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 15 毫秒
1.
C. Zhang  H. Wu  X. Li  H. Shi  F. Wei  G. Zhu 《Plant pathology》2013,62(6):1378-1383
During 2009–2010, a total of 323 isolates of Xanthomonas oryzae pv. oryzae were obtained from rice with symptoms of bacterial leaf blight (BLB) in four provinces (Zhejiang, Jiangsu, Anhui and Hubei) in China. These isolates were tested for baseline sensitivity to zinc thiazole, a novel bactericide with strong antibacterial activity against Xanthomonas. The sampled pathogenic population had similar sensitivity to zinc thiazole (0·1–16·8 mg L?1) in all four regions and over the whole two‐year study period. The baseline sensitivity was distributed as a unimodal curve with a mean EC50 value of 6·79 ± 1·61 mg L?1. The risk of mutation to resistance of zinc thiazole in X. oryzae pv. oryzae was further evaluated in vitro and in vivo. Twelve zinc thiazole‐resistant mutants were obtained through ultraviolet (UV) irradiation, culturing on zinc thiazole‐amended nutrient agar (NA) plates, and culturing on zinc thiazole‐treated rice plants. These zinc thiazole‐resistant mutants had resistance factors (RF = EC50 value of a mutant / EC50 value of the wildtype parent of this mutant) of 12·4 to 186·1 with a mean RF value of 44·1. Mutants obtained via UV irradiation, culturing on NA plates and culturing on rice plants had mean RF values of 51·8, 24·5 and 14·4, respectively. All mutants showed decreases in resistance to zinc thiazole after 20 successive transfers on bactericide‐free media or 10 successive inoculation–reisolations on bactericide‐free rice plants. No significant difference was found in bacterial growth and sensitivity to bismerthiazol between zinc thiazole‐resistant mutants and their parents. However, a significant decrease was observed in the pathogenicity of zinc thiazole‐resistant mutants compared with their parents, especially for mutants obtained via UV irradiation.  相似文献   

2.
Marker‐free transgenic tobacco (Nicotiana tabacum) lines containing a chitinase (ChiC) gene isolated from Streptomyces griseus strain HUT 6037 were produced by Agrobacterium‐mediated transformation. One marker‐free transgenic line, TC‐1, was retransformed with the wasabi defensin (WD) gene, isolated from Wasabia japonica. Of the retransformed shoots, 37% co‐expressed the ChiC/WD genes, as confirmed by western and northern analyses. Southern blot analysis showed that no chromosomal rearrangement was introduced between the first and the second transformation. Transgenic lines either expressing ChiC or WD, or co‐expressing both genes were challenged with Fusarium oxysporum f.sp. nicotianae (Fon). Assessment of in vitro plant survival in the presence of Fon showed that transgenic lines co‐expressing both genes had significantly enhanced protection against the fungus (infection indices 0·0–1.·2) compared with corresponding isogenic lines expressing either of the genes (infection indices 2·5–9·8). Whole‐plant infection indices in transgenic lines were significantly related (r = 0·93, P < 0·01) to the extent of root colonization of the host, which ranged from 2·1% to 11·3% in lines co‐expressing both genes, and from 16·8% to 37·7% in lines expressing just one of the genes (compared with 86·4% in non‐transformed controls). Leaf extracts of transgenic lines also inhibited mycelial growth of Fon in vitro and caused hyphal abnormalities.  相似文献   

3.
Isolates of Penicillium spp. were collected regularly from 2001 to 2003 from the surfaces of apple fruit pre- and postharvest, and from the atmosphere of orchards and storage rooms in France. Penicillium spp. were not detected from the atmosphere of conventional orchards, while their density did not exceed 50 spores m−3 in the atmosphere of organically managed orchards. Penicillium spp. were seldom detected on apple surfaces in the orchard. The density of Penicillium on apples increased from 10 to 50 spores cm−2 after 1 month in storage to 300–400 spores cm−2 after 6 months. The level of airborne Penicillium increased by up to 2 × 104 and 2·5 × 103 spores m−3 within nondisinfected and previously disinfected warehouses, respectively. Penicillium expansum (30–62%) and P. solitum (6–45%) were the most prevalent species on apple or in storage rooms. Other species of Penicillium isolated included P. commune, P. verrucosum, P. chrysogenum, P. rugulosum and P. digitatum. Apple fruit were also surveyed for wounds and the number of open lenticels using the sulphur dioxide test. The incidence of wounding at harvest varied from 12 to 36%, depending on cultivar and locality. When apples were inoculated at harvest by either aqueous or aerial inoculum of P. expansum, the decay incidence was constantly higher than the incidence of wounding. The number of open lenticels per cm2 of apple surface varied from 0·5 on cv. Boskoop to 4·4 on cv. Golden Delicious. An average of 13 and 2·1% of lenticels, respectively, were infected when they were inoculated by P. expansum and P. verrucosum. Cultivars of apple fruit that showed a greater number of open lenticels, combined with a large diameter varying from 100 to 200 µm, were more susceptible to P. expansum.  相似文献   

4.
Ammonium molybdate was tested as a potential fungicide for use in apples (cv Golden Delicious) against blue and grey mould, important post‐harvest diseases of pome fruits. In tests in vivo at 20 °C, ammonium molybdate (15 mM ) reduced lesion diameters of Penicillium expansum, Botrytis cinerea and Rhizopus stolonifer by 84%, 88% and 100% respectively. When apples treated with ammonium molybdate were stored at 1 °C for three months, a significant reduction in severity and incidence of P expansum and B cinerea was observed in both years of study (1998 and 1999). In the second year of the experiment the reduction in disease severity was greater than 88% for both pathogens, and the level of control was similar to, or greater than, that observed with the fungicide imazalil. When ammonium molybdate was applied as a pre‐harvest treatment, a significant reduction in blue mould decay was observed after three months in cold storage. In vitro, ammonium molybdate greatly inhibited spore germination of P expansum and B cinerea, although better inhibition was obtained against grey mould. Ammonium dimolybdate, sodium molybdate and potassium molybdate were also tested in vitro in comparison with ammonium molybdate as inhibitors of spore germination, but only ammonium molybdate inhibited spore germination by more than 50%. © 2001 Society of Chemical Industry  相似文献   

5.
Root and stem extracts of Fumaria parviflora showed strong nematicidal activity against Meloidogyne incognita in in vitro and in planta experiments. Phytochemical screening of F. parviflora revealed the presence of seven classes of bioactive compounds (alkaloids, flavonoids, glycosides, tannins, saponins, steroids and phenols). Quantitative determination of the plant extracts showed the highest percentages of alkaloids (0·9 ± 0·04) and saponins (1·3 ± 0·07) in the roots and total phenolic contents in the stem (16·75 ± 0·07 μg dry g?1). The n‐hexane, chloroform, ethyl acetate and methanol extracts of roots and stems at concentrations of 3·12, 6·24, 12·5, 25·0 and 50·0 mg mL?1, significantly inhibited hatching and increased mortality of second‐stage juveniles (J2s) compared with water controls. Percentage J2 mortality and hatch inhibition were directly related to exposure time. In pot trials with tomato cv. Rio Grande, root and stem extracts at concentrations of 1000, 2000 and 3000 ppm, applied as soil drenches, significantly reduced the number of galls, galling index, eggs masses, eggs and reproduction factor compared with the water control. Regardless of concentration, all the extracts significantly increased the host plant growth parameters studied. The n‐hexane extracts from the roots and stem were the most active, followed by the methanol ones, at all concentrations. The in vitro and in planta results suggest that extracts from the roots and stem of F. parviflora may be potential novel nematicides.  相似文献   

6.
Bioassays using pellets of agar, thatch-agar and turfgrass-agar were developed using benzimidazole-sensitive Penicillium expansum Link, to detect the fungicide methyl benzimidazol-2-ylcarbamate (MBC) which is the major fungitoxic degradation product of benomyl [methyl 1-(butylcarbamoyl)benzimidazol-2-ylcarbamate] in thatch and turfgrass clippings. These bioassays were used to estimate the amount of fungicide that was biologically available and hence, by subtraction from that applied, the amount that remained bound and biologically unavailable. The limit of quantitation was 0·5 mg kg−1. From 19·9% to 93·2% of the applied fungicide was bound by thatch and 46·2% to 56·9% was bound to turfgrass clippings depending on the concentrations used. In-vitro degradation studies showed that MBC had a half life of approximately 2·5 weeks at 23°C in non-sterilized thatch.  相似文献   

7.
Downy blight, caused by Peronophythora litchii, is an important disease of lychee (litchi) plants in China. The in vitro sensitivities of various asexual stages of P. litchii to the three carboxylic acid amide (CAA) fungicides dimethomorph, flumorph and pyrimorph were studied with four single‐sporangium isolates. None of the three fungicides affected zoospore discharge from sporangia, but they strongly inhibited mycelial growth (mean EC50 values of 0·075, 0·258 and 0·115 mg L?1, respectively); sporangial production (mean EC50 values of 0·085, 0·315 and 0·150 mg L?1, respectively); germination of cystospores (mean EC50 values of 0·140, 0·150 and 0·645 mg L?1, respectively); and germination of sporangia (mean EC50 values of 0·203, 0·5 and 0·743 mg L?1, respectively). As mycelial growth was the most sensitive stage to dimethomorph and pyrimorph, it was chosen to test baseline sensitivities to the three fungicides. In 2007, from 131 isolates collected in Fujian, Guangdong and Guangxi provinces, 127, 116 and 113 isolates were used to establish baseline sensitivity for dimethomorph, flumorph and pyrimorph respectively. Isolates from different provinces exhibited similar baseline sensitivity to the same fungicide. Baseline sensitivities to dimethomorph, flumorph and pyrimorph were distributed as unimodal curves, with mean EC50 values of 0·082 (± 0·01), 0·282 (± 0·047), and 0·115 (± 0·032) mg L?1, respectively. This information will serve as a baseline for tracking future changes in sensitivities of P. litchii populations to these three CAA fungicides.  相似文献   

8.
Alternaria leaf blight (ALB), caused by Alternaria dauci, is one of the most damaging foliar diseases of carrot worldwide. The aim of this study was to compare different methods for evaluating levels of carrot resistance to ALB. Three techniques were investigated by comparison with a visual disease assessment control: in vivo conidial germination, a bioassay based on a drop‐inoculation method, and in planta quantification of fungal biomass by quantitative PCR (Q‐PCR). Three carrot cultivars showing different degrees of resistance to A. dauci were used, i.e. a susceptible cultivar (Presto) and two partially resistant genotypes (Texto and Bolero), challenged with an aggressive or a very aggressive isolate of A. dauci. Both partially resistant genotypes produced a higher mean number of germ tubes per conidium (up to 3·42±0·35) than the susceptible one (1·26±0·18). The drop‐inoculation results allowed one of the partially resistant genotypes (Bolero, log10(S+1) = 1·34±0·13) to be distinguished from the susceptible one (1·90±0·13). By contrast, fungal growth measured by Q‐PCR clearly differentiated the two partially resistant genotypes with log10(I) values of 2·77±0·13 compared to the susceptible cultivar (3·65±0·13) at 15 days post‐inoculation. This result was strongly correlated (r2 = 0·91) with the disease severity index scored at the same date. Data obtained with the different assessment methods strongly suggest that the Texto and Bolero genotypes have different genetic resistance sources.  相似文献   

9.
Sclerotinia stem rot (SSR), caused by Sclerotinia sclerotiorum, is a major disease of soybean in Canada. Laboratory and greenhouse experiments were conducted to evaluate potential effectiveness of cell suspensions, cell‐free culture filtrates and broth cultures of Bacillus subtilis strain SB24 for suppression of SSR. The SB24 cell suspensions and cell‐free culture filtrates significantly reduced mycelial growth of S. sclerotiorum by 50 to 75% and suppressed sclerotial formation by > 90%. The severity on soybean was negatively correlated (r < ?0·84, P < 0·01) to the concentrations of cell suspension, cell‐free culture filtrate and broth culture applied. The cell suspension and broth culture preparations significantly (P < 0·01) reduced SSR severity by 45 to 90% at concentrations ranging from 5 × 106 to 109 CFU mL?1. The most effective concentration was 5 × 108 CFU mL?1 for all three preparations, reducing the severity by 60 to 90%. The B. subtilis SB24 was most effective in reducing disease severity when applied ≤ 24 h before plant inoculation with S. sclerotiorum and a significant effectiveness was observed up to 15 days after plant inoculation. The population density of B. subtilis on soybean leaves decreased by 1·5 to 2·5 log units over 15 days under field conditions, and by 0·8 log units over 5 weeks under control conditions. The decrease in population density was significantly correlated with rainfall in the field (r < ?0·93, P < 0·01), suggesting that the biocontrol bacteria may be washed away by rain.  相似文献   

10.
Ppdfn1 is a defensin gene previously identified in peach (Prunus persica). The biological role of Ppdfn1 was investigated by analysing its expression profile in leaves, flowers and fruits, either inoculated with the Monilinia laxa fungal pathogen or mock‐inoculated. Ppdfn1 expression was highest in flowers and, in fruits, did not vary upon M. laxa inoculation. To characterize the PpDFN1 antifungal activity, the recombinant mature peptide was expressed in Escherichia coli and purified; recombinant PpDFN1 displays antifungal activity against Botrytis cinerea, M. laxa and Penicillium expansum, with IC50 values of 15·1, 9·9 and 1·1 μg mL?1, respectively. Treatment of fungal hyphae with FITC‐labelled PpDFN1 indicated that the peptide is not internalized by fungal hyphae, but localizes on their external cell surface. At this site, PpDFN1 is capable of membrane destabilization and permeabilization, as demonstrated by SYTOX Green fluorescence uptake by the treated mycelia. Using artificial lipid monolayers, it was shown that PpDFN1 interacts with sphingolipid‐containing membranes; however the strongest interaction occurs with monolayers composed of lipids extracted from sensitive fungi, such as Pexpansum. These data suggest that the lipid composition of fungal membranes is of key relevance for defensin specificity.  相似文献   

11.
The objective of this study was to develop a reliable and high throughput screening method to evaluate the response of St. Augustine grass (Stenotaphrum secundatum) genotypes to the grey leaf spot (GLS) caused by Magnaporthe oryzae infection. Whole plant, detached stolon and detached leaf assays under growth chamber conditions were compared to field conditions on eight commercial and nine advanced breeding lines of St. Augustine grass. Disease was assessed using two variables, lesion size (LS) and overall plant disease severity (SEV). LS and SEV were highly correlated for field and growth chamber screening methods using the whole plant assay (LS r2 = 0·79; SEV r2 = 0·83; P 0·001), the detached stolon assay (LS r2 = 0·75; SEV r2 = 0·72; P 0·001), and the detached leaf assay (LS r2 = 0·46; SEV r2 = 0·60; P 0·001). Genotypic variation for resistance in 17 St. Augustine grass genotypes was identified using all screening methods for LS (P < 0·05) and SEV (P < 0·05). The rank‐sum method was used to classify St. Augustine grass genotypes into highly resistant (HR), resistant (R), moderately resistant (MR), moderately susceptible (MS), susceptible (S) and highly susceptible (HS) classes based on the rank‐sum values of LS and SEV. Two introduced African polyploids used as parents, and two F1 interploid progeny obtained using an in vitro embryo rescue technique, were classified as highly resistant (HR), or resistant (R), across all screening methods.  相似文献   

12.
Control of Avena fatua (L.) (wild oat) with diclofop methyl applied at 0·7 kg ha?1 at the two-leaf stage and difenzoquat at 0·84 kg ha?1 at the four-leaf stage in wheat (Triticum aestivum L.) under field conditions was good and not affected when either of these herbicides was mixed with 3,6-dichloropicolinic acid as the monoethanolamine salt at 0·14, 0·20 or 0·30 kg ha?1. In the glasshouse, mixtures containing 3,6-dichloropicolinic acid at rates as high as 0·6 kg ha?1 also did not affect control of A. fatua. When barban at 0·35 kg ha?1, or flamprop methyl at 0·56 kg ha?1 was mixed with similar rates of 3,6-dichloropicolinic acid and applied at the two-leaf and four-leaf stage of A. fatua respectively, a reduction in control of A. fatua (antagonism) occurred under both field and glasshouse conditions. The herbicides for control of A. fatua did not influence the fresh weight suppression of C. arvense shoots obtained in the glasshouse with 3,6-dichloropico-colinic acid at 0·3 kg ha?1. Early tolerance of wheat (cv. Neepawa) was acceptable with all mixtures. Wheat yields with diclofop methyl or difenzoquat alone or in mixture with 3,6-dichloropicolinic acid were increased over the yields from the A. fatua-infested control.  相似文献   

13.
To elucidate the fate of flupyrazofos [O,O-diethyl O-(1-phenyl-3-trifluoromethyl-5-pyrazoyl)phosphorothionate] in soil, an aerobic soil metabolism study was carried out for 60 days with [14C]flupyrazofos applied at a concentration of 0·38 μg g-1 to a loamy soil. The material balance ranged from 103·5% to 86·9% and the half-life of [14C]flupyrazofos was calculated to be 13·6 days. The metabolites identified during the study were 1-phenyl-3-trifluoromethyl-5-hydroxypyrazole (PTMHP) and O,O-diethyl O-(1-phenyl-3-trifluoromethyl-5-pyrazoyl)phosphate (flupyrazofos oxon), with maximum levels of 9·8% and 1·6% of applied radiocarbon, respectively. Evolved [14C]carbon dioxide accounted for up to 5·3% of applied radiocarbon and no volatile products were detected during the study. Non-extractable 14C-residue reached 31·6% of applied material at 60 days after treatment and radiocarbon was distributed almost evenly in humin, humic acid and fulvic acid fraction. © 1998 Society of Chemical Industry  相似文献   

14.
The yeast Pichia anomala strain K was selected in Belgium from the apple surface for its antagonistic activity against post-harvest diseases of apples. The efficacy of this strain against P. expansum was evaluated in the laboratory in three scenarios designed to mimic practical conditions, with different periods of incubation between biological treatment, wounding of fruit surface, and pathogen inoculation. Higher protection levels and higher final yeast densities were obtained when the applied initial concentration was 1 × 108 cfu ml−1 than when it was only 1 × 105 cfu ml−1. The protection level correlated positively with the yeast density determined in wounds and was influenced by apple surface wetness. In orchard trials spanning two successive years, biological treatment against P. expansum, based on a powder of P. anomala strain K (1 × 107 cfu ml−1), β-1,3-glucans (YGT 2 g l−1), and CaCl2.2H20 (20 g l−1), was applied to apples pre- or post-harvest under practical conditions and its effect compared with standard chemical treatments. The first year, the highest reduction (95.2%) against blue decay was obtained by means of four successive fungicide treatments and the next-highest level (87.6%) with pre-harvest high-volume spraying of the three-component mixture 12 days before harvest. The second year, the best results were obtained with post-harvest Sumico (carbendazim 25% and diethofencarb 25%) treatment and post-harvest biological treatment, both by dipping the apples, 88.3 and 56.3% respectively. A density threshold of 1 × 104 cfu cm−2 of strain K on the apple surface seemed to be required just after harvest for high protective activity, whatever the method and time of application. In the case of pre-harvest biological treatments, variations in meteorological conditions between the 2 years may have considerably affected strain K population density and its efficacies.  相似文献   

15.
An RP-HPLC method for the quantitative determination of chlorotoluron in technical products and formulations has been developed, using as column μBondapak C18, 250×4·6 (ID) mm, as eluent methanol+water+acetic acid (60+40+0·1 by volume), with detection by UV at 243 nm. Recoveries were 99·3–100%, RSD (n=5)<0·47%. © of SCI.  相似文献   

16.
Common scab is one of the most important soil‐borne diseases of potato and is difficult to control. Selection of potato breeding lines for resistance to common scab is also cumbersome due to environmental factors influencing symptom development and an erratic spatial distribution of the scab pathogens (Streptomyces spp.) in the field. The bacterial phytotoxin thaxtomin A, which causes scab symptoms, can be used to screen large numbers of potato seedlings for tolerance in vitro, but few studies have investigated whether the results correspond to resistance to common scab observed in the field. In this study, 120 F1 potato progeny from a single cross were screened in vitro by exposing the seedlings to thaxtomin A added to the culture medium. Eighteen genotypes were selected based on high sensitivity or tolerance using shoot growth as the criterion, multiplied in vitro, and tested for resistance to common scab caused by S. turgidiscabies and S. scabies in a glasshouse and in three different fields. Evaluation of ca. 6500 tubers showed that the 18 potato genotypes differed in scab indices and disease severity (P < 0·0001). The relative shoot height in vitro (thaxtomin A used at 0·5 μg mL?1) and the scab index in the field showed significant correlation (rs = ?0·463, P = 0·0528, n = 18), also consistent with the results obtained under controlled conditions in the glasshouse. Hence, the in vitro bioassay may be used to discard scab‐susceptible genotypes and elevate the overall levels of common scab resistance in the potato breeding populations.  相似文献   

17.
NMR and UV spectroscopy and molecular modeling methods were applied to probe the interaction of the two imidazolinones, imazethapyr (5-ethyl-2-(4-isopropyl-4-methyl-5-oxo-2-imidazolin-2-yl)nicotinic acid) and its structural isomer CL 303,135 (5-ethyl-3-(4-isopropyl-4-methyl-5-oxo-2-imidazolin-2-yl)picolinic acid), with metal ions. Both the imidazolinones inhibit the enzyme acetohydroxyacid synthase (AHAS) in vitro. However, while imazethapyr is a herbicide that is used widely in agriculture, CL303,135 does not exhibit herbicidal activity. Imazethapyr and CL303,135 exhibited considerable differences in their interactions with metals. In the metal complex of imazethapyr, the carboxyl moiety binds strongly and the pyridine nitrogen binds weakly with metals. In the case of CL303,135, both the pyridine nitrogen and the carboxyl group that are positioned ortho to each other participated strongly in the binding and were found to act together as a strong bidentate ligand to a metal ion. Both of the imidazolinones form predominantly 2:1 complexes with multivalent metal ions. However, imazethapyr binds two orders-of-magnitude more weakly (1·0×109 M -2) with metal ions compared to CL303,135 (1·7×1011 M -2). The interactions of the model compounds, nicotinic acid and picolinic acid, with metals were examined similarly. It was concluded that the strong affinity of CL303,135 for metals compared to imazethapyr may affect its absorption from soil into plants, or its translocation in plants, thereby explaining the differences in herbicidal activity of imazethapyr and CL303,135. © 1997 SCI.  相似文献   

18.
A study of rice diseases in Cambodia from 2005 to 2007 showed widespread occurrence of diseases caused by Acidovorax avenae subsp. avenae, Burkholderia gladioli, B. cepacia and Pantoea ananatis. This is the first report of these pathogens in Cambodia. Additionally, a pseudomonad causing a widespread disease similar to sheath brown rot (caused by Pseudomonas fuscovaginae) was isolated. The studied strains were pathogenic to rice cvs Sen Pidau and IR 66, producing similar, though slightly less severe, symptoms to those observed in the field. Based on comparative 16S rDNA gene sequence analysis, combined with cell wall fatty acid analysis and metabolic profiles, the isolated strains were allocated to the genus Pseudomonas. The novel species were differentiated from Pseudomonas fuscovaginae and P. putida by their inability to metabolize d ‐fructose, d ‐galactose, d ‐galactonic acid lactone, d ‐galacturonic acid, d ‐glucosaminic acid, d ‐glucuronic acid, p‐hydroxy phenylacetic acid, d ‐saccharic acid and urocanic acid. The major fatty acids were C16:0, summed feature 3 (C16:1ω7c and C16:1ω6c) and summed feature 8 (C18:1ω7c), representing 80% of the total. Partial 16S rRNA gene sequences (1460 bp) were identical, except for two nucleotide changes amongst the six strains. Alignment of the causal strains within type‐culture databases revealed similarities of 99·7% with Pseudomonas parafulva AJ 2129T, 99·2% with P. fulva IAM 1592T, 98·9% with P. plecoglossicidia FPC 951T, and 98·1% with P. fuscovaginae MAFF 301177T. On the basis of data from this polyphasic study, it is proposed that the unknown strains isolated from rice represent a novel species of the genus Pseudomonas.  相似文献   

19.
The toxicity of a promising new insecticide, imidacloprid, was evaluated against several susceptible and resistant strains of German cockroach and house fly. Imidacloprid rapidly immobilized German cockroaches followed by a period of about 72 h during which some cockroaches recovered. After 72 h there was no further recovery. Imidacloprid-treated houseflies were immobilized more slowly than treated cockroaches, with the maximum effect observed after 72 h, and there was no recovery. Based upon 72-h LD50 values imidacloprid was moderately toxic to German cockroaches (LD50 values were 6–8 ng mg-1) and had only low toxicity to house flies (LD50 140 ng mg-1). Piperonyl butoxide (PBO) blocked the observed recovery in German cockroaches. PBO also greatly enhanced the 72-h LD50 of imidacloprid from 43- to 59-fold in cockroaches and 86-fold in house flies. Two strains of German cockroach (Baygon-R and Pyr-R) showed >4-fold cross-resistance to imidacloprid. This cross-resistance could not be suppressed by PBO, suggesting that P450 monooxygenase-mediated detoxication is not responsible for this cross-resistance. Variation in the level of synergism observed with PBO (between strains) suggests the ‘basal’ level of monooxygenase-mediated detoxication of imidacloprid is quite variable between strains of German cockroach. The AVER and LPR strains of house fly showed significant cross-resistance to imidacloprid. PBO reduced the level of cross-resistance in AVER from >4·2-fold to 0·5-fold (i.e. the AVER strain LD50 was half that of the susceptible strain when both were treated with PBO), but PBO did not suppress the cross-resistance in LPR. These data suggest monooxygenases are the mechanism responsible for cross-resistance to imidacloprid in AVER, but not in the LPR strain. © of SCI.  相似文献   

20.
Isolates of Colletotrichum sublineolum were collected from different sorghum‐producing regions of Ethiopia and divided into five groups based on their geographic origin. The growth rate of 50 isolates showed considerable variation: 1·7–5·8 mm day?1, mean 3·3 mm day?1. However, the isolates displayed little variation in colony colour and colony margin, except for isolates from the north, which were different from the others. Amplified fragment length polymorphism analysis of 102 isolates revealed much greater variations among the different groups. Dice similarity coefficients ranged from 0·32 to 0·96 (mean 0·78). Cluster analysis and principal coordinate analysis revealed a differentiation of the isolates according to their geographic origin, and both methods clearly indicated a genetic separation between the southern, the eastern and the other isolates. Analysis of molecular variance (amova ) indicated a high level of genetic variation both among (42%) and within (58%) the C. sublineolum sampling sites in Ethiopia. The amova also indicated a high level of genetic differentiation (FST = 0·42) and limited gene flow (Nm = 0·343). The results of this study confirmed the presence of a highly diverse pathogen, which is in agreement with the existence of diverse host genotypes and widely ranging environmental conditions in sorghum‐producing regions of the country. Such diversity should be taken into account in future breeding programmes to achieve an effective and sustainable disease management strategy.  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