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 203 毫秒
1.
At extremely low values of moisture pressure (?96.4 MPa; aw 0.50), the spores of xerotolerant streptomycetes (Streptomyces odorifer and S. rubiginosohelvolus) germinated, their germ lengths increased, and lateral branching of the mycelium was observed after 5 days of incubation in a thin layer of agarized nutrient medium. At ?22.6 MPa (aw 0.86), the mycelium begins to branch after a 2-day incubation; over a 5-day incubation at ?2.8 MPa (aw 0.98), it goes through a reproduction cycle, which culminates in spore formation. The mathematical model approach enabled us to elucidate the behavioral patterns of Streptomyces spores in a thin layer of agarized nutrient medium at low humidity levels. The dynamics of spore germination is governed by the exponential law, which allows calculation of the average duration of the period a before spore germination, as well as the time needed for 50% of viable spores to germinate.  相似文献   

2.
Certain fatty acids in the C5 to C18 range, at concentrations as low as 10−5 M, were found to inhibit the germinations of spores of the sensitive fern, Onoclea sensibilis L. The addition of gametophytic culture flltrates of the bracken fern, Pteridium aquilinum (L.) Kuhn, containing antheridogen A, was found to overcome this inhibition and allow the spores to germinate and the gametophytes to develop in a normal fashion. Some fatty acids were found to increase the antheridium-inducing potency of antheridogen A as much as 10-fold. An effect similar to this may promote the diecious reproduction of ferns.  相似文献   

3.
Summary Acids like hydrogen fluoride, hydrazoic and fluoroacetic have been shown to prevent the germination of ascospores of N. tetrasperma when dormant spores are treated. On the other hand, propionate, cysteine and others are ineffective when used in this way. When activated ascospores were treated, much lower concentrations of the acids were sufficient to poison the spores. As in other systems, these substances are most effective at a pH below their pKa.The kinetics of uptake of fluoride by dormant ascospores were studied and shown to be very different from those reported for cations. However, P32 was not absorbed by dormant ascospores, even at pH 1.5.Respiratory inhibition by azide and fluoroacetate occurred immediately after the spores were activated, but in the case of 5-nitro-2-furfuryl methyl ether no effect was observed until just before germination occurred.These results suggest that a permeability barrier exists in the dormant ascospore which disappears upon germination. Moreover, the dormant spore seems to be permeable to acids of small size but impermeable to those possessing more than 3 methylene groups or of equivalent size.This work was made possible by a grant from the Michigan-Memorial Phoenix Project of the University of Michigan to whom the authors would like to express their gratitude.  相似文献   

4.
团扇蕨孢子发生和发育的显微观察   总被引:1,自引:0,他引:1  
利用光学显微镜对膜蕨科(Hymenophyllaceae)团扇蕨(Gonocormus minutus(Blume) Bosch)孢子的发生和发育进行了观察。研究结果表明:团扇蕨孢子为多边圆形,三裂缝不明显,外壁表面光滑,周壁薄,紧贴外壁表面,由周壁形成乳头状或颗粒状纹饰。在外壁形成后期,孢子表面和囊腔中出现大量小球;在周壁形成时期,孢子表面和周围出现较多小球体;小球和小球体参与孢子壁的形成。团扇蕨绒毡层为混合型,内层为周原质团绒毡层;外层为腺质型绒毡层。本文为膜蕨科系统演化和发育生物学研究提供依据。  相似文献   

5.
Botrytis cinerea and Puccinia recondita spores adhere strongly to polystyrene microtiter plates coincident with germination. We developed assays for inhibition of spore adhesion in 96-well microtiter plates by using sulforhodamine B staining to quantify the adherent spores. In both organisms, fungicides that inhibited germination strongly inhibited spore adhesion, with 50% effective concentrations (EC50s) comparable to those for inhibition of germination. In contrast, fungicides that acted after germination in B. cinerea inhibited spore adhesion to microtiter plates only at concentrations much higher than their EC50s for inhibition of mycelial growth. Similarly, in P. recondita the ergosterol biosynthesis inhibitors myclobutanil and fenbuconazole acted after germination and did not inhibit spore adhesion. The assays provide a rapid, high-throughput alternative to traditional spore germination assays and may be applicable to other fungi.  相似文献   

6.
Protease activities were detected in quiescent and germinating spores of the ostrich fern (Matteuccia struthiopteris [L.] Todaro). Peak endopeptidase, aminopeptidase, and carboxypeptidase activities were detected 12 to 24 hours after spores began imbibing under light. There was a correlation between activities of proteases, the onset of a decline in levels of soluble protein, and an increase in levels of free amino acids. The earliest visible event of spore germination, breakage of the spore coat and protrusion of a rhizoid cell, was observed after peak protease activity, 48 to 72 hours after the start of imbibition. Results of this study demonstrate similarities in the pattern of protease activities during germination of ostrich fern spores to those of some seeds.  相似文献   

7.
Freshwater and marine sediments often harbor reservoirs of plant diaspores, from which germination and establishment may occur whenever the sediment falls dry. Therewith, they form valuable records of historical inter- and intraspecific diversity, and are increasingly exploited to facilitate diversity establishment in new or restored nature areas. Yet, while ferns may constitute a considerable part of a vegetation’s diversity and sediments are known to contain fern spores, little is known about their longevity, which may suffer from inundation and - in sea bottoms - salt stress. We tested the potential of ferns to establish from a sea or lake bottom, using experimental studies on spore survival and gametophyte formation, as well as a spore bank analysis on sediments from a former Dutch inland sea. Our experimental results revealed clear differences among species. For Asplenium scolopendrium and Gymnocarpium dryopteris, spore germination was not affected by inundated storage alone, but decreased with rising salt concentrations. In contrast, for Asplenium trichomanes subsp. quadrivalens germination decreased following inundation, but not in response to salt. Germination rates decreased with time of storage in saline water. Smaller and less viable gametophytes were produced when saline storage lasted for a year. Effects on germination and gametophyte development clearly differed among genotypes of A. scolopendrium. Spore bank analyses detected no viable spores in marine sediment layers. Only two very small gametophytes (identified as Thelypteris palustris via DNA barcoding) emerged from freshwater sediments. Both died before maturation. We conclude that marine, and likely even freshwater sediments, will generally be of little value for long-term storage of fern diversity. The development of any fern vegetation on a former sea floor will depend heavily on the deposition of spores onto the drained land by natural or artificial means of dispersal.  相似文献   

8.
9.
Few vertebrates are known to consume ferns regularly. Several species of mammals consume leaves to some extent but the consumption of fern spores is much rare. In Galicia (Northwest Spain) we studied the seasonal variation in the consumption of Culcita macrocarpa fertile leaves (i.e. with spores) in two populations (Capelada and Eume), assessed whether consumption rate increased with fern population size, and evaluated whether the consumer was a spore predator or a spore disperser. Consumption began in December and finished by mid February, and occurred before spore release, which happened later in Capelada than in Eume, probably influenced by differences in altitude. The consumer was identified as Apodemus sylvaticus by DNA analysis of its droppings and by capture of live animals. Throughout Galicia there was a significant increase in fern consumption rate as the population size of C. macrocarpa increased. Germination tests from droppings were carried out in 14 dishes but only in two dishes 1% and 0.3% of the spores germinated. Our results suggest that woodmouse can disperse spores of C. macrocarpa, although most of the spores were digested.  相似文献   

10.
High-precision measurements of size changes of individual bacterial spores based on ellipse fitting to bright-field images recorded with a digital camera were employed to monitor the germination of Bacillus spores with a precision of ∼5 nm. To characterize the germination of individual spores, we recorded bright-field and phase-contrast images and found that the timing of changes in their normalized intensities coincided, so the bright-field images can be used to characterize spore size and refractility changes during germination. The major conclusions from this work were as follows. (i) The sizes of germinating B. cereus spores were nearly unchanged until Trelease, the time of the completion of CaDPA (a 1:1 chelate of Ca2+ and dipicolinic acid [DPA]) release after addition of nutrient germinants. (ii) The minor axis of germinating B. cereus spores rapidly increased by ∼50 nm in a few seconds right after Trelease, while the major axis was slightly decreased or unchanged. Both the minor and major axes remained unchanged for a further 30 to 45 s and then increased by 100 to 200 nm by Tlys, the time of completion of cortex lysis. (iii) Individual spores in a population showed significant heterogeneity in the timing of germination events, such as Trelease and Tlys, but also variation in size changes during germination. (iv) Bacillus subtilis wild-type spores, B. subtilis spores lacking the cortex-lytic enzyme CwlJ, and wild-type Bacillus megaterium spores showed similar kinetics of size changes during nutrient germination. The size increases in germinating spores probably result from uptake of water and cortex lysis after completion of CaDPA release.  相似文献   

11.
Antheridogen activity in the fern Ceratopteris thalictroides (L.) Brongn.   总被引:1,自引:0,他引:1  
Spores of the homosporous fern Ceratopteris thalictroides , in multispore culture, initially produce spatulate gametophytes bearing only antheridia (males) and cordate gametophytes bearing both antheridia and archegonia (hermaphrodites). When multispore cultures are sampled, the ratio of male to hermaphroditic gametophytes is a constant for each population examined. Four possible causes of such a sex ratio (cytoplasmic inheritance, nuclear inheritance, incipient heterospory and an antheridogen) are investigated. Evidence presented indicates that an antheridogen causes the existence of two gametophyte types, while one or more cytoplasmic units are the probable cause of the sex ratio. The activity of the antheridogen is to cause potentially hermaphroditic plants to become male. This activity was elucidated in monospore culture. Populational differences in antheridogen activity are also demonstrated. The significance of antheridogens is discussed in relation to the mating system of these plants.  相似文献   

12.
Berberine, an alkaloid originally extracted from the plant Coptis chinensis and other herb plants, has been used as a pharmacological substance for many years. The therapeutic effect of berberine has been attributed to its interaction with nucleic acids and blocking cell division. However, levels of berberine entering individual microbial cells minimal for growth inhibition and its effects on bacterial spores have not been determined. In this work the kinetics and levels of berberine accumulation by individual dormant and germinated spores were measured by laser tweezers Raman spectroscopy and differential interference and fluorescence microscopy, and effects of berberine on spore germination and outgrowth and spore and growing cell viability were determined. The major conclusions from this work are that: (1) colony formation from B. subtilis spores was blocked ~ 99% by 25 μg/mL berberine plus 20 μg/mL INF55 (a multidrug resistance pump inhibitor); (2) 200 μg/mL berberine had no effect on B. subtilis spore germination with L-valine, but spore outgrowth was completely blocked; (3) berberine levels accumulated in single spores germinating with ≥ 25 μg/mL berberine were > 10 mg/mL; (4) fluorescence microscopy showed that germinated spores accumulated high-levels of berberine primarily in the spore core, while dormant spores accumulated very low berberine levels primarily in spore coats; and (5) during germination, uptake of berberine began at the time of commitment (T1) and reached a maximum after the completion of CaDPA release (Trelease) and spore cortex lysis (Tlysis).  相似文献   

13.
Germination and outgrowth are critical steps for returning Bacillus subtilis spores to life. However, oxidative stress due to full hydration of the spore core during germination and activation of metabolism in spore outgrowth may generate oxidative DNA damage that in many species is processed by apurinic/apyrimidinic (AP) endonucleases. B. subtilis spores possess two AP endonucleases, Nfo and ExoA; the outgrowth of spores lacking both of these enzymes was slowed, and the spores had an elevated mutation frequency, suggesting that these enzymes repair DNA lesions induced by oxidative stress during spore germination and outgrowth. Addition of H2O2 also slowed the outgrowth of nfo exoA spores and increased the mutation frequency, and nfo and exoA mutations slowed the outgrowth of spores deficient in either RecA, nucleotide excision repair (NER), or the DNA-protective α/β-type small acid-soluble spore proteins (SASP). These results suggest that α/β-type SASP protect DNA of germinating spores against damage that can be repaired by Nfo and ExoA, which is generated either spontaneously or promoted by addition of H2O2. The contribution of RecA and Nfo/ExoA was similar to but greater than that of NER in repair of DNA damage generated during spore germination and outgrowth. However, nfo and exoA mutations increased the spontaneous mutation frequencies of outgrown spores lacking uvrA or recA to about the same extent, suggesting that DNA lesions generated during spore germination and outgrowth are processed by Nfo/ExoA in combination with NER and/or RecA. These results suggest that Nfo/ExoA, RecA, the NER system, and α/β-type SASP all contribute to the repair of and/or protection against oxidative damage of DNA in germinating and outgrowing spores.  相似文献   

14.
Several alcohols and other solvents inhibit germination of spores of the fern, Onoclea sensibilis L. The inhibition is reversible when spores are transferred to solvent-free media. The effectiveness of different solvents, measured by the concentration needed to inhibit germination by 50%, increases with their lipid solubility. The activity of straight-chain alcohols from methanol through n-heptanol is highly correlated with lipid solubility, whereas the correlation is weaker for several other solvents. The results indicate that some lipophilic site in the spore is important in germination.  相似文献   

15.
Pattern of 3H-uridine incorporation into RNA of spores of Onocleasensibilis imbibed in complete darkness (non-germinating conditions)and induced to germinate in red light was followed by oligo-dTcellulose chromatography, gel electrophoresis coupled with fluorographyand autoradiography. In dark-imbibed spores, RNA synthesis wasinitiated about 24 h after sowing, with most of the label accumulatingin the high mol. wt. poly(A)RNA fraction. There was noincorporation of the label into poly(A) + RNA until 48 h aftersowing. In contrast, photo-induced spores began to synthesizeall fractions of RNA within 12 h after sowing and by 24 h, incorporationof 3H-uridine into RNA of irradiated spores was nearly 70-foldhigher than that into dark-imbibed spores. Protein synthesis,as monitored by 3H-arginine incorporation into the acid-insolublefraction and by autoradiography, was initiated in spores within1–2 h after sowing under both conditions. Autoradiographicexperiments also showed that the onset of protein synthesisin the cytoplasm of the germinating spore is independent ofthe transport of newly synthesized nuclear RNA. One-dimensionalsodium dodecyl sulphate-polyacrylamide gel electrophoresis of35S-methionine-labelled proteins revealed a good correspondencebetween proteins synthesized in a cell-free translation systemdirected by poly(A) +RNA of dormant spores and those synthesizedin vivo by dark-imbibed and photo-induced spores. These resultsindicate that stored mRNAs of O. sensibilis spores are functionallycompetent and provide templates for the synthesis of proteinsduring dark-imbibition and germination. Key words: Onoclea sensibilis, fern spore germination, gene expression, protein synthesis, sensitive fern, stored mRNA  相似文献   

16.
When exposed to nutrient or nonnutrient germinants, individual Bacillus spores can return to life through germination followed by outgrowth. Laser tweezers, Raman spectroscopy, and either differential interference contrast or phase-contrast microscopy were used to analyze the slow dipicolinic acid (DPA) leakage (normally ∼20% of spore DPA) from individual spores that takes place prior to the lag time, Tlag, when spores begin rapid release of remaining DPA. Major conclusions from this work with Bacillus subtilis spores were as follows: (i) slow DPA leakage from wild-type spores germinating with nutrients did not begin immediately after nutrient exposure but only at a later heterogeneous time T1; (ii) the period of slow DPA leakage (ΔTleakage = TlagT1) was heterogeneous among individual spores, although the amount of DPA released in this period was relatively constant; (iii) increases in germination temperature significantly decreased T1 times but increased values of ΔTleakage; (iv) upon germination with l-valine for 10 min followed by addition of d-alanine to block further germination, all germinated spores had T1 times of less than 10 min, suggesting that T1 is the time when spores become committed to germinate; (v) elevated levels of SpoVA proteins involved in DPA movement in spore germination decreased T1 and Tlag times but not the amount of DPA released in ΔTleakage; (vi) lack of the cortex-lytic enzyme CwlJ increased DPA leakage during germination due to longer ΔTleakage times in which more DPA was released; and (vii) there was slow DPA leakage early in germination of B. subtilis spores by the nonnutrients CaDPA and dodecylamine and in nutrient germination of Bacillus cereus and Bacillus megaterium spores. Overall, these findings have identified and characterized a new early event in Bacillus spore germination.  相似文献   

17.
Multispore cultures of Ceratopteris thalictroides (L.) Brongn. gametophytes contain plants of two distinct morphologies, hermaphrodite and male (Fig. 3). The male gametophyte has previously been shown to be induced by an antheridogen system. In the absence of the antheridogen all gametophytes become hermaphroditic. It is shown that the hermaphroditic gametophytes are first to develop in multispore cultures and are also the only source of antheridogen. Thus, since males are only produced in the presence of antheridogen, an explanation of the occurrence of two morphologies is evident. That is, the first gametophytes to develop do so in the absence of antheridogen, become hermaphrodites and produce antheridogen to which later developing gametophytes respond by becoming males. This does not explain why there is a range in development times for the gametophytes. Three possible controlling factors of development are investigated in multispore cultures: spore size, time of germination, and gametophyte growth. Data collected on these three factors are statistically analyzed. Analyses were done to determine the relative importance (singly or in combination) of these factors in predicting the developmental potential of a gametophyte. It was found that spore size most accurately indicated (ca 75% of the time) future gametophyte development.  相似文献   

18.
Short exposures of Bacillus spores to nutrient germinants can commit spores to germinate when germinants are removed or their binding to the spores'' nutrient germinant receptors (GRs) is inhibited. Bacillus subtilis spores were exposed to germinants for various periods, followed by germinant removal to prevent further commitment. Release of spore dipicolinic acid (DPA) was then measured by differential interference contrast microscopy to monitor germination of multiple individual spores, and spores did not release DPA after 1 to 2 min of germinant exposure until ∼7 min after germinant removal. With longer germinant exposures, percentages of committed spores with times for completion of DPA release (Trelease) greater than the time of germinant removal (Tb) increased, while the time TlagTb, where Tlag represents the time when rapid DPA release began, was decreased but rapid DPA release times (ΔTrelease = TreleaseTlag) were increased; Factors affecting average Trelease values and the percentages of committed spores were germinant exposure time, germinant concentration, sporulation conditions, and spore heat activation, as previously shown for commitment of spore populations. Surprisingly, germination of spores given a 2nd short germinant exposure 30 to 45 min after a 1st exposure of the same duration was significantly higher than after the 1st exposure, but the number of spores that germinated in the 2nd germinant exposure decreased as the interval between germinant exposures increased up to 12 h. The latter results indicate that spores have some memory, albeit transient, of their previous exposure to nutrient germinants.  相似文献   

19.
High-level heat resistance of spores of Bacillus thermoamylovorans poses challenges to the food industry, as industrial sterilization processes may not inactivate such spores, resulting in food spoilage upon germination and outgrowth. In this study, the germination and heat resistance properties of spores of four food-spoiling isolates were determined. Flow cytometry counts of spores were much higher than their counts on rich medium (maximum, 5%). Microscopic analysis revealed inefficient nutrient-induced germination of spores of all four isolates despite the presence of most known germination-related genes, including two operons encoding nutrient germinant receptors (GRs), in their genomes. In contrast, exposure to nonnutrient germinant calcium-dipicolinic acid (Ca-DPA) resulted in efficient (50 to 98%) spore germination. All four strains harbored cwlJ and gerQ genes, which are known to be essential for Ca-DPA-induced germination in Bacillus subtilis. When determining spore survival upon heating, low viable counts can be due to spore inactivation and an inability to germinate. To dissect these two phenomena, the recoveries of spores upon heat treatment were determined on plates with and without preexposure to Ca-DPA. The high-level heat resistance of spores as observed in this study (D120°C, 1.9 ± 0.2 and 1.3 ± 0.1 min; z value, 12.2 ± 1.8°C) is in line with survival of sterilization processes in the food industry. The recovery of B. thermoamylovorans spores can be improved via nonnutrient germination, thereby avoiding gross underestimation of their levels in food ingredients.  相似文献   

20.
Spores of Anemia mexicana Klotzsch and Anemia phyllitidis (L.) Swartz were tested comparatively to investigate the effects of various treatments on spore germination and early gametophyte development in light and darkness. The optimum pH for induction of spore germination is approximately 6. Both species have a minimum 8 hour light insensitive preinduction phase for spore germination. An additional 8 to 12 hours of light are needed to induce 50% germination in A. phyllitidis while at least 24 hours of light are needed for A. mexicana spores. A. phyllitidis has greater sensitivity to the four gibberellic acids tested (GA3, GA4, GA7, and GA13) than A. mexicana for induction of spore germination in darkness. In both species the greatest response was observed with GA4 and GA7. GA13 was clearly the least effective. Gametophytes of each species are 100 times more sensitive to their own antheridiogen than to the antheridiogen of the other species. AMO-1618 (1 millimolar), fenarimol (1 mm), and ancymidol (0.1 mm) had essentially no effect on light-induced germination. The latter two did, however, inhibit gametophyte development.  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