首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The heavy use of fertilizers in agricultural lands can result in significant nitrate (NO3) loadings to the aquatic environment. We hypothesized that biological denitrification in agricultural ditches and streams could be enhanced by adding elemental sulfur (So) to the sediment layer, where it could act as a biofilm support and electron donor. Using a bench-scale stream mesocosm with a bed of So granules, we explored NO3 removal fluxes as a function of the effluent NO3 concentrations. With effluent NO3 ranging from 0.5 mg N L−1 to 4.1 mg N L−1, NO3 removal fluxes ranged from 228 mg N m−2 d−1 to 708 mg N m−2 d−1. This is as much as 100 times higher than for agricultural drainage streams. Sulfate (SO42−) production was high due to aerobic sulfur oxidation. Molecular studies demonstrated that the So amendment selected for Thiobacillus species, and that no special inoculum was required for establishing a So-based autotrophic denitrifying community. Modeling studies suggested that denitrification was diffusion limited, and advective flow through the bed would greatly enhance NO3 removal fluxes. Our results indicate that amendment with So is an effective means to stimulate denitrification in a stream environment. To minimize SO42− production, it may be better to place So deeper in the sediment layer.  相似文献   

2.
3.
The purpose of this study is to investigate the nitrogen removal performance of the anaerobic ammonium oxidation (Anammox) process and the microbial community that enables the Anammox system to function well at ambient temperatures. A reactor with a novel spiral structure was used as the gas-solid separator. The reactor was fed with synthetic inorganic wastewater composed mainly of NH4+-N and NO2-N, and operated for 92 days. Stable nitrogen removal rates (NRR) of 16.3 and 17.5 kg-N m−3 d−1 were obtained at operating temperatures of 33 ± 1 and 23 ± 2 °C, respectively. To our knowledge, such a high NRR at ambient temperatures has not been reported previously. In addition, the experiments presented herein confirm that high influent NO2-N concentration of 460 mg L−1 did not noticeably inhibit the Anammox activity. Furthermore, the freshwater Anammox bacterium KU2, which was identified as the dominant bacterial species in the consortium by 16S rRNA gene analysis, is considered to be responsible for the stable nitrogen removal performance at ambient temperatures.  相似文献   

4.
Climatic influence (global warming and decreased rainfall) could lead to an increase in the ecological and toxicological effects of the pollution in aquatic ecosystems, especially contamination from agricultural nitrate (NO3) fertilizers. Physicochemical properties of the surface waters and sediments of four selected sites varying in NO3 concentration along La Rocina Stream, which feeds Marisma del Rocio in Doñana National Park (South West, Spain), were studied. Electrical conductivity, pH, content in macro and microelements, total organic carbon and nitrogen, and dissolved carbon and nitrogen were affected by each sampling site and sampling time. Contaminant NO3 in surface water at the site with the highest NO3 concentration (ranged in 61.6-106.6 mg L−1) was of inorganic origin, most probably from chemical fertilizers, as determined chemically (90% of the total dissolved nitrogen from NO3) and by isotopic analysis of δ15N-NO3. Changes in seasonal weather conditions and hydrological effects at the sampling sites were also responsible for variations in some biological activities (dehydrogenase, β-glucosidase, arylsulphatase, acid phosphatase and urease) in sediments, as well as in the production of the greenhouse gases CO2, CH4 and N2O. Both organic matter and NO3 contents influenced rates of gas production. Increased NO3 concentration also resulted in enhanced levels of potential denitrification measured as N2O production. The denitrification process was affected by NO3 contamination and the rainfall regimen, increasing the greenhouse gases emissions (CO2, CH4 and especially N2O) during the driest season in all sampling sites studied.  相似文献   

5.
Nitrate (NO3) loss from agriculture to shallow groundwater and transferral to sensitive aquatic ecosystems is of global concern. Denitrifying bioreactor technology, where a solid carbon (C) reactive media intercepts contaminated groundwater, has been successfully used to convert NO3 to di-nitrogen (N2) gas. One of the challenges of groundwater remediation research is how to track denitrification potential spatially and temporally within reactive media and subsoil. First, using δ15N/δ18O isotopes, eight wells were divided into indicative transformational processes of ‘nitrification’ or ‘denitrification’ wells. Then, using N2/argon (Ar) ratios these wells were divided into ‘low denitrification potential’ or high denitrification potential’ categories. Secondly, using falling head tests, the saturated hydraulic conductivity (Ksat) in each well was estimated, creating two groups of ‘slow’ (0.06 m day−1) and ‘fast’ (0.13 m day−1) wells, respectively. Thirdly, two ‘low denitrification potential’ wells (one fast and one slow) with high NO3 concentration were amended with woodchip to enhance denitrification. Water samples were retrieved from all wells using a low flow syringe to avoid de-gassing and analysed for N2/Ar ratio using membrane inlet mass spectrometry. Results showed that there was good agreement between isotope and chemical (N2/Ar ratio and dissolved organic C (DOC)) and physio-chemical (dissolved oxygen, temperature, conductivity and pH) parameters. To explain the spatial and temporal distribution of NO3 and other parameters on site, the development of predictive models using the available datasets for this field site was examined for NO3, Cl, N2/Ar and DOC. Initial statistical analysis was directed towards the testing of the effect of woodchip amendment. The analysis was formulated as a repeated measures analysis of the factorial structure for treatment and time. Nitrate concentrations were related to Ksat and water level (p < 0.0001 and p = 0.02, respectively), but did not respond to woodchip addition (p = 0.09). This non-destructive technique allows elucidation of denitrification potential over time and could be used in denitrifying bioreactor technology to assess denitrification hotspots in reactive media, while developing a NO3 spatial and temporal predictive model for bioreactor site specific conditions.  相似文献   

6.
Wang B  Lan CQ 《Bioresource technology》2011,102(10):5639-5644
Biomass productivity of 350 mg DCW L−1 day−1 with a final biomass concentration of 3.15 g DCW L−1 was obtained with Neochloris oleoabundans grown in artificial wastewater at sodium nitrate and phosphate concentrations of 140 and 47 mg L−1, respectively, with undetectable levels of residual N and P in effluents. In secondary municipal wastewater effluents enriched with 70 mg N L−1, the alga achieved a final biomass concentration of 2.1 g DCW L−1 and a biomass productivity of 233.3 mg DCW L−1 day−1. While N removal was very sensitive to N:P ratio, P removal was independent of N:P ratio in the tested range. These results indicate that N. oleoabundans could potentially be employed for combined biofuel production and wastewater treatment.  相似文献   

7.
Hydrogenotrophic denitrification was demonstrated using hydrogen generated from anoxic corrosion of metallic iron. For this purpose, a mixture of hydrogenated water and nitrate solution was used as reactor feed. A semi-batch reactor with nitrate loading of 2000 mg m−3 d−1 and hydraulic retention time (HRT) of 50 days produced effluent with nitrate concentration of 0.27 mg N L−1 (99% nitrate removal). A continuous flow reactor with nitrate loading of 28.9 mg m−3 d−1 and HRT of 15.6 days produced effluent with nitrate concentration of ∼0.025 mg N L−1 (95% nitrate removal). In both cases, the concentration of nitrate degradation by-products, viz., ammonia and nitrite, were below detection limits. The rate of denitrification in the reactors was controlled by hydrogen availability, and hence to operate such reactors at higher nitrate loading rates and/or lower HRT than reported in the present study, hydrogen concentration in the hydrogenated water must be significantly increased.  相似文献   

8.
The Iberian Peninsula encompasses more than 80% of the species richness of European aquatic ranunculi. The floristic diversity of the phytocoenosis characterised by aquatic Ranunculus and the main physical–chemical factors of the water were studied in 43 localities of the central Iberian Peninsula. Four aquatic Ranunculus communities are found in most of the aquatic environments. These are species-poor and have an uneven distribution: three species of Batrachium are heterophyllous and their communities are distributed in different aquatic ecosystems on silicated substrates; one species is homophyllous and its community occurs in various aquatic ecosystems with carbonated waters. In the Mediterranean climate, Ranunculus species are present in different habitats, as shown by the results of all the statistical analyses. Ranunculus trichophyllus communities occur in base-rich waters with a high buffering capacity (2273.44 ± 794.57 mg CaCO3 L−1) and a high concentration of cations (Ca2+, 121 ± 33.12 mg L−1; Mg2+, 71.64 ± 82.77 mg L−1), nitrates (2.89 ± 4.80 mg L−1), ammonium (2.19 ± 1.36 mg L−1) and sulphates (216.25 ± 218.54 mg L−1). Ranunculus penicillatus communities are found in flowing waters with a high concentration of phosphates (0.48 ± 0.6 mg L−1) and intermediate buffering capacity (683.66 ± 446.76 mg CaCO3 L−1). Both Ranunculus pseudofluitans and Ranunculus peltatus communities grow in waters with low buffering capacity (R. pseudofluitans, 385.91 ± 209.2 mg CaCO3 L−1; R. peltatus, 263.3 ± 180.36 mg CaCO3 L−1), and a low concentration of cations (R. pseudofluitans: Ca2+, 12.57 ± 9.42 mg L−1; Mg2+, 3.42 ± 1.67 mg L−1; R. peltatus: Ca2+, 15 ± 18.26 mg L−1; Mg2+, 6.26 ± 8.89 mg L−1) and nutrients (R. pseudofluitans: nitrates, 0.23 ± 0.2 mg L−1; phosphates, 0.09 ± 0.1 mg L−1; R. peltatus: nitrates, 0.19 ± 0.21 mg L−1; phosphates, 0.09 ± 0.12 mg L−1); the first in flowing waters, the latter in still waters.  相似文献   

9.
We studied microbial N2 production via anammox and denitrification in the anoxic water column of a restored mining pit lake in Germany over an annual cycle. We obtained high-resolution hydrochemical profiles using a continuous pumping sampler. Lake Rassnitzer is permanently stratified at ca. 29 m depth, entraining anoxic water below a saline density gradient. Mixed-layer nitrate concentrations averaged ca. 200 μmol L−1, but decreased to zero in the anoxic bottom waters. In contrast, ammonium was <5 μmol L−1 in the mixed layer but increased in the anoxic waters to ca. 600 μmol L−1 near the sediments. In January and October, 15N tracer measurements detected anammox activity (maximum 504 nmol N2 L−1 d−1 in 15NH4+-amended incubations), but no denitrification. In contrast, in May, N2 production was dominated by denitrification (maximum 74 nmol N2 L−1 d−1). Anammox activity in May was significantly lower than in October, as characterized by anammox rates (maximum 6 vs. 16 nmol N2 L−1 d−1 in incubations with 15NO3), as well as relative and absolute anammox bacterial cell abundances (0.56% vs. 0.98% of all bacteria, and 2.7×104 vs. 5.2×104 anammox cells mL−1, respectively) (quantified by catalyzed reporter deposition-fluorescence in situ hybridization (CARD-FISH) with anammox bacteria-specific probes). Anammox bacterial diversity was investigated with anammox bacteria-specific 16S rRNA gene clone libraries. The majority of anammox bacterial sequences were related to the widespread Candidatus Scalindua sorokinii/brodae cluster. However, we also found sequences related to Candidatus S. wagneri and Candidatus Brocadia fulgida, which suggests a high anammox bacterial diversity in this lake comparable with estuarine sediments.  相似文献   

10.
This study investigated the anaerobic degradation of tetrachlorobisphenol-A (TCBPA) in sediment samples collected at three sites along the Erren River in southern Taiwan. TCBPA anaerobic degradation half-lives (t1/2) in the sediment were 12.6, 16.9 and 21.7 d at concentrations of 50, 100, and 250 ??g g−1, respectively. TCBPA (50 ??g g−1) anaerobic degradation half-lives (t1/2) in the sediment were 10.1, 11.8, 11.0, 11.6, 10.8, 9.1, 8.5, 18.2, 19.3, and 16.1 d by the addition of yeast extract (5 mg l−1), cellulose (0.96 mg l−1), sodium chloride (1%), brij 30 (130 mg l−1), brij 35 (43 mg l−1), rhamnolipid (55 ??M), surfactin (91 ??M), phthalic esters (2 mg l−1), nonylphenol (2 mg l−1), and heavy metals (2 mg l−1), respectively. The degradation rate of TCBPA was enhanced by the addition of yeast extract, cellulose, sodium chloride, brij 30, brij 35, rhamnolipid, or surfactin. However, it was inhibited by the addition of phthalic esters, nonylphenol, or heavy metals. Also noted was the presence of dichlorobisphenol-A and bisphenol-A, two intermediate products resulting from the anaerobic degradation of TCBPA accumulated in the sediments.  相似文献   

11.
The metabolic capability of denitrifying sludge to oxidize ammonium and p-cresol was evaluated in batch cultures. Ammonium oxidation was studied in presence of nitrite and/or p-cresol by 55 h. At 50 mg/L NH4+-N and 76 mg/L NO2-N, the substrates were consumed at 100% and 95%, respectively, being N2 the product. At 50 mg/L NH4+-N and 133 mg/L NO2-N, the consumption efficiencies decreased to 96% and 70%, respectively. The increase in nitrite concentration affected the ammonium oxidation rate. Nonetheless, the N2 production rate did not change. In organotrophic denitrification, the p-cresol oxidation rate was slower than ammonium oxidation. In litho-organotrophic cultures, the p-cresol and ammonium oxidation rates were affected at 133 mg/L NO2-N. Nonetheless, at 76 mg/L NO2-N the denitrifying sludge oxidized ammonium and p-cresol, but at different rate. Finally, this is the first work reporting the simultaneous oxidation of ammonium and p-cresol with the production of N2 from denitrifying sludge.  相似文献   

12.
Zhang J  Zhou J  Liu J  Chen K  Liu L  Chen J 《Bioresource technology》2011,102(7):4807-4814
The immediate precursor of L-ascorbic acid, or vitamin C, is 2-keto-l-gulonic acid (2-KLG). This is commonly produced commercially by Ketogulonicigenium vulgare and Bacillus megaterium, using corn steep liquor powder (CSLP) as an organic nitrogen source. In this study, the effects of the individual CSLP components (amino acids, vitamins, and metal elements) on 2-KLG production were evaluated, with the aim of developing a complete, chemically defined medium for 2-KLG production. Forty components of CSLP were analyzed, and key components were correlated to biomass, 2-KLG productivity, and consumption rate of L-sorbose. Glycine had the greatest effect, followed by serine, biotin, proline, nicotinic acid, and threonine. The combination of 0.28 g L−1 serine, 0.36 g L−1glycine, 0.18 g L−1 threonine, 0.28 g L−1proline, 0.19 g L−1nicotinic acid, and 0.62 mg L−1biotin in a chemically defined medium produced the highest maximum biomass concentration (4.2 × 109 cfu mL−1), 2-KLG concentration (58 g L−1), and yield (0.76 g g−1) after culturing for 28 h.  相似文献   

13.
The formation of aerobic granules with low organic loading synthetic wastewater (150-200 mg L−1 of influent COD, acetate/propionate = 1/3) at low aeration rate (0.6 cm s−1 of superficial gas velocity) had been investigated in the anaerobic/oxic/anoxic SBR. Aerobic granules with smooth surface and compact structure were successfully obtained after 50 days. However, these aerobic granules were unstable when the d(0.9) of granules increased to more than 1 mm. The results suggested that the aerobic granules with small diameter (smaller than 1000 μm) were more favorable for treating the low substrate loading wastewater at the low aeration rate. The cycle test revealed that most of the influent COD was removed at the anaerobic stage. The effluent concentrations of N-NH4+ and P-PO43− were lower than 1 mg L−1, and the effluent concentration of nitrate gradually decreased with the granulation. Phosphate accumulating organisms were found to utilize O2 or NOx as electron acceptor for phosphorus removal in the study. Simultaneous nitrogen and phosphorus removal occurred inside the granules.  相似文献   

14.
Cytidine (cyt) and adenosine (ado) react with cis-[L2Pt(μ-OH)]2(NO3)2 (L = PMe3, PPh3) in various solvents to give the nucleoside complexes cis-[L2Pt{cyt(− H),N3N4}]3(NO3)3 (L = PMe3, 1),cis-[L2Pt{cyt(− H),N4}(cyt,N3)]NO3 (L = PPh3, 2), cis-[L2Pt{ado(− H),N1N6}]2(NO3)2 (L = PMe3, 3) and cis-[L2Pt{ado(− H),N6N7}]NO3 (L = PPh3, 4). When the condensation reaction is carried out in solution of nitriles (RCN, R = Me, Ph) the amidine derivatives cis-[(PPh3)2PtNH=C(R){cyt(− 2H)}]NO3 (R = Me, 5a; R = Ph, 5b) and cis-[(PPh3)2PtNH=C(R){ado(− 2H)}]NO3 (R = Me, 6a: R = Ph, 6b) are quantitatively formed. The coordination mode of these nucleosides, characterized in solution by multinuclear NMR spectroscopy and mass spectrometry, is similar to that previously observed for the nucleobases 1-methylcytosine (1-MeCy) and 9-methyladenine (9-MeAd). The cytotoxic properties of the new complexes, and those of the nucleobase analogs, cis-[(PPh3)2PtNH=C(R){1-MeCy(− 2H)}]NO3 (R = Me, 7a: R = Ph, 7b), cis-[(PPh3)2PtNH=C(R){9-MeAd(− 2H)}]NO3 (R = Me, 8a: R = Ph, 8b) have been investigated in a wide panel of human cancer cells. Interestingly, whereas the Pt(II) nucleoside complexes (1-4) did not show appreciable cytotoxicity, the corresponding amidine derivatives (7a, 7b, 8a, 8b, 5b, and 6b) exhibited a significant in vitro antitumor activity.  相似文献   

15.
Wong BT  Lee DJ 《Bioresource technology》2011,102(12):6673-6679
The inhibitory effects of 90-189 mg l−1 of sulfide and 25-75 mg-N l−1 of nitrate on methanogenesis were investigated in a mixed methanogenic culture using butyrate as carbon source. In the initial phase of 90 mg l−1 S2− test, autotrophic denitrification of nitrate occurred with sulfide as the electron donor. Then the sulfate-reducing strains converted the produced sulfur back to sulfide via heterotrophic oxidation pathway. Methanogenesis was not markedly inhibited when 90 mg l−1 of sulfide was dosed alone. When 25-75 mg-N l−1 of nitrate was presented, initiation of methanogenesis was seriously delayed. Nitrogen oxides (NOx), the intermediates for nitrate reduction via denitrification pathway, inhibited methanogenesis. The 90 mg l−1 of sulfide favored heterotrophic dissimilatory nitrate reduction to ammonia (DNRA) pathway for nitrate reduction. Possible ways of maximizing methane production from an organic carbon-rich wastewater with high levels of sulfide and nitrate were discussed.  相似文献   

16.
The effect of non-ionic surfactants on the biofiltration of methane (CH4) was analyzed. Two biofilters (BF) treating CH4 were operated for one year at fixed CH4 concentration of 4.8 g m−3 and air flow rate of 0.25 m−3 h−1. Three polyoxyethylenes (Brijs), and 3 mono polyoxyethylenesorbitans (Tweens) were added to the nutrient solution at a concentration of 0.5% (w/w). Without surfactant, CH4 conversion had an average level of 35%, with Brijs the CH4 conversion varied between 38% and 46%, and with Tweens between 43% and 48%. The non-ionic surfactants decreased the biomass accumulation in the packed bed due to their detergent character. Biofilters were operated in a range of nitrogen concentration in the nutrient solution from 0.5 to 2 gN L−1 using Tween 20 at a concentration of 0.5% (w/w). The ECmax observed in this study, 45 g m−3 h−1, occurred when the nitrogen concentration was 1 gN L−1.  相似文献   

17.
The complexes of 2-[2-(diphenylphosphoryl)prop-2-yl]-1,8-naphthyridine (L) with lanthanide nitrates Ln(NO3)3 (Ln = Nd, Eu, Lu) were investigated to elucidate the coordination ability of a novel type of potentially tridentate ligands - phosphorylalkyl substituted naphthyridines. The X-ray crystal structures of [NdL3]3+ · 3(NO3) · MeCN (1), [EuL3]3+ · 3(NO3) · [Eu(NO3)3 · 4H2O] · MeCN (2), and [LuL3]3+ · 3(NO3) · [Lu(NO3)3 · 3H2O] · 2 MeCN · 0.5 H2O (3) are reported together with their IR and Raman spectra. All the compounds studied contain isostructural [LnL3]3+ cations and three NO3 counterions. Coordination of each L appears to be O,N,N tridentate-cyclic and coordination number of Ln is nine. Vibrational spectra of 1-3 are also compared with that of free ligand and model compounds.  相似文献   

18.
Salt marshes near urban, industrial and mining areas are often affected both by heavy metals and by eutrophic water. The aim of this study was to assess and evaluate the main processes involved in the decrease of nitrate concentration in pore water of mine wastes flooded with eutrophic water, considering the presence or absence of plant rhizhosphere. Basic (pH ∼ 7.8) carbonated loam mine wastes and free-carbonated acidic (pH ∼ 6.2) sandy-loam mine wastes were collected from polluted coastal salt marshes of SE Spain which regularly receive nutrient-enriched water. The wastes were put in pots and flooded for 15 weeks with eutrophic water (dissolved organic carbon ∼26 mg L−1, PO43− ∼23 mg L−1, NO3 ∼180 mg L−1). Three treatments were assayed for each type of waste: pots with Sarcocornia fruticosa, pots with Phragmites australis and unvegetated pots. Soluble organic carbon, nitrate, soluble Cd, Pb and Zn, pH and Eh were monitored. But the 2nd day of flooding, nitrate concentrations had decreased between 70% and 90% (equivalent to 1.01-1.12 g N-NO3 m−2 day−1) with respect to the content in the water used for flooding, except in unvegetated pots with acidic wastes. Denitrification was the main mechanism associated with the removal of nitrate. The role of vegetation in improving the rhizospheric environment was relevant in the acidic wastes because higher sand content, lower pH and higher soluble metal concentrations might strongly hinder microbial activity Hence, revegetation of salt marshes polluted by acidic sandy mining wastes might improve the capacity of this type of environment to act as a green filter against excessive nitrate contents flowing through them.  相似文献   

19.
A mixed bacterial culture was acclimated to the removal of high nitrate-N concentrations (100–750 mg NO3 -N L−1) from salty wastewaters. The experiments were carried out under anoxic conditions in the presence of 0.5, 1.5 and 3% (w/v) NaCl at different temperatures. The acclimated mixed bacterial culture was attached to quartz sand and zeolite. Denitrification was monitored in a continuous-flow bioreactor at different hydraulic retention times (HRT). Nitrate removal with cells attached to quartz sand and zeolite was completed at HRT of 167 h and 25 h respectively. Then brine denitrification with bacterial cells attached to zeolite was monitored for 85 days. Under the increased nitrate loading rate, nitrate removal was above 90%. Furthermore, during denitrification, not more than 0.5 mg NO2 -N L−1 could be produced. It can be concluded that nitrate removal with the cells attached to zeolite is economically and operationally a promising solution to denitrification of brine wastewaters.  相似文献   

20.
Reservoirs are intrinsically linked to the rivers that feed them, creating a river–reservoir continuum in which water and sediment inputs are a function of the surrounding watershed land use. We examined the spatial and temporal variability of sediment denitrification rates by sampling longitudinally along an agriculturally influenced river–reservoir continuum monthly for 13 months. Sediment denitrification rates ranged from 0 to 63 μg N2O g ash free dry mass of sediments (AFDM)−1 h−1 or 0–2.7 μg N2O g dry mass of sediments (DM)−1 h−1 at reservoir sites, vs. 0–12 μg N2O gAFDM−1 h−1 or 0–0.27 μg N2O gDM−1 h−1 at riverine sites. Temporally, highest denitrification activity traveled through the reservoir from upper reservoir sites to the dam, following the load of high nitrate (NO3-N) water associated with spring runoff. Annual mean sediment denitrification rates at different reservoir sites were consistently higher than at riverine sites, yet significant relationships among theses sites differed when denitrification rates were expressed per gDM vs. per gAFDM. There was a significant positive relationship between sediment denitrification rates and NO3-N concentration up to a threshold of 0.88 mg NO3 -N l−1, above which it appeared NO3-N was no longer limiting. Denitrification assays were amended seasonally with NO3-N and an organic carbon source (glucose) to determine nutrient limitation of sediment denitrification. While organic carbon never limited sediment denitrification, all sites were significantly limited by NO3-N during fall and winter when ambient NO 3-N was low.  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-23

京公网安备 11010802026262号