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 281 毫秒
1.
Cucumber downy mildew caused byPseudoperonospora cubensis (Berk. and Curt.) Rostov. limits crop production in Shandong Province of China. Since management of downy mildew is strongly dependent on fungicides, a rational design of control programs requires a good understanding of the fungicide resistance phenomenon in field populations of the pathogen. A total of 106 and 97 isolates ofP. cubensis were obtained in 2006 and 2007, respectively. The EC50 values for the growth of all the 106 isolates collected in 2006 were 0.0063–0.0688μg ml−1 (average: 0.0196±0.0048μg ml−1) azoxystrobin and these were therefore considered sensitive isolates. However, 57 field isolates ofP. cubensis of the 97 collected in 2007 with EC50 values that ranged from 0.609 to >51.2μg ml−1 were considered resistant to azoxystrobin. Fragments of the fungicide-targeted mitochondrial cytochromeb gene from total pathogen DNA were amplified using polymerase chain reaction and their sequences analyzed to elucidate the molecular mechanism of resistance. A single point mutation (GGT to GCT) in the cytochromeb gene, resulting in substitution of glycine by alanine at position 143, was found in the three selected azoxystrobin-resistant isolates of downy mildew. This substitution in cytochromeb exhibited different resistance levels, with the resistance factor from 21.15 to greater than 2618.9. In addition, the different resistance levels seemed to appear within 1 year (between 2006 and 2007). Therefore, growers of Shandong Province in China now are faced with a challenge in managing the azoxystrobin resistance in cucumber downy mildew. http://www.phytoparasitica.org posting March 10, 2008.  相似文献   

2.
A total of 108 isolates of cucurbit powdery mildew (CPM) fungi, 78Golovinomyces cichoracearum (Gc) and 30Podosphaera xanthii (Px), originating from nine Czech regions and 22 districts of the Czech Republic were collected in the years 2001–2004. These isolates were screened for tolerance and/or resistance to the three frequently used fungicides (fenarimol, dinocap, benomyl). Fungicide sensitivity was determined by a modified leaf-disc bioassay with five concentrations. Fungicide efficacy differed significantly: fenarimol was the most effective and all isolates of both CPM were controlled by the recommended concentration (36μg a.i. ml−1). Some isolates expressed resistance (profuse sporulation) or tolerance (limited sporulation) to lower concentrations (9.6 and 18μg a.i. ml−1). Specific temporal variation in tolerance/resistance was observed, with some isolates ofGc from 2002 evincing tolerant or resistant reactions to these low fenarimol concentrations, but isolates with similar reactions were not detected during the years 2003–2004. Dinocap showed decreasing efficacy during all 3 years. A shift to more tolerant reactions in the CPM populations was detected forGc in 2001–2002, and forPx in 2001 and 2004. Benomyl was found to be ineffective, because the majority of screened isolates (88%Gc and 97%Px) belonged to the highly resistant strains, with resistant reaction on the recommended concentration (250μg a.i. ml−1) and tolerance or resistance on higher concentrations (500 and 1000μg a.i. ml−1). Sensitivity differed between the CPM species. Whereas practically allPx isolates (except one from 2003) were resistant, 12% ofGc isolates from the years 2001–2003 showed sensitive and/or tolerant reactions. In 2004, only benomyl-resistantGc strains were detected. Variation in tolerance/resistance was detected to all screened fungicides during the course of this study at some repeatedly sampled locations. http://www.phytoparasitica.org posting June 12, 2008.  相似文献   

3.
Three commercial formulations of strobilurins, viz., azoxystrobin, kresoxim-methyl, and trifloxystrobin were evaluated for their efficacy against pearl millet downy mildew disease caused by Sclerospora graminicola. In vitro studies revealed inhibition of S. graminicola sporulation, zoospore release, and zoospore motility at 0.1-2 μg ml−1 of all the three fungicides. The fungicides were evaluated for phytotoxic effects on seed quality parameters and for their effectiveness against downy mildew disease by treating pearl millet by: (1) seed dressing, (2) seed dressing followed by foliar spray, and (3) also by foliar spray alone. The highest non-phytotoxic concentrations of 5, 10, and 10 μg ml−1 for azoxystrobin, trifloxystrobin, and kresoxim-methyl, respectively, were selected for further studies. Under greenhouse conditions, these fungicides showed varying degrees of protection against downy mildew disease. Among the three fungicides, azoxystrobin proved to be the best by offering disease protection of 66%. Further, seed treatment along with foliar application of these fungicides to diseased plants showed enhanced protection against the disease to 93, 82, and 62% in treatments of azoxystrobin, kresoxim-methyl and trifloxystrobin respectively. Foliar spray alone provided significant increase in disease protection levels of 91, 79, and 59% in treatments of azoxystrobin, kresoxim-methyl, and trifloxystrobin, respectively. Disease curative activity of azoxystrobin was higher compared to trifloxystrobin and kresoxim-methyl. Tested fungicides showed weaker translaminar activity, as the disease inhibition was marginal when applied on adaxial leaf surface. Partial systemic activity of azoxystrobin was evident by root uptake, while trifloxystrobin and kresoxim-methyl showed lack of systemic action in pearl millet. A trend in protection against downy mildew disease similar to greenhouse results was evident in the field trials. Grain yield was significantly increased in all strobilurin fungicide treatments over control and maximum increase in yield of 1673 kg ha−1 was observed in combination treatments of seed treatment and foliar spray with azoxystrobin.  相似文献   

4.
BACKGROUND: Cucurbit powdery mildew elicited by Podosphaera fusca (Fr.) U Braun & N Shishkoff limits crop production in Spain. Disease control is largely dependent on fungicides such as sterol demethylation inhibitors (DMIs). Fungicide resistance is an increasing problem in this pathogen. To overcome such risk, it is necessary to design rational control programmes based upon knowledge of field resistance. The aim of this study was to investigate the state of DMI sensitivity of Spanish P. fusca populations and provide tools for improved disease management. RESULTS: Using a leaf‐disc assay, sensitivity to fenarimol, myclobutanil and triadimenol of 50 isolates of P. fusca was analysed to determine discriminatory concentrations between sensitive and resistant isolates. As no clearly different groups of isolates could be identified, discriminatory concentrations were established on the basis of maximum fungicide field application rate, 100 mg L?1 for the three fungicides tested. Subsequently, a survey of DMI resistance was carried out in different provinces located in the south central area of Spain during the cucurbit growing seasons in 2002, 2003 and 2004. Examination of a collection of 250 isolates revealed that 23% were resistant to fenarimol and 7% to triadimenol, the provinces of Almería, Badajoz and Murcia being the locations with the highest frequencies of resistance. By contrast, no resistance to myclobutanil was found. CONCLUSION: Results show that fenarimol and, to a lesser extent, triadimenol have become less efficient for controlling cucurbit powdery mildew in Spain. These are important observations that should lead to reconsideration of the current disease management programmes. Copyright © 2010 Society of Chemical Industry  相似文献   

5.
Tan spot, caused by the fungus Pyrenophora tritici-repentis (Ptr), is a disease that has become more prevalent and intense in wheat crops in Argentina in recent years. Failure to control the disease with strobilurin fungicides, which were once effective, has been observed in different zones where wheat is grown. However, whether or not true resistance is present in the pathogen population in the region is not scientifically confirmed. This study evaluated the sensitivity of numerous Ptr isolates to representative QoI fungicides used in Argentina through in vitro and in planta assays, as well as through molecular analysis. Eighty-two monosporic isolates obtained in different locations in the north and south of Buenos Aires province in 2014, 2016, and 2018 were tested to determine sensitivity to selected QoI fungicides in conidial germination and mycelial inhibition assays, as well as in molecular analysis. Conidial germination was not inhibited at 1 µg/ml of azoxystrobin, trifloxystrobin, and pyraclostrobin. On the other hand, mycelial growth was inhibited by 59%, 56%, and 86% at 100 µg/ml of azoxystrobin, trifloxystrobin, and pyraclostrobin, respectively. The molecular analysis detected the G143A mutation in the cytb gene of all the 82 Ptr isolates, but the F129L and G137R substitutions were not present. This study documents the G143A mutation conferring QoI resistance in Ptr in South America. The findings of this study are key for future decisions regarding use of fungicide and rotation in the region.  相似文献   

6.
Stem rot caused by Lasiodiplodia theobromae is an important postharvest disease of papaya in Brazil, responsible for reducing the quality and quantity of fruits. Fungicide use is one of the main disease management measures. However, there are no estimates available of pathogen sensitivity to commonly employed fungicides. Therefore, the EC50 from 120 isolates of L. theobromae from northeastern Brazil, representative of six populations of the pathogen, was estimated in vitro for fungicides of the methyl benzimidazole carbamates—MBC (benomyl and thiabendazole) and demethylation-inhibiting—DMI (imazalil, prochloraz, tebuconazole) groups. Mycelial growth on fungicide-free media and virulence on papaya fruits of the MBC-sensitive and non-sensitive isolates were compared. For MBCs, 8.4% of isolates were non-sensitive to fungicides. For the remaining 91.6%, the mean EC50 ranged from 0.002 to 0.13 μg ml−1 and 0.36 to 1.27 μg ml−1 for benomyl and thiabendazole, respectively. For DMIs, the mean EC50 range for imazalil was 0.001 to 2.27 μg ml−1, 0.04 to 1.75 μg ml−1 for prochloraz, and 0.14 to 4.05 μg ml−1 for tebuconazole. The EC50 values of non-sensitive isolates were significantly (P≤0.05) higher those for the sensitive isolates for each of the DMI fungicides. Differences (P≤0.05) were found in the levels of sensitivity to DMI fungicides among the isolate populations associated with orchards. The populations from two orchards were less sensitive to DMIs. No solid evidence was found for fitness costs relating to MBC non-sensitive isolates because mycelial growth in fungicide-free media and virulence on papaya fruits were similar to those of sensitive isolates.  相似文献   

7.
Cercospora species cause cercospora leaf blight (CLB) and purple seed stain (PSS) on soybean. Because there are few resistant soybean varieties available, CLB/PSS management relies heavily upon fungicide applications. Sensitivity of 62 Argentinian Cercospora isolates to demethylation inhibitor (DMI), methyl benzimidazole carbamate (MBC), quinone outside inhibitor (QoI), succinate dehydrogenase inhibitor (SDHI) fungicides, and mancozeb was determined in this study. All isolates were sensitive to difenoconazole, epoxiconazole, prothioconazole, tebuconazole, and cyproconazole (EC50 values ranged from 0.006 to 2.4 µg/ml). In contrast, 51% of the tested isolates were sensitive (EC50 values ranged from 0.003 to 0.2 µg/ml), and 49% were highly resistant (EC50 > 100 µg/ml) to carbendazim. Interestingly, all isolates were completely resistant to azoxystrobin, trifloxystrobin, and pyraclostrobin, and insensitive to boscalid, fluxapyroxad, and pydiflumetofen (EC50 > 100 µg/ml). The G143A mutation was detected in 82% (53) of the QoI-resistant isolates and the E198A mutation in 97% (31) of the carbendazim-resistant isolates. No apparent resistance mutations were detected in the succinate dehydrogenase genes (subunits sdhB, sdhC, and sdhD). Mancozeb completely inhibited mycelial growth of the isolates evaluated at a concentration of 100 µg/ml. All Argentinian Cercospora isolates were sensitive to the DMI fungicides tested, but we report for the first time resistance to QoI and MBC fungicides. Mechanism(s) other than fungicide target-site modification may be responsible for resistance of Cercospora to QoI and MBC fungicides. Moreover, based on our results and on the recent introduction of SDHI fungicides on soybean in Argentina, Cercospora species causing CLB/PSS are insensitive (naturally resistant) to SDHI fungicides. Insensitivity must be confirmed under field conditions.  相似文献   

8.
Up to 2005 the sensitivity of Stemphylium vesicarium (Wallr.) Simm., the causal agent of pear brown spot, to the strobilurin fungicides kresoxim-methyl, trifloxystrobin and pyraclostrobin was still comparable with baseline values associated with good efficacy in the field. During 2006, the first resistant isolates were detected in two commercial pear orchards in the Emilia-Romagna region (Italy), one of which was affected by considerable control failure linked to strobilurin treatments as demonstrated in a field trial. In vitro sensitivity tests with 0.5 mg l−1 of kresoxim-methyl, trifloxystrobin and pyraclostrobin showed that in the population collected in the orchard with control failure the conidial germination was greater than 90% compared to an untreated control both in 2006 and in 2007, i.e. 1 year after the suspension of strobilurin applications. In the other orchard, where only a few symptomatic fruits were found and the strobilurins were still in use, the conidial germination was lower, about 50% in 2006 and 25% in 2007. The molecular analysis of mitochondrial cytochrome b gene of some monospore isolates with different levels of sensitivity confirmed the presence of the mutation causing G143A substitution in all the resistant isolates. In conclusion, both in vitro tests and molecular analysis confirmed the first occurrence of Stemphylium vesicarium resistance to all strobilurin fungicides tested.  相似文献   

9.
The baseline sensitivity of Botrytis cinerea to propamidine and assessment of the risk of propamidine resistance in vitro are presented in this article. The baseline sensitivities of 41 wild-type strains were distributed as a unimodal curve with EC50 values of mycelial growth ranging from 0.182 to 1.460 μg ml−1, with a mean of 0.79 ± 0.27 μg ml−1. A total of 10 resistant mutants, obtained from one parental strain, were induced by UV irradiation and selected for resistance to propamidine with an average frequency of 1.98 × 10−9 and 0.025 respectively. These mutants were divided into three classes of resistant phenotypes with low (LR), moderate (MR) and high (HR) levels of resistance, determined by the EC50 values of 5.0–15.0 μg ml−1, 15.1–75.0 μg ml−1 and more than 75.0 μg ml−1 respectively. Neither positive cross-resistance nor negative cross-resistance was detected between propamidine and the fungicides, benzimidazole carbendazim, anilino-pyrimidine pyrimethanil, dicarboximide iprodione or procymidone. All 10 propamidine-resistant mutants showed reduced mycelial growth in vitro, sporulation, spore germination and pathogenicity when compared with the parental strain. These studies demonstrated that propamidine possesses a low risk of resistance developing. However, as B. cinerea is a high-risk pathogen, appropriate precautions against resistance development should be taken.  相似文献   

10.
Qo inhibitor (QoI) fungicides are used to control gray blight caused by Pestalotiopsis longiseta in Japanese tea cultivation. However, field isolates of P. longiseta highly resistant to QoI fungicides were found in 2008, resulting in failure of QoI fungicidal control. This resistance was attributed to a mutation in the cytochrome b gene (cytb) in which alanine was substituted for glycine at position 143 (G143A). In 2009–2010, we detected field isolates that had an intermediate reaction between sensitive and resistant isolates in a preliminary assay. These isolates showed intermediate sensitivity to azoxystrobin and kresoxim-methyl on PDA plates. The intermediate reaction to azoxystrobin was also confirmed on detached tea leaves. Consequently, they were considered moderately resistant to QoI fungicides. Nucleotide sequencing of cytb showed that moderate resistance correlated with a single point mutation; leucine was substituted for phenylalanine at amino acid position 129 (F129L). Sequence analysis also revealed two types of cytb, with or without an intron between codons 131 and 132, in P. longiseta. F129L and G143A mutations were detected in both types of cytb according to their QoI resistance. This result suggests that G143A and F129L mutations have each occurred at least twice in the P. longiseta population.  相似文献   

11.
The baseline sensitivity ofFusarium graminearum Schwade [teleomorph =Gibberella zeae (Schweinitz) Petch] to the fungicide JS399-19 (development code no.) [2-cyano-3-amino-3-phenylacrylic acetate] and the assessment of risk to JS399-19 resistancein vitro are presented. The mean EC50 values for JS399-19 inhibiting mycelial growth of three populations of wild-typeF. graminearum isolates were 0.102±0.048, 0.113±0.035 and 0.110±0.036 μg ml−1, respectively. Through UV irradiation and selection for resistance to the fungicide, we obtained a total of 76 resistant mutants derived from five wild-type isolates ofF. graminearum with an average frequency of 1.71 × 10−7% and 3.5%, respectively. These mutants could be divided into three categories of resistant phenotypes with low (LR), moderate (MR) and high (HR) level of resistance, determined by the EC50 values of 1.5–15.0 μg ml−1, 15.1–75.0 μg ml−1 and more than 75.0 μg ml−1, respectively. There was no positive cross-resistance between JS399-19 and fungicides belonging to other chemical classes, such as benzimidazoles, ergosterol biosynthesis inhibitors and strobilurins, suggesting that JS399-19 presumably has a new biochemical mode of action. Although the resistant mutants appeared to have comparable pathogenicity to their wild-type parental isolates, they showed decreased mycelial growth on potato-sucrose-agar plates and decreased sporulation capacity in mung bean broth. Nevertheless, most of the resistant mutants possessed fitness levels comparable to their parents and had MR or HR levels of resistance. As these studies yielded a high frequency of laboratory resistance inF. graminearum, appropriate precautions against resistance development in natural populations should be taken into account. http://www.phytoparasitica.org posting August 7, 2008.  相似文献   

12.
Quinone outside inhibitors (QoIs) and succinate dehydrogenase inhibitors (SDHIs) are major groups of agricultural fungicides. However, resistance to some of these fungicides has been reported in a Japanese population of Puccinia horiana, the causal agent of chrysanthemum white rust disease. Because their mechanisms are not well understood, we investigated the existence of mutations in QoI and SDHI target protein-encoding genes. Eight out of nine isolates from cultivated chrysanthemum carried L275F and L299F amino acid substitutions in cytochrome b, the target protein of QoIs. These isolates showed 23- and 17-fold higher EC50 values for the QoI fungicides azoxystrobin and kresoxim-methyl, respectively, in basidiospore germination inhibitory tests, while they were hypersensitive to another QoI, famoxadone. All nine isolates were resistant to SDHI oxycarboxin and carried the I88F substitution in SdhC. This substitution was orthologous to the SdhC-I86F substitution found in some Brazilian isolates of the soybean rust fungus, Phakopsora pachyrhizi, showing reduced sensitivity to some SDHIs. Although the rarity of wild-type sensitive isolates, the subsequent limited number of comparisons between wild types and mutants, and a difficulty in applying reverse genetic analysis to this obligate parasite, are obstacles in making definitive conclusions, L275F and L299F in cytochrome b and SdhC-I88F are suspected to be responsible for the different patterns of sensitivity to QoI and for oxycarboxin-resistance in P. horiana, respectively.  相似文献   

13.
BACKGROUND: A single nucleotide polymorphism in the mitochondrial cytochrome b gene confers resistance to strobilurin (QoI) fungicides in phytopathogenic fungi. Recent studies have revealed worrying levels of resistance to strobilurins in Podosphaera fusca (Fr.) U Braun & N Shishkoff comb. nov. [ = Sphaerothecafusca (Fr.) S Blumer], the main causal agent of cucurbit powdery mildew in Spain. In the present study the underlying resistance mechanism to QoI fungicides in the Spanish populations of P. fusca was investigated. RESULTS: Analysis of the Q(o) domains of cytochrome b in a collection of isolates revealed that none of the typical mutations conferring resistance to QoI, including the G143A and F129L substitutions, was present in the QoI-resistant isolates. Moreover, although different amino acid polymorphisms were observed in the two regions spanning the Q(o) site, none of them consistently distinguished QoI-resistant from QoI-sensitive strains. Exposure to salicylhydroxamic acid (SHAM), a specific inhibitor of alternative oxidase, in the presence of trifloxystrobin did not have any effect on QoI resistance, ruling out alternative respiration as the mechanism of resistance. Sensitivity tests to a battery of respiration inhibitors revealed high levels of cross-resistance to all Qo-inhibitors tested but not to Qi-inhibitors, these features resembling those of a target-site-based resistance. CONCLUSIONS: The results indicate that the mechanism responsible for QoI resistance in P. fusca is not linked to typical mutations in cytochrome b gene and that the absence of the G143A substitution cannot be explained by an intron following codon 143. These are important observations, especially in relation to the possible molecular diagnosis of resistance.  相似文献   

14.
BACKGROUND: Cercospora leaf spot (CLS), caused by the fungus Cercospora beticola, is the most serious foliar disease of sugar beet (Beta vulgaris L.) worldwide. Disease control is mainly achieved by timely fungicide applications. In 2011, CLS control failures were reported in spite of application of quinone outside inhibitor (QoI) fungicide in several counties in Michigan, United States. The purpose of this study was to confirm the resistant phenotype and identify the molecular basis for QoI resistance of Michigan C. beticola isolates. RESULTS: Isolates collected in Michigan in 1998 and 1999 that had no previous exposure to the QoI fungicides trifloxystrobin or pyraclostrobin exhibited QoI EC50 values of ?0.006 µg mL?1. In contrast, all isolates obtained in 2011 exhibited EC50 values of > 0.92 µg mL?1 to both fungicides and harbored a mutation in cytochrome b (cytb) that led to an amino acid exchange from glycine to alanine at position 143 (G143A) compared with baseline QoI‐sensitive isolates. Microsatellite analysis of the isolates suggested that QoI resistance emerged independently in multiple genotypic backgrounds at multiple locations. A real‐time PCR assay utilizing dual‐labeled fluorogenic probes was developed to detect and differentiate QoI‐resistant isolates harboring the G143A mutation from sensitive isolates. CONCLUSION: The G143A mutation in cytb is associated with QoI resistance in C. beticola. Accurate monitoring of this mutation will be essential for fungicide resistance management in this pathosystem. Copyright © 2012 Society of Chemical Industry  相似文献   

15.
Anthracnose, caused by Colletotrichum gloeosporioides, is one of the most important diseases in grape-growing regions worldwide. In Jiangsu Province of China, quinone-outside inhibitor fungicides (QoIs) have been extensively sprayed as disease control for more than 10 years. A spore germination assay of 64 isolates obtained from 32 commercial vineyards was used to assess isolate sensitivity to azoxystrobin and 62 were found to be resistant to azoxystrobin. The biological fitness of QoI-resistant (QoIR) isolates was significantly lower than the sensitive isolates (QoIS) in terms of mycelial growth and conidiation. Nucleotide sequence alignment of CgCytb genes from the QoIR and QoIS isolates revealed that two point mutations (F129L and G143A) are involved in the QoI resistance. Isolates with the G143A mutation expressed high resistance to azoxystrobin, whereas isolates carrying the F129L mutation exhibited moderate resistance. Positive cross-resistance was observed between azoxystrobin and kersoxim-methyl, pyraclostrobin, or benzothiostrobin, but not with fluazinam. This study provides important information for management of QoIR populations of C. gloeosporioides in the field.  相似文献   

16.
Mutants of Phytophthora infestans with high resistance to the amidocarbamates iprovalicarb and benthiavalicarb and to the cyanoimidazole cyazofamid were isolated after UV-mutagenesis and selection on media containing one of the above fungicides. In vitro fungitoxicity tests showed that all resistant strains presented a highly reduced sensitivity to both cyazofamid and to the amidocarbamates. Cross-resistance studies with other oomycete fungicides from different chemical groups showed that the mutation(s) for resistance to iprovalicarb (IPV), benthiavalicarb (BVC) and cyazofamid (CZF) also greatly reduced the sensitivity of mutant strains to the phenylamide metalaxyl, acetamide cymoxanil, morpholine dimethomorph, benzamide zoxamide and to chlorothalonil. A lower reduction of sensitivity of mutant strains to the strobilurins azoxystrobin, kresoxim-methyl, pyraclostrobin and trifloxystrobin, azolones famoxadone and fenamidone and to antimycin A was observed. A resistance correlation was not apparent for the dithiocarbamate propineb and phenylpyridinamine fluazinam. Studies of fitness parameters in the wild-type and mutant strains of P. infestans showed that most resistant isolates had significantly reduced sporulation and sporangial germination, but not in the differentiation of sporangia into zoospores. Pathogenicity tests on tomato seedlings showed that most resistant isolates were significantly less pathogenic compared to the wild-type parent strain. However, experiments on the stability of the resistant phenotypes did not show a reduction in resistance when the mutants were grown for more than eleven generations on inhibitor-free medium. This is believed to be the first report of high level multi-drug resistance in fungal pathogens to chemically unrelated fungicides inhibiting different sites of cellular pathway.  相似文献   

17.
Wild type (WT) field isolates of Bremia lactucae failed to germinate in vitro or infect lettuce leaves in the presence of CAA (carboxylic acid amide) fungicides. Minimal inhibitory concentrations (MIC) for mandipropamid, dimethomorph, benthiavalicarb and iprovalicarb were 0.005, 0.5, 0.5 and 5 μg ml−1, respectively. Mutagenesis experiments showed that spores exposed to EMS (ethyl methane sulphonate) or UV irradiation (254 nm) could infect lettuce leaves in the presence of up to 100 μg ml−1 CAA. The proportion of infected leaves relative to the number of spores inoculated (infection frequency) was inversely related to the concentration of CAA used, ranging between 0 and 160 per 1 × 106 spores. Resistant mutants (RM) lost their resistance within 1–14 reproduction cycles on CAA-treated plants. Crosses were made between RMxWT isolates and RMxRM isolates with an attempt to obtain stable homozygous resistant off-springs. Such crosses yielded few resistant but unstable progeny isolates. Mutagenic treatments given to hybrid isolates also failed to produce stable resistance. Previous gene sequencing data showed that stable resistance to CAAs is based on a single SNP in the cellulose synthase 3 (CesA3) gene of Plasmopara viticola. Therefore, we sequenced a 582 bp DNA fragment of Ces3A of WT, RM and hybrid isolates of B.lactucae. No mutation in this gene fragment was found. We conclude that mutagenic agents like EMS or UV may induce resistance to CAA in Bremia lactucae but this resistance is not stable and not linked to mutations in CesA3 gene.  相似文献   

18.
BACKGROUND: It is possible that a single nucleotide polymorphism (SNP) (G143A mutation) in the cytochrome b gene could confer resistance to quinone outside inhibiting (QoI) fungicides (strobilurins) in rice blast fungus because this mutation caused a high level of resistance to fungicides such as azoxystrobin in Pyricularia grisea Sacc. and other fungal plant pathogens. The aim of this study was to survey Magnaporthe oryzae B Couch sp. nov. isolates in Japan for resistance to QoIs, and to try to develop molecular detection methods for QoI resistance. RESULTS: A survey on the QoI resistance among M. oryzae isolates from rice was conducted in Japan. A total of 813 single‐spore isolates of M. oryzae were tested for their sensitivity to azoxystrobin using a mycelial growth test on PDA. QoI fungicide resistance was not found among these isolates. The introduction of G143A mutation into a plasmid containing the cytochrome b gene sequence of rice blast fungus was achieved by site‐directed mutagenesis. Molecular diagnostic methods were developed for identifying QoI resistance in rice blast fungus using the plasmid construct. CONCLUSION: As the management of rice blast disease is often dependent on chemicals, the rational design of control programmes requires a proper understanding of the fungicide resistance phenomenon in field populations of the pathogen. Mutation of the cytochrome b gene of rice blast fungus would be specifically detected from diseased leaves and seeds using the molecular methods developed in this study. Copyright © 2009 Society of Chemical Industry  相似文献   

19.
The sterol biosynthesis inhibitors bromuconazole and difenoconazole and tank mixes of each fungicide with captan were applied to apples and evaluated as controls for moldy-core and fruit decay caused by Alternaria alternata. Effectiveness of a mixture of bromuconazole and captan in controlling colonization by the fungus was also evaluated. Decay formation by A. alternata on mature detached fruits was partially inhibited by bromuconazole at 0.5 μg ml−1 and was completely inhibited at 50 μg ml−1; it was significantly affected by either bromoconazole at 5 μg ml−1 or captan at 1,250 μg ml−1, and was completely inhibited by their mixture. In general, three foliar applications of bromuconazole or difenoconazole in the field, during the bloom period, reduced the numbers of infected fruits by 40–60% compared with untreated control trees. However, tank mixes of either fungicide with captan improved control of moldy-core in fruits at harvest. Tank mixtures of bromuconazole and captan also significantly reduced the percentage of fruits colonized by A. alternata when sampled at various days after full bloom. Artificial inoculations in the orchard at full bloom did not change the inhibitory effects of the tank mixtures. Large-scale demonstration trials in commercial orchards supported these findings. The inhibitory effects of tank mixes on decay development in detached fruits, and on moldy-core in the field indicate that a control programme based on mixtures of either bromuconazole or difenoconazole with captan during the bloom period can effectively reduce moldy-core on Red Delicious apples.  相似文献   

20.
Indoor and field experiments were conducted to evaluate the efficacy of applying the mycoparasite Coniothyrium minitans to the aerial parts of rapeseed plants at the flowering stage to control sclerotinia diseases caused by Sclerotinia sclerotiorum. Under controlled conditions, a petal inoculation technique was used to determine the effect of conidial suspensions of C. minitans on suppression of sclerotinia leaf blight. Results showed that C. minitans was effective in inhibiting infection initiated by ascospores of S. sclerotiorum on flower petals by restricting mycelial growth of the pathogen. Suppression of lesion development was related to the conidial concentration of C. minitans, with larger lesions at low concentration (5×103conidia ml−1), but smaller lesions at high concentration (5×104 conidia ml−1 or higher). When C. minitans-treated rapeseed leaves were inoculated with mycelia of S. sclerotiorum, C. minitans failed to prevent infection of leaves, but caused a significant reduction in number of sclerotia produced on the diseased leaves. No significant difference in efficacy was detected between the two isolates of C. minitans, LRC 2137 and Chy-1, on the two rapeseed cultivars, Westar (spring type) and Zhongyou 821 (winter type). Results of field trials showed a significant reduction of stem rot of rapeseed in four (1997, 1999, 2003 and 2004) out of five years by aerial application of C. minitans, compared with controls. No significant difference in suppressive efficacy was observed between the treatments of C. minitans (106 conidia ml−1), C. minitans (106 conidia ml−1) + benomyl (50 μg ml−1) and benomyl (100 μg ml−1) in 2003, and between the treatments of C. minitans (106 conidia ml−1), C. minitans (106 conidia ml−1) + vinclozolin (100 μg ml−1) and vinclozolin (500 μg ml−1) in 2004. Sclerotia of S. sclerotiorum collected from diseased plants in plots treated with C. minitans in 1999, 2000 and 2003, or with C. minitans + benomyl in 2003 were infected by C. minitans at frequencies ranging from 21.3 to 54.5%. This study concludes that aerial spraying of C. minitans is an effective method for controlling sclerotinia diseases of rapeseed.  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