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 62 毫秒
1.
Baseline sensitivity and efficacy of thifluzamide in Rhizoctonia solani   总被引:1,自引:0,他引:1  
Thifluzamide is a SDHI (succinate dehydrogenase inhibitor) fungicide, which interferes with succinate ubiquinone reductase in the mitochondrial electron transport chain of fungi. Presently, jinggangmycin is the major fungicide extensively used for the control of rice sheath blight caused by Rhizoctonia solani and resistance to jinggangmycin was first reported to occur in China. A total of 128 isolates of R. solani from Anhui Province of China were characterised for the baseline sensitivity to thifluzamide. The isolates were very sensitive to thifluzamide and the baseline sensitivity curve was unimodal with an average EC50 value of 0.058 ± 0.012 µg mL?1. However, EC50 values of boscalid (another SDHI fungicide) for inhibition of mycelial growth of 22 arbitrarily selected R. solani isolates ranged from 1.89 to 2.68 µg mL?1. Thifluzamide applied at 110 µg mL?1 exhibited excellent protective and curative activity against rice sheath blight and provided 81.1–91.0% protective or curative control efficacy. In field trials in 2010 and 2011, control efficacies of thifluzamide at 82 g.a.i ha?1 15 and 30 days after second application were 84.2% and 86.7%, respectively, suggesting excellent activity against sheath blight. There was a statistically significant difference in the efficacy between thifluzamide and boscalid or jinggangmycin. These results suggested that thifluzamide should be a good alternative fungicide to jinggangmycin for the control of rice sheath blight.  相似文献   

2.
Resistance risk assessment for fludioxonil in Stemphylium solani   总被引:1,自引:0,他引:1       下载免费PDF全文
An outbreak of grey leaf spot caused by Stemphylium solani was observed on tomato in Shandong Province of China in recent years and brought huge economical losses. Fludioxonil is a phenylpyrrole fungicide with strong antifungal activity against S. solani. To evaluate the risk of S. solani developing fludioxonil resistance, a total of 145 field isolates were examined for sensitivity to fludioxonil by measuring mycelial growth. The baseline sensitivity was distributed as a unimodal curve with a mean EC50 value of 0.0659 (±0.0170) µg mL?1. Five mutants with high resistance to fludioxonil (RF > 1000) were obtained by successively selecting on fludioxonil‐amended plates in the laboratory. All the resistant mutants associated with strongly reduced fitness in mycelial growth, sporulation and pathogenicity. Fludioxonil had positive cross‐resistance with procymidone and iprodione, but there was no cross‐resistance with other fungicides including boscalid, fluazinam, azoxystrobin and flusilazole. Based on the current results, resistance risk of S. solani to fludioxonil could be moderate. This is the first report of baseline sensitivity of S. solani to fludioxonil and its risk assessment. In order to delay the resistance development, it is recommended that fludioxonil can be used as one component of the mixture or fungicides with different modes of action should be alternatively used for this disease management.  相似文献   

3.
Grey mould, caused by Botrytis cinerea Pers ex Fr., is one of the most common diseases of tomato worldwide. Fludioxonil belongs to the phenylpyrrole fungicides, which have high activity against B. cinerea. The sensitivity of fludioxonil was evaluated on the basis of the level of inhibition of mycelium growth in 274 B. cinerea isolates collected from different locations (untreated with this fungicide) in Henan Province, China. The EC50 values for fludioxonil ranged from 0.0033 to 0.0415 mg/l, and the average EC50 values were 0.0156 ± 0.0078 mg/l. Three fludioxonil‐resistant mutants were obtained by subculturing fludioxonil‐sensitive wild‐type isolates on continuously increasing fludioxonil concentrations. For the cross‐resistance assay, fludioxonil revealed positive cross‐resistance with procymidone but did not reveal cross‐resistance with pyrimethanil, boscalid and trifloxystrobin. Mycelial growth, conidial production, hyphal dry weight and pathogenicity were diminished significantly between the fludioxonil‐resistant mutants and their sensitive wild‐type parental isolates. This study shows for the first time that fludioxonil‐resistant isolates of B. cinerea are still not present in Henan Province because this fungicide is an attractive and effective fungicide for chemical control. Recommendations can be made to growers to use fludioxonil to control grey mould and to consider the potential moderate resistance risk of using this fungicide.  相似文献   

4.
Sclerotinia stem rot caused by Sclerotinia sclerotiorum is one of the most important diseases in oilseed rape‐growing areas of China. To determine the frequency of resistance of field isolates of S. sclerotiorum to carbendazim and dimethachlone, a total of 556 isolates from 10 different regions of Henan Province were obtained between 2015 and 2016. The frequency of isolates with a high‐resistance phenotype and a moderate‐resistance phenotype to carbendazim was 69.2% and 10.8%, respectively. However, S. sclerotiorum isolates resistant to dimethachlone were not detected. The baseline sensitivity of S. sclerotiorum to dimethachlone was distributed as a unimodal curve with a mean EC50 value of 0.39 ± 0.09 μg ml?1 for the inhibition of mycelial growth. Four dimethachlone‐resistant mutants were obtained from 20 wild‐type isolates induced by exposure to increasing concentrations of the fungicide in vitro. The mutants showed high levels of resistance to dimethachlone, with resistance factors that ranged from 179 to 323. Positive cross‐resistance occurred between dimethachlone and procymidone, iprodione, and fludioxonil; however, no cross‐resistance was observed for carbendazim and boscalid. The fitness of the dimethachlone‐resistant mutants was significantly lower than that of the wild‐type isolates, as measured by mycelial growth, hyphal dry weight, sclerotium number and dry weight, and pathogenicity. Additionally, based on osmotic tests, the inhibition of mycelial growth caused by NaCl applied at different concentrations was significantly higher for the dimethachlone‐resistant mutants than for their wild‐type parents.  相似文献   

5.
The sensitivity of Alternaria solani isolates to the fungicides mancozeb and chlorothalonil was evaluated, to determine if inadequate disease management by these fungicides could be attributed to reduced sensitivity of A. solani isolates to these fungicides. The sensitivity of 60 isolates of A. solani was assessed using the inhibition of radial mycelial growth (RG) method, using fungicide concentrations of 0, 1.0, 10, 100, 500 and 1000 μg a.i ml?1 medium. EC50 was calculated for each isolate and fungicide combination. The EC50 values of different A. solani isolates to mancozeb ranged from 9.05 to 712.65 μg ml?1. EC50 values of different isolates to chlorothalonil ranged from 4.25 to 849.4 μg ml?1. The percentage of isolates with reduced sensitivity was 46.7 and 53.3% for mancozeb and chlorothalonil, respectively. Results of the in vivo tests demonstrated decline in disease control by the two fungicides with the reduced-sensitivity isolates compared to the sensitive ones.  相似文献   

6.
Isolates of Phytophthora cactorum resistant to the systemic fungicide metalaxyl were obtained by exposing them to sequentially increased concentrations of metalaxyl. A linear relationship was observed between the concentrations of metalaxyl and percentage inhibition of mycelial growth of P. cactorum. The stability of metalaxyl-resistant isolates 150R and 250R was confirmed after six serial transfers on corn meal agar without fungicide. The in vitro metalaxyl-resistant isolate (Ph10) was less aggressive on apple rootstocks compared with the Ph07 isolated from metalaxyl-treated trees and the Ph03 isolated from untreated trees. Metalaxyl-resistant and sensitive isolates remained sensitive to the chemically unrelated fungicide fosetyl-Al at high concentration (600 μg/ml), to mancozeb, and to a mixture of metalaxyl + mancozeb. Significant differences in resistance to metalaxyl existed among P. cactorum field isolates.  相似文献   

7.
One hundred and eighty isolates of Rhizoctonia solani AG1‐IA, the causal agent of rice sheath blight, were obtained from six locations in southern China. The genetic structure of R. solani isolates was investigated using random amplified polymorphic DNA (RAPD) markers, and a considerable genetic variation among R. solani isolates was observed. Most of the genetic diversity was distributed within populations, rather than among them. The distribution pattern of the genetic variation of R. solani appears to be the result of high gene flow (Nm) and low‐genetic differentiation among populations. The aggressiveness of R. solani was visually assessed by rice seedlings of five different cultivars in the glasshouse. All isolates tested were found to induce significantly different levels of disease severity, reflecting considerable variation in aggressiveness. The isolates were divided into highly virulent, moderately virulent and weakly virulent groups, and the moderately virulent isolates were dominant in R. solani population. No significant correlation was observed among the genetic similarity, pathogenic aggressiveness and geographical origins of the isolates. Information obtained from this study may be useful for breeding for improved resistance to sheath blight.  相似文献   

8.
This study assessed the fenhexamid sensitivity of 143 Botrytis cinerea isolates collected from greenhouse strawberries in five regions of China between 2012 and 2013, and identified four isolates with moderate levels of resistance: two from the Xinjiang Uygur Autonomous Region and two from Hebei Province. The baseline fenhexamid sensitivity of B. cinerea exhibited a unimodal distribution with a mean EC50 value of 0.20 ± 0.10 μg/ml (SD). The EC50 values of the fenhexamid‐resistant isolates ranged from 0.05 to 0.40 μg/ml. Molecular analysis of the fenhexamid target gene erg27 revealed that the resistant isolates collected from Xinjiang (163‐6 and 163‐22) contained three mutations that led to amino acid changes (V365A, E368D and A378T) known to be associated with fenhexamid resistance, but that the isolates from Hebei lacked any mutations, indicating that an alternative mechanism could be responsible for their resistance. Most of the biological characteristics of the fenhexamid‐resistant isolates, such as mycelial growth, sclerotia production and pathogenicity, did not significantly differ from those of the sensitive ones ( .05), but it was noted that some of the resistant isolates exhibited reduced rates of sporulation and spore germination. In addition, the resistant isolates exhibited lower osmotic sensitivity than the sensitive ones. The study found no evidence of cross‐resistance with other fungicides, but that there was negative cross‐resistance with procymidone, iprodione, carbendazim and pyraclostrobin, which indicates that the inclusion of these fungicides within an integrated pest management (IPM) programme could help to minimize the risk of fenhexamid resistance developing in B. cinerea.  相似文献   

9.
Botrytis cinerea, the causal agent of grey mould in a broad range of crops, is considered a high‐risk plant pathogen for fungicide resistance development. The use of fungicide mixtures, particularly combinations with synergistic activity, can be a useful tactic to counteract resistance build‐up in pathogen populations. The present study aimed to investigate the effects of different ratios of two‐way mixtures of carbendazim, iprodione, kresoxim‐methyl, tebuconazole and penconazole on four B. cinerea isolates that were sensitive or resistant to benzimidazoles, dicarboximides and strobilurins. The isolates that were resistant to benzimidazoles and strobilurins had E198A and G143A mutations in β‐tubulin and cytochrome b genes, respectively. The mixtures had different effects on each of the isolates in vitro but, in 13 combinations, the synergistic effect was observed against all or three isolates. In greenhouse experiments, 11 fungicide combinations used in decreased (EC75) concentrations showed the maximum control efficiency. The two follow‐up greenhouse experiments using six selected combinations revealed they were highly effective against additional isolates with various fungicide resistance profiles. The identified mixtures‐ratios have potential for use in grey mould management programs in the greenhouse.  相似文献   

10.
Pepper Phytophthora blight (PPB), caused by Phytophthora capsici, is an important disease of pepper in China. The extensive application of metalaxyl has resulted in widespread resistance to this fungicide in field. This study has evaluated the activities of several fungicides against the mycelial growth and sporangium germination of metalaxyl‐sensitive and metalaxyl‐resistant P. capsici isolates by determination of EC50 values. The results showed that the novel carboxylic acid amide (CAA) fungicide mandipropamid exhibited excellent inhibitory activity against PPB both in vitro and in vivo, with averagely EC50 values of 0.075 and 0.004 μg/ml in mycelial growth and sporangium germination, respectively, and over 88% efficacy in controlling PPB. The other three CAA fungicides also provided over 70% efficacy in controlling PPB. The mycelial growth was less sensitive to quinone outside inhibitor (QoI) fungicides azoxystrobin and trifloxystrobin than that of sporangium germination in P. capsici isolates. However, azoxystrobin and trifloxystrobin provided over 80% efficacy in controlling PPB. It was noted that propamocarb and cymoxanil did not exhibit activity against the mycelial growth or sporangium germination of P. capsici isolates in the in vitro tests, with over 70% efficacy in controlling PPB. The new fungicide mixture 62.5 g/l fluopicolide + 625 g/l propamocarb (trade name infinito, 687.5 g/l suspension concentrate (SC)) produced over 88% efficacy in controlling PPB caused by both metalaxyl‐sensitive and metalaxyl‐resistant isolates. The data of this study also proved that there was obviously no cross‐resistance between metalaxyl and the other tested fungicides. Therefore, these fungicides should be good alternatives to metalaxyl for the control of PPB and management of metalaxyl resistance.  相似文献   

11.
Rice leaves with bacterial blight or bacterial leaf streak symptoms were collected in southern China in 2007 and 2008. Five hundred and thirty‐four single‐colony isolates of Xanthomonas oryzae pv. oryzae and 827 single‐colony isolates of Xanthomonas oryzae pv. oryzicola were obtained and tested on plates for sensitivity to streptomycin. Four strains (0.75%) of X. oryzae pv. oryzae isolated from the same county of Province Yunnan were resistant to streptomycin, and the resistance factor (the ratio of the mean median effective concentration inhibiting growth of resistant isolates to that of sensitive isolates) was approximately 226. The resistant isolate also showed streptomycin resistance in vivo. In addition to resistant isolates, isolates of less sensitivity were also present in the population of X. oryzae pv. oryzae from Province Yunnan. However, no isolates with decreased streptomycin‐sensitivity were obtained from the population of X. oryzae pv. oryzicola. Mutations in the rpsL (encoding S12 protein) and rrs genes (encoding 16S rRNA) and the presence of the strA gene accounting for streptomycin resistance in other phytopathogens or animal and human pathogenic bacteria were examined on sensitive and resistant strains of X. oryzae pv. oryzae by polymerase chain reaction amplification and sequencing. Neither the presence of the strA gene nor mutations in the rpsL or rrs were found, suggesting that different resistance mechanisms are involved in the resistant isolates of X. oryzae pv. oryzae.  相似文献   

12.
13.
Summary The in-vitro effect of the seed dressing fungicide tetramethyl thiuram disulphide (TMTD, Thiram) and its degradation product sodium dimethyl dithiocarbamate (NaDDC) on the sensitivity of Rhizobium spp. and Azotobacter chroococcum isolates was studied at different pH values. The Azotobacter chroococcum isolates were more sensitive than Rhizobium spp. isolates at all pH values. The sensitivity of these isolates to TMTD and NaDDC had a maximum at pH 7, and it decreased on either side of this value. The Rhizobium isolates from urid (Phaseolus radiatus) were resistant to TMTD at pH 7 but proved sensitive to its degradation product. A difference between the power of the sensitive and resistant isolates to reduce TMTD was revealed.  相似文献   

14.
Succinate dehydrogenase inhibitor (SDHI) fungicides constitute a relatively recent fungicide class registered for the treatment of grey mould on grapevine in Italy. The sensitivity profile to a novel compound fluopyram was established for a set of 203 Botrytis cinerea isolates collected from Sicilian vineyards within 2009–2012 prior its introduction into market. In addition, its performances were compared in in vitro and in vivo assays with other registered SDHI fungicide boscalid, to evaluate their frequency distributions EC50 values and cross‐resistance patterns. Results of the article showed that EC50 values for fluopyram ranged from 0.05 to 1.98 µg mL?1. Although EC50 values of boscalid ranged from 0.01 to 89.52 µg mL?1, no cross‐resistance relationship was observed between the two fungicides (r = 0.003; P = 0.964) within our B. cinerea population. On further confirming these data, boscalid failed in controlling grey mould infections when boscalid‐resistant isolates were inoculated on grape berries whereas fluopyram exhibited a good efficacy against the same isolates. This study represents the first report on the baseline sensitivity to fluopyram within B. cinerea population from Sicilian table grape vineyards in Italy, and it clearly shows the lack of cross‐resistance in vitro and in vivo between fluopyram and boscalid for the field pathogen isolates. These results provided useful information for managing of fungicide resistance suggesting that fluopyram could be a valid alternative to boscalid for the control of grey mould of table grape.  相似文献   

15.
Chang WL  Kao CY  Wu CT  Huang AH  Wu JJ  Yang HB  Cheng HC  Sheu BS 《Helicobacter》2012,17(3):210-215
Backgrounds: The levofloxacin resistance caused by gyrA gene mutation is rising rapidly to limit wide application for Helicobacter pylori eradication. We investigated whether gemifloxacin has a superior antimicrobial activity to levofloxacin against H. pylori. Materials and Methods: Forty‐four consecutive clinical H. pylori isolates with levofloxacin resistance and 80 randomly selected levofloxacin‐sensitive controls were tested for gemifloxacin sensitivity by E‐test. The resistance to levofloxacin or gemifloxacin was defined as minimal inhibitory concentration (MIC) >1 mg/L. The clinical features and GyrA mutation patterns checked by direct sequencing were also analyzed to assess its association with the H. pylori gemifloxacin resistance. Results: All levofloxacin‐sensitive H. pylori isolates were sensitive to gemifloxacin. Eight strains (18.2%) resistant to levofloxacin could be still sensitive to gemifloxacin. Gemifloxacin achieved a 5‐time lower in MIC levels against levofloxacin‐resistant isolates. Nearly all levofloxacin‐resistant isolates (97.7%, 43/44) had GyrA mutation at amino acid position 87 or 91. Double mutation sites may play dual roles in quinolone resistance, as N87K plus H57Y or D91N plus V77A mutations showed high‐level resistance to both quinolones; whereas D91Y plus A97V or D91N plus A97V mutations showed low level levofloxacin resistance to become sensitive to gemifloxacin. In H. pylori isolates with single N87K, D91Y or D91N mutation, near 20% was gemifloxacin‐sensitive and levofloxacin‐resistant. The gemifloxacin‐resistant rate of H. pylori was higher in patients with gastric ulcer than in those without (p <.05). Conclusion: Gemifloxacin is superior to levofloxacin in antimicrobial activity against clinical H. pylori isolates, and even overcome some levofloxacin resistance.  相似文献   

16.
Mutants of Ustilago maydis (DC.) Corda, resistant to the piperidine fungicide fenpropidin, were isolated in a mutation frequency of 3.2 × 10–5, after UV‐irradiation and selection on media containing 75 μg/ml fenpropidin. Genetic analysis with 15 such mutant isolates resulted in the identification of two unlinked chromosomal loci, U/fpd‐1 and U/fpd‐2. The U/fpd mutations are responsible for moderate resistance levels to fenpropidin (Rf: 42–56 or 15 based on effective concentration causing a 50% reduction in the growth rate (EC50) or minimal inhibitory concentration (MIC) values, respectively). Haploid strains carrying both U/fpd mutations do not exhibit higher levels of resistance to fenpropidin, indicating no additivity of gene effect between non‐allelic genes. Cross‐resistance studies with other Sterol Biosynthesis inhibitors (SBIs) showed that the U/fpd‐mutant isolates exhibited a positive cross‐resistance to the piperidine piperalin and to the related morpholine fungicides fenpropimorph and tridemorph, but not to the inhibitors of C‐14 demethylase and squalene epoxidase. Crosses between mutants carrying the U/fpd‐genes with compatible isolates carrying the U/fpm or U/tdm mutations, which have been identified in previous genetic studies for resistance to morpholine fungicides fenpropimorph and tridemorph, yielded, with the exception of U/fpd‐2 × U/fpm‐2 crosses, a large number of recombinants with wild‐type sensitivity, indicating that the mutant genes involved were not allelic. Analysis of progeny from crosses between U/fpd‐2 and U/fpm‐2 mutants yielded no recombinants with wild‐type sensitivity, but a 1 : 1 progeny segregation was observed at the MIC for the U/fpd‐2 isolates, indicating that these genes are alleles of the same locus. A study of the fitness of fenpropidin‐resistant isolates showed that the U/fpd mutations do not affect the phytopathogenic fitness‐determining characteristics such as growth in liquid culture and pathogenicity on young corn plants.  相似文献   

17.
Sustainable disease management depends on the ability to monitor the development of fungicide resistance in pathogen populations. A point mutation resulting in an alteration (F200Y) at codon 200 of the target protein β‐tubulin leads to a moderate level of resistance to carbendazim in Botrytis cinerea. Although traditional methods remain a cornerstone in detection of fungicide resistance, molecular methods that do not require the isolation of pathogens, can detect the presence of resistance alleles at low frequencies, and require less time and labour than traditional methods. In this study, we present an efficient, rapid, and highly specific method for detecting the moderately carbendazim‐resistant isolates in B. cinerea based on loop‐mediated isothermal amplification (LAMP). By using specific LAMP primers, we detected the resistance‐conferring mutation underlying β‐tubulin F200Y. The concentrations of LAMP components and LAMP parameters were optimised, resulting in reaction temperatures and times of 61–65°C and 45 min, respectively. The feasibility of the LAMP assay was verified by assaying the diseased samples with artificial inoculation in the different hosts. The LAMP assay developed in the current study was specific, stable, repeatable and sensitive, and was successfully applied for detection of moderately carbendazim‐resistant isolates of B. cinerea in plant samples.  相似文献   

18.
19.
In continuation of our previous research on the development of novel pyrazole‐4‐carboxamide with potential antifungal activity, compound SCU2028 , namely N‐[2‐[(3‐chlorophenyl)amino]phenyl]‐3‐(difluoromethyl)‐1‐methyl‐1H‐pyrazole‐4‐carboxamide, was synthesized by new method, structurally characterized by IR, HR‐ESI‐MS, 1H‐ and 13C‐NMR spectra and further identified by single‐crystal X‐ray diffraction. In pot tests, compound SCU2028 showed good in vivo antifungal activity against Rhizoctonia solani (R. solani) and IC50 value of it was 7.48 mg L?1. In field trials, control efficacy of compound SCU2028 at 200 g.a.i. ha?1 was 42.30 % on the 7th day after the first spraying and 68.10 % on the 14th day after the second spraying, only slightly lower than that of thifluzamide (57.20 % and 71.40 %, respectively). Further in vitro inhibitory activity showed inhibitory ability of compound SCU2028 was 45‐fold higher than that of bixafen and molecular docking of compound SCU2028 to SDH predicted its binding orientation in the active site of the target protein SDH. These results suggested that compound SCU2028 was a potential fungicide for control of rice sheath blight.  相似文献   

20.
Rhizoctonia solani is an important plant pathogen for a number of crops and maintaining an extensive collection of reference isolates is important in understanding relationships of this pathogen with multiple hosts. Current long‐term storage methods typically call for frequent transfer increasing the risk of changes in morphological, physiological or virulence characteristics. Cryopreservation using storage in liquid nitrogen (LN) was evaluated to examine the potential for storage of a R. solani culture collection containing 106 isolates (primarily from sugar beet). Cultures were stored on autoclaved barley grains in the vapour phase of LN. After 60 days, 5 years and 10 years in storage, all isolates were tested for viability by calculating the percentage of barley grains from which R. solani mycelia grew. Five years after initial storage, all isolates except one had no change in viability. After 10 years in storage, 67 of 106 isolates had no significant decrease in viability, 39 of 106 isolates had a significant decrease in viability but only 9 isolates had less than 10% growth, with 4 having no growth. A subset of isolates stored for 10 years were tested for pathogenicity on a susceptible (FC901) and resistant (FC703) sugar beet germplasm. All isolates tested maintained approximately the same level of virulence that they had prior to storage on both germplasms. This indicates that cryogenic methods are suitable for the preservation or storage of R. solani culture collections, although efficacy may vary with individual isolates.  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