Journal articles on the topic 'Herbicide resistance'

To see the other types of publications on this topic, follow the link: Herbicide resistance.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Herbicide resistance.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Elezovic, Ibrahim, Dragana Bozic, and Sava Vrbnicanin. "Weed resistance to herbicides states: Causes and possibilities of preventive resistance." Pesticidi 18, no. 1 (2003): 5–21. http://dx.doi.org/10.2298/pif0301005e.

Full text
Abstract:
Resistance occurs as a result of heritable changes to biochemical processes that enable plant survival when treated with a herbicide. Resistance can result from changes to the herbicides target site such that binding of the herbicide is reduced, or over-expression of the target site may occur. Alternatively, there may be a reduction in the amount of herbicide that reaches the target enzyme through detoxication, sequestration, or reduced absorption of herbicide. Finally, the plant may survive through the ability to protect plant metabolism from toxic compounds produced as a consequence of herbicide action. Herbicide-resistant weeds were predicted shortly after the introduction of herbicides. During the 1970s, many, additional important weed species (e.g., Amaranthus spp., Chennpodium spp., Erigeron canadensis Kochia scoparia, Solanum nigrum, Panicum crus-galli, Senecio vulgaris, Poa annua) were reported to be resistant to triazine herbicides and several other herbicides. Over the last 10 years and now ALS-herbicide-resistant weeds account for the greatest number of resistant species and probably the largest area affected by resistance. In contrast to triazine resistance target-site-based resistance to the ALS-inhibiting herbicides can be conferred by a number of different point mutations. Differences occur in target-site cross-resistance among the different chemical classes of herbicides that inhibit ALS. The differences are related to particular amino acid substitutions that occur within the binding region. Indeed, six different substitutions of Ala, Arg, Glu, Leu, Ser, or Tri for Pro 173 have been observed in different weed populations.
APA, Harvard, Vancouver, ISO, and other styles
2

Rigon, Carlos A. G., Todd A. Gaines, Anita Küpper, and Franck E. Dayan. "Metabolism-Based Herbicide Resistance, the Major Threat Among the Non-Target Site Resistance Mechanisms." Outlooks on Pest Management 31, no. 4 (August 1, 2020): 162–68. http://dx.doi.org/10.1564/v31_aug_04.

Full text
Abstract:
Evolution of resistance to pesticides is a problem challenging the sustainability of global food production. Resistance to herbicides is driven by the intense selection pressure imparted by synthetic herbicides on which we rely to manage weeds. Target-site resistance (TSR) mechanisms involve changes to the herbicide target protein and provide resistance only to herbicides within a single mechanism of action. Non-target site resistance (NTSR) mechanisms reduce the quantity of herbicide reaching the target site and/or modify the herbicide. NTSR mechanisms include reduced absorption and/or translocation, increased sequestration, and enhanced metabolic degradation. Of these diverse mechanisms contributing to NTSR, metabolism-based herbicide resistance represents a major threat because it can impart resistance to herbicides from varied chemical classes across any number of mechanisms of action.
APA, Harvard, Vancouver, ISO, and other styles
3

Gaines, Todd A., Stephen O. Duke, Sarah Morran, Carlos A. G. Rigon, Patrick J. Tranel, Anita Küpper, and Franck E. Dayan. "Mechanisms of evolved herbicide resistance." Journal of Biological Chemistry 295, no. 30 (May 19, 2020): 10307–30. http://dx.doi.org/10.1074/jbc.rev120.013572.

Full text
Abstract:
The widely successful use of synthetic herbicides over the past 70 years has imposed strong and widespread selection pressure, leading to the evolution of herbicide resistance in hundreds of weed species. Both target-site resistance (TSR) and nontarget-site resistance (NTSR) mechanisms have evolved to most herbicide classes. TSR often involves mutations in genes encoding the protein targets of herbicides, affecting the binding of the herbicide either at or near catalytic domains or in regions affecting access to them. Most of these mutations are nonsynonymous SNPs, but polymorphisms in more than one codon or entire codon deletions have also evolved. Some herbicides bind multiple proteins, making the evolution of TSR mechanisms more difficult. Increased amounts of protein target, by increased gene expression or by gene duplication, are an important, albeit less common, TSR mechanism. NTSR mechanisms include reduced absorption or translocation and increased sequestration or metabolic degradation. The mechanisms that can contribute to NTSR are complex and often involve genes that are members of large gene families. For example, enzymes involved in herbicide metabolism–based resistances include cytochromes P450, GSH S-transferases, glucosyl and other transferases, aryl acylamidase, and others. Both TSR and NTSR mechanisms can combine at the individual level to produce higher resistance levels. The vast array of herbicide-resistance mechanisms for generalist (NTSR) and specialist (TSR and some NTSR) adaptations that have evolved over a few decades illustrate the evolutionary resilience of weed populations to extreme selection pressures. These evolutionary processes drive herbicide and herbicide-resistant crop development and resistance management strategies.
APA, Harvard, Vancouver, ISO, and other styles
4

Vencill, William K., Robert L. Nichols, Theodore M. Webster, John K. Soteres, Carol Mallory-Smith, Nilda R. Burgos, William G. Johnson, and Marilyn R. McClelland. "Herbicide Resistance: Toward an Understanding of Resistance Development and the Impact of Herbicide-Resistant Crops." Weed Science 60, SP1 (2012): 2–30. http://dx.doi.org/10.1614/ws-d-11-00206.1.

Full text
Abstract:
Development of herbicide-resistant crops has resulted in significant changes to agronomic practices, one of which is the adoption of effective, simple, low-risk, crop-production systems with less dependency on tillage and lower energy requirements. Overall, the changes have had a positive environmental effect by reducing soil erosion, the fuel use for tillage, and the number of herbicides with groundwater advisories as well as a slight reduction in the overall environmental impact quotient of herbicide use. However, herbicides exert a high selection pressure on weed populations, and density and diversity of weed communities change over time in response to herbicides and other control practices imposed on them. Repeated and intensive use of herbicides with the same mechanisms of action (MOA; the mechanism in the plant that the herbicide detrimentally affects so that the plant succumbs to the herbicide; e.g., inhibition of an enzyme that is vital to plant growth or the inability of a plant to metabolize the herbicide before it has done damage) can rapidly select for shifts to tolerant, difficult-to-control weeds and the evolution of herbicide-resistant weeds, especially in the absence of the concurrent use of herbicides with different mechanisms of action or the use of mechanical or cultural practices or both.
APA, Harvard, Vancouver, ISO, and other styles
5

Jones, Eric A. L., and Micheal D. K. Owen. "Investigating the Efficacy of Selected Very-Long-Chain Fatty Acid-Inhibiting Herbicides on Iowa Waterhemp (Amaranthus tuberculatus) Populations with Evolved Multiple Herbicide Resistances." Agronomy 11, no. 3 (March 21, 2021): 595. http://dx.doi.org/10.3390/agronomy11030595.

Full text
Abstract:
Very long chain fatty acid (VLCFA)-inhibiting herbicides (Herbicide group (HG) 15) have been applied to corn and soybean fields in Iowa since the 1960s. The VLCFA-inhibiting herbicides are now applied more frequently to control multiple herbicide-resistant (MHR) waterhemp (Amaranthus tuberculatus Moq. J.D. Sauer) populations that are ubiquitous across the Midwest United States as resistance to the VLCFA-inhibiting herbicides is not widespread. Waterhemp has evolved multiple resistances to herbicides from seven sites of action (HG 2, 4, 5, 9, 14, 15, and 27), and six-way herbicide-resistant populations have been confirmed. Thus, the objective of this study was to determine if selected Iowa waterhemp populations are less sensitive to VLCFA-inhibiting herbicides when additional herbicide resistance traits have evolved within the selected population. Dose–response assays were conducted in a germination chamber to determine the efficacy of three selected VLCFA-inhibiting herbicides (acetochlor, S-metolachlor, and flufenacet) on selected Iowa MHR waterhemp populations. An herbicide-susceptible, three-way, four-way, and five-way herbicide-resistant waterhemp population responded to the herbicide treatments differently; however, several of the four-way and five-way herbicide-resistant populations exhibited resistance ratios greater than 1 when treated with acetochlor and S-metolachlor. Selected four-way herbicide-resistant waterhemp populations from Iowa were subjected to a dose–response assay in the field using the same VLCFA-inhibiting herbicides, and all herbicides achieved control greater than 80% at the maximum labeled rate. The results of the experiments provide evidence that some MHR waterhemp populations may exhibit decreased susceptibility the VLCFA-inhibiting herbicides, but generally, these herbicides remain efficacious on Iowa MHR waterhemp populations.
APA, Harvard, Vancouver, ISO, and other styles
6

Dong, Huirong, Yong Huang, and Kejian Wang. "The Development of Herbicide Resistance Crop Plants Using CRISPR/Cas9-Mediated Gene Editing." Genes 12, no. 6 (June 12, 2021): 912. http://dx.doi.org/10.3390/genes12060912.

Full text
Abstract:
The rapid increase in herbicide-resistant weeds creates a huge challenge to global food security because it can reduce crop production, causing considerable losses. Combined with a lack of novel herbicides, cultivating herbicide-resistant crops becomes an effective strategy to control weeds because of reduced crop phytotoxicity, and it expands the herbicidal spectrum. Recently developed clustered regularly interspaced short palindromic repeat/CRISPR-associated protein (CRISPR/Cas)-mediated genome editing techniques enable efficiently targeted modification and hold great potential in creating desired plants with herbicide resistance. In the present review, we briefly summarize the mechanism responsible for herbicide resistance in plants and then discuss the applications of traditional mutagenesis and transgenic breeding in cultivating herbicide-resistant crops. We mainly emphasize the development and use of CRISPR/Cas technology in herbicide-resistant crop improvement. Finally, we discuss the future applications of the CRISPR/Cas system for developing herbicide-resistant crops.
APA, Harvard, Vancouver, ISO, and other styles
7

Walsh, Michael J., Stephen B. Powles, Brett R. Beard, Ben T. Parkin, and Sally A. Porter. "Multiple-herbicide resistance across four modes of action in wild radish (Raphanus raphanistrum)." Weed Science 52, no. 1 (February 2004): 8–13. http://dx.doi.org/10.1614/ws-03-016r.

Full text
Abstract:
Populations of wild radish were collected from two fields in the northern Western Australian wheatbelt, where typical herbicide-use patterns had been practiced for the previous 17 seasons within an intensive crop production program. The herbicide resistance status of these populations clearly established that there was multiple-herbicide resistance across many herbicides from at least four modes of action. One population exhibited multiple-herbicide resistance to the phytoene desaturase (PDS)–inhibiting herbicide diflufenican (3.0-fold), the auxin analog herbicide 2,4-D (2.2-fold), and the photosystem II–inhibiting herbicides metribuzin and atrazine. Another population was found to be multiply resistant to the acetolactate synthase–inhibiting herbicides, the PDS-inhibiting herbicide diflufenican (2.5-fold), and the auxin analog herbicide 2,4-D amine (2.4-fold). Therefore, each population has developed multiple-herbicide resistance across several modes of action. The multiple resistance status of these wild radish populations developed from conventional herbicide usage in intensive cropping rotations, indicating a dramatic challenge for the future control of wild radish.
APA, Harvard, Vancouver, ISO, and other styles
8

Manalil, Sudheesh, Roberto Busi, Michael Renton, and Stephen B. Powles. "Rapid Evolution of Herbicide Resistance by Low Herbicide Dosages." Weed Science 59, no. 2 (June 2011): 210–17. http://dx.doi.org/10.1614/ws-d-10-00111.1.

Full text
Abstract:
Herbicide rate cutting is an example of poor use of agrochemicals that can have potential adverse implications due to rapid herbicide resistance evolution. Recent laboratory-level studies have revealed that herbicides at lower-than-recommended rates can result in rapid herbicide resistance evolution in rigid ryegrass populations. However, crop-field-level studies have until now been lacking. In this study, we examined the impact of low rates of diclofop on the evolution of herbicide resistance in a herbicide-susceptible rigid ryegrass population grown either in a field wheat crop or in potted plants maintained in the field. Subsequent dose–response profiles indicated rapid evolution of diclofop resistance in the selected rigid ryegrass lines from both the crop-field and field pot studies. In addition, there was moderate level of resistance in the selected lines against other tested herbicides to which the population has never been exposed. This resistance evolution was possible because low rates of diclofop allowed substantial rigid ryegrass survivors due to the potential in this cross-pollinated species to accumulate all minor herbicide resistance traits present in the population. The practical lesson from this research is that herbicides should be used at the recommended rates that ensure high weed mortality to minimize the likelihood of minor herbicide resistance traits leading to rapid herbicide resistance evolution.
APA, Harvard, Vancouver, ISO, and other styles
9

Shaner, Dale L. "Lessons Learned From the History of Herbicide Resistance." Weed Science 62, no. 2 (June 2014): 427–31. http://dx.doi.org/10.1614/ws-d-13-00109.1.

Full text
Abstract:
The selection of herbicide-resistant weed populations began with the introduction of synthetic herbicides in the late 1940s. For the first 20 years after introduction, there were limited reported cases of herbicide-resistant weeds. This changed in 1968 with the discovery of triazine-resistant common groundsel. Over the next 15 yr, the cases of herbicide-resistant weeds increased, primarily to triazine herbicides. Although triazine resistance was widespread, the resistant biotypes were highly unfit and were easily controlled with specific alternative herbicides. Weed scientists presumed that this would be the case for future herbicide-resistant cases and thus there was not much concern, although the companies affected by triazine resistance were somewhat active in trying to detect and manage resistance. It was not until the late 1980s with the discovery of resistance to Acetyl Co-A carboxylase (ACCase) and acetolactate synthase (ALS) inhibitors that herbicide resistance attracted much more attention, particularly from industry. The rapid evolution of resistance to these classes of herbicides affected many companies, who responded by first establishing working groups to address resistance to specific classes of herbicides, and then by formation of the Herbicide Resistance Action Committee (HRAC). The goal of these groups, in cooperation with academia and governmental agencies, was to act as a forum for the exchange of information on herbicide-resistance selection and to develop guidelines for managing resistance. Despite these efforts, herbicide resistance continued to increase. The introduction of glyphosate-resistant crops in the 1995 provided a brief respite from herbicide resistance, and farmers rapidly adopted this relatively simple and reliable weed management system based on glyphosate. There were many warnings from academia and some companies that the glyphosate-resistant crop system was not sustainable, but this advice was not heeded. The selection of glyphosate resistant weeds dramatically changed weed management and renewed emphasis on herbicide resistance management. To date, the lesson learned from our experience with herbicide resistance is that no herbicide is invulnerable to selecting for resistant biotypes, and that over-reliance on a weed management system based solely on herbicides is not sustainable. Hopefully we have learned that a diverse weed management program that combines multiple methods is the only system that will work for the long term.
APA, Harvard, Vancouver, ISO, and other styles
10

Toubou, Elisavet, Vassiliki Vindena, Christos A. Damalas, and Spyridon D. Koutroubas. "Weed control practices and awareness of herbicide resistance among cereal farmers of northern Greece." Weed Technology 34, no. 6 (July 20, 2020): 909–15. http://dx.doi.org/10.1017/wet.2020.79.

Full text
Abstract:
AbstractKnowledge of weed control practices and farmers’ awareness of herbicide resistance could be a basis for improving weed management programs with respect to herbicide resistance, but research on this topic is limited. This study reports current weed control practices and levels of awareness of herbicide resistance among cereal farmers of northern Greece. Face-to-face interviews were conducted with 250 cereal farmers of Evros district, based on a structured questionnaire. Most farmers (82.8%) used herbicides in cereal production, with one application per growing season. Farmers appeared divided with respect to using the same herbicide each year; the majority of the farmers (90.8%) applied crop rotation. Almost half of the farmers (47.2%) did not know what herbicide resistance is, but most farmers (75.1%) felt herbicide resistance would be a problem for them. According to their answers on nine knowledge questions about herbicide resistance, 66.8% of the farmers had good knowledge, and 33.2% had poor knowledge. Almost seven in 10 farmers (69.8%) did not consider herbicide resistance when purchasing an herbicide for use, and only 40.4% were willing to change common weed control practices to prevent herbicide resistance. Awareness of herbicide resistance did not differ by sex; poor awareness levels increased with advanced age, low education levels, and small farm size. Farmers who used chemical weed control had higher awareness levels of herbicide resistance than farmers who never used herbicides. Farmers who were keeping records of herbicide applications, those who observed low efficacy of herbicides in their field, and those who applied crop rotation had high awareness levels of herbicide resistance, whereas farmers who used the same herbicide each year had poor awareness. Findings shed light on inter-relationships between farmers’ awareness of herbicide resistance and current weed control practices that could be useful for targeted extension education.
APA, Harvard, Vancouver, ISO, and other styles
11

Dear, BS, GA Sandral, and NE Coombes. "Change in stomatal resistance and water use in subterranean clover (Trifolium subterraneum L.) in response to broadleaf herbicides." Australian Journal of Agricultural Research 47, no. 4 (1996): 625. http://dx.doi.org/10.1071/ar9960625.

Full text
Abstract:
The effect of 5 broadleaf herbicides on the water use and stomatal resistance of 2 cultivars of subterranean clover (Trifolium subterraneum L.) was examined in a glasshouse study. The herbicide treatments 2,4-DB, MCPA, bromoxynil, MCPA+terbutryn, and MCPA+diuron were applied at 6 rates at 2 times (14 May, 14 June) to plants at 2 leaf stages (3-4 and 8-10 leaves). Each of the herbicides reduced water use by the clover within 24 h, the size of the reduction increasing with the rate of herbicide applied. The herbicide treatments MCPA+terbutryn, MCPA+diuron, and bromoxynil caused the largest reductions (44-52%) in total water use over the 30-day period when applied at the recommended rate, and MCPA and 2,4-DB the least reduction (16-22%). Stomatal resistance increased substantially within 2 days of application of each of the herbicides. The magnitude of the change differed with herbicide and increased with herbicide rate. The effect of the herbicides on stomatal resistance declined 10-20 days after herbicide application in all treatments except 2,4-DB, but stomatal resistance of all herbicide-treated plants was still higher than the control 30 days after herbicide application. The herbicides LICPA+terbutryn and LlCPA+diuron and bromoxynil caused the largest increase in stomatal resistance and 2,4-DB the least. Stomatal resistance was found to be highly negatively correlated with daily water use by the clover plants at 2 days (r = -0.84, P < 0.01) and 30 days (r = -0.88, P < 0.01) after herbicide application. All of the herbicides reduced the LA1 of the plants, the effect increasing as the herbicide rate increased. Herbicide and herbicide rate had the largest effect on both water use and stomatal resistance; the effect of cultivar, leaf stage, and spraying time accounted for a relatively small proportion of the variance. The findings support the hypothesis that some broadleaf herbicides can result in a water-saving effect in subterranean clover swards through increasing stomatal resistance and decreasing the LAI, thereby potentially reducing moisture stress during seed set.
APA, Harvard, Vancouver, ISO, and other styles
12

Leon, Ramon G., and J. Bryan Unruh. "Turfgrass Herbicides: Mechanisms of Action and Resistance Management." EDIS 2015, no. 7 (October 9, 2015): 4. http://dx.doi.org/10.32473/edis-ag398-2015.

Full text
Abstract:
Herbicides are an effective tool for controlling weeds in turfgrass; however, weeds can become resistant to herbicides and create significant problems for turfgrass production. The best way to combat herbicide resistance is to rotate herbicides with different mechanisms of action (MOA) because using herbicides with different MOAs makes it more likely that weeds resistant to one herbicide will encounter an herbicide to which they are not resistant. This 4-page fact sheet focuses on how to create an herbicide program that uses different MOAs to manage resistant weeds. Written by Ramon G. Leon and Bryan Unruh, and published by the UF Department of Agronomy, August 2015. SS-AGR-394/AG398: Turfgrass Herbicides: Mechanisms of Action and Resistance Management (ufl.edu)
APA, Harvard, Vancouver, ISO, and other styles
13

Gressel, Jonathan, and Lee A. Segel. "Negative Cross Resistance; a Possible Key to Atrazine Resistance Management: A Call for Whole Plant Data." Zeitschrift für Naturforschung C 45, no. 5 (May 1, 1990): 470–73. http://dx.doi.org/10.1515/znc-1990-0528.

Full text
Abstract:
Many photosystem II inhibiting herbicides still inhibit this process in triazine-resistant plants; i.e. they have no cross resistance with atrazine. Five- to twenty-fold lower concentrations of phenolic type herbicidcs and 5-fold less of the active ingredient of pyridate and half as much ioxynil are required to inhibit thylakoid PS II in atrazine-resistant biotypes than in sensitive biotypes; i.e., they even show “negative cross resistance”. Negative cross resistance may be the major reason that atrazine resistance did not evolve where herbicide mixtures were used, when the mixed herbicide (usually a non-PS II inhibiting acetanilide) also controlled triazine-sensitivc weeds. Mathematical modeling in principle allows quantification of the very low field levels of herbicides possessing negative cross resistance that could be mixed with atrazine that would stop or delay the evolution of resistant populations without affecting the maize crop. There are few available actual dose response curves of atrazine-resistant vs. susceptible weeds at the whole plant level for herbicidcs exerting negative cross resistance. Thus, “real situation” modeling cannot be done. Data acquisition is called for so that the model can be extrapolated from the thylakoid to the field.
APA, Harvard, Vancouver, ISO, and other styles
14

Broster, J. C., J. E. Pratley, R. H. L. Ip, L. Ang, and K. P. Seng. "Cropping practices influence incidence of herbicide resistance in annual ryegrass (Lolium rigidum) in Australia." Crop and Pasture Science 70, no. 1 (2019): 77. http://dx.doi.org/10.1071/cp18355.

Full text
Abstract:
Herbicide resistance is a common occurrence in southern Australia. The evolution of herbicide resistance is influenced by the selection pressure placed on the weed species controlled by that herbicide. Results from resistance screening of ~4500 annual ryegrass (Lolium rigidum Gaud.) samples were entered in a GIS database, together with several agricultural parameters used in the Australian Bureau of Statistics Agricultural Surveys. This allowed a study of the associations between mode of action of resistance, geographic distribution of resistance across southern Australia, and farming practices employed in particular regions. Cultivation was negatively associated with resistances in acetyl-CoA carboxylase (ACCase)-inhibiting cyclohexanedione and acetolactate synthase (ALS)-inhibiting herbicides. Higher proportions of wheat sown were associated with higher incidences of resistance. ACCase-inhibiting aryloxyphenoxypropionate and cyclohexanedione and ALS-inhibiting resistances were higher in those shires where soils were predominantly acidic. This study demonstrates the association between farm practice and the evolution of herbicide resistance. The analysis provides reinforcement to the principle of rotating chemical modes of action with non-chemical weed control measures to minimise the risk of herbicide resistance evolution in any farming system.
APA, Harvard, Vancouver, ISO, and other styles
15

Délye, Christophe. "Weed resistance to acetyl coenzyme A carboxylase inhibitors: an update." Weed Science 53, no. 5 (October 2005): 728–46. http://dx.doi.org/10.1614/ws-04-203r.1.

Full text
Abstract:
Herbicides targeting grass plastidic acetyl coenzyme A carboxylase (ACC) are effective selective graminicides. Their intensive use worldwide has selected for resistance genes in a number of grass weed species. Biochemistry and molecular biology have been the means of determining the herbicidal activity and selectivity toward crop plants of ACC-inhibiting herbicides. In recent years, elucidation of the tridimensional structure of ACC and identification of five amino acid residues within the ACC carboxyl transferase domain that are critical determinants for herbicide sensitivity shed light on the basis of ACC-based resistance to herbicides. However, metabolism-based resistance to ACC-inhibiting herbicides is much less well known, although this type of resistance seems to be widespread. A number of genes thus endow resistance to ACC-inhibiting herbicides, with the possibility for various resistance genes that confer dominant resistance at the herbicide field rate to accumulate within a single weed population or plant. This, together with a poor knowledge of the genetic parameters driving resistance, renders the evolution of resistance to ACC-inhibiting herbicides unpredictable. Future research should consider developing tactics to slow the spread of resistance. For this purpose, it is crucial that our understanding of metabolism-based resistance improves rapidly because this mechanism is complex and can confer resistance to herbicides with different target sites.
APA, Harvard, Vancouver, ISO, and other styles
16

Beckie, Hugh J., Michael B. Ashworth, and Ken C. Flower. "Herbicide Resistance Management: Recent Developments and Trends." Plants 8, no. 6 (June 8, 2019): 161. http://dx.doi.org/10.3390/plants8060161.

Full text
Abstract:
This review covers recent developments and trends in herbicide-resistant (HR) weed management in agronomic field crops. In countries where input-intensive agriculture is practiced, these developments and trends over the past decade include renewed efforts by the agrichemical industry in herbicide discovery, cultivation of crops with combined (stacked) HR traits, increasing reliance on preemergence vs. postemergence herbicides, breeding for weed-competitive crop cultivars, expansion of harvest weed seed control practices, and advances in site-specific or precision weed management. The unifying framework or strategy underlying these developments and trends is mitigation of viable weed seeds into the soil seed bank and maintaining low weed seed banks to minimize population proliferation, evolution of resistance to additional herbicidal sites of action, and spread. A key question going forward is: how much weed control is enough to consistently achieve the goal of low weed seed banks? The vision for future HR weed management programs must be sustained crop production and profitability with reduced herbicide (particularly glyphosate) dependency.
APA, Harvard, Vancouver, ISO, and other styles
17

Mithila, J., J. Christopher Hall, William G. Johnson, Kevin B. Kelley, and Dean E. Riechers. "Evolution of Resistance to Auxinic Herbicides: Historical Perspectives, Mechanisms of Resistance, and Implications for Broadleaf Weed Management in Agronomic Crops." Weed Science 59, no. 4 (December 2011): 445–57. http://dx.doi.org/10.1614/ws-d-11-00062.1.

Full text
Abstract:
Auxinic herbicides are widely used for control of broadleaf weeds in cereal crops and turfgrass. These herbicides are structurally similar to the natural plant hormone auxin, and induce several of the same physiological and biochemical responses at low concentrations. After several decades of research to understand the auxin signal transduction pathway, the receptors for auxin binding and resultant biochemical and physiological responses have recently been discovered in plants. However, the precise mode of action for the auxinic herbicides is not completely understood despite their extensive use in agriculture for over six decades. Auxinic herbicide-resistant weed biotypes offer excellent model species for uncovering the mode of action as well as resistance to these compounds. Compared with other herbicide families, the incidence of resistance to auxinic herbicides is relatively low, with only 29 auxinic herbicide-resistant weed species discovered to date. The relatively low incidence of resistance to auxinic herbicides has been attributed to the presence of rare alleles imparting resistance in natural weed populations, the potential for fitness penalties due to mutations conferring resistance in weeds, and the complex mode of action of auxinic herbicides in sensitive dicot plants. This review discusses recent advances in the auxin signal transduction pathway and its relation to auxinic herbicide mode of action. Furthermore, comprehensive information about the genetics and inheritance of auxinic herbicide resistance and case studies examining mechanisms of resistance in auxinic herbicide-resistant broadleaf weed biotypes are provided. Within the context of recent findings pertaining to auxin biology and mechanisms of resistance to auxinic herbicides, agronomic implications of the evolution of resistance to these herbicides are discussed in light of new auxinic herbicide-resistant crops that will be commercialized in the near future.
APA, Harvard, Vancouver, ISO, and other styles
18

Ilchenko, A. S., B. F. Varenyk, N. P. Lamary, and S. I. Karapira. "Inheriting the resistance of sunflower to tribenuron methyl under insufficient humidification in the southern Steppe of Ukraine." Agricultural Science and Practice 10, no. 2 (November 1, 2023): 38–45. http://dx.doi.org/10.15407/agrisp10.02.038.

Full text
Abstract:
Aim. This study aimed to investigate the inheritance of resistance to sulfonylurea herbicides in sunflower under conditions of insufficient humidification in the southern Steppe of Ukraine. Methods. Isolation, castration, hybridization, herbicide application (Granstar Pro 75 % w.g., containing tribenuron methyl as the active substance (a.s.)), evaluation of herbicide resistance, and statistical analysis of the acquired data. Results. The investigation into the inheritance of resistance to sulfonylurea herbicides involved the utilization of both resistant (SURES-1, OS 1099 V, OS 2017 V) and non-resistant (Od 1002 B, Od 1318 V, OS 1295 V) sunflower genotypes. Through crossing, four F1 hybrid combinations were generated, namely OS 2017 V × OS 1099 V, SURES-1 × Od 1002 B, SURES-1 × Od 1318 V, and Od 1318 V x OS 1295 V. Subsequent treatment of F1 plants with the herbicide Granstar Pro 75 % w.g. revealed that three combinations (OS 2017 V × OS 1099 V, SURES-1 × Od 1002 B, SURES-1 × Od 1318 V) exhibited complete resistance to the herbicide. In the second generation, following herbicide treatment, the hybrid combinations SURES-1 × Od 1002 B and SURES-1 × Od 1318 V displayed segregation into resistant and non-resistant plants. Conversely, the plants in the combination OS 2017 V × OS 1099 V maintained complete resistance to the herbicidal effects. Conclusions. The investigation, conducted in the challenging climatic conditions of the southern Steppe of Ukraine, demonstrated complete resistance to sulfonylurea herbicides in three hybrid combinations of both F1 and F2 generations. Notably, the combination OS 2017 V × OS 1099 V exhibited uniform resistance throughout the second generation, devoid of segregation. Moreover, the results of F2 segregation analysis in the SURES-1 × Od 1002 B and SURES-1 × Od 1318 V populations indicated that resistance to tribenuron methyl is primarily governed by the presence of a dominant gene allele. These findings offer valuable insights for the development of sunflower hybrids with enhanced herbicide resistance, particularly in regions with adverse climatic conditions.
APA, Harvard, Vancouver, ISO, and other styles
19

Torra, Joel, José M. Montull, Isabel M. Calha, María D. Osuna, Joao Portugal, and Rafael de Prado. "Current Status of Herbicide Resistance in the Iberian Peninsula: Future Trends and Challenges." Agronomy 12, no. 4 (April 13, 2022): 929. http://dx.doi.org/10.3390/agronomy12040929.

Full text
Abstract:
The evolution of herbicide resistance in weeds has emerged as one of the most serious threats to sustainable food production systems, which necessitates the evaluation of herbicides to determine their efficacy. The first herbicide resistance case in the Iberian Peninsula was reported about 50 years ago, wherein Panicum dichotomiflorum was found to be resistant (R) to atrazine in Spanish maize fields. Since then, herbicide resistance has evolved in 33 weed species, representing a total of 77 single-herbicide-resistance cases in this geographic area: 66 in Spain and 11 in Portugal. Changes in agricultural practices, namely the adoption of non-tillage systems and the increased use of herbicides, led to the selection of weed biotypes resistant to a wide range of herbicides. Nowadays the most important crops in Spain and Portugal (maize, winter cereals, rice, citrus, fruits, and olive orchards) are affected, with biotypes resistant to several mechanisms of action (MoAs), namely: ALS inhibitors (20 species), ACCase inhibitors (8 species), PS II inhibitors (18 species), and synthetic auxin herbicides (3 species). More recently, the fast increase in cases of resistance to the EPSPS-inhibiting herbicide glyphosate has been remarkable, with 11 species already having evolved resistance in the last 10 years in the Iberian Peninsula. The diversity of resistance mechanisms, both target-site and non-target-site, are responsible for the resistance to different MoAs, involving point mutations in the target site and enhanced rates of herbicide detoxification, respectively. More serious are the 13 cases reported with multiple-herbicide resistance, with three cases of resistance to three–four MoAs, and one case of resistance to five MoAs. Future research perspectives should further study the relationship between management strategies and the occurrence of TSR and NTSR resistance, to improve their design, develop monitoring and diagnostic tools for herbicide resistance, and deepen the study of NTSR resistance.
APA, Harvard, Vancouver, ISO, and other styles
20

Gherekhloo, Javid, Saeid Hassanpour-bourkheili, Parvin Hejazirad, Sajedeh Golmohammadzadeh, Jose G. Vazquez-Garcia, and Rafael De Prado. "Herbicide Resistance in Phalaris Species: A Review." Plants 10, no. 11 (October 21, 2021): 2248. http://dx.doi.org/10.3390/plants10112248.

Full text
Abstract:
Weeds, such as Phalaris spp., can drastically reduce the yield of crops, and the evolution of resistance to herbicides has further exacerbated this issue. Thus far, 23 cases of herbicide resistance in 11 countries have been reported in Phalaris spp., including Phalaris minor Retz., Phalaris paradoxa L., and Phalaris brachystachys L., for photosystem II (PS-II), acetyl-CoA carboxylase (ACCase), and acetolactate synthase (ALS)-inhibiting herbicides. This paper will first review the cases of herbicide resistance reported in P. minor, P. paradoxa, and P. brachystachys. Then, the mechanisms of resistance in Phalaris spp. are discussed in detail. Finally, the fitness cost of herbicide resistance and the literature on the management of herbicide-resistant weeds from these species are reviewed.
APA, Harvard, Vancouver, ISO, and other styles
21

Pinho, Camila Ferreira de, Jessica Ferreira Lourenço Leal, Amanda dos Santos Souza, Gabriella Francisco Pereira Borges de Oliveira, Claudia de Oliveira, Ana Claudia Langaro, Aroldo Ferreira Lopes Machado, Pedro Jacob Christoffoleti, and Luiz Henrique Saes Zobiole. "First evidence of multiple resistance of Sumatran Fleabane (Conyza sumatrensis (Retz.) E.Walker) to five- mode-of-action herbicides." Australian Journal of Crop Science, no. 13(10):2019 (October 20, 2019): 1688–97. http://dx.doi.org/10.21475/ajcs.19.13.10.p1981.

Full text
Abstract:
Herbicide resistance is the evolutionary response of weeds to the selection pressure caused by repeated application of the same active ingredient. It can result from two different mechanisms, known as target site resistance (TSR) and non-target site resistance (NTSR). In addition to single-herbicide resistance, multiple resistance can occur due to herbicides selection or accumulation of resistance genes by cross-pollination. The aim of this research was to investigate the suspect of multiple herbicide resistance of Sumatran Fleabane (Conyza sumatrensis (Retz.) E.Walker) to herbicides frequently used as a burndown application. Dose-responses in a whole-plant assay were carried out to investigate multiple-resistance of Sumatran fleabane to paraquat, saflufenacil, diuron, 2,4-D and glyphosate. Results indicated that the resistance index (ratio R/S) based on herbicide rate to cause 50% mortality (LD50) were 25.51, 1.39, 7.29, 1.84 and 7.55 for paraquat, saflufenacil, diuron, 2,4-D and glyphosate, respectively. Based on herbicide rate required to cause a 50% reduction in plant growth (GR50), the resistant index were 51.83, 14.10, 5.05, 3.96 and 32.90 for the same herbicides, respectively. Our results confirmed multiple resistance of Conyza sumatrensis from Paraná-Brazil to herbicides from five-mode of-action. This was the first report of Conyza sumatrensis resistant to 2,4-D and the first case of Conyza sumatrensis presenting multiple resistant to herbicides from five- mode of-action in the world.
APA, Harvard, Vancouver, ISO, and other styles
22

Peterson, Dallas E. "The Impact of Herbicide-Resistant Weeds on Kansas Agriculture." Weed Technology 13, no. 3 (September 1999): 632–35. http://dx.doi.org/10.1017/s0890037x00046315.

Full text
Abstract:
Herbicides are important components of weed management programs for most Kansas farmers. Monocropping systems and repeated use of the same or similar herbicides in some areas of the state have resulted in the development of herbicide-resistant weeds. The development of herbicide-resistant weed populations can have an immediate and a long-term effect on the cost, implementation, and effectiveness of weed control programs. In Kansas, resistance to triazine herbicides has been confirmed in kochia (Kochia scoparia), redroot pigweed, common waterhemp (Amaranthus rudis), Palmer amaranth (Amaranthus palmeri), and downy brome (Bromus tectorum) populations, and resistance to acetolactate synthase (ALS)-inhibiting herbicides has been confirmed in kochia, Russian thistle (Salsola kali), common waterhemp, Palmer amaranth, common cocklebur (Xanthium strumarium), shattercane (Sorghum bicolor), and common sunflower (Helianthus annum). The frequency and distribution of herbicide resistance varies among species. Producers who experience herbicide resistance problems adjust their weed control program accordingly. Producers that have not encountered an herbicide resistance problem tend to continue with a successful herbicide program until it fails. The recommended management strategies for herbicide-resistant weed populations include an integrated system of crop rotation, rotation of herbicide modes of action, tank-mixes of herbicides with different modes of action, and cultivation. The greatest direct cost to the producer occurs during the first year of poor weed control. The first response to an herbicide failure often is to reapply the same herbicide that has worked well previously. By the time the producer realizes that the treatment is not going to work, it usually is too late for any other remedial action. Consequently, the farmer experiences reduced crop production from weed competition, high herbicide costs, and a tremendous increase in the seed bank. The increase in seed bank may cost the farmer the most in the long run because the increased weed pressure often requires an intensified control program for several years.
APA, Harvard, Vancouver, ISO, and other styles
23

Vrbničanin, Sava. "Weed resistance to herbicides." Acta herbologica 29, no. 2 (2020): 79–96. http://dx.doi.org/10.5937/actaherb2002079v.

Full text
Abstract:
Weed resistance to herbicides represents the acquired resistance of individuals to complete the life cycle and leave offspring in the conditions of extended exposure to the same herbicide, i.e. herbicides of the same mechanism of action to which they were sensitive at the beginning of the application. Based on the herbicide resistance mechanisms, all processes can be grouped as follows: target-site resistance, non-target-site resistance, cross-resistance and multiple-resistance. Currently, herbicide resistance has been reported in 514 cases (species x site of action) worldwide, in 262 weed species (152 dicotyledons, 110 monocotyledons). Many of those biotypes are resistant to als inhibitors, PS II inhibitors, EPSPS inhibitors and ACC-ase inhibitors. The higher degree of resistance to als inhibitors has been confirmed in the following weed species: Amaranthus retroflexus, Sorghum halepense, Ambrosia artemisiifolia and Helianthus annuus.
APA, Harvard, Vancouver, ISO, and other styles
24

Beckie, Hugh J. "Herbicide-Resistant Weeds: Management Tactics and Practices." Weed Technology 20, no. 3 (September 2006): 793–814. http://dx.doi.org/10.1614/wt-05-084r1.1.

Full text
Abstract:
In input-intensive cropping systems around the world, farmers rarely proactively manage weeds to prevent or delay the selection for herbicide resistance. Farmers usually increase the adoption of integrated weed management practices only after herbicide resistance has evolved, although herbicides continue to be the dominant method of weed control. Intergroup herbicide resistance in various weed species has been the main impetus for changes in management practices and adoption of cropping systems that reduce selection for resistance. The effectiveness and adoption of herbicide and nonherbicide tactics and practices for the proactive and reactive management of herbicide-resistant (HR) weeds are reviewed. Herbicide tactics include sequences and rotations, mixtures, application rates, site-specific application, and use of HR crops. Nonherbicide weed-management practices or nonselective herbicides applied preplant or in crop, integrated with less-frequent selective herbicide use in diversified cropping systems, have mitigated the evolution, spread, and economic impact of HR weeds.
APA, Harvard, Vancouver, ISO, and other styles
25

Rauch, Traci A., Donald C. Thill, Seth A. Gersdorf, and William J. Price. "Widespread Occurrence of Herbicide-Resistant Italian Ryegrass (Lolium multiflorum) in Northern Idaho and Eastern Washington." Weed Technology 24, no. 3 (September 2010): 281–88. http://dx.doi.org/10.1614/wt-d-09-00059.1.

Full text
Abstract:
Persistent use of herbicides has resulted in the selection of many herbicide-resistant weeds worldwide. A survey of 75 fields in the Palouse region of the inland Pacific Northwest was conducted to determine the extent of Italian ryegrass resistance to grass herbicides commonly used in winter wheat-cropping systems. Plants grown from collected seed samples were tested for resistance to diclofop, clodinafop, quizalofop, tralkoxydim, sethoxydim, clethodim, pinoxaden, triasulfuron, mesosulfuron, flucarbazone, imazamox, and flufenacet/metribuzin. Averaged across herbicide families within a herbicide group, some level of resistance was exhibited in 73, 31, and 31% of the populations to the aryloxyphenoxypropionates, cyclohexanediones, and phenylpyrazoline herbicides, respectively, and 39, 53, and 55% of the populations to the sulfonylureas, sulfonylaminocarbonyltriazolinone, and imidazolinone herbicides, respectively. Twelve percent of the populations showed some level of resistance to flufenacet/metribuzin. Cross-resistance to all acetyl coenzyme A carboxylase-inhibiting (group 1) herbicides was observed in 12% of the populations, whereas 25% of the populations were cross-resistant to all acetolactate synthase-inhibiting (group 2) herbicides tested. Of all the populations tested, 7% exhibited multiple resistance to at least one herbicide within all three groups tested. Only 5% of populations were completely susceptible to all 12 herbicides tested. These results indicate that herbicide-resistant Italian ryegrass populations are now common across much of the Palouse region in northern Idaho and eastern Washington.
APA, Harvard, Vancouver, ISO, and other styles
26

Délye, Christophe, Arnaud Duhoux, Fanny Pernin, Chance W. Riggins, and Patrick J. Tranel. "Molecular Mechanisms of Herbicide Resistance." Weed Science 63, SP1 (February 2015): 91–115. http://dx.doi.org/10.1614/ws-d-13-00096.1.

Full text
Abstract:
Resistance to herbicides occurs in weeds as the result of evolutionary adaptation (Jasieniuk et al. 1996). Basically, two types of mechanisms are involved in resistance (Beckie and Tardif 2012; Délye 2013). Target-site resistance (TSR) is caused by changes in the tridimensional structure of the herbicide target protein that decrease herbicide binding, or by increased activity (e.g., due to increased expression or increased intrinsic activity) of the target protein. Nontarget-site resistance (NTSR) is endowed by any mechanism not belonging to TSR, e.g., reduction in herbicide uptake or translocation in the plant, or enhanced herbicide detoxification (reviewed in Délye 2013; Yuan et al. 2007).
APA, Harvard, Vancouver, ISO, and other styles
27

Broster, J. C., and J. E. Pratley. "A decade of monitoring herbicide resistance in Lolium rigidum in Australia." Australian Journal of Experimental Agriculture 46, no. 9 (2006): 1151. http://dx.doi.org/10.1071/ea04254.

Full text
Abstract:
Charles Sturt University commenced herbicide resistance monitoring in 1991. A random survey in 1991 to determine the level of resistance in annual ryegrass (Lolium rigidum) to selective herbicides across the south-west slopes region of New South Wales found that 30% of samples were resistant to at least 1 herbicide. A subsequent survey of commercially available ryegrass seed found that 58% of these samples were resistant to at least 1 herbicide. As a result of these findings, a commercial testing service was established and has since received samples from a large proportion of the southern Australian cropping belt. Seventy-seven percent of samples tested were resistant to Group AI, 40% to Group B and 22% to Group AII herbicides. Lower levels of resistance were found to Group D (8%), Group C (1%) and Group M (0.4%) herbicides. The correlation between resistance in Group AI and AII herbicides was lower than expected given that these herbicides are considered to have the same mode of action. Within the Group AI herbicides the observed response of the samples was consistent across herbicide formulations. Resistance to clethodim varied from observed responses to other Group AII herbicides. The variation in resistance levels (and degree of multiple resistance) in each Australian state is discussed in relation to environmental conditions and cultural practices. The size of this dataset allows for the analysis of the relationships present among herbicide resistant annual ryegrass.
APA, Harvard, Vancouver, ISO, and other styles
28

Paril, Jefferson F., and Alexandre J. Fournier-Level. "instaGraminoid, a Novel Colorimetric Method to Assess Herbicide Resistance, Identifies Patterns of Cross-Resistance in Annual Ryegrass." Plant Phenomics 2019 (April 23, 2019): 1–13. http://dx.doi.org/10.34133/2019/7937156.

Full text
Abstract:
Herbicide resistance in agricultural weeds is a global problem with an increasing understanding that it is caused by multiple genes leading to quantitative resistance. These quantitative patterns of resistance are not easy to decipher with mortality assays alone, and there is a need for straightforward and unbiased protocols to accurately assess quantitative herbicide resistance. instaGraminoid—a computer vision and statistical analysis package—was developed as an automated and scalable method for quantifying herbicide resistance. The package was tested in rigid ryegrass (Lolium rigidum), the most noxious and highly resistant weed in Australia and the Mediterranean region. This method provides quantitative measures of the degree of chlorosis and necrosis of individual plants which was shown to accurately reflect herbicide resistance. We were able to reliably characterise resistance to four herbicides with different sites of action (glyphosate, sulfometuron, terbuthylazine, and trifluralin) in two L. rigidum populations from Southeast Australia. Cross-validation of the method across populations and herbicide treatments showed high repeatability and transferability. Significant positive correlations in resistance of individual plants were observed across herbicides, which suggest either the accumulation of herbicide-specific resistance alleles in single genotypes (multiple stacked resistance) or the presence of general broad-effects resistance alleles (cross-resistance). We used these quantitative estimates of cross-resistance to simulate how resistance development under an herbicide rotation strategy is likely to be higher than expected.
APA, Harvard, Vancouver, ISO, and other styles
29

Gherekhloo, Javid, Mostafa Oveisi, Eskandar Zand, and Rafael De Prado. "A Review of Herbicide Resistance in Iran." Weed Science 64, no. 4 (December 2016): 551–61. http://dx.doi.org/10.1614/ws-d-15-00139.1.

Full text
Abstract:
Continuous use of herbicides has triggered a phenomenon called herbicide resistance. Nowadays, herbicide resistance is a worldwide problem that threatens sustainable agriculture. A study of over a decade on herbicides in Iran has revealed that herbicide resistance has been occurring since 2004 in some weed species. Almost all the results of these studies have been published in national scientific journals and in conference proceedings on the subject. In the current review, studies on herbicide resistance in Iran were included to provide a perspective of developing weed resistance to herbicides for international scientists. More than 70% of arable land in Iran is given over to cultivation of wheat, barley, and rice; wheat alone covers nearly 52%. Within the past 40 years, 108 herbicides from different groups of modes of action have been registered in Iran, of which 28 are for the selective control of weeds in wheat and barley. Major resistance to ACCase-inhibiting herbicides has been shown in some weed species, such as winter wild oat, wild oat, littleseed canarygrass, hood canarygrass, and rigid ryegrass. With respect to the broad area of wheat crop production and continuous use of herbicides with the sole mechanism of action of ACCase inhibition, the provinces of West Azerbaijan, Tehran, Khorasan, Isfahan, Markazi, and Semnan are at risk of resistance development. In addition, because of continuous long-term use of tribenuron-methyl, resistance in broadleaf species is also being developed. Evidence has recently shown resistance of turnipweed and wild mustard populations to this herbicide. Stable monitoring of fields in doubtful areas and providing good education and training for technicians and farmers to practice integrated methods would help to prevent or delay the development of resistance to herbicides.
APA, Harvard, Vancouver, ISO, and other styles
30

Liu, Xiangying, Shihai Xiang, Tao Zong, Guolan Ma, Lamei Wu, Kailin Liu, Xuguo Zhou, and Lianyang Bai. "Herbicide resistance in China: a quantitative review." Weed Science 67, no. 6 (August 23, 2019): 605–12. http://dx.doi.org/10.1017/wsc.2019.46.

Full text
Abstract:
AbstractThe widespread, rapid evolution of herbicide-resistant weeds is a serious and escalating agronomic problem worldwide. During China’s economic boom, the country became one of the most important herbicide producers and consumers in the world, and herbicide resistance has dramatically increased in the past decade and has become a serious threat to agriculture. Here, following an evidence-based PRISMA (preferred reporting items for systematic reviews and meta-analyses) approach, we carried out a systematic review to quantitatively assess herbicide resistance in China. Multiple weed species, including 26, 18, 11, 9, 5, 5, 4, and 3 species in rice (Oryza sativa L.), wheat (Triticum aestivum L.), soybean [Glycine max (L.) Merr.], corn (Zea mays L.), canola (Brassica napus L.), cotton (Gossypium hirsutum L.)., orchards, and peanut (Arachis hypogaea L.) fields, respectively, have developed herbicide resistance. Acetolactate synthase inhibitors, acetyl-CoA carboxylase inhibitors, and synthetic auxin herbicides are the most resistance-prone herbicides and are the most frequently used mechanisms of action, followed by 5-enolpyruvylshikimate-3-phosphate synthase inhibitors and protoporphyrinogen oxidase inhibitors. The lack of alternative herbicides to manage weeds that exhibit cross-resistance or multiple resistance (or both) is an emerging issue and poses one of the greatest threats challenging the crop production and food safety both in China and globally.
APA, Harvard, Vancouver, ISO, and other styles
31

Ngow, Zachary, Trevor K. James, and Christopher E. Buddenhagen. "A herbicide resistance risk assessment for weeds in maize in New Zealand." New Zealand Plant Protection 74, no. 1 (November 23, 2021): 78–86. http://dx.doi.org/10.30843/nzpp.2021.74.11738.

Full text
Abstract:
Despite an extensive history of research into herbicide resistance in New Zealand maize, some aspects remain understudied. Herbicide resistance was first detected in New Zealand in the 1980s in maize crops, with atrazine resistance in Chenopodium album L. and Persicaria maculosa Gray. Since then, Chenopodium album has also developed resistance to dicamba, and in the last five years Digitaria sanguinalis (L.) Scop. populations have been reported to be resistant to nicosulfuron. Here we estimate the risk of herbicide resistance arising in 39 common maize weeds. A list of weeds associated with maize was generated, omitting uncommon weeds and those that grow outside of the maize growing season. Weeds were ranked for their risk of evolving herbicide resistance with a scoring protocol that accounts for the specific herbicides used in New Zealand maize. Seven weed species were classified as having a high risk of developing herbicide resistance: Echinochloa crus-galli (L.) P.Beauv., Chenopodium album, Eleusine indica (L.) Gaertn., Xanthium strumarium L., Amaranthus powellii S.Watson, Solanum nigrum L. and Digitaria sanguinalis. Seventeen species were classed as moderate risk, and 15 were low risk. Herbicide classes associated with more resistant species were classed as high risk,these included acetohydroxy acid synthase inhibitors and photosystem-II inhibitors. Synthetic auxins had a moderate risk but only two herbicides in this class (dicamba and clopyralid) are registered for maize in New Zealand. Other herbicide mode-of-action groups used in maize were low risk. We recommend outreach to farmers regarding weed-control strategies that prevent high-risk species from developing resistance. High-risk herbicide groups should be monitored for losses of efficacy. Resistance surveys should focus on these species and herbicides.
APA, Harvard, Vancouver, ISO, and other styles
32

Owen, Mechelle J., Neree J. Martinez, and Stephen B. Powles. "Herbicide resistance in Bromus and Hordeum spp. in the Western Australian grain belt." Crop and Pasture Science 66, no. 5 (2015): 466. http://dx.doi.org/10.1071/cp14293.

Full text
Abstract:
Random surveys conducted in the Western Australian (WA) grain belt have shown that herbicide-resistant Lolium rigidum and Raphanus raphanistrum are a widespread problem across the cropping region. In 2010, a random survey was conducted to establish the levels of herbicide resistance for common weed species in crop fields, including the minor but emerging weeds Bromus and Hordeum spp. This is the first random survey in WA to establish the frequency of herbicide resistance in these species. For the annual grass weed Bromus, 91 populations were collected, indicating that this species was present in >20% of fields. Nearly all populations were susceptible to the commonly used herbicides tested in this study; however, a small number of populations (13%) displayed resistance to the acetolactate synthase-inhibiting sulfonylurea herbicides. Only one population displayed resistance to the acetyl-coenzyme A carboxylase-inhibiting herbicides. Forty-seven Hordeum populations were collected from 10% of fields, with most populations being susceptible to all herbicides tested. Of the Hordeum populations, 8% were resistant to the sulfonylurea herbicide sulfosulfuron, some with cross-resistance to the imidazolinone herbicides. No resistance was found to glyphosate or paraquat, although resistance to these herbicides has been documented elsewhere in Australia for Hordeum spp. (Victoria) and Bromus spp. (Victoria, South Australia and WA).
APA, Harvard, Vancouver, ISO, and other styles
33

Green, Jerry M. "Evolution of Glyphosate-Resistant Crop Technology." Weed Science 57, no. 1 (February 2009): 108–17. http://dx.doi.org/10.1614/ws-08-030.1.

Full text
Abstract:
New and improved glyphosate-resistant (GR) crops continue to be rapidly developed. These crops confer greater crop safety to multiple glyphosate applications, higher rates, and wider application timings. Many of these crops will also have glyphosate resistance stacked with traits that confer resistance to herbicides with other modes of actions to expand the utility of existing herbicides and to increase the number of mixture options that can delay the evolution of GR weeds. Some breeding stacks of herbicide resistance traits are currently available, but the trend in the future will be to combine resistance genes in molecular stacks. The first example of such a molecular stack has a new metabolically based mechanism to inactivate glyphosate combined with an active site-based resistance for herbicides that inhibit acetolactate synthase (ALS). This stack confers resistance to glyphosate and all five classes of ALS-inhibiting herbicides. Other molecular stacks will include glyphosate resistance with resistance to auxin herbicides and herbicides that inhibit acetyl coenzyme A carboxylase (ACCase) and 4-hydroxyphenyl pyruvate dioxygenase (HPPD). Scientists are also studying a number of other herbicide resistance transgenes. Some of these new transgenes will be used to make new multiple herbicide-resistant crops that offer growers more herbicide options to meet their changing weed management needs and to help sustain the efficacy of glyphosate.
APA, Harvard, Vancouver, ISO, and other styles
34

Owen, Mechelle J., and Stephen B. Powles. "Distribution and frequency of herbicide-resistant wild oat (Avena spp.) across the Western Australian grain belt." Crop and Pasture Science 60, no. 1 (2009): 25. http://dx.doi.org/10.1071/cp08178.

Full text
Abstract:
In 2005, a random survey was conducted across 14 million hectares of the Western Australian grain belt to establish the frequency and distribution of herbicide-resistant wild oat (Avena spp.) in cropping fields. In total, 677 cropping fields were visited, with wild oat populations collected from 150 fields. These wild oat populations were screened with several herbicides commonly used to control this weed. Most of the wild oat populations (71%) were found to contain individuals resistant to the ACCase-inhibiting herbicide diclofop-methyl. Resistance to other ACCase-inhibiting herbicides was markedly lower. Herbicides of alternative modes of action were effective on all wild oat populations. Overall, wild oat resistance to diclofop-methyl was found to be widespread across the Western Australian grain belt, but resistance to other herbicides was relatively low. Therefore, through diversity in herbicide use and with cultural management, it is possible to maintain wild oat populations at a low level and/or minimise herbicide resistance evolution.
APA, Harvard, Vancouver, ISO, and other styles
35

Zhu, Guangtao, Hao Wang, Haitao Gao, Ying Liu, Jun Li, Zhike Feng, and Liyao Dong. "Multiple Resistance to Three Modes of Action of Herbicides in a Single Italian Ryegrass (Lolium multiflorum L.) Population in China." Agronomy 13, no. 1 (January 10, 2023): 216. http://dx.doi.org/10.3390/agronomy13010216.

Full text
Abstract:
Italian ryegrass (Lolium multiflorum L.), a cross-pollinated grass, is gradually becoming a predominant weed in wheat fields in China and is evolving resistance to many groups of herbicides. The aim of this study is to determine the resistance levels of a single L. multiflorum population from a wheat field in Henan Province China, to three modes of action (MoAs) of herbicides and to further characterize the potential resistance mechanisms. This L. multiflorum population evolved multiple herbicide resistances to pyroxsulam [acetolactate synthase (ALS)], pinoxaden [acetyl-CoA carboxylase (ACCase)] and isoproturon [photosystem II (PSII)]. Target-site resistance (TSR) mutations (Pro-197-Gln, Pro-197-Thr, and Trp-574-Leu) and non-target-site resistance (NTSR) mediated by cytochrome P450 monooxygenase (CYP450) genes were associated with pyroxsulam resistance. Pinoxaden resistance was conferred by two TSR mutations, which referred to a rare Ile-2041-Val mutation and a common Ile-1781-Leu mutation but with two different nucleotide substitutions (CTA/TTA). CYP450- and glutathione-S-transferase (GST)-mediated resistances were the main resistance mechanisms for this multiple herbicide-resistant (MHR) population to the PSII inhibitor isoproturon. This is the first case of a single L. multiflorum population evolving multiple resistance to three herbicide MoAs (ALS, ACCase and PSII) in China. Diverse resistance mechanisms including TSR and NTSR mean L. multiflorum exhibits a high degree of resistance plasticity.
APA, Harvard, Vancouver, ISO, and other styles
36

Wyse, Donald L. "Future Impact of Crops with Modified Herbicide Resistance." Weed Technology 6, no. 3 (September 1992): 665–68. http://dx.doi.org/10.1017/s0890037x00036009.

Full text
Abstract:
The development of crop cultivars with resistance to selected herbicides has the potential to impact environmental quality, food safety, consumers, and crop producers in either a positive or negative manner. The technology that makes it possible to develop herbicide-resistant crops is neither good nor bad, it is rather how the products of this technology are used that will determine whether or not the introduction of herbicide-resistant crops is ultimately a good or bad decision. The introduction of herbicide-resistant crops will have diverse impacts leading to redundancy, diversity, and confusion in crop production systems. Often the introduction of herbicide-resistant cultivars will have the same impact on cropping systems as the introduction of a new herbicide that has the same mode-of-action and use pattern of herbicides already in use. This may add diversity of herbicide options for a given crop but will cause redundancy of product use over several years. This redundancy could lead to weed resistance and water quality concerns. Confusion at the user level will exist because not all cultivars of a crop will be resistant to the herbicide; this could be the major deterrent to widespread adoption of herbicide-resistant crops. Steps must be taken to provide information to crop producers that will insure that herbicide-resistant crops are used effectively and safely. Weed scientists will determine whether this technology will be used to improve food safety, water quality, crop production systems, and farmer profitability or have a negative impact on agriculture and the whole of society.
APA, Harvard, Vancouver, ISO, and other styles
37

Travlos, Ilias. "Evaluation of Herbicide-Resistance Status on Populations of Littleseed Canarygrass (Phalaris Minor Retz.) from Southern Greece and Suggestions for their Effective Control." Journal of Plant Protection Research 52, no. 3 (July 1, 2012): 308–13. http://dx.doi.org/10.2478/v10045-012-0050-3.

Full text
Abstract:
Evaluation of Herbicide-Resistance Status on Populations of Littleseed Canarygrass (Phalaris MinorRetz.) from Southern Greece and Suggestions for their Effective ControlIn 2010, a survey was conducted in the wheat fields of a typical cereal-producing region of Greece to establish the frequency and distribution of herbicide-resistant littleseed canarygrass (Phalaris minorRetz.). In total, 73 canarygrass accessions were collected and screened in a field experiment with several herbicides commonly used to control this weed. Most of the weed populations were classed as resistant (or developing resistance) to the acetyl-CoA varboxylase (ACCase)-inhibiting herbicide diclofop, while resistance to clodinafop was markedly lower. The results of the pot experiments showed that some of the canary populations were found to have a very high level of diclofop resistance (resistance index up to 12.4), while cross resistance with other herbicides was also common. The levels of resistance and cross resistance patterns among populations varied along with the different amounts and times of selection pressure. Such variation indicated either more than one mechanism of resistance or different resistance mutations in these weed populations. The population which had the highest diclofop resistance level, showed resistance to all aryloxyphenoxypropinate (APP) herbicides applied and non-ACCase inhibitors. Alternative ACCase-inhibiting herbicides, such as pinoxaden remain effective on the majority of the tested canarygrass populations, while the acetolactate synthase (ALS)-inhibiting herbicide mesosulfuron + iodosulfuron could also provide some solutions. Consequently, there is an opportunity to effectively control canarygrass by selecting from a wide range of herbicides. It is the integration of agronomic practices with herbicide application, which helps in effective management ofP. minorand particularly its resistant populations.
APA, Harvard, Vancouver, ISO, and other styles
38

Powles, Stephen B., and Todd A. Gaines. "Exploring the Potential for a Regulatory Change to Encourage Diversity in Herbicide Use." Weed Science 64, SP1 (September 2016): 649–54. http://dx.doi.org/10.1614/ws-d-15-00070.1.

Full text
Abstract:
An overreliance on herbicides in several important grain- and cotton-producing regions of the world has led to the widespread evolution of herbicide-resistant weed populations. Of particular concern are weed populations that exhibit simultaneous resistance to multiple herbicides (MHR). Too often, herbicides are the only tool used for weed control. We use the term herbicide-only syndrome (HOS) for this quasi-addiction to herbicides. Growers and their advisers focus on herbicide technology, unaware of or ignoring basic evolutionary principles or the necessary diversity provided by other methods of weed control. Diversity in weed control practices disrupts resistance evolution. Significant challenges exist to implementing diversity, including how to address information so that producers choose to alter existing behaviors (HOS) and take calculated risks by attempting new and more complex strategies. Herbicide resistance management in the long term will require creativity in many sectors, including roles for growers, industry, researchers, consultants, retailers, and regulators. There can be creativity in herbicide registration and regulation, as exemplified by the recent U.S. Environmental Protection Agency program that encourages herbicide registrants to register products in minor crops. We propose one idea for a regulatory incentive to enable herbicide registrants in jurisdictions such as the United States to receive an extended data exclusivity period in exchange for not developing one new herbicide in multiple crops used together in rotation, or for implementing stewardship practices such as robust mixtures or limitations on application frequency. This incentive would provide a mechanism to register herbicides in ways that help to ensure herbicide longevity. Approaches based only on market or financial incentives have contributed to the current situation of widespread MHR. Our suggestion for regulatory creativity is one way to provide both financial and biological benefits to the registering company and to the overall stakeholder community by incentivizing good resistance management.
APA, Harvard, Vancouver, ISO, and other styles
39

Patton, Aaron J., Daniel V. Weisenberger, and Geoff P. Schortgen. "2,4-D–Resistant Buckhorn Plantain (Plantago lanceolata) in Managed Turf." Weed Technology 32, no. 2 (February 1, 2018): 182–89. http://dx.doi.org/10.1017/wet.2017.98.

Full text
Abstract:
AbstractA population of buckhorn plantain with suspected resistance to 2,4-D was identified in central Indiana following 30 yr of 2,4-D–containing herbicide applications. Our objectives were to (1) confirm and quantify the level of herbicide resistance in the buckhorn plantain population using dose–response experiments and (2) find alternative herbicides that could be used to control this population. Greenhouse experiments were conducted to quantify the dose–response of resistant (R) and susceptible (S) biotypes of buckhorn plantain to both 2,4-D and triclopyr, two synthetic auxin herbicides from different chemical families. The R biotype was ≥6.2 times less sensitive to 2,4-D than the S biotype. The efficacy of triclopyr was similar on both the R and S biotypes of buckhorn plantain, suggesting the absence of cross-resistance to this herbicide. This is the first report of 2,4-D resistance in buckhorn plantain and the first report of 2,4-D resistance in turf. The resistance mechanism was limited to within a chemical family (phenoxycarboxylic acid) and did not occur across all WSSA Group 4 synthetic auxin herbicides, as the pyridinecarboxylic acid herbicides clopyralid and triclopyr and the arylpicolinate herbicide halauxifen-methyl provided control in our experiments.
APA, Harvard, Vancouver, ISO, and other styles
40

Dyer, W. E. "Herbicide-resistant crops : A weed scientist’s perspective." Comptes rendus 75, no. 4 (April 12, 2005): 71–77. http://dx.doi.org/10.7202/706073ar.

Full text
Abstract:
Herbicide-resistant crops offer a potentially valuable alternative strategy for weed management. If used appropriately, they may promote the use of agrichemicals more environmentally benign than the herbicides they replace, and provide producers with additional tools for controlling weeds. However, the controversy surrounding the development and use of these cultivars may limit and eventually prevent their widespread adoption. Concerns include: overuse of herbicides, escape of herbicide resistance genes from resistant cultivars into weedy relatives, genetic modifications for resistance conferring weediness to the cultivar (i.e. volunteer plants in subsequent crops), potential pleiotropic effects of genetic modifications for resistance, and selection of new herbicide-resistant weeds in the new herbicide regime. Of these concerns, the potential for selecting new resistant weeds may have the highest likelihood of affecting the long-term success of herbicide-resistant crops.
APA, Harvard, Vancouver, ISO, and other styles
41

Derr, Jeffrey F., Joseph C. Neal, and Prasanta C. Bhowmik. "Herbicide resistance in the nursery crop production and landscape maintenance industries." Weed Technology 34, no. 3 (June 2020): 437–46. http://dx.doi.org/10.1017/wet.2020.40.

Full text
Abstract:
AbstractWeed management is an important issue for nursery crop and Christmas tree producers, as well as for those maintaining turfgrass or ornamental species in landscape plantings. PRE and POST herbicides are important weed management tools for these industries. Reports of herbicide-resistant weeds increased from fewer than 100 cases in 1985 to nearly 500 cases globally in 2019, including ones found in turfgrass or ornamental systems. The evolution, persistence, and management of herbicide-resistant weeds are an ongoing educational process. We must keep our stakeholders aware of improved weed control technology and provide them information on resistant weeds. A symposium at the 2019 Weed Science Society of America meeting was conducted with presentations and discussions by invited speakers in relation to current research and potential management strategies for resistant weeds in turfgrass, landscape ornamental, and nursery crops. To prepare for the symposium, a survey was prepared for nursery producers and landscapers on the issues of herbicide-resistant weeds and offsite movement of herbicides used to control herbicide-resistant weeds. Overall, most respondents felt herbicide-resistant weeds are a serious problem and most had personally observed herbicide resistance on properties they maintain. Resistance to glyphosate was the herbicide cited by most respondents, followed by resistance to triazine herbicides. Most felt their weed-control costs had increased because of resistant weeds. Approximately 20% of respondents had their operation affected by drift of herbicides from nearby farm fields, with most reporting no damage from spray or vapor drift, but a few reported greater than 50% of the crop damaged.
APA, Harvard, Vancouver, ISO, and other styles
42

Harrison, Howard F. "Developing Herbicide-Tolerant Crop Cultivars: Introduction." Weed Technology 6, no. 3 (September 1992): 613–14. http://dx.doi.org/10.1017/s0890037x00035909.

Full text
Abstract:
In recent years considerable research in the private and public sectors has been directed toward introducing herbicide tolerance into normally susceptible crop species (9). Interest in developing herbicide-tolerant crop cultivars, clones, or hybrids (HTCs)3has been spurred by the reduction in the rate of discovery of new herbicidal compounds, the rising expense of developing new herbicides, and new tools of biotechnology that greatly increased our ability to develop HTC genotypes. Potential benefits of developing HTCs include: a) an increased margin of safety with which herbicides can be used with subsequent reduced crop losses due to herbicide injury, b) reduced risk of crop damage from residual herbicides from rotational crops, and c) introduction of new herbicides for use on normally susceptible crops. The last objective can be considered to be similar to breeding for resistance to diseases or insects. The most serious weed problems for a crop can be solved by developing crop tolerance to herbicides that control the weeds. This approach is particularly promising for minor crops for which new herbicide development is essentially lacking. However, the reluctance of herbicide manufacturers to register their products for minor crops may impede this approach. By developing tolerance to nontoxic, nonpolluting herbicides that are suitable for conservation tillage, the negative environmental effects of weed control can be reduced.
APA, Harvard, Vancouver, ISO, and other styles
43

Stokłosa, Agnieszka, and Jacek Kieć. "The level of wild oat resistance to ACC-ase inhibitors in South-Eastern Poland." Acta Agrobotanica 59, no. 2 (2012): 263–74. http://dx.doi.org/10.5586/aa.2006.082.

Full text
Abstract:
In the years 2003-2004 two experiments: laboratory and field, were carried out to detect the level of wild oat herbicide resistance. Weeds were collected from fields in south-eastern Poland, on the basis of survey conducted between farmers who have problems with wild oat herbicidal fighting. Two herbicides were tested: fenoxaprop- P-ethyl and diclofop methyl, both ACC-ase inhibitors. The laboratory test was established by modified Letouze et al. (1997) method. After 6-day growth on each herbicide water solution, the number of alive wild oat seedlings was assesed. In the field experiment plants in 3-5 leaf stage were sprayed with field dose of each herbicide. Two weeks after spraying the percentage of plants damage and in July the percentage of flowering plants were assesed. The wild oat plants that survive in laboratory test in ≥50% and plants that flowered in ≥50% in field conditions were recognized as resistant. On that basis wild oat resistace to fenoxaprop-P-ethyl and to diclofop methyl was stated on 17% and 28% fields from overall number of 18 fields, respectively.
APA, Harvard, Vancouver, ISO, and other styles
44

Holt, Jodie S., and Homer M. Lebaron. "Significance and Distribution of Herbicide Resistance." Weed Technology 4, no. 1 (March 1990): 141–49. http://dx.doi.org/10.1017/s0890037x00025148.

Full text
Abstract:
Herbicide-resistant weed species have become widespread in recent years. Fifty-five weed species, including 40 dicots and 15 grasses, are known to have biotypes resistant to the triazine herbicides. One or more resistant species have arisen in 31 states of the United States, four provinces of Canada, 18 countries in Europe, and Israel, Japan, Australia, and New Zealand. Resistance to other classes of herbicides is more restricted in distribution and recent in detection but is becoming more widespread. Trifluralin resistance has spread in the southeastern United States and has been detected in Canada, while 11 species with biotypes resistant to paraquat have been reported around the world. Diclofop-methyl-resistant weed species are problems in cereal production in Australia and have been found in Oregon, South Africa, and the United Kingdom. Resistance to the substituted ureas also is present in the United Kingdom, West Germany, and Hungary. Within the last 2 yr, biotypes of at least four weed species resistant to the sulfonylurea herbicides have arisen following several annual applications of these herbicides in wheat. Some resistant biotypes have multiple resistance to different classes of herbicides, which greatly exacerbates the threat of resistance. Herbicide resistance has reached the level where more concerted efforts are needed in research, education, and development of effective management strategies to preserve herbicides as essential tools of agricultural technology.
APA, Harvard, Vancouver, ISO, and other styles
45

Beckie, H. J., L. M. Hall, F. J. Tardif, and G. Séguin-Swartz. "Acetolactate synthase inhibitor-resistant stinkweed (Thlaspi arvense L.) in Alberta." Canadian Journal of Plant Science 87, no. 4 (October 1, 2007): 965–72. http://dx.doi.org/10.4141/cjps06019.

Full text
Abstract:
Two stinkweed populations from southern and central Alberta were not controlled by acetolactate synthase (ALS)-inhibiting herbicides in 2000. This study reports on their cross-resistance to ALS-inhibiting herbicides, molecular basis of resistance, and inheritance of resistance. Both putative herbicide-resistant biotypes responded similarly to increasing doses of the herbicides. The biotypes were highly resistant to ethametsulfuron and exhibited a low level of resistance to metsulfuron and imazethapyr. However, both biotypes were not resistant to florasulam, a triazolopyrimidine ALS inhibitor, or sulfometuron, a non-selective sulfonylurea ALS inhibitor. The cross-resistance pattern was consistent with the confirmed target-site mutation. Sequence analysis of the ALS gene detected a Pro197Leu mutation in both biotypes. Similar to many other ALS inhibitor-resistant weed biotypes, resistance was conferred by a single dominant gene. This study confirms the first global occurrence of herbicide resistance in this species. Key words: ALS-inhibitor resistance, ALS sequence, herbicide resistance, target-site mutation
APA, Harvard, Vancouver, ISO, and other styles
46

Förster, Britta, Peter B. Heifetz, Anita Lardans, John E. Boynton, and Nicholas W. Gillham. "Herbicide Resistance and Growth of D1 Ala251 Mutants in Chlamydomonas." Zeitschrift für Naturforschung C 52, no. 9-10 (October 1, 1997): 654–64. http://dx.doi.org/10.1515/znc-1997-9-1013.

Full text
Abstract:
We elucidated the effects of substituting seven amino acids for Ala at residue 251 of the Chlamydomonas reinhardtii D1 protein on herbicide resistance and photoautotrophic growth. Ala251 has been suggested to play a key role in the structural integrity and function of the stromal loop between transmembrane helices IV and V of D1 and has previously been shown to affect resistance to “classical” PSII specific herbicides. Sensitive and rapid microtiter assays were employed to compare herbicide resistance and photoautotrophic growth in the various mutants. Substitution of Ala251 by Ile, Leu or Val conferred resistance to the PSII herbicides atrazine, bromacil and metribuzin but not to DCMU, and impaired photoautotrophic growth in high and low light. Compared to an otherwise isogenic wildtype strain, the lie and Val mutants exhibited nearly identical levels of herbicide resistance and reduced growth while the Leu mutant had even slower growth and higher levels of herbicide resistance. In contrast Cys, Pro, Ser and Gly mutants were phenotypically indistinguishable from wildtype in terms of herbicide sensitivity and photoautotrophic doubling times. Collectively the seven Ala251 mutations differed markedly from an Ala mutant (dr-1) at the well characterized Ser264 D1 residue in terms of herbicide resistance and photoautotrophic growth
APA, Harvard, Vancouver, ISO, and other styles
47

Brenly-Bultemeier, Traci L., Jeff Stachler, and S. Kent Harrison. "Confirmation of Shattercane (Sorghum bicolor) Resistance to ALS-Inhibiting Herbicides in Ohio." Plant Health Progress 3, no. 1 (January 2002): 2. http://dx.doi.org/10.1094/php-2002-1021-01-rs.

Full text
Abstract:
A population of shattercane (Sorghum bicolor (L.) Moench) located in Fairfield County, Ohio, was investigated for herbicide resistance after it persisted in a field that had been treated repeatedly with herbicides that inhibit acetolactate synthase (ALS). Herbicide bioassays confirmed cross-resistance of the suspected resistant (R) population to the ALS inhibitors nicosulfuron, primisulfuron, and imazethapyr. Herbicide doses required to reduce R shattercane shoot dry weight 50% (i.e., the GR50 values) were > 35,000, > 40,000, and 34,215 g ai/ha for nicosulfuron, primisulfuron, and imazethapyr, respectively. In contrast, GR50 values for the same herbicides applied to a susceptible (S) shattercane population from an adjacent county were 0.185, 0.025, and 0.038 g/ha, respectively. The high levels of resistance exhibited by the R population suggest that the resistance mechanism is due to one or more alterations in ALS, the herbicide target site. Effective management of ALS herbicide-resistant shattercane will require an integrated strategy designed to isolate the R population and deplete its soil seed bank while minimizing herbicide selection pressure. Accepted for publication 14 October 2002. Published 21 October 2002.
APA, Harvard, Vancouver, ISO, and other styles
48

Wuerffel, R. Joseph, Julie M. Young, Joseph L. Matthews, and Bryan G. Young. "Characterization of PPO-Inhibitor–Resistant Waterhemp (Amaranthus tuberculatus) Response to Soil-Applied PPO-Inhibiting Herbicides." Weed Science 63, no. 2 (June 2015): 511–21. http://dx.doi.org/10.1614/ws-d-14-00108.1.

Full text
Abstract:
Waterhemp resistance to foliar applications of protoporphyrinogen oxidase (PPO)–inhibiting herbicides has become increasingly disconcerting given the widespread distribution of glyphosate resistance. Fortunately, soil-residual PPO-inhibiting herbicides remain efficacious in waterhemp populations resistant to PPO-inhibiting herbicides; however, these herbicides should theoretically select for the resistant biotype as herbicide concentrations diminish in the soil. Accordingly, the objectives of this research were twofold: (1) evaluate the efficacy of three PPO-inhibiting herbicides, foliar- and soil-applied, on PPO-resistant (PPO-R) and PPO-susceptible (PPO-S) waterhemp, and (2) investigate the differential effects of PPO-inhibiting herbicides on an R biotype and an S biotype during several discrete developmental events relevant to soil–residual herbicide activity (i.e., radicle protrusion, radicle elongation, and waterhemp emergence). Greenhouse and growth chamber experiments indicated that the R biotype was least sensitive to the diphenylether herbicide fomesafen, followed by sulfentrazone and flumioxazin; however, fomesafen pluss-metolachlor improved soil-residual efficacy over fomesafen alone. Growth stage considerably influenced the R : S ratio, decreasing from 38× to 3.4×, when comparing ratios generated from foliar applications and soil-residual applications measuring radicle protrusion, respectively. Overall, this research supports the use of full soil-residual herbicide rates, reinforcing the importance of best management practices to manage the spread of herbicide resistance.
APA, Harvard, Vancouver, ISO, and other styles
49

Stankiewicz-Kosyl, Marta, Agnieszka Synowiec, Małgorzata Haliniarz, Anna Wenda-Piesik, Krzysztof Domaradzki, Danuta Parylak, Mariola Wrochna, et al. "Herbicide Resistance and Management Options of Papaver rhoeas L. and Centaurea cyanus L. in Europe: A Review." Agronomy 10, no. 6 (June 18, 2020): 874. http://dx.doi.org/10.3390/agronomy10060874.

Full text
Abstract:
Corn poppy (Papaver rhoeas L.) and cornflower (Centaurea cyanus L.) are two overwintering weed species found in crop fields in Europe. They are characterised by a similar life cycle, similar competitive efforts, and a spectrum of herbicides recommended for their control. This review summarises the biology and herbicide resistance phenomena of corn poppy and cornflower in Europe. Corn poppy is one of the most dangerous dicotyledonous weeds, having developed herbicide resistance to acetolactate synthase inhibitors and growth regulators, especially in Mediterranean countries and Great Britain. Target site resistance to acetolactate synthase inhibitors dominates among herbicide-resistant poppy biotypes. The importance of non-target site resistance to acetolactate synthase inhibitors in this species may be underestimated because non-target site resistance is very often associated with target site resistance. Cornflower, meanwhile, is increasingly rare in European agricultural landscapes, with acetolactate synthase inhibitors-resistant biotypes only listed in Poland. However, the mechanisms of cornflower herbicide resistance are not well recognised. Currently, herbicides mainly from acetolactate synthase and photosystem II inhibitors as well as from synthetic auxins groups are recommended for the control of both weeds. Integrated methods of management of both weeds, especially herbicide-resistant biotypes, continue to be underrepresented.
APA, Harvard, Vancouver, ISO, and other styles
50

Owen, Micheal D. K. "Diverse Approaches to Herbicide-Resistant Weed Management." Weed Science 64, SP1 (September 2016): 570–84. http://dx.doi.org/10.1614/ws-d-15-00117.1.

Full text
Abstract:
Herbicides have been the principal means of weed control in developed countries for approximately 50 yr because they are the most cost-effective method. Such general use of herbicides has resulted in weed resistance to herbicides, which continues to be a growing problem. Within the past decade, the evolution of resistance to the once-dominant herbicide glyphosate has resulted in major concerns about the future ability to control weeds in many crop systems. Moreover, many weed species have evolved resistance to multiple mechanisms of herbicide action. Given the dearth of new herbicides with novel mechanisms of action, it appears inevitable that weed management programs will need to be supplemented by the use of tactics other than herbicides. However, the inclusion of more diversity for weed management also introduces complexity, cost, and time constraints to current crop production systems. This paper describes broadly the considerations, opportunities, and constraints of diverse weed management tactics to address the burgeoning problems with herbicide resistance.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography