首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
N2 fixation by bacteria in associative symbiosis with washed roots of 13 Poaceae and 8 other noncultivated plant species in Finland was demonstrated by the acetylene reduction method. The roots most active in C2H2 reduction were those of Agrostis stolonifera, Calamagrostis lanceolata, Elytrigia repens, and Phalaris arundinacea, which produced 538 to 1,510 nmol of C2H4·g−1 (dry weight)· h−1 when incubated at pO2 0.04 with sucrose (pH 6.5), and 70 to 269 nmol of C2H4· g−1 (dry weight)·h−1 without an added energy source and unbuffered. Azospirillum lipferum, Enterobacter agglomerans, Klebsiella pneumoniae, and a Pseudomonas sp. were the acetylene-reducing organisms isolated. The results demonstrate the presence of N2-fixing organisms in associative symbiosis with plant roots found in a northern climatic region in acidic soils ranging down to pH 4.0.  相似文献   

2.
Nitrogenase activity in mangrove forests at two locations in the North Island, New Zealand, was measured by acetylene reduction and 15N2 uptake. Nitrogenase activity (C2H2 reduction) in surface sediments 0 to 10 mm deep was highly correlated (r = 0.91, n = 17) with the dry weight of decomposing particulate organic matter in the sediment and was independent of light. The activity was not correlated with the dry weight of roots in the top 10 mm of sediment (r = −0.01, n = 13). Seasonal and sample variation in acetylene reduction rates ranged from 0.4 to 50.0 μmol of C2H4 m−2 h−1 under air, and acetylene reduction was depressed in anaerobic atmospheres. Nitrogen fixation rates of decomposing leaves from the surface measured by 15N2 uptake ranged from 5.1 to 7.8 nmol of N2 g (dry weight)−1 h−1, and the mean molar ratio of acetylene reduced to nitrogen fixed was 4.5:1. Anaerobic conditions depressed the nitrogenase activity in decomposing leaves, which was independent of light. Nitrogenase activity was also found to be associated with pneumatophores. This activity was light dependent and was probably attributable to one or more species of Calothrix present as an epiphyte. Rates of activity were generally between 100 and 500 nmol of C2H4 pneumatophore−1 h−1 in summer, but values up to 1,500 nmol of C2H4 pneumatophore−1 h−1 were obtained.  相似文献   

3.
Succinate-limited continuous cultures of an Azorhizobium caulinodans strain were grown on ammonia or nitrogen gas as a nitrogen source. Ammonia-grown cells became oxygen limited at 1.7 μM dissolved oxygen, whereas nitrogen-fixing cells remained succinate limited even at dissolved oxygen concentrations as low as 0.9 μM. Nitrogen-fixing cells tolerated dissolved oxygen concentrations as high as 41 μM. Succinate-dependent oxygen uptake rates of cells from the different steady states ranged from 178 to 236 nmol min−1 mg of protein−1 and were not affected by varying chemostat-dissolved oxygen concentration or nitrogen source. When equimolar concentrations of succinate and β-hydroxybutyrate were combined, oxygen uptake rates were greater than when either substrate was used alone. Azide could also used alone as a respiratory substrate regardless of nitrogen source; however, when azide was added following succinate additions, oxygen uptake was inhibited in ammonia-grown cells and stimulated in nitrogen-fixing cells. Use of 25 mM succinate in the chemostat resevoir at a dilution rate of 0.1 h−1 resulted in high levels of background respiration and nitrogenase activity, indicating that the cells were not energy limited. Lowering the reservoir succinate to 5 mM imposed energy limitation. Maximum succinate-dependent nitrogenase activity was 1,741 nmol of C2H4h−1 mg (dry weight)−1, and maximum hydrogen-dependent nitrogenase activity was 949 nmol of C2H4 h−1 mg (dry weight)−1. However, when concentration of 5% (vol/vol) hydrogen or greater were combined with succinate, nitrogenase activity decreased by 35% in comparison to when succinate was used alone. Substitution of argon for nitrogen in the chemostat inflow gas resulted in “washout,” proving that ORS571 can grow on N2 and that there was not a nitrogen source in the medium that could substitute.  相似文献   

4.
Acetate dominated the extracellular pool of volatile fatty acids (VFAs) in the hindgut fluid of Reticulitermes flavipes, Zootermopsis angusticollis, and Incisitermes schwarzi, where it occurred at concentrations of 57.9 to 80.6 mM and accounted for 94 to 98 mol% of all VFAs. Small amounts of C3 to C5 VFAs were also observed. Acetate was also the major VFA in hindgut homogenates of Schedorhinotermes lamanianus, Prorhinotermes simplex, Coptotermes formosanus, and Nasutitermes corniger. Estimates of in situ acetogenesis by the hindgut microbiota of R. flavipes (20.2 to 43.3 nmol · termite−1 · h−1) revealed that this activity could support 77 to 100% of the respiratory requirements of the termite (51.6 to 63.6 nmol of O2 · termite−1 · h−1). This conclusion was buttressed by the demonstration of acetate in R. flavipes hemolymph (at 9.0 to 11.6 mM), but not in feces, and by the ability of termite tissues to readily oxidize acetate to CO2. About 85% of the acetate produced by the hindgut microbiota was derived from cellulose C; the remainder was derived from hemicellulose C. Selective removal of major groups of microbes from the hindgut of R. flavipes indicated that protozoa were primarily responsible for acetogenesis but that bacteria also functioned in this capacity. H2 and CH4 were evolved by R. flavipes (usually about 0.4 nmol · termite−1 · h−1), but these compounds represented a minor fate of electrons derived from wood dissimilation within R. flavipes. A working model is proposed for symbiotic wood polysaccharide degradation in R. flavipes, and the possible roles of individual gut microbes, including CO2-reducing acetogenic bacteria, are discussed.  相似文献   

5.
Removal of the plant hormone ethylene (C2H4) is often required by horticultural storage facilities, which are operated at temperatures below 10°C. The aim of this study was to demonstrate an efficient, biological C2H4 removal under such low-temperature conditions. Peat-soil, acclimated to degradation of C2H4, was packed in a biofilter (687 cm3) and subjected to an airflow (~73 ml min−1) with 2 ppm (μl liter−1) C2H4. The C2H4 removal efficiencies achieved at 20, 10, and 5°C, respectively, were 99.0, 98.8, and 98.4%. This corresponded to C2H4 levels of 0.022 to 0.032 ppm in the biofilter outlet air. At 2°C, the average C2H4 removal efficiency dropped to 83%. The detailed temperature response of C2H4 removal was tested under batch conditions by incubation of 1-g soil samples in a temperature gradient ranging from 0 to 29°C with increments of 1°C. The C2H4 removal rate was highest at 26°C (0.85 μg of C2H4 g [dry weight]−1 h−1), but remained at levels of 0.14 to 0.28 μg of C2H4 g (dry weight)−1 h−1 at 0 to 10°C. At 35 to 40°C, the C2H4 removal rate was negligible (0.02 to 0.06 μg of C2H4 g [dry weight]−1 h−1). The Q10 (i.e., the ratio of rates 10°C apart) for C2H4 removal was 1.9 for the interval 0 to 10°C. In conclusion, the present results demonstrated microbial C2H4 removal, which proceeded at 0 to 2°C and produced a moderately psychrophilic temperature response.  相似文献   

6.
The transformation of 1,2,4-trichlorobenzene (1,2,4-TCB) at initial concentrations in nano- and micromolar ranges was studied in batch experiments with Burkholderia sp. strain PS14. 1,2,4-TCB was metabolized from nano- and micromolar concentrations to below its detection limit of 0.5 nM. At low initial 1,2,4-TCB concentrations, a first-order relationship between specific transformation rate and substrate concentration was observed with a specific affinity (a0A) of 0.32 liter · mg (dry weight)−1 · h−1 followed by a second one at higher concentrations with an aoA of 0.77 liter · mg (dry weight)−1 · h−1. This transition from the first-order kinetics at low initial 1,2,4-TCB concentrations to the second first-order kinetics at higher 1,2,4-TCB concentrations was shifted towards higher initial 1,2,4-TCB concentrations with increasing cell mass. At high initial concentrations of 1,2,4-TCB, a maximal transformation rate of approximately 37 nmol · min−1 · mg (dry weight)−1 was measured, irrespective of the cell concentration.  相似文献   

7.
The specific growth rate is a key control parameter in the industrial production of baker’s yeast. Nevertheless, quantitative data describing its effect on fermentative capacity are not available from the literature. In this study, the effect of the specific growth rate on the physiology and fermentative capacity of an industrial Saccharomyces cerevisiae strain in aerobic, glucose-limited chemostat cultures was investigated. At specific growth rates (dilution rates, D) below 0.28 h−1, glucose metabolism was fully respiratory. Above this dilution rate, respirofermentative metabolism set in, with ethanol production rates of up to 14 mmol of ethanol · g of biomass−1 · h−1 at D = 0.40 h−1. A substantial fermentative capacity (assayed offline as ethanol production rate under anaerobic conditions) was found in cultures in which no ethanol was detectable (D < 0.28 h−1). This fermentative capacity increased with increasing dilution rates, from 10.0 mmol of ethanol · g of dry yeast biomass−1 · h−1 at D = 0.025 h−1 to 20.5 mmol of ethanol · g of dry yeast biomass−1 · h−1 at D = 0.28 h−1. At even higher dilution rates, the fermentative capacity showed only a small further increase, up to 22.0 mmol of ethanol · g of dry yeast biomass−1 · h−1 at D = 0.40 h−1. The activities of all glycolytic enzymes, pyruvate decarboxylase, and alcohol dehydrogenase were determined in cell extracts. Only the in vitro activities of pyruvate decarboxylase and phosphofructokinase showed a clear positive correlation with fermentative capacity. These enzymes are interesting targets for overexpression in attempts to improve the fermentative capacity of aerobic cultures grown at low specific growth rates.  相似文献   

8.
Cut carnations (Dianthus caryophyllus L. cv. `Improved White Sim') were exposed to ultra high purity 14C2H4 (20 μl/1) during flower opening and senescence to study its incorporation and metabolism. During treatment precautions were taken to exclude inhibitory volatiles from rubber serum stoppers which were identified as CS2 and COS. As with the pea seedling (Nature 1975, 255:144-147), cut carnations incorporated 14C2H4 into ethanol-soluble tissue metabolites and oxidized the hormone to 14CO2. Oxidation increased from 0.5 to 3 dpm · mg dry wt−1·6 hr−1 during the period of flower opening and early petal wilt. As severe petal wilt set in, and the ovary increased in size and dry weight, oxidation increased to a peak of nearly 29 dpm · mg dry wt−1·6 hr−1. Concomitant with this peak was a similar rise in the rate of 14C2H4 incorporation into the petals, peduncle, bracts, and sepals. Much higher rates of incorporation were found for the reproductive and receptacle tissues. Incorporation into these tissues steadily increased during flower opening reaching a peak of over 160 dpm · mg dry wt−1 · 6 hr−1 just before full bloom. This peak preceded a peak of endogenous ethylene production while the 14C2H4 oxidation peak followed it.  相似文献   

9.
Pigeon peas (Cajanus cajan) were grown in large soil columns (90-cm length by 30-cm diameter) and inoculated with four different strains of cowpea rhizobia, which varied with respect to hydrogen uptake activity (Hup). Despite the profuse liberation of H2 from Hup- nodules in vitro, H2 gas was not detected in any of the soil columns. When H2 was injected into the columns, the rates of consumption were highest in the treatments (including control) containing Hup- nodules (218 and 177 nmol · h−1 · cm−2) and lowest in the Hup+ treatments (158, 92, and 64 nmoles · h−1 · cm−2). In situ H2 uptake rates in small soil cores at fixed distances from the nodules decreased exponentially with distance from the nodule (R2 = 0.99). This decrease in H2 consumption was associated with a similar decrease in numbers of H2-oxidizing chemolithotrophic bacteria as determined by the most-probable-number method. On the basis of two equations derived separately upon diffusive theory (Fix's Law) and kinetic theory (Michaelis-Menten), the empirically derived rate constants and coefficients indicated that all of the H2 emitted from Hup- nodules would be consumed by H2-oxidizing bacteria within a 3- to 4.5-cm radius of the nodule surface. It is concluded that H2 is not lost from the soil-plant ecosystem during N2 fixation in C. cajan but is conserved by H2-oxidizing bacteria.  相似文献   

10.
The gene cassette (camA+ camB+ camC) encoding a cytochrome P-450cam variant was integrated into the nonessential gene pcpM of the pentachlorophenol degrader Sphingobium chlorophenolicum ATCC 39723 by homologous recombination. The recombinant strain could degrade hexachlorobenzene at a rate of 0.67 nmol · mg (dry weight)−1 · h−1, and intermediate pentachlorophenol was also identified.  相似文献   

11.
There are only a few examples of microbial conversion of picric acid (2,4,6-trinitrophenol). None of the organisms that have been described previously is able to use this compound as a sole source of carbon, nitrogen, and energy at high rates. In this study we isolated and characterized a strain, strain CB 22-2, that was able to use picric acid as a sole source of carbon and energy at concentrations up to 40 mM and at rates of 1.6 mmol · h−1 · g (dry weight) of cells−1 in continuous cultures and 920 μmol · h−1 · g (dry weight) of cells−1 in flasks. In addition, this strain was able to use picric acid as a sole source of nitrogen at comparable rates in a nitrogen-free medium. Biochemical characterization and 16S ribosomal DNA analysis revealed that strain CB 22-2 is a Nocardioides sp. strain. High-pressure liquid chromatography and UV-visible light data, the low residual chemical oxygen demand, and the stoichiometric release of 2.9 ± 0.1 mol of nitrite per mol of picric acid provided strong evidence that complete mineralization of picric acid occurred. During transformation, the metabolites detected in the culture supernatant were the [H]-Meisenheimer complexes of picric acid and 2,4-dinitrophenol (H-DNP), as well as 2,4-dinitrophenol. Experiments performed with crude extracts revealed that H-DNP formation indeed is a physiologically relevant step in picric acid metabolism.  相似文献   

12.
Free-living and surface-associated microbial communities in sand-packed columns perfused with groundwater were compared by examination of compositional and functional characteristics. The composition of the microbial communities was assessed by bulk DNA extraction, PCR amplification of 16S ribosomal DNA fragments, separation of these fragments by denaturing gradient gel electrophoresis, and sequence analysis. Community function was assessed by measurement of β-glucosidase and aminopeptidase extracellular enzyme activities. Free-living populations in the aqueous phase exhibited a greater diversity of phylotypes than populations associated with the solid phase. The attached bacterial community displayed significantly greater β-glucosidase and aminopeptidase enzyme activities per volume of porous medium than those of the free-living community. On a per-cell basis, the attached community had a significantly higher cell-specific aminopeptidase enzyme activity (1.07 × 10−7 nmol cell−1 h−1) than the free-living community (5.02 × 10−8 nmol cell−1 h−1). Conversely, the free-living community had a significantly higher cell-specific β-glucosidase activity (1.92 × 10−6 nmol cell−1 h−1) than the surface-associated community (6.08 × 10−7 nmol cell−1 h−1). The compositional and functional differences observed between these two communities may reflect different roles for these distinct but interacting communities in the decomposition of natural organic matter or biodegradation of xenobiotics in aquifers.  相似文献   

13.
Degradation of dimethyl sulfide and methanethiol in slurries prepared from sediments of minerotrophic peatland ditches were studied under various conditions. Maximal aerobic dimethyl sulfide-degrading capacities (4.95 nmol per ml of sediment slurry · h−1), measured in bottles shaken under an air atmosphere, were 10-fold higher than the maximal anaerobic degrading capacities determined from bottles shaken under N2 or H2 atmosphere (0.37 and 0.32 nmol per ml of sediment slurry · h−1, respectively). Incubations under experimental conditions which mimic the in situ conditions (i.e., not shaken and with an air headspace), however, revealed that aerobic degradation of dimethyl sulfide and methanethiol in freshwater sediments is low due to oxygen limitation. Inhibition studies with bromoethanesulfonic acid and sodium tungstate demonstrated that the degradation of dimethyl sulfide and methanethiol in these incubations originated mainly from methanogenic activity. Prolonged incubation under a H2 atmosphere resulted in lower dimethyl sulfide degradation rates. Kinetic analysis of the data resulted in apparent Km values (6 to 8 μM) for aerobic dimethyl sulfide degradation which are comparable to those reported for Thiobacillus spp., Hyphomicrobium spp., and other methylotrophs. Apparent Km values determined for anaerobic degradation of dimethyl sulfide (3 to 8 μM) were of the same order of magnitude. The low apparent Km values obtained explain the low dimethyl sulfide and methanethiol concentrations in freshwater sediments that we reported previously. Our observations point to methanogenesis as the major mechanism of dimethyl sulfide and methanethiol consumption in freshwater sediments.  相似文献   

14.
Bacterial biomass, metabolic condition, and activity were measured over a 16-month period in the surface sediments of the following four field sites with differing dissolved organic matter regimes: a woodlot spring seep, a meadow spring seep, a second-order stream, and a third-order stream. Total bacterial biomass was measured by lipid phosphate and epifluorescence microscopic counts (EMC), and viable biomass was measured by 14C most probable number, EMC with 2-(p-iodophenyl)-3-(p-nitrophenyl)-5-phenyl tetrazolium chloride reduction, and ATP. Bacterial metabolic condition was determined from the percentage of respiring cells, poly-β-hydroxybutyrate concentrations, and adenylate energy charge. Activity measures included 14C-lipid synthesis, 32P-phospholipid synthesis, the rate of uptake of algal lysate dissolved organic carbon, and respiration, from which biosynthesis was calculated (dissolved organic carbon uptake corrected for respiration). Total bacterial biomass (from EMC) ranged from 0.012 to 0.354 μg of C/mg of dry sediment and was usually lowest in the third-order stream. The percentage of cells respiring was less than 25% at all sites, indicating that most bacteria were dormant or dead. Adenylate energy charge was measured only in the third-order stream and was uniformly low. Poly-β-hydroxybutyrate concentrations were greater in the woodlot spring seep than in the second- and third-order streams. Uptake of algal lysate dissolved organic carbon ranged from undetectable levels to 166 mg of C · m−2 · h−1. Little community respiration could be attributed to algal lysate metabolism. Phospholipid synthesis ranged from 0.006 to 0.354 pmol · mg of dry sediment−1 · h−1. Phospholipid synthesis rates were used to estimate bacterial turnover at the study sites. An estimated 375 bacterial generations per year were produced in the woodlot spring seep, and 67 per year were produced in the third-order stream.  相似文献   

15.
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   

16.
Most probable numbers (MPNs) of methanogens in various salt marsh and estuarine sediments were determined with an anaerobic, habitat-simulating culture medium with 80% H2 plus 20% CO2 as substrate. Average MPNs for the short Spartina (SS) marsh sediments of Sapelo Island, Ga., were maximal at the 5- to 7-cm depth (1.2 × 107/g of dry sediment). Populations decreased to approximately 880/g of dry sediment at the 34- to 36-cm depth. There was no significant difference between summer and winter populations. In tall Spartina (TS) marsh sediments, average populations were maximal (1.2 × 106/g of dry sediment) in the upper 0- to 2-cm zone; populations from the 5- to 36-cm zones were similar (average of 9 × 104/g of dry sediment). Methanogenic populations for TS sediments of James Island Creek marsh, Charleston, S.C., were similar (average of 3 × 106/g of dry sediment) for all depths tested (0 to 22 cm), which was comparable to the trend observed for TS sediments at Sapelo Island, Ga. Sediment grab samples collected along a transect of James Island Creek and its adjacent Spartina marsh had MPNs that were approximately 20 times greater for the region of Spartina growth (average of 106/g of dry sediment) compared with the channel (approximately 5 × 104 methanogens per g of dry sediment). A similar trend was found at Pawley's Island marsh, S.C., but populations were approximately one order of magnitude lower. In vitro rates of methanogenesis with SS sediments incubated under 80% H2-20% CO2 showed that the 5- to 7-cm region exhibited maximal activity (51 nmol of CH4 g−1 h−1), which was greater than rates for sediments above and below this depth. SS sediment samples (5 to 7 cm) incubated under 100% N2 and supplemented with formate exhibited rates of methanogenesis similar to those generated by samples under 80% H2-20% CO2. Replacing the N2 atmosphere with H2 resulted in an eightfold decrease in the rate of methanogenesis. In vitro methanogenic activity by TS salt marsh sediments, incubated under 80% H2-20% CO2, was similar for all depths tested (0 to 22 cm). TS sediment samples (0 to 7 cm) supplemented with formate and incubated under 100% N2 had greater rates of methanogenesis compared with unsupplemented samples.  相似文献   

17.
Formate-dependent proton reduction to H2 (HCOO + H2O → HCO3 + H2) has been reported for hyperthermophilic Thermococcus strains. In this study, a hyperthermophilic archaeon, Thermococcus onnurineus strain NA1, yielded H2 accumulation to a partial pressure of 1 × 105 to 7 × 105 Pa until the values of Gibbs free energy change (ΔG) reached near thermodynamic equilibrium (−1 to −3 kJ mol−1). The bioenergetic requirement for the metabolism to conserve energy was demonstrated by ΔG values as small as −5 kJ mol−1, which are less than the biological minimum energy quantum, −20 kJ mol−1, as calculated by Schink (B. Schink, Microbiol. Mol. Biol. Rev. 61:262-280, 1997). Considering formate as a possible H2 storage material, the H2 production potential of the strain was assessed. The volumetric H2 production rate increased linearly with increasing cell density, leading to 2,820 mmol liter−1 h−1 at an optical density at 600 nm (OD600) of 18.6, and resulted in the high specific H2 production rates of 404 ± 6 mmol g−1 h−1. The H2 productivity indicates the great potential of T. onnurineus strain NA1 for practical application in comparison with H2-producing microbes. Our result demonstrates that T. onnurineus strain NA1 has a highly efficient metabolic system to thrive on formate in hydrothermal systems.  相似文献   

18.
Trichloroethylene (TCE) was removed from soils by using a wheat rhizosphere established by coating seeds with a recombinant, TCE-degrading Pseudomonas fluorescens strain that expresses the tomA+ (toluene o-monooxygenase) genes from Burkholderia cepacia PR123(TOM23C). A transposon integration vector was used to insert tomA+ into the chromosome of P. fluorescens 2-79, producing a stable strain that expressed constitutively the monooxygenase at a level of 1.1 nmol/min · mg of protein (initial TCE concentration, 10 μM, assuming that all of the TCE was in the liquid) for more than 280 cell generations (36 days). We also constructed a salicylate-inducible P. fluorescens strain that degraded TCE at an initial rate of 2.6 nmol/min · mg of protein in the presence of 10 μM TCE [cf. B. cepacia G4 PR123(TOM23C), which degraded TCE at an initial rate of 2.5 nmol/min · mg of protein]. A constitutive strain, P. fluorescens 2-79TOM, grew (maximum specific growth rate, 0.78 h−1) and colonized wheat (3 × 106 CFU/cm of root) as well as wild-type P. fluorescens 2-79 (maximum specific growth rate, 0.77 h−1; level of colonization, 4 × 106 CFU/cm of root). Rhizoremediation of TCE was demonstrated by using microcosms containing the constitutive monooxygenase-expressing microorganism, soil, and wheat. These closed microcosms degraded an average of 63% of the initial TCE in 4 days (20.6 nmol of TCE/day · plant), compared to the 9% of the initial TCE removed by negative controls consisting of microcosms containing wild-type P. fluorescens 2-79-inoculated wheat, uninoculated wheat, or sterile soil.  相似文献   

19.
In the intestinal tracts of animals, methanogenesis from CO2 and other C1 compounds strictly depends on the supply of electron donors by fermenting bacteria, but sources and sinks of reducing equivalents may be spatially separated. Microsensor measurements in the intestinal tract of the omnivorous cockroach Blaberus sp. showed that molecular hydrogen strongly accumulated in the midgut (H2 partial pressures of 3 to 26 kPa), whereas it was not detectable (<0.1 kPa) in the posterior hindgut. Moreover, living cockroaches emitted large quantities of CH4 [105 ± 49 nmol (g of cockroach)−1 h−1] but only traces of H2. In vitro incubation of isolated gut compartments, however, revealed that the midguts produced considerable amounts of H2, whereas hindguts emitted only CH4 [106 ± 58 and 71 ± 50 nmol (g of cockroach)−1 h−1, respectively]. When ligated midgut and hindgut segments were incubated in the same vials, methane emission increased by 28% over that of isolated hindguts, whereas only traces of H2 accumulated in the headspace. Radial hydrogen profiles obtained under air enriched with H2 (20 kPa) identified the hindgut as an efficient sink for externally supplied H2. A cross-epithelial transfer of hydrogen from the midgut to the hindgut compartment was clearly evidenced by the steep H2 concentration gradients which developed when ligated fragments of midgut and hindgut were placed on top of each other—a configuration that simulates the situation in vivo. These findings emphasize that it is essential to analyze the compartmentalization of the gut and the spatial organization of its microbiota in order to understand the functional interactions among different microbial populations during digestion.  相似文献   

20.
The kinetic parameters Km, Vmax, Tt (turnover time), and v (natural velocity) were determined for H2 and acetate conversion to methane by Wintergreen Lake sediment, using short-term (a few hours) methods and incubation temperatures of 10 to 14°C. Estimates of the Michaelis-Menten constant, Km, for both the consumption of hydrogen and the conversion of hydrogen to methane by sediment microflora averaged about 0.024 μmol g−1 of dry sediment. The maximal velocity, Vmax, averaged 4.8 μmol of H2 g−1 h−1 for hydrogen consumption and 0.64 μmol of CH4 g−1 h−1 for the conversion of hydrogen to methane during the winter. Estimated natural rates of hydrogen consumption and hydrogen conversion to methane could be calculated from the Michaelis-Menten equation and estimates of Km, Vmax, and the in situ dissolved-hydrogen concentration. These results indicate that methane may not be the only fate of hydrogen in the sediment. Among several potential hydrogen donors tested, only formate stimulated the rate of sediment methanogenesis. Formate conversion to methane was so rapid that an accurate estimate of kinetic parameters was not possible. Kinetic experiments using [2-14C]acetate and sediments collected in the summer indicated that acetate was being converted to methane at or near the maximal rate. A minimum natural rate of acetate conversion to methane was estimated to be about 110 nmol of CH4 g−1 h−1, which was 66% of the Vmax (163 nmol of CH4 g−1 h−1). A 15-min preincubation of sediment with 5.0 × 10−3 atm of hydrogen had a pronounced effect on the kinetic parameters for the conversion of acetate to methane. The acetate pool size, expressed as the term Km + Sn (Sn is in situ substrate concentration), decreased by 37% and Tt decreased by 43%. The Vmax remained relatively constant. A preincubation with hydrogen also caused a 37% decrease in the amount of labeled carbon dioxide produced from the metabolism of [U-14C]valine by sediment heterotrophs.  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