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 novel process was developed to induce a simultaneous oxidation of ammonia and denitrification in a single system consisting of two chambers separated by a cation exchange membrane. One was an anoxic chamber and the other was an aerobic chamber. The maximum mass flux via the membrane was calculated as 0.83 mg NH4+-N/m2 s in a batch test when the initial concentration of NH4+ was 700 mg N/L. And it was observed that NO3 and NO2 moved via the membrane in a reverse direction when NH4+ was transported. When the system was operated in a continuous mode by feeding a wastewater containing glucose and NH4+, it was observed that soluble chemical oxygen demand and NH4+ were simultaneously removed showing 99% and 71  86% of efficiency, respectively. Denitrification occurred in the anoxic chamber and nitrification was carried out in the aerobic chamber.  相似文献   

2.
The objective of this study is to develop a technique to remove ammonium ion from water intended for potable purposes. An ion exchange method is used with a selective ion exchanger, a natural cation zeolite, clinoptilolite. Glass columns (Fig. 1) are used for laboratory experiments. These experiments show that the NH4+ exchange capacity is very small compared to its total capacity 2.17 meq g−1; its value depends essentially on the NH4+ initial concentration and less on the Ca2+ concentration in the influent water. Figure 3 illustrates the practical exchange capacity relative to the initial concentration of ammonium ion for a soft water (Ca2+ = 35–50 mg l−1). We were particularly interested in waters weak in ammonium ion concentration (NH4+ = 1–3 mg l−1). In this case and for 1 and 2 mg l−1 NH4+ concentration in water, the practical capacity is only 0.06 and 0.108 meq g−1 respectively. The leakage is smaller than the ECC limit (European Community Council) for drinking waters (NH4+ 0.5 mg l−1) and the treated volume of water to breakthrough, defined at 0.5 mg l−1 of NH4+, is 720 BV (BV = bed volume) in both cases.In another way Fig. 6 shows that hard waters (due to Ca2+ ions) are more difficult to treat than soft waters. The practical capacity is smaller than before and the NH4+-leakage is greater. To lessen NH4+-leakage to less than 0.5 mg l−1 for soft waters down-flow and up-flow, regeneration is used. Figure 7 shows that up-flow regeneration is more attractive than down-flow regeneration.Cycle reproducibility (Figs 4 and 5) shows that the regeneration conditions satisfied our requirements: in this case, the salt consumption is 180 eq of salt per eq of NH4+ eliminated. This prompted us to try to reuse the regenerant (with NH4+ ion). An increase of NH4+-leakage is noticed in the presence of an NH4+-residual in the regenerant. This increase is more significant with down-flow regeneration.After these laboratory experiments, we carried out a semi-industrial pilot-plant. Our objective was first to verify the laboratory results and secondly to study clinoptilolite behaviour relative to the time it was used. Two plexiglass columns comprise the pilot-plant shown in Fig. 9; soft water is used for these experiments. The first column is regenerated with fresh salt solution. The cycles obtained, considering their initial NH4+-concentration, are reproduced in Fig. 10. For 2 mg l−1 NH4+ in the influent water, the leakage is about 0.2 mg l−1 and the treated volume to breakthrough (0.5 mg l−1 of NH4+) is about 750 BV. The second column is regenerated with a recycled solution. The quality of the cycles decreases with the number of reuse of the regenerant as shown in Fig. 11. Nevertheless, it is interesting to note that after 3 reuses, the performance decrease is only 25% and the leakage, although it increases is smaller than 0.5 mg l−1.Pilot results allowed us to propose a treatment of 30,000 m3 day−1; the cost per cubic meter water treated, relative to NH4+-removal, is about 0.165 FF (0.033 US $) for a plant and 0.77 FF (0.014 US $) for the same plant at the seaside. Using two serial columns decreased the cost by about 40–50%.  相似文献   

3.
Attached-cell reactors using a bed of granular material for wastewater treatment develop a high biomass concentration which allows an important reduction of the required residence time (Jeris et al., 1977; Elmaleh, 1982). In nitrification of ammonia containing wastewater, oxygen is currently the limiting substrate; in theory, 4.18 g of oxygen are required per 1 g of nitrogen (Painter, 1970). Oxygen can be added with hydrogen peroxide (Grigoropolou, 1980; Seropian, 1980; Yahi et al., 1982) which is nevertheless expensive and it seems better to transfer oxygen from a gas phase, i.e. air, to the liquid phase through a fixed bed (Charpentier, 1976).Two attached-cell reactors (Fig. 1) were operated in parallel for nitrification of ammonia containing synthetic wastewater (Table 2). Air was upflowed through a granular packing (Table 1) maintained in fixed bed while the liquid influent was injected at co- or counter-current.
1. (1) Owing to the high oxygen transfer properties of the system and to the fact that the thickness of biofilm is always less than 100 μm, the whole process was not limited by oxygen concentration of which remained larger than 7 mg l−1 (Fig. 2a) (Bungay et al., 1969). Oxidised nitrogen ammonia is completely converted into nitrate (Fig. 2b). Experimental conditions are given in Table 3.
2. (2) The plot of ammonia conversion against air superficial velocity shows a maximum (Fig. 3) after which conversion decreases rapidly by overloading of the packing (Prost, 1965). Experimental conditions are given in Table 4.
3. (3) Process efficiency decreases when superficial upflow velocity is increased (Fig. 4).
4. (4) Complete abatement of inlet pollution is reached when nitrogen concentration is less than 25 mg l−1 (Fig. 5) which corresponds to a volumetric loading up to 0.6 kg N (NH4+) m−3 day−1.
Moreover, the experimental data were fitted to a model based on classical assumptions (Roques, 1980; Grady, 1982; Atkinson and Fowler, 1974; Grasmick et al., 1979; Grasmick, 1982; Harremoes, 1976, 1978; Jennings et al., 1976; Williamson and MacCarty, 1976); i.e. zero order intrinsic kinetics and diffusion transport (Table 5), and recently developed (Grasmick, 1982; Rodrigues et al., 1984). This model provides, particularly, a very easy method to check its own use—in reaction regime and in diffusion regime—when time spans or inlet concentration are changed; experimental results can indeed be plotted in such a way that straight lines are obtained (Table 6). Figures 6 and 7 show the data obtained with the counter-current nitrification reactor when respectively inlet concentration and time spans are varied. The plotted straight lines show that the overall reaction is zero order and that, therefore, the biofilm is fully penetrated. A critical time span θc and a critical inlet concentration Cc, for which complete conversion is achieved, are then calculated, θc is theoretically proportional to C1 which is verified in Fig. 8. The straight line θc vs C1 can then be used in reactor design.  相似文献   

4.
Yu Tian  Yaobin Lu 《Water research》2010,44(20):6031-6040
Nutrient release is reported as one of the main disadvantage of sludge reduction induced by aquatic worm. In this study, a Static Sequencing Batch Worm Reactor (SSBWR) was proposed with novel structure of perforated panels, combined aeration system and cycle operation. Effective simultaneous nitrification and denitrification were obtained owing to the stratified sludge layer containing aerobic and anoxic microzone formed on each carrier during most of the operation time in the SSBWR, which created suitable conditions for remarkable sludge reduction and nutrient removal. The results showed that the total nitrogen (TN) concentration, NO3?–N + NO2?–N concentration and NH4+–N release could be reduced by 67.5%, 98.5% and 63.0%, respectively. And the soluble chemical oxygen demand (sCOD) released by sludge predation was also proved to provide a carbon source for denitrification leading to carbon release control and substantial cost savings. A schematic diagram of the stratified sludge layer and the mass balance of the nitrification–denitrification cycle were given, providing further insight into the nutrient (sCOD and nitrogen compounds) transformation during the worm predation in the SSBWR. For the mixed sludge liquid of 3000 mg TSS/L, 30 mg/L sCOD and 40 mg/L NO3?–N, the NO3?–N and NO2?–N came close to zero, and the sludge concentration, NH4+–N release and sCOD release was reduced by 33.6%, 63.0% and 72.5%, respectively, during 48 h’ predation.  相似文献   

5.
The pathways of N in aerobic farm waste treatment systems are discussed in relation to the dissolved oxygen (DO) and pH of the mixed liquor. The change in pH, DO, oxygen uptake rate and nitrogen balance were monitored under steady, and non-steady, state conditions in an oxidation ditch treating undiluted pig waste. A kinetic analysis of the mass balance for nitrogen allowed an interpretation of the fate of nitrogen under different prevailing conditions. Undesirable accumulations of nitrite were noted in the presence of high levels of free NH3 and HNO2. The process was self-promoting and was encouraged by the influx of raw waste. Concentrations of 500 mg 1−1 NO2-N and 1200 mg 1−1 NO3-N were the maximum values observed and were considered to be the concentrations at which product inhibition arrested nitrifying activity. Attainment of these levels prevented complete nitrification despite an adequate retention time. pH and DO were inversely related probably through nitrification, but pH appeared to be lowered by accumulation of nitrite and nitrate anions, and thus by the balance between nitrification and denitrification. Considerable N loss through denitrification was found to occur despite apparently aerobic mixed liquors. At low DO simultaneous nitrification-denitrification could eliminate 90 per cent of the soluble-N. NH3 desorption in laboratory cultures was found to be first order in free NH3 but was not a significant mode of N loss under field conditions.  相似文献   

6.
Previous studies conducted at Castle Lake, California suggested that phytoplankton growth could be limited by low levels of dissolved inorganic nitrogen during the summer growing season. We have used the isotopes H14CO3, 13NH4+, 15NO3, and 13NO3 as tracers to determine the rates of inorganic carbon and nitrogen uptake by the natural phytoplankton community and their response to nitrogen enrichment. Experimental designs followed the procedures of several bioassay techniques which have been developed to investigate algal nutrient deficiency.The results indicate that the phytoplankton were nitrogen deficient throughout the growing season. Ammonium uptake was rate-limited by the external concentration of ammonium. Nitrate uptake was regulated by the availability of nitrate and molybdenum, and potentially by the assimilation of ammonium. Although N-enrichment enhanced photosynthesis in short-term experiments during the early portion of the growing season, the photosynthetic response was initially negative during midseason, and a positive response did not occur until 2–3 days had elapsed. This suggests that a lack of stimulation, or even inhibition, of inorganic carbon uptake upon nitrogen enrichment does not necessarily preclude N-deficiency.  相似文献   

7.
Overland flow treatment of municipal and industrial wastewater has been proposed as an economical and effective method of removing pollutants. Properly designed and manipulated nitrification-denitrification in this technique could remove a significant amount of N.Applications of wastewater containing NH4 −N to a simulated overland flow model led to the disappearance of NH4+ −N and the formation of nitrate. The N balance in simulated overland flow system was estimated by using labeled 15N. The amount of N removed in the system depends upon denitrification rates. The results of this study indicated N was absorbed by soil and applied NH4+ −N was assimilated by the vegetation. The absorbed NH4+ −N on the aerated surface soil mass was nitrified and converted to oxidized forms of N. The nitrate formed diffused downward to the reduced zone during subsequent wastewater applications. Some of this nitrate was denitrified to gaseous forms of N or was reduced to organic forms by assimilatory processes. Thus, the net loss of N in an overland flow system was less than would have been predicted from non-labeled N mass balance calculations.  相似文献   

8.
An analysis of long-term historical records of the concentrations of major ions, TDS and nutrients for 20 river sites in Latvia is reported. Periods of water quality observations ranged from 15 to 43 years. A study of the quarterly adjusted time series showed that characteristic features of the data are non-normal distributions, seasonality, serial correlation and presence of significant trends that are mostly positive. The application of state-of-the-art software, based on non-parametric statistics such as the Seasonal Kendall slope estimator and the Seasonal Hodges-Lehmann estimator, made it possible to investigate these water quality records more accurately than other methods allow.Typical seasonal variations and concentration-discharge relationships were analyzed for different constituents. It was shown that fertilizer application and marsh land reclamation can cause widespread and intensive river water quality changes. Concentration increases of as much as 5–10 times that of background values were detected for NO3, Cl, Na+ + K+ and SO42−. The main water quality changes took place in the 1960s and the early 1970s when fertilizer applications and reclamation works increased. After that, concentration increases for constituents other than NO3 and SO42− were statistically insignficiant. The significant increases for NO3 and SO42− were probably due to the additional impact of increased atmospheric deposition. The results of long-term changes of river loads entering the Baltic Sea and the Gulf of Riga from Latvian territory are examined.  相似文献   

9.
Ion-exclusion chromatography of PO43− in wastewaters has been investigated on a cation exchange resin in the H+-form. Phosphate ion was chromatographed by ion-exclusion as H3PO4 formed by cation exchange, and monitored with a flow coulometric detector based on the electrochemical reaction of H+ ion from H3PO4 and p-benzoquinone at a constant applied potential (+0.45 V vs Ag-AgI) or a conductometric detector. A reasonable separation of PO43 from strong acid anions (Cl SO42− and NO3) and CO32− as coexisting anions could be accomplished by isocratic elution with 60% (v/v) acetone-water. The calibration graph for PO43− with the flow coulometric detector was linear over the concentration range 1–150 ppm (slope = 0.982), but not with the conductometric detector. The agreement of the analytical results of PO43− between the ion-exclusion chromatography with the flow coulometric detector and the colorimetry (molybdenum-blue method) was excellent for some industrial wastewater and domestic sewage samples.  相似文献   

10.
Using bulk deposition, throughfall, stemflow, soil infiltration, runoff water, litterfall data, ion mass budgets were calculated for a catchment area and for mature spruce and pine stands on it. The ions considered in mass balances were Na+, K+, Ca2+, Mg2+, SO42−, NO3, NH4+, HCO3, and H+. Corresponding fluxes for the budgets were calculated as an average for 6 years of studies (1995–2000). Annual input–output balances of all nutrients were positive at the plot-scale, so that leaching into soil water was less than the corresponding deposition load. Deposition of Ca, Mg, Na and S into soil by precipitation exceeded input through litterfall. A proton budget approach shows that the main soil buffering process is retention of sulphate, which clearly exceeds weathering. At the catchment-scale, input–output analysis shows essential output of cations due to weathering from the soil. A distinct change in input–output balance of sulphate during study period was evident. The retention of sulphur has been replaced by its release from the catchment area.  相似文献   

11.
Activated sludge from a domestic sewage works was enriched with nitrifying bacteria by running a laboratory fermenter on ammonia-supplemented sewage. This enriched culture was used to determine respirometrically the kinetics of microbial nitrification. It was demonstrated that the reaction fits the Michaelis-Menten model for temperatures from 10 to 35°C, having a temperature optimum at 15°C (K3 0.72 mg 1−1 NH3). Nitrification is unaffected by high dissolved oxygen concentration 38 mg 1−1 O2 at 30°C) after acclimatisation. Nitrite concentrations > 20 mg 1−1 are inhibitory to the reaction.  相似文献   

12.
Factors affecting the denitrification rate in two water-sediment systems   总被引:1,自引:0,他引:1  
Effects of temperature, oxygen and nitrate concentrations of the overlying water, and the thickness of the sediment layer on the rate of denitrification in the sediment were investigated in two water-sediment systems, A and B. At 4°C, denitrification started after a prolonged lag period in contrast to nitrification which did not occur significantly. At 15°C, and particularly at 25°C, both processes proceeded readily. The disappearance of NO2 - N from the overlying water was more rapidly than that of NO3 - N.The denitrification rate was slightly reduced by increasing the dissolved oxygen concentration in the overlying water from 0 to approximately 2 mgl−1. A further rise of the dissolved oxygen concentration had no further decreasing effect on the denitrification rate.The denitrification rate in sediment was dependent on the nitrate concentration in the overlying water approximating first order kinetics at lower concentrations, gradually becoming independent of the nitrate concentration at higher nitrate contents (zero order kinetics).When starting with a nitrate-nitrogen concentration of 25.2 mgl−1, a sediment layer of 7 mm with A and 14 mm with B was roughly found to be involved in denitrification.Denitrification rates found in the present laboratory experiments were supposed to be considerably lower than those occurring under natural conditions as additional mechanisms for the transport of nitrate into sediments occurred in natural environments.  相似文献   

13.
Continuous flow stirred reactors were used to evaluate the maximum denitrification specific removal rates for influent solutions made from NH4NO3, CaNO3, KNO3 and UO2 fuel fabrication waste water. Nitrate substrate concentrations ranged from 0.01 to 20 kg NO3/m3. Values for Umax (maximum specific substrate removal rate per unit mass of microorganisms per unit time, days−1) were determined using graphical solutions to the Lineweaver-Burk equations. For NH4NO3 solutions at nitrate substrate concentrations <6 kg NO3/m3 the value for Umax was found to be 1.73 days−1. At nitrate substrate concentrations >6 kg NO3/m3 a nonlinear relationship was observed in the Lineweaver-Burk plots indicating nitrate substrate inhibition. Specific removal rates at nitrate concentrations >6 kg NO3/m3 averaged <1.0 days−1. Ammonia toxicity may also have occurred as the pH of the mixed liquor was near 8. Methanol concentrations as high as 11.6 kg CH3OH/m3 did not inhibit denitrification rates. The highest specific removal rates recorded (3.13 ± 0.56 days−1) were with influents made from UO2 fuel fabrication waste water.  相似文献   

14.
Influence of high NaCl and NH4Cl salt levels on methanogenic associations   总被引:3,自引:0,他引:3  
The effect of high levels of NaCl and NH4Cl on the activity and attachment of methanogenic associations in semi-continuous flow-through reactor systems has been evaluated. Two well-functioning reactors received shock concentrations of NaCl and NH4Cl while two other reactors were adapted to increasing levels of the salts during a period of 45 days. The methanogenic associations, grown on a medium containing mainly acetate and ethanol, were found to be more resistant to NaCl and NH4Cl than previously reported. Initial inhibition occurred at shock treatments of 30 gl−1 for both salts. The reactors which were gradually exposed to increasing levels of the salts, adapted well and their tolerance levels surpassed those of the non-trained counterparts. Initial inhibition and fifty percent inhibition was observed at 65 and 95 gl−1 respectively for adaptation to NaCl. Initial inhibition for the reactor adapting to NH4Cl occurred at 30 gl−1 and a 50% inhibition was observed at 45 gl−1 of NH4Cl. For the reactors receiving NH4Cl, the free ammonia-N should be kept below a concentration of 80–100 mg l−1 for optimal performance. The bacterial populations in the reactors consisted mostly out of Methanosarcina (> 99% of the biomass)  相似文献   

15.
The export and concentration of inorganic nitrogen and total phosphorus from 34 watersheds in a northwestern Iowa lake district were measured during March 1971–August 1973. Annual nutrient losses were approximately 0.35 kg ha−1 P. 6.7 kg ha−1 NO3-N, and 1.0 kg ha−1 NH3-N. A statistical analysis of the relationship between land-use and plant nutrients was used to determine differences among streams. Animal units in feedlots were significantly correlated with phosphorus and ammonia nitrogen (mg l−1 and kg ha−1 yr−1). Nitrate nitrogen was negatively correlated with the percentage of watershed in marshland. Tile drainage and surface runoff from grasslands, feedlots, cornfields, and soybean fields were analyzed for nitrogen and phosphorus in spring 1974: mean values are given.  相似文献   

16.
Previous experiments carried out with the laboratory TOD meter Ionics 225 of the DOW Chèmical made it possible (after a high temperature catalytic action) to characterize the stable forms of organic and inorganic carbon and nitrogen (NH4+, NO2, NO3), and the principal cations (Na+, K+, Ca2+, Mg2+) in the course of the total oxygen demand (TOD) measurement.The object of this study is firstly to compare the oxidation capability of different techniques of organic pollution (particularly the COD and TOD) in relation to the constituent elements of the organic matter C, N, P, S, and to calculate the possible interferences of the inorganic compounds at the time of TOD test.These investigations warrant the application of this technique to measure the amount of organic pollution in relatively mineralized conditions (Industrial wastewater, sea-water…). The present publication is concerned more with the study of the transformation of the organic and inorganic sulphur forms (S2−, SO32−. SO42−) in the course of the TOD measurement.The study of the oxidizability of the organic sulphur compound type CxHyOzS, made it possible to establish a specific relation with a ratio of 0–50 mg of organic sulphur l−1, between the oxygen demand of this element [TOD (S)] and its concentration (TOD (S) = 0.97 [S]).These tests showed a partial oxidation of the sulphur to SO2 and SO3 as the literature claimed. On the other hand, the oxidation of the same compounds during the COD tests varies greatly and although it is not possible to establish a correlation between these two measurements, as applies in the case of organic nitrogen, nevertheless these experiments showed a greater reliability of the TOD compared with the COD in the oxidation of organic matter in general. We then carried out experiments on the different mineral forms of sulphur in order to distinguish the possible effects and to recommend simple improvements.A relative study on sulphate ions had been carried out with standard solutions which have the same TOD (the basic TOD is obtained using potassium phthalate acid) and the same increasing concentration of the salt M2SO4 type. The experiments showed that the basic TOD decreases when the concentration of sulphate ions is increased (Fig. 3). Therefore, the interference is negative and taking into consideration the specific oxygen demand of the cation, we can propose an evaluation of this interference (ΔTOD (SO42−) = 0.203 [SO42−]). The same experiments have been conducted with a salt of M2SO3 type and similar results obtained (Fig. 5).The specific interference of the sulphite ion is negative and can be estimated by the following equation (ΔTOD (SO32−) = 0.132 [SO32−]). In both cases, we have to note that the transformation of these inorganic anions occurs between those relative to the theoretical dissociation reaction corresponding to the appearance of the oxide SO2 and SO3. For sulphurous on the contrary, the interference is positive and therefore corresponds to an extra oxygen demand (Fig. 8).The experiments were conducted directly with the M2S salts (M representing K or Na) in aqueous solution.The evaluation of this interference had been made in the consideration of two concentration ranges of the sulphurous ions (0–35 mg S2− l−1): TOD (S2−) = 0.4 [S2−] and (35–100 mg S2− l−1): TOD (S2−) = 1.2 [S2−] − 30.Therefore this study confirms a better oxidation of the organic matter by TOD test in comparison with COD test.But sulphate and sulfite have a negative interference in the TOD measurement, whereas sulphurous is positive.The evaluation model of these interferences allows a correction to be made of the TOD value or to verify TOD measurement of organic pollution obtained by this technique.  相似文献   

17.
Ozone reacts with free aqueous chlorine when present as hypochlorite ion (OCl) with a second order rate constant of 120 ± 15 M−1 s−1 at 20°C. About 77% of the chlorine reacts to produce Cl and 23% is oxidized to ClO3. No ClO4 is formed. Conversion of chlorine to monochloramine reduces the ozone reaction rate to 26 ± 4 M−1 s−1, independent of pH, NH2Cl is transformed quantitatively to NO3 and Cl by O3. Rate data for other chloramines are also presented. The direct reaction of ozone with chlorine accounts for a significant amount of the chlorine and ozone demand found when the two oxidants are used in combination under water works conditions.  相似文献   

18.
Rainwater was collected at the Portuguese west coast between September 2008 and September 2009, and analysed for pH, conductivity, and Cl, NO3, SO42−, and NH4+ concentrations. Results of rainwater chemical composition were compared with those obtained at the same site between 1986 and 1989. In both collection periods rainwater was predominantly (≈ 80%) associated to oceanic air masses. The rainwater concentration of H+ was in the same range as twenty years ago. A clear decrease of non sea salt sulphate (NSS-SO42−) was observed in 2008-2009 relatively to 1986-1989, not only in samples with origin in central and northern Europe, but also in samples from Atlantic Ocean and Mediterranean. This decrease indicates that SO2 emissions were reduced, which may be due to a lower content of sulphur in oil by-products. A decrease was also observed in NH4+ concentration in 2008-2009. On the contrary an increase of NO3 concentration was observed for samples of all origins in 2008-2009 relatively to 1986-1989, which was particularly high (more than 3 fold) for samples with origin in Atlantic Ocean, suggesting the incorporation of this ion by rainout at the sampling site. The contribution of local sources is indeed suggested by the moderate negative correlation of NH4+, NO3 and NSS-SO42− with rainwater volume. The high increase of NO3 concentration can be attributed to the increase of local vehicular and industrial emissions in the sampling area.  相似文献   

19.
Emission of nitrous oxide (N2O) during biological wastewater treatment is of growing concern since N2O is a major stratospheric ozone-depleting substance and an important greenhouse gas. The emission of N2O from a lab-scale granular sequencing batch reactor (SBR) for partial nitrification (PN) treating synthetic wastewater without organic carbon was therefore determined in this study, because PN process is known to produce more N2O than conventional nitrification processes. The average N2O emission rate from the SBR was 0.32 ± 0.17 mg-N L−1 h−1, corresponding to the average emission of N2O of 0.8 ± 0.4% of the incoming nitrogen load (1.5 ± 0.8% of the converted NH4+). Analysis of dynamic concentration profiles during one cycle of the SBR operation demonstrated that N2O concentration in off-gas was the highest just after starting aeration whereas N2O concentration in effluent was gradually increased in the initial 40 min of the aeration period and was decreased thereafter. Isotopomer analysis was conducted to identify the main N2O production pathway in the reactor during one cycle. The hydroxylamine (NH2OH) oxidation pathway accounted for 65% of the total N2O production in the initial phase during one cycle, whereas contribution of the NO2 reduction pathway to N2O production was comparable with that of the NH2OH oxidation pathway in the latter phase. In addition, spatial distributions of bacteria and their activities in single microbial granules taken from the reactor were determined with microsensors and by in situ hybridization. Partial nitrification occurred mainly in the oxic surface layer of the granules and ammonia-oxidizing bacteria were abundant in this layer. N2O production was also found mainly in the oxic surface layer. Based on these results, although N2O was produced mainly via NH2OH oxidation pathway in the autotrophic partial nitrification reactor, N2O production mechanisms were complex and could involve multiple N2O production pathways.  相似文献   

20.
A number of denitrifying bacteria were isolated from activated sludge and drinking water. These bacteria were tested for the synthesis of the dissimilatory nitrate reductase under aerobic conditions (dissolved oxygen concentration above 4 mg · l−1). The synthesis of this enzyme varied from total repression by oxygen in some bacteria, especially those isolated from drinking water, until a nearly non oxygen-repressed synthesis in other bacteria (strains 15 and N4). The effect of the dissolved oxygen concentration during growth of the bacteria on the synthesis of the dissimilatory nitrate reductase in cells of strain 15 was studied more extensively. A considerable repression of the enzyme synthesis was obtained when the dissolved oxygen concentration was relatively high (approx 15 mg·l−1). Addition of chlorate to the growth medium of strain 15 (using NH+4-N as nitrogen source) also resulted in a serious repression of the nitrate reductase synthesis during aerobic growth (dissolved oxygen above 4 mg·l−1). The dissimilatory nitrate reductase of aerobically grown cells of strains 15 and N4 was found to be mainly localized in the membrane fraction.  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