one mechanism.

That different mechanisms of resistance to ACCase-inhibiting herbicides can occur in

populationsof Avena isshownbyacomparisonof Avenasterilis biotypeSAS l whichhas

a mutant ACCase enzyme (Chapter 4) andA.fatuabiotype UM-l which possess alterations

in membrane properties (Devine et al. 1993). Other mechanisms, such as enhanced

metabolism of diclofop a¡e also possible as, has been documented in some Inlium rigidum

biotypes. For example, L. rigidumbiotype SLR31 was able to detoxify diclofop acid at

about 1.5 times of the rate of the susceptible biotype (Holtum et al. l99l). However, to

date, enhanced metabolism of diclofop has not been observed in biotypes of wild oat.

To determine whether non-target site mechanisms co-exist with resistant ACCase in biotype

NAS 4, uptake, tanslocation, and metabolism of ll4c]diclofop-methyl as well as the ability

of plasma membrane to recover from diclofop-induced membrane depolarization by diclofop

acid were investigated.

101

6.2 Materials and Methods

I;.:l¡ I[ tfrilf F,i.jf.å Í-li:iii¡,nY; ;: ühi!'il:fi;ìi'i T fii: fiüf; #li]_.

Most materials and methods are as described in Chapter 2. Additional methods are

described below.

6.2.1 Effects of the Cytochrome P450 Inhibitor Tetcyclasis on Diclofop

Metabolism

To demonstrate the effects of tetcyclasis on the metabolism of diclofop-methyl, wild oat

biotypes NAS 4, SAF 19 and wheat plants were grown in hydroponic solution as described

in section 2.2.3. Tetcyclasis (20 ttM) was added to the nutrient solution 24 h prior to

treatment with I pL of 5 mM ll4C]diclofop-methyl applied to the leaf axil. Plants were

harvested 24 h following herbicide treatment, ground, radioactivity extracted and the

metabolites examined by HPLC. The experiments were a completely randomized block

design with four replicates.

6.2.2 Effects of the Cytochrome P450 Inhibitor Tetcyclasis on Plant

Growth

To determine the effects of tetcyclasis on the growth response of wild oat biotypes, resistant

(NAS 4 and SAS 1), susceptible (SAF 19) wild oat biotypes, and wheat cv. Halberd were

cultured using hydroponics in I L rectangular plastic trays. Each tray contained a single

row of each biotype (six plants/row). The nutrient solution was replaced with solution

containing 20 pM tetcyclasis when plants were at the two-leaf stage. After t¡eatment with

20 pM tetcyclasis for 24 h, a I pL droplet of l0 mM diclofop-methyl (commercial

formulation) was placed on the leaf axil of the two-leaf stage plants. Seven days following

herbicide application, shoot and root length of individual plants tvere measured. Dry weight

of shoot and root were also determined. The experiments were a completely randomized

block design with four replicates.

To determine the growth response of wild oat biotypes to various rates of diclofop-methyl in

the presence and absence of tetcyclasis, susceptible SAF 19 and resistant NAS 4 were

t02

grown in hydroponic culnue and treated with 20 pM tetcyclasis as described above. After

treatment with 20 ¡rM tetcyclasis for 24h, a I ¡tL droplet of 0, 2.5, 5 and l0 mM diclofop-

methyl (commercial formulation) were placed on the leaf æril of the two-leaf stage plants.

Seven days following herbicide application, shoot and root length of individual plants were

measured. Dry weight of shoot and root were also determined. The experiments were a

completely randomized block design with four rreplicates.

6.2.3 Inhibition of ACCase by Diclofop in Single Lines of NAS 4

To test the hypothesis that two mechanisms of resistance to diclofop are present in each

individual of the NAS 4 population, seeds from ten different plants of biotype NAS 4 which

had survived 52 g haloxyfop-ethoxyethyl h¿-l were separately collected and used for

ACCase measurements. Two lines of NAS 4, resistant SAS I and susceptible SAF 19

biotypes were arranged in each assay with four replicates.

6.3 Results

6.3.1 Uptake and Translocation of ¡ugfDiclofop-Methyl

Labelled diclofop-methyl was absorbed rapidly when applied to the leaf axil of two-leaf

stage wild oat seedlings (Figure 6.1) with no differences between the two biotypes. More

than 9OVo of the applied herbicide was observed in the stem following application of

herbicide onto the leaf axil of two-leaf seedlings of both biotypes. The treated area (stem)

contained the major fraction of the ¡l4C1diclofop at atl harvest times (Figure 6.2B). Very

little radioactivity was translocated to the leaf tips and roots of either biotype (Figure 6.2C-

D). Translocation of radiolabel from the treated area to the meristematic region was similar

and reached a maximum of l2%o and lïVo respectively in the susceptible SAF 19 and

resistant NAS 4 biotypes (Figure 6.24). Therefore, uptake and translocation do not

contribute to resistance to diclofop in this biotype. This result is simila¡ to that observed for

biotype SAS I in Chapter 4, however, more radioactivity appeared in the leaves of both

biotypes in that experiment. This may be because the plants in the experiment described

here were growing more slowly.

103

e 1oo.l

-ÈÈcÉ Bo-cl-o-860èa

b40.¡.t-I820.-€GI

ú0

0 44872Tme after application (h)

96

Figure 6.1 Absorption of [l4C]d'clofop-methyl by susceptible SAF 19 (O) and resistant

NAS 4 (O) wild oat biotypes. ll4c]diclofop-methyl was applied to the leaf axil of the two-

leaf stage seedlings. Points are the mean * standard error of four replicate experiments.

104

0612182/4896

Time after application (h)

Figure 6.24-D Distribution of radioactivity in the meristem (A), stem (B), leaf (C) and

root (D) from ¡lag¡diclofop-methyl deposited on the leaf axil of twoJeaf stage seedlings of

the susceptible SAF 19 (O) and resistant NAS 4 (O) wild oat biotypes. Points a¡e the mean

* standard error of five replicate experiments.

20

10

6.04 100

-3Bo-dË60€¡a

:40B\

;20i.l.r

Ëoclo.i

H10ú

0

10

0

A

Hù-o

B

c

D

105

6.3.2 Metabolism of [14C]nictofop-Methyl

In susceptible (SAF 19) and resistant (NAS 4) biotypes, ¡l4c¡diclofop-methyl was

degraded to diclofop acid at an equal rate @igure 6.3A). Radioactivity rapidly accumulated

in the ph¡otoxic form, diclofop acid, with 6OVo of the label in this form t h after treatment

(Figure 6.38). Conversion of diclofop-methyl to diclofop acid was similar in both biotypes

(Figure 6.38), however, the conversion of diclofop acid to the other metabolites was more

rapid in the resistant biotype compared to the susceptible. From 6 to 48 h after treatment,

the susceptible tissue contained between 5 to lsEo more of the radiolabel as ¡laç¡diclofop

acid than did the resistant tissue. The decline in diclofop acid was accompanied by an

increase in more polar metabolites in both biotypes. These metabolites were evident at 6 h

after the treatment in the resistant biotype and were produced more rapidly by this biotype

(Figure 6.3C). By 72 h after treatment, the content of ll4C]diclofop acid had decreased to

11 and l4%o in biotypes NAS 4 and SAF 19, respectively (Figure 5.38) with no differences

between the two biotypes. However, the initial rate of production of polar metabolites wÍrs

1.5 times faster in the resistant biotype compared to the susceptible biotype (Figure 6.3C).

The HPLC elution profiles for the two biotypes at24 h after treatment were similar except

the resistant biotype, NAS 4, had more polar metabolites eluting between 4 to 7 min than

did the susceptible (Figure 6.44-8).

It is clear that the resistant biotype is able to convert diclofop acid to more polar compounds

at a faster rate compared to the susceptible biotype. There are two reactions which may be

involved in this conversion. In wheat, diclofop is predominantly metabolized via an aryl-

hydroxylation reaction catalyzed by a cytochrome P450 (McFadden et aI. 1989, Zimmerlin

and Durst 1992). The arylhydroxylated diclofop is then conjugated to sugars and other

compounds (Shimabukuro et aI. 1979). In some susceptible grass species, diclofop is

predominately metabolized via an ester linkage to a glucose moiety (Donald and

Shimabukuro 1980).

106

100

80

60

40

20

0100

80

60

40

20

0100

80

60

40

20

¡t23IclL9XÉ)

-I.laGI-o-Croñ

-.-.-JIGIoEcúLfr

-Ho.-I--€.-L-çr).-â

072

Time after application (h)

Figure 6.34-C Distribution of radioactivity in diclofop-methyl (A), diclofop acid @)

and other metabolites (C) up to 72 h after trearment with I ¡rL of 5 mM [14C]¿iclofop-

methyl in susceptible SAF 19 (O) and resistant NAS 4 (O) wild oat biotypes. Points are

the mean + standard error of three replicates.

0 u 48

B

c

-r.

107

I.--Ira-LGIliù¡

ol¡LG

t¡.l-¡Ic!o.lE6lú

A

susceptible wild oat{.

B

resistant wild oat ,F

0 5 10 15 20 2s

Elution time (min)

30 35

Figure 6.4.4'-8 HPLC elution profiles of extracts from susceptible SAF 19 (A) and

resistant NAS 4 (B) wild oat biotypes}4hafter exposure to [14C]diclofop-methyl dissolved

in a commercial herbicide formulation. ¡l4g¡¿tclofop acid eluted at the retention time of 26

min (*).

108

6.3.3 Effect of Tetcyclasis on Diclofop Metabolism

To test for involvement of cytochrome P450 enzymes in metabolism of diclofop in wild oat

biotype NAS 4, tetcyclasis, a broad-spectn¡m cytochrome P450 inhibitor @rear et al. 1991,

Mougin et al. 199L, Christopher 1993), was applied in combination with diclofop.

Tetcyclasis was applied24hbefore application of diclofop-methyl to wild oat plants. By

24 h following diclofop-methyl application, the susceptible biotype SAF 19 contained

4L.8Vo of the extactable radioactivity as diclofop acid in plants not treated with tetcyclasis

(Tabte 6.1). Pre-teatment with tetcyclasis did not affect metabolism of diclofop acid in this

biotype with4l.l7o of the extractable herbicide as diclofop acid. Conversely, control plants

of the resistant biotype NAS 4 had only 30.17o of the extractable radioactivity as diclofop

acid. Pre-treating these plants with tetcyclasis significantly reduced metabolism of diclofop

acid such that 39.l%o of the extracted radioactivity uras in this form 24 h after treatment

(Table 6.1). In wheat, which can rapidly metabolize diclofop acid to polar metabolites, the

level of diclofop acid in this tissue was unaffected by treatment with tetcyclasis.

An investigation of the HPLC profiles of the susceptible and resistant biotypes showed a

single obvious difference in metabolites (Figure 6.4A-8). The resistant biotype contained

considerably more metabolites eluting between 4 and 7 min than did the susceptible. These

metabolites are much reduced in the resistant biotype after treatment with tetcyclasis such

that there was little difference between susceptible and resistant biotypes in the amount of

radioactivity eluting in the first seven minutes (Figure 6.5). Treatment with tetcyclasis did

not affect the amounts of any other polar metabolites in either biotype. Although tetcyclasis

did not inhibit the overall rate of metabolism of diclofop acid in wheat plants, there was a

significant difference in the metabolites produced in wheat (Figure 6.64-8). In the

tetcyclasis-treated wheat plants, a reduction in the amount of metabolites eluting between 2

and 5 min and an increase in a metabolite with a retention time of 16 min were observed.

109

Table 6.1 The amount of radiolabel as ¡l4ç¡diclofop acid in the susceptible SAF 19

and resistant NAS 4 wild oatbioqpes at24hafter treatment with 5 mlvt ¡149¡¿¡"lofop-

methyl t 20 FM tetcyclasis. Data a¡e the mean I standa¡d deviation of five replicate

experiments.

TreaEnent Biotpe Diclofop acid(Vo)

Diclofop-methyl SAF 19

NAS 4

wheat

Diclofop-methyl + Tetcyclasis SAF 19

NAS 4

wheat

41.8 f 3.3

30.1 + 2.8

to.2 + 2.o

41.1 r 3.5

39.1 r 3.9

11.4 t 3.0

110

rl.A

susceptible wild oat

{qB

resistant wild oat

€.l-:.¡a-LeÉL-.lÊLcl

-.l.l

-IcËo

OIEcÉ

ú

10 15 20 25 30 35

Elution time (min)

Figure 6.54-8 HPLC elution profiles of ll4c]diclofop-methyl and metabolites in

extracts from susceptible SAF 19 (A) and resistant NAS 4 @) wild oat biotypes 24 h after

treatment with ¡l+gldiclofop-methyl. Prior to treatment with diclofop-methyl, the plants

had been exposed to 20 pM tetcyclasis for 24 h. ¡t+çrdt"lofop acid eluted at the retention

time of 26 min (*).

05

lll

A

*

_^

B

{.

r.lËra-LcËLt¡.!lÊLG

È.-a-!Ð

Iclc.-EÉB

ú

05 10 15 20 25

Elution time (min)30 35

Figure 6.64.8 HPLC elution profiles of ll4C]d'clofop-methyl and metabolites in

extracts from wheat cv. Halberd 24 h after exposure to ll4c]d'clofop-methyl (A) or

diclofop-methyl plus 20 ¡rM tetcyclasis (B). ¡t+g¡diclofop acid eluted at the retention time

of 26 min (*).

tt2

6.3.4 Effects of letcyclasis on Plant Growth

Application of 20 pM tetcyclasis to the nutrient solution of hydroponically-grown plants

partially inhibited diclofop acid metabolism in the resistant biotype NAS 4 (Table 6.1). To

test effects on whole plant $owth, tetcyclasis was applied to the roots of hydroponically-

grown plants in combination with diclofop methyl (topically applied to the axil) to determine

whether inhibition of diclofop acid metabolism leads to an increase in phytotoxic effects of

herbicide application. Three experiments were conducted, the first compared biotype

NAS 4 with wheat cv. Halberd which can rapidly metabolize diclofop (Table 6.1).

Diclofop-methyl, applied alone, had no effect on shoot length or shoot dry weight of

biotype NAS 4 (Table 6.2). In wheat, diclofop-methyl application reduced shoot length but

the shoot dry weight was not affected. Tetcyciasis (20 pM), applied alone via the nutrient

solution of hydroponically grown plants, decreased the shoot length of both species but

shoot dry weight was not significantly different from the control (Table 6.2). Tetcyclasis,

applied alone, had no effect on the length and dry weight of roots in either species.

Diclofop-methyl (10 mM), applied alone, had only a small effect on shoot growth in both

NAS 4 and wheat since they a¡e both resistant to the herbicide. Tetcyclasis, when applied

with herbicide, was able to increase the phytotoxic effect of diclofop-methyl on dry weight

and shoot length in wild oat biotype NAS 4 and on shoot length in wheat cv. Halberd,

however, shoot dry weight of wheat when treated with the combination of diclofop-methyl

and tetcyclasis was not significantly reduced. Although there w¿¡s no mortality caused by

the application of diclofop-methyl and tetcyclasis, the shoot length and dry weight of wild

oat biotype NAS 4 and shoot length of wheat were significantly reduced (Table 6.2). Thus,

tetcyclasis does increase the ph¡otoxicity of diclofop-methyl in wild oat biotype NAS 4.

The second experiment compared biotype NAS 4 with a susceptible biotype SAF 19 and

SAS I, a biotype with a highly resistant ta¡get enzyme ACCase, but no increase in diclofop-

methyl metabolism (Chapter 4). In this experiment (Table 6.3), diclofop-methyl alone had

no effect on biotypes NAS 4 and SAS 1 but caused a substantial reduction in shoot and root

dry weight of biotype SAF 19. Tetcyclasis alone reduced shoot length of all three biotypes

113

but root length of the three biotypes were r¡naffected. A combination of diclofop-methyl and

tetcyclasis had a synergistic effect on reduction of shoot and root length and shoot and root

dry weight of biotype NAS 4. Tetcyclasis in combination with diclofop-methyl also

reduced shoot dry weight in this experiment, however, this reduction uras not significantly

greater than treatment with tetcyclasis. In contrast, treatment with diclofop-methyl plus

tetcyclasis showed no synergistic effects on shoot and root dry weight in the susceptible

biotype. The inhibition of dry weight of shoot and root in the susceptible biotype when

treated with the combination of two chemicals was the result of diclofop-methyl alone.

Diclofop-methyl and tetcyclasis combined had a synergistic effect on shoot and root length

and shoot dry weight in biotype SAS l.

The third experiment examined the responses of susceptible biotype SAF 19 and resistant

biotype NAS 4 to different concentrations of diclofop-methyl when applied with tetcyclasis.

Diclofop-methyl applied at all concentrations did not affect total dry weight of plants of the

resistant biotype, however, treatments as low as 2.5 mIVl reduced the dry weight of the

susceptible biotype to about 40Vo of. the control (Figure 6.7). Diclofop-methyl applied at

l0 mM reduced dry weight of the susceptible SAF 19 to less than 20Vo of the control.

Application of tetcyclãsis in combination with diclofop-methyl did not decrease total dry

weight of susceptible SAF 19 over that acheived by diclofop-methyl applied alone. In

contrast, tetcyclasis applied in combination with diclofop-metþl had a synergistic effect on

total dry weight of the resistant NAS 4 biotype. At 10 mM diclofop-metþI, in the presence

of tetcyclasis, total dry weight was reducedby 4OVo compared to no reduction in the absence

of tetcyclasis @igure 6.7).

These in vitro and in vivo experiments demonstrate that enhanced diclofop metabolism

contributes to diclofop-methyl resistance in biotype NAS 4. Tetcyclasis did not increase

phytotoxicity of diclofop, as determined by plant dry weight, in wheat or susceptible

biotype SAF 19 where it does not decrease diclofop metabolism. However, biotype NAS 4

did not become fully susceptible upon treaünent with tetcyclasis as this biotype also contains

a diclofop-resistant ACCase.

t14

Table 6.2 Effects of diclofop-methyl and tetcyclasis alone and in combination on shoot

and root length (cm) and shoot and root dry maner (mg/planQ of wild oat biotype NAS 4

and wheat cv. Halberd grown in hydroponic culture. Values are the means of four

replicates.

Biotype Control Diclofopmethyl Tetcyclasis Diclofopmethyl +Tetcyclasis

Wild oat biotvoe NAS 4

Shoot length 27 a* 24a r8b 14c

Root length 14a 15a 13a 13a

Shoot dry weight lM a 94a 88a 58b

Root dry weight 44 a 45a 47a 34b

Wheat cv. Halberd

Shoot length 27a 22b 2tb l8c

Root length 13a 14a 12a 12a

Shoot dry weight 143 a I33 a ll0a 102 a

Root dry weight 33 a 3tb 4la 29b

*Values within a row followed with the same letter a¡e not significantly different at p <0.05

according to Duncan's multiple range test.

115

Table 63 Effects of diclofop-methyl and teæyclasis alone and in combination on shoot

and root lenglh (cm) and shoot and root dry matter (mg/plant) of susceptible SAF 19,

resistant SAS I and NAS 4 wild oat biotypes grown in hydroponic culture. Values are the

means of fotu replicates.

Biotpe Contol Diclofopmethyl Teûcyclasis Diclofopmethyl + Tetcyclasis

Susceptible biotvpe SAF 19

Shoot length

Root length

Shoot dry weight

Root dry weight

Resistant biotvoe SAS I

Shoot length

Root length

Shoot dry weight

Root dry weight

29 a* 13b

19a lla

96a 55b

49a 2tb

30a 28a

19a 19a

lO7 a 97a

46a 38b

3la 30a

2la 2Oa

109 a l0l ab

48a 4la

Resistant biotype NAS 4

Shoot length

Root lengfh

Shoot dry weight

Root dry weight

18 ab

17a

62 ab

45a

20b

19a

83a

47a

20b

2ta

84 ab

t2b

10a

40b

20b

14c

t7b

59b

43 ab

l6c

17b

65b

49a 30b

*Values within a ro\ry followed with the same letter are not signifrcantly different at p <0.05

according to Duncan's multiple range test.

116

120

-o¡rI-toICHo

èa

I--è0a-e)Þ

LE-cÉ

-oFr

80

40

050 10

Diclofop-methyl (mM)

Figure 6.7 Total dry weight of susceptible SAF 19 (O, O) and resistant NAS 4 (1, tr)

wild oat biotypes grown in hydroponic culture when treated with diclofop-methyl alone

(closed symbols) and in the presence (open symbols) of 20 ttM tetcyclasis (Tet). Diclofopr

tetcyclasis data are a%o of. the tetcyclasis alone to give lOÙVo values. Point symbols are the

mean t standard error of four replicates.

- Tet

+ Tet

+ Tet

- Tet

t17

6.3.5 Inhibition of ACCase Activity by Diclofop Acid in Single Lines of

Biotype NAS 4

Biotype NAS 4 has two mechanisms of resistance to diclofop-methyl, a less-sensitive

ACCase (Chapter 3) and enhanced metabolism of diclofop (Figure 6.3). To establish

whether both mechanistns are present within all individuals of the population, seed from ten

individual plants were collected and each family tested for the presence of a diclofop-

resistant ACCase. These lines were compared to biotype SAF 19, with a sensitive ACCase,

and biotype SAS 1, with a highly-resistant ACCase. As expected, ACCase from SAS 1

biotype exhibited 50-fold less sensitivity to diclofop inhibition compared to that of

susceptible SAF 19 biotype. This level of resistance is simila¡ to reported in Chapter 3

when ACCase from biotype SAS 1 was compared to that of susceptible biotype SAS 2.

Concentrations of diclofop acid giving 507o inhibition (I5g) of ACCase activity from the ten

families of biotype NAS 4 ranged from 1.2 to 3.0 pM. The I59 of ACCase from these NAS

4 lines were between 6 to 15 times greater than that of the susceptible (Table 6.5). In

previous studies, ACCase extracted from the whole population of NAS 4 biotype was found

to be six-fold more resistant to diclofop than the susceptible (Table 3.6). These small

differences of R/S values between the single lines and bulk population suggested that the

population of NAS 4 is homogeneous and all individuals contain a resistant ACCase. The

6- to l5-fold resistance of ACCase to diclofop in the single lines and in the bulk population

of NAS 4 biotype is probably insufficient to account for the high degree of resistance to

diclofop-methyl evident at the whole plant level. For example, biotype NAS 4 shows

greater resistance to diclofop-methyl than biotype SAS I despite possessing an ACCase 3-

to 8-fold less resistant. It is likely that resistance to diclofop-methyl in this biotype involves

two mechanisms, a mutant ACCase enzyme and enhanced metabolism to detoxify the

herbicide, and these two mechanisms of resistance co-exist in all individuals.

118

Table 6.5 Concentration of diclofop acid giving 507o inhibition of ACCase activity (Iso)

from the susceptible SAF 19, resistant SAS I and l0lines of resistant NAS 4 (#1410) wild

oat biotypes. Ratios are the value for the resistant divided by the value for the susceptible

biotype. ACCase activity was measured in partially purified extracts of the shoot

meristematic regions. Data are the mean t standard error of four replicates.

Biotype Iso 0¡¡¡) R/S Ratio

Susceptible SAF 19

Resistant SAS 1

Resistant NAS 4

#l

#2

#3

H

#5

#6

#7

#8

#9

#10

0.20 r 0.04

9.96 t 1.1

2.48 + 0.5

3.01 r 0.7

2.08 + 0.6

1.96 + 0.5

1.38 r 0.6

1.28 + 0.5

1.19 r 0.6

1.98 + 0.4

1.82 + 0.6

2.11 r 0.6

49.8

12.4

15.0

10.4

9.8

6.9

mean (#1+10)

6.4

5.9

9.9

9.1

10.5

1.93 + 0.18 9.6 + 2.6

119

6.3.6 Effect of Diclofop Acid on Plasma Membrane Potential

The ability of diclofop acid to depolarize plasma membrane potentials in etiolated coleoptile

cells of susceptible SAF 19 and resistant NAS 4 wild oat biotypes was examined. Plasma

membrane potentials of untreated coleoptile cells from susceptible and resistant biotypes

were between -110 to -l15 mV (Figure 6.8). Diclofop acid (50 pM) depolarized plasma

membrane potentials from both biotypes within 15 min to between -50 and -60 mV (Figure

6.3A-8). After removal of herbicide, plasma membrane potentials from both biotypes

stayed depolarized for at least 20 min. The resistant biotype NAS 4 was unable to repolarize

the plasma membrane potential following diclofop acid-induced depolarization.

6.4 Discussion

rWild oat biotype NAS 4 shows high level resistance to diclofop-methyl, but possesses only

a moderately-resistant ACCase (Chapter 3). Studies revealed that there were no differences

in uptake and translocation of diclofop-methyl between susceptible and resistant biotypes

(Figures 6.L and 6.2). Nor does this biotype possess a mechanism which allows rapid

repolarization of the plasma membrane potential (Figure 6.8). Biotype NAS 4 possesses an

ACCase with 6-fold resistance to diclofop acid (Chapter 3). In addition, this biotype

exhibits an increase in the rate of diclofop-methyl metabolism (Figure 6.34-C). A similar

result was reported for L. rigidum biotype SLR 3l which exhibited a l.5-fold increase in

metabolism of diclofop-methyl (Holtum et al. 1991). In vivo studies with whole plants

show that the level of diclofop-methyl resistance in NAS 4 biotype is reduced in the

presence of the cytochrome P450 inhibitor, tetcyclasis, whereas tetcyclasis has little effect

on the response of the susceptible biotype to diclofop-methyl (Table 6.3). The rate of

metabolism of diclofop acid in biotype NAS 4 decreased in the presence of tetcyclasis.

'When plants were treated with tetcyclasis 24 h prior to application of ¡l4c1diclofop-methyl,

metabolism of diclofop in the resistant biotype was reduced to the level of the susceptible

biotype (Table 6.1, Figures 6.5 and 6.7). These results suggest that the enhanced

metabolism of diclofop-methyl observed in biotype NAS 4 is probably due to a cytochrome

t20

-120

-E-cl.---I(¡)

-ofe)-tcl¡i

--IIq)|=À

.100

-E0

-60

-120

-100

-80

-60

0

00 10 20

Time (min)30

Figure 6.84-B The effect of 50 pM diclofop acid on plasma membrane potentials in

coleoptile cells of the susceptible SAF 19 (A) and resistant NAS 4 (B) biotypes. Arrows,

addition (+) and removal C) of diclofop acid (DA). The data shown is from a single

coleoptile and is representative of ten coleoptiles each of the susceptible and resistant

biotypes.

+DA

üA

DA

,I

- +DA

ü B

.DA

+

,I

tzt

P450 monooxygenase. Although wheat contains a P450 eni¿)me able to rapidly metabolize

diclofop-methyl (Zimmerlin and Drust 1990 and 1992) and tercyclasis was able to enhance

the phytotoxicity of chlorotoluron towa¡d wheat (Mougin et al. 1991), diclofop metabolism

in wheat cv. Halberd was not affected by tetcyclasis (Table 6.1). It may be that tetcyclasis

is not an effective inhibitor of diclofop metabolism in wheat or the concentration used was

not sufflrcient to inhibit the detoxification of diclofop in wheat. Specific herbicide-inhibition

interactions were evident in L. rigidum where tetcyclasis does increase chlorotoluron

phytotoúcity in biotype WLR 2 but does not affect the phytotoxicity of chlorsulfuron in

biotype SLR 3l (Christopher et al. 1994). Nevertheless, more work has to be done to

establish the direct involment of P450 enzymes in enhanced metabolism of diclofop-methyl

in biotype NAS 4.

It is likely that two mechanisms of resistance are present in each individual of the NAS 4

population (Table 6.4). The combined contributions of enhanced metabolism and mutant

ACCase enzyme to herbicide resistance may be required to explain the high degree of

resistance at the whole plant level. Hence, it is probable that both mechanisms contribute to

diclofop-methyl resistance in this biotype. Resistance to diclofop-methyl endowed by a

resistant ACCase tn A. sterilis and Z. multiforum is conferred by a single, nuclear encoded,

incompletely dominant gene (Barr et al.1992, Betts et al.1992). In case of biotype NAS 4

which has two mechanisms responsible for resistance, there must be more than one gene

involved. Definite proof of a number of genes conferring resistance in this biotype has yet

to be determined.

Resistant ACCase is a mechanism frequently found in species resistant to ACCase-inhibiting

herbicides, whereas non-target site resistance mechanisms, such as enhanced metabolism of

herbicide, a¡e seldom observed. However, a 1.5 fold increase in the rate of metabolism of

diclofop-methyl has been observed in one biotype of L. rigidum (Holtum et at. 1991) which

probably confer at least partial resistance in this biotype on its own. Where resistant weeds

have at least two mechanisms of resistance, target site-based resistance will most often be

the most significant mechanism (Devine and Shimabukuro 1994). Therefore, biotype

t22

NAS 4 not only has an insensitive target ACCase, albeit not sufficient to confer the high

level of resistance to diclofop-methyl at the whole plant, but it also has an increased capacity

to metabolize diclofop-methyl. Together, these two mechanisms enable the NAS 4 plants to

resist high rates of the herbicide. This is the first evidence for wild oat having two

mechanisms of resistance (target and non-target site based) to ACCase-inhibiting herbicides.

ln L. rigiduzr, biotype SLR 3l has three mechanisms of resistance, an enhanced metabolism

to detoxify diclofop-methyl, an alteration of plasma membrane, and resistant ACCase

(Häusler et al. 1991, Holn¡m et al. 1991, Tardif and Powles 1994). However, orùy a L27o

subset of this biotype contained the insensitive ACCase (Tardif and Powles 1994). In

contrast, it is likely that both mechanisms of resistance are present in all individuals of

biotype NAS 4.

t23

Chapter 7

Non-Target Site Mechanism Conferring Resistance to ACCase-

Inhibiting Herbicide in Biotype NAS 14

7.1 Introduction

As described in Chapter 3, resistant wild oat biotlpe NAS 14 has a low level of resistance to

ACCase-inhibiting herbicides and yet contains a sensitive target ACCase enzyme

(Chapter 3). Thus, resistance to herbicides in this biotype must involve mechanisms other

than changes at the,target site. For example, changes to the leaf cuticle may reduce herbicide

uptake by leaves, or herbicide may be prevented from reaching the target site by reduced

translocation, enhanced herbicide detoxification or increased sequestration.

Reduced absorption or translocation of ACCase-inhibiting herbicides have not been reported

to be mechanisms of herbicide resistance. Foliar absorption of diclofop-metþl occurred

readily in susceptible and resistant plants with little difference between them (Donald and

Shimabukuro 1980, Dahroug and Muller 1990, Holtum et al. 1991, Shimabukuro and

Hoffer 1991, Devine et rl. 1992). Once in the plant, translocation of radioactivity from

[1aC]diclofop-methyl out of the treated area is limited (Brezeanu et al. 1976, Boldt and

Putnam 1980, Hall et al. 1982, Jacobson and Shimabukuro 1982, Baker and Chamel 1990).

Although reduced uptake and ranslocation have not been reported to be major factors in the

selective action of diclofop-methyl and other APP herbicides in the earlier studies, they could

contribute to low-level resistance in biotype NAS 14.

To determine the mechanism conferring low level resistance to ACCase-inhibiting herbicides

in biotype NAS 14, studies of uptake, translocation and metabolism of diclofop-methyl were

conducted. The effect of diclofop acid on the plasma membrane potential was also

investigated.

124

7.2 Materials and Methods

7.2.! Uptake, Translocation and Metabotism of ll4C]pictofop-Methyl

The protocol for uptake, translocation and metabolism of herbicide in both susceptible

SAF 19 and resistant NAS 14 wild oat biotypes are described in Chapter 2.

7.2.2 Absorption of lr4c]Diclofop-Methyl to Leaf Cuticle

To determine the amount of herbicide absorbed into the leaf cuticle, a 1 pL drop of 5 mM

[1aC]diclofop-methyl was smeared on the adaxial surface of the second leaf of two-leaf

seedlings. External radioactivity was removed by washing treated leaves in 5 mL of

methanol: H2O: Triton X-100 (20:79.99:0.01 by volume) and blotted dry. The radiolabel

in the cuticle of treated leaf was subsequently removed by washing in 4 mL of chloroform

for 2 min (Tucker et al. 1994). Radioactivity in both the methanol and chloroform washes

was determined by LSS. Plant tissue was oxidized in the Biological Sample Oxidizer.

7.3 Results

7.3.1 Uptake and Metabolism of ll4c]Diclofop-Methyl

The absorption of llaC]diclofop-methyl of biotype NAS 14 was similar to that of the

susceptible SAF 19 (Figure 7.1). At 6 h following application,4T and45Vo of the applied

radioactivity had been absorbed by susceptible and resistant biotypes, respectively and

herbicidal absorption was at a ma¡cimum (80Vo) by 36 h after application in bothbiotypes.

ll4c]diclofop-methyl was metabolized at an equal rate in both susceptible and resistant

biotypes (Figure 7.2 A). Radioactivity rapidly accumulated in diclofop acid with about 55Vo

of radiolabel as diclofop acid by t h after treatment (Figure 7.28). The amount of

[laC]diclofop acid, the phytotoxic form, reached a maximum at 6 h after application and

contributed 81 and 86Vo of the total radioactivity in susceptible SAF 19 and resistant NAS 14

bioqpes, respectively. Diclofop acid was converted to more polar metabolites from 6 h after

application at an equal rate in both biotypes (Figure 7.2C).

t25

100

EO

60

40

20

Eq).--ÈÈGI

-clIo€Ê-io

èe

J.I.I-Icgo.-EcÉ

ú0

24 4E

Time after application (h)

72

Figure 7.1 Uptake of l14c]d'clofop-methyl in the susceptible SAF 19 (O) and resistant

NAS 14 (O) wild oat biotypes when applied to the leaf axil of the two-leaf stage seedlings

cultued in soil.

0

126

100

80

60attù¡H40Ëõ20-aI

ãoË 1ooft

:80èr

;60a-

840G,

€206lLr0; roooËBoÊ

.Ë 60

ô40

20

00 u48

Time after application (h)

72

Figure 7.2A-C Percentage of radioactivity extracted as diclofop-methyl (A), diclofop

acid (B) and turther metabolites (C) in the susceptible SAF 19 (O) and resistant NAS 14 (O)

wild oat biotypes from 0 to 72h after treatment. Each point is the mean of three replicate

experiments. Vertical bars represent the standa¡d erors and a¡e sometimes obscured by

symbols.

A

B

c

r27

7.3.2 Translocation of ll4C]nictofop-Methyl

Herbicide was absorbed into the leaves of the susceptible and resistant biotypes at the same

rate @igrue 7.1) and metabolism of diclofop in both biotypes was similar (Figure 7.zA-C).

One other possible mechanism of resistance is that the herbicide may be prevented from

reaching the target site. Translocation of tlaCldiclofop-methyl was examined when

herbicide was applied on the leaf axil. By 6 h after application, 9.5 and 12.8 7o of the

radioactivity absorbed was found in the meristematic zone of resistant and susceptible

biotypes, respectively and this decreasedto 5Vo by 96 h after treatnent (Figure 7.3A). More

than 807o of the radioactivity remained in the stem of both susceptible and resistant biotypes

over the time course of the experiment (Figure 7.38). More radioactivity was translocated to

the leaves of resistant NAS 14 than that of susceptible by 96 h after the ¡lag¡*clofop-methyl

had been applied to the leaf axil of two-leaf seedlings (Figure 7.3C). This may be a result of

the growth of the second leaf and third leaf of NAS 14 biotype being less affected by

herbicide application than the growth of the second leaf of the susceptible SAF 19 biotype

where the thftd leaf did not appear. The distribution of radioactivity in the stem and the root

were similar for both biotypes (Figure 7.3 B, D).

The small but nevertheless significant difference in the amount of radioactivity in the

meristem of the two biotypes (p<0.05) (Figure 7.3A) may be sufficient to explain the low

level of resistance in biotype NAS 14, however, more evidence is required to support this

conclusion.

7.3.3 Absorption of Herbicide to Leaf Cuticle

Although there are no differences in herbicide absorption into the leaves of the two biotypes,

more herbicide may be bound by the leaf cuticle of the resistant plants and therefore not

available for translocation to the meristem. To test this hypothesis, an experiment was

performed where herbicide was smeared onto the leaf surface and after harvest the leaf was

successively washed with water/methanol, to remove unabsorbed herbicide, and

chloroform, to remove the cuticular waxes. No significant differences in the amount of

t28

0061218z 4896

Time after application (h)

Figure 7.34.D Distribution of radioactivity from ¡l4çldiclofop-methyl in the shoot

meristematic region (A), stem (B), leaf (C) and root (D) in susceptible SAF 19 (O) and

resistant NAS 4 (O) when applied to the leaf axil of nvo-leaf stage seedlings.

20

10

100

50

0€),¡É

ÉË

-È¡l--GI!¡¡

o9CHo

Èa

È.-.!lttIÉtlo.-

rÉcË

ú

0

10

0

10

A

h

or9B

t29

[laC]diclofopmethyl retained on the leaf surface or in the cuticle of susceptible and resistant

biotypes was observed (Figure 7.4). Radioactivity from [l4C]dictofop-methyl was

distributed similarly in the external fractions of treated leaves. By 48 h after application,

4OVo of radioactivity was recovered in the \ilater/methanol wash (Figure 7.4A) and 57o of

radioactivity was extracted from the cuticle by chloroform @igure 7.48). The remainder of

the radioactivity was found within the leaf. These experiments indicate that passage of

diclofop-methyl through the cuticle does not appear to be impeded in the resistant biotype.

7.3.4 ACCase Activity fz Vivo

Previous experiments (Chapter 4, Figure 4.7) have shown that it is possible to treat Avena

seedlings with diclofop-methyl and then extract ACCase to determine the in vivo effect of

herbicide on ACCase activity. This technique was employed with biotype NAS 14 in which

a 1 ttJ- droplet of 5 mM diclofop-methyl was applied to the leaf a¡ril of 2-leaf seedlings. By

6,71 and 72 h following application, ACCase activity in the crude extract of meristematic

region from both biotypes was assayed. ACCase activity in both biotypes was inhibited

following application of the herbicide, however, the resistant biotype NAS 14 had more

extractable ACCase activity 6 h after herbicide application than did the susceptible biotype

SAF 19 (Figure 7.5). Some diclofop-methyl reached the meristems in both biotypes as

ACCase extracted from treated plants were inhibited by 45 and 35Vo in the susceptible and

resistant biotypes, respectively. As both biotypes contain sensitive ACCase as shown in

Table 3.10, ACCase activity from both biotypes remained inhibited 24 a¡d 72 h following

herbicide application. However, the greater inhibition of ACCase in vivo in the susceptible

biotype would suggest that more herbicide reaches the meristem in this biotype compared to

the susceptible. The result supports the conclusions from the translocation experiments in

that slightly less herbicide appears to be translocated to the meristematic regions and reach

the active site in the resistant biotype.

130

100

EO

60

40

20ELq)

oIq)L

CHo

èa

-.!lal

-c)cË

.-EGú

0

10

0100

80

60

40

20

06 12u

Time after application (h)

4E

Figure 7.4A-C The amount of [lag¡¿iclofop in methanol wash (A), chloroform wash

(B) and plant tissue (C) of susceptible SAF 19 (tr) and resistant NAS 14 (f) wild oat

biotypes following treatment with 5 mM ¡t+6¡¿iclofop-methyl. Data a¡e the mean of five

replicate experiments. Vertical bars represent the standard errors of the means.

A

B

TIc

I

131

u48Time after application (h)

72

Figure 7.5 Extractable ACCase activity from the meristematic region of susceptible (O)

SAF 19 and resistant (O) NAS 14 wild oat biotypes after treatment with I ttl. of 5 mM

diclofop-methyl to the leaf axil of twoJeaf stage seedlings. Data presented as percentage of

ACCase from untreated plants. Vertical ba¡s are the standard deviations of four replicates.

100

EO

60

40

20

-o¡rIHcI

tto

Èa

ìa-

.-rCJGI(¡)aÀ6tOU

00

132

7.3.5 Effect of Diclofop on Plasma Membrane Potential

The plasma membrane potential of coleoptile cells from resistant biotype NAS 14 was

depolarized by diclofop acid in the same manner as that of the susceptible biotype (Figure

7.64-B). In the presence of 50 pM diclofop acid, the electrochemical plasma membrane

potential in susceptible SAF 19 and resistant NAS 14 biotypes raprdly depolarized from -110

to -120 mV after a lag period of I to 5 min (Figure 7.6). Following herbicide removal, the

plasma membrane potential from both biotypes remained depolarized with a potential

between -50 and -60 mV. No recovery in polarity was observed in either biotypes for up to

60 min after herbicide removal (data not shown).

7.4 Discussion

The low level of diclofop resistance evident in biotype NAS 14 (Table 3.10) is not due to

possession of an insensitive ACCase, an alteration in the response of the plasma membrane

to diclofop acid, reduced herbicide absorption or an enhanced rate of herbicide metabolism

(Figures 7.1, 7.3 and 7.5). The only difference observed between these biotypes is a

reduction in herbicide content of the shoot meristematic zone following application of

diclofop-methyl to the leaf axil of two-leaf plants @gure 7 .2A). This difference in herbicide

distribution is reflected in measurements of ACCase activity in the shoot meristematic region

at 6 h following herbicide application, which showed ACCase was inhibitedby 45Vo in the

susceptible biotype and by 35Vo in the resistant biotype (Figure 7.5). This indicates that

herbicide might not reach the ACCase in the meristem as readily in the resistant biotype

compared to the susceptible.

The presence of diclofop-methyl in the shoot meristem is crucial to its herbicidal action. For

example, application of diclofop-methyl on or below the meristematic region caused death of

wild oat and barley whereas application to the leaves did not (Friesen et al. 1976).

Translocation of diclofop-methyl in plants is limited and usually less than SVo of ndioactivity

absorbed is translocated out of the treated area (Boldt and Putman 1980, Donald and

Shimabukuro 1980, Hall et al. 1982, Jacobson and Shimabukuro 1982, Dahroug and Muller

133

-100

-80

-60

-120

-140

-120

.100

-80

-Ha-6ta-

--to)9oÈ(Ð-lI(Ëli¡-Tte)¡ià

0

-60

00 10 20

Time (min)30

Figure 7.6^-B The effect of 50 ¡rM diclofop acid on plasma membrane potential of

coleoptile cells of the susceptible SAF 19 (A) and resistant NAS 14 (B) wild oat biotypes.

Arows, addition (+) and removal G) of diclofop acid (DA). The data shown is from a

single coleoptile and is representative of ten coleoptiles each of the susceptible and resistant

biotypes.

+DA

ü A

.DA

,

+DA

ü B

DA

)I I

t34

1990, Baker and Shamel 1990). However, translocation of diclofop-methyl to the shoot

meristem can be increased by placement of the herbicide. A greater amount of diclofop-

methyl and its metabolites were found in the meristematic zone when the herbicide was

placed closer to the leaf base of wild oat seedlings (Walter et al. 1980). In general, little of

the applied herbicide reaches the shoot meristem and therefore, small decreases in

translocation may have a large impact on plant suntival.

In conclusion, the small difference in translocation of diclofop between the susceptible and

resistant biotypes may be suffrcient to confer the low level of resistance to herbicides in the

NAS 14 biotype, however, at present this cannot be substantiated.

135

Chapter I

Investigations of Alterations of Membrane Resporæes to Diclofop

as Mechanism of Resistance in \ryild Oat

8.1 Introduction

As presented in the previous chapters, resistance to APP and CHD herbicides in the range of

biotypes of wild oat studied can involve a modified forrr of ACCase enzyme, enhanced rate

of herbicide metabolism or retarded translocation of herbicide. However, membrane

depolarization by diclofop has been proposed to be an alternative mechanism of resistance to

ACCase-inhibiting herbicides (Shimabukuro and Hoffer 1992). Initially, diclofop was

presumed to act as a protonophore in the plasma membrane (rWright and Shimabukuro

1987). Later, this hypothesis was revised by Shimabukuro and Hoffer (1992) who

suggested that the collapse of the proton motive force in response to diclofop acid is due to

the specific interaction of diclofop with a plasma membrane protein. APP and CHD

herbicides have the ability to depolarize plasma membrane potentials in a number of

susceptible and resistant species (Lucas et al. 1984, rWright and Shimabukuro t987,

DiTomaso et al. 1991, Häusler et al. 1991, Holtum et al. 1991, Shimabukuro and Hoffer

t992, Devine et al. L993, DiTomaso 1993, Dotray et al. 1993). However, in some

diclofop-resistant weed biotypes, repolarization of the plasma membrane potentials occurred

following removing herbicide from the treatment solution (Häusler et al. 1991, Holtum et al.

199I, Shimabukuro and Hoffer l992,Devine et al. 1993). Despite these studies, it is still

unclear how membrane repolarization following herbicide removal in the resistant biotypes is

related to resistance under field conditions since plasma membrane potentials from resistant

crops (e.g. wheat and pea) are also depolarized by diclofop acid (Wright and Shimabukuro

1987, DiTomaso et al. 1991, DiTomaso 1993). In addition, a full recovery of membrane

polarity has been observed in both susceptible corn and resistant pea root cells following

herbicide removal (DiTomaso et al. 1991).

136

A number of membrane-related phenomena have been investigated in resistantAvenafatua,

Loliutn rigidum and crop species. It has been found that diclofqp-induced depolarization

and subsequent repolarization of the membrane potential in susceptible and resistant I.

rigídumis pH dependent (DiTomaso 1993, Holtum et al. 1994). Differences in the ability of

susceptible and resistant L. rigidum and A. fatua roots to acidify the external media were

also noted (Häusler et al. 1991, Devine et al. 1993). The roots of the resistant L. rigidum

biotypes exhibited a decreased rate of acidification of the external solution relative to the

susceptible biotypes (Häusler et al. l99l).

In this chapter, the effects of diclofop acid on depolarization and repolarization of plasma

membrane potentials in coleoptiles of five resistant biotypes of wild oat were investigated.

In addition, other membrane-related experiments such as the effect of pH on recovery of the

plasma membrane potential following removal of diclofop acid and the ability of intact roots

of these biotypes to acidify the external solution were also conducted to ascertain whether an

altered membrane response is a mechanism of resistance in these resistant wild oat biotypes.

8.2 Materials and Methods

8.2.1 Effect of pH on Membrane Response to Diclofop Acid

Etiolated coleoptiles of susceptible and resistant wild oat biotypes were used. The

experiments \ryere conducted in the same manner as described in section 2.7. The basal

solution was replaced with 50 pM diclofop acid for 15 min. The herbicidal solution was

then replaced with the basal solution buffered by 2.5 mM MES-Tris (pH 5.6, 5.8, 6.0,6.2,

6.4,6.5). Five to six etiolated coleoptiles were used for each treatment.

E.2.2 Effect of Diclofop Acid on Acidification of External Solution

One day after root emergence, germinated seeds of the resistant (SAS 1, NAS 4, NAF 6,

NAF 14, SAF 14, NAF 8, NAS 6, SAF 34, V/AF 12) and susceptible (SAS 2, SAF 19)

biotypes of wild oat were placed on filter paper Ekwip@ R6 (Robert Bryce, NSW, Australia)

soaked with distilled water. Filter papers were rolled up, wrapped with plastic sheet and

r37

placed vertically in a 500-mI beaker. Plants were kept in the growth room under the same

conditions as described in section 2.2.2. For each biotype, roots of l0 intact wild oat

seedlings (300450 mg F'W) at the two-leaf stage were washed with distilled water before

being placed in a 10-rnl. beaker containing 4mL of basal medium (1 rnlvf KzSO¿ and I rnlvf

CaSO¿). Root-induced pH changes in solution urere measured by an Action mini pH

combination AglAgCl elecrode connected to a Beckman pH meter. Technical grade diclofop

acid was dissolved in acetone and diluted with the treatment solution to obtain a 50 pM

herbicide solution with<l%o acetone. After a I h treatnent, the basal solution was replaced

with the herbicide solution. After I h exposure to herbicide, the herbicide solution was

replaced with the basal solution. Changes in pH were recorded at room temperature.

Experiments were repeated three times and the data were pooled.

8.3 Results

8.3.1 Effect of Diclofop Acid on Plasma Membrane Potential

The response to diclofop acid of plasma membrane potentials from etiolated coleoptiles cells

of resistant and susceptible biotypes are shown in Figure 8.14-F. Initial membrane

potentials in coleoptile cells from susceptible and resistant wild oat were between -100 and

-120 mV. In the presence of 50 ttJvf diclofop acid, rapid depolarization of the membrane

potential occurred reaching a minimum potential of between -50 and -70 mV within 15 min.

Upon removal of herbicide, membrane potentials from all biotypes remained depolarized

(Figure 8.14-F). No recovery of plasma membrane potential was evident in any of these

resistant wild oat biotypes.

A correlation between recovery of the plasma membrane potential following diclofop acid-

induced depolarization and resistance to APP herbicides was reported in some resistant

biotypes of L. rigidurn (Häusler et al. 1991, Holtum et al. 1991, Shimabukuro and Hoffer

1992, DiTomaso 1993) andA. fatua (Devine et al. 1993). These resistant biotypes had the

ability to recover from diclofop acid induced depolarization whereas the susceptible biotypes

did not. To verify that plasma membrane potential could recover in a resistant L. rigidum

138

biotype and to check that (in the same system) resistant wild oat biotypes could not, the

experiments u,ere also conducted with L. rigidum. In this study, coleoptiles of resistant l.

rigidum biotype SLR 3l showed a recovery of membrane potentials whereas membrane

potentials remained depolarized in the susceptible biotype VIR 1 following removal of

diclofop acid from the bathing solution (Figure 8.2). This result confirms the observations

of Holtum et al. (1991) and provides additional evidence that this recovery phenomenon

does not (rccur in the resistant wild oat biotypes studied here.

8.3.2 Effects of pH on Plasma Membrane Repolarization

The experiments conducted in section 8.3.1 were performed in Higinbotham's solution

(Higinbotha¡n et al. 1970) at pH 5.8. The recovery of membrane polarity following removal

of ACCase-inhibiting herbicide has been reported in resistant biotypes of A. fatua (Devine et

al. 1993) and l,. rigidum (Häusler et al. 1991, Holn¡m et al. 1991, Shimabukuro and Hoffer

1992) at pH 5.7. However, a susceptible L. rigidum biotype also showed membrane

repolarization upon removal of herbicide when the basal solution was buffered at pH 6.0

(DiTomaso 1993). In this case, pH seems to be an important factor in membrane

repolarization. To determine whether pH has an effect on membrane response of wild oat

biotypes, the repolarization solution was buffered with 2.5 mM MES, at a range of pH from

5.6 to 6.5. Coleoptile cells of two susceptible and two resistant biotypes were examined.

At pH 5.6, none of the biotypes demonstrated any recovery of plasma membrane potential

following removal of the herbicide (Figure 8.3). As the pH increased partial repolarization

of the membrane potential was observed in all biotypes and at pH 6.5 full repolarization of

the membrane potential was observed in all biotypes. There were some slight differences

between biotypes with biotype NAF 8 showing repolarization of the membrane potential at a

lower pH than the other biotypes. This biotype had reached 50Vo repolarization at pH 6.0

compared with pH 6.2 f.or biotype SAF 19 and SAS 1 and at pH 6.3 for biotype SAS 2

(Figure 8.3).

139

-120

-80

-120

-100

-80

.60

-12.0

-120

.100

.80

.60

-100

0

0

-40

0---I¡-6t

--H(¡)+¡oÈ()-t(Ë¡r¡-It()tià

-120

-100

-80

.60

-120

-80

-60

-100

-80

-60

00

001020 300

Time (min)10 20 30

Figure 8.14.F The effect of 50 pM diclofop acid on plasma membrane potentials in

coleoptile cells of resistant biotypes NAF 8 (A), SAF 14 (B), SAF 34 (C), WAF 12 (D),

NAS 6 (E) and susceptible biotype SAF 19 (F). Anows, addition (+) and removal G) of

diclofop acid (DA). The data shown is from a single coleoptile and is representative of ten

coleoptiles each of the susceptible and resistant biotypes.

trrttll

+DA

B

DA

,Itrrlllll

+A

DA

7

rrttllll

+DA.DA

c

artrttlll

+DAü D

DA

+DA

.DA

,rrrttlt

+DA

.DA

aI

140

-120

-100 +DA

A

susceptible L rigidum

.DA

ü B

resistant L rigìdum+DA

10 20

Time (min)30

Figure 8.24-B The effect of 50 pM diclofop acid on plasma membrane potentials in

coleoptile cells of the susceptible VLR I (A) and resistant SLR 31 (B) ¿. rigidum. Arrows,

addition (+) and removal (-) of diclofop acid (DA). The data shown is from a single

coleoptile and is representative of five coleoptiles each of the susceptible and resistant

biotypes.

-It-cg.--Iq)IoÈÉ)-LcÉ¡rÊÅIIc)l!(a

-E0

.60

40

0

DA

100

-80

-60

.40

00

100

80

60

40

20

0

-Ho.-€N.-LG-oÈé)E

CHo

èa

Lé)

oCJê)ú

r4l

0 5.6 5.8 6.0 6.4 6.6

Figure E.3 Repolarization of plasma membrane potentials in cells of etiolated coleoptiles

of susceptible SAF 19 (tr) and SAS 2 (O) and resistant NAF 8 (I), SAS I (O) wild oat

biotypes as a function of pH of the bathing solution during the recovery phase.

Repolarization is presented as a percentage of the fuIl depolarization. Data are the mean tstandard error and calculated on steady-state potentials measured I h afrer the removal of 50

¡tM diclofop acid.

6.2

pH

142

From the results in Figure 8.3, it is likely that pH affects the ability of plasma membraries to

repolarize following removal of diclofop acid. To confirm this observation, single

coleoptiles of three resistant and nvo susceptible biotypes were bathed in 50 pM diclofop

acid for 15 min prior to placement in a herbicide-free solution buffered with 2.5 mM MES at

pH 6.5. After 20 min, the sa¡ne coleoptile was treated a second time with diclofop acid for

15 min and the herbicide was replaced with a buffered solution at pH 5.8. The results

indicated that an initial diclofop acid-induced depolarization of the membrane potential was

reversible in the herbicide-free solution buffered at pH 6.5 @igures 8.4 to 8.5). The second

exposure to diclofop acid caused a similar depolarization to the first, but the membrane

potential did not recover in a herbicide-free solution buffered at pH 5.8. Biotype NAF 8

showed a full recovery of membrane depolarization within 20 min when the herbicide

solution was replaced with solution buffered at pH 6.5 whereas in the other biotypes

recovery of membrane polarity was only 60to80Vo of the initial value.

8.3.3 Acidification of the External Solution by Roots of Intact Plants

The ability of roots of wild oat biotypes to acidify the external medium was examined by

placing ten wild oat plants at the two-leaf stage in an unbuffered solution. The pH of the

external solution was decreased by about I unit within I h (Figure 8.6). There were no

differences in the rate of acidification in the susceptible SAS 2 biotype compared to the

resistant SAS I biotype. The pH of the solution slowly decreased when the roots of

susceptible and resistant plants were exposed to 50 ¡tM diclofop acid, decreasing by about

0.2 pH units over I h. Following herbicide treatment, the roots were washed and

transferred to a herbicide-free solution. The roots of both biotypes continued to acidify the

extemal solution, however the rate of acidification was much reduced compared to the initial

acidification @gure 8.6).

Similar results were obtained from a range of other susceptible and resistant biotypes of wild

oat (Table 8.1). The roots of all biotypes tested readily acidified the pH of the external

solution which decreased by 2 to 2.5 pH units g-1 root FW over a I h period. Addition of

t43

-120

120

20 30 40 50 60 70 80

Time (min)

Figure 8.44-B Electrochemical potentials measured in coleoptile cells of the susceptible

wild oat biotypes SAF 19 (A) and SAS 2 (B). Arows, addition (+) and removal (-) of

diclofop acid (DA) from the bathing solution. Following removal of diclofop the solution

was buffered at either pH 6.5 or pH 5.8. The data shown is a single coleoptile and is

representative of five coleoptiles each of the two susceptible wild oat biotypes.

.100

.80

-60

-tT-GIa-

--té)IoÈq)-TcÉ¡r¡-:Ic)Hà

0

100

-80

-60

0100

+DA+

A

6.5

DA 5.8

DA

,rrrtttttllttttll

+DA

B+DA

pH 6.5pH 5.8

,+ DA

ltllttlttltlttlt

+DA

pH 5.8

+

A

6.5

+DA

pH 6.5

pH 5.8

.DA

B

+DA DA

c

5.86.5

t4

-t2x)

.100

-80

0

120

-120

20 30 40 50 60 70 80

Time (min)

Figure 8.54-C Electrochemical potentials measured in coleoptile cells of three resistant

wild oat biotypes NAS 4 (A), SAS I (B) and NAF 8 (C). Arrows, addition (+) and removal

C) of diclofop acid (DA) from the bathing solution. Following removal of diclofop acid the

solution was buffered at either pH 6.5 or pH 5.8. The data shown is a single coleoptile and

is representative of five coleoptiles each of the th¡ee resistant wild oat biotypes.

-60

100

-80

-60

0

-LI

-cla-€-I€)t¡oÈ(¡)tcÉL¡IIq)

-100

-80

-60

0100

t45

diclofop acid to the solution resulted in a reduced rate of acidification for all biotypes in a

range of 0.3 to 0.6 pH units g-1 root F\M h-1. When roots were transferred to herbicide-

free solution, acidification of the external solution was about the sarne, 0.4 to 0.7 pH

units g-1 root F-\M ¡-1, as in the presence of diclofop acid. These results indicated that the

roots of susceptible and resistant biotypes had an equal ability to acidify the pH of external

solution and diclofop acid prevented the acidification of the external medium in all biotypes.

The roots of the susceptible L. rigidumbiotype rü,R 1 acidified the external medium atl.27

pH units g-1 root F'\M h-l Gable 8.1). In contrast, roots of the resistant biotype SLR 3l

acidified the external solution at a much slower rate of 0.16 pH units g:1 root FlV h-l. These

results a¡e similar to those reported previously (Häusler et al. 1991, DiTomaso 1993).

Addition of 50 pM diclofop acid reduced acidification in both biotypes and caused an

alkalization of the medium in biotype SLR 31. Upon removal of the herbicide, acidification

of the susceptible biotype VLR 1 remained inhibited whereas acidification resumed in the

resistant biotype SLR 31 to the same rate as that prior to teatment.

8.4 Discussion

Membrane repolarization following diclofop acid-induced depolarization was observed in the

resistant L. rigidum biotype SLR 3l but not in the susceptible biotype VLR I (Figure 8.2A-

B). Similarly, the ability of the diclofop-resistant biotypes to overcome the effects of

diclofop acid on plasma membrane depolarization were reported for resistant biotypes of

A. fatua (Devine et al. 1993) and L. rigidum (Häusler et al. 1991, Holtum et al. 1991,

Shimabukuro and Hoffer 1992, DiTomaso 1993). It has been suggested that this recovery

of membrane polarity following the removal of herbicide is associated with resistance to

ACCase-inhibiting herbicides (Häusler et al. 1991, Devine et al. 1993). This phenomenon

did not occur in the resistant wild oat biotypes examined here where the membrane remained

depolarized following removal of diclofop acid in both susceptible and resistant biotypes

(Chapters 4 to 7 and Figure 8.14-E). Clearly, repolarization of the plasma membrane

t46

7.0

l̂- oa

120

Time (min)

Asusceptible wild oat

Bresistant wild oat

180

6.0

HÊÈ 0

5.0

7.0

6.0

5.0

+DA

+DA

Al- pe

00 60 u0

Figure 8.6 Acidification of an unbuffered solution (1 mM CaSO4, I mM K2SOa) by the

roots of ten seedlings of susceptible SAS 2 (A) and resistant SAS 1 (B) wild oat biotypes.

Anows, addition (+) and removal (-) of 50 pM diclofop acid (DA). The data shown is a

representative of three measurements of the susceptible and resistant biotypes.

t47

Table E.l Acidification (A pH g-1 root fresh weight h-l) of an unbuffered solution (lmM

CaSO¿, knlvf KzSO¿)bytherootsof susceptibleandresistantbiotypesof wildoat. ThepH

was measured prior to addition of 50 ¡rM diclofop acid, after addition of the herbicide, and

following removal of the herbicide. Values are the mean of t standard deviation of th¡ee

mea$¡tements

Biotype  pH g-t root FW h-l

Control +Diclofopacid -Diclofopacid

Susceotible biowoes

SAS 2

SAF 19

Resistant biotvoes

SAS 1

NAF 6

NAS 4

NAS 14

NAF 8

NAS 6

SAF 14

SAF 34

rwAF 12

Inlium ripi.dum

VLR I (susceptible)

SLR 31 (resistant)

2.47 + 0.22 0.30 + 0.08 0.45 r 0.04

2.4t + 0.24 4.28 + 0.12 0.54 r 0.20

2.25 + O.t8 0.46 r 0.09 0.55 + 0.05

2.16 r 0.09 0.25 + 0.11 0.72 + o.tt

2.44 t 0.lo 0.67 r 0.08 0.45 r 0.14

2.47 + O.to o.42 + O.20 0.62 + 0.15

2.05 + O.t2 0.37 ! 0.14 0.54 t 0.14

2.35 + 0.t6 0.43 + 0.13 0.69 r 0.18

2.r3 + O.20 0.33 + 0.10 0.71 t 0.17

2.22 + 0.t8 0.50 r 0.12 0.58 + 0.17

2.14 + O.05 0.42 + 0.05 0.39 + 0.18

t.27 + O.22 0.08 + 0.01 0.2t + 0.14

0.16+0.19 +O.49t0.14 0.21 +. o.tg

148

potential following removal of the herbicide in these wild oat biotypes is not correlated with

resistrnce to ACCase-inhibiting herbicides.

The kinetics of depolarization and repolarization of plasma membrane potential depend on

the herbicide, the herbicide concentration, the biotype and the pH of the bathing solution

(Holtum et at. 1994). It is clear from this study and previous reports that pH is an important

determinant for the recovery of the plasma membrane potential following removal of

diclofop acid (DiTomaso 1993, Holtum et al. 1994). It was proposed that the differences in

membrane repolarization following removal of herbicide between susceptible, VLR 1, and

resistant, SLR 31, biotypes of L. rigidum were due to differences in acidification of the cell

wall by these biotypes (DiTomaso 1993). Therefore, strict control of pH by buffer should

eliminate the differences between these biotypes. Recovery of membrane polarity is

dependent on pH rather than buffering capacity as the resistant biotype SLR 3l will recover

from diclofop acid-induced depolarization in a solution buffered at pH 5.8 whereas the

susceptible biotype VLR 1 will not (data not shown). In this study the effects of pH on

recovery of the membrane potential following removal of diclofop acid were examined

(Figure 8.3) and it is also apparent that different biotypes display different pH profiles for

the recovery. Therefore, the differences observed in recovery of the plasma membrane

potential between resistant and susceptible biotypes of L. rigidum (Häusler et al. 1991,

Figure 8.2) and A. fatua (Devine et al. 1993) might be attributable to different pH profiles

for recovery between these biotypes. It has long been known that depolarization of the

membrane potential by diclofop acid is pH dependent (Lucas et al. 1984, Wright and

Shimabukuro 1987, DiTomaso 1993), however, this study demonsüates that recovery from

depolarization is also pH dependent (Figure 8.3). Plasma membrane potentials of coleoptile

cells from two susceptible and three resistant wild oat biotypes depolarized by diclofop acid

at pH 5.8 recover from depolarization upon herbicide removat if the pH is increased to pH

6.5, but not at pH 5.8 (Figures 8.4 and 8.5). Hence, depolarization of the plasma

membrane potential by ACCase-inhibiting herbicides is probably not lethal to plants and this

t49

phenomenon may be only transitory following field application of these herbicides

(DiTomaso et al. 1991, DiTomaso 1993).

Two hypotheses have been proposed to explain the diclofop acid-induced depolarization of

the plasma membrane potential. Either, that diclofop acid acts as a protonophore (Wright

and Shimabukuro 1987), or diclofop acid interacts with a plasma membrane protein to

induce an influx of protons (Shimabukuro and Hoffer 1992). Diclofop acid and other

ACCase-inhibiting herbicides are weak acids with a pKa of between 3.5 and 4.6 (Dotray et

al. 1993). At the pH prevalent in the cell wall, pH 5.5 (Sterling 1994), a higher proportion

of these herbicides will be in a protonated, lipid soluble form than at the pH of the

cytoplasm, pH 7.5. The herbicides will therefore tend to move from the apoplast to the

cytoplasm, lose a proton and become trapped in the cytoplasm. This will lead to an influx of

protons into the cell and a dissipation of the plasma membrane potential. Increasing the

external pH to 6.5 following removal of the herbicide will tend to mobilize some of the

protonated diclofop acid from the cell membrane to the external solution where it will be

deprotonated. Such a scenario could easily explain the response of repolarization to pH

observed in this study (Figure 8.3). Differences between biotypes in the pH required for

recovery of the plasma membrane potential may reflect differences in the amount of diclofop

acid which partitions into the plasma membrane or in the micro-environment surrounding the

plasma membrane.

If there is a mechanism(s) of resistance which give rise to the repolarization of plasma

membrane potential at pH 5.7 in some resistant biotypes of Z. rigidum (Häusler et al. 1991)

then it does not operate in the resistant Avena biotypes examined here. To date, the

molecular basis of repolarization as a resistance mechanism is unknown, but have been

postulated to involve sequestration of the herbicide (Holtum et aJ. 1994). Correlated with

this mechanism(s) in L. rigidurn biotypes is a reduced ability to acidify the external medium

(Häusler et al. 1991). For example, the roots of susceptible biotypes VLR I and SLR 2

decreased the pH of the external medium by 1.7 to 2.4 pH units, whereas the diclofop-

resistant biotypes SLR 31, WLR 96, VLR 6 and NLR 2 showed an acidification of less than

150

0.2 pH units (Häusler et al. 1991). DiTomaso (1993) reported that roots of the susceptible

biotype VLR I were able to acidify the external medium ^t4.5

times the rate of the resistant

biotype SLR 31. This has been proposed to explain both the membrane repolarization and

resistance in one biotype of. L. rigidum, SLR 31, (DiTomaso 1993). Briefly, this

hypothesis suggests that the reduced acidification leads to a higher pH in the apoplast. At

this higher pH, more of the herbicide is in the unprotonated form and unable to enter the cell.

Under such conditions, biotype SLR 3l should be resistant to all weak acid herbicides,

however, this is clearly not the case as most of the SLR 31 population is susceptible to

sethoxydim (Tardif and Powles 1994) and sulfometuron-methyl (Christopher et al. l99l).

In contrast to the situation observed with ¿. rigidum, there were no differences in

acidification of the external medium between resistant, UM-l, and susceptible, UM-5,

biotypes of. A. fatua which differed in thei¡ ability to recover from diclofop acid-induced

membrane depolarization (Devine et al. 1993). However, in this case the resistant biotype

was able to recommence acidification following treatment with diclofop acid, whereas the

susceptible biotype could not (Devine et al. 1993). Similar results were obtained here with

resistant and susceptible biotypes of Z. rigidum (Table 8.1). Addition of 50 ¡tM diclofop

acid to the medium bathing the roots reduced acidification by both biotypes, however, only

the resistant biotype SLR 3l was able to recommence acidification at pre-treatment rates

following removal of the herbicide. In contrast, none of the Avena biotypes examined here

were able to recommence acidification at the pre-treatment rates following removal of

diclofop acid (Table 8.1). A simple explanation for the diclofop acid-induced reduction in

acidification would be proton movement into the cells as the protonated form of diclofop

acid. Alternatively, diclofop acid may be having effects on membrane components which

reduce net proton efflux. The fact that susceptible biotypes of Z. rigidum (Table 8.1) andA.

føtuø (Devine et al. 1993) are unable to recommence acidification following treaünent with

diclofop acid supports the second rather than the first hypothesis. In common with the

membrane repolarization phenomenon, rapid re-establishment of acidification following

treatment with diclofop acid is not required for resistance to the ACCase-inhibiting

herbicides, even when active site changes are not involved (See Chapter 7).

151

In conclusion, the mechanism(s) which give rise to the repolarization of the plasma

membrane potential and the re-establishment of acidification by intact roots following

diclofop acid treatment in some resistant biotypes of L. rtgidum md A. fatua do not operate

in the resistantz4,v¿n¿ biotypes examined here. Clea¡ly, more work is required to understand

these phenomema, however, they are not always correlated with resistance to ACCase-

inhibiting herbicides.

152

Chapter 9

General Conclusions

This project has been conducted to elucidate the mechanisms endowing resistance to

ACCase-inhibiting herbicides in wild oat (Avena spp.) populations in Australia. Studies

with ten putative resistant wild oat biotypes, which had been exposed to different herbicide

treatments in the field, showed resistance to one or more APP or CHD herbicides. The

patterns of resistance to APP and CHD herbicides observed in wild oat could be classified

into three goups; 1) High to moderate resistance to APP herbicides and low-level resistance

to CHD herbicides, 2) Moderate resistance to both APP and CHD herbicides and 3) Low-

level resistance to both classes of herbicides. \When the response of the target site ACCase

to herbicides was examined, it proved in most biotypes to be resistant to at least one CHD or

APP herbicide. With a few exceptions the patterns of resistance at the ACCase level

mi¡rored resistance at the whole plant level (Chapter 3).

Four biotypes with different patterns of resistance were selected for further examination of

mechanisms of resistance. Biotypes SAS 1 and NAS 4 are highly resistant to APP

herbicides, biotype NAF 6 is moderately resistant to both APP and CHD herbicides, and

biotype NAS 14 is slightly resistant to both groups of herbicides. The results indicated that

resistance to ACCase-inhibiting herbicides can be conferred by both target site and non-

target site mechanisms (Chapters 4 to 7).

Target site-based resistance is by far the most common mechanism of herbicide resistance

documented. Almost all cases of PS tr-inhibitor resistance in weeds a¡e due to a single

change in the Dl protein of photosystem II such that the triazine herbicides no longer inhibit

photosystem tr (Gronwald 1994). Similarly, resistance to Als-inhibiting herbicides is

mostly the result of changes to the target site, ALS, which renders the enzyme resistant to

these herbicides (Saari et al. 1994). Target site resistance to the ACCase-inhibiting

herbicides is also common (Devine and Shimabukuro 1994). This mechanism has also been

153

widely observed in this study (Chapter 3). For both biotypes SAS I and NAF 6, a resistant

ACCase was the only mechanism observed (Chapters 4 and 5). Experiments examining

uptake, translocation and metabolism of ¡tagtd'clofop-methyl showed no differences

between resistant and susceptible biotypes. Likewise, no differences in uptake or

metabolism of ¡l+ç¡r"*oxydim were observed between susceptible SAS 2 and resistant

NAF 6 (Chapter 5). These two resistant biotypes were not identical. Biotype SAS 1 was

highly resistant to APP herbicides, but only slightly resistant to CHD herbicides, whereas

biotype NAF 6 was moderately resistant to both classes of herbicides. Frequently, target

site resistance to APP herbicides results in resistance to the CHD herbicides (Devine and

Shimabukuro 1994), however, this does not always prevail as has been observed in a

diclofop-methyl resistant biotype of L. multiþrum (Gronwald et al. 1992). V/ith the

exception of biotype NAF 6, none of the biotypes of wild oat with target site-based

resistance studied in this thesis have sEong resistance to the CHD herbicides (Chapter 3).

Clearly, there are a number of potential mutations within ACCase which can endow

resistance to ACCase-inhibiting herbicides. This has been observed in maize cell cultures

selected with sethoxydim or haloxyfop where th¡ee mutant ACCase alleles of ACCase

structural gene (Accl), Accl-S, AccI-HI and AccI-H2, have been identified (Marshall et al.

1992). The Accl-S exhibited high resistance to sethoxydim and haloxyfop. The AccI-Hl

allele exhibited resistance only to haloxyfop and the Accl-H2 showed high resistance to

haloxyfop and intermediate resistance to sethoxydim. A similar situation is likely in the wild

oat biotypes studied here with several possible different resistant alleles of ACCase, based

on responses of ACCase to herbicides (Chapter 3). Simila¡ly, resistance to AlS-inhibiting

herbicides can be conferred by many possible mutations within ALS in plants (reviewed in

Saa¡i et al. 1994). Mutations in the ALS gene conferring resistance have been related to 24

different mutations at 10 sites. In contrast, a single change ser 264 to gly in the D1 protein

of PS tr has been observed to provide target-site resistance to PS ll-inhibiting herbicides in

higher plants (Gronwald 1994). The large number of possible mutations within ALS, and

probably ACCase, may account in part for the relative rapid rate of development of

resistance to the ALS- and ACCase-inhibiting herbicides. Resistance to these herbicides can

154

appear in as little as 3 or 4 applications (Tardif et al. 1993, Devine and Shimabukuro 1994,

Maxwell and Mortimer 1994), whereas resistance to PS tr-inhibiting herbicides only appears

after l0 or more applications of herbicide (Ryan 1970, Burnet et al. 1991). Therefore, the

more rapid development of resistance to the ACCase and AlS-inhibiting herbicides might be

due to an increased nu¡rrber of viable mutations within the target sites.

Biotype NAS 4 has target site resistance to APP herbicides, however, in this case moderate

resistance at the enzyme level appears insufFlcient to explain the high-level resistance at the

whole plant level. This biotype contains a second mechanism of resistance, enhanced

metabolism of diclofop. The seeds used in these experiments were collected from a number

of plants so it is possible that different mechanisms of resistance were present in different

individuals of this population. This hypothesis was tested by examining the response of

ACCase to diclofop acid in the progeny from ten individual resistant plants (Chapter 6). The

result showed that ten lines of biotype NAS 4 exhibited about the same level of resistance at

the ACCase as the whole population (Chapter 3). This means that resistance at the ACCase

level is quite uniform within the NAS 4 population and therefore the two mechanisms of

resistance coexist within each individual. Enhanced metabolism of diclofop has been

observed in a biotype of Lolium rigidum which showed a 1.5 fold increased capacity to

metabolize diclofop (Holtum et al. 1991). In addition, enhanced metabolism has been

reported in biotypes of L. rigidurn resistant to AlS-inhibiting (Christopher et al. 1992) and

PS tr-inhibiting herbicides (Burnet et al. 1993a and 1993b). In these last cases, enhanced

metabolism due to cytochrome P450 monooxygenases has been implicated. Studies with the

cytochrome P450 inhibitor, tetcyclasis, revealed that this compound can both decrease

diclofop acid metabolism in biotype NAS 4 and increase diclofop-methyl phytotoxicity

(Chapter 6). Following pre-treatment with tetcyclasis, the amount of diclofop acid

remaining in biotype NAS 4 24hafter treatment with diclofop-methyl, was increased to the

same level as that found in the susceptible biotype SAF 19. Therefore, it is likely that

enhanced metabolism of diclofop acid in biotype NAS 4 is endowed by a cytochrome P450

monooxygenase.

155

Resistance as a result of a target site mutation and enhanced metabolism has been reported

for a tnazine-resistant biotype of Brachepodium distachyon '(Gressel et al. 1983), a

sulfonylurea-resistant biotype of L. rigidun (Cbristopher et al. 1992), a diclofop-methyl

resistant biotype of L. rigidurø (Preston et al. unpublished) and in small subsets of diclofop-

methyl resistant (Tardif and Powles 1994) and sulfonylurea-resistant (Burnet et al. 1994)

populations of Z. rigidum. In all of these cases, the target site resistance is the most

powerful resistance mechanism. Multiple mechanisms of resistance to the same herbicide

have been reported on a few occasions, most often in biotypes of L. rígidum. lt was noted

that herbicide treatment of a large, polymorphic population results in the survival of

individuals that possess one or more (different) mechanisms conferring the ability to

withstand the dose of herbicide applied (Powles and Matthews 1992). rù/ith obligate

outcrossing species like Z. rigidum, there is gene flow among the survivors resulting in

exchange of different resistance genes and their accumulation in the next generation.

However, dependent upon genetic variation, the size of a population exposed to herbicide

selection, and the ability and efficiency of cross-pollination, there can be enrichment of a

number of different resistance mechanisms (Powles and Matthews 1992). Although wild

oat is self-pollinated species, outcrossing can occur at rates up to l2vo (Imam and Allard

1965, Chancellor 1976, Bickelmann 1993). Hence, the genetic variability between and

within natural populations of wild oat can be maintained. It is, therefore, an indication of

biological diversity that two mechanisms of resistance, both contributing to resistance, have

appeared in a single biotype of wild oat.

Resistance to diclofop in biotype NAS 14 is not due to a change in the ACCase target site

nor is it due to enhanced herbicide detoxification. Biotype NAS 14 did demonstate reduced

translocation of radioactivity to the meristematic region from leaf-applied [t+C]diclofop-

methyl relative to the susceptible biotype. This result was confirmed by measuring the

activity of ACCase in the meristematic region of susceptible and resistant biotypes following

application of 5 mM diclofop-methyl to the leaf axil. As shown in Chapter 3, biotype NAS

14 has a herbicide-sensitive ACCase, however, the extractable ACCase activity in the

meristematic region was greater in this biotype following application of diclofop-methyl than

156

in the susceptible biotype SAF 19. This indicates that herbicide moves more slowly to the

meristematic region in the resistant biotype. It can be difficult to determine the mechanism of

resistance in a biotype that has a low level of resistance, however, the small difference in

translocation observed here may be sufficient to account for the low level of resistance

observed in this biotype. Reduced herbicide translocation is a mechanism of resistance in

some paraquat-resistant biotypes (reviewed in Preston 1994), however, this mechanism has

not been documented as a resistance mechanism in weeds for any other herbicides.

As summarised above, three mechanisms of resistance were found in resistant wild oat

biotypes (Chapters 3 to 7), however, the modified target site, ACCase, was the most

common phenomenon. The results in Chapters 4 to 7 may be used to indicate possible

resistance mechanisms operating in other resistant biotypes. Resistance in biotypes NAF 8,

NAS 2 and SAF 14 that are highly resistant to diclofop-metþl in the field but exhibit a

moderate-level of resistance at the enzyme level might involve more than one mechanism as

found in biotype NAS 4. However, these biotypes are less resistant than biotype NAS 4

and, therefore, the resistant ACCase may be the most important, or sole mechanism

operating. Biotype WAF 12 has low-level-resistance to diclofop-methyl like biotype NAS

14, however, ACCase from this biotype is sensitive to diclofop acid. This biotype might

have a similar non-target site mechanism conferring resistance. The obligate cross-pollinated

L. rigidum is capable of evolving multiple mechanisms to overcome herbicides (Holtum and

Powles 1991). As has been demonstrated here, the self-pollinated Avena species can also

exhibit diversity in resistance mechanisms.

Some resistant weeds have the ability to re-establish the plasma membrane potential

following treatment with diclofop acid (Häusler et al. 1991, Holtum et al. 1991, Devine et

al. 1993). This phenomenon was not observed in the ten wild oat biotypes studied here.

The recovery of membrane polarity is depended upon the pH of bathing solution

(Chapter 8). Both susceptible and resistant wild oat biotypes exhibited membrane

repolarization when the herbicide solution was replaced with the herbicide-free solution

buffered at pH 6.5. Although repolarization of plasma membrane potentials has been

t57

documented as a mechanism of resistance to ACCase-inhibiting herbicides in some resistant

biotypes of. A. fana and Z. rigidum, this phenomenon was not appeared in the resistant

biotypes of A. fatua and.á. sterilis studied here. However, it is still unclea¡ how the ability

to restore the membrane potential to its normal state is correlated with the selective action of

diclofop acid between susceptible and resistant plants. What is clear is that this rapid

repolarization of the plasma membrane potential is not a requirement for resistance to

ACCase-inhibiting herbicides.

The number of applications and classes of herbicides cannot be used to predict the patterns

and level of resistance to APP and"/or CHD herbicides in wild oat. For example, biotype

NAS 4 had been exposed to two applications of tralkoxydim but it is sensitive to CHD

herbicides at the whole plant and the enzyme level (Chapter 3). In contrast, biotype NAF 6

had never been exposed to CHD herbicides but is moderately resistant to CHD herbicides

(Chapter 3). Biotype NAS 14 had been treated with four applications of diclofop-methyl

and two of fluazifop-butyl, however, it has developed lowlevel resistance to both APP and

CHD herbicides whereas biotype SAS 1 with five applications of APP herbicides has high-

level resistance to APP herbicides.

Wild oat seeds can survive in the soil for up to nine years (Jones t976, Gwynne and Murray

1985, Miller and Nalewaja 1990). Therefore, once resistant wild oat are in a field, the

resistance genes can remain for a considerable period. There are few alternative herbicides

to diclofop-methyl, fenoxaprop-ethyl and talkoxydim, for the selective control of wild oat in

cereals. With the development of target site-based resistance to the ACCase-inhibiting

herbicides, these alternative herbicides will be relied on to control wild oat. Wild oat can

also develop resistance due to non-target site mechanisms such as enhanced metabolism, for

example biotype NAS 4, and decreased translocation, for example biotype NAS 14. Such

non-target site mechanisms, as has been documented for Z. rigidum and A. myosuroides

(Hall et al. 1994) can lead to cross and multiple resistance across chemical classes.

Enhanced metabolism, in particular, has been demonstrated as a mechanism endowing cross

and multiple resistance in these two species. Such a scenario appearing in wild oat may

158

mea¡r that herbicides will not be an effective means sf ço¡trslling this species in the future.

Already there are at least 15 populations of triallate-resistant wild oat exhibiting cross-

resistance to difenzoquat (O'Dovovan et al. 1994) and more cases of cross and multiple

resistance could be expected.

159

Literature Cited

Adkins, S.}V., Simpson, G.M. and Naylor, J.M. (1984a). The Physiological basis of seed

dormancy rnAvenafatuø. fr.. Action of nitrogeneous compounds. Physiol. Plant.

6O:227-233.

Adkins, S.W., Simpson, G.M. and Naylor, J.M. (1984b). The physiological basis of seed

dormancy rn Avenafatw.lY. Altemative respiration and nitogeneous compounds.

Physiol. Plant. 60: 234-238.

Adkins, S.'W., Naylor, J.M. and Simpson, G.M. (1984c). The physiological basis of seed

dormancy in Avena fatua. V. Action of ethanol and other organic compounds.

Physiol. Plant. 62: 18-24.

Agenbag, G.A. and DeVilliers, O.T. (1989). The effect of nitrogen fertilizers on

germination and seedling emergence of wild oat (l4,. fatua L.) seed in different soil

types. V/eed Res. 29: 239-245.

Aguero-Alvarado, R., Appleby, A.P. and Armstrong, D.J. (1991). Antagonism of

haloxyfop activity in tall fescue (Festuca arundinacea) by dicamba and bentazon.

Weed Sci.39: 1-5.

Alban, C., Baldet, P. and Douce, R. (1994). Localization and cahracterization of two

structurally different forms of acetyl-CoA carboxylase in young pea leaves, of

which one is sensitive to aryloxyphenoxypropionate herbicides. Biochem. J. 300:

557-565.

Al-Feel, W., Chirala, S.S. and Wakil, S.J. (1992). Cloning of the yeast FAS3 gene and

primary structure of yeast acetyl-coenzyme A carboxylase. Proc. Natl. Acad. Sci.

USA 89: 4534-4538.

160

Anderson, R.N. and Gronwald, J.W. (1987). Noncytoplasmic inheritance of atrazine

tolerance in velvetleaf (Abutílon theophrastí). Weed Sci. 35: 496498.

Asa¡e-Boamah, N.K. and Fletcher, R.A. (1983). Physiological and cytological effects of

BAS 9052 OH on corn (7za mays) seedlings. Weed Sci. 31: 49-55.

Ashton, F.M. and Monaco, T.J. (1991). Aryloxyphenoxys, benzoics, bipyridiliums and

pyrazoliums. .In Weed Science Principles and Practices, (F.M. Ashton and T.J.

Monaco, Eds.), A Wiley-Interscience Publication, NY, PP. 172-lg}.

Ashton, A.R., Jenkin, C.D.L. and Whitfeld, P.R. (1994). Molecular cloning of two

different cDNAs for maize acteyl CoA carboylase. Plant Mol. Biol. A:3549.

Baker, E.A. and Chamel, A.R. (1990). Herbicide penetration across isolated and intact leaf

cuticles. Pestic. Sci. 29: 187-196.

Ba¡r, 4.R., Mansooji, A.M., Holtum, J.A.M. and Powles, S.B. (1992). The inheritance

of herbicide resistance inAvena sterílis ssp.ludoviciana,biotype SAS l. Proc. lst

International Weed Contol Congress, Melbourne, Australia, pp. 7 0-7 5.

Ba¡rentine, W.L., Snipes, C.E. and Smeda, R.J. (1992). herbicide resistance confirmed in

johnsongrass biotypes. Miss. Agric. Forest. Exp. Stn. Res. Rep. 17: l-5.

Beckett, T.H., Stroller, E.'W. and Bode, L.E. (1992). Quizalofop and sethoxydim activity

as affected by adjuvants and ammonium fertilizers. rü/eed Sci. 40: 12-19.

Betts, K.J., Ehlke, N.J., Wyse, D.L., Gronwald, J.W. and Somers, D.A. (1992).

Mechanism of inheritance of diclofop resistance in Italian ryegrass (Lolium

multiþrum). V/eed Sci. 40: 184-189.

Bickelmann, U. (1993). On fatuiods and hybrids of Avena spp. and their significance for

seed production. I. Flowering behaviour of oats and origin of fatuoids and hybrids.

Plant Va¡ieties and Seeds 6:21-26.

161

Bishop, T., Powles, S.B. and Cornic, G. (1987). Mechanism of paraquat resistance in

Hordeum glaucuni.Il. Paraquat uptake and translocation.' Aust. J. Plant Physiol.

t4: 539-547.

Black, M. (1959). Dormancy studies in seed of Avena fatua. I. Possible role of

germination inhibitors. Can. J. Bot. 37:393402.

Boldt, L.D. and Barrett, M. (1991). Effects of diclofop and haloxyfop on lipid synthesis in

corn(7*amoys) and bean (Phaseolus vulgaris). rüVeed Sci. 39: 143-148.

Boldt, P.F. and Putnam, A.R. (1980). Selectivity mechanisms for foliar application of

diclofop-methyl. I. Retention, absorption, translocation and volatility. S/eed Sci.

28: 474477.

Boutsalis, P., Holtum, J.A.M. and Powles, S.B. (1990). Diclofop-methyl resistance in a

biotype of wild oat (Avena fatua L.). Proc. 9th Australian V/eed Conf., Adelaide,

Australia, pp. 216-219.

Boutsalis, P. and Powles, S.B. (1995a). Resistance of dicot weeds to acetolactate synthase

(Als)-inhibiting herbicides in Australia. T9Veed Res. in press.

Boutsalis, P. and Powles, S.B. (1995b). Inheritance and mechanism of resistance to

herbicides inhibiting acetolactate synthase in Sonchus oleraceus L. Theor. Appl.

Genet. in press.

Brezeanu, 4.G., Davis, D.G. and Shimabukuro, R.H. (1976). Ultrastructural effects and

translocation of metþl-2-(4-(2,4 -diclorophenoxy) phenoxy) propanoate in wheat

(Triticum aesticum) and wild oat(Avenafana). Can. J. Bot.54: 2038-2M8.

Browse, J., Roughan, P.G. and Slack, C.R. (1981). Light control of fatty acid synthesis

and diurnal fluctuations of fatty acid composition in leaves. Biochem. 1.196:347-

354.

t62

Browse, J. and Somerville, C. (1991). Glycerolipid synthesis: biochemistry and

regulation. Annu. Rev. Plant Physiol. Plant Mol. Biol. 42: 267-506.

Brun, V.S. (1965). The effects of fresh water storage on the germination of certain weed

seeds.'Weeds 31: 38-39.

Buhler, D.D., Swisher, B.A. and Burnside, O.C. (1985). Behavior of l4C haloxyfop-

methyl in intact plants and cell cultures. V/eed Sci. 33: 291-299.

Burnet, M.V/.M., Hildebran, O.8., Holtum, J.A.M. and Powles, S.B. (1991). Amitrole,

triazine, substituted urea and metribuzin resistance in a biotype of rigid ryegrass

(Inlium rigidum). Weed Sci. 39: 317-323.

Burnet, M.V/.M., Loveys, 8.R., Holtum, J.A.M. and Powles, S.B. (1993a). A

mechanism of chlorotoluron resistance in Lolium rigidum. Planta 190: 182-189.

Burnet, M.W.M., Loveys, 8.R., Holtum, J.A.M. and Powles, S.B. (1993b). Increased

detoxification is a mechanism of simazine resistance in l-olium rigidum. Pestic.

Biochem. Physiol. 46: 2O7 -2L8.

Burnet, M.W.M., Christopher, J.T., Holtum, J.A.M. and Powles, S.B. (1994).

Identification of two mechaisms of sulfonylurea resistance within one population of

rigid ryegrass (Lolium rigidum) using a selective germination medium. rù/eed Sci.

42:468-473.

Burton, J.D., Gronwald, J.W., Somers, D.4., Connelly, J.4., Gengenbach, B.G. and

Wyse, D.L. (1987). Inhibition of plant acetyl coenzyme A carboxylase by the

herbicides sethoxydim and haloxyfop. Biochem. Biophys. Res. Commun. 14E:

to39-t34r'..

Burton, J.D., Gronwald, J.W., Somers, D.4., Gengenbach, B.G. and Wyse, D.L.

(1989). Inhibition of corn acetyl-CoA carboxylase by cyclohexane-dione and

aryloxyphenoxypropionate herbicides. Pestic. Biochem. Physiol. 34: 7 6-85.

163

Burton, J.D., Gronwald, J.W., Keith, R.4., Somers, D.4., Gengenbach, B.G. and'lVyse,

D.L. (1991). Kinetics of inhibition of acetyl-coenzyme A carboxylase by

sethoxydim and haloxyfop. Pestic. Biochem. Physiol. 39: 100-109.

Cairns, A.L.P. and DeVilüers, O.T" (1986). Breaking dormancy of. AvenafatusL. seed by

treatment with ammonia. V/eed Res. 26: t9l-197.

Carrpbell, J.R. and Penner, D. (1985). Sethoxydim metabolism in monocotyledonous and

dicotyledonous plants. V/eed Sci. 33: 771-773.

Carr, J.C., Davies¡ L.G., Cobb, A.H. and Pallett, K.E. (1985). The metabolic activity of

fluazifop acid in excised apical meristem sections. Proc. British Crop Prot. Conf.

-Weeds: 155-L62.

Catanzaro, C.J., Burton, J.D. and Kroch, W.A. (1993). Graminicide resistance of acetyl-

CoA carboxylase from ornamental grasses. Pestic. Biochem. Physiol. 45: 147-

153.

Chancellor, R.J. (1976). Seed behaviour. /n V/ild Oats in lVorld Agriculture, (D.P.

Jones, Ed.), The Whitefriars Press, Ltd., London, pp. 65-87.

Chandrasena, N.R. and Sagar, G.R. (1987). Effect of fluazifop-butyl on the chlorophyll

content, fluorescence and chloroplast ultrastructure of Elymus repens (L.) Gould

leaves. Weed Res.27: lO3-112.

Chandrasena, N.R. and Sagar, G.R. (1989). Fluazifop toxicity to quackgrass (Agropyron

repens) as influenced by some application factors and site of application. Weed Sci.

37:79O-796.

Cha¡les, D.J., Hasegawa, P.M. and Cherry, J.H. (1986). Characterization of acetyl-CoA

carboxylase in the seed of two soybean genotypes. Phytochem. 25: 55-59.

tg

Chauvel, B. (1991). Polymorphisme génétique et sélection de la resistance aux urées

substitutées chez Alopecurus myosuroÍd¿s Huds. Ph.D. Thesis, University of

Paris-Orsay.

Cho, H., V/idholm, J.M. and Slife, F.W. (1986). Effects of haloxyfop on corn (Zea

mays) and soybean (Glycine max) cell suspension cultures. Weed Sci.34: 496-

501.

Christopher, J.T. (1993). Mechanisms of resistance to herbicide which inhibit acetolactate

synthase in annual ryegrass (Lolium rigidum). Ph.D. Thesis, University of

Adelaide, Austalia.

Christopher, J.T., Powles, S.B.and Holtum, J.A.M. (1992). Resistance to acetolactate

synthase-inhibiting herbicides in annual ryegfass (Inlium rigidum) involves at least

two mechanisms. Plant Physiol. 100: 1909-L9I3.

Christopher, J.T., Powles, S.8., Liljegren, D.R. and Holtum, J.A.M. (1991). Cross-

resistance to herbicides in annual ryegrass (Lolium rigidum).II. Chlorsulfuron

resistance involves a wheat-like detoxification system. Plant Physiol. 95: 1036-

to43.

Christopher, J.T., Preston, C. and Powles, S.B. (1994). Malathion antagonizes

metabolism-based chlorsulfuron resistance in l-olium rigidum. Pestic. Biochem.

Physiol. 49: 172-182.

Cotterman, J.C. and Sarri, L.L. (1992). Rapid metabolic inactivation is the basis for cross-

resistance to chlorsulfuron in diclofop-methyl-resistant rigid ryegrass (Lolium

rigidum) biotype SR4/84. Pestic. Biochem. Physiol. 43:182-192.

Coupland, D. (1994). Resistance to the auxin analog herbicides. /n Herbicide Resistance in

Plants: Biochemistry and Biology, (S.B. Powles and J.A.M. Holtum, Eds.), CRC

Press Inc., Boca Raton, EL, pp.l7l-214.

l6s

Coupland, D., Lutman, P.J.W. and Health, C. (1990). Uptake, translocation and

metabolism of mecoprop in a sensitive and a resistant biotype of Stellaria media.

Pestic. Biochem. Physiol. 36: 6l'67.

Dahroug, S. and Muller, F. (1990). Response of susceptible and tolerant plants to

diclofop-methyl. J. Plant Dis. Prot. 97: 154-167.

Devine, M.D., Maclsaac, S.4., Romano, M.L. and Hall, J.C. (1992). Investigation of the

mechanism of diclofop resistance in two biotypes of. Averufatua. Pesttc. Biochem.

Physiol. 42:88-96.

Devine, M.D., Marles, M.A.S. and Hall, L.M. (1991). Inhibition of acetolactate synthase

in susceptible and resistant biotypes of Stellaria media. Pestic. Sci. 31: 273-280.

Devine, M.D., Hall, J.C., Romano, M.L., Marles, M.A.S., Thompson, L.W. and

Shimabukuro, R.H. (1993). Diclofop and fenoxaprop resistance in wild oat is

associated with an altered effect on the plasma membrane electrogenic potential.

Pestic. Biochem. Physiol. 45: 167-177.

Devine, M.D. and Shimabukuro, R.H. (1994). Resistance to acetyl coenzyme A

carboxylase inhibiting herbicides. /z Herbicide Resistance in Plants: Biochemistry

and Biology, (S.8. Powles and J.A.M. Holtum, Eds.), CRC Press Inc., Boca

Raton, FL, pp. 14l-169.

Dicks, J.W., Slater, J.lW. and Bewick, D.W. (1985). PP 005- The R-enantiomer of

fluazifop-butyl. Proc. British Crop Prot. Conf. -V/eeds: 271-280.

DiTomaso, J.M., Brown, P.H., Stowe, 4.E., Linscott, D.L. and Kochian, L.V. (1991).

Effects of diclofop and diclofop-methyl on membrane potentials in roots of intact

oat, maize and pea seedlings. Plant Physiol. 95: 1063-1069.

DiTomaso, J.M. (1993). Membrane response to diclofop is pH dependent and is regulated

by the protonated form of the herbicide in roots of pea and resistant and susceptible

rigid ryegrass. Plant Physiol. 102: L33l-1336.

166

Donald, W. \V. and Shimabukuro, R. H. (1980). Selectivity of diclofop-methyl between

wheat and wild oat: growth and herbicide metabolism. Physiol. Plant. 49:459-

464.

Dotray, P.4., DiTomaso, J.M., Gronwald, J.}V., V/yse, D.L. and Kochian, L.V. (1993).

Effects of acetyl-coenzyme A carboxylase inhibitors on root cell transmembrane

electric potentials in graminicide-tolerant and -susceptible corn (kamays L.). Plant

Physiol. 103:919-924.

Duke, S.O. and Kenyon, W.H. (1988). Polycyclic alkanoic acids. fn Herbicides:

Chemistry, Degradation, Mode of Action. Vol. 3, (P.C. Kearney and D.D.

Kaufman, Eds.), Marcel-Dekker, NY, pp. 71-116.

Egli, M.4., Gengenbach,8.G., Gronwald, J.'W., Somers, D.A. and Wyse, D.L. (1993).

Cha¡acterization of maize acetyl coenzyme A carboxylase. Plant Physiol. 101:

499-506.

Evenson, K.J., Gronwald, J.rW. and Wyse, D.L. (1994.) Purification and characterization

of acetyl-coenzyme A carboxylase from diclofop-resistant and -suscepttble l-olium

multiþrum. Plant Physiol. 105: 671-680.

Finlayson, S. and Dennis, D.T. (1983). Acetyl-coenzyme A carboxylase from developing

endosperm of Ricinus communis. I. Isolation and characterization. Archv.

Biochem. Biophys. 255: 576-585.

Frear, D.S., Swanson, H.R. and Thalaker, F.W. (1991). Induced microsomal oxidation

of diclofop, triasulfuron, chlorsulfuron and linuron in wheat. Pestic. Biochem.

Physiol. 4t 274-287.

Friesen, H.4., O'Sullivan, P.A. and Vanden Born, Vf.H. (1976). HOE 23408, a new

selective herbicide for wild oats and green foxtail in wheat and barley. Can. J.

Plant Sci. 56: 567-578.

t67

Fuerst, E.P., Nakatani, H.Y., Dodge, 4.D., Penner, D. and Arntzen, C.J. (1985)

Paraquat resistance ln Conyza. Plant Phys iol. 77 : 984-989:

Gerwick, B.C. (1988). Potential mechanisms for bentazon antagonism with haloxyfop.

Weed Sci. 36: 286-290.

Genryick, B.C., Jackson, L.4., Handly, J., Gray, N.R. and Russell, J.W. (1988). Pre-

emergence and post-emergence activities of the (R) and (S) enantiomers of

haloxyfop. Weed Sci. 36: 453456.

Gressel, J., Regev, Y., Malkin, S. and Kleifeld, Y. (1983). Characterization of a s-

triazine-resistant biotype of Brachypodium distachyon V/eed Sci. 31: 450456.

Gronwald, J.W. (1991). Lipid biosynthesis inhibitors. V/eed Sci. 39: 43549.

Gronwald, J.rW. (1994). Resistance to photosystem tr inhibiting herbicides. /n Herbicide

Resistance in Plants: Biochemistry and Biology (S.B. Powles and J.A.M. Holtum,

Eds.) CRC Press Inc., Boca Raton, EL, pp.27-60.

Gronwald, J.W., Eberlein, C.V., Betts, K.J., Baerg, R.J., Ehlke, N.J. and Wyse, D.L.

(1992). Mechanism of diclofop resistance in an Italian ryegrass (Lolium

multiþrum Lam.) biotype. Pestic. Biochem. Physiol. 44: 126-139.

Gronwald, J.W., Pfister, K. and Steinback, K.E. (1989). Atrazine resistance in velvetleaf

(Abutilon theophrastd) due to enhanced atrazine detoxification. Pestic. Biochem.

Physiol. 34: 149-163.

Gwynne, D.C. and Murray, R.B. (1985). Grass weeds- wild oats. /n V/eed Biology and

Control in Agriculture and Horticulture, (D.C. Gwynne and R.B. Murray, Eds.),

Batsford Academic and Educational, London, pp. 99-1L2.

Hadfield, S.T., rù/ilson, 4., Kuet, S.F. and Mason, R. (1994). The metabolism of

tralkoxydim in field grown spring wheat and maize plant cell suspension culture.

Pestic. Sci. 40: 193-20[..

168

Hall, C., Edgington, L.V. and Switzer, C.M. (19S2). Translocation of different 2,4-D,

bentazon, diclofop or diclofop-methyl combinations in oat (Avena sativa) and

soybean (Glycine mtu\). lVeed Sci. 30: 676-682.

Hall, L.M., Holtum, J.A.M. and Powles, S.B. (1994). Mechanisms responsible for cross

resistance and multiple resistance. /n Herbicide Resistance in Plants: Biochemistry

and Biology, (S.8. Powles and J.A.M. Holtum, Eds.), CRC Press Inc., Boca

Raton, FL, pP. 243-26L.

Hall, L.M., Moss, S.R. and Powles, S.B. (1993). Towa¡ds an understanding of

resistance to APP herbicides in Alopecurus rnyosuroides. Froc. l4th Asian Pacific

Weed Science Society, Brisbane, Australia, pp. 299-301.

Harker, K.N. and Blackshaw, R.E. (1991). Influence of growth stage and broadleaf

herbicides on Ealkoxydim activity. rÙVeed Sci. 39: 650-659.

Harwood, J.L. (1983). Fatty acid metabolism. Annu. Rev. Plant Physiol. 39: 155-162.

Harwood, J.L. (1989). The properties and importance of acetyl-coenzyme A carboxylase

in plants. Proc. British Crop Prot. Conf.-'Weeds: 155-162.

Harwood, J.L., Ridley, S.M. and Walker, K.A. (1989). Herbicides inhibiting lipid

synthesis. In Herbicide and Plant Metabolism, (4.D. Dodge, Ed.), Cambridge

University Press, NY, pp. 73-96.

Harwood, J.L., Walker, K.4., Abulnaja, D. and Ridley, S.M. (1987). Herbicides

affecting lipid metabolism. Proc. British Crop Prot. Conf. -'Weeds: 159-169.

Häusler, R.E., Holtum, J.A.M. and Powles, S.B. (1991). Cross-resistance to herbicides

in annual ryegrass (I-olium rigidum).IV. Correlation between membrane effects

and resistance to graminicides. Plant Physiol. 97: 1035-1043.

169

Heap, I.M., Murray, B.G., Leoppsþ, H.A. and Morrison, I.N. (1993). Resistance to

aryloxyphenoxypropionate and cyclohexanedione herbicides in wild oat (Avena

latua). Weed Sci.41: 232-238.

Heap, J. and Knight, R. (1982). A population of ryegrass tolerant to the herbicide

diclofop-methyl. J. Aust. Agric. Sci.48: 156-157.

Heap, I.M. and Knight, R. (1986). The occurrence of herbicide cross resistance in a

populationof annualryegrass, Loliumrigídum, resistanttodiclofop-methyl. Aust.

J. Agric. Res.37: 149-156.

Heap, I.M. and Knight, R. (1990). Variation in herbicide cross-resistance among

population of annual ryegrass (Lolium rtgidum) resistant to diclofop-methyl. Aust.

J. Agric. Res. 41: L2l-128.

Hendley, P., Dicks, J.'W., Monaco, T.J., Slyfield, S.M., Tummon, O.J. and Ba¡rett, J.C.

(1985). Translocation and metabolism of pyridinyloxyphenoxy-propionate

herbicides in rhizomatous quackgrass (Agropyron repens). rJVeed Sci. 33: ll-24.

Higinbotham, N., Graves, J.S. and Davis, R.F. (1970). Evidence for an electrogenic ion

transport pump in cells of higher plants. J. Membr. Biol. 3: 2lO-222.

Hoagland, D.R. and Arnon, D.L (1938). The water-culture method for growing plants

without soil. Calif. Agric. Exp. Stn. Circ.347; l-32.

Hoerauff, R.A. and Shimabukuro, R.H. (1979). The response of resistant and susceptible

plants to diclofop-methyl. Weed Res. 19: 293-299.

Holt, J.S., Powles, S.B. and Holtum, J.A.M. (1993). Mechanisms and agronomic aspects

of herbicide resistance. Annu. Rev. Plant Physiol. Plant Mol. Biol. 44: 203-229.

Holtum, J.A.M. and Powles, S.B. (1991). Annual ryegrass: an abundance of resistance, a

plethora of mechanisms. Proc. Brighton Crop Prot. Conf. -'Weeds: l07l-1077.

t70

Holtum, J.A.M., Matthews, J.M., Häusler, R.E., Liljegren, D.A. and Powles, S.B.

(1991). Cross-resistance to herbicides in annual ryegrass (Loliutn rigidum).Itr. On

the mechanism of resistance to diclofop-methyl. Plant Physiol. 97: lO2Gl034.

Holtum, J.A.M., Häusler, R.E., Devine, M.D. and Powles, S.B. (1994). Recovery of

transmembrane potentials in plants resistant to aryloxyphenoxypropanoate

herbicide: a phenomenon awaiting explanation. Weed Sci. 42: 293-301.

Hosaka, H., Hideo, I., and ishikawa, H. (1984). Response of monocotyledons to BAS

9052 OH. Weed Sci.32: 28-32.

Hosaka, H., Inaba, H., Satoh, 4., Tanoue, T. (1987). Selectivity mechanisms of

sethoxydim between red fescue (Festuca rubra) and tall fescue (Festuca

arundinaceø). Weed Res. (Japan) 32 255-262.

Howard, J.L. and Ridley, S.M. (1990). Acetyl-CoA carboxylase: a rapid novel assay

procedure used in conjunction with the preparation of enzyme from maize leaves.

FrEBS Lett. 261: 26I-264.

Hutson, D.H. and Roberts, T.R. (1987). Wild oat herbicides. In Progress in Pesticide

Biochemistry and Toxicology. (D.H. Hutson and T.R. Roberts, Eds.), John Wiley

& Sons Ltd., NY, pp. t69-197.

Imam, A.G. and Allard, R.W. (1965). Population studies in predominantly self-pollinated

species. VI. Genetic variability between and within natural population of wild oat in

differing habitats in California. Genetics 51: 49-52.

Ishikawa, H., Yamada, S., Hosaka, H., Kawana, T., Okunuki, S. and Kohara, K.

(1985). Herbicidal properties of sethoxydim for the control of gramineous weeds.

J. Pestic. Sci. 10: 187-193.

Ishihara, K., Shiotani, H., Soeda, Y., Ono, S. (1988). Fate of the herbicide sethoxydim in

sugar beet. J. Pestic. Sci. 13: 231-237.

17t

Itoh, K. and Miyahara, M. (1984). Inheritance of paraquat resistance in Erigeron

phitadetphicus L. Weed Res. (Japan) 29: 3Ot-307 .

Islam, A.K.M.R. and Powles, S.B. (1988). Inheritance of resistance to paraquat in badey

grass Hordeum glaucum Steud. Weed Res. 28: 393'397.

Jacobson, A. and Shimabukuro, R.H. (1982). The absorption and translocation of

diclofop-methyl and amitrole in wheat and oat roots. Physiol. Plant. 54:3440.

Jasieniuk, M., Brtlé-Babel, A.L. and Morrison, I.N. (1994). Inheritance of trifluralin

resistance in green foxtail (Setaria viridis). V/eed Sci. 42: 123-127 .

Jensen, K.I.N. and Caseley, J.C. (1990). Antagonistic effects of 2,4-D amine and

bentazone on control of Avenafatuawithtralkoxydim. Weed Res. 30: 389-395.

Jordan, D.L., York, A.C. and Corbin, F.T. (1989). Effect of ammonium sulfate and

bentazon on sethoxydim absorption. V/eed Technol. 3:674-677.

Joseph, O.O., Hobbs, S.L.A. and Jana, S. (1990). Diclofop resistance in wild oat (Avena

fatua). Weed Sci. 38: 475479.

Kannangara, C.G. and Stumpf, P.K. (1972). Fat metabolism in higher plants. A

prokaryotic type acetyl CoA ca¡boxylase in spinach chloroplasts. Arch. Biochem.

Biophys. 152: 83-91.

Kell, J.J., Meggitt, W.F. and Penner, D. (1984). Absorption, translocation and activity of

fluazifop-butyl as influenced by plant growth stage and environment. V/eed Sci.

32: 143-149.

Kobek, K., Forcke, M., Litchtenthaler, H.K., Retzlaff, G. and Wurzer, B. (1988).

Inhibition of fatty acid biosynthesis in isolated chloroplasts by cycloxydim and

other cyclohexane-1, 3-diones. Physiol. Plant. 72: 492498.

172

Köcher, H., Kellner, H.M., Lotzsch, K. and Dorn, E. (1982). Mode of action and

metabolic fate of the herbicide fenoxaprop-ethyl, HOE 33171. Proc. British Crop

Prot. Conf. -Weeds: 341'347.

Konishi, T. and Sasaki, Y. (1994). Compartmentalization of wo forms of acetyl-CoA

carboxylase in plants and the origin of their tolerance toward herbicides. Proc.

Natl. Acad. Sci. USA 91: 3598-3601.

Leach, G.E., Kirkwood, R.C. and Ma¡shall, G. (1993). The basis of resistance displayed

to fluazifop-butyl by biotypes of. Elucine indica. Proc. Brighton Crop Prot. Conf.

-'Weeds: 2OI-207.

Lefsrud, C. and Hall, J.C. (1989). Basis for sensitivity differences among crabgrass, oat,

and wheat to fenoxaprop-ethyl. Pestic. Biochem. Physiol. 34:218-227.

Lewis, J. (1961). The influences of water level, soil depth and type on the survival of crop

and weed seeds. Proc. Int. Seed Test Assoc. 26: 68-85.

Li, S.J. and Cronan, J.E. (1992a). The gene encoding the biotin carboxylase subunit of

Escherichia coli acetyl-CoA carboxylase. J. Biol. Chem. 267:855-863.

Li, S.J. and Cronan, J.E. (1992b). The genes encoding the two carboxyltransferase

subunits of. Escherichia coli acetyl-CoA carboxylase. J. Biol. Chem. 267 16841-

16847.

Lichtenthaler, H.K. and Kobek, K. (1987). Inhibition by sethoxydim of pigment

accumulation and fatty acid biosynthesis in chloroplasts of Avena seedlings. Z.

Naturforsch 42c: 127 5 -127 9.

López-Casillas, F., Bai, D.H., Luo, X. Kong, I.S., Hermodson, M.A. and Kim, K.H.

(1988). Stn¡cture of the coding sequence of acetyl-coerøyme A carboxylase. Proc.

Natl. Acad. Sci. USA 85: 5784-5788.

t73

Lucas, \M.J., lVilson, C. and Wright, J.P. (1984). Perbutation of Chara plasmalemma

transport function by 214(2',4'-dichlorophenoxy)phenoxy'lpropionic acid. Plant

Physiol. 74:61-66.

Mansooji, 4.M., Holtum, J.A.M., Boutsalis, P., Matthews, J.M. and Powles, S.B.

(1992). Resistance to aryloxyphenoxypropionate herbicides in two wild oats

species (Avenafatua andAvena sterilis). lVeed Sci. 40: 599-605.

Mansooji, A.M. (1994). Herbicide resistance in wild oats, Avend spp. Ph.D. Thesis,

Universþ of Adelaide, Ausralia.

Mallory-Smith, C.4., Thill, D.C., Dial, M.J. and Zemet¡a, R.S. (1990). Inheritance of

sulfonylurea herbicide resistance tn l-actuca spp. V/eed Technol. 4:787-79O.

Marles, M.A.S., Devine, M.D. and Hall, J.C. (1993). Herbicide resistance in Setaria

viridis conferred by a less sensitive form of acetyl coenzyme A carboxylase. Pestic.

Biochem. Physiol. 46: 7 -L4.

Marshall, L.C., Somers, D.4., Dotray, P.D., Gengenbach, B.G., 'Wyse, D.L. and

Gronwald, J.W. (1992). Allelic mutations in acetyl-coenzyme A carboxylase

confer herbicide tolerance in maize. Theor. Appl. Genet. 83:435442.

Marorder, H.L., Prego, I.A. and Cairoli, M.A. (1987). Behaviour of l4C haloxyfop in

common bermudagrass (Cynodon dacryIon) stolons. Weed Sci. 35: 599-603.

Marshall, D.R. and Jain, S.K. (1970). Seed predation in the population dynamics of Avena

fatw andAvenabarbata. Ecology 51: 886-891.

Martin, R.J., McMillian, M.G. and Cook, J.B. (1988). Survey of farm management

practices of the northern wheat belt of New South Wales. Aus. J. Exp. Agric. 28:

499-509.

r74

Manwell, B.D. and Mortimer, A.M. (1994). Selection for herbicide resistance. In

Herbicide Resistance in Plants: Biochemistry and Biology, (S.8. Powles and

J.A.M. Holtum, Eds.), CRC Press Inc., Boca Raton, FL' pp' 1-26'

McAlister, F.M., Holtum, J.A.M. and Powles, S.B. (1995). Dinitroaniline herbicide

resistance in rigid ryegfass (Lolium rigidum). Iffeed Sci. 43: 55-62.

McFadden, J.J., Frear, D.S. and Mansager, E.R. (1989). Aryl hydroxylation of diclofop

by a cytochrome P-450 dependent monooxygenase from wheat. Pestic. Biochem.

Physiol. 34:92-100.

Medd, R.W. and Pandey, S. (1990). Estimating the cost of wild oats (Avena spp.) in the

Australian wheat indusry. Plant Prot. Quaterþ 5: t42'144.

Medd, R.W., McMillan, M.G. and Cook, A.S. (1992). Spray-topping of wild oats (Avena

spp.) in wheat with selective herbicides. Plant Prot. Quarterly 7:62'65.

Miller, S.D. and Nalewaja, J.D. (1990). Influence of burial depth on wild oats (Avena

fatua) seed longevity. V/eed Technol. 4:514-517.

Moss, S.R. (1990). Herbicide cross-resistance in slender foxtail (Alopecurus

myosuroides). Weed Sci. 38: 492496.

Mougin, C., Polge, N., Scalla, R. and Cabanne, F. (1991). Interactions of va¡ious

agrochemicals with cytochrome P450-dependent monoxygenases of wheat cells.

Pestic. Biochem. Physiol. 40: 1-11.

Mueller, T.C., William, W.'W. and Barrett, M. (1989). Antagonism of Johnsongrass

(Sorghum halepense) control with fenoxaprop, haloxyfop and sethoxydimby 2,4-

D. Weed Technol. 3: 86-89.

Mueller, T.C., Ba¡rett, M. and'Witt, W.W. (1990). A basis for the antagonistic effect of

2,4-D on haloxyfop-metþl toxicity to johnsongrass (Sorghum halepense). rWeed

Sci.38: 103-107.

175

Naka¡nura, Y. and Yamada, M. (1975). Fatty acid synthesis by spinach chloroplasts. I.

Properties of fatty acid synthesis from acetate. Plant Cell Physiol. 16z 139-149.

Nakamura, Y. and Yamada, M. (1979). The light-dependent step of de novo synthesis of

long chain fatty acids in spinach chloroplasts. Plant Sci. I-ett. 14:29I-295.

Naylor, J.M. and Simpson, G.M. (1961). Dormancy studies in seeds of Avena fatua.n.

A gibberellin-sensitive inhibitory mechanism in the embryo. Can. J. Bot. 39: 281-

295.

Nikolau, B.J. and Hawke, J.C; (1934). Purification and cha¡acterization of maize leaf

acetyl-coenzyme A carboxylase. Arch. Biochem. Biophys. 228:86-96.

O'Donavan, J.T., Sharma, M.P., Harker, K.N., Maurice, D., Baig, M. and Blackshaw,

R.E. (1994). Wild oat (Avena fatua) populations resistant to triallate are also

resistant to difenzoquat. Weed Sci. 42: 195-199.

Parker,'W.B., Marshall, L.C., Burton, J.D., Somers, D.4., Wyse, D.L., Gronwald, J.W.

and Gengenbach, B.G. (1990a). Dominant mutations causing alterations in acetyl

coenzyme A carboxylase confer tolerance to cyclohexanedione and

aryloxyphenoxypropionate herbicides in maize. Proc. Natl. Acad. Sci. USA 87:

7175-7r79.

Parker, Vf.B., Somers, D.4.,'Wyse, D.L., Keith, R.4., Burton, J.D., Gronwald, J.W.

and Gengenbach, B.G. (1990b), Selection and cha¡acterization of sethoxydim-

tolerant maize tissue cultures. Plant Physiol.92: t22O-1225.

Paterson, J.G. (1976). The distribution and ecology of wild oats (Avena spp.) in Western

Australia. J. Appl. Ecol. 13: 257-26r''

Paterson, J.G., Boyd, V/.J.R. and Goodchild, N.A. (1976). Effect of temperature and

depth of burial on the persistence of seed of. AvenafatuaL. in'Western Australia. J.

Appl. Ecol. 13: 841-847.

176

Peters, N.C.B. (1986). Factors affecting sssdling emergence of different strains of. Avena

fatw L. Weed Res. 26: 29-38.

Powles, S.B. (1986). Appearance of a biotype of the weed, Hordeum glaucum Steud.,

resistant to herbicide Paraquat. $[eed Res. 26: t67-172.

Powles, S.B. and Matthews, J.M. (1992). Multiple herbicide resistance in annual ryegrass

(Lotium rigidum): a driving force for the adoption of the integrated weed

mangement. /n Resistance '91: Acheivements and Developments in Combating

Pesticide Resistance, (I. Denholm, A.L. Devonshire and D.W. Hollomon, Eds.),

Elsevier, NY, pp. 75-87.

Powles, S.8., Tucker, E.S. and Morgan, T.R. (1989). A capeweed (Arctotheca calendula)

biotype in Australia resistant to bipyridyl herbicides. rWeed Sci. 37: 60-62.

Preston, C. (1994). Resistance to photosystem I inhibiting herbicides. In Herbicide

Resistance in Plants: Biochemistry and Biology (S.8. Powles and J.A.M. Holtum,

Eds.), CRC Press Inc., Boca Raton, FL, pp. 6l-82.

Preston, C., Holtum, J.A.M. and Powles, S.B. (1992). On the mechanism of resistance to

paraquat in Hordeum glaucum and I1. leporinum. Delayed inhibition of

photosynthetíc 02 evolution after paraquat application. Plant Physiol. 100: 630-

636.

Preston, C., Balachandran, S. and Powles, S.B. (L994). Investigation of mechanisms of

resistance to bipyridyl herbicides in Arthotheca calendula (L.) Levyns. Plant Cell

Environ. 17 : lll3-1123.

Purba, E. (1993). Factors influencing the development of resistance to the bipyridyl

herbicides in Australia. Ph.D. Thesis, University of Adelaide, Austalia.

Purba, E., Preston, C. and Powles, S.B. (1993a). Inheritance of bipyridyl herbicide

resistance in Anhotheca calendula and Hordeum leporinum. Theor. Appt. Genet.

87:598-602.

t77

Purba, E., Preston, C. and Powles, S.B. (1993b). Paraquat resistance in a biotype of

Vulpia bromoides (L.) S.F. Gray. lVeed Res. 33: 409413.

Quail, P.H. and Carter, O.G. (1968). Survival and seasonal germination of seeds of Avena

Íatua andÁ. ludoviciana. Aust. J. Agric. Res. 19: 721'729.

Ratterman, D.M. and Balke, N.E. (1988). Herbicide disruption of proton gradient

development and maintainance by plasmalemma and tonoplast vesicles from oat

root. Pectic. Biochem. Physiol. 3l:22I-236.

Ratterman, D.M. and Balke, N.E. (1989). Diclofop-methyl increases the proton

permeability of isolated oat root tonoplast. Plant Physiol. 91: 756-765.

Rendina, A.R. and Felts, J.M. (1988). Cyclohexanedione herbicides are selective and

potent inhibitors of acetyl-CoA carboxylase from grasses. Plant Physiol. 86: 983-

986.

Rendina, 4.R., Felts, J.M., Beaudoin, J.D., Craig-Kennard, 4.C., Look, L.L., Paraskos,

S.L. and Hangenah, J.A. (1988). Kinetic characterization, stereoselectivity, and

species selectivity of the inhibition of plant acetyl-CoA carboxylase by the

aryloxyphenoxypropionic acid herbicides. Arch. Biochem. Biophys. 265: 219-

225.

Rendina, A.R., Beaudoin, J.D., Craig-Kennard, A.C. and Breen, M.K. (1989). Kinetic

of inhibition of acetyl-coenzymeA carboxylase by aryloxyphenoxypropionate and

cyclohexanedione graminicides. Proc. British Crop Prot. Conf. -'Weeds: 163-172.

Rendina, 4.R., Craig-Kennard, 4.C., Beaudoin, J.D. and Breen, M.K. (1990).

Inhibition of acetyl-coenzyme A carboxylase by two classes of grass-selective

herbicides. J. Agric. Food Chem.38: 1282-87.

Ritter, R.J. (1989). The spread of triazine-resistant weeds. In Understanding Herbicide

Resistance in W'eeds. Sandoz Crop Prot. Corp., Illinois, pp.16-20.

178

Rgyneberg, T., Balke, N.E. and Lund-Høie, K. (1994). Cycloxydim absorption by

suspension-culn¡red velvetleaf (Abutilon theophrastí Medic.) cells. 'Weed Res. 34:

l-9.

Ryan, G.F. (1970). Resistance of corn¡non groundsel to simazine and atrazine. Weed Sci.

18: 614-616.

Saari, L.L., CotteÍnan, J.C. and Primiani, M.M. (1990). Mechanism of sulfonylurea

herbicide resistance in the broadleaf weed, Kochia scoparia. Plant Physiol.

93:55-61.

Saari, L.L., Cottennan, J.C., Smith, lV.F. and Primiani, M.M. (1992). Sulfonylurea

herbicide resistance in common chickweed, perennial rye$ass and Russian thistle.

Pestic. Biochem. Physiol. 42: 110-1 18.

Saari, L.L., Cottennan, J.C. and Thill, D.C. (1994). Resistance to acetolactate synthase

inhibiting herbicides. /n Herbicide Resistance in Plants: Biochemistry and Biology,

(S.8. Powles and J.A.M. Holtum, Eds.), CRC Press Inc., Boca Raton, FL, Pp.

83-139.

Sasaki, Y., Hakamada, K., Suama, Y., Nagano, Y., Furusawa, I., Matsuno, R. (1993).

Chloroplast-encoded protein as a subunit of acetyl-CoA carboxylase in pea plant.

J. Biol. Chem. 268: 25118-25123.

Schmitzer, P.R., Eilers, R.J. and Cséke, C. (1993). Lack of cross-resistance of

imazaquin-resistantX¿nthium strumoriurn acetolactate synthase to flumetsulam and

cilorimuron. Plant Physiol. 103: 281-283.

Secor, J. and Cséke, C. (1988). Inhibition of acetyl-CoA carboxylase activity by

haloxyfop and tralkoxydim. Plant Physiol. 86: L0-t2.

Secor, J., Cséke, C. and Owen, V/.J. (1989). The discovery of the selective inhibition of

acetyl coenzyme A carboxylase activity by two classes of graminicides. Proc.

British Crop Prot. Conf. -Weeds: 145-154.

t79

Shaaltiel, Y., Chua, N.H., Gepatein, S. and Gressel, J. (1988). Dominant pleiotropy

controls enzymes co-segragating with paraquat resistance'in Conyza bonariensis.

Theor. Appl.Genet. 75: 850-856.

Shimabukuro, R.H., Walsh, W.C. and Hoerauf, R.A. (1979). Metabolism and selectivity

of diclofop-methyl in wild oat and wheat. J. Agric. Food Chem.27:615-620.

Shimabukuro, R.H. (1990). Selectivity and mode of action of the Post emergence herbicide

diclofop-methyl. PGRSA Quarterly lt: 37 -54.

Shimabukuro, R.H. and Hoffer, B.L. (1991). Metabolism of diclofop-methyl in

susceptible and resistant biotypes of Lolium rigidum. Pestic. Biochem. Physiol.

39:251-260.

Shimabukuro, R.H. and Hoffer, B.L. (L992). Effect of diclofop on membrane potentials

of herbicide-resistant and -susceptible annual ryegrass root tips. Plant Physiol. 98:

l4t5-1422.

Shimabukuro, R.H., Walsh, V/.C. and Jacobson, J.A. (1987). Aryl-Ø-glucoside of

diclofop: A detoxification product in wheat shoots and wild oat suspension culture.

J. Agric. Food Chem.35:393-397.

Shoaf, A.R. and Carlson, W.C. (1986). Analytical techniques to measure sethoxydim and

breakdown products. Weed Sci. 34: 745-751.

Shoaf, A.R. and Carlson, W.C. (1992). Stability of sethoxydim and its degradation

products in solution, in soil, and on surfaces. V/eed Sci. 40: 384-389.

Shooler, 4.8., Bell, A.R., Nalewaja, J.D. (1972). Inheritance of siduron tolerance in

foxtail barley. 'Weed Sci. 20: t67-169.

Shorrosh, B.S., Dixon, A.D. and Ohlrogge, J.B. (1994). Molecular cloning,

cha¡acterization, and elicitation of acetyl-CoA carboxylase from alfafa. Proc. Natl.

Acad. Sci. USA 9l: 4323-4327.

180

Smeda, R.J., Barrentine, W.L. and Snipes, C.E. (1993). Johnsongrass (Sorghum

halepense L. Pers.) resistance to post emergence grass herbicides. V/SSA Abstract

33: 53.

Smeda, R.J. and Vaughn, K.C. (1994). Resistance to dinitroaniline herbicides. fn

Herbicide Resistance in Plants: Biochemistry and Biology, (S.8. Powles and

J.A.M. Holtum, Eds.), CRC Press Inc., Boca Raton, FL, pp. 215-228.

Smith, A.M. and Vanden Born, W.H. 1992. Ammonium sulfate increases efficacy of

sethoxydim through increased absorption and translocation. Weed Sci. 40: 351-

358.

Stanger, C.E. and Appleby, A.P. (1989). Italian ryegr¿¡ss (I-olium multiþrum ) accessions

tolerant to diclofop. V/eed Sci. 37: 350-352.

Sterling, T.M. (1994). Mechanisms of herbicide absorption across plant membranes and

accumulation in plant cells. Weed Sci. 42:263-276.

Stoltenberg, D.E., Gronwald, J.W., Wyse, D.L., Burton, J.D., Somers, D.A. and

Gengenbach, B.G. (1939). Effect of sethoxydim and haloxyfop on acetyl

coenzyme A carboxylase activity in Festuca species. Weed Sci. 37 : 512-516.

Stoltenberg, D.E. and Vfeiderholt, R.J. (1993). Giant foxtail (Setaria faberi Herrm.)

resistance to acetyl-CoA carboxylase inhibitors. WSSA Abstract 33: 183.

Stonebridge, W.C. (1981). Selective post-emergence grass weed control in broad-leaf

arable crops. Outl. Agric. 10: 385-392.

Swisher, B.A. and Corbin, F.T. (1982). Behavior of BAS-90520H in soybean and

Johnsongrass plant and cell cultures. \Veed Sci. 30: 640-650.

Takai, T., Yokohama, C., Wada, K. and Tanabe, T. (1988). Primary srtucture of chicken

liver acetyl CoA carboxylase deduced from cDNA sequence. J. Biol. Chem. 263:

2651-2657.

181

Tal, 4., Romano, M.L.Stephenson, G.R., Schwan, A.L. and Hall, J.C. (1993).

Glutathione conjugation: a detoxification pathway for fenoxaprop-ethyl in barley,

crabgrass, oat, and wheat. Pestic. Biochem. Physiol. 46: 190-199.

Tanaka, Y., Chisaka, H. and Saka, H. (1986). Movement of paraquat in resistant and

susceptible biotypes of Erigeron Philadelphiczs and E. canadensis. Physiol. Plant.

66: 605-608.

Tardif, F.J., Holtum, J.A.M. and Powles, S.B. (1993). Occurrence of a herbicide-

resistant acetyl-coenzyme A carboxylase mutant in annual ryegrass (Lolium

rigidum) selected by sethoxydim. Planta 190: 176-181.

Tardif, F.J. and Powles, S.B. (1993). Target-site-based resistance to herbicides inhibiting

acetyl-CoA carboxylase. Proc. Brighton Crop Prot. Conf. -Weeds: 533-540.

Ta¡dif, F.J. and Powles, S.B. (1994). Herbicide multiple-resistance in a Inlium rigidum

biotype is endowed by multiple mechanisms: isolation of a subset with resistant

acetyl-CoA carboxylase. Physiol. Plant. 9L: 488494.

Thomas, H. and Jones, D.D. (1976). Origin and identification of weed species of Avena.

/n V/ild Oats in lVorld Agriculture, (D.P. Jones, Ed.), Agricultural Resea¡ch

Council, London, pp. 1-18..

Thurston, J.M. (1951). Some experiments and field observations on the germination of

wild oats (A. fatua and A. ludoviciana) seeds in soil and the emergence of

seedlings. Ann. Appl. Biol.38: 812-832.

Thurston, J.M. (1959). A comparative study of the growth of wild oats ( Avenafatua L.

andAvena ludovicianø Dur.) and of cultivated cereals with varied nitrogen supply.

Ann. Appl. Biol. 47:7t6-793.

Thurston, J.M. (1966). Survival of seeds of wild oats (Avena fatua and A. Iudoviciana)

and cha¡lock (Sinøpsr,s ravensis) in soil under leys. rWeed Res. 6: 67-80.

t82

Todd, B.G. and Stobbe, E.H. (1980). The basis of the antagonistic effect of 2,4-D on

diclofop-methyl toxicity to wild oat (Avenafatua). lVeed Sci. 2E: 371-377.

Tritter, S.4., Holl, F.B. and Todd, B.G. (1987). Diclofop-methyl and diclofoP uptake in

oat (Avena sativa L.) protoplasts. Can. J. Plant Sci. 67: 2t5-223.

Tucker, E.S. and Powles, S.B. (1991). A biotype of hare barley (Hordeum leporinum)

resistant to paraquat and diquat. Weed Sci.39: 159-162.

Tucker, T.4., Langeland, K.A. and Corbin, F.T. (1994). Absorption and translocation of

t{C-imazapyr and l4C-glyphosate in alligatorweed (Altemanthera philoxeroides).

Weed Technol. 8:32-36.

Turnham, E. and Northcote, D.H. (1983). Changes in the activity of acetyl-CoA

carboxylase during rape-seed formation. Biochem. I . 212: 223 -229 .

Uchiyama, M., Washio, N., Ikai, T., Igarashi, H. and Suzuki, K. (1986). Stereospecific

responses to (R)-(+)-and (S)-(l-quizalofop-ethyl in tissues of several plants.

Pestic. Sci. 11: 459-467.

Watkins, F.B. (1971). Efffects of annual dressings of nitrogen fertilizer on wild oat

infestations. Weed Res. 11: 292-3OI.

Walker, K.4., Ridley, S.M., Lewis, T. and Harwood, J.L. (1989). Action of aryloxy-

phenoxy carboxylic acid on lipid metabolism. Rev. Weed Sci. 4: 7l-84.

'Walter, H., Koch, W. and Muller, F. (1980). EinfluB des Applikationsortes auf die

Penetration und Translokation von diclofop-methyl bei Flughafer (AvenafatuaL.).

Weed Res. 20: 325-331.

Weber, 4., Fischer, E., VonBraitz, H.S. and Lüttge, U. (1988). The effects of the

herbicide sethoxydim on transport processes in sensitive and tolerant grass species.

I. Effects on the electrical membrane potential and alanine uptake. Z. Naturforsch

43c:249-256.

183

Weber, A. and Lünge, U. (1988). The effects of the herbicide sethoxydim on transport

processes in sensitive and tolerant grass species. tr. Effects on membrane-bound

redox systems in plant cells. Z. Naturforsch. 43c: 257-263.

V/halley, R.D.B. and Burfrtt, J.M; (1972). Ecotypic variation ín A. fatuaL., A. sterilis

L.(A. Iudoviciana), and A. barbata Pott. in New South V/ales and southern

Queensland. Aust. J. Agric. Res.23: 799-810.

Wilson, B.J. (1972). Studies of the fate of Avena fatua seeds on cereal stubble, as

influenced by autumn treatment. hoc. British Crop Prot. Conf. -'Weeds: 242-247.

\I/ilson, R.G. (1988). Biology of weed seeds in soil. /n V/eed Management in

Agroecosystem: Ecological Approaches, (M.4. Altien and M. Liebman, Eds.),

CRC PRess, Boca Raton, FL, pp. 25-39.

V/ilson, B.J. and Cussans, G.rW. (1975). A study of the population dynamics of Avena

fatua L. as influenced by straw burning, seed shedding and cultivation. 'Weed Res.

15:249-258.

Wink, O., Dorn, E. and Beyermann, K. (1984). Metabolism of the herbicide Hoe 33171in

soybean. J. Agric. Food Chem.32 187-192.

Worthing, C.R. and Hance, R.J. (1991). The Pesticide Mannual: A World Compendium.

The British Crop Protection Council, Unwin Brothers Ltd., Old'Woking, Surrey,

UK, 1140p.

rWright, J.P. and Shimabukuro, R.H. (1987). Effects of diclofop and diclofop-methyl on

the membrane potentials of wheat and oat coleoptiles. Plant Physiol. 85: 188-193.

Zimmerlin, A. and Durst, F. (1990). Xenobiotic metabolism in plants: aryl hydroxylation

of diclofop by a cytochrome P-450 erxzyme from wheat. Phytochem.29: 1729-

t732.

184

Zimmerlin, A. and Durst, F. (1992). Aryl hydroxylation of the herbicide diclofop by a

wheat cytochfome P450 monoxygenase. Plant Physiol. 1ü): 874-881.

185

Appendix

The following l0 tables (Tables A.l-10) show the herbicide use histories of the ten resistant

biotlpes of wild oat.

Table 4.1 History of herbicide application of field from which wild oat biotype SAS 1

was collected

Year Crop Herbicide use

1981 wheat diclofop-methyl

1982 pasture none

1983 medic trifluralin, fluazifop-butyl

1984 rapeseed trifluralin, fluazifop-butyl

1985 wheat triallate, diclofop-methyl

1986 wheat diclofop-methyl

1987 clover trifluralin, fluazifop-butyl

1988 wheat glyphosate

1989 clover trifluralin, haloxyfop-ethoxyethyl (failed)

186

Table 4.2 History of herbicide application of field from which wild oat biotype NAS 4

was collected

Year Crop Herbicide use

1980 wheat diclofopmethyl

1981 wheat diclofopmethyl

1982 lupin nifluralin

1983 wheat diclofopmethyl

1984 wheat diclofopmethyl

1985 wheat diclofopmethyl

1986 pea trifluralin, fluazifop-butyl

1987 wheat fralkoxydim

1988 wheat talkoxydim

1989 canola trifluralin, haloxyfop-ethoxyethyl

1990 wheat fenoxaprop-ethyl (tailed)

187

Table 4.3 History of herbicide application of field from which wild oat biotype NAF 6

was collected

Year Crop Herbicide use

1988 pasture none

1989 wheat flamprop-methyl

1990 wheat fenoxaprop-ethyl

L99l wheat fenoxaprop-ethyl (failed)

Table 4.4 History of herbicide application of field from which wild oat biotype NAS 14

was collected

Year Crop Herbicide use

t982 wheat none

1983 wheat none

1984 wheat none

1985 lupin simazine, fluazifop-butyl

1986 wheat diclofop-methyl

1987 wheat diclofop-methyl

1988 lupin simazine, fluazifop-butyl

1989 wheat diclofopmethyl (failed)

188

Table AS History of herbicide application of field from which wild oat biotype NAF 8

was collected

Year Crop Herbicide use

1988 pasture

lupin

wheat

wheat

1989

1990

1991

Table 4.6 History of herbicide application of field from which wild oat biotype NAS 2

was collected

Year Crop Herbicide use

none

fluazifop-butyl

diclofopmethyl

diclofopmethyl (tailed)

1980

1981

1982

1983

1984

1985

1986

1987

1988

1989

wheat

wheat

lupins

wheat

wheat

wheat

pea

wheat

wheat

canola

wheat

diclofop-methyl

diclofop-methyl

tifluralin

diclofop-methyl

diclofop-methyl

diclofop-methyl

trifl uralin, fl uazifop-butyl

tralkoxydim

tralkoxydim

trifluralin, haloxyfop-ethoxyethyl

talkoxydim (failed)1990

r89

Table 4.7 History of herbicide apptication of field from which wild oat biotype NAS 6

was collected

Year Crop Herbicide use

1988 pasture none

1989 wheat flamprop-methyl

1990 wheat fenoxapropethyl

1991 wheat fenoxaprop-ethyl (tailed)

Table 4.8 History of herbicide application of field from which wild oat biotype SAF 14

was collected

Yea¡ Crop Herbicide use

1984 barley diclofop-methyl

1985 rapeseed diclofop-methyl

1986 wheat diclofopmethyl

L987 bean simazine, fluazifop-butyl

1988 wheat diclofop-methyl

1989 pea fluazifop-butyl, glyphosate

1990 wheat fenoxaprop-ethyl (failed)

190

Table 4.9 History of herbicide application of field from which wild oat biotype SAF 34

was collected

Year Crop Herbicide use

l98l pasture none

t982 titicale diclofopmethYl

1983 pasture none

1984 wheat diclofop-methyl

1985 pasture none

1986 wheat diclofop-methyl

t987 beans fluazifopbutyl

1988 wheat diclofop-methyl

1989 barley tralkoxydim

1990 wheat diclofop-methyl (failed)

191

Table 4.10 History of herbicide application of field from which wild oat biotype V/AF 12

was collected

Year Crop Herbicide use

1985 pastue none

1986 pasture none

1987 pasture none

1988 wheat tiallaæ

1989 wheat diclofopmethyl

1990 wheat diclofopmethyl

t99t wheat triasulfuron