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.
Bacillus sp. strain SG-1 is a marine bacterial species isolated from a near-shore manganese sediment sample. Its mature dormant spores promote the oxidation of Mn2+ to MnO2. By quantifying the amounts of immobilized and oxidized manganese, it was established that bound manganese was almost instantaneously oxidized. When the final oxidation of manganese by the spores was partly inhibited by NaN3 or anaerobiosis, an equivalent decrease in manganese immobilization was observed. After formation of a certain amount of MnO2 by the spores, the oxidation rate decreased. A maximal encrustment was observed after which no further oxidation occurred. The oxidizing activity could be recovered by reduction of the MnO2 with hydroxylamine. Once the spores were encrusted, they could bind significant amounts of manganese, even when no oxidation occurred. Purified spore coat preparations oxidized manganese at the same rate as intact spores. During the oxidation of manganese in spore coat preparations, molecular oxygen was consumed and protons were liberated. The data indicate that a spore coat component promoted the oxidation of Mn2+ in a biologically catalyzed process, after adsorption of the ion to incipiently formed MnO2. Eventually, when large amounts of MnO2 were allowed to accumulate, the active sites were masked and further oxidation was prevented.  相似文献   

2.
The geochemical cycling of cobalt (Co) has often been considered to be controlled by the scavenging and oxidation of Co(II) on the surface of manganese [Mn(III,IV)] oxides or manganates. Because Mn(II) oxidation in the environment is often catalyzed by bacteria, we have investigated the ability of Mn(II)-oxidizing bacteria to bind and oxidize Co(II) in the absence of Mn(II) to determine whether some Mn(II)-oxidizing bacteria also oxidize Co(II) independently of Mn oxidation. We used the marine Bacillus sp. strain SG-1, which produces mature spores that oxidize Mn(II), apparently due to a protein in their spore coats (R.A. Rosson and K. H. Nealson, J. Bacteriol. 151:1027-1034, 1982; J. P. M. de Vrind et al., Appl. Environ. Microbiol. 52:1096-1100, 1986). A method to measure Co(II) oxidation using radioactive 57Co as a tracer and treatments with nonradioactive (cold) Co(II) and ascorbate to discriminate bound Co from oxidized Co was developed. SG-1 spores were found to oxidize Co(II) over a wide range of pH, temperature, and Co(II) concentration. Leucoberbelin blue, a reagent that reacts with Mn(III,IV) oxides forming a blue color, was found to also react with Co(III) oxides and was used to verify the presence of oxidized Co in the absence of added Mn(II). Co(II) oxidation occurred optimally around pH 8 and between 55 and 65°C. SG-1 spores oxidized Co(II) at all Co(II) concentrations tested from the trace levels found in seawater to 100 mM. Co(II) oxidation was found to follow Michaelis-Menten kinetics. An Eadie-Hofstee plot of the data suggests that SG-1 spores have two oxidation systems, a high-affinity-low-rate system (Km, 3.3 × 10-8 M; Vmax, 1.7 × 10-15 M · spore-1 · h-1) and a low-affinity-high-rate system (Km, 5.2 × 10-6 M; Vmax, 8.9 × 10-15 M · spore-1 · h-1). SG-1 spores did not oxidize Co(II) in the absence of oxygen, also indicating that oxidation was not due to abiological Co(II) oxidation on the surface of preformed Mn(III,IV) oxides. These results suggest that some microorganisms may directly oxidize Co(II) and such biological activities may exert some control on the behavior of Co in nature. SG-1 spores may also have useful applications in metal removal, recovery, and immobilization processes.  相似文献   

3.
Dormant spores of the marine Bacillus sp. strain SG-1 catalyze the oxidation of manganese(II), thereby becoming encrusted with insoluble Mn(III,IV) oxides. In this study, it was found that the Mn(II)-oxidizing activity could be removed from SG-1 spores using a French press and recovered in the supernatant following centrifugation of the spores. Transmission electron microscopy of thin sections of SG-1 spores revealed that the ridged outermost layer was removed by passage through the French press, leaving the remainder of the spore intact. Comparative chemical analysis of this layer with the underlying spore coats suggested that this outer layer is chemically distinct from the spore coat. Taken together, these results indicate that this outer layer is an exosporium. Previous genetic analysis of strain SG-1 identified a cluster of genes involved in Mn(II) oxidation, the mnx genes. The product of the most downstream gene in this cluster, MnxG, appears to be a multicopper oxidase and is essential for Mn(II) oxidation. In this study, MnxG was overexpressed in Escherichia coli and used to generate polyclonal antibodies. Western blot analysis demonstrated that MnxG is localized to the exosporium of wild-type spores but is absent in the non-oxidizing spores of transposon mutants within the mnx gene cluster. To our knowledge, Mn(II) oxidation is the first oxidase activity, and MnxG one of the first gene products, ever shown to be associated with an exosporium.  相似文献   

4.
The marine Bacillus sp. strain SG-1 forms spores that oxidize manganese(II) as a result of the activities of uncharacterized components of its spore coat. Nucleotide sequence analysis of chromosomal loci previously identified through insertion mutagenesis as being involved in manganese oxidation identified seven possible genes (designated mnxA to mnxG) in what appears to be an operon. A potential recognition site for the sporulation, mother-cell-specific, RNA polymerase sigma factor, sigmaK, was located just upstream of the cluster, and correspondingly, measurement of beta-galactosidase activity from a Tn917-lacZ insertion in mnxD showed expression at mid-sporulation to late sporulation (approximately stage IV to V of sporulation). Spores of nonoxidizing mutants appeared unaffected with respect to their temperature and chemical resistance properties and germination characteristics. However, transmission electron microscopy revealed alterations in the outermost spore coat. This suggests that products of these genes may be involved in the deposition of the spore coat structure and/or are spore coat proteins themselves. Regions of the deduced protein product of mnxG showed amino acid sequence similarity to the family of multicopper oxidases, a diverse group of proteins that use multiple copper ions to oxidize a variety of substrates. Similar regions included those that are involved in binding of copper, and the addition of copper at a low concentration was found to enhance manganese oxidation by the spores. This suggests that the product of this gene may function like a copper oxidase and that it may be directly responsible for the oxidation of manganese by the spores.  相似文献   

5.
Mature spores of the marine Bacillus sp. strain SG-1 bind and oxidize manganese(II), thereby becoming encrusted with a manganese(IV) oxide. Both the function and mechanism of this oxidation are unknown, although evidence suggests that spore coat proteins are involved. To further study this phenomenon, methods of genetic analysis were developed for SG-1. By a modified protoplast transformation procedure, SG-1 was transformed (approximately 100 transformants per micrograms of DNA) with several different plasmids of gram-positive origin. Transposon Tn917, delivered on the temperature-sensitive plasmid pLTV1, was used to generate mutants of SG-1. Conditions were established that allowed 98% plasmid loss and insertions to be recovered at a frequency of 10(-3). Each mutant was found to be the result of a single insertion event. Restriction analysis of 27 mutants that do not oxidize manganese but still sporulate localized 17 of the insertions within two regions of the chromosome (termed Mnx regions), and a physical map of these regions was generated. Analysis of 18 transposon integrants in which manganese oxidation was unaffected revealed random transposon integration, with none of their insertions mapping within the Mnx regions. The Mnx regions were cloned from wild-type SG-1, and the largest region, carried on the lactococcal plasmid pGK13, was used to complement in trans one of the nonoxidizing mutants. These results demonstrate that the Mnx regions encode factors that are required for the oxidation of manganese, and this represents the first report identifying genes involved in bacterial manganese oxidation.  相似文献   

6.
Aims:  To determine roles of coats in staining Bacillus subtilis spores, and whether spores have membrane potential.
Methods and Results:  Staining by four dyes and autofluorescence of B. subtilis spores that lack some ( cotE , gerE ) or most ( cotE gerE) coat protein was measured. Wild-type, cotE and gerE spores autofluorescenced and bound dyes, but cotE gerE spores did not autofluorescence and were stained only by two dyes. A membrane potential-sensitive dye DiOC6(3) bound to dormant Bacillus megaterium and B. subtilis spores. While this binding was abolished by the protonophore FCCP, DiOC6(3) bound to heat-killed spores, but not to dormant B. subtilis cotE gerE spores. However, DiOC6(3) bound well to all germinated spores.
Conclusions:  The autofluorescence of dormant B. subtilis spores and the binding of some dyes are due to the coat. There is no membrane potential in dormant Bacillus spores, although membrane potential is generated when spores germinate.
Significance and Impact of the Study:  The elimination of the autofluorescence of B. subtilis spores may allow assessment of the location of low abundance spore proteins using fluorescent reporter technology. The dormant spore's lack of membrane potential may allow tests of spore viability by assessing membrane potential in germinating spores.  相似文献   

7.
Spores of marine Bacillus sp. strain SG-1 are capable of oxidizing Mn(II) and Co(II), which results in the precipitation of Mn(III, IV) and Co(III) oxides and hydroxides on the spore surface. The spores also bind other heavy metals; however, little is known about the mechanism and capacity of this metal binding. In this study the characteristics of the spore surface and Cu(II) adsorption to this surface were investigated. The specific surface area of wet SG-1 spores was 74.7 m2 per g of dry weight as measured by the methylene blue adsorption method. This surface area is 11-fold greater than the surface area of dried spores, as determined with an N2 adsorption surface area analyzer or as calculated from the spore dimensions, suggesting that the spore surface is porous. The surface exchange capacity as measured by the proton exchange method was found to be 30.6 μmol m−2, which is equal to a surface site density of 18.3 sites nm−2. The SG-1 spore surface charge characteristics were obtained from acid-base titration data. The surface charge density varied with pH, and the zero point of charge was pH 4.5. The titration curves suggest that the spore surface is dominated by negatively charged sites that are largely carboxylate groups but also phosphate groups. Copper adsorption by SG-1 spores was rapid and complete within minutes. The spores exhibited a high affinity for Cu(II). The amounts of copper adsorbed increased from negligible at pH 3 to maximum levels at pH >6. Their great surface area, site density, and affinity give SG-1 spores a high capability for binding metals on their surfaces, as demonstrated by our experiments with Cu(II).  相似文献   

8.
Bacterial spores are renowned for their longevity, ubiquity, and resistance to environmental insults, but virtually nothing is known regarding whether these metabolically dormant structures impact their surrounding chemical environments. In the present study, a number of spore-forming bacteria that produce dormant spores which enzymatically oxidize soluble Mn(II) to insoluble Mn(IV) oxides were isolated from coastal marine sediments. The highly charged and reactive surfaces of biogenic metal oxides dramatically influence the oxidation and sorption of both trace metals and organics in the environment. Prior to this study, the only known Mn(II)-oxidizing sporeformer was the marine Bacillus sp. strain SG-1, an extensively studied bacterium in which Mn(II) oxidation is believed to be catalyzed by a multicopper oxidase, MnxG. Phylogenetic analysis based on 16S rRNA and mnxG sequences obtained from 15 different Mn(II)-oxidizing sporeformers (including SG-1) revealed extensive diversity within the genus Bacillus, with organisms falling into several distinct clusters and lineages. In addition, active Mn(II)-oxidizing proteins of various sizes, as observed in sodium dodecyl sulfate-polyacrylamide electrophoresis gels, were recovered from the outer layers of purified dormant spores of the isolates. These are the first active Mn(II)-oxidizing enzymes identified in spores or gram-positive bacteria. Although extremely resistant to denaturation, the activities of these enzymes were inhibited by azide and o-phenanthroline, consistent with the involvement of multicopper oxidases. Overall, these studies suggest that the commonly held view that bacterial spores are merely inactive structures in the environment should be revised.  相似文献   

9.
In a seawater medium containing 0.005% yeast extract and 0.005% peptone, a marine bacillus, SG-1, sporulated only when associated with solid surfaces. The spores (rather than the vegetative cells) were responsible for the oxidation of manganese, and the degree of sporulation was determined by the surface area available rather than by the chemical nature of the clay or silica surface used.  相似文献   

10.
Methionine residues in α/β-type small, acid-soluble spore proteins (SASP) of Bacillus species were readily oxidized to methionine sulfoxide in vitro by t-butyl hydroperoxide (tBHP) or hydrogen peroxide (H2O2). These oxidized α/β-type SASP no longer bound to DNA effectively, but DNA binding protected α/β-type SASP against methionine oxidation by peroxides in vitro. Incubation of an oxidized α/β-type SASP with peptidyl methionine sulfoxide reductase (MsrA), which can reduce methionine sulfoxide residues back to methionine, restored the α/β-type SASP’s ability to bind to DNA. Both tBHP and H2O2 caused some oxidation of the two methionine residues of an α/β-type SASP (SspC) in spores of Bacillus subtilis, although one methionine which is highly conserved in α/β-type SASP was only oxidized to a small degree. However, much more methionine sulfoxide was generated by peroxide treatment of spores carrying a mutant form of SspC which has a lower affinity for DNA. MsrA activity was present in wild-type B. subtilis spores. However, msrA mutant spores were no more sensitive to H2O2 than were wild-type spores. The major mechanism operating for dealing with oxidative damage to α/β-type SASP in spores is DNA binding, which protects the protein’s methionine residues from oxidation both in vitro and in vivo. This may be important in vivo since α/β-type SASP containing oxidized methionine residues no longer bind DNA well and α/β-type SASP-DNA binding is essential for long-term spore survival.  相似文献   

11.
Bacterial spores are renowned for their longevity, ubiquity, and resistance to environmental insults, but virtually nothing is known regarding whether these metabolically dormant structures impact their surrounding chemical environments. In the present study, a number of spore-forming bacteria that produce dormant spores which enzymatically oxidize soluble Mn(II) to insoluble Mn(IV) oxides were isolated from coastal marine sediments. The highly charged and reactive surfaces of biogenic metal oxides dramatically influence the oxidation and sorption of both trace metals and organics in the environment. Prior to this study, the only known Mn(II)-oxidizing sporeformer was the marine Bacillus sp. strain SG-1, an extensively studied bacterium in which Mn(II) oxidation is believed to be catalyzed by a multicopper oxidase, MnxG. Phylogenetic analysis based on 16S rRNA and mnxG sequences obtained from 15 different Mn(II)-oxidizing sporeformers (including SG-1) revealed extensive diversity within the genus Bacillus, with organisms falling into several distinct clusters and lineages. In addition, active Mn(II)-oxidizing proteins of various sizes, as observed in sodium dodecyl sulfate-polyacrylamide electrophoresis gels, were recovered from the outer layers of purified dormant spores of the isolates. These are the first active Mn(II)-oxidizing enzymes identified in spores or gram-positive bacteria. Although extremely resistant to denaturation, the activities of these enzymes were inhibited by azide and o-phenanthroline, consistent with the involvement of multicopper oxidases. Overall, these studies suggest that the commonly held view that bacterial spores are merely inactive structures in the environment should be revised.  相似文献   

12.
Interactions between spores of Bacillus anthracis and macrophages are critical for the development of anthrax infections, as spores are thought to use macrophages as vehicles to disseminate in the host. In this study, we report a novel mechanism for phagocytosis of B. anthracis spores. Murine macrophage-like cell line RAW264.7, bone marrow-derived macrophages, and primary peritoneal macrophages from mice were used. The results indicated that activation of the classical complement pathway (CCP) was a primary mechanism for spore phagocytosis. Phagocytosis was significantly reduced in the absence of C1q or C3. C3 fragments were found deposited on the spore surface, and the deposition was dependent on C1q and Ca(2+). C1q recruitment to the spore surface was mediated by the spore surface protein BclA, as recombinant BclA bound directly and specifically to C1q and inhibited C1q binding to spores in a dose-dependent manner. C1q binding to spores lacking BclA (ΔbclA) was also significantly reduced compared with wild-type spores. In addition, deposition of both C3 and C4 as well as phagocytosis of spores were significantly reduced when BclA was absent, but were not reduced in the absence of IgG, suggesting that BclA, but not IgG, is important in these processes. Taken together, these results support a model in which spores actively engage CCP primarily through BclA interaction with C1q, leading to CCP activation and opsonophagocytosis of spores in an IgG-independent manner. These findings are likely to have significant implications on B. anthracis pathogenesis and microbial manipulation of complement.  相似文献   

13.
Bacillus subtilis spores that germinated poorly with saturating levels of nutrient germinants, termed superdormant spores, were separated from the great majority of dormant spore populations that germinated more rapidly. These purified superdormant spores (1.5 to 3% of spore populations) germinated extremely poorly with the germinants used to isolate them but better with germinants targeting germinant receptors not activated in superdormant spore isolation although not as well as the initial dormant spores. The level of β-galactosidase from a gerA-lacZ fusion in superdormant spores isolated by germination via the GerA germinant receptor was identical to that in the initial dormant spores. Levels of the germination proteins GerD and SpoVAD were also identical in dormant and superdormant spores. However, levels of subunits of a germinant receptor or germinant receptors activated in superdormant spore isolation were 6- to 10-fold lower than those in dormant spores, while levels of subunits of germinant receptors not activated in superdormant spore isolation were only ≤ 2-fold lower. These results indicate that (i) levels of β-galactosidase from lacZ fusions to operons encoding germinant receptors may not be an accurate reflection of actual germinant receptor levels in spores and (ii) a low level of a specific germinant receptor or germinant receptors is a major cause of spore superdormancy.  相似文献   

14.
An analysis of metabolism by measurement of respiratory quotient values indicates that reduced substances, such as lipids and/or amino acids, are the primary respiratory substrates of dormant Dictyostelium discoideum spores. The spores appear to consume both reduced substances and carbohydrates during the swelling stage of germination. The respiration of emerged myxamoebae is again dominated by the consumption of reduced substances. The pool of trehalose remains largely intact during heat-induced activation and also during postactivation lag. The initiation of spore swelling is accompanied by a decrease in the trehalose pool; the majority of trehalose is consumed before late spore swelling. Upon placing heat-activated spores under restrictive environmental conditions, swelling and trehalose hydrolysis are both prevented. Release from these conditions results in rapid swelling and hydrolysis of trehalose. Trehalase, the enzyme responsible for trehalose breakdown, is present in dormant spores at basal levels. This preformed enzyme is responsible for the hydrolysis of trehalose even though there is a significant increase in trehalase activity with the emergence of myxamoebae. RNA and protein synthesis inhibitors do not prevent trehalose hydrolysis or spore swelling. It is concluded that oxidation of reduced substances occurs in dormant, activated, and swollen spores, as well as in emerged myxamoebae of D. discoideum. Carbohydrate utilization dominates over the oxidation of reduced substances only during the swelling stage of germination.  相似文献   

15.
Thermograms of the exosporium-lacking dormant spores of Bacillus megaterium ATCC 33729, obtained by differential scanning calorimetry, showed three major irreversible endothermic transitions with peaks at 56, 100, and 114 degrees C and a major irreversible exothermic transition with a peak at 119 degrees C. The 114 degrees C transition was identified with coat proteins, and the 56 degrees C transition was identified with heat inactivation. Thermograms of the germinated spores and vegetative cells were much alike, including an endothermic transition attributable to DNA. The ascending part of the main endothermic 100 degrees C transition in the dormant-spore thermograms corresponded to a first-order reaction and was correlated with spore death; i.e., greater than 99.9% of the spores were killed when the transition peak was reached. The maximum death rate of the dormant spores during calorimetry, calculated from separately measured D and z values, occurred at temperatures above the 73 degrees C onset of thermal denaturation and was equivalent to the maximum inactivation rate calculated for the critical target. Most of the spore killing occurred before the release of most of the dipicolinic acid and other intraprotoplast materials. The exothermic 119 degrees C transition was a consequence of the endothermic 100 degrees C transition and probably represented the aggregation of intraprotoplast spore components. Taken together with prior evidence, the results suggest that a crucial protein is the rate-limiting primary target in the heat killing of dormant bacterial spores.  相似文献   

16.
Dormant spores of Bacillus megaterium contained no detectable reduced nicotinamide adenine dinucleotide (NADH) or reduced nicotinamide adenine dinucleotide phosphate (NADPH) despite significant levels of the oxidized forms of these nucleotides (NAD and NADP). During the first minutes of spore germination there was rapid accumulation of NADH and NADPH. However, this accumulation followed the fall in optical density that is characteristic of the initiation of spore germination. Accumulation of NADH and NADPH early in germination was not blocked by fluoride or cyanide, and it occurred even when germination was carried out in the absence of an exogenous source of reducing power. In addition to pyridine nucleotide reduction, de novo synthesis also began early in germination as the pyridine nucleotide levels increased to those found in growing cells. Midlog-phase cells grown in several different media had 20 to 35 times as much total pyridine nucleotide as did dormant spores. However, as growth and sporulation proceeded, the NADH plus NAD level fell four- to fivefold whereas the NADPH plus NADP level fell by a lesser amount. From min 10 of spore germination until midway through sporulation the value for the ratio of NADH/NAD is about 0.1 (0.03 to 0.18) while the ratio of NADPH/ANDP is about 1.4 (0.3 to 2.4). Comparison of these ratios in log-phase versus stationary phase (sporulation) growth in all three growth media tested did not reveal any common pattern of changes.  相似文献   

17.
Mitochondria from dormant spores of the fungus Botryodiplodia theobromae did not contain extractable cyctochrome c oxidase (EC 1.9.3.1) activity; however, this enzyme activity was elaborated rapidly after 150 min of the 240-min germination sequence. The absence of cytochrome c oxidase activity in the dormant spores apparently is not an artifact caused by spore disruption and fractionation procedures, transient enzyme instability, or insensitivity of the enzyme assay. Mitochondria from dormant spores of three other phylogenetically diverse genera of fungi were observed to contain readily detectable quantities of cytochrome c oxidase, suggesting that the absence of the enzyme in B. theobromae may be relatively novel. The elaboration of cytochrome c oxidase activity in germinating spores was abolished by cycloheximide if the drug was added at or before 95 min of germination, but development of enzyme activity was initially insensitive to inhibitors of the mitochondrial genetic system, chloramphenicol or ethidium bromide. Incubation of spores in both ethionine and S-2-aminoethyl-l-cysteine reduced the amount of extracted cytochrome c oxidase activity. Elaboration of enzyme activity was severely retarded by cerulenin, an inhibitor of fatty acid biosynthesis and of spore germination. This enzyme activity developed in water-incubated or 1% Tween 80-incubated spores in which only the cytoplasmic ribosomes are functional in translation of a stored nuclear messenger RNA. The results of this study show that cytoplasmic (but not mitochondrial) ribosome function is required for development of this enzyme activity during spore germination, and they suggest that a portion of the cytochrome c oxidase enzyme or some other protein required for its activity is synthesized de novo upon germination.  相似文献   

18.
Abstract RasG protein levels in dormant and germinating spores of Dictyostelium discoideum strains JC1 and SG1 were estimated by Western blotting. Ras Glevels were very low in dormant spores and remained low during the lag period, regardless of whether spores were heat activated or treated with autoactivator during the early stages of spore germination. RasG levels increased late during spore swelling just prior to the emergence stage of germination. These data are consistent with a requirement for RasG during vegetative growth.  相似文献   

19.
Previous investigators using the extent of uptake of the weak base methylamine to measure internal pH have shown that the pH in the core region of dormant spores of Bacillus megaterium is 6.3 to 6.5. Elevation of the internal pH of spores by 1.6 U had no significant effect on their degree of dormancy or their heat or ultraviolet light resistance. Surprisingly, the rate of methylamine uptake into dormant spores was slow (time for half-maximal uptake, 2.5 h at 24 degrees C). Most of the methylamine taken up by dormant spores was rapidly (time for half-maximal uptake, less than 3 min) released during spore germination as the internal pH of spores rose to approximately 7.5. This rise in internal spore pH took place before dipicolinic acid release, was not abolished by inhibition of energy metabolism, and during germination at pH 8.0 was accompanied by a decrease in the pH of the germination medium. Also accompanying the rise in internal spore pH during germination was the release of greater than 80% of the spores K+ and Na+. The K+ was subsequently reabsorbed in an energy-dependent process. These data indicate (i) that between pH 6.2 and 7.8 internal spore pH has little effect on dormant spore properties, (ii) that there is a strong permeability barrier in dormant spores to movement of charged molecules and small uncharged molecules, and (iii) that extremely early in spore germination this permeability barrier is breached, allowing rapid release of internal monovalent cations (H+, Na+, and K+).  相似文献   

20.
The role of osmotic pressure in the germination of Nosema algerae spores   总被引:1,自引:0,他引:1  
Both the lag period and the time required for the filament and sporoplasm to emerge from Nosema algerae spores were prolonged when germination occurred under hyperosmotic conditions. Polyethylene glycol (PEG) and sucrose inhibited germination, first by preventing eversion of the filament, and then at higher concentrations by preventing stimulation. The size of the spore cases decreased by about 21% following germination, indicating an elastic spore wall and turgor pressure in the dormant spores. Increased pressure during germination was indicated by less osmotically-induced shrinkage in stimulated than in dormant spores and by higher concentration of solutes in the homogenates of germinated than ungerminated spores. These results are consistent with the hypothesis of a pressure increase during germination that is caused by an endogenous increase in solute concentration.  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