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.
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   

2.
The capacity for dissimilatory reduction of NO3 to N2 (N2O) and NH4+ was measured in 15NO3-amended marine sediment. Incubation with acetylene (7 × 10−3 atmospheres [normal]) caused accumulation of N2O in the sediment. The rate of N2O production equaled the rate of N2 production in samples without acetylene. Complete inhibition of the reduction of N2O to N2 suggests that the “acetylene blockage technique” is applicable to assays for denitrification in marine sediments. The capacity for reduction of NO3 by denitrification decreased rapidly with depth in the sediment, whereas the capacity for reduction of NO3 to NH4+ was significant also in deeper layers. The data suggested that the latter process may be equally as significant as denitrification in the turnover of NO3 in marine sediments.  相似文献   

3.
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   

4.
Microbial Fe reduction in acetate- and succinate-containing enrichment cultures initiated with an estuarine sediment inoculum was studied. Fe reduction was unaffected when SO42− reduction was inhibited by MoO42−, indicating that both processes could occur independently. Bacterially produced sulfide precipitated as FeS but was not completely responsible for Fe reduction. The separation of oxidized Fe particles from bacteria by dialysis tubing demonstrated that direct bacterial contact was necessary for Fe reduction. Fe reduction in cultures amended with NO3 was delayed until NO3 and NO2 were removed. However, bacterial attachment to oxidized Fe particles in NO3-amended cultures occurred early during growth in a manner similar to NO3-free cultures. During late stages of growth, bacteria not attached to Fe particles became pale and swollen, while attached cells remained bright blue when examined by 4′,6-diamidine-2-phenylindole epifluo-rescence microscopy. The presence of added oxidized Mn had no effect on Fe reduction. The results suggested that enzymatic Fe reduction was responsible for reducing Fe in these cultures even in the presence of sulfide and that cells incapable of Fe reduction became unhealthy when Fe(III) was the only available electron acceptor.  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.
Dissimilatory reduction of NO2 to N2O and NH4+ by a soil Citrobacter sp. was studied in an attempt to elucidate the physiological and ecological significance of N2O production by this mechanism. In batch cultures with defined media, NO2 reduction to NH4+ was favored by high glucose and low NO3 concentrations. Nitrous oxide production was greatest at high glucose and intermediate NO3 concentrations. With succinate as the energy source, little or no NO2 was reduced to NH4+ but N2O was produced. Resting cell suspensions reduced NO2 simultaneously to N2O and free extracellular NH4+. Chloramphenicol prevented the induction of N2O-producing activity. The Km for NO2 reduction to N2O was estimated to be 0.9 mM NO2, yet the apparent Km for overall NO2 reduction was considerably lower, no greater than 0.04 mM NO2. Activities for N2O and NH4+ production increased markedly after depletion of NO3 from the media. Amendment with NO3 inhibited N2O and NH4+ production by molybdate-grown cells but not by tungstate-grown cells. Sulfite inhibited production of NH4+ but not of N2O. In a related experiment, three Escherichia coli mutants lacking NADH-dependent nitrite reductase produced N2O at rates equal to the wild type. These observations suggest that N2O is produced enzymatically but not by the same enzyme system responsible for dissimilatory reduction of NO2 to NH4+.  相似文献   

7.
Aquaspirillum magnetotacticum MS-1 grew microaerobically but not anaerobically with NO3 or NH4+ as the sole nitrogen source. Nevertheless, cell yields varied directly with NO3 concentration under microaerobic conditions. Products of NO3 reduction included NH4+, N2O, NO, and N2. NO2 and NH2OH, each toxic to cells at 0.2 mM, were not detected as products of cells growing on NO3. NO3 reduction to NH4+ was completely repressed by the addition of 2 mM NH4+ to the growth medium, whereas NO3 reduction to N2O or to N2 was not. C2H2 completely inhibited N2O reduction to N2 by growing cells. These results indicate that A. magnetotacticum is a microaerophilic denitrifier that is versatile in its nitrogen metabolism, concomitantly reducing NO3 by assimilatory and dissimilatory means. This bacterium appears to be the first described denitrifier with an absolute requirement for O2. The process of NO3 reduction appears well adapted for avoiding accumulation of several nitrogenous intermediates that are toxic to cells.  相似文献   

8.
Microzonation of denitrification was studied in stream sediments by a combined O2 and N2O microsensor technique. O2 and N2O concentration profiles were recorded simultaneously in intact sediment cores in which C2H2 was added to inhibit N2O reduction in denitrification. The N2O profiles were used to obtain high-resolution profiles of denitrification activity and NO3 distribution in the sediments. O2 penetrated about 1 mm into the dark-incubated sediments, and denitrification was largely restricted to a thin anoxic layer immediately below that. With 115 μM NO3 in the water phase, denitrification was limited to a narrow zone from 0.7 to 1.4 mm in depth, and total activity was 34 nmol of N cm−2 h−1. With 1,250 μM NO3 in the water, the denitrification zone was extended to a layer from 0.9 to 4.8 mm in depth, and total activity increased to 124 nmol of N cm−2 h−1. Within most of the activity zone, denitrification was not dependent on the NO3 concentration and the apparent Km for NO3 was less than 10 μM. Denitrification was the only NO3-consuming process in the dark-incubated stream sediment. Even in the presence of C2H2, a significant N2O reduction (up to 30% of the total N2O production) occurred in the reduced, NO3-free layers below the denitrification zone. This effect must be corrected for during use of the conventional C2H2 inhibition technique.  相似文献   

9.
Denitrification by Chromobacterium violaceum   总被引:2,自引:0,他引:2       下载免费PDF全文
One host (Rana catesbiana)-associated and two free-living mesophilic strains of bacteria with violet pigmentation and biochemical characteristics of Chromobacterium violaceum were isolated from freshwater habitats. Cells of each freshly isolated strain and of strain ATCC 12472 (the neotype strain) grew anaerobically with glucose as the sole carbon and energy source. The major fermentation products of cells grown in Trypticase soy broth (BBL Microbiology Systems, Cockeysville, Md.) supplemented with glucose included acetate, small amounts of propionate, lactate, and pyruvate. The final cell yield and culture growth rate of each strain cultured anaerobically in this medium increased approximately twofold with the addition of 2 mM NaNO3. Final growth yields increased in direct proportion to the quantity of added NaNO3 over the range of 0.5 to 5 mM. Each strain reduced NO3, producing NO2, NO, and N2O. NO2 accumulated transiently. With 2 mM NaNO3 in the medium, N2O made up 85 to 98% of the N product recovered with each strain. N-oxides were recovered in the same quantity and distribution whether 0.01 atm (ca. 1 kPa) of C2H2 (added to block N2O reduction) was present or not. Neither N2 production nor gas accumulation was detected during NO3 reduction by growing cells. Cell growth in media containing 0.5 to 5 mM NaNO2 in lieu of NaNO3 was delayed, and although N2O was produced by the end of growth, NO2 -containing media did not support growth to an extent greater than did medium lacking NO3 or NO2. The data indicate that C. violaceum cells ferment glucose or denitrify, terminating denitrification with the production of N2O, and that NO2 reduction to N2O is not coupled to growth but may serve as a detoxification mechanism. No strain detectably fixed N2 (reduced C2H2).  相似文献   

10.
A sensitive NO2 biosensor that is based on bacterial reduction of NO2 to N2O and subsequent detection of the N2O by a built-in electrochemical N2O sensor was developed. Four different denitrifying organisms lacking NO3 reductase activity were assessed for use in the biosensor. The relevant physiological aspects examined included denitrifying characteristics, growth rate, NO2 tolerance, and temperature and salinity effects on the growth rate. Two organisms were successfully used in the biosensor. The preferred organism was Stenotrophomonas nitritireducens, which is an organism with a denitrifying pathway deficient in both NO3 and N2O reductases. Alternatively Alcaligenes faecalis could be used when acetylene was added to inhibit its N2O reductase. The macroscale biosensors constructed exhibited a linear NO2 response at concentrations up to 1 to 2 mM. The detection limit was around 1 μM NO2, and the 90% response time was 0.5 to 3 min. The sensor signal was specific for NO2, and interference was observed only with NH2OH, NO, N2O, and H2S. The sensor signal was affected by changes in temperature and salinity, and calibration had to be performed in a system with a temperature and an ionic strength comparable to those of the medium analyzed. A broad range of water bodies could be analyzed with the biosensor, including freshwater systems, marine systems, and oxic-anoxic wastewaters. The NO2 biosensor was successfully used for long-term online monitoring in wastewater. Microscale versions of the NO2 biosensor were constructed and used to measure NO2 profiles in marine sediment.  相似文献   

11.
A pilot-scale field experiment demonstrated that a one-time amendment of emulsified vegetable oil (EVO) reduced groundwater U(VI) concentrations for 1 year in a fast-flowing aquifer. However, little is known about how EVO amendment stimulates the functional gene composition, structure, and dynamics of groundwater microbial communities toward prolonged U(VI) reduction. In this study, we hypothesized that EVO amendment would shift the functional gene composition and structure of groundwater microbial communities and stimulate key functional genes/groups involved in EVO biodegradation and reduction of electron acceptors in the aquifer. To test these hypotheses, groundwater microbial communities after EVO amendment were analyzed using a comprehensive functional gene microarray. Our results showed that EVO amendment stimulated sequential shifts in the functional composition and structure of groundwater microbial communities. Particularly, the relative abundance of key functional genes/groups involved in EVO biodegradation and the reduction of NO3, Mn(IV), Fe(III), U(VI), and SO42− significantly increased, especially during the active U(VI) reduction period. The relative abundance for some of these key functional genes/groups remained elevated over 9 months. Montel tests suggested that the dynamics in the abundance, composition, and structure of these key functional genes/groups were significantly correlated with groundwater concentrations of acetate, NO3, Mn(II), Fe(II), U(VI), and SO42−. Our results suggest that EVO amendment stimulated dynamic succession of key functional microbial communities. This study improves our understanding of the composition, structure, and function changes needed for groundwater microbial communities to sustain a long-term U(VI) reduction.  相似文献   

12.
A more sensitive analytical method for NO3 was developed based on the conversion of NO3 to N2O by a denitrifier that could not reduce N2O further. The improved detectability resulted from the high sensitivity of the 63Ni electron capture gas chromatographic detector for N2O and the purification of the nitrogen afforded by the transformation of the N to a gaseous product with a low atmospheric background. The selected denitrifier quantitatively converted NO3 to N2O within 10 min. The optimum measurement range was from 0.5 to 50 ppb (50 μg/liter) of NO3 N, and the detection limit was 0.2 ppb of N. The values measured by the denitrifier method compared well with those measured by the high-pressure liquid chromatographic UV method above 2 ppb of N, which is the detection limit of the latter method. It should be possible to analyze all types of samples for nitrate, except those with inhibiting substances, by this method. To illustrate the use of the denitrifier method, NO3 concentrations of <2 ppb of NO3 N were measured in distilled and deionized purified water samples and in anaerobic lake water samples, but were not detected at the surface of the sediment. The denitrifier method was also used to measure the atom% of 15N in NO3. This method avoids the incomplete reduction and contamination of the NO3 -N by the NH4+ and N2 pools which can occur by the conventional method of 15NO3 analysis. N2O-producing denitrifier strains were also used to measure the apparent Km values for NO3 use by these organisms. Analysis of N2O production by use of a progress curve yielded Km values of 1.7 and 1.8 μM NO3 for the two denitrifier strains studied.  相似文献   

13.
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   

14.
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   

15.
Denitrification in San Francisco Bay Intertidal Sediments   总被引:23,自引:17,他引:6       下载免费PDF全文
The acetylene block technique was employed to study denitrification in intertidal estuarine sediments. Addition of nitrate to sediment slurries stimulated denitrification. During the dry season, sediment-slurry denitrification rates displayed Michaelis-Menten kinetics, and ambient NO3 + NO2 concentrations (≤26 μM) were below the apparent Km (50 μM) for nitrate. During the rainy season, when ambient NO3 + NO2 concentrations were higher (37 to 89 μM), an accurate estimate of the Km could not be obtained. Endogenous denitrification activity was confined to the upper 3 cm of the sediment column. However, the addition of nitrate to deeper sediments demonstrated immediate N2O production, and potential activity existed at all depths sampled (the deepest was 15 cm). Loss of N2O in the presence of C2H2 was sometimes observed during these short-term sediment incubations. Experiments with sediment slurries and washed cell suspensions of a marine pseudomonad confirmed that this N2O loss was caused by incomplete blockage of N2O reductase by C2H2 at low nitrate concentrations. Areal estimates of denitrification (in the absence of added nitrate) ranged from 0.8 to 1.2 μmol of N2 m−2 h−1 (for undisturbed sediments) to 17 to 280 μmol of N2 m−2 h−1 (for shaken sediment slurries).  相似文献   

16.
Until recently, denitrification was thought to be the only significant pathway for N2 formation and, in turn, the removal of nitrogen in aquatic sediments. The discovery of anaerobic ammonium oxidation in the laboratory suggested that alternative metabolisms might be present in the environment. By using a combination of 15N-labeled NH4+, NO3, and NO2 (and 14N analogues), production of 29N2 and 30N2 was measured in anaerobic sediment slurries from six sites along the Thames estuary. The production of 29N2 in the presence of 15NH4+ and either 14NO3 or 14NO2 confirmed the presence of anaerobic ammonium oxidation, with the stoichiometry of the reaction indicating that the oxidation was coupled to the reduction of NO2. Anaerobic ammonium oxidation proceeded at equal rates via either the direct reduction of NO2 or indirect reduction, following the initial reduction of NO3. Whether NO2 was directly present at 800 μM or it accumulated at 3 to 20 μM (from the reduction of NO3), the rate of 29N2 formation was not affected, which suggested that anaerobic ammonium oxidation was saturated at low concentrations of NO2. We observed a shift in the significance of anaerobic ammonium oxidation to N2 formation relative to denitrification, from 8% near the head of the estuary to less than 1% at the coast. The relative importance of anaerobic ammonium oxidation was positively correlated (P < 0.05) with sediment organic content. This report of anaerobic ammonium oxidation in organically enriched estuarine sediments, though in contrast to a recent report on continental shelf sediments, confirms the presence of this novel metabolism in another aquatic sediment system.  相似文献   

17.
Heterotrophic bacteria, yeasts, fungi, plants, and animal breath were investigated as possible sources of N2O. Microbes found to produce N2O from NO3 but not consume it were: (i) all of the nitrate-respiring bacteria examined, including strains of Escherichia, Serratia, Klebsiella, Enterobacter, Erwinia, and Bacillus; (ii) one of the assimilatory nitrate-reducing bacteria examined, Azotobacter vinelandii, but not Azotobacter macrocytogenes or Acinetobacter sp.; and (iii) some but not all of the assimilatory nitrate-reducing yeasts and fungi, including strains of Hansenula, Rhodotorula, Aspergillus, Alternaria, and Fusarium. The NO3-reducing obligate anaerobe Clostridium KDHS2 did not produce N2O. Production of N2O occurred only in stationary phase. The nitrate-respiring bacteria produced much more N2O than the other organisms, with yields of N2O ranging from 3 to 36% of 3.5 mM NO3. Production of N2O was apparently not regulated by ammonium and was not restricted to aerobic or anaerobic conditions. Plants do not appear to produce N2O, although N2O was found to arise from some damaged plant tops, probably due to microbial growth. Concentrations of N2O above the ambient level in the atmosphere were found in human breath and appeared to increase after a meal of high-nitrate food.  相似文献   

18.
A microscale biosensor for acetate, propionate, isobutyrate, and lactate is described. The sensor is based on the bacterial respiration of low-molecular-weight, negatively charged species with a concomitant reduction of NO3 to N2O. A culture of denitrifying bacteria deficient in N2O reductase was immobilized in front of the tip of an electrochemical N2O microsensor. The bacteria were separated from the outside environment by an ion-permeable membrane and supplied with nutrients (except for electron donors) from a medium reservoir behind the N2O sensor. The signal of the sensor, which corresponded to the rate of N2O production, was proportional to the supply of the electron donor to the bacterial mass. The selectivity for volatile fatty acids compared to other organic compounds was increased by selectively enhancing the transport of negatively charged compounds into the sensor by electrophoretic migration (electrophoretic sensitivity control). The sensor was susceptible to interference from O2, N2O, NO2, H2S, and NO3. Interference from NO3 was low and could be quantified and accounted for. The detection limit was equivalent to about 1 μM acetate, and the 90% response time was 30 to 90 s. The response of the sensor was not affected by changes in pH between 5.5 and 9 and was also unaffected by changes in salinity in the range of 2 to 32‰. The functioning of the sensor over a temperature span of 7 to 30°C was investigated. The concentration range for a linear response was increased five times by increasing the temperature from 7 to 19.5°C. The life span of the biosensor varied between 1 and 3 weeks after manufacturing.  相似文献   

19.
Aerobic N2O production was studied in nitrifying humus from urea-fertilized pine forest soil. Acetylene and nitrapyrin inhibited both NH4+ oxidation and N2O production, indicating that N2O production was closely associated with autotrophic NH4+ oxidation. N2O production was enhanced by low soil pH; it was negligible above pH 4.7. When soil pH decreased from 4.7 to 4.1, the relative amount of N2O-N produced from NH4+-N oxidized increased exponentially to 20%. There was also some evidence that N2O formation was stimulated by salts (potassium sulfate and sodium phosphates). The maximum rate of N2O-N production was 0.17 μg of N2O-N per g of soil per h. When humus was treated with NO2, N2O evolved immediately, indicating chemical formation, but no N2O was formed on the addition of NO3. The amount of N2O-N evolved was 0.6 to 4.6% of NO2-N added. A high concentration of NO2 and low soil pH enhanced chemical production of N2O. There was no accumulation of NO2 during nitrification. The calculations indicated that chemical formation of N2O was not the main source of N2O during NH4+ oxidation. After the addition of inhibitors of NH4+ oxidation the soils contained NO3, but no N2O was produced. The results suggest that enhanced autotrophic NH4+ oxidation is a potential source of N2O in fertilized acid forest soil.  相似文献   

20.
Phylogenetically diverse species of bacteria can catalyze the oxidation of ferrous iron [Fe(II)] coupled to nitrate (NO3) reduction, often referred to as nitrate-dependent iron oxidation (NDFO). Very little is known about the biochemistry of NDFO, and though growth benefits have been observed, mineral encrustations and nitrite accumulation likely limit growth. Acidovorax ebreus, like other species in the Acidovorax genus, is proficient at catalyzing NDFO. Our results suggest that the induction of specific Fe(II) oxidoreductase proteins is not required for NDFO. No upregulated periplasmic or outer membrane redox-active proteins, like those involved in Fe(II) oxidation by acidophilic iron oxidizers or anaerobic photoferrotrophs, were observed in proteomic experiments. We demonstrate that while “abiotic” extracellular reactions between Fe(II) and biogenic NO2/NO can be involved in NDFO, intracellular reactions between Fe(II) and periplasmic components are essential to initiate extensive NDFO. We present evidence that an organic cosubstrate inhibits NDFO, likely by keeping periplasmic enzymes in their reduced state, stimulating metal efflux pumping, or both, and that growth during NDFO relies on the capacity of a nitrate-reducing bacterium to overcome the toxicity of Fe(II) and reactive nitrogen species. On the basis of our data and evidence in the literature, we postulate that all respiratory nitrate-reducing bacteria are innately capable of catalyzing NDFO. Our findings have implications for a mechanistic understanding of NDFO, the biogeochemical controls on anaerobic Fe(II) oxidation, and the production of NO2, NO, and N2O in the environment.  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