首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Littleseed canarygrass (Phalaris minor Retz.), a troublesome weed of wheat in India, has evolved multiple herbicide resistance across three modes of action: photosynthesis at the photosystem II site A, acetyl‐coA carboxylase (ACCase), and acetolactate synthase inhibition. The multiple herbicide‐resistant (MHR) populations had a low level of sulfosulfuron resistance but a high level of resistance to clodinafop and fenoxaprop (ACCase inhibitors). Some of the populations had GR50 (50% growth reduction) values for clodinafop that were 11.7‐fold greater than that of the most susceptible population. The clodinafop‐resistant populations also showed a higher level of cross‐resistance to fenoxaprop (fop group) but a low level of cross‐resistance to pinoxaden (den group). Although clodinafop and pinoxaden are from two different chemical families (fop and den groups), their same site of action is responsible for cross‐resistance behavior. The populations that were resistant to four groups of herbicides (phenylureas, sulfonylurea, aryloxyphenoxypropionate, and phenylpyrazolin) were susceptible to the triazine (metribuzin and terbutryn) and dinitroaniline (pendimethalin) herbicides. The P. minor populations that were resistant to the aryloxyphenoxypropionate and phenylurea herbicides were effectively controlled by the sulfonylurea herbicide, sulfosulfuron. In the fields infested with P. minor that was resistant to clodinafop, a sulfosulfuron application (25 g ha?1) increased the wheat yield by 99.2% over that achieved using the recommended rate of clodinafop (60 g ha?1). However, the evolution of multiple resistance against the four groups is a threat to wheat production. To prevent the spread of MHR P. minor populations, as well as the extension of multiple resistance to new chemicals, concerted efforts in developing and implementing a sound, integrated weed management program are needed. The integrated approach, consisting of crop and herbicide rotation with cultural and mechanical weed control tactics, should be considered as a long‐term resistance management strategy that will help to sustain wheat productivity and farmers' income.  相似文献   

2.
BACKGROUND: The repeated use of acetyl‐coenzyme A carboxylase (ACCase) inhibiting herbicides to control grass weeds has selected for resistance in Lolium spp. populations in Italy. The efficacy of pinoxaden, a recently marketed phenylpyrazoline herbicide, is of concern where resistance to ACCase inhibitors has already been ascertained. ACCase mutations associated with pinoxaden resistance were investigated, and the cross‐resistance pattern to clodinafop, haloxyfop, sethoxydim, clethodim and pinoxaden was established on homo/heterozygous plants for four mutant ACCase alleles. RESULTS: Seven different mutant ACCase alleles (1781‐Leu, 1999‐Leu, 2041‐Asn, 2041‐Val, 2078‐Gly, 2088‐Arg and 2096‐Ala) and 13 combinations with two types of mutation were detected in the pinoxaden‐resistant plants. The 1781‐Leu allele appears to confer a dominant resistance to pinoxaden, clodinafop, haloxyfop, sethoxydim and clethodim at 60 g AI ha?1. The 2041‐Asn and 2041‐Val alleles are associated with dominant or partially dominant resistance to FOPs, no substantial resistance to DIMs and a moderate resistance to pinoxaden. The 2088‐Arg allele endows a partially dominant resistance to clodinafop, sethoxydim and most likely to pinoxaden. In addition, non‐target‐site resistance mechanisms seem to be involved in pinoxaden resistance. CONCLUSION: Almost all the ACCase mutations selected in the field by other ACCase inhibitors are likely to confer resistance to pinoxaden. Although pinoxaden is sometimes able to control FOP‐resistant populations, it should not be considered as a sustainable ACCase resistance management tool. The presence of non‐ACCase‐based resistance mechanisms that could confer resistance to herbicides with different modes of action further complicates the resistance management strategies. Copyright © 2011 Society of Chemical Industry  相似文献   

3.
BACKGROUND: Repeated use of acetyl‐CoA carboxylase (ACCase) inhibitors, especially fenoxaprop and clodinafop, since the late 1980s has selected for resistance in Alopecurus myosuroides Huds. (black‐grass) in France. We investigated whether resistance to pinoxaden, a phenylpyrazoline ACCase inhibitor to be marketed in France, was present in French black‐grass populations. We investigated pinoxaden resistance conferred by five mutant ACCase isoforms. Using 84 French black‐grass field samples, we also compared the frequencies of other mechanisms endowing resistance to fenoxaprop, clodinafop or pinoxaden. RESULTS: ACCase mutant isoforms Leu‐1781, Gly‐2078 and, likely, Cys‐2027 conferred cross‐resistance to pinoxaden, while isoform Asn‐2041 possibly conferred moderate resistance. Other mechanisms of resistance to fenoxaprop, clodinafop and pinoxaden were detected in 99, 68 and 64% of the samples investigated, respectively. Cross‐ or multiple resistance to fenoxaprop or clodinafop and pinoxaden was not systematically observed, suggesting a diversity of mechanisms exist. CONCLUSION: Pinoxaden resistance was observed before pinoxaden release in France. Only a fraction of the mechanisms endowing fenoxaprop or clodinafop resistance also confer pinoxaden resistance. Pinoxaden resistance was likely mostly selected for by ACCase inhibitors, and, in some cases, possibly by herbicides with other modes of action. This illustrates the necessity to use metabolisable herbicides cautiously where black‐grass has evolved non‐target‐site‐based resistance. Copyright © 2009 Society of Chemical Industry  相似文献   

4.
L Pan  J Li  T Zhang  D Zhang  L Y Dong 《Weed Research》2015,55(6):609-620
Beckmannia syzigachne (American sloughgrass) is a competitive grass weed found in China. Fenoxaprop‐P‐ethyl is widely used for control of this species in China. Resistance to fenoxaprop‐P‐ethyl in B. syzigachne has been reported to be conferred by an isoleucine(Ile)‐1781‐leucine(Leu) substitution in the gene encoding the herbicide target, acetyl‐CoA carboxylase (ACCase). In this study, three mutations were detected by derived cleaved amplified polymorphic sequence (dCAPS) method in fenoxaprop‐P‐ethyl‐resistant B. syzigachne populations: Ile‐1781‐Leu in population JCWL‐R, Ile‐2041‐Asn in JCJT‐R and Gly‐2096‐Ala in JYJD‐R. The data indicated they were genetically homogeneous (homozygous mutant) at the ACCase locus. The use of cytochrome P450 inhibitors was shown to slightly reduce the GR50 value of fenoxaprop‐P‐ethyl‐resistant populations, from which we inferred a combination of target‐site resistance (TSR) and non‐target‐site resistance (NTSR) was involved in fenoxaprop‐P‐ethyl‐resistance. We characterised the cross‐resistance patterns to ACCase inhibitors in B. syzigachne. The plants in the JCWL‐R population were highly resistant to all tested APPs (aryloxyphen‐oxypropionates), sethoxydim and pinoxaden, and moderately resistant to clethodim. The plants in the JCJT‐R population were highly resistant to fluazifop‐P‐butyl, clodinafop‐propargyl, cyhalofop‐butyl, metamifop and pinoxaden; moderately resistant to haloxyfop‐R‐methyl, quizalofop‐P‐ethyl and sethoxydim; and sensitive to clethodim. The plants in the JYJD‐R population were highly resistant to clodinafop‐propargyl, metamifop and pinoxaden; moderately resistant to haloxyfop‐R‐methyl, cyhalofop‐butyl, quizalofop‐P‐ethyl, fluazifop‐P‐butyl and sethoxydim; and sensitive to clethodim. If resistance to ACCase inhibitors is present in B. syzigachne populations in the field, then our results indicate that clethodim should be used. While we demonstrated the cross‐resistance patterns of TSR resulting from three mutations in B. syzigachne, we also demonstrated that NTSR plays a role in resistance, which will complicate weed management.  相似文献   

5.
Wild barley (Hordeum spontaneum) is one of the most troublesome weed species in winter wheat (Triticum aestivum) in Iran. Two bioassay experiments were conducted in order to study the response of wild barley and wheat to different herbicides and to study the efficacy of pre‐emergence (PRE), postemergence (POST), and PRE followed by POST applications of sulfosulfuron on wild barely. Moreover, the degradation of sulfosulfuron was studied by liquid chromatography coupled with tandem mass spectrometry (LC‐MS/MS). The results showed that wild barley was highly tolerant to clodinafop‐propargyl and its dry weight was reduced by only 15%, compared to the control, at the recommended dose (64 g ai ha?1). Sulfosulfuron reduced the wild barley biomass by ≤50% at the highest dose (90 g ai ha?1) in the first bioassay but by not more than 20% and 12% at the recommended dose (22 g ai ha?1) in the first and second bioassay, respectively. Significant differences were found among the application methods of sulfosulfuron, with the POST application being the least effective method. In contrast to the POST application, wild barley was severely injured by the PRE application of sulfosulfuron, with an ED50 dose of 7.3 g ai ha?1. The degradation study showed that wild barley can metabolize sulfosulfuron that is applied POST, but at a lower rate than wheat. By 4 h after application, wild barley had metabolized 26% of the sulfosulfuron, compared to 46% by wheat. In conclusion, wild barley can metabolize the recommended dose of sulfosulfuron that is applied POST; thus, the PRE application of sulfosulfuron or other integrated methods should be considered for the effective control of wild barley in wheat.  相似文献   

6.
The intensive use of the acetolactate synthase (ALS)‐inhibiting herbicides, imazethapyr, penoxsulam and bispyribac‐sodium, in imidazolinone‐resistant (Clearfield) rice increases the risk of the evolution of ALS‐resistant barnyardgrass. In 2009, imazethapyr failed to control barnyardgrass that was collected from a field in Arkansas, USA, following the failure of the herbicide in 2008. A greenhouse experiment was conducted to confirm and document the level of resistance of the biotype against three ALS‐inhibiting herbicides that currently are labeled in rice. The level of control of the resistant biotype at the labeled rate of bispyribac‐sodium of 35 g ai ha?1 was 10%, penoxsulam at 22 g ai ha?1 was 0% and imazethapyr at 70 g ai ha?1 was 25%. The level of mortality of the susceptible biotype was 100% with all the herbicides at the labeled rate. The dose needed to kill 50% of the resistant plants was 49 g ha?1 of bispyribac‐sodium, 254 g ha?1 of penoxsulam and 170 g ha?1 of imazethapyr. For the susceptible biotype, bispyribac‐sodium at 6 g ha?1, penoxsulam at 10 g ha?1 and imazethapyr at 12 g ha?1 killed 50% of the treated plants. Based on these findings, it was confirmed that a barnyardgrass population has evolved cross‐resistance to three ALS‐inhibiting herbicides in rice culture in Arkansas. Furthermore, an experiment was conducted to determine if the ALS‐resistant biotype could be controlled using other mechanisms of action. The results indicated that propanil, a photosystem II inhibitor, and quinclorac, a synthetic auxin, failed to control the resistant biotype at the labeled rates, whereas all the other evaluated herbicides provided effective control of both biotypes.  相似文献   

7.
The level of resistance and patterns of cross-resistance to clodinafop, sethoxydim, and pinoxaden were examined in 12 putative resistant and one susceptible populations of winter wild oat (Avena ludoviciana) collected from Fars Province, in the southwest of Iran. The responses of biomass and length of coleoptiles to the increasing dosages of the three herbicides were determined in both whole-plant and seed bioassays. In the whole-plant bioassay, all 12 putative resistant populations were found to be resistant to clodinafop with resistance ratios (R/S) ranging from 1.76 to >47.04. Most clodinafop-resistant populations exhibited low levels of cross-resistance to sethoxydim. Three highly sethoxydim-resistant populations, F2, S2, and ES4, were slightly resistant to clodinafop. Six populations (M1, M2, F2, S2, S4, and ES4) showed high cross-resistance to pinoxaden with R/S values as large as 10.73 to 40.29. A highly clodinafop-resistant population, M2, was more sensitive to pinoxaden than the susceptible population. The results of the seed bioassay resembled those obtained from the whole-plant experiment suggesting seed bioassay as an inexpensive, rapid method for screening-resistant genotypes.  相似文献   

8.
The rapid range expansion of naturalized Italian ryegrass (Lolium multiflorum Lam.) in farmland is a serious problem in Fukuroi city in Shizuoka Prefecture, Japan. Glyphosate has been used to control Italian ryegrass in the levees of rice paddy fields and wheat fields for ~20 years, but this weed in Fukuroi city is poorly controlled by glyphosate. In order to elucidate the level of resistance to glyphosate in Italian ryegrass populations, seed bioassays and a foliar application experiment, using seeds collected from 16 wild populations in and around Fukuroi city and from three susceptible cultivars, were conducted. For the susceptible cultivars and one population from a site where glyphosate had not been applied for >10 years, the shoot length in the seed bioassays was greatly suppressed at a glyphosate concentration of 10 mg ai L?1 and no seedling survived after the foliar application of glyphosate at a rate of 2.3 kg ai ha?1. Nine wild populations from levees in the southern part of Fukuroi city showed vigorous shoot growth at a glyphosate concentration of 10 mg ai L?1 and had at least a 78% survival rate after the application of glyphosate at 2.3 kg ai ha?1. Four wild populations from levees in the northern part of Fukuroi city showed a slight suppression of the shoot growth as a result of the glyphosate treatment and their survival rates ranged from 20 to 64%. The results suggested that resistance to glyphosate has evolved in the wild populations of Italian ryegrass that are growing on the levees. This is the first report of a glyphosate‐resistant weed in Japan.  相似文献   

9.
Phalaris minor, the most serious weed in wheat in north‐western India, has developed extensive isoproturon resistance due to continuous isoproturon use. For its control, alternative herbicides (flufenacet, metribuzin and sulfosulfuron) at different application rates and timing were evaluated in wheat. In addition, herbicide carryover risk onto rotational crops (sorghum; maize and green gram, Vigina radiata) was also assessed. Isoproturon at 1 and 2 kg a.i. ha?1 provided only 10.5% and 51.8%P. minor control respectively. Of the other herbicides, early post‐emergent [15–21 days after sowing (DAS)] flufenacet at 180–480 g a.i. ha?1 provided acceptable control of P. minor, but failed to control broad‐leaved weeds and was phytotoxic to the wheat crop. Metribuzin at 210 g a.i. ha?1 was effective in controlling both Phalaris and dicotyledonous weeds. Mixtures of both flufenacet and metribuzin at reduced rates were better than flufenacet for weed control and grain yield. The efficacy of flufenacet and metribuzin was drastically reduced with later growth stages of P. minor (four to five leaf). Whereas sulfosulfuron at 25–30 g a.i. ha?1, applied either early post‐emergence (19 DAS) or post‐emergence (30–42 DAS), was quite effective. Overall, sulfosulfuron was the most effective treatment with regard to weed control and crop yield. However, maize and sorghum grown in rotation after harvest of sulfosulfuron‐treated wheat plots showed 65–73% crop biomass inhibition. The residual effect of sulfosulfuron was also noticed on Trianthema portulacastrum (Horse purslane), causing 73.5% dry matter reduction. By contrast, no carryover damage with flufenacet was observed on maize, sorghum and green gram. Glasshouse pot experiments and field trials investigating crop sensitivity to pre‐plant applications of sulfosulfuron found the decreasing order: sorghum > maize > green gram. The risk of carryover onto rotational crops should be considered when choosing alternative herbicides for P. minor control in wheat.  相似文献   

10.
Experiments were conducted to (i) evaluate the efficacy of propanil formulations available in Sri Lanka in controlling Echinochloa crus‐galli; (ii) study the seedling growth of propanil‐resistant (R) and ‐susceptible (S) biotypes of the weed under different temperatures; (iii) quantify the level of resistance in R biotypes and; (iv) to suggest alternative control measures for R biotypes. Field studies showed that retail propanil formulations (36% a.i., EC) applied at 2.7 kg a.i. ha?1 gave less than 30% control of E. crus‐galli collected from several locations of the north dry zone of Sri Lanka. Chemical analysis revealed that there was no adulteration of propanil formulations at the retailer level. Growth studies conducted in controlled environments indicated that per cent germination and seedling growth of R and S biotypes were similar at the day/night temperature regimes imposed. However, per cent germination for plants grown under a 34/31°C (day/night) regime was 27–29% higher compared to those grown at 28/24°C. At the higher temperature regime, R and S biotypes reached the 2–3 leaf stage five days earlier, and the 4–5 leaf stage seven days earlier. The ED50 values from the dose–response experiments indicated that the R biotype was four times more resistant to propanil than susceptible ones. The resistance index (RI) did not vary significantly under different temperature regimes. Quinclorac (25% a.i., SC) applied at 200 g a.i. ha?1 and bispyribac‐sodium (10% a.i., SC) applied at 30 g a.i. ha?1 (recommended dosages) successfully controlled propanil‐resistant biotypes of E. crus‐galli. Conversely, oxadiazon and propanil (8% and 23% a.i., EC, respectively) applied at 280 + 805 g a.i. ha?1 did not result in satisfactory control.  相似文献   

11.
In this study, the results of two-year investigations on herbicide resistance in silky bent grass (Apera spica-venti) populations are presented. Two populations of A. spica-venti were sown in a winter wheat field at the Ihinger Hof Research Station near Stuttgart in Germany in October 2008. Whole-plant bioassays conducted with both populations in the greenhouse before the field trial was set up revealed that population A was strongly resistant to acetolactate synthase (ALS)-inhibitors, whereas population B was sensitive to this group of herbicides. Each block was treated with isoproturon, fenoxaprop-P-ethyl, sulfosulfuron and meso-iodosulfuron at the recommended field dose. One treatment remained unsprayed. It was found that the average number of silky bent grass panicles was significantly higher in the second year (2010) with 343 panicles m? 2 in the untreated control compared to 44 panicles m? 2 in the first year (2009). Efficacy of both ALS-inhibitors was significantly reduced in the resistant population compared to the sensitive population. Grass-weed plants surviving treatments with ALS-inhibitors produced the same number of seeds as the untreated plants. It was found that germination rate of seeds from the resistant population was more than threefold higher than from the sensitive population. Grain yield was equal in all treatments and populations in 2009. In 2010, applications of isoproturon and fenoxaprop-P-ethyl resulted in higher grain yields, with a mean of 4.6 t ha? 1 compared to an average of 3.9 t ha? 1 in the plots treated with sulfosulfuron and meso-iodosulfuron and 4.3 t ha? 1 in the untreated plots. However, these differences were not statistically significant. The sensitive population of the second generation (2009/2010) was approximately 20% more tolerant to ALS-inhibitors than the sensitive population of the first generation (2008/2009) which indicates introgression of herbicide resistance traits already after one year. These results clearly show that herbicide resistance to ALS-inhibitors in silky bent grass is likely to spread rapidly causing significant economic losses. Therefore, management strategies need to be developed and tested to prevent and overcome herbicide resistance in European cereal production systems.  相似文献   

12.
The influence of application volume on the efficacy of clodinafop‐propargyl and fenoxaprop‐P‐ethyl on cultivated oats (Avena sativa) was studied in the glasshouse. Both herbicides were more efficacious when applied in 75 L ha?1 than in 300 L ha?1, with 11002 and 11006 nozzles respectively. However, when the same two volume rates were created by varying the speed of a 11003 nozzle, clodinafop‐propargyl efficacy was not affected by application volume, whereas fenoxaprop‐P‐ethyl remained more efficacious at 75 L ha?1. This suggests that in the first experiment, fenoxaprop‐P‐ethyl efficacy was affected by changes in both spray quality and concentration, whereas only the former influenced clodinafop‐propargyl efficacy. The hypothesis that the fenoxaprop‐P‐ethyl formulation was more influenced by concentration than clodinafop‐propargyl was supported by dynamic surface tension studies and measures of active ingredient retention by oat plants. Within the practically relevant concentration range considered, surface activity of clodinafop‐propargyl remained low, with little influence on herbicide retention. In contrast, depending on the concentration, surface activity of fenoxaprop‐P‐ethyl was below or above levels that were critical for its retention. Although these differences may not be as apparent in the field as in the glasshouse, our study certainly indicates that both clodinafop‐propargyl and fenoxaprop‐P‐ethyl are herbicides favoured by low application volumes.  相似文献   

13.
The reduction of herbicide applications is a main research priority in recent years. One way to achieve this goal is by using adjuvants that can increase the efficacy of foliar‐applied herbicides by reducing surface tension. Previous studies have shown that the surface tension of distilled water decreases under the influence of a magnetic field. In order to compare the effects of a magnetic field and Frigate in clodinafop‐propargyl and cycloxydim in controlling wild oat and evaluating the surface tension, a dose–response greenhouse experiment was conducted by using 0, 8, 16, 32, 48, and 64 g ai ha?1 of clodinafop‐propargyl and 0, 15, 30, 60, 90, and 120 g ai ha?1 of cycloxydim with Frigate and/or by passing them through a magnetic field. The magnetic field and Frigate caused a significant reduction in the surface tension of the herbicide solutions. But, Frigate was more effective in reducing the surface tension of the herbicide solutions, compared with the magnetic field. The magnetic field and Frigate increased the efficacy of clodinafop‐propargyl and cycloxydim remarkably. Frigate increased the efficacy of the herbicides more than did the magnetic field. The magnetic field and Frigate collectively had more of an effect on the herbicides' efficacy than when they were applied individually. The magnetic field and Frigate were more effective in increasing the efficacy of clodinafop‐propargyl than that of cycloxydim.  相似文献   

14.
Echinochloa colona is the most common grass weed of summer fallows in the grain‐cropping systems of the subtropical region of Australia. Glyphosate is the most commonly used herbicide for summer grass control in fallows in this region. The world's first population of glyphosate‐resistant E. colona was confirmed in Australia in 2007 and, since then, >70 populations have been confirmed to be resistant in the subtropical region. The efficacy of alternative herbicides on glyphosate‐susceptible populations was evaluated in three field experiments and on both glyphosate‐susceptible and glyphosate‐resistant populations in two pot experiments. The treatments were knockdown and pre‐emergence herbicides that were applied as a single application (alone or in a mixture) or as part of a sequential application to weeds at different growth stages. Glyphosate at 720 g ai ha?1 provided good control of small glyphosate‐susceptible plants (pre‐ to early tillering), but was not always effective on larger susceptible plants. Paraquat was effective and the most reliable when applied at 500 g ai ha?1 on small plants, irrespective of the glyphosate resistance status. The sequential application of glyphosate followed by paraquat provided 96–100% control across all experiments, irrespective of the growth stage, and the addition of metolachlor and metolachlor + atrazine to glyphosate or paraquat significantly reduced subsequent emergence. Herbicide treatments have been identified that provide excellent control of small E. colona plants, irrespective of their glyphosate resistance status. These tactics of knockdown herbicides, sequential applications and pre‐emergence herbicides should be incorporated into an integrated weed management strategy in order to greatly improve E. colona control, reduce seed production by the sprayed survivors and to minimize the risk of the further development of glyphosate resistance.  相似文献   

15.
八种除草剂对小麦田三种抗精噁唑禾草灵杂草的生物活性   总被引:3,自引:0,他引:3  
为筛选能有效防除抗精噁唑禾草灵杂草的除草剂,采用温室盆栽法测定菵草(敏感S_w、抗性R_w)、日本看麦娘(敏感S_r、抗性R_r)、耿氏硬草(敏感S_y、抗性R_y)对精噁唑禾草灵的抗性水平,并研究了8种除草剂对这3种抗性杂草的生物活性。结果显示,3种杂草抗性种群对精噁唑禾草灵的抗性指数均大于33.7,已达高抗水平。3种杂草抗性种群均对同类乙酰辅酶A羧化酶类抑制剂唑啉草酯和炔草酯产生了11.6~56.5倍不等的高水平抗性。对部分乙酰乳酸合成酶类抑制剂产生了2.0~4.8倍的低水平抗性,氟唑磺隆对3种杂草抗性种群防效较差,GR_(90)为67.31~114.39g(a.i.)/hm~2;啶磺草胺仅对Rr种群防效较好,GR_(90)为4.67 g(a.i.)/hm~2;甲基二磺隆对3种杂草抗性种群防效均较好,但对Rr种群存在抗性风险,已出现2.0倍低水平抗性;磺酰磺隆对Ry和Rr种群防效较好;丙苯磺隆对Rr种群防效好。细胞分裂抑制剂氟噻草胺对3种杂草抗性种群防效均最好,在田间推荐剂量120 g(a.i.)/hm~2下可达90%以上的防效。  相似文献   

16.
Biological characterization of sulfoxaflor, a novel insecticide   总被引:1,自引:0,他引:1  
BACKGROUND: The commercialization of new insecticides is important for ensuring that multiple effective product choices are available. In particular, new insecticides that exhibit high potency and lack insecticidal cross‐resistance are particularly useful in insecticide resistance management (IRM) programs. Sulfoxaflor possesses these characteristics and is the first compound under development from the novel sulfoxamine class of insecticides. RESULTS: In the laboratory, sulfoxaflor demonstrated high levels of insecticidal potency against a broad range of sap‐feeding insect species. The potency of sulfoxaflor was comparable with that of commercial products, including neonicotinoids, for the control of a wide range of aphids, whiteflies (Homoptera) and true bugs (Heteroptera). Sulfoxaflor performed equally well in the laboratory against both insecticide‐susceptible and insecticide‐resistant populations of sweetpotato whitefly, Bemisia tabaci Gennadius, and brown planthopper, Nilaparvata lugens (Stål), including populations resistant to the neonicotinoid insecticide imidacloprid. These laboratory efficacy trends were confirmed in field trials from multiple geographies and crops, and in populations of insects with histories of repeated exposure to insecticides. In particular, a sulfoxaflor use rate of 25 g ha?1 against cotton aphid (Aphis gossypii Glover) outperformed acetamiprid (25 g ha?1) and dicrotophos (560 g ha?1). Sulfoxaflor (50 g ha?1) provided a control of sweetpotato whitefly equivalent to that of acetamiprid (75 g ha?1) and imidacloprid (50 g ha?1) and better than that of thiamethoxam (50 g ha?1). CONCLUSION: The novel chemistry of sulfoxaflor, its unique biological spectrum of activity and its lack of cross‐resistance highlight the potential of sulfoxaflor as an important new tool for the control of sap‐feeding insect pests. Copyright © 2010 Society of Chemical Industry  相似文献   

17.
Glyphosate is one of the most commonly used broad‐spectrum herbicides over the last 40 years. Due to the widespread adoption of glyphosate‐resistant (GR) crop technology, especially corn, cotton and soybean, several weed species have evolved resistance to this herbicide. Research was conducted to confirm and characterize the magnitude and mechanism of glyphosate resistance in two GR common ragweed ( A mbrosia artemisiifolia L.) biotypes from Mississippi, USA. A glyphosate‐susceptible (GS) biotype was included for comparison. The effective glyphosate dose to reduce the growth of the treated plants by 50% for the GR1, GR2 and GS biotypes was 0.58, 0.46 and 0.11 kg ae ha?1, respectively, indicating that the level of resistance was five and fourfold that of the GS biotype for GR1 and GR2, respectively. Studies using 14 C‐glyphosate have not indicated any difference in its absorption between the biotypes, but the GR1 and GR2 biotypes translocated more 14 C‐glyphosate, compared to the GS biotype. This difference in translocation within resistant biotypes is unique. There was no amino acid substitution at codon 106 that was detected by the 5‐enolpyruvylshikimate‐3‐phosphate synthase gene sequence analysis of the resistant and susceptible biotypes. Therefore, the mechanism of resistance to glyphosate in common ragweed biotypes from Mississippi is not related to a target site mutation or reduced absorption and/or translocation of glyphosate.  相似文献   

18.
The expansion of atrazine‐resistant Chenopodium album (common lambsquarters) since the 1980s has forced New Zealand's maize‐growers to use an additional postemergence herbicide application. The frequent use of dicamba for this has selected for a common lambsquarters population with reduced sensitivity to dicamba. Initial greenhouse experiments with seeds that had been collected from the plants that survived field applications of dicamba showed that these plants could tolerate ≤1.2 kg ha?1, fourfold the recommended rate. These dicamba‐resistant plants were morphologically distinct from the susceptible population. The leaves of the resistant plants were less dentate and a lighter shade of green. The resistant plants were shorter, had a lower biomass and growth rate and flowered ≤19 days earlier than the susceptible plants. When grown together in various density ratios, the average biomass of both the susceptible and the resistant plants increased as the number of susceptible plants decreased in the mixture. The field experiments demonstrated that the resistant population tolerated dicamba at ≤2.4 kg ha?1, eightfold the recommended rate. Postemergence applications of bromoxynil, pyridate, nicosulfuron and mesotrione effectively controlled both populations. Nicosulfuron and mesotrione provided long‐term residual control, with nicosulfuron being more effective on the grass weeds. High rates of dicamba damaged the maize plants, resulting in an increased weed cover and reduced grain yield. The number of viable common lambsquarters seeds in the soil seed bank at the end of the growing season declined in the treatments in which common lambsquarters was controlled effectively.  相似文献   

19.
BACKGROUND: Studies were conducted to elucidate the mechanism of glufosinate resistance in an Italian ryegrass population. RESULTS: Glufosinate rates required to reduce growth by 50% (GR50) were 0.15 and 0.18 kg AI ha?1 for two susceptible populations C1 and C2 respectively, and 0.45 kg AI ha?1 for the resistant population MG, resulting in a resistance index of 2.8. Ammonia accumulation after glufosinate treatment was on average 1.5 times less for the resistant population than for the susceptible populations. The glufosinate concentrations (µM ) required to reduce the glutamine synthetase (GS) enzyme activity by 50% (I50) were 31 and 137 for C1 and C2 respectively, and 2432 for the resistant population MG. One amino acid substitution in the plastidic GS2 gene, aspartic acid for asparagine at position 171, was identified in the resistant population. CONCLUSIONS: This is the first report of glufosinate resistance in a weed species that involves an altered target site. Copyright © 2012 Society of Chemical Industry  相似文献   

20.
Sprangletop (Leptochloa chinensis L. Nees) is a serious grass weed in direct‐seeded rice cropping systems in Thailand. One population of sprangletop, BLC1, was found to be resistant to fenoxaprop‐p‐ethyl at 62‐fold the concentration of a susceptible biotype, SLC1. This study elucidated the inheritance of resistance to fenoxaprop‐p‐ethyl in this sprangletop BLC1 genotype. The reaction to the herbicide at 0.12–2.4 mg ai L?1 was determined in the seedlings of self‐pollinated resistant BLC1, susceptible SLC1 and SLC1 that had been allowed to cross‐pollinate with BLC1. At 0.24 mg ai L?1, all the seedlings of SLC1 were killed, while 99% of BLC1 survived, along with 5% of the cross‐pollinated SLC1 seedlings, which were considered to be putative F1 hybrids. The root and shoot lengths of the F1 hybrids in 0.24 mg ai L?1 of fenoxaprop‐p‐ethyl, relative to those in the absence of the herbicide, were close to or the same as the resistant parent, indicating that the resistance is a nearly complete to complete dominant trait. One‐hundred‐and‐forty‐one of the F2‐derived F3 families were classified by their response to the herbicide at 0.24 and 0.48 mg ai L?1 into 39 homozygous susceptible : 72 segregating : 30 homozygous resistant, fitted with a 1:2:1 ratio at χ2 = 1.21 and P = 0.56, indicating that the resistance to fenoxaprop‐p‐ethyl in the sprangletop BLC1 genotype is controlled by a single gene.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号