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 46 毫秒
1.
Although previous research has demonstrated that NO3 inhibits microbial Fe(III) reduction in laboratory cultures and natural sediments, the mechanisms of this inhibition have not been fully studied in an environmentally relevant medium that utilizes solid-phase, iron oxide minerals as a Fe(III) source. To study the dynamics of Fe and NO3 biogeochemistry when ferric (hydr)oxides are used as the Fe(III) source, Shewanella putrefaciens 200 was incubated under anoxic conditions in a low-ionic-strength, artificial groundwater medium with various amounts of NO3 and synthetic, high-surface-area goethite. Results showed that the presence of NO3 inhibited microbial goethite reduction more severely than it inhibited microbial reduction of the aqueous or microcrystalline sources of Fe(III) used in other studies. More interestingly, the presence of goethite also resulted in a twofold decrease in the rate of NO3 reduction, a 10-fold decrease in the rate of NO2 reduction, and a 20-fold increase in the amounts of N2O produced. Nitrogen stable isotope experiments that utilized δ15N values of N2O to distinguish between chemical and biological reduction of NO2 revealed that the N2O produced during NO2 or NO3 reduction in the presence of goethite was primarily of abiotic origin. These results indicate that concomitant microbial Fe(III) and NO3 reduction produces NO2 and Fe(II), which then abiotically react to reduce NO2 to N2O with the subsequent oxidation of Fe(II) to Fe(III).  相似文献   

2.
A lithotrophic freshwater Beggiatoa strain was enriched in O2-H2S gradient tubes to investigate its ability to oxidize sulfide with NO3 as an alternative electron acceptor. The gradient tubes contained different NO3 concentrations, and the chemotactic response of the Beggiatoa mats was observed. The effects of the Beggiatoa sp. on vertical gradients of O2, H2S, pH, and NO3 were determined with microsensors. The more NO3 that was added to the agar, the deeper the Beggiatoa filaments glided into anoxic agar layers, suggesting that the Beggiatoa sp. used NO3 to oxidize sulfide at depths below the depth that O2 penetrated. In the presence of NO3 Beggiatoa formed thick mats (>8 mm), compared to the thin mats (ca. 0.4 mm) that were formed when no NO3 was added. These thick mats spatially separated O2 and sulfide but not NO3 and sulfide, and therefore NO3 must have served as the electron acceptor for sulfide oxidation. This interpretation is consistent with a fourfold-lower O2 flux and a twofold-higher sulfide flux into the NO3-exposed mats compared to the fluxes for controls without NO3. Additionally, a pronounced pH maximum was observed within the Beggiatoa mat; such a pH maximum is known to occur when sulfide is oxidized to S0 with NO3 as the electron acceptor.  相似文献   

3.
Cystathionine β-synthase (CBS) is a pyridoxal phosphate-dependent enzyme that catalyzes the condensation of homocysteine with serine or with cysteine to form cystathionine and either water or hydrogen sulfide, respectively. Human CBS possesses a noncatalytic heme cofactor with cysteine and histidine as ligands, which in its oxidized state is relatively unreactive. Ferric CBS (Fe(III)-CBS) can be reduced by strong chemical and biochemical reductants to Fe(II)-CBS, which can bind carbon monoxide (CO) or nitric oxide (NO), leading to inactive enzyme. Alternatively, Fe(II)-CBS can be reoxidized by O2 to Fe(III)-CBS, forming superoxide radical anion (O2˙̄). In this study, we describe the kinetics of nitrite (NO2) reduction by Fe(II)-CBS to form Fe(II)NO-CBS. The second order rate constant for the reaction of Fe(II)-CBS with nitrite was obtained at low dithionite concentrations. Reoxidation of Fe(II)NO-CBS by O2 showed complex kinetic behavior and led to peroxynitrite (ONOO) formation, which was detected using the fluorescent probe, coumarin boronic acid. Thus, in addition to being a potential source of superoxide radical, CBS constitutes a previously unrecognized source of NO and peroxynitrite.  相似文献   

4.
Experiments demonstrated that Beggiatoa could induce a H2S-depleted suboxic zone of more than 10 mm in marine sediments and cause a divergence in sediment NO3 reduction from denitrification to dissimilatory NO3 reduction to ammonium. pH, O2, and H2S profiles indicated that the bacteria oxidized H2S with NO3 and transported S0 to the sediment surface for aerobic oxidation.  相似文献   

5.
BioDeNOx is an integrated physicochemical and biological process for the removal of nitrogen oxides (NOx) from flue gases. In this process, the flue gas is purged through a scrubber containing a solution of Fe(II)EDTA2−, which binds the NOx to form an Fe(II)EDTA·NO2− complex. Subsequently, this complex is reduced in the bioreactor to dinitrogen by microbial denitrification. Fe(II)EDTA2−, which is oxidized to Fe(III)EDTA by oxygen in the flue gas, is regenerated by microbial iron reduction. In this study, the microbial communities of both lab- and pilot-scale reactors were studied using culture-dependent and -independent approaches. A pure bacterial strain, KT-1, closely affiliated by 16S rRNA analysis to the gram-positive denitrifying bacterium Bacillus azotoformans, was obtained. DNA-DNA homology of the isolate with the type strain was 89%, indicating that strain KT-1 belongs to the species B. azotoformans. Strain KT-1 reduces Fe(II)EDTA·NO2− complex to N2 using ethanol, acetate, and Fe(II)EDTA2− as electron donors. It does not reduce Fe(III)EDTA. Denaturing gradient gel electrophoresis analysis of PCR-amplified 16S rRNA gene fragments showed the presence of bacteria closely affiliated with members of the phylum Deferribacteres, an Fe(III)-reducing group of bacteria. Fluorescent in situ hybridization with oligonucleotide probes designed for strain KT-1 and members of the phylum Deferribacteres showed that the latter were more dominant in both reactors.  相似文献   

6.
We screened soybean rhizobia originating from three germplasm collections for the ability to grow anaerobically in the presence of NO3 and for differences in final product formation from anaerobic NO3 metabolism. Denitrification abilities of selected strains as free-living bacteria and as bacteroids were compared. Anaerobic growth in the presence of NO3 was observed in 270 of 321 strains of soybean rhizobia. All strains belonging to the 135 serogroup did not grow anaerobically in the presence of NO3. An investigation with several strains indicated that bacteria not growing anaerobically in the presence of NO3 also did not utilize NO3 as the sole N source aerobically. An exception was strain USDA 33, which grew on NO3 but failed to denitrify. Dissimilation of NO3 by the free-living cultures proceeded without the significant release of intermediate products. Nitrous oxide reductase was inhibited by C2H2, but preceding steps of denitrification were not affected. Final products of denitrification were NO2, N2O, or N2; serogroups 31, 46, 76, and 94 predominantly liberated NO2, whereas evolution of N2 was prevalent in serogroups 110 and 122, and all three were formed as final products by strains belonging to serogroups 6 and 123. Anaerobic metabolism of NO3 by bacteroid preparations of Bradyrhizobium japonicum proceeded without delay and was evident by NO2 accumulation irrespective of which final product was formed by the strain as free-living bacteria. Anaerobic C2H2 reduction in the presence of NO3 was observed in bacteroid preparations capable of NO3 respiration but was absent in bacteria that were determined to be deficient in dissimilatory nitrate reductase.  相似文献   

7.
A recent study (D. C. Cooper, F. W. Picardal, A. Schimmelmann, and A. J. Coby, Appl. Environ. Microbiol. 69:3517-3525, 2003) has shown that NO3 and NO2 (NOx) reduction by Shewanella putrefaciens 200 is inhibited in the presence of goethite. The hypothetical mechanism offered to explain this finding involved the formation of a Fe(III) (hydr)oxide coating on the cell via the surface-catalyzed, abiotic reaction between Fe2+ and NO2. This coating could then inhibit reduction of NOx by physically blocking transport into the cell. Although the data in the previous study were consistent with such an explanation, the hypothesis was largely speculative. In the current work, this hypothesis was tested and its environmental significance explored through a number of experiments. The inhibition of ~3 mM NO3 reduction was observed during reduction of a variety of Fe(III) (hydr)oxides, including goethite, hematite, and an iron-bearing, natural sediment. Inhibition of oxygen and fumarate reduction was observed following treatment of cells with Fe2+ and NO2, demonstrating that utilization of other soluble electron acceptors could also be inhibited. Previous adsorption of Fe2+ onto Paracoccus denitrificans inhibited NOx reduction, showing that Fe(II) can reduce rates of soluble electron acceptor utilization by non-iron-reducing bacteria. NO2 was chemically reduced to N2O by goethite or cell-sorbed Fe2+, but not at appreciable rates by aqueous Fe2+. Transmission and scanning electron microscopy showed an electron-dense, Fe-enriched coating on cells treated with Fe2+ and NO2. The formation and effects of such coatings underscore the complexity of the biogeochemical reactions that occur in the subsurface.  相似文献   

8.
Microbial communities have the potential to control the biogeochemical fate of some radionuclides in contaminated land scenarios or in the vicinity of a geological repository for radioactive waste. However, there have been few studies of ionizing radiation effects on microbial communities in sediment systems. Here, acetate and lactate amended sediment microcosms irradiated with gamma radiation at 0.5 or 30 Gy h−1 for 8 weeks all displayed NO3 and Fe(III) reduction, although the rate of Fe(III) reduction was decreased in 30-Gy h−1 treatments. These systems were dominated by fermentation processes. Pyrosequencing indicated that the 30-Gy h−1 treatment resulted in a community dominated by two Clostridial species. In systems containing no added electron donor, irradiation at either dose rate did not restrict NO3, Fe(III), or SO42− reduction. Rather, Fe(III) reduction was stimulated in the 0.5-Gy h−1-treated systems. In irradiated systems, there was a relative increase in the proportion of bacteria capable of Fe(III) reduction, with Geothrix fermentans and Geobacter sp. identified in the 0.5-Gy h−1 and 30-Gy h−1 treatments, respectively. These results indicate that biogeochemical processes will likely not be restricted by dose rates in such environments, and electron accepting processes may even be stimulated by radiation.  相似文献   

9.
The contribution of the biochemical pathways nitrification, denitrification, and dissimilatory NO3 reduction to NH4+ (DNRA) to the accumulation of NO2 in freshwaters is governed by the species compositions of the bacterial populations resident in the sediments, available carbon (C) and nitrogen (N) substrates, and environmental conditions. Recent studies of major rivers in Northern Ireland have shown that high NO2 concentrations found in summer, under warm, slow-flowing conditions, arise from anaerobic NO3 reduction. Locally, agricultural pollutants entering rivers are important C and N sources, providing ideal substrates for the aquatic bacteria involved in cycling of N. In this study a range of organic C compounds commonly found in agricultural pollutants were provided as energy sources in 48-h incubation experiments to investigate if the chemical compositions of the pollutants affected which NO3 reduction pathway was followed and influenced subsequent NO2 accumulation. Carbon stored within the sediments was sufficient to support DNRA and denitrifier populations, and the resulting NO2 peak (80 μg of N liter−1 [approximate]) observed at 24 h was indicative of the simultaneous activities of both bacterial groups. The value of glycine as an energy source for denitrification or DNRA appeared to be limited, but glycine was an important source of additional N. Glucose was an efficient substrate for both the denitrification and DNRA pathways, with a NO2 peak of 160 μg of N liter−1 noted at 24 h. Addition of formate and acetate stimulated continuous NO2 production throughout the 48-h period, caused by partial inhibition of the denitrification pathway. The formate treatment resulted in a high NO2 accumulation (1,300 μg of N liter−1 [approximate]), and acetate treatment resulted in a low NO2 concentration (<100 μg of N liter−1).  相似文献   

10.
Role of Chemotaxis in the Ecology of Denitrifiers   总被引:4,自引:2,他引:2       下载免费PDF全文
A modification of the Adler capillary assay was used to evaluate the chemotactic responses of several denitrifiers to nitrate and nitrite. Strong positive chemotaxis was observed to NO3 and NO2 by soil isolates of Pseudomonas aeruginosa, Pseudomonas fluorescens, and Pseudomonas stutzeri, with the peak response occurring at 10−3 M for both attractants. In addition, a strong chemoattraction to serine (peak response at 10−2 M), tryptone, and a soil extract, but not to NH4+, was observed for all denitrifiers tested. Chemotaxis was not dependent on a previous growth on NO3, NO2, or a soil extract, and the chemoattraction to NO3 occurred when the bacteria were grown aerobically or anaerobically. However, the best response to NO3 was usually observed when the cells were grown aerobically with 10 mM NO3 in the growth medium. Capillary tubes containing 103 M NO3 submerged into soil-water mixtures elicited a significant chemotactic response to NO3 by the indigenous soil microflora, the majority of which were Pseudomonas spp. A chemotactic strain of P. fluorescens also was shown to survive significantly better in aerobic and anaerobic soils than was a nonmotile strain of the same species. Both strains had equal growth rates in liquid cultures. Thus, chemotaxis may be one mechanism by which denitrifiers successfully compete for available NO3 and NO2, and which may facilitate the survival of naturally occurring populations of some denitrifiers.  相似文献   

11.
The role of NO3 and NO2 in the induction of nitrite reductase (NiR) activity in detached leaves of 8-day-old barley (Hordeum vulgare L.) seedlings was investigated. Barley leaves contained 6 to 8 micromoles NO2/gram fresh weight × hour of endogenous NiR activity when grown in N-free solutions. Supply of both NO2 and NO3 induced the enzyme activity above the endogenous levels (5 and 10 times, respectively at 10 millimolar NO2 and NO3 over a 24 hour period). In NO3-supplied leaves, NiR induction occurred at an ambient NO3 concentration of as low as 0.05 millimolar; however, no NiR induction was found in leaves supplied with NO2 until the ambient NO2 concentration was 0.5 millimolar. Nitrate accumulated in NO2-fed leaves. The amount of NO3 accumulating in NO2-fed leaves induced similar levels of NiR as did equivalent amounts of NO3 accumulating in NO3-fed leaves. Induction of NiR in NO2-fed leaves was not seen until NO3 was detectable (30 nanomoles/gram fresh weight) in the leaves. The internal concentrations of NO3, irrespective of N source, were highly correlated with the levels of NiR induced. When the reduction of NO3 to NO2 was inhibited by WO42−, the induction of NiR was inhibited only partially. The results indicate that in barley leaves NiR is induced by NO3 directly, i.e. without being reduced to NO2, and that absorbed NO2 induces the enzyme activity indirectly after being oxidized to NO3 within the leaf.  相似文献   

12.
Studies were carried out to elucidate the nature and importance of Fe3+ reduction in anaerobic slurries of marine surface sediment. A constant accumulation of Fe2+ took place immediately after the endogenous NO3 was depleted. Pasteurized controls showed no activity of Fe3+ reduction. Additions of 0.2 mM NO3 and NO2 to the active slurries arrested the Fe3+ reduction, and the process was resumed only after a depletion of the added compounds. Extended, initial aeration of the sediment did not affect the capacity for reduction of NO3 and Fe3+, but the treatments with NO3 increased the capacity for Fe3+ reduction. Addition of 20 mM MoO42− completely inhibited the SO42− reduction, but did not affect the reduction of Fe3+. The process of Fe3+ reduction was most likely associated with the activity of facultative anaerobic, NO3-reducing bacteria. In surface sediment, the bulk of the Fe3+ reduction may be microbial, and the process may be important for mineralization in situ if the availability of NO3 is low.  相似文献   

13.
Mining-impacted sediments of Lake Coeur d'Alene, Idaho, contain more than 10% metals on a dry weight basis, approximately 80% of which is iron. Since iron (hydr)oxides adsorb toxic, ore-associated elements, such as arsenic, iron (hydr)oxide reduction may in part control the mobility and bioavailability of these elements. Geochemical and microbiological data were collected to examine the ecological role of dissimilatory Fe(III)-reducing bacteria in this habitat. The concentration of mild-acid-extractable Fe(II) increased with sediment depth up to 50 g kg−1, suggesting that iron reduction has occurred recently. The maximum concentrations of dissolved Fe(II) in interstitial water (41 mg liter−1) occurred 10 to 15 cm beneath the sediment-water interface, suggesting that sulfidogenesis may not be the predominant terminal electron-accepting process in this environment and that dissolved Fe(II) arises from biological reductive dissolution of iron (hydr)oxides. The concentration of sedimentary magnetite (Fe3O4), a common product of bacterial Fe(III) hydroxide reduction, was as much as 15.5 g kg−1. Most-probable-number enrichment cultures revealed that the mean density of Fe(III)-reducing bacteria was 8.3 × 105 cells g (dry weight) of sediment−1. Two new strains of dissimilatory Fe(III)-reducing bacteria were isolated from surface sediments. Collectively, the results of this study support the hypothesis that dissimilatory reduction of iron has been and continues to be an important biogeochemical process in the environment examined.  相似文献   

14.
We examined nitrate-dependent Fe2+ oxidation mediated by anaerobic ammonium oxidation (anammox) bacteria. Enrichment cultures of “Candidatus Brocadia sinica” anaerobically oxidized Fe2+ and reduced NO3 to nitrogen gas at rates of 3.7 ± 0.2 and 1.3 ± 0.1 (mean ± standard deviation [SD]) nmol mg protein−1 min−1, respectively (37°C and pH 7.3). This nitrate reduction rate is an order of magnitude lower than the anammox activity of “Ca. Brocadia sinica” (10 to 75 nmol NH4+ mg protein−1 min−1). A 15N tracer experiment demonstrated that coupling of nitrate-dependent Fe2+ oxidation and the anammox reaction was responsible for producing nitrogen gas from NO3 by “Ca. Brocadia sinica.” The activities of nitrate-dependent Fe2+ oxidation were dependent on temperature and pH, and the highest activities were seen at temperatures of 30 to 45°C and pHs ranging from 5.9 to 9.8. The mean half-saturation constant for NO3 ± SD of “Ca. Brocadia sinica” was determined to be 51 ± 21 μM. Nitrate-dependent Fe2+ oxidation was further demonstrated by another anammox bacterium, “Candidatus Scalindua sp.,” whose rates of Fe2+ oxidation and NO3 reduction were 4.7 ± 0.59 and 1.45 ± 0.05 nmol mg protein−1 min−1, respectively (20°C and pH 7.3). Co-occurrence of nitrate-dependent Fe2+ oxidation and the anammox reaction decreased the molar ratios of consumed NO2 to consumed NH4+ (ΔNO2/ΔNH4+) and produced NO3 to consumed NH4+ (ΔNO3/ΔNH4+). These reactions are preferable to the application of anammox processes for wastewater treatment.  相似文献   

15.
The effects of several photosynthetic inhibitors and uncouplers of oxidative phosphorylation on NO3 and NO2 assimilation were studied using detached barley (Hordeum vulgare L. cv Numar) leaves in which only endogenous NO3 or NO2 were available for reduction. Uncouplers of oxidative phosphorylation greatly increased NO3 reduction in both light and darkness, while photosynthetic inhibitors did not.

The NO2 concentration in the control leaves was very low in both light and darkness; 98% or more of the NO2 formed from NO3 was further assimilated in control leaves. More NO2 accumulated in the leaves in light and darkness in the presence of photosynthetic inhibitors. Of this NO2, 94% or more was further assimilated. It appears that metabolites, either external or internal to the chloroplast, capable of reducing NADP (which, in turn, could reduce ferredoxin via NADP reductase) might support NO2 reduction in darkness and light when photosynthetic electron flow is inhibited by photosynthetic inhibitors.

Nitrite assimilation was much more sensitive to uncouplers in darkness than in light: in darkness, 74% or more of NO2 formed from NO3 was further assimilated, whereas in light, 95% or more of the NO2 was further assimilated.

  相似文献   

16.
Nitrite oxidation is the second step of nitrification. It is the primary source of oceanic nitrate, the predominant form of bioavailable nitrogen in the ocean. Despite its obvious importance, nitrite oxidation has rarely been investigated in marine settings. We determined nitrite oxidation rates directly in 15N-incubation experiments and compared the rates with those of nitrate reduction to nitrite, ammonia oxidation, anammox, denitrification, as well as dissimilatory nitrate/nitrite reduction to ammonium in the Namibian oxygen minimum zone (OMZ). Nitrite oxidation (⩽372 nM NO2 d−1) was detected throughout the OMZ even when in situ oxygen concentrations were low to non-detectable. Nitrite oxidation rates often exceeded ammonia oxidation rates, whereas nitrate reduction served as an alternative and significant source of nitrite. Nitrite oxidation and anammox co-occurred in these oxygen-deficient waters, suggesting that nitrite-oxidizing bacteria (NOB) likely compete with anammox bacteria for nitrite when substrate availability became low. Among all of the known NOB genera targeted via catalyzed reporter deposition fluorescence in situ hybridization, only Nitrospina and Nitrococcus were detectable in the Namibian OMZ samples investigated. These NOB were abundant throughout the OMZ and contributed up to ∼9% of total microbial community. Our combined results reveal that a considerable fraction of the recently recycled nitrogen or reduced NO3 was re-oxidized back to NO3 via nitrite oxidation, instead of being lost from the system through the anammox or denitrification pathways.  相似文献   

17.
Steady state cultures of Anabaena flos-aquae were established over a wide range of phosphate-limited growth rates while N was supplied as either NH3, NO3, or N2 gas. At growth rates greater than 0.03 per hour, rates of gross and net carbon fixation were similar on all N sources. However, at lower growth rates (<0.03 per hour) in the NO3 and N2 cultures, gross photosynthesis greatly exceeded net photosynthesis. The increase in photosynthetic O2 evolution with growth rate was greatest when N requirements were met by NO3 and least when met by NH3. These results were combined with previously reported measurements of cellular chemical composition, N assimilation, and acetylene reduction (Layzell, Turpin, Elrifi 1985 Plant Physiol 78: 739-745) to construct empirical models of carbon and energy flow for cultures grown at 30, 60, and 100% of their maximal growth rate on all N sources. The models suggested that over this growth range, 89 to 100% of photodriven electrons were allocated to biomass production in the NH3 cells, whereas only 49 to 74% and 54 to 90% were partitioned to biomass in the NO3-and N2-grown cells, respectively. The models were used to estimate the relative contribution of active, maintenance, and establishment costs associated with NO3 and N2 assimilation over the entire range of growth rates. The models showed that the relative contribution of the component costs of N assimilation were growth rate dependent. At higher growth rates, the major costs for NO3 assimilation were the active costs, while in N2-fixing cultures the major energetic requirements were those associated with heterocyst establishment and maintenance. It was concluded that compared with NO3 assimilation, N2 fixation was energetically unfavorable due to the costs of heterocyst establishment and maintenance, rather than the active costs of N2 assimilation.  相似文献   

18.
Filamentous sulfur bacteria of the genus Thioploca occur as dense mats on the continental shelf off the coast of Chile and Peru. Since little is known about their nitrogen, sulfur, and carbon metabolism, this study was undertaken to investigate their (eco)physiology. Thioploca is able to store internally high concentrations of sulfur globules and nitrate. It has been previously hypothesized that these large vacuolated bacteria can oxidize sulfide by reducing their internally stored nitrate. We examined this nitrate reduction by incubation experiments of washed Thioploca sheaths with trichomes in combination with 15N compounds and mass spectrometry and found that these Thioploca samples produce ammonium at a rate of 1 nmol min−1 mg of protein−1. Controls showed no significant activity. Sulfate was shown to be the end product of sulfide oxidation and was observed at a rate of 2 to 3 nmol min−1 mg of protein−1. The ammonium and sulfate production rates were not influenced by the addition of sulfide, suggesting that sulfide is first oxidized to elemental sulfur, and in a second independent step elemental sulfur is oxidized to sulfate. The average sulfide oxidation rate measured was 5 nmol min−1 mg of protein−1 and could be increased to 10.7 nmol min−1 mg of protein−1 after the trichomes were starved for 45 h. Incorporation of 14CO2 was at a rate of 0.4 to 0.8 nmol min−1 mg of protein−1, which is half the rate calculated from sulfide oxidation. [2-14C]acetate incorporation was 0.4 nmol min−1 mg of protein−1, which is equal to the CO2 fixation rate, and no 14CO2 production was detected. These results suggest that Thioploca species are facultative chemolithoautotrophs capable of mixotrophic growth. Microautoradiography confirmed that Thioploca cells assimilated the majority of the radiocarbon from [2-14C]acetate, with only a minor contribution by epibiontic bacteria present in the samples.  相似文献   

19.
Phosphate-limited chemostat cultures were used to study cell growth and N assimilation in Anabaena flos-aquae under various N sources to determine the relative energetic costs associated with the assimilation of NH3, NO3, or N2. Expressed as a function of relative growth rate, steady state cellular P contents and PO4 assimilation rates did not vary with N-source. However, N-source did alter the maximal PO4-limited growth rate achieved by the cultures: the NO3 and N2 cultures attained only 97 and 80%, respectively, of the maximal growth rate of the NH3 grown cells. Cellular biomass and C contents did not vary with growth rate, but changed with N source. The NO3-grown cells were the smallest (627 ± 34 micromoles C · 10−9 cells), while NH3-grown cells were largest (900 ± 44 micromoles C · 10−9 cells) and N2-fixing cells were intermediate (726 ± 48 micromoles C · 10−9 cells) in size. In the NO3-and N2-grown cultures, N content per cell was only 57 and 63%, respectively, of that in the NH3-grown cells. Heterocysts were absent in NH3-grown cultures but were present in both the N2 and NO3 cultures. In the NO3-grown cultures C2H2 reduction was detected only at high growth rates, where it was estimated to account for a maximum of 6% of the N assimilated. In the N2-fixing cultures the acetylene:N2 ratio varied from 3.4:1 at lower growth rates to 3.0:1 at growth rates approaching maximal.

Compared with NH3, the assimilation of NO3 and N2 resulted either in a decrease in cellular C (NO3 and N2 cultures) or in a lower maximal growth rate (N2 culture only). The observed changes in cell C content were used to calculate the net cost (in electron pair equivalents) associated with growth on NO3 or N2 compared with NH3.

  相似文献   

20.
Nitrogen (N) is an essential nutrient in the sea and its distribution is controlled by microorganisms. Within the N cycle, nitrite (NO2) has a central role because its intermediate redox state allows both oxidation and reduction, and so it may be used by several coupled and/or competing microbial processes. In the upper water column and oxygen minimum zone (OMZ) of the eastern tropical North Pacific Ocean (ETNP), we investigated aerobic NO2 oxidation, and its relationship to ammonia (NH3) oxidation, using rate measurements, quantification of NO2-oxidizing bacteria via quantitative PCR (QPCR), and pyrosequencing. 15NO2 oxidation rates typically exhibited two subsurface maxima at six stations sampled: one located below the euphotic zone and beneath NH3 oxidation rate maxima, and another within the OMZ. 15NO2 oxidation rates were highest where dissolved oxygen concentrations were <5 μM, where NO2 accumulated, and when nitrate (NO3) reductase genes were expressed; they are likely sustained by NO3 reduction at these depths. QPCR and pyrosequencing data were strongly correlated (r2=0.79), and indicated that Nitrospina bacteria numbered up to 9.25% of bacterial communities. Different Nitrospina groups were distributed across different depth ranges, suggesting significant ecological diversity within Nitrospina as a whole. Across the data set, 15NO2 oxidation rates were decoupled from 15NH4+ oxidation rates, but correlated with Nitrospina (r2=0.246, P<0.05) and NO2 concentrations (r2=0.276, P<0.05). Our findings suggest that Nitrospina have a quantitatively important role in NO2 oxidation and N cycling in the ETNP, and provide new insight into their ecology and interactions with other N-cycling processes in this biogeochemically important region of the ocean.  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