Open access peer-reviewed chapter

Weed Resistance to Herbicides

Written By

Sava Vrbničanin, Danijela Pavlović and Dragana Božić

Submitted: 31 October 2016 Reviewed: 17 February 2017 Published: 04 October 2017

DOI: 10.5772/67979

From the Edited Volume

Herbicide Resistance in Weeds and Crops

Edited by Zvonko Pacanoski

Chapter metrics overview

2,905 Chapter Downloads

View Full Metrics

Abstract

Unfortunately, herbicide resistance developed shortly after the introduction of the herbicides 2,4‐D in 1957. According the herbicide resistance mechanisms, all processes can be grouped as follows: target‐site resistance, non‐target‐site resistance, cross‐resistance and multiple‐resistance. Target‐site resistance is generally due to a single or several mutations in the gene encoding the herbicide‐target enzyme, which, in turn, decreases the affinity for herbicide binding to that enzyme. Non‐target‐site resistance is caused by mechanisms that reduce the amount of herbicidal active compound before it can attack the plant through the reduced absorption or altered translocation, increased herbicide sequestration or enhanced herbicide metabolism. Cross‐resistance means that a single‐resistance mechanism causes resistance to several herbicides with some mode of action. Multiple‐resistance is a situation where two or more resistance mechanisms are present within the same plant, often due to sequential selection by herbicides with different modes of action. Currently, herbicide resistance has been reported in 478 weed biotypes (252 weed species) in 67 countries. Many of those biotypes are resistant to acetolactate synthase (ALS) inhibitors, PS II inhibitors, ACC‐ase inhibitors and EPSPS inhibitors. Strategy for herbicide‐resistance weed management must involve all the available preventive, cultural, mechanical and chemical measures for effective, safe and cost‐effective weed control.

Keywords

  • weed
  • herbicide
  • resistance
  • management

1. Introduction and general overview of resistance

Since the introduction of 2,4‐D as a first selective herbicide in 1947, herbicides have had a major positive impact on weed management in all over the world. Unfortunately, herbicide resistance developed shortly after the introduction of the herbicides. The phenomenon of resistance can be defined as the decreased response of a species’ population to herbicide [1]. It is also defined as a ‘survival of a segment of the population of a weed species following an herbicide dose lethal to the normal population’ [2]. In addition, resistance can be defined as ‘the inherited ability to survive treatment by a herbicide’ [3], or it is a ‘phenomen which occurs as a result of heritable changes to biochemical processes that enable weed species survival when treated with a herbicide’ [4].

Weed resistance to herbicides is a normal and predictable outcome of natural selection. In that context, rare mutations that confer herbicide resistance exist in wild/weed populations before any herbicide introduction. These mutations increase over time after each herbicide application until they become predominant at what time the weed population is called resistant [5]. The first confirmed herbicide‐resistant weed species was Senecio vulgaris that had developed resistance to PS II inhibitors (atrazine and simazine) after the herbicides had been applied once or twice annually for 10 years [6]. Therefore, about 30‐resistant weed populations have been confirmed within the first decade, mostly in N. America and W. Europe [7]. Some weed species, such as Lolium rigidum, Echinochloa crus‐galli var. crus‐galli, Poa annua, Alopecurus myosuroides, Echinochloa colona, Eleusine indica, Amaranthus sp., etc. have a high affinity to develop resistance especially due to their congenital genetic variability. Additionally, herbicides of different chemical groups and different modes of action (e.g. sulfonylurea and synthetic auxins) can greatly differ in their risk levels for resistance. On the other hand, different chemical groups with the same mode of action such as herbicide inhibitors of acetolactate/acetohydroxyacid synthase (ALS/AHAS) (sulfonylurea, pyrimidinyl(thio)benzoate, sulfonylaminocarbonyl‐triazolinone, imidazolinone) can also be distinguished in their risk level for resistance.

Currently, herbicide resistance has been reported in 478 weed biotypes (252 weed species) in 67 countries. Many of those biotypes are resistant to ALS inhibitors, B/2 (97 dicots + 62 monocots), PS II inhibitors (C1/5 = 51 + 23, C2/7 = 10 + 18, C3/6 = 3 + 1), ACC‐ase inhibitors, A/1 (48 monocots) and EPSPS inhibitors, G (19 + 17). The highest number of confirmed resistant weed species belongs to the families: Poaceae (80 species), Asteraceae (39), Brassicaceae (22), Cyperaceae (12), Amaranthaceae (11), Scrophulariaceae (9), Chenopodiaceae (8), Alismataceae (7), Polygonaceae (7) and Caryophyllaceae (6). According to the number of active ingredients (a.i.), those four sites of action participate in the next relation: 50 a.i. from ALS inhibitors, 24 a.i. from PS II inhibitors, 15 a.i. from ACC‐ase inhibitors and 2 a.i. from EPSPS inhibitors. Atrazine (PS II inhibitors) is an active ingredient, which was confirmed by the greatest number of weed resistant species (66), the second is imazethapyr (44), followed by tribenuron‐methyl (43), imazamox (37), chlorsulfuron (36) metsulfuron‐methyl (35), glyphosate (34), iodosulfuron‐methyl‐sodium (33), fenoxaprop‐P‐ethyl (31), simazine (31), bensulfuron‐methyl (29), thifensulfuron‐methyl (27), fluazifop‐P‐bityl (25), pyrazosulfuron‐ethyl (25), etc. In relation to herbicide‐resistant weeds by county and site of action top 10 counties are the United States, Australia, Canada, France, Brazil, China, Spain, Israel, Japan and Germany [7] (Table 1).

Country Total number of resistant weed species Number of resistant weed species according to the site of action
ALS ACC‐ase PS II EPSPS
USA 156 51 15 26 16
Australia 84 25 12 7 13
Canada 64 25 4 12 5
France 48 16 6 22 2
Brazil 42 19 6 4 8
China 41 14 8 1 2
Spain 37 8 2 18 5
Israel 36 12 6 12 2
Japan 36 21 2 1 3
Germany 32 10 5 13 0

Table 1.

Top 10 countries with the most number of confirmed resistant weed species.

In Serbia, study of weed resistance to herbicides started in the 1990s with resistance of Amaranthus retroflexus and Chenopodium hybridum to PS II inhibitors (atrazine) [810]. Until today, in Serbia, as a small county with less than 3 million ha arable lands, in the last 15 years, eight herbicide‐resistant weed species were confirmed: A. retroflexus, Setaria viridis, C. hybridum and Abutilon theophrasti to PS II inhibitors, as well as A. retroflexus, E. crus‐galli, Datura stramonium, Chenopodium album and Sorghum halepense to ALS inhibitors [1119]. According to the herbicide resistance mechanisms, all processes can be grouped as follows: target‐site resistance, non‐target‐site resistance, cross‐resistance and multiple‐resistance [2022].

Target‐site resistance (TSR) is generally due to a single or several mutations in the gene encoding the herbicide‐target enzyme, which, in turn, decreases the affinity for herbicide binding to that enzyme. Most, but not all cases of resistance to herbicide ALS inhibitors, ACC‐ase, triazine, dinitroaniline etc. are due to modifications of the site of action of the herbicide. In addition, gene overproduction (amplification) is the most recently identified herbicide resistance mechanism, for example, EPSPS gene amplification correlates with glyphosate resistance in Amaranthus palmeri and Kochia scoparia [2325], and causes resistance by increasing the production of the target enzyme, effectively diluting the herbicide in relation to the target site (Figure 1).

Figure 1.

The route of the herbicide after the application, and the possible mechanisms of resistance in plant. After application: (1) herbicide absorption/penetration, (2) translocation, (3) accumulate at the target protein location, and (4) binding to the target protein, (5) disruption of the biosynthesis pathways or cell structures, and/or generation of cytotoxic molecules. NTSR mechanisms: (A) reduction in herbicide penetration, (B) altered translocation of the herbicide away from the target protein, (C) enhanced detoxification of the herbicide, or (D) enhanced neutralization of cytotoxic molecules generated by herbicide action. TSR mechanisms: (E) target protein overproduction, and/or (F) structural mutations that modify the 3D structure and electrochemical properties of the target protein. Structural mutations can have no, moderate or strong negative effects on the stability of herbicide binding to the target protein, which results in (F‐a) no, (F‐b) moderate or (F‐c) marked reduction in herbicide sensitivity at the protein level, respectively; or can (F‐d) increase the stability of herbicide binding to the target protein, which results in an increase in herbicide sensitivity at the protein level (downloaded from Ref. [32]).

Non‐target‐site resistance(NTSR) is caused by mechanisms that reduce the amount of herbicidal active compound before it can attack the plant. Reduced absorption (penetration) or altered translocation, increased herbicide sequestration or enhanced herbicide metabolism (detoxification) can cause resistance due to the restriction of herbicide movement where the herbicide does not reach its site of action in sufficient concentration to cause plant mortality. Active vacuolar or cell walls sequestration can keep the herbicide from the site of action leading to resistance. For example, vacuolar herbicide sequestration correlates with glyphosate resistance in Conyza canadensis, Lolium sp. etc. [26, 27] (Figure 1). Finally, the biochemical reactions that detoxify herbicides can be grouped into four major categories: oxidation, reduction, hydrolysis and conjugation [28].

Cross‐resistance(CR) means that a single‐resistance mechanism causes resistance to several herbicides. CR can be conferred by a single gene or by two or more genes influencing a single mechanism. There are two types of CR: target‐site cross‐resistance (TS‐CR) and non‐target‐site cross‐resistance (NTS‐CR). The most common type of CR is TS‐CR where an altered target site confers resistance to many or all of the herbicides that inhibit the same enzyme, for example, Trp‐574‐Leu amino acid substitution within the ALS gene was found in two populations of Cyperus iria after exposition to bispyribac‐sodium, halosulfuron, imazamox and penoxsulam [29]. On the other hand, NTS‐CR is type of herbicide resistance in which a mechanism other than resistant enzyme target sites is involved (e.g. reduced absorption, translocation, or enhanced herbicide detoxification) [30].

Multiple‐resistance is a situation where two or more resistance mechanisms are present within the same plant, often due to sequential selection by herbicides with different modes of action (e.g. resistance of Lolium sp. populations to glyphosate and ACC‐ase inhibitors, as well as resistance to glyphosate and ALS inhibitors were confirmed by multiple‐resistance [31]).

Advertisement

2. Weed resistance to herbicides photosystem II inhibitors, triazines

The triazine herbicides were discovered in the J.R. Geigy Ltd. laboratories, an international chemical company founded in 1952 and based in Basel, Switzerland [33]. Generally, in the latter half of the twentieth century, triazines have played a significant role in the promotion of the crop production. Atrazine is one of the most used triazine herbicides in agriculture for control of annual monocots (Setaria sp., E. crus‐galli, Digitaria sanguinalis) and dicot weed species (Amaranthus sp., Chenopodium sp., Cirsium arvense, D. stramonium, Sonchus sp., Xanthium strumarium, etc.) and is the most widely used herbicide in maize, orchards and sorghum crops. Triazines specifically inhibit photosystem II (PS II) in plants and in all organisms with oxygen‐evolving photosystems. Generally, they prevent electron transfer by displacing plastoquinone (QB) from a specific binding site on the D1 protein subunit of PS II [34, 35].

The intensive use of triazines resulted in two important cases: appearance of atrazine‐resistant weed species, leading to the increased use of herbicide mixtures or alternative herbicides. The first confirmed atrazine‐resistant weed species [6] helped identify the herbicide‐binding D1 protein in PS II. After the psbA gene was found and sequenced [36], the psbA gene from an atrazine‐tolerant Amaranthus was then sequenced [37]. Based on their findings, the resistance is due to a chloroplast genome mutation of the psbA gene, which codes the D1 protein. The molecular analysis showed that resistance is due to the substitution of serine 264 to glycine (Ser‐264‐Gly) in many weed species [3842]. The substituted urea herbicides, as PS II inhibitors [43] also bind in a niche on the D1 protein, but not at the identical site as the triazines.

A schematic diagram of the folding of the herbicide‐binding site on the D1 protein [44], updated with further amino acids in triazine resistance, is given in Figure 2 [45]. From total of 345 amino acids in the D1 protein, around 60 are part of the herbicide and QB‐binding site. Arrows indicate possible mutations (such as Val‐219, Ala‐251, Phe‐255, Gly‐256, Ser‐264 and Leu‐275) in herbicide‐resistant plants and algae or amino acids tagged by herbicides azidoderivatives (Met‐214 by azidoatrazine) [45].

Figure 2.

The amino acid sequence of the herbicide‐binding D1 protein in the PS II. This work of Michel and Deisenhofer was honoured with the Nobel Prize. (downloaded from Ref. [45]).

Currently, resistance to herbicides that target photosynthesis at PS II has been documented in 74 weed species for triazines (C1/5 group), 28 in C2/7 and only 4 in C3/6 according to the data in the Table 2 [7]. Except the usual amino acid substitution Ser‐264‐Gly in the D1 protein, reduced absorption, translocation and/or detoxification have been reported very often for resistance to triazines in many weed species (Table 3).

Herbicide groups according the site of action Number of weed species Total
1970–1979 1980–1989 1990–1999 2000–2009 2010–2016
Inhibitors PS II (C1/5) 20 32 11 7 4 74
Inhibitors PS II (C2/7) 1 6 13 4 4 28
Inhibitors PS II (C3/6) / / 1 3 / 4
Inhibitors ESPSP enzyme / / 2 18 16 36
Inhibitors AHAS enzyme / 11 62 53 33 159
Inhibitors ACC‐ase / 5 21 14 8 48
Total 21 54 110 99 65

Table 2.

The first confirmed cases of weed species that have developed resistance to different herbicides site of action according to decades.

Weed species Mechanisms
Amino acid substitution Other mechanisms of resistance
Abutilon theophrasti Medic. Detoxification [4549]
Alopecurus myosuroides Huds. Detoxification [5052]
Amaranthus tuberculatus Moq. Sauer. Detoxification [53, 54]
Amaranthus retroflexus L. Detoxification [55]
Amaranthus hybridus L. Ser‐264‐Gly [37]
Amaranthus powellii S. Wats. Detoxification [56]
Bromus tectorum L. Detoxification [57]
Brassica napus L. Ser‐264‐Gly [58]
Chenopodium album L. Ser‐264‐Gly, Ala‐251‐Val, Leu‐218‐Val [41, 59] Detoxification [55, 56, 60, 61]
Echinochloa crus‐galli (L.) P. Beauv. Reduced absorption and translocation, detoxification [60, 62, 63]
Kochia scoparia (L.) Schr. Val‐219‐Ile [64]
Lolium rigidum Gaudin. Detoxification [6567]
Poa annua L. Ser‐264‐Gly, Val‐219‐Ile [40, 6870]
Portulaca oleracea L. Ser‐264‐Thr [71]
Solanum nigrum L. Ser‐264‐Gly [72]
Senecio vulgaris L. Asn‐266‐Thr [73]
Vulpia bromoides (L.) S.F.Gray. Ser‐264‐Gly [42]

Table 3.

Confirmed mechanisms of resistance to herbicide PS II inhibitors in some weed species.

However, diverse chemical groups of herbicides PS II inhibitors (according to HRAC: C1—triazineas, triazinones, triazolinone, pyridazinones, phenyl‐carbametes, uracils; C2—amides, ureas; C3—benzothiadiazinones, nitriles, phenyl pyridazines) bind to overlapping, but not identical sites on the D1 protein [43]. Several different amino acid substitutions that confer resistance to herbicide PS II inhibitors have been identified in or near the QB‐binding niche such as: Ser‐264‐Thr in Portulaca oleracea [71], Ser‐264‐Gly and Val‐219‐Ile in P. annua and K. scoparia [64, 68, 70], Asn‐266‐Thr in S. vulgaris [73] as well as Ser‐264‐Gly, Ala‐251‐Val and Leu‐218‐Val in C. album [41, 59]. In addition, dependence of herbicides, interaction between herbicides, specific amino acid substitution, varying levels of cross or negative cross‐resistance have been reported for different mutations in the D1 protein [64]. Resistance ratios for P. oleracea a Ser‐264‐Thr mutant were 8 and >6 for linuron and diuron, respectively; >800 for atrazine; and >20 for terbacil. Linuron resistant P. oleracea was negatively cross‐resistant to pyridate and bentazon (0.75 and 0.5, respectively) [71].

Advertisement

3. Weed resistance to herbicide ALS inhibitors

Herbicide inhibitors of acetoacetate synthase (ALS) and acetohydroxyacid synthase (AHAS) belong to several chemical classes: sulfonylurea (SU), triazolopyrimidines (TPs), pyrimidinyl(thio) benzoates, sulfonylaminocarbonyltriazolinones, imidazolinones (IMIs). The first commercial SU herbicide was chlorsulfuron, which was introduced by DuPont in 1982 for weed control in small grain crops. The SUs are highly active herbicides, effective at use rates as low as 2 g a.i. ha−1 [74]. Almost simultaneously, researchers at American Cyanamid discovered a structurally distinct family of herbicides, the IMIs, which were also shown to inhibit the ALS enzyme [75]. Since then, three additional chemical classes of ALS inhibitors have been discovered. Those products provide both pre‐emergent and post‐emergent control of many serious monocot and dicot weed species in many crops.

ALS is the first enzyme in the branched‐chain amino acid pathway, which catalyzes the first steps in amino acid biosynthesis such as valine, leucine and isoleucine [76]. The first case of resistance to ALS inhibitors (chlorsulfuron) was reported within 5 years after the introduction of SU herbicides, in 1987 in the United States [77]. Herbicide‐resistant weed evolution is more common for ALS inhibitors compared to herbicides of other groups. Currently, 159 weed species have evolved resistance to ALS‐inhibiting herbicides [7] according to decades that could be seen in Table 2. Weed resistance to ALS inhibitors is due to an alteration of the gene encoding the ALS enzyme. The positions in ALS from various sources (plant, yeast, bacteria) where mutations are known to confer resistance to one or more herbicides distributed across the α, β and γ domain of the protein (Figure 3) [78]. Weed species or genera with high incidence of target‐site ALS resistance include Amaranthus spp., K. scoparia and Papaver rhoeas, among others. Studies have shown that mutations of eight amino acid residues are known to be involved in causing weed resistance: Ala‐122, Pro‐197, Ala‐205, Asp‐376, Arg‐377, Trp‐574, Ser‐653 and Gly‐654 (Table 4). The most different amino acid substitutions in α‐domain at position Pro‐197 have been linked in confirmed weed‐resistant species such as: K. scoparia (Pro‐197‐Ser/Thr/Leu/Ala/Gln/Arg), Descurainia sophia (Pro‐197‐Ser/Thr/Leu/Ala/His/Tyr), P. rhoeas (Pro‐197‐Ser/Thr/Leu/Ala/His/Arg), L. rigidum (Pro‐197‐Ser/Leu/Ala/Gln/Arg), Apera spica‐venti (Pro‐197‐Ser/Thr/Ala/Asn), etc. Also, the substitution of Trp‐574‐Leu confers resistance to several weed species (A. retroflexus, C. iria, D. sophia, C. canadensis, K. scoparia, P. annua etc.) and the levels of resistance are all high against SUs, IMIs and TPs (cross‐resistance) [29, 7983]. Generally, the low number of confirmed weeds resistant to ALS inhibitors is due to altered enzyme activity, reduced translocation and detoxification. Additionally, many weed populations resistant to ALS inhibitors have developed multiple‐resistance to other chemical classes with different modes of action (e.g. auxinic herbicides, EPSPS inhibitors, ACC‐ase inhibitors) [3184, 85].

Figure 3.

ALS mutations conferring herbicide resistance. Arrows point to positions in the sequences of ALS from different sources (plant, yeast, bacteria) where spontaneous or induced mutations result in an herbicide‐insensitive enzyme. Colours designate substitutions occurring in more than one species (downloaded from Ref. [78]).

Weed species Mechanism of resistance
Amino acid substitutions Other mechanisms of resistance
Amaranthus retroflexus L. Ala‐122‐Thr, Pro‐197‐Leu, Ala‐205‐Val, Asp‐376‐Glu, Arg‐377‐His, Trp‐574‐Leu, Ser‐653‐Thr [80]
Amaranthus powellii S. Warts. Ala‐122‐Thr, Asp‐376‐Glu, Arg‐377‐His, Ser‐653‐Thr
Amaranthus hybridus L. Ala‐122‐Thr, Asp‐376‐Glu, Arg‐377‐His, Ser‐653‐Asn Detoxification [86]
Amaranthus blitoides S. Wats Pro‐197‐Ser, Arg‐377‐His
Amaranthus tuberculatus (Moq.) Sauer Arg‐377‐His, Ser‐653‐Asn/Thr Altered enzyme activity, detoxification [87, 88]
Amaranthus palmeri (S.) Warts. Arg‐377‐His, Ser‐653‐Asn Altered enzyme activity [87, 89]
Ambrosia artemisiifolia L. Arg‐377‐His
Ambrosia trifida L. Arg‐377‐His Reduced translocation, detoxification, sequestration [90]
Alopecurus aequalis Sobol. Pro‐197‐Thr, Arg‐377‐His
Alopecurus myosuroides Huds. Pro‐197‐Thr, Arg‐377‐His
Anthemis cotula L. Pro‐197‐Ser/Thr/Leu/Gln
Apera spica‐venti (L.) P.B. Ala‐122‐Val, Pro‐197‐Ser/Thr/Ala/Asn, Arg‐377‐His, Trp‐574‐Leu/Met
Avena fatua L. Ser‐653‐Asn/Thr
Bromus tectorum L. Pro‐197‐Ser
Capsella bursa‐pastoris (L.) Med. Pro‐197‐Ser/Thr/Leu/His
Camelina microcarpa Andrz. Arg‐377‐His
Conyza canadensis (L.) Cronq. Pro‐197‐Ser/Ala, Ala‐205‐Val, Asp‐376‐Glu, Trp‐574‐Leu
Cyperus difformis L. Pro‐197‐Ser/Ala/His
Cyperus iria L., C. escculentus L. Trp‐574‐Leu
Descurainia sophia (L.) Webb. Pro‐197‐Ser/Thr/Leu/Ala/His/Tyr, Asp‐376‐Glu, Arg‐377‐Leu
Echinochloa crus‐galli L. Ala‐122‐Thr/Val, Arg‐377‐His Detoxification [91]
Echinochloa phyllopogon (Stapf) Koss Arg‐377‐His Detoxification [92]
Galium aparine L. Trp‐574‐Gly
Galium spurium L. Asp‐375‐Glu, Trp‐574‐Leu, Ser‐653‐Asn
Helianthus annuus L. Pro‐197‐Leu, Ala‐205‐Val Altered enzyme activity [90]
Kochia scoparia (L.) Schrad. Pro‐197‐Ser/Thr/Leu/Ala/Gln/Arg, Asp‐376‐Glu, Trp‐574‐Leu Altered enzyme activity [93]
Lactuca serriola L. Pro‐197‐Thr/His
Lamium amplexicaule L. Pro‐197‐Arg
Lolium perenne L. Asp‐376‐Glu
Lolium rigidum Gaud. Pro‐197‐Ser/Leu/Ala/Gln/Arg, Trp‐574‐Leu Detoxification altered enzyme activity [94, 95]
Myosoton aquaticum (L.) Moench. Pro‐197‐Ser/Glu Detoxification [96]
Papaver rhoeas L. Pro‐197‐Ser/Thr/Leu/Ala/His/Arg, Trp‐574‐Leu
Poa annua L. Ala‐205‐Phe, Trp‐574‐Leu
Polygonum convolvulus L. Trp‐574‐Leu
Raphanus raphanistrum L. Ala‐122‐Try, Pro‐197‐Ser/Thr/Ala/His, Asp‐376‐Glu, Trp‐574‐Leu
Schoenoplectus juncoides Roxb. Pro‐197‐Ser/Leu/His, Asp‐376‐Glu, Trp‐574‐Leu
Schoenoplectus mucronatus (L.) Palla Pro‐197‐His, Trp‐574‐Leu
Senecio vulgaris L. Pro‐197‐Ser/Leu
Setaria viridis (L.) Beauv. Ser‐653‐Asn/Thr/Ile, Gly‐654‐Asp Altered enzyme activity [97]
Sinapis arvensis L. Pro‐197‐Ser, Asp‐376‐Glu, Trp‐574‐Leu
Sisymbrium orientale Torn. Pro‐197‐Ile, Trp‐574‐Leu
Solanum ptycanthum Dunn Ala‐122‐Thr, Ala‐205‐Val
Sonchus asper (L.) Mill. Pro‐197‐Leu
Sorghum bicolor (L.) Moench Altered enzyme activity [98, 99]
Sorghum halepense (L.) Pers. Trp‐574‐Leu [100]
Stellaria media (L.) Vill. Pro‐197‐Gln, Trp‐574‐Leu Altered enzyme activity [101]
Thlaspi arvense L. Pro‐197‐Leu
Xanthium strumarium L. Ala‐122‐Thr, Ala‐205‐Val, Trp‐574‐Leu Altered enzyme activity [102]

Table 4.

Confirmed mechanisms of resistance to herbicide ALS inhibitors in some weed species.

Amino acid substitution in weed‐resistant species to ALS inhibitors downloaded from HRAC [103].

Advertisement

4. Weed resistance to herbicides ACC‐ase inhibitors

Herbicides acetyl‐CoenzymeA carboxylase (ACC‐ase) inhibitors are aryloxyphenoxy‐propionates (APPs/FOPs), cyclohexanediones (CHDs/DIMs) and phenylpyrazoline. The first herbicide ACC‐ase inhibitors commercialized in 1975 [104]. They are used as foliar herbicides to control monocot weed species in dicot crops and some of them even in cereals or in rice. The mode of action of these herbicides is inhibition of fatty acid biosynthesis through blocking of the acetyl‐CoenzymeA carboxylase [105]. Inhibition of lipid biosynthesis can explain the reduction of growth, increase in permeability of membrane and the ultrastructural effects commonly observed. In living organisms, ACC‐ase exists in two different types: multi‐subunit type and multi‐functional type with 17–51 kDa (prokaryote) and 220–280 kDa (eukaryote) in size, respectively [106]. In dicot plants, the enzyme is structurally distinguished from the enzyme of monocots which contains four regions (biotin carboxylase, biotin carboxy carrier protein, carboxyl‐transferase α and β), while in dicots, they are encoded on separate proteins.

The frequent use of FOPs and DIMs has resulted in the development of resistance to ACC‐ase inhibitors in some monocot species in many countries in the world. Currently, 48 weed species have evolved resistance to these herbicides [7]. By decades, dynamics of the confirmation of the first cases of resistant weed species to the ACC‐ase can be seen in Table 2. Generally, mechanisms of resistance to ACC‐inhibiting herbicides can be divided in two categories: ACC‐related and metabolism‐based. Target‐site resistance to ACC‐ase inhibitors due to the herbicides binding to the carboxyl‐transferase region within the ACC‐ase enzyme results in amino acid substitution in that region (Figure 4) [107, 108]. Weed species or genera with high affinity of target‐site ACC‐ase resistance are A. myosuroides, Avena sp., Bechmannia syzigachne, E. crus‐galli, Lolium sp., etc. Most commonly amino acid substitution such as Ile‐1781‐Leu, Trp‐1999‐Cys, Trp‐2027‐Cys, Ile‐2041‐Asn, Asp‐2078‐Gly, Cys‐2088‐Arg, Gly‐2096‐Ser was confirmed in monocot resistant populations of weed species [109116]. Amino acid substitutions such as Asp‐2078‐Gly and Cys‐2088‐Arg usually provide strong level of resistance to all ACC‐ase (FOPs, DIMs, pinoxaden) inhibitors [81]. Moreover, altered enzyme activity, gene expression and detoxification were very often included in weed resistance to ACC‐ase inhibiting herbicides (Table 5). Also, in some population of weed species such as A. myosuroides [117], E. crus‐galli [118], L. rigidum [111] and Lolium perenne [85], target and non‐target multiple‐resistance, which involves ACC‐ase and ALS inhibitors or ACC‐ase and EPSPS inhibitors, was confirmed.

Figure 4.

Single amino acid mutations in acetyl‐CoA carboxylase in monocot‐resistant weed populations (downloaded from Ref. [108]).

Weed species Mechanism of resistance
Amino acid substitutions Other mechanisms of resistance
Alopecurus aequalis Sobol. Ile‐1781‐Leu [119]
Alopecurus myosuroides Huds. Ile‐1781‐Leu, Trp‐2027‐Cys, Ile‐2041‐Asn, Asp‐2078‐Gly, Gly‐2096‐Ala [110, 112] Detoxification, gene expression [120122]
Avena fatua L. Ile‐1781‐Leu, Trp‐1999‐Cys,Trp‐2027‐Cys, Ile‐2041‐Asn, Asp‐2078‐Gly, Cys‐2088‐Arg, Gly‐2096‐Ser [109, 114] Detoxification [123, 124]
Avena sterilis L. Ile‐1781‐Leu, Trp‐1999‐Cys, Trp‐2027‐Cys, Ile‐2041‐Asn, Asp‐2078‐Gly, Cys‐2088‐Arg [115] Detoxification [123, 125]
Bechmannia syzigachne (Steud.) Fernald Ile‐1781‐Leu, Ile‐2041‐Asn, Asp‐2078‐Gly [126, 127]
Echinochloa crus‐galli (L.) Beauv. Ile‐1781‐Leu [128] Altered enzyme activity, gene expression [129]
Echinochloa colona (L.) Link. altered enzyme activity [130]
Eleusine indica (L.) Gaertn Asp‐2078‐Gly, Thr‐1805‐Ser [131]
Hordeum glaucum (Steud.) Tzvelev Ile‐1781‐Leu, Gly‐2096‐Ala [132]
Hordeum leporinum (Link) Arcang. Ile‐1781‐Leu, Gly‐2096‐Ala [132] Detoxification, altered enzyme activity [133]
Lolium multiflorum Lam. Ile‐1781‐Leu (Ile‐418‐Leu), Cys‐2088‐Arg [134, 135] Detoxification [136]
Lolium perenne L. ssp. multiflorum Lam. Ile‐1781‐Leu, Trp‐2027‐Cys, Ile‐2041‐Asn, Asp‐2078‐Gly [137, 138]
Lolium rigidum Gaud. Ile‐1781‐Leu, Ile‐2041‐Asn, Asp‐2078‐Gly, Cys‐2088‐Arg, Gly‐2096‐Ala, Trp‐2027‐Cys [111, 116] Detoxification [139, 140]
Lolium sp. Ile‐1781‐Leu, Trp‐1999‐Cys, Ile‐2041‐Asn/Val, Asp‐2078‐Gly, Cys‐2088‐Arg, Gly‐2096‐Ala [113]
Pseudosclerochloa kengiana Trp‐1999‐Ser [141]
Setaria viridis L. Beauv. Ile‐1780‐Leu [142] Altered enzyme activity [143]
Sorghum halepense (L.) Pers. Ile‐2041‐Asn [144] Altered enzyme activity [145]

Table 5.

Confirmed mechanisms of resistance to herbicide ACC‐ase inhibitors in some weed species.

Advertisement

5. Weed resistance to herbicide EPSPS inhibitors, glyphosate

Glyphosate was discovered and developed as a non‐selective herbicide by Chemical Company Monsanto in 1974. N‐(phosphonometil) glycine, the active ingredient in glyphosate, is a derivate of the amino acid glycine and phosphonic acid. It’s mode of action in relation to the enzyme EPSPS (5‐enolpyruvylshikimate‐3‐phosphate synthase) preventing the biosynthesis of the aromatic amino acids required for the production of growth regulators, anthocyanins, phenolics and proteins [146]. The site of action is located in the chloroplast and it was confirmed 8 years after glyphosate introduction [146]. Broad weed spectrum (annual and perennial, monocots and dicots), high efficacy, lack of soil activity and low mammalian toxicity are key characteristics that make glyphosate the world’s most widely used herbicide [147]. Because glyphosate is inherently non‐selective, selectivity has often been achieved by placement and timing, for example, as a pre‐plant or pre‐emergence herbicide for the control of weeds in no‐till systems and for turf‐grass renovation [104]. The introduction of genetically modified glyphosate resistant crops in the United States and other parts of the world [148] has led to enormous increase of glyphosate use on arable land (cotton, canola, corn, wheat, sugar beets, potatoes, etc.) as a post‐emergence herbicide.

Glyphosate‐resistant weeds were not found during the first 15 years of glyphosate use (1972–1997). Based on the resistance risk criteria for assessing the risk of developing weed resistance to glyphosate, it was estimated that the glyphosate has low risk for the evolution of weed resistance [149]. However, in the last 19 years (1998–2016), glyphosate resistance in 36 weed species was confirmed and according to the decades, it looks like this: 2 (first decade), 18 (second) and 16 species (the last, third) (Table 2) [7]. Mechanism of glyphosate resistance to weed species includes target‐site mutation, target‐site gene amplification/expression, active vacuole sequestration, limited cellular uptake and a rapid necrosis response [21].

In a number of cases of confirmed weed resistance to glyphosate, the resistance was based on some different mechanisms which include non‐target‐site (limited absorption and translocation, vacuolar sequestration) and target‐site resistance (amino acid substitution, ESPSP gene expression/amplification, altered enzyme activity) (Table 6). Generally, usually confirmed cases of weed resistance to glyphosate were due to reduced absorption and translocation of the herbicide. Further, cDNA sequence analysis of the EPSPS gene indicated that resistance to glyphosate was based on substitution of proline with serine (Pro‐106‐Ser), alanine (Pro‐106‐Ala), threonine (Pro‐106‐Thr), or leucine (Pro‐106‐Leu)) at the position 106 of the EPSPS protein in many weed species (Amaranthus tuberculatus, E. indica, E. colona, L. rigidum, Lolium multiflorum, P. annua).

Weed species Mechanisms
Amino acid substitution Other mechanisms of resistance
Amaranthus tuberculatus Moq. Sauer. Pro‐106‐Ser [150, 151] Gene expression [150, 151]
Amaranthus palmeri S. Wats. Gene expression [23, 24, 152]
Abutilon theophrasti Medic. Reduced absorption and translocation [153, 154]
Cirsium arvense (L.) Scop. Reduced absorption and translocation [155]
Conyza canadensis (L.) Cronq. Reduced absorption and translocation, detoxification, vacuole sequestration [26, 156160]
Conyza bonariensis (L.) Cronq. Reduced absorption and translocation [161]
Chenopodium album L. Reduced absorption and translocation [162]
Cyperus esculentus L. Reduced absorption and translocation [163]
Cyperus rotundus L. Detoxification [164]
Eleusina indica (L.) Gaertn. Pro‐106‐Ser/Thr [165167]
Echinochloa colona (L.) Link. Pro‐106‐Ser [168]
Kochia scoparia (L.) Schr. Detoxification, gene expression [25]
Lolium rigidum Gaudin. Pro‐106‐Ser/Ala/Thr/Leu [169, 170] Reduced absorption and translocation, altered enzyme activity [171177]
Lolium multiflorum Lam. Pro‐106‐Ser [178] Reduced absorption and translocation [178, 179]
Lolium perenne L. ssp. multiflorum Lam. Gene expression [180]
Poa annua L. Pro‐106‐Ala [181] Reduced translocation [181]
Sorghum halepense (L.) Pers. Reduced absorption and translocation [182]

Table 6.

Confirmed mechanisms of resistance to EPSPS inhibitor in some weed species.

Advertisement

6. Management strategies for herbicide‐resistant weeds

Strategy for herbicide‐resistance weed management must involve all the available preventive, cultural, mechanical and chemical measures for effective, safe and cost‐effective weed control [183]: (a) survey of present weed flora; (b) preventing weed seed production and reduction of weed seed in the soil seed‐bank; (c) prevention of the movement of seeds and vegetative propagules from field to field or from field margins (or lost field) to field; (d) keep arable and non‐arable land as weed free as possible; (e) sowing pure crop seeds; (f) growing competitive crops that can suppress weeds; (g) destruction of weed seeds in post‐harvest materials (e.g. Integrated Harrington Seed Destructor); (h) use mechanical and physical measures where appropriate; (i) using herbicides with different modes of action, tank mixtures and sequential applications; (j) use of recommended herbicide rate for certain number of weed populations; (k) adopting crop rotations that allow use of herbicides of alternative mode of action; (l) intensify research and professional communication and grower education programs and (m) publish guidelines for managing anti‐resistant strategy.

The state government sectors, universities and research institutes, technology development centres, farmers and other relevant stakeholders were called to proactively address emerging weed resistance problems and to develop cost‐effective resistance‐management strategy and practices that support effective weed control.

Advertisement

Acknowledgments

This work was supported by the Ministry of Education, Science and Technological Development of Republic of Serbia.

References

  1. 1. LeBaron HM, Gressel J. Herbicide Resistance in Plants. Wiley Interscience NYP; 1982
  2. 2. Penner D. Herbicide action and metabolism. In: Turf Weeds and their Control. Madison, WI: American Society of Agronomy and Crop Science Society of America; 1994. pp. 37–70
  3. 3. Gressel J. Why get resistance? It can be prevented or delayed. In: Caseley JC, Cussans GW, Atkin RK, editors. Herbicide Resistance in Weeds and Crops. Butterworth‐Heinemann Ltd; 1991. pp. 1–27
  4. 4. Preston C, Mallory‐Smith CA. Biochemical mechanisms, inheritance, and molecular genetics of herbicide resistance in weeds. In: Powles SB, Shanel DL, editors. Herbicide Resistance in World Grains. Boca Raton, FL: CRC Press; 2001. pp. 23–61
  5. 5. Heap I. Herbicide resistant weeds. In: Primentel D, Peshin R, editors. Integrated Pest Management. Pesticide Problems. Vol. 3. Springer; 2014. pp. 281–303
  6. 6. Ryan GF. Resistance of common groundsel to simazine and atrazine. Weed Science. 1970;18:614–616
  7. 7. Heap I. The International Survey of Herbicide Resistant Weeds [Internet]. Available from: http://www.weedscience.org [Accessed: 3 January 2017]
  8. 8. Janjić V, Stanković‐Kalezić R, Radivojević Lj, Marisavljević D, Jovanović Lj, Ajder S. Resistance of Amaranthus retroflexus L. and Chenopodium alabum L. to atrazine (In Serbian). Acta biologica Iugoslavica, Series G: Acta Herbologica. 1994;3(1):63–73
  9. 9. Janjić V, Stanković‐Kalezić R, Radivojević Lj, Ajder S, Marisavljević D, Jovanović LJ. Distribution of atrazine resistant population of Amaranthus retroflexus L. (In Serbian). Acta Biologica Iugoslavica, Series G: Acta Herbologica. 1994;3(1):23–31
  10. 10. Janjić V, Stanković‐Kalezić R, Radivojević Lj, Marisavljević D, Ajder S, Jovanović LJ. Detection some triazine resistant weeds using chlorophyll fluorescence. In: Proceedings of the International Symposium on Weed and Crop Resistance to Herbicides. Cordoba, Spain: 1995. p. 134
  11. 11. Malidža G, Rajković M, Vrbničanin S, Božić D. Cross‐resistance of Sorghum halepense to ALS inhibitors in Serbia and implications for resistance management. Book of Abstracts of the VII Congress on Plant Protection: Integrated Plant Protection Knowledge—Based Step Towards Sustainable Agriculture, Forestry and Landscape Architecture. 2014:143–144
  12. 12. Malidža G, Rajković M, Vrbničanin S, Božić D. Identification and distribution of ALS resistant Sorghum halepense populations in Serbia. Book of Abstracts of the 17th European Weed Research Society Symposium; 2015:115
  13. 13. Pavlović D, Vrbničanin S, Božić D, Fischer A. Morphophysiological traits and triazine sensitivity in Chenopodium album L. Pest Management Science. 2008;64(2):101–107
  14. 14. Božić D, Vrbničanin S, Barać M, Stefanović L. Determination of Jonsongrass (Sorghum halepense (L.) Pers.) level of sensitivity to nicosulfuron. Maydica. 2007;52(3):271–277
  15. 15. Pavlovic D, Vrbnicanin S, Reinhardt C. Crop response to glyphosate trimesium sulphosate. Notulae Botanicae Horti Agrobotanici Cluj‐Napoca. 2013;41(2):582–589
  16. 16. Bozic D, Barac M, Saric‐Krsmanovic M, Pavlovic D, Ritz C, Vrbnicanin V. Common cocklebur (Xanthium strumarium) response to nicosulfuron. Notulae Botanicae Horti Agrobotanici Cluj‐Napoca. 2015;43(1):186–191
  17. 17. Pavlović D, Vrbničanin S, Reinhard C, Marisavljević D. Morpho‐anatomical response of glyphosate‐resistant and –susceptible maize to glyphosate trimesium. In: Proceedings of the International Symposium on Current Trends in Plant Protection. Belgrade, Serbia; 2012. pp. 188–191
  18. 18. Pavlović D, Vrbničanin S, Elezović I, Marisavljević D, Božić D. Physiological parameters as a basis for differentiating between resistant and susceptible populations of Chenopodium album L. to atrazine (in Serbian). Pesticidi i fitomedicina. 2005;20(4):241–247
  19. 19. Bozic D, Pavlovic D, Bregola V, Di Loreto A, Bosi S, Vrbnicanin S. Gene flow from Herbicide‐Resistant sunflower hybrids to weedy sunflower. Journal of Plant Diseases and Protection. 2015;122(4):183–188
  20. 20. Heap I, LeBaron H. Introduction and overview of resistance. In: Powles SB, Shanel DL, editors. Herbicide Resistance in World Grains. Boca Raton, Florida: CRC Press; 2001. pp. 1–22
  21. 21. Sammons RD, Gaines TA. Glyphosate resistance: State of knowledge. Pest Management Science. 2014;70:1367–1377
  22. 22. Yu Q, Powles S. Metabolism‐based herbicide resistance and Cross‐Resistance in crop weeds: A threat to herbicide sustainability and global crop production. Plant Physiology. 2014;166:1106–1118
  23. 23. Gaines TA, Zhang W, Wang D, Bukun B, Chisholm ST, Shaner DL, Nissen SJ, Patzoldt WL, Tranel PJ, Culpepper AS, Grey TL, Webster TM, Vencill WK, Sammons RD, Jiang J, Preston C, Leach JE, Westra P. Gene amplification confers glyphosate resistance in Amaranthus palmeri. Proceedings of the National Academy of Sciences. 2010;107(3):1029–1034
  24. 24. Vila‐Aiub MM, Goh SS, Gaines TA, Han H, Busi R, Yu Q, Powles SB. No fitness cost of glyphosate resistance endowed by massive EPSPS gene amplification in Amaranthus palmeri. Planta. 2014;239(4):793–801
  25. 25. Godar AS, Stahlman PW, Jugulam M, Dille JA. Glyphosate‐Resistant Kochia (Kochia scoparia) in Kansas: EPSPS gene copy number in relation to resistance levels. Weed Science. 2015;63:587–595
  26. 26. Ge X, D’Avignon DA, Ackermana JJH, Sammons RD. Rapid vacuolar sequestration: The horseweed glyphosate resistance mechanism. Pest Management Science. 2010;66:345–348
  27. 27. Ge X, D’Avignon DA, Ackerman JJH, Collavo A, Sattin M, Ostrander EL. Hall EL, Sammons RD, Preston C. Vacuolar glyphosate‐sequestration correlates with glyphosate resistance in ryegrass (Lolium spp.) from Australia, South America, and Europe: A 31P NMR investigation. Journal of Agriculture and Food Chemistry. 2012;60:1243–1250
  28. 28. Dekker JH, Duke SO. Herbicide‐Resistant Field Crops [Internet]. Iowa State University, Digital Repository. 1995. Available from: http://lib.dr.iastate.edu/agron_pubs (Accessed: 5 December 2016)
  29. 29. Riar DS, Tehranchian P, Norsworthy JK, Nandula V, McElroy S, Srivastava V, Chen S, Bond JA, Scott RC. Acetolactate synthase‐inhibiting, herbicide‐resistant rice flatsedge (Cyperus iria): Cross‐resistance and molecular mechanism of resistance. Weed Science. 2015;63:748–757
  30. 30. LeBaron HM, Utah H. Distribution and management of Triazine‐Resistant weeds. In: LeBaron HM, McFarland EJ, Burnside CO, editors. The Triazine Herbicides 50 years Revolutionizing Agriculture. Elsevier; 2008. pp. 119–133
  31. 31. Collavo A, Sattin M. First glyphosate‐resistant Lolium spp. biotypes found in a European annual arable cropping system also affected by ACCase and ALS resistance. Weed Science. 2014;54:325–334
  32. 32. Délye C, Jasieniuk M, Le Corre V. Deciphering the evolution of herbicide resistance in weeds. Trends in Genetics. 2013;29(11):649–658
  33. 33. Müller G. History of the discovery and development of triazine herbicides 13. In: LeBaron HM, McFarland EJ, Burnside CO, editors. The Triazine Herbicides 50 years Revolutionizing Agriculture. Elsevier; 2008. pp. 13–31
  34. 34. Good NE. Inhibitors of the hill reaction. Plant Physiology. 1961;36:788–803
  35. 35. Trebst A. The mode of action of triazine herbicides in plants. In: LeBaron HM, McFarland EJ, Burnside CO, editors. The Triazine Herbicides 50 years Revolutionizing Agriculture. Elsevier; 2008. pp. 101–111
  36. 36. Zurawski G, Bohnert HJ, Whitfield PR, Bottomley W. Nucleotide sequence of the gene for the Mr 32 000 thylakoid membrane protein from Spinacia oleraceaand Nicotiana debneyipredicts a totally conserved primary translation product of Mr 38,950. Proceedings of the National Academy of Sciences United State of America. 1982;79:7699–7703
  37. 37. Hirschberg J, McIntosh L. Molecular basis of herbicide resistance in Amaranthus hybridus. Science. 1983;222(4630):1346–1349
  38. 38. Morden CW, Golden SS. A psbA genes indicate common ancestry of prochlorophytes and chloroplasts. Nature. 1989;337:382–385
  39. 39. Van Oorschot JC. Chloroplastic resistance of weeds to triazines in Europe. In: Caseley JC, Cussans WG, Atkin KR, editors. Herbicide Resistance in Weeds and Crops. Boston: Butterworth‐Heinemann; 1991. pp. 87–102
  40. 40. Kelly ST, Coats GE, Luthe DS. Mode of resistance of Triazine‐Resistant annual bluegrass (Poa annua). Weed Technology. 1999;13:747–752
  41. 41. Mechant E, De Marez T, Hermann O, Bulcke R. Resistance of Chenopodium album to photosystem II‐inhibitors. Communications in Agricultural and Applied Biological Sciences. 2008;73(4):913–917
  42. 42. Ashworth MB, Han H, Knell G, Powles SB. Identification of triazine‐resistant Vulpia bromoides. Weed Technology. 2016;30:456–463
  43. 43. Gronwald JW. Resistance to photosystem II inhibiting herbicides. In: Powles SB, Holtum JAM, editors. Herbicide Resistance in Plants: Biology and Biochemistry. Boca Raton, Florida: Lewis Publishers; 1994. pp. 27–60
  44. 44. Trebst A. The three dimensional structure of the herbicide binding niche on the reaction center polypeptide of photosystem II. Z. Naturforschg. 1987;42c:742–750
  45. 45. Gronwald JW, Anderson RN, Yee C. Atrazine resistance in velvetleaf (Abutilon theophrasti) due to enhanced atrazine detoxification. Pesticide Biochemistry and Physiology. 1989;34:149–163
  46. 46. Stowe AE, DiTomaso JM. Evidence for increased herbicide detoxification in triazine tolerant velvetleaf. In Proceedings of the 43rd Annual Meeting of the NorthEastern; Weed Science Society; 1989. pp. 31–35
  47. 47. Anderson MP, Gronwald JW. Atrazine resistance in a velvetleaf (Abutilon theophrasti) biotype due to enhanced glutathione S‐Transferase activity. Plant Physiology. 1991;96: 104–109
  48. 48. Plaisance KL, Anderson MP, Gronwald JW. Glutathione S‐transferase isoforms in atrazine‐resistant and susceptible biotypes of Abutilon theophrasti. In: Proceedings of the International Symposium on Weed and Crop Resistance to Herbicides, Cordoba Spain; 1995. p. 99
  49. 49. Pavlovic D, Vrbnicanin S, Bozic D, Simoncic A. Abutilon theophrasti Medic. Population responses to atrazine. Journal Central European Agriculture. 2007;8(4):435–442
  50. 50. Hall LM, Moss SR, Powles SB. Mechanism of resistance to chlorotoluron in two biotypes of the grass weed Alopecurus myosuroides. Pesticide Biochemistry and Physiology. 1995;53:180–192
  51. 51. Reade HPJ, Belfield LJ, Cobb HA. Rapid tests for herbicide resistance in black‐grass based on elevated gluthatione S‐transferase activity and abundance. Proceedings of the Brighton crop Protection Conference‐Weeds. 1999;184–190
  52. 52. Milner LJ, Reade PHJ, Cobb HA. An investigation of gluthatione S‐transferase activity in Alopecurus myosuroides Huds. (black‐grass) in the field. Proceedings of the Brighton crop Protection Conference‐Weeds. 1999;173–178
  53. 53. Patzoldt WL, Dixon BS, Tranel PJ. Triazine resistance in Amaranthus tuberculatus (Moq) Sauer. that is not site‐of‐action mediated. Pest Management Science. 2003;59:1134–1142
  54. 54. Huffman J, Hausman EN, Hager GA, Riechers ED, Tranel JP. Genetics and inheritance of nontarget‐site resistances to atrazine and mesotrione in a Waterhemp (Amaranthus tuberculatus) population from Illinois. Weed Science. 2015;63:799–809
  55. 55. Pavlovic D, Vrbnicanin S, Elezovic I, Jovanovic LJ, Marisavljevic D. Alterations in amount of chlorophyll as indicator of resistance for Chenopodium album L. and Amaranthus retroflexus L. to atrazine. Journal of Plant Disease and Protection. 2006;131–138
  56. 56. Khan US, Warwick IS, Marriage BP. Atrazin metabolism in resistant and susceptible biotypes of Chenopodium album L., Chenopodium strictum Roth. and Amaranthus powelli S. wats. Weed Research. 1985;25:33–37
  57. 57. Menendez J, Bastida F, De Prado R. Resistance to chlortoluron in a downy brome (Bromus tectorum) biotype. Weed Science. 2006;54:237–245
  58. 58. Sundby C, Chow WS, Anderson JM. Effects on photosystem II function, photoinhibition and plant performance of the spontaneous mutation of serine 264 in the photosystem II reaction center D1 protein in triazine‐resistant Brassica napus L. Plant Physiology. 1993;103(1):105–113
  59. 59. Thiel H, Varrelmann M. Identification of a new PSII target site psbA mutation leading to D1 amino acid Leu218Val exchange in the Chenopodium album D1 protein and comparison to cross‐resistance profiles of known modifications at positions 251 and 264. Pest Management Science. 2014;70(2):278–285
  60. 60. Giebel J, Stachecki S, Praczyk T. The level of polyamines as an indicator of resistance or susceptibility of Chenopodium album to atrazine. Proceedings of the Brighton Crop Protection Conference‐Weeds. 1999;163–166
  61. 61. Giebel J, Praczyk T. Modification of IAA‐oxidase activity by factor present in leaves of Chenopodium album and Echinochloa crus‐galli in relation to susceptible‐resistant plant response to triazine. Journal of Plant Protection Research. 2002;42(3):271–278
  62. 62. Carey III FV, Duke OS, Hoaland ER, Talbert ER. Resistance mechanism of propanil resistant Barnyard grass I: Adsorption, translocation and site of action studies. Pesticide Biochemistry and Physiology. 1995;52:182–189
  63. 63. Hoagland ER, Norsworty KJ, Carey F, Talbert ER. Metabolically based resistance to the herbicide propanil in Echinochloa species. Weed Science. 2004;52:475–486
  64. 64. Mengistu LW, Christoffers MJ, Lym RG. A psbA mutation in Kochia scoparia (L) Schrad from railroad rights‐of‐way with resistance to diuron, tebuthiuron and metribuzin. Pest Management Science. 2005;61:1035–1042
  65. 65. Burnet MWM, Loveys BR, Holtum JAM, Powles SB. A mechanism of chlorotoluron resistance in Lolium rigidum. Planta 1993;190:182–189
  66. 66. Burnet MWM, Loveys BR, Holtum JAM, Powles SB. Increased detoxification is a mechanism of simazine resistance in Lolium rigidum. Pesticide Biochemistry and Physiology. 1993;46:207–218
  67. 67. De Prado R, De Prado JL, Menendez J. Resistance to substituted urea herbicides in Lolium rigidum biotypes. Pesticide Biochemistry and Physiology. 1997;57:126–136
  68. 68. Mengistu LW, Mueller, GW, Liston A, Barker RE. A psbA Mutation (valine219 to isoleucine) in Poa annua resistant to metribuzin and diuron. Pest Management Science. 2000;56:209–217
  69. 69. Perry DH, McElroy JS, Dane F, van Santen E, Walker RH. Triazine‐resistant annual bluegrass (Poa annua) populations with Ser 264 mutation are resistant to amicarbazone. Weed Science. 2012;60:355–359
  70. 70. Svyantek AW, Aldahir P, Chen S, Flessner ML, McCullough PE, Sidhu SS, McElroy JS. Target and nontarget resistance mechanisms induce annual bluegrass (Poa annua) resistance to atrazine, amicarbazone, and diuron. Weed Technology. 2016;30(3):773–782
  71. 71. Masabni JG, Zandstra BH. A serine‐to‐threonine mutation in linuron‐resistant Portulaca oleracea. Weed Science. 1999;47:393–400
  72. 72. Stankiewicz M, Gadamski G, Gawronski WS. Genetic variation and phylogenetic relationships of triazine‐resistant and triazine‐susceptible biotypes of Solanum nigrum—analysis using RAPD markers. Weed Research. 2001;41:287–300
  73. 73. Park KW, Mallory‐Smith CA. A psbA mutation (Asn266 to Thr) in Senecio vulgaris L. confers resistance to several PS II‐inhibiting herbicides. Pest Management Science. 2006;62(9):880–885
  74. 74. Beyer EM, Duffy MJ, Hay JV, Schlueter DD. Sulfonylureas. In: Kearney PC, Kaufman DD, editors. Herbicides: Chemistry, Degradation and Mode of Action. New York: Dekker; 1988. pp. 117–189
  75. 75. Shaner DL, Anderson PC, Stidham MA. Imidazolinones: Potent inhibitors of acetohydroxyacid synthase. Plant Physiology. 1984;76:545–546
  76. 76. Devine MD, Duke SO, Fedtke C. Physiology of Herbicide Action. Englewood Cliffs, NJ: Prentice Hall; 1993
  77. 77. Shaner DL. Herbicide resistance: Where are we? How did we get here? Where are we going? Weed Technology. 1992;9:850–856
  78. 78. Pang SS, Duggleby RG, Schowen RL, Guddat LW. The crystal structure of Klebsiella pneumoniae acetolactate synthase with enzyme‐bound cofactor and with an unusual intermediate. The Journal of Biological Chemistry. 2004;279:2242–2253
  79. 79. Thomson MK. Acetohydroxyacid synthase inhibitors (AHAS/ALS). In: Krämer W, Schirmer U, editors. Modern Crop Protection Compounds. Wiley‐VCH; 2007. pp. 27–45
  80. 80. Scarabel L, Varotto S, Sattin M. European biotype of Amaranthus retroflexus cross‐resistant to ALS inhibitors and response to alternative herbicides. Weed Research. 2007;47:527–533
  81. 81. Beckie HJ, Tardif FJ. Herbicide cross resistance in weeds. Crop Protection. 2012;35:15–28
  82. 82. McElroy JS, Flessner ML, Wang Z, Dane F, Walker RH, Wehtje GR. A Trp574 to Leu amino acid substitution in the ALS gene of annual bluegrass (Poa annua) is associated with resistance to ALS‐Inhibiting herbicides. Weed Science. 2013;61:21–25
  83. 83. Deng W, Yanga Q, Zhang Y, Jiao H, Mei Y, Li X, Zheng M. Cross‐resistance patterns to acetolactate synthase (ALS)‐inhibiting herbicides of flixweed (Descurainia sophia L.) conferred by different combinations of ALS isozymes with a Pro‐197‐Thr mutation or a novel Trp‐574‐Leu mutation. Pesticide Biochemistry and Physiology. 2017;136:41–45
  84. 84. Preston C, Dolman FC, Boutsalis P. Multiple resistance to acetohydroxyacid synthase–inhibiting and auxinic herbicides in a population of oriental mustard (Sisymbrium orientale). Weed Science. 2013;61:185–192
  85. 85. Liu M, Hulting AG, Mallory‐Smith C. Characterization of multiple herbicide‐resistant Italian ryegrass (Lolium perenne ssp. multiflorum) populations from winter wheat fields in Oregon. Weed Science. 2016;64:331–338
  86. 86. Manley BS, Wilson HP, Hines TH. Characterization of imidazolinone‐resistant smooth pigweed (Amaranthus hybridus). Weed Technology. 1998;12(4):575–584
  87. 87. Sprague CL, Stoller EW, Wax LM, Horak MJ. Palmer amaranth (Amaranthus palmeri) and common waterhemp (Amaranthus rudis) resistance to selected ALS‐inhibiting herbicides. Weed Science. 1997;45(2):192–197
  88. 88. Guo J, Riggins CW, Hausman NE, Hager AG, Riechers DE, Davis AS, Tranel PJ. Nontarget‐Site resistance to ALS inhibitors in waterhemp (Amaranthus tuberculatus). Weed Science. 2015;63:399–407
  89. 89. Burgos NR, Kuk YI, Talbert RE. Amaranthus palmeri resistance and differential tolerance to Amaranthus palmeri and Amaranthus hybridus to ALS‐inhibitor herbicides. Pest Management Science. 2001;57:449–457
  90. 90. Zelaya IA, Owen MDK. Evolved resistance to acetolactate synthase‐inhibiting herbicides in common sunflower (Helianthus annuus), giant ragweed (Ambrosia trifida) and shattercane (Sorghum bicolor) in Iowa. Weed Science. 2004;52(4):538–548
  91. 91. Riar DS, Norsworthy JK, Bond JA, Bararpour MT, Wilson MJ, Scott RC. Resistance of Echinochloa crus‐galli populations to acetolactate synthase‐inhibiting herbicides. International Journal of Agronomy. 2012;893–953
  92. 92. Fischer AJ, Bayer DE, Carriere MD, Ateh CM, Yim KO. Mechanisms of resistance to Bispyribac‐Sodium in an Echinochloa phyllopogon accession. Pesticide Biochemistry and Physiology. 2000;68:156–165
  93. 93. Saari L, Cotterman J, Primiani M. Mechanism of sulfonylurea herbicide resistance in the broadleaf weed, Kochia scoparia. Plant Physiology. 1990;93:55–61
  94. 94. Shimabukuro RH, Hoffer BL. Metabolism of diclofop‐methyl in susceptible and resistant biotypes of Lolium rigidum. Pesticide Biochemistry and Physiology. 1991;39:251–260
  95. 95. Christopher JT, Powles SB, Joseph AM, Holtum JAM. Resistance to acetolactate Synthase‐Inhibiting herbicides in annual ryegrass (Lolium rigidum) involves at least two mechanisms. Plant Physiology. 1992;100:1909–1913
  96. 96. Liu W, Wu C, Guo W, Du L, Yuan G, Wang J. Resistance mechanisms to an acetolactate synthase (ALS) inhibitor in water starwort (Myosoton aquaticum) populations from China. Weed Science. 2015;63:770–780
  97. 97. Volenberg DS, Stoltenberg DE, Boerboom CM. Green foxtail (Setaria viridis) resistance to acetolactate synthase inhibitors. Phytoprotection. 2002;83:99–109
  98. 98. Anderson DD, Nissen SJ, Martin AR, Roeth FW. Mechanism of primisulfuron resistance in a shattercane (Sorghum bicolor) biotype. Weed Science. 1998;46(2):158–162
  99. 99. Lee DC, Martin RA, Roeth WF, Johnson EB, Lee JD. Comparison of ALS inhibitors resistance and allelic interactions in shattercane accessions. Weed Science. 1999;47(3):275–281
  100. 100. Hernández MJ, León R, Fischer AJ, Gebauer M, Galdames R, Figueroa R. Target‐site resistance to nicosulfuron in johnsongrass (Sorghum halepense) from Chilean corn fields. Weed Science. 2015;63:631–640
  101. 101. Devine MD, Marles MAS, Hall LM. Inhibition of acetolactate synthase in susceptible and resistant biotypes of Stellaria media. Pesticide Science. 1991;31:273–280
  102. 102. Lee JM, Owen MDK. Comparison of acetolactate synthase enzyme inhibition among resistant and susceptible Xanthium strumarium biotypes. Weed Science. 2000;48(3):286–290
  103. 103. Tranel PJ, Wright TR, Heap IM. Mutations in Herbicide‐resistant Weeds to ALS Inhibitors [Internet]. Available from: http://www.weedscience.com [Accessed: 10 January 2017]
  104. 104. Monaco TJ, Weller SC, Ashton FM. Weed Science–Principles and Practices. 4th ed. John Wiley and Sons, Inc.; 2002
  105. 105. Gronwald JW, Eberlein CV, Betts KJ, Baerg RJ, Ehlke NJ, Wyse DL. Mechanism of diclofop resistance in an Italian ryegrass (Lolium multiflorum Lam.) biotype. Pesticide Biochemistry and Physiology. 1992;44:126–139
  106. 106. Kukorelli G, Reisinger P, Pinke G. ACCase inhibitor herbicides—selectivity, weed resistance and fitness cost: A review. International Journal of Pest Management. 2013;59(3):165–173
  107. 107. Zhang H, Tweel B, Tong L. Molecular basis for the inhibition of the carboxyltransferase domain of acetylcoenzyme‐A carboxylase by haloxyfop and dielofop. Proceedings of the National Academy of Sciences United State of America. 2004;101:5910–5915
  108. 108. Jang SR, Marjanovic J, Gornicki P. Resistance to herbicides caused by single amino acid mutations in acetyl‐CoA carboxylase in resistant populations of grassy weeds. New Phytologist. 2013;197:1110–1116
  109. 109. Christoffers MJ, Berg ML, Messersmith CG. An isoleucine to leucine mutation in acetyl‐CoA carboxylase confers herbicide resistance in wild oat. Genome. 2002;45:1049–1056
  110. 110. Moss SR, Cocker KM, Brown AC, Hall L, Field LM. Characterisation of target‐site resistance to ACCase‐inhibiting herbicides in the weed Alopecurus myosuroides (black‐grass). Pest Management Science. 2003;59:190–201
  111. 111. Yu Q, Collavo A, Zheng MQ, Owen M, Sattin M, Powles SB. Diversity of Acetyl‐Coenzyme a carboxylase mutations in resistant Lolium populations: Evaluation using clethodim. Plant Physiology. 2007;145:547–558
  112. 112. Délye C, Matéjicek A, Michel S. Cross‐resistance patterns to ACCase‐inhibiting herbicides conferred by mutant ACCase isoforms in Alopecurus myosuroides Huds. (black‐grass), re‐examined at the recommended herbicide field rate. Pest Management Science. 2008;64:1179–1186
  113. 113. Scarabel L, Panozzo S, Varotto S, Sattin M. Allelic variation of the ACCase gene and response to ACCase‐inhibiting herbicides in pinoxaden‐resistant Lolium spp. Pest Management Science. 2011;67:932–941
  114. 114. Beckie HJ, Warwick SI, Sauder CA. Basis for herbicide resistance in Canadian populations of wild oat (Avena fatua). Weed Science. 2012;60(1):10–18
  115. 115. Papapanagiotou AP, Paresidou MI, Kaloumenos NS. Eleftherohorinos IG. ACCase mutations in Avena sterilis populations and their impact on plant fitness. Pesticide Biochemistry and Physiology. 2015;123:40–48
  116. 116. Saini RK, Malone J, Preston C, Gill G. Target enzyme‐based resistance to clethodim in Lolium rigidum populations in Australia. Weed Science. 2015;63:946–953
  117. 117. Keshtkar E, Mathiassen SK, Moss SR, Kudsk P. Resistance profile of herbicide‐resistant Alopecurus myosuroides (black‐grass) populations in Denmark. Crop Protection. 2015;69:83–89
  118. 118. Iwakami S, Hashimoto M, Matsushima K, Watanabe H, Hamamura K, Uchino A. Multiple‐herbicide resistance in Echinochloa crus‐galli var. formosensis, an allohexaploid weed species, in dry‐seeded rice. Pesticide Biochemistry and Physiology. 2015;119:1–8
  119. 119. Guo W, Liu W, Li L, Yuan G, Du L, Wang J. Molecular basis for resistance to Fenoxaprop in Shortawn Foxtail (Alopecurus aequalis) from China. Weed Science 2015;63 416–424
  120. 120. Cocker KM, Moss SR, Coleman JOD. Multiple mechanisms of resistance to fenoxaprop‐P‐ethyl in UK and other European populations of herbicide‐resistant Alopecurus myosuroides (black‐grass). Pesticide Biochemistry and Physiology. 1999;65:169–180
  121. 121. Letouzé A, Gasquez J. Inheritance of fenoxaprop‐P‐ethyl resistance in a blackgrass (Alopecurus myosuroides Huds.) population. Theoretical and Applied Genetics. 2001;103: 288–296
  122. 122. Petit C, Duhieu B, Boucansaud K, Délye C. Complex genetic control of non‐target‐site‐based resistance to herbicides inhibiting acetyl‐coenzyme A carboxylase and acetolactate‐synthase in Alopecurus myosuroides Huds. Plant Science. 2010;178:501–509
  123. 123. Cocker KM, Coleman JOD, Blair AM, Clarke JH, Moss SR. Biochemical mechanisms of cross‐resistance to aryloxyphenoxypropionate and cyclohexanedione herbicides in populations of Avena spp. Weed Research. 2000;40:323–334
  124. 124. Ahmad‐Hamdani MS, Yu Q, Han H, Cawthray GR, Wang SF, Powles SB. Herbicide resistance endowed by enhanced rates of herbicide metabolism in wild oat (Avena spp.). Weed Science. 2013;61:55–62
  125. 125. Maneechote C, Preston C, Powles SB. A diclofop‐methyl‐resistant Avena sterilis biotype with a herbicide‐resistant acetyl‐coenzyme A carboxylase and enhanced metabolism of diclofop‐methyl. Pesticide Science. 1997;49:105–114
  126. 126. Li L, Bi Y, Liu W, Yuan G, Wang J. Molecular basis for resistance to fenoxaprop‐p‐ethyl in American sloughgrass (Beckmannia syzigachne Steud.). Pesticide Biochemistry and Physiology. 2013;105:118–121
  127. 127. Tang W, Zhou Z, Zhang Y, Chen J. Resistance of American sloughgrass (Bechmannia syzigachne) populations to ACCase‐inhibiting herbicides involves three different target site mutations from China. Pesticide Biochemistry and Physiology. 2015;124:93–96
  128. 128. Huan ZB, Jin T, Zhang SY, Wang JX. Cloning and sequence analysis of plastid acetyl‐CoA carboxylase cDNA from two Echinochloa crusgalli biotypes. Journal of Pesticide Science. 2011;36(4):461–466
  129. 129. Huan Z, Xu Z, Lv D, Wang J. Determination of ACCase Sensitivity and gene expression in Quizalofop‐Ethyl‐Resistant and ‐Susceptible barnyardgrass (Echinochloa crus‐galli) biotypes. Weed Science. 2013;61:537–542
  130. 130. Wright AA, Nandula VK, Grier L, Showmaker KC, Bond JA, Peterson DG, Ray JD, Shaw DR. Characterization of Fenoxaprop‐P‐Ethyl–Resistant junglerice (Echinochloa colona) from Mississippi. Weed Science. 2016;64:588–595
  131. 131. McCullough PE, McElroy JS, Yu J, Zhang H, Miller TB, Chen S, Johnston CR, Czarnota MA. ALS‐resistant spotted spurge (Chamaesyce maculata) confirmed in Georgia. Weed Science. 2016;64:216–222
  132. 132. Shergill LS, Malone J, Boutsalis P, Preston C, Gill G. Target‐Site point mutations conferring resistance to ACCase‐inhibiting herbicides in smooth barley (Hordeum glaucum) and hare barley (Hordeum leporinum). Weed Science. 2015;63:408–415
  133. 133. Matthews N, Powles SB, Preston P. Mechanisms of resistance to acetyl‐coenzyme A carboxylase‐inhibiting herbicides in a Hordeum leporinum population. Pest Management Science. 2000;56:441–447
  134. 134. White GM, Moss SR, Karp A. Differences in the molecular basis of resistance to the cyclohexanedione herbicide sethoxydim in Lolium multiflorum. Weed Research. 2005;45:440–448
  135. 135. Kaundun SS, Hutchings SJ, Dale RP, McIndoe E. Broad resistance to ACCase inhibiting herbicides in a ryegrass population is due only to a cysteine to arginine mutation in the target enzyme. PLoS ONE. 2012;7(6):e39759
  136. 136. Cocker KM, Northcroft DS, Coleman JOD, Moss SR. Resistance to ACCase‐inhibiting herbicides and isoproturon in UK populations of Lolium multiflorum mechanisms of resistance and implications for control. Pest Management Science. 2001;57:587–597
  137. 137. Martins BAB, Sánchez‐Olguín E, Perez‐Jones A, Hulting AG, Mallory‐Smith C. Alleles contributing to ACCase‐resistance in an Italian ryegrass (Lolium perenne ssp. multiflorum) population from oregon. Weed Science. 2014;62:468–473
  138. 138. Collavo A, Panozzo S, Allegri A, Sattin M. A “Stressed” Alfalfa‐Based cropping system leads to the selection of Quizalofop‐Resistant Italian ryegrass (Lolium perenne ssp. multiflorum). Weed Science. 2016;64:683–694
  139. 139. Preston C. Inheritance and linkage of metabolism‐based herbicide cross‐resistance in rigid ryegrass (Lolium rigidum). Weed Science. 2003;51:4–12
  140. 140. Busi R, Vila‐Aiub MM, Powles SB. Genetic control of a cytochrome P450 metabolism based herbicide resistance mechanism in Lolium rigidum. Heredity. 2011;106:817–824
  141. 141. Yuan G, Liu W, Bi V, Du L, Guo W, Wang J. Molecular basis for resistance to ACCase‐inhibiting herbicides in Pseudosclerochloa kengiana populations. Pesticide Biochemistry and Physiology. 2015;119:9–15
  142. 142. Déley C, Wang T, Darmency H. An isoleucine‐leucine substitution in chloroplastic acetyl‐CoA carboxylase from green foxtail (Setaria viridis L. Beauv.) is responsible for resistance to the cyclohexanedione herbicide sethoxydim. Planta. 2002;214:421–427
  143. 143. Marles MAS, Devine MD, Hall CJ. Herbicide resistance in Setari aviridis conferred by sell sensitive forms of acetyl coenzyle A carboxylase. Pesticide Biochemistry and Physiology. 1993;46:7–14
  144. 144. Scarabel L, Panozzo S, Savoia W, Sattin M. Target‐site ACCase‐resistant johnsongrass (Sorghum halepense) selected in summer dicot crops. Weed Technology. 2014;28(2):307–315
  145. 145. Burke IC, Burton JD, York AC, Cranmer J, Wilcut JW. Mechanism of resistance to clethodim in a johnsongrass (Sorghum halepense) biotype. Weed Science. 2006;54(3):401–406
  146. 146. Steinrucken HC, Amrhein N. The herbicide glyphosate is a potent inhibitor of 5‐enolpyruvylshikimic acid‐3‐phosphate synthase. Biochemical and Biophysical Research Communications. 1980;94:1207–1212
  147. 147. Baylis AD. Why glyphosate is a global herbicide: Strengths, weaknesses and prospects. Pest Management Science. 2000;56:299–308
  148. 148. James C, Krattiger AF. Global Review of the Field Testing and Commercialization of Transgenic Plants, 1986 to 1995: The First Decade of Crop Biotechnology. ISAAA Briefs No. 1. Ithaca, NY: ISAAA; 1996. p. 31
  149. 149. Bradshaw LD, Padgette SR, Kimball SL, Wells BH. Perspectives on glyphosate resistance. Weed Technology. 1997;11:189–198
  150. 150. Nandula KV, Ray DJ, Ribeiro ND, Pan Z, Reddy NK. Glyphosate resistance in tall waterhemp (Amaranthus tuberculatus) from Mississippi is due to both altered target‐site and nontarget‐site mechanisms. Weed Science. 2013;61:374–383
  151. 151. Chatham AL, Bradley WK, Kruger RG, Martin RJ, Owen DKM, Peterson ED, Mithila J, Tranel JP. A multistate study of the association between glyphosate resistance and EPSPS gene amplification in waterhemp (Amaranthus tuberculatus). Weed Science. 2015;63:569–577
  152. 152. Mohseni‐Moghadam M, Schroeder J, Ashigh J. Mechanism of resistance and inheritance in glyphosate resistant palmer amaranth (Amaranthus palmeri) populations from New Mexico, USA. Weed Science. 2013;61:517–525
  153. 153. Young BG, Knepp AW, Wax LM, Hart SE. Glyphosate translocation in common lambsquarters (Chenopodium album) and velvetleaf (Abutilon theophrasti) in response to ammonium sulfate. Weed Science. 2003;51:151–156
  154. 154. Krga I, Pavlović D, Andelković A, Ðurović S, Marisavljević D. The resistance of orchard weed populations to glyphosate (In Serbian: Otpornost korovskih populacija u voćnjacima na glifosat). Plant Protection. 2013;64(3):125–133
  155. 155. Boerboom MC, Wzese LD. Influence of glyphosate concentration on glyphosate absorption and translocation in Canada Thistle (Cirsium arvense). Weed Science. 1988;36: 291–295
  156. 156. Bourque J, Chen YS, Heck G, Hubmeier C, Reynolds T, Tran M, Ratliff PG, Sammans D. Investigation into glyphosate‐resistant horseweed (Conyza canadensis): Resistance mechanism studies. Book of Abstract of the Weed Science Society of America. 2002;42:65
  157. 157. Feng PC, Tran M, Chiu T, Sammons RD, Heck GR, CaJacob CA. Investigations into glyphosate resistant horseweed (Conyza canadensis): retention, uptake, translocation and metabolism. Weed Science. 2004;52:498–505
  158. 158. Koger CH, Reddy KN. Role of absorption and translocation in the mechanism of glyphosate resistance in horseweed (Conyza canadensis). Weed Science. 2005;53:84–89
  159. 159. Dinelli G, Marotti I, Bonetti A, Minelli M, Catizone P, Barnes J. Physiological and molecular insight on the mechanisms of resistance to glyphosate in Conyza canadensis (L.) Cronq biotipes. Pesticide Biochemistry and Physiology. 2006;86:30–41
  160. 160. Pavlovic D, Reinhardt FC, Bozic D, Vrbnicanin S. Determination of Conyza canadensis levels of sensitivity to glyphosate trmesium sulphsate. International Journal of Agriculture and Biology. 2013;15(6):1091–1097
  161. 161. Dinelli G, Marotti I, Bonetti A, Catiyone P, Urbano JM, Barnes J. Physiological and molecular mechanisms of glyphosate resistance in Conyza bonariensis (L.) Cronq biotypes from Spain. Weed Research. 2008;48:257–265
  162. 162. Yerka MK, De Leon N, Stoltenberg DE. Pollen‐mediated gene flow in common lambsquarters (Chenopodium album). Weed Science. 2012;60(4):600–606
  163. 163. Tehranchian P, Riar DS, Norsworthy JK, Nandula V, McElroy S, Chen S, Scott RC. ALS‐resistant smallflower umbrella sedge (Cyperus difformis) in arkansas rice: Physiological and molecular basis of resistance. Weed Science. 2015;63:561–568
  164. 164. Wang YC. Effect of glyphosate on aromatic amino acid metabolism in purple nutsedge (Cyperus rotundus). Weed Technology. 2001;15(4):628–635
  165. 165. Baerson SR, Rodriguez DJ, Tran M, Feng Y, Biest NA, Dill GM. Glyphosate‐Resistant goosegrass. Identification of a mutation in the target enzyme 5‐Enolpyruvylshikimate‐3‐Phosphate synthase. Plant Physiology. 2002;129:1265–1275
  166. 166. Ng CH, Wickneswari R, Salmijan S, Teng YT, Ismail BS. Gene polymorphisms in glyphosate‐resistant and susceptible biotypes of Eleusine indica from Malaysia. Weed Research. 2003;43:108–115
  167. 167. Kaundun SS, Zelaya IA, Dale RP, Lycett AJ, Carter P, Sharples KR, McIndoe E. Importance of the P106S target‐site mutation in conferring resistance to glyphosate in a goosegrass (Eleusine indica) population from the Philippines. Weed Science. 2008;56:637–646
  168. 168. Alarcón‐Reverte R, García A, Urzúa J, Fischer AJ. Resistance to glyphosate in junglerice (Echinochloa colona) from California. Weed Science. 2013;61:48–54
  169. 169. Wakelin AM, Preston CA. Target‐site mutation is present in a glyphosate‐resistant Lolium rigidum population. Weed Research. 2006;46:432–440
  170. 170. Kaundun SS, Dale RP, Zelaya IA, Dinelli G, Marotti I, McIndoe E, Cairns A. A novel P106L mutation in EPSPS and an unknown mechanism(s) act additively to confer resistance to glyphosate in a South African Lolium rigidum population. Journal of Agricultural and Food Chemistry. 2011;59(7):3227–3233
  171. 171. Feng PCC, Pratley JE, Bohn JA. Resistance to glyphosate in Lolium rigidum. II Uptake, translocation and metabolism. Weed Science. 1999;47:412–415
  172. 172. Simarmata M, Kaufmann JE, Penner D. Potential basis of glyphosate resistance in California rigid ryegrass (Lolium rigidum). Weed Science. 2003;51:678–682
  173. 173. Lorraine‐Colwill DF, Powles SB, Hawkes TR, Hollinshead PH, Warner SAJ, Preston C. Investigations into the mechanism of glyphosate resistance in Lolium rigidum. Pesticide Biochemistry and Physiology. 2003;74:62–72
  174. 174. Wakelin AM, Lorraine‐Colwill DF, Preston C. Glyphosate resistance in four different populations of Lolium rigidum is associated with reduced translocation of glyphosate to meristematic zones. Weed Research. 2004;44:453–459
  175. 175. Adu‐Yeboah P, Malone MJ, Gill G, Preston C. Reduced glyphosate translocation in two glyphosate‐resistant populations of rigid ryegrass (Lolium rigidum) from fence lines in South Australia. Weed Science. 2014;62:4–10
  176. 176. Baerson SR, Rodriguez DJ, Biest NA, Tran M, You J, Kreuger RW, Dill GM, Pratley JE, Gruys KJ. Investigating the mechanism of glyphosate resistance in rigid ryegrass (Lolium rigidum). Weed Science. 2002;50:721–730
  177. 177. Pavlović D, Reinhardt C, Elezović I, Vrbničanin S. Identification of glyphosate resistance in Lolium rigidum Gaudin. Pesticides and Phytomedicine. 2011;26(4):393–399
  178. 178. Perez‐Jones A, Park KW, Polge N, Colquhoun J, Mallory‐Smith CA. Investigating the mechanisms of glyphosate resistance in Lolium multiflorum. Planta. 2007;226:395–404
  179. 179. Michitte P, De Prado R, Espinoza N, Ruiz‐Santaella JP, Gaurit C. Mechanisms of resistance to glyphosate in a Ryegrass (Lolium multiflorum) biotype from Chile. Weed Science. 2007;55(5):435–440
  180. 180. Salas AR, Scott CR, Dayan EF, Burgos RN. EPSPS Gene Amplification in glyphosate‐resistant Italian ryegrass (Lolium perenne ssp. multiflorum) populations from Arkansas (United States). Journal of Agricultural and Food Chemistry. 2015;63:5885–5893
  181. 181. Cross BR, McCarty BL, Tharayil N, McElroy SJ, Chen S, McCullough EM, Powell AB, Bridges CW. A Pro106 to Ala substitution is associated with resistance to glyphosate in annual bluegrass (Poa annua). Weed Science. 2015;63:613–622
  182. 182. Riar DS, Norsworthy JK, Johnson DB, Scott RC, Bagavathiannan M. Glyphosate resistance in a johnsongrass (Sorghum halepense) biotype from arkansas. Weed Science. 2011;59:299–304
  183. 183. Norsworthy JK, Ward SM, Shaw DR, Lewellyn RS, Nichols RL, Webster TM, Bradley KW, Frisvold G, Powles SB, Burgos NR, Witt WW, Barrett M. Reducing the risks of herbicide resistance: Best management practices and recommendations. Weed Science 2012;12(Special Issue):31–62

Written By

Sava Vrbničanin, Danijela Pavlović and Dragana Božić

Submitted: 31 October 2016 Reviewed: 17 February 2017 Published: 04 October 2017