You are on page 1of 191

INFLUENCE OF APPLICATION PLACEMENT, RESISTANCE

GENOTYPE, AND PPO-INHIBITING HERBICIDE ON THE PPO-


RESISTANCE PHENOTYPE IN WATERHEMP
by
Jesse Andrew Haarmann

A Dissertation
Submitted to the Faculty of Purdue University
In Partial Fulfillment of the Requirements for the degree of

Doctor of Philosophy

Department of Botany and Plant Pathology


West Lafayette, Indiana
May 2023
THE PURDUE UNIVERSITY GRADUATE SCHOOL
STATEMENT OF COMMITTEE APPROVAL

Dr. William G. Johnson, Chair


Department of Botany and Plant Pathology

Dr. Bryan G. Young


Department of Botany and Plant Pathology

Dr. Joshua R. Widhalm


Department of Horticulture and Landscape Architecture

Dr. Robert E Pruitt


Department of Botany and Plant Pathology

Dr. Patrick J. Tranel


Department of Crop Sciences
University of Illinois Urbana – Champaign

Approved by:
Dr. Tesfaye Mengiste
Dedicated to my wife Brittany
ACKNOWLEDGMENTS

Completion of my degree would not have been possible without the help and support of

several groups of people. First, I would like to thank my advisor Dr. Bill Johnson for allowing me

to stay on at Purdue for my PhD degree. Also, thank you to my graduate committee, Dr. Bryan

Young, Dr. Josh Widhalm, Dr. Bob Pruitt, and Dr. Pat Tranel for the guidance, advisement, and

use of facilities. I also want to thank Julie Young for all of the assistance with logistics and

planning, methods advisement, and supply acquisition. To all my fellow graduate students and the

undergraduate employees at Purdue, both past and present, especially Cade Hayden, Dr. Joe Ikley,

Dr. Cara McCauley, Connor Hodgskiss, Stephanie Desimini, Dr. Nick Steppig, Marcelo Zimmer,

Dr. Ben Westrich, Dr. Matthew Osterholt, Lucas Maia, Rose Vagedes, and Claudia Bland, thank

you for all of your help over the years, and for your friendship. A big thank you also goes to Dr.

Mearaj Shaikh and the rest of the Widhalm lab for assistance with extractions and PCR.

To my wife, Brittany, thank you for remaining by my side during these challenging years.

To my parents, Tom and Janet Haarmann, thank you for supporting me through my never-ending

education and inspiring a passion for lifelong learning and achievement and a love of agriculture.

I owe so much of who I am today to your encouragement and guidance. I am truly honored and

humbled to have received the support and friendship of so many wonderful people over the last

four years.

4
TABLE OF CONTENTS

LIST OF TABLES .......................................................................................................................... 8


LIST OF FIGURES ...................................................................................................................... 10
ABSTRACT.................................................................................................................................. 13
A REVIEW OF THE LITERATURE ................................................................. 15
1.1 PPO is an Important Herbicide Target .............................................................................. 15
1.2 Trifludimoxazin, a New PPO-Inhibiting Herbicide .......................................................... 16
1.3 PPO-Inhibitor Resistance Mutations................................................................................. 19
1.4 Herbicide Resistance Genotyping Assays ........................................................................ 22
1.5 Herbicide Dose Responses ................................................................................................ 24
1.6 Biologically Effective Dose .............................................................................................. 26
1.7 Herbicide Effectiveness PRE vs POST............................................................................. 28
1.8 Natural Selection in Weed Populations ............................................................................ 29
1.9 Organelle Targeting of Plant Enzymes ............................................................................. 30
1.10 PPO-Pathway Control and Flux ..................................................................................... 32
1.11 Summary......................................................................................................................... 36
1.12 Justification of Research................................................................................................. 37
1.13 Literature Cited ............................................................................................................... 37
ARE PRE-APPLIED PPO INHIBITORS SELECTING FOR NEW
RESISTANCE MUTATIONS IN WATERHEMP (AMARANTHUS TUBERCULATUS)? .... 52
2.1 Abstract ............................................................................................................................. 52
2.2 Introduction ....................................................................................................................... 53
2.3 Materials and Methods ...................................................................................................... 57
2.4 Results and Discussion ..................................................................................................... 60
2.4.1 Flumioxazin ............................................................................................................... 60
2.4.2 Fomesafen .................................................................................................................. 61
2.4.3 Oxadiazon .................................................................................................................. 61
2.4.4 Saflufenacil ................................................................................................................ 63
2.4.5 Sulfentrazone ............................................................................................................. 63
2.4.6 Trifludimoxazin ......................................................................................................... 64

5
2.4.7 Variance in Observed Resistance .............................................................................. 64
2.4.8 Practical Implications ................................................................................................ 66
2.5 Literature Cited ................................................................................................................. 68
DOES TRIFLUDIMOXAZIN SELECT FOR RESISTANT WATERHEMP
(AMARANTHUS TUBERCULATUS) SIMILARLY TO OTHER PPO-INHIBITING
HERBICIDES? ........................................................................................................................... 86
3.1 Abstract ............................................................................................................................. 86
3.2 Introduction ....................................................................................................................... 87
3.3 Materials and Methods ...................................................................................................... 91
3.3.1 Site Selection and Experimental Design.................................................................... 91
3.3.2 Data Collection and Analysis .................................................................................... 93
3.4 Results and Discussion ..................................................................................................... 95
3.4.1 PRE Herbicide Efficacy ............................................................................................. 95
3.4.2 PRE Resistance Selection .......................................................................................... 96
3.4.3 POST Efficacy ......................................................................................................... 100
3.4.4 POST Resistance Selection...................................................................................... 101
3.5 Acknowledgements ......................................................................................................... 104
3.6 Literature Cited ............................................................................................................... 104
INVESTIGATING THE POTENTIAL CO-OCCURRENCE OF TARGET-SITE
AND NON-TARGET-SITE RESISTANCE IN WATERHEMP (AMARANTHUS
TUBERCULATUS) POPULATIONS ....................................................................................... 122
4.1 Abstract ........................................................................................................................... 122
4.2 Introduction ..................................................................................................................... 123
4.3 Materials and Methods .................................................................................................... 127
4.3.1 Plant Materials ......................................................................................................... 127
4.3.2 Spray Applications and Preparations ....................................................................... 128
4.3.3 Experimental Design, Data Collection, and Analysis ............................................. 129
4.4 Results and Discussion ................................................................................................... 129
4.4.1 Response to Detoxification Inhibitors Alone .......................................................... 129
4.4.2 PPO-Resistance Response ....................................................................................... 131
4.5 Literature Cited ............................................................................................................... 136

6
INVESTIGATIONS OF PPO-INHIBITOR RESISTANCE IN WATERHEMP
(AMARANTHUS TUBERCULATUS) LACKING KNOWN RESISTANCE-CONFERRING
MUTATIONS ......................................................................................................................... 148
5.1 Abstract ........................................................................................................................... 148
5.2 Introduction ..................................................................................................................... 149
5.3 Materials and Methods .................................................................................................... 152
5.3.1 Preliminary Purifying Screen .................................................................................. 152
5.3.2 Gene Sequencing and PCR Reactions ..................................................................... 155
5.3.3 Dose-Response Assays ............................................................................................ 156
5.3.4 Gene Expression Assay ........................................................................................... 157
5.3.5 Lipid Peroxidation Assay ........................................................................................ 158
5.4 Results and Discussion ................................................................................................... 159
5.4.1 Preliminary Purifying Screen .................................................................................. 159
5.4.2 Target-Site Gene Sequencing .................................................................................. 160
5.4.3 Dose-Response Assays ............................................................................................ 161
5.4.4 Gene-Expression Assay ........................................................................................... 164
5.4.5 Lipid-Peroxidation Assay ........................................................................................ 165
5.5 Acknowledgements ......................................................................................................... 167
5.6 Literature Cited ............................................................................................................... 167
VITA ........................................................................................................................................... 187

7
LIST OF TABLES

Table 2.1. Discovery of novel PPO-inhibitor resistance mutations in waterhemp and Palmer
amaranth. ....................................................................................................................................... 76
Table 2.2. Herbicides used for PRE dose-response experiments on susceptible and resistant
waterhemp biotypes with ΔG210 and R128G mutations.............................................................. 77
Table 2.3. Response of ΔG210, R128G and susceptible biotypes of waterhemp to PRE application
of PPO-inhibiting herbicides in the greenhouse, as measured by seedling emergence (LD50) and
shoot biomass (GR50). ................................................................................................................... 78
Table 2.4. Resistance ratios between susceptible, R128G, and ΔG210 waterhemp biotypes for six
PPO-inhibiting herbicides applied PRE in the greenhouse........................................................... 79
Table 3.1. Site characteristics for field trials conducted in 2020 and 2021a. .............................. 111
Table 3.2. Trial initiation and tissue collection dates for PRE and POST field trials at
Throckmorton and Davis field sites in 2020 and 2021. .............................................................. 112
Table 3.3. Weekly rainfall totals for Davis and Throckmorton field locations in 2020 and 2021.
..................................................................................................................................................... 113
Table 3.4. Control of waterhemp recorded 35 days after PRE herbicide application at
Throckmorton and Davis field sites in 2020 and 2021. .............................................................. 114
Table 3.5. Waterhemp emergence at the time of each tissue collection following PRE herbicide
treatments at Throckmorton and Davis Field Sites in 2020 and 2021. ....................................... 115
Table 3.6. Frequency of the first waterhemp to emerge after PRE PPO-inhibitor application that
are heterozygous or homozygous for the ΔG210 mutation in 2020 and 2021 at Throckmorton and
Davis Field Sites. ........................................................................................................................ 116
Table 3.7. Zygosity percentages of ΔG210 in the first waterhemp to emerge after treatment with
PRE PPO-inhibitor application at Throckmorton and Davis field sites in 2020 and 2021......... 117
Table 3.8. Herbicide soil properties of PPO inhibitors used in the current experiments. ........... 118
Table 3.9. Waterhemp control and survival assessed 14 days after POST PPO-inhibitor treatment
at Throckmorton and Davis field sites in 2020 and 2021. .......................................................... 119
Table 3.10. Frequency of surviving waterhemp that have at least one ΔG210 allele after treatment
with POST herbicide. .................................................................................................................. 120
Table 3.11. Zygosity percentages of the ΔG210 mutation in waterhemp surviving treatment with
POST PPO-inhibitor treatments at Throckmorton and Davis field sites in 2020 and 2021. ...... 121
Table 4.1 Waterhemp control after treatment with a detoxification inhibitor followed by fomesafen
application at 0, 20, or 60 g ai ha-1 accessed 14 days after treatment. ........................................ 143

8
Table 4.2. Waterhemp height and biomass after treatment with a detoxification inhibitor followed
by fomesafen application at 0, 20, or 60 g ai ha-1 accessed 14 days after treatment................... 144
Table 4.3. Waterhemp height and biomass after treatment with a detoxification inhibitor followed
by fomesafen application at 0, 20, or 60 g ai ha-1 accessed 14 days after treatment................... 145
Table 5.1. Phenotypic assignment for each plant sprayed with fomesafen at 14 days after treatment.
..................................................................................................................................................... 174
Table 5.2. Amino acid polymorphisms observed in isolated waterhemp PPX2 genes that survived
a dose of fomesafen. ................................................................................................................... 175
Table 5.3. Amino acid polymorphisms observed in isolated waterhemp PPX1 genes that survived
a dose of fomesafen. ................................................................................................................... 176
Table 5.4. Comparison of GR50 and LD50 values and R:S ratios calculated from waterhemp
biomass and survival at 14 days after fomesafen treatment for one susceptible population (Des),
two resistant populations with ΔG210 (Car and Was), and four populations displaying a delayed
regrowth resistance phenotype (Fran, IA-340, IA-358, and IA-369). ........................................ 177

9
LIST OF FIGURES

Figure 1.1. Names and structure of PPO-inhibiting herbicide molecules commonly mentioned in
the literature. ................................................................................................................................. 50
Figure 1.2. Tetrapyrrole synthesis pathway as presented in (Czarnecki et al. 2012) Abbreviations:
ALAD, ALA dehydratase; CAO, Chl a oxygenase; CBR, chlorophyll b reductase; ChlS,
chlorophyll synthase; CPO, coproporphyrinogen III oxidase; DVR, divinyl protochlorophyllide
reductase; FeCh, Fe chelatase; FLU, flourescent; GluRS, glutamyl-tRNA synthetase; GluTR,
glutamyl-tRNA reductase; GluTRBP, GluTR binding protein; GSAT, glutamate-1-semialdehyde
aminotransferase; HBS, hydroxymethylbilane synthase; HCAR, 7-hydroxymethyl chlorophyll a
reductase; HO, heme oxygenase; MgCh, Mg chelatase; MTF, Mg protoporphyrin IX
methyltransferase; PBS, phytochromobilin synthase; POR, light dependent NADPH-
protochlorophyllide oxidoreductase; PPOX, protoporphyrinogen IX oxidase; UROD,
uroporphyrinogen III decarboxylase; UROM, uroporphyrinogen III methyltransferase; UROS,
uroporphyrinogen III synthase. Colors represent different organelle localizations. ..................... 51
Figure 2.1. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to
flumioxazin applied preemergence. Emergence and biomass were evaluated 14 days after
treatment and normalized to a percentage of the untreated for each biotype. .............................. 80
Figure 2.2. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to
fomesafen applied preemergence. Emergence and biomass were evaluated 14 days after treatment
and normalized to a percentage of the untreated for each biotype. .............................................. 81
Figure 2.3. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to
oxadiazon applied preemergence. Emergence and biomass were evaluated 14 days after treatment
and normalized to a percentage of the untreated for each biotype. .............................................. 82
Figure 2.4. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to
saflufenacil applied preemergence. Emergence and biomass were evaluated 14 days after treatment
and normalized to a percentage of the untreated for each biotype. .............................................. 83
Figure 2.5. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to
sulfentrazone applied preemergence. Emergence and biomass were evaluated 14 days after
treatment and normalized to a percentage of the untreated for each biotype. .............................. 84
Figure 2.6. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to
trifludimoxazin applied preemergence. Emergence and biomass were evaluated 14 days after
treatment and normalized to a percentage of the untreated for each biotype. .............................. 85
Figure 4.1. Isolated lesions on waterhemp leaves occurring after NBD-Cl application. Photo taken
at 48 h after treatment. ................................................................................................................ 146
Figure 4.2. The IL-WAS waterhemp population was sprayed with detoxification inhibitor
treatments followed by 60 g ha-1 of fomesafen. Plants portrayed in the photographs are replications
of the same treatment. ................................................................................................................. 147

10
Figure 5.1. Waterhemp control 14 days after fomesafen application for susceptible, ΔG210
resistant, and unknown cause resistant waterhemp from 6 populations. Populations are pooled due
to lack of significance in ANOVA. Means were separated using Tukeys HSD at α=0.05. Presence
of ΔG210 was confirmed by qPCR assay of leaf tissue collected after waterhemp survival.
Assignment to putative novel resistance mechanism group was made if the plant was controlled
less than susceptible plants and was confirmed as being absent of the ΔG210 mutation via qPCR
assay of leaf tissue collected after waterhemp survival. Assignment to susceptible group was made
at waterhemp death or individual waterhemp control was similar to the susceptible waterhemp
population and was confirmed as being absent of the ΔG210 mutation via qPCR assay of leaf tissue
collected after waterhemp survival. ............................................................................................ 178
Figure 5.2. Phylogenetic tree of selected Miseq PPX2 sequences and multiple Genbank accessions
of waterhemp and closely related species. Accessions shown in red are more closely related to
Palmer amaranth and redroot pigweed than wild type waterhemp accessions. .......................... 179
Figure 5.3. Comparison of plant symptomatology at 24 hours after treatment with fomesafen at 20
g ha-1. Colored boxes represent different types of populations with red indicating ΔG210 resistant,
yellow indicating susceptible, and blue indicating delayed regrowth phenotype populations. .. 180
Figure 5.4. Comparison of plant symptomology between populations at 3d after treatment with 7
rates of fomesafen. Red arrows in each frame point to the 20 g ha-1 and 63 g ha-1 rate of fomesafen
treatments. ................................................................................................................................... 181
Figure 5.5. Comparison of plant symptomology between populations at 14d after treatment with 8
rates of fomesafen. Red arrows in each frame point to the 20 g ha-1 and 63 g ha-1 rate of fomesafen
treatments. ................................................................................................................................... 182
Figure 5.6. Dose-response curves of susceptible (Des) ΔG210 resistant (Car and Was), and putative
novel R mechanism populations (Fran, IA-340, IA-358, and IA-369) to fomesafen. Aboveground
biomass was harvested 14 days after fomesafen treatment and normalized to a percentage of the
untreated for each population...................................................................................................... 183
Figure 5.7. Dose-response curves of susceptible (Des) ΔG210 resistant (Car and Was), and putative
novel R mechanism populations (Fran, IA-340, IA-358, and IA-369) to fomesafen. Waterhemp
survival was assessed 14 days after fomesafen treatment. ......................................................... 184
Figure 5.8. Mean gene expression of protoporphyrinogen oxidase 1 (PPX1) and 2 (PPX2) in
susceptible (Des) ΔG210 resistant (Car and Was), and four resistant populations with a putative
novel mechanism (Fran, IA-340, IA-358, and IA-369) prior to fomesafen treatment as determined
by qPCR assays. Expression values for each target are normalized to that of elongation factor-1
alpha and are expressed as percentages of the susceptible population. Means of each population
are indicated by X symbol. Overall F test P values were 0.077 and 0.0438 for PPX1 and PPX2
respectively. Expression differences between Was and IA-358 PPX2 were the only mean
differences using Tukey’s HSD (α = 0.05). ................................................................................ 185
Figure 5.9. Lipid peroxidation as indicated by MDA content of mature leaf tissue (top) and
meristem/young leaf tissue (bottom) in known susceptible (Des), ΔG210 resistant (Car) and two
resistant populations with putative novel mechanism (Fran and IA-369) following treatment with
fomesafen at 20 g ha-1. Significant differences of Fran and IA-369 populations compared to Des

11
and Car populations are denoted by * and † symbols respectively using Fishers protected LSD
(α = 0.05)..................................................................................................................................... 186

12
ABSTRACT

The PPO inhibitors are a valuable group of herbicides that provide soil-residual and foliar

control of glyphosate-resistant Amaranthus species. The ΔG210 mutation in the PPX2 gene

confers PPO-inhibitor resistance and has been present in the Midwest for more than a decade. Until

recently, PPO-inhibitor resistance in waterhemp was attributable to just the ΔG210 mutation in the

PPX2 gene, but recently, several new PPO-resistant biotypes have been discovered in waterhemp

and Palmer amaranth. A possible explanation is a change in PPO-inhibitor use patterns and

commonly used active ingredients.

Research was conducted to directly compare the ΔG210 mutation with the recently

discovered R128G mutation to PPO inhibitors applied PRE. The greatest resistance observed was

to fomesafen with a 20- to 37-fold resistance ratio. Other herbicides tested had resistance ratios

less than 9 for both mutations. Sulfentrazone was the only herbicide for which the R128G mutation

conferred greater resistance than did the ΔG210 mutation with a 0.48 ΔG210 to R128G ED50 ratio

(4.2 vs 8.8 R/S ratios). Overall, the data do not support our hypothesis that the R128G mutation

was selected for by soil-applied PPO inhibitors. We conclude that the R128G mutation in

waterhemp is not more robust than the ΔG210 mutation with respect to conferring resistance to

PPO inhibitors applied preemergence. Furthermore, there is no evidence that the utility of PPO

inhibitors applied preemergence will diminish any further as a result of the R128G mutation

increasing in frequency.

A set of field trials was conducted to investigate how a new PPO inhibitor, trifludimoxazin,

will select for resistant biotypes in the field. Plants that emerged through or survived a PPO-

inhibitor application were genotyped for the ΔG210 mutation. Overall, a greater number of

resistant plants survived the foliar herbicide applications than emerged through soil applications.

13
Trifludimoxazin did not increase the frequency of PPO-resistant individuals when applied to soil,

but when applied to foliage, increased the frequency of PPO-resistant individuals by 2.5- to 2.6-

fold, similar to other PPO inhibitors applied to foliage.

Several waterhemp populations have been identified that have PPO-inhibitor resistance not

completely explained by known target-site mutations. Malathion and NBD-Cl, known inhibitors

of cytochrome P450 and GST, were coapplied with fomesafen to partially reverse resistance to

fomesafen. While there was some support for the initial hypothesis, that fomesafen detoxification

is contributing to overall resistance, the lack of consistency between herbicide rates and the

noticeable phytotoxic response to malathion and NBD-Cl renders it impossible to rule out additive

effects from the multiple applied xenobiotics. We conclude that our methodology was insufficient

to address our hypothesis.

Another set of waterhemp populations had a resistance phenotype in the absence of target-

site mutations. After a purifying screen and confirmed absence of target site resistance mutations,

the progeny generation of four waterhemp populations were subjected to dose-response, target-

site expression, and lipid-peroxidation assays. Models for lethal dose (LD) indicate a less robust

R phenotype than ΔG210, with LD50 ratios of only 1.9- to 6.1- vs 19- to 27-fold resistance in the

ΔG210 populations. A target site expression experiment and lipid peroxidation experiment were

inconclusive, but provided some evidence of increased target-site expression or increased

antioxidant capacity as causal mechanisms, although no mechanisms have been fully ruled out.

These data add to the body of literature that suggests that many factors contribute to PPO-inhibitor

resistance. With greater utilization of PPO-inhibitors in the future, the complexity of resistance

will likely increase also.

14
A REVIEW OF THE LITERATURE

1.1 PPO is an Important Herbicide Target

Protoporphyrinogen IX oxidase (PPO)-inhibiting herbicides are an important class of

herbicides used worldwide. First introduced in the 1960s, these herbicides provide preemergence

(PRE) and postemergence (POST) weed control in a variety of crops. Major chemical classes

include diphenylethers, N-Phenyl-imides, N-Phenyl-triazolinones, and N-Phenyl-oxadiazolones,

and thiadiazoles (Hao et al. 2011) (Figure 1.1).

PPO enzymes are present in nearly all life including humans (Maneli et al. 2003).

Deficiencies in PPO enzyme activity in humans result in a pathological condition known as

variegate porphyria. Symptoms include sensitivity to sunlight and surface tissue damage which

appear to be analogous to phytotoxic herbicide symptoms (Ramanujam and Anderson 2015).

PPO inhibitors target the last common step in heme and chlorophyll biosynthesis,

protoporphyrinogen IX oxidase (Matringe et al. 1989) (Figure 1.2). More detail about specific

pathway enzymes will occur later in this review. Phytotoxicity and plant death occurs via

generation of reactive oxygen species following herbicide application. Inhibition of PPO leads to

buildup of the substrate protoporphyrinogen (protogen). Protogen is actively transported out of the

plastid where it is enzymatically oxidized to protoporphyrin (Proto) by non-specific oxidases

(Jacobs and Jacobs 1993, Lee et al. 1993). Proto rings are potent photo dynamic molecules which

absorb photons and discharge energy. While in the cytosol, energy absorbed by proto is discharged

into oxygen radicals which initiate a cascade of lipid peroxidation (Jacobs and Jacobs 1993,

Matringe et al. 1989). Lipid peroxidation leads to the destruction of cell membranes and eventual

cell death.

15
Despite their prolonged period of use, resistance to PPO inhibitors is relatively uncommon

in terms of number of resistant species (Heap 2023). The first documented incidence of resistance

was not until 2001 with waterhemp [Amaranthus tuberculatus (Moq.) J.D.Sauer] (Heap 2023,

Patzoldt et al. 2006, Shoup et al. 2003), followed shortly thereafter by common ragweed (Ambrosia

artemissifolia L.) (Heap 2023, Rousonelos et al. 2012). In the past 10 years, there have been 10

new species documented as resistant to PPO inhibitors for 14 resistant species worldwide (Heap

2023). The most prolific and problematic resistant species, particularly in the United States, have

been waterhemp and Palmer amaranth (Amaranthus palmeri S. Wats). Increased incidence of

PPO-inhibitor resistance in these species is primarily due to the increased use of PPO inhibitors in

crops after widespread glyphosate resistance. Further discussion about resistance to PPO inhibitors

can be found later in this review.

1.2 Trifludimoxazin, a New PPO-Inhibiting Herbicide

PPO-inhibiting herbicides are a place of renewed herbicide development interest. With the

introduction of glyphosate-resistant crops in 1996, weed control methods shifted away from

multiple products applied PRE and POST to nearly exclusively glyphosate (Young 2006).

Exclusive use of glyphosate led to a nearly complete halt of investment by the agro-chemical

industry into new herbicide chemistries (Dayan 2019, Duke 2012). As a result, a new herbicide

mechanism of action has not been released in over 25 years. At present, 15 years after the

widespread incidence of glyphosate resistance, there are no new mechanisms of action on the

immediate horizon for corn and soybean use (Dayan 2019). With the increased incidence of

glyphosate resistance, pesticide manufacturers have resorted to releasing existing compounds that

have activity on known herbicide targets that can be effective on glyphosate-resistant weed species.

PPO is a common target site for this purpose (Hao et al. 2011, Meazza et al. 2004). One example

16
is saflufenacil. Saflufenacil is a PPO inhibitor that was brought to market by BASF in 2010 as a

solution to glyphosate resistance in horseweed (Conyza canadensis) (Grossmann et al. 2010,

Kraehmer et al. 2014).

Newer PPO-inhibiting compounds are currently in development. These new PPO-

inhibiting compounds putatively control PPO-inhibitor resistant biotypes (Armel et al. 2017,

Steppig 2022, Witschel et al. 2021). More is currently known about BASF’s compound

trifludimoxazin than other similar PPO inhibitors under development. Trifludimoxazin has PRE

and POST activity and is active on major Midwest weed species including waterhemp, Palmer

amaranth, and giant ragweed (Steppig 2022). It does not have crop selectivity when applied POST,

so use in crops is limited to PRE only at this time. Recent research has confirmed that foliar

applications of trifludimoxazin to PPO-inhibitor resistant waterhemp and Palmer amaranth does

not result in a dose-response shift compared to a susceptible, that is typical of fomesafen or

saflufenacil (Steppig 2022). An in vitro experiment conducted by Porri et al. (2022a) demonstrated

that trifludimoxazin in vitro had only a 10- to 100-fold difference in affinity between resistant and

susceptible PPO2 enzymes, whereas other PPO inhibitors tested had a 10,000+ fold difference in

affinity between susceptible and resistant PPO2 enzymes.

A similar phenomenon occurs with Photosystem II (PSII) inhibitors which all target the Qb

binding site of the D1 protein (Devine et al. 1992). Prior to separation into three herbicide site of

action groups, it was documented that triazine-resistant weed biotypes were also resistant to

triazinone and uracil herbicides, partially resistant to uracils and amides, and more susceptible to

nitrophenols, phenols, and bentazon (Pfister and Arntzen 1979). These classifications correspond

to separation into WSSA groups 5, 6, and 7. If trifludimoxazin binds differently to the target

17
enzymes than other PPO inhibitors, then a reclassification of PPO-inhibiting compounds may be

warranted.

It remains unclear how PRE applications of trifludimoxazin will influence emergence of

PPO-resistant weeds. Previous research by Wuerffel et al. (2015c) indicated that applications of

PPO-inhibiting herbicides PRE to a PPO-inhibitor resistant waterhemp population still provided

some control. However, the length of residual activity is compromised. The reason for partial

control is possibly that herbicide concentrations in the soil where seeds are germinating are

sufficient to control resistant biotypes, but as the herbicide dissipates through the soil profile, the

concentration decreases to a point where resistant seedlings can survive, but susceptible seedlings

cannot survive, which is known as a discriminating dose. Inclusion of an alternative mode of action

may partially reduce selection for resistant plants, but the alternative mode of action must be more

persistent or less soluble in order to be present in sufficient concentration for control. (Mansfield

2021, Westrich 2022, Wuerffel et al. 2015c). Despite this theory, Mansfield (2021) found that

increasing concentration ratio of S-metolachlor did not reduce selection pressure for PPO-resistant

waterhemp even though the S-metolachlor would have been present in sufficient quantity to

control PPO-resistant waterhemp that was the first to emerge. In the same study, saflufenacil

applications increased the frequency of homozygous resistant waterhemp, but not total number of

resistant waterhemp. The mechanism for this phenomenon was not discussed in the thesis. A

trifludimoxazin formulation currently under development is a premix with saflufenacil in a 2:1

saflufenacil to trifludimoxazin ratio. Herbicide premixes typically increase control spectrum and

length of residual in addition to being a resistance management tool (Beckie and Reboud 2009,

Kraehmer et al. 2014). The interaction of saflufenacil combinations with trifludimoxazin or other

PPO inhibitors and the resulting genotypes of emerging plants is unknown.

18
1.3 PPO-Inhibitor Resistance Mutations

Resistance to PPO-inhibiting herbicides was first identified in a Kansas waterhemp

population in 2001 (Shoup et al. 2003). Since confirmation of resistance in waterhemp, PPO-

resistance has been identified in 14 species (Heap 2023). Of those species, waterhemp, common

ragweed, and Palmer amaranth have been subjected to the most research. PPO-inhibitor resistance

in all three species is conferred by mutations in the PPX2 gene (Giacomini et al. 2017, Patzoldt et

al. 2006, Rousonelos et al. 2012, Salas et al. 2016).

Patzolt et al. (2006) elucidated that deletion of glycine at the 210th position conferred

resistance in waterhemp in a bacterial functional complementation assay. This mutation was an

unexpected finding as a deletion event that maintains the reading frame is relatively unusual. The

210th position is part of a short sequence repeat, so deletion of this amino acid is likely a result of

slippage of the DNA replication machinery (Dayan et al. 2010, Patzoldt et al. 2006). Protein-ligand

modeling suggests that deletion of glycine at the 210th position results in the loss of a hydrogen

bond and causes a partial uncoiling of the alpha-8 helix (Dayan et al. 2010). Partial uncoiling

results in a several fold increase in the size of the binding pocket which causes a loss of affinity

for PPO inhibitors without a loss of affinity for protogen.

Common ragweed resistance was also evolved with a mutation in PPX2, but via a R98L

mutation rather than a glycine deletion at position 210 (Rousonelos et al. 2012). The 98th position

was predicted to be a site of herbicide resistance, because the negative charge of the arginine

residue is critical in substrate binding in the target site (Heinemann et al. 2007, Koch et al. 2004).

Analysis of sequence homology of common ragweed and other species indicates that a ΔG210 or

analogous mutation is very unlikely in common ragweed, because it does not have the same short

sequence repeats that would allow a homologous deletion. Given that Palmer amaranth has a very

similar PPX2 sequence and the same intense selection pressures of waterhemp, researchers
19
predicted that it was only a matter of time before Palmer amaranth evolved the same resistance

mutation (Riggins and Tranel 2012). Indeed, Palmer amaranth was confirmed as resistant to PPO

inhibitors in Arkansas via the same ΔG210 mutation (Salas et al. 2016).

PPO-resistant waterhemp and Palmer amaranth are increasingly common (Copeland et al.

2018, Salas et al. 2016, Wuerffel et al. 2015b). However, it was observed that the ΔG210 mutation

did not explain a large number of apparent resistance cases that were being investigated by

diagnostic labs and Extension agents (Copeland et al. 2018, Giacomini et al. 2017, Salas-Perez et

al. 2017). Giacomini et al. (2017) confirmed two new resistance mutations, R128G and R128M,

at the 128th position (equivalent to the 98th position in common ragweed) of PPX2 in Palmer

amaranth. Sequence comparisons of waterhemp and Palmer amaranth indicated that the same R128

mutations found in Palmer amaranth are highly unlikely in waterhemp. The palmer amaranth wild

type codon is AGG whereas the waterhemp wild type codon is AGA. Both codons encode an

arginine, however, a single substitution in the Palmer amaranth sequence can yield ATG, encoding

methionine or GGG, encoding a glycine. While R128G is possible in waterhemp with a single

substitution (AGA to GGA), an R128M mutation or a, R128G mutation with the same codon

would require 2 nucleotide substitutions and is therefore highly unlikely. Interestingly, Nie et al.

(2019) sequenced PPX2 from waterhemp populations across the Midwest and found AGG, GGA,

GGG, AAA, and ATA codons conferring R128G, R128I, and R128K mutations. Only the R128G

and R128I were found to be resistant in a bacterial functional complementation assay. These

unlikely codon sequences as well as unique phenotypes from the respective populations strongly

suggests interspecies hybridization and gene flow with Palmer amaranth and tumble pigweed.

Given the multitude of possibilities for R128 mutations, determining exactly which mutations

confer resistance is of interest. Porri et al. (2022a) tested PPO inhibitors in vitro with all possible

20
substitutions at the 128th position of PPX2. Several mutations resulted in a non-functional protein,

and several mutations resulted in significant resistance. Most recently, a G399A mutation has been

discovered in Palmer amaranth (Rangani et al. 2019). Plant assays indicate that that it confers an

11 to 16-fold resistance to the herbicide fomesafen. Structural modeling indicates that this

resistance SNP causes stearic hindrance of PPO-inhibitor binding.

The ΔG210 mutation has been found to be the most abundant resistance mutation in

waterhemp across the Midwest (80 to 100%) (Nie et al. 2019) whereas Palmer amaranth has a

much lower incidence of the ΔG210 mutation and a higher frequency of other resistance

conferring mutations, such as R128 and G399 (approximately 50 to 70%) (Copeland et al. 2018,

Rangani et al. 2019, Salas-Perez et al. 2017). Reasons for differences in mutation frequency are

unclear. Perhaps there is an unknown fitness penalty in Palmer amaranth but not waterhemp,

leading to its lower frequency. It could also be a result of the near-simultaneous evolution of the

ΔG210 and R128 resistance mutations in Palmer amaranth. PPO-inhibitor resistance was

confirmed in waterhemp in 2001 but not until 2016 in Palmer amaranth, followed by confirmation

of R128 mutations in 2018. Some have suggested that particular PPO-inhibitor chemical families,

relative potencies, or application timings such as PRE or POST select for particular resistance

mutations (Wu et al. 2020). Non-target-site resistance also contributes to PPO-inhibitor resistance.

Varanasi et al. (2018) reported a Palmer amaranth population that is resistant to PPO inhibitors

and does not have any target-site mutations.

Until recently, all known mutations conferring resistance to PPO inhibitors were due to

mutations in the PPX2 gene. Goosegrass [Eleusine indica (L.) Gaertn.], conversely, has evolved

resistance to oxadiazon in turfgrass settings via an A212T mutation in the PPX1 gene (Bi et al.

2019). The fact that the mutation is in PPX1 is unusual and poses many questions due to the fact

21
that both goosegrass and oxadiazon are unique in terms of weed biology and herbicidal

characteristics. Of the few PPO inhibitors labeled for use in turfgrass, only oxadiazon is labeled

for control of grass weeds (McElroy and Martins 2013). Even broad-spectrum PPO inhibitors used

in major row crops do not control grass weeds PRE or POST particularly well (Loux et al. 2022).

Research is lacking on which factors control grass weed selectivity for PPO-inhibiting herbicides.

Oxadiazon is one of the few PPO inhibitors labeled for weed control in turf and is used at a massive

use rate of over 1 kg ha-1 (Anonymous, Ronstar Flo® Herbicide label). Oxadiazon also does not

have a row crop label, so its activity on Amaranthus spp. or common ragweed both resistant and

susceptible, is unknown. Oxadiazon appears to bind differently to PPO2 than other PPO inhibitors

(Rangani et al. 2019). Rangani et al. (2019) included oxadiazon in in vitro PPO2 assays. Oxadiazon

had a resistance factor of 100- to 1000-fold less than that of diphenyl ether herbicides for the

ΔG210 mutation and had a resistance factor of 1 for the R128L mutation. Thus, it remains

unknown if this PPX1 mutation is a result of goosegrass biology, oxadiazon/target-site interactions,

or a combination of the two factors.

1.4 Herbicide Resistance Genotyping Assays

Rapid identification of resistance is critical for effective weed management decisions

(Burgos et al. 2013). Traditional resistance assays are either discriminating dose or dose-response

assays (Burgos et al. 2013, Heap 2020). These assays, particularly full dose-response assays, can

provide robust evidence of resistance. The problem with these assays, however, is the length of

time required to conduct them. First, seeds from a suspected resistant plant must be collected. Then

the seeds must be grown and sprayed in a controlled environment. From the first instance of

suspected resistance to confirmation can take as much as 9 months to a year.

22
Molecular assays offer an alternative to whole plant assays, provided that the particular

resistance mechanism has been identified. Several genotyping assays have been developed that

detect a particular mutation in the DNA, such as single nucleotide polymorphisms (SNP) or

deletion/insertion mutations. A general method known as PCR-RFLP is the most broadly

applicable method (Burgos et al. 2013). A specific method known as derived cleaved amplified

polymorphic sequences (dCAPS) assay is a common resistance assay, and it has been developed

for all of the PPO-resistance mutations in Amaranthus species (Giacomini et al. 2017, Patzoldt et

al. 2006, Rangani et al. 2019). A dCAPS assay functions by allele-specific primers, cleaving a

DNA PCR product. The cleaved fragments are separated via gel electrophoresis. The presence of

particular bands indicates the presence or absence of a mutation. Shortcomings of dCAPS assays

are that they require gel electrophoresis which is relatively low throughput, gel bands can be

ambiguous, and the assay does not reliably distinguish between homozygous and heterozygous

mutants (Wuerffel et al. 2015a).

TaqMan® is a quantitative polymerase chain reaction (qPCR)-based assay that can reliably

distinguish between homozygous and heterozygous mutants. TaqMan® assays function through

binding of a fluorescent probe to a PCR-amplified product. The fluorescence of the probe is

quantified in a special machine hence the name qPCR. TaqMan® assays have been developed for

the ΔG210 mutations in waterhemp and Palmer amaranth. The major disadvantage for the qPCR

assay is access to a qPCR machine which can be costly (Kaundun et al. 2020). Most Recently, A

new PCR-RFLP method named derived Polymorphic Amplified cleaved sequence (DPACS) assay

has been developed for the ΔG210 mutation (Kaundun et al. 2020). Benefits of this system are that

it is broadly applicable to a wider range of species due to longer primer sequences, and it does not

require costly machinery. The same disadvantages of dCAPS are present in this method, so the

23
qPCR-based assay is the most preferred. Other methods for detecting mutations include

pyrosequencing, ELISA, next-generation sequencing, and many others (Barres et al. 2016). The

only other genotyping assay used to my knowledge is Illumina MISEQ Next generation

sequencing (Nie et al. 2019). This method was used to identify novel R128 mutations in waterhemp.

This method potentially could be used to assay resistance for diagnostic purposes if samples are

submitted frequently enough to fill a 96-well tray with samples, otherwise the single sample would

be cost-prohibitive.

Genotype assays are useful only insofar as the frequency of that particular genotype. In the

case of waterhemp and Palmer amaranth, PPO-inhibitor resistance can be caused by 2 or 3 different

mutations, respectively (Nie et al. 2019, Rangani et al. 2019). Therefore, to confirm a resistance

genotype, up to three assays must be run. In addition, a particular genotype is also not necessarily

a phenotype. The presence of a novel resistance mechanism, such as enhanced detoxification or

an uncharacterized polymorphism, renders each genotyping assay less useful. An assay that could

simultaneously detect multiple mutations would be beneficial to laboratories and diagnosticians

for more robust and efficient resistance confirmation. Also of interest would be a mobile assay that

does not require DNA extraction or PCR so that the assay can be more accessible to growers and

agronomists.

1.5 Herbicide Dose Responses

The gold standard for determining plant resistance to herbicides is the whole plant dose-

response assay (Burgos et al. 2013, Seefeldt et al. 1995). A typical herbicide dose-response assay

treats a plant or group of plants to a known dose of herbicide usually expressed as g ai ha-1. The

treatment structure should encompass the entire range of responses from no visible effect to

complete plant death. Data are typically analyzed using log logistic models to derive an ED50 value.

24
ED50 values generated can be lethal dose, but are usually growth reduction expressed as a

percentage of the untreated control. Ultimately, the statistic of interest is the R/S ratio, or the ratio

of ED50 of the resistant population compared to the susceptible population. One of the major

drawbacks of the whole plant dose-response assay is how many factors can influence not only the

ED values, but also the R/S ratio. Multiple research groups report a large variety of resistance

magnitudes to PPO inhibitors in waterhemp and Palmer amaranth (Lillie et al. 2020, Patzoldt et al.

2005, 2006, Rangani et al. 2019, Salas-Perez et al. 2017, Salas et al. 2016, Shoup et al. 2003,

Steppig 2022, Wuerffel et al. 2015b). Explanations for the range of ED50 values are primarily

greenhouse conditions at the time of application and plant size. PPO-inhibitor activity is influenced

heavily by temperature and light intensity, so time of year and sunshine on the day of application

can drastically influence PPO-inhibitor activity (Fausey and Renner 2001, Hatterman-Valenti et

al. 2011, Wichert et al. 1992).

Plant size is the other major influence on dose response (Lillie et al. 2020, Soltani et al.

2016). Plant size has a known influence on herbicide sensitivity because larger plants have greater

detoxification capabilities via CyP450 and GST as well as greater capabilities of absorbing ROS

(Dayan et al. 2019). With increasing plant size also comes an associated increase in leaf area index

(LAI), defined as the ratio of leaf area per unit of ground area covered. For contact herbicides such

as PPO inhibitors, plant coverage is a critical application parameter necessary for an effective

application (Franca et al. 2020). Herbicides applications are broadcast evenly across an area with

little regard for plant size parameters like leaf area index, nodes, or biomass. As a result, a larger

weed is getting a smaller relative dose than a smaller weed. Researchers in other plant protection

disciplines, such as viticulture pathology have experimented with variable dosing based on target

canopy characteristics (Siegfried et al. 2007). Siegfried et al. (2007) was able to develop a LAI

25
adapted dosage for grape fungicides that allows the application of fungicide at reduced rates, that

results in equal disease control as full labeled rates in season long spray programs.

1.6 Biologically Effective Dose

Biologically effective dose (BED) is generally defined as the minimum dose of a chemical

required to elicit a particular response. LD50 and GR50 are examples of biologically effective dose.

BED is the culmination of multiple processes from the molecular to the community level,

beginning with target enzyme kinetics. Enzyme inhibitors mimic an enzyme substrate or product

and block the normal catalysis reaction either transiently or permanently. Substrates and inhibitors

bind to the enzyme through a combination of hydrogen bonds, hydrophilic or hydrophobic

interactions, and charge-charge interactions. A small change in the target-site binding pocket can

have profound changes on BED. For example, a single hydrogen bond has a binding energy of 1

to 3 kcal mol-1, which equates to a 10-fold change in the kinetic parameter ki and a tenfold change

in BED (Fersht et al. 1985, Pace et al. 2014). Most cases of target-site herbicide resistance occur

through amino acid changes in the protein structure that disrupt the molecular interactions listed

above. (Dayan et al. 2010, 2014, Gaines et al. 2020, Patzoldt et al. 2006). Also influencing kinetic

parameters is the abundance of substrate and product interfering with inhibitor binding, and the

quantity of the target enzyme (Strelow et al. 2004). Glyphosate resistance in Palmer amaranth and

waterhemp are well-known examples of herbicide resistance through target-site abundance

(Gaines et al. 2010, Lorentz et al. 2014). To my knowledge, there has been no incidence of

herbicide resistance through altered substrate or product abundances.

Enzyme inhibitors must reach the target enzyme in a quantity necessary to be effective

which is the relevance of absorption, translocation, and detoxification processes in herbicide action.

Absorption and translocation of herbicides is a popular area of study and has profound impacts on

26
herbicide efficacy. The focus of many research groups is to improve absorption and translocation

with manipulation of droplet characteristics and the addition of adjuvants and tank mix partners

(Butts et al. 2018, Franca et al. 2020, Nandula et al. 2007). Other researchers have sought to

understand the environmental factors that influence absorption and translocation such as

temperature, humidity, and water potential, and how those factors interact with the composition of

plant cuticles and vasculature. (Hatterman-Valenti et al. 2011, Matzenbacher et al. 2014).

Herbicide detoxification is another critical aspect of herbicide action, particularly for

resistance and crop selectivity. The primary mechanisms of herbicide detoxification is through

CyP450 and glutathione-S-transferases (GST). These pathways are utilized by both crops and

weeds to survive a herbicide application. While there are several exceptions, detoxification is the

primary means of crop selectivity for herbicides. Detoxification is also a problematic means to

herbicide resistance in several weed species and is anticipated to be a primary concern for herbicide

resistance in the future (Gaines et al. 2020, Jugulam and Shyam 2019, Scarabel et al. 2015).

Herbicide interception is the last determination of BED. Larger leaves will intercept a

greater quantity of herbicide, and overlapping canopies from crop or high weed density will result

in less herbicide interception per plant. Several groups have quantified biologically effective doses

of herbicides and acknowledge that there is a shift with weed size, but no research to date has

related the shift in plant dose response with any parameter related to increasing size (Barker and

Dayan 2020, Hager et al. 2003, Soltani et al. 2016, Steckel et al. 1997). While the dose parameter

of g ha-1 is useful for describing a field scale application, it is inadequate for describing plant

physiology. A model that explains an increase in dose required to control weeds increasing in size

is justified so that future research can more accurately describe plant/herbicide physiology. Also,

emerging technologies will allow for more precise sensing of plant canopies and detecting a pest

27
species. Precision technologies have already been developed that enable a spray boom to spray

herbicide when a weed is detected and to be shut off in the absence of a weed as well as variable

rate application technologies (Hong et al. 2012, Ruixiu Sui et al. 2003, Wen et al. 2019). If a similar

technology were developed that could sense the size or ground cover of a weed, then an adjusted

dose could be applied both higher and lower than the regular labeled rate to enable the most

effective control while using a minimum amount of product.

1.7 Herbicide Effectiveness PRE vs POST

Weed populations resistant to a given herbicide site of action can still be partially managed

with PRE-applications of the respective herbicide group. (Boe 2019, Mansfield 2021, Westrich

2022, Wuerffel et al. 2015c) Length of residual activity is shortened and the frequency of resistant

plants is increased depending on species, herbicide active ingredient, and herbicide rate. It is

unknown if the partial control is a result of relative dose of the herbicide overcoming the resistance

mechanism, or if there is a difference in plant physiology that is altering the susceptibility of the

target plants. Seeds, seedlings, and established plants have very distinctive metabolic activities

compared to established plants (Silva et al. 2017). In addition, germinating seeds are not fully

developed, particularly in the case of plastids, which are where several major herbicide target

enzymes are located (Pyke 1999). Rather than chloroplasts, germinating seeds have proplastids

which develop into chloroplasts. While all shoot apical meristems contain proplastids,

comparatively little of a mature plant’s cells contain proplastids. Research is lacking if the different

metabolic states of emerged plants and seedlings causes a major metabolic shift that explains

partial control, or if relative dose is what is driving the partial control response of PRE-herbicides

on resistant populations.

28
1.8 Natural Selection in Weed Populations

Modern agricultural practices place immense selection pressure on herbicide resistance

evolution. While currently the ΔG210 mutation is the most common PPO-inhibitor resistance

mutation in waterhemp, Palmer amaranth populations are much more diverse in PPO-resistance

mutations (Copeland et al. 2018, Nie et al. 2019, Wuerffel et al. 2015a). It remains unclear if other

mutations are mere novelties or if one of them provides a sufficient fitness benefit that enables it

to overtake ΔG210 as the most widespread resistance mutation. Such a replacement is certainly

possible, especially in Amaranthus species.

Previously, triazine resistance in waterhemp was identified as being caused by a target-site

mutation in the pbsA gene coding for the D1 protein (Foes et al. 1998). More recently, a GST-

based detoxification was determined to be the most common mechanism of triazine resistance (Ma

et al. 2013, Patzoldt et al. 2003, Vennapusa et al. 2018). The S264G mutation associated with

triazine target-site resistance carries a known fitness penalty (Anderson et al. 1996). The spread of

this resistance mechanism is further limited by the fact that the pbsA gene is plastid-encoded, which

presumably limits the spread via only maternal inheritance (Murphy and Tranel 2019). While not

completely analogous to PPO-inhibitor resistance, the previous example demonstrates that a single

resistance mutation or mechanism is likely not the end of resistance evolution.

Some speculate that various resistance mutations are the product of particular selection

pressures, such as PRE vs POST applications or different chemical families (Wu et al. 2020). Wu

et al. (2020) sprayed PPO-inhibitor-resistant Palmer amaranth plants in a greenhouse with full

labeled rates of fomesafen and saflufenacil and conducted PPX2/inhibitor modeling. The

resistance genotypes were the three known PPX2 mutations (ΔG210, R128G, and G399A). Their

conclusion was that fomesafen selects for more diverse mutations because of weaker target-site

binding and that saflufenacil binding to the target site is more robust to the known resistance
29
mutations. Further experiments are necessary to solidify those conclusions, and the selection

pressure of PRE vs POST applications on different resistance genotypes is widely understudied.

Fitness costs are a nebulous subject in the weed science discipline (Keshtkar et al. 2019,

Vila-Aiub et al. 2009). Multiple studies have sought to identify fitness costs of herbicide resistance

traits with limited success in both findings and experimental execution (Vila-Aiub et al. 2009, Wu

et al. 2018). To date, no fitness penalty for PPO-inhibitor resistance has been identified at the

whole plant or community level (Murphy and Tranel 2019, Wu et al. 2018). However, a decrease

in PPO2 activity has been documented for enzymes with resistance mutations. Dayan et al. 2010

reported 10 fold decrease in kcat (turnover number) for PPO2 enzymes with the ΔG210 mutation

(Dayan et al. 2010). (Porri et al. 2022a) similarly reported less enzyme activity in vitro for resistant

enzymes with multiple substitutions in the R128 position. Other groups using in vitro assays are

seeing similar results (Porri et al. 2022b, Rangani et al. 2019).

The apparent lack of fitness cost may be a consequence of multiple isoforms (PPO1 and

PPO2) in shoot tissues (Watanabe et al. 2001). The resistant enzyme may have enough activity to

prevent the toxic accumulation of proto in the cytosol, but in the absence of PPO-inhibiting

compounds, the normal pathway needs are met by PPX1. PPX1 as the main metabolic enzyme

may be a reason that resistance to PPX1 has only been recently documented in goosegrass (Bi et

al. 2019). Any compromise in PPO1 metabolic efficiency may be heavily selected against in nature.

1.9 Organelle Targeting of Plant Enzymes

Subcellular localization of multiple forms of plant isoenzymes play a role in several

herbicide modes of action. For example, plants have two isoforms of glutamine synthetase, the

enzymatic target of the herbicide glufosinate (Castro-Rodríguez et al. 2011, Takano et al. 2020).

Glufosinate primarily targets the plastidial isoform (GS2) (Takano and Dayan 2020). GS2 is also

30
known to have a resistance mutation that confers glufosinate resistance in Lolium perenne L. ssp.

multiflorum (Avila-Garcia et al. 2012). The other isoform, GS1 is localized to the cytosol. the GS

isozymes have different metabolic roles, and their activities vary with plant development in

different organs and cell types (Castro-Rodríguez et al. 2011).

ACCase inhibitor selectivity is determined by the presence of an insensitive ACCase

enzyme in the chloroplast. All plants have ACCase enzymes in both the cytosol and the chloroplast.

The cytosolic enzyme for all plants is approximately 220 kd and is of eukaryotic origin. Grasses

have the 220 kd protein localized to both the plastid and the cytosol, whereas most dicots contain

a 35-kd isoform of ACCase of prokaryotic origin and localized to the plastid (Konishi et al. 1996).

ACCase inhibitors only inhibit the cytosol-localized isoform, and the plastid-localized isoform

remains functional for dicots. Grasses are susceptible to ACCase inhibitors because their single

eukaryotic version of the enzyme is localized to both the cytosol and chloroplasts (Konishi et al.

1996). Evolution of chloroplast targeting was likely necessary to maintain normal pathway

function (Konishi et al. 1996).

PPO inhibitors also have two isoenzymes; PPO1, which is targeted to the plastid, and PPO2

which is targeted to the mitochondria (Lermontova et al. 1997). PPO-inhibiting compounds inhibit

both PPO1 and PPO2 with relatively equal affinity, but PPO1 is understood as the primary target

(Dayan et al. 2018, Matringe et al. 1992). Despite PPO1 being the primary herbicide target, the

most widespread resistance mechanisms to PPO inhibitors are mutations in the mitochondrial

isoform of the gene, PPX2 (Nie et al. 2019, Patzoldt et al. 2006, Rangani et al. 2019). Therefore,

the PPO2 enzyme clearly plays a major role in the mechanism of action that has not been fully

investigated. Dual targeting of PPO2 to both the mitochondria and the plastid in resistant species

may contribute to the herbicide resistance mechanism. Dual targeting of PPO2 is known to occur

31
in spinach and Arabidopsis (Watanabe et al. 1998, Zybailov et al. 2008). Evidence for Arabidopsis

PPO2 dual targeting is from LCMS identification of the enzyme from chloroplast extracts

(Zybailov et al. 2008). In spinach, dual targeting occurs via multiple in-frame initiation codons

leading to synthesized proteins of different size being targeted to different organelles. The different

size proteins are due to a 30 amino acid extension to the 5’ sequence. If synthesis begins at the

first initiation codon, the long form of the gene is transcribed, and a 57 kd protein is synthesized

and targeted to the plastid. If transcription begins at the second initiation codon, a 55 kd protein is

synthesized and targeted to the mitochondria (Watanabe et al. 2001).

Several plant species exhibit multiple in-frame initiation codons separated by

approximately 30 aa in their PPX2 sequence including potato, maize, waterhemp, and Palmer

amaranth, so it seems likely that PPO2 is dual targeted to both the plastid and mitochondria in all

plants. Experimental evidence is needed to confirm that PPO2 is truly dual-targeted in all plants.

Common ragweed gene sequence has not been fully determined at the 5’ end, so it is unknown if

common ragweed PPO2 protein has multiple in-frame initiation codons (Rousonelos et al. 2012).

Proteins can be targeted in other ways, however. Some proteins rely on ambiguous signals that

export proteins different compartments with equal affinity (Kunze and Berger 2015). Dual

targeting to multiple plant organelles is more common than previously thought, with many genes

being targeted to both organelles in a sample (Dayan et al. 2014, Xu et al. 2013). The evolutionary

function of this phenomenon is not well understood.

1.10 PPO-Pathway Control and Flux

The fact that the PPX2 gene has evolved most resistance mutations thus far indicates that

the PPO2 enzyme is more involved in the PPO-inhibitor mechanism of action than previously

thought. Knowledge is lacking about the interaction of the two enzymes in situ, and whether

32
presence of resistant PPO enzyme in one organelle or the other is sufficient to confer resistance to

PPO inhibitors.

Most isoenzymes have some distinct tissue expression, subcellular localization, or

functional difference, such as the examples mentioned previously. However, plant PPOs seem to

be unique in that they are both expressed in the plastid and apparently perform the same function

(Watanabe et al. 2001). The normal PPO1 is present as well as the dual targeted mitochondrial

isoform which appears to be redundant in the plastid. It is plausible that PPO enzymes facilitate

the partitioning of tetrapyrrole precursors into their final end product (Sobotka et al. 2008). Proto

can go on to be catalyzed into heme via the enzymes ferrochelatase 1 (FC1) and ferrochelatase 2

(FC2), or chlorophyll via Mg chelatase (MGC) (Hey et al. 2016, Yaronskaya et al. 2002).

Ferrochelaase 1 and 2 are homodimers whereas Mg-chelatase is a heterotrimer encoded by 3 genes

(Axelsson et al. 2006, Scharfenberg er al. 2015).

Intuitively, MGC is localized to the plastid to produce chlorophyll, however the

relationship between FC1 and FC2 is more complex. FC2 is localized to the plastid. Heme

produced by FC2 is associated with hemoproteins related to photoprotection and photosynthetic

machinery. Whereas FC1 is associated with maintenance role hemoproteins (Espinas et al. 2016).

There is some disagreement in the literature regarding the localization of FC1, however the most

recent evidence suggests that it is localized to both the mitochondria and the plastid, but mostly

limited to the chloroplast (Chow et al. 1997, Dyer 2018, Lister et al. 2001, Masuda et al. 2003). It

is unknown to what degree flux is partitioned into each of the downstream enzymes and in what

organelles.

Jacobs and Jacobs (1995) are often cited in weed science literature where they

demonstrated that isolated chloroplasts export up to 50% of detectable protogen in the absence of

33
an inhibitor which suggests that a large portion of pathway flux is processed through the

mitochondria. This could be evidence of pathway partitioning. They suggested that protogen and

proto are transported out of the plastid and toward the mitochondria via a porphyrin carrier protein

similar to what has been identified in animals and humans (Hamza and Dailey 2012). However,

this was prior to the knowledge of dual targeting of PPO2 to both organelles which begs the

question of why the plant would export highly toxic compounds if the remainder of the pathway

is localized in the plastid, including downstream steps from PPO2.

A porphyrin transporter has been identified in Arabidopsis that supports the hypothesis that

large quantities of proto and protogen are exported from the chloroplast. AtABC1 encodes a

plastid-targeted ABC transporter (Møller et al. 2001). Arabidopsis mutants with a defect in the

gene ABC1 accumulate proto and have deficiencies in light signaling. Overexpression of this gene

in wild type Arabidopsis reduces proto accumulation after flumioxazin application (Møller et al.

2001). Homologous genes have not been well characterized in weed species. A homologous gene

in weed or crop species could contribute to PPO-inhibitor selectivity or resistance. ABC

transporters are known to contribute to the metabolism of xenobiotics. One such transporter has

been implicated in the glyphosate-resistance mechanism in horseweed (Ge et al. 2010, Tani et al.

2015). Plastid-targeted ABC transporters have also been patented by industry for conferring

herbicide resistance in crops.

PPO inhibitors cause buildup of protoporphyrinogen which leaks out and is converted to

protoporphyrin outside of the plastid and causes phytotoxicity (Jacobs and Jacobs 1993). Resistant

PPO enzymes prevent protoporphyrinogen buildup, but metal chelatase enzymes must process the

protoporphyrin to prevent too much from accumulating, or phytotoxicity would still occur (Kim

et al. 2014, Yaronskaya et al. 2002). Mg- and Fe-chelatases are the next step in chlorophyll and

34
heme biosynthesis. Overexpression of these enzymes are known to confer PPO resistance in crop

plants (Kim et al. 2014). Overexpression in weed species would possibly assist in prevention of

phytotoxicity by eliminating excess protoporphyrin and preventing production of ROS.

The compound 2,2’ bipyridine has been shown to be an iron chelator (Hendry and Stobart

1978). Treating barley plants with 2,2’ bipyridine has been shown to cause a decline in the cell’s

heme pool, increase flux through the porphyrin synthesis pathway, and increase synthesis of Mg-

chelated intermediates (Duggan and Gassman 1974). Such a compound may be useful in

combination with PPO inhibitors. PPO-inhibitor action depends on the accumulation of proto, so

an increase in pathway flux would theoretically render plants more susceptible to PPO inhibitors.

Its utility could be more useful in PPX2-resistant weeds. If localization to the mitochondria is an

important factor in PPO resistance, then there must be a diversion of the PPO pathway toward

heme synthesis because only the heme portion of the pathway occurs in the mitochondria. An

increase in heme synthesis as a result of PPO-inhibitor application coupled with blockage of iron

chelatase activity from iron chelation induced by 2,2-bipyridine may render PPO-inhibitor-

resistant biotypes susceptible.

Herbicide synergism as a method to overcome resistance has occurred before with target-

site triazine resistance. Co-application of HPPD-inhibiting herbicides such as mesotrione in

combination with atrazine results in synergistic effects. Even in triazine resistant biotypes, the

treated plants exhibit PSII inhibitor symptomology rather than characteristic bleaching that occurs

with HPPD-inhibiting herbicides (Walsh et al. 2012, Woodyard et al. 2009). The synergism occurs

because HPPD is an enzyme in the pathway for synthesizing plastoquinone (Ma et al. 2013). The

inhibition of 4 HPPD causes a depletion of plastoquinone which shifts the kinetic parameters of

atrazine binding to the Qb binding site (Woodyard et al. 2009). If the previously theorized

35
mechanism of PPO-pathway flow is accurate, then the combination of 2,2-bipyridine with PPO

inhibitors would be considered herbicide synergy and would be a viable innovation for the control

of resistant biotypes.

1.11 Summary

PPO inhibitors are among the most versatile and valuable chemical weed control tools in

global agriculture. Resistance to these compounds, especially in Amaranthus species threatens

their long-term utility. While the genetic basis for target-site resistance to PPO inhibitors is well

characterized, the variability of the resistance phenotype is noteworthy. Several likely contributing

factors have been identified in the weed science and plant physiology literature such as PPO-

inhibitor chemical family, variability in the PPO-pathway expression and flux, and interaction of

the two isoenzyme target sites. First, knowledge of how novel PPO inhibitors such as

trifludimoxazin affect control and selection for resistant biotypes will aid in the management of

problematic populations. Second, variability in how the PPO pathway is expressed and how much

pathway flux is utilized can vary by species, population, and plant life stage. Predicting and

manipulating pathway abundance and throughput can potentially assist with the control of resistant

populations by overwhelming or overcoming the resistance mechanism. Finally, PPO inhibitors

target two different isoenzymes. The presumed primary herbicide target is the chloroplast localized

isoform. However, this understanding is greatly complicated by the fact that most target-site

mutations conferring PPO-inhibitor resistance are in the gene for mitochondrial isoform, which

may be targeted to both the chloroplast and mitochondria. Knowledge of exactly which protein is

targeted by PPO-inhibiting herbicides and in which organelles can aid in the future development

of more specific inhibitor molecules.

36
1.12 Justification of Research

The use of PPO-inhibiting herbicides has increased as a response to glyphosate resistance

and will continue to increase with growing utilization of PRE and POST PPO-inhibiting herbicides.

Selection for novel PPO-inhibitor-resistant biotypes is occurring in waterhemp and Palmer

amaranth and threatens the long-term utility of PPO inhibitors. Given the recent proliferation of

novel resistant genotypes, a comparison of novel resistance biotypes with ΔG210 biotype is

justified to determine the cause of selection and to anticipate the loss of utility of PPO-inhibiting

herbicides as new biotypes spread. Novel PPO inhibitors, particularly trifludimoxazin, are

currently under development that are able to control PPO-inhibitor-resistant Amaranthus species.

We currently do not know how the use of these products, both PRE and POST, will influence the

selection for known and unknown resistance mechanisms. Understanding how resistance will

develop for new herbicide products is crucial for pesticide stewardship, product value, and benefit

to the end users. Finally, non-target-site resistance is a growing threat to the utility of PPO-

inhibiting herbicides. Previous research indicates that it may be present as a partial resistance

mechanism or an independent cause of resistance to PPO inhibitors in waterhemp. Investigations

into particular populations are justified to confirm the presence or absence of non-target-site

resistance mechanisms to PPO inhibitors in waterhemp. These investigations of PPO-inhibiting

herbicides and PPO-resistant biotypes will help to preserve the utility of PPO inhibitors, predict

future resistance developments, and contribute to broader knowledge about resistance evolution.

1.13 Literature Cited

Anonymous (2021) Ronstar® FLO herbicide product label. Cary, NC: Bayer Environmental
Science

37
Anderson DD, Higley LG, Martin AR, Roeth FW (1996) Competition between triazine-resistant
and -susceptible common waterhemp (Amaranthus rudis). Weed Sci 44:853–859

Armel GR, Hanzlik K, Witschel M, Hennigh DS, Bowe S, Simon A, Liebl R, Mankin L (2017)
Trifludimoxazin: a new PPO inhibitor that controls PPO resistant weed biotypes. In
Proceedings of the Weed Science Society of America Annual Meeting. Tuscon, AZ:
Weed Science Society of America
Avila-Garcia WV., Sanchez-Olguin E, Hulting AG, Mallory-Smith C (2012) Target-site mutation
associated with glufosinate resistance in Italian ryegrass (Lolium perenne L. ssp. multiflorum).
Pest Manag Sci 68:1248–1254

Axelsson E, Lundqvist J, Sawicki A, Nilsson S, Schröder I, Al-Karadaghi S, Willows RD, Hansson


M (2006) Recessiveness and dominance in barley mutants deficient in Mg-chelatase subunit
D, an AAA protein involved in chlorophyll biosynthesis. Plant Cell 18:3606–3616

Barker A, Dayan F (2020) The physiological basis of differential resistance to PPO-inhibiting


herbicides used pre- and post-emergent. in Proceedings of the Weed Science Society of
America Annual Meeting. Maui, HI: Weed Science Society of America

Barres B, Micoud A, Corio-Costet MF, Debieu D, Fillinger S, Walker AS (2016) Trends and
challenges in pesticide resistance detection. Trends Plant Sci 21:834-853

Beckie HJ, Reboud X (2009) Selecting for weed resistance: Herbicide rotation and mixture. Weed
Technol 23:363–370

Bi B, Wang Q, Coleman JJ, Porri A, Peppers JM, Patel JD, Betz M, Lerchl J, McElroy JS (2019)
A novel mutation A212T in chloroplast Protoporphyrinogen Oxidase (PPO1) confers
resistance to PPO inhibitor oxadiazon in Eleusine indica. Pest Manag Sci:76:1786–1794

Boe JE (2019) Establishing the value of ALS-inhibiting herbicides in fields with confirmed weed
resistance to ALS-inhibiting herbicides. MS thesis. West Lafayette IN: Purdue University.
190 p

Burgos NR, Tranel PJ, Streibig JC, Davis VM, Shaner D, Norsworthy JK, Ritz C (2013) Review:
confirmation of resistance to herbicides and evaluation of resistance levels. Weed Sci 61:4–
20

38
Butts TR, Samples CA, Franca LX, Dodds DM, Reynolds DB, Adams JW, Zollinger RK, Howatt
KA, Fritz BK, Clint Hoffmann W, Kruger GR (2018) Spray droplet size and carrier volume
effect on dicamba and glufosinate efficacy. Pest Manag Sci 74:2020–2029

Castro-Rodríguez V, García-Gutiérrez A, Canales J, Avila C, Kirby EG, Cánovas FM (2011) The


glutamine synthetase gene family in Populus. BMC Plant Biol 11:119

Chow KS, Singh DP, Roper JM, Smith AG (1997) A single precursor protein for ferrochelatase-I
from Arabidopsis is imported in vitro into both chloroplasts and mitochondria. J Biol Chem
272:27565–27571

Copeland JD, Giacomini DA, Tranel PJ, Montgomery GB, Steckel LE (2018) Distribution of
PPX2 mutations conferring PPO-inhibitor resistance in Palmer amaranth populations of
Tennessee. Weed Technol 32:592–596

Czarnecki O, Gläßer C, Chen JG, Mayer KFX, Grimm B (2012) Evidence for a contribution of
ALA synthesis to plastid-to-nucleus signaling. Front Plant Sci 3:236

Dayan FE (2019) Current status and future prospects in herbicide discovery. Plants (Basel,
Switzerland) 8:341

Dayan FE, Barker A, Dayan L, Ravet K (2019) The role of antioxidants in the protection of plants
against inhibitors of protoporphyrinogen oxidase. React Oxyg Species 7:55–63

Dayan FE, Barker A, Tranel PJ (2018) Origins and structure of chloroplastic and mitochondrial
plant protoporphyrinogen oxidases: implications for the evolution of herbicide resistance.
Pest Manag Sci 74:2226–2234

Dayan FE, Daga PR, Duke SO, Lee RM, Tranel PJ, Doerksen RJ (2010) Biochemical and
structural consequences of a glycine deletion in the α-8 helix of protoporphyrinogen oxidase.
Biochim Biophys Acta - Proteins Proteomics 1804:1548–1556

Dayan FE, Owens DK, Tranel PJ, Preston C, Duke SO (2014) Evolution of resistance to phytoene
desaturase and protoporphyrinogen oxidase inhibitors - state of knowledge. Pest Manag Sci
70:1358–1366

Devine M, Duke SO, Fedtke C, ed (1992) Physiology of herbicide action. Englewood Cliffs, New
Jersey, USA: PTR Prentice Hall 441 p

39
Duggan J, Gassman M (1974) Induction of porphyrin synthesis in etiolated bean leaves by
chelators of iron. Plant Physiol 53:206–215

Duke SO (2012) Why have no new herbicide modes of action appeared in recent years? Pest
Manag Sci 68:505–512

Dyer WE (2018) Exploiting weed seed dormancy and germination requirements through
agronomic practices 1. Weed Sci 43:498–503

Espinas NA, Kobayashi K, Sato Y, Mochizuki N, Takahashi K, Tanaka R, Masuda T (2016)


Allocation of heme is differentially regulated by ferrochelatase isoforms in arabidopsis cells.
Front Plant Sci 7:1326

Fausey JC, Renner KA (2001) Environmental effects on CGA-248757 and flumiclorac


efficacy/soybean tolerance. Weed Sci 49:668–674

Fersht AR, Shi JP, Knill-Jones J, Lowe DM, Wilkinson AJ, Blow DM, Brick P, Carter P, Waye
MMY, Winter G (1985) Hydrogen bonding and biological specificity analysed by protein
engineering. Nature 314:235–238

Foes MJ, Tranel PJ, Wax LM, Stoller EW (1998) A biotype of common waterhemp (Amaranthus
rudis) resistant to triazine and ALS herbicides. Weed Sci 46:514–520

Franca LX, Dodds DM, Butts TR, Kruger GR, Reynolds DB, Mills JA, Bond JA, Catchot AL,
Peterson DG (2020) Evaluation of optimal droplet size for control of Palmer amaranth
(Amaranthus palmeri) with acifluorfen. Weed Technol 34:511–519

Gaines TA, Duke SO, Morran S, G Rigon CA, Tranel PJ, Küpper A, Dayan FE (2020) Mechanisms
of evolved herbicide resistance J Biol Chem 295:10307–10330

Gaines TA, Zhang W, Wang D, Bukun B, Chisholm ST, Shaner DL, Nissen SJ, Patzoldt WL,
Tranel PJ, Culpepper AS, Grey TL, Webster TM, Vencill WK, Sammons RD, Jiang J, Preston
C, Leach JE, Westra P (2010) Gene amplification confers glyphosate resistance in
Amaranthus palmeri. Proc Natl Acad Sci U S A 107:1029–1034

Ge X, André d’Avignon D, Ackerman JJH, Douglas Sammons R (2010) Rapid vacuolar


sequestration: The horseweed glyphosate resistance mechanism. Pest Manag Sci 66:345–348

40
Giacomini DA, Umphres AM, Nie H, Mueller TC, Steckel LE, Young BG, Scott RC, Tranel PJ
(2017) Two new PPX2 mutations associated with resistance to PPO-inhibiting herbicides in
Amaranthus palmeri. Pest Manag Sci 73:1559–1563

Grossmann K, Niggeweg R, Christiansen N, Looser R, Ehrhardt T (2010) The herbicide


saflufenacil (kixorTM) is a new inhibitor of protoporphyrinogen IX oxidase activity. Weed Sci
58:1–9

Hager AG, Wax LM, Bollero G a, Stoller EW (2003) Influence of diphenylether herbicide
application rate and timing on common waterhemp (Amaranthus rudis) control in soybean
(Glycine max). Weed Technol 17:14–20

Hamza I, Dailey HA (2012) One ring to rule them all: Trafficking of heme and heme synthesis
intermediates in the metazoans Biochim Biophys Acta 1823:1617–1632

Hao GF, Zuo Y, Yang S-G, Yang G-F (2011) Protoporphyrinogen oxidase inhibitor: An ideal
target for herbicide discovery. Chim Int J Chem 65:961–969

Hatterman-Valenti H, Pitty A, Owen M (2011) Environmental effects on velvetleaf (Abutilon


theophrasti) epicuticular wax deposition and herbicide absorption. Weed Sci 59:14–21

Heap I (2020) Criteria for confirmation of herbicide-resistant weeds with specific emphasis on
confirming low level resistance

Heap I (2023) International survey of herbicide resistant weeds. http://www.weedscience.org.


Accessed March 25, 2023

Heinemann IU, Diekmann N, Masoumi A, Koch M, Messerschmidt A, Jahn M, Jahn D (2007)


Functional definition of the tobacco protoporphyrinogen IX oxidase substrate-binding site.
Biochem J 402:575–80

Hendry GAF, Stobart AK (1978) Effect of 2,2′-bipyridyl on porphyrin synthesis in etiolated and
light-treated barley leaves. Phytochemistry 17:671–674

Hey D, Ortega-Rodes P, Fan T, Schnurrer F, Brings L, Hedtke B, Grimm B (2016) Transgenic


tobacco lines expressing sense or antisense FERROCHELATASE 1 RNA show modified
ferrochelatase activity in roots and provide experimental evidence for dual localization of
ferrochelatase 1. Plant Cell Physiol 57:2576–2585

41
Hong S, Minzan L, Zhang Q (2012) Detection system of smart sprayers: Status, challenges, and
perspectives. Int J Agric Biol Eng 5:10–23

Jacobs JM, Jacobs NJ (1993) Porphyrin accumulation and export by isolated barley (Hordeum
vulgare) plastids: Effect of diphenyl ether herbicides. Plant Physiol 101:1181–1188

Jacobs JM, Jacobs NJ (1995) Terminal enzymes of heme biosynthesis in the plant plasma
membrane. Arch Biochem Biophys 323:274–278

Jugulam M, Shyam C (2019) Non-target-site resistance to herbicides: recent developments. Plants


8:417

Kaundun SS, Hutchings S, Marchegiani E, Rauser R, Jackson L V (2020) A derived polymorphic


amplified cleaved sequence assay for detecting the Δ210 PPX2L codon deletion conferring
target‐site resistance to protoporphyrinogen oxidase‐inhibiting herbicides. Pest Manag Sci
76:789–796

Keshtkar E, Abdolshahi R, Sasanfar H, Zand E, Beffa R, Dayan FE, Kudsk P (2019) Assessing
fitness costs from a herbicide-resistance management perspective: A review and insight.
Weed Sci 67:137–148

Kim JG, Back K, Lee HY, Lee HJ, Phung TH, Grimm B, Jung S (2014) Increased expression of
Fe-chelatase leads to increased metabolic flux into heme and confers protection against
photodynamically induced oxidative stress. Plant Mol Biol 86:271–287

Koch M, Breithaupt C, Kiefersauer R, Freigang J, Huber R, Messerschmidt A (2004) Crystal


structure of protoporphyrinogen IX oxidase: a key enzyme in haem and chlorophyll
biosynthesis. EMBO J 23:1720–1728

Konishi T, Shinohara K, Yamada K, Sasaki Y (1996) Acetyl-CoA carboxylase in higher plants:


Most plants other than gramineae have both the prokaryotic and the eukaryotic forms of this
enzyme. Plant Cell Physiol 37:117–122

Kraehmer H, Van Almsick A, Beffa R, Dietrich H, Eckes P, Hacker E, Hain R, John Strek H,
Stuebler H, Willms L (2014) Herbicides as weed control agents: State of the Art: II. Recent
achievements. Plant Physiol 166:1132–1148

42
Kunze M, Berger J (2015) The similarity between N-terminal targeting signals for protein import
into different organelles and its evolutionary relevance. Front Physiol 6:259

Lee HJ, Duke M V., Duke SO (1993) Cellular localization of protoporphyrinogen-oxidizing


activities of etiolated barley (Hordeum vulgare L.) leaves: Relationship to mechanism of
action of protoporphyrinogen oxidase-inhibiting herbicides. Plant Physiol 102:881–889

Lermontova I, Kruse E, Mock HP, Grimm B (1997) Cloning and characterization of a plastidal
and a mitochondrial isoform of tobacco protoporphyrinogen IX oxidase. Proc Natl Acad Sci
U S A 94:8895–8900

Lillie KJ, Giacomini DA, Tranel PJ (2020) Comparing responses of sensitive and resistant
populations of Palmer amaranth (Amaranthus palmeri) and waterhemp (Amaranthus
tuberculatus var. rudis) to PPO inhibitors. Weed Technol. 34: 140–146. doi:
10.1017/wet.2019.84

Lister R, Chew O, Rudhe C, Lee MN, Whelan J (2001) Arabidopsis thaliana ferrochelatase-I and
-II are not imported into Arabidopsis mitochondria. FEBS Lett 506:291–295

Lorentz L, Gaines TA, Nissen SJ, Westra P, Strek HJ, Dehne HW, Ruiz-Santaella JP, Beffa R
(2014) Characterization of glyphosate resistance in Amaranthus tuberculatus populations. J
Agric Food Chem 62:8134–8142

Loux MM, Doohan D, Dobbels AF, Johnson WG, Young BG, Legleiter TR, Hager A (2022) Weed
Control Guide for Ohio, Indiana and Illinois

Ma R, Kaundun SS, Tranel PJ, Riggins CW, McGinness DL, Hager AG, Hawkes T, Mc EI,
Riechers DE (2013) Distinct detoxification mechanisms confer resistance to mesotrione and
atrazine in a population of waterhemp. Plant Physiol 163:363–377

Maneli MH, Corrigall A V, Klump HH, Davids LM, Kirsch RE, Meissner PN (2003) Kinetic and
physical characterisation of recombinant wild-type and mutant human protoporphyrinogen
oxidases. Biochim Biophys Acta 1650:10–21

Mansfield BC (2021) Characterization of protoporphyrinogen oxidase (PPO) herbicide resistance


in tall waterhemp (Amaranthus tuberculatus). MS thesis. West Lafayette IN: Purdue
University. 158 p

43
Masuda T, Suzuki T, Shimada H, Ohta H, Takamiya K (2003) Subcellular localization of two
types of ferrochelatase in cucumber. Planta 217:602–609

Matringe M, Camadro JM, Labbe P, Scalla R (1989) Protoporphyrinogen oxidase as a molecular


target for diphenyl ether herbicides. Biochem J 260:231–235

Matringe M, Camadroll JM, Block MA, Joyards J, Scallall R, Labbe P, Douces R (1992)
Localization within chloroplasts of protoporphyrinogen oxidase, the target enzyme for
diphenylether-like herbicides J Biol Chem 267:4646–4651

Matzenbacher FO, Vidal RA, Merotto Jr. A, Trezzi MM (2014) Environmental and physiological
factors that affect the efficacy of herbicides that inhibit the enzyme protoporphyrinogen
oxidase: a literature review. Planta Daninha 32:457–463

McElroy JS, Martins D (2013) Use of herbicides on turfgrass. Planta Daninha 31:455–467

Meazza G, Bettarini F, La Porta P, Piccardi P, Signorini E, Portoso D, Fornara L (2004) Synthesis


and herbicidal activity of novel heterocyclic protoporphyrinogen oxidase inhibitors. Pest
Manag Sci 60:1178–1188

Møller SG, Kunkel T, Chua NH (2001) A plastidic ABC protein involved in intercompartmental
communication of light signaling. Genes Dev 15:90–103

Murphy BP, Tranel PJ (2019) Target-site mutations conferring herbicide resistance. Plants 8:382

Nandula VK, Poston DH, Reddy KN, Koger CH (2007) Formulation and adjuvant effects on
uptake and translocation of clethodim in bermudagrass (Cynodon dactylon) Weed Sci 55:6–
11

Nie H, Mansfield BC, Harre NT, Young JM, Steppig NR, Young BG (2019) Investigating target‐
site resistance mechanism to the PPO‐inhibiting herbicide fomesafen in waterhemp and
interspecific hybridization of Amaranthus species using next generation sequencing. Pest
Manag Sci 75:3235–3244

Pace CN, Fu H, Fryar KL, Landua J, Trevino SR, Schell D, Thurlkill RL, Imura S, Scholtz JM,
Gajiwala K, Sevcik J, Urbanikova L, Myers JK, Takano K, Hebert EJ, Shirley BA, Grimsley
GR (2014) Contribution of hydrogen bonds to protein stability. Protein Sci 23:652–661

44
Park J, Ahn YO, Nam J-W, Hong M-K, Song N, Kim T, Yu G-H, Sung S-K (2018) Biochemical
and physiological mode of action of tiafenacil, a new protoporphyrinogen IX oxidase-
inhibiting herbicide. Pestic Biochem Physiol 152:38–44

Patzoldt WL, Dixon BS, Tranel PJ (2003) Triazine resistance in Amaranthus tuberculatus (Moq)
Sauer that is not site-of-action mediated. Pest Manag Sci 59:1134–1142

Patzoldt WL, Hager AG, McCormick JS, Tranel PJ (2006) A codon deletion confers resistance to
herbicides inhibiting protoporphyrinogen oxidase. Proc Natl Acad Sci U S A 103:12329–
12334

Patzoldt WL, Tranel PJ, Hager AG (2005) A waterhemp (Amaranthus tuberculatus) biotype with
multiple resistance across three herbicide sites of action. Weed Sci 53:30–36

Pfister K, Arntzen CJ (1979) The mode of action of photosystem II-specific inhibitors in herbicide-
resistant weed biotypes Z Naturforsch 34:996–1009

Porri A, Betz M, Seebruck K, Knapp M, Johnen P, Witschel M, Aponte R, Liebl R, Tranel PJ,
Lerchl J (2022a) Inhibition profile of trifludimoxazin towards PPO2 target site mutations.
Pest Manag Sci 79:507–519

Porri A, Noguera MM, Betz M, Sälinger D, Brändle F, Bowe SJ, Lerchl J, Meyer L, Knapp M,
Roma-Burgos N (2022b) Can double PPO mutations exist in the same allele and are such
mutants functional? Pest Manag Sci 78:2258–2264

Pyke KA (1999) Plastid division and development. Plant Cell 11:549–556

Ramanujam VMS, Anderson KE (2015) Porphyria diagnostics-part 1: A brief overview of the


porphyrias. Curr Protoc Hum Genet:17.20.1–17.20.26

Rangani G, Salas-Perez RA, Aponte RA, Knapp M, Craig IR, Mietzner T, Langaro AC, Noguera
MM, Porri A, Roma-Burgos N (2019) A novel single-site mutation in the catalytic domain of
Protoporphyrinogen Oxidase IX (PPO) confers resistance to PPO-inhibiting herbicides. Front
Plant Sci 10:568

Riggins CW, Tranel PJ (2012) Will the Amaranthus tuberculatus resistance mechanism to PPO-
inhibiting herbicides evolve in other Amaranthus species? Int J Agron 2012:305764

45
Rousonelos SL, Lee RM, Moreira MS, VanGessel MJ, Tranel PJ (2012) Characterization of a
common ragweed (Ambrosia artemisiifolia) population resistant to ALS- and PPO-inhibiting
herbicides. Weed Sci 60:335–344

Ruixiu Sui, J. Alex Thomasson, Jeffrey L. Willers, F. Paul Lee, Rui Wang (2003) Variable-rate
spray system dynamic evaluation. in Proceedings of the 2003 ASAE Annual Meeting. Las
Vegas, NV: American Society of Agricultural and Biological Engineers

Salas-Perez RA, Burgos NR, Rangani G, Singh S, Paulo Refatti J, Piveta L, Tranel PJ,
Mauromoustakos A, Scott RC (2017) Frequency of gly-210 deletion mutation among
protoporphyrinogen oxidase inhibitor–resistant Palmer amaranth (Amaranthus palmeri)
populations. Weed Sci 65:718–731

Salas RA, Burgos NR, Tranel PJ, Singh S, Glasgow L, Scott RC, Nichols RL (2016) Resistance
to PPO-inhibiting herbicide in Palmer amaranth from Arkansas. Pest Manag Sci 72:864–869

Scarabel L, Pernin F, Délye C (2015) Occurrence, genetic control and evolution of non-target-site
based resistance to herbicides inhibiting acetolactate synthase (ALS) in the dicot weed
Papaver rhoeas. Plant Sci 238:158–169

Scharfenberg M, Mittermayr L, Von Roepenack-Lahaye E, Schlicke H, Grimm B, Leister D,


Kleine T (2015) Functional characterization of the two ferrochelatases in Arabidopsis
thaliana. Plant Cell Environ 38:280–298

Seefeldt SS, Jensen JE, Patrick Fuerst E (1995) Log-logistic analysis of herbicide dose-response
relationships. Weed Technol 9:218–227

Shoup DE, Al-Khatib K, Peterson DE (2003) Common waterhemp (Amaranthus rudis) resistance
to protoporphyrinogen oxidase-inhibiting herbicides. Weed Sci 51:145–150

Siegfried W, Viret O, Huber B, Wohlhauser R (2007) Dosage of plant protection products adapted
to leaf area index in viticulture. Crop Prot 26:73–82

Silva AT, Ligterink W, Hilhorst HWM (2017) Metabolite profiling and associated gene expression
reveal two metabolic shifts during the seed-to-seedling transition in Arabidopsis thaliana.
Plant Mol Biol 95:481–496

46
Sobotka R, McLean S, Zuberova M, Hunter CN, Tichy M (2008) The C-terminal extension of
ferrochelatase is critical for enzyme activity and for functioning of the tetrapyrrole pathway
in Synechocystis strain PCC 6803. J Bacteriol 190:2086–2095

Soltani N, Nurse RE, Sikkema PH (2016) Biologically effective dose of glyphosate as influenced
by weed size in corn. Can J Plant Sci 96:455–460

Steckel GJ, Wax LM, Simmons FW, Phillips II WH (1997) Glufosinate efficacy on annual weeds
is influenced by rate and growth stage. Weed Technol 11:484–488

Steppig NR (2022) Evaluation of trifludimoxazin, a new protoporphyrinogen oxidase- inhibiting


herbicide, for use in soybean. PhD dissertation West Lafayette, IN: Purdue University. 172p

Strelow J, Dewe W, Iversen PW, Brooks HB, Radding JA, McGee J, Weidner J (2004) Mechanism
of action assays for enzymes. Assay Guidance Manual. Eli Lilly & Company and the National
Center for Advancing Translational Sciences

Takano HK, Beffa R, Preston C, Westra P, Dayan FE (2020) A novel insight into the mode of
action of glufosinate: how reactive oxygen species are formed. Photosynth Res 144:361–372

Takano HK, Dayan FE (2020) Glufosinate‐ammonium: a review of the current state of knowledge.
Pest Manag Sci 76:3911–3925

Tani E, Chachalis D, Travlos IS (2015) A glyphosate resistance mechanism in Conyza canadensis


involves synchronization of EPSPS and ABC-transporter genes. Plant Mol Biol Report
33:1721–1730

Varanasi VK, Brabham C, Norsworthy JK (2018) Confirmation and characterization of non-target


site resistance to fomesafen in Palmer amaranth (Amaranthus palmeri). Weed Sci 66:702–
709

Vennapusa AR, Faleco F, Vieira B, Samuelson S, Kruger GR, Werle R, Jugulam M (2018)
Prevalence and mechanism of atrazine resistance in waterhemp (Amaranthus tuberculatus)
from Nebraska. Weed Sci 66:595–602

Vila-Aiub MM, Neve P, Powles SB (2009) Fitness costs associated with evolved herbicide
resistance alleles in plants. New Phytol 184:751–767

47
Walsh MJ, Stratford K, Stone K, Powles SB (2012) Synergistic effects of atrazine and mesotrione
on susceptible and resistant wild radish (Raphanus raphanistrum) populations and the
potential for overcoming resistance to triazine herbicides. Weed Technol 26:341–347

Watanabe N, Che FS, Iwano M, Takayama S, Nakano T, Yoshida S, Isogai A (1998) Molecular
characterization of photomixotrophic tobacco cells resistant to protoporphyrinogen oxidase-
inhibiting herbicides 1

Watanabe N, Che FS, Iwano M, Takayama S, Yoshida S, Isogai A (2001) Dual targeting of spinach
protoporphyrinogen oxidase II to mitochondria and chloroplasts by alternative use of two in-
frame initiation codons. J Biol Chem 276:20474–20481

Wen S, Zhang Q, Yin X, Lan Y, Zhang J, Ge Y (2019) Design of plant protection UAV variable
spray system based on neural networks. Sensors 19:1112

Westrich BC (2022) Influence of mesotrione, ALS-inhibitor resistance, and self incompatability


on giant ragweed management in soybean. PhD dissertation West Lafayette, IN: Purdue
University. 189 p

Wichert RA, Bozsa R, Talbert RE, Oliver LR (1992) Temperature and relative humidity effects
on diphenylether herbicides. Weed Technol 6:19–24

Witschel M, Aponte R, Armel G, Bowerman P, Mietzner T, Newton T, Porri A, Simon A, Seitz T


(2021) Tirexor®—design of a new resistance breaking protoporphyrinogen IX oxidase
inhibitor. Recent Highlights Discov Optim Crop Prot Prod:501–509

Woodyard AJ, Hugie JA, Riechers DE (2009) Interactions of mesotrione and atrazine in two weed
species with different mechanisms for atrazine resistance. Weed Sci 57:369–378

Wu C, Davis AS, Tranel PJ (2018) Limited fitness costs of herbicide-resistance traits in


Amaranthus tuberculatus facilitate resistance evolution. Pest Manag Sci 74:293–301

Wu C, Goldsmith MR, Pawlak J, Feng P, Smith S, Navarro S, Perez-Jones A (2020) Differences


in efficacy, resistance mechanism and target protein interaction between two PPO inhibitors
in Palmer amaranth (Amaranthus palmeri). Weed Sci 68:105–115

48
Wuerffel RJ, Young JM, Lee RM, Tranel PJ, Lightfoot DA, Young BG (2015a) Distribution of
the ΔG210 protoporphyrinogen oxidase mutation in Illinois waterhemp (Amaranthus
tuberculatus) and an improved molecular method for detection. Weed Sci 63:839–845

Wuerffel RJ, Young JM, Matthews JL, Young BG (2015b) Characterization of PPO-inhibitor–
resistant waterhemp (Amaranthus tuberculatus) response to soil-applied PPO-inhibiting
herbicides. Weed Sci 63:511–521

Wuerffel RJ, Young JM, Tranel PJ, Young BG (2015c) Soil-residual protoporphyrinogen oxidase–
inhibiting herbicides influence the frequency of associated resistance in waterhemp
(Amaranthus tuberculatus). Weed Sci 63:529–538

Xu L, Carrie C, Law SR, Murcha MW, Whelan J (2013) Acquisition, conservation, and loss of
dual-targeted proteins in land plants. Plant Physiol 161:644–662

Yaronskaya EB, Rassadina V V., Averina NG (2002) Regulation of magnesium chelatase activity
during excessive accumulation of porphyrins in green barley leaves. Russ J Plant Physiol
49:771–776
Young BG (2006) Changes in herbicide use patterns and production practices resulting from
glyphosate-resistant crops. Weed Technol 20:301–307
Zybailov B, Rutschow H, Friso G, Rudella A, Emanuelsson O, Sun Q, Van Wijk KJ (2008) Sorting
signals, N-terminal modifications and abundance of the chloroplast proteome. PLoS One
3:e1994

49
Diphenyl ether N-Phenyl-Triazolinone

Fomesafen Lactofen Acifluorfen Sulfentrazone Carfentrazone

N-Phenyl-imide N-Phenyl-
50

oxadiazolinone

Flumioxazin Trifludimoxazin Saflufenacil Butafenacil Oxadiazon

Figure 1.1. Names and structure of PPO-inhibiting herbicide molecules commonly mentioned in the literature.
Figure 1.2. Tetrapyrrole synthesis pathway as presented in (Czarnecki et al. 2012) Abbreviations:
ALAD, ALA dehydratase; CAO, Chl a oxygenase; CBR, chlorophyll b reductase; ChlS,
chlorophyll synthase; CPO, coproporphyrinogen III oxidase; DVR, divinyl protochlorophyllide
reductase; FeCh, Fe chelatase; FLU, flourescent; GluRS, glutamyl-tRNA synthetase; GluTR,
glutamyl-tRNA reductase; GluTRBP, GluTR binding protein; GSAT, glutamate-1-semialdehyde
aminotransferase; HBS, hydroxymethylbilane synthase; HCAR, 7-hydroxymethyl chlorophyll a
reductase; HO, heme oxygenase; MgCh, Mg chelatase; MTF, Mg protoporphyrin IX
methyltransferase; PBS, phytochromobilin synthase; POR, light dependent NADPH-
protochlorophyllide oxidoreductase; PPOX, protoporphyrinogen IX oxidase; UROD,
uroporphyrinogen III decarboxylase; UROM, uroporphyrinogen III methyltransferase; UROS,
uroporphyrinogen III synthase. Colors represent different organelle localizations.

51
ARE PRE-APPLIED PPO INHIBITORS SELECTING
FOR NEW RESISTANCE MUTATIONS IN WATERHEMP
(AMARANTHUS TUBERCULATUS)?

Short title: ΔG210 vs R128G

2.1 Abstract

The PPO inhibitors are a valuable group of herbicides that provide soil and foliar control of

glyphosate-resistant Amaranthus species. The ΔG210 mutation in the waterhemp PPX2 gene

confers PPO-inhibitor resistance and has been common in the Midwest for more than a decade.

While most PPO-inhibitor resistance in waterhemp is attributable to the ΔG210 mutation, R128

mutations have recently been identified in several populations of waterhemp throughout the

Midwest. Given the already widespread distribution and the broad cross resistance of the ΔG210

mutation, it is unclear why R128 mutations would evolve so much later in waterhemp. We

hypothesized that either the soil-applied PPO inhibitor use pattern or specific active ingredients

that are applied primarily preemergence are selecting for novel mutations conferring resistance to

PPO inhibitors. We conducted a series of dose-response experiments with flumioxazin,

sulfentrazone, fomesafen, oxadiazon, saflufenacil and trifludimoxazin applied preemergence to

greenhouse flats containing waterhemp seeds homozygous for ΔG210, R128G, and susceptible

PPX2 alleles. Both the R128G and ΔG210 mutations exhibited resistance for all herbicides tested

including trifludimoxazin. The greatest resistance ratio was for fomesafen (20- to 37-fold)

followed by saflufenacil (10 to 15-fold). All other herbicides had resistance ratios less than 9 for

both mutations. Sulfentrazone was the only herbicide for which the R128G mutation conferred

greater resistance (R/S = 8.8) than did the ΔG210 mutation (R/S = 4.2). The R128G mutation in

52
waterhemp was not more robust than the ΔG210 mutation with respect to conferring resistance to

PPO inhibitors applied preemergence; and this research does not support a hypothesis that the

R128G mutation was selected using the soil-applied PPO inhibitors. Furthermore, there is no

evidence that the commercial efficacy of PPO inhibitors applied preemergence will diminish any

further as a result of the R128G mutation increasing in frequency, relative to the what has been

realized with the ΔG210 mutation.

Nomenclature: Waterhemp, Amaranthus tuberculatus (Moq.) J.D.Sauer, flumioxazin, fomesafen,

oxadiazon, saflufenacil, sulfentrazone, trifludimoxazin

Key words: PPO-inhibitor resistance, target-site resistance, ΔG210 mutation, R128G mutation,

resistance selection

2.2 Introduction

Protoporphyrinogen IX oxidase (PPO)-inhibiting herbicides are an important class of

herbicides used worldwide that provide preemergence (PRE) and postemergence (POST) weed

control in a variety of crops (Hao et al. 2011). Despite the fact that they have been used since the

1960s, resistance to PPO inhibitors is relatively uncommon in terms of the number of resistant

species (Heap 2023). The first documented incidence of resistance was not until 2001, in

waterhemp [Amaranthus tuberculatus (Moq.) J.D.Sauer] (Heap 2023, Shoup et al. 2003) followed

shortly thereafter by common ragweed (Ambrosia artemisiifolia L.) (Rousonelos et al. 2012). In

the past 10 years, there have been 9 new species documented as resistant to PPO inhibitors and a

total of 13 resistant species worldwide (Heap 2023). The most prolific and problematic resistant

species in the United States have been waterhemp and Palmer amaranth (Amaranthus palmeri S.

Wats) (Van Wychen 2022). Increased incidence of PPO-inhibitor resistance in these species is
53
primarily due to the increased use of PPO inhibitors in crops as a result of widespread glyphosate

resistance (Green and Owen 2010, Legleiter et al. 2009, Young 2006).

Plants have two nuclear genes (PPX1 and PPX2) that encode for two distinct PPO enzymes

(PPO1 and PPO2) (Lermontova et al. 1997). PPO1 is localized to the chloroplast while PPO2 is

dual targeted to both the plastid and the mitochondria (Watanabe et al. 2001). Prior to the evolution

of resistance, there has been a widespread assumption that PPO1 is the primary target of PPO

inhibitors (Dayan et al. 2018, Lee et al. 1993). However, mutations in the PPX2 gene actually

confer the vast majority of PPO-inhibitor resistance in several species (Giacomini et al. 2017,

Mendes et al. 2020, Patzoldt et al. 2006, Rousonelos et al. 2012, Salas et al. 2016). Table 2.1

describes the PPX2 resistance mutations in waterhemp and Palmer amaranth that have been

discovered to date. Patzoldt et al. (2006) determined that a glycine deletion at the 210th position

(ΔG210) conferred resistance in waterhemp. Different models suggest that deletion of glycine at

the 210th position results in altered geometry of the binding pocket resulting in reduced inhibitor

binding affinity (Dayan et al. 2010, Noguera et al. 2021). Palmer amaranth was confirmed as

resistant to PPO inhibitors in Arkansas via the same ΔG210 mutation (Salas et al. 2016). Shortly

thereafter, Giacomini et al. (2017) confirmed two new resistance mutations, R128G and R128M,

at the 128th position of PPO2 in Palmer amaranth (equivalent to R98 in common ragweed).

Interestingly, Nie et al. (2019) sequenced PPX2 from waterhemp populations across the

Midwest and found multiple codons, previously unknown to be present in waterhemp, coding for

R128G, R128I, and R128K. Only the R128G and R128I were found to be resistant in bacterial

functional complementation assays. These unlikely codon sequences indicate that waterhemp

PPX2 wild type alleles are more diverse than previously observed. Most recently a G399A

54
mutation was found in an Arkansas Palmer amaranth population, indicating that there is still active

selection for new PPO-inhibitor resistant biotypes (Rangani et al. 2019).

Most instances of herbicide resistance in Amaranthus species are conferred by single

mutations or at least single mechanisms that are widespread across the United States (Murphy and

Tranel 2019). There are several known instances of multiple mutations or mechanisms of

resistance to the same herbicide being present in the same species (Bell et al. 2013, Chatham et al.

2015, Patzoldt and Tranel 2007, Schultz et al. 2015, Vennapusa et al. 2018). In all cases, there is

a known fitness benefit of one of the two existing resistance mutations in either the presence or

absence of herbicide selection (Cahoon et al. 2022, Patzoldt and Tranel 2007, Vila-Aiub et al. 2009,

Wu et al. 2018).

Why new mutations are being selected despite the relative abundance of the ΔG210

mutation has not been elucidated. One possible explanation is a fitness penalty conferred by the

ΔG210 mutation. A fitness penalty is the loss of competitiveness of the resistant biotype in the

absence of herbicide selection, leading to eventual decline of allele frequency in the population.

To date, there is no evidence of a fitness penalty conferred by the ΔG210 mutation at the whole

organism level (Wu et al. 2018).

Potentially, these new PPX2 mutations are accumulating in the same plants and having

additive effects. In a survey of Palmer amaranth for PPO-inhibitor resistance, the most highly

resistant populations possessed two or even three resistance alleles in the same population

(Noguera et al. 2021). Other modeling and in vitro experiments indicate that PPX2 double mutants

are more resistant than single mutants (Porri et al. 2022a). However, double mutants appear to

have a severe fitness penalty compared to single mutants (Porri et al. 2022b). For example,

Noguera et al. (2021) only observed 32 plants, out of thousands that were genotyped, with 2

55
resistance mutations in the same allele. Porri et al. (2022b) found that ΔG210 and R128G double

mutant alleles can only exist as a heterozygote because of a severe loss of enzyme activity.

Therefore, with the present set of known resistance mutations, mutation accumulation is not greatly

contributing to selection of novel resistance mutations.

A potential explanation for why new resistance mutations are being selected is the change

in PPO-inhibitor use patterns. Prior to the commercialization of glyphosate-resistant soybeans in

1996, POST PPO inhibitors such as lactofen, acifluorfen, and fomesafen were applied to

approximately 25% of soybean hectares (USDA NASS 1994). As growers increasingly adopted

multiple glyphosate applications as their primary weed control practice, glyphosate resistance

began to emerge in key weed species and spread throughout the entire United States from

approximately 2005 to 2010 (Legleiter and Bradley 2008, Young 2006). As a result, growers

resumed use of PPO inhibitors for control of glyphosate-resistant species, but with greater adoption

of PRE applications (USDA NASS 2020, Legleiter et al. 2009). PRE applications are significant

because weed seedlings receive a greater relative herbicide dose, and the PPO-inhibiting herbicide

active ingredients are different for PRE products and POST products. Flumioxazin, sulfentrazone,

and saflufenacil are the most common PRE herbicides used in soybean, with nationwide usage

between 9% and 20% for each compound, and even greater regional usage for some active

ingredients (USDA NASS 2020). Fomesafen is still a commonly used herbicide used both POST

and PRE on 20% of soybean hectares, and is the only herbicide that has been used extensively in

both the 1990s and the present (USDA NASS 1994, USDA NASS 2020).

Herbicide resistance is not the complete insensitivity to the herbicide, but rather a shift of

the dose-response curve. The potency of an herbicide is typically quantified using the ED50 statistic,

the dose of herbicide required to reduce growth by 50% (Burgos et al. 2013). The amount of curve

56
shift, or magnitude of resistance, is quantified with an R/S ratio. R:S ratios are calculated by

dividing the ED50 of the resistant biotype by the ED50 of the susceptible biotype. Each PPO

inhibitor has a unique R/S ratio for any given herbicide, which can vary by population. Reported

resistance ratios can be very different between research groups depending on variables measured,

experimental design, and growing conditions during the experiments. However, one consistent

observation is that the magnitude of resistance to lactofen and acifluorfen is typically much greater

than that of other PPO inhibitors (Lillie et al. 2020, Patzoldt et al. 2005, Shoup et al. 2003, Wuerffel

et al. 2015a). PPO-inhibitor resistance was first confirmed in waterhemp in 2001, shortly after a

period when lactofen, acifluorfen, and fomesafen were the most common PPO inhibitors in use

(Shoup et al. 2003). In contrast, R128 mutations were identified in waterhemp and Palmer

amaranth during a period when sulfentrazone, flumioxazin, saflufenacil, and fomesafen were the

most widely used PPO-inhibiting herbicides. To date there has not been a published direct

comparison of PPO-inhibitor resistance conferred by the ΔG210 and R128 mutations in either a

PRE or POST experiment for any PPO-inhibiting herbicides other than fomesafen. The objective

of this research was to determine if PPO inhibitors applied PRE select for the R128G more so than

for the ΔG210 mutation in waterhemp, which would be indicated by a greater magnitude of

resistance.

2.3 Materials and Methods

A series of greenhouse dose-response experiments were conducted to compare the PPO-

inhibitor resistance phenotypes conferred by the ΔG210 and R128G mutations in waterhemp for

soil-applied herbicides. The herbicides flumioxazin, fomesafen, saflufenacil, sulfentrazone,

trifludimoxazin, and oxadiazon were used to characterize PPO-inhibitor resistant (PPO-R)

57
waterhemp response to common soil-applied PPO inhibitors as well as unique PPO inhibitors with

unknown effects on resistant biotypes.

Biotypes homozygous for ΔG210, R128G, and susceptible PPX2 alleles were used. The

two resistant biotypes were sourced from a single population of waterhemp from Gibson County,

IN collected in 2016. Multiple individuals with homozygous alleles for ΔG210 and R128G were

identified via sanger sequencing and allowed to cross pollinate in isolation to increase seed

quantity and obtain pure lines with the desired genotype. The susceptible population was sourced

from a field population in Vigo County, IN collected in 2016 as part of a multi-state waterhemp

survey. Greenhouse experiments as well as genotyping data have confirmed that all individuals

are susceptible to the herbicide fomesafen and do not contain any known resistance-conferring

mutations (Mansfield 2021).

Seeds of each biotype were scarified with a 10% sodium hypochlorite solution. After

drying completely, 200 seeds were placed in a vial, then stored at 4C until planting within 24 hours.

The seeds were planted in square 10 cm by 10 cm plastic pots filled with 450 ml of pulverized

sandy-loam soil (pH 7, 3.2% OM) wetted to field capacity. The seeds were covered with another

50 ml of soil which corresponded to seed coverage of approximately 5 mm. Pots were again

watered to field capacity. To protect against soil borne pathogens, all pots were treated with a 10

ml drench of mefonoxam (Subdue Maxx, Syngenta Crop Protection, LLC, Greensboro, NC) at a

rate of 1 mg mefonoxam pot-1 (Wuerffel et al. 2015a).

Six PPO-inhibiting herbicides that are commonly used for soil-residual control were

applied to each pot using a track-mounted research sprayer (Generation III Research Sprayer,

DeVries Manufacturing, Hollandale, MN) calibrated to deliver 140 L ha-1 of spray solution at 207

kPa with an even-fan XR8002 nozzle (TeeJet Technologies, Glendale Heights, IL). Each herbicide

58
was applied at six rates in addition to a 0 g ai ha-1 rate in a log3 dosing structure. Rates were selected

based on previous research by Bi et al. (2019), Lillie et al. (2020), Steppig (2022), and Wuerffel

et al. (2015a). Herbicides and doses utilized for each experiment are shown in Table 2.2. After

application, herbicide treatments were incorporated into the soil profile via overhead irrigation

calibrated to deliver 1 cm of water. Soil moisture was maintained by placing pots into individual

plastic dishes and sub-irrigating with 80 ml of water when surface dryness occurred approximately

every other day (Harder et al. 2012, Lillie et al. 2020, Wuerffel et al. 2015a).

Waterhemp emergence was enumerated at 7 and 14 days after treatment (DAT).

Waterhemp control was assessed at 7 and 14 DAT on a 0 to 100% scale with 0 indicating no

reduction in waterhemp growth and emergence and 100 indicating complete waterhemp control.

Also at 14 DAT, waterhemp aboveground biomass was clipped at the soil surface and dried at 60C

prior to weighing. Data were converted to a percentage of the non-treated and used to fit non-linear

regression models using the DRC package in R studio 4.1.2 (Knezevic et al. 2007, Ritz et al. 2015).

Emergence data were used to fit Weibul type I models (Equation 1).

𝑓(𝑥) = 0 + (𝑑 − 0)𝑒𝑥𝑝 (−𝑒𝑥𝑝(𝑏(𝑙𝑜𝑔(𝑥) − 𝑒))) [1]

where b is the slope of the curve, , d is the upper asymptote, and e is the herbicide rate required to

produce 50% emergence reduction (i.e. LD50 value), using the drc package in R software v. 3.6.2

(Ritz et al. 2015). Biomass data were used to fit log logistic models (Equation 2).
𝑑
𝑓(𝑥) = [2]
1+exp (𝑏(log(𝑥)−log (𝑒)))

where b is the slope of the curve, d is the upper asymptote, and e is the herbicide rate required to

produce 50% biomass reduction (i.e. GR50 value), also using the drc package in R software v. 3.6.2

(Ritz et al. 2015). For both models, data were pooled over experimental runs after validating lack

of run interactions with ANOVA.

59
2.4 Results and Discussion

Calculated LD50 (emergence) and GR50 (biomass) values are shown in Table 2.3. Based on

weed emergence, the R128G and ΔG210 biotypes produced a significant resistance ratio for all

herbicides (p<0.05), with RG210/S ratios between 2.9 and 20 and RR128/S ratios between 2.5 and 37

(Table 2.4). Conversely, not all resistance ratios were significant for the biomass data with RG210/S

ratios of 1.1 to 3.9 and RR128/S ratios from 0.9 to 5.2. The GR50 values are the preferred statistic to

report resistance in POST experiments because growth reduction is a direct measure of herbicidal

activity. On the other hand, GR is an inferior way to measure PRE activity compared to LD because

it is not a direct measure of herbicide activity like LD. The following discussion will focus on LD

models because they are the most powerful and indicative of the true herbicide response observed

during these experiments. GR models generally showed similar trends, but with less power and

some spurious results that do not correlate with emergence or visual control data.

2.4.1 Flumioxazin

The two resistance mutations conferred a similar, low level of resistance to flumioxazin at

2.9 to 3.8-fold (Figure 2.1, Table 2.4). Fitted LD50 values were 3.9, 11, and 15 g ha-1 from the

susceptible, ΔG210, and R128G biotypes, respectively (Table 2.3). Our resistance ratio for ΔG210

is similar to that observed by Patzoldt et al. (2005), slightly lower than Lillie et al. (2020) and

substantially lower than Wuerffel et al. (2015a). These results fail to support our hypothesis

because flumioxazin is one of the most commonly used PRE herbicides in soybean production. If

flumioxazin was a major contributor to R128 mutation evolution and spread, then we would expect

a greater magnitude of resistance from R128 mutations toward flumioxazin. Because the two

mutations confer relatively equal resistance and flumioxazin is widely used in locations where

60
resistance is evolving, we find it very unlikely that the new mutations have been preferentially

selected by flumioxazin more so than other PPO herbicides applied PRE or POST.

2.4.2 Fomesafen

Waterhemp resistance to fomesafen was the greatest of all herbicides with 20 to 37-fold

resistance ratios (Figure 2.2, Table 2.4). The resistance ratio of 20 for the ΔG210 mutation is

slightly lower than that observed by Lillie et al. (2020) and Wuerffel et al. (2015a), who evaluated

PRE activity of fomesafen, but greater than Patzolt et al. (2005), Shoup et al. (2003), and Mansfield

(2021) who evaluated only POST activity. Given that fomesafen is used as both a PRE and POST

herbicide, had the greatest observed resistance in the present study, and is widely in U.S. soybean

production, we conclude that selection for R128 mutations was likely influenced by fomesafen

applications and not by other PRE PPO inhibitors (NASS 2020). This is somewhat surprising since

precedent in herbicide resistance literature would predict that a single resistance mutation would

dominate a geographical area, and the ΔG210 biotype predated the R128 biotype by a decade (Nie

et al. 2019, Shoup et al. 2003). Due to a hormetic response in the R128G biotype observed only

with fomesafen applications, it is difficult to determine the true difference in resistance between

ΔG210 and R128G biotypes for this herbicide. However, previous experiments (Steppig 2022,

Steppig et al. 2017) demonstrate that fomesafen elicits greater resistance response from R128G

than ΔG210 in POST dose-response experiments.

2.4.3 Oxadiazon

Both waterhemp biotypes had a significant resistance ratio in response to oxadiazon with

resistance ratios of 7.4 for ΔG210 and 2.5 for R128G (Figure 2.3, Table 2.4). The ΔG210 mutation

did confer a greater resistance to oxadiazon than the R128G mutation (2.9-fold). This result is

61
similar to PPO2 enzyme kinetic data that showed in vitro that PPO2 with the ΔG210 mutation had

a resistance ratio of 151, whereas a R128L mutation had a resistance factor of only 1 (Rangani et

al. 2019). However, the current experiment used waterhemp with an R128G mutation whereas the

in vitro experiment utilized an R128L mutation, so the resistance factor of R128G in vitro is not

entirely known.

Oxadiazon is a PRE PPO-inhibiting herbicide used for control of annual grasses in turfgrass

(Anonymous 2021). Oxadiazon is not used in soybean or labeled for control of waterhemp, so the

effect of PPO-inhibitor resistance mutations in the PPX2 gene on oxadiazon efficacy has not been

previously tested. It was included in this experiment because goosegrass [Eleusine indica (L.)

Gaertn.], a labeled target weed, has evolved resistance to oxadiazon (McElroy et al. 2017).

Goosegrass resistance to oxadiazon is unique in that it involves a mutation in the PPX1 gene rather

than the PPX2 gene, as has been observed in other species (Bi et al. 2019). It is unknown if

evolution of PPX1 mediated resistance is a consequence of goosegrass biology, oxadiazon-enzyme

interaction, or a combination of other factors.

These results may provide some insight as to why goosegrass evolved resistance via a

PPX1 mutation. The evolution of the ΔG210 mutation is only likely in a minority of species and

may not be possible in goosegrass (Dayan et al. 2018). Furthermore, our observed resistance to

oxadiazon conferred by the R128G mutation was low level (2.5-fold). A PPX1 mutation may have

been necessary to confer a sufficient selective advantage. If this is the case, then we can expect an

increase in resistance conferring mutations in the PPX1 gene in Amaranthus or other weedy

species as growers increase the use of previously under-utilized PPO inhibitors or novel PPO

inhibitors that are soon to be commercialized. (Porri et al. 2022a, Witschel et al. 2021).

62
2.4.4 Saflufenacil

Resistance to saflufenacil was second only to fomesafen with resistance ratios of 10 to 14

(Table 2.4, Figure 2.4). These results are very different from Steppig (2022) who only observed

R/S ratios of 1.3 to 1.9 in an identical PRE experiment and ratios of 2.0 to 2.4 in a POST

experiment using the same populations. The only other experiment utilizing a saflufenacil dose

response is Wu et al. (2020) who reported a 2.6 to 3-fold resistance in a POST experiment on

Palmer amaranth. However, their model was fitted with visual control estimates and did not have

a sufficiently low dose range to avoid extrapolation of a large portion of the dose-response curve.

2.4.5 Sulfentrazone

Resistance to sulfentrazone was 4.2-fold for ΔG210 and 8.8 for R128G (Table 2.4, Figure

2.5). Sulfentrazone is the only herbicide that appears to support our hypothesis in that it is the only

herbicide where R128G conferred a greater magnitude of resistance than ΔG210 (R128G/ΔG210

ratio of 2.1). Such an observation is odd considering that the R128 mutations are also present in

Palmer amaranth and at a greater frequency than in waterhemp, yet growers in the Mid-South

region of the United States use very little sulfentrazone. We conclude that greater observed

resistance to sulfentrazone by plants with the R128G mutation is likely more of a coincidence than

a major influence in the selection for the R128G mutation. Especially considering that the

maximum observed resistance ratio was 8.8 for sulfentrazone compared to 20 to 30 for fomesafen.

Despite being an unlikely cause of resistance selection, the observed difference in resistance is

potentially meaningful for resistance selection and efficacy in the field.

63
2.4.6 Trifludimoxazin

Both R biotypes exhibited resistance to trifludimoxazin of 2.4 to 3.5-fold (Figure 2.6, Table

2.4). This observation was similar to Steppig (2022) who observed resistance ratios of 1.5 to 2.5

in an identical experiment. This observation was unexpected, because trifludimoxazin is purported

to control resistant biotypes equally well as susceptible biotypes due to unique target enzyme

binding properties (Porri et al. 2022a, Steppig 2022, Witschel et al. 2021). An in-vitro experiment

by Porri et al. (2022a) showed that ΔG210 PPO2 protein had only a 7-fold resistance (IC50) to

trifludimoxazin compared to the wild type compared to a resistance factor of over 1000 for

saflufenacil, lactofen, and fomesafen. This translates to ΔG210 and R128G waterhemp biotypes

not having resistance in a POST dose-response experiment (Steppig 2022). Our observed

resistance to trifludimoxazin is very similar to flumioxazin which has a very similar molecular

structure (Asher et al. 2021, Shaner 2014). Given these apparently contradictory results, it is

unclear how field applications of trifludimoxazin will affect the spread of PPO resistance. If

trifludimoxazin truly does not elicit a resistance response then field applications may not select for

target-site resistance just like a herbicide of a different site of action (Wuerffel et al. 2015b).

2.4.7 Variance in Observed Resistance

Direct comparisons between resistance bioassay results from different research groups,

especially for resistance ratios, requires scrutiny of the differences in the experimental methods

and plant materials. First, exact experimental conditions can never be replicated. Plant response to

herbicides can vary due to differences in growing medium, greenhouse temperature, time of year,

and light regime (Schafer et al. 2012, Wichert et al. 1992). Further, the plant response to herbicides

is so variable that comparisons in terms of absolute dose, such as g ha-1, are nearly meaningless.

64
This is especially true when comparing field experiments to greenhouse experiments and when

using a recommended herbicide use rate as a benchmark in a greenhouse experiment.

Because magnitude of resistance is a more stable indicator of resistance phenotype, the

most valuable parameter for evaluating resistance is the R/S ratio (Burgos et al. 2013). Even so,

R/S ratios reported in the literature are quite variable and difficult to compare, especially for PPO-

inhibiting herbicides (Lillie et al. 2020, Rangani et al. 2019, Shoup et al. 2003, Steppig 2022,

Wuerffel et al. 2015a). Even our own methods repeated with the same populations and similar

dose structure resulted in very different observed R/S ratios (Steppig 2022). The factors that

contribute to a R/S ratio are the enzyme kinetic effects conferred by the mutation of interest, as

well as quantitative traits that influence herbicide efficacy of both resistant and susceptible plants.

In an ideal scenario, dose-response assays would be conducted on near isogenic lines that are

homozygous resistant and susceptible. However, in wild, obligate outcrossing species such as

waterhemp and Palmer amaranth, generating isogenic lines is impossible in any reasonable amount

of time. The next best option, which is still intensive, is selecting resistant and susceptible F1

individuals from a cross and using F3 seeds homozygous at the gene of interest for experiments.

Such a strategy would reduce plant to plant variance, but population level variance would still be

present. However, producing these lines is superior because it isolates the effect of the resistance

mutation and controls for minor, quantitative genes that could influence herbicide susceptibility.

Because of the difficulty of generating pure lines, published results are often a comparison of an

unrelated susceptible population compared to the resistant population.

The lack of consistency in reported resistance to PPO inhibitors suggests that there is some

underlying covariate has contributed to the magnitude of resistance and has a variable influence

by population and/or environment. To date, there has been no evidence of either reactive oxygen

65
species (ROS) scavenging or PPO-protein expression level as contributors to PPO resistance

(Mansfield 2021, Barker and Dayan 2020, Montgomery et al. 2021). One potential contributing

factor is herbicide detoxification. Soybean varieties are known to have differential PPO-inhibitor

metabolism capabilities (Dayan et al. 1997). Waterhemp and Palmer amaranth have also been

shown to be resistant in the absence of target-site resistance mutations indicating the high

likelihood of metabolic resistance (Noguera et al. 2021, Obenland et al. 2019, Varanasi et al. 2018).

There are several examples of target-site and non-target-site resistance co-occurring in the same

plants, so it is possible that herbicide detoxification is a major contributor to population level PPO-

inhibitor response (Alcántara-de la Cruz et al. 2016, Hatami et al. 2016, Kaundun 2010, Rey-

Caballero et al. 2017).

2.4.8 Practical Implications

One point of consideration in resistance selection is the number of obstacles before

reproduction. The most noticeable cases of herbicide resistance are to POST applications, because

they are often the final selection event for a weed in a field. In contrast, PRE herbicides will usually

have a POST herbicide application of a different mode of action prior to a weed being able to

reproduce. Current recommendations for growers are to utilize herbicide programs in which a PRE

herbicide is followed by a planned POST herbicide in combination with an overlapping residual

herbicide (Norsworthy et al. 2012). For troublesome Palmer amaranth and waterhemp populations,

a zero-tolerance approach is recommended where growers should not allow a single weed to

produce seed in their fields where resistant plant are either confirmed or suspected (Norsworthy et

al. 2014). Given these current resistance management recommendations, the risk of herbicide

resistance to PRE herbicides broadly has been low because of effective POST herbicide

technologies. However, this current state is fragile and actively degrading. To date there is
66
confirmation of glufosinate resistant Palmer amaranth and several reports of herbicide failure and

resistance to dicamba and 2,4-D in waterhemp (Bobadilla et al. 2022, Priess et al. 2022, Shergill

et al. 2018).

The PPO inhibitors used PRE have maintained a high level of utility when used on PPO-

inhibitor resistant populations, especially when used in combination with other modes of action

(Falk et al. 2006, Umphres et al. 2018, Wuerffel et al. 2015b). Based on the present study, there is

little evidence that the utility of PPO inhibitors applied PRE will diminish any further as the R128

mutation increases in frequency. Furthermore, waterhemp or other plants that have both mutations

may not pose a serious threat to PPO-inhibitor effectiveness either. A heterozygous plant will

likely exhibit an intermediate response, and a double mutant allele with both the R128 and ΔG210

mutation carries a significant fitness penalty if the enzyme is functional at all (Noguera et al. 2021,

Porri et al. 2022b, Steppig et al. 2017).

To mitigate PPO-inhibitor resistance, a farmer should exchange diphenyl-ether herbicides

(lactofen, fomesafen, and acifluorfen) for other effective POST and overlapping residual

applications, particularly herbicides from HRAC Groups 4, 10, and 15. While switching to non-

PPO-inhibiting herbicides is ideal to stop the spread of resistance, PRE PPO inhibitors can still be

part of a weed management program and can help to reduce total seedbank populations. In

conclusion, the R128G mutation in waterhemp is not substantially different from the ΔG210

mutation with respect to resistance. PPO inhibitors are still partially effective PRE and growers

must steward these products to prolong their utility in weed management programs. Further

investigations are warranted to explain trifludimoxazin resistance selection and the variability of

PPO-inhibitor resistance.

67
2.5 Literature Cited

Anonymous (2021) Ronstar® FLO herbicide product label. Cary, NC: Bayer Environmental
Science
Alcántara-de la Cruz R, Fernández-Moreno PT, Ozuna C V., Rojano-Delgado AM, Cruz-Hipolito
HE, Domínguez-Valenzuela JA, Barro F, de Prado R (2016) Target and non-target site
mechanisms developed by glyphosate-resistant hairy beggarticks (Bidens pilosa L.)
populations from Mexico. Front Plant Sci 7:1492

Asher BS, Dotray PA, Liebl RA, Keeling JW, Ritchie GD, Udeigwe TK, Reed JD, Keller KE,
Bowe SJ, Aldridge RB, Simon A (2021) Vertical mobility and cotton tolerance to
trifludimoxazin, a new protoporphyrinogen oxidase-inhibiting herbicide, in three West Texas
soils. Weed Technol 35:144–148

Barker A, Dayan F (2020) The physiological basis of differential resistance to PPO-inhibiting


herbicides used pre- and post-emergent. in Proceedings of the Weed Science Society of
America Annual Meeting. Maui, HI: Weed Science Society of America

Bell MS, Hager AG, Tranel PJ (2013) Multiple resistance to herbicides from four site-of-action
groups in waterhemp (Amaranthus tuberculatus). Weed Sci 61:460–468

Bi B, Wang Q, Coleman JJ, Porri A, Peppers JM, Patel JD, Betz M, Lerchl J, McElroy JS (2019)
A novel mutation A212T in chloroplast Protoporphyrinogen Oxidase (PPO1) confers
resistance to PPO inhibitor oxadiazon in Eleusine indica. Pest Manag Sci:76:1786–1794

Bobadilla LK, Giacomini DA, Hager AG, Tranel PJ (2022) Characterization and inheritance of
dicamba resistance in a multiple-resistant waterhemp (Amaranthus tuberculatus) population
from Illinois. Weed Sci 70:4–13

Burgos NR, Tranel PJ, Streibig JC, Davis VM, Shaner D, Norsworthy JK, Ritz C (2013) Review:
confirmation of resistance to herbicides and evaluation of resistance levels. Weed Sci 61:4–
20

Cahoon CW, Jordan DL, Tranel PJ, Riggins C, Seagroves R, Inman M, Everman W, Leon R,
Professor A, Professor A, Professor Emeritus R, Cahoon Jr CW (2022) In-field assessment
of EPSPS amplification on fitness cost in mixed glyphosate-resistant and -sensitive
populations of Palmer amaranth (Amaranthus palmeri). Weed Sci 70:1–20
68
Chatham LA, Wu C, Riggins CW, Hager AG, Young BG, Roskamp GK, Tranel PJ (2015) EPSPS
gene amplification is present in the majority of glyphosate-resistant Illinois waterhemp
(Amaranthus tuberculatus) populations. Weed Technol 29:48–55

Dayan FE, Barker A, Tranel PJ (2018) Origins and structure of chloroplastic and mitochondrial
plant protoporphyrinogen oxidases: implications for the evolution of herbicide resistance.
Pest Manag Sci 74:2226–2234

Dayan FE, Daga PR, Duke SO, Lee RM, Tranel PJ, Doerksen RJ (2010) Biochemical and
structural consequences of a glycine deletion in the α-8 helix of protoporphyrinogen oxidase.
Biochim Biophys Acta - Proteins Proteomics 1804:1548–1556

Dayan FE, Weete JD, Duke SO, Hancock HG (1997) Soybean (Glycine max) cultivar differences
in response to sulfentrazone. Weed Sci 45:634–641

Falk JS, Shoup DE, Al-Khatib K, Peterson DE (2006) Protox-resistant common waterhemp
(Amaranthus rudis) response to herbicides applied at different growth stages. Weed Sci
54:793–799

Giacomini DA, Umphres AM, Nie H, Mueller TC, Steckel LE, Young BG, Scott RC, Tranel PJ
(2017) Two new PPX2 mutations associated with resistance to PPO-inhibiting herbicides in
Amaranthus palmeri. Pest Manag Sci 73:1559–1563

Green JM, Owen MDK (2010) Herbicide-resistant crops: Utilities and limitations for herbicide-
resistant weed management. J Agric Food Chem 59:5819–5829

Hao GF, Zuo Y, Yang S-G, Yang G-F (2011) Protoporphyrinogen oxidase inhibitor: An ideal
target for herbicide discovery. Chim Int J Chem 65:961–969

Harder DB, Nelson KA, Smeda RJ (2012) Management options and factors affecting control of a
common waterhemp (Amaranthus rudis) biotype resistant to protoporphyrinogen oxidase-
inhibiting herbicides . Int J Agron 2012:1–7

Hatami ZM, Gherekhloo J, Rojano-Delgado AM, Osuna MD, Alcántara R, Fernández P,


Sadeghipour HR, de Prado R (2016) Multiple mechanisms increase levels of resistance in
rapistrum rugosum to ALS herbicides. Front Plant Sci 7:169

69
Heap I (2023) International survey of herbicide resistant weeds. http://www.weedscience.org.
Accessed March 25, 2023

Kaundun SS (2010) An aspartate to glycine change in the carboxyl transferase domain of acetyl
CoA carboxylase and non-target-site mechanism(s) confer resistance to ACCase inhibitor
herbicides in a Lolium multiflorum population. Pest Manag Sci 66:1249–1256

Knezevic SZ, Streibig JC, Ritz C (2007) Utilizing R software package for dose-response studies:
The concept and data analysis. Weed Technol 21:840–848

Lee HJ, Duke M V., Duke SO (1993) Cellular localization of protoporphyrinogen-oxidizing


activities of etiolated barley (Hordeum vulgare L.) leaves: Relationship to mechanism of
action of protoporphyrinogen oxidase-inhibiting herbicides. Plant Physiol 102:881–889

Legleiter TR, Bradley KW (2008) Glyphosate and multiple herbicide resistance in common
waterhemp (Amaranthus rudis) populations from Missouri. Weed Sci 56:582–587

Legleiter TR, Bradley KW, Massey RE (2009) Glyphosate-resistant waterhemp (Amaranthus


rudis) control and economic returns with herbicide programs in soybean. Weed Technol
23:54–61

Lermontova I, Kruse E, Mock HP, Grimm B (1997) Cloning and characterization of a plastidal
and a mitochondrial isoform of tobacco protoporphyrinogen IX oxidase. Proc Natl Acad Sci
USA 94:8895–8900

Lillie KJ, Giacomini DA, Tranel PJ (2020) Comparing responses of sensitive and resistant
populations of Palmer amaranth (Amaranthus palmeri) and waterhemp (Amaranthus
tuberculatus var. rudis) to PPO inhibitors. Weed Technol. 34: 140–146. doi:
10.1017/wet.2019.84

Mansfield BC (2021) Characterization of protoporphyrinogen oxidase (PPO) herbicide resistance


in tall waterhemp (Amaranthus tuberculatus). MS thesis. West Lafayette IN: Purdue
University. 158 p

McElroy JS, Head WB, Wehtje GR, Spak D (2017) Identification of goosegrass (Eleusine indica)
biotypes resistant to preemergence-applied oxadiazon. Weed Technol 31:675–681

70
Mendes RR, Takano HK, Adegas FS, Oliveira RS, Gaines TA, Dayan FE (2020) R128L target site
mutation in PPO2 evolves in wild poinsettia (Euphorbia heterophylla) with cross-resistance
to PPO-inhibiting herbicides. Weed Sci. 68:437–444

Murphy BP, Tranel PJ (2019) Target-site mutations conferring herbicide resistance. Plants 8:382

Nie H, Mansfield BC, Harre NT, Young JM, Steppig NR, Young BG (2019) Investigating target‐
site resistance mechanism to the PPO‐inhibiting herbicide fomesafen in waterhemp and
interspecific hybridization of Amaranthus species using next generation sequencing. Pest
Manag Sci 75:3235–3244

Noguera MM, Rangani G, Heiser J, Bararpour T, Steckel LE, Betz M, Porri A, Lerchl J,
Zimmermann S, Nichols RL, Roma-Burgos N (2021) Functional PPO2 mutations: co-
occurrence in one plant or the same ppo2 allele of herbicide-resistant Amaranthus palmeri in
the US mid-south. Pest Manag Sci 77:1001–1012

Norsworthy JK, Griffith G, Griffin T, Bagavathiannan M, Gbur EE (2014) In-field movement of


glyphosate-resistant Palmer amaranth (Amaranthus palmeri) and its impact on cotton lint
yield: Evidence supporting a zero-threshold strategy. Weed Sci 62:237–249
Norsworthy JK, Ward SM, Shaw DR, Llewellyn RS, Nichols RL, Webster TM, Bradley KW,
Frisvold G, Powles SB, Burgos NR, Witt WW, Barrett M (2012) Reducing the risks of
herbicide resistance: best management practices and recommendations. Weed Sci 60:31–62

Obenland OA, Ma R, O’Brien SR, Lygin A V., Riechers DE (2019) Carfentrazone-ethyl resistance
in an Amaranthus tuberculatus population is not mediated by amino acid alterations in the
PPO2 protein. PLoS One 14:e0215431

Patzoldt WL, Hager AG, McCormick JS, Tranel PJ (2006) A codon deletion confers resistance to
herbicides inhibiting protoporphyrinogen oxidase. Proc Natl Acad Sci USA 103:12329–34

Patzoldt WL, Tranel PJ (2007) Multiple ALS mutations confer herbicide resistance in waterhemp
(Amaranthus tuberculatus). Weed Sci 55:421–428

Patzoldt WL, Tranel PJ, Hager AG (2005) A waterhemp (Amaranthus tuberculatus) biotype with
multiple resistance across three herbicide sites of action. Weed Sci 53:30–36

71
Porri A, Betz M, Seebruck K, Knapp M, Johnen P, Witschel M, Aponte R, Liebl R, Tranel PJ,
Lerchl J (2022a) Inhibition profile of trifludimoxazin towards PPO2 target site mutations.
Pest Manag Sci 79:507–519

Porri A, Noguera MM, Betz M, Sälinger D, Brändle F, Bowe SJ, Lerchl J, Meyer L, Knapp M,
Roma-Burgos N (2022b) Can double PPO mutations exist in the same allele and are such
mutants functional? Pest Manag Sci 78:2258–2264

Priess GL, Norsworthy JK, Godara N, Mauromoustakos A, Butts TR, Roberts TL, Barber T (2022)
Confirmation of glufosinate-resistant Palmer amaranth and response to other herbicides.
Weed Technol 36:368–372

Rangani G, Salas-Perez RA, Aponte RA, Knapp M, Craig IR, Mietzner T, Langaro AC, Noguera
MM, Porri A, Roma-Burgos N (2019) A novel single-site mutation in the catalytic domain of
Protoporphyrinogen Oxidase IX (PPO) confers resistance to PPO-inhibiting herbicides. Front
Plant Sci 10:568

Rey-Caballero J, Menéndez J, Osuna MD, Salas M, Torra J (2017) Target-site and non-target-site
resistance mechanisms to ALS inhibiting herbicides in Papaver rhoeas. Pestic Biochem
Physiol 138:57–65

Ritz C, Baty F, Streibig JC, Gerhard D (2015) Dose-response analysis using R. PLoS One 10(12):
e0146021. https://doi.org/10.1371/journal.pone.0146021

Rousonelos SL, Lee RM, Moreira MS, VanGessel MJ, Tranel PJ (2012) Characterization of a
common ragweed (Ambrosia artemisiifolia) population resistant to ALS- and PPO-inhibiting
herbicides. Weed Sci 60:335–344

Salas RA, Burgos NR, Tranel PJ, Singh S, Glasgow L, Scott RC, Nichols RL (2016) Resistance
to PPO-inhibiting herbicide in Palmer amaranth from Arkansas. Pest Manag Sci 72:864–869

Schafer JR, Hallett SG, Johnson WG (2012) Response of giant ragweed (Ambrosia trifida),
horseweed (Conyza canadensis), and common lambsquarters (Chenopodium album) biotypes
to glyphosate in the presence and absence of soil microorganisms. Weed Sci 60:641–649

72
Schultz JL, Chatham LA, Riggins CW, Tranel PJ, Bradley KW (2015) Distribution of herbicide
resistances and molecular mechanisms conferring resistance in Missouri waterhemp
(Amaranthus rudis Sauer) populations. Weed Sci 63:336–345

Shaner DL, ed (2014) Herbicide Handbook. 10th edn. Lawrence, KS: Weed Science Society of
America. 513 p

Shergill LS, Barlow BR, Bish MD, Bradley KW (2018) Investigations of 2,4-D and multiple
herbicide resistance in a Missouri waterhemp (Amaranthus tuberculatus) population. Weed
Sci 66:386–394

Shoup DE, Al-Khatib K, Peterson DE (2003) Common waterhemp (Amaranthus rudis) resistance
to protoporphyrinogen oxidase-inhibiting herbicides. Weed Sci 51:145–150

Steppig NR (2022) Evaluation of trifludimoxazin, a new protoporphyrinogen oxidase- inhibiting


herbicide, for use in soybean. PhD dissertation. West Lafayette, IN: Purdue University. 172p

Steppig NR, Mansfield BC, Nie H, Young JM, Young BG (2017) Presence of an alternative
mechanism of resistance to PPO-inhibiting herbicides in tall waterhemp populations from
Indiana, Illinois, Iowa, Missouri, and Minnesota. Proceedings of the 2017 North Central
Weed Science Society Annual Meeting. St. Louis, MO: North Central Weed Science
Society
Umphres AM, Steckel LE, Mueller TC (2018) Control of protoporphyrinogen oxidase inhibiting
herbicide resistant and susceptible Palmer amaranth (Amaranthus palmeri) with soil-applied
protoporphyrinogen oxidase–inhibiting herbicides. Weed Technol 32:95–100

[USDA-NASS] US Department of Agriculture, National Agricultural Statistics Service (2020)


Agricultural chemical use program.
https://www.nass.usda.gov/Surveys/Guide_to_NASS_Surveys/Chemical_Use/. Accessed:
March 15, 2022

[USDA NASS] US Department of Agriculture, National Agricultural Statistics Service (1994)


Field crops chemical use survey.
http://usda.mannlib.cornell.edu/usda/nass/AgriChemUsFC//1990s/1994/AgriChemUsFC-
03-16-1994.txt. Accessed December 15, 2017

73
Van Wychen L (2022) 2022 Survey of the most common and troublesome weeds in broadleaf
crops, fruits & vegetables in the United States and Canada. Weed Science Society of America
National Weed Survey Dataset. Available: http://wssa.net/wp-content/uploads/2022 Weed-
Survey Broadleaf crops.xlsx

Varanasi VK, Brabham C, Norsworthy JK (2018) Confirmation and characterization of non-target


site resistance to fomesafen in Palmer amaranth (Amaranthus palmeri). Weed Sci 66:702–
709

Vennapusa AR, Faleco F, Vieira B, Samuelson S, Kruger GR, Werle R, Jugulam M (2018)
Prevalence and mechanism of atrazine resistance in waterhemp (Amaranthus tuberculatus)
from Nebraska. Weed Sci 66:595–602

Vila-Aiub MM, Neve P, Powles SB (2009) Fitness costs associated with evolved herbicide
resistance alleles in plants. New Phytol 184:751–767

Watanabe N, Che FS, Iwano M, Takayama S, Yoshida S, Isogai A (2001) Dual targeting of spinach
protoporphyrinogen oxidase II to mitochondria and chloroplasts by alternative use of two in-
frame initiation codons. J Biol Chem 276:20474–20481

Wichert RA, Bozsa R, Talbert RE, Oliver LR (1992) Temperature and relative humidity effects
on diphenylether herbicides.Weed Technol 6:19–24

Witschel M, Aponte R, Armel G, Bowerman P, Mietzner T, Newton T, Porri A, Simon A, Seitz T


(2021) Tirexor®—design of a new resistance breaking protoporphyrinogen IX oxidase
inhibitor. Recent Highlights Discov Optim Crop Prot Prod:501–509

Wu C, Davis AS, Tranel PJ (2018) Limited fitness costs of herbicide-resistance traits in


Amaranthus tuberculatus facilitate resistance evolution. Pest Manag Sci 74:293–301

Wu C, Goldsmith MR, Pawlak J, Feng P, Smith S, Navarro S, Perez-Jones A (2020) Differences


in efficacy, resistance mechanism and target protein interaction between two PPO inhibitors
in Palmer amaranth (Amaranthus palmeri). Weed Sci 68:105–115

Wuerffel RJ, Young JM, Matthews JL, Young BG (2015a) Characterization of PPO-inhibitor–
resistant waterhemp (Amaranthus tuberculatus) response to soil-applied PPO-inhibiting
herbicides. Weed Sci 63:511–521

74
Wuerffel RJ, Young JM, Tranel PJ, Young BG (2015b) Soil-residual protoporphyrinogen oxidase–
inhibiting herbicides influence the frequency of associated resistance in waterhemp
(Amaranthus tuberculatus). Weed Sci 63:529–538

Young BG (2006) Changes in herbicide use patterns and production practices resulting from
glyphosate-resistant crops. Weed Technol 20:301–307

75
Table 2.1. Discovery of novel PPO-inhibitor resistance mutations in waterhemp and Palmer amaranth.
Mutation Species Sample Collected Author(s) Place of Collection
ΔG210 Waterhemp 2001 Patzoldt et al. 2006, Kansas
Shoup et al. 2003
ΔG210 Palmer 2011 Salas et al. 2016 Lawrence County, Arkansas
Amaranth
R128G and R128M Palmer 2016 Giacomini et al. 2017 Shelby County, Tennessee, Lauderdale
76

Amaranth County, Tennessee, and Woodruff County,


Arkansas
G399A Palmer 2015 Rangani et al. 2019 Mississippi County, Arkansas
Amaranth
R128G and R128I Waterhemp 2016 Nie et al. 2019 14 counties across Illinois, Indiana, Minnesota,
and Missouri
Table 2.2. Herbicides used for PRE dose-response experiments on susceptible and resistant waterhemp biotypes with ΔG210 and
R128G mutations.
Common name Trade name Manufacturer Manufacturer location Lower ratea Upper rate

--------g ai ha-1--------

Fomesafen Flexstar® Syngenta Crop Protection LLC Greensboro, NC 2.47 600

Flumioxazin Valor® SX Valent USA Corp. Walnut Creek, CA 0.61 150


77

Saflufenacil Sharpen® BASF Research Triangle Park, NC 0.61 150

Sulfentrazone Spartan® FMC Corp Philadelphia PA 2.88 700

Trifludimoxazin Tirexor® BASF Research Triangle Park, NC 0.31 75

Oxadiazon Ronstar® Flo Bayer Environmental Sciences Cary, NC 28.0 6800


a
All herbicides were applied at 6 rates in addition to a nontreated check in a log3 dosing structure.
Table 2.3. Response of ΔG210, R128G and susceptible biotypes of waterhemp to PRE
application of PPO-inhibiting herbicides in the greenhouse, as measured by seedling emergence
(LD50) and shoot biomass (GR50).
Emergence Biomass

Herbicide Biotype LD50a std err GR50b std err

--------------------g ai ha-1--------------------

Flumioxazin ΔG210 11.4 2.9 6.2 3.2

R128G 15.1 4.2 11.3 3.7

Susceptible 3.9 0.8 2.8 0.7

Fomesafen ΔG210 662.9 214.8 87.4 24.4

R128G 1199.3 338.1 199.9 118.2

Susceptible 32.6 7.4 22.4 7.6

Oxadiazon ΔG210 1413.3 337.3 92.4 30.0

R128G 474.0 80.7 192.8 37.2

Susceptible 191.0 34.1 66.3 11.9

Sulfentrazone ΔG210 75.9 17.2 17.0 4.4

R128G 161.3 52.5 25.6 7.2

Susceptible 18.4 4.5 11.1 2.2

Saflufenacil ΔG210 186.6 61.2 7.0 1.6

R128G 127.4 53.1 21.3 4.0

Susceptible 12.8 4.0 4.1 0.7

Trifludimoxazin ΔG210 10.2 1.9 6.4 0.5

R128G 7.1 1.6 5.2 0.4

Susceptible 2.9 0.5 6.0 0.3


a
LD50 value was fitted with a 3-parameter Weibull type I model.
b
GR50 value was fitted with a 3-parameter log logistic model.

78
Table 2.4. Resistance ratios between susceptible, R128G, and ΔG210 waterhemp biotypes for
six PPO-inhibiting herbicides applied PRE in the greenhouse.
Emergence (LD50) Biomass (GR50)

Herbicide RR128/RG210a RG210/S RR128/S RR128/RG210a RG210/S RR128/S

Flumioxazin 1.3 2.9* 3.8* 1.9 2.2 4.0*

Fomesafen 1.8 20.3* 36.8* 2.3* 3.9 8.9

Oxadiazon 0.3* 7.4* 2.5* 2.1* 1.4 2.9*

Sulfentrazone 2.1* 4.2* 8.8* 1.5 1.5 2.3

Saflufenacil 0.7 14.6* 10.0* 3.1* 1.7 5.2*

Trifludimoxazin 0.7 3.5* 2.4* 0.8 1.1 0.9


a
abbreviations RG210- ΔG210, RR128 – R128G, S- Susceptible
b
ratio values with a * indicate a ratio significantly different than 1.0 at α = 0.05

79
Waterhemp emergence (% of nontreated)

Adjusted dry weight (% of nontreated)


S
1
0
0
0
80

Flumioxazin rate (g ai ha-1) Flumioxazin rate (g ai ha-1)

Figure 2.1. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to flumioxazin applied preemergence.
Emergence and biomass were evaluated 14 days after treatment and normalized to a percentage of the untreated for each biotype.
Adjusted dry weight (% of nontreated)
Waterhemp emergence (% of nontreated)
81

Fomesafen rate (g ai ha-1) Fomesafen rate (g ai ha-1)

Figure 2.2. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to fomesafen applied preemergence.
Emergence and biomass were evaluated 14 days after treatment and normalized to a percentage of the untreated for each biotype.
Waterhemp emergence (% of nontreated)

Adjusted dry weight (% of nontreated)


82

Oxadiazon rate (g ai ha-1) Oxadiazon rate (g ai ha-1)

Figure 2.3. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to oxadiazon applied preemergence.
Emergence and biomass were evaluated 14 days after treatment and normalized to a percentage of the untreated for each biotype.
Adjusted dry weight (% of nontreated)
Waterhemp emergence (% of nontreated)
83

Saflufenacil rate (g ai ha-1) Saflufenacil rate (g ai ha-1)

Figure 2.4. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to saflufenacil applied preemergence.
Emergence and biomass were evaluated 14 days after treatment and normalized to a percentage of the untreated for each biotype.
Waterhemp emergence (% of nontreated)

Adjusted dry weight (% of nontreated)


84

Sulfentrazone rate (g ai ha-1) Sulfentrazone rate (g ai ha-1)

Figure 2.5. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to sulfentrazone applied preemergence.
Emergence and biomass were evaluated 14 days after treatment and normalized to a percentage of the untreated for each biotype.
Waterhemp emergence (% of nontreated)

Adjusted dry weight (% of nontreated)


85

Trifludimoxazin rate (g ai ha-1) Trifludimoxazin rate (g ai ha-1)

Figure 2.6. Dose-response curves of the ΔG210, R128G, and susceptible waterhemp biotypes to trifludimoxazin applied preemergence.
Emergence and biomass were evaluated 14 days after treatment and normalized to a percentage of the untreated for each biotype.
DOES TRIFLUDIMOXAZIN SELECT FOR RESISTANT
WATERHEMP (AMARANTHUS TUBERCULATUS) SIMILARLY TO
OTHER PPO-INHIBITING HERBICIDES?

Short title: Trifludimoxazin ΔG210

3.1 Abstract

Resistance to PPO-inhibiting herbicides in waterhemp [Amaranthus tuberculatus (Moq.)

J.D.Sauer] is primarily conferred by the ΔG210 mutation in the PPX2 gene. Soil applications of

PPO inhibitors provide partial control of resistant populations at the cost of reduced length of

residual activity while selecting for a greater frequency of PPO-inhibitor-resistant (PPO-R)

individuals. Trifludimoxazin is a PPO-inhibiting herbicide currently under development that has

soil and foliar activity and has been documented to control PPO-R waterhemp (ΔG210 and R128G)

in foliar applications. Our objective was to determine if applications of trifludimoxazin select for

PPO-R individuals when applied PRE or POST in a similar manner as do other PPO inhibitors,

and determine if combinations of trifludimoxazin with other PPO inhibitors can reduce selection

for PPO-R individuals. Separate PRE and POST field experiments were conducted in 2020 and

2021 at two Indiana locations. Trifludimoxazin, saflufenacil, and fomesafen were applied PRE or

POST at rates of 12.5, 25, and 263 g ai ha-1, respectively, along with all two- and three-way

combinations of those herbicides and a no herbicide control. Leaf tissue from the first 25 plants

emerging in each plot in the PRE experiment were collected, and also from 25 surviving plants in

each plot from the POST experiment. Following DNA extraction, qPCR assays were performed to

test for ΔG210 mutation on all plant tissue samples. Overall, a greater number of PPO-R plants

survived the foliar herbicide applications than emerged through soil applications. Trifludimoxazin

86
did not increase the frequency of PPO-resistant individuals when applied to soil, but when applied

to foliage, increased the frequency of PPO-resistant individuals by 2.5 to 2.6-fold, similar to other

PPO inhibitors applied to foliage. Despite selection for PPO-R individuals, herbicide combinations

increased the length of residual waterhemp control. Therefore, fewer waterhemp plants survived,

which reduces the reliance on subsequent herbicide applications or other non-chemical practices

to control surviving waterhemp plants.

Nomenclature: Waterhemp, Amaranthus tuberculatus (Moq.) J.D.Sauer, fomesafen, saflufenacil,

trifludimoxazin

Key words: PPO-inhibitor resistance, target-site resistance, ΔG210 mutation, resistance selection

3.2 Introduction

Waterhemp [Amaranthus tuberculatus (Moq.) J.D.Sauer] is a summer annual weed native

to the midwestern United States that is capable of producing over 500,000 seeds per plant

(Heneghan and Johnson 2017). Waterhemp is particularly troublesome because of its propensity

to evolve resistance to several different herbicide site-of-action groups (Heap 2023, Schultz et al.

2015). Confirmed cases of herbicide resistance in waterhemp occurred in the 1990’s to acetolactate

synthase (ALS) inhibitors and Photosystem II (PSII) inhibitors (Heap 2023). In the 2000s,

resistance to PPO inhibitors and glyphosate was confirmed (Legleiter and Bradley 2008, Shoup et

al. 2003). Given the widespread reliance on glyphosate for weed control, growers returned to

utilizing soil-residual herbicides (also known as preemergence or PRE herbicide) and added tank

mix partners to glyphosate, or substituted other herbicides with glyphosate to control resistant

species. The primary substituted herbicides were glufosinate and PPO-inhibiting herbicides such

87
as lactofen and fomesafen (Green and Owen 2010, Legleiter et al. 2009, Takano and Dayan 2020,

Young 2006).

Despite an increased reliance on PPO-inhibiting herbicides in soybean production, PPO-

inhibitor resistance has been confirmed in waterhemp since 2001, and is common across soybean

growing areas in both waterhemp and Palmer amaranth (Amaranthus palmeri S. Wats) (Copeland

et al. 2018, Heap 2023, Nie et al. 2019, Shoup et al. 2003). Both species have evolved similar

mutations in the PPX2 gene, which encodes one of the target enzymes, protoporphyrinogen

oxidase II (PPO2) (Lermontova et al. 1997, Matringe et al. 1989). In waterhemp and Palmer

amaranth, a glycine deletion at the 210th position (ΔG210) and multiple substitutions at the R218

position were shown to confer resistance to lactofen and other PPO inhibitors (Giacomini et al.

2017, Nie et al. 2019, Patzoldt et al. 2006, Salas et al. 2016). Palmer amaranth has also evolved

resistance via a glycine substitution at the 399th position (G399A) (Rangani et al. 2019). Recently,

a Palmer amaranth population lacking target-site mutations demonstrated resistance to fomesafen

(Varanasi et al. 2018). The authors suggested that increased fomesafen detoxification was the

mechanism of resistance. Therefore, non-target-site resistance may be a future concern for the

long-term viability of PPO-inhibiting herbicides.

Despite widespread resistance to PPO inhibitors, several PPO-inhibiting herbicides,

particularly flumioxazin, sulfentrazone, saflufenacil, and fomesafen, are used for PRE residual

control of Amaranthus species on 10 to 21% of soybean hectares each, (NASS 2020). Soil-applied

PPO inhibitors are used for management of PPO-R populations, although at the cost of reduced

length of residual control and an increased frequency of resistant individuals that emerge through

the soil-residual herbicide layer (Falk et al. 2006, Umphres et al. 2018, Wuerffel et al. 2015c). A

common theory is that soil-applied PPO-inhibiting herbicides control resistant individuals via the

88
increased relative dose that germinating seeds encounter coupled with the relatively low level of

resistance conferred by the ΔG210 mutation. As the herbicide dissipates or degrades in the soil

profile, the segregating weed population is exposed to a discriminating dose of the PPO-inhibiting

herbicide, a dose that is lethal to a susceptible individual and non-injurious or at least non-lethal

to a resistant individual. For resistance selection to occur, the discriminating dose must occur in

the presence of an imbibing seed or emerging radicle, and there must not be a lethal dose of another

herbicide present in the soil as well. Thus, the increase in frequency of PPO-R waterhemp after a

soil-residual PPO-inhibitor application appears to be less than that of other weed species and

herbicide combinations for weed resistance.

Westrich (2022) and Boe (2019) assayed giant ragweed (Ambrosia trifida L.) and waterhemp,

respectively, for ALS-inhibitor resistance after germination through a soil-residual application of

ALS-inhibiting herbicides. The frequency of ALS resistance in survivors increased to nearly 100%,

whereas Wuerffel et al. (2015c) and Mansfield (2021) observed that the frequency of PPO-R

increased by only 10 to 30% after PPO-inhibitor application. The resistance selection was variable

by environmental conditions and did not always rise to statistical significance.

The differential response of various resistant biotypes after a herbicide application appears

to be dependent on the magnitude of resistance. Previous research suggests that different PPO-

inhibiting herbicides elicit different resistance magnitudes at the whole-plant and target enzyme

level (Lillie et al. 2020, Patzoldt et al. 2005, Porri et al. 2022, Rangani et al. 2019, Shoup et al.

2003, Wuerffel et al. 2015b). Therefore, different PPO-inhibiting herbicide active ingredients

could potentially select for resistant individuals more or less strongly than other herbicides.

Limited comparisons have been made between fomesafen and other residual PPO inhibitors, but

89
in the experiments, there were not enough treatments of other herbicides to draw firm conclusions

(Wuerffel et al. 2015c, Mansfield 2021)

Trifludimoxazin is a PPO-inhibiting herbicide under development by BASF that is shown to

have high levels of activity on PPO-resistant biotypes of waterhemp and Palmer amaranth in field

research, greenhouse research, and in vitro (Porri et al. 2022, Witschel et al. 2021, Steppig 2022).

Dose-response experiments on greenhouse grown waterhemp and Palmer amaranth with target-

site resistance (ΔG210 and R128G) demonstrated that foliar applications of trifludimoxazin to

waterhemp and Palmer amaranth populations did not result in a dose-response shift between

susceptible and resistant populations (Steppig et al. 2022). This result is potentially explained by

the maintained affinity of trifludimoxazin to the resistant PPO enzymes (Porri et al. 2022).

Trifludimoxazin in vitro has only a 7-fold difference in affinity (IC50) between ΔG210 and

susceptible PPO2 enzymes, whereas other PPO inhibitors tested had over 1000-fold difference in

affinity between susceptible and resistant PPO2 enzymes. Conversely, results from Chapter 2

demonstrated that soil applications of trifludimoxazin caused a 2.4 to 3.5-fold dose-response shift,

which was a similar response to that of flumioxazin in the very same populations of waterhemp

tested by Steppig (2022).

Given these apparently contradictory results, it is unclear if trifludimoxazin applied PRE or

POST will select for resistant individuals in the same way as do other PPO inhibitors. POST

applications are also relevant because resistance to foliar applications of trifludimoxazin in

soybean are envisioned for future commercialization (Witschel et al. 2021). Furthermore, little has

been published regarding how foliar applications of PPO inhibitors select for resistance.

Presumably, selection for resistance is quite high, but there are some nuances to consider. In order

for resistance selection to occur, the treated plants must be exposed to a discriminating dose.

90
However, several circumstances can arise that would lead to reduced selection. For example, if the

herbicide application is compromised by environmental conditions or weeds that are so large so as

to be uncontrollable, the S plants would be more likely to survive and reduce selection for

resistance. Conversely, a herbicide rate that is so high that it controls even the R individuals would

not select for resistance either. Control of low-level resistant individuals is commonly

recommended in that growers are advised to spray plants that are less than 10cm in height and with

the maximum labeled herbicide rate (Norsworthy et al. 2012). How PPO-resistant waterhemp

populations will respond to applications of trifludimoxazin alone or in combination with other

PPO inhibitors is unclear. Therefore, experiments were conducted to quantify trifludimoxazin

selection for the ΔG210 mutation in segregating field populations of waterhemp after soil and

foliar applications. The secondary objective was to investigate if combinations of trifludimoxazin

with other PPO inhibitors influences resistance selection.

3.3 Materials and Methods

3.3.1 Site Selection and Experimental Design

Field experiments were conducted in 2020 and 2021 to evaluate trifludimoxazin selection

for target-site resistance in field populations of waterhemp. The field sites were intentionally

chosen based on low frequency of PPO-R waterhemp at Throckmorton Purdue Agricultural Center

(Throckmorton) near Lafayette, IN (40.27N, 86.88W), and high frequency of PPO-R waterhemp

at Davis Purdue Agricultural Center (Davis) near Farmland, IN (40.26, 85.15W) based on previous

genotyping and experimental evidence (Mansfield 2021, Steppig 2022). Soil properties at each

location are described in Table 3.1.

91
A set of two independent experiments were conducted at each field location and year in

close proximity. One trial evaluated waterhemp emergence through PRE PPO-inhibiting

herbicides in soybean while the other evaluated waterhemp survival after a POST application of

PPO-inhibiting herbicides in a non-crop area. The two trials consisted of an identical set of

herbicide treatments; the only difference being that the PRE trials received a PPO-inhibiting

herbicide application PRE at soybean planting, while the POST trials received a PPO-inhibiting

herbicide with crop oil concentrate (Prime oil® Winfield Solutions, LLC, St. Paul, MN) added at

1% v/v applied POST when waterhemp was 10 to 20 cm in height.

Herbicide treatments consisted of trifludimoxazin (Tirexor®, BASF Corporation, Research

Triangle Park, NC) at 12.5 g ai ha-1, Saflufenacil (Sharpen®, BASF Corporation, Research

Triangle Park, NC) at 25 g ai ha-1, fomesafen (Flexstar®, Syngenta Crop Protection, LLC,

Greensboro, NC) at 263 g ai ha-1, as well as all two and three-way combinations of those herbicides

and a non-treated control. Application rates for the herbicides applied in mixture were the same as

the herbicides applied alone. Rates of saflufenacil and fomesafen were selected based on previous

research conducted by Wuerffel et al. (2015c) and Mansfield (2021). The rate of trifludimoxazin

was chosen because the anticipated commercial product will be formulated in a 2:1 ratio of

saflufenacil to trifludimoxazin (Findley et al. 2020). All herbicide treatments were applied to plots

measuring 3m in width by 9m in length using a 2m-handheld spray boom equipped with four

XR8002 nozzles (TeeJet® Spraying Systems, Wheaton, IL) propelled using compressed CO2 and

calibrated to deliver 140 L ha-1 at a pressure of 207 kPa.

Both trials received a burndown of paraquat (Gramoxone®, Syngenta Crop Protection LLC,

Greensboro,NC) in late May to remove unwanted vegetation including winter annual weeds and

early-emerging waterhemp. Immediately after burndown, PRE trials were planted with soybean

92
variety Stine 32EA12, featuring resistance to glufosinate and 2,4-D, (Enlist E3®, Corteva

Agriscience, Indianapolis, IN) at a rate of 340,000 seeds ha-1 in 76-cm row spacing. POST trials

were initiated in late June or early July when the subsequent cohort of emerged waterhemp reached

a targeted average height of 15 to 20 cm. The middle 2m of plot were sprayed, thus leaving an

untreated strip between plots.

3.3.2 Data Collection and Analysis

For PRE experiments, visual assessments of soil-residual control of waterhemp were

conducted weekly from 2 to 6 weeks after the PRE application using a 0 to 100% scale (100

corresponds to complete waterhemp control; 0 corresponds to no inhibition of waterhemp

emergence). Waterhemp emergence was also assessed by counting emerged plants in two quadrats

(0.25 m2 ) per plot at the time of waterhemp tissue collections. Waterhemp tissue was collected

once for each treatment at one of two collection timings for subsequent DNA extraction and RT-

qPCR assays for the ΔG210 mutation. (Table 3.2). The decision to collect tissue was made when

either 25 waterhemp plants per plot had emerged in a treatment or at the final evaluation of the

season, whichever occurred first. For tissue sampling, the largest and likely oldest 25 plants were

cut at the ground level from each plot, placed in a bag, then stored at -20C until DNA extraction

and subsequent RT-qPCR assays. The first plants to emerge were chosen because those are the

most likely to have experienced herbicide exposure. A POST application of glufosinate (Liberty®,

BASF Corporation, Research Triangle Park, NC) at 660 g ai ha-1 was made immediately after

waterhemp tissue collection in the PRE trials to remove emerged weeds in an attempt to assess a

second cohort of waterhemp emergence. Lack of rainfall in June and July in both trial years

coupled with soybean canopy closure, resulted in an inadequate second emergence cohort of

waterhemp. Therefore, only data from the single collection timing was collected, but later

93
evaluation timings were affected by the glufosinate application. The only affected data presented

in this manuscript is waterhemp emergence at the collection 2 timing. In all site years, treatments

of no herbicide, trifludimoxazin, saflufenacil, and trifludimoxazin+saflufencacil received POST

glufosinate, so the presented % emergence is reduced from collection 1 to collection 2. Because

the vegetation was removed, emergence % values are calculated based on emergence in the no

herbicide treatment at collection 1.

Visual assessments of waterhemp control were taken at 7, 14, and 21 days after POST

treatments using a similar 0 to 100% scale. Waterhemp survival following POST treatments was

quantified by counting surviving waterhemp plants in two 0.25m2 quadrats per plot at 14 days after

herbicide treatment (DAT). Waterhemp tissue was collected at 14 to 16 DAT at which time a total

of 25 surviving plants per plot were clipped at the soil surface, placed in a plastic bag, and stored

at -20C until DNA extraction and RT-qPCR assays. To reduce bias of a particular plant size or age,

plants were randomly selected without regard for size or injury at the time of collection.

Waterhemp DNA was extracted using a modified cetyltrimethylammonium bromide

(CTAB) protocol originally designed by Saghai-Maroof et al. (1984). A TaqMan® SNP genotyping

assay developed by Wuerffel et al. (2015a), modified by Giacomini et al. (2017), and sourced from

ABI (Applied Biosystems Inc. Grand Island, NY) was used to genotype each sample as

homozygous susceptible, homozygous resistant, or heterozygous based on fluorescence of specific

fluorophores in the presence of a PCR product generated with forward primer (5′-

TGATTATGTTATTGACCCTTTTGTTGCG) and reverse primer (5′ GAGGGAGTATAA

TTTATTTACAACCTCCAGAA). For each sample, a 10-µl reaction was prepared with 4.8 µl of

nuclease free nano-pure water, 2 µl of GoTaq Flexi buffer, 1.2 µl of 25 mM MgCl2, 0.4 µl of 10

mM dNTP, 0.5 µl of 20X primers and TaqMan® probes, 0.1 µl of GoTaq Flexi polymerase (5 U

94
µl−1), and 1 µl of genomic DNA. Reactions were amplified using a CFX384 RT-PCR detection

system (Bio-Rad Laboratories, Hercules, CA 94547) with the following cycle conditions: 2 min at

95 C; 39 cycles of 95 C for 15 s and 60 C for 1 min; followed by a plate read after every cycle.

Experiments were arranged in a randomized complete block design with four replications.

All data were subjected to analysis of variance (ANOVA) using PROC GLIMMIX in SAS 9.4

(SAS Institute; Cary, NC) using a one-way ANOVA for herbicide treatment with other model

fixed effects of location, and year to determine experimental interactions. Data were square root

or arcsin square root transformed when appropriate to better conform to model assumptions and

combined over site, and/or years when the effects did not interact with herbicide treatment.

Transformed means were separated using Tukey’s honestly significant difference (HSD) (α = 0.05)

and presented as non-transformed means for ease of interpretation.

3.4 Results and Discussion

3.4.1 PRE Herbicide Efficacy

All site years received an activating rainfall within 6 d of herbicide application (Table 3.3).

Waterhemp control at 35 DAT is presented by year and pooled over site due to a significant

treatment by year interaction (Table 3.4). In 2020, a dry period occurring after trial initiation and

herbicide activation (7 to 28 DAT) resulted in reduced waterhemp emergence and extended

residual control with the exception of trifludimoxazin alone at both locations. All of the herbicide

treatments except trifludimoxazin provided 83 to 98% control of waterhemp. Trifludimoxazin

resulted in 52% control of waterhemp. In 2021, greater amounts of activating rainfall occurring

during days 0 to 6, and rainfall events occurring between 21 and 27 DAT at both locations resulted

in reduced waterhemp control for both saflufenacil and trifludimoxazin alone in comparison to

95
fomesafen and herbicide mixture treatments. In contrast, fomesafen and combinations of

fomesafen with trifludimoxazin and saflufenacil provided greater control. Due to rate selection,

fomesafen containing treatments provided the greatest length of residual control.

Waterhemp emergence data showed similar trends as control data (Table 3.5). The number

of emerged waterhemp after trifludimoxazin treatment in all site years and saflufenacil at the Davis

location was not reduced in comparison to no herbicide treatment. Waterhemp emergence was the

lowest (0 to 19% of non-treated) at all site years for fomesafen herbicide combinations. Even at

later evaluation timings, waterhemp emergence in fomesafen combination treatments remained

low at 1 to 20% emergence of the non-treated (Table 3.5).

3.4.2 PRE Resistance Selection

The field locations had different resistance selection intensities. At the Davis location, all

treatments with the exception of trifludimoxazin, increased the frequency of ΔG210 from 34% in

the untreated plot to between 58 and 70% (Table 3.6). The frequency of resistance after

trifludimoxazin application was an intermediate level similar to both the no herbicide treatment

and the other single or two-way combination herbicide treatments. Conversely, at the

Throckmorton location, only saflufenacil applied alone increased the frequency of resistance from

18% in the no herbicide treatment to 48%. These results are similar to Wuerffel et al. (2015c) in

that resistance selection after PPO-inhibitor application does not always increase.

Trifludimoxazin, fomesafen, and saflufenacil+fomesafen had similar frequencies of

heterozygous individuals compared to the non-treated while saflufenacil, trifludimoxazin+

saflufenacil, trifludimoxazin+fomesafen, and trifludimoxazin+saflufenacil+fomesafen increased

the frequency of heterozygous individuals by 15% to 21% (Table 3.7). In 2021, saflufenacil,

fomesafen, and saflufenacil+fomesafen increased the frequency of homozygous resistant

96
genotype/total samples by 11% to 20 %. In 2020, the increase was not significant for any treatment.

An increase in homozygous and heterozygous individuals with ΔG210 out of total collected plants

was expected because they are the two component parts of the overall frequency of resistance.

Out of the total number of resistant plants, the ratio of homozygous individuals/total

resistant individuals for fomesafen and saflufenacil+fomesafen increased from 6% in the non-

treated to 37% and 38%, respectively, in 2021 (Table 3.7). Other herbicide treatments were

intermediate, but statistically similar to both the no herbicide treatment and the fomesafen and

fomesafen+saflufenacil treatments, ranging from 14% to 21%. In 2020, there was no significant

change in homozygous waterhemp frequency between herbicide treated and non-treated treatments.

An increase in homozygous/total resistant ratio likely occurred for a couple of reasons. The

first is that PPO resistance via the ΔG210 mutation is an incompletely dominant trait (Patzoldt et

al. 2006). Incomplete dominance in the context of herbicide resistance means that a heterozygous

individual has a resistant phenotype, but the homozygous individual has a more resilient phenotype.

The inconsistent selection for more homozygous individuals is likely from the exposure of

germinating seedlings to a discriminating dose that still controls heterozygous individuals but does

not control homozygous individuals. The inconsistency is expected, because the selection for

overall resistance is inconsistent and relatively weak, so the selection for homozygous resistant

plants would be even more so. Another potential reason is the observation that not all plants

carrying the PPO-resistance mutation necessarily have a resistant phenotype (Wu et al. 2020).

Such an observation may be part of the incompletely dominant nature of the allele or perhaps there

are uncharacterized differences in pathway expression patterns or interactions of PPO1 and PPO2

target enzymes contributing to the phenomenon (Dayan et al. 2018, Watanabe et al. 2001)

97
While selection for resistance was inconsistent even for labeled PPO inhibitors,

trifludimoxazin generally did not increase the frequency of resistant individuals or any of the

genotype ratios. The lack of resistance selection is consistent with previous research that shows

that trifludimoxazin has a smaller resistance ratio than other PPO inhibitors at the target site and

whole plant level (Porri et al. 2022, Steppig 2022).

Some potential considerations that should be explored given the experimental conditions

and rate selection is whether the trifludimoxazin was adequately activated or completely washed

out of the soil profile at herbicide activation to where germinating waterhemp was not exposed at

all, or secondly, whether there was a consistent frequency of resistance emerging throughout the

season. We find the lack of exposure to trifludimoxazin to be unlikely for several reasons. The

first is that at all site years, there was at least 14 d of waterhemp control, indicating that emergence

in the treated area was slowed at least for some period of time (data not shown). The second is the

comparison with saflufenacil alone. In 2021, saflufenacil provided less control than

trifludimoxazin yet selected for an increased frequency of resistance. Finally, there is little

apparent association with waterhemp control and resistance selection. Overall resistance selection

was influenced by location interactions, whereas waterhemp control was influenced by year

interactions. Even where genotypic frequencies were influenced by year interactions, there was no

pattern of length of residual control and resistance selection.

Our other consideration was the resistance frequency of emerging waterhemp throughout

the growing season. Resistant individuals can potentially be more or less likely to emerge at

different times of the season either from independent selection or pleiotropic effects of the

resistance mutation (Kremer and Lotz 1998, Owen et al. 2011, Shergill et al. 2015). If differential

emergence of resistant individuals were occurring, then the interpretation of the entire experiment

98
would be severely complicated. We tested this possibility in 2021 by collecting newly emerged

waterhemp every 2 weeks in an area of the nontreated plots. There was no pattern of differential

emergence of resistant and susceptible individuals throughout the season (data not shown). While

differential emergence and dormancy of waterhemp and Palmer amaranth can vary by adapted

geography and management practice, there is no documented differences in dormancy or

emergence associated with herbicide resistance in these species (Jha and Norsworthy 2009,

Sosnoskie et al. 2013, Wu and Owen 2014).

We conclude that PRE applications of trifludimoxazin do not select for ΔG210 waterhemp

as much as do other PPO inhibitors. The lack of selection likely comes from a combination of the

physical properties of trifludimoxazin in soil and the lesser magnitude of resistance to

trifludimoxazin in comparison to other PPO inhibitors (Table 3.8) (Asher et al. 2021, Porri et al.

2022). Since trifludimoxazin is used at a lower rate than other PPO inhibitors, and is relatively

hydrophobic, the trifludimoxazin may become unavailable, i.e. adsorbed to soil, more readily at

low soil volumetric water content than more water soluble herbicides such as fomesafen and

saflufenacil. Future research should investigate the influence of herbicide properties on availability

in soil solution and influence on the herbicide dose that weed seedlings receive.

Our second objective, evaluating if trifludimoxazin in combination with other PPO

inhibitors could reduce resistance selection, could not be fully tested. Because the residual control

provided by fomesafen was much greater than that of trifludimoxazin and saflufenacil, the latter

two herbicides were likely not available in the soil any longer and did not contribute to waterhemp

control. Wuerffel et al. (2015c) observed the same shortcoming with combinations of S-

metolochlor and fomesafen because of the equal or lesser persistence of S-metolachlor at the

chosen rate ratio. However, since trifludimoxazin and saflufenacil had similar residual control, the

99
treatment of trifludimoxazin+saflufenacil can allow a partial evaluation of the objective. We

observed that the addition of trifludimoxazin to saflufenacil does not reduce selection for ΔG210

waterhemp. To reduce selection for resistance, the alternative herbicide must have the same or

greater length of availability in the soil (Wrubel and Gressel 1994). Our results, as well as

Mansfield (2021), demonstrate that an alternative herbicide does not alleviate resistance selection

despite having equal or greater available dose in the soil. Perhaps a hormetic effect on germination

is stimulating germination of resistant individuals at non-lethal rates and, therefore, increasing the

frequency of resistant individuals. Hormetic effects of herbicides have been documented for PPO-

inhibitor applications and other herbicides, but it is unclear truly what is happening at the seed

germination and physiological level (Chapter 2, Chandran et al. 1999, Goggin and Powles 2014,

Wuerffel et al. 2015b).

3.4.3 POST Efficacy

Waterhemp control was pooled across sight years and waterhemp survival analyzed

separately by location (Table 3.9). The location interaction likely occurred from a higher baseline

frequency of resistance leading to greater waterhemp survival. While there were similar trends in

waterhemp control at all sight years, the control ratings are based on both waterhemp injury and

waterhemp death. Trifludimoxazin and trifludimoxazin combinations with other herbicides

provided 79% to 87% control of waterhemp. Which was greater than saflufenacil and fomesafen

alone which provided only 45% to 56% control of waterhemp. Similar trends were present in

waterhemp survival data at the Throckmorton site with 13% to 22% waterhemp survival compared

to as much as 52% to 65% survival after fomesafen and saflufenacil application. The Davis

location had no differences in waterhemp survival between any of the herbicide treatments. In

100
contrast to PRE experiments, trifludimoxazin-containing treatments had the greatest efficacy when

applied POST, whereas trifludimoxazin residual activity was lower than that of other herbicides.

3.4.4 POST Resistance Selection

For overall resistance selection (Table 3.10), all treatments, with the exception of

saflufenacil in 2020, increased the frequency of ΔG210 waterhemp by at least 31% in 2020 and

39% in 2021. In 2020, the frequency of ΔG210 for trifludimoxazin was not different than in other

treatments. In 2021, treatments containing saflufenacil selected for a greater frequency of ΔG210

than trifludimoxazin treatment by at least 21%. The cause of such a greater resistance selection in

2021, especially for saflufenacil, is likely because of less herbicide efficacy for all treatments

observed in 2020. Less overall efficacy means that all plants are more likely to survive, including

susceptible plants, thus diluting the resistance selection effects.

For each individual genotype, there were different interaction effects that are difficult to

interpret. Heterozygous/total samples had a three-way interaction of location, year, and herbicide

treatment (Table 3.11), but homozygous/ total samples and homozygous/ total resistant only had a

two-way interaction of location and treatment. The location interaction for homozygous ratios is

possibly a result of the existing frequency of resistance. If the baseline allele frequency is not high

enough, it is possible that there were insufficient numbers of homozygous resistant individuals to

get consistent, powerful, and interpretable results. According to Hardy Weinberg equilibrium

calculations and observed genotypic ratios of untreated plots at the Throckmorton location, only

1% of plants were homozygous for ΔG210. Therefore, a selection event is more likely to result in

differences. The three-way interaction for the heterozygous genotype ratio likely comes from the

combined effects of year that influenced overall resistance selection, and the location effect that

influenced the homozygous ratios.

101
There were no differences in heterozygous frequency for any herbicide treatments with the

exception of Davis in 2021 where saflufenacil and saflufenacil+trifludimoxazin selected for a

greater percentage of heterozygous waterhemp than trifludimoxazin alone. Trifludimoxazin

increased the percentage of heterozygous waterhemp only at the Throckmorton site in 2021.

For frequency of homozygous/ total samples, at the Davis location, only fomesafen increased

the percentage of homozygous individuals from 6% in the non-treated to 24% after treatment with

fomesafen. At the Throckmorton location, treatments containing trifludimoxazin and

trifludimoxazin+saflufenacil increased the frequency of homozygous individuals from 1% in the

non-treated, to 15 to 20% after herbicide treatment. This increase caused a nearly identical increase

in the ratio of homozygous/total resistant plants.

In contrast to the PRE experiments, trifludimoxazin did cause significant selection for

ΔG210 in waterhemp when applied POST. However, there is some evidence that the selection is

weaker than that caused by other PPO inhibitors. Primarily, the 2021 trial year showed that

trifludimoxazin selected for ΔG210 waterhemp less than saflufenacil and saflufenacil

combinations. Also, the frequency of heterozygous individuals did not significantly increase after

trifludimoxazin application in three out of four site years, but there was an observed increase in

the frequency of homozygous resistant waterhemp. Similarly to what was previously discussed

with PRE activity, an increase in homozygous individuals is consistent with an incompletely

dominant trait (Patzoldt et al. 2006). Perhaps only homozygous individuals had a strong enough

phenotype to confer resistance to trifludimoxazin. The observation that trifludimoxazin selects for

resistance is inconsistent with Steppig et al. (2022), who observed that target-site mutations in

waterhemp and Palmer amaranth did not cause a dose-response shift. A potential explanation for

102
this discrepancy is that there is a true dose-response shift, but the greenhouse and plant growth

conditions as well as dose-response modeling failed to detect the small effect.

While combination with other PPO inhibitors can increase efficacy and reduce the number

of surviving plants, the frequency of resistant individuals is similar for PPO-inhibitor combination

treatments with or without trifludimoxazin (Table 3.10). Since trifludimoxazin is so much more

effective than other herbicides and it still selects for increased resistance, the marginal increase in

control from combination with other PPO inhibitors is a positive attribute at minimal cost.

Potential limitations to the current dataset are the potential presence of other target-site

resistance alleles, or other small effect genes that could co-select with herbicide resistance. While

we did not directly assay for the presence of R128 or G399A alleles, we find it extremely unlikely

that they are present in any frequency that would influence our results. First, a 2016 survey of

Midwest U.S. waterhemp and other internal experiments using the subject populations have failed

to detect any target-site resistance allele other than ΔG210 (Nie et al. 2019). Secondly, our results

in both site years are consistent with previous research about resistance selection after PPO-

inhibitor application (Wuerffel et al. 2015c). There were no site years, treatments, or plots that

indicated any anomaly in control or unexplained resistance.

However, we cannot rule out the presence of secondary-partial resistance mechanisms that

may be present. With the comparatively small resistance selection that occurs from PPO-inhibitor

application PRE, trifludimoxazin and other soil-applied PPO inhibitors are at increased risk of

developing a new resistance mechanism or a partial secondary resistance mechanism that will

increase its ability to emerge through a herbicide layer. Non-target-site resistance has already been

reported in Palmer amaranth (Varanasi et al. 2018). Furthermore, the most recent cases of herbicide

resistance in waterhemp are to herbicides with heavy soil-applied usage i.e. HPPD inhibitors and

103
VLCFA inhibitors via herbicide detoxification (Ma et al. 2013, Murphy et al. 2021, Strom et al.

2020). While trifludimoxazin has unique structural properties that allow it to bind to PPO2 (Porri

et al. 2022) and overcome resistance mutations, there are still many structural moieties that could

potentially be avenues for cross resistance, particularly with flumioxazin and saflufenacil because

of their structural similarity (Dimaano et al. 2020, Jugulam and Shyam 2019).

Overall, we conclude that trifludimoxazin exerts a selection pressure for ΔG210 waterhemp

that is less than other PPO inhibitors, but still potentially meaningful for the spread of the ΔG210

mutation. While co-application of trifludimoxazin with other PPO inhibitors can increase weed

control both PRE and POST, the presence of another PPO-inhibiting herbicide may cause further

selection for resistance. Therefore, trifludimoxazin can be used to increase waterhemp control and

reduce the amount of waterhemp survivorship. However, management of escaped waterhemp

plants with effective POST herbicide options or other tactics must be available and implemented

to reduce the shift toward greater ΔG210 frequencies in the weed populations.

3.5 Acknowledgements

This research received no specific grant from any funding agency, commercial or not-for-

profit sectors. No conflicts of interest have been declared.

3.6 Literature Cited

Asher BS, Dotray PA, Liebl RA, Keeling JW, Ritchie GD, Udeigwe TK, Reed JD, Keller KE,
Bowe SJ, Aldridge RB, Simon A (2021) Vertical mobility and cotton tolerance to
trifludimoxazin, a new protoporphyrinogen oxidase-inhibiting herbicide, in three West Texas
soils. Weed Technol 35:144–148

104
Boe JE (2019) Establishing the value of ALS-inhibiting herbicides in fields with confirmed weed
resistance to ALS-inhibiting herbicides. MS thesis. West Lafayette IN: Purdue University.
190 p

Chandran RS, Singh M, Salihu S (1999) Thiazopyr stimulates hairy beggarticks (Bidens pilosa)
germination. Weed Technol 13:576–580

Copeland JD, Giacomini DA, Tranel PJ, Montgomery GB, Steckel LE (2018) Distribution of
PPX2 mutations conferring PPO-inhibitor resistance in Palmer amaranth populations of
Tennessee. Weed Technol 32:592–596

Dayan FE, Barker A, Tranel PJ (2018) Origins and structure of chloroplastic and mitochondrial
plant protoporphyrinogen oxidases: implications for the evolution of herbicide resistance.
Pest Manag Sci 74:2226–2234

Dimaano NG, Yamaguchi T, Fukunishi K, Tominaga T, Iwakami S (2020) Functional


characterization of cytochrome P450 CYP81A subfamily to disclose the pattern of cross-
resistance in Echinochloa phyllopogon. Plant Mol Biol 102:403–416

Falk JS, Shoup DE, Al-Khatib K, Peterson DE (2006) Protox-resistant common waterhemp
(Amaranthus rudis) response to herbicides applied at different growth stages. Weed Sci
54:793–799

Findley D, Youmans C, Bowe S (2020) Tirexor (trifludimoxazin): next generation burndown


update. in Proceedings of the 2020 Weed Science Society of America Annual Meeting. Maui,
HI: Weed Science Society of America

Giacomini DA, Umphres AM, Nie H, Mueller TC, Steckel LE, Young BG, Scott RC, Tranel PJ
(2017) Two new PPX2 mutations associated with resistance to PPO-inhibiting herbicides in
Amaranthus palmeri. Pest Manag Sci 73:1559–1563

Goggin DE, Powles SB (2014) Fluridone: a combination germination stimulant and herbicide for
problem fields? Pest Manag Sci 70:1418–1424

Green JM, Owen MDK (2010) Herbicide-resistant crops: Utilities and limitations for herbicide-
resistant weed management. J Agric Food Chem 59:5819–5829

105
Heap I (2023) International survey of herbicide resistant weeds. http://www.weedscience.org.
Accessed March 25, 2023

Heneghan JM, Johnson WG (2017) The growth and development of five waterhemp (Amaranthus
tuberculatus) populations in a common garden. Weed Sci 65:247–255

Jha P, Norsworthy JK (2009) Soybean canopy and tillage effects on emergence of Palmer amaranth
(Amaranthus palmeri) from a natural seed bank. Weed Sci 57:644–651

Jugulam M, Shyam C (2019) Non-target-site resistance to herbicides: recent developments. Plants


8:417

Kremer E, Lotz LAP (1998) Germination and emergence characteristics of triazine-susceptible


and triazine-resistant biotypes of Solanum nigrum. J Appl Ecol 35:302–310

Legleiter TR, Bradley KW (2008) Glyphosate and multiple herbicide resistance in common
waterhemp (Amaranthus rudis) populations from Missouri. Weed Sci 56:582–587

Legleiter TR, Bradley KW, Massey RE (2009) Glyphosate-resistant waterhemp (Amaranthus


rudis) control and economic returns with herbicide programs in soybean. Weed Technol
23:54–61

Lermontova I, Kruse E, Mock HP, Grimm B (1997) Cloning and characterization of a plastidal
and a mitochondrial isoform of tobacco protoporphyrinogen IX oxidase. Proc Natl Acad Sci
U S A 94:8895–8900

Lillie KJ, Giacomini DA, Tranel PJ (2020) Comparing responses of sensitive and resistant
populations of Palmer amaranth (Amaranthus palmeri) and waterhemp (Amaranthus
tuberculatus var. rudis) to PPO inhibitors. Weed Technol. 34: 140–146. doi:
10.1017/wet.2019.84

Ma R, Kaundun SS, Tranel PJ, Riggins CW, McGinness DL, Hager AG, Hawkes T, Mc EI,
Riechers DE (2013) Distinct detoxification mechanisms confer resistance to mesotrione and
atrazine in a population of waterhemp. Plant Physiol 163:363–377

Mansfield BC (2021) Characterization of protoporphyrinogen oxidase (PPO) herbicide resistance


in tall waterhemp (Amaranthus tuberculatus). MS thesis. West Lafayette IN: Purdue
University. 158 p

106
Matringe M, Camadro JM, Labbe P, Scalla R (1989) Protoporphyrinogen oxidase as a molecular
target for diphenyl ether herbicides. Biochem J 260:231–235

Murphy BP, Beffa R, Tranel PJ (2021) Genetic architecture underlying HPPD-inhibitor resistance
in a Nebraska Amaranthus tuberculatus population. Pest Manag Sci 77:4884–4891

Nie H, Mansfield BC, Harre NT, Young JM, Steppig NR, Young BG (2019) Investigating target‐
site resistance mechanism to the PPO‐inhibiting herbicide fomesafen in waterhemp and
interspecific hybridization of Amaranthus species using next generation sequencing. Pest
Manag Sci 75:3235–3244

Norsworthy JK, Ward SM, Shaw DR, Llewellyn RS, Nichols RL, Webster TM, Bradley KW,
Frisvold G, Powles SB, Burgos NR, Witt WW, Barrett M (2012) Reducing the risks of
herbicide resistance: best management practices and recommendations. Weed Sci 60:31–62

Owen MJ, Michael PJ, Renton M, Steadman KJ, Powles SB (2011) Towards large-scale prediction
of Lolium rigidum emergence. II. correlation between dormancy and herbicide resistance
levels suggests an impact of cropping systems. Weed Res 51:133–141

Papiernik SK, Koskinen WC, Barber BL (2012) Low Sorption and fast dissipation of the herbicide
saflufenacil in surface soils and subsoils of an eroded prairie landscape. J Agric Food Chem
60:10936–10941

Patzoldt WL, Hager AG, McCormick JS, Tranel PJ (2006) A codon deletion confers resistance to
herbicides inhibiting protoporphyrinogen oxidase. Proc Natl Acad Sci U S A 103:12329–34

Patzoldt WL, Tranel PJ, Hager AG (2005) A waterhemp (Amaranthus tuberculatus) biotype with
multiple resistance across three herbicide sites of action. Weed Sci 53:30–36

Porri A, Betz M, Seebruck K, Knapp M, Johnen P, Witschel M, Aponte R, Liebl R, Tranel PJ,
Lerchl J (2022) Inhibition profile of trifludimoxazin towards PPO2 target site mutations. Pest
Manag Sci 79:507–519

Rangani G, Salas-Perez RA, Aponte RA, Knapp M, Craig IR, Mietzner T, Langaro AC, Noguera
MM, Porri A, Roma-Burgos N (2019) A novel single-site mutation in the catalytic domain of
protoporphyrinogen oxidase IX (PPO) Confers Resistance to PPO-inhibiting herbicides.
Front Plant Sci 10:568

107
Saghai-Maroof MA, Soliman KM, Jorgensen RA, Allard RW (1984) Ribosomal DNA spacer-
length polymorphisms in barley: mendelian inheritance, chromosomal location, and
population dynamics. Proc Natl Acad Sci U S A 81:8014–8018

Salas RA, Burgos NR, Tranel PJ, Singh S, Glasgow L, Scott RC, Nichols RL (2016) Resistance
to PPO-inhibiting herbicide in Palmer amaranth from Arkansas. Pest Manag Sci 72:864–869

Schultz JL, Chatham LA, Riggins CW, Tranel PJ, Bradley KW (2015) Distribution of herbicide
resistances and molecular mechanisms conferring resistance in Missouri waterhemp
(Amaranthus rudis Sauer) populations. Weed Sci 63:336–345

Shaner DL, ed (2014) Herbicide Handbook. 10th edn. Lawrence, KS: Weed Science Society of
America. 513 p

Shergill LS, Fleet B, Preston C, Gill G (2015) Incidence of herbicide resistance, seedling
emergence, and seed persistence of smooth barley (Hordeum glaucum) in South Australia.
Weed Technol 29:782–792

Shoup DE, Al-Khatib K, Peterson DE (2003) Common waterhemp (Amaranthus rudis) resistance
to protoporphyrinogen oxidase-inhibiting herbicides. Weed Sci 51:145–150

Sosnoskie LM, Webster TM, Culpepper AS (2013) Glyphosate resistance does not affect Palmer
amaranth (Amaranthus palmeri) seedbank longevity. Weed Sci 61:283–288

Steppig NR (2022) Evaluation of trifludimoxazin, a new protoporphyrinogen oxidase- inhibiting


herbicide, for use in soybean. Ph.D dissertation. West Lafayette, IN: Purdue University. 172p

Strom SA, Hager AG, Seiter NJ, Davis AS, Riechers DE (2020) Metabolic resistance to S-
metolachlor in two waterhemp (Amaranthus tuberculatus) populations from Illinois, USA.
Pest Manag Sci 76:3139–3148

Takano HK, Dayan FE (2020) Glufosinate‐ammonium: a review of the current state of knowledge.
Pest Manag Sci 76:3911–3925

Umphres AM, Steckel LE, Mueller TC (2018) Control of protoporphyrinogen oxidase inhibiting
herbicide resistant and susceptible Palmer amaranth (Amaranthus palmeri ) with soil-applied
protoporphyrinogen oxidase–inhibiting herbicides. Weed Technol 32:95–100

108
[USDA-NASS] US Department of Agriculture, National Agricultural Statistics Service (2020)
Agricultural Chemical Use Program.
https://www.nass.usda.gov/Surveys/Guide_to_NASS_Surveys/Chemical_Use/. Accessed:
January 15, 2022

Varanasi VK, Brabham C, Norsworthy JK (2018) Confirmation and characterization of non-target


site resistance to fomesafen in Palmer amaranth (Amaranthus palmeri). Weed Sci 66:702–
709

Watanabe N, Che FS, Iwano M, Takayama S, Yoshida S, Isogai A (2001) Dual targeting of spinach
protoporphyrinogen oxidase II to mitochondria and chloroplasts by alternative use of two in-
frame initiation codons. J Biol Chem 276:20474–20481

Westrich BC (2022) Influence of mesotrione, ALS-inhibitor resistance, and self incompatability


on giant ragweed management in soybean. Ph.D dissertation. West Lafayette, IN: Purdue
University. 189 p

Witschel M, Aponte R, Armel G, Bowerman P, Mietzner T, Newton T, Porri A, Simon A, Seitz T


(2021) Tirexor®—design of a new resistance breaking protoporphyrinogen IX oxidase
inhibitor. Recent Highlights Discov Optim Crop Prot Prod:501–509

Wrubel RP, Gressel J (1994) Are herbicide mixtures useful for delaying the rapid evolution of
resistance? a case study. Weed Technol 8:635–648

Wu C, Goldsmith MR, Pawlak J, Feng P, Smith S, Navarro S, Perez-Jones A (2020) Differences


in efficacy, resistance mechanism and target protein interaction between two PPO inhibitors
in Palmer amaranth (Amaranthus palmeri). Weed Sci 68:105–115

Wu C, Owen MDK (2014) When is the best time to emerge: reproductive phenology and success
of natural common waterhemp (Amaranthus rudis) cohorts in the Midwest United States?
Weed Sci 62:107–117

Wuerffel RJ, Young JM, Lee RM, Tranel PJ, Lightfoot DA, Young BG (2015a) Distribution of
the ΔG210 protoporphyrinogen oxidase mutation in Illinois waterhemp (Amaranthus
tuberculatus) and an improved molecular method for detection. Weed Sci 63:839–845

109
Wuerffel RJ, Young JM, Matthews JL, Young BG (2015b) Characterization of PPO-inhibitor–
resistant waterhemp (Amaranthus tuberculatus) response to soil-applied PPO-inhibiting
herbicides. Weed Sci 63:511–521

Wuerffel RJ, Young JM, Tranel PJ, Young BG (2015c) Soil-residual protoporphyrinogen oxidase–
inhibiting herbicides influence the frequency of associated resistance in waterhemp
(Amaranthus tuberculatus). Weed Sci 63:529–538

Young BG (2006) Changes in herbicide use patterns and production practices resulting from
glyphosate-resistant crops. Weed Technol 20:301–307

110
Table 3.1. Site characteristics for field trials conducted in 2020 and 2021a.

Soil Properties

Locationb Sand Silt Clay Texture OM pH CEC

----------%---------- % mEq 100 g-1 soil

Davis 15 42 43 SiC 3.5 6.3 16.0

Throckmorton 41 38 21 L 2.3 6.5 14.3


a
Abbreviations: CEC, cation exchange capacity; Davis, Davis Purdue Agriculture Center; L,
loam; Throckmorton, Throckmorton Purdue Agricultural Center; OM, organic matter; SiC, silty
clay.
b
GPS coordinates for field locations: Davis (40.25N, -86.88W) and Throckmorton (40.29N,
-86.90W)

111
Table 3.2. Trial initiation and tissue collection dates for PRE and POST field trials at Throckmorton and Davis field sites in 2020 and 2021.
Collection 1a Collection 2b

Date of Days since Date of Days since

Trial Year Site Application date collection application collection application

PREc 2020 Throckmorton 5/26/2020 7/3/2020 38 7/31/2020 66

Davis 5/26/2020 7/9/2020 44 7/31/2020 66

2021 Throckmorton 5/24/2021 6/28/2021 35 7/21/2021 59

Davis 5/25/2021 6/29/2021 36 7/13/2021 49


112

POST 2020 Throckmorton 7/2/2020 7/15/2020 14

Davis 7/4/2020 7/17/2020 14

2021 Throckmorton 6/17/2021 7/3/2021 16

Davis 6/22/2021 7/6/2021 14


a
Collection 1 was triggered when multiple treatments had at least 25 plants per plot large enough for tissue collection for PRE trials and at
14 days after treatment for POST trials.
b
Collection 2 occurred when the remainder of treatments had 25 plants per plot large enough for tissue collection or at the last evaluation
of the season; whichever occurred first.
c
Glufosinate was sprayed after tissue was collected for PRE treatments which removed existing vegetation and affected emergence data
counted at the time of collection 2.
Table 3.3. Weekly rainfall totals for Davis and Throckmorton field locations in 2020 and 2021.

Throckmorton Davis

2020 2021 2020 2021

Rainfall Interval Precipitation

DATab ------------------------cm------------------------

-7 to -1 0.86 1.19 0.91 0.03

0 to 6 0.36 2.69 1.65 3.35

7 to 13 0.33 0.18 0.08 0.36

14 to 20 0.79 0.28 0.38 0.43

21 to 27 0.81 1.22 0.66 2.24

28 to 34 3.28 8.05 3.99 0.13

35 to 41 1.83 1.17 3.68 2.36

42 to 48 1.62 3.73 1.50 3.66

49 to 55 0.63 5.21 0.00 3.23

56 to 62 3.23 0.00 7.57 0.66

a
Abbreviations: DAT, days after treatment.
b
Numbering begins (day 0) on the day of PRE herbicide application for each site year

113
Table 3.4. Control of waterhemp recorded 35 days after PRE herbicide application at
Throckmorton and Davis field sites in 2020 and 2021.
Waterhemp control 35 d after applicationa

Treatment 2020 2021

----------------%---------------

Trifludimoxazin 52 d 58 c

Saflufenacil 83 c 41 c

Fomesafen 91 abc 89 ab

Trifludimoxazin+Saflufenacil 86 bc 83 b

Triflufimoxazin+Fomesafen 96 ab 91 ab

Saflufenacil+Fomesafen 98 a 95 ab

Trifludimoxazin+Saflufenacil+Fomesafen 97 ab 97 a

No Herbicide 0- 0-

a
Means within a column followed by the same letter are not significantly different according to
Tukey’s HSD test at α=0.05.

114
Table 3.5. Waterhemp emergence at the time of each tissue collection following PRE herbicide treatments at Throckmorton and Davis Field Sites
in 2020 and 2021.
Waterhemp emergence recorded at collection 1a Waterhemp emergence recorded at collection 2b

Davis Throckmorton Davis Throckmorton

Treatmentc 2020 2021 2020 2021 2020 2021 2020 2021

-----------------------------------------------------------%-------------------------------------------------------------

Trifludimoxazin 117 a 63 ab 41 ab 42 ab 0b 5 ab 0b 21 ab

Saflufenacil 69 a 57 abc 11 bc 32 bc 0b 4 ab 1 ab 25 ab

Fomesafen 16 b 13 cd 12 bc 1d 22 a 0b 9a 1c
115

Tri+Saf 166 a 30 bcd 11 bc 14 bcd 0b 7 ab 11 a 41 a

Tri+Fom 19 b 8d 2c 4 cd 15 a 16 a 5 ab 7 bc

Saf+Fom 18 b 7d 0c 3 cd 15 a 16 a 2 ab 2c

Tri+Saf+Fom 18 b 3d 1c 2d 20 a 14 a 3 ab 1c

No Herbicide 100 a 100 a 100 a 100 a 0b 7 ab 0b 10 abc

a
Means within a column followed by the same letter are not significantly different according to Tukey’s HSD test at α=0.05.
b
Treatments that were collected at the first timing were sprayed with glufosinate to remove existing vegetation including the no herbicide control.
Percentages calculated for collection 2 timing are based on emergence in the no herbicide control observed at collection 1.
c
Abbreviations: Tri, trifludimoxazin; Saf, saflufenacil; Fom, fomesafen.
Table 3.6. Frequency of the first waterhemp to emerge after PRE PPO-inhibitor application that
are heterozygous or homozygous for the ΔG210 mutation in 2020 and 2021 at Throckmorton and
Davis Field Sites.
Frequency of waterhemp with one or more

ΔG210 allelea

Treatment Davis Throckmorton

----------------------%-----------------------

Trifludimoxazin 46 bc 29 ab

Saflufenacil 59 ab 48 a

Fomesafen 60 ab 28 ab

Trifludimoxazin+Saflufenacil 69 a 39 ab

Triflufimoxazin+Fomesafen 63 ab 39 ab

Saflufenacil+Fomesafen 67 ab 29 ab

Trifludimoxazin+Saflufenacil+Fomesafen 70 a 27 ab

No Herbicide 34 c 18 b
a
Means within a column followed by the same letter are not significantly different according to
Tukey’s HSD test at α=0.05.

116
Table 3.7. Zygosity percentages of ΔG210 in the first waterhemp to emerge after treatment with PRE PPO-inhibitor application at
Throckmorton and Davis field sites in 2020 and 2021.
Homozygous Resistant/Total Homozygous

Resistanta Resistant/Total Collected Heterozygous /Total

Treatmentb 2020 2021 2020 2021 collected

---------------------------------------------------%-------------------------------------------------

Trifludimoxazin 17 ns 21 ab 4 ns 11 ab 31 ab

Saflufenacil 19 21 ab 11 13 a 42 a
117

Fomesafen 12 38 a 5 16 a 33 ab

Tri+Saf 19 14 ab 8 9 ab 45 a

Tri+Fom 14 21 ab 8 12 ab 41 a

Saf+Fom 6 37 a 4 22 a 35 ab

Tri+Saf+Fom 11 19 ab 6 13 ab 39 a

No Herbicide 14 6b 3 2b 24 b

a
Means within a column followed by the same letter are not significantly different according to Tukey’s HSD test at α=0.05.
b
Abbreviations: Tri, trifludimoxazin; Saf, saflufenacil; Fom, fomesafen
Table 3.8. Herbicide soil properties of PPO inhibitors used in the current experiments.
Herbicide Chemical family Water solubility Koca Kdb,c Soil half-life

mg L-1 mL g-1 day

fomesafen diphenylether 600,000 (salt at pH 7, 25 C) 60 1.11 to 12.76 100

saflufenacil N-phenyl-imide 2100 (pH 7) 9 to 56 0.02 to 0.2 1 to 36


118

trifludimoxazin N-phenyl-imide 1.78 315 to 692 0.52 12-383


a
Koc: soil/water partition coefficient - defined as the tendency of the herbicide to bind to soil by organic matter. A small value
means the herbicide will less likely be adsorbed to the soil and thus more mobile in the soil.
b
Kd: soil sorption index - defined as the ratio of the herbicide amount in soil compared to the amount in water. A small value
means a greater herbicide concentration in water.
c
References: Papiernik et al. (2012); Shaner (2014); Asher et al. (2021).
Table 3.9. Waterhemp control and survival assessed 14 days after POST PPO-inhibitor treatment
at Throckmorton and Davis field sites in 2020 and 2021.
Waterhemp survival

Treatment Waterhemp controla Davis Throckmorton

--------------------- %-------------------

Trifludimoxazin 79 ab 53 b 17 c

Saflufenacil 56 cd 56 ab 65 ab

Fomesafen 45 d 56 b 52 b

Trifludimoxazin+Saflufenacil 87 a 37 b 13 c

Triflufimoxazin+Fomesafen 82 a 38 b 21 c

Saflufenacil+Fomesafen 68 bc 54 b 34 bc

Trifludimoxazin+Saflufenacil+Fomesafen 87 a 34 b 22 c

No Herbicide 0- 100 a 100 a


a
Means within a column followed by the same letter are not significantly different according to
Tukey’s HSD test at α=0.05.

119
Table 3.10. Frequency of surviving waterhemp that have at least one ΔG210 allele after
treatment with POST herbicide.

Frequency of resistancea

Treatment 2020 2021

-----------------%----------------

Trifludimoxazin 55 a 64 b

Saflufenacil 45 ab 85 a

Fomesafen 54 a 82 ab

Trifludimoxazin+Saflufenacil 54 a 79 ab

Triflufimoxazin+Fomesafen 58 a 70 ab

Saflufenacil+Fomesafen 64 a 87 a

Trifludimoxazin+Saflufenacil+Fomesafen 66 a 87 a

No Herbicide 22 b 25 c
a
Means within a column followed by the same letter are not significantly different according to
Tukey’s HSD test at α=0.05.

120
Table 3.11. Zygosity percentages of the ΔG210 mutation in waterhemp surviving treatment with POST PPO-inhibitor treatments at
Throckmorton and Davis field sites in 2020 and 2021.
Homozygous Resistant/ Homozygous

Total Resistanta Resistant/Total Collected Heterozygous/Total collected

Davis Throckmorton

Treatmentb Davis Throckmorton Davis Throckmorton 2020 2021 2020 2021

-------------------------------------------------------------------%------------------------------------------------------------

Trifludimoxazin 19 ns 22 a 13 ab 15 a 39 ab 48 bc 47 ab 49 a

Saflufenacil 14 9 ab 11 ab 5 ab 41 ab 77 a 41 ab 69 a
121

Fomesafen 29 16 ab 24 a 9 ab 44 ab 53 abc 42 ab 66 a

Tri+Saf 12 21 a 10 ab 15 a 40 ab 77 a 49 a 54 a

Tri+Fom 15 9 ab 12 ab 7 ab 40 ab 56 abc 64 a 57 a

Saf+Fom 22 9 ab 19 ab 6 ab 61 a 65 ab 53 a 74 a

Tri+Saf+Fom 21 30 a 17 ab 20 a 60 a 71 ab 37 ab 62 a

No Herbicide 17 2b 6b 1b 16 b 31 c 20 b 13 b
a
Means within a column followed by the same letter are not significantly different according to Tukey’s HSD test at α=0.05
b
Abbreviations: Tri, trifludimoxazin; Saf, saflufenacil; Fom, fomesafen.
INVESTIGATING THE POTENTIAL CO-
OCCURRENCE OF TARGET-SITE AND NON-TARGET-SITE
RESISTANCE IN WATERHEMP (AMARANTHUS TUBERCULATUS)
POPULATIONS

Short title: Fomesafen and detox inhibitors

4.1 Abstract

Widespread glyphosate resistance in waterhemp [Amaranthus tuberculatus (Moq.)

J.D.Sauer] during the last decade resulted in increased use of protoporphyrinogen IX oxidase (PPO)

herbicides in soybean. A 2016 field survey across several Midwestern states discovered two PPO-

resistant (R) waterhemp populations that exhibited an exceptionally high magnitude of fomesafen

resistance (16 to 17-fold) in comparison to other R populations (4 to 10-fold). This greater

resistance was not explained by presence of novel mutations or R allele zygosity. Furthermore,

non-target-site resistance, particularly CyP450 and GST based detoxification, has become

increasingly important for other herbicide groups. We hypothesized that enhanced fomesafen

detoxification was contributing to the increased resistance in these select PPO-R waterhemp

populations. Our objective was to reduce the PPO-inhibitor resistance phenotype using CyP450

and GST inhibitors (malathion and NBD-Cl, respectively) in combination with fomesafen

applications. Waterhemp populations named IN-RAN and IA-340 were the susceptible

populations. IL-BRO and IN-PIKE were 6 to 9X resistant. Finally, IL-WAS and IN-DUB were 16

to 17X resistant. Plants of each population were sprayed with NBD-CL, Malathion, NBD-Cl

followed by Malathion, or neither inhibitor prior to being sprayed with fomesafen at rates of 0, 20,

or 60 g ai ha-1. Malathion and NBD-Cl applications without fomesafen resulted in stunting that

was variable by population. At the 60 g ha-1 rate, malathion + NBD-Cl increased control of IL-

122
WAS by 20 to 26 percentage points in comparison to no inhibitor and NBD-Cl alone. While this

statistic appears to support our hypothesis, the lack of consistency between high and low herbicide

rates, and the noticeable phytotoxic response to malathion and NBD-Cl does not exclude potential

additive effects from the multiple applied xenobiotics. We conclude that our methodology was

insufficient to address our hypothesis because of the presence of many unvalidated assumptions

about the action of detoxification inhibitors applied to foliage as well as the issue of phytotoxicity

of malathion and NBD-Cl by themselves.

Nomenclature: Waterhemp, Amaranthus tuberculatus (Moq.) J.D. Sauer, fomesafen, NBD-Cl, 4-

Chloro-7-nitro-2,1,3-benzoxadiazole, Malathion

Key words: PPO inhibitor resistance, target-site resistance, ΔG210 mutation, Non-target-site

resistance, herbicide metabolism, Glutathione-S-transferase, Cytochrome P450

4.2 Introduction

One of the most troublesome characteristics of waterhemp [Amaranthus tuberculatus (Moq.)

J.D. Sauer] is its propensity to evolve herbicide resistance (Tranel et al. 2011). To date, waterhemp

has evolved resistance to herbicides from seven sites of action including PPO-inhibiting herbicides

(Heap 2023, Shoup et al. 2003). The first case of PPO-inhibitor resistance was confirmed in

waterhemp in 2001 and is primarily caused by a codon deletion at the 210th position of the PPX2

gene encoding one of the target enzymes protoporphyrinogen IX oxidase 2 (PPO2) (Patzoldt et al.

2006, Shoup et al. 2003). Mutations at the 128th position of the PPX2 gene (R128G and R128I)

also confer resistance in waterhemp, but populations with this mutation are less widespread (Nie

et al. 2019).

123
Palmer amaranth (Amaranthus Palmeri S. Wats), a closely related species, has also evolved

resistance via the ΔG210 mutation, as well as R128G, R128M, and G399A mutations (Giacomini

et al. 2017, Rangani et al. 2019, Salas et al. 2016). Other species that have evolved PPO-inhibitor

resistance include common ragweed (Ambrosia artemisiifolia L.) (Rousonelos et al. 2012), wild

poinsettia (Euphorbia heterophylla L.) (Mendes et al. 2020), and redroot pigweed (Amaranthus

retroflexus L.) (Huang et al. 2020) via R128 (or equivalent) mutations in PPX2 as well as

goosegrass [Eleusine indica (L.) Gaertn.] via an A212T mutation in the PPX1 gene (Bi et al. 2019).

Resistance to PPO inhibitors that is not explained by target-site mutations has also been

documented in waterhemp and Palmer amaranth to the herbicides carfentrazone-ethyl and

fomesafen, respectively (Obenland et al. 2019, Varanasi et al. 2018). Both authors claim that

resistance is explained by enhanced detoxification of the herbicide.

One of the most concerning trends in herbicide resistance evolution, particularly in

Amaranthus species, is the growing prevalence of non-target-site resistance (Jugulam and Shyam

2019). Waterhemp has evolved resistance to HRAC groups 5, 15, and 27 via enhanced

detoxification by Phase I and Phase II detoxification enzymes (Lygin et al. 2018, Patzoldt et al.

2003, Strom et al. 2020, Vennapusa et al. 2018). While multiple enzymes can contribute to phase

I and phase II xenobiotic metabolism, the most commonly implicated enzymes are cytochrome

P450 (P450) and glutathione S-transferase (GST) respectively. Other weed species have evolved

detoxification-based resistance to nearly all known herbicide mode of action groups (Heap 2023,

Jugulam and Shyam 2019). A potential threat of weeds with non-target-site resistance is broad

cross resistance to multiple modes of action. Potentially, a single P450 or GST enzyme could be

upregulated that is able to metabolize multiple herbicide molecules (Dimaano and Iwakami 2021).

A single P450 gene from Echinochloa phyllopogon overexpressed in a bacterial system was able

124
to metabolize herbicides from 6 different site of action groups (Dimaano et al. 2020). Such an

ability portends that cross resistance will accelerate and current resistance management practices

are potentially worsening metabolic resistance evolution.

The resistance phenotype to PPO inhibitors is variable in terms of population and active

ingredient response. Resistance to lactofen is typically reported at between 20 to 80 fold and up to

280 fold (Lillie et al. 2020, Patzoldt et al. 2005, Shoup et al. 2003, Wuerffel et al. 2015).

Resistances to other PPO inhibitors, most notably, a similar diphenyl-ether herbicide, fomesafen,

are typically reported at lower levels. Fomesafen resistance is typically reported at 6 to 15 fold,

but sometimes approaches 20- or 30-fold (Lillie et al. 2020, Patzoldt et al. 2005, Shoup et al. 2003,

Wuerffel et al. 2015). Resistances to other PPO inhibitors are reported at 5 to 30 fold, depending

on the lab group and experiment. From several studies, it is clear that the ΔG210 mutation confers

the greatest resistance to lactofen, but also confers meaningful cross resistance to other PPO

inhibitors. The wide range of reported resistance ratios is difficult to explain. Potential contributing

factors are application conditions, growth conditions and plant size, population genetic purity, and

evaluation period (Hatterman-Valenti et al. 2011, Lillie et al. 2020, Salas et al. 2016, Wichert et

al. 1992). However, these do not fully explain the range in resistance responses observed by so

many research groups. In unpublished research, we have evaluated many Midwest populations for

PPO-inhibitor resistance and observed R/S ratios of between 4.4X and 17X despite all populations

being fixed or nearly fixed for the ΔG210 mutation. The only remaining factor that can contribute

to the observed resistance variance is background genetic effects. Background genetic effects can

take the form of more sensitive susceptible populations and more resistant R populations.

According to our unpublished PPO-resistance screen, increases in GR50 account for the variable

resistance rather than variance in S populations.

125
Target-site resistance is typically a monogenic, Mendelian trait (Murphy and Tranel 2019).

However, recent experiments and increasing awareness of non-target-site resistance is causing a

paradigm shift toward all herbicide resistance as a quantitative trait (Kreiner et al. 2021, Murphy

et al. 2021). While single locus target-site alleles are a large effect contributor, they are a single

component to an overall resistance phenotype. Multigenic resistance has been documented in

several species for many herbicide site of action groups, particularly target-site and enhanced

detoxification occurring in the same plant (Alcántara-de la Cruz et al. 2016, Hatami et al. 2016,

Kaundun 2010, Rey-Caballero et al. 2017). A large proportion of glyphosate resistance in

waterhemp throughout the Midwest is explained by polymorphisms in genes throughout the

genome and not just related to EPSP synthase (Kreiner et al. 2021). A recent Palmer amaranth

survey observed high phenotypic variance in PPO-resistance (Noguera et al. 2021). Populations

were clustered into five response groups in which multiple target-site resistance alleles were

associated with greater resistance. The authors credited much of the increasing resistance with

accumulation of target-site resistance alleles; however, they did not fully rule out the potential

accumulation of non-target-site resistance alleles. Notably, populations predominantly from

Mississippi carried no target-site alleles, but had resistant individuals (Noguera et al. 2021). Given

the increasing incidence of non-target-site resistance and the unexplained variance in PPO-

inhibitor resistance observed in waterhemp, we hypothesized that increased detoxification is

contributing to the overall resistance phenotype to fomesafen in waterhemp populations. Our

objective was to determine if co-applying a P450 inhibitor, a GST inhibitor, or both in combination

with fomesafen would cause a partial reversal of fomesafen resistance.

126
4.3 Materials and Methods

4.3.1 Plant Materials

Greenhouse experiments were conducted in the Fall of 2021 to investigate co-occurrence

of target-site and non-target-site resistance occurring in waterhemp. The subject waterhemp

populations were selected from a 2016 PPO-resistance survey (Nie et al. 2019). The selected

populations were all previously subjected to full dose-response characterization and genotyping

assays for the ΔG210 mutation. There were two populations of each desired response range of

susceptible, highly resistant, and moderately resistant (unpublished data). Susceptible populations

had a GR50 of 2.1 to 2.7 g ai ha-1 of fomesafen and are designated as IA-340 and IN-RAN.

Moderately resistant populations were named IL-BRO1 and IN-PIKE1 and had a GR50 of 18 and

25 g ai ha-1 (6.7 and 9.3 fold resistant) respectively. Highly resistant populations were named IN-

DUB and IL-WAS and had GR50 values of 43 and 45 g ai ha-1 (16- and 17-fold resistant). In all of

the resistant populations, at least 95% of individual plants were either heterozygous or

homozygous for the ΔG210 mutation and were confirmed as being absent of any R128 mutations.

Seeds of each population were sown for germination in flats containing a peat moss-based

potting mix. At approximately 1 wk after sowing, seedlings at the 1-leaf stage were transplanted

into 164 ml cone-tainers (Ray Leach SC-10 Super Cell Cone-tainers; Stuewe & Sons, Tangent,

OR) filled with a mixture of two parts peatmoss-based potting mix and one part sand. Pots were

watered as needed and fertilized weekly with a micro- and macronutrient fertilizer (Jack’s Classic

Professional 20-20-20, JR Peters Inc., Allentown PA) throughout the course of the experiments.

Greenhouse environmental conditions included 16:8 light/dark photoperiods, where natural light

was supplemented with high-pressure sodium bulbs delivering 1100 μmol m-2 s-1 photon flux

during daylight hours, and day/night temperatures of 30 and 25C, respectively.

127
4.3.2 Spray Applications and Preparations

Applications of malathion and NBD-Cl (4-Chloro-7-nitro-2,1,3-benzoxadiazole) were

applied to inhibit cytochrome P450 and glutathione-S-transferase respectively (Kreuz and Fonné-

Pfister 1992, Ricci et al. 2005). Malathion (Spectracide Malathion Insect Spray Concentrate,

Spectrum Group, United Industries Corporation, St. Louis MO) was applied 2h prior to fomesafen

application at a rate of 1500 g ai ha-1 plus a non-ionic surfactant (Activator 90, Loveland Products,

Inc., Greeley, CO) at 0.25% v/v similar to methodologies described previously (Oliveira et al.

2018, Varanasi et al. 2018). NBD-Cl 98% powder (CAS No. 10199-89-0) was dissolved in DMSO

and applied 48 h prior to fomesafen application at a rate of 270 g ha-1 (Cummins et al. 2013,

Varanasi et al. 2018).

NBD-CL is highly soluble in DMSO and acetone, but highly insoluble in water. The

following procedure was developed by trial and error to create a sprayable mixture composed of

mostly water. NBD-CL powder was dissolved into a quantity of DMSO corresponding to 1% of

final spray volume, followed by mixing with COC at 1% v/v and Tween20 (2% v/v stock solution)

at 2.5% v/v. The solvent and surfactant solution was decanted into a spray bottle, to which, water

was added and shaken vigorously for several minutes to create a final volume to spray at 140 L ha-
1
. The final product was a sprayable mixture of partially dissolved, partially suspended NBD-CL.

Fomesafen (Flexstar®, Syngenta Crop Protection, LLC, Greensboro, NC) was applied at

the 5 to 7 leaf stage at a rate 0, 20, or 60 g ai ha-1 with 1% (v/v) crop oil concentrate (Prime Oil®,

Winfield Solutions, LLC, St. Paul, MN). Herbicide and detoxification-inhibitor applications were

made using a track-mounted research sprayer (Generation III Research Sprayer, DeVries

Manufacturing, Hollandale MN) calibrated to deliver 140 L ha-1 at 207 kPa via an even flat fan

XR8002E (TeeJet Technologies, Glendale Heights, IL) nozzle.

128
4.3.3 Experimental Design, Data Collection, and Analysis

Each population received one of twelve treatment regimens from the three-factor factorial

arrangement. Factors consisted of: 1. NBD-CL or blank DMSO-solvent solution; 2. Malathion or

no malathion; 3. Fomesafen applied at 0, 20, or 60 g ai ha-1. Plants receiving only the blank DMSO-

adjuvant solution served as the control. The experiment utilized a randomized complete block

design with 8 replications and was repeated once in time.

Visual estimates of control and plant height were recorded at 7 and 14 d after treatment

(DAT). Control estimates were taken on a 0 to 100% scale with 100% = complete plant death and

0% = no injury. At 14 DAT plants were clipped at the soil surface and oven dried at 60C until dry

biomass was recorded. Biomass data were normalized to a percentage based on the adjuvant-

solvent only treatment for each population. Data were subjected to analysis of variance using

PROC GLIMMIX in SAS 9.4 (SAS Institute; Cary, NC), with means separated using Tukey’s

HSD (α = 0.05). Waterhemp population, herbicide rate, and detoxification inhibitor treatment were

considered fixed effects while rep nested within run was considered a random factor.

4.4 Results and Discussion

4.4.1 Response to Detoxification Inhibitors Alone

Applications of detoxification inhibitors were made so that they conformed to previously

published research while still meeting our own experimental objectives. Published experiments

utilizing malathion in combination with herbicides typically use 1500 or 2000 g ha-1 followed by

some groups with a soil drench of malathion solution 2 d after application (Ma et al. 2013,

Obenland et al. 2019, Oliveira et al. 2018, Shyam et al. 2021, Varanasi et al. 2018). Malathion

application is a straightforward process since the compound is readily available as a formulated

129
concentrate for spray application. Multiple research groups have used malathion to successfully

demonstrate P450 based herbicide metabolism (Kreuz and Fonné-Pfister 1992, Ma et al. 2013,

Oliveira et al. 2018, Shyam et al. 2021, Tardif and Powles 1999).

In contrast, NBD-Cl has a relatively short history in the peer reviewed literature. The

property of being a non-reversable inhibiter of GST, along with other analogues, was first

demonstrated by Ricci et al. (2005). The first use in studying herbicide resistance was for

demonstrating GST based resistance to chlorotoluron and fenoxaprop-p-ethyl in blackgrass

(Alopecurus myosuroides Huds.) (Cummins et al. 2013). NBD-Cl applied to resistant blackgrass

prior to herbicide application completely or almost completely reversed resistance to the herbicides.

Other groups have utilized NBD-Cl with mixed results. Some groups successfully reversed the

resistance phenotype of the subject population, others had a phenotype reversal in some, but not

all populations, lastly some groups did not see a phenotype reversal effect. (Ma et al. 2016, Rangani

et al. 2021, Shyam et al. 2021, Strom et al. 2020, Varanasi et al. 2018).

One immediate observation that we made was the difficulty required to prepare the NBD-

Cl for spraying. NBD-Cl can be readily dissolved in a small volume of acetone or DMSO, however

water added to the solvent solution causes precipitation. We wanted to avoid spraying a primarily

solvent based spray solution because of phytotoxicity concerns, so maintaining a mostly dissolved

solution required oil adjuvants and surfactant. A description of the methods used by other groups

for spraying NBD-Cl in the literature was almost entirely absent. Only Cummins et al. (2013)

provided mixing and spraying procedures in their manuscript in which they applied 1600 L ha-1 of

carrier volume. We attempted to replicate the mixing procedure with the exception of reducing

final spray volume, but were unsuccessful. A spray volume of approximately 1600 L ha-1 was

required to obtain a sprayable solution of NBD-CL at 270 g ha-1. We observed that applying 1600

130
L ha-1 with our chosen adjuvants caused severe phytotoxicity and even caused some waterhemp

lethality in a preliminary experiment (data not shown). The toxicity was not due to the NBD-Cl

but rather the spray mix components, because other plants sprayed with only the water-solvent-

adjuvant solution using the same volume had equal injury (data not shown). While some other

groups mention the chosen solvent used with no other preparation details, most methods entirely

omit mixing procedures for NDB-CL (Ma et al. 2016, Shyam et al. 2021, Varanasi et al. 2018).

Our final developed mixing procedure was a semi-dissolved, low volume (140 L ha-1)

mixture that did cause observable plant injury. Within 24 hours after spraying, NBD-Cl treated

plants developed small lesions on the treated leaf surface (Figure 4.1). Newly emerged leaves were

unaffected, but treated plants lacked vigor and had a general unhealthy appearance. Long term

injury from NBD-CL, malathion, and NBD-Cl+malathion was observable at 14 d after application

(Table 4.1). The injury was variable by population, but generally manifested as height reduction

and stunting (Table 4.2). The cause of waterhemp injury is unclear. Since GSTs and their precursor

molecules and pathways are involved in a variety of plant processes, perhaps inhibiting waterhemp

GST and P450 created a metabolic drain that reduced growth (Dixon et al. 2002, Xu et al. 2015).

4.4.2 PPO-Resistance Response

Waterhemp control is presented as two-way factorial of population and detoxification

inhibitor treatment, and separated by fomesafen rate due to significant three-way interaction of

rate, population, and detoxification inhibitor. The interaction occurred due to variable effects of

detoxification inhibitors alone on each population, as well as differing effects of detoxification

inhibitors at the two herbicide rates. Control of S populations at 14 d after fomesafen application

was much greater than R populations at both the 20 and 60 g ha-1 rate (Table 4.1). Fomesafen alone

provided 86 to 88 % control at 20 g ha-1 and 98 to 100% control at 60 g ha-1. In resistant waterhemp

131
populations, the previously characterized moderately and highly resistant characterizations were

not apparent. The 20 g ha-1 rate of fomesafen provided 35 to 40% control of IL-BRO1, IN-PIKE,

and IL-WAS compared to 63% control of IN-DUB, the most resistant waterhemp population

according to previous dose-response experiments. A similar trend was present at the 60 g ha-1 rate

where control of IL-WAS was 54% compared to control of IN-DUB of 76%. Height and biomass

means for each population also indicate that IN-DUB did not exhibit the resistant phenotype as

was previously characterized, evidenced by the lack of differences between the other 3 resistant

populations (Table 4.3). A possible reason for why the previous resistance characterization was

not reproduced in the present study is the experimental conditions. The original experiments were

part of a multi-state survey of PPO-resistant waterhemp in which dozens of populations were

screened and sprayed at the same time. Perhaps variance in germination or growth rate contributed

to greater than usual plant size differences leading to the appearance of a more or less resistant

phenotype (Mansfield 2021).

There were few differences between fomesafen alone and fomesafen preceded by

detoxification inhibitor treatment. At 60 g fomesafen ha-1, control of IL-WAS treated with both

malathion and NBD-CL was increased to 73% compared to 47 to 54% from NBD-CL and no

detoxification inhibitor. Because IL-WAS was previously characterized as highly resistant and

there was a significant increase in control, there is some evidence that fomesafen detoxification is

contributing to the overall resistance (Figure 4.2). However, biomass and height data indicate that

the observation in IL-WAS was not unique.

Data for height and biomass are separated by herbicide rate but combined over population

and detoxification inhibitor due to insignificant interactions of population and detoxification

inhibitor. (Table 4.2, Table 4.3). The lack of a significant population by detoxification inhibitor

132
term is countervailing evidence that some waterhemp populations are more rapidly metabolizing

fomesafen. Similar numerical trends were present for each individual population and no individual

population responded in terms of height or biomass for any of the detoxification inhibitor

treatments. At the 20 g ha-1 rate, NBD-CL and fomesafen alone treatments had greater height by

1.9 to 3.2 cm and greater biomass by 8 to 11 percentage points compared to malathion and

malathion+NBD-Cl treatments. This observation may indicate that all waterhemp reduces

phytotoxicity to some extent via partial metabolism of fomesafen. However, the presence of

similar biomass and height reductions in the absence of fomesafen treatment indicates that the

increased phytotoxicity is an additive effect of multiple applied xenobiotics. Interestingly, the

application of NBD-CL created a significant height and biomass reduction when applied alone,

but NBD-CL applied prior to fomesafen application appeared to have a safening or antagonistic

effect.

With the assumption that malathion and NDB-Cl truly inhibited P450 and GST enzymes in

the treated plants, there is some weak evidence that P450 enzymes are contributing to partial

fomesafen tolerance in all waterhemp populations and the IL-WAS population may be

metabolizing fomesafen to a greater extent than other populations. Fomesafen and other diphenyl-

ether metabolism in soybean occurs via GST mediated ether bond cleavage (Andrews et al. 1997,

Frear et al. 1983). To our knowledge, there is no documented evidence of P450-mediated

metabolism of diphenyl-ether herbicides in crops or weeds. However, metabolism of saflufenacil

and sulfentrazone in soybean is through P450-mediated demethylation (Dayan et al. 1997, Wu et

al. 2019). Also, resistance mechanisms to herbicides in weeds is not always the same pathway as

in the associated crops (Lygin et al. 2018, Strom et al. 2020). P450 enzymes as a group are capable

133
of double bond oxidation, ring structure oxidations, and heteroatom dealkylations (Dimaano and

Iwakami 2021).

The methodology and conclusions in our research rely on the previously stated assumption

that these P450 and GST inhibitors were truly inhibiting their respective enzymes and having no

secondary herbicidal effects. The first part, inhibition of P450 and GST is unverifiable with our

chosen methods, and previous literature indicates that it is not a reliable assumption in all

circumstances (Ma et al. 2016, Oliveira et al. 2018, Strom et al. 2020). P450 and GST are a diverse

set of enzymes that are likely not going to be universally inhibited by a single molecule. Several

groups apply multiple known P450 inhibitors as separate treatments with malathion, such as

piperonyl butoxide and amitrole (Oliveira et al. 2018, Varanasi et al. 2018). At least one of these

treatments would commonly not inhibit resistance conferring P450 adequately for phenotype

reversal. Plant tissue may also play a role. Strom et al. (2020) successfully exposed waterhemp

seedlings to NBD-Cl and documented differential S-metolachlor metabolism. However, NBD-Cl

did not successfully inhibit corn GSTs well enough to reduce metabolism of S-metolachlor. It is

unclear if the exposure event was not adequate for absorption, or if some enzyme level physiology

is different in corn vs waterhemp such as expression or binding characteristics. With the lack of

precedent in the literature, NBD-Cl treatment rates have likely not been optimized (Cummins et

al. 2013, Ma et al. 2016).

The second assumption about no secondary herbicidal effects is verifiably false. We cannot

conclusively say whether the increased waterhemp control is from the additive general

phytotoxicity of malathion followed by fomesafen or if the inhibition of P450 enzymes caused

reduced fomesafen detoxification and therefore increased herbicide activity. Since the central

catalytic molecule of P450s is heme, potentially the inhibition of P450s is causing a secondary

134
effect on the PPO pathway. Perturbations in chemical pathways can often result in herbicide

synergy (Takano et al. 2020, Walsh et al. 2012, Woodyard et al. 2009).

When studying herbicide interactions, there are specific statistical approaches for

determining herbicide synergy or additivity. The most common methods are isobole analysis and

Colby’s method. Both approaches use statistics to determine if plant response exceeds that of the

theoretical additive response, with isobole analysis being more robust, but more complex in

experimental design and modeling. Formal additivity/synergy analysis may be appropriate for

resistance phenotype reversal, especially in the case of partially phytotoxic chemicals like NBD-

Cl and malathion (Cedergreen 2014, Cedergreen et al. 2008, Colby 1967).

In comparing our research to others with similar methodology, we see a disturbing lack of a

consistent standard for what is considered adequate resistance phenotype reversal. Several authors

conclude metabolic resistance from any statistical increase in control compared to no P450 or GST

inhibitor. This conclusion of metabolic resistance seems to be reached even in the presence of

either a similar effect occurring in the susceptible, or with a complete lack of susceptible or target-

site resistant population for comparison. We suggest that a consistent standard be implemented

where either a particular benchmark increase in efficacy be achieved or have supplemental


14
experiments that measure degradation of herbicide and/or increase in metabolites with C or

chromatography-based experiments.

Several groups have successfully demonstrated a compelling resistance phenotype reversal

where co-application of detoxification inhibitors with herbicide provides a complete or nearly

complete phenotype reversal (Brabham et al. 2019, Chen et al. 2020, Cummins et al. 2013, Oliveira

et al. 2018, Strom et al. 2020). Others have measured parent herbicide disappearance or performed

genetic assays as a supplemental experiment with the co-application of detoxification inhibitors

135
and herbicides (Chen et al. 2020, Ma et al. 2013, Nandula et al. 2020, Oliveira et al. 2018, Rangani

et al. 2021, Strom et al. 2020). However, others present an unclear increase in control in

unbalanced experiments and no supplemental experiments (González-Torralva and Norsworthy

2021, Obenland et al. 2019, Varanasi et al. 2018, 2019). Some of the less interpretable results are

repeatedly cited as a clear-cut phenomenon, when the conclusions are much less conclusive. The

assumptions required to interpret detoxification inhibitor experiments, the confounding factor of

phytotoxic additivity from multiple chemical applications, as well as the lack of consistent

methodology leads us to the conclusion that these types of experiments should not be standalone

experiments. While there are limitations, our opinion is that this methodology still has merit, but

only as a preliminary hypothesis building experiment.

In conclusion, the hypothesis that increased detoxification is contributing to the overall

resistance phenotype to fomesafen in waterhemp populations was not fully testable with our

chosen methodology. Despite the methodological flaws, it is unlikely that fomesafen detoxification

is meaningfully contributing to the overall resistance phenotype to fomesafen. The greater value

of this manuscript is a call for increased materials and methods transparency, a more consistent

standard for diagnosing detoxification-based resistance, and more robust supplemental

experiments that directly measure herbicide degradation or concentrations in treated plants.

4.5 Literature Cited

Alcántara-de la Cruz R, Fernández-Moreno PT, Ozuna C V., Rojano-Delgado AM, Cruz-Hipolito


HE, Domínguez-Valenzuela JA, Barro F, de Prado R (2016) Target and non-target site
mechanisms developed by glyphosate-resistant hairy beggarticks (Bidens pilosa L.)
populations from Mexico. Front Plant Sci 7:1492

136
Andrews CJ, Skipsey M, Townson JK, Morris C, Jepson I, Edwards R (1997) Glutathione
transferase activities toward herbicides used selectively in soybean. Pest Manag Sci 51:213–
222

Bi B, Wang Q, Coleman JJ, Porri A, Peppers JM, Patel JD, Betz M, Lerchl J, McElroy JS (2019)
A novel mutation A212T in chloroplast protoporphyrinogen oxidase (PPO1) confers
resistance to PPO inhibitor oxadiazon in Eleusine indica. Pest Manag Sci 76:1786–1794

Brabham C, Norsworthy JK, Houston MM, Varanasi VK, Barber T (2019) Confirmation of S -
metolachlor resistance in Palmer amaranth (Amaranthus palmeri). Weed Technol 33:720–
726

Cedergreen N (2014) Quantifying synergy: a systematic review of mixture toxicity studies within
environmental toxicology. PLoS One 9:1-12
Cedergreen N, Christensen AM, Kamper A, Kudsk P, Mathiassen SK, Streibig JC, Sorensen H
(2008) A review of independent action compared to concentration addition as reference
models for mixtures of compounds with different molecular target sites. Environ Toxicol
Chem 27:1621-1632
Chen W, Wu L, Wang J, Yu Q, Bai L, Pan L (2020) Quizalofop-p-ethyl resistance in Polypogon
fugax involves glutathione S-transferases. Pest Manag Sci 76:3800–3805

Colby SR (1967) Calculating synergistic and antagonistic responses of herbicide combinations.


Weeds 15:20-22
Cummins I, Wortley DJ, Sabbadin F, He Z, Coxon CR, Straker HE, Sellars JD, Knight K, Edwards
L, Hughes D, Kaundun SS, Hutchings SJ, Steel PG, Edwards R (2013) Key role for a
glutathione transferase in multiple-herbicide resistance in grass weeds. Proc Natl Acad Sci U
S A 110:5812–5817

Dayan FE, Weete JD, Duke SO, Hancock HG (1997) Soybean (Glycine max) cultivar differences
in response to sulfentrazone. Weed Sci 45:634–641

Dimaano NG, Iwakami S (2021) Cytochrome P450-mediated herbicide metabolism in plants:


current understanding and prospects. Pest Manag Sci 77:22–32

137
Dimaano NG, Yamaguchi T, Fukunishi K, Tominaga T, Iwakami S (2020) Functional
characterization of cytochrome P450 CYP81A subfamily to disclose the pattern of cross-
resistance in Echinochloa phyllopogon. Plant Mol Biol 102:403–416

Dixon DP, Lapthorn A, Edwards R (2002) Plant glutathione transferases. Genome Biol
3:reviews3004.1–reviews3004.10

Frear DS, Swanson HR, Mansager ER (1983) Acifluorfen metabolism in soybean: diphenylether
bond cleavage and the formation of homoglutathione, cysteine, and glucose conjugates. Pestic
Biochem Physiol 20:299–310

Giacomini DA, Umphres AM, Nie H, Mueller TC, Steckel LE, Young BG, Scott RC, Tranel PJ
(2017) Two new PPX2 mutations associated with resistance to PPO-inhibiting herbicides in
Amaranthus palmeri. Pest Manag Sci 73:1559–1563

González-Torralva F, Norsworthy JK (2021) Understanding resistance mechanisms to trifluralin


in an Arkansas Palmer amaranth population. Genes 2021, Vol 12, Page 1225 12:1225

Hatami ZM, Gherekhloo J, Rojano-Delgado AM, Osuna MD, Alcántara R, Fernández P,


Sadeghipour HR, de Prado R (2016) Multiple mechanisms increase levels of resistance in
rapistrum rugosum to ALS Herbicides. Front Plant Sci 7:169

Hatterman-Valenti H, Pitty A, Owen M (2011) Environmental effects on velvetleaf (Abutilon


theophrasti) epicuticular wax deposition and herbicide absorption. Weed Sci 59:14–21

Heap I (2023) International survey of herbicide resistant weeds. http://www.weedscience.org.


Accessed March 25, 2023

Huang Z, Cui H, Wang C, Wu T, Zhang C, Huang H, Wei S (2020) Investigation of resistance


mechanism to fomesafen in Amaranthus retroflexus L. Pestic Biochem Physiol 165:104560

Jugulam M, Shyam C (2019) Non-target-site resistance to herbicides: recent developments. Plants


8:417

Kaundun SS (2010) An aspartate to glycine change in the carboxyl transferase domain of acetyl
CoA carboxylase and non-target-site mechanism(s) confer resistance to ACCase inhibitor
herbicides in a Lolium multiflorum population. Pest Manag Sci 66:1249–1256

138
Kreiner JM, Tranel PJ, Weigel D, Stinchcombe JR, Wright SI (2021) The genetic architecture and
population genomic signatures of glyphosate resistance in Amaranthus tuberculatus. Mol
Ecol 30: 5373-5389

Kreuz K, Fonné-Pfister R (1992) Herbicide-insecticide interaction in maize: Malathion inhibits


cytochrome P450-dependent primisulfuron metabolism. Pestic Biochem Physiol 43:232–240

Lillie KJ, Giacomini DA, Tranel PJ (2020) Comparing responses of sensitive and resistant
populations of Palmer amaranth (Amaranthus palmeri) and waterhemp (Amaranthus
tuberculatus var. rudis) to PPO inhibitors. Weed Technol. 34: 140–146.

Lygin A V., Kaundun SS, Morris JA, Mcindoe E, Hamilton AR, Riechers DE (2018) Metabolic
pathway of topramezone in multiple-resistant waterhemp (Amaranthus tuberculatus) differs
from naturally tolerant maize. Front Plant Sci 9:1644

Ma R, Evans AF, Riechers DE (2016) Differential responses to preemergence and postemergence


atrazine in two atrazine-resistant waterhemp populations. Agron J 108:1196–1202

Ma R, Kaundun SS, Tranel PJ, Riggins CW, McGinness DL, Hager AG, Hawkes T, Mc EI,
Riechers DE (2013) Distinct detoxification mechanisms confer resistance to mesotrione and
atrazine in a population of waterhemp. Plant Physiol 163:363–377

Mansfield BC (2021) Characterization of protoporphyrinogen oxidase (PPO) herbicide resistance


in tall waterhemp (Amaranthus tuberculatus). MS thesis. West Lafayette IN: Purdue
University. 158 p

Mendes RR, Takano HK, Adegas FS, Oliveira RS, Gaines TA, Dayan FE (2020) R128L target site
mutation in PPO2 evolves in wild poinsettia (Euphorbia heterophylla) with cross-resistance
to PPO-inhibiting herbicides. Weed Sci. 68:437–444

Murphy BP, Beffa R, Tranel PJ (2021) Genetic architecture underlying HPPD-inhibitor resistance
in a Nebraska Amaranthus tuberculatus population. Pest Manag Sci 77:4884–4891

Murphy BP, Tranel PJ (2019) Target-site mutations conferring herbicide resistance. Plants 8:382

Nandula VK, Giacomini DA, Lawrence BH, Molin WT, Bond JA (2020) Resistance to clethodim
in Italian ryegrass (Lolium perenne ssp. multiflorum) from Mississippi and North Carolina.
Pest Manag Sci 76:1378–1385

139
Nie H, Mansfield BC, Harre NT, Young JM, Steppig NR, Young BG (2019) Investigating target‐
site resistance mechanism to the PPO‐inhibiting herbicide fomesafen in waterhemp and
interspecific hybridization of Amaranthus species using next generation sequencing. Pest
Manag Sci 75:3235–3244

Noguera MM, Rangani G, Heiser J, Bararpour T, Steckel LE, Betz M, Porri A, Lerchl J,
Zimmermann S, Nichols RL, Roma-Burgos N (2021) Functional PPO2 mutations: co-
occurrence in one plant or the same ppo2 allele of herbicide-resistant Amaranthus palmeri in
the US mid-south. Pest Manag Sci 77:1001–1012

Obenland OA, Ma R, O’Brien SR, Lygin A V., Riechers DE (2019) Carfentrazone-ethyl resistance
in an Amaranthus tuberculatus population is not mediated by amino acid alterations in the
PPO2 protein. PLoS One 14:e0215431

Oliveira MC, Gaines TA, Dayan FE, Patterson EL, Jhala AJ, Knezevic SZ (2018) Reversing
resistance to tembotrione in an Amaranthus tuberculatus (var. rudis) population from
Nebraska, USA with cytochrome P450 inhibitors. Pest Manag Sci 74:2296–2305

Patzoldt WL, Dixon BS, Tranel PJ (2003) Triazine resistance in Amaranthus tuberculatus (Moq)
Sauer that is not site-of-action mediated. Pest Manag Sci 59:1134–1142

Patzoldt WL, Hager AG, McCormick JS, Tranel PJ (2006) A codon deletion confers resistance to
herbicides inhibiting protoporphyrinogen oxidase. Proc Natl Acad Sci U S A 103:12329–34

Patzoldt WL, Tranel PJ, Hager AG (2005) A waterhemp (Amaranthus tuberculatus) biotype with
multiple resistance across three herbicide sites of action. Weed Sci 53:30–36

Rangani G, Noguera M, Salas-Perez R, Benedetti L, Roma-Burgos N (2021) Mechanism of


resistance to S-metolachlor in Palmer amaranth. Front Plant Sci 12:652581

Rangani G, Salas-Perez RA, Aponte RA, Knapp M, Craig IR, Mietzner T, Langaro AC, Noguera
MM, Porri A, Roma-Burgos N (2019) A novel single-site mutation in the catalytic domain of
protoporphyrinogen oxidase IX (PPO) confers resistance to PPO-inhibiting herbicides. Front
Plant Sci 10:568

140
Rey-Caballero J, Menéndez J, Osuna MD, Salas M, Torra J (2017) Target-site and non-target-site
resistance mechanisms to ALS inhibiting herbicides in Papaver rhoeas. Pestic Biochem
Physiol 138:57–65

Ricci G, De Maria F, Antonini G, Turella P, Bullo A, Stella L, Filomeni G, Federici G, Caccuri


AM (2005) 7-Nitro-2,1,3-benzoxadiazole derivatives, a new class of suicide inhibitors for
glutathione S-transferases. J Biol Chem 280:26397–26405

Rousonelos SL, Lee RM, Moreira MS, VanGessel MJ, Tranel PJ (2012) Characterization of a
common ragweed (Ambrosia artemisiifolia) population resistant to ALS- and PPO-inhibiting
herbicides. Weed Sci 60:335–344

Salas RA, Burgos NR, Tranel PJ, Singh S, Glasgow L, Scott RC, Nichols RL (2016) Resistance
to PPO-inhibiting herbicide in Palmer amaranth from Arkansas. Pest Manag Sci 72:864–869

Shoup DE, Al-Khatib K, Peterson DE (2003) Common waterhemp (Amaranthus rudis) resistance
to protoporphyrinogen oxidase-inhibiting herbicides. Weed Sci 51:145–150

Shyam C, Borgato EA, Peterson DE, Dille JA, Jugulam M (2021) Predominance of metabolic
resistance in a six-way-resistant Palmer amaranth (Amaranthus palmeri) population. Front
Plant Sci 11:2162

Strom SA, Hager AG, Seiter NJ, Davis AS, Riechers DE (2020) Metabolic resistance to S-
metolachlor in two waterhemp (Amaranthus tuberculatus) populations from Illinois, USA.
Pest Manag Sci 76:3139–3148

Takano HK, Beffa R, Preston C, Westra P, Dayan FE (2020) Glufosinate enhances the activity of
protoporphyrinogen oxidase inhibitors. Weed Sci 68:324–332

Tardif FJ, Powles SB (1999) Effect of malathion on resistance to soil-applied herbicides in a


population of rigid ryegrass (Lolium rigidum). Weed Sci 47:258–261

Tranel PJ, Riggins CW, Bell MS, Hager AG (2011) Herbicide resistances in Amaranthus
tuberculatus: A call for new options. J Agric Food Chem 59:5808–5812

Varanasi VK, Brabham C, Korres NE, Norsworthy JK (2019) Nontarget site resistance in Palmer
amaranth [Amaranthus palmeri (S.) Wats.] confers cross-resistance to protoporphyrinogen
oxidase-inhibiting herbicides. Weed Technol 33:349–354

141
Varanasi VK, Brabham C, Norsworthy JK (2018) Confirmation and characterization of non-target
site resistance to fomesafen in Palmer amaranth (Amaranthus palmeri). Weed Sci 66:702–
709

Vennapusa AR, Faleco F, Vieira B, Samuelson S, Kruger GR, Werle R, Jugulam M (2018)
Prevalence and mechanism of atrazine resistance in waterhemp (Amaranthus tuberculatus)
from Nebraska. Weed Sci 66:595–602

Walsh MJ, Stratford K, Stone K, Powles SB (2012) Synergistic effects of atrazine and mesotrione
on susceptible and resistant wild radish (Raphanus raphanistrum) populations and the
potential for overcoming resistance to triazine herbicides. Weed Technol 26:341–347

Wichert RA, Bozsa R, Talbert RE, Oliver LR (1992) Temperature and relative humidity effects
on diphenylether herbicides. Weed Technol 6:19–24

Woodyard AJ, Hugie JA, Riechers DE (2009) Interactions of mesotrione and atrazine in two weed
species with different mechanisms for atrazine resistance. Weed Sci 57:369–378

Wu C, Liu X, Wu X, Dong F, Xu J, Zheng Y, Zheng Y (2019) Simultaneous determination of


saflufenacil and three metabolites in five agriculture products using liquid chromatography—
tandem mass spectrometry. J Food Biochem 43:e12778

Wuerffel RJ, Young JM, Matthews JL, Young BG (2015) Characterization of PPO-inhibitor–
resistant waterhemp (Amaranthus tuberculatus) response to soil-applied PPO-inhibiting
herbicides. Weed Sci 63:511–521

Xu J, Wang XY, Guo WZ (2015) The cytochrome P450 superfamily: key players in plant
development and defense. J Integr Agric 14:1673–1686

142
Table 4.1 Waterhemp control after treatment with a detoxification inhibitor followed by
fomesafen application at 0, 20, or 60 g ai ha-1 accessed 14 days after treatment.
Waterhemp controla
Population Detoxification inhibitor 0 g ai ha-1 20 g ai ha-1 60 g ai ha-1
-----------------%-----------------
IN-RAN Both 12 abc 94 a 97 ab
Malathion 4c 67 bcd 100 a
NBD-Cl 23 ab 87 a 98 ab
None 0c 88 a 98 ab
IA-340 Both 2c 91 a 97 ab
Malathion 2c 93 a 100 a
NBD-Cl 3c 87 a 98 ab
None 0c 86 ab 100 a
IL-BRO1 Both 5 bc 47 efg 70 c-f
Malathion 3c 49 d-g 51 fg
NBD-Cl 1c 36 g 54 efg
None 0c 40 fg 58 d-g
IN-PIKE Both 3c 48 d-g 65 c-g
Malathion 2c 45 efg 66 c-g
NBD-Cl 3c 39 fg 59 d-g
None 0c 37 g 67 c-f
IL-WAS Both 6 abc 47 efg 73 cde
Malathion 0c 47 efg 59 d-g
NBD-Cl 1c 38 g 47 g
None 2c 35 g 54 fg
IN-DUB Both 25 a 75 abc 77 cd
Malathion 26 a 67 bcd 80 bc
NBD-Cl 13 abc 58 c-f 76 cd
None 0c 63 cde 76 cd
a
Means within a column followed by the same letter are not significantly different
according to Tukey’s HSD test at α=0.05.

143
Table 4.2. Waterhemp height and biomass after treatment with a detoxification inhibitor
followed by fomesafen application at 0, 20, or 60 g ai ha-1 accessed 14 days after treatment.
Heighta Biomassb

Detoxification 0 g ai 20 g ai 60 g ai 0 g ai 20 g ai 60 g ai

inhibitorc ha-1 ha-1 ha-1 ha-1 ha-1 ha-1

---------------cm--------------- ----------------%----------------

Both 23.1 b 12.2 b 9.3 b 84 c 34 b 28 ns

Malathion 24.4 ab 13.0 b 9.9 ab 94 ab 36 b 28

NBD-Cl 23.8 b 15.4 a 11.0 a 87 bc 45 a 34

None 26.0 a 14.9 a 10.2 ab 100 a 44 a 27


a
Means within a column followed by the same letter are not significantly different according
to Tukey’s HSD test at α=0.05.
b
log transformation was used for mean separation and presented as untransformed means for
ease of interpretation.
c
Effect of detoxification inhibitor presented and pooled over population due to insignificant
two-way interaction of detoxification inhibitor and population.

144
Table 4.3. Waterhemp height and biomass after treatment with a detoxification inhibitor
followed by fomesafen application at 0, 20, or 60 g ai ha-1 accessed 14 days after treatment.
Heighta Biomassb

Populationc 0 g ai ha-1 20 g ai ha-1 60 g ai ha-1 0 g ai ha-1 20 g ai ha-1 60 g ai ha-1

---------------cm--------------- ----------------%----------------

IN-RAN 21.6 b 6.3 c 1.8 c 77 b 14 c 9b

IA-340 26.2 a 7.4 c 2.3 c 93 a 15 c 7b

IL-BRO1 26.6 a 18.4 a 16.6 a 90 a 49 ab 42 a

IN-PIKE 25.7 a 18.5 a 14.3 a 95 a 58 a 41 a

IL-WAS 26.0 a 19.6 a 15.7 a 95 a 56 a 39 a

IN-DUB 19.9 b 12.9 b 9.7 b 98 a 46 b 37 a


a
Means within a column followed by the same letter are not significantly different according to
Tukey’s HSD test at α=0.05.
b
Log transformation was used for mean separation and presented as untransformed means for ease
of interpretation.
c
Effect of population presented and pooled over detoxification inhibitor due to insignificant two-
way interaction of detoxification inhibitor and population.

145
Figure 4.1. Isolated lesions on waterhemp leaves occurring after NBD-Cl application. Photo taken
at 48 h after treatment.

146
147

Figure 4.2. The IL-WAS waterhemp population was sprayed with detoxification inhibitor treatments followed by 60 g ha-1 of fomesafen.
Plants portrayed in the photographs are replications of the same treatment. Scale bar shown on the left of each photograph is equal to
12.6 cm
INVESTIGATIONS OF PPO-INHIBITOR RESISTANCE
IN WATERHEMP (AMARANTHUS TUBERCULATUS) LACKING
KNOWN RESISTANCE-CONFERRING MUTATIONS

5.1 Abstract

Protoporphyrinogen oxidase (PPO)-inhibitor resistance in Amaranthus species has evolved

beyond just the presence of the ΔG210 mutation. Multiple target-site mutations, double mutants,

and non-target-site resistance have been identified. We have identified several populations of

waterhemp [Amaranthus tuberculatus (Moq.) J.D.Sauer] with individual plants that display a

resistance phenotype to fomesafen that do not have the ΔG210 mutation. Our objective was to

determine if non-target-site resistance was present at low frequency in the observed populations.

A purifying selection and preliminary screen were conducted to identify resistant waterhemp

individuals from these populations and perform a seed increase. Resistant plants from four

populations were identified that did not have ΔG210 or R128 mutations. Full-length PPX1 and

PPX2 genes were sequenced, and there were no mutations that were likely to be causal for the

putative novel resistance mechanism. Progeny were grown and subjected to a dose-response assay.

At onset of plant symptom development from fomesafen, the populations with a putative novel

resistance mechanism displayed a unique phenotype. The putative novel resistant individuals

displayed initial phytotoxicity symptoms indistinguishable from the S through 5 to 7 d after

treatment, but eventually exhibited plant regrowth and survival (delayed regrowth phenotype). In

contrast, plants with the ΔG210 mutation had noticeably less phytotoxicity immediately upon

herbicide symptom onset. Models for lethal dose (LD) indicate a less robust R phenotype than

148
ΔG210, with LD50 ratios of only 1.9- to 6.1-fold vs 19- to 27-fold resistance in the ΔG210

populations. A target site expression experiment and lipid peroxidation experiment were

inconclusive, but provided some evidence of increased target-site expression or increased

antioxidant capacity as causal mechanisms, although no mechanisms have been fully ruled out.

This unique mechanism may be a component of a multigenic resistance complex, multiple

resistance from other sites of action, or a precursor to a more robust resistance phenotype.

5.2 Introduction

Protoporphyrinogen IX oxidase (PPO)-inhibiting herbicides are an important tool for weed

management in global agriculture (Hao et al. 2011). Despite their usage in multiple crops since the

1970s, PPO-inhibitor resistance is relatively uncommon, with 14 resistant species globally (Heap

2023). Most of these species have been confirmed as resistant in the past 10 years due to the

increased use of PPO inhibitors as a response to glyphosate resistance, particularly in waterhemp

[Amaranthus tuberculatus (Moq.) J.D.Sauer] and Palmer amaranth (Amaranthus Palmeri S. Wats)

(Legleiter et al. 2009, Young 2006). Waterhemp is a problematic weed native to the Midwest

capable of producing over 500,000 seeds per plant (Heneghan and Johnson 2017). It is one of the

most troublesome and most problematic weeds in the Midwest region primarily because of its

propensity to evolve herbicide resistance. To date, resistance to herbicides targeting seven different

sites of action have been confirmed in waterhemp (Heap 2023). With the evolution of glyphosate

resistance in waterhemp and other key species, growers responded by increasing usage of

alternative herbicides and use patterns, including PPO inhibitors applied both PRE and POST

(NASS 2020).

Resistance to PPO inhibitors is conferred primarily by target-site mutations. There are two

PPO isoforms in plants. PPO1 is encoded by the PPX1 gene and is localized to the chloroplast.

149
PPO2 is encoded by the PPX2 gene and is localized to the mitochondria and the chloroplast

(Lermontova et al. 1997, Watanabe et al. 2001). The first species confirmed as resistant was

waterhemp in 2001 (Shoup et al. 2003). The mechanism of resistance was later confirmed to be a

deletion of a glycine codon at the 210th position in the PPX2 gene (ΔG210), leading to a change in

binding site geometry (Dayan et al. 2010, Porri et al. 2022). Other resistance-conferring mutations

in the PPX2 gene include R128 (or equivalent) in waterhemp, Palmer amaranth, redroot pigweed

(Amaranthus retroflexus L.), common ragweed (Ambrosia artemisiifolia L.), and wild poinsettia

(Euphorbia heterophylla L.) (Giacomini et al. 2017, Jiang et al. 2021, Mendes et al. 2020, Nie et

al. 2019, Rousonelos et al. 2012). Palmer amaranth has also evolved resistance via a G399A

mutation in PPX2 (Rangani et al. 2019). In goosegrass [Eleusine indica (L.) Gaertn.], An A212T

mutation in the PPX1 gene has been shown to confer resistance to oxadiazon (Bi et al. 2019).

PPO-inhibitor resistance is not limited to target-site mutations. A survey of Palmer amaranth

accessions from around the mid-south revealed multiple populations, particularly in Mississippi,

that had no target-site resistance alleles, yet had a PPO-inhibitor-resistant phenotype (Noguera et

al. 2021). Another Palmer amaranth population from Arkansas was also shown to be resistant to

PPO inhibitors, but had no target-site mutations (Varanasi et al. 2018). In waterhemp, a population

susceptible to fomesafen, and resistant to carfentrazone-ethyl contained no resistance mutation in

the PPX2 gene, indicating the possibility of non-target-site resistance as well (Obenland et al.

2019).

In a recent survey of PPO-inhibitor-resistant waterhemp, we have observed multiple

populations that have a low frequency of plants exhibiting a resistant phenotype that is not

explained by ΔG210 or R128 mutations (Mansfield 2021, Nie et al. 2019). While the populations

had individuals possessing ΔG210 alleles, genotyping of individual plants did not explain the

150
resistant phenotype of multiple plants. The consistent and repeatable observation of these plants

in multiple experiments indicates that no genotyping errors occurred.

Potential causes for the observed phenotype are previously uncharacterized target-site

mutations, increased expression of the target enzymes, or non-target-site related responses (Gaines

et al. 2020). Non-target-site resistance is most often the evolved ability to metabolize the herbicide

via phase I or phase II enzymes (Dimaano and Iwakami 2021, Gaines et al. 2020, Jugulam and

Shyam 2019). However, other mechanisms are possible including herbicide sequestration, reduced

uptake or translocation, or ROS scavenging to prevent cell damage (Dayan et al. 2019, Gaines et

al. 2020). Non-target-site resistance is increasingly common in many weed species, including

waterhemp. Non-target-site resistance has been confirmed to HRAC groups 5, 15, and 27 in

waterhemp, while in other species, non-target-site resistance has been confirmed for nearly every

herbicide site of action group (Heap 2023, Ma et al. 2013, Strom et al. 2020). A potentially

disastrous scenario is a resistance event that confers cross-resistance to herbicides of multiple

modes of action. Dimaano et al. (2020) demonstrated a single CyP450 enzyme capable of

metabolizing multiple herbicide groups when expressed in Escherichia coli. Many herbicides of

different chemical families possess similar moieties that may be recognizable by a single

overexpressed CyP450 or GST enzyme. For example, fomesafen is a diphenyl-ether herbicide. In

addition to other diphenyl-ether PPO inhibitors, 13 herbicides from 4 different sites of action have

ether linkages of phenyl- or heterocyclic-ring structures (Shaner 2014).

Investigations of resistance mechanisms in our surveyed populations have been relatively

narrow. Nie et al. (2018) sequenced the R128 region of multiple populations and confirmed the

absence of any mutations in that region of the PPX2 gene. Mansfield (2021) investigated the

potential for ROS scavenging as a potential contributor to PPO-inhibitor resistance and did not

151
observe any consistent trends. However, given the low frequency of these anomalous resistant

individuals, it is possible that a measurable effect was diluted with susceptible or ΔG210

individuals. The objectives of this research are to isolate and confirm PPO-resistant individuals in

select populations of waterhemp from the 2016 survey, confirm the presence or absence of

potential resistance conferring target-site mutations, perform a purifying selection, and

characterize the resistance response in a more purified population.

5.3 Materials and Methods

5.3.1 Preliminary Purifying Screen

A greenhouse experiment was conducted in the spring of 2022 to investigate anomalous

resistant individuals in select populations of waterhemp sourced from a 2016 survey of PPO-

inhibitor resistance and an Indiana field population with suspected multiple resistance to several

herbicide groups. The populations from the survey were sourced from Cerro Gordo (IA-340)

(43.17.122, -93.20.368), Chickasaw (IA-369) (43.04.682, -92.38.010), and Wayne (IA-358)

(40.754844, -93.286675) Counties in Iowa, USA. The Indiana field population was sourced from

near Francesville, IN (Fran) (40.91, -86.89). The four subject populations were compared to a

known PPO-inhibitor susceptible population sourced from near Desoto, IL (Des) and a known

resistant population with nearly 100% frequency of the ΔG210 mutation sourced from near Carlyle,

IL (Car).

Seeds of each population were treated with 20% sodium hypochlorite solution for 10

minutes and sown for germination in flats containing a peat moss-based potting mix (Sunshine®

Mix#5; Sun Gro Horticulture, Agawa, MA). Approximately 1 week after sowing, seedlings at the

1 leaf stage were transplanted into 10 cm square pots filled with a mixture of two parts peatmoss-

152
based potting mix and one part sand. Pots were watered as needed and fertilized weekly with a

micro- and macronutrient fertilizer (Jack’s Classic Professional 20-20-20, JR Peters Inc.,

Allentown PA) throughout the course of the experiments. Greenhouse environmental conditions

included 16:8 light/dark photoperiods, where natural light was supplemented with high-pressure

sodium bulbs delivering 1100 μmol m-2 s-1 photon flux during daylight hours, and day/night

temperatures of 30 and 25C, respectively.

Plants were grown until the 7 to 9 leaf stage, at which point, plants from each population

were treated with one of four rates of fomesafen (Flexstar®, Syngenta Crop Protection, LLC,

Greensboro, NC) with crop oil concentrate (Prime Oil®, Winfield Solutions, LLC, St. Paul, MN)

at 1% v/v. Herbicide treatments were applied using a track-mounted research sprayer (Generation

III Research Sprayer, DeVries Manufacturing, Hollandale, MN) calibrated to deliver 140 L ha-1 at

207 kPa via an even flat fan XR8002E (TeeJet Technologies, Glendale Heights, IL) nozzle. Rates

of fomesafen were 0, 20, 60, or 100 g ai ha-1 in the first experimental run. Due to greater than

desired phytotoxicity, the 100 g ha-1 rate was replaced with 40 g ha-1 in the second run. Each

treatment had 10 replications in each experimental run.

Visual estimates of waterhemp control were recorded as well as a qualitative assignment

of a resistant (R) or susceptible (S) phenotype at 7 and 14 days after treatment. At 14 days after

treatment, leaf tissue from all surviving plants was collected for DNA extraction and qPCR assays

for the ΔG210 mutation. Based on phenotypic assignment, (R or S) as well as PCR genotype data,

plants were separated into 3 groups of susceptible (ΔG210 absent and S phenotype), ΔG210

resistant (ΔG210 present), and putative novel R mechanism (ΔG210 absent and R phenotype).

Data were subjected to ANOVA and control means were separated using Tukey’s HSD (α=0.05).

153
DNA was extracted using a modified cetyltrimethylammonium bromide (CTAB) protocol

originally designed by Saghai-Maroof et al. (1984). A TaqMan® SNP genotyping assay developed

by Wuerffel et al. (2015a), modified by Giacomini et al. (2017), and sourced from ABI (Applied

Biosystems Inc. Grand Island, NY) was used to genotype each sample as homozygous susceptible,

homozygous resistant, or heterozygous based on fluorescence of specific fluorophores in the

presence of a PCR product generated with forward primer (5′- TGATTATGTT

ATTGACCCTTTTGTTGCG) and reverse primer (5′ GAGGGAGTATAATTTATTTA

CAACCTCCAGAA). For each sample, a 10 µl reaction was prepared with 4.8 µl of nuclease free

nano-pure water, 2 µl of GoTaq Flexi buffer, 1.2 µl of 25 mM MgCl2, 0.4 µl of 10 mM dNTP, 0.5

µl of 20X primers and TaqMan® probes, 0.1 µl of GoTaq Flexi polymerase (5 U µl−1), and 1 µl

of genomic DNA. Reactions were amplified using a CFX384 RT-PCR detection system (Bio-Rad

Laboratories, Hercules, CA) with the following cycle conditions: 2 min at 95 C; 39 cycles of 95 C

for 15 s and 60 C for 1 min; followed by a plate read after every cycle.

All plants with the putative novel R mechanism as well as two non-treated plants from each

population were repotted into four-liter pots filled with the same 2:1 potting mixture. Each

population was isolated to its own greenhouse room with no other flowering stage Amaranthus

plants. Plants were fertilized to encourage growth and allowed to randomly mate with plants of the

same population. If any of the non-treated plants were male, they were discarded, but females were

allowed to be pollinated and produce seed. By chance, only female plants were acquired from the

IA-358 population that had no ΔG210 alleles and had the putative novel R mechanism, so those

plants were pollinated by the IA-340 population. Therefore, all progeny of the IA-358 population

are actually a one-way cross of IA-340 males and IA-358 females. At seed maturity, plants were

harvested, dried, threshed, and sieved to obtain seed from each plant. The same quantity of seed

154
from each plant (excluding non-treated) was blended to create the progeny populations used for

subsequent experiments.

5.3.2 Gene Sequencing and PCR Reactions

Prior to fomesafen application, an immature leaf was collected from each plant and placed

into a 1.5ml tube, flash frozen in liquid N2 and stored at -80C until RNA extraction was performed.

Total RNA was extracted from all plants having the putative novel R mechanism as well as one

known resistant plant from the Car population, one known susceptible plant from each subject

population and the known S population, and the female untreated plants from each subject

population. RNA was extracted using the RNeasy plant mini kit (Qiagen, Hilden, Germany) then

cleaned using RNA Clean and Concentrator kit (Zymo Research, Irvine, CA). A cDNA library for

each plant was created using Protoscript® First Strand cDNA Synthesis kit (New England Biolabs,

Ipswich, MA). Each cDNA library was used as a template for PCR reactions to amplify the coding

sequence for the PPX1 and PPX2 genes. Each reaction consisted of 25 µl Q5® Hot Start High-

Fidelity 2X Master Mix (New England Biolabs, Ipswich, MA), 19 µl of nuclease free water, 1 ul

of template cDNA, and 2.5 µl each of forward and reverse primers diluted to 10 µM (Integrated

DNA Technologies, Skokie, IL). Primer sequences are as follows: PPX1 (forward, 5′-

ATGAGTGCGATGGCGTTATCGAGC-3′; and reverse, 5′- CTACTTATCTTTGTACTGTG

AGAGGAAATCC-3′); PPX2 (forward, 5′- ATGGTAATTCAATCCATTACCCACCTT

TCACC-3′; and reverse, 5′- TTACGCGGTCTTCTCATCCATCTTCACG-3′). PCR reactions

were performed in a SimpliAmp thermal cycler (Applied Biosystems, Waltham, MA) with the

following conditions: 95 C for 30 s; 30 cycles of 98 C for 10 s, 60 C for 30 s, and 72 C for 50s;

and 72 C for 2 min. A sample of each PCR product was placed on a gel and confirmed to have a

single product of between 1600 and 1700bp. The PCR product was purified using Wizard® SV

155
gel and PCR Clean-Up System (Promega, Madison, WI). Purified PCR products were submitted

to Purdue Genomics core for sequencing using Illumina MiSeq system (Illumina, San Diego, CA).

WideSeq libraries were constructed using the Illumina DNA Prep Kit, followed by sequencing on

a 500 cycle MiSeq cassette. Obtained sequences were aligned in Benchling (https://benchling.com)

with previously submitted waterhemp and Palmer amaranth PPX1 and PPX2 sequences to identify

any amino acid codon polymorphisms.

5.3.3 Dose-Response Assays

Dose-response experiments were conducted in the fall of 2022 on the progeny populations

and compared to the same known R and S populations described previously, as well as a second

known PPO-inhibitor resistant population sourced from Washington County, IL (Was) (38.432254,

-89.628843). Seeds were grown as described previously with the exception of being transplanted

into 164 ml cone-tainers (Ray Leach SC-10 Super Cell Cone-tainers; Stuewe & Sons, Tangent,

OR) rather than 10 cm square pots. Fomesafen was applied as described previously at the 5 to 7

leaf stage at one of seven fomesafen rates in addition to a nontreated treatment in a half-log dose

structure. The S (Des) population received rates from 0.2 to 200 g ha-1 while all other populations

received 0.63 to 632 g ha-1. Visual assessment of waterhemp control was taken at 3, 7, and 14 days

after treatment. Also, at 14 days after treatment, aboveground tissue was clipped at the soil surface

and dried at 60C for dry biomass measurement. Biomass data were converted to a % of the non-

treated control for model fitting. Control data were used to assign a binary value of 0 or 1 for alive

or dead to model lethal dose. Data were used to fit 3 parameter log logistic models (Equation I):
𝑑−𝑐
𝑓(𝑥) = [1]
1+exp (𝑏(log(𝑥)−log (𝑒)))

156
where b is the slope of the curve, c is the lower asymptote, d is the upper asymptote, and e is the

herbicide rate required to produce 50% effective dose, which were GR50 or LD50 values. Models

were fitted using the drc package in R 3.5.2 (Knezevic et al. 2007, Ritz et al. 2015). Data were

pooled over experimental runs after validating lack of run interactions with ANOVA (α = 0.05).

Calculated GR50 and LD50 values were used to quantify resistance for each population. Statistical

comparisons were made using the compParm function within the drc package.

5.3.4 Gene Expression Assay

To test for expression differences of target-site genes between populations, qPCR assays

were conducted similar to Montgomery et al. (2020). Young leaf tissue was collected from 12

plants of each population utilized in the dose-response experiments. After tissue collection,

waterhemp plants were sprayed with fomesafen at 20 g ha-1 and assigned control ratings at 14 days

after treatment. Tissue was used for RNA extraction and creation of cDNA libraries as described

previously. The cDNA was used as template for qPCR gene expression assays. Primer sequences

utilized were identical to Montgomery et al. (2020) and were as follows: PPX1 (forward, 5’ -

TGCTTACTATGGCGGTTGAC-3’ ; and reverse, 5’ -CCGTAGGTTTGGAAGGAACA-3’ );

PPX2 (forward, 5’ -TGTGGTTGTCACTGCTCCA-3’ ; and reverse, 5’ -GGGGATAAG

AACTCCGAAGC-3’); EF1-alpha (forward, 5’ -CACGCTTTGCTTGCTTTCACTCT TG-3’ ;

and reverse, 5’ - CGATTTCATCGTACCTAGCCTTGGAGTAC -3’). The primers were designed

for Palmer amaranth, but all three primer sets were confirmed to also be an exact match for

waterhemp based on the obtained PPX1 and PPX2 sequences as well as the publicly available

waterhemp reference genomes. Each qPCR reaction consisted of 5 μl iTaq Universal SYBR Green

Supermix (Bio-Rad, Hercules, CA), 1 μl forward primer and 1 μl reverse primer, diluted to 10 μM

(Integrated DNA Technologies), 1 μl template cDNA, and 2 μl water. Assays were conducted with

157
a Quant Studio 6 flex real time PCR system with the following thermoprofile: 95 C for 30 s, 35

cycles of 95C for 15 s and 60C for 30 s followed by a melt curve analysis. The experiment utilized

6 biological replicates and 3 technical replicates. Relative expression was calculated using the

ΔΔCT method. Expression of each target-site gene was normalized to the S population and mean

fold expression differences were calculated. Data were subjected to ANOVA and means were

separated using Tukey’s HSD (α=0.05).

5.3.5 Lipid Peroxidation Assay

To quantify lipid peroxidation, an assay of malondialdehyde, a byproduct of lipid

peroxidation, was conducted based on methods described by Harre et al. (2018) and Mansfield

(2021), which were adapted from methods by Zhang and Kirkham (1996). Waterhemp plants from

Fran, IA-369, Des, and Car were grown as described previously. Plants were sprayed at the 5 to 7

leaf stage with fomesafen at 20 g ai ha-1. At 0, 12, 24, and 48h after fomesafen treatment, plants

were segmented into three component tissue types of stem, mature leaf, and meristem/immature

leaf. Each tissue type was placed in an aluminum foil envelope, flash frozen in liquid N2 and stored

at -80C until MDA assay was conducted.

Trichloroacetic acid (TCA) solution (0.1%) was added to fresh waterhemp tissue ground in

liquid N2 with a mortar and pestle at approximately 1 ml per 0.1 g of tissue. Stem and leaf samples

typically had a mass of approximately 0.1 g whereas meristem samples only had mass of

approximately 0.05g. Therefore, stem and leaf sample vials received 1 ml of TCA solution and

meristem sample vials received 0.5 ml of TCA solution. Exact mass of tissue in each sample was

recorded for later calculations. Tissue-TCA mixture was vortexed vigorously and centrifuged at

10,000 g for 10 minutes at room temperature. The supernatant was collected and 200 µl was added

to 800 µl of 0.5% thiobarbituic acid dissolved in 20% TCA solution. Samples were placed in a

158
95C water bath for 30 minutes, then rapidly chilled in an ice bath for 5 minutes. After cooling,

samples were centrifuged at 10,000 g for 10 minutes. Supernatant was placed in a 350 µl

polypropylene 96 well plate. Absorbance at 532 and 600 nm was recorded with a Multiscan™ Sky

spectrophotometer (Thermo Scientific, Waltham, MA). Nonspecific absorption at 600 nm was

subtracted from the reading at 532 nm and the MDA concentration was calculated using an

extinction coefficient of 155 mM-1 cm-1. The experiment was repeated twice and utilized a split

plot design where main plot was collection timing and sub plot was population. Each tissue type

and herbicide rate were analyzed separately as a two-way ANOVA with population and harvest

timing as main effects and replication nested within experimental run as a random effect. Means

were separated using Fishers protected LSD (α = 0.05).

5.4 Results and Discussion

5.4.1 Preliminary Purifying Screen

To identify resistant individuals, waterhemp was treated with one of four fomesafen rates.

After assigning each plant to a resistance group of ΔG210, putative novel R mechanism, and

susceptible, waterhemp population was not a significant term in the model, so waterhemp control

is pooled across populations. The number of waterhemp assigned to each group from each

population is shown in Table 5.1. Control increased with increasing herbicide rates (Figure 5.1).

At the 20 and 40 g ha-1 fomesafen rates, individuals with putative novel R mechanism were

controlled less than susceptible individuals, but more than ΔG210 individuals indicating an

intermediate resistance phenotype. At higher herbicide rates, high levels of control and waterhemp

lethality resulted in fewer differences between treatments.

159
5.4.2 Target-Site Gene Sequencing

From the individuals with putative novel R mechanism in the screening, a number of the

most resistant plants were kept for seed production and gene sequencing. The sequences used for

basis of comparison were from a number of published sequences from Palmer amaranth and

waterhemp (Giacomini et al. 2017, Patzoldt et al. 2006, Rangani et al. 2019, Varanasi et al. 2018),

as well as a known susceptible and known ΔG210 from the Des and Car populations. From the

nearly full-length target gene sequences obtained, a number of deduced amino acid polymorphisms

were observed (Table 5.2, Table 5.3). However, none of the polymorphisms explained the resistant

phenotype in either PPX1 or PPX2 genes. Special attention was given to the R128, and G398

(equivalent to Palmer amaranth G399) regions of PPX2 as well as the A235 region (equivalent to

A212 in goosegrass) of PPX1. All of the observed amino acid polymorphisms were present in

known susceptible plants from the current experiment, present in a previous sequence submission,

or the mutation was an extremely similar functioning residue which is less likely to cause

alterations to the protein.

An interesting observation was made regarding the sequence conservation of the PPX2 gene.

Similar to the findings of Nie et al. (2019), there were distinct types of PPX2 alleles that had a

specific subset of mutations that were also present in other Amaranthus species, but not other

waterhemp accessions (Figure 5.2). In fact, many of the observed mutations recorded in Table 5.2

are actually wild-type Palmer amaranth amino acid residues. The most conserved feature of the

novel waterhemp allele is a glycine insertion at position 251. This glycine insertion was apparently

a prerequisite to the majority of the other observed mutations. The same pattern was not observed

with PPX1. There were drastically fewer mutated positions overall compared to PPX2, and very

few of those mutations matched Palmer amaranth or other Amaranthus species’ PPX1 sequences.

Additionally, nearly all of the plants with Palmer amaranth-type alleles were heterozygous for the
160
wild-type allele and the Palmer amaranth allele. Presence of the Palmer amaranth allele did not

appear to be related to our putative novel R phenotype. There were several resistant individuals

that were homozygous wild type and several known susceptible plants that were heterozygous.

This observation indicates a recent invasion of the novel genotype that may not be related to PPO-

inhibitor resistance. The different alleles could be from the recent hybridization between common

and tall waterhemp subspecies (Kreiner et al. 2021, Waselkov et al. 2020, Waselkov and Olsen

2014). Another potential contributor may be spread of resistant ALS alleles which are linked to the

PPX2 gene (Tranel et al. 2017). An experiment utilizing bioinformatics such as a genome wide

associative study could reveal a broader invasion and evolutionary picture related to the spread of

herbicide resistance.

5.4.3 Dose-Response Assays

A dose-response assay of the progeny generation of the four putative novel R mechanism

populations compared to a known susceptible and two known resistant populations was conducted.

At 24 hours after application, there was an apparent absence of a resistant phenotype in the four

subject populations whereas the two known resistant populations displayed the characteristic lack

of immediate phytotoxicity (Figure 5.3). The subject populations incurred an equal amount of

initial phytotoxicity as the susceptible population and that lasted 5 to 7 days after treatment (Figure

5.4). However, by 14 days after treatment, substantial regrowth occurred in the subject populations,

but not the susceptible population, particularly at the 20 and 60 g ha-1 dose (Figure 5.5). The initial

phytotoxicity followed by regrowth resulted in an insignificant R/S ratio for the three IA

populations when fitted with biomass data (0.6- to 1.2- fold) (Figure 5.6, Table 5.4). Only the

Francesville population had a significant R/S ratio of 1.8- fold while the known resistant

populations were 4.9- and 9.7- fold resistant. Even though Francesville R/S ratio rose to statistical

161
significance, this observed resistance phenotype may not necessarily be meaningful in a field

setting. However, with models fitted using survival data, there is a drastic increase in the observed

R/S ratio for all populations, particularly the four subject populations (Figure 5.7, Table 5.4). R/S

ratios ranged from 1.9- to 8.6- fold in the subject populations while known resistant populations

were 19- and 27- fold resistant. Francesville had the greatest resistance of subject populations in

the waterhemp survival model as well.

With the unique phenotype of the subject populations, the insignificant R/S ratios are not

entirely unexpected. First, these populations are known to have a resistant phenotype that is less

robust than the ΔG210, so the maximum R/S ratio we could have observed was less than what was

observed in the known resistant population (4.9-fold). Secondly, the experiment was not designed

to capture the observed differences. The regrowth phenotype that we observed in the dose-response

experiment was not an expected response. Target-site resistance mutations cause a noticeable lack

of phytotoxicity (under our treatment conditions) that allows them to keep growing while

susceptible individuals suffer growth reduction. Unlike target-site resistant populations, the subject

populations suffered the same initial growth reduction as susceptible plants and only had 7 to 10

days to regenerate biomass. Biomass accumulation can be affected by many factors including

experimental design choices and growing conditions (Burgos et al. 2013, Schafer et al. 2012).

Perhaps a superior design for the dose-response experiment could have implemented a longer

evaluation window.

Even though GR50 is the preferred resistance metric for POST herbicide dose responses, it

does not adequately capture the resistance that we observed (Burgos et al. 2013). On the other

hand, LD50 is another common metric for quantifying resistance and arguably the most important.

In a field setting, surviving the herbicide application is the most important objective because weeds

162
have the ability to recover from defoliation and phytotoxicity in order to produce seed (Baysinger

and Sims 1992, Hartzler and Battles 2001, Mager et al. 2006). In these same populations, plants

that incurred 85 to 90% injury were observed to grow to 1 meter or more and produce large

quantities of seed. One of the mechanisms of glyphosate resistance in giant ragweed (Ambrosia

trifida L.) is a triggered rapid necrosis response (RR) where ROS generation causes defoliation

and prevents glyphosate translocation (Van Horn et al. 2018). Harre et al. (2017) observed that

R/S ratios of RR giant ragweed were greater than non-RR giant ragweed in a 21 day experiment.

Presumably, the observed magnitude of resistance would have been smaller if plants were not

given enough time to recover from the defoliation.

Out of the four populations, Fran, followed by IA-369, had the most robust resistance,

whereas IA-358 barely showed any resistance. The differences in populations are impossible to

know without further studies, but we speculate that genetic background is contributing to our

results. Fran is a population that is currently being investigated for resistance to multiple sites of

action, including HPPD inhibitors, dicamba, and atrazine (Bland et al. 2022). Perhaps genes

implicated in resistance to other chemistries are playing a role in the resistance observed in the

Fran population. Field histories of the IA populations are unknown, so we have less insight into

those populations. However, since the IA-358 is not a pure population, but rather a hybrid of IA-

358 and IA-340, perhaps the outcross with another population disrupted the resistance phenotype

because the trait is potentially recessive, less dominant, or multigenic.

Even with the more robust quantification of resistance with LD50 values, the magnitude of

resistance is still relatively weak. It is possible that the resistance trait is part of a multigenic

resistance complex or is a precursor to a more robust resistance mechanism. Secondary partial

resistance to PPO inhibitors in Amaranthus species has been theorized by several authors

163
(Montgomery et al. 2020, Noguera et al. 2021, Wuerffel et al. 2015b) and multiple mechanisms of

resistance have been documented in the same redroot pigweed individuals (Cao et al. 2022). The

ΔG210 mutation was present in 3 of the 4 populations during the initial screening experiment and

PPX2 sequence data indicates high heterozygosity and recent invasion. Perhaps these populations

were recently invaded by a multigenetic PPO-resistant biotype. As a select few plants reproduce,

the ΔG210 and secondary resistance mechanism would segregate producing the few individual

resistant plants observed in these populations.

5.4.4 Gene-Expression Assay

A constitutive gene expression assay of PPX1 and PPX2 genes was conducted to rule out

potential target-site resistance not caused by a target-site mutation. There were no differences in

expression of PPX2 between the susceptible population and any other population (Figure 5.8).

However, the IL-WAS population, previously characterized as being highly resistant in

comparison to other populations with ΔG210, had significantly greater expression than IA-358 by

2.2 fold.

The F test for PPX1 expression was not significant at the typical α of 0.05 (P=0.0771).

However, there is some indication of increased PPX1 expression in the IA-369 population.

Expression of IA-369 ranged from 0.7 to 4 fold relative to Des population with an average of 2.0

fold. The other subject populations had an average PPX1 expression of 1.1- to 1.2- fold. Other

groups that have assayed PPX1 and PPX2 gene expression have reported a high variance similar

to the current experiment, but no others have reported any remote indication of increased

expression (Borgato 2022, Montgomery et al. 2020, Varanasi et al. 2018). Both glyphosate and

glufosinate resistance in Amaranthus species is caused by increased expression of the respective

target-site genes of EPSPS and GS2 (Carvalho-Moore et al. 2022, Chatham et al. 2015, Gaines et

164
al. 2010). The increased expression of each respective target-site in resistant plants is upwards of

10-fold or greater. In the case of glyphosate, there is also a strong linear relationship between

EPSPS expression and resistance phenotype (Gaines et al. 2010). In the present study, we observed

only a 2-fold increase in one of the target-site genes, which did not have a linear relationship with

waterhemp control (data not shown). Also, there was no evidence of increased expression in the

other three populations with a similar phenotype. Despite these inconsistencies, there could be

multigenic factors that are contributing to the resistance phenotype. For instance, the total target

enzyme content is not necessarily the same as expression. Secondly, the interaction of PPX1 and

PPX2 at the subcellular level is not fully characterized, especially in the context of herbicide action

(Dayan et al. 2018). We find that this experiment is inconclusive and further studies should

investigate the potential contribution of increased PPX1 expression in IA-369 population toward

PPO-inhibitor resistance.

5.4.5 Lipid-Peroxidation Assay

A malondialdehyde (MDA) assay was conducted as an objective measure of phytotoxicity

in the period after fomesafen application. Malondialdehyde is a byproduct of lipid peroxidation

and can be used as a direct measure of damage from PPO-inhibitor activity. Mature leaf tissue

accumulated more MDA than meristem and young leaf tissue (Figure 5.9). There was no

meaningful change in the detected level of stem MDA content, so that data is excluded from

analysis. Des and Car MDA contents were different at all treatment collection timings. At 12 hours

after treatment, IA-369 and Fran had more MDA content than Car by 23 to 37%, but less MDA

than Des by 26 to 32% in both tissue types. Later collection timings indicated MDA content in

Fran and IA-369 approaching that of Des by 24 HAT in meristem tissue and by 48 HAT in leaf

tissue. Overall, results indicate less rapid lipid peroxidation as shown by MDA content in

165
Francesville and IA-369 populations in the first 12 to 24 hours after fomesafen application, but the

phytotoxicity eventually reaches similar levels as Des. This less rapid MDA accumulation is a new

finding that was not noticeable by visual assessment. The delayed lipid peroxidation relative to the

susceptible may allow the plants time to induce a survival response that allows for survival and

regrowth. Results indicate that despite having a similar appearance in the first hours after

fomesafen treatment, there is a difference in physiological response.

Further research is required to determine the cause of these observations, but some

inferences can be drawn from the pattern of MDA accumulation. First, fomesafen metabolism is

likely not contributing to the resistance phenotype. If herbicide metabolism were occurring, then

the reduced MDA content would have persisted through the entire evaluation period, rather than

eventually rising to similar levels as the susceptible. Second, the mechanism must be a process

that delays phototoxicity. Delayed phytotoxicity is consistent with increased expression of a

susceptible target enzyme. Given that there was some evidence of increased expression in one of

the tested populations (IA-369), this hypothesis merits further investigation. ROS scavenging

could also contribute to less rapid phytotoxicity. Antioxidant enzymes are part of plants’ normal

metabolism (Gill and Tuteja 2010). Dayan et al. (2019) demonstrated that increased antioxidant

activity in plants can confer PPO-inhibitor resistance. Mansfield (2021) investigated increased

antioxidant activity in several populations, including the parent generations of the IA-369, IA-340,

and IA-358 populations. Results indicated no evidence of antioxidant activity contributing to PPO

resistance. However, given the number of populations tested and the low frequency of the resistant

phenotype in the subject populations, the experiment was likely not adequately designed to identify

and quantify differences in ROS in segregating populations. Experiments would have been better

suited for the progeny generation such as we described previously.

166
In this manuscript, we present a report of multiple waterhemp populations with a

significant but low-level fomesafen resistance phenotype. We found no evidence of causal target-

site mutations in either the PPX1 or PPX2 genes, but observed inconclusive evidence of increased

expression of PPX1 in one of the subject populations that warrants further investigation. A lipid

peroxidation assay indicated that Francesville and IA-369 populations incurred less lipid

peroxidation within the first 24 hours after treatment but eventually reached similar levels as the

susceptible population. Future experiments should utilize omics approaches, such as RNA seq, to

develop leads for further physiological experiments.

5.5 Acknowledgements

We want to thank Purdue Genomics core for target gene sequencing.

5.6 Literature Cited

Baysinger JA, Sims BD (1992) Giant Ragweed (Ambrosia trifida) control in soybean (Glycine
max). Weed Technol 6:13–18

Bi B, Wang Q, Coleman JJ, Porri A, Peppers JM, Patel JD, Betz M, Lerchl J, McElroy JS (2019)
A novel mutation A212T in chloroplast protoporphyrinogen oxidase (PPO1) confers
resistance to PPO inhibitor oxadiazon in Eleusine indica. Pest Manag Sci 76:1786–1794
Bland CR, Young BG, Johnson WG, (2022) Characterization of six herbicide resistance traits in
an Indiana waterhemp population. Proceedings of the 2022 North Central Weed Science
Society Annual Meeting. St.Louis, MO: North Central Weed Science Society
Borgato EA (2022) Exploring the mechanisms of Palmer amaranth resistance to
protoporphyrinogen oxidase-inhibitor herbicides, dynamics of gender, and management
strategies. PhD dissertation. Manhattan KS: Kansas State University

Burgos NR, Tranel PJ, Streibig JC, Davis VM, Shaner D, Norsworthy JK, Ritz C (2013) Review:
confirmation of resistance to herbicides and evaluation of resistance levels. Weed Sci 61:4–
20

167
Cao Y, Huang H, Wei S, Lan Y, Li W, Sun Y, Wang R, Huang Z (2022) Target gene mutation and
enhanced metabolism confer fomesafen resistance in an Amaranthus retroflexus L.
population from China. Pestic Biochem Physiol 188:105256

Carvalho-Moore P, Norsworthy JK, González-Torralva F, Hwang JI, Patel JD, Barber LT, Butts
TR, McElroy JS (2022) Unraveling the mechanism of resistance in a glufosinate-resistant
Palmer amaranth (Amaranthus palmeri) accession. Weed Sci 70:370–379

Chatham LA, Wu C, Riggins CW, Hager AG, Young BG, Roskamp GK, Tranel PJ (2015) EPSPS
gene amplification is present in the majority of glyphosate-resistant Illinois waterhemp
(Amaranthus tuberculatus) populations. Weed Technol 29:48–55

Dayan FE, Barker A, Dayan L, Ravet K (2019) The role of antioxidants in the protection of plants
against inhibitors of protoporphyrinogen oxidase. React Oxyg Species 7:55–63

Dayan FE, Barker A, Tranel PJ (2018) Origins and structure of chloroplastic and mitochondrial
plant protoporphyrinogen oxidases: implications for the evolution of herbicide resistance.
Pest Manag Sci 74:2226–2234

Dayan FE, Daga PR, Duke SO, Lee RM, Tranel PJ, Doerksen RJ (2010) Biochemical and
structural consequences of a glycine deletion in the α-8 helix of protoporphyrinogen oxidase.
Biochim Biophys Acta - Proteins Proteomics 1804:1548–1556

Dimaano NG, Iwakami S (2021) Cytochrome P450-mediated herbicide metabolism in plants:


current understanding and prospects. Pest Manag Sci 77:22–32

Dimaano NG, Yamaguchi T, Fukunishi K, Tominaga T, Iwakami S (2020) Functional


characterization of cytochrome P450 CYP81A subfamily to disclose the pattern of cross-
resistance in Echinochloa phyllopogon. Plant Mol Biol 102:403–416

Gaines TA, Duke SO, Morran S, Rigon CAG, Tranel PJ, Küpper A, Dayan FE (2020) Mechanisms
of evolved herbicide resistance. J Biol Chem 295:10307–10330

Gaines TA, Zhang W, Wang D, Bukun B, Chisholm ST, Shaner DL, Nissen SJ, Patzoldt WL,
Tranel PJ, Culpepper AS, Grey TL, Webster TM, Vencill WK, Sammons RD, Jiang J, Preston
C, Leach JE, Westra P (2010) Gene amplification confers glyphosate resistance in
Amaranthus palmeri. Proc Natl Acad Sci U S A 107:1029–1034

168
Giacomini DA, Umphres AM, Nie H, Mueller TC, Steckel LE, Young BG, Scott RC, Tranel PJ
(2017) Two new PPX2 mutations associated with resistance to PPO-inhibiting herbicides in
Amaranthus palmeri. Pest Manag Sci 73:1559–1563

Gill SS, Tuteja N (2010) Reactive oxygen species and antioxidant machinery in abiotic stress
tolerance in crop plants. Plant Physiol Biochem 48:909–930

Hao GF, Zuo Y, Yang S-G, Yang G-F (2011) Protoporphyrinogen oxidase inhibitor: An ideal
target for herbicide discovery. Chim Int J Chem 65:961–969

Harre NT, Nie H, Jiang Y, Young BG (2018) Differential antioxidant enzyme activity in rapid-
response glyphosate-resistant Ambrosia trifida. Pest Manag Sci 74:2125–2132

Harre NT, Nie H, Robertson RR, Johnson WG, Weller SC, Young BG (2017) Distribution of
herbicide-resistant giant ragweed (Ambrosia trifida) in Indiana and characterization of
distinct glyphosate-resistant biotypes. Weed Sci 65:699–709

Hartzler RG, Battles BA (2001) Reduced fitness of velvetleaf (Abutilon theophrasti) surviving
glyphosate. Weed Technol 15:492–496

Heap I (2023) International survey of herbicide resistant weeds. http://www.weedscience.org.


Accessed March 25, 2023

Heneghan JM, Johnson WG (2017) The growth and development of five waterhemp (Amaranthus
tuberculatus) populations in a common garden. Weed Sci 65:247–255

Jiang X, Ju Q, Guo W, Li L, Qu C, Qu M (2021) Target-site basis for fomesafen resistance in


redroot pigweed (Amaranthus retroflexus) from China. Weed Sci 69:290–299

Jugulam M, Shyam C (2019) Non-target-site resistance to herbicides: recent developments. Plants


8:417

Knezevic SZ, Streibig JC, Ritz C (2007) Utilizing R software package for dose-response studies:
The concept and data analysis. Weed Technol 21:840–848

Kreiner JM, Tranel PJ, Weigel D, Stinchcombe JR, Wright SI (2021) The genetic architecture and
population genomic signatures of glyphosate resistance in Amaranthus tuberculatus. Mol
Ecol 30: 5373-5389

169
Legleiter TR, Bradley KW, Massey RE (2009) Glyphosate-resistant waterhemp (Amaranthus
rudis) control and economic returns with herbicide programs in soybean. Weed Technol
23:54–61

Lermontova I, Kruse E, Mock HP, Grimm B (1997) Cloning and characterization of a plastidal
and a mitochondrial isoform of tobacco protoporphyrinogen IX oxidase. Proc Natl Acad Sci
U S A 94:8895–8900

Ma R, Kaundun SS, Tranel PJ, Riggins CW, McGinness DL, Hager AG, Hawkes T, Mc EI,
Riechers DE (2013) Distinct detoxification mechanisms confer resistance to mesotrione and
atrazine in a population of waterhemp. Plant Physiol 163:363–377

Mager HJ, Young BG, Preece JE (2006) Characterization of compensatory weed growth. Weed
Sci 54:274–281

Mansfield BC (2021) Characterization of protoporphyrinogen oxidase (PPO) herbicide resistance


in tall waterhemp (Amaranthus tuberculatus). MS thesis. West Lafayette IN: Purdue
University. 158 p

Mendes RR, Takano HK, Adegas FS, Oliveira RS, Gaines TA, Dayan FE (2020) Arg-128-Leu
target-site mutation in PPO2 evolves in wild poinsettia (Euphorbia heterophylla) with cross-
resistance to PPO-inhibiting herbicides. Weed Sci 68:437–444

Montgomery JS, Giacomini DA, Tranel PJ (2020) Molecular confirmation of resistance to PPO
inhibitors in Amaranthus tuberculatus and Amaranthus palmeri, and isolation of the G399A
PPO2 substitution in A. palmeri. Weed Technol:35:99–105

Nie H, Mansfield BC, Harre NT, Young JM, Steppig NR, Young BG (2019) Investigating target‐
site resistance mechanism to the PPO‐inhibiting herbicide fomesafen in waterhemp and
interspecific hybridization of Amaranthus species using next generation sequencing. Pest
Manag Sci 75:3235–3244

Noguera MM, Rangani G, Heiser J, Bararpour T, Steckel LE, Betz M, Porri A, Lerchl J,
Zimmermann S, Nichols RL, Roma-Burgos N (2021) Functional PPO2 mutations: co-
occurrence in one plant or the same PPO2 allele of herbicide-resistant Amaranthus palmeri
in the US mid-south. Pest Manag Sci 77:1001–1012

170
Obenland OA, Ma R, O’Brien SR, Lygin A V., Riechers DE (2019) Carfentrazone-ethyl resistance
in an Amaranthus tuberculatus population is not mediated by amino acid alterations in the
PPO2 protein. PLoS One 14:e0215431

Patzoldt WL, Hager AG, McCormick JS, Tranel PJ (2006) A codon deletion confers resistance to
herbicides inhibiting protoporphyrinogen oxidase. Proc Natl Acad Sci U S A 103:12329–34

Porri A, Betz M, Seebruck K, Knapp M, Johnen P, Witschel M, Aponte R, Liebl R, Tranel PJ,
Lerchl J (2022) Inhibition profile of trifludimoxazin towards PPO2 target site mutations. Pest
Manag Sci 79:507–519

Rangani G, Salas-Perez RA, Aponte RA, Knapp M, Craig IR, Mietzner T, Langaro AC, Noguera
MM, Porri A, Roma-Burgos N (2019) A novel single-site mutation in the catalytic domain of
Protoporphyrinogen Oxidase IX (PPO) confers resistance to PPO-inhibiting herbicides. Front
Plant Sci 10:568

Ritz C, Baty F, Streibig JC, Gerhard D (2015) Dose-Response Analysis Using R. PLoS One 10(12):
e0146021. https://doi.org/10.1371/journal.pone.0146021

Rousonelos SL, Lee RM, Moreira MS, VanGessel MJ, Tranel PJ (2012) Characterization of a
common ragweed (Ambrosia artemisiifolia) population resistant to ALS- and PPO-inhibiting
herbicides. Weed Sci 60:335–344

Saghai-Maroof MA, Soliman KM, Jorgensen RA, Allard RW (1984) Ribosomal DNA spacer-
length polymorphisms in barley: mendelian inheritance, chromosomal location, and
population dynamics. Proc Natl Acad Sci U S A 81:8014–8018

Schafer JR, Hallett SG, Johnson WG (2012) Response of giant ragweed (Ambrosia trifida),
horseweed (Conyza canadensis), and common lambsquarters (Chenopodium album)
biotypes to glyphosate in the presence and absence of soil microorganisms. Weed Sci
60:641–649
Shaner DL, ed (2014) Herbicide Handbook. 10th edn. Lawrence, KS: Weed Science Society of
America. 513 p
Shoup DE, Al-Khatib K, Peterson DE (2003) Common waterhemp (Amaranthus rudis) resistance
to protoporphyrinogen oxidase-inhibiting herbicides. Weed Sci 51:145–150

171
Strom SA, Hager AG, Seiter NJ, Davis AS, Riechers DE (2020) Metabolic resistance to S-
metolachlor in two waterhemp (Amaranthus tuberculatus) populations from Illinois, USA.
Pest Manag Sci 76:3139–3148

Tranel PJ, Wu C, Sadeque A (2017) Target-site resistances to ALS and PPO inhibitors are linked
in waterhemp (Amaranthus tuberculatus). Weed Sci 65:4–8

[USDA-NASS] US Department of Agriculture, National Agricultural Statistics Service (2020)


Agricultural chemical use program.
https://www.nass.usda.gov/Surveys/Guide_to_NASS_Surveys/Chemical_Use/. Accessed:
May 15, 2022

Van Horn CR, Moretti ML, Robertson RR, Segobye K, Weller SC, Young BG, Johnson WG,
Schulz B, Green AC, Jeffery T, Lespérance MA, Tardif FJ, Sikkema PH, Hall JC, McLean
MD, Lawton MB, Sammons RD, Wang D, Westra P, Gaines TA (2018) Glyphosate
resistance in Ambrosia trifida: Part 1. Novel rapid cell death response to glyphosate. Pest
Manag Sci 74:1071–1078

Varanasi VK, Brabham C, Norsworthy JK (2018) Confirmation and characterization of non-target


site resistance to fomesafen in Palmer amaranth (Amaranthus palmeri). Weed Sci 66:702–
709

Waselkov KE, Olsen KM (2014) Population genetics and origin of the native North American
agricultural weed waterhemp (Amaranthus tuberculatus; Amaranthaceae). Am J Bot
101:1726–1736

Waselkov KE, Regenold ND, Lum RC, Olsen KM (2020) Agricultural adaptation in the native
North American weed waterhemp, Amaranthus tuberculatus (Amaranthaceae). PLoS ONE
15(9): e0238861. https://doi.org/10.1371/journal.pone.0238861

Watanabe N, Che FS, Iwano M, Takayama S, Yoshida S, Isogai A (2001) Dual targeting of spinach
protoporphyrinogen oxidase II to mitochondria and chloroplasts by alternative use of two in-
frame initiation codons. J Biol Chem 276:20474–20481

172
Wuerffel RJ, Young JM, Lee RM, Tranel PJ, Lightfoot DA, Young BG (2015a) Distribution of
the ΔG210 protoporphyrinogen oxidase mutation in Illinois waterhemp (Amaranthus
tuberculatus) and an improved molecular method for detection. Weed Sci 63:839–845

Wuerffel RJ, Young JM, Matthews JL, Young BG (2015b) Characterization of PPO-inhibitor–
resistant waterhemp (Amaranthus tuberculatus) response to soil-applied PPO-inhibiting
herbicides. Weed Sci 63:511–521

Young BG (2006) Changes in herbicide use patterns and production practices resulting from
glyphosate-resistant crops. Weed Technol 20:301–307

Zhang J, Kirkham MB (1996) Antioxidant responses to drought in sunflower and sorghum


seedlings. New Phytol 132:361–373

173
Table 5.1. Phenotypic assignment for each plant sprayed with fomesafen at 14 days after treatment.
Waterhemp Resistance Group Assignments After Treatment With Four Rates of Fomesafen

ΔG210a Unknownb Susceptiblec

20 g 40 ga 60 g 100 g 20 g 40 g 60 g 100 g 20 g 40 g 60 g 100 g

Population ha-1 ha-1 ha-1 ha-1 ha-1 ha-1 ha-1 ha-1 ha-1 ha-1 ha-1 ha-1

------------------------------------------------No. of plants---------------------------------------------------------

Car 17 10 18 10 0 0 0 0 3 0 2 0

Des 0 0 0 0 0 0 0 0 20 10 20 10
174

Fran 13 9 14 7 3 1 2 0 4 0 4 3

IA-340 0 0 0 0 3 3 0 0 17 7 20 10

IA-358 2 1 0 0 3 0 1 0 15 9 19 10

IA-369 1 1 0 0 9 4 5 1 10 5 15 9
a
Presence of ΔG210 was confirmed by qPCR assay of leaf tissue collected after waterhemp survival.
b
Assignment to putative novel R mechanism group was made if the plant was controlled less than susceptible plants and
was confirmed as being absent of the ΔG210 mutation via qPCR assay of leaf tissue collected after waterhemp survival
c
Assignment to susceptible group was made at waterhemp death or individual waterhemp control was similar to the
susceptible waterhemp population and was confirmed as being absent of the ΔG210 mutation via qPCR assay of leaf
tissue collected after waterhemp survival.
Table 5.2. Amino acid polymorphisms observed in isolated waterhemp PPX2 genes that survived a dose of fomesafen.
previous
a
Position IA-369 (16) IA-358 (2) IA-340 (5) Fran (5) Known S Total submissions similar residue
----------------------% of individuals---------------------- -------No.-----
G251 insb 75 100 60 80 3 29 present
H170R 68.8 100 40 60 3 25 present yes
T358S 62.5 0 40 40 3 21 present
Q269H 56.3 100 20 60 2 21 present
E38D 43.8 100 40 40 1 17 present yes
M279I 25 100 0 80 1 15 present yes
A172V 37.5 50 20 40 2 15
D453E 43.8 100 0 20 0 10 yes
T358N 25 100 0 60 0 9
R261H 12.5 100 0 40 1 9 yes
N431K 25 100 0 40 1 9 present
175

K194R 12.5 100 0 40 0 7 yes


T41S 18.8 0 20 0 1 5
S42T 18.8 0 20 0 1 5
K258N 18.8 0 20 0 1 5
I257V 18.8 0 20 0 1 5 yes
S68N 6.3 50 0 20 0 4 present
N473K 0 50 40 0 1 4
S476C 6.3 50 0 20 0 3 yes
P260T 12.5 0 0 20 0 3
L307S 18.8 0 0 0 0 3
L166F 0 0 20 20 1 3
I304T 6.3 0 20 0 1 3
E249D 12.5 0 0 20 0 3 yes
a
Number in parentheses indicates total number of resistant individuals sequenced for that population.
b
Abbreviation: ins, insertion
Table 5.3. Amino acid polymorphisms observed in isolated waterhemp PPX1 genes that
survived a dose of fomesafen.
IA-369 IA-358
a
Position (16) (2) IA-340 (5) Fran (5) Known S Total
-------------------------%------------------------ -------No.-------
V498L 56.3 50 60 100 3 25
P338L 18.8 50 60 60 1 11
L196I 18.8 0 60 0 1 7
D67E 0 0 0 20 2 3
N55D 0 0 0 20 1 2
R107C 12.5 0 0 0 0 2
P172T 0 0 40 0 0 2
S202P 6.3 0 0 20 0 2
K254R 12.5 0 0 0 0 2
P279L 0 50 0 0 0 2
N325D 12.5 0 0 0 0 2
G36R 0 50 0 0 0 1
T48S 6.3 0 0 0 0 1
S52A 0 0 20 0 0 1
A54T 0 0 0 0 1 1
T61K 0 0 0 0 1 1
S62N 0 0 20 0 0 1
K68T 0 0 20 0 0 1
R107H 6.3 0 0 0 0 1
D141Y 0 0 0 0 0 1
L144I 6.3 0 0 0 0 1
P172I 0 0 20 0 0 1
D179Y 0 0 0 0 0 1
F183I 0 0 0 20 0 1
G195R 0 50 0 0 0 1
P274L 0 0 0 0 0 1
V350I 6.3 0 0 0 0 1
a
Number in parentheses indicates total number of resistant individuals sequenced
for that population.

176
Table 5.4. Comparison of GR50 and LD50 values and R:S ratios calculated from waterhemp
biomass and survival at 14 days after fomesafen treatment for one susceptible population
(Des), two resistant populations with ΔG210 (Car and Was), and four populations displaying
a delayed regrowth resistance phenotype (Fran, IA-340, IA-358, and IA-369).
Population GR50 (Std err) R/S ratioa LD 50 (Std err) R/S ratio

g ai ha-1 g ai ha-1

Des 3.0 (0.4) - 20.7 (2.4) -

Fran 5.3 (1.1) 1.8* 177.5 (37.6) 8.6*

IA-340 3.6 (0.6) 1.2 75.5 (13.5) 3.6*

IA-358b 1.9 (0.4) 0.6 39.2 (5.0) 1.9*

IA-369 3.2 (0.6) 1.1 98.6 (16.3) 4.8*

Car 14.8 (3.2) 4.9* 403.9 (103.1) 19.5*

Was 29.5 (7.4) 9.7* 562.1 (139.5) 27.1*


a
Ratios with a * indicate ratio is different than 1 at α=0.05.
b
IA-358 population is actually a single cross hybrid of IA-358 and IA-340 due to lack of males
that were selected as resistant.

177
ΔG210 Unknown Resistance Susceptible
a a
100 bc ab acb
c
90 c
cd
Waterhemp Control 14 DAT (%)

de
80
ef
f
70
g
60

50

40

30

20

10

0
20 40 60 100
Fomesafen Rate (g ai ha-1)

Figure 5.1. Waterhemp control 14 days after fomesafen application for susceptible, ΔG210
resistant, and unknown cause resistant waterhemp from 6 populations. Populations are pooled due
to lack of significance in ANOVA. Means were separated using Tukeys HSD at α=0.05. Presence
of ΔG210 was confirmed by qPCR assay of leaf tissue collected after waterhemp survival.
Assignment to putative novel resistance mechanism group was made if the plant was controlled
less than susceptible plants and was confirmed as being absent of the ΔG210 mutation via qPCR
assay of leaf tissue collected after waterhemp survival. Assignment to susceptible group was made
at waterhemp death or individual waterhemp control was similar to the susceptible waterhemp
population and was confirmed as being absent of the ΔG210 mutation via qPCR assay of leaf tissue
collected after waterhemp survival.

178
Figure 5.2. Phylogenetic tree of selected Miseq PPX2 sequences and multiple Genbank
accessions of waterhemp and closely related species. Accessions shown in red are more closely
related to Palmer amaranth and redroot pigweed than wild type waterhemp accessions.

179
180

Figure 5.3. Comparison of plant symptomatology at 24 hours after treatment with fomesafen at 20 g ha-1. Colored boxes represent
different types of populations with red indicating ΔG210 resistant, yellow indicating susceptible, and blue indicating delayed regrowth
phenotype populations.
181

Figure 5.4. Comparison of plant symptomology between populations at 3d after treatment with 7 rates of fomesafen. Red arrows in each
frame point to the 20 g ha-1 and 63 g ha-1 rate of fomesafen treatments.
182

Figure 5.5. Comparison of plant symptomology between populations at 14d after treatment with 8 rates of fomesafen. Red arrows in
each frame point to the 20 g ha-1 and 63 g ha-1 rate of fomesafen treatments.
Adjusted dry weight (% of nontreated)

Fomesafen rate (g ai ha-1)

Figure 5.6. Dose-response curves of susceptible (Des) ΔG210 resistant (Car and Was), and
putative novel R mechanism populations (Fran, IA-340, IA-358, and IA-369) to fomesafen.
Aboveground biomass was harvested 14 days after fomesafen treatment and normalized to a
percentage of the untreated for each population

183
Waterhemp Survival (%)

Fomesafen rate (g ai ha-1)

Figure 5.7. Dose-response curves of susceptible (Des) ΔG210 resistant (Car and Was), and
putative novel R mechanism populations (Fran, IA-340, IA-358, and IA-369) to fomesafen.
Waterhemp survival was assessed 14 days after fomesafen treatment.

184
Figure 5.8. Mean gene expression of protoporphyrinogen oxidase 1 (PPX1) and 2 (PPX2) in
susceptible (Des) ΔG210 resistant (Car and Was), and four resistant populations with a putative
novel mechanism (Fran, IA-340, IA-358, and IA-369) prior to fomesafen treatment as
determined by qPCR assays. Expression values for each target are normalized to that of
elongation factor-1 alpha and are expressed as percentages of the susceptible population. Means
of each population are indicated by X symbol. Overall F test P values were 0.077 and 0.0438 for
PPX1 and PPX2 respectively. Expression differences between Was and IA-358 PPX2 were the
only mean differences using Tukey’s HSD (α = 0.05).

185
180

160

MDA Content (nmol g-1 tissue) 140

120
*
100 †*
80

60

40
Car Des
20
Fran IA-369
0
0 10 20 30 40 50 60
Hours After Treatment
180
Car
160 Des
Fran
140
MDA Content (nmo g-1 tissue)

IA-369
120

100 †
80 †* †
60

40

20

0
0 10 20 30 40 50 60
Hours After Treatment

Figure 5.9. Lipid peroxidation as indicated by MDA content of mature leaf tissue (top) and
meristem/young leaf tissue (bottom) in known susceptible (Des), ΔG210 resistant (Car) and two resistant
populations with putative novel mechanism (Fran and IA-369) following treatment with fomesafen at 20
g ha-1. Significant differences of Fran and IA-369 populations compared to Des and Car populations are
denoted by * and † symbols respectively using Fishers protected LSD (α = 0.05).

186
VITA

Jesse A. Haarmann
EDUCATION
Doctor of Philosophy in Weed Science — Purdue University May 2023
Dissertation: Influence of Application Placement, Resistance Genotype, and PPO-
Inhibiting Herbicide on the PPO-Resistance Phenotype in Waterhemp
Advisor: William G. Johnson, Ph.D. GPA:3.81

Master of Science in Weed Science – Purdue University May 2019


Thesis: The Effect of Herbicide Respray Treatments and Timings on Regrowth of Four
Weed Species.
Advisor: William G. Johnson, Ph.D. GPA:3.96

Bachelor of Science– University of Illinois Urbana-Champaign May 2017


Major: Crop Sciences GPA:3.70

RELAVENT SKILLS

Technical

• Field site management for trial initiation, implementation, and maintenance of desired
pest populations
• Knowledge of Midwest agricultural practices and industry wide product offerings
• Use of pesticide research application equipment including hand boom and ATV
sprayer
• Operation of agronomic research and large-scale machinery including planters,
combines, sprayers, and tillage equipment
• Greenhouse research and plant production methods including use of track-mounted
research sprayer, plant propagation, seed germination, pollination, and seed
production.
• Laboratory techniques including nucleic acid extraction, confocal microscopy, gel
electrophoresis, and PCR
• Experimental design and data analysis using SAS and R statistical software
• Research software and equipment including ARM and Harvest Master Weigh Bucket
System
• Technical writing for academic and extension publications
• Cooperation with faculty, graduate students, undergraduate workers, and industry
cooperators
• Experience conducting research in a variety of agronomic disciplines including: weed
science, plant pathology, soil fertility, and corn and soybean variety trials
• Obtained an Indiana Pesticide Applicator License.

187
Communication
• Published research in peer-reviewed academic journals
• Collaboration with individuals internal and external to Purdue: university faculty,
industry representatives, graduate students, and interns for the design and implementation
of research trials and demonstrations
• Presented research findings via oral and poster presentations at scientific meetings
• Presented research findings at grower extension meetings
• Demonstrated ability to work in a team environment

Leadership and Management

• Management of off-site research locations for Purdue Herbicide Evaluation Program and
student projects
• Responsible for field site harvest, trial termination, and data reports for primary field sites
• Decision making regarding protocol design, trial implementation, and data collection in
order to balance workload, optimize weed density, and effectively navigate space
limitations
• Lead plot tours for academic and industry cooperators
• Mentoring of new graduate students and interns on weed science research, writing, and
statistical analysis

PROFESSIONAL EXPERIENCE

Purdue University, West Lafayette, IN


Graduate Research Associate Aug. 2019 - Present
• Manage over 40 field trials at 5 Indiana field locations.
• Prepare, organize, and maintain field sites.
• Establish field experiments and apply treatments in a timely manner.
• Coordinate logistics for labor and equipment transport.
• Collect and compile experimental data for data reports to clients.

Purdue University, West Lafayette, IN


Graduate Research Assistant May 2017 - Aug. 2019
• Conduct field, greenhouse, and laboratory research related to herbicide efficacy and weed
control.
• Analyze data using SAS and R statistical software.
• Organize and present research project data for professional society meetings and
scientific journals.
• Make applications and collect data for industry contract research trials.

BASF, Seymour, IL
Midwest Research Farm Intern Apr. 2016 - Nov. 2016

188
• Manage applications and data collection for ten agronomic research trials.
• Assist in management of various field research trials.
• Guide customers through plot tours.
• Perform site maintenance and upkeep.

Becks Superior Hybrids, Effingham, IL


Practical Farm Research Intern May 2015 - Aug. 2015
• Assist in the management of agronomic research trials.
• Ensure proper timing of applications for intern projects.
• Assist with location establishment and maintenance.
AgriGold, Fairfield, IL
Sales Intern May 2014 - Aug. 2014
• Complete seed returns for all customers of two district sales managers.
• Erect plot and field signs for brand promotion in the area.
• Generate sales leads for DSMs by building relationships with area farmers.

Haarmann Family Farm, Effingham, IL


Farm Assistant 2004-2018
• Learn basic farm operation and economics.
• Assist in the decision making of agronomic related decisions.
• Operate and maintain large-scale farm equipment.

TEACHING EXPERIENCE

Teaching Assistant, BTNY 304 – Introductory Weed Science (Fall 2018 and Fall 2021)
• Prepare and present lab demonstrations for spray application technology, herbicide
movement in soil, and herbicide symptomology.
• Plant and maintain weed specimens for identification quizzes.
• Grade exams and assignments.

REFEREED JOURNAL ARTICLES


Haarmann JA, Young BG, Johnson WG (2021) Control of Palmer amaranth (Amaranthus
palmeri) regrowth following failed applications of glufosinate and fomesafen. Weed Technology.
35: 464–470. doi: 10.1017/wet.2021.16 (Citations:3)

Haarmann JA, Young BG, Johnson WG (2020) Control of waterhemp (Amaranthus


tuberculatus) regrowth after failed applications of glufosinate or fomesafen. Weed Technology.
34: 794–800. doi: 10.1017/wet.2020.58 (Citations:3)

EXTENSION PUBLICATION
Haarmann JA, Whitford F, Johnson WG, Delucchi TF, Zimmer M, Young BG, Steckel L,
Davis V. (2022) A Soybean Postemergence Herbicide Failed to Perform; Is Retreatment
Warranted? Purdue Extension WS-57.

189
EXTENSION AND OUTREACH PRESENTATIONS
• Throckmorton Purdue Agricultural Center – Corteva Field Day July 11th, 2022
• Throckmorton Purdue Agricultural Center – Weed Science Field Day June 30th, 2022
• Purdue 3 Minute Thesis Competition Mar. 4th, 2022
• Throckmorton Purdue Agricultural Center – Weed Science Field Day June 24th, 2021
• Botany Research Showcase – Grad Student Lightning Talk Nov. 13th, 2019
• Feldun Purdue Agricultural Center – Pasture and Forage Field Day Aug. 7th, 2019
• Throckmorton Purdue Agricultural Center – Weed Science Field Day June 20th, 2019
• Graduate Thesis Defense Seminar Apr. 10th, 2019
• Throckmorton Purdue Agricultural Center – Weed Science Field Day July 2nd, 2018
• Throckmorton Purdue Agricultural Center – Weed Science Field Day June 29th, 2017

CONFERENCE PROCEEDINGS
Haarmann, J., Young, B., Johnson, W. (2022, December) Confirmation of PPO-Inhibitor
Resistant Waterhemp (Amaranthus tuberculatus) That Does Not Have Resistance-
Conferring Target-Site Mutations. Paper presented at the annual meeting of the North Central
Weed Science Society, St. Louis, Missouri.
Haarmann, J., Young, B., Johnson, W. (2022, December) Is Detoxification Contributing to PPO-
inhibitor Resistance in Waterhemp (Amaranthus tuberculatus)? Poster presented at the
annual meeting of the North Central Weed Science Society, St. Louis, Missouri.
Haarmann, J., Young, B., Johnson, W. (2022, February) Investigating the Potential Co-
Occurrence of Target-Site and Non-Target-Site Resistance to PPO Inhibitors in the Same
Populations of Waterhemp (Amaranthus tuberculatus). Poster presented at the annual
meeting of the Weed Science Society of America, Vancouver, Canada.
Haarmann, J., Young, B., Johnson, W. (2021, December) The Effect of ΔG210 and R128G
Mutations in Waterhemp (Amaranthus tuberculatus) on the Resistance Phenotype for Soil-
Applied PPO Inhibitors. Paper presented at the annual meeting of the North Central Weed
Science Society, Grand Rapids, Michigan
Haarmann, J., Young, B., Johnson, W. (2021, December) Soil-Applied PPO Inhibitors Select for
Fewer Resistant Individuals than Foliar Applications. Poster presented at the annual meeting
of the North Central Weed Science Society, Grand Rapids, Michigan
Haarmann, J., Young, B., Johnson, W. (2020, December) The Effect of Trifludimoxazin on the
Frequency of the ΔG210 Target-Site Mutation in Field Populations of Waterhemp. Poster
presented at the annual meeting of the North Central Weed Science Society, Virtual meeting.
Haarmann, J., Young, B., Johnson, W. (2019, December) Is Waterhemp More Difficult to
Control Following a Sublethal Glufosinate Application? Paper presented at the annual
meeting of the North Central Weed Science Society, Columbus, Ohio.
Haarmann, J., Young, B., Johnson, W. (2018, December) Efficacy of Soybean Herbicide Respray
Applications on Palmer amaranth and Waterhemp. Paper presented at the annual meeting of
the North Central Weed Science Society, Milwaukee, Wisconsin.
Haarmann, J., Young, B., Johnson, W. (2018, December) POST Herbicide Respray Efficacy as
Influenced by Severity of Waterhemp Injury from Prior Herbicide Applications. Poster
presented at the annual meeting of the North Central Weed Science Society, Milwaukee
Wisconsin.

190
Haarmann, J., Young, B., Johnson, W. (2017, December) Strategies for Control of Waterhemp
that Survived a Post Contact Herbicide. Paper presented at the annual meeting of the North
Central Weed Science Society, St. Louis, Missouri.
Haarmann, J., Young, B., Johnson, W. (2017, December) Strategies for Control of Palmer
Amaranth That Survived a Post Contact Herbicide. Poster presented at the annual meeting
of the North Central Weed Science Society, St. Louis, Missouri.
Petersen, W., Haarmann, J., Johnson, W. (2017, December) Methods to Control Giant Ragweed
Populations Following Survival of a POST Herbicide Treatment. Paper presented at the
annual meeting of the North Central Weed Science Society, St. Louis, Missouri.
Haarmann, J. Marshal, J. (2016, November) Evaluating the Visual Effects of Oxidative Stress on
Strobilurin Fungicide Treated Soybeans. Poster presented at the annual meeting of the
Students in Agronomy, Soil, and Environmental Sciences, Phoenix, Arizona.

HONORS AND AWARDS


• Purdue College of Agriculture Graduate Student Spotlight 2022
• 2nd place – Poster Contest, WSSA meetings 2022
• 1st place – Graduate Team, NCWSS Weeds Contest 2018, 2019, 2021, 2022
• 1 place – Graduate Individual, NCWSS Weeds Contest
st
2022
• 3rd place – Graduate Individual, NCWSS Weeds Contest 2018, 2019, 2021
• 2nd place – Poster Contest, NCWSS meetings 2017, 2021
• University of Illinois Dean’s List Fall 2015, Spring 2016, Fall 2016

PROFESSIONAL INVOLVEMENT
• North Central Weed Science Society 2017 – Present
• Weed Science Society of America 2020 – Present
• Purdue Botany and Plant Pathology Graduate Student Organization 2017 – Present
• IlliDell of Alpha Gamma Sigma Professional Agriculture Fraternity 2013 – 2017
• University of Illinois Field and Furrow Club 2013 – 2017

191

You might also like