You are on page 1of 46

JBC Papers in Press. Published on May 19, 2020 as Manuscript REV120.

013572
The latest version is at https://www.jbc.org/cgi/doi/10.1074/jbc.REV120.013572

Mechanisms of evolved herbicide resistance


Todd A. Gaines,1 Stephen O. Duke,2* Sarah Morran,1 Carlos A. G. Rigon,1 Patrick J. Tranel,3 Anita
Küpper,4 and Franck E. Dayan1

From the 1Agricultural Biology Department, Colorado State University, Fort Collins, CO 80523, USA;
2
National Center for Natural Products Research, School of Pharmacy, University of Mississippi, 38677
USA; 3Department of Crop Sciences, University of Illinois, Urbana, IL 61801, USA; 4Bayer AG,
CropScience Division, Frankfurt am Main, Germany

Running title: Mechanisms of evolved herbicide resistance

*To whom correspondence should be addressed: Stephen O. Duke: National Center for Natural Products
Research, School of Pharmacy, University of Mississippi, 38677 USA; sduke@olemiss.edu; Tel. +1
(662) 832-1594.

Keywords: Cytochrome P450, glutathione-S-transferase, herbicide metabolism, reduced translocation,

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


target-site resistance, non-target-site resistance, cross-resistance, multiple resistance, plant evolution,
selection pressure

ABSTRACT the individual level to produce higher resistance


The widely successful use of synthetic herbicides levels. The vast array of herbicide-resistance
over the past 70 years has imposed strong and mechanisms for generalist (NTSR) and specialist
widespread selection pressure, leading to the (TSR and some NTSR) adaptations that have
evolution of herbicide resistance in hundreds of evolved over a few decades illustrate the
weed species. Both target-site resistance (TSR) and evolutionary resilience of weed populations to
non-target-site resistance (NTSR) mechanisms extreme selection pressures. These evolutionary
have evolved to most herbicide classes. TSR often processes drive herbicide and herbicide-resistant
involves mutations in genes encoding the protein crop development and resistance management
targets of herbicides, affecting the binding of the strategies.
herbicide either at or near catalytic domains or in
regions affecting access to them. Most of these Plants that are not wanted at a particular time
mutations are non-synonymous single-nucleotide and/or place (weeds) have been managed mostly
polymorphisms, but polymorphisms in more than with synthetic herbicides for more than 70 years.
one codon or entire codon deletions have also Before that time, weeds were largely controlled
evolved. Some herbicides bind multiple proteins, with laborious manual weeding and often
making TSR mechanisms more difficult to evolve. environmentally damaging tillage. Adoption of
Increased amounts of protein target, by increased synthetic herbicides reduced the cost and
gene expression or by gene duplication, is an increased the efficacy of weeding, thereby
important, albeit less common, TSR mechanism. contributing to the yield increases and efficiency
NTSR mechanisms include reduced absorption or of agriculture seen since the middle of the last
translocation and increased sequestration or century. However, as with antibiotics, the utility of
metabolic degradation. The mechanisms that can synthetic weed killers is being threatened by wide-
contribute to NTSR are complex and often involve spread evolution of resistance to most chemical
genes that are members of large gene families. For classes of herbicides that act on most of the more
example, enzymes involved in herbicide than 25 molecular targets of current commercial
metabolism–based resistances include cytochromes herbicides (1).
P450, glutathione-S-transferases, glucosyl and The main goal of this review is to provide an
other transferases, aryl acylamidase, and others. update on the rapidly evolving topic of
Both TSR and NTSR mechanisms can combine at mechanisms of evolved herbicide resistance in

1
weeds, as there has been no recent comprehensive mechanisms include reduced herbicide uptake and
review on this topic. We hope that this review will translocation, increased herbicide sequestration,
inspire plant molecular biologists and biochemists and enhanced degradation or metabolism of the
to determine more clearly how these resistance herbicide to less-toxic compounds. On the other
mechanisms evolve and the biochemical and hand, TSR mechanisms alter the amino acid
physiological changes in weeds imparted by sequence and/or expression level of the target
resistance mutations. Such information will be enzyme, reducing the herbicide’s ability to inhibit
useful in resistance management and in design of the enzyme or requiring a greater herbicide
herbicide molecules for which evolution of concentration to achieve adequate inhibition.
resistance is more problematic for weeds. We Under intense selection pressure from highly
provide discussions of the implications of effective herbicides, all possible mechanisms
herbicide resistance mechanisms for the conferring a greater chance of survival and
development of new herbicides and herbicide- reproduction to the individual may be selected.
resistant crops. More than one mechanism may be operating to
We summarize the wide array of resistance confer resistance, including combinations of TSR
mechanisms that weeds have evolved to survive and NTSR mechanisms. Several resistance
the intense selection pressure imparted by mechanisms can co-exist within a species, within
commercial herbicides. Herbicide resistance a population, and even within a single individual.

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


mechanisms can be broadly divided into two Different resistance mechanisms combine through
categories, referred to as target-site resistance cross-pollination between individuals. Species
(TSR) mechanisms and non-target-site resistance with high levels of cross-pollination are more
(NTSR) mechanisms. Herbicide efficacy is likely to accumulate diverse resistance
generally dependent on how much of the herbicide mechanisms (to a single herbicide, or to multiple
enters a plant cell and how long its active form herbicides), and this process can occur more
remains available to interact with its site of action rapidly than in self-pollinated species.
(also called target-site). A full understanding of
the mechanism of resistance to a herbicide requires Target site mechanisms
understanding that herbicide’s mechanism of
action (MOA). MOA of herbicides is not A single nucleotide mutation in the gene
discussed in detail in this review, but we have encoding a protein bound by a herbicide can result
provided a summary of the molecular targets and in a single amino acid change, disrupting the
MOAs of the herbicides mentioned in this review ability of the herbicide to bind to the protein
(Table 1). There are 26 molecular target sites of without disabling the enzyme function. Generally,
the more than 260 commercial herbicide active there are few amino acids in or near the herbicide
ingredients that are recognized by the Herbicide binding-site of most target-site proteins where an
Resistance Action Committee, an industry amino acid substitution will result in TSR. Most
organization that monitors herbicide resistance target-site mutations occur in or near the
(2). Of these target sites, resistance has evolved herbicide-binding site, but some mutations occur
globally (in 92 crops in 70 countries) to 167 elsewhere in the protein structure. Target-site
herbicides representing about 23 of these targets, mutations are identified by the amino acid and its
with 512 weed species evolving resistance to one position in the protein, numbered from the
of more herbicides (3). This review will not protein’s start codon. In some cases, the mutation
provide an encyclopedic elaboration for each of can confer very high-level resistance, and in other
these cases but will give examples of different cases, the mutation confers lower level (but
mechanisms of resistance that often cross significant) resistance. Some target-site mutations
herbicide classes. reduce normal enzymatic function, and other
NTSR mechanisms include all mechanisms mutations retain nearly full enzymatic function. In
that reduce the concentration of active herbicide addition to single nucleotide substitutions, whole-
remaining available to interact with the target site codon deletions can also reduce herbicide binding
protein, as well as mechanisms that allow the plant to the target-site enzyme. It is important to note
to cope with inhibition of the target site. NTSR that the same molecular mechanism (e.g., a single

2
nucleotide polymorphism (SNP) leading to an Additional psbA resistance-imparting mutations
amino acid change) forms the basis of resistance include Val219Ile, Asn266Thr, Phe255Ile, and
for many different herbicide sites of action. Ala251Val.
TSR mechanisms are specialist mechanisms, The dinitroaniline herbicides (such as
specific to a single site of action. Whether a trifluralin and oryzalin) bind to plant tubulin
specific target-site mutation that confers resistance protein and disrupt meristem development by
to a given herbicide also confers resistance to depolymerizing microtubules (1). The first
different chemical families within the same site- reported mutation affecting binding of
of-action group varies, depending on how the dinitroaniline herbicides to plant microtubules was
specific herbicides interact with the target protein. found in an -tubulin gene transcript from
TSR can also be due to increased expression of the goosegrass (Eleusine indica) that encoded a
target-site gene, producing more enzyme than can Thr239Ile substitution (6). This substitution
be substantially inhibited by typical herbicide conferred resistance to trifluralin and oryzalin
application rates. Increased gene expression can be when transformed in maize, demonstrating that
due to regulatory changes increasing transcription this mutation was the molecular basis of
and/or increased genomic copy number of the dinitroaniline herbicide resistance (6).
target-site gene, also resulting in increased Substitutions of Val202Thr (7) and Arg243 to Met
transcription. or Lys (8) have been reported for dinitroaniline

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


resistance in annual ryegrass (Lolium rigidum). A
Single nucleotide polymorphism substitution of Met268Thr has also been identified
Non-synonymous SNPs imparting resistance in E. indica, and both Thr239Ile and Leu136Phe
to a herbicide target site is the most common have been identified in green foxtail (Setaria
mechanism of TSR. These mutations may be viridis) (9), demonstrating that substitutions at
within or in the proximity of the catalytic domain multiple amino acid positions in a target site gene
of an enzyme and affect a herbicide’s ability to can confer resistance, depending on the molecular
compete for the binding of a substrate, or these structure of the target site protein and where the
mutations may affect other domains of enzymes herbicide binds to the protein.
and proteins. Acetyl CoA carboxylase (ACCase; EC
6.4.1.2) is a key enzyme for fatty acid biosynthesis
Mutations affecting herbicide binding pathways. The plastidic form of ACCase in
The first discovered target-site mutation was grasses is inhibited by the
for the photosystem II (PSII)-inhibiting aryloxyphenoxypropionate (APP),
herbicides, which compete with plastoquinone for cyclohexanedione (CHD), and phenylpyrazoline
binding on the D1 protein encoded by the psbA (PPZ - pinoxaden is the only member) herbicide
gene and thereby inhibit PSII electron transport chemical families (1). Eight mutations have been
(4). Amino acid substitutions in the psbA gene reported, all contained within the carboxyl
typically confer high-level resistance to a single transferase domain of the ACCase enzyme.
chemical family, but not to herbicides from other Known mutations occur at seven positions:
families or groups. For example, the single amino Ile1781Leu or Val, Trp1999Cys or Leu,
acid change Ser264Gly confers high level Trp2027Cys, Ile2041Asn or Val, Asp2078Gly,
resistance to the triazine herbicides in numerous Cys2088Arg, and Gly2096Ala or Ser (10).
species around the world (e.g., (5)) but moderate Substitutions at positions Ile1781 and Asp2078
to no resistance to the other families of PSII have been reported most often, and these
inhibitors, including the triazinones, which are in substitutions confer resistance to APP, CHD, and
the same group as the triazines. The substitution of PPZ herbicides. Ile1781 is within the binding site
Gly for Ser at position 264 prevents triazine for the three ACCase herbicide chemical families,
binding, but also compromises plastoquinone explaining the resistance pattern to all three
binding and impairs photosynthesis, resulting in a classes. Asp2078 is not within the binding site but
strong fitness penalty (4). A Ser264Thr occurs next to Ile1781, so the substitution of Gly
substitution confers resistance to triazines and for Asp at 2078 can cause a large effect on
ureas, but not to the nitrile or triazinone families. resistance level due to the substantial change in the

3
structure of the binding site. Mutations at positions key enzyme in the synthesis of branched chain
Ile2041 and Gly2096 confer resistance only to amino acids. Several chemical families have ALS
APP herbicides, while mutations at Trp2027 can as their site of action, including the sulfonylureas
confer resistance to APP and PPZ herbicides (11). (SUs) and imidazolinones (IMIs). Twenty-one
combinations of weed species by ALS inhibitor
Somatic mutations resistance-endowing amino acid substitutions
Target site mutations have also been identified have been reported to date, with 127 total unique
in the invasive aquatic weed Hydrilla verticillata (by species) occurrences of the different
conferring resistance to fluridone, an inhibitor of substitutions (3). Resistance-imparting
phytoene desaturase (PDS; EC 1.3.99.31), an substitutions at Pro197 have been reported most
ezyme essential for carotenoid synthesis (12). H. frequently, followed by mutations at Trp574. As
verticillata is a dioecious plant (male and female with ACCase target site mutations, some of the
flowers are on separate plants) and only the female ALS mutations confer very high-level resistance,
form was introduced to the United States. and the resistance spectrum across chemical
Consequently, resistance has evolved through the families varies by mutation. General patterns are
selection of a mutation in meristematic tissue that that the Trp574 mutation confers resistance to SUs
was able to regenerate into whole plants and and IMIs, the Ser653 mutation confers resistance
spread to other lakes (13). The specific mutations to IMIs but not SUs, and the Pro197 mutation

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


in PDS are Arg304 to Ser, Cys, or His. These confers resistance to SUs but not IMIs (depending
mutations have arisen through somatic variation, on the specific amino acid substitution; some
with the consequence that H. verticillata Pro197 mutations do confer resistance to IMIs).
populations within a confined body of water The known ALS resistance mutations do not occur
normally contain a single resistance mutation. The at the substrate binding site, but instead occur at
three mutations confer cross-resistance to the amino acid positions where the ALS herbicides
PDS-inhibiting herbicide norflurazon, but also can bind to and block an access channel within the
confer negative cross-resistance (i.e., increased enzyme through which the ALS substrates must
sensitivity) to three different PDS-inhibitors move (16). This is the biochemical reason why
(beflubutamid, picolinafen, and diflufenican) (14). ALS mutations generally have little effect on the
This finding further emphasizes that the effects of normal catalytic activity, in contrast to mutations
amino acid substitutions are dependent on the in the psbA gene that significantly affect normal
target site protein structure and vary across biochemical function. The SUs and IMIs bind to
different herbicide site of action groups. partially overlapping sites in the ALS enzyme, but
An interesting aspect of somatic mutations have different modes of binding (Figure 1). This is
imparting resistance to herbicides is that the lack important from an evolutionary perspective,
of sexual reproduction may make the resistance because rotating different ALS chemical families
trait more stable in a population than in species in the field will likely select for the same mutation,
dependent on sexual reproduction due to a lack of while rotating among different PSII-inhibiting
genetic recombination. Consequently, the three chemical families may select for different
distinct populations with variable resistance to mutations. In the case of ALS-inhibiting
fluridone have remained in the same abundance in herbicides, rotations to other ALS-inhibiting
some of these lakes despite an 8-year period with chemical families provide the same selection
no fluridone selection pressure (15). While the pressure from an evolutionary perspective and are
aquatic environment in which this case evolved not functional in slowing the evolution of
may be partly responsible for this to occur, some resistance. In support of this concept, resistance to
terrestrial weeds also can reproduce vegetatively, ALS-inhibiting herbicides has occurred rapidly
suggesting the possibility of somatic mutations following introduction of these herbicides.
creating herbicide resistance in these species. Rotations among different PSII-inhibiting
chemical families may in some cases slow the
Mutations affecting access to target site evolution of TSR, because the amino acid
Acetolactate synthase (ALS; EC 2.2.1.6) (also substitution conferring resistance to one family
called acetohydroxyacid synthase or AHAS) is a does not confer resistance to another family.

4
The rarer a mutation is within a population inhibitor required to decrease reaction rate to half
prior to herbicide selection, the longer it will take of the uninhibited value) for glyphosate. This
for the mutation to be selected and reach a high increase in Ki for EPSPS with the mutation is the
frequency within the population. A study of a reason for resistance, as more glyphosate is
herbicide-susceptible L. rigidum population found required to inhibit an equivalent amount of
that SU target site resistance allele frequency enzyme; however, the structural change also
within previously untreated populations was as increases the Michaelis constant for PEP (K m, the
high as 1.2 × 10-4, and IMI target site resistance substrate concentration required for effective
allele frequency was as high as 5.8 × 10-5 (17). catalysis to occur).
Similarly, a mutagenesis experiment in An EPSPS mutant with a higher Km requires
Arabidopsis found ALS resistance (SU and IMI) higher PEP concentration to achieve the same
at a frequency of 3.2 × 10-5 in progeny of M1 lines reaction velocity as the wild-type EPSPS with
(first generation after chemical mutagenesis) and normal Km. This means that the selectivity factor
with no detectable glyphosate resistance in for PEP binding over glyphosate binding is
250,000 M1 progeny screened (18). These affected, due to changed affinity for PEP, and a
experimental results are corroborated by the higher PEP concentration is necessary under
relatively fast initial evolution of resistance to normal conditions to maintain the same reaction
ALS-inhibitors following their introduction, and rate. This effectively reduces the catalytic activity

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


the relatively slower initial evolution of resistance of EPSPS, a possible reason why this target site
to glyphosate following its introduction (3). mutation may be rarer within populations
The relatively higher number of mutations compared to ALS target site mutations (which do
imparting resistance to various classes of ALS- not generally affect the Km or catalytic activity of
inhibiting herbicides in comparison to other ALS). A mutation at Thr102Ser conferred
herbicide groups is due to the fact that these glyphosate resistance in tridax daisy (Tridax
molecules do not compete with the substrate for procumbens) (21). This mutation imparted lower
the catalytic domain of the enzyme. Instead, they affinity to glyphosate, but also higher affinity for
block the opening of the channel leading to the PEP.
catalytic domain (Figure 1). Consequently,
mutations imparting resistance often have no Multiple nucleotide polymorphisms
impact on the kinetic properties of ALS. A concomitant mutation, generated by point
Glyphosate inhibits 5-enolpyruvylshikimate- mutation of the maize EPSPS and commercialized
3-phosphate synthase (EPSPS; EC 2.5.1.19), a key as GA21 glyphosate-resistant maize (22), of both
enzyme in the shikimate pathway, and mutations Thr102Ile and Pro106Ser (TIPS) resulted in
in the EPSPS gene have been reported in weeds. structural changes that retained high affinity for
Most described EPSPS target site mutations in PEP and made the enzyme insensitive to
weeds are located at the Pro106 residue, using the glyphosate inhibition (23) (Figure 2). The first
Arabidopsis numbering system, from the start of known naturally-occurring case of this TIPS
the mature enzyme (reviewed by 19). Known double mutation evolved in E. indica (24), in
resistance mutations in weeds include Pro106 to which the Pro106Ser had previously evolved. This
Ser, Thr, Ala, or Leu (19). Glyphosate inhibits double mutation confers a much higher level of
EPSPS by competing with the normal substrate glyphosate resistance, and maintains the affinity
phosphoenolpyruvate (PEP) for binding to the for PEP at a similar level as the wild-type enzyme
enzyme (Figure 2). Glyphosate binding is almost (25), although the TIPS mutation was found to
irreversible, so once bound, that unit of EPSPS is have a high fitness cost in E. indica (26). The
blocked. The Pro106 residue is not directly double TIPS mutation has also been reported in
involved in a molecular interaction with hairy beggarticks (Bidens pilosa) from Mexico
glyphosate or PEP, but it provides part of the (27). A double Thr102Ile and Pro106Thr (TIPT)
molecular structure at the active site (20), and mutation was found in greater beggarticks (Bidens
changing the Pro106 to a different residue changes subalternans) from Brazil, a tetraploid species in
the spacing in the active site. This increases the which the TIPT mutation was found in only one of
inhibitory constant (Ki, the concentration of the two genomes (28). Recently, a triple amino

5
acid substitution in EPSPS was found in smooth kochia (Bassia scoparia), changing a GGT (Gly)
pigweed (Amaranthus hybridus) in Argentina, at amino acid position 127 to an AAT (Asn) (33).
with a Thr102Ile, Ala103Val, and Pro106Ser This double mutation is located in the conserved
(TAP-IVS) allele conferring high resistance to degron region II of the Aux/IAA protein and
glyphosate (29,30). This stepwise evolution of confers resistance to dicamba.
mutations is an excellent example of how From an evolutionary perspective, rotating
herbicide resistance is rapid evolution, as many different synthetic auxin herbicide chemical
gene families over evolutionary time have evolved families may be an effective resistance
by this same process involving the incremental management practice, as TSR mechanisms may be
accumulation of mutations that alter and improve highly specific to certain chemical families and
enzymatic activity and efficiency. When a may not necessarily confer resistance across
herbicide with a single site of action is used different families.
repeatedly, evolutionary processes leading to
higher resistance levels and more efficient Codon deletion affecting topology of target site
resistance mechanisms are expected to occur. A codon deletion is the removal of three
nucleotides from the coding sequence of a gene.
Receptor/Co-receptor interactions The coding frame is unaffected by a codon
Synthetic auxin herbicides mimic the deletion, and a single amino acid is removed from

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


endogenous auxin hormone indole-3-acetic acid the protein encoded by the allele carrying the
(IAA) and deregulate growth and development codon deletion. Codon deletions conferring
processes. Synthetic auxins bind to a receptor herbicide resistance are rare relative to single
protein (auxin F-box, or AFB) and to a co-receptor nucleotide substitutions, with the only known
protein (Aux/IAA) to deregulate gene expression example to date applying to herbicides that inhibit
controlling plant growth, as well as binding to protoporphyrinogen oxidase (PPO; EC 1.3.3.4), a
other auxin-binding proteins. In the case of key enzyme in chlorophyll and heme biosynthesis.
synthetic auxin herbicides, mutations that reduce Common waterhemp (Amaranthus
herbicide binding to auxin-binding proteins, AFB tuberculatus) resistant to PPO inhibitors
proteins, or Aux/IAA proteins (the sites of action (including lactofen, fomesafen, and others) had a
for synthetic auxins) could have roles in conferring three-nucleotide deletion in its PPX2 gene, which
resistance in weeds. For example, assays of auxin- encodes a PPO enzyme (34). It is thought that the
binding protein preparations isolated from auxinic three-nucleotide deletion was fostered by its
herbicide-susceptible and resistant wild mustard occurrence within a region containing bi-repeats
(Sinapsis arvensis) found similar binding for the of three nucleotides. These short simple repeat
normal substrate indole-3-acetic acid (IAA), but (SSR) regions are typically associated with
differences in binding were found for several insertion and deletion events. Most plant species
auxinic herbicides, and these differences do not have a SSR region in the homologous
correlated with the whole-plant resistance location of PPX2, suggesting they are not
phenotypes (31). predisposed for the same mutation (35). Palmer
Multiple auxin-binding proteins interact with amaranth (Amaranthus palmeri), however, is one
native auxins and synthetic auxin herbicides, and of the few weeds with a homologous SSR region
evidence indicates that some mutations confer (36) and, in fact, resistance to PPO inhibitors due
resistance specifically to certain chemical families to the same codon deletion was subsequently
of synthetic auxins. For example, mutations in the documented in this species (37). The codon
Arabidopsis TIR1 homolog, an AFB protein, deletion, which specifically results in removal of a
confer resistance to the picolinate class of glycine at position 210 was functionally validated
synthetic auxins such as picloram but not to 2,4-D to confer resistance using a transgenic E. coli
(32). system (34) and also has been subjected to
Aux/IAA transcriptional repressors are co- biochemical and structural analysis (38). Deletion
receptors for synthetic auxin herbicides and part of of glycine 210 is proposed to partially unravel an
the target site complex. A double-nucleotide TSR helix adjacent to the PPO active site, raising the
substitution was discovered in the IAA16 gene of Ki for the herbicide and enlarging the active site

6
cavity (Figure 3). This structural change confers sethoxydim and quizalofop had 2- to 3-fold higher
broad cross resistance to PPO-inhibiting ACCase enzyme activity relative to an ACCase-
herbicides at both the enzyme and whole-plant susceptible population (43). The I50 (herbicide
level, although resistance magnitudes vary among concentration required to inhibit 50% of enzyme
herbicides by up to 10-fold (39). Efforts are activity in vitro) was similar between resistant and
underway to develop new PPO-inhibiting susceptible populations, but the higher activity
herbicides that overcome the Gly210 PPO deletion was maintained across a range of herbicide
(40). To date, a codon deletion has not been concentrations in vitro. However, the study did not
reported to confer herbicide resistance in any weed determine whether the increased ACCase
species other than A. tuberculatus and A. palmeri. enzymatic activity was due to ACCase gene
duplication or up-regulation of ACCase
Increased expression of target site genes transcription.

Up-regulation Target site gene duplication


Baerson et al. (41) found elevated EPSPS Gene duplication, the heritable replication of a
expression (2.5 to 3 times higher) in glyphosate- coding segment of DNA resulting in one or more
resistant L. rigidum. The EPSPS protein extracted additional gene copies within the genome of an
from resistant and susceptible individuals was organism (44), is a common process in the

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


equally sensitive to glyphosate inhibition, but evolutionary history of plants and is vital for
basal enzyme activity was higher in the resistant generating genomic diversity (45). Increased gene
population. The authors concluded that higher expression at the mRNA level is the immediate
EPSPS mRNA expression was resulting in higher result of gene duplication, and as mutations
EPSPS production, but they were uncertain accumulate in duplicated gene copies over time,
whether the magnitude of over-expression duplicated gene copies can begin to have
accounted for the observed resistance level. No variations in function or acquire new functions
evidence was found to indicate EPSPS gene (44). The term gene amplification is also
duplication in the glyphosate resistant population frequently used synonymously with gene
using DNA blot hybridization, despite the duplication. Gene amplification has been defined
observed 2.5 to 3 times higher EPSPS expression in some literature as the non-heritable replication
(41). In glyphosate-resistant populations of of a segment of DNA, such as in cases of cancer
horseweed (Conyza canadensis) and hairy tumors where amplified gene copies are not
fleabane (C. bonariensis), basal EPSPS mRNA inherited in the progeny of the individual, and gene
expression was 2-fold higher than in glyphosate- duplication indicates the heritable replication of a
susceptible populations when measured using segment of DNA (44).
northern blots. The initial examples of EPSPS gene
Other examples of target site gene up- duplication related to glyphosate resistance were
regulation have been reported. In three ALS- obtained in cell culture studies. The involvement
resistant barleygrass (Hordeum leporinum) of gene duplication in glyphosate-resistant A.
populations from Western Australia, ALS enzyme palmeri (46) was the first case identified in a weed
activity was 3-fold higher than in an ALS- population (Figure 4). The EPSPS gene in a
susceptible population (42). In addition, the three glyphosate-resistant A. palmeri population was
populations also had a mutation at Pro-197 duplicated from 4- to over 100-fold relative to a
resulting in a serine substitution. While this susceptible population. Expression of EPSPS
mutation confers ALS resistance on its own, the mRNA and EPSPS protein corresponded with the
increased ALS expression may also contribute to increased genomic EPSPS copy number. The
the observed resistance level. It is not known if the increased EPSPS expression means that more
up-regulation of ALS protein activity is due to EPSPS is present at the target-site than the typical
altered gene expression regulation, gene applied concentration of glyphosate can inhibit
duplication, and/or reduced enzyme turnover rates. (19). Enough uninhibited EPSPS remains
A johnsongrass (Sorghum halepense) population following typical glyphosate applications such that
resistant to the ACCase-inhibiting herbicides the plant can survive. In the studied A. palmeri

7
population, duplicate copies of the EPSPS gene available evidence suggests that higher EPSPS
appeared to be present on all chromosomes (2n = copy number confers higher glyphosate resistance
34). in A. palmeri (e.g., (50), L. multiflorum (59), and
In a fascinating example of molecular genetic B. scoparia (60). Continued observation and
variation leading to adaptation, the duplicated monitoring of EPSPS copy number in populations
EPSPS gene was contained within a >300 kb over time will be needed to determine whether
replicon containing multiple additional open EPSPS copy number may increase following
reading frames for other genes and various types continued glyphosate selection pressure.
of repetitive DNA elements (47). The replicon was Recently, a population of glyphosate-resistant
found to have a circular structure existing outside spiny amaranth (Amaranthus spinosus) was
the chromosome, termed an extra-chromosomal reported in which EPSPS gene duplication and
circular DNA (eccDNA) (48). The eccDNA can sequence data revealed that the EPSPS gene in
attach to the chromosomes to be transmitted both glyphosate-resistant A. spinosus individuals is
at mitosis and at meiosis, explaining the observed identical to glyphosate-resistant A. palmeri EPSPS
variation in heritability of EPSPS gene copy (61) and not glyphosate-susceptible A. spinosus.
number in A. palmeri (49,50). The eccDNA This result indicates that the EPSPS gene has
sequence across glyphosate-resistant A. palmeri transferred through inter-specific cross-
populations from across the U.S. was nearly pollination. The two species are most closely

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


identical, suggesting the possibility of a single related to each other among Amaranthus species
origin of the eccDNA for glyphosate resistance (62), and the two species can hybridize and
followed by seed- and pollen-mediated gene flow produce fertile hybrids (63). This is important
(47,51). EPSPS gene duplication has been from an evolutionary perspective, because a
reported in many glyphosate-resistant weed genetic trait that may be extremely rare within
species. Glyphosate resistance via EPSPS gene populations (such as the eccDNA containing
duplication can be viewed as an example of EPSPS) has a selective advantage in multiple
convergent evolution across these diverse plant species that are under similar selection
species (52). Populations of B. scoparia and A. environments (repeated exposure to glyphosate).
tuberculatus have fewer duplicated EPSPS Presumably genetic transfer between A. palmeri
copies than A. palmeri, on the range of 4 to 10-fold and A. spinosus occurs at a low level in wild
(Table 2). populations, but the selective advantage of the
Despite having a lower quantity of duplicated glyphosate resistance trait that initially evolved in
EPSPS copies, A. tuberculatus populations have A. palmeri enabled transfer of this trait, plus
a similarly high level of resistance as the A. additional, unknown linked genetic traits, to A.
palmeri populations (Table 2). Both EPSPS spinosus. This genetic transfer may have
mRNA expression and EPSPS protein levels have evolutionary implications beyond loss of
been reported to have a linear correlation with glyphosate effectiveness in A. spinosus.
EPSPS genomic copy number in B. scoparia and While EPSPS gene duplication in A. palmeri
A. tuberculatus populations (57,58), as well as in has occurred via eccDNA, in B. scoparia a tandem
Italian ryegrass (Lolium multiflorum populations) gene duplication has occurred (60). Several repeat
(59). In the case of L. multiflorum, EPSPS protein units were identified in a bacterial artificial
expression level correlated very well with chromosome (BAC) assembly of the duplicated
population level resistance; as EPSPS protein locus, including a 56.1 kb repeat containing seven
expression increased, so did the dose required to predicted genes, and a 32.7 kb repeat containing
achieve 50% reduction in plant growth (59). four predicted genes (64). Some of these co-
Comparisons of resistance level across species are duplicated genes showed similar increased
problematic, as the various reports were conducted expression as EPSPS. The border of the duplicated
under different experimental conditions and region contained a mobile genetic element that
estimated LD50 and GR50 (concentrations causing may have provided the initial DNA break to
50% mortality and growth reduction, respectively) initiate the tandem duplication process. The
parameters are not directly comparable. However, potential evolutionary consequences of
duplicating and over-expressing other genes in

8
addition to EPSPS needs further exploration, as multiflorum with a three-fold difference in
does the relative stability of duplicated EPSPS glyphosate susceptibility and reduced absorption
inheritance for both tandem duplication and in the less sensitive biotype (75). When reduced
eccDNA mechanisms (65). absorption is implicated, it is most often only one
In addition to EPSPS gene duplication, the contributing factor to the overall resistance
target-site gene ACCase was found to have 5-7 mechanism. For example, resistance to glyphosate
fold higher gene copy number in a large crabgrass in A. tuberculatus biotypes was due to both
(Digitaria sanguinalis) population resistant to five reduced absorption and a herbicide-resistance
ACCase inhibitor herbicides, resulting in 3- to 9- allele of the glyphosate enzyme target EPSPS (76).
fold higher ACCase transcript abundance (66).
Reduced translocation and vacuolar
Non-target site mechanisms sequestration
Many foliar-applied systemic herbicides rely
Reduced absorption on translocation through the phloem for optimal
To be effective, herbicides must be absorbed activity. These herbicides must cross the cuticle
into cells of plants through the roots, in the case of barrier and enter the cells of mature source leaves
soil-applied herbicides, or from the leaves in the (symplast). This transport can involve active (i.e.,
case of foliar-applied herbicides (Figure 5). protein-mediated) and/or passive diffusion

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


Menendez et al. (67) provide an excellent processes. Once inside the symplast, systemic
description of the factors involved in foliar herbicides translocate from source leaves to
absorption and root absorption of herbicides. younger sink leaves via the phloem, often along
Differences in root absorption of herbicides with the movement of photosynthetic sugars.
between species have been attributed to root Herbicide resistance due to reduced translocation
morphology differences (68). There are no cases occurs when the herbicide is retained in source
of evolved resistance to soil-applied herbicides leaves and prevented from translocating to the
due to reduced root absorption. Early work on growing points (Figure 5). Mechanisms that trap
differential foliar absorption of herbicides between the herbicide in source leaves (e.g., through
species was attributed mainly to differences in sequestration within vacuoles or leaf trichomes) or
cuticle thickness and/or composition (68), but the prevent its normal movement to the growing
number and/or structures of leaf trichomes and points across membrane barriers (through altered
hairs have also been implicated (69). Hirsute activity of active membrane transporters) will
leaves are covered with hairy trichomes that can reduce the total amount of herbicide translocated,
retain spray droplets better than smooth, hairless thus conferring resistance. Reduced absorption
or glandless cuticles, thereby facilitating across the cuticle and reduced translocation out of
absorption. Other leaves have lysigenous glands source leaves sometimes work in concert.
involved in the production and storage of oily Reduced translocation of glyphosate is the
secondary metabolites that can compartmentalize most prominent example of this NTSR mechanism
lipophilic herbicides, preventing them from (65). In these plants, the amount of glyphosate
reaching their site of action (70). Differences in delivered to the meristems is lower than what is
foliar absorption of herbicides between plants necessary to be phytotoxic. Reduced glyphosate
have been attributed to leaf anatomical features translocation was first demonstrated in
rather than any biochemical differences. glyphosate-resistant L. rigidum from Australia,
Decreased absorption is not a common NTSR where glyphosate moved to the edges of treated
mechanism, but has been reported with resistance leaves, and less glyphosate translocated to the
of common sunflower (Helianthus annuus) to meristems, relative to glyphosate-susceptible L.
imazethapyr and chlorimuron (71), prickly lettuce rigidum (77). Glyphosate-resistant C. canadensis
(Lactuca serriola) to 2,4-D, annual bluegrass (Poa had reduced translocation as well (78). This is due
annua) to atrazine (72), and L. multiflorum and S. to differences in cellular distribution of glyphosate
halepense to glyphosate (73,74). No differences and subsequent phloem loading and translocation.
were found in cuticular wax amount per unit area In these biotypes, glyphosate enters the
of leaf surface between two biotypes of L. symplasm of source leaves normally but cannot

9
translocate to the meristems because it is rapidly and the finding of Ge et al. (79) demonstrating
sequestered within the vacuole (79) (Figure 4). vacuole sequestration of glyphosate, intact
The vacuole sequestration process is temperature- protoplasts of paraquat-resistant and -susceptible
dependent, with less sequestration occurring in C. L. rigidum were isolated. The paraquat
canadensis under colder temperatures (80). The concentration was measured, as a method to
dependence on temperature suggests the determine if the paraquat concentration within
involvement of active membrane transporters. single cells was higher in paraquat-resistant
Considerable research effort with both microarray protoplasts than paraquat-susceptible protoplasts
and transcriptomic sequencing (81,82) has (85). Paraquat-resistant L. rigidum protoplasts
searched for a specific ABC transporter gene contained more paraquat than susceptible
suspected to be responsible for the vacuolar protoplasts. Thus, the paraquat is likely
sequestration, but no causative specific gene has sequestered in the vacuole, in a similar process as
yet been confirmed. reported for vacuolar sequestration of glyphosate.
The effect of lower temperatures on the Reduced paraquat translocation attributed to
reduced translocation mechanism has also been vacuole sequestration has also been reported in L.
reported in S. halepense and L. multiflorum (83). multiflorum from California (86) and in two
The temperature effect supports the hypothesis Conyza spp. from California (87).
that vacuole sequestration is an active process and Characterization of a paraquat-resistant

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


restricts glyphosate translocation by preventing Arabidopsis mutant revealed a mutation in a gene
normal glyphosate movement into the phloem. called Paraquat Resistant 1 (PAR1), a putative
Reduced translocation may result from other amino acid transporter (88). The mutant, par1, had
NTSR mechanisms in A. palmeri, A. tuberculatus, similar cellular paraquat absorption as the normal
and S. halepense (19). In these reports, reduced PAR1 genotype, but had reduced paraquat
cellular glyphosate absorption across the plasma concentration in the chloroplast. Overexpressing
membrane may be associated with altered rate of the PAR1 homologue in rice resulted in paraquat
an active glyphosate transport process (Figure 4). hypersensitivity, and silencing the rice PAR1
Reduced cellular uptake is not predicted to alter resulted in paraquat resistance, revealing that
translocation at the whole plant level, and reduction in paraquat transport into the chloroplast
generally no changes in glyphosate translocation can confer resistance. Natural polymorphic
have been observed in A. palmeri. Finally, reduced variations in a gene (LHR1) for a plasma
chloroplast absorption of glyphosate has been membrane-localized polyamine transporter in
suggested (19) (Figure 4). This phenotype is Arabidopsis caused variations in uptake of
difficult to measure, as isolating intact chloroplasts paraquat that correlated with sensitivity to
is technically difficult and may have unknown paraquat (89). Endogenous polyamines have been
effects on the physiological ability of the isolated shown to play a role in paraquat-resistant E. indica
chloroplasts to exhibit any reduced absorption. To (90). Evolved paraquat resistance in weeds has not
date, reduced chloroplast glyphosate absorption as yet been attributed to selection for orthologs of
a glyphosate resistance mechanism has not been PAR1 or LHR1.
experimentally demonstrated. Reduced translocation of 2,4-D was observed
Reduced translocation of paraquat from in a 2,4-D-resistant population of L. serriola,
treated source leaves to sink leaves has also been relative to a susceptible population (91), as well as
reported for paraquat-resistant L. rigidum (84). In in 2,4-D resistant wild radish (Raphanus
this population, there was no difference in the raphanistrum) (92). Reduced dicamba
interaction of paraquat with the target site translocation was found in a dicamba resistant B.
Photosystem I complex (indicating a lack of TSR), scoparia population (93). Naturally-occurring
and no differences were found in the antioxidant auxins such as IAA are polar and readily
systems of superoxide dismutase or ascorbate translocate in the phloem, so it is reasonable that a
peroxidase (increased antioxidant activity may synthetic auxin such as 2,4-D or dicamba also
allow a plant to tolerate the free radicals generated requires adequate translocation from the
by diversion of electrons from Photosystem I). application site to the target site in growing
Because the population had reduced translocation meristems. Changes in auxin transport from cell to

10
cell via active transporters could play a role in Sumatran fleabane (Conyza sumatrensis) biotype
reduced 2,4-D translocation. Reduced from southern Brazil (96). The symptoms after
translocation of 2,4-D or dicamba would 2,4-D application are similar to those observed in
presumably reduce the total concentration glyphosate-resistant A. trifida (i.e., necrosis in
achieved at the target site to low enough levels to older leaves and absent in younger leaves,
enable survival. followed by regrowth from meristems). The 2,4-
An unusual way for a plant to achieve reduced D-resistant biotype was also resistant to
translocation of a herbicide to the meristems has glyphosate, but not by the phoenix phenomenon.
been described as the phoenix phenomenon, which Rapid necrosis symptoms did not occur with six
is the result of reduced translocation caused by other auxinic herbicides (e.g., picloram). ROS
more rapid action of the herbicide (Figure 6). One accumulation was much higher within 30 min after
of the assets of glyphosate as a herbicide is that it 2,4-D treatment than in a susceptible accession
acts slowly, allowing it to translocate to and kill and remained much higher for more than 7 h. The
meristematic tissues. With herbicides that act resistant biotype did not show the normal epinasty
rapidly, translocation from treated plant organs is symptom (leaves bending downward) caused by
limited because of the rapid action of the 2,4-D, probably due to rapid cell death caused by
herbicide. Giant ragweed (Ambrosia trifida) ROS inhibiting herbicide translocation.
evolved a rapid response to glyphosate that

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


prevents translocation of the herbicide to Metabolic alterations
meristems (94,95). Light- and/or sucrose- Plants contain large numbers of genes
dependent rapid withering and desiccation of encoding enzymes that perform biochemical
treated foliage occurs, followed by regrowth of the reactions for the synthesis of secondary
plant from meristems that were not contacted by metabolites and for detoxifying xenobiotic
the foliar spray – hence the “phoenix compounds (e.g., herbicides) (97).
phenomenon”. None of the other known Serendipitously, some members of these gene
mechanisms of glyphosate resistance were found families can also detoxify herbicides. The
(95). Reactive oxygen species (ROS) accumulated selective action of many herbicides (i.e., they
within 30 min of treatment, only in the older leaves control weeds without damaging crops) often
of the resistant biotype of the weed. Treatment of depends on relatively rapid metabolism of the
the leaf tissue with exogenous phenylalanine and active ingredients into harmless breakdown
tyrosine (two aromatic amino acids that are products in crops compared to weeds. This
products of the shikimate pathway) prevents the biochemical feature (differential rates of
rapid effect, indicating that the effect may be detoxification) has been repeatedly exploited for
associated with EPSPS inhibition or to selective chemical weed control. Herbicide
physiological blocking of the rapid response by the detoxification is generally divided into three
amino acids. The mechanism for accelerated phases (Figure 5). Phase I involves addition of a
action of glyphosate in this biotype is unknown, functional group to the herbicide by oxidation,
but it might be explained by rapid cessation of reduction, or hydrolysis, often mediated by
carbon fixation due a rapid deregulation of the cytochrome P450 monooxygenases (P450; EC
shikimate pathway removing enough erythrose-4- 1.6.2.4). Phase II involves more complex changes
phosphate and phosphoenolpyruvate from the C3 to a herbicide, such as conjugation, either to
carbon fixation pathway to stop it. In another glutathione mediated by glutathione-S-
species with a rapid response to glyphosate, sugar transferases (GST; EC 2.5.1.18), or to glucose
beet (Beta vulgaris), cessation of carbon fixation mediated by glucosyltransferases (GT; EC 2.4).
is rapid, and the symptoms are similar to those of Note that phase II enzymes such as GSTs and GTs
the resistant A. trifida. Alternatively, perception of can directly detoxify some herbicides without
glyphosate in mature, green plant cells of this depending on phase I activation. The final step in
resistant biotype may trigger a rapid cell death plants (phase III) involves compartmentalization
defense mechanism. of the herbicide metabolites in the vacuole or
Rapid necrosis followed by regrowth was also incorporation into cell walls (Figure 7). In general,
identified as a resistance mechanism to 2,4-D in a the same genes and biochemical mechanisms for

11
herbicide detoxification often exist in weeds that herbicide groups, or if multiple distinct metabolic
are related to crops, with the critical difference mechanisms are present. The key concept is that
being that expression of these genes is lower in the some enhanced metabolism mechanisms,
weeds. Thus, there is evolutionary potential to including P450 and GST, can confer broad-
select for increased expression and/or mutations of spectrum resistance, with known examples of
these key genes in weeds, enabling enhanced cross-resistance due to a single mechanism and
herbicide metabolism to confer resistance. multiple resistance due to accumulation of
Prominent examples of herbicides that are multiple, distinct mechanisms. From an
selective due to differential detoxification between evolutionary biology perspective, enhanced
crops and weeds include the selective ALS and metabolism can be considered a broad-spectrum,
ACCase inhibitors. Some of these, including generalist adaptive response. Critically, enhanced
chlorsulfuron, diclofop-methyl, and fenoxaprop- metabolism mechanisms can also combine with
P-ethyl, are used to control grass weeds in wheat. other mechanisms including TSR and reduced
For these herbicides, rapid metabolism and crop translocation to confer higher levels of resistance.
safety in wheat involves detoxification pathways
including P450, GT, and GST. For example, Cytochrome P450-mediated herbicide metabolism
metabolism of the ACCase inhibitor diclofop- Cytochrome P450 monooxygenases are
methyl to non-toxic metabolites in wheat occurs membrane-bound proteins localized in the

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


via P450-mediated aryl hydroxylation, followed endoplasmic reticulum and are one of the largest
by glucose conjugation, likely mediated by GT gene families in all organisms (101). These
(e.g.,(98)). These wheat-like metabolic pathways enzymes have crucial roles in the synthesis of
also exist in grass weeds that have considerable hormones, lipids, and metabolism of endogenous
intra-specific variation for herbicide and exogenous substances (Figure 8). The number
susceptibility. In susceptible weed populations the and diversity of P450s in plants is higher than in
detoxification rates are generally too slow to other organisms. While in humans there are
prevent herbicide phytotoxicity. However, some approximately 54 P450 genes (0.1% of genome),
weed species such as blackgrass (Alopecurus in plants the number is higher, with 246 in
myosuroides) in Europe, and L. rigidum in Arabidopsis (1% of genome) and 328 in rice (0.5-
Australia have evolved wheat-like rapid 1% of genome). Weeds have been found to have
detoxification pathways (e.g., 99,100). These high diversity in P450 genes, with 917, 323, and
weed populations have long histories of repeated 277 in barnyardgrass (Echinochloa crus-galli), C.
exposure to wheat-selective herbicides and have canadensis, and L. rigidum, respectively. The
evolved enhanced metabolism-based resistance. greater diversity of the P450 genes in plants has
Enhanced herbicide metabolism is especially likely evolved for chemical defense, mainly for
problematic from a weed management standpoint degradation of many xenobiotics and by chance
because the detoxification systems that confer these P450 genes also detoxify herbicides (102).
resistance to one herbicide can sometimes have These genes have gained importance in agriculture
activity on other herbicides with the same or as the basis for herbicide selectivity in crops due
unrelated sites of action. The term cross-resistance to natural metabolic processes. For example,
is defined as when a single mechanism (such as different P450 monooxygenases metabolize ALS
enhanced metabolism) confers resistance to more inhibitors in rice (103), cinmethylin in wheat
than one herbicide with the same or of a different (104), 4-hydroxyphenylpyruvate dioxygenase
site of action group. The term multiple resistance (HPPD; EC 1.13.11.27) inhibitors in maize (105),
is defined as when multiple, distinct mechanisms and other mode-of-action herbicides in different
have combined within an individual (or crops (106), at a faster rate than weeds. The
population), and the individual (or population) is implication is that genetic variation exists in plants
resistant to herbicides from more than one site of for the expression level of P450 genes and crops
action group. When considering enhanced have higher expression of P450 genes that
metabolism, it may often be difficult to distinguish metabolize herbicides. It should be noted that the
phenotypically whether a single metabolic herbicide discovery process often identifies
mechanism is conferring cross-resistance to other candidate herbicides that are metabolized faster in

12
some weeds than in crops; these herbicides would achieved in late watergrass (Echinochloa
not be effective and are not commercialized. phyllopogon) by first inducing P450 activity with
However, the wide use of herbicides for weed a sub-lethal herbicide dose, demonstrating P450
control has selected biotypes with the same ability activity in metabolizing bispyribac-sodium,
to inactivate herbicide molecules, threatening fenoxaprop-p-ethyl, and thiobencarb (114). E.
weed management worldwide. The most phyllopogon populations are also resistant to
important reactions catalyzed by these enzymes clomazone and penoxsulam via enhanced
are either aryl- or alkyl hydroxylation, the first step oxidative metabolism from P450 activity
in the metabolism of xenobiotics (107). In general, (115,116). Expression of several P450 genes was
P450s insert molecular oxygen on a herbicide measured using quantitative PCR in a metabolism-
molecule to be more reactive or more soluble based resistant E. phyllopogon population relative
using an electron from NADPH P450 reductase. to a susceptible population, and substantial
As a result, herbicide molecules are metabolized induction of the P450 genes occurred following
to products with reduced or modified sub-lethal herbicide treatment (117). Analysis of
phytotoxicity in weeds with metabolism-based twelve candidate P450 genes from resistant E.
resistance mechanisms. phyllopogon revealed that CYP81A12 and
Extensive work with P450 inhibitors provides CYP81A21 had the highest transcription levels in
strong evidence for the role of various P450 the resistant biotype (118). These P450 genes were

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


isozymes in herbicide metabolism. Application of transformed into Arabidopsis, and transgenic
a P450 inhibitor prior to herbicide exposure will plants exhibited resistance to bensulfuron-methyl
block P450 activity. When the resistance and penoxsulam (118). Recently, P450 genes were
mechanism is due to enhanced metabolism by cloned from E. phyllopogon and transformed into
P450s, inhibition of the P450 activity enables the both Arabidopsis and E. coli (119). The ability to
herbicide to reach the target site in a high enough assay the membrane-bound P450 proteins in
concentration to cause normal phytotoxicity. P450 bacteria is a major advance in available tools to
inhibitors have been used in various resistant functionally validate candidate resistance genes
weeds to test for P450 herbicide metabolism. In L. from weeds. The E. phyllopogon CYP81As
rigidum populations resistant to multiple metabolized 18 herbicides from 13 different
herbicides through metabolism-based chemical classes. The recombinant expression of
mechanisms, the P450 inhibitor malathion restores CYP81As in E. coli metabolized different
chlorsulfuron activity (108). The herbicide herbicides by demethylation or hydroxylation
amitrole, structurally similar to the P450 inhibitor reactions in unrelated herbicide groups belonging
1-aminobenzotriazole, restores diclofop-methyl to ALS, ACCase, PDS, PSII, PPO, HPPD, and 1-
activity. However, the P450 inhibitors have deoxy-D-xylulose-5-phosphate synthase inhibitors
specificity for P450 isoforms. Pre-treatment with (119), demonstrating the incredible potential of
piperonyl butoxide (PBO), but not malathion, P450 enzymes to metabolize herbicides and confer
increased control of resistant P. annua with cross-resistance.
fenoxaprop (109), indicating that PBO inhibited The generally broad substrate recognition of
specific P450 enzymes critical for fenoxaprop plant P450 enzymes poses a threat for weed
metabolism. Although P450 genes found in the management due to potentially unpredictable
past decade were mostly related to metabolism of cross-resistance patterns. While CYP81A12 and
ACCase, ALS, and PSII inhibitors, more recently CYP81A21 can metabolize ALS inhibitors through
P450 inhibitors have inhibited the P450 demethylation, the same genes are also involved in
metabolism of several classes of herbicides (i.e., cross resistance to ACCase inhibitors such as
HPPD, PPO, synthetic auxins, and carotenoid diclofop-methyl, tralkoxydim, and pinoxaden
synthesis inhibitors) (110-113). Critically, P450s through hydroxylation (120). The same
have potential to metabolize herbicides from CYP81A12 and CYP81A21 genes, as well as
different mode of action groups, creating CYP81A15 and CYP81A24, confer resistance to
unpredictable patterns of cross-resistance. clomazone (121). Other enzymes from this P450
Successful isolation of microsomes with family, such as CYP81A6, were found to
herbicide metabolism activity from weeds was metabolize bentazon, sulfonylureas, and also

13
quinclorac in rice. The P450 enzymes of family enzymes and which substrates (herbicides) they
CYP81A appear to be “super P450s” able to may degrade. Biochemical modeling predicting
metabolize different chemical classes within the tertiary structure of known P450s from weeds
different herbicide mechanisms of action in the would enable docking simulations with different
same weed species. However, other P450 enzymes herbicides and anticipate classes of herbicides
of families CYP72, CYP71, CYP70, and CYP96 more likely to succumb to metabolic degradation.
have also been found to metabolize different Transcriptional regulation of P450 genes is
herbicides. another area where more research is needed.
Fewer examples of enhanced P450 Although a trans-element was proposed to control
metabolism have been identified in eudicot species the expression of both CYP81A12 and CYP81A21
(typically broadleaf species). The reasons may be genes in the resistant E. phyllopogon (118), we
biological, such as lower P450 activity or fewer lack studies to functionally demonstrate
P450 genes in eudicots, or the reason may be due transcription factors, activator elements, repressor
to less research into enhanced herbicide elements, or epigenetic modifications that may be
metabolism for eudicots. Recently, reports of regulating P450 gene expression in metabolic
metabolism-based herbicide resistance in eudicots resistant weeds.
are increasing. More rapid initial metabolism of
chlorimuron was found in an ALS-resistant A. Glutathione-S-transferases (GST)

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


hybridus population (122). A survey of ALS- The enzyme super-family of GSTs are
resistant A. palmeri in Georgia found that more involved in herbicide detoxification by
than half of the populations had no ALS target site conjugating glutathione to the herbicide molecule,
mutations and exhibited enhanced ALS herbicide rendering the herbicide non-toxic. This
metabolism (123). Resistance to HPPD inhibitors conjugation reaction can occur directly to the
has been reported in A. tuberculatus (124,125), A. active herbicide or following the activity of other
palmeri (110,126) and recently in R. raphanistrum enzymes such as P450s. GST activity was first
(127). The resistant biotypes metabolized HPPD identified on triazine herbicides, including
inhibitors at a faster rate than the susceptible atrazine in corn (130). Additional herbicides
populations through hydroxylation reactions, known to be glutathione-conjugated by GSTs
indicating a P450 role in the resistance mechanism include chloroacetamides (such as alachlor and
in the broadleaf species (110,127,128). Recently, metolachlor), sulfonylureas (such as chlorimuron
A. tuberculatus biotypes from Nebraska with ethyl), diphenylethers (such as fluorodifen), and
resistance to tembotrione and 2,4-D, as well as A. aryloxyphenoxypropionates (such as fenoxaprop-
palmeri from Arkansas with resistance to ethyl) (131). E. phyllopogon populations from
fomesafen, were controlled when a P450 inhibitor California are resistant to fenoxaprop-ethyl via
was applied (111,113,129). As research efforts enhanced GST-mediated glutathione conjugation
into metabolism-based resistance expand, more (132), as are A. myosuroides populations from the
examples of broadleaf species with enhanced UK (133). Not all herbicides within a given
herbicide metabolism will likely be reported, and chemical group are equally susceptible to GST-
the candidate genes in eudicot species can be mediated glutathione conjugation. For example,
identified. while GST acts directly on the ACCase inhibitor
More work is needed to improve fenoxaprop ethyl (133), structurally similar
understanding of P450 evolution and regulation ACCase inhibitor herbicides, such as diclofop-
for herbicide resistance. The role of gene copy methyl, have no chemical features that can be
number variation for cytochrome P450 should be glutathione-conjugated and are instead able to be
studied, especially in polyploid weeds, as a ring-hydroxylated by P450s (98). The chemical
pathway to generate novel allelic variation and structure of the herbicide determines whether GST
increased expression. Chromosome duplication in is able to conjugate the herbicide with glutathione.
polyploid species may generate more potential Recently a GST (AmGSTF1), from multiple
variation for the evolution of metabolic resistance herbicide resistant A. myosuroides with enhanced
through P450s. Little is known about the structure- GST activity, was cloned and expressed in
activity relationship between the many P450s transgenic Arabidopsis (134). The transgenic lines

14
had increased resistance to chlorotoluron, alachlor, EC 3.5.1.13), an enzyme that hydrolyzes propanil
and atrazine, and also had increased antioxidant to 3,4-dinitroaniline, a non-phytotoxic compound.
flavonoid and anthocyanin content. In additional Studies of a wild-type rice and a rice mutant with
experiments, a GST inhibitor (4-chloro-7-nitro- no AA activity indicated that the normal role of
benzoxadiazole) synergized with the herbicides this enzyme in plants is in nitrogen metabolism
fenoxaprop and clodinafop in multiple-resistant A. relating to asparagine. Weedy rice (feral forms of
myosuroides populations and restored herbicide cultivated rice, often called red rice) is also
activity (134), providing additional evidence for tolerant to propanil by the same mechanism.
the role of increased GST expression as a Among Oryza species, tolerance to propanil
resistance mechanism. A. myosuroides correlates with AA activity (142). Resistance to
populations in Europe were found to be resistant propanil by elevated AA activity has evolved in
to flufenacet via enhanced GST-mediated junglerice (Echinochloa colona) (143) and E.
metabolism (135). Populations of Lolium spp. crus-galli (144). The AA inhibitors anilofos,
resistant to flufenacet were found to conjugate piperophos, and carbaryl synergize with propanil
glutathione to flufenacet for metabolic activity in propanil-resistant E. crus-galli (144).
detoxification, and overall GST activity was Not all evolved resistance to propanil is by
increased in protein extracts (136). enhanced AA activity, as resistance to propanil in
rice sedge (Cyperus difformis) is a single amino

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


Glycosyl transferases (GT) and other transferases acid change of the D1 protein of PSII (145).
The enzyme family of GT have roles in Phase
II of herbicide metabolism, via conjugation of Aldo-keto reductase
glucose to herbicide metabolites after initial Phase Some plant species, especially legumes,
I modification (typically hydroxylation or metabolize glyphosate to aminomethyl
demethylation) (137). For example, monocot phosphonic acid (AMPA) and glyoxylate, whereas
species (most often grasses) are tolerant to others (grasses in particular) apparently have little
synthetic auxins in part because glycosylation of capacity for this degradation pathway or any other
hydroxylated rings of auxinic herbicides tends to transformation of the herbicide (146). Neither the
be irreversible, highlighting the importance of GT enzyme nor the gene for the glyphosate
in protecting a plant from herbicide activity (69). oxidoreductase (GOX) that is considered
To date, no reports are known of herbicides for responsible for glyphosate degradation in plants
which the first step of metabolism involves GT; have been identified. AMPA is very weakly
however, increased GT expression may be phytotoxic, so sufficiently rapid degradation of
necessary as a secondary biochemical step for glyphosate to AMPA should provide resistance.
metabolic resistance in weeds (138,139). Because glyphosate is a very slow-acting
On the other hand, sensitive eudicots tend to herbicide, evolution of such a resistance
utilize other transferases, such as amino acid mechanism would seem likely. However,
transferases, to catalyze a reversible conjugation numerous studies have found no differences in
of amino acid residues to auxinic herbicides, but a glyphosate degradation in glyphosate-resistant
herbicidally active metabolite may be recovered weeds. Pan et al. (147), however, recently found
upon hydrolysis (140). that the mechanism of glyphosate resistance of an
Similarly, several other classes of transferases E. colona biotype involves elevated levels of aldo-
may be involved in Phase II metabolism. keto reductase (AKR; EC 1.1) due to upregulation
However, enzymes such as malonyl-transferase of two AKR genes, resulting in more rapid
typically impart herbicide selectivity, but have not metabolism of glyphosate to AMPA. Earlier,
been demonstrated to be directly involved in overexpression of an AKR gene from rice provided
evolved herbicide resistance (141). glyphosate resistance to tobacco, and silencing this
gene in rice caused hypersensitivity to glyphosate
Aryl acylamidase (148). Whether AKRs are the GOX enzymes that
The mechanism of natural tolerance of rice cause accumulation of AMPA in other weeds and
(Oryza spp.) to the PSII inhibitor herbicide crops remains to be determined (146).
propanil is high levels of aryl acylamidase (AA;

15
-Cyanoalanine synthase gene types, many of which exist in plants as gene
An example of the broad-spectrum resistance families. This means that identifying the specific
that can be conferred by enhanced metabolic genes involved in a particular case of resistance
activity is quinclorac resistance in E. phyllopogon. can be difficult. Gene family members can be
Quinclorac is a unique auxin-type herbicide difficult to distinguish using traditional
because it has activity on grasses. Normally, most sequencing techniques and can be subject to
synthetic auxin herbicides have excellent activity specific expression regulation at various
on eudicot species but no or very little activity on developmental stages or following herbicide
grasses. Quinclorac stimulates ethylene synthesis treatments. Sequence variation among gene family
in sensitive grass species, which results in members occurs normally and plays an important
accumulation of cyanide and subsequent toxicity evolutionary role in functional diversification, so
(149). The enzyme β-cyanoalanine synthase (β- mutations altering substrate specificity for
CAS; EC 4.4.1.9) detoxifies the cyanide generated herbicide detoxification may also occur. The most
upon ethylene synthesis, but normally β-CAS important gene families for NTSR characterized to
activity is insufficient to protect against the toxic date are P450s and GSTs. Unraveling the complex
accumulation of cyanide. Surprisingly, E. mechanisms involved in NTSR and understanding
phyllopogon populations in California exhibited their relationship to overall plant stress response
quinclorac resistance before quinclorac had been pathways is a clear priority for future research

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


used commercially (150), due to their selection (151). To that end, ongoing research is utilizing
history with multiple other herbicides and next generation sequencing technology and
resulting enrichment for enhanced metabolism transcriptomics to help unravel the molecular and
mechanisms. These quinclorac-resistant genetic basis of NTSR, particularly for
populations were found to produce less ethylene metabolism-based resistance (138,152). An
after quinclorac treatment than susceptible improved understanding of the evolution and
populations. The P450 inhibitor malathion regulation of NTSR could guide an additional
synergized with quinclorac on the resistant classification system for herbicides, old and new,
populations, suggesting that enhanced P450 according to their metabolism or sequestration by
activity detoxified quinclorac and inhibition of this specific NTSR mechanisms.
P450 metabolism enabled quinclorac to reach
phytotoxic concentrations. The mechanism is even Importance of genomics and transcriptomics in
more complicated, however, as β-CAS activity understanding NTSR
was 2- to 3-fold higher in resistant populations. Understanding the molecular and genetic basis
The malathion treatment did not change ethylene of resistance mechanisms can help explain
production in the resistant nor the susceptible patterns of cross-resistance across different
populations, but malathion treatment did inhibit β- herbicide modes of action due to a single
CAS activity in the resistant population. mechanism. Target site mutations are specific for
Therefore, malathion is also an inhibitor of β-CAS herbicides within the same mode of action. NTSR
activity, in addition to P450 activity. β-CAS mechanisms, however, can confer resistance
functions to process accumulated cyanide and across dissimilar modes of action. For example,
metabolize it to non-toxic forms. The enhanced similarities in the resistance mechanism could
activity of this protective mechanism in the provide an explanation for the concomitant
resistant population was inhibited by malathion, evolution of paraquat resistance following
demonstrating that enhanced β-CAS activity can recurrent selection with low glyphosate rates in L.
confer metabolic resistance to quinclorac. An rigidum (153). A more thorough understanding of
additional, unknown mechanism makes the cross-resistance patterns across different modes of
resistant populations less sensitive to the action would be highly beneficial for developing
stimulation of ethylene production by quinclorac. improved herbicide rotation recommendations,
avoiding repeated selection with herbicides that
NTSR Summary are vulnerable to shared resistance mechanisms
The diverse mechanisms that can contribute to despite having different modes of action.
NTSR are complex and involve several different

16
“Omics” approaches (genomics,
transcriptomics, proteomics, and metabolomics) Genetics of herbicide resistance
hold great potential for making rapid advances in
our understanding of NTSR mechanisms (154). Of The inheritance and genetic basis of different
these omics approaches, transcriptomics have resistance mechanisms has evolutionary
provided the most insights thus far. It is generally implications, such as how rapidly selection can
thought that NTSR is often mediated by increased occur and how frequent mechanisms may be in
expression of one or more genes (e.g., a gene unselected populations. As described above, many
encoding a P450 or an ABC transporter) and, herbicide-resistance traits, particularly TSR,
consequently, a transcriptomics approach is require a change of a single enzyme and,
ideally suited to identify such genes (155). One of consequently, are inherited as single-gene traits.
the downfalls of a transcriptomic approach, Furthermore, herbicide-resistance alleles often act
however, is that it usually yields several “false in an additive to dominant fashion. Because of this
positives,” requiring significant downstream single-gene, additive-to-dominant nature of many
research to validate candidate genes. herbicide resistances, their evolution can occur
Consequently, scientists typically opt to select the rapidly, and they can spread effectively by both
most obvious candidates (such as genes annotated seed and pollen (162).
as coding potential metabolism enzymes) for The additive-to-dominant nature of many

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


follow-up study, reducing the likelihood that novel herbicide-resistance traits means that the
NTSR genes will be discovered. Additionally, resistance allele often will be selected in
identification of a gene that is overexpressed and heterozygous plants, particularly in cases where
confers herbicide resistance provides insight into the magnitude of resistance is high. For example,
the physiology of the resistance mechanism, but single ALS amino acid changes can confer
not directly into the evolution of the mechanism. resistance ratios of 100-fold or more, relative to
Specifically, transcriptomics is not well suited to sensitive biotypes (e.g., (163)). In such cases,
determine the molecular change that caused heterozygotes and homozygous-resistant plants
overexpression of the gene. will have similar responses to typical field doses
Draft genomes have recently been obtained for of the herbicide, even if the resistance allele is not
some of our most important herbicide-resistant completely dominant. Resistance in this case can
weeds, including A. tuberculatus (156), C. be considered “functionally” dominant. In
canadensis (157), B. scoparia (64), R. contrast, single amino acid changes to EPSPS
raphanistrum (158), E. crus-galli (159), and E. confer very modest levels of resistance (typically
indica (160). Although some of these genomes are less than 10-fold (19)). In E. indica, an EPSPS
still highly fragmented, those of A. tuberculatus, target-site mutation was additive (164).
C. canadensis, and E. crus-galli have N50s greater Consequently, in this case, plants heterozygous for
than 1.5 Mb (N50 is the length of a an EPSPS mutation potentially could be controlled
computationally-assembled contiguous sequence by full recommended rates of glyphosate and using
for which 50% of the sequences in an assembly are reduced rates would foster resistance evolution.
longer and 50% are shorter; longer N50 indicates Known examples of herbicide resistance
a better assembly). Such high-quality genomes inherited as a recessive trait are restricted to TRS
enable identification of candidate resistance genes to herbicides that bind tubulin (107) and to a few
through approaches such as genetic mapping of cases of resistance to auxinic herbicides (165). As
quantitative trait loci (QTL) in biparental mapping expected, because of the recessive nature of these
populations and genome-wide association studies traits, they have been observed almost exclusively
(GWAS) in existing populations (161). Such in self-pollinated weed species. However,
genomic approaches could identify herbicide- outcrossed weeds that have evolved recessive
resistance genes even if they are not resistance include yellow starthistle (Centaurea
overexpressed, as well as potentially identify solstitialis) resistant to some auxinic herbicides
trans-acting factors that cause overexpression of (166) and L. rigidum resistant to dinitroaniline
genes identified through transcriptomics herbicides (167). Population-genetics theory
approaches. predicts that in an outcrossed population the

17
probability of a resistant plant occurring for a does not follow a single-gene model. In fact, the
recessive mutation in a reasonably sized field (e.g., EPSPS copy number appears to be unstable in
30 ha) is essentially nil, even with a relatively high genetic transmission, as segregating F2 A. palmeri
mutation rate (1 x 10-6 gametes per locus per progeny were highly variable for EPSPS copy
generation) and high weed density (500 m-2) (162). number, from less than the parental copy number
So how were C. solstitialis and L. rigidum able to to greater than the sum of both parents (50).
evolve recessive resistance? In the case of C. NTSR traits, such as enhanced herbicide
solstitialis, it was reported that the species is not metabolism and reduced translocation, sometimes
completely outcrossed, and even the very low show inheritance consistent with a single-gene
selfing rate (<0.1%) would be sufficient for model. Examples include atrazine resistance in
generating a few plants with the homozygous- velvetleaf (Abutilon theophrasti) and A.
recessive mutation (166). Lolium spp., although tuberculatus, paraquat resistance in L. rigidum,
predominantly self-incompatible (female flowers glyphosate resistance in C. canadensis, and 2,4-D
on the plant reject pollen shed from the same plant, resistance in oriental mustard (Sisymbrium
requiring pollination from a different individual), orientale) (171-174). More often, however NTSR
also can exhibit some degree of self-pollination appears to be mediated by multiple genes.
(168). In addition, dinitroaniline herbicide The multigenic nature of NTSR often has been
resistance in L. rigidum was not completely described as “creeping” resistance, in which the

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


recessive; heterozygous plants exhibited some magnitude of resistance slowly increases over
resistance to low herbicide doses (167). repeated herbicide selection, presumably due to
The discussion of herbicide-resistance the stacking of small-effect resistance alleles in the
inheritance so far has included only nuclear genes. population (175). Empirical support for creeping
An important exception to nuclear inheritance of resistance is provided by low-dose selection of
herbicide resistance is maternal inheritance of experimental populations. Such studies, conducted
resistance to PSII inhibitors, reported over 40 with different species and different herbicides,
years ago (169). The basis for this maternal have consistently demonstrated the ability of
inheritance is now known to be a herbicide- plants to evolve increased herbicide resistance (in
insensitive D1 protein, which is encoded by a some cases by many fold) after only two or three
chloroplastic gene and, therefore, transmitted only generations (176-178). That such a shift in
maternally in most species (5). For species that are herbicide response was due to the accumulation of
predominantly self-pollinated, the lack of pollen- multiple alleles was supported by a follow-up
mediated dispersal of this resistance trait is study on one of the selected populations that
probably of limited consequence. In outcrossing indicated at least three loci were involved. It is also
species, however, dissemination of nuclear- noteworthy that when similar selection protocols
inherited traits would be favored over maternally were followed on an outcrossed weed (L. rigidum)
inherited traits. This may explain why nuclear- and selfed weeds (wild oat, Avena fatua), the
encoded NTSR to atrazine is more common in A. selfed weed’s shift was only about 5% of that of
tuberculatus (a dioecious—and, therefore, the outcrossed weed (178). This is consistent with
outcrossed—species) than is maternally inherited the idea that the resistance evolution occurred by
atrazine TSR (170). However, TSR to PSII stacking together diverse alleles initially
inhibitors also is known to incur a significant distributed among different plants.
fitness penalty (4), and this also would disfavor the The evolution of NTSR also could involve
evolution of this resistance. epigenetic factors (179). For example, changes in
Inheritance of glyphosate resistance conferred the methylation status of a gene could affect its
by EPSPS duplication in some cases is observed to expression. Epigenetic involvement in evolved
be consistent with a single-gene model. This herbicide resistance is an understudied topic,
apparent single-gene inheritance occurs because although differential methylation of EPSPS
the multiple EPSPS copies are found together in recently was reported between glyphosate-
the genome as tandem repeats (60). In other cases, resistant and sensitive biotypes of C. canadensis
notably in A. palmeri, the EPSPS copies are (180). It is not yet clear, however, if this
present as eccDNA and, consequently, inheritance differential methylation contributes to resistance.

18
Predicting and mitigating herbicide resistance sub-genome is known to contain higher nucleotide
requires understanding not only the genetics of the substitution rates than the early meadow-grass
resistance traits, but also the origins of the genetic (Poa infirma) progenitor sub-genome (186), and
diversity that fuels resistance evolution (181). We this expression dominance toward a more dynamic
are now moving into an era in which population sub-genome may represent a resistance evolution
genomics approaches should be able to shed process present in this species. Despite these
increased light on, for example, the importance of examples just described, studies are limited on the
new mutations vs. standing genetic variation in the influence of ploidy on resistance mechanisms, and
evolutionary history of herbicide resistance (182). further research is needed to understand the
complexities of herbicide resistance and evolution
Influences of ploidy in polyploid weeds.
Understanding resistance mechanisms in
polyploid species is complicated by the presence Implications for herbicide-resistant crops
of multiple genomes and associated regulatory The wide range of mechanisms of evolved
processes, ranging from whole-genome to single resistance to herbicides by weeds provides a
allele, that are not a consideration in diploids. In wealth of potential genes for production of
particular, allele dosage, or the relative herbicide-resistant crops. However, not since the
contribution of orthologous or homoeologous production of triazine-resistant canola by

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


genes (homoeologs are genes with homologous backcrossing canola with an evolved triazine-
sequence/function that are present in the different resistant weed (187), has a gene from a weed been
genomes of a polyploid) in a polyploid individual, used to produce a commercial herbicide-resistant
can influence the effect of target site mutations in crop. The rapid evolution of resistance to ALS and
weed species. ACCase inhibitors by single nucleotide
The influence of poidy has been documented polymorphisms provided a clue that crops resistant
for herbicide resistance in multiple species. An to these herbicides could be generated by
ACCase inhibitor-resistant population of mutagenesis and breeding. This approach has been
hexaploid A. fatua had three different mutations in especially successful in producing imidazolinone-
ACCase (at positions 1781, 2078, and 2088) that resistant rice, wheat, oilseed rape, sorghum, maize,
segregated independently, indicating one mutation sunflower, and sugarbeet. Imidazolinone-resistant
present for each homoeologous loci (183). When crops have been the only significant commercial
occurring alone, each allele conferred a lower success with this technology, even though crops
ACCase resistance level than the same mutation resistant to ACCase inhibitors and sulfonylurea
confers in diploid species. The mutations showed ALS inhibitors have been generated by such
an additive effect when co-expressed in vitro, mutation breeding approaches (188).
suggesting in single mutant lines the contribution Relatively few herbicide-resistance genes
of susceptible ACCase alleles were diluting the have been utilized to produce commercial,
effect of the resistance allele. Similar allele dose transgenic, herbicide-resistant crops, and most of
effects were observed in IMI-resistant hexaploid them have been of bacterial origin. These include
wheat (184). Resistance mutations on ALS genes bacterial genes for a glyphosate-resistant form of
on two of the three sub-genomes (referring to one EPSPS and those that degrade glyphosate,
of the multiple genomes within a polyploid) glufosinate, 2,4-D, bromoxynil, and dicamba. A
resulted in higher resistance than when the non-bacterial exception is a TIPS form of EPSPS
mutation was present on only one sub-genome. that was engineered from a plant EPSPS gene and
More recently, an allotetraploid P. annua is used in some glyphosate-resistant maize
population was identified with resistance to cultivars (22). It took fifteen more years after use
glyphosate via a Pro106Leu mutation in EPSPS as a transgene for the TIPS form of EPSPS to
and 7-fold increased EPSPS copy number (185). evolve in a weed (24). A bacterial gene with a
Interestingly, both the target site mutation and single nucleotide polymorphism encoding an
duplicated EPSPS genes were present in isoxaflutole-resistant HPPD is being used as a
homoeologs only from the supine bluegrass (Poa transgene for isoxaflutole-resistant crops (189).
supina) progenitor sub-genome. The P. supina However, another company is developing crops

19
resistant to triketone HPPD inhibitor herbicides pressure over a wide area can overcome any
through use of a mutant triketone-resistant HPPD perceived genetic barrier.
gene from oats (190).
For the most part, researchers developing Implications for herbicide discovery
transgenic herbicide-resistant crops became
involved with crops resistant to a particular Decades of non-diversified chemical weed
herbicide before a weed had evolved resistance to control and the resistance challenges that arose
that herbicide or the genes for weed resistance had from it have taught us several hard lessons about
been discovered. A major consideration in the importance of herbicide stewardship. Research
development of herbicide-resistant crops is the into the evolution of resistance provides valuable
strong benefit of the transgene providing insights for future herbicide discovery, as certain
resistance to only one herbicide or class of resistance mechanisms are more prevalent within
herbicides (e.g., glyphosate or imidazolinones) certain modes of action than others. These findings
controlled by the same company. The very high can be used to pro-actively influence the
cost of developing and obtaining regulatory development of less resistance-prone herbicides.
approval of such a crop can preclude production of For TSR, the rate and likelihood at which a
a crop that is resistant to products belonging to a point mutation conferring resistance occurs is
competing company. If transgenic, herbicide- determined by both the target protein and the

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


resistant crops continue to be developed, the inhibitor. Many herbicidal molecules compete
burgeoning information on the molecular biology with substrates that are crucial for plant survival
of resistance to herbicides evolved by weeds might by emulating their volume, surface geometry, and
influence the choice of future transgenes. For binding properties. Compounds such as
example, targeted mutagenesis on a codon glyphosate and transition state analogues such as
involved in resistance to certain PDS inhibitors glufosinate tightly fit into the active site of the
identified a double mutation not likely to occur in target enzyme (competitive binding) where they
nature that imparts even greater resistance to these form strong hydrogen bridges. Point mutations in
herbicides (14). Interestingly, this mutation this highly conserved area would render the
provided negative cross-resistance to other classes enzyme dysfunctional. In contrast, herbicides
of PDS inhibitors. This may be particularly useful inhibiting ACCase, ALS, PPO, or PDS adopt
in managing volunteer plants as well as weeds that different binding modes than the substrate so that
may evolve resistance by selection for this a mutation can have differential effects on
mutation. inhibitor and substrate binding. The enzyme’s
What we know about the evolutionary biology reaction would still take place despite an amino
of herbicide-resistant weeds will influence what acid change in the inhibitor binding domain –
herbicides are chosen for future herbicide-resistant where genetic diversity within and among species
crops and how resulting crops are managed. An is more likely. However, changes can come at a
impetus to the development of glyphosate- cost of a reduced turnover rate (kcat) for the
resistant crops in the early 1990s was the false substrate, as seen with the Gly210 PPX2 deletion.
view that evolution of weeds resistant to Therefore, the vast majority of resistance-
glyphosate would be almost impossible because conferring point mutations are found in
more than one codon in EPSPS would have to hydrophobic binding sites away from the active
mutate simultaneously to provide robust TSR reaction site. ALS point mutations, for example,
(191). The unparalleled success of glyphosate- are found in the non-conserved terminal ends of
resistant crops and the resulting massive selection the ALS sequences (192). Not all point mutations
pressure for glyphosate resistance helped to confer resistance to all compounds from the same
overcome the difficulty that weeds had in evolving chemical class. For example, while the 1781
resistance. Thus, even if a herbicide is selected mutation in A. myosuroides confers resistance to
because mutations imparting resistance are less all ACCase inhibitors, the 2027 mutation confers
likely to evolve, resistance management must be resistance to aryloxyphenoxypropionates, but not
part of the stewardship of future herbicide- to cyclohexanediones (10). The closer the
resistant crops because sustained, strong selection mutation is to the active site, the more likely it is

20
for a point mutation to negatively affect the strategic positions of the molecules. A second
enzyme’s catalytic rate but also to confer major pathway of metabolic degradation is the
resistance to all molecules of a site of action rather nucleophilic attack of electrophilic centers in the
than to just one chemical class. inhibitor molecule, such as through GSTs. Such
Sometimes, several homologues for the same soft regions can be circumvented by the
target exist in a plant. Resistance via point introduction of heteroatoms such as oxygen at the
mutations would be less likely to evolve if these respective place of the molecule. However, these
isoforms were expressed at the same time, non- changes can alter a molecule’s binding properties.
redundant in function, and can be inhibited by the Once Phase I metabolism has taken place, phase II
same molecule. In the case of a point mutation in metabolism, such as glycosylation, cannot be
one of the target enzyme isoforms, the inhibition prevented. The biggest challenge in pesticide
of the remaining enzymes would then still lead to discovery lies in creating a biologically active
the death of the plant. Alternatively, one herbicide molecule that is toxicologically benign to non-
may inhibit two different target proteins (e.g., target organisms and is degraded in the
auxinic herbicides). The probability of an environment as well as by non-target organisms.
individual to evolve two coincidental, independent Normalized by use frequency, strong
and catalytically competent point mutations in differences in the rates of herbicide resistance
each target enzyme is equal to the product of the evolution between different modes of action have

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


frequencies of the single mutation. Hence, been observed (196). Glyphosate has been such an
molecules that fully inhibit several targets that excellent herbicide because it tightly occupies the
individually would lead to plant death, are active site of the EPSPS-S3P complex only,
preferred. Likewise, target site mutations are less outcompeting the substrate PEP. Its low
likely to occur if they need to be homozygous to dissociation rate allows for almost irreversible
confer resistance (recessive inheritance). In that inhibition of the enzyme. Heteroatoms and
case an identical mutation would need to be phosphonates that cannot be oxidized leave it inert
present in both the paternal and maternal alleles. for metabolism by most metabolic enzymes.
They are also less likely to spread if they impart a Hence, resistant plants have generally had to
fitness penalty for the plant, and if compensatory “resort” to the fairly complicated and comparably
mutations rarely occur, as in the case of TSR to slow-to-evolve mechanism of EPSPS gene
PSII inhibitors (193,194). Duplication of the gene duplication and even slower-to-evolve double
for the herbicide target protein, likely guided by codon mutations in EPSPS to survive. In the end,
transposons or non-homologous recombination, is any type of persistent selection pressure will be
a resistance mechanism that is particularly eventually met by adaptation, and the potential
difficult to predict in herbicide discovery number of resistance mechanisms are many. Even
strategies. if all currently known resistance mechanisms
The physicochemical properties of a molecule could be addressed, new, unpredictable
define how it binds the protein target, as well as mechanisms may evolve, such as the “phoenix
strongly influence uptake, translocation, and phenomenon” mechanism of glyphosate and 2,4-
degradation pathway before it reaches the target D resistance. Therefore, it is futile to set a goal of
site. For example, a major pathway of herbicide finding a “resistance-proof” herbicide.
metabolism is through oxidation via P450s, which Nevertheless, the considerations listed in Table 3
can add hydroxyl groups to a molecule. This should be considered in herbicide discovery
reaction is more difficult for small and polar strategies. Weed biology, herbicide application
molecules, which allow for strong hydrogen rates and herbicide mixtures all influence the type
bridges, inhibiting the active site of the target only. and rate at which resistance mechanisms evolve
In contrast, hydrophobic parts of the molecule, like (197). We can aim to delay the evolution of
aryl or alkyl moieties, are more prone to oxidation, resistance to newly developed molecules by
as there is a positive correlation between choosing sensible targets and optimized
lipophilicity and KM for P450-catalyzed reactions molecules. We must recognize, however, that
(195). Prevention of this metabolic degradation doing so will not replace the necessity to practice
can be achieved by introduction of halogens into integrated resistance management.

21
Conflict of interest: The authors declare that they have no conflicts of interest with the contents of this
article.
Author contributions: All authors contributed substantially to the writing of this review.

References
1. Dayan, F. E., Barker, A., Bough, R., Ortiz, M., Takano, H., and Duke, S. O. (2019) Herbicide
Mechanisms of Action and Resistance. In Comprehensive Biotechnology (Moo-Young, M. ed.), 3rd
Ed., Pergamon, Oxford. pp 36-48
2. Herbicide Resistance Action Committee (2020) HRAC Mode of Action Classification 2020.
https://hracglobal.com/tools/hrac-mode-of-action-classification-2020-map. Accessed April 22,
2020.
3. Heap, I. (2020) The international survey of herbicide resistant weeds. Available on-line:
www.weedscience.com. Accessed Feb 25, 2020.
4. Gronwald, J. W. (1994) Resistance to photosystem II inhibiting herbicides. In Herbicide Resistance
in Plants: Biology and Biochemistry (Powles, S. B., and Holtum, J. A. M. eds.). pp 27-60

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


5. Hirschberg, J., and McIntosh, L. (1983) Molecular basis of herbicide resistance in Amaranthus
hybridus. Science 222, 1346-1349
6. Anthony, R. G., Waldin, T. R., Ray, J. A., Bright, S. W. J., and Hussey, P. J. (1998) Herbicide
resistance caused by spontaneous mutation of the cytoskeletal protein tubulin. Nature 393, 260-263
7. Chen, J., Chu, Z., Han, H., Goggin, D. E., Yu, Q., Sayer, C., and Powles, S. B. (2020) A Val-202-
Phe α-tubulin mutation and enhanced metabolism confer dinitroaniline resistance in a single Lolium
rigidum population. Pest Manag. Sci. 76, 645-652
8. Chu, Z., Chen, J., Nyporko, A., Han, H., Yu, Q., and Powles, S. (2018) Novel α-tubulin mutations
conferring resistance to dinitroaniline herbicides in Lolium rigidum. Front. Plant Sci. 9
9. Delye, C., Menchari, Y., Michel, S., and Darmency, H. (2004) Molecular bases for sensitivity to
tubulin-binding herbicides in green foxtail. Plant Physiol. 136, 3920-3932
10. Kaundun, S. S. (2014) Resistance to acetyl-CoA carboxylase-inhibiting herbicides. Pest Manag.
Sci. 70, 1405-1417
11. Beckie, H. J., and Tardif, F. J. (2012) Herbicide cross resistance in weeds. Crop Protect. 35, 15-28
12. Michel, A., Arias, R. S., Scheffler, B. E., Duke, S. O., Netherland, M. D., and Dayan, F. E. (2004)
Somatic mutation-mediated evolution of herbicide resistance in the nonindigenous invasive plant
hydrilla (Hydrilla verticillata). Molec. Ecol. 13, 3229-3237
13. Dayan, F. E., and Netherland, M. D. (2005) Hydrilla, the perfect aquatic weed, becomes more
noxious than ever. Outlooks Pest Manag. 16, 277-282
14. Arias, R. S., Dayan, F. E., Michel, A., Howell, J. L., and Scheffler, B. E. (2006) Characterization of
a higher plant herbicide-resistant phytoene desaturase and its use as a selectable marker. Plant
Biotechnol. J. 4, 263-273
15. Netherland, M. D., and Jones, D. (2015) Fluridone-resistant hydrilla (Hydrilla verticillata) is still
dominant in the Kissimmee chain of lakes. Invasive Plant Sci. Manag. 8, 212-218
16. McCourt, J. A., Pang, S. S., King-Scott, J., Guddat, L. W., and Duggleby, R. G. (2006) Herbicide-
binding sites revealed in the structure of plant acetohydroxyacid synthase. Proc. Natl. Acad. Sci.
USA 130, 569–573
17. Preston, C., and Powles, S. B. (2002) Evolution of herbicide resistance in weeds: Initial frequency
of target site-based resistance to acetolactate synthase-inhibiting herbicides in Lolium rigidum.
Heredity 88, 8-13
18. Jander, G., Baerson, S. R., Hudak, J. A., Gonzalez, K. A., Gruys, K. J., and Last, R. L. (2003)
Ethylmethanesulfonate saturation mutagenesis in Arabidopsis to determine frequency of herbicide
resistance. Plant Physiol. 131, 139-146

22
19. Sammons, D. R., and Gaines, T. A. (2014) Glyphosate resistance: State of knowledge. Pest Manag.
Sci. 70, 1367-1377
20. Healy-Fried, M. L., Funke, T., Priestman, M. A., Han, H., and Schonbrunn, E. (2007) Structural
basis of glyphosate tolerance resulting from mutations of Pro(101) in Escherichia coli 5-
enolpyruvylshikimate-3-phosphate synthase. J. Biol. Chem. 282, 32949-32955
21. Li, J., Peng, Q., Han, H., Nyporko, A., Kulynych, T., Yu, Q., and Powles, S. (2018) Glyphosate
resistance in Tridax procumbens via a novel EPSPS Thr-102-Ser substitution. J. Agric. Food Chem.
66, 7880-7888
22. Sidhu, R. S., Hammond, B. G., Fuchs, R. L., Mutz, J.-N., Holden, L. R., George, B., and Olson, T.
(2000) Glyphosate-tolerant corn:  The composition and feeding value of grain from glyphosate-
tolerant corn is equivalent to that of conventional corn (Zea mays L.). J. Agric. Food Chem. 48,
2305-2312
23. Funke, T., Yang, Y., Han, H., Healy-Fried, M., Olesen, S., Becker, A., and Schönbrunn, E. (2009)
Structural basis of glyphosate resistance resulting from the double mutation Thr97 → Ile and Pro101
→ Ser in 5-enolpyruvylshikimate-3-phosphate synthase from Escherichia coli. J. Biol. Chem. 284,
9854-9860
24. Yu, Q., Jalaludin, A., Han, H., Chen, M., Sammons, R. D., and Powles, S. B. (2015) Evolution of a
double amino acid substitution in the 5-enolpyruvylshikimate-3-phosphate synthase in Eleusine

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


indica conferring high-level glyphosate resistance. Plant Physiol. 167, 1440-1447
25. Dill, G. M. (2005) Glyphosate-resistant crops: history, status and future. Pest Manag. Sci. 61, 219-
224
26. Han, H., Vila-Aiub, M. M., Jalaludin, A., Yu, Q., and Powles, S. B. (2017) A double EPSPS gene
mutation endowing glyphosate resistance shows a remarkably high resistance cost. Plant Cell
Environ. 40, 3031-3042
27. Alcántara-de la Cruz, R., Fernández-Moreno, P. T., Ozuna, C. V., Rojano-Delgado, A. M., Cruz-
Hipolito, H. E., Domínguez-Valenzuela, J. A., Barro, F., and 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
28. Takano, H. K., Fernandes, V. N. A., Adegas, F. S., Oliveira Jr, R. S., Westra, P., Gaines, T. A., and
Dayan, F. E. (2020) A novel TIPT double mutation in EPSPS conferring glyphosate resistance in
tetraploid Bidens subalternans. Pest Manag. Sci. 76, 95-102
29. Perotti, V. E., Larran, A. S., Palmieri, V. E., Martinatto, A. K., Alvarez, C. E., Tuesca, D., and
Permingeat, H. R. (2019) A novel triple amino acid substitution in the EPSPS found in a high-level
glyphosate resistant Amaranthus hybridus population from Argentina. Pest Manag. Sci. 75, 1242-
1251
30. García, M. J., Palma-Bautista, C., Rojano-Delgado, A. M., Bracamonte, E., Portugal, J., Alcántara-
de la Cruz, R., and De Prado, R. (2019) The triple amino acid substitution TAP-IVS in the EPSPS
gene confers high glyphosate resistance to the superweed Amaranthus hybridus. Internat. J. Mol.
Sci. 20, 2396
31. Webb, S. R., and Hall, J. C. (1995) Auxinic herbicide-resistant and herbicide-susceptible wild
mustard (Sinapis arvensis L.) biotypes: Effect of auxinic herbicides on seedling growth and auxin-
binding activity. Pestic. Biochem. Physiol. 52, 137-148
32. Walsh, T. A., Neal, R., Merlo, A. O., Honma, M., Hicks, G. R., Wolff, K., Matsumura, W., and
Davies, J. P. (2006) Mutations in an auxin receptor homolog AFB5 and in SGT1b confer resistance
to synthetic picolinate auxins and not to 2,4-dichlorophenoxyacetic acid or indole-3-acetic acid in
Arabidopsis. Plant Physiol. 142, 542-552
33. LeClere, S., Wu, C., Westra, P., and Sammons, R. D. (2018) Cross-resistance to dicamba, 2,4-D,
and fluroxypyr in Kochia scoparia is endowed by a mutation in an AUX/IAA gene. Proc. Nat. Acad.
Sci. USA 115, E2911-E2920

23
34. Patzoldt, W. L., Hager, A. G., McCormick, J. S., and Tranel, P. J. (2006) A codon deletion confers
resistance to herbicides inhibiting protoporphyrinogen oxidase. Proc. Nat. Acad. Sci. USA 103,
12329-12334
35. Dayan, F. E., Barker, A., and Tranel, P. J. (2018) Origins and structure of chloroplastic and
mitochondrial plant protoporphyrinogen oxidases: implications for the evolution of herbicide
resistance. Pest Manag. Sci. 74, 2226-2234
36. Riggins, C. W., and Tranel, P. J. (2012) Will the Amaranthus tuberculatus resistance mechanism to
PPO-inhibiting herbicides evolve in other Amaranthus species? Internat. J. Agron. 2012, 305764
37. Salas, R. A., Burgos, N. R., Tranel, P. J., Singh, S., Glasgow, L., Scott, R. C., and Nichols, R. L.
(2016) Resistance to PPO-inhibiting herbicide in Palmer amaranth from Arkansas. Pest Manag.
Sci. 72, 864-869
38. Dayan, F. E., Daga, P. R., Duke, S. O., Lee, R. M., Tranel, P. J., and Doerksen, R. J. (2010)
Biochemical and structural consequences of a glycine deletion in the alpha-8 helix of
protoporphyrinogen oxidase. Biochim. Biophys. Act. Protein Proteom. 1804, 1548-1556
39. Patzoldt, W. L., Tranel, P. J., and Hager, A. G. (2005) A waterhemp (Amaranthus tuberculatus)
biotype with multiple resistance across three herbicide sites of action. Weed Sci. 53, 30-36
40. Armel, G. R., Nielson, R. L., Liebl, R. A., Bowe, S., Hennigh, D. S., Francis, I. K., Oostlander, M.
D., and Ramos, R. A. (2018) Trifludimoxazin: a global perspective on a versatile PPO herbicide.

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


Weed Sci. Soc. Amer. Abstr., 196
41. Baerson, S. R., Rodriguez, D. J., Biest, N. A., Tran, M., You, J. S., Kreuger, R. W., Dill, G. M.,
Pratley, J. E., and Gruys, K. J. (2002) Investigating the mechanism of glyphosate resistance in rigid
ryegrass (Lolium ridigum). Weed Sci. 50, 721-730
42. Yu, Q., Nelson, J. K., Zheng, M. Q., Jackson, M., and Powles, S. B. (2007) Molecular
characterisation of resistance to ALS-inhibiting herbicides in Hordeum leporinum biotypes. Pest
Manag. Sci. 63, 918-927
43. Bradley, K. W., Wu, J. R., Hatzios, K. K., and Hagood, E. S. (2001) The mechanism of resistance
to aryloxyphenoxypropionate and cyclohexanedione herbicides in a Johnsongrass biotype. Weed
Sci. 49, 477-484
44. Innan, H., and Kondrashov, F. (2010) The evolution of gene duplications: classifying and
distinguishing between models. Nat. Rev. Genet. 11, 97-108
45. Flagel, L. E., and Wendel, J. F. (2009) Gene duplication and evolutionary novelty in plants. New
Phytol. 183, 557-564
46. Gaines, T. A., Zhang, W., Wang, D., Bukun, B., Chisholm, S. T., Shaner, D. L., Nissen, S. J.,
Patzoldt, W. L., Tranel, P. J., Culpepper, A. S., Grey, T. L., Webster, T. M., Vencill, W. K.,
Sammons, R. D., Jiang, J. M., Preston, C., Leach, J. E., and Westra, P. (2010) Gene amplification
confers glyphosate resistance in Amaranthus palmeri. Proc. Nat. Acad. Sci. USA 107, 1029-1034
47. Molin, W. T., Wright, A. A., Lawton-Rauh, A., and Saski, C. A. (2017) The unique genomic
landscape surrounding the EPSPS gene in glyphosate resistant Amaranthus palmeri: a repetitive
path to resistance. BMC Genom. 18, 91
48. Koo, D.-H., Molin, W. T., Saski, C. A., Jiang, J., Putta, K., Jugulam, M., Friebe, B., and Gill, B. S.
(2018) Extrachromosomal circular DNA-based amplification and transmission of herbicide
resistance in crop weed Amaranthus palmeri. Proc. Nat. Acad. Sci. USA 115, 3332-3337
49. Giacomini, D. A., Westra, P., and Ward, S. M. (2019) Variable inheritance of amplified EPSPS
gene copies in glyphosate-resistant Palmer amaranth (Amaranthus palmeri). Weed Sci. 67, 176-182
50. Gaines, T. A., Shaner, D. L., Ward, S. M., Leach, J. E., Preston, C., and Westra, P. (2011)
Mechanism of resistance of evolved glyphosate-resistant Palmer amaranth (Amaranthus palmeri).
J. Agric. Food Chem. 59, 5886-5889
51. Molin, W. T., Wright, A. A., VanGessel, M. J., McCloskey, W. B., Jugulam, M., and Hoagland, R.
E. (2017) Survey of the genomic landscape surrounding the 5‐ enolpyruvylshikimate‐ 3‐
phosphate synthase (EPSPS) gene in glyphosate-resistant Amaranthus palmeri from geographically
distant populations in the United States. Pest Manag. Sci. 74, 1109-1117

24
52. Patterson, E. L., Pettinga, D. J., Ravet, K., Neve, P., and Gaines, T. A. (2018) Glyphosate
resistance and EPSPS gene duplication: Convergent evolution in multiple plant species. J. Hered.
109, 117-125
53. Ngo, T. D., Malone, J. M., Boutsalis, P., Gill, G., and Preston, C. (2018) EPSPS gene amplification
conferring resistance to glyphosate in windmill grass (Chloris truncata) in Australia. Pest Manag.
Sci. 74, 1101-1108
54. Adu-Yeboah, P., Malone, J. M., Fleet, B., Gill, G., and Preston, C. (2020) EPSPS gene
amplification confers resistance to glyphosate resistant populations of Hordeum glaucum Stued
(northern barley grass) in South Australia. Pest Manag. Sci. 76, 1214-1221
55. Malone, J. M., Morran, S., Shirley, N., Boutsalis, P., and Preston, C. (2016) EPSPS gene
amplification in glyphosate‐ resistant Bromus diandrus. Pest Manag. Sci. 72, 81-88
56. Chen, J., Huang, H., Zhang, C., Wei, S., Huang, Z., Chen, J., and Wang, X. (2015) Mutations and
amplification of EPSPS gene confer resistance to glyphosate in goosegrass (Eleusine indica).
Planta 242, 859-868
57. Lorentz, L., Gaines, T. A., Nissen, S. J., Westra, P., Strek, H., Dehne, H. W., Ruiz-Santaella, J. P.,
and Beffa, R. (2014) Characterization of glyphosate resistance in Amaranthus tuberculatus
populations. J. Agric. Food Chem. 62, 8134-8142
58. Wiersma, A. T., Gaines, T. A., Preston, C., Hamilton, J. P., Giacomini, D., Buell, C. R., Leach, J.

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


E., and Westra, P. (2015) Gene amplification of 5-enol-pyruvylshikimate-3-phosphate synthase in
glyphosate-resistant Kochia scoparia. Planta 241, 463-474
59. Salas, R. A., Dayan, F. E., Pan, Z., Watson, S. B., Dickson, J. W., Scott, R. C., and Burgos, N. R.
(2012) EPSPS gene amplification in glyphosate-resistant Italian ryegrass (Lolium perenne ssp.
multiflorum) from Arkansas. Pest Manag. Sci. 68, 1223-1230
60. Jugulam, M., Niehues, K., Godar, A. S., Koo, D.-H., Danilova, T., Friebe, B., Sehgal, S., Varanasi,
V. K., Wiersma, A., Westra, P., Stahlman, P. W., and Gill, B. S. (2014) Tandem amplification of a
chromosomal segment harboring EPSPS locus confers glyphosate resistance in Kochia scoparia.
Plant Physiol. 166, 1200-1207
61. Nandula, V. K., Wright, A. A., Bond, J. A., Ray, J. D., Eubank, T. W., and Molin, W. T. (2014)
EPSPS amplification in glyphosate-resistant spiny amaranth (Amaranthus spinosus): a case of gene
transfer via interspecific hybridization from glyphosate-resistant Palmer amaranth (Amaranthus
palmeri). Pest Manag. Sci. 70, 1902-1909
62. Wassom, J. J., and Tranel, P. J. (2005) Amplified fragment length polymorphism-based genetic
relationships among weedy Amaranthus species. J. Hered. 96, 410-416
63. Gaines, T. A., Ward, S. M., Bukun, B., Preston, C., Leach, J. E., and Westra, P. (2012) Interspecific
hybridization transfers a previously unknown glyphosate resistance mechanism in Amaranthus
species. Evolution. Applic. 5, 29-38
64. Patterson, E. L., Saski, C. A., Sloan, D. B., Tranel, P. J., Westra, P., and Gaines, T. A. (2019) The
draft genome of Kochia scoparia and the mechanism of glyphosate resistance via transposon-
mediated EPSPS tandem gene duplication. Genome Biol. Evol. 11, 2927-2940
65. Gaines, T. A., Patterson, E. L., and Neve, P. (2019) Molecular mechanisms of adaptive evolution
revealed by global selection for glyphosate resistance. New Phytol. 223, 1770-1775
66. Laforest, M., Soufiane, B., Simard, M.-J., Obeid, K., Page, E., and Nurse, R. E. (2017) Acetyl-CoA
carboxylase overexpression in herbicide resistant large crabgrass (Digitaria sanguinalis). Pest
Manag. Sci. 73, 2227–2235
67. Menendez, J., Rojano-Delgado, M. A., and De Prado, R. (2014) Differences in herbicide uptake,
translocation, and distribution as sources of herbicide resistance in weeds. Am. Chem. Soc. Symp.
Ser. 1171, 141-157
68. Hess, F. E. (1985) Herbicide absorption and translocation and their relationship to plant tolerance
and susceptibility. In Weed Physiology (Duke, S. O. ed.), CRC Press, Boca Raton, FL. pp 191-214
69. Devine, M., Duke, S. O., and Fedtke, C. (1992) Physiology of Herbicide Action, Prentice Hall,
Englewood Cliffs, New Jersey

25
70. Stegink, S. J., and Vaughn, K. C. (1988) Norflurazon (SAN-9789) reduces abscisic acid levels in
cotton seedlings: A glandless isoline is more sensitive than its glanded counterpart. Pestic.
Biochem. Physiol. 31, 269-275
71. White, A. D., Owen, M. D. K., Hartzler, R. G., and Cardina, J. (2002) Common sunflower
resistance to acetolactate synthase–inhibiting herbicides. Weed Sci. 50, 432-437
72. Svyantek, A. W., Aldahir, P., Chen, S., Flessner, M. L., McCullough, P. E., Sidhu, S. S., and
McElroy, J. S. (2016) Target and nontarget resistance mechanisms induce annual bluegrass (Poa
annua) resistance to atrazine, amicarbazone, and diuron. Weed Technol. 30, 773-782
73. Michitte, P., De Prado, R., Espinoza, N., Ruiz-Santaella, J. P., and Gauvrit, C. (2007) Mechanisms
of resistance to glyphosate in a ryegrass (Lolium multiflorum) biotype from Chile. Weed Sci. 55,
435-440
74. Vila-Aiub, M. M., Balbi, M. C., Distefano, A. J., Fernandez, L., Hopp, E., Yu, Q., and Powles, S.
B. (2012) Glyphosate resistance in perennial Sorghum halepense (Johnsongrass), endowed by
reduced glyphosate translocation and leaf uptake. Pest Manag. Sci. 68, 430-436
75. Nandula, V. K., Reddy, K. N., Poston, D. H., Rimando, A. M., and Duke, S. O. (2008) Glyphosate
tolerance mechanism in Italian ryegrass (Lolium multiflorum) from Mississippi. Weed Sci. 56, 344-
349
76. Nandula, V. K., Ray, J. D., Ribeiro, D. N., Pan, Z., and Reddy, K. N. (2013) Glyphosate resistance

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


in tall waterhemp (Amaranthus tuberculatus) from Mississippi is due to both altered target-site and
nontarget-site mechanisms. Weed Sci. 61, 374-383
77. Lorraine-Colwill, D. F., Powles, S. B., Hawkes, T. R., Hollinshead, P. H., Warner, S. A. J., and
Preston, C. (2002) Investigations into the mechanism of glyphosate resistance in Lolium rigidum.
Pestic. Biochem. Physiol. 74, 62-72
78. Feng, P. C. C., Tran, M., Chiu, T., Sammons, R. D., Heck, G. R., and CaJacob, C. A. (2004)
Investigations into glyphosate-resistant horseweed (Conyza canadensis): Retention, uptake,
translocation, and metabolism. Weed Sci. 52, 498-505
79. Ge, X., d'Avignon, D. A., Ackerman, J. J. H., and Sammons, R. D. (2010) Rapid vacuolar
sequestration: The horseweed glyphosate resistance mechanism. Pest Manag. Sci. 66, 345-348
80. Ge, X., d'Avignon, D. A., Ackerman, J. J. H., Duncan, B., Spaur, M. B., and Sammons, R. D.
(2011) Glyphosate-resistant horseweed made sensitive to glyphosate: low-temperature suppression
of glyphosate vacuolar sequestration revealed by P-31 NMR. Pest Manag. Sci. 67, 1215-1221
81. Peng, Y., Abercrombie, L. L. G., Yuan, J. S., Riggins, C. W., Sammons, R. D., Tranel, P. J., and
Stewart, C. N. (2010) Characterization of the horseweed (Conyza canadensis) transcriptome using
GS-FLX 454 pyrosequencing and its application for expression analysis of candidate non-target
herbicide resistance genes. Pest Manag. Sci. 66, 1053-1062
82. Yuan, J. S., Abercrombie, L. L. G., Cao, Y., Halfhill, M. D., Zhou, X., Peng, Y., Hu, J., Rao, M. R.,
Heck, G. R., Larosa, T. J., Sammons, R. D., Wang, X., Ranjan, P., Johnson, D. H., Wadl, P. A.,
Scheffler, B. E., Rinehart, T. A., Trigiano, R. N., and Stewart, C. N., Jr. (2010) Functional
genomics analysis of horseweed (Conyza canadensis) with special reference to the evolution of
non-target-site glyphosate resistance. Weed Sci. 58, 109-117
83. Vila-Aiub, M. M., Gundel, P. E., Yu, Q., and Powles, S. B. (2013) Glyphosate resistance in
Sorghum halepense and Lolium rigidum is reduced at suboptimal growing temperatures. Pest
Manag. Sci. 69, 228-232
84. Yu, Q., Cairns, A., and Powles, S. B. (2004) Paraquat resistance in a population of Lolium rigidum.
Function. Plant Biol. 31, 247-254
85. Yu, Q., Huang, S., and Powles, S. (2010) Direct measurement of paraquat in leaf protoplasts
indicates vacuolar paraquat sequestration as a resistance mechanism in Lolium rigidum. Pestic.
Biochem. Physiol. 98, 104-109
86. Brunharo, C. A. C. G., and Hanson, B. D. (2017) Vacuolar sequestration of paraquat is involved in
the resistance mechanism in Lolium perenne L. spp. multiflorum. Front. Plant Sci. 8, 1485

26
87. Moretti, M., and Hanson, B. (2017) Reduced translocation is involved in resistance to glyphosate
and paraquat in Conyza bonariensis and Conyza canadensis from California. Weed Res. 57, 25-34
88. Li, J., Mu, J., Bai, J., Fu, F., Zou, T., An, F., Zhang, J., Jing, H., Wang, Q., Li, Z., Yang, S., and
Zuo, J. (2013) PARAQUAT RESISTANT1, a Golgi-localized putative transporter protein, is
involved in intracellular transport of paraquat. Plant Physiol. 162, 470-483
89. Fujita, M., Fujita, Y., Iuchi, S., Yamada, K., Kobayashi, Y., Urano, K., Kobayashi, M.,
Yamaguchi-Shinozaki, K., and Shinozaki, K. (2012) Natural variation in a polyamine transporter
determines paraquat tolerance in Arabidopsis. Proc. Natl. Acad. Sci. USA 109, 6343-6347
90. Luo, Q., Wei, J., Dong, Z., Shen, X., and Chen, Y. (2019) Differences of endogenous polyamines
and putative genes associated with paraquat resistance in goosegrass (Eleusine indica L.). PLoS
ONE 14, e0216513
91. Riar, D. S., Burke, I. C., Yenish, J. P., Bell, J., and Gill, K. (2011) Inheritance and physiological
basis for 2,4-D resistance in prickly lettuce (Lactuca serriola L.). J. Agric. Food Chem. 59, 9417-
9423
92. Goggin, D. E., Cawthray, G. R., and Powles, S. B. (2016) 2,4-D resistance in wild radish: reduced
herbicide translocation via inhibition of cellular transport. J. Experiment. Bot. 67, 3223-3235
93. Pettinga, D. J., Ou, J., Patterson, E. L., Jugulam, M., Westra, P., and Gaines, T. A. (2018) Increased
chalcone synthase (CHS) expression is associated with dicamba resistance in Kochia scoparia. Pest

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


Manag. Sci. 74, 2306-2315
94. Moretti, M. L., Van Horn, C. R., Robertson, R., Segobye, K., Weller, S. C., Young, B. G., Johnson,
W. G., Sammons, R. D., Wang, D., Ge, X., d'Avignon, D. A., Gaines, T. A., Westra, P., Green, A.,
Jeffery, T., Lesperance, M., Tardif, F., Sikkema, P., Hall, J. C., McLean, M., Lawton, M., and
Schulz, B. (2018) Glyphosate resistance in Ambrosia trifida: Part 2. Rapid response physiology and
non‐ target‐ site resistance. Pest Manag. Sci. 74, 1079-1088
95. Van Horn, C. R., Moretti, M. L., Robertson, R. R., Segobye, K., Weller, S. C., Young, B. G.,
Johnson, W. G., Schulz, B., Green, A. C., Jeffery, T., Lesperance, M., Tardif, F., Sikkema, P., Hall,
J. C., McLean, M., Lawton, M., Sammons, R. D., and Gaines, T. (2018) Glyphosate resistance in
Ambrosia trifida: Part 1. Novel rapid cell death response to glyphosate. Pest Manag. Sci. 74, 1071-
1078
96. de Queiroz, A. R. S., Delatorre, C. A., Lucio, F. R., Rossi, C. V. S., Zobiole, L. H. S., and Merotto,
A. (2020) Rapid necrosis: a novel plant resistance mechanism to 2,4-D. Weed Sci. 68, 6-18
97. Yuan, J. S., Tranel, P. J., and Stewart, C. N. (2007) Non-target-site herbicide resistance: a family
business. Trends Plant Sci. 12, 6-13
98. Zimmerlin, A., and Durst, F. (1992) Aryl hydroxylation of the herbicide diclofop by a wheat
cytochome-P-450 monooxygenase - substrate-specificity and physiological activity. Plant Physiol.
100, 874-881
99. Hall, L. M., Moss, S. R., and Powles, S. B. (1995) Mechanism of resistance to chlorotoluron in two
biotypes of the grass weed Alopecurus myosuroides. Pestic. Biochem. Physiol. 53, 180-192
100. Christopher, J. T., Powles, S. B., Liljegren, D. R., and Holtum, J. A. M. (1991) Cross-resistance to
herbicides in annual ryegrass (Lolium rigidum). II. Chlorsulfuron resistance involves a wheat-like
detoxification system. Plant Physiol. 95, 1036-1043
101. Werck-Reichhart, D., and Feyereisen, R. (2000) Cytochromes P450: a success story. Genom. Biol.
1, 1-9
102. Werck-Reichhart, D., Hehn, A., and Didierjean, L. (2000) Cytochromes P450 for engineering
herbicide tolerance. Trends Plant Sci. 5, 116-123
103. Deng, F., and Hatzios, K. K. (2003) Characterization of cytochrome P450-mediated bensulfuron-
methyl O-demethylation in rice. Pestic. Biochem. Physiol. 74, 102-115
104. Busi, R., Dayan, F. E., Francis, I., Goggin, D., Lerchl, J., Porri, A., Powles, S. B., Sun, C., and
Beckie, H. J. (2020) Cinmethylin controls multiple herbicide-resistant Lolium rigidum and its
wheat selectivity is P450-based. Pest Manag. Sci., https://doi.org/10.1002/ps.5798

27
105. Grossmann, K., and Ehrhardt, T. (2007) On the mechanism of action and selectivity of the corn
herbicide topramezone: a new inhibitor of 4-hydroxyphenylpyruvate dioxygenase. Pest Manag. Sci.
63, 429-439
106. Siminszky, B. (2006) Plant cytochrome P450-mediated herbicide metabolism. Phytochem. Rev. 5,
445-458
107. Powles, S. B., and Yu, Q. (2010) Evolution in action: Plants resistant to herbicides. Annu. Rev.
Plant Biol. 61, 317-347
108. Christopher, J. T., Preston, C., and Powles, S. B. (1994) Malathion antagonizes metabolism-based
chlorsulfuron resistance in Lolium rigidum. Pestic. Biochem. Physiol. 49, 172-182
109. Wang, H. C., Li, J., Lv, B., Lou, Y. L., and Dong, L. Y. (2013) The role of cytochrome P450
monooxygenase in the different responses to fenoxaprop-P-ethyl in annual bluegrass (Poa annua
L.) and short awned foxtail (Alopecurus aequalis Sobol.). Pest. Biochem. Physiol. 107, 334-342
110. Küpper, A., Peter, F., Zöllner, P., Lorentz, L., Tranel, P. J., Beffa, R., and Gaines, T. A. (2018)
Tembotrione detoxification in 4-hydroxyphenylpyruvate dioxygenase (HPPD) inhibitor-resistant
Palmer amaranth (Amaranthus palmeri S. Wats.). Pest Manag. Sci. 74, 2325-2334
111. Figueiredo, M. R. A., Leibhart, L. J., Reicher, Z. J., Tranel, P. J., Nissen, S. J., Westra, P.,
Bernards, M. L., Kruger, G. R., Gaines, T. A., and Jugulam, M. (2018) Metabolism of 2,4-
dichlorophenoxyacetic acid contributes to resistance in a common waterhemp (Amaranthus

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


tuberculatus) population. Pest Manag. Sci. 74, 2356-2362
112. Lu, H., Yu, Q., Han, H., Owen, M. J., and Powles, S. B. Non-target-site resistance to PDS-
inhibiting herbicides in a wild radish (Raphanus raphanistrum) population. Pest Manag. Sci., DOI:
10.1002/ps.5733
113. Varanasi, V. K., Brabham, C., and Norsworthy, J. K. (2018) Confirmation and characterization of
non–target site resistance to fomesafen in Palmer amaranth (Amaranthus palmeri). Weed Sci. 66,
702-709
114. Yun, M. S., Yogo, Y., Miura, R., Yamasue, Y., and Fischer, A. J. (2005) Cytochrome P-450
monooxygenase activity in herbicide-resistant and -susceptible late watergrass (Echinochloa
phyllopogon). Pest. Biochem. Physiol. 83, 107-114
115. Yasuor, H., Osuna, M. D., Ortiz, A., Saldain, N. E., Eckert, J. W., and Fischer, A. J. (2009)
Mechanism of resistance to penoxsulam in late watergrass (Echinochloa phyllopogon (Stapf)
Koss.). J. Agric. Food Chem. 57, 3653-3660
116. Yasuor, H., Zou, W., Tolstikov, V. V., Tjeerdema, R. S., and Fischer, A. J. (2010) Differential
oxidative metabolism and 5-ketoclomazone accumulation are involved in Echinochloa phyllopogon
resistance to clomazone. Plant Physiol. 153, 319-326
117. Iwakami, S., Uchino, A., Kataoka, Y., Shibaike, H., Watanabe, H., and Inamura, T. (2014)
Cytochrome P450 genes induced by bispyribac-sodium treatment in a multiple-herbicide resistant
biotype of Echinochloa phyllopogon. Pest Manag. Sci. 70, 549-558
118. Iwakami, S., Endo, M., Saika, H., Okuno, J., Nakamura, N., Yokoyama, M., Watanabe, H., Toki,
S., Uchino, A., and Inamura, T. (2014) Cytochrome P450 CYP81A12 and CYP81A21 are
associated with resistance to two acetolactate synthase inhibitors in Echinochloa phyllopogon.
Plant Physiol. 165, 618-629
119. Dimaano, N. G., Yamaguchi, T., Fukunishi, K., Tominaga, T., and 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
120. Iwakami, S., Kamidate, Y., Yamaguchi, T., Ishizaka, M., Endo, M., Suda, H., Nagai, K., Sunohara,
Y., Toki, S., Uchino, A., Tominaga, T., and Matsumoto, H. (2019) CYP81A P450s are involved in
concomitant cross-resistance to acetolactate synthase and acetyl-CoA carboxylase herbicides in
Echinochloa phyllopogon. New Phytol. 221, 2112-2122
121. Guo, F., Iwakami, S., Yamaguchi, T., Uchino, A., Sunohara, Y., and Matsumoto, H. (2019) Role of
CYP81A cytochrome P450s in clomazone metabolism in Echinochloa phyllopogon. Plant Sci. 283,
321-328

28
122. Manley, B. S., Hatzios, K. K., and Wilson, H. P. (1999) Absorption, translocation, and metabolism
of chlorimuron and nicosulfuron in imidazolinone-resistant and -susceptible smooth pigweed
(Amaranthus hybridus). Weed Technol. 13, 759-764
123. Vencill, W. K., Li, X., and Grey, T. L. (2013) Multiple mechanisms of Palmer amaranth
(Amaranthus palmeri) resistance to ALS-inhibiting herbicides. Proc. Weed Sci. Soc. Am. 53, 363
124. Hausman, N. E., Singh, S., Tranel, P. J., Riechers, D. E., Kaundun, S. S., Polge, N. D., Thomas, D.
A., and Hager, A. G. (2011) Resistance to HPPD-inhibiting herbicides in a population of
waterhemp (Amaranthus tuberculatus) from Illinois, United States. Pest Manag. Sci. 67, 258-261
125. McMullan, P. M., and Green, J. M. (2011) Identification of a tall waterhemp (Amaranthus
tuberculatus) biotype resistant to HPPD-inhibiting herbicides, atrazine, and thifensulfuron in Iowa.
Weed Technol. 25, 514-518
126. Wilde, T., Beffa, R. S., Kleven, T., Philbrook, B., and Strek, H. (2013) HPPD resistance testing in
the USA - Preliminary bioassay results. Proc. Weed Sci. Soc. Am. 53, 131
127. Lu, H., Yu, Q., Han, H., Owen, M. J., and Powles, S. B. (2020) Evolution of resistance to HPPD-
inhibiting herbicides in a wild radish population via enhanced herbicide metabolism. Pest Manag.
Sci. 76, 1929-1937
128. Ma, R., McGinness, D., Hausman, N. E., Tranel, P. J., Hager, A., Kaundun, S. S., Hawkes, T., Vail,
G. D., and Riechers, D. E. (2012) Investigation of resistance mechanisms to mesotrione in a

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


waterhemp (Amaranthus tuberculatus) population from Illinois. Proc. Weed Sci. Soc. Am. 52, 412
129. Oliveira, M. C., Gaines, T. A., Dayan, F. E., Patterson, E. L., Jhala, A. J., and Knezevic, S. Z.
(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
130. Shimabukuro, R. H., Frear, D. S., Swanson, H. R., and Walsh, W. C. (1971) Glutathione
conjugation: enzymatic basis for atrazine resistance in corn. Plant Physiol. 47, 10-14
131. Cummins, I., Dixon, D. P., Freitag-Pohl, S., Skipsey, M., and Edwards, R. (2011) Multiple roles for
plant glutathione transferases in xenobiotic detoxification. Drug Metabol. Rev. 43, 266-280
132. Bakkali, Y., Ruiz-Santaella, J. P., Osuna, M. D., Wagner, J., Fischer, A. J., and De Prado, R. (2007)
Late watergrass (Echinochloa phyllopogon): Mechanisms involved in the resistance to fenoxaprop-
p-ethyl. J. Agric. Food Chem. 55, 4052-4058
133. Cummins, I., Moss, S., Cole, D. J., and Edwards, R. (1997) Glutathione transferases in herbicide-
resistant and herbicide-susceptible black-grass (Alopecurus myosuroides). Pestic. Sci. 51, 244-250
134. Cummins, I., Wortley, D. J., Sabbadin, F., He, Z., Coxon, C. R., Straker, H. E., Sellars, J. D.,
Knight, K., Edwards, L., Hughes, D., Kaundun, S. S., Hutchings, S. J., Steel, P. G., and Edwards,
R. (2013) Key role for a glutathione transferase in multiple-herbicide resistance in grass weeds.
Proc. Nat. Acad. Sci. USA 110, 5812-5817
135. Dücker, R., Zöllner, P., Parcharidou, E., Ries, S., Lorentz, L., and Beffa, R. (2019) Enhanced
metabolism causes reduced flufenacet sensitivity in black-grass (Alopecurus myosuroides Huds.)
field populations. Pest Manag. Sci. 75, 2996-3004
136. Dücker, R., Zöllner, P., Lümmen, P., Ries, S., Collavo, A., and Beffa, R. (2019) Glutathione
transferase plays a major role in flufenacet resistance of ryegrass (Lolium spp.) field populations.
Pest Manag. Sci. 75, 3084-3092
137. Nandula, V. K., Riechers, D. E., Ferhatoglu, Y., Barrett, M., Duke, S. O., Dayan, F. E., Goldberg-
Cavalleri, A., Tétard-Jones, C., Wortley, D. J., Onkokesung, N., Brazier-Hicks, M., Edwards, R.,
Gaines, T., Iwakami, S., Jugulam, M., and Ma, R. (2019) Herbicide metabolism: Crop selectivity,
bioactivation, weed resistance, and regulation. Weed Sci. 67, 149-175
138. Gaines, T. A., Lorentz, L., Figge, A., Herrmann, J., Maiwald, F., Ott, M. C., Han, H., Busi, R., Yu,
Q., Powles, S. B., and Beffa, R. (2014) RNA-Seq transcriptome analysis to identify genes involved
in metabolism-based diclofop resistance in Lolium rigidum. The Plant J. 78, 865-876
139. Duhoux, A., Carrère, S., Gouzy, J., Bonin, L., and Délye, C. (2015) RNA-Seq analysis of rye-grass
transcriptomic response to an herbicide inhibiting acetolactate-synthase identifies transcripts linked
to non-target-site-based resistance. Plant Mol. Biol. 87, 473-487

29
140. Scheel, D., and Sandermann, H. (1981) Metabolism of 2,4-dichlorophenoxyacetic acid in cell
suspension cultures of soybean (Glycine max L.) and wheat (Triticum aestivum L.). Planta 152,
253-258
141. Sandermann, H. J., Haas, M., Messner, B., Pflumacher, S., Schroder, P., and Wetzel, A. (1997) The
role of glucosyl and malonyl conjugation in herbicide selectivity. In Regulation of Enzymatic
Systems Detoxifying Xenobiotics in Plants (Hatzios, K. K. ed.), Kluwer Academic, Dordrecht, The
Netherlands. pp 211–231
142. Chen, J. J., and Matsunaka, S. (1990) The propanil hydrolyzing enzyme aryl acylamidase in the
wild rices of genus Oryza. Pestic. Biochem. Physiol. 38, 26-33
143. Leah, J. M., Caseley, J. C., Riches, C. R., and Valverde, B. (1994) Association between elevated
activity of aryl acylamidase and propanil resistance in Jungle-rice, Echinochloa colona. Pestic. Sci.
42, 281-289
144. Hirase, K., and Hoagland, R. E. (2006) Characterization of aryl acylamidase activity from propanil-
resistant barnyardgrass (Echinochloa crus-galli [L.] Beauv.). Weed Biology and Management 6,
197-203
145. Pedroso, R. M., Al-Khatib, K., Alarcón-Reverte, R., and Fischer, A. J. (2016) A psbA mutation
(Val219 to Ile) causes resistance to propanil and increased susceptibility to bentazon in Cyperus
difformis. Pest Manag. Sci. 72, 1673-1680

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


146. Duke, S. O. (2011) Glyphosate degradation in glyphosate-resistant and -susceptible crops and
weeds. J. Agric. Food Chem. 59, 5835-5841
147. Pan, L., Yu, Q., Han, H., Mao, L., Nyporko, A., Fan, L., Bai, L., and Powles, S. (2019) Aldo-keto
reductase metabolizes glyphosate and confers glyphosate resistance in Echinochloa colona. Plant
Physiol. 181, 1519-1534
148. Vemanna, R. S., Vennapusa, A. R., Easwaran, M., Chandrashekar, B. K., Rao, H., Ghanti, K.,
Sudhakar, C., Mysore, K. S., and Makarla, U. (2017) Aldo-keto reductase enzymes detoxify
glyphosate and improve herbicide resistance in plants. Plant Biotechnol. J. 15, 794-804
149. Grossmann, K. (2010) Auxin herbicides: current status of mechanism and mode of action. Pest
Manag. Sci. 66, 113-120
150. Yasuor, H., Milan, M., Eckert, J. W., and Fischer, A. J. (2012) Quinclorac resistance: a concerted
hormonal and enzymatic effort in Echinochloa phyllopogon. Pest Manag. Sci. 68, 108-115
151. Délye, C., Gardin, J. A. C., Boucansaud, K., Chauvel, B., and Petit, C. (2011) Non-target-site-based
resistance should be the centre of attention for herbicide resistance research: Alopecurus
myosuroides as an illustration. Weed Res. 51, 433-437
152. Gardin, J., Beffa, R., Gouzy, J., Carrere, S., and Delye, C. (2013) A transcriptomics-based approach
enables the first identification of candidate genes involved in non-target-ste-based resistance to
herbicides in a grass weed (Alopecurus myosuroides). Proc. Global Herbic. Resist. Challenge, 68
153. Busi, R., and Powles, S. B. (2011) Reduced sensitivity to paraquat evolves under selection with low
glyphosate doses in Lolium rigidum. Agron. Sustain. Develop. 31, 525-531
154. Maroli, A. S., Gaines, T. A., Foley, M. E., Duke, S. O., Doğramacı, M., Anderson, J. V., Horvath,
D. P., Chao, W. S., and Tharayil, N. (2018) Omics in weed science: A perspective from genomics,
transcriptomics, and metabolomics approaches. Weed Sci. 66, 681-695
155. Giacomini, D. A., Gaines, T., Beffa, R., and Tranel, P. J. (2018) Optimizing RNA-seq studies to
investigate herbicide resistance. Pest Manag. Sci. 74, 2260-2264
156. Kreiner, J. M., Giacomini, D. A., Bemm, F., Waithaka, B., Regalado, J., Lanz, C., Hildebrandt, J.,
Sikkema, P. H., Tranel, P. J., Weigel, D., Stinchcombe, J. R., and Wright, S. I. (2019) Multiple
modes of convergent adaptation in the spread of glyphosate-resistant Amaranthus tuberculatus.
Proc. Natl. Acad. Sci. USA 116, 21076-21084
157. Laforest, M., Martin, S. L., Bisaillon, K., Soufiane, B., Meloche, S., and Page, E. (2020) A
chromosome-scale draft sequence of the Canada fleabane genome. Pest Manag. Sci.,
https://doi.org/10.1002/ps.5753

30
158. Moghe, G. D., Hufnagel, D. E., Tang, H., Xiao, Y., Dworkin, I., Town, C. D., Conner, J. K., and
Shiu, S.-H. (2014) Consequences of whole genome triplication as revealed by comparative
genomic analyses of the wild radish Raphanus raphanistrum and three other brassicaceae species.
Plant Cell 26, 1925-1937
159. Guo, L., Qiu, J., Ye, C., Jin, G., Mao, L., Zhang, H., Yang, X., Peng, Q., Wang, Y., Jia, L., Lin, Z.,
Li, G., Fu, F., Liu, C., Chen, L., Shen, E., Wang, W., Chu, Q., Wu, D., Wu, S., Xia, C., Zhang, Y.,
Zhou, X., Wang, L., Wu, L., Song, W., Wang, Y., Shu, Q., Aoki, D., Yumoto, E., Yokota, T.,
Miyamoto, K., Okada, K., Kim, D.-S., Cai, D., Zhang, C., Lou, Y., Qian, Q., Yamaguchi, H.,
Yamane, H., Kong, C.-H., Timko, M. P., Bai, L., and Fan, L. (2017) Echinochloa crus-galli
genome analysis provides insight into its adaptation and invasiveness as a weed. Nat. Commun. 8,
1031
160. Zhang, H., Hall, N., Goertzen, L. R., Bi, B., Chen, C. Y., Peatman, E., Lowe, E. K., Patel, J., and
McElroy, J. S. (2019) Development of a goosegrass (Eleusine indica) draft genome and application
to weed science research. Pest Manag. Sci. 75, 2776-2784
161. Jamann, T. M., Balint-Kurti, P. J., and Holland, J. B. (2015) QTL mapping using high-throughput
sequencing. In Plant Functional Genomics: Methods and Protocols (Alonso, J. M., and Stepanova,
A. N. eds.), Springer New York, New York, NY. pp 257-285
162. Jasieniuk, M., Brûlé-Babel, A. L., and Morrison, I. N. (1996) The evolution and genetics of

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


herbicide resistance in weeds. Weed Sci. 44, 176-193
163. Foes, M. J., Liu, L., Tranel, P. J., Wax, L. M., and Stoller, E. W. (1998) A biotype of common
waterhemp (Amaranthus rudis) resistant to triazine and ALS herbicides. Weed Sci. 46, 514-520
164. Huffman, J. L., Riggins, C. W., Steckel, L. E., and Tranel, P. J. (2016) The EPSPS Pro106Ser
substitution solely accounts for glyphosate resistance in a goosegrass (Eleusine indica) population
from Tennessee, United States. J. Integr. Agric. 15, 1304-1312
165. Jugulam, M., and Godar, A. S. (2013) Understanding genetics of herbicide resistance in weeds:
implications for weed management. Adv. Crop Sci. Technol. 1, 1-3
166. Sabba, R. P., Ray, I. M., Lownds, N., and Sterling, T. M. (2003) Inheritance of resistance to
clopyralid and picloram in yellow starthistle (Centaurea solstitialis L.) is controlled by a single
nuclear recessive gene. J. Hered. 94, 523-527
167. Chen, J., Lu, H., Han, H., Yu, Q., Sayer, C., and Powles, S. (2019) Genetic inheritance of
dinitroaniline resistance in an annual ryegrass population. Plant Sci. 283, 189-194
168. Fearon, C. H., Hayward, M. D., and Lawrence, M. J. (1983) Self-incompatibility in ryegrass.
Heredity 50, 35-45
169. Machado, V. S., Bandeen, J. D., Stephenson, G. R., and Lavigne, P. (1978) Uniparental inheritance
of chloroplast atrazine tolerance in Brassica campetris. Can. J. Plant Sci. 58, 977-981
170. Patzoldt, W. L., Dixon, B. S., and Tranel, P. J. (2003) Triazine resistance in Amaranthus
tuberculatus (Moq) Sauer that is not site-of-action mediated. Pest Manag. Sci. 59, 1134-1142
171. Huffman, J., Hausman, N. E., Hager, A. G., Riechers, D. E., and Tranel, P. J. (2015) Genetics and
inheritance of nontarget-site resistances to atrazine and mesotrione in a waterhemp (Amaranthus
tuberculatus) population from Illinois. Weed Sci. 63, 799-809
172. Andersen, R. N., and Gronwald, J. W. (1987) Noncytoplasmic inheritance of atrazine tolerance in
velvetleaf (Abutilon theophrasti). Weed Sci. 35, 496-498
173. Dang, H. T., Malone, J. M., Boutsalis, P., Krishnan, M., Gill, G., and Preston, C. (2018) Reduced
translocation in 2,4-D-resistant oriental mustard populations (Sisymbrium orientale L.) from
Australia. Pest Manag. Sci. 74, 1524-1532
174. Zelaya, I. A., Owen, M. D. K., and VanGessel, M. J. (2004) Inheritance of evolved glyphosate
resistance in Conyza canadensis (L.) Cronq. Theoret. Applied Gen. 110, 58-70
175. Gressel, J. (2009) Evolving understanding of the evolution of herbicide resistance. Pest Manag. Sci.
65, 1164-1173

31
176. Vieira, B. C., Luck, J. D., Amundsen, K. L., Werle, R., Gaines, T. A., and Kruger, G. R. (2020)
Herbicide drift exposure leads to reduced herbicide sensitivity in Amaranthus spp. Scient. Rep. 10,
2146
177. Tehranchian, P., Norsworthy, J. K., Powles, S., Bararpour, M. T., Bagavathiannan, M. V., Barber,
T., and Scott, R. C. (2017) Recurrent sublethal-dose selection for reduced susceptibility of Palmer
amaranth (Amaranthus palmeri) to dicamba. Weed Sci. 65, 206-212
178. Busi, R., Girotto, M., and Powles, S. B. (2016) Response to low-dose herbicide selection in self-
pollinated Avena fatua. Pest Manag. Sci. 72, 603-608
179. Markus, C., Pecinka, A., Karan, R., Barney, J. N., and Merotto Jr, A. (2018) Epigenetic regulation
– contribution to herbicide resistance in weeds? Pest Manag. Sci. 74, 275-281
180. Margaritopoulou, T., Tani, E., Chachalis, D., and Travlos, I. (2018) Involvement of epigenetic
mechanisms in herbicide resistance: The case of Conyza canadensis. Agriculture 8, 17
181. Hawkins, N. J., Bass, C., Dixon, A., and Neve, P. (2019) The evolutionary origins of pesticide
resistance. Biol. Rev. 94, 135-155
182. Kreiner, J. M., Stinchcombe, J. R., and Wright, S. I. (2018) Population genomics of herbicide
resistance: Adaptation via evolutionary rescue. Annu. Rev. Plant Biol. 69, 611-635
183. Yu, Q., Ahmad-Hamdani, M. S., Han, H., Christoffers, M. J., and Powles, S. B. (2013) Herbicide
resistance-endowing ACCase gene mutations in hexaploid wild oat (Avena fatua): insights into

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


resistance evolution in a hexaploid species. Heredity 110, 220-231
184. Hanson, B. D., Shaner, D. L., Westra, P., and Nissen, S. J. (2006) Response of selected hard red
wheat lines to imazamox as affected by number and location of resistance genes, parental
background, and growth habit. Crop Sci. 46, 1206-1211
185. Brunharo, C. A. d. C. G., Morran, S., Martin, K., Moretti, M. L., and Hanson, B. D. (2019) EPSPS
duplication and mutation involved in glyphosate resistance in the allotetraploid weed species Poa
annua L. Pest Manag. Sci. 75, 1663-1670
186. Chen, S., McElroy, J. S., Dane, F., and Goertzen, L. R. (2016) Transcriptome assembly and
comparison of an allotetraploid weed species, annual bluegrass, with its two diploid progenitor
species, Poa supina Schrad and Poa infirma Kunth. The Plant Gen. 9,
doi:10.3835/plantgenome2015.3806.0050
187. Beversdorf, W. D., and Kott, L. S. (1987) Development of triazine resistance in crops by classical
plant breeding. Weed Sci. 35, 9-11
188. Duke, S. O. (2014) Biotechnology: Herbicide-Resistant Crops. In Encyclopedia of Agriculture and
Food Systems (Van Alfen, N. ed.), Elsevier, San Diego. pp 94-116
189. Dreesen, R., Capt, A., Oberdoerfer, R., Coats, I., and Pallett, K. E. (2018) Characterization and
safety evaluation of HPPD W336, a modified 4-hydroxyphenylpyruvate dioxygenase protein, and
the impact of its expression on plant metabolism in herbicide-tolerant MST-FGØ72-2 soybean.
Regulat. Toxicol. Pharmacol. 97, 170-185
190. Hawkes, T. R., Langford, M. P., Viner, R., Blain, R. E., Callaghan, F. M., Mackay, E. A., Hogg, B.
V., Singh, S., and Dale, R. P. (2019) Characterization of 4-hydroxyphenylpyruvate dioxygenases,
inhibition by herbicides and engineering for herbicide tolerance in crops. Pestic. Biochem. Physiol.
156, 9-28
191. Bradshaw, L. D., Padgette, S. R., Kimball, S. L., and Wells, B. H. (1997) Perspectives on
glyphosate resistance. Weed Technol. 11, 189-198
192. Tranel, P. J., and Wright, T. R. (2002) Resistance of weeds to ALS-inhibiting herbicides: what have
we learned? Weed Sci. 50, 700–712
193. Ahrens, W. H., and Stoller, E. W. (1983) Competition, growth rate, and CO2 fixation in triazine-
susceptible and -resistant smooth pigweed (Amaranthus hybridus). Weed Sci. 31, 438-444
194. Conard, S. G., and Radosevich, S. R. (1979) Ecological fitness of Senecio vulgaris and Amaranthus
retroflexus biotypes susceptible or resistant to atrazine. J. Appl. Ecol., 171-177
195. de Montellano, P. R. O. (2015) Substrate oxidation by cytochrome P450 enzymes. In Cytochrome
P450, Springer. pp 111-176

32
196. Kniss, A. R. (2017) Genetically engineered herbicide-resistant crops and herbicide resistant weed
evolution in the United States. Weed Sci. 66, 260-273
197. Powles, S. B., and Holtum, J. (1994) Herbicide Resistance in Plants: Biology and Biochemistry,
CRC Press, Boca Raton, FL

Downloaded from http://www.jbc.org/ by guest on May 19, 2020

33
Table 1. Herbicides mentioned in the text with their molecular targets and mechanisms of resistance.
Only the mechanisms provided in the text are listed. Additional herbicides to which resistance has
evolved are found in ref. (3).

Inhibited process/molecular target Herbicide chemical Mechanism(s) of


class/herbicide1 resistance
Amino acid synthesis
ALS; branched chain amino acids Imidazolinones
imazethapyr TSR and NTSR
imazaquin TSR
Sulfonylureas
benzulfuron TRS and NTSR
chlorimuron TSR and NTSR
chlorsulfuron TSR and NTSR
Pyrimidinyl benzolates
bispyribac TSR and NTSR
Triazolopyrimidine

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


penoxsulam NTSR
EPSPS; shikimate pathway function glyphosate TSR and NTSR
Glutamine synthase glufosinate NTSR
Auxin mimics
F-box proteins Phenoxycarboxylates
2,4-D TSR and NTSR
dicamba TSR and NTSR
quinclorac NTSR
Picolinates
picloram TSR
Carotenoid synthesis
HPPD isoxaflutole NTSR
Triketones
mesotrione NTSR
tembotrione NTSR
Lycopene cyclase amitrole NTSR
Phytoene desaturase Diphenyl heterocycles
fluridone TSR
Phenyl-ethers
beflubutamid no resistance
diflufenican no resistance
picolinafen no resistance
N-Phenyl heterocycles
norflurazon TSR
Deoxy-D-xyulose phosphate synthase clomazone NTSR
Cell division
-Tubulin Dinitroanilines
oryzalin TSR
trifluralin TRS
Chlorophyll synthesis
PPO Diphenyl ethers
lactofen TSR
fomesafen NTSR

34
Fatty acid synthesis
ACCase Aryloxyphenoxypropionates
diclofop-methyl TSR and NTSR
fenoxaprop-P-ethyl TSR and NTSR
clodinofop NTSR
quizalofop TSR
Cyclohexanediones
sethoxydim TSR
tralkoxydim TSR and NTSR
pinoxaden NTSR
Fatty acid thioesterase cinmethylin NTSR
Very long-chain fatty acid synthases thiobencarb NTSR
Chloroacetamides
alachlor NTSR
metolachlor NTSR
Photosynthesis
Photosystem II D1 protein Amides

Downloaded from http://www.jbc.org/ by guest on May 19, 2020


propanil TSR and NTSR
Nitriles
bromoxynil NTSR
Triazines
atrazine TSR and NTSR
Ureas
chlorotoluron NTSR
bentazon NTSR
Photosystem I energy diversion paraquat NTSR

1
the chemical class of herbicides underlined are not provided, as they are the only representatives
of their chemical class

35
Table 2. EPSPS gene duplication and glyphosate resistance level (LD50, dose required to cause 50%
mortality) reported in glyphosate-resistant weed species. Note that some values for LD50 were measured
in different populations and reported in different studies than EPSPS copy number [adapted from (19)].
EPSPS Relative Genomic LD50 (R/S)
Species Copy Number Range
Bassia scoparia 3–9 2-8
Chloris truncata 32-48(53) 2.4-8.7
Hordeum glaucum 9-11(54) 2.8-6.6
(55)
Bromus diandrus 10-36 4.7
Amaranthus spinosus 26-37 5
Amaranthus tuberculatus 2-8 5-19
Lolium multiflorum 15 – 25 12-13
Amaranthus palmeri 2 – 160 15-40
Eleusine indica 28(56) -

Downloaded from http://www.jbc.org/ by guest on May 19, 2020

36
Table 3. Features of an ideal herbicide from a resistance management perspective.

- Inhibits several different targets


- Inhibits several homologous and non-redundant target genes
- Fits tightly and deep into enzyme’s active site
- Only fits into enzyme’s active site
- Does not bind to hydrophobic pockets around the active site in case the natural substrate does not
contain any lipophilic parts
- Builds strong hydrogen bridges
- Small polar molecule with low structural complexity
- Low lipophilicity
- Non-reversable binding, strong enzyme affinity (low KM or P50 values)
- High volume overlap with substrate for good competition

Downloaded from http://www.jbc.org/ by guest on May 19, 2020

37
Downloaded from http://www.jbc.org/ by guest on May 19, 2020
Figure 1. Crystal structure of Arabidopsis ALS. A) view of the homodimer with chain A (gold) and chain
B (slate colors) (Adapted from McCourt et al. (16)). The herbicide imazaquin is located at the entrance of
the channel leading to the catalytic domain of the enzyme. B) Closer view of the interface between the two
ALS monomers with imazaquin positioned at the entrance. The area in red highlights the position of Ser653
imparting resistance to imidazolines but not to sulfonylureas.

38
Downloaded from http://www.jbc.org/ by guest on May 19, 2020
Figure 2. Interaction of glyphosate with EPSP synthase. A) Interaction between shikimate-3-
phosphate (S3P) and glyphosate (GLY) (yellow dotted line) within the catalytic domain of EPSP
synthase. B) Location of the TIPS double mutation in glyphosate-resistant EPSP synthase relative
to the S3P-glyphosate complex. Leucine is shown in pink and serine is shown in slate. C) Mutation-
induced structural changes in EPSP synthase. In the ternary complex, the mutations cause a shift
of the Cα atom of Gly96 toward the phosphonate moiety of glyphosate, seen most drastically in
the TIPS enzyme (pink), thereby narrowing the inhibitor binding site (residue numbers are for E.
coli EPSPS, and is equivalent to Gly101, Thr102, and Pro106 in plants) (Adapted from Funke et
al., (23)).

39
Downloaded from http://www.jbc.org/ by guest on May 19, 2020
Figure 3. View of the catalytic domain of PPO. The porphyrin substrate is centered on top of α-helix 8
(green) and stabilized by several interactions with residues lining the pocket. The yellow spheres represent
the position of Gly210, the deletion of which confers TSR. The two groups of pink spheres represent
Arg128 and Gly399, which can be substituted to impart TSR.

40
Downloaded from http://www.jbc.org/ by guest on May 19, 2020
Figure 4. Summary of the mechanisms of resistance to glyphosate. Observed (regular font) and
putative (italicized font) glyphosate resistance mechanisms. Glyphosate (red circles) crosses the plasma
membrane (blue) to enter the cytoplasm and is transported into the chloroplast (green) to the target-site
enzyme, EPSPS, in herbicide-sensitive plants. Expression of EPSPS variants with 1-3 amino acid
differences can confer resistance to the herbicide (19, 26, 29). Target gene duplication of EPSPS produces
more EPSPS protein that remains sensitive to glyphosate, requiring proportionally more glyphosate to cause
complete inhibition of the extra enzyme (52). Other routes to resistance include sequestration in the vacuole
and enhanced metabolism by aldo-keto reductases (79, 80, 147). Altered import/export from the chloroplast
and/or cytoplasm may also alter glyphosate effectiveness, but these mechanisms remain hypothetical and
have not been documented in weeds. The phoenix phenomenon shown in Figure 6 results in reduced
translocation and involves a currently unknown mechanism triggering cell death upon glyphosate
application.

41
Downloaded from http://www.jbc.org/ by guest on May 19, 2020
Figure 5. Summary of non-target-site resistance (NTSR). Plants can evolve resistance to a herbicide by
reducing its absorption, altering its translocation and/or sequestration, developing a rapid necrosis of the
foliage (phoenix phenomenon), or via degradation of the active ingredient through the phases I, II and III
of metabolism.

42
Downloaded from http://www.jbc.org/ by guest on May 19, 2020
Figure 6. The phoenix phenomenon in plants treated with glyphosate. Both giant ragweed (Ambrosia
trifida) biotypes were sprayed with 0.7 kg ha-1 glyphosate. Glyphosate-susceptible A. trifida at 2 days (A)
and 21 days (B) after glyphosate treatment, behaving like most plants treated with glyphosate. Growth stops
but no injury is observed for the first few days. Glyphosate-resistant A. trifida at 2 days (C) and 21 days
(D) after glyphosate treatment. In plants exhibiting the phoenix phenomenon, older leaves dessicate very
rapidly, trapping most of the glyphosate in dead tissues, and the new shoots emerge undamaged from the
glyphosate treatment. Cover image from (95) with permission from John Wiley.

43
Downloaded from http://www.jbc.org/ by guest on May 19, 2020
Figure 7. Summary of herbicide metabolism in a plant cell. Herbicides are normally taken through the
three phases of metabolism as they are detoxified by plant cells. Typically, phase I introduces small
functional groups on the structure of the active ingredient, phase II attaches a number of water-soluble
metabolites via the action of several types of transferases, and phase III moves the conjugated metabolites
to the vacuole (or the cell wall) for compartmentalization and further degradation. Active transport
sometimes requires ABC transporters (or other transporter types) to move the herbicide metabolites across
membranes.

44
Downloaded from http://www.jbc.org/ by guest on May 19, 2020
Figure 8. Examples of the reactions catalyzed by plant cytochrome P450 monooxygenases involved
in herbicide metabolism. Functional groups either on the substrates of P450 monooxygenases or on the
products of the reactions catalyze by these enzymes are shown in red.

45
Mechanisms of evolved herbicide resistance
Todd A Gaines, Stephen O. Duke, Sarah Morran, Carlos A. G. Rigon, Patrick J Tranel,
Anita Küpper and Franck E Dayan
J. Biol. Chem. published online May 19, 2020

Access the most updated version of this article at doi: 10.1074/jbc.REV120.013572

Alerts:
• When this article is cited
• When a correction for this article is posted

Click here to choose from all of JBC's e-mail alerts

Downloaded from http://www.jbc.org/ by guest on May 19, 2020

You might also like