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.
The population composition and biogeochemistry of sulfate-reducing bacteria (SRB) in the rhizosphere of the marsh grass Spartina alterniflora was investigated over two growing seasons by molecular probing, enumerations of culturable SRB, and measurements of SO42− reduction rates and geochemical parameters. SO42− reduction was rapid in marsh sediments with rates up to 3.5 μmol ml−1 day−1. Rates increased greatly when plant growth began in April and decreased again when plants flowered in late July. Results with nucleic acid probes revealed that SRB rRNA accounted for up to 43% of the rRNA from members of the domain Bacteria in marsh sediments, with the highest percentages occurring in bacteria physically associated with root surfaces. The relative abundance (RA) of SRB rRNA in whole-sediment samples compared to that of Bacteria rRNA did not vary greatly throughout the year, despite large temporal changes in SO42− reduction activity. However, the RA of root-associated SRB did increase from <10 to >30% when plants were actively growing. rRNA from members of the family Desulfobacteriaceae comprised the majority of the SRB rRNA at 3 to 34% of Bacteria rRNA, with Desulfobulbus spp. accounting for 1 to 16%. The RA of Desulfovibrio rRNA generally comprised from <1 to 3% of the Bacteria rRNA. The highest Desulfobacteriaceae RA in whole sediments was 26% and was found in the deepest sediment samples (6 to 8 cm). Culturable SRB abundance, determined by most-probable-number analyses, was high at >107 ml−1. Ethanol utilizers were most abundant, followed by acetate utilizers. The high numbers of culturable SRB and the high RA of SRB rRNA compared to that of Bacteria rRNA may be due to the release of SRB substrates in plant root exudates, creating a microbial food web that circumvents fermentation.  相似文献   

2.
An investigation of the terminal anaerobic processes occurring in polluted intertidal sediments indicated that terminal carbon flow was mainly mediated by sulfate-reducing organisms in sediments with high sulfate concentrations (>10 mM in the interstitial water) exposed to low loadings of nutrient (equivalent to <102 kg of N · day−1) and biochemical oxygen demand (<0.7 × 103 kg · day−1) in effluents from different pollution sources. However, in sediments exposed to high loadings of nutrient (>102 kg of N · day−1) and biochemical oxygen demand (>0.7 × 103 kg · day−1), methanogenesis was the major process in the mediation of terminal carbon flow, and sulfate concentrations were low (≤2 mM). The respiratory index [14CO2/(14CO2 + 14CH4)] for [2-14C]acetate catabolism, a measure of terminal carbon flow, was ≥0.96 for sediment with high sulfate, but in sediments with sulfate as little as 10 μM in the interstitial water, respiratory index values of ≤0.22 were obtained. In the latter sediment, methane production rates as high as 3 μmol · g−1 (dry weight) · h−1 were obtained, and there was a potential for active sulfate reduction.  相似文献   

3.
Microbial sulfate reduction is an important metabolic activity in petroleum hydrocarbon (PHC)-contaminated aquifers. We quantified carbon source-enhanced microbial SO42− reduction in a PHC-contaminated aquifer by using single-well push-pull tests and related the consumption of sulfate and added carbon sources to the presence of certain genera of sulfate-reducing bacteria (SRB). We also used molecular methods to assess suspended SRB diversity. In four consecutive tests, we injected anoxic test solutions (1,000 liters) containing bromide as a conservative tracer, sulfate, and either propionate, butyrate, lactate, or acetate as reactants into an existing monitoring well. After an initial incubation period, 1,000 liters of test solution-groundwater mixture was extracted from the same well. Average total test duration was 71 h. We measured concentrations of bromide, sulfate, and carbon sources in native groundwater as well as in injection and extraction phase samples and characterized the SRB population by using fluorescence in situ hybridization (FISH) and denaturing gradient gel electrophoresis (DGGE). Enhanced sulfate reduction concomitant with carbon source degradation was observed in all tests. Computed first-order rate coefficients ranged from 0.19 to 0.32 day−1 for sulfate reduction and from 0.13 to 0.60 day−1 for carbon source degradation. Sulfur isotope fractionation in unconsumed sulfate indicated that sulfate reduction was microbially mediated. Enhancement of sulfate reduction due to carbon source additions in all tests and variability of rate coefficients suggested the presence of specific SRB genera and a high diversity of SRB. We confirmed this by using FISH and DGGE. A large fraction of suspended bacteria hybridized with SRB-targeting probes SRB385 plus SRB385-Db (11 to 24% of total cells). FISH results showed that the activity of these bacteria was enhanced by addition of sulfate and carbon sources during push-pull tests. However, DGGE profiles indicated that the bacterial community structure of the dominant species did not change during the tests. Thus, the combination of push-pull tests with molecular methods provided valuable insights into microbial processes, activities, and diversity in the sulfate-reducing zone of a PHC-contaminated aquifer.  相似文献   

4.
We used an H2-purging culture vessel to replace an H2-consuming syntrophic partner, allowing the growth of pure cultures of Syntrophothermus lipocalidus on butyrate and Aminobacterium colombiense on alanine. By decoupling the syntrophic association, it was possible to manipulate and monitor the single organism's growth environment and determine the change in Gibbs free energy yield (ΔG) in response to changes in the concentrations of reactants and products, the purging rate, and the temperature. In each of these situations, H2 production changed such that ΔG remained nearly constant for each organism (−11.1 ± 1.4 kJ mol butyrate−1 for S. lipocalidus and −58.2 ± 1.0 kJ mol alanine−1 for A. colombiense). The cellular maintenance energy, determined from the ΔG value and the hydrogen production rate at the point where the cell number was constant, was 4.6 × 10−13 kJ cell−1 day−1 for S. lipocalidus at 55°C and 6.2 × 10−13 kJ cell−1 day−1 for A. colombiense at 37°C. S. lipocalidus, in particular, seems adapted to thrive under conditions of low energy availability.  相似文献   

5.
The kinetics for the reduction of sulfate alone and for concurrent uranium [U(VI)] and sulfate reduction, by mixed and pure cultures of sulfate-reducing bacteria (SRB) at 21 ± 3°C were studied. The mixed culture contained the SRB Desulfovibrio vulgaris along with a Clostridium sp. determined via 16S ribosomal DNA analysis. The pure culture was Desulfovibrio desulfuricans (ATCC 7757). A zero-order model best fit the data for the reduction of sulfate from 0.1 to 10 mM. A lag time occurred below cell concentrations of 0.1 mg (dry weight) of cells/ml. For the mixed culture, average values for the maximum specific reaction rate, Vmax, ranged from 2.4 ± 0.2 μmol of sulfate/mg (dry weight) of SRB · h−1) at 0.25 mM sulfate to 5.0 ± 1.1 μmol of sulfate/mg (dry weight) of SRB · h−1 at 10 mM sulfate (average cell concentration, 0.52 mg [dry weight]/ml). For the pure culture, Vmax was 1.6 ± 0.2 μmol of sulfate/mg (dry weight) of SRB · h−1 at 1 mM sulfate (0.29 mg [dry weight] of cells/ml). When both electron acceptors were present, sulfate reduction remained zero order for both cultures, while uranium reduction was first order, with rate constants of 0.071 ± 0.003 mg (dry weight) of cells/ml · min−1 for the mixed culture and 0.137 ± 0.016 mg (dry weight) of cells/ml · min−1 (U0 = 1 mM) for the D. desulfuricans culture. Both cultures exhibited a faster rate of uranium reduction in the presence of sulfate and no lag time until the onset of U reduction in contrast to U alone. This kinetics information can be used to design an SRB-dominated biotreatment scheme for the removal of U(VI) from an aqueous source.  相似文献   

6.
A method was developed to follow bacterial nitrate reduction in freshwater sediments by using common high-performance liquid chromatographic equipment. The low detection limit (14 pmol) of the method enabled us to study concentration profiles and reaction kinetics under natural conditions. Significant nitrate concentrations (1 to 27 μM) were observed in the sediment of Lake Vechten during the nonstratified period; the concentration profiles showed a successive depletion of oxygen, nitrate, and sulfate with depth. The profiles were restricted to the upper 3 cm of the sediment which is rich in organics and loosely structured. Nitrate reduction in the sediment-water interface followed first-order reaction kinetics at in situ concentrations. Remarkably high potential nitrate-reducing activity was observed in the part of the sediment in which nitrate did not diffuse. This activity was also observed throughout the whole year. Estimates of Km varied between 17 and 100 μM and Vmax varied between 7.2 and 36 μmol cm−3 day−1 for samples taken at different depths. The diffusion coefficient of nitrate ([10 ± 0.4] × 10−6 cm2 s−1) across the sediment-water interface was estimated by a constant-source technique and applied to a mathematical model to estimate the net nitrate reduction during the nonstratified period. In this period, observed nitrate reduction rates by the model, 0.2 to 0.4 mmol m−2 day−1, were lower than those found for oxygen (27 mmol m−2 day−1) and sulfate (0.4 mmol m−2 day−1). During the summer stratification, nitrate was absent in the sediment and reduction could not be estimated by the model.  相似文献   

7.
Soda lakes are naturally occurring highly alkaline and saline environments. Although the sulfur cycle is one of the most active element cycles in these lakes, little is known about the sulfate-reducing bacteria (SRB). In this study we investigated the diversity, activity, and abundance of SRB in sediment samples and enrichment cultures from a range of (hyper)saline soda lakes of the Kulunda Steppe in southeastern Siberia in Russia. For this purpose, a polyphasic approach was used, including denaturing gradient gel electrophoresis of dsr gene fragments, sulfate reduction rate measurements, serial dilutions, and quantitative real-time PCR (qPCR). Comparative sequence analysis revealed the presence of several novel clusters of SRB, mostly affiliated with members of the order Desulfovibrionales and family Desulfobacteraceae. We detected sulfate reducers and observed substantial sulfate reducing rates (between 12 and 423 μmol/dm3 day−1) for most lakes, even at a salinity of 475 g/liter. Enrichments were obtained at salt saturating conditions (4 M Na+), using H2 or volatile fatty acids as electron donors, and an extremely halophilic SRB, strain ASO3-1, was isolated. Furthermore, a high dsr gene copy number of 108 cells per ml was detected in a hypersaline lake by qPCR. Our results indicate the presence of diverse and active SRB communities in these extreme ecosystems.  相似文献   

8.
Values of Δ34S (, where δ34SHS and indicate the differences in the isotopic compositions of the HS and SO42− in the eluent, respectively) for many modern marine sediments are in the range of −55 to −75‰, much greater than the −2 to −46‰ ɛ34S (kinetic isotope enrichment) values commonly observed for microbial sulfate reduction in laboratory batch culture and chemostat experiments. It has been proposed that at extremely low sulfate reduction rates under hypersulfidic conditions with a nonlimited supply of sulfate, isotopic enrichment in laboratory culture experiments should increase to the levels recorded in nature. We examined the effect of extremely low sulfate reduction rates and electron donor limitation on S isotope fractionation by culturing a thermophilic, sulfate-reducing bacterium, Desulfotomaculum putei, in a biomass-recycling culture vessel, or “retentostat.” The cell-specific rate of sulfate reduction and the specific growth rate decreased progressively from the exponential phase to the maintenance phase, yielding average maintenance coefficients of 10−16 to 10−18 mol of SO4 cell−1 h−1 toward the end of the experiments. Overall S mass and isotopic balance were conserved during the experiment. The differences in the δ34S values of the sulfate and sulfide eluting from the retentostat were significantly larger, attaining a maximum Δ34S of −20.9‰, than the −9.7‰ observed during the batch culture experiment, but differences did not attain the values observed in marine sediments.Dissimilatory SO42− reduction is a geologically ancient, anaerobic, energy-yielding metabolic process during which SO42−-reducing bacteria (SRB) reduce SO42− to H2S while oxidizing organic molecules or H2. SO42− reduction is a dominant pathway for organic degradation in marine sediments (23) and in terrestrial subsurface settings where sulfur-bearing minerals dominate over Fe3+-bearing minerals. For example, at depths greater than 1.5 km below land surface in the fractured sedimentary and igneous rocks of the Witwatersrand Basin of South Africa, SO42− reduction is the dominant electron-accepting process (3, 26, 46, 48, 61).The enrichment of 32S in biogenic sulfides, with respect to the parent SO42−, imparted by SRB, is traceable through the geologic record (10, 54). The magnitude of the Δ34S (= δ34Spyrite − δ34Sbarite/gypsum, where δ34Spyrite and δ34Sbarite/gypsum are the isotopic compositions of pyrite and barite or gypsum) increases from −10‰ in the 3.47-billion-year-old North Pole deposits to −30‰ in late-Archaean deposits (55), to −75‰ in Neoproterozoic to modern sulfide-bearing marine sediments (13).The kinetic isotopic enrichment, ɛ34S, deduced from trends in the δ34S values of SO42− and HS in batch culture microbial SO42− reduction experiments using the Rayleigh relationship, ranges from −2‰ to −46‰ (6, 7, 11, 17, 22, 27, 28, 30, 31, 38, 39). The variation in ɛ34S values has been attributed to the SO42− concentration, the type of electron donor and its concentration, the SO42− reduction rate per cell (csSRR) (22), temperature, and species-specific isotope enrichment effects. In these laboratory experiments, doubling times are on the order of hours and csSRRs range from to 0.1 to 18 fmol cell−1 h−1 (7, 12, 17, 22, 30, 32, 39, 40).Experiments performed during the 1960s found that the magnitude of ɛ34S was inversely proportional to the csSRR for organic electron donors (16, 31, 38, 39) when SO42− was not limiting. More-recent batch culture experiments on 3 psychrophilic (optimum growth temperature, <20°C) and mesophilic (optimum growth temperature, between 20°C and 45°C) SRB strains (7) and on 32 psychrophilic to thermophilic SRB strains (22), however, have failed to reproduce such a relationship. In 2001, Canfield (11) reported an inverse correlation between ɛ34S and reduction rate using a flowthrough sediment column and demonstrated that ɛ34S values of approximately −35 to −40‰ were produced when organic substrates added by way of amendment were limited with respect to SO42−. Because it was not possible to readily evaluate changes in biomass in the sediment column with changes in temperature or substrate provision rate, it was inferred that changes in ɛ34S were related to changes in the csSRRs. More recently, Canfield et al. (12) observed a 6‰ variation in ɛ34S values related to the temperature of the batch culture experiments relative to the optimum growth temperature. The few early experiments that were performed using H2 as the electron donor yielded ɛ34S values ranging from −3 to −19‰ (22, 38, 39), which appear to correlate with the csSRR (39). Hoek et al. (32) also found that the ɛ34S values for the thermophilic SO42− reducer Thermodesulfatator indicus increased from between −1.5‰ and −10‰ in batch cultures with high H2 concentrations to between −24‰ and −37‰ in batch cultures grown under H2 limitation with respect to SO42−. Detmers et al. (22) found that the average ɛ34S of SRB that oxidize their organic carbon electron donor completely to CO2 averaged −25‰, versus −9.5‰ for SRB that release acetate during their oxidation of their organic carbon electron donor. Detmers et al. (22) speculated that the greater free energy yield per mole of SO42− from incomplete carbon oxidation relative to that for complete carbon oxidation promotes complete SO42− reduction and hinders isotopic enrichment due to isotopic exchange of the intracellular sulfur species pools.None of these experiments, however, have yielded ɛ34S factors capable of producing the Δ34S values of −55 to −75‰ observed in the geological record from ∼1.0 billion years ago to today. Various schemes have been hypothesized, and observations that involve either the disproportionation of S2O32− (36), the disproportionation of S0 produced by oxidation of either H2S or S2O32− (15), or the disproportionation of SO32− (29) have been made. Attribution of the increasing Δ34S values recorded for Achaean to Neoproterozoic sediments to the increasing role of H2S oxidative pathways makes sense in the context of increasing O2 concentrations in the atmosphere (14) but is not consistent with the lack of significant fractionation observed during oxidative reactions (29). To explain the Δ34S values of −55 to −77‰ reported to occur in interstitial pore waters from 100- to 300-m-deep, hypersulfidic ocean sediments (51, 64, 67), where the presence of a S-oxidative cycle is unlikely, an alternative, elaborate model of the SO42−-reducing pathway has been proposed by Brunner and Bernasconi (9). This model attributes the large Δ34S values to a multistep, reversible reduction of SO32− to HS involving S3O62− and S2O32− (20, 25, 41, 42, 52, 66). The conditions under which the maximum ɛ34S values might be expressed are a combination of elevated HS concentrations, electron donor limitations, nonlimiting SO42− concentrations, and a very low csSRR. The csSRR for subsurface environments has been estimated from biogeochemical-flux modeling to be 10−6 to 10−7 fmol cell−1 h−1 (23), with a corresponding cell turnover rate greater than 1,000 years (37).Batch and chemostat culture systems, despite low growth rates, cannot completely attain a state of zero growth with constant substrate provision and therefore do not accurately reflect the in situ nutritional states of microbes in many natural settings. Retentostats, or recycling fermentor vessels, recycle 100% of biomass to the culturing vessel, allowing experimenters to culture microbial cells to a large biomass with a constant nutrient supply rate until the substrate supply rate itself becomes the growth-limiting factor and cells enter a resting state in which their specific growth rate approaches zero and they carry on maintenance metabolism (1, 47, 53, 58, 59, 62, 63). Utilizing this approach, Colwell et al. (18) were able to obtain a cell-specific respiration rate of 7 × 10−4 fmol of CH4 cell−1 h−1 for a mesophilic marine methanogen, a rate that is comparable to that estimated for methanogenic communities in deep marine sediments off the coast of Peru (49).In this study, the conditions that Brunner and Bernasconi (9) hypothesized would lead to the large Δ34S values seen in nature were recreated in the laboratory by limiting the electron donor supply rate with respect to the SO42− supply rate in a retentostat vessel. The S isotopic enrichment by a resting culture of Desulfotomaculum putei at an extremely low csSRR was compared to that of a batch culture experiment to determine whether the ɛ34S values produced under the former conditions approach the Δ34S seen in nature.  相似文献   

9.
The relationship between growth rate and rRNA content in a marine Synechococcus strain was examined. A combination of flow cytometry and whole-cell hybridization with fluorescently labeled 16S rRNA-targeted oligonucleotide probes was used to measure the rRNA content of Synechococcus strain WH8101 cells grown at a range of light-limited growth rates. The sensitivity of this approach was sufficient for the analysis of rRNA even in very slowly growing Synechococcus cells (μ = 0.15 day−1). The relationship between growth rate and cellular rRNA content comprised three phases: (i) at low growth rates (<~0.7 day−1), rRNA cell−1 remained approximately constant; (ii) at intermediate rates (~0.7 − 1.6 day−1), rRNA cell−1 increased proportionally with growth rate; and (iii) at the highest, light-saturated rates (>~1.6 day−1), rRNA cell−1 dropped abruptly. Total cellular RNA (as measured with the nucleic acid stain SYBR Green II) was well correlated with the probe-based measure of rRNA and varied in a similar manner with growth rate. Mean cell volume and rRNA concentration (amount of rRNA per cubic micrometer) were related to growth rate in a manner similar to rRNA cell−1, although the overall magnitude of change in both cases was reduced. These patterns are hypothesized to reflect an approximately linear increase in ribosome efficiency with increasing growth rate, which is consistent with the prevailing prokaryotic model at low growth rates. Taken together, these results support the notion that measurements of cellular rRNA content might be useful for estimating in situ growth rates in natural Synechococcus populations.  相似文献   

10.
Hadas  Ora  Pinkas  Riki 《Hydrobiologia》1992,235(1):295-301
Monomictic Lake Kinneret is stratified during summer and autumn, resulting in a hypolimnion rich in H2S (3–7 mg 1–1). In winter and spring every year a bloom of dinoflagallate Peridinium gatunense produces an average biomass of 150000 ton wet weight. Part of this biomass sinks to the hypolimnion and sediments where it is decomposed and mineralized, with some of the mineralization due to the activity of sulfate-reducing bacteria (SRB). The sulfate-reduction potential of the upper sediment layer at the deepest part of the lake (42 m) was measured. The activity of the enzyme arylsulfatase was also monitored. Rates of sulfate-reduction ranged from a minimum of 12 nmoles SOf4 p2–-reduced cm–3 day–1 in December before lake overturn to a maximum of 1673 nmoles SOf4 p2– reduced cm–3 day–1 in July during stratification. These rates are considerably higher than those recorded from other freshwater lakes in the world and are probably limited more by the availability of organic matter than by sulfate concentrations.  相似文献   

11.
Propionate consumption was studied in syntrophic batch and chemostat cocultures of Syntrophobacter fumaroxidans and Methanospirillum hungatei. The Gibbs free energy available for the H2-consuming methanogens was <−20 kJ mol of CH4−1 and thus allowed the synthesis of 1/3 mol of ATP per reaction. The Gibbs free energy available for the propionate oxidizer, on the other hand, was usually >−10 kJ mol of propionate−1. Nevertheless, the syntrophic coculture grew in the chemostat at steady-state rates of 0.04 to 0.07 day−1 and produced maximum biomass yields of 2.6 g mol of propionate−1 and 7.6 g mol of CH4−1 for S. fumaroxidans and M. hungatei, respectively. The energy efficiency for syntrophic growth of S. fumaroxidans, i.e., the biomass produced per unit of available Gibbs free energy was comparable to a theoretical growth yield of 5 to 12 g mol of ATP−1. However, a lower growth efficiency was observed when sulfate served as an additional electron acceptor, suggesting inefficient energy conservation in the presence of sulfate. The maintenance Gibbs free energy determined from the maintenance coefficient of syntrophically grown S. fumaroxidans was surprisingly low (0.14 kJ h−1 mol of biomass C−1) compared to the theoretical value. On the other hand, the Gibbs free-energy dissipation per mole of biomass C produced was much higher than expected. We conclude that the small Gibbs free energy available in many methanogenic environments is sufficient for syntrophic propionate oxidizers to survive on a Gibbs free energy that is much lower than that theoretically predicted.  相似文献   

12.
The coexistence of sulfate-reducing bacteria (SRB) and methanogenic archaea (MA) in anaerobic biofilms developed in sewer inner pipe surfaces favors the accumulation of sulfide (H2S) and methane (CH4) as metabolic end products, causing severe impacts on sewerage systems. In this study, we investigated the time course of H2S and CH4 production and emission rates during different stages of biofilm development in relation to changes in the composition of microbial biofilm communities. The study was carried out in a laboratory sewer pilot plant that mimics a full-scale anaerobic rising sewer using a combination of process data and molecular techniques (e.g., quantitative PCR [qPCR], denaturing gradient gel electrophoresis [DGGE], and 16S rRNA gene pyrotag sequencing). After 2 weeks of biofilm growth, H2S emission was notably high (290.7 ± 72.3 mg S-H2S liter−1 day−1), whereas emissions of CH4 remained low (17.9 ± 15.9 mg COD-CH4 liter−1 day−1). This contrasting trend coincided with a stable SRB community and an archaeal community composed solely of methanogens derived from the human gut (i.e., Methanobrevibacter and Methanosphaera). In turn, CH4 emissions increased after 1 year of biofilm growth (327.6 ± 16.6 mg COD-CH4 liter−1 day−1), coinciding with the replacement of methanogenic colonizers by species more adapted to sewer conditions (i.e., Methanosaeta spp.). Our study provides data that confirm the capacity of our laboratory experimental system to mimic the functioning of full-scale sewers both microbiologically and operationally in terms of sulfide and methane production, gaining insight into the complex dynamics of key microbial groups during biofilm development.  相似文献   

13.
Sulfate Reduction in Freshwater Sediments Receiving Acid Mine Drainage   总被引:24,自引:8,他引:16       下载免费PDF全文
One arm of Lake Anna, Va., receives acid mine drainage (AMD) from Contrary Creek (SO42− concentration = 2 to 20 mM, pH = 2.5 to 3.5). Acid-volatile sulfide concentrations, SO42− reduction rates, and interstitial SO42− concentrations were measured at various depths in the sediment at four stations in four seasons to assess the effects of the AMD-added SO42− on bacterial SO42− reduction. Acid-volatile sulfide concentrations were always an order of magnitude higher at the stations receiving AMD than at a control station in another arm of the lake that received no AMD. Summer SO42− reduction rates were also an order of magnitude higher at stations that received AMD than at the control station (226 versus 13.5 mmol m−2 day−1), but winter values were inconclusive, probably due to low sediment temperature (6°C). Profiles of interstitial SO42− concentrations at the AMD stations showed a rapid decrease with depth (from 1,270 to 6 μM in the top 6 cm) due to rapid SO42− reduction. Bottom-water SO42− concentrations in the AMD-receiving arm were highest in winter and lowest in summer. These data support the conclusion that there is a significant enhancement of SO42− reduction in sediments receiving high SO42− inputs from AMD.  相似文献   

14.
Dynamics of Bacterial Sulfate Reduction in a Eutrophic Lake   总被引:22,自引:13,他引:9       下载免费PDF全文
Bacterial sulfate reduction in the surface sediment and the water column of Lake Mendota, Madison, Wis., was studied by using radioactive sulfate (35SO42−). High rates of sulfate reduction were observed at the sediment surface, where the sulfate pool (0.2 mM SO42−) had a turnover time of 10 to 24 h. Daily sulfate reduction rates in Lake Mendota sediment varied from 50 to 600 nmol of SO42− cm−3, depending on temperature and sampling date. Rates of sulfate reduction in the water column were 103 times lower than that for the surface sediment and, on an areal basis, accounted for less than 18% of the total sulfate reduction in the hypolimnion during summer stratification. Rates of bacterial sulfate reduction in the sediment were not sulfate limited at sulfate concentrations greater than 0.1 mM in short-term experiments. Although sulfate reduction seemed to be sulfate limited below 0.1 mM, Michaelis-Menten kinetics were not observed. The optimum temperature (36 to 37°C) for sulfate reduction in the sediment was considerably higher than in situ temperatures (1 to 13°C). The response of sulfate reduction to the addition of various electron donors metabolized by sulfate-reducing bacteria in pure culture was investigated. The degree of stimulation was in this order: H2 > n-butanol > n-propanol > ethanol > glucose. Acetate and lactate caused no stimulation.  相似文献   

15.
Rapid Methane Oxidation in a Landfill Cover Soil   总被引:28,自引:5,他引:28       下载免费PDF全文
Methane oxidation rates observed in a topsoil covering a retired landfill are the highest reported (45 g m−2 day−1) for any environment. This microbial community had the capacity to rapidly oxidize CH4 at concentrations ranging from <1 ppm (microliters per liter) (first-order rate constant [k] = −0.54 h−1) to >104 ppm (k = −2.37 h−1). The physiological characteristics of a methanotroph isolated from the soil (characteristics determined in aqueous medium) and the natural population, however, were similar to those of other natural populations and cultures: the Q10 and optimum temperature were 1.9 and 31°C, respectively, the apparent half-saturation constant was 2.5 to 9.3 μM, and 19 to 69% of oxidized CH4 was assimilated into biomass. The CH4 oxidation rate of this soil under waterlogged (41% [wt/vol] H2O) conditions, 6.1 mg liter−1 day−1, was near rates reported for lake sediment and much lower than the rate of 116 mg liter−1 day−1 in the same soil under moist (11% H2O) conditions. Since there are no large physiological differences between this microbial community and other CH4 oxidizers, we attribute the high CH4 oxidation rate in moist soil to enhanced CH4 transport to the microorganisms; gas-phase molecular diffusion is 104-fold faster than aqueous diffusion. These high CH4 oxidation rates in moist soil have implications that are important in global climate change. Soil CH4 oxidation could become a negative feedback to atmospheric CH4 increases (and warming) in areas that are presently waterlogged but are projected to undergo a reduction in summer soil moisture.  相似文献   

16.
The steady-state effect of 2,5,2′,5′-tetrachlorobiphenyl (TCBP) on the green alga Selenastrum capricornutum was investigated in a P-limited two-stage chemostat system. The partition coefficient of this polychlorinated biphenyl congener was 5.9 × 104 in steady-state cultures. At a cellular TCBP concentration of 12.2 × 10−8 ng · cell−1, growth rate was not affected. However, photosynthetic capacity (Pmax) was significantly enhanced by TCBP (56 × 10−9 μmol of C · cell−1 · h−1 versus 34 × 10−9 μmol of C · cell−1 · h−1 in the control). Photosynthetic efficiency, or the slope of the photosynthesis-irradiance curve, was also significantly higher. There was little difference in the cell chlorophyll a content, and therefore the difference in these photosynthetic characteristics was the same even when they were expressed on a per-chlorophyll a basis. Cell C content was higher in TCBP-containing cells than in TCBP-free cells, but approximately 36% of the C fixed by cells with TCBP was not incorporated as cell C. The maximum P uptake rate was also enhanced by TCBP, but the half-saturation concentration appeared to be unaffected.  相似文献   

17.
The distribution of sulphate-reducing bacteria (SRB) in the sediments of the Colne River estuary, Essex, UK covering different saline concentrations of sediment porewater was investigated by the use of quantitative competitive PCR. Here, we show that a new PCR primer set and a new quantitative method using PCR are useful tools for the detection and the enumeration of SRB in natural environments. A PCR primer set selective for the dissimilatory sulphite reductase gene (dsr) of SRB was designed. PCR amplification using the single set of dsr-specific primers resulted in PCR products of the expected size from all 27 SRB strains tested, including Gram-negative and positive species. Sixty clones derived from sediment DNA using the primers were sequenced and all were closely related with the predicted dsr of SRB. These results indicate that PCR using the newly designed primer set are useful for the selective detection of SRB from a natural sample. This primer set was used to estimate cell numbers by dsr selective competitive PCR using a competitor, which was about 20% shorter than the targeted region of dsr. This procedure was applied to sediment samples from the River Colne estuary, Essex, UK together with simultaneous measurement of in situ rates of sulphate reduction. High densities of SRB ranging from 0.2 ? 5.7 × 108 cells ml? 1 wet sediment were estimated by the competitive PCR assuming that all SRB have a single copy of dsr. Using these estimates cell specific sulphate reduction rates of 10? 17 to 10? 15 mol of SO4 2 ? cell? 1 day? 1 were calculated, which is within the range of, or lower than, those previously reported for pure cultures of SRB. Our results show that the newly developed competitive PCR technique targeted to dsr is a powerful tool for rapid and reproducible estimation of SRB numbers in situ and is superior to the use of culture-dependent techniques.  相似文献   

18.
A kinetic model has been developed and kinetic parameters of anaerobic degradation of glycerol, an abundant by-product of biofuel manufacturing, by a consortium of sulfate reducing bacteria (SRB) in a closed system have been determined. The following main species of SRB has been identified in the consortium: Desulfovibrio baarsii, Desulfomicrobium sp., and Desufatomaculum sp. The proposed model included processes of glycerol degradation, sulfate reduction, and inhibition by metabolic products, as well as effects of pH and temperature. The suggested equation for the anaerobic glycerol degradation was based on Edward and Andrew’s equation. The following kinetic parameters of the anaerobic glycerol degradation were obtained for the initial glycerol concentration from 0.15 to 4 ml/l and sulfate concentration of 2760 mg/l at 22°C: maximum specific growth rate of SRB μmax = 0.56 day−1, economic coefficient of ashless biomass from glycerol of 0.08 mol SRB/mol COC, and yield of ashless biomass from sulfate of 0.020 mol SRB/mol SO4. It was shown that the optimum molar ratio of $ {{C_{Gl} } \mathord{\left/ {\vphantom {{C_{Gl} } {C_{SO_4 } }}} \right. \kern-\nulldelimiterspace} {C_{SO_4 } }} $ {{C_{Gl} } \mathord{\left/ {\vphantom {{C_{Gl} } {C_{SO_4 } }}} \right. \kern-\nulldelimiterspace} {C_{SO_4 } }} for SRB growth was 0.8. Initial boundary concentration of inhibition by undissociated hydrogen sulfide was 70 mg/l. Dependence of the specific growth rate of bacteria on the temperature was approximated by the Arrhenius equation in the temperature range of 20–30°C with the goodness of fit R2 = 0.99.  相似文献   

19.
Bacterial sulfate reduction activity (SRA) was measured in surface sediments and slurries from three sites in the Great Salt Lake (Utah, USA) using radiolabeled 35S-sulfate. High rates of sulfate reduction (363 ± 103 and 6,131 ± 835 nmol cm-3 d-1) were measured at two sites in the moderately hypersaline southern arm of the lake, whereas significantly lower rates (32 ± 9 nmol cm-3 d-1) were measured in the extremely hypersaline northern arm. Bacterial sulfate reduction was strongly affected by salinity and showed an optimum around 5-6% NaCl in the southern arm and an optimum of around 12% NaCl in the more hypersaline northern arm of the lake. High densities of sulfate-reducing bacteria (SRB) ranging from 2.2 × 107 to 6.7 × 108 cells cm-3 were determined by a newly developed tracer MPN-technique (T-MPN) employing sediment media and 35S-sulfate. Calculation of specific sulfate reduction rates yielded values comparable to those obtained in pure cultures of SRB. However, when using a conventional MPN technique with synthetic media containing high amounts of Fe(II), the numbers of SRB were underestimated by 1-4 orders of magnitude as compared to the T-MPN method. Our results suggest that high densities of slightly to moderately halophilic and extremely halotolerant SRB are responsible for the high rates of sulfate reduction measured in Great Salt Lake sediments.  相似文献   

20.
Pseudomonas aeruginosa strain NB1 uses chloromethane (CM) as its sole source of carbon and energy under nitrate-reducing and aerobic conditions. The observed yield of NB1 was 0.20 (±0.06) (mean ± standard deviation) and 0.28 (±0.01) mg of total suspended solids (TSS) mg of CM−1 under anoxic and aerobic conditions, respectively. The stoichiometry of nitrate consumption was 0.75 (±0.10) electron equivalents (eeq) of NO3 per eeq of CM, which is consistent with the yield when it is expressed on an eeq basis. Nitrate was stoichiometrically converted to dinitrogen (0.51 ± 0.05 mol of N2 per mol of NO3). The stoichiometry of oxygen use with CM (0.85 ± 0.21 eeq of O2 per eeq of CM) was also consistent with the aerobic yield. Stoichiometric release of chloride and minimal accumulation of soluble metabolic products (measured as chemical oxygen demand) following CM consumption, under anoxic and aerobic conditions, indicated complete biodegradation of CM. Acetylene did not inhibit CM use under aerobic conditions, implying that a monooxygenase was not involved in initiating aerobic CM metabolism. Under anoxic conditions, the maximum specific CM utilization rate (k) for NB1 was 5.01 (±0.06) μmol of CM mg of TSS−1 day−1, the maximum specific growth rate (μmax) was 0.0506 day−1, and the Monod half-saturation coefficient (Ks) was 0.067 (±0.004) μM. Under aerobic conditions, the values for k, μmax, and Ks were 10.7 (±0.11) μmol of CM mg of TSS−1 day−1, 0.145 day−1, and 0.93 (±0.042) μM, respectively, indicating that NB1 used CM faster under aerobic conditions. Strain NB1 also grew on methanol, ethanol, and acetate under denitrifying and aerobic conditions, but not on methane, formate, or dichloromethane.  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