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.
Uptake of [14C]fenarimol (30 μM) by mycelium of wild-type Aspergillus nidulans was characterized by a rapid initial accumulation during the first 10 min of incubation with the fungicide and a subsequent gradual release with time. Uptake appeared to be the result of influx and efflux. Influx of fenarimol could not be inhibited by low temperature, anaerobiosis, starvation of mycelium, or incubation with several respiratory inhibitors and is, therefore, a passive process. Under identical test conditions efflux activity was severely inhibited and should, therefore, be regarded as an energy-dependent mechanism. After prolonged incubation (90 min) an equilibrium between influx and efflux was established, resulting in an energy-dependent permeability barrier, since uptake could instantaneously be enhanced by addition of oligomycin or N,N′-dicyclohexylcarbodiimide. It also indicates that efflux activity is inducible; this hypothesis is supported by the observation that pretreatment of mycelium with unlabeled fungicide prevented subsequent uptake of [14C]fenarimol. Uptake by fenarimol-resistant mutants J146, M193, and R264 of A. nidulans, all possessing the imaB gene for resistance, was relatively low and almost constant in time. In this case, uptake appeared to be considerably enhanced by low temperature, anaerobiosis, starvation of mycelium, and incubation with respiratory inhibitors. Low uptake by these mutants is ascribed to a higher energy-dependent efflux activity for fenarimol compared with the wild-type strain. Upon inhibition of the barrier activity, net uptake resulted from remaining passive influx, which in that case may be as high as in the wild-type strain. The results suggest that both wild-type and fenarimol-resistant mutants possess an energy-dependent efflux mechanism with different efficiencies to excrete fenarimol and probably other chemicals to which cross-resistance or collateral sensitivity is present.  相似文献   

2.
Mycelial uptake of [14C]fenarimol (10 μg/ml) by 20 fenarimol-resistant mutants of Aspergillus nidulans was compared with uptake by wild-type strain 003. Uptake of the fungicide during the initial 10 min of incubation was significantly lower in all mutant strains than in the wild-type strain indicating that resistance is related with reduced uptake. Upon prolonged incubation a gradual decrease of accumulated radioactivity in the wild-type strain was observed. A few mutants displayed resistance to unrelated chemicals such as p-fluorophenylalanine or d-serine; this phenomenon appeared not to be due to a decreased uptake of the corresponding natural amino acids. Incorporation of [3H]adenine and [14C]leucine by mycelium of mutant M193 was hardly inhibited after 5 hr of incubation with the fungicide, whereas a distinct effect was found with the wild-type strain. At this time also fungitoxicity to the wild-type strain became apparent. Probably, this effect is indirectly caused by inhibition of ergosterol biosynthesis. Mycelium of mutant M193 incorporated [14C]acetate slightly less effectively than the wild-type strain. After 2 hr of incubation with this radiochemical leakage of [14C]acetate metabolites from mycelium of the mutant strain was observed. This indicates that resistance might be correlated with increased excretion of fungal metabolites, which in turn may be related with reduced fitness of fenarimol-resistant mutants.  相似文献   

3.
Metabolism of imazalil (1-[2-(2,4-dichlorophenyl)-2-(2-propenyloxy)ethyl]-1H-imidazole) inPenicillium italicum isolates with a wild-type sensitivity and with various degrees of resistance to sterol demethylation inhibitors was studied in liquid cultures. The metabolite 1-[2(2,4-dichlorophenyl)-2-(2,3-dihydroxypropyloxy)ethyl]-1H-imidazole (R42243) was detected in the culture filtrate after prolonged incubation. The metabolism occurred in the propenyl side chain of imazalil probably through epoxidation and hydratation. This is the first report of such a conversion of imazalil in fungi. R42243 was much less toxic toP. italicum than imazalil. Therefore, the metabolism can be regarded as a detoxification step. Both wild-type and resistant isolates metabolized imazalil, but metabolism by resistant isolates was faster than by the wild-type isolate. This is probably caused by a relatively strong inhibition of growth of the wild-type isolate by the fungicide. Results indicate that the detoxification of imazalil does not operate as a mechanism of resistance. This conclusion was confirmed by the fact that resistant isolates showed cross-resistance to miconazole and R42243, which had a similar structure as imazalil except for the propenyl side chain.  相似文献   

4.
抗戊唑醇禾谷丝核菌的渗透压敏感性及相对渗率变化研究   总被引:11,自引:7,他引:4  
经药剂筛选和UV诱导,分别获得了抗戊唑醇的禾谷丝核菌Rhizoctonia cerealis菌株,比较了其与敏感菌株对渗透压的敏感性和经不同浓度戊唑醇处理后的相对渗率变化。在相同的渗透压环境下,敏感菌株对渗透压较抗性菌株更为敏感。不同抗性菌株间对渗透压的敏感性在低渗和中渗环境下差异不大,在高渗环境下有一定的差异;而敏感菌株只对低渗透压(1%葡萄糖)和高渗透压(8%葡萄糖)相对较为敏感。用不同浓度的戊唑醇处理后,与敏感菌株相比,抗性菌株能在较短的时间里(1 h)渗出更多的电解质,但最终前者的相对渗率要明显高于后者。用高浓度的戊唑醇处理抗性菌株后,其相对渗率明显高于低浓度处理。  相似文献   

5.
赣南脐橙绿霉病菌对常用杀菌剂抗性监测   总被引:2,自引:0,他引:2  
 本文研究了来自赣南7个县的柑橘绿霉病菌(Penicillium digitatum)种群对该地区常用杀菌剂抑霉唑、咪鲜胺、甲基硫菌灵和百可得的抗性频率、抗性水平和对抑霉唑的抗性分子机制。结果表明:病菌对抑霉唑和咪鲜胺存在基本一致的抗性;2011和2012年病菌种群对抑霉唑和咪鲜胺的抗性频率分别为82%和90%,平均抗性倍数为51.5倍,抗性分子机制均属于IMZ-R3,即CYP51B基因启动子区发生199 bp插入的突变;病菌种群对甲基硫菌灵的抗性频率分别为82%和91%;病菌种群对百可得均表现敏感。本研究为采后柑橘病害防治药剂选择提供了科学的依据。  相似文献   

6.
Imazalil differentially inhibited dry weight increase of 10-hour-old germlings of wild-type and DMI-resistant isolates ofPenicillium italicum in liquid malt cultures. EC50 values ranged from 0.005 to 0.27 μg ml?1. In all isolates ergosterol constituted the major sterol (over 95% of total sterols) in the absence of the fungicide. Therefore, DMI-resistance cannot be associated to a deficiency of the C-14 demethylation enzyme in the ergosterol biosynthetic pathway. Imazalil treatment at concentrations around EC50 values for inhibition of mycelial growth resulted in a decrease in ergosterol content and a simultaneous increase in 24-methylene-24,25-dihydrolanosterol content in all isolates. A correlation existed between the imazalil concentration necessary to induce such changes in sterol composition and the EC50 values for inhibition of mycelial growth of the different isolates. The reason for the differential effects of imazalil on sterol composition in the variousP. italicum isolates may be due to decreased accumulation of the fungicide in the mycelium and to other yet non-identified mechanisms of resistance.  相似文献   

7.
The acylalanine fungicides CGA 29212 and metalaxyl inhibit colony growth of Phytophthora megasperma f. sp. medicaginis at much lower concentrations than structurally related chloroacetanilide herbicides. Metolachlor, among the latter, shows the highest antifungal activity, followed by propachlor. Dimethachlor and alachlor are only weakly inhibitory. A metalaxyl- and CGA 29212-resistant strain of P. megasperma f. sp. medicaginis shows cross-resistance to metolachlor and propachlor, but levels of resistance are much lower than observed with CGA 29212 and metalaxyl. Cross-resistance does not occur to dimethachlor and alachlor. All compounds except metalaxyl inhibit uptake of [3H]uridine by mycelium, propachlor being most effective. Effects are similar with both a metalaxyl-sensitive and a -resistant strain. CGA 29212, metalaxyl, and metolachlor inhibit incorporation of [3H]uridine into RNA by mycelium of the sensitive strain at concentrations not inhibitory to uptake. Metalaxyl slightly affects incorporation by mycelium of the resistant strain; the other compounds have a more pronounced effect but only at concentrations inhibitory to uptake. Metalaxyl, CGA 29212, and propachlor do not induce leakage of radioactivity from mycelium of both strains when added at high concentrations to cultures previously incubated with [3H]uridine; under these conditions incorporation by mycelium of the metalaxyl-resistant strain is significantly more affected by CGA 29212 and propachlor than by metalaxyl. Endogenous RNA polymerase activity of isolated nuclei from a metalaxyl-sensitive strain is inhibited by CGA 29212, metalaxyl, and metolachlor but not by propachlor, dimethachlor, and alachlor. Neither compound has any effect on endogenous RNA polymerase activity of isolated nuclei from a metalaxyl-resistant strain. CGA 29212 and metolachlor evidently have a metalaxyl-type of action. The presence of cross-resistance of the metalaxyl-resistant strain to propachlor also indicates a metalaxyl-type of action for this compound, although this could not be confirmed by an inhibitory effect of propachlor on endogenous RNA polymerase activity. In addition to a metalaxyl-type of action, CGA 29212, metolachlor, and propachlor have a second one that is also present in dimethachlor and alachlor, which lack the metalaxyl-type of action. The second mechanism of action, involving inhibition of [3H]uridine uptake, is most prominent with propachlor and might be related to the primary mechanism of action in plants of the chloroacetanilide herbicides.  相似文献   

8.
Applying thiabendazole to potato seed tubers affected with silver scurf caused by Helminthosporium solani sensitive to thiabendazole decreased the severity of disease on progeny tubers at harvest, but about 50% of the isolates from these were resistant to the fungicide. The disease was not decreased when samples of the progeny tubers were treated with thiabendazole and planted in the following year, and the incidence of resistant isolates increased. Resistant isolates continued to be present when tubers were planted in the next 2 years without fungicide treatment. Treatment with a mixture of thiabendazole and imazalil also decreased the disease and fewer isolates were resistant than when treated with thiabendazole alone, although the proportion increased after treatment with the mixture in the following year. When seed tubers were infected with thiabendazole-resistant H. solani , silver scurf on progeny tubers was not affected by thiabendazole applied to the seed tubers but was decreased by the mixture of thiabendazole and imazalil. Imazalil was equally effective against H. solani sensitive or resistant to thiabendazole.
Some isolates of H. solani had grey aerial mycelium and of 516 of these isolates obtained in 4 years 29% were resistant to the fungicide. Other isolates produced small, black colonies and their frequency increased with thiabendazole treatment of seed tubers. Of 244 of these isolates, 62% were resistant.  相似文献   

9.
3-Phenylindole is an antimicrobial compound active towards many fungi and gram-positive bacteria. At 5 μg/ml it inhibits growth of Aspergillus niger. Higher concentrations (50 μg/ml) also suppress spore germination; they do not kill the fungus. Dry weight of the fungus still increases for 1 or 2 days after fungicide treatment. The toxicant has no effect on O2 uptake even at higher concentrations (100 μg/ml). The compound markedly affects composition of the lipid fraction of A. niger inducing a decrease in phospholipid concentration with a coincident increase in free fatty acids. Sterol pattern and sterol concentration were not affected. Antifungal activity was reversed by phospholipids added to the medium. 3-Phenylindole induced a slight leakage of 32P-labeled compounds from the treated cells under growth conditions but not under nongrowth conditions. A strain of A. niger resistant to 3-phenylindole had the same phospholipid and sterol pattern as the wild type, but the level of both components was higher (40–60%). The 3-phenylindole-resistant strain showed resistance to triarimol and pimaricin. The wild type and the resistant strain both took up 3-phenylindole quite rapidly and accumulated it in the mycelium. 3-Phenylindole possibly interferes with phospholipid function in cell membranes, although the specific site of action has not yet been elucidated.  相似文献   

10.
Genetic control and mechanisms of resistance to tebuconazole, a sterol C14-demethylation inhibitor, were investigated in the phytopathogenic fungusNectria haematococca. Resistant mutants have been selected from the laboratory, following UV irradiation. They have been characterized through genetic crosses and mutations in at least three genes were found to be responsible for resistance. The genesTeb1, Teb2, andTeb3were clearly identified, a fourth gene calledTeb4could be hypothesized. Mutations at lociTeb2andTeb3induced pleiotropic effects such as reduced sporulation and growth rate, mycelium pigmentation (Teb2), or altered ascospore viability (Teb3). The resistance levels determined by mutations in the different genes were relatively low (below 10). When associated in double mutants, the additive effect was recorded. Cross-resistance toward other sterol C14-demethylation inhibitors was observed in all the resistant strains, except in theTeb4-carrying strain; moreover, for some C14-demethylation inhibitors hypersensitivity was expressed. A constitutive energy-dependent efflux seemed implicated in the mechanism of resistance for theTeb1-carrying strain and probably also for theTeb2andTeb3-carrying strains. However, theTeb4-carrying strain exhibited a kinetic of fungicide uptake similar to that of the wild-type strain. The sterol profile of theTeb4-carrying strain was similar to that of all the other resistant mutants and wild-type strains. Thus the resistance mechanism induced by mutation at theTeb4locus has not been found yet.  相似文献   

11.

Sensitivity and inherent resistance risk of Alternaria solani to fludioxonil, cross-resistance profiles and the potential implications of resistance mutations on fitness parameters were investigated. Fludioxonil was highly effective against a wild type A. solani field strain both in vitro (EC50?=?0.05 μg/mL) and in preventive applications on artificially inoculated tomato fruit. Mutants with low [Resistance factor (Rf): 15 based on EC50], medium (Rf: 150–300) and high (Rf: > 1000) levels of phenylpyrrole resistance were isolated from the wild type strain at high frequencies following mutagenesis with UV irradiation and selection on fludioxonil containing medium. Resistant isolates retained their resistance levels even after 9 subcultures on fungicide-free growth medium while they could express their resistant phenotypes in planta. Investigation of cross-resistance relationships showed that fludioxonil resistance mutations also reduce the sensitivity of mutant strains to the aromatic hydrocarbon fungicide quintozene as well as the dicarboximides iprodione and vinclozolin. No cross-resistance was observed between fludioxonil and fungicides with different modes of action such as the sterol biosynthesis inhibitors (DMIs) imazalil and flusilazole and the carboxamide boscalid. All fludioxonil resistant isolates were more sensitive to the anilinopyrimidine pyrimethanil, while only two isolates were less sensitive to the QoI pyraclostrobin compared to the wild-type strain. Study of fitness determining parameters showed that resistance mutation(s) had no adverse effects on mycelial growth, conidial germination and sensitivity to osmotic stress while they had a pleiotropic effect on virulence and conidia production in resistant mutants. Results of the present study indicate that fludioxonil is a highly effective fungicide against A. solani, while the risk of resistance development to this fungicide is considered to be medium making fludioxonil an ideal alternative to high risk fungicides such as boscalid and pyraclostrobin whose performance against early blight has already been compromised by resistance development.

  相似文献   

12.
Imazalil differentially inhibited dry weight increase of 10-hour-old germlings of wild-type and DMI-resistant isolates ofPenicillium italicum in liquid malt cultures. EC50 values ranged from 0.005 to 0.27 g ml–1. In all isolates ergosterol constituted the major sterol (over 95% of total sterols) in the absence of the fungicide. Therefore, DMI-resistance cannot be associated to a deficiency of the C-14 demethylation enzyme in the ergosterol biosynthetic pathway. Imazalil treatment at concentrations around EC50 values for inhibition of mycelial growth resulted in a decrease in ergosterol content and a simultaneous increase in 24-methylene-24,25-dihydrolanosterol content in all isolates. A correlation existed between the imazalil concentration necessary to induce such changes in sterol composition and the EC50 values for inhibition of mycelial growth of the different isolates. The reason for the differential effects of imazalil on sterol composition in the variousP. italicum isolates may be due to decreased accumulation of the fungicide in the mycelium and to other yet non-identified mechanisms of resistance.Imazalil remt differentieel de toename in drooggewicht van 10-uur-oude gekiemde sporen van wild-type en DMI-resistente isolaten vanPenicillium italicum in vloeistofcultures van moutextract. De EC50 waarden voor groei van de verschillende isolaten lopen uiteen van 0,005 tot 0,27 g ml–1. In afwezigheid van het fungicide is in alle isolaten ergosterol het belangrijkste sterol (meer dan 95% van het totaal). DMI-resistentie kan daarom niet in verband staan met deficiëntie van het C-14 demethyleringsenzym in de ergosterol biosynthese. Imazalilbehandeling van mycelium bij concentraties rond de EC50 waarde voor groeiremming, resulteerde bij alle isolaten in een afname van het ergosterolgehalte en een gelijktijdige toename van het gehalte aan 24-methyleen-24,25-dihydrolanosterol. Er bestaat dus een nauwe correlatie tussen de imazalilconcentratie die noodzakelijk is om vergelijkbare veranderingen in sterolsamenstelling te induceren en de EC50 waarde voor remming van myceliumgroei van de verschillende isolaten. De differentiële effecten van imazalil op de sterolsamenstelling van de verschillendeP. italicum isolaten kunnen worden veroorzaakt door verminderde accumulatie van het fungicide in het mycelium en door andere, nog niet geïdentificeerde resistentiemechanismen.  相似文献   

13.
Fenarimol-resistant laboratory isolates of Penicillium italicum exhibit a high degree of cross-resistance to triazoles, a low degree of cross-resistance to imazalil and a low degree of negatively-correlated cross-resistance to morpholines. The cross-resistance to triazoles is ascribed to a differential accumulation level of these SBI fungicides in wild-type and fenarimol-resistant isolates. As in the case of fenarimol, the relatively low accumulation level of triazoles in resistant isolates is probably caused by the operation of a constitutive, energy-dependent efflux from mycelium, which hinders the triazoles from reaching their site of action in sterol biosynthesis. The accumulation level of imazalil and morpholines did not differ among wild-type and fenarimol-resistant isolates. This may be related to the fact that these fungicides at physiological pHs occur in a protonated form. The lack of a differential accumulation level can be the reason for the deviating cross-resistance patterns of these fungicides.  相似文献   

14.
The fitness of twelve fenarimol-resistant mutants of Aspergillus nidulans was tested with respect to spore germination, germ tube elongation, hyphal growth and sporulation. Half of the strains tested were isolated on triarimolcontaining media. The other strains were selected on imazalil-or cycloheximide-containing media (Van Tuyl, 1977).A number of mutant strains produced spores with unimpaired viability on synthetic medium. In other strains a reduction in spore viability was noticed. The rate of germ tube elongation of all resistant mutants was significantly lower than that of the wild-type strain. Mutant strains with a low degree of resistance always had an almost normal mycelial growth rate, whereas growth of some of the strains with a relatively high degree of resistance was significantly slower. Spore production on malt agar was highest in the wild-type strain and was found to be lower in fenarimol-resistant mutants. In most of the mutant strains tested a high degree of cross-resistance between fenarimol, imazalil and triadimefon was established; in some of them cross-resistance to these chemicals was low or even absent.Possible implications of the results described for the chance of development of resistance in phytopathogenic fungi to sterol biosynthesis-inhibiting fungicides are discussed.  相似文献   

15.
The fine structure and sterol composition of wild-type and triazole-resistant laboratory strains of Ustilago avenae was investigated by electron microscopic and biochemical methods. The growth rate of the mutants was only slightly affected by a fungicide (triadimefon) concentration of about 0.1 mg/ml, whereas the wild-type cells were completely inhibited. Biochemically the sterol composition of wild-type and triazole-resistant strains did not differ. In freeze-fracture electron microscopy no ultrastructural differences were observed between the different untreated strains (wild and resistant). Filipin labeling allowed the localization of ergosterol in the plasmalemma (PF and EF). Generally, wild-type samples and mutants exhibited a clear pattern of filipin-sterol (FS-) complexes. These results are in accord with the biochemical experiments. Neither a modification of the sterol composition nor an altered localization of sterols seemed to be the prime cause of resistance in U. avenae mutants. Alternative explanations for the resistance mechanism are discussed.  相似文献   

16.
When the petioles of detached tobacco leaves (10–17 cm2) were incubated in aqueous solutions containing [14C]metalaxyl, uptake of the fungicide was dependent on the temperature and photoperiod. Detached leaves took up 78% more [14C]metalaxyl at 26°C than at 16°C. The rate of uptake in the light at 21°C was linear, but after an additional 20h in the dark, there was only twice as much fungicide in the leaves. Different sized leaves contained the same amount of fungicide per cm2 area. Uptake by detached leaves of the 14C-labelled anilide lactones ofurace and RE-26940 [2-methoxy-N-(tetrahydro-2-oxo-3-thienyl)acet-2′,6′-xylidide] was similar to that of metalaxyl. At the concentration of metalaxyl (66 ng ml?1) that controlled blue mould (Peronospora tabacina) on detached tobacco leaves, the amount of fungicide in the leaves was found to be 7.25 ng. Autoradiography showed that the distribution of [14C]metalaxyl in detached leaves after incubation for 23h was uniform, although higher concentrations of the label were present in the smaller veins of the leaves.  相似文献   

17.
The inherent resistance risk forMonilinia fructicola against sterol-biosynthesis inhibitors (SBIs) was estimated inin vitro andin vivo laboratory studies. Several mutant strains were selected on media amended with the triazole fungicides penconazole, etaconazole or the morpholine fungicide fenpropimorph.The potential forM. fructicola to develop resistance to the triazoles or to the morpholines was similar.The level of resistance attained did not differ for the two classes of fungicides after a single cycle of treatment with nitrosoguanidine (NTG). Attemps to select mutants with a higher level of resistance to penconazole after successive mutagenic treatments were successful. Most of the mutants were less fit than wild-type strains. Mutants with a low level of resistance had an almost normal mycelial growth rate, whereas growth of mutants with a higher level of resistance was significantly reduced. Spore production was highest in the wild-type strains, similar to the latter in a few resistant strains and less in most others. Only one mutant with an intermediate level of resistance could successfully compete in a mixed population with a wild type strain during successive infection cycles on peaches. Resistance was not stable in highly resistant mutants. Cross resistance to the inhibitors of 14-methylsterol demethylation (DMIs) tested was confirmedin vitro andin vivo for all mutant strains. One DMI-resistant mutant was also resistant to fenpropimorph and two fenpropimorph-resistant mutants were resistant to penconazole.  相似文献   

18.
The phenylpyrrole fungicide fenpiclonil accumulates to high levels in mycelium of Fusarium sulphureum and artificial liposomes. The accumulation is a result of a physicochemical distribution of the fungicide over lipid-like material of the mycelium and the medium. Accumulation is reversible, as the fungicide can easily be removed from the mycelium by washing with water. Fenpiclonil is not metabolized by the fungus during incubation for 24 h. The fungicide affects neither membrane fluidity in artificial liposomes nor amino acid accumulation in bacterial vesicles. Thus, accumulation of the fungicide does not seem to affect the functioning of membranes. Fenpiclonil induces the accumulation of the membrane-potential probe tetra-phenylphosphonium bromide by F. sulphureum. Since accumulation of this probe probably reflects the cumulative potential over all cell membranes, the increased accumulation is more likely a result of changes of potentials over membranes of intracellular organelles rather than plasma membrane hyperpolarization. Hence, the previously described effects of fenpiclonil on amino acid and saccharide uptake cannot be explained by plasma membrane hyperpolarization.  相似文献   

19.
The San Roman strain of the southern cattle tick, Boophilus microplus, collected from Mexico was previously reported to have a high level of resistance to the organophosphate acaricide coumaphos. An oxidative detoxification mechanism was suspected to contribute to coumaphos resistance in this tick strain, as coumaphos bioassay with piperonyl butoxide (PBO) on larvae of this resistant strain resulted in enhanced coumaphos toxicity, while coumaphos assays with PBO resulted in reduced toxicity of coumaphos in a susceptible reference strain. In this study, we further analyzed the mechanism of oxidative metabolic detoxification with synergist bioassays of coroxon, the toxic metabolite of coumaphos, and the mechanism of target-site insensitivity with acetylcholinesterase (AChE) inhibition kinetics assays. Bioassays of coroxon with PBO resulted in synergism of coroxon toxicity in both the San Roman and the susceptible reference strains. The synergism ratio of PBO on coroxon in the resistant strain was 4.5 times that of the susceptible strain. The results suggested that the cytP450-based metabolic detoxification existed in both resistant and susceptible strains, but its activity was significantly enhanced in the resistant strain. Comparisons of AChE activity and inhibition kinetics by coroxon in both susceptible and resistant strains revealed that the resistant San Roman strain had an insensitive AChE, with a reduced phosphorylation rate, resulting in a reduced bimolecular reaction constant. These data indicate a mechanism of coumaphos resistance in the San Roman strain that involves both insensitive AChE and enhanced cytP450-based metabolic detoxification.  相似文献   

20.
The spread of fungicide resistant and/or tolerant phytopathogenic fungi is an important factor affecting crop protection. However, the mechanisms of fungal response to fungicide application are not entirely characterised. In particular, the contribution of previously known resistance factors and the final influence of fungicide treatments on metabolism of surviving mycelia (e.g. mycotoxin increased release and biosynthesis potentially causing contamination of the crops) merit investigation, in order to improve future molecular diagnostics of fungicide resistant strains. The performed experiments have shown distinct expression changes for homologs of a known fungicide resistance factor Flr1 (yeast; DHA1 family of major facilitator superfamily transporters) after azole application in cultured fusaria. Two distantly related homologs of that gene were selected, based on the unsupervised clustering and phylogenetic analysis of transporter sequences. One of these (FGSG_02865), was found to occur across the Fusarium sambucinum complex (F. graminearum, F. culmorum, F. cerealis) and was upregulated starting 24 h after fungicide treatments. This delayed response may point to possible involvement of DHA1 antiporters in a generalised response to stress resulting from fungicide treatment. Additional expression profiling was conducted for the mycotoxin biosynthetic genes (trichothecene and zearalenone gene clusters) in strains of Fusarium sambucinum complex cereal pathogens. The expression changes, when subjected to treatment with the fungicides (flusilazole, carbendazim), show that even an effective treatment (in this study, the benzimidazole fungicide carbendazim) applied to the grown mycelium, can result in enhanced activation of mycotoxin biosynthetic genes in fungal cells which survive the treatment. Our results suggest that increased mycotoxin contamination can be strongly influenced not only by the amount or the type of antifungal compound, but also the timing of fungicide exposition (stage of infection).  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