ArticlePDF AvailableLiterature Review

Predicting the unpredictable: The regulatory nature and promiscuity of herbicide cross resistance

Wiley
Pest Management Science
Authors:

Abstract and Figures

The emergence of herbicide‐resistant weeds is a significant threat to modern agriculture. Cross resistance, a phenomenon where resistance to one herbicide confers resistance to another, is a particular concern due to its unpredictability. Nontarget‐site (NTS) cross resistance is especially challenging to predict, as it arises from genes that encode enzymes that do not directly involve the herbicide target site and can affect multiple herbicides. Recent advancements in genomic and structural biology techniques could provide new venues for predicting NTS resistance in weed species. In this review, we present an overview of the latest approaches that could be used. We discuss the use of genomic and epigenomics techniques such as ATAC‐seq and DAP‐seq to identify transcription factors and cis‐regulatory elements associated with resistance traits. Enzyme/protein structure prediction and docking analysis are discussed as an initial step for predicting herbicide binding affinities with key enzymes to identify candidates for subsequent in vitro validation. We also provide example analyses that can be deployed toward elucidating cross resistance and its regulatory patterns. Ultimately, our review provides important insights into the latest scientific advancements and potential directions for predicting and managing herbicide cross resistance in weeds. This article is protected by copyright. All rights reserved.
Content may be subject to copyright.
Review
Received: 14 July 2023 Revised: 14 August 2023 Accepted article published: 16 August 2023 Published online in Wiley Online Library:
(wileyonlinelibrary.com) DOI 10.1002/ps.7728
Predicting the unpredictable: the regulatory
nature and promiscuity of herbicide cross
resistance
Lucas K Bobadilla and Patrick J Tranel
*
Abstract
The emergence of herbicide-resistant weeds is a signicant threat to modern agriculture. Cross resistance, a phenomenon where
resistance to one herbicide confers resistance to another, is a particular concern owing to its unpredictability. Nontarget-site (NTS)
cross resistance is especially challenging to predict, as it arises from genes that encode enzymes that do not directly involve the
herbicide target site and can affect multiple herbicides. Recent advancements in genomic and structural biology techniques could
provide new venues for predicting NTS resistance in weed species. In this review, we present an overview of the latest
approaches that could be used. We discuss the use of genomic and epigenomics techniques such as ATAC-seq and DAP-
seq to identify transcription factors and cis-regulatory elements associated with resistance traits. Enzyme/protein structure
prediction and docking analysis are discussed as an initial step for predicting herbicide binding afnities with key enzymes
to identify candidates for subsequent in vitro validation. We also provide example analyses that can be deployed toward elu-
cidating cross resistance and its regulatory patterns. Ultimately, our review provides important insights into the latest scien-
tic advancements and potential directions for predicting and managing herbicide cross resistance in weeds.
© 2023 The Authors. Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry.
Keywords: cross resistance; herbicide metabolism; cis-regulatory elements; transcription factors
1 INTRODUCTION
Herbicide resistance is a phenomenon that has been documented
during the last decades as the main challenge for weed manage-
ment. Multiple cases of point mutation at the herbicide site-of-action
(SoA, target-site resistance) have been identied and can provide a
high resistance level.
1
However, a plant can become resistant with-
out any mutation at the herbicide site of action; this resistance type
is referred to as nontarget-site resistance (NTS).
2
For the last several
years, unraveling the mechanisms of NTS herbicide resistance has
been a primary goal for weed scientists. However, only recently
have signicant advances been made toward identifying key
enzyme players, mainly as a consequence of the advent of weed
genomic resources.
3,4
In many cases, NTS-resistant weed populations are identied as
cross resistant, able to survive herbicides from the same and dif-
ferent SoA groups. These populations are challenging to study
owing to the potential involvement of numerous genes for NTS
resistance traits.
57
For example, even when a major player in
the detoxication of a herbicide is identied, it is typically
unknown if the same encoded enzyme can reduce the toxicity
of other herbicides or if cross resistance is caused by other genes.
From a practical standpoint, the unpredictable nature of cross
resistance makes it difcult to select the best herbicide combina-
tion to mix or rotate to mitigate resistance evolution.
8
Conse-
quently, NTS herbicide resistance is viewed as one of the greatest
threats to the sustainability of chemical weed management. But
what if it is possible to make the unpredictable predictable?
We believethat recent scientic advances have now made it possi-
ble to begin systematically unravelling the genetic underpinnings
of NTS resistance and cross resistance.
Current genomics resources for weeds, several of which have been
achieved by the International Weed Genomics Consortium, are
growing and will serve as the basis for prediction tools for better
weed management decisions. However, weed scientists still lack
important datasets such as regulome and epigenome maps. Such
datasets would allow weed scientists to understand complex adap-
tation patterns in weeds, such as NTS resistance and climate change
adaptation. The present review aims to shed light on new frontiers
for weed scientists regarding the regulatory machinery of cros s resis-
tance, how this information could be generated and how it could be
used to build prediction models based on gene regulatory networks.
2 MECHANISMS OF HERBICIDE CROSS
RESISTANCE
The primary chemical strategy for managing herbicide resistance
is to use a different herbicide. However, this strategy can be
*Correspondence to: PJ Tranel, Department of Crop Sciences, University of
Illinois, 1201 W Gregory Dr, Urbana, IL 61801, USA. E-mail: tranel@illinois.edu
Department of Crop Sciences, University of Illinois, Urbana, IL, USA
© 2023 The Authors. Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry.
This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any
medium, provided the original work is properly cited.
1
rendered ineffective by multiple resistance and cross resistance.
The Herbicide Resistance Action Committee (www.hracglobal.
com) considers multiple herbicide resistance to be the existence
of two or more distinct mechanisms, each of which provides resis-
tance to separate herbicides. By contrast, cross resistance is
dened as the presence of a single mechanism that confers resis-
tance to different herbicides.
In some cases, the distinction between multiple and cross resis-
tance is clear: a plant that has increased expression of an enzyme
that can detoxify one herbicide and a target-site mutation confer-
ring resistance to another herbicide would be an example of mul-
tiple resistance, whereas a plant that has increased expression of
an enzyme that can detoxify multiple herbicides would be an
example of cross resistance. However, with an increased appreci-
ation for the complexity of herbicide resistance in some cases, it is
becoming apparent that the distinction between multiple and
cross resistance is not always clear. For example, in cases where
resistance to a single herbicide is mediated by increased expres-
sion of two or more enzymes working together to detoxify the
herbicide, and they are able to singularly or jointly detoxify
another herbicide, should that be considered multiple resistance
or cross resistance? Does the classication of this case change if
the two or more enzymes are independently regulated or if they
are regulated by the same transcription factor? If a plant was
found to be resistant to two different herbicides as a result of
increased expression of two different enzymes, each of which
acted on only one of the herbicides, presumably the plant would
be considered to have multiple resistance. But if it was later deter-
mined that the two enzymes were upregulated due to a single
genetic change (e.g., increased expression of a transcription fac-
tor), should the plant be reclassied as being cross resistant? For
the purpose of this review, we dene cross resistance as resis-
tance to herbicides other than the one that selected the
resistance.
There are several mechanisms by which cross resistance can
occur, and they differ in their complexity and predictability
(Table 1). The simplest is target-site cross resistance;
9
if a herbicide
selects for a mutation that reduces binding afnity between the
herbicide and its target site, the same mutation often also will
confer resistance to other herbicides that have the same target
site. For example, mutations in the gene encoding acetolactate
synthase often confer cross resistance to multiple herbicides that
target this enzyme. This type of cross resistance is, at some level,
predictable because it affects only herbicides with the same bind-
ing site. NTS resistance, by contrast, is less predictable and often
occurs as a result of increased expression of one or more genes,
which results in less herbicide reaching the site of action or allows
the plant to ameliorate the effects of herbicide inhibition of the
target site.
1
NTS is less predictable than TS resistance because
(i) a single gene can affect herbicides across different SoA groups
and (ii) it can involve multiple genes, each potentially affecting
different herbicides (Table 1). Although NTS resistance potentially
could be caused by any of a vast array of genes, the end result
often is increased metabolism (detoxication) of the herbicide.
Because herbicide metabolism is such an important component
of NTS, it is the focus of this review. However, several of the
approaches which we discuss also could be applicable to other
NTS resistance mechanisms.
Herbicide metabolism can be divided into three phases
1,10,11
:
Phase 1 involves the introduction of small functional groups on
herbicide molecules through catalysis by key enzymes such as
cytochrome P450s (CYP450s); Phase 2 involves the conjugation
of catalyzed herbicides to common plant metabolites via
transferases such as glutathione-S-transferases (GSTs) and
UDP-glycosyltransferases (UGTs); and Phase 3 involves the
sequestration of metabolites into vacuoles and/or cell wall mate-
rial, potentially through ATP-binding cassette (ABC) transporters.
Cross resistance may be caused by a single enzyme or by the
combination of multiple ones participating in any of the three
phases of herbicide metabolism. For example, single CYP450s
and GSTs have been shown to metabolize herbicides from multi-
ple SoA groups.
1217
Additionally, multiple genes encoding differ-
ent enzymes can be co-expressed (and potentially co-regulated),
where each enzyme could play a role in the detoxication process
of multiple herbicides. For instance, a study identied that expres-
sion of a GST and ABC transporter were co-induced by multiple
Table 1. Some potential mechanisms of herbicide cross resistance and their predictability
Mechanism Number of genes Can confer resistance to: Example Predictability
Site of action 1 Herbicides with the same
site of action
A mutation in the gene encoding acetolactate
synthase selected by one Group 2 herbicide confers
cross resistance to other Group 2 herbicides.
High
Detoxication
enzyme
1 Herbicides with similar
functional groups
A cytochrome P450 that can detoxify a Group 1
herbicide is also able to detoxify a Group 2 herbicide.
Shared genomic
location
>1 Potentially any herbicide Genes for two cytochrome P450s are located near each
other on the same chromosome: selection for
increased expression of one (e.g., by chromatin
remodeling) results in increased expression of the
other. Each cytochrome P450 can detoxify different
herbicides.
Shared gene
expression
network
>1 Potentially any herbicide A cytochrome P450 and a GST are in a shared
regulatory network. Selection for increased
expression of one (e.g., by increased expression of a
transcription factor) results in increased expression
of the other due to a shared transcription factor
binding site. The P450 and GST can detoxify
different herbicides.
Low
www.soci.org LK Bobadilla, PJ Tranel
wileyonlinelibrary.com/journal/ps © 2023 The Authors.
Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry. Pest Manag Sci 2023
2
15264998, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ps.7728, Wiley Online Library on [30/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
herbicides from different SoA groups.
18
The multigenic nature of
NTS cross resistance also has been suggested by several transcrip-
tomics studies, which typically reveal combinations of CYP450s,
GSTs, ABC transporters and UGTs being constitutively over-
expressed in resistant biotypes.
19
Another example of how co-
expressed genes could lead to cross resistance was suggested
by a previous study in which a multiple-resistant Amaranthus
tuberculatus population contained genomic hotspots of differen-
tially expressed genes.
6
If such a cluster of differentially expressed
genes included multiple NTS genes, selection by a herbicide for
increased expression of one could lead to cross resistance to
herbicides conferred by another.
Selection for NTS cross resistance is often observed in elds
where an indirect selection occurs (e.g., cross resistance had
already evolved to a herbicide never sprayed in that eld). This
phenomenon also was observed in antibiotic resistance, where
>50% of resistance cases that show cross resistance evolved
under a single antibiotic pressure.
20
As mentioned previously,
multiple enzymes and other molecules are known to have some
involvement in NTS metabolic resistance; however, their
co-induction and regulation still require elucidation. Additionally,
phytohormones related to plant stress, including salicylic acid and
jasmonic acid, need to be further examined regarding their
involvement in metabolic resistance and how they can affect
the co-regulation and co-expression of key NTS genes.
2124
Further molecular structural characterization and herbicide/
enzyme afnity estimations also are needed to identify metabolic
potentials to avoid cross resistance to novel herbicides. Although
it is well-known that enzyme promiscuity can account for herbi-
cide cross resistance, a recent study conrmed that co-regulation
of NTS genes also can lead to cross resistance.
25
There is much to
be learned from investigation of how NTS genes are regulated,
26
and it is our opinion that exploration of expression networks
involving NTS genes will be necessary to more fully predict NTS
cross resistance.
3 APPROACHES FOR IDENTIFICATION OF
REGULATORY ELEMENTS AND
CROSS-RESISTANCE PREDICTION
A fundamental question in biology is how complex gene expres-
sion patterns are regulated. Very few NTS resistance studies inves-
tigate the regulatory nature of NTS genes and their co-expression
patterns. This leaves questions open about the involvement of
master regulators, such as cis-regulatory elements and transcrip-
tion factors. Identifying these key regulators is essential, so predic-
tions about cross resistance based on expression proles can be
inferred. To achieve this, weed scientists must not only collect
gene expression data, but also generate epigenomic and
transcription-factor-binding-site data to establish the foundation
for constructing regulatory networks for cross resistance. Further-
more, developing predictive afnity models for key enzymes is
essential for identifying major genes encoding enzymes that
could potentially metabolize herbicide molecules. To validate
the accuracy of the afnity prediction, the generation of metabo-
lomics and in vitro validation assays is necessary (Fig. 1). These sets
of experiments would form the basis of a systems biology
approach to characterize metabolic cross resistance.
27
By inte-
grating these layers into the constructed regulatory networks,
one could determine which enzymes are co-regulated in a
specic weed species, and utilize their enzyme afnity and
metabolomics information to provide better-informed herbicide
recommendations. In the following sections, we discuss several
approaches and analyses that could be employed to achieve
this goal.
3.1 Herbicide and enzyme afnity predictions
The identication of key enzymes or proteins involved in herbi-
cide metabolism that can metabolize other herbicides is essential
for predicting cross resistance. However, predicting whether an
enzyme can metabolize a particular molecule can be challenging
and requires experimental validation. In vitro assays, such as
microsomal, recombinant enzyme and radio-labeled substrate
assays, can evaluate an enzyme's ability to metabolize a specic
molecule.
28
These assays have been widely used for drug
enzyme interactions and molecule discoveries, but can be chal-
lenging and require specialized measurement equipment, such
as mass spectrometry.
29,30
Alternatively, protein structures can be used to predict the afn-
ity of a specic molecule for an enzyme. The gold standard for
identifying an enzyme/protein structure is via crystallography.
However, this method can be challenging, time-consuming and
unsuitable for high-throughput analysis.
31
Instead, molecule
docking using homology structure prediction can, as a starting
point, help identify how a specic molecule, such as a CYP450 or
a GST, could bind to a herbicide molecule.
Recent advances in articial intelligence and machine-
learning algorithms, such as AlphaFold2 and the Google appli-
cation, ProteInfer, have made protein structure and function
prediction via homology and computational modeling much
more accessible.
32,33
Once the protein structure has been pre-
dicted, docking analysis combined with binding pocket predic-
tions can be used to predict the potential afnity between the
enzyme and multiple herbicides. In fact, some of the earliest
enzyme-docking predictions based on protein structure were
made using the D1 protein and photosystem II inhibitor herbi-
cides.
34
Herbicide-protein afnity predictions can be imple-
mented into a herbicide-resistance gene regulatory network
and used to select candidates for prediction validation via
in vitro or in vivo assays (Fig. 1). The predictions also could gen-
erate a database focused on herbicideenzyme binding afnity
and enzyme structure to help weed scientists make better deci-
sions regarding weed management. With time, these initial
predictions should be curated via in vivo and in vitro validations
to ensure data quality.
A few recent studies in weed science have utilized protein struc-
ture prediction with docking simulations. For instance, Ha et al.
12
applied a docking analysis for herbicide afnity identication
using the identied CYP81A in Echinochloa phyllopogon and its
rice homologs. GSTs and UGTs also should be analyzed for their
afnity with herbicide molecules. Plant GSTs contain N-terminal
and C-terminal structures, and the C-terminal located at the cata-
lytic region is highly variable regarding its sequence, making plant
GSTs capable of accepting much larger and more diverse
substrates than those of mammals.
35
In order to illustrate the approach for herbicide/enzyme afnity,
we predicted the structure of CYP81A10v7 via AlphaFold2, which
was previously identied as the culprit for cross resistance in
Lolium rigidum,
13
and ran a docking simulation analysis for some
of the herbicides identied in the same study that could be
metabolized by CYP81A10v7. Glyphosate also was included to
illustrate how a molecule not metabolized by this specic
CYP450 would behave compared to the other herbicide
Herbicide cross resistance www.soci.org
Pest Manag Sci 2023 © 2023 The Authors.
Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry.
wileyonlinelibrary.com/journal/ps
3
15264998, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ps.7728, Wiley Online Library on [30/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
molecules. Briey, we ran AlphaFold2
32
using the monomer
model and used the CYP81A10v7 protein sequence as input.
Ligand-binding pocket predictions were made using P2Rank to
obtain the coordinates around heme binding and for docking
prediction.
36
The heme co-factor molecule was added into
the CYP450 structure inside the predicted binding pocket. We
then used the generated structure as the receptor in Autodock
Vina to predict the binding afnity at the identied binding
pocket with the herbicides chlorsulfuron, atrazine, mesotrione,
triuralin, and glyphosate.
3741
As expected, multiple docking
modeswithstrongafnity (average of ∼−7.2 kcal mol
1
)were
identied for the herbicides that CYP81A10v7 previously
was identied to be able to metabolize, with all docking within
the heme-binding region with a similar pattern (Fig. 2).
However, a lower afnity was identied for glyphosate
(2.323 kcal mol
1
), indicating that this specic CYP450 would
probably not be able to hydroxylate glyphosate as efciently as
the other herbicides.
Databases also are available for some enzymes that could be
used as a starting point to build binding predictions. For instance,
the Cazy database (http://www.cazy.org) contains enzyme struc-
ture and binding information for carbohydrate-active enzymes
such as UGTs. For CYP450s, a recent database named PCPD was
created, which contains information about structure and ligand
docking for 181 plant CYP450s.
39
These databases could be used
as an initial step and as a repository for future weed species'
CYP450 structure and ligand information.
Even though AlphaFold2 is very good at predicting protein
structure, there are still limitations with complex proteins and in
building the correct conformation of some regions of the
Figure 1. Novel applications for the identication of regulatory elements and chromatin structure. DAP-seq can be used as a high-throughput tool for
genome-wide identication of transcription factor binding sites. In parallel, ATAC-seq and RNA-seq could be applied to identify accessible chromatin
and novel enhancers in different herbicide-resistant biotypes. Once these data are generated, they can be applied to build gene regulatory network
models for nontarget-site (NTS) resistance that, in conjunction with metabolomics and herbicide/enzyme afnity studies, could be used to predict
chances of cross resistance and apply better herbicide-resistance management decisions. Created with BioRender.com.
www.soci.org LK Bobadilla, PJ Tranel
wileyonlinelibrary.com/journal/ps © 2023 The Authors.
Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry. Pest Manag Sci 2023
4
15264998, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ps.7728, Wiley Online Library on [30/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
structure. However, the eld of protein structure prediction and
molecular docking is rapidly evolving, which may provide more
precise and better predictions in the near future. Despite the lim-
itations, protein structure prediction and docking analysis can
serve as a quick rst step for identifying the cross-resistance
potential of some enzymes and narrowing down candidates for
further validation. In conjunction with gene expression informa-
tion, one can build a regulatory network to identify key NTS
enzymes that are co-expressed and the array of herbicides that
can be metabolized by co-expressed enzymes.
3.2 Discovery and characterization of regulatory
elements
The elucidation of how gene expression is regulated in
response to abiotic stress and other developmental cues in
plants requires identifying cis-regulatory elements (CREs) and
transcription factor (TF) binding sites in their native chromatin
environment.
CREs are noncoding regions of DNA that play a major role in reg-
ulating the transcription of nearby genes. CREs are essential com-
ponents of gene regulatory networks that govern processes such
as cell differentiation, fate determination, and responses to biotic
and abiotic stressors.
4244
These elements consist of short DNA
motifs, typically 430 bp in length, which serve as binding sites
for sequence-specic transcription factors and modulate gene
expression by facilitating TF binding and transcriptional activa-
tion. TFs are proteins that can regulate and modulate gene
expression by binding to specic CREs controlling the transcrip-
tion of genes into mRNA molecules. They interact with other reg-
ulatory proteins, such as co-activators and co-repressors, to
activate or repress gene expression.
44
The identication of CREs can help to identify the regulatory
nature of duplicated genes and identify master regulators within
genomic hotspots (Fig. 3). For instance, CYP701 enzymes encoded
from duplicated genes in rice signicantly differ in their hydroxyl-
ation capabilities, where gene duplication favors drift to the pro-
miscuity of one of the paralogs and negative selection on the
other.
45
Thus, identifying CREs of duplicated genes and compar-
ing them can lead to identifying genes that can encode enzymes
that are more promiscuous regarding metabolic capabilities.
One of the most used methodology for identifying in vivo CREs and
TF binding sites of interest is chromatin immunoprecipitation-
sequencing (ChIP-seq).
46
However, this method requires either
antibodies for the TFs or transgenic plants expressing epitope-
tagged TFs, which makes it impractical for high-throughput
analysis.
An alternative technique becoming increasingly attractive
for genome-wide characterization of CREs and TFs in plants is
the assay for transposase-accessible chromatin sequencing
(ATAC-seq).
47
This method utilizes a hyperactive transposase
enzyme to insert sequencing adapters into regions of open
chromatin, allowing the identication of open chromatin
regions, DNA accessibility, and transcription factor binding sites
in the genome. ATAC-seq is widely used to study epigenetic
changes associated with various biological processes, such as
development, differentiation and environmental stimuli.
48
Figure 2. Herbicide/enzyme afnity prediction. The protein structure of cytochrome P450 81A (CYP81A) from Lolium rigidum was predicted using Alpha-
Fold2. The predicted structure was used in docking analysis for herbicides previously identied to be metabolized by CYP81A. The far-right panel shows
the average docking distance (in Å) from the heme co-factor.
Figure 3. (a) Differential regulatory machinery in gene duplication
events. In the case of an interspersed duplication, the duplicated gene
can contain a different core promoter and cis-regulatory elements that
could modulate its expression. A similar situation could occur with a tan-
dem duplication event in which a novel cis-regulatory element could
modulate both or only one of the duplicated genes. (b) Genomic hotspot
containing a cluster of differentially expressed genes related to nontarget-
site (NTS) resistance. The hypothetical regulatory model indicates a distal
cis-regulatory element that enhances the expression of multiple NTS
genes. Created with BioRender.com.
Herbicide cross resistance www.soci.org
Pest Manag Sci 2023 © 2023 The Authors.
Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry.
wileyonlinelibrary.com/journal/ps
5
15264998, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ps.7728, Wiley Online Library on [30/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
For ATAC-seq, Tn5 transposase is used to cut open chromatin,
leaving a staggered nick, followed by ligation of high-throughput
sequencing adapters to these regions. The nick is then repaired,
leaving a 9-bp duplication, and pair-end sequencing is performed
to ensure good coverage and unique alignments at the acquired
open regions.
49
ATAC-seq is now routinely being used for the
identication of CREs and motifs in animal and model plant
genomes, including at the level of single cells.
50
However, plants
pose an additional layer of complexity as a result of the Tn5 trans-
posase used in ATAC-seq being fully capable of accessing mito-
chondrial and chloroplast genomes, leading to a sequestration
of transposase and lower sequence output. To address this chal-
lenge, multiple methods have been developed.
51
DNA afnity purication sequencing (DAP-seq) is a method that
allows for the identication of DNA-binding proteins, transcrip-
tion factors and proteinDNA interactions by using recombinant
proteins to capture specic DNA sequences in vitro, followed by
high-throughput sequencing of the puried DNA. It is commonly
used to study the regulatory networks of transcription factors and
other DNA-binding proteins and to identify novel DNA-binding
proteins.
52
In order to conduct DAP-seq, expression vectors are created and
expressed in vitro to yield an afnity tag, such as HaloTag, fused to
TFs. To create a genomic DNA library, plant tissues are sonicated
to fragment the genomic DNA, and sequencing adapters are
ligated to the resulting DNA fragments. The HaloTag-TF is then
incubated with the DNA library, and the resulting TF/DNA com-
plexes are puried using HaloTag ligand-conjugated magnetic
beads. The unbound DNA is removed through washing, and
the TF-bound DNA is eluted and amplied via PCR for library
construction. This approach allows for consistent and precise
TF binding site determination and also can be adapted to exam-
ine the effect of DNA modicationsonTFbindingviathe
ampDAP-seq approach.
52
The choice between ATAC-seq and DAP-seq depends on the sci-
entic question being addressed. ATAC-seq is preferred when
investigating changes in chromatin accessibility and regulatory
regions in response to various stimuli or biological conditions.
DAP-seq, on the other hand, is useful when identifying DNA-
binding proteins, characterizing TF binding sites, or elucidating
regulatory networks.
44
ATAC-seq provides more efcient ways
for identifying CREs for a multitude of species and even different
cell types. However, without previous knowledge of TF sequence
motifs, regions identied from ATAC-seq cannot be validated. This
gap can be lled by DAP-seq to identify TF binding sites and serve
as a basis for ATAC-seq data interpretation (Fig. 1).
Recently, a new method for characterizing the cistrome
(complete set of regulatory elements in a genome) and chromatin
proling, called MNase-dened cistrome-occupancy analysis
(MOA-seq), was proposed as a one-step assay for identifying puta-
tive TF-binding sites within accessible chromatin regions.
53
How-
ever, MOA-seq has only been tested for maize, whereas DAP-seq
and ATAC-seq have been conducted for multiple species. Another
point to consider is distal CREs that can affect their target genes
via chromatin interaction.
44,54
Genome-wide studies utilizing
chromosome conformation capture approaches showed that
chromatin regions with similar epigenomic landscapes could
physically interact via phase separation.
55
Studies utilizing
high-throughput chromosome conformation capture (HI-C) tech-
nologies can be conducted to elucidate this gene expression
regulation via distal chromatin interaction.
5658
For instance, a
study in rice utilized HI-C technology to identify gene chromatin
loops and enhancer activities that could modulate gene expres-
sion.
59
Recent reference genomes are now routinely assembled
accompanied by HI-C, facilitating the identication of distal regu-
latory elements.
60
This approach is being used by the Interna-
tional Weeds Genomics Consortium to assemble the reference
genomes from multiple weed species.
61
These approaches have been applied to build cistromes and to
characterize chromatin states during different physiological
stages and stresses. For example, the cistrome was built for
Arabidopsis, where 1812 TFs and their binding sites were
characterized.
52
DAP-seq was used to characterize MADS-box
transcription factors in apple trees to investigate dormancy regu-
lation.
62
In studies on abiotic stresses, an interesting approach
involved combining ATAC-seq and RNA-seq to understand the
dynamics of chromatin accessibility in apple trees in response to
drought.
63
Another study also used the same approach to investi-
gate heat stress adaptation in rice, where available ATAC-seq and
RNA-seq data were used to identify key TFs for heat stress
responses.
64
In both DAP-seq and ATAC-seq experiments, careful
experimental design, appropriate controls, validation using alter-
native methods and rigorous data analyses are crucial to address
and mitigate potential pitfalls that can lead to, for example, false
positives.
In order to demonstrate a potential application of these data-
sets, we predicted the regulatory similarities of all CYP450s within
the Amaranthus tuberculatus genome. CYP450s were identied
through BLAST and hidden-Markov approaches to identify their
conserved domains, as described previously.
65
We utilized avail-
able cistrome and motif databases from Arabidopsis thaliana to
identify potential regulators and their similarities of each
CYP450.
52,66
We extracted the promoter region of each CYP450s
and performed a regulatory enrichment prediction analysis based
on motifs extracted from the database.
67
Briey, the regulation
prediction tool of PlantRegMap was used to identify enriched
motifs. Extracted promoter sequences from each of 185 CYP450s
identied in the A. tuberculatus genome were used as input. The
promotor region was dened as from 2000 bp before to 100 bp
after the transcription start, and we used a binding site prediction
threshold of 0.05.
Of the 185 CYP450s, 54 contained enriched motifs for 16 TFs
(Fig. 4). CYP450s containing motifs from the enriched TFs were
grouped to identify those with putative similar regulatory pat-
terns. Further analyses of each TF revealed they were members
of the families CPP, NAC, TCP and MIKC MADS-box. Results indi-
cated that MIKC MADS-box TF were the major regulators of
CYP450s in A. tuberculatus, consistent with their role as regulators
of CYP450s in other species.
68,69
These methods could be useful for weed scientists to improve
our understanding of the regulatory nature of NTS resistance.
Identifying commonalities in the co-regulation of NTS genes
could help to identify master regulators for herbicide stress
responses and mechanisms of enhanced metabolic resistance. It
is important to note that our example used the A. thaliana motif
database which, even though a model organism, will not perfectly
represent all the potential TF binding sites in A. tuberculatus.In
the future, weed scientists should build regulome and motif
database within their species of interest for more precise predic-
tions. Also, predicting gene regulation solely based on the pres-
ence of a motif in a promoter sequence can lead to false
positives, and other factors such as chromatin structure and
modulation will need to be considered as well. Another impor-
tant factor to consider is that promoter sequences can contain
www.soci.org LK Bobadilla, PJ Tranel
wileyonlinelibrary.com/journal/ps © 2023 The Authors.
Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry. Pest Manag Sci 2023
6
15264998, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ps.7728, Wiley Online Library on [30/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
multiple overlapping motifs, making it difcult to determine
which motif is responsible for the observed regulatory effect.
Lastly, gene regulation often involves the combinatorial action
of multiple regulatory elements. Predicting the effects of indi-
vidual motifs may not provide the full regulatory landscape of
a certain gene. Nevertheless, the example that we presented
illustrates the potential applications which can be done with
current available resources and how they could be further
developed.
Together with enzyme afnity predictions and metabolomics
studies, this system biology approach could provide a foundation
for predicting cross resistance and understanding NTS resistance
in weeds. Beyond the scope of characterizing NTS cross-resistance
regulation, this approach could also be used by weed scientists to
investigate weediness traits and species adaptation to climate
change, among other applications.
3.3 Gene regulatory networks
Gene regulatory networks (GRNs) are complex systems of inter-
acting genes and their regulatory elements, such as TFs and
CREs, that determine when and where genes are expressed in
an organism.
70
GRNs provide a systems biology perspective that
investigates the interactions among different components of
the plant system, including genes, proteins, metabolites and
environmental factors, to understand the underlying mecha-
nisms governing plant growth, development and response to
stress.
71,72
Studying GRNs is critical for comprehending how genes work
together to produce complex phenotypes, such as herbicide
stress responses, and how alterations in these networks can result
in resistance. To understand NTS resistance, GRNs can be
developed by collecting transcriptomics data to build a gene
co-expression network. Weighted co-expression analysis can
examine the global expression patterns for an NTS resistance phe-
notype.
73
Directionality can then be added to the co-expression
network to indicate which genes regulate other genes using
previously described methods for characterizing CREs and TFs.
This resource can serve as the core structure of the GRN, which
can later be incremented with other data types such as metabolo-
mics, proteomics and herbicide/enzyme afnity assays.
NTS resistance is an evolutionary phenomenon that can change
the topology of the herbicide response GRN. Understanding how
Figure 4. Regulatory prediction analysis via motif enrichment for Amaranthus tuberculatus cytochrome P450s (CYP450s). (a) All enriched transcription
factors (based on the Arabidopsis thaliana transcription factor database) and their respective families. Ribbons indicate the number of CPY450s predicted
to be regulated by that specic transcription factor. Ribbons are color-coded according to the transcription factor family. (b) Dendrogram of CPY450s in
the A. tuberculatus genome clustered according to their regulatory similarities, indicating groups of CYP450s that might be co-regulated.
Herbicide cross resistance www.soci.org
Pest Manag Sci 2023 © 2023 The Authors.
Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry.
wileyonlinelibrary.com/journal/ps
7
15264998, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ps.7728, Wiley Online Library on [30/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
the evolution of a trait affects a GRN is critical to predicting
how this evolution could occur.
74
These evolutionary patterns
can be used to identify subnetworks and understand resistance
mechanisms and their regulation. This systems biology approach
could help to predict novel venues for resistance, allowing weed
scientists to build predictive models for novel resistance types.
Generating data on regulatory elements, for example with
ATAC-seq, can help compare GRNs and similar NTS resistance
phenotypes across species to identify commonalities across mul-
tiple weed species.
The application of GRNs to understand abiotic stress has been
proposed numerous times.
75,76
For example, a study used RNA-
seq, ChIP-seq and bisulte sequencing to study the GRN varia-
tions of barley under abiotic stress.
77
Selection and evolution are
critical elements for the stability of a GRN, and abiotic stresses
can serve as signicant variables in it. Herbicides can exert strong
selection pressure, making them optimal candidates for testing
variabilities in GRN topologies in an abiotic stress context. Besides
the application towards understanding the regulatory nature of
herbicide resistance, GRNs also could be applied to expand our
knowledge of how herbicide safeners increase crop tolerance to
herbicides.
4 PULLING IT ALL TOGETHER
The Herbicide Resistance Action Committee has grouped herbi-
cides by mode-of-action, which makes it easy to select herbi-
cides for the purpose of mitigating TS herbicide resistance. We
envision an app that would utilize much more extensive data-
sets, such as those described in this review, to guide herbicide
selection also for mitigating NTS herbicide resistance. To mini-
mize the chance of herbicide-resistance evolution, herbicide
mix or rotation partners to manage any given weed species
should not:
(1) Have the same target site.
(2) Be metabolized by the same enzyme.
(3) Be metabolized by enzymes that are different but share a
common regulatory element.
(4) Be metabolized by enzymes that are different but are proxi-
mal in the genome.
Unsurprisingly, obtaining such datasets is not trivial; however, it
is feasible. We also acknowledge that, as stated before, our review
is focused on metabolic resistance despite our awareness of other
NTS resistance mechanisms. For example, we mentioned but did
not discuss cross resistance that could arise from acquired ability
to mitigate toxicity arising from herbicidal inhibition of its target
site (e.g. elevated antioxidant activity). Perfect prediction of cross
resistance will likely never be possible, but the steps and
approaches described herein would be a signicant advance.
We also fully acknowledge that much of what we are proposing
is model-based, and validation of modeled outcomes is required.
Nevertheless, generation of such datasets also could serve as the
starting point for implementing multi-scale models to under-
stand weed adaptation and herbicide-resistance evolution.
For instance, the CROPS In silico initiative focuses on using
multi-scale modeling for crop improvement and generating
new hypotheses for targeted engineering.
78
Although these
goals may seem utopian, weed scientists and others must work
together towards them to increase our predictability for the evo-
lution and adaptation of pests.
5 CONCLUSIONS
Weed and pest management scientists are delving into genomics
by creating reference genomes for multiple species. Moreover,
researchers could utilize protein structural computational model-
ing to develop prediction models for pesticide afnity with key
enzymes involved with NTS resistance. As a result, researchers
would be able to anticipate cross resistance and better under-
stand the regulatory nature of NTS resistance. For a more compre-
hensive understanding, researchers should prioritize generating
high-quality data to identify the regulation of crucial genes such
as CYP450s and GSTs by characterizing TFs and CREs. Investment
and time towards those analyses will allow weed scientists to
build predictive tools for herbicide management and rotation
recommendations based on cross-resistance potential. Advance-
ment in this area will enable weed scientists to elucidate the
unpredictability and promiscuity of NTS resistance and provide
genomics-based management recommendations to reduce the
occurrence of cross resistance.
AUTHOR CONTRIBUTIONS
LKB and PJT were involved in the conceptualization, writing and
editing of this paper.
ACKNOWLEDGEMENTS
This work was partially supported by the USDA National Institute
of Food and Agriculture (grant no. 2020-67013-31854).
CONFLICT OF INTEREST
The authors declare no conicts of interest.
DATA AVAILABILITY STATEMENT
Analyses conducted in this paper used publicly available data.
REFERENCES
1 Gaines TA, Duke SO, Morran S, Rigon CAG, Tranel PJ, Küpper A et al.,
Mechanisms of evolved herbicide resistance. J Biol Chem 295:
1030710330 (2020).
2 Rigon CA, Gaines TA, Küpper A and Dayan FE, Metabolism-based her-
bicide resistance, the major threat among the non-target site resis-
tance mechanisms. Outlooks Pest Manage 31:162168 (2020).
3 Maroli AS, Gaines TA, Foley ME, Duke SO, DoğramacıM, Anderson JV
et al., Omics in weed science: a perspective from genomics, tran-
scriptomics, and metabolomics approaches. Weed Sci 66:681695
(2018).
4 Tranel PJ and Horvath DP, Molecular biology and genomics: new tools
for weed science. Bioscience 59:207215 (2009).
5 Bobadilla LK, Giacomini DA, Hager AG and Tranel PJ, Characterization
and inheritance of dicamba resistance in a multiple-resistant water-
hemp (Amaranthus tuberculatus) population from Illinois. Weed Sci
70:413 (2022).
6 Giacomini DA, Patterson EL, Küpper A, Beffa R, Gaines TA and Tranel PJ,
Coexpression clusters and allele-specic expression in metabolism-
based herbicide resistance. Genome Biol Evol 12:22672278 (2020).
7 Moretti ML, Bobadilla LK and Hanson BD, Cross-resistance to diquat in
glyphosate/paraquat-resistant hairy eabane (Conyza bonariensis)
and horseweed (Conyza canadensis) and conrmation of 2,4-D resis-
tance in Conyza bonariensis.Weed Technol 35:554559 (2021).
8 Yanniccari M, Gigón R and Larsen A, Cytochrome P450 herbicide
metabolism as the main mechanism of cross-resistance to ACCase-
and ALS-inhibitors in Lolium spp. populations from Argentina: a
molecular approach in characterization and detection. Front Plant
Sci 11:600301 (2020).
www.soci.org LK Bobadilla, PJ Tranel
wileyonlinelibrary.com/journal/ps © 2023 The Authors.
Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry. Pest Manag Sci 2023
8
15264998, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ps.7728, Wiley Online Library on [30/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
9 Murphy BP and Tranel PJ, Target-site mutations conferring herbicide
resistance. Plan Theory 8:382 (2019).
10 Jugulam M and Shyam C, Non-target-site resistance to herbicides:
recent developments. Plan Theory 8:417 (2019).
11 Suzukawa AK, Bobadilla LK, Mallory-Smith C and Brunharo CA, Non-tar-
get-site resistance in Lolium spp. globally: a review, front.Plant Sci
11:609209 (2020).
12 Ha W, Yamaguchi T, Iwakami S, Sunohara Y and Matsumoto H, Compar-
ison of herbicide specicity of CYP81A cytochrome P450s from rice
and a multiple-herbicide resistant weed, Echinochloa phyllopogon.
Pest Manage Sci 78:42074216 (2022).
13 Han H, Yu Q, Beffa R, González S, Maiwald F, Wang J et al., Cytochrome
P450 CYP81A10v7 in Lolium rigidum confers metabolic resistance to
herbicides across at least ve modes of action. Plant J 105:7992
(2021).
14 Torra J, Montull JM, Taberner A, Onkokesung N, Boonham N and
Edwards R, Target-site and non-target-site resistance mechanisms
confer multiple andcross-resistance to ALS andACCase inhibiting her-
bicides in Lolium rigidum from Spain. Front Plant Sci 12:625138 (2021).
15 Reade JP, Milner LJ and Cobb AH, A role for glutathione S-transferases
in resistance to herbicides in grasses. Weed Sci 52:468474 (2004).
16 Franco-Ortega S, Goldberg-Cavalleri A, Walker A, Brazier-Hicks M,
Onkokesung N and Edwards R, Non-target site herbicide resistance
is conferred by two distinct mechanisms in black-grass (Alopecurus
myosuroides). Front Plant Sci 12:636652 (2021).
17 Hu T, He S, Yang G, Zeng H, Wang G, Chen Z et al., Isolation and char-
acterization of a rice glutathione S-transferase gene promoter regu-
lated by herbicides and hormones. Plant Cell Rep 30:539549 (2011).
18 Pang S, Duan L, Liu Z, Song X, Li X and Wang C, Co-induction of a
glutathione-S-transferase, a glutathione transporter and an ABC
transporter in maize by xenobiotics. PLoS One 7:e40712 (2012).
19 Giacomini DA, Gaines T, Beffa R and Tranel PJ, Optimizing RNA-seq
studies to investigate herbicide resistance. Pest Manage Sci 74:
22602264 (2018).
20 Lázár V, Nagy I, Spohn R, CsörgőB, Györkei Á, Nyerges Á et al., Genome-
wide analysis captures the determinants of the antibiotic cross-
resistance interaction network. Nat Commun 5:4352 (2014).
21 Khatami SA, Barmaki M, Alebrahim MT and Bajwa AA, Salicylic acid pre-
treatment reduces the physiological damage caused by the herbi-
cide mesosulfuron-methyl+ iodosulfuron-methyl in wheat (Triticum
aestivum). Agronomy 12:3053 (2022).
22 Khatooni M, Karimmojeni H, Zali AG and Tseng T-M, Salicylic acid
enhances tolerance of Valeriana ofcinalis L. to bentazon herbicide.
Ind Crops Prod 177:114495 (2022).
23 Ma LY, Zhang SH, Zhang JJ, Zhang AP, Li N, Wang XQ et al., Jasmonic
acids facilitate the degradation and detoxication of herbicide
isoproturon residues in wheat crops (Triticum aestivum). Chem Res
Toxicol 31:752761 (2018).
24 Kaya A and Doganlar ZB, Exogenous jasmonic acid induces stress
tolerance in tobacco (Nicotiana tabacum) exposed to imazapic.
Ecotoxicol Environ Saf 124:470479 (2016).
25 Suda H, Kubo T, Yoshimoto Y, Tanaka K, Tanaka S, Uchino A et al., Tran-
scriptionally linked simultaneous overexpression of P450 genes for broad-
spectrum herbicide resistance. Plant Physiol 192:30173029 (2023).
26 Wang J, Lian L, Qi J, Fang Y, Nyporko A, Yu Q et al., Metabolic resistance
to acetolactate synthase inhibitors in Beckmannia syzigachne: identi-
cation of CYP81Q32 and its transcription regulation. Plant J 115:
317334 (2023).
27 Veenstra TD, Omics in systems biology: current progress and future
outlook. Proteomics 21:2000235 (2021).
28 Rao VS, Srinivas K, Sujini GN and Kumar GN, Protein-protein interaction
detection: methods and analysis. Int J Proteomics 2014:147648
(2014).
29 Kato H, Computational prediction of cytochrome P450 inhibition and
induction. Drug Metab Pharmacokinet 35:3044 (2020).
30 Olofsson SK and Cars O, Optimizing drug exposure to minimize selec-
tion of antibiotic resistance. Clin Infect Dis 45:129136 (2007).
31 Maveyraud L and Mourey L, Protein X-ray crystallography and drug dis-
covery. Molecules 25:1030 (2020).
32 Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O et al.,
Highly accurate protein structure prediction with AlphaFold. Nature
596:583589 (2021).
33 Sanderson T, Bileschi ML, Belanger D and Colwell LJ, ProteInfer, deep
neural networks for protein functional inference. eLife 12:e80942
(2023).
34 Sobolev V and Edelman M, Modeling the quinone-B binding site of
the photosystem-II reaction center using notions of complemen-
tarity and contact-surface between atoms. Proteins 21:214225
(1995).
35 Nianiou-Obeidat I, Madesis P, Kissoudis C, Voulgari G, Chronopoulou E,
Tsaftaris A et al., Plant glutathione transferase-mediated stress toler-
ance: functions and biotechnological applications. Plant Cell Rep 36:
791805 (2017).
36 Krivák R and Hoksza D, P2Rank: machine learning based tool for rapid
and accurate prediction of ligand binding sites from protein struc-
ture. J Cheminf 10:39 (2018).
37 Trott O and Olson AJ, AutoDock Vina: improving the speed and accu-
racy of docking with a new scoring function, efcient optimization,
and multithreading. J Comput Chem 31:455461 (2010).
38 Eberhardt J, Santos-Martins D, Tillack AF and Forli S, Autodock vina
1.2.0: new docking methods, expanded force eld, and python bind-
ings. J Chem Inf Model 61:38913898 (2021).
39 Wang H, Wang Q, Liu Y, Liao X, Chu H, Chang H et al., PCPD: plant
cytochrome P450 database and web-based tools for structural
construction and ligand docking. Synth Syst Biotechnol 6:102109
(2021).
40 Matthews S, Belcher JD, Tee KL, Girvan HM, McLean KJ, Rigby SE et al.,
Catalytic determinants of alkene production by the cytochrome
P450 peroxygenase OleTJE. J Biol Chem 292:51285143 (2017).
41 Hammerer L, Winkler CK and Kroutil W, Regioselective biocatalytic
hydroxylation of fatty acids by cytochrome P450s. Catal Lett 148:
787812 (2018).
42 Galli M, Feng F and Gallavotti A, Mapping regulatory determinants in
plants. Front Genet 11:591194 (2020).
43 Marand AP, Zhang T, Zhu B and Jiang J, Towards genome-wide predic-
tion and characterization of enhancers in plants. Biochim Biophys
Acta, Gene Regul Mech 1860:131139 (2017).
44 Schmitz RJ, Grotewold E and Stam M, Cis-regulatory sequences in
plants: their importance, discovery, and future challenges. Plant Cell
34:718741 (2022).
45 Werck-Reichhart D, Promiscuity, a driver of plant cytochrome P450
evolution? Biomolecules 13:394 (2023).
46 Park PJ, ChIPseq: advantages and challenges of a maturing technol-
ogy. Nat Rev Genet 10:669680 (2009).
47 Grandi FC, Modi H, Kampman L and Corces MR, Chromatin accessibility
proling by ATAC-seq. Nat Protoc 17:15181552 (2022).
48 Bajic M, Maher KA and Deal RB, Identication of open chromatin regions
in plant genomes using ATAC-seq, in Plant Chromatin Dynamics.
Methods in Molecular Biology, Chromatin Dynamics, ed. by Bemer M
and Baroux C. Methods in Molecular, Humana Press, New York, Vol.
1675, pp. 183201 (2018).
49 Li Z, Schulz MH, Look T, Begemann M, Zenke M and Costa IG, Identi-
cation of transcription factor binding sites using ATAC-seq. Genome
Biol 20:121 (2019).
50 Marand AP and Schmitz RJ, Single-cell analysis of cis-regulatory ele-
ments. Curr Opin Plant Biol 65:102094 (2022).
51 Lu Z, Hofmeister BT, Vollmers C, DuBois RM and Schmitz RJ, Combining
ATAC-seq with nuclei sorting for discovery of cis-regulatory regions
in plant genomes. Nucleic Acids Res 45:e41 (2017).
52 O'Malley RC, Huang SC, Song L, Lewsey MG, Bartlett A, Nery JR et al.,
Cistrome and epicistrome features shape the regulatory DNA land-
scape. Cell 165:12801292 (2016).
53 Savadel SD, Hartwig T, Turpin ZM, Vera DL, Lung P-Y, Sui X et al., The
native cistrome and sequence motif families of the maize ear. PLoS
Genet 17:e1009689 (2021).
54 Louwers M, Bader R, Haring M, van Driel R, de Laat W and Stam M, Tis-
sue- and expression levelspecic chromatin looping at maize b1
epialleles. Plant Cell 21:832842 (2009).
55 Ricci WA, Lu Z, Ji L, Marand AP, Ethridge CL, Murphy NG et al., Wide-
spread long-range cis-regulatory elements in the maize genome.
Nat Plants 5:12371249 (2019).
56 Forcato M, Nicoletti C, Pal K, Livi CM, Ferrari F and Bicciato S, Compar-
ison of computational methods for Hi-C data analysis. Nat Methods
14:679685 (2017).
57 Fraser J, Williamson I, Bickmore WA and Dostie J, An overview of
genome organization and how we got there: from FISH to Hi-C.
Microbiol Mol Biol Rev 79:347372 (2015).
58 Van Berkum NL, Lieberman-Aiden E, Williams L, Imakaev M, Gnirke A,
Mirny LA et al., Hi-C: a method to study the three-dimensional archi-
tecture of genomes. J Visualized Exp 39:e1869 (2010).
Herbicide cross resistance www.soci.org
Pest Manag Sci 2023 © 2023 The Authors.
Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry.
wileyonlinelibrary.com/journal/ps
9
15264998, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ps.7728, Wiley Online Library on [30/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
59 Dong Q, Li N, Li X, Yuan Z, Xie D, Wang X et al., Genome-wide Hi-C anal-
ysis reveals extensive hierarchical chromatin interactions in rice.
Plant J 94:11411156 (2018).
60 Lightfoot DJ, Jarvis DE, Ramaraj T, Lee R, Jellen EN and Maughan PJ,
Single-molecule sequencing and Hi-C-based proximity-guided
assembly of amaranth (Amaranthus hypochondriacus) chromo-
somes provide insights into genome evolution. BMC Biol 15:74
(2017).
61 Montgomery JS, Morran S, MacGregor DR, McElroy JS, Neve P, Neto C
et al., The International Weed Genomics Consortium: Community
resources for weed genomics research. bioRxiv:202307 (2023).
doi:10.1101/2023.07.19.549613
62 da Silveira FV, Severing E, Lai X, Estevan J, Farrera I, Hugouvieux V et al.,
Unraveling the role of MADS transcription factor complexes in apple
tree dormancy. New Phytol 232:20712088 (2021).
63 Wang S, He J, Deng M, Wang C, Wang R, Yan J et al., Integrating ATAC-
seq and RNA-seq reveals the dynamics of chromatin accessibility
and gene expression in apple response to drought. Int J Mol Sci 23:
11191 (2022).
64 Qiu F, Zheng Y, Lin Y, Woldegiorgis ST, Xu S, Feng C et al., Integrated
ATAC-seq and RNA-seq data analysis to reveal osbZIP14 function in
rice in response to heat stress. Int J Mol Sci 24:5619 (2023).
65 Li Y and Wei K, Comparative functional genomics analysis of cyto-
chrome P450 gene superfamily in wheat and maize. BMC Plant Biol
20:93 (2020).
66 Jin J, Tian F, Yang D-C, Meng Y-Q, Kong L, Luo J et al., PlantTFDB 4.0:
toward a central hub for transcription factors and regulatory interac-
tions in plants. Nucleic Acids Res 45:D1040D1045 (2017).
67 Tian F, Yang D-C, Meng Y-Q, Jin J and Gao G, PlantRegMap: charting
functional regulatory maps in plants. Nucleic Acids Res 48:D1104
D1113 (2020).
68 Fei X, Shi Q, Qi Y, Wang S, Lei Y, Hu H et al., ZbAGL11, a class D MADS-
box transcription factor of Zanthoxylum bungeanum, is involved in
sporophytic apomixis. Hortic Res 8:23 (2021).
69 GreenupAG,SasaniS,OliverSN,TalbotMJ,DennisES,HemmingMNet al.,
ODDSOC2 is a MADS box oral repressor that is down-regulated by ver-
nalization in temperate cereals. Plant Physiol 153:10621073 (2010).
70 Krouk G, Lingeman J, Colon AM, Coruzzi G and Shasha D, Gene
regulatory networks in plants: learning causality from time and per-
turbation. Genome Biol 14:17 (2013).
71 Cramer GR, Urano K, Delrot S, Pezzotti M and Shinozaki K, Effects of abi-
otic stress on plants: a systems biology perspective. BMC Plant Biol
11:163 (2011).
72 Kitano H, Systems biology: a brief overview. Science 295:16621664 (2002).
73 Langfelder P and Horvath S, WGCNA: an R package for weighted corre-
lation network analysis. BMC BMC Bioinf 9:113 (2008).
74 Jones DM and Vandepoele K, Identication and evolution of gene reg-
ulatory networks: insights from comparative studies in plants. Curr
Opin Plant Biol 54:4248 (2020).
75 Castelán-Muñoz N, Herrera J, Cajero-Sánchez W, Arrizubieta M, Trejo C,
García-Ponce B et al., MADS-box genes are key components of
genetic regulatory networks involved in abiotic stress and plastic
developmental responses in plants. Front Plant Sci 10:853 (2019).
76 Hong-bo S, Plant gene regulatory network system under abiotic stress.
Acta Biol Szeged 50:19 (2006).
77 Xu Q, Huang S, Guo G, Yang C, Wang M, Zeng X et al., Inferring regula-
tory element landscapes and gene regulatory networks from inte-
grated analysis in eight hulless barley varieties under abiotic stress.
BMC Genomics 23:110 (2022).
78 Marshall-Colon A, Long SP, Allen DK, Allen G, Beard DA, Benes B et al.,
Crops in silico: generating virtual crops using an integrative and
multi-scale modeling platform. Front Plant Sci 8:786 (2017).
www.soci.org LK Bobadilla, PJ Tranel
wileyonlinelibrary.com/journal/ps © 2023 The Authors.
Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry. Pest Manag Sci 2023
10
15264998, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/ps.7728, Wiley Online Library on [30/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
... Identifying how DE genes are regulated could aid in identifying the actual expression modulator. Understanding the regulatory nature of the involved genes is key to building further predictions for crossresistance and creating strategies to minimize it (Bobadilla and Tranel 2023). ...
... There is still a need to better understand this phenomenon whereby a large region can have multiple genes with altered expression. However, the presence of cis-regulatory elements can play an important role in altering the expression of multiple, proximal genes (Bobadilla and Tranel 2023). ...
Article
Full-text available
Waterhemp ( Amaranthus tuberculatus (Moq.) Sauer) is one of the most troublesome weeds in the US. An A. tuberculatus population (CHR) was identified in Illinois, US, as resistant to herbicides from six different site-of-action groups. Recently, the same population was also recognized as dicamba resistant. This study aimed to identify key resistance genes and the putative dicamba resistance mechanism in A. tuberculatus via transcriptomics analysis. Multiple differentially expressed genes and co-expression gene modules were identified as associated with dicamba resistance. Specifically, genes encoding glutathione-S-transferases, ABC transporters, peroxidases, and UDP-glycosyltransferases were identified. Results indicated enhanced oxidative stress tolerance as the primary mechanism for reducing dicamba toxicity. Results also point to potential glycosylation via UDP-glycosyltransferases and conjugation via glutathione-S-transferases of dicamba and its byproducts. This is the first transcriptomics characterization of dicamba resistance in A. tuberculatus. Multiple non-target-site resistance genes were identified, indicating a cross-resistance pattern in the CHR population leading to a putative-enhanced oxidative stress response. Regions of multiple differentially expressed genes (i.e., genomic hotspots) across the A. tuberculatus genome corroborate with previous results and potentially add to the complexity of non-target-site resistance traits.
... Herbicide resistance is a big and growing challenge for sustainable agriculture. 1,2 As of today, 272 weed species are reported to have evolved herbicide resistance globally. 3 In the Western Hemisphere, after the introduction of glyphosate-tolerant crops in the late 1990s, yield-robbing weeds such as waterhemp (Amaranthus tuberculatus) and Palmer amaranth (A. ...
Article
Full-text available
Epyrifenacil is a novel PPO‐inhibiting herbicide discovered and developed by Sumitomo Chemical. Epyrifenacil belongs to the pyrimidinedione chemical class and has a unique three‐ring structure. It is systemically active on a broad range of weeds including grass weeds and some target‐site‐based PPO‐inhibitor resistant broadleaf weeds. Its systemic action is mediated by a phloem movement of the active form of epyrifenacil. In addition, epyrifenacil's vapor action is sufficiently low to not cause an off‐target movement to nontarget sensitive crops. It is expected that epyrifenacil will contribute to global food production in the near future. © 2024 The Author(s). Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry.
... These mechanisms may include genes that are members of large gene families, including genes encoding enzymes involved in herbicide degradation [11,12,19]. It is evident that cross-resistant biotypes, resistant to different herbicides [20][21][22][23], and multi-resistant weed biotypes, whose resistance to different herbicides is caused by more than one mechanism [22,24,25], are particularly dangerous. The spread of herbicide resistance among weeds reduces the effectiveness of crop protection and in general casts doubt on the prospects of further using the chemical method of weed control. ...
Article
Systemic acquired acclimation and resistance are vital physiological mechanisms, essential for plants to survive challenging conditions, including herbicide stress. Harmonizing this adaptation involves a series of complex communication pathways. Hydrogen peroxide (H2O2) metabolism might play pivotal roles in orchestrating weeds’ acclimation and defense responses. In the context of herbicide resistance, the interaction between H2O2 and key stress signaling pathways is crucial in understanding weed physiology and developing effective management strategies. This dynamic interplay might significantly influence how weeds develop resistance to the various challenges posed by herbicides. Moreover, the production and eradication of H2O2 can be highly compartmentalized, depending on the type of herbicide exposure. Till date there have been no studies aiming to explore/discuss these possibilities. Therefore, in this mini-review, our objective is to delve into the potentialities and recent advancements regarding H2O2-mediated signaling of transcriptomic changes during herbicide stress.
Article
Full-text available
Weed management in agriculture is hampered by inefficient intensive methods, such as monoculture, deep plowing, and herbicides, leading to health and environmental problems. Furthermore, the prevalence of herbicide-resistant weed ecotypes in the Mediterranean, particularly in France (with over 61 ecotypes), Spain (41), and Italy (37), is a major concern, with a significant proportion of herbicides in the region. In this study, we examined the benefits of adopting agroecology as a sustainable approach for weed management in the Mediterranean region. Agroecology offers a variety of techniques and practices to improve sustainability and weed management, while preserving ecological balance and biodiversity. However, solving these challenges is multifactorial and depends on local specificities, predominant weed species, crops, sowing dates, and pedo-climatic factors. In addition, this study included a systematic analysis of agroecological weed management in Mediterranean countries, assessing the effectiveness of existing practices, and identifying areas requiring further exploration in agroecosystems. A bibliometric analysis was also included to assess the literature on agroecology and weed management quantitatively, identifying major trends, influential studies, and research gaps. The bibliometric analysis highlighted the importance of alternative herbicides in Mediterranean “weed” (with a link strength of 44), “agroecology” (22), and “biodiversity” (16). Italy has the strongest collaboration network, with a link strength of 61, followed by Turkey (44), and France (42). Using specific keywords to agroecological practices for weed management in Scopus, France worked the most in this context (around 25% of studies), followed by Spain (17%) and Italy (17%), while all other countries contributed to less than 40% of studies carried out in the Mediterranean context. Clearly, it is imperative to foster collaboration between Mediterranean countries to develop effective and sustainable weed control strategies. Understanding the challenges of herbicide-resistant weeds, exploring their reasons and mechanisms, and using systematic studies and bibliometric analyses will help to develop effective strategies for managing weeds in the Mediterranean. Agroecological management favors effective control, while promoting healthy and sustainable ecosystems, preserving biodiversity, and ensuring long-term food security.
Preprint
Full-text available
The International Weed Genomics Consortium is a collaborative group of researchers focused on developing genomic resources for the study of weedy plants. Weeds are attractive systems for basic and applied research due to their impacts on agricultural systems and capacity to swiftly adapt in response to anthropogenic selection pressures. Our goal is to use genomic information to develop sustainable and effective weed control methods and to provide insights about biotic and abiotic stress tolerance to assist crop breeding. Here, we outline resources under development by the consortium and highlight areas of research that will be impacted by these enabling resources.
Article
Full-text available
Broad-spectrum herbicide resistance (BSHR), often linked to weeds with metabolism-based herbicide resistance, poses a threat to food production. Past studies have revealed that overexpression of catalytically-promiscuous enzymes explains BSHR in some weeds; however, the mechanism of BSHR expression remains poorly understood. Here, we investigated the molecular basis of high-level resistance to diclofop-methyl in BSHR late watergrass (Echinochloa phyllopogon) found in the US, which cannot be solely explained by the overexpression of promiscuous cytochrome P450 monooxygenases CYP81A12/21. The BSHR late watergrass line rapidly produced two distinct hydroxylated-diclofop-acids, only one of which was the major metabolite produced by CYP81A12/21. RNA-seq and subsequent RT-qPCR-based segregation screening identified the transcriptionally-linked overexpression of a gene, CYP709C69, with CYP81A12/21 in the BSHR line. The gene conferred diclofop-methyl resistance in plants and produced another hydroxylated-diclofop-acid in yeast (Saccharomyces cerevisiae). Unlike CYP81A12/21, CYP709C69 showed no other herbicide-metabolizing function except for a presumed clomazone-activating function. The overexpression of the three herbicide-metabolizing genes was also identified in another BSHR late watergrass in Japan, suggesting a convergence of BSHR evolution at the molecular level. Synteny analysis of the P450 genes implied that they are located at mutually independent loci, which supports the idea that a single trans-element regulates the three genes. We propose that transcriptionally-linked simultaneous overexpression of herbicide-metabolizing genes enhances and broadens the metabolic resistance in weeds. The convergence of the complex mechanism in BSHR late watergrass from two countries suggests that BSHR evolved through co-opting a conserved gene-regulatory system in late watergrass.
Article
Full-text available
Transcription factors (TFs) play critical roles in mediating the plant response to various abiotic stresses, particularly heat stress. Plants respond to elevated temperatures by modulating the expression of genes involved in diverse metabolic pathways, a regulatory process primarily governed by multiple TFs in a networked configuration. Many TFs, such as WRKY, MYB, NAC, bZIP, zinc finger protein, AP2/ERF, DREB, ERF, bHLH, and brassinosteroids, are associated with heat shock factor (Hsf) families, and are involved in heat stress tolerance. These TFs hold the potential to control multiple genes, which makes them ideal targets for enhancing the heat stress tolerance of crop plants. Despite their immense importance, only a small number of heat-stress-responsive TFs have been identified in rice. The molecular mechanisms underpinning the role of TFs in rice adaptation to heat stress still need to be researched. This study identified three TF genes, including OsbZIP14, OsMYB2, and OsHSF7, by integrating transcriptomic and epigenetic sequencing data analysis of rice in response to heat stress. Through comprehensive bioinformatics analysis, we demonstrated that OsbZIP14, one of the key heat-responsive TF genes, contained a basic-leucine zipper domain and primarily functioned as a nuclear TF with transcriptional activation capability. By knocking out the OsbZIP14 gene in the rice cultivar Zhonghua 11, we observed that the knockout mutant OsbZIP14 exhibited dwarfism with reduced tiller during the grain-filling stage. Under high-temperature treatment, it was also demonstrated that in the OsbZIP14 mutant, the expression of the OsbZIP58 gene, a key regulator of rice seed storage protein (SSP) accumulation, was upregulated. Furthermore, bimolecular fluorescence complementation (BiFC) experiments uncovered a direct interaction between OsbZIP14 and OsbZIP58. Our results suggested that OsbZIP14 acts as a key TF gene through the concerted action of OsbZIP58 and OsbZIP14 during rice filling under heat stress. These findings provide good candidate genes for genetic improvement of rice but also offer valuable scientific insights into the mechanism of heat tolerance stress in rice.
Article
Full-text available
Predicting the function of a protein from its amino acid sequence is a long-standing challenge in bioinformatics. Traditional approaches use sequence alignment to compare a query sequence either to thousands of models of protein families or to large databases of individual protein sequences. Here we introduce ProteInfer, which instead employs deep convolutional neural networks to directly predict a variety of protein functions - EC numbers and GO terms - directly from an unaligned amino acid sequence. This approach provides precise predictions which complement alignment-based methods, and the computational efficiency of a single neural network permits novel and lightweight software interfaces, which we demonstrate with an in-browser graphical interface for protein function prediction in which all computation is performed on the user's personal computer with no data uploaded to remote servers. Moreover, these models place full-length amino acid sequences into a generalised functional space, facilitating downstream analysis and interpretation. To read the interactive version of this paper, please visit https://google-research.github.io/proteinfer/.
Article
Full-text available
Plant cytochrome P450 monooxygenases were long considered to be highly substrate-specific, regioselective and stereoselective enzymes, in this respect differing from their animal counterparts. The functional data that have recently accumulated clearly counter this initial dogma. Highly promiscuous P450 enzymes have now been reported, mainly in terpenoid pathways with functions in plant adaptation, but also some very versatile xenobiotic/herbicide metabolizers. An overlap and predictable interference between endogenous and herbicide metabolism are starting to emerge. Both substrate preference and permissiveness vary between plant P450 families, with high promiscuity seemingly favoring retention of gene duplicates and evolutionary blooms. Yet significant promiscuity can also be observed in the families under high negative selection and with essential functions, usually enhanced after gene duplication. The strategies so far implemented, to systematically explore P450 catalytic capacity, are described and discussed.
Article
Full-text available
Background The cis-regulatory element became increasingly important for resistance breeding. There were many DNA variations identified by resequencing. To investigate the links between the DNA variations and cis-regulatory element was the fundamental work. DNA variations in cis-regulatory elements caused phenotype variations in general. Results We used WGBS, ChIP-seq and RNA-seq technology to decipher the regulatory element landscape from eight hulless barley varieties under four kinds of abiotic stresses. We discovered 231,440 lowly methylated regions (LMRs) from the methylome data of eight varieties. The LMRs mainly distributed in the intergenic regions. A total of 97,909 enhancer-gene pairs were identified from the correlation analysis between methylation degree and expression level. A lot of enriched motifs were recognized from the tolerant-specific LMRs. The key transcription factors were screened out and the transcription factor regulatory network was inferred from the enhancer-gene pairs data for drought stress. The NAC transcription factor was predicted to target to TCP, bHLH, bZIP transcription factor genes. We concluded that the H3K27me3 modification regions overlapped with the LMRs more than the H3K4me3. The variation of single nucleotide polymorphism was more abundant in LMRs than the remain regions of the genome. Conclusions Epigenetic regulation is an important mechanism for organisms to adapt to complex environments. Through the study of DNA methylation and histone modification, we found that many changes had taken place in enhancers and transcription factors in the abiotic stress of hulless barley. For example, transcription factors including NAC may play an important role. This enriched the molecular basis of highland barley stress response.
Article
Full-text available
Chemical herbicides are the most common method of weed control in crops, but they can also negatively affect the host crops, such as wheat (Triticum aestivum L.). The damage caused to the crop plants is often temporary and minor, but sometimes, it can be more substantial, requiring remedial measures. Salicylic acid (SA) is a plant hormone widely used to promote plant growth and to mitigate oxidative stress through its exogenous application. We evaluated the role of exogenously applied SA (as a pre-treatment) in ameliorating the oxidative damage caused by the herbicide mesosulfuron-methyl + iodosulfuron-methyl in wheat plants. The herbicide disrupted the physiological function of plants by affecting several enzymatic antioxidants. The hydrogen peroxide (H2O2) and malondialdehyde (MDA) contents increased at herbicide concentrations higher than 18 g ai ha−1 compared with the untreated control. However, the SA decreased the H2O2 and MDA contents compared with plants that were not treated with SA prior to the herbicide application. The activity of superoxide dismutase (SOD) and polyphenol oxidase (PPO) enzymes increased with increasing rates of the herbicide, as well as over time, regardless of the SA treatment. The activity of catalase (CAT) increased up to the herbicide rate of 18 g ai ha−1 and then decreased at the higher rates, while SA pre-treatment enhanced the CAT activity. The activities of ascorbate peroxidase, peroxidase, and glutathione-S-transferase enzymes generally increased in response to the herbicide application and SA pre-treatment, but fluctuated across different days of sampling following the herbicide application. Herbicide stress also induced high levels of proline production in wheat leaves as compared with the untreated control, while SA pre-treatment decreased the proline contents. Overall, the pre-treatment with different concentrations of SA mitigated the herbicide damage to the physiological functions by regulating the enzymatic antioxidants.
Article
Full-text available
Drought resistance in plants is influenced by multiple signaling pathways that involve various transcription factors, many target genes, and multiple types of epigenetic modifications. Studies on epigenetic modifications of drought focus on DNA methylation and histone modifications, with fewer on chromatin remodeling. Changes in chromatin accessibility can play an important role in abiotic stress in plants by affecting RNA polymerase binding and various regulatory factors. However, the changes in chromatin accessibility during drought in apples are not well understood. In this study, the landscape of chromatin accessibility associated with the gene expression of apple (GL3) under drought conditions was analyzed by Assay for Transposase Accessible Chromatin with high-throughput sequencing (ATAC-seq) and RNA-seq. Differential analysis between drought treatment and control identified 23,466 peaks of upregulated chromatin accessibility and 2447 peaks of downregulated accessibility. The drought-induced chromatin accessibility changed genes were mainly enriched in metabolism, stimulus, and binding pathways. By combining results from differential analysis of RNA-seq and ATAC-seq, we identified 240 genes with higher chromatin accessibility and increased gene expression under drought conditions that may play important functions in the drought response process. Among them, a total of nine transcription factor genes were identified, including ATHB7, HAT5, and WRKY26. These transcription factor genes are differentially expressed with different chromatin accessibility motif binding loci that may participate in apple response to drought by regulating downstream genes. Our study provides a reference for chromatin accessibility under drought stress in apples and the results will facilitate subsequent studies on chromatin remodelers and transcription factors.
Article
Frequent herbicide use selects for herbicide resistance in weeds. Cytochrome P450s are important detoxification enzymes responsible for herbicide resistance in plants. We identified and characterized a candidate P450 gene (BsCYP81Q32) from the problematic weed Beckmannia syzigachne to test whether it conferred metabolic resistance to the acetolactate synthase (ALS)-inhibiting herbicides mesosulfuron-methyl, bispyribac-sodium, and pyriminobac-methyl. Transgenic rice overexpressing BsCYP81Q32 was resistant to the three herbicides. Equally, rice overexpressing the rice ortholog gene OsCYP81Q32 was more resistant to mesosulfuron-methyl. Conversely, an OsCYP81Q32 gene knockout generated using CRISPR/Cas9 enhanced mesosulfuron-methyl sensitivity in rice. Overexpression of the BsCYP81Q32 gene resulted in enhanced mesosulfuron-methyl metabolism in transgenic rice seedlings via O-demethylation. The major metabolite, demethylated mesosulfuron-methyl, was chemically synthesized and displayed reduced herbicidal effect in plants. Moreover, a transcription factor (BsTGAL6) was identified and shown to bind a key region in the BsCYP81Q32 promoter for gene activation. Inhibition of BsTGAL6 expression by salicylic acid treatment in B. syzigachne plants reduced BsCYP81Q32 expression and consequently changed the whole plant response to mesosulfuron-methyl. Sequence polymorphisms in an important region of the BsTGAL6 promoter may explain the higher expression of BsTGAL6 in resistant versus susceptible B. syzigachne plants. Collectively, our study reveals the evolution of an herbicide-metabolizing and resistance-endowing P450 and its transcription regulation in an economically important weedy plant species.