首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Citrate and malate were poorer sources of exogenous carbon than several hexose, pentose, or disaccharide sugars for supporting macroconidial germination by Fusarium solani at high conidial density (1 X 10(5) condia/ml). Only citrate, however, failed to block chlamydospore morphogenesis to a degree comparable to glucose or other readily used sugars. Mostly immature chlamydospores were formed in the presence of citrate. At low conidial density (5 X 10(3) conidia/ml), exogenous carbon-independent macroconidial germination and subsequent rapid chalmydospore formation on germ tubes was not inhibited by ammonium or nitrate nitrogen. The citrate-phosphate buffered, low pH (4.0) medium of Cochrane induced more immature chlamydospore formation by F. solani than a pH 6.0 medium, but few mature chlamydospores were formed in either medium. Condensation of hyphal cytoplasm into developing chlamydospores, a character typical of chlamydospore formation, did not occur extensively and macroconidia, hyphae, and immature chlamydospores stained deeply with Sudan III, suggesting lipid biosynthesis. This inhibition of chlamydospore maturation may be due partly to nitrogen deficiency imposed by the high C:N ratio of the medium and to the presence of citrate. Only vesiculate hyphal cells were formed by F. solani f. sp. phaseoli in both media. Field soils to which the clone of F. solani used is indigenous had mean pH values ranging from 5.2 to 6.0.  相似文献   

2.
The effects of medium, isolate, temperature, light and pH on the formation and germination of Septoria tritici secondary conidia were tested. Of the six media tested, the malt–yeast extract agar was the best and generated 1.82 × 109 conidia/plate. The ten isolates tested showed different ability of conidia production. Darkness significantly reduced conidial formation and enhanced the transition of intermediates. The conidial germination and germ tube growth was strongly inhibited at 30°C. The suitable pH for conidial budding in malt–yeast broth (MYB) was between 5 and 9. At pH 2, 10 and 11, almost no new conidia were formed. The number of conidia reached 1.27 × 108 conidia/ml after 7 days in MYB, significantly more than that in potato dextrose broth, wheat leaf extract and H2O.  相似文献   

3.
Transfer ribonucleic acid (tRNA) methylases were studied during the germination of spores in Neurospora crassa. The total methylase capacity and base specific tRNA methylase activities were determined in extracts from cells harvested at various stages of germination. Germinated conidia have a 65% higher methylase capacity than ungerminated conidia. Three predominant methylase activities were found in the extracts, and the relative amount of each activity was different at the various stages. Enzymes from vegetative cells catalyzed significant hypermethylation of tRNA from conidia, whereas conidial enzymes were much less active on tRNA from vegetative cells. The results indicate differences in the tRNA methylase content and tRNA species of conidia and vegetative cells.  相似文献   

4.
Penicillium paneum is an important contaminant of cereal grains which is able to grow at low temperature, low pH, high levels of carbon dioxide, and under acid conditions. P. paneum produces mycotoxins, which may be harmful to animals and humans. We found that conidia in dense suspensions showed poor germination, suggesting the presence of a self-inhibitor. A volatile compound(s) produced by these high-density conditions also inhibited mycelial growth of different species of fungi belonging to a variety of genera, suggesting a broad action range. The heat-stable compound was isolated by successive centrifugation of the supernatant obtained from spore suspensions with a density of 10(9) conidia ml(-1). By using static headspace analyses, two major peaks were distinguished, with the highest production of these metabolites after 22 h of incubation at 25 degrees C and shaking at 140 rpm. Gas chromatography coupled with mass spectra analysis revealed the compounds to be 3-octanone and 1-octen-3-ol. Notably, only the latter compound appeared to block the germination process at different developmental stages of the conidia (swelling and germ tube formation). In this study, 1-octen-3-ol influenced different developmental processes during the P. paneum life cycle, including induction of microcycle conidiation and inhibition of spore germination. Therefore, the compound can be considered a fungal hormone during fungal development.  相似文献   

5.
The ascomycetous fungus Fusarium graminearum is an important plant pathogen causing Fusarium head blight disease of wheat and barley. To understand early developmental stages of this organism, we followed the germination of macroconidia microscopically to understand the timing of key events. These events, recorded after suspension of spores in liquid germination medium, included spore swelling at 2h, germination tube emergence and elongation from conidia at 8h and hyphal branching at 24h. To understand changes in gene expression during these developmental changes, RNA was isolated from spores and used to interrogate the F. graminearum Affymetrix GeneChip. RNAs corresponding to 5813 genes were detected in fresh spores and 5146, 5249 and 5993, respectively, in spores incubated in germination medium after 2, 8 or 24h (P<0.001). Gene expression data were used to predict the cellular and physiological state of each developmental stage for known processes. Predictions were confirmed microscopically for several previously unreported developmental events such as manifestation of peroxisomes in fresh spores and nuclear division resulting in binuclear cells within macroconidia prior to spore germination. Knowledge of stage-specific gene expression and changes in gene expression levels between developmental stages are an important first step to understanding the molecular mechanisms responsible for spore germination and development.  相似文献   

6.
The ascomycetous fungus Fusarium graminearum is an important plant pathogen causing Fusarium head blight disease of wheat and barley. To understand early developmental stages of this organism, we followed the germination of macroconidia microscopically to understand the timing of key events. These events, recorded after suspension of spores in liquid germination medium, included spore swelling at 2h, germination tube emergence and elongation from conidia at 8h and hyphal branching at 24h. To understand changes in gene expression during these developmental changes, RNA was isolated from spores and used to interrogate the F. graminearum Affymetrix GeneChip. RNAs corresponding to 5813 genes were detected in fresh spores and 5146, 5249 and 5993, respectively, in spores incubated in germination medium after 2, 8 or 24h (P<0.001). Gene expression data were used to predict the cellular and physiological state of each developmental stage for known processes. Predictions were confirmed microscopically for several previously unreported developmental events such as manifestation of peroxisomes in fresh spores and nuclear division resulting in binuclear cells within macroconidia prior to spore germination. Knowledge of stage-specific gene expression and changes in gene expression levels between developmental stages are an important first step to understanding the molecular mechanisms responsible for spore germination and development.  相似文献   

7.
Thind , K. S., and G. S. Rawla . (Panjab U., Chandigarh, India.) A new fungus on Cyperus iria. Amer. Jour. Bot. 48(10): 859–862. Illus. 1961—Duosporium cyperi gen. et sp. nov., collected on Cyperus iria Linn., is characterized by the production of 2 types of conidia, macroconidia and microconidia. The macroconidia are straight, cylindrical, triseptate, with their middle cells dark brown, long, thick-walled, and end cells much paler or subhyaline, short, thin-walled, the middle and the end cells being equally wide. The microconidia are nonseptate, spherical, dark brown, prominently verrucose, and produced apically on the conidiophores like the macroconidia; occasionally also formed as secondary conidia on germinating 3-septate macroconidia in culture. The formation of 2 distinct types of conidia in Duosporium and the features of its macroconidia characteristically differentiate it from its close relative Helminthosporium, although Duosporium. cyperi is not unique among fungi in producing 2 types of conidia.  相似文献   

8.
The effects of temperature (4–20°C), relative humidity (RH, 0–100%), pH (3–7), availability of nutrients (0–5 g/l sucrose) and artificial light (0–494 μmol/m2/s) on macroconidial germination of Fusarium graminearum were studied. Germ tubes emerged between 2 and 6 h after inoculation at 100% RH and 20°C. Incubation in light (205 ± 14 μmol/m/s) retarded the germination for approximately 0.5 h in comparison with incubation in darkness. The times required for 50% of the macroconidia to germinate were 3.5 h at 20°C, 5.4 h at 14°C and 26.3 h at 4°C. No germination was observed after an incubation period of 18 h at 20°C in darkness at RH less than 80%. At RH greater than 80%, germination increased with humidity. Germination was observed when macroconidia were incubated in glucose (5 g/l) or sucrose (concentration range from 2.5 × 10?4 to 5 g/l) whereas no germination was observed when macroconidia were incubated in sterile deionized water up to 22 h. Macroconidia germinated quantitatively within 18 h at pH 3–7. Repeated freezing (?15°C) and thawing (20°C) water agar plates with either germinated or non‐germinated macroconidia for up to five times did not prevent fungal growth after thawing. However, the fungal growth rate of mycelium was negatively related to the number of freezing events the non‐germinated macroconidia experienced. The fungal growth rate of mycelium was not significantly affected by the number of freezing events the germinated spores experienced. Incubation of macroconidia at low humidity (0–53% RH) suppressed germination and decreased the viability of the spores.  相似文献   

9.
Colletotrichum acutatum may develop one or more secondary conidia after conidial germination and before mycelial growth. Secondary conidia formation and germination were influenced by conidia concentration. Concentrations greater than 1x105 conidia/mL were associated with germination decrease and with secondary conidia emergence. Secondary conidia can form either alone or simultaneously with germ tubes and appressoria. Confocal analysis showed numerous lipid bodies stored inside ungerminated conidia, which diminished during germ tube and appressoria formation, with or without secondary conidia formation. They were also reduced during secondary conidia formation alone. While there was a decrease inside germinated conidia, lipid bodies appeared inside secondary conidia since the initial stages. Intense vacuolization inside primary germinated conidia occurred at the same time as the decrease in lipid bodies, which were internalized and digested by vacuoles. During these events, small acidic vesicles inside secondary conidia were formed. Considering that the conidia were maintained in distilled water, with no exogenous nutrients, it is clear that ungerminated conidia contain enough stored lipids to form germ tubes, appressoria, and the additional secondary conidia replete with lipid reserves. These results suggested a very complex and well-balanced regulation that makes possible the catabolic and anabolic pathways of these lipid bodies.  相似文献   

10.
Summary This study compares the development of shoot apical meristems of white spruce somatic and zygotic embryos during germination. In mature somatic embryos, the functional part of the shoot apical meristem was bi-layered. After partial drying, a normal shoot meristem was formed from these two cell layers during germination. Other cells within the meristem were vacuolated and separated by intercellular air spaces. In the absence of the partial drying treatment, somatic embryos enlarged in size primarily due to vacuolation of cells and the formation of large intercellular air spaces. A majority of these somatic embryos failed to form a functional shoot apical meristem. Compared with somatic embryos, the shoot apical meristem of a mature zygotic embryo was well organized with a densely cytoplasmic apical layer. The cells within the meristem were tightly packed. Judging from the cell profiles during germination, all cells within the meristem of the zygotic embryo took part in the formation of the vegetative shoot apical meristem.  相似文献   

11.
Beauveria bassiana colonizes insect hosts initially through a yeast phase, which is common in some artificial liquid cultures, but not reported on artificial solid media. We describe a yeast-like phase for B. bassiana isolate 447 (ATCC 20872) on MacConkey agar and its virulence toward Diatraea saccharalis and Tetranychus urticae. The yeast-like cells of B. bassiana developed by budding from germinating conidia after 24-h incubation. Cells were typically 5-10 microm and fungal colonies were initially circular and mucoid, but later were covered with mycelia and conidia. Ability to produce yeast-like cells on MacConkey medium was relatively common among different B. bassiana isolates, but growth rate and timing of yeast-like cell production also varied. Metarhizium anisopliae and Paecilomyces spp. isolates did not grow as yeast-like cells on MacConkey medium. Yeast-like cells of B. bassiana 447 were more virulent against D. saccharalis than conidia when 10(7)cells/ml were used. At 10(8)cells/ml, the estimated mean survival time was 5.4 days for the yeast suspension and 7.7 days for the conidial suspension, perhaps due to faster germination. The LC(50) was also lower for yeast than conidial suspensions. Yeast-like cells and conidia had similar virulence against T. urticae; the average mortalities with yeast-like cells and conidia were, respectively, 42.8 and 45.0%, with 10(7)cells/ml, and 77.8 and 74.4%, with 10(8)cells/ml. The estimated mean survival times were 3.6 and 3.9 for yeast and conidial suspensions, respectively. The bioassay results demonstrate the yeast-like structures produced on MacConkey agar are effective as inoculum for B. bassiana applications against arthropod pests, and possibly superior to conidia against some species. Obtaining well-defined yeast phase cultures of entomopathogenic hyphomycetes may be an important step in studies of the biology and nutrition, pathogenesis, and the genetic manipulation of these fungi.  相似文献   

12.
Ergothioneine (EGT) is a histidine derivative with sulfur on the imidazole ring and a trimethylated amine; it is postulated to have an antioxidant function. Although EGT apparently is only produced by fungi and some prokaryotes, it is acquired by animals and plants from the environment, and is concentrated in animal tissues in cells with an EGT transporter. Monobromobimane derivatives of EGT allowed conclusive identification of EGT by LC/MS and the quantification of EGT in Colletotrichum graminicola and Neurospora crassa conidia and mycelia. EGT concentrations were significantly (α=0.05) higher in conidia than in mycelia, with approximately 17X and 5X more in C. graminicola and N. crassa, respectively. The first EGT biosynthetic gene in a fungus was identified by quantifying EGT in N. crassa wild type and knockouts in putative homologs of actinomycete EGT biosynthetic genes. NcΔEgt-1, a strain with a knockout in gene NCU04343, does not produce EGT, in contrast to the wild type. To determine the effects of EGT in vivo, we compared NcΔEgt-1 to the wild type. NcΔEgt-1 is not pleiotropically affected in rate of hyphal elongation in Vogel's medium either with or without ammonium nitrate and in the rate of germination of macroconidia on Vogel's medium. The superoxide-producer menadione had indistinguishable effects on conidial germination between the two strains. Cupric sulfate also had indistinguishable effects on conidial germination and on hyphal growth between the two strains. In contrast, germination of NcΔEgt-1 conidia was significantly more sensitive to tert-butyl hydroperoxide than the wild type; germination of 50% (GI(50)) of the NcΔEgt-1 conidia was prevented at 2.7 mM tert-butyl hydroperoxide whereas the GI(50) for the wild type was 4.7 mM tert-butyl hydroperoxide, or at a 1.7X greater concentration. In the presence of tert-butyl hydroperoxide and the fluorescent reactive oxygen species indicator 5-(and-6)-carboxy-2',7'-dichlorodihydrofluorescein diacetate, significantly (P=0.0002) more NcΔEgt-1 conidia fluoresced than wild type conidia, indicating that EGT quenched peroxides in vivo. While five to 21-day-old conidia of both strains germinated 100%, NcΔEgt-1 conidia had significantly (P<0.001) diminished longevity. Linear regression analysis indicates that germination of the wild type declined to 50% in 35 days, in comparison to 25 days for the NcΔEgt-1, which is equivalent to a 29% reduction in conidial life span in the NcEgt-1 deletion strain. Consequently, the data indicate that endogenous EGT helps protect conidia during the quiescent period between conidiogenesis and germination, and that EGT helps protect conidia during the germination process from the toxicity of peroxide but not from superoxide or Cu(2+). Based on an in silico analysis, we postulate that NcEgt-1 was acquired early in the mycota lineage as a fusion of two adjacent prokaryotic genes, that was then lost in the Saccharomycotina, and that NcEgt-1 catalyzes the first two steps of EGT biosynthesis from histidine to hercynine to hercynylcysteine sulfoxide.  相似文献   

13.
Helminthosporium euphorbiae is a pathogen of the weedEuphorbia heterophylla, which causes severe losses in soybean (Glycine max) crops. The fungus causes leaf loss and affects germination, making it a promising biocontrol agent for this weed. In order to start a breeding program for this species, four isolates were examined for number of nuclei in the conidia and hyphae and nuclear behavior at different cultivation stages. The conidia were multinucleated with about 20 nuclei per conidium, and 5 to 7 nuclei were observed in the hyphae compartments. The high number of nuclei makes the genetic manipulation of this species diffucult, so the protoplast formation is an alternative for obtaining cells with a reduced number of nuclei. Thus the experimental conditions for the production and regeneration of protoplasts inH. euphorbiae were determined by assessing three enzymatic complexes and seven osmotic stabilizers. The efficiency of formation and regeneration frequency of the protoplasts varied depending on isolates, stabilizers and enzyme mixture used. The number of nuclei estimated per protoplast was reduced to 1 to 6, depending on the stage of mycelial growth during the protoplast formation process.  相似文献   

14.
We investigated acid-base permeability properties of electrically resistive monolayers of alveolar epithelial cells (AEC) grown in primary culture. AEC monolayers were grown on tissue culture-treated polycarbonate filters. Filters were mounted in a partitioned cuvette containing two fluid compartments (apical and basolateral) separated by the adherent monolayer, cells were loaded with the pH-sensitive dye 2',7'-bis(2-carboxyethyl)-5(6)-carboxyfluorescein, and intracellular pH was determined. Monolayers in HCO-free Na(+) buffer (140 mM Na(+), 6 mM HEPES, pH 7.4) maintained a transepithelial pH gradient between the two fluid compartments over 30 min. Replacement of apical fluid by acidic (6.4) or basic (8.0) buffer resulted in minimal changes in intracellular pH. Replacement of basolateral fluid by acidic or basic buffer resulted in transmembrane proton fluxes and intracellular acidification or alkalinization. Intracellular alkalinization was blocked > or =80% by 100 microM dimethylamiloride, an inhibitor of Na(+)/H(+) exchange, whereas acidification was not affected by a series of acid/base transport inhibitors. Additional experiments in which AEC monolayers were grown in the presence of acidic (6.4) or basic (8.0) medium revealed differential effects on bioelectric properties depending on whether extracellular pH was altered in apical or basolateral fluid compartments bathing the cells. Acid exposure reduced (and base exposure increased) short-circuit current from the basolateral side; apical exposure did not affect short-circuit current in either case. We conclude that AEC monolayers are relatively impermeable to transepithelial acid/base fluxes, primarily because of impermeability of intercellular junctions and of the apical, rather than basolateral, cell membrane. The principal basolateral acid exit pathway observed under these experimental conditions is Na(+)/H(+) exchange, whereas proton uptake into cells occurs across the basolateral cell membrane by a different, undetermined mechanism. These results are consistent with the ability of the alveolar epithelium to maintain an apical-to-basolateral (air space-to-blood) pH gradient in situ.  相似文献   

15.
The levels of glucosamine and galactosamine were determined in conidia, germinating conidia, and vegetative mycelia of Neurospora crassa. In the vegetative mycelia about 90% of the amino sugars were shown to be components of the cell wall. The remaining 10% of the amino sugars were tentatively identified as the nucleotide sugars uridine diphospho-2-acetamido-2-deoxy-D-glucose and uridine diphospho-2-acetamido-2-deoxy-D-galactose. Conidia and vegetative mycelia contained about the same levels of glucosamine. During the first 9 h after the initiation of germination, the total glucosamine content had increased 3.1-fold, whereas the residual dry weight of the culture had increased 7.7-fold. This led to a drop in the glucosamine concentration from 100 mumol/g of residual dry weight to 42 mumol/g. During this time, all of the conidia had germinated and the surface area of the new germ tubes had increased to 10 times that of the conidia. Either germ tubes were initially produced without glucosamine-containing polymers, or these polymers (probably chitin) were deposited only at low densities in the germ tube cell walls. The chitin precursor uridine diphospho-2-acetamido-2-deoxy-D-glucose was present at all times during conidial germination. Conida contained very low levels of galactosamine. During germination, galactosamine could not be detected until the culture had reached a cell density of about 0.6 mg of residual dry weight per ml of growth medium. This was observed regardless of the time required to reach this cell density or the fold increase in dry weight. The accumulation of galactosamine-containing polymers does not appear to be necessary for germ tube formation. The levels of soluble galactosamine (uridine diphospho-2-actamido-2-deoxy-D-galatose) were very low in conidia and increased during germination at the same time that galactosamine appeared in the cellular polymers. In addition, under certain culture conditions, the appearance of galactosamine and the increase in the glucosamine concentration occurred simultaneously.  相似文献   

16.
The kinetics of the development of the mycelial form of Sporothrix schenckii from yeast cells and conidia in a minimal basal medium with glucose at pH 4.0 and 25 °C were established. Germ tube formation was used as the index of germination for both yeast cells and conidia. Yeast cells were first observed to develop germ tubes after 3 h of incubation, reaching 92±5%, after 12 h of incubation. Germ tubes were first detected in conidia after 9 h of incubation, and 12 h after inoculation 92±6% of the conidia had germ tubes. After 24 h of incubation, fully developed, sporulating mycelia were observed from both yeast cells and conidia. A delay in germ tube formation from yeast cells was observed when But2cAMP(10 mM) and But2cGMP (10 mM) were added to the medium. Also the addition of caffeine, a cyclic nucleotide phosphodiesterase inhibitor, inhibited the yeast to mycelial transition. Conidial germination into the mycelial form was also inhibited when cAMP, But2cAMP and caffeine were added to the medium. These results suggest the possible involvement of cyclic nucleotides in the control of dimorphism in S. schenckii.  相似文献   

17.
Penicillium paneum is an important contaminant of cereal grains which is able to grow at low temperature, low pH, high levels of carbon dioxide, and under acid conditions. P. paneum produces mycotoxins, which may be harmful to animals and humans. We found that conidia in dense suspensions showed poor germination, suggesting the presence of a self-inhibitor. A volatile compound(s) produced by these high-density conditions also inhibited mycelial growth of different species of fungi belonging to a variety of genera, suggesting a broad action range. The heat-stable compound was isolated by successive centrifugation of the supernatant obtained from spore suspensions with a density of 109 conidia ml−1. By using static headspace analyses, two major peaks were distinguished, with the highest production of these metabolites after 22 h of incubation at 25°C and shaking at 140 rpm. Gas chromatography coupled with mass spectra analysis revealed the compounds to be 3-octanone and 1-octen-3-ol. Notably, only the latter compound appeared to block the germination process at different developmental stages of the conidia (swelling and germ tube formation). In this study, 1-octen-3-ol influenced different developmental processes during the P. paneum life cycle, including induction of microcycle conidiation and inhibition of spore germination. Therefore, the compound can be considered a fungal hormone during fungal development.  相似文献   

18.
The adhesion and germination of conidia of nine fungal species were assessed on polycarbonate membranes or on the skin of apple fruit in sealed glass bottles injected or not injected with acetate esters. Adhesion was determined after dislodging conidia from surfaces using a sonication probe. Adhesion and germination of conidia of Aspergillus flavus, Aspergillus fumigatus, Aspergillus niger, Penicillium citrinum, Penicillium claviforme, or Trichoderma sp. on membranes after 48 h were not increased in a 1.84 microg mL(-1) headspace of butyl acetate (BA), ethyl acetate, hexyl acetate, 2-methylbutyl acetate, pentyl acetate, or propyl acetate. Adhesion and germination of Botrytis cinerea, Penicillium expansum, and Penicillium roquefortii conidia were stimulated by all esters. Only conidia of B. cinerea and P. expansum exhibited increased adhesion and germination on the skin of apple fruit in bottles exposed to 0.92 microg mL(-1) of BA. Only conidia of B. cinerea and P. expansum produced decay in inoculated puncture wounds on fruit. Freshly made puncture wounds or 24-h-old puncture wounds in fruit were more adhesive than the unpunctured skin of fruit to conidia of B. cinerea or P. expansum. Fresh wounds were more adhesive to both fungi than 24-h-old puncture wounds. The skin and wounds of fruit were as adhesive to B. cinerea conidia as they were to P. expansum conidia. A 4-h exposure to 1.43 microg mL(-1) of BA increased adhesion of B. cinerea and P. expansum conidia in 24-h-old wounds. Results suggest that acetate-ester stimulation most likely is not a rare phenomenon in the fungi. For nutrient-dependent decay pathogens of apple fruit, acetate esters may be an alternative chemical cue used to maintain adhesion of conidia to wound surfaces.  相似文献   

19.
Asexual spores (conidia) are the infectious propagules of many pathogenic fungi, and the capacity to sense the host environment and trigger conidial germination is a key pathogenicity determinant. Germination of conidia requires the de novo establishment of a polarised growth axis and consequent germ tube extension. The molecular mechanisms that control polarisation during germination are poorly understood. In the dimorphic human pathogenic fungus Penicillium marneffei, conidia germinate to produce one of two cell types that have very different fates in response to an environmental cue. At 25 degrees C, conidia germinate to produce the saprophytic cell type, septate, multinucleate hyphae that have the capacity to undergo asexual development. At 37 degrees C, conidia germinate to produce the pathogenic cell type, arthroconidiating hyphae that liberate uninucleate yeast cells. This study shows that the p21-activated kinase pakA is an essential component of the polarity establishment machinery during conidial germination and polarised growth of yeast cells at 37 degrees C but is not required for germination or polarised growth at 25 degrees C. Analysis shows that the heterotrimeric G protein alpha subunit GasC and the CDC42 orthologue CflA lie upstream of PakA for germination at both temperatures, while the Ras orthologue RasA only functions at 25 degrees C. These findings suggest that although some proteins that regulate the establishment of polarised growth in budding yeast are conserved in filamentous fungi, the circuitry and downstream effectors are differentially regulated to give rise to distinct cell types.  相似文献   

20.
This paper presents results on the effect of light, temperature and substrate during spore formation on the germinability of conidia in Colletotrichum falcatum. Light seems to have no effect on the germination of conidia unless the cultures were exposed to a high intensity of light during sporulation, in which case the spores showed a reduced germination and an increased appressoria formation. Conidia produced at temperatures higher than the optimum showed better germination and less appressoria formation than the spores produced at the temperature optimum for the growth and sporulation of the fungus. A similar increase in germination was also observed in conidia obtained from inoculated sugarcane leaves as compared to those produced on culture media. The light type virulent isolates of C. falcatum showed greater sensitivity to all these treatments than the dark type weakly pathogenic 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