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.
Microbial communities in marine hydrothermal sediments (0 to 30 cm deep) in an inlet of Kodakara-Jima Island, Kagoshima, Japan, were studied with reference to environmental factors, especially the presence of amino acids. The study area was shallow, and the sea floor was covered with sand through which hot volcanic gas bubbled and geothermally heated water seeped out. The total bacterial density increased with depth in the sediments in parallel with a rise in the ambient temperature (80(deg)C at the surface and 104(deg)C at a depth of 30 cm in the sediments). As estimated by most-probable-number studies, hyperthermophilic sulfur-dependent heterotrophs growing at 90(deg)C dominated the microbial community (3 x 10(sup7) cells (middot) g of sediment(sup-1) at a depth of 30 cm in the sediments), followed in abundance by hyperthermophilic sulfur-dependent facultative autotrophs (3.3 x 10(sup2) cells (middot) g of sediment(sup-1)). The cooler sandy or rocky floor surrounding the hot spots was covered with white bacterial mats which consisted of large Beggiatoa-like filaments. Both the total organic carbon content, most of which was particulate (75% in the surface sediments), and the amino acid concentration in void seawater in the sediments decreased with depth. Amino acids, both hydrolyzable and free, constituted approximately 23% of the dissolved organic carbon in the surface sediments. These results indicate that a lower amino acid concentration is probably due to consumption by dense populations of hyperthermophilic sulfur-dependent heterotrophs, which require amino acids for their growth and thus create a gradient of amino acid concentration in the sediments. The role of primary producers, which supply essential amino acids to sustain this microbial community, is also discussed.  相似文献   

2.
Plasmid transfer rates for the conjugative plasmid R388::Tn1721 from Pseudomonas cepacia (donor) to Pseudomonas fluorescens (recipient) on agar media, in broth, and in microcosms containing sterile or nonsterile soil, in the presence or absence of germinating pea seeds, were determined. Donors, recipients, and transconjugants were enumerated on selective media after 1 day on agar or in broth culture and over a 7-day period in soil or pea spermosphere microcosms. Donor and recipient growth rates and plasmid transfer rate constants [(gamma), where (gamma) = transconjugants (middot) (donors (middot) recipients)(sup-1) (middot) h(sup-1)] were calculated for three initial parental densities (10(sup4), 10(sup6), or 10(sup8) CFU/g or ml) in each system. For all initial density levels, values of (gamma) in agar and broth matings were higher than those in soil or in the pea spermosphere-rhizosphere microcosms. Values of (gamma) were not influenced by the pea spermosphere or by sterile or nonsterile conditions of the soil. However, (gamma) values in microcosm experiments were inversely related to initial parental density and were directly related to donor growth rates. Values of (gamma) averaged 4 x 10(sup-10), 4 x 10(sup-12), and 3 x 10(sup-14) when initial donor and recipient cell densities were 10(sup4), 10(sup6), and 10(sup8) CFU/g, respectively. These results suggest that the plasmid transfer rate constant is independent of parental cell density only when parental growth is not limited. In a resource-limited environment, intra- or interspecific competition may reduce the transfer rate by limiting parental growth.  相似文献   

3.
An exploratory study carried out in Pyrenean and Alpine lakes shows that a rich, active microbial community lives in the slush layers of the winter cover of such lakes in spite of the low temperature and the seasonal occurrence of the habitat. Bacteria were very diverse in morphology, with filaments reaching up to 100 (mu)m long; flagellates, both autotrophic (chrysophytes, cryptophytes, dinoflagellates, and volvocales) and heterotrophic, and ciliates were abundant, reaching biovolume values up to 2.7 x 10(sup6) (mu)m(sup3) ml(sup-1). Species composition was very variable, with dominance depending on date and depth. Although many species were typical of lake plankton communities, some were restricted to the slush, for instance the predatory ciliates Dileptus sp. and Lacrymaria sp., and others were restricted to the surface pools, such as the snow algae Chlamydomonas nivalis. Microbial biomasses and usually bacterial and algal activities were greater in the slush layers than in the lake water. Photosynthesis rate in the upper cover layers reached values up to 0.5 (mu)g of C liter(sup-1) h(sup-1), and high bacterial activities up to 226 pmol of leucine incorporated liter(sup-1) h(sup-1) and 25 pmol of thymidine incorporated liter(sup-1) h(sup-1) were measured. For most species, lake water flooding the ice and snow cover could provide an inoculum. Differential growth depending on the environmental conditions (nutrients, organic matter, light) of a particular slush layer could provide dominance of different groups or species. However, there was no obvious colonizing mechanism for those species not appearing either in plankton or in communities on top of the snowpack.  相似文献   

4.
The metabolism of atmospheric methane in a forest soil was studied by radiotracer techniques. Maximum (sup14)CH(inf4) oxidation (163.5 pmol of C cm(sup-3) h(sup-1)) and (sup14)C assimilation (50.3 pmol of C cm(sup-3) h(sup-1)) occurred at the A(inf2) horizon located 15 to 18 cm below the soil surface. At this depth, 31 to 43% of the atmospheric methane oxidized was assimilated into microbial biomass; the remaining methane was recovered as (sup14)CO(inf2). Methane-derived carbon was incorporated into all major cell macromolecules by the soil microorganisms (50% as proteins, 19% as nucleic acids and polysaccharides, and 5% as lipids). The percentage of methane assimilated (carbon conversion efficiency) remained constant at temperatures between 5 and 20(deg)C, followed by a decrease at 30(deg)C. The carbon conversion efficiency did not increase at methane concentrations between 1.7 and 1,000 ppm. In contrast, the overall methane oxidation activity increased at elevated methane concentrations, with an apparent K(infm) of 21 ppm (31 nM CH(inf4)) and a V(infmax) of 188 pmol of CH(inf4) cm(sup-3) h(sup-1). Methane oxidizers from soil depths with maximum methanotrophic activity respired approximately 1 to 3% of the assimilated methane-derived carbon per day. This apparent endogenous respiration did not change significantly in the absence of methane. Similarly, the potential for oxidation of atmospheric methane was relatively insensitive to methane starvation. Soil samples from depths above and below the zone with maximum atmospheric methane oxidation activity showed a dramatic increase in the turnover of the methane assimilated (>20 times increase). Physical disturbance such as sieving or mixing of soil samples decreased methane oxidation and assimilation by 50 to 58% but did not alter the carbon conversion efficiency. Ammonia addition (0.1 or 1.0 (mu)mol g [fresh weight](sup-1)) decreased both methane oxidation and carbon conversion efficiency. This resulted in a dramatic decrease in methane assimilation (85 to 99%). In addition, ammonia-treated soil showed up to 10 times greater turnover of the assimilated methane-derived carbon (relative to untreated soil). The results suggest a potential for microbial growth on atmospheric methane. However, growth was regulated strongly by soil parameters other than the methane concentration. The pattern observed for metabolism of atmospheric methane in soils was not consistent with the physiology of known methanotrophic bacteria.  相似文献   

5.
J. Kim  G. Rhee 《Applied microbiology》1997,63(5):1771-1776
The growth dynamics of polychlorinated biphenyl (PCB)-dechlorinating microorganisms were determined for the first time, along with those of sulfate reducers and methanogens, by using the most-probable-number technique. The time course of Aroclor 1248 dechlorination mirrored the growth of dechlorinators; dechlorination ensued when the dechlorinating population increased by 2 orders of magnitude from 2.5 x 10(sup5) to 4.6 x 10(sup7) cells g of sediment(sup-1), at a specific growth rate of 6.7 day(sup-1) between 2 and 6 weeks. During this period, PCB-dechlorinating microorganisms dechlorinated Aroclor 1248 at a rate of 3.9 x 10(sup-8) mol of Cl g of sediment(sup-1) day(sup-1), reducing the average number of Cl molecules per biphenyl from 3.9 to 2.8. The growth yield was 4.2 x 10(sup13) cells mol of Cl dechlorinated(sup-1). Once dechlorination reached a plateau, after 6 weeks, the number of dechlorinators began to decrease. On the other hand, dechlorinators inoculated into PCB-free sediments decreased over time from their initial level, suggesting that PCBs are required for their selective enrichment. The numbers of sulfate reducers and methanogens increased in both PCB-free and contaminated sediments, showing little difference between them. The maximum population size of sulfate reducers was about an order of magnitude higher than that of dechlorinators, whereas that of methanogens was slightly less. Unlike those of dechlorinators, however, numbers of both sulfate reducers and methanogens remained high even when dechlorination ceased. The results of this study imply that PCB concentrations may have to exceed a certain threshold to maintain the growth of PCB dechlorinators.  相似文献   

6.
Microbiological growth parameters, including limiting factors, kinetics, and minimal cell densities were assessed for subsurface microbiological communities collected with rock from an area proposed for a nuclear waste repository at Yucca Mountain, Nevada. Phospholipid fatty acid analysis revealed that approximately 104–105 viable cells per gram of dry rock are extant, and water availability was shown to be the primary factor limiting microbial growth in situ. Phosphate and carbon limitation, however, also suppress final cell densities by at least one order of magnitude under saturated conditions. Despite these limiting factors, significant growth of aerobic chemoheterotrophic microorganisms was shown to occur in unconcentrated simulated groundwater with or without addition of a reduced carbon source (7 × 107 and 8 × 106 planktonic cells/mL, respectively), indicating that when water becomes available in the repository environment, microbial growth will ensue. Organisms that were isolated from stationary cultures grown from Yucca Mountain rock in concentrated and unconcentrated simulated groundwaters showed significant 16S rDNA sequence divergence from reference organisms. Different (but related) organisms were isolated from concentrated and unconcentrated groundwater-grown cultures. Generally, as experimental conditions approached those expected to be encountered in situ, the organisms isolated were more divergent from reference organisms. Organisms that were isolated have metabolic properties that could allow them to be active and grow within the repository environment if water availability is sufficient.  相似文献   

7.
Twenty-six subsurface samples were collected from a borehole at depths of 173.3 to 196.8 m in the saturated zone at the Hanford Site in south-central Washington State. The sampling was performed throughout strata that included fine-grained lacustrine (lake) sediments, a paleosol (buried soil) sequence, and coarse-grained fluvial (river) sediments. A subcoring method and tracers were used to minimize and quantify contamination to obtain samples that were representative of subsurface strata. Sediment samples were tested for total organic carbon, inorganic carbon, total microorganisms by direct microscopic counts, culturable aerobic heterotrophs by plate counts, culturable anaerobes by most-probable-number enumeration, basal respiration rates, and mineralization of (sup14)C-labeled glucose and acetate. Total direct microscopic counts of microorganisms were low, ranging from below detection to 1.9 x 10(sup5) cells g (dry weight)(sup-1). Culturable aerobes and anaerobes were below minimum levels of detection in most samples. Direct microscopic counts, basal respiration rates, and (sup14)C-glucose mineralization were all positively correlated with total organic carbon and were highest in the lacustrine sediments. In contrast to previous subsurface studies, these saturated-zone samples did not have higher microbial abundance and activities than unsaturated sediments sampled from the same borehole, the fine-textured lacustrine sediment had higher microbial numbers and activities than the coarse-textured fluvial sands, and the paleosol samples did not have higher biomass and activities relative to the other sediments. The results of this study expand the subsurface microbiology database to include information from an environment very different from those previously studied.  相似文献   

8.
Budding methylotrophic bacteria resembling Hyphomicrobium spp. were counted for 12 months in a German sewage treatment plant by most-probable-number (MPN) methods. Influent samples contained up to 2 x 10(sup4) cells ml(sup-1), activated sludge consistently contained 1 x 10(sup5) to 5 x 10(sup5) cells ml(sup-1), and the effluent contained 1 x 10(sup3) to 4 x 10(sup3) cells ml(sup-1). The receiving lake had only 2 to 12 cells ml(sup-1). Six morphological groups with different growth requirements could be observed among 1,199 pure cultures that had been isolated from MPN dilutions. With dot blot DNA hybridizations, 671 isolates were assigned to 30 hybridization groups (HGs) and 84 could not be classified. Only HG 22 hybridized with a known species, Hyphomicrobium facilis IFAM B-522. Fourteen HGs (HGs 8 to 20 and HG 22) were specific for the lake; most others occurred only in the treatment plant. HGs 1, 3, and 26 were found in the activated sludge tank throughout the year, and HGs 27 and 28 were found for most of the year. In summary, it was demonstrated that bacteria with nearly identical and specific morphologies and nutritional types showed a high level of genetic diversity, although they were isolated under the same conditions and from the same treatment plant or its receiving lake. A directional exchange of these genetically different populations was possible but less significant, as was shown by the establishment of distinct populations in specific stations.  相似文献   

9.
Y. Konishi  S. Asai    N. Yoshida 《Applied microbiology》1995,61(10):3617-3622
The growth kinetics of Thiobacillus thiooxidans on elemental sulfur in batch cultures at 30(deg)C and pH 1.5 was studied by measuring the time courses of the concentration of adsorbed cells on sulfur, the concentration of free cells suspended in liquid medium, and the amount of sulfur oxidized. As the elemental sulfur was oxidized to sulfate ions, the surface concentration of adsorbed cells per unit mass of sulfur approached a maximum value (maximum adsorption capacity of sulfur particles) whereas the concentration of free cells continued to increase with time. There was a close relationship between the concentrations of free and adsorbed cells during the microbial sulfur oxidation, and the two cell concentrations were well correlated by the Langmuir isotherm with adsorption equilibrium constant K(infA) and maximum adsorption capacity X(infAm) of 2.10 x 10(sup-9) ml per cell and 4.57 x 10(sup10) cells per g, respectively. The total concentration of free and adsorbed cells increased in parallel with the amount of sulfate formed. The total growth on elemental sulfur gave a characteristic growth curve in which a linear-growth phase followed the period of an initial exponential phase. The batch rate data collected under a wide variety of inoculum levels (about 10(sup5) to 10(sup8) cells per ml) were consistent with a kinetic model assuming that the growth rate of adsorbed bacteria is proportional to the product of the concentration, X(infA), of adsorbed cells and the fraction, (theta)(infV), of adsorption sites unoccupied by cells. The kinetic and stoichiometric parameters appearing in the model were estimated from the experimental data, and the specific growth rate, (mu)(infA), and growth yield, Y(infA), were 2.58 day(sup-1) and 2.05 x 10(sup11) cells per g, respectively. The proposed model and the parameter values allowed us to predict quantitatively the surface attachment of T. thiooxidans cells on elemental sulfur and the bacterial growth in both initial exponential and subsequent linear phases. The transition from exponential to linear growth was a result of two competing factors: an increase in the adsorbed-cell concentration, X(infA), permitted a decrease in the unoccupied-site fraction, (theta)(infV).  相似文献   

10.
Vol. 61, no. 3, p. 1138, column 1, line 4 from bottom: "(A(inf600) = 0.1 = 8 x 10(sup8) cells ml(sup-1))" should read "(A(inf600) = 0.1 = 1.8 x 10(sup8) cells ml(sup-1))." [This corrects the article on p. 1138 in vol. 61.].  相似文献   

11.
Bacterium-specific viruses have attracted much interest in aquatic microbial ecology because they have been shown to be about 10 times more abundant than planktonic bacteria. So far most of the studies of interactions of planktonic bacteria and viruses have been done in marine environments, and very little is known about these interactions in lakes. Therefore, we studied phage proliferation in Lake Constance, a large mesotrophic lake in Germany. We enumerated bacteria and quantified the fraction of bacteria with mature intracellular phage particles and the number of free viruses by transmission electron microscopy. Between the end of March and early August 1992, peaks of bacterial abundance were followed in 1 to 2 weeks by peaks in the fraction of bacteria containing visible phage particles (0 to 1.7%) and in the number of free viruses (1 x 10(sup7) to 4 x 10(sup7) ml(sup-1)). We estimated that 1 to 17% +/- 12% of all bacteria were phage infected, implying that phage-induced mortality was <34% +/- 24% of total mortality. A direct comparison between phage-induced mortality, the net decrease of bacterial numbers, and bacterial growth rates indicated that phage-induced mortality accounted for <11% of total bacterial mortality during the phytoplankton spring bloom and 18 to 21% following the bloom. Estimated burst sizes ranged from 21 to 121 phages. Phage production rates of 0.5 x 10(sup6) to 2.5 x 10(sup6) ml(sup-1) day(sup-1) accounted for 70 to 380% of the observed net increase rates of free phages, implying high rates of simultaneous phage decay. The cyclic dynamics between bacteria and phages and the varying size structure of the intracellular mature phage particles suggested that phage infection was important in structuring the bacterial host assemblage during the study period.  相似文献   

12.
Fluorescently stained viruses were used as probes to label, identify, and enumerate specific strains of bacteria and cyanobacteria in mixed microbial assemblages. Several marine virus isolates were fluorescently stained with YOYO-1 or POPO-1 (Molecular Probes, Inc.) and added to seawater samples that contained natural microbial communities. Cells to which the stained viruses adsorbed were easily distinguished from nonhost cells; typically, there was undetectable binding of stained viruses to natural microbial assemblages containing >10(sup6) bacteria ml(sup-1) but to which host cells were not added. Host cells that were added to natural seawater were quantified with 99% (plusmn) 2% (mean (plusmn) range) efficiency with fluorescently labeled virus probes (FLVPs). A marine bacterial isolate (strain PWH3a), tentatively identified as Vibrio natriegens, was introduced into natural microbial communities that were either supplemented with nutrients or untreated, and changes in the abundance of the isolate were monitored with FLVPs. Simultaneously, the concentrations of viruses that infected strain PWH3a were monitored by plaque assay. Following the addition of PWH3a, the concentration of viruses infecting this strain increased from undetectable levels (<1 ml(sup-1)) to 2.9 x 10(sup7) and 8.3 x 10(sup8) ml(sup-1) for the untreated and nutrient-enriched samples, respectively. The increase in viruses was associated with a collapse in populations of strain PWH3a from ca. 30 to 2% and 43 to 0.01% of the microbial communities in untreated and nutrient-enriched samples, respectively. These results clearly demonstrate that FLVPs can be used to identify and quantify specific groups of bacteria in mixed microbial communities. The data show as well that viruses which are present at low abundances in natural aquatic viral communities can control microbial community structure.  相似文献   

13.
Roots of Phaseolus vulgaris L. were incubated with dilute suspensions (1 x 10(sup3) to 3 x 10(sup3) bacteria ml(sup-1)) of an antibiotic-resistant indicator strain of Rhizobium leguminosarum bv. phaseoli in mineral medium and washed four times by a standardized procedure prior to quantitation of adsorption (G. Caetano-Anolles and G. Favelukes, Appl. Environ. Microbiol. 52:371-376, 1986). The population of rhizobia remaining adsorbed on roots after washing was homogeneous, as indicated by the first-order course of its desorption by hydrodynamic shear. Rhizobia were maximally active for adsorption in the early stationary phase of growth. The process leading to adsorption was rapid, without an initial lag, and slowed down after 1 h. Adsorption of the indicator strain at 10(sup3) bacteria ml(sup-1) was inhibited to different extents in the presence of 10(sup3) to 10(sup8) antibiotic-sensitive competitor rhizobia ml(sup-1). After a steep rise above 10(sup4) bacteria ml(sup-1), inhibition by heterologous competitors in the concentration range of 10(sup5) to 10(sup7) bacteria ml(sup-1) was markedly less than by homologous strains, while at 10(sup8) bacteria ml(sup-1) it approached the high level of inhibition by the latter. At 10(sup7) bacteria ml(sup-1), all of the heterologous strains tested were consistently less inhibitory than homologous competitors (P < 0.001). These differences in competitive behavior indicate that in the process of adsorption of R. leguminosarum bv. phaseoli to its host bean roots, different modes of adsorption occur and that some of these modes are specific for the microsymbiont (as previously reported for the alfalfa system [G. Caetano-Anolles and G. Favelukes, Appl. Environ. Microbiol. 52:377-381, 1986]). Moreover, whereas the nonspecific process occurred either in the absence or in the presence of Ca(sup2+) and Mg(sup2+) ions, expression of specificity was totally dependent on the presence of those cations. R. leguminosarum bv. phaseoli bacteria adsorbed in the presence of Ca(sup2+) and Mg(sup2+) were more resistant to desorption by shear forces than were rhizobia adsorbed in their absence. These results indicate that (i) symbiotic specificity in the P. vulgaris-R. leguminosarum bv. phaseoli system is expressed already during the early process of rhizobial adsorption to roots, (ii) Ca(sup2+) and Mg(sup2+) ions are required by R. leguminosarum bv. phaseoli for that specificity, and (iii) those cations cause tighter binding of rhizobia to roots.  相似文献   

14.
The activity and distribution of CH(inf4)-oxidizing bacteria (MOB) in flooded rice (Oryza sativa) soil microcosms was investigated. CH(inf4) oxidation was shown to occur in undisturbed microcosms by using (sup14)CH(inf4), and model calculations indicated that almost 90% of the oxidation measured had taken place at a depth where only roots could provide the O(inf2) necessary. Slurry from soil planted with rice had an apparent K(infm) for CH(inf4) of 4 (mu)M and a V(infmax) of 0.1 (mu)mol g (dry weight)(sup-1) h(sup-1). At a depth of 1 to 2 cm, there was no significant difference (P > 0.05) in numbers of MOB between soil from planted and nonplanted microcosms (mean, 7.7 x 10(sup5) g [fresh weight](sup-1)). Thus, the densely rooted soil at 1 to 2 cm deep did not represent rhizospheric soil with respect to the number of MOB. A significantly increased number of MOB was found only in soil immediately around the roots (1.2 x 10(sup6) g [fresh weight](sup-1)), corresponding to a layer of 0.1 to 0.2 mm. Plant-associated CH(inf4) oxidation was shown in a double chamber with carefully washed intact rice plants. Up to 90% of the CH(inf4) supplied to the root compartment was oxidized in the plants. CH(inf4) oxidation on isolated roots was higher and had a larger variability than that in soil slurries. Roots had an apparent K(infm) for CH(inf4) of 6 (mu)M and a V(infmax) of 5 (mu)mol g (dry weight)(sup-1) h(sup-1). The average number of MOB in homogenized roots was larger than on the rhizoplane and increased with plant age. MOB also were found in surface-sterilized roots and basal culms, indicating the ability of these bacteria to colonize the interior of roots and culms.  相似文献   

15.
Induction of high-affinity iron transport during root colonization by Pseudomonas fluorescens Pf-5 (pvd-inaZ) was examined in lupine and barley growing in microcosms. P. fluorescens Pf-5 (pvd-inaZ) contains a plasmid carrying pvd-inaZ; thus, in this strain, ice nucleation activity is regulated by pyoverdin production. Lupine or barley plants were grown for 18 or 8 days, respectively, in soil amended with 2% calcium carbonate and inoculated with P. fluorescens Pf-5 (pvd-inaZ) at a density of 4 x 10(sup8) CFU g (dry weight) of soil(sup-1). A filter paper blotting technique was used to sample cells from the rhizosphere in different root zones, and then the cells were resuspended for enumeration and measurement of ice nucleation activity. The population density of P. fluorescens Pf-5 (pvd-inaZ) in the rhizosphere decreased by one order of magnitude in both lupine and barley over time. The ice nucleation activity ranged from -3.4 to -3.0 log ice nuclei CFU(sup-1) for lupine and -3.0 to -2.8 log ice nuclei CFU(sup-1) for barley, was similar in all root zones, and did not change over time. An in vitro experiment was conducted to determine the relationship between ice nucleation activity and pyoverdin production in P. fluorescens Pf-5 (pvd-inaZ). An ice nucleation activity of approximately -3.0 log ice nuclei CFU(sup-1) was measured in the in vitro experiment at 25 to 50 (mu)M FeCl(inf3). By using the regression between ice nucleation activity and pyoverdin production determined in vitro and assuming a P. fluorescens Pf-5 (pvd-inaZ) population density of 10(sup8) CFU g of root(sup-1), the maximum possible pyoverdin accumulation by P. fluorescens Pf-5 (pvd-inaZ) in the rhizosphere was estimated to be 0.5 and 0.8 nmol g of root(sup-1) for lupine and barley, respectively. The low ice nucleation activity measured in the rhizosphere suggests that nutritional competition for iron in the rhizosphere may not be a major factor influencing root colonization by P. fluorescens Pf-5 (pvd-inaZ).  相似文献   

16.
Three pyoverdines, Pf-A, Pf-B, and Pf-C, were purified with copper-chelate Sepharose and Sephadex G-15 columns from Pseudomonas fluorescens 2-79, and the yields (per 100 ml of culture supernatant) were 2.8, 21.6, and 3.2 mg, respectively. The absorption and fluorescence spectra of these pyoverdines were strongly pH dependent. Characteristic changes in the maximal absorbance wavelengths were observed when Fe(sup3+) or Cu(sup2+) was added. The addition of Cu(sup2+) shifted the pyoverdine Pf-B absorbance spectrum so that it exhibited a single peak at 410 nm but did not give rise to a new absorbance maximum at approximately 460 nm, which appeared when Fe(sup3+) was added. Fluorescence quenching experiments revealed that the forward reaction rate constant with pyoverdines was much higher with Cu(sup2+) (10(sup4) to 10(sup5) M(sup-1) s(sup-1)) than with Fe(sup3+) (10(sup2) M(sup-1) s(sup-1)). However, Cu(sup2+)-pyoverdine complexes were completely dissociated by EDTA at a low concentration (0.1 mM), while the level of Fe(sup3+)-pyoverdine complex dissociation at the same EDTA concentration was relatively low. The dissociation of Fe(sup3+)-pyoverdine complexes was EDTA concentration dependent. Formation of free pyoverdine was observed when the three types of Fe(sup3+)-pyoverdine complexes were incubated separately with P. fluorescens 2-79 cells, thus demonstrating that pyoverdines Pf-A, Pf-B, and Pf-C mediate iron transport.  相似文献   

17.
The effects of nitrate availability and the presence of Glyceria maxima on the composition and activity of the dissimilatory nitrate-reducing bacterial community were studied in the laboratory. Four different concentrations of NO(inf3)(sup-), 0, 533, 1434, and 2,905 (mu)g of NO(inf3)(sup-)-N g of dry sediment(sup-1), were added to pots containing freshwater sediment, and the pots were then incubated for a period of 69 days. Upon harvest, NH(inf4)(sup+) was not detectable in sediment that received 0 or 533 (mu)g of NO(inf3)(sup-)-N g of dry sediment(sup-1). Nitrate concentrations in these pots ranged from 0 to 8 (mu)g of NO(inf3)(sup-)-N g of dry sediment(sup-1) at harvest. In pots that received 1,434 or 2,905 (mu)g of NO(inf3)(sup-)-N g of dry sediment(sup-1), final concentrations varied between 10 and 48 (mu)g of NH(inf4)(sup+)-N g of dry sediment(sup-1) and between 200 and 1,600 (mu)g of NO(inf3)(sup-)-N g of dry sediment(sup-1), respectively. Higher input levels of NO(inf3)(sup-) resulted in increased numbers of potential nitrate-reducing bacteria and higher potential nitrate-reducing activity in the rhizosphere. In sediment samples from the rhizosphere, the contribution of denitrification to the potential nitrate-reducing capacity varied from 8% under NO(inf3)(sup-)-limiting conditions to 58% when NO(inf3)(sup-) was in ample supply. In bulk sediment with excess NO(inf3)(sup-), this percentage was 44%. The nitrate-reducing community consisted almost entirely of NO(inf2)(sup-)-accumulating or NH(inf4)(sup+)-producing gram-positive species when NO(inf3)(sup-) was not added to the sediment. The addition of NO(inf3)(sup-) resulted in an increase of denitrifying Pseudomonas and Moraxella strains. The factor controlling the composition of the nitrate-reducing community when NO(inf3)(sup-) is limited is the presence of G. maxima. In sediment with excess NO(inf3)(sup-), nitrate availability determines the composition of the nitrate-reducing community.  相似文献   

18.
Recent studies on Northern Ireland rivers have shown that summer nitrite (NO(inf2)(sup-)) concentrations greatly exceed the European Union guideline of 3 (mu)g of N liter(sup-1) for rivers supporting salmonid fisheries. In fast-flowing aerobic small streams, NO(inf2)(sup-) is thought to originate from nitrification, due to the retardation of Nitrobacter strains by the presence of free ammonia. Multiple regression analyses of NO(inf2)(sup-) concentrations against water quality variables of the six major rivers of the Lough Neagh catchment in Northern Ireland, however, suggested that the high NO(inf2)(sup-) concentrations found in the summer under warm, slow-flow conditions may result from the reduction of NO(inf3)(sup-). This hypothesis was supported by field observations of weekly changes in N species. Here, reduction of NO(inf3)(sup-) was observed to occur simultaneously with elevation of NO(inf2)(sup-) levels and subsequently NH(inf4)(sup+) levels, indicating that dissimilatory NO(inf3)(sup-) reduction to NH(inf4)(sup+) (DNRA) performed by fermentative bacteria (e.g., Aeromonas and Vibrio spp.) is responsible for NO(inf2)(sup-) accumulation in these large rivers. Mechanistic studies in which (sup15)N-labelled NO(inf3)(sup-) in sediment extracts was used provided further support for this hypothesis. Maximal concentrations of NO(inf2)(sup-) accumulation (up to 1.4 mg of N liter(sup-1)) were found in sediments deeper than 6 cm associated with a high concentration of metabolizable carbon and anaerobic conditions. The (sup15)N enrichment of the NO(inf2)(sup-) was comparable to that of the NO(inf3)(sup-) pool, indicating that the NO(inf2)(sup-) was predominantly NO(inf3)(sup-) derived. There is evidence which suggests that the high NO(inf2)(sup-) concentrations observed arose from the inhibition of the DNRA NO(inf2)(sup-) reductase system by NO(inf3)(sup-).  相似文献   

19.
A chemically defined medium for the growth of Leuconostoc mesenteroides was developed. This medium contained lactose, Mn(sup2+), Mg(sup2+), 12 amino acids, eight vitamins, adenine, uracil, and Tween 80. We showed the beneficial effect of aerobic conditions on growth and that potassium phosphate (135 mM) is a suitable buffer. The growth rate in this medium was 0.85 (plusmn) 0.10 h(sup-1) for the six strains examined, and cell densities up to 3.5 x 10(sup9) CFU/ml were attained.  相似文献   

20.
A minimal growth medium containing only nine amino acids and vitamins as the sole carbon and energy sources allowed the growth of Pyrococcus abyssi GE 5, a novel hyperthermophilic sulfur-metabolizing archaeon isolated from deep-sea hydrothermal vents. The generation time in this medium was about 40 min, and cell densities up to 5 x 10(sup8) cells ml(sup-1) were attained. These results are similar to those obtained previously with complex proteinaceous media.  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