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.
Excessive NH(3) production in the rumen is a major nutritional inefficiency in ruminant animals. Experiments were undertaken to compare the rates of NH(3) production from different substrates in ruminal fluid in vitro and to assess the role of asaccharolytic bacteria in NH(3) production. Ruminal fluid was taken from four rumen-fistulated sheep receiving a mixed hay-concentrate diet. The calculated rate of NH(3) production from Trypticase varied from 1.8 to 19.7 nmol mg of protein(-1) min(-1) depending on the substrate, its concentration, and the method used. Monensin (5 micro M) inhibited NH(3) production from proteins, peptides, and amino acids by an average of 28% with substrate at 2 mg/ml, compared to 48% with substrate at 20 mg/ml (P = 0.011). Of the total bacterial population, 1.4% grew on Trypticase alone, of which 93% was eliminated by 5 micro M monensin. Many fewer bacteria (0.002% of the total) grew on amino acids alone. Nineteen isolates capable of growth on Trypticase were obtained from four sheep. 16S ribosomal DNA and traditional identification methods indicated the bacteria fell into six groups. All were sensitive to monensin, and all except one group (group III, similar to Atopobium minutum), produced NH(3) at >250 nmol min(-1) mg of protein(-1), depending on the medium, as determined by a batch culture method. All isolates had exopeptidase activity, but only group III had an apparent dipeptidyl peptidase I activity. Groups I, II, and IV were most closely related to asaccharolytic ruminal and oral Clostridium and Eubacterium spp. Group V comprised one isolate, similar to Desulfomonas piger (formerly Desulfovibrio pigra). Group VI was 95% similar to Acidaminococcus fermentans. Growth of the Atopobium- and Desulfomonas-like isolates was enhanced by sugars, while growth of groups I, II, and V was significantly depressed by sugars. This study therefore demonstrates that different methodologies and different substrate concentrations provide an explanation for different apparent rates of ruminal NH(3) production reported in different studies and identifies a diverse range of hyper-ammonia-producing bacteria in the rumen of sheep.  相似文献   

2.
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   

3.
The origin of cell nitrogen and amino acid nitrogen during growth of ruminal cellulolytic bacteria in different growth media was investigated by using 15NH3. At high concentrations of peptides (Trypticase, 10 g/liter) and amino acids (15.5 g/liter), significant amounts of cell nitrogen of Fibrobacter succinogenes BL2 (51%), Ruminococcus flavefaciens 17 (43%), and Ruminococcus albus SY3 (46%) were derived from non-NH3-N. With peptides at 1 g/liter, a mean of 80% of cell nitrogen was from NH3. More cell nitrogen was formed from NH3 during growth on cellobiose compared with growth on cellulose in all media. Phenylalanine was essential for F. succinogenes, and its 15N enrichment declined more than that of other amino acids in all species when amino acids were added to the medium.  相似文献   

4.
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   

5.
The anaerobic ammonia-oxidizing activity of the planctomycete Candidatus “Brocadia anammoxidans” was not inhibited by NO concentrations up to 600 ppm and NO2 concentrations up to 100 ppm. B. anammoxidans was able to convert (detoxify) NO, which might explain the high NO tolerance of this organism. In the presence of NO2, the specific ammonia oxidation activity of B. anammoxidans increased, and Nitrosomonas-like microorganisms recovered an NO2-dependent anaerobic ammonia oxidation activity. Addition of NO2 to a mixed population of B. anammoxidans and Nitrosomonas induced simultaneous specific anaerobic ammonia oxidation activities of up to 5.5 mmol of NH4+ g of protein−1 h−1 by B. anammoxidans and up to 1.5 mmol of NH4+ g of protein−1 h−1 by Nitrosomonas. The stoichiometry of the converted N compounds (NO2/NH3 ratio) and the microbial community structure were strongly influenced by NO2. The combined activity of B. anammoxidans and Nitrosomonas-like ammonia oxidizers might be of relevance in natural environments and for technical applications.  相似文献   

6.
When ruminal fluid from a cow fed timothy hay was serially diluted (10-fold increments into anaerobic broth containing 15 mg ml−1 Trypticase), the low dilutions (≤10−6) had optical densities greater than 2.0 and ammonia concentrations greater than 100 mM. The optical densities and ammonia concentrations of the 10−8 and 10−9 dilutions were very low, but large cocci were observed in the 10−8 dilution. The large cocci were isolated and identified by 16S rDNA sequencing as Megasphaera elsdenii. The freshly isolated strain (JL1) grew well on Trypticase, but less than 4% of the amino acid nitrogen in Trypticase was converted to ammonia. Optical density and ammonia production were twice as great if Casamino acids were provided, and similar results were obtained with seven other strains (B159, AW106, YT91, LC1, T81, J1, and YZ70). Specific activities of deamination (based on Casamino acids) of the eight strains ranged from 100 (strain JL1) to 325 (strain B159) nmol mg protein−1 min−1. None of the strains could utilize branched-chain amino acids as an energy source for growth, but specific activities of branched-chain amino acid deamination ranged from 15 to 65 nmol mg protein−1 min−1. All eight of the M. elsdenii strains grew well in the presence of 5 μM monensin, and only two of the strains were strongly inhibited by 20 μM monensin. On the basis of these results, it appears that M. elsdenii is deficient in peptidase activity and can utilize only a few amino acids. Some M. elsdenii strains produced ammonia and branched-chain volatile fatty acids nearly as fast as obligate amino acid-fermenting ruminal bacteria, but the extent of this production was at least fourfold lower. Because all of the strains could tolerate 5 μM monensin, it is unlikely that this feed additive would significantly inhibit M. elsdenii in vivo. Received: 12 December 2001 / Accepted: 5 February 2002  相似文献   

7.
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   

8.
Anammox and denitrification mediated by bacteria are known to be the major microbial processes converting fixed N to N2 gas in various ecosystems. Codenitrification and denitrification by fungi are additional pathways producing N2 in soils. However, fungal codenitrification and denitrification have not been well investigated in agricultural soils. To evaluate bacterial and fungal processes contributing to N2 production, molecular and 15N isotope analyses were conducted with soil samples collected at six different agricultural fields in the United States. Denitrifying and anammox bacterial abundances were measured based on quantitative PCR (qPCR) of nitrous oxide reductase (nosZ) and hydrazine oxidase (hzo) genes, respectively, while the internal transcribed spacer (ITS) of Fusarium oxysporum was quantified to estimate the abundance of codenitrifying and denitrifying fungi. 15N tracer incubation experiments with 15NO3 or 15NH4+ addition were conducted to measure the N2 production rates from anammox, denitrification, and codenitrification. Soil incubation experiments with antibiotic treatments were also used to differentiate between fungal and bacterial N2 production rates in soil samples. Denitrifying bacteria were found to be the most abundant, followed by F. oxysporum based on the qPCR assays. The potential denitrification rates by bacteria and fungi ranged from 4.118 to 42.121 nmol N2-N g−1 day−1, while the combined potential rates of anammox and codenitrification ranged from 2.796 to 147.711 nmol N2-N g−1 day−1. Soil incubation experiments with antibiotics indicated that fungal codenitrification was the primary process contributing to N2 production in the North Carolina soil. This study clearly demonstrates the importance of fungal processes in the agricultural N cycle.  相似文献   

9.
Nitrogen fixation (diazotrophy) has recently been demonstrated in several methanogenic archaebacteria. To compare the process in an archaebacterium with that in eubacteria, we examined the properties of diazotrophic growth and nitrogenase activity in Methanosarcina barkeri 227. Growth yields with methanol or acetate as a growth substrate were significantly lower in N2-grown cultures than in NH4+-grown cultures, and the culture doubling times were increased, indicating that diazotrophy was energetically costly, as it is in eubacteria. Growth of nitrogen-fixing cells was inhibited when molybdenum was omitted from the medium; addition of 10 nM molybdate stimulated growth, while 1 μM molybdate restored maximum diazotrophic growth. Omission of molybdenum did not inhibit growth of ammonia-grown cells. Tungstate (100 μM) strongly inhibited growth of molybdenum-deficient diazotrophic cells, while ammonia-grown cells were unaffected. The addition of 100 nM vanadate or chromate did not stimulate diazotrophic growth of molybdenum-starved cells. These results are consistent with the presence of a molybdenum-containing nitrogenase in M. barkeri. Acetylene, the usual substrate for assaying nitrogenase activity, inhibited methanogenesis by M. barkeri and consequently needed to be used at a low partial pressure (0.3% of the headspace) when acetylene reduction by whole cells was assayed. Whole cells reduced 0.3% acetylene to ethylene at a very low rate (1 to 2 nmol h−1 mg of protein−1), and they “switched off” acetylene reduction in response to added ammonia or glutamine. Crude extracts from diazotrophic cells reduced 10% acetylene at a rate of 4 to 5 nmol of C2H4 formed h−1 mg of protein−1 when supplied with ATP and reducing power, while extracts of Klebsiella pneumoniae prepared by the same procedures had rates 100-fold higher. Acetylene reduction by extracts required ATP and was completely inhibited by 1 mM ADP in the presence of 5 mM ATP. The low rates of C2H2 reduction could be due to improper assay conditions, to switched-off enzyme, or to the nitrogenase's having lower activity towards acetylene than towards dinitrogen.  相似文献   

10.
The inducibility and kinetics of the NO3, NO2, and NH4+ transporters in roots of wheat seedlings (Triticum aestivum cv Yercora Rojo) were characterized using precise methods approaching constant analysis of the substrate solutions. A microcomputer-controlled automated high performance liquid chromatography system was used to determine the depletion of each N species (initially at 1 millimolar) from complete nutrient solutions. Uptake rate analyses were performed using computerized curve-fitting techniques. More precise estimates were obtained for the time required for and the extent of the induction of each transporter. Up to 10 and 6 hours, respectively, were required to achieve apparent full induction of the NO3 and NO2 transporters. Evidence for substrate inducibility of the NH4+ transporters requiring 5 hours is presented. The transport of NO3 was mediated by a dual system (or dual phasic), whereas only single systems were found for transport of NO2 and NH4+. The Km values for NO3, NO2, and NH4+ were, respectively, 0.027, 0.054, and 0.05 millimolar. The Km for mechanism II of NO3 transport could not be defined in this study as it exhibited only apparent first order kinetics up to 1 millimolar.  相似文献   

11.
Ricinus communis L. plants were grown in nutrient solutions in which N was supplied as NO3 or NH4+, the solutions being maintained at pH 5.5. In NO3-fed plants excess nutrient anion over cation uptake was equivalent to net OH efflux, and the total charge from NO3 and SO42− reduction equated to the sum of organic anion accumulation plus net OH efflux. In NH4+-fed plants a large H+ efflux was recorded in close agreement with excess cation over anion uptake. This H+ efflux equated to the sum of net cation (NH4+ minus SO42−) assimilation plus organic anion accumulation. In vivo nitrate reductase assays revealed that the roots may have the capacity to reduce just under half of the total NO3 that is taken up and reduced in NO3-fed plants. Organic anion concentration in these plants was much higher in the shoots than in the roots. In NH4+-fed plants absorbed NH4+ was almost exclusively assimilated in the roots. These plants were considerably lower in organic anions than NO3-fed plants, but had equal concentrations in shoots and roots. Xylem and phloem saps were collected from plants exposed to both N sources and analyzed for all major contributing ionic and nitrogenous compounds. The results obtained were used to assist in interpreting the ion uptake, assimilation, and accumulation data in terms of shoot/root pH regulation and cycling of nutrients.  相似文献   

12.
The effect of nitrogen form (NH4-N, NH4-N + NO3, NO3) on nitrate reductase activity in roots and shoots of maize (Zea mays L. cv INRA 508) seedlings was studied. Nitrate reductase activity in leaves was consistent with the well known fact that NO3 increases, and NH4+ and amide-N decrease, nitrate reductase activity. Nitrate reductase activity in the roots, however, could not be explained by the root content of NO3, NH4-N, and amide-N. In roots, nitrate reductase activity in vitro was correlated with the rate of nitrate reduction in vivo. Inasmuch as nitrate reduction results in the production of OH and stimulates the synthesis of organic anions, it was postulated that nitrate reductase activity of roots is stimulated by the released OH or by the synthesized organic anions rather than by nitrate itself. Addition of HCO3 to nutrient solution of maize seedlings resulted in a significant increase of the nitrate reductase activity in the roots. As HCO3, like OH, increases pH and promotes the synthesis of organic anions, this provides circumstantial evidence that alkaline conditions and/or organic anions have a more direct impact on nitrate reductase activity than do NO3, NH4-N, and amide-N.  相似文献   

13.
We investigated autotrophic anaerobic ammonium-oxidizing (anammox) biofilms for their spatial organization, community composition, and in situ activities by using molecular biological techniques combined with microelectrodes. Results of phylogenetic analysis and fluorescence in situ hybridization (FISH) revealed that “Brocadia”-like anammox bacteria that hybridized with the Amx820 probe dominated, with 60 to 92% of total bacteria in the upper part (<1,000 μm) of the biofilm, where high anammox activity was mainly detected with microelectrodes. The relative abundance of anammox bacteria decreased along the flow direction of the reactor. FISH results also indicated that Nitrosomonas-, Nitrosospira-, and Nitrosococcus-like aerobic ammonia-oxidizing bacteria (AOB) and Nitrospira-like nitrite-oxidizing bacteria (NOB) coexisted with anammox bacteria and accounted for 13 to 21% of total bacteria in the biofilms. Microelectrode measurements at three points along the anammox reactor revealed that the NH4+ and NO2 consumption rates decreased from 0.68 and 0.64 μmol cm−2 h−1 at P2 (the second port, 170 mm from the inlet port) to 0.30 and 0.35 μmol cm−2 h−1 at P3 (the third port, 205 mm from the inlet port), respectively. No anammox activity was detected at P4 (the fourth port, 240 mm from the inlet port), even though sufficient amounts of NH4+ and NO2 and a high abundance of anammox bacteria were still present. This result could be explained by the inhibitory effect of organic compounds derived from biomass decay and/or produced by anammox and coexisting bacteria in the upper parts of the biofilm and in the upstream part of the reactor. The anammox activities in the biofilm determined by microelectrodes reflected the overall reactor performance. The several groups of aerobic AOB lineages, Nitrospira-like NOB, and Betaproteobacteria coexisting in the anammox biofilm might consume a trace amount of O2 or organic compounds, which consequently established suitable microenvironments for anammox bacteria.  相似文献   

14.
Changes in the concentrations of NH4+ and amides during the growth of suspension cultures of rose (Rosa cv. Paul's Scarlet) cells were examined. When cells were grown in medium possessing only NO3 as a nitrogen source, the concentrations of NH4+ and amides increased to 4.0 × 10−1 and 5.9 micromoles per gram fresh weight, respectively. The amounts of both constituents declined during the later stages of growth. When a trace amount of NH4+ was added to the NO3 base starting medium, the concentration of NH4+ in the cells was increased to 7.0 × 10−1 micromoles per gram fresh weight.  相似文献   

15.
Ammonia secretion by the collecting duct (CD) is critical for acid-base homeostasis and, when defective, causes distal renal tubular acidosis (dRTA). The Rhesus protein RhCG mediates NH3 transport as evident from cell-free and cellular models as well as from Rhcg-null mice. Here, we investigated in a Rhcg mouse model the metabolic effects of Rhcg haploinsufficiency, the role of Rhcg in basolateral NH3 transport, and the mechanisms of adaptation to the lack of Rhcg. Both Rhcg+/+ and Rhcg+/− mice were able to handle an acute acid load, whereas Rhcg−/− mice developed severe metabolic acidosis with reduced ammonuria and high mortality. However, chronic acid loading revealed that Rhcg+/− mice did not fully recover, showing lower blood HCO3 concentration and more alkaline urine. Microperfusion studies demonstrated that transepithelial NH3 permeability was reduced by 80 and 40%, respectively, in CDs from Rhcg−/− and Rhcg+/− mice compared with controls. Basolateral membrane permeability to NH3 was reduced in CDs from Rhcg−/− mice consistent with basolateral Rhcg localization. Rhcg−/− responded to acid loading with normal expression of enzymes and transporters involved in proximal tubular ammoniagenesis but reduced abundance of the NKCC2 transporter responsible for medullary accumulation of ammonium. Consequently, tissue ammonium content was decreased. These data demonstrate a role for apical and basolateral Rhcg in transepithelial NH3 transport and uncover an incomplete dRTA phenotype in Rhcg+/− mice. Haploinsufficiency or reduced expression of RhCG may underlie human forms of (in)complete dRTA.  相似文献   

16.
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.
This study aimed at the investigation of the mechanisms of acidogenesis, which is a key process during anaerobic digestion. To expose possible bottlenecks, specific activities of the key enzymes of acidification, such as acetate kinase (Ack, 0.23–0.99 U mg−1 protein), butyrate kinase (Buk, < 0.03 U mg−1 protein) and butyryl-CoA:acetate-CoA transferase (But, 3.24–7.64 U mg−1 protein), were determined in cell free extracts of biogas reactor content from three different biogas reactors. Furthermore, the detection of Ack was successful via Western blot analysis. Quantification of corresponding functional genes encoding Buk (buk) and But (but) was not feasible, although an amplification was possible. Thus, phylogenetic trees were constructed based on respective gene fragments. Four new clades of possible butyrate-producing bacteria were postulated, as well as bacteria of the genera Roseburia or Clostridium identified. The low Buk activity was in contrast to the high specific But activity in the analysed samples. Butyrate formation via Buk activity does barely occur in the investigated biogas reactor. Specific enzyme activities (Ack, Buk and But) in samples drawn from three different biogas reactors correlated with ammonia and ammonium concentrations (NH3 and NH4+-N), and a negative dependency can be postulated. Thus, high concentrations of NH3 and NH4+-N may lead to a bottleneck in acidogenesis due to decreased specific acidogenic enzyme activities.  相似文献   

19.
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   

20.
Vibrio gazogenes ATCC 29988 growth and prodigiosin synthesis were studied in batch culture on complex and defined media and in chemostat cultures on defined medium. In batch culture on complex medium, a maximum growth rate of 0.75 h−1 and a maximum prodigiosin concentration of 80 ng of prodigiosin · mg of cell protein−1 were observed. In batch culture on defined medium, maximum growth rates were lower (maximum growth rate, 0.40 h−1), and maximum prodigiosin concentrations were higher (1,500 ng · mg of protein−1). In batch culture on either complex or defined medium, growth was characterized by a period of logarithmic growth followed by a period of linear growth; on either medium, prodigiosin biosynthesis was maximum during linear growth. In batch culture on defined medium, the initial concentration of glucose optimal for growth and pigment production was 3.0%; higher levels of glucose suppressed synthesis of the pigment. V. gazogenes had an absolute requirement for Na+; optimal growth occurred in the presence of 100 mM NaCl. Increases in the concentration of Na+ up to 600 mM resulted in further increases in the concentration of pigment in the broth. Prodigiosin was synthesized at a maximum level in the presence of inorganic phosphate concentrations suboptimal for growth. Concentrations of KH2PO4 above 0.4 mM caused decreased pigment synthesis, whereas maximum cell growth occurred at 1.0 mM. Optimal growth and pigment production occurred in the presence of 8 to 16 mg of ferric ion · liter−1, with higher concentrations proving inhibitory to both growth and pigment production. Both growth and pigment production were found to decrease with increased concentrations of p-aminobenzoic acid. The highest specific concentration of prodigiosin (3,480 ng · mg protein−1) was observed in chemostat cultures at a dilution rate of 0.057 h−1. The specific rate of prodigiosin production at this dilution rate was approximately 80% greater than that observed in batch culture on defined medium. At dilution rates greater than 0.057 h−1, the concentration of cells decreased with increasing dilution rate, resulting in a profile comparable to that expected for linear growth kinetics. No explanation could be found for the linear growth profiles obtained for both batch and chemostat cultures.  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