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 0 毫秒
1.
The behavior of dimethyl phthalate (DMP) from municipal solid waste (MSW) in the leachate and refuse of two simulated landfill bioreactors was compared. In one reactor, the leachate was circulated between a landfill and a methanogenic reactor, while the other reactor was operated using direct recirculation of the leachate. The results revealed that the original concentration of DMP in the refuse was approximately 3.3 μg g−1, and the concentration decreased greatly during decomposition of the waste in both reactors. The major loss of DMP from the landfill occurred in an active methanogenic environment in the later period, while the environment was acidic due to a high concentration of chemical oxygen demand (COD), volatile fatty acids (VFA), and contained a large volume of biologically degradable material (BDM) during the early stage. In addition, a high correlation was found between the residual DMP concentrations and the BDM of the refuse in both systems. Circulating the leachate between the landfill and a methanogenic reactor resulted in an increase in the biodegradability of MSW and the degree of waste stabilization. Furthermore, the removal of DMP was enhanced 14% in the landfill that was operated in conjunction with the methanogenic reactor when compared to the landfill in which there was direct leachate recirculation.  相似文献   

2.
The removal of phthalic acid diesters (PAEs) in municipal solid waste (MSW) from two simulated landfill reactors was compared. The results showed that the original concentrations of dimethyl phthalate (DMP), dibutyl phthalate (DBP) and dioctyl phthalate (DOP) in the refuse were 3.3 μg g−1, 18.5 μg g−1 and 0.8 μg g−1, respectively. The concentrations of DMP and DBP in both leachate and refuse decreased greatly during decomposition of the waste in both reactors. The major loss of PAEs from the landfill occurred during an active methanogenic environment with a low concentration of volatile fatty acids (VFA) in the later period. In addition, strong correlations were found between the residual DMP, DBP concentrations and the biologically degradable material (BDM) of the refuse. Finally, PAEs degraded more rapidly in the landfill that was operated in conjunction with the methanogenic reactor when compared to the landfill with direct leachate discharge.  相似文献   

3.
In this paper the well-known problem of optimal input design is considered. In particular, the focus is on input design for the estimation of kinetic parameters in bioreactors. The problem is formulated as follows: given the model structure (f,g), which is assumed to be affine in the input, and the specific parameter of interest theta;(k) find a feedback law that maximizes the sensitivity of the model output to the parameter under different flow conditions in the bioreactor and, possibly, minimize the input or state costs. Analytical solutions to these problems are presented. As an example a bioreactor with a biomass that grows according to the well-known Monod kinetics is considered.  相似文献   

4.
In this study, the anoxic oxidation of arsenite (As(III)) linked to chemolithotrophic denitrification was shown to be feasible in continuous bioreactors. Biological oxidation of As(III) was stable over prolonged periods of operation ranging up to 3 years in continuous denitrifying bioreactors with granular biofilms. As(III) was removed with a high conversion efficiency (>92%) to arsenate (As(V)) in periods with high volumetric loadings (e.g., 3.5–5.1 mmol As L day?1). The maximum specific activity of sampled granular sludge from the bioreactors was 0.98 ± 0.04 mmol As(V) formed g?1 VSS day?1 when determined at an initial concentration of 0.5 mM As(III). The microbial population adapted to high influent concentrations of As(III) up to 5.2 mM. However, the As(III) oxidation process was severely inhibited when 7.6–8.1 mM As(III) was fed. Activity was restored upon lowering the As(III) concentration to 3.8 mM. Several experimental strategies were utilized to demonstrate a dependence of the nitrate removal on As(III) oxidation as well as a dependence of the As(III) removal on nitrate reduction. The molar stoichiometric ratio of As(V) formed to nitrate removed (corrected for endogenous denitrification) in the bioreactors approximated 2.5, indicating complete denitrification was occurring. As(III) oxidation was also shown to be linked to the complete denitrification of NO to N2 gas by demonstrating a significantly enhanced production of N2 beyond the background endogenous production in a batch bioassay spiked with 3.5 mM As(III). The N2 production also corresponded closely to the expected stoichiometry of 2.5 mol As(III) mol?1 N2–N for complete denitrification. Biotechnol. Bioeng. 2010;105: 909–917. © 2009 Wiley Periodicals, Inc.  相似文献   

5.
A pure culture of Pseudomonas fluorescens was used as a model system to study the kinetics of denitrification. An exponentially growing culture was harvested and resuspended in an anoxic acetate solution buffered with K/Na phosphate at pH values of 6.6, 7.0, 7.4, and 7.8. The temperature was kept at 28 degrees C in all assays. Nitrate pulses of approximately 0.2 mg N/L caused nitrite to accumulate due to a faster rate of nitrate reduction over nitrite reduction. The rate of nitrate reduction was observed to depend on its concentration as predicted by the Michaelis-Menten equation. At nonlimiting nitrate concentrations, nitrite reduction was described by the same equation. Otherwise, nitrite reduction also depended on nitrate concentration. Consequently, nitrate and nitrite reductions compete with each other for the oxidation of common electron donors. A kinetic model for nitrate competitive inhibition of nitrite reduction is proposed. The model was used to interpret the nitrate and nitrite profiles observed at the four pH values: the optimum pH value was 7.0 in both cases; the affinity for nitrite was also not affected by the medium pH in the range of values 6.6 to 7.4 (K(mNO(3) ) = 0.04 mg N/L); the affinity for nitrite was also not affected by the medium pH in the range of values 6.6 to 7.4 (K(mNO(2) ) = 0.06 mg N/L), but it decreased sharply for the pH value of 7.8. Although the ratio between the two maximum reduction rates (V(max NO(2) )/V(max NO(3) )) is constant, nitrite accumulation depends on the medium pH value. Therefore, the regulation mechanism that shifts the electron flow between the two terminal reductases is readily reversible and does not change their relative maximum reduction rates. (c) 1995 John Wiley & Sons, Inc.  相似文献   

6.
Denitrification beds are a simple and relatively inexpensive technology for removing nitrate from point source discharges. To date, operational beds have used wood media as the carbon source, as it provides a sustained nitrate removal rate (2-10 g N m−3 of media d−1) while maintaining permeability. In pilot-scale (2.9 m−3) denitrification beds receiving municipal wastewater effluent dosed with KNO3, we looked at improving nitrate removal by using alternative carbon media (maize cobs) and increasing bed temperature through passive solar heating. The influence of flow regime (horizontal-point, horizontal-diffuse, downflow and upflow) on short-circuit flow was also investigated.The long-term nitrate removal rate (21.8 g N m−3 d−1) of the maize cob beds over the 15-month period of the trial was 2-11-fold higher than sustained removal rates reported by other researchers for wood-based beds. While passive solar heating raised the mean bed temperature by 3.4 °C, it did not cause a measurable increase in the nitrate removal rate due to the variability in the removal rate exceeding the expected increase due to temperature.Horizontal flow had more short-circuiting than vertical flow. Short-circuiting in the horizontal flow was attributed to flow being concentrated near the top surface due to the buoyancy effect of warmer water. Greater short-circuiting in the solar heated horizontal and upflow beds than in the corresponding unheated beds was attributed to the buoyancy effect being more pronounced in the solar heated beds.Overall, downflow was deemed the most effective of the four tested flow regimes. It provided the highest increase in bed temperature due to solar heating, had the highest nitrate removal rate in the latter part of the trial and had more plug-flow characteristics. While passive solar heating raised bed temperature, we were unable to demonstrate a significant difference (at 95% CL) in nitrate removal rate between the unheated and solar heated beds because of the high variability in nitrate removal rate and the increase in short-circuiting in the solar heated horizontal and upflow beds.  相似文献   

7.
The denitrifying bacterium Alcaligenes xylosoxidans synthesises two azurins (Az), which are termed Az I and Az 2. Both function as effective electron donors to copper nitrite reductase (NiR) in vitro. As a first step towards identifying the physiological relevance of these electron transfer proteins in the denitrification process, the gene (azuA) encoding Az I was characterised and its expression with respect to denitrification determined. We show that the azuA gene from A. xylosoxidans is monocistronic and its expression is increased when cells are grown under denitrifying conditions in the presence of nitrate or nitrite. The expression pattern of azuA was similar, though not identical, to that of the monocistronic nirK gene, which encodes copper NiR, and is in accord with both gene products being synthesised when the bacterium denitrifies. Recombinant Az I was exported to the periplasm of the heterologous host Escherichia coli, was synthesised at very high levels (80 mg purified protein per litre) and was fully loaded with copper. Electron donation from reduced recombinant Az to NiR was indistinguishable from the activity determined with the native protein. Taken together, these findings indicate that in A. xylosoxidans azuA expression is coordinated with denitrification and recombinant Az I is processed and matured in the periplasm of E. coli in the same way it is in A. xylosoxidans.  相似文献   

8.
9.
Cd1-型亚硝酸盐还原酶脱氮工程菌的构建与表达   总被引:1,自引:0,他引:1  
目的用基因工程技术将铜绿假单胞菌PAO1中nirS基因定向克隆至表达载体pQE-30上,使nirS基因得到高效表达。方法根据GenBank公布的nirS碱基序列和表达载体pQE-30的多克隆位点设计引物,以铜绿假单胞菌PAO1的基因组DNA为模板,应用PCR技术扩增目的片段nirS;之后经过BamH I和HindⅢ双酶切,定向克隆到pQE-30上,化学转化DH5α,构建含有重组质粒的转化子pQE30-nirS-DH5α;经酶切和测序鉴定,扩增产物的碱基序列与GenBank公布的序列完全吻合,再将重组质粒pQE30-nirS转化表达菌株SG13009构建脱氮基因工程菌pQE30-nirS-SG-13009(PNS)。最后用SDS-PAGE(IPTG浓度:0.05mmol/L;诱导温度:25℃;诱导时间:2~3h)和His-tag in-gel Stain鉴定cd1-型亚硝酸盐还原酶的分子量与特异性。结果构建了高效表达cd1-型亚硝酸盐还原酶的脱氮基因工程菌PNS。结论Cd1-型亚硝酸盐还原酶能够在脱氮基因工程菌PNS中得到正确和高效的表达。  相似文献   

10.
Presently, the wastewater treatment practices can be significantly improved through the introduction of new microbial treatment technologies. To meet increasingly stringent discharge standards, new applications and control strategies for the sustainable removal of ammonium from wastewater have to be implemented. Partial nitrification to nitrite was reported to be technically feasible and economically favorable, especially when wastewater with high ammonium concentrations or low C/N ratios is treated. For successful implementation of the technology, the critical point is how to maintain partial nitrification of ammonium to nitrite. Partial nitrification can be obtained by selectively inhibiting nitrite oxidizing bacteria through appropriate regulation of the system’s DO concentration, microbial SRT, pH, temperature, substrate concentration and load, operational and aeration pattern, and inhibitor. The review addressed the microbiology, its consequences for their application, the current status regarding application, and the future developments.  相似文献   

11.
The great spatial and temporal variability of nitrogen (N) processing introduces large uncertainties for quantifying N cycles in large scales, e.g. a watershed scale, and hence challenges the present techniques in measuring ecosystem N mass balance. The dual isotopes of nitrate (δ18O and δ15N) integrate signals for both nitrate sources and N processing, making them promising for studies on large scale N cycling. Here, the dual isotopes, as well as some ion tracers, from a subtropical river in south China were reported to identify the main nitrate sources and to assess the possible occurrence and degree of denitrification in the context of monsoon climate. Our results indicated that nitrification of reduced fertilizer N in soil zones was the main nitrate source, with sewage and manure as another important source in dry winter. Seasonal changes of denitrification was apparent by the ~1:2 enrichment of 18O and 15N from April to August, and suggested to occur over the watershed rather than in the river. The lowest denitrification (10%) occurred in April, when the fertilizer application was strongest and the monsoon rainfall abruptly increased, causing enhancement of leaching. The highest denitrification (48%) took place in August due to the high soil temperature and moisture. In December, denitrification was significant (26%) perhaps due to the high enough temperature for microbial activities, whereas the low soil moisture appeared to limit the degree of denitrification. This study suggests that the seasonal variations in denitrification should be taken into account when estimating regional N mass balance.  相似文献   

12.
In batch experiments, we studied the isotope fractionation of nitrogen and oxygen during denitrification of two bacterial strains (Azoarcus sp. strain DSM 9056 and Pseudomonas pseudoalcaligenes strain F10). Denitrification experiments were conducted with succinate and toluene as electron donor in three waters with a distinct oxygen isotope composition. Nitrate consumption was observed in all batch experiments. Reaction rates for succinate experiments were more than six times higher than those for toluene experiments. Nitrogen and oxygen isotopes became progressively enriched in the remaining nitrate pool in the course of the experiments; the nitrogen and oxygen isotope fractionation varied between 8.6–16.2 and 4.0–7.3‰, respectively. Within this range, neither electron donors nor the oxygen isotope composition of the medium affected the isotope fractionation process. The experimental results provide evidence that the oxygen isotope fractionation during nitrate reduction is controlled by a kinetic isotope effect which can be quantified using the Rayleigh model. The isotopic examination of nitrite released upon denitrification revealed that nitrogen isotope fractionation largely follows the fractionation of the nitrate pool. However, the oxygen isotope values of nitrite are clearly influenced by a rapid isotope equilibration with the oxygen of the ambient water. Even though this equilibration may in part be due to storage, it shows that under certain natural conditions (re-oxidation of nitrite) the nitrate pool may also be indirectly affected by an isotope equilibration.  相似文献   

13.
14.
The effect of different electron donors on the pathway and kinetics of nitrate reduction in a sulfide-acclimated mixed, mesophilic (35 degrees C) methanogenic culture was investigated. A mixture of dextrin and peptone, glucose, propionate, acetate, and H(2)/CO(2) were used as substrates at an initial chemical oxygen demand of 1,500 mg/L and the initial nitrate concentration ranged between 0 and 300 mg N/L. The fastest nitrate reduction was observed in the H(2)/CO(2) and acetate-fed cultures. In the case of propionate, nitrate reduction was the slowest followed by partial recovery of methanogenesis and accumulation of volatile fatty acids due to inhibition as a result of accumulation of denitrification intermediates. Similarly, accumulation of nitrite and nitric oxide and partial or complete inhibition of methanogenesis was observed in the H(2)/CO(2)-fed cultures. Methanogenesis completely recovered in the dextrin/peptone-, glucose-, and acetate-fed cultures at all nitrate levels. Denitrification was the dominant pathway of nitrate reduction in the propionate-, acetate-, and H(2)/CO(2)-fed cultures regardless of the COD/N value. However, both denitrification and dissimilatory nitrate reduction to ammonia (DNRA) were observed in the dextrin/peptone- and glucose-fed cultures and the degree of predominance of either of the two pathways was a function of the COD/N value. Therefore, the type of electron donor used affected both the nitrate reduction pathway and kinetics, as well as the recovery of fermentation and/or methanogenesis in the mixed methanogenic culture.  相似文献   

15.
16.
Denitrification beds are a simple approach for removing nitrate (NO3) from a range of point sources prior to discharge into receiving waters. These beds are large containers filled with woodchips that act as an energy source for microorganisms to convert NO3 to nitrogen (N) gases (N2O, N2) through denitrification. This study investigated the biological mechanism of NO3 removal, its controlling factors and its adverse effects in a large denitrification bed (176 m × 5 m × 1.5 m) receiving effluent with a high NO3 concentration (>100 g N m−3) from a hydroponic glasshouse (Karaka, Auckland, New Zealand). Samples of woodchips and water were collected from 12 sites along the bed every two months for one year, along with measurements of gas fluxes from the bed surface. Denitrifying enzyme activity (DEA), factors limiting denitrification (availability of carbon, dissolved organic carbon (DOC), dissolved oxygen (DO), temperature, pH, and concentrations of NO3, nitrite (NO2) and sulfide (S2−)), greenhouse gas (GHG) production - as nitrous oxide (N2O), methane (CH4), carbon dioxide (CO2) - and carbon (C) loss were determined. NO3-N concentration declined along the bed with total NO3-N removal rates of 10.1 kg N d−1 for the whole bed or 7.6 g N m−3 d−1. NO3-N removal rates increased with temperature (Q10 = 2.0). In laboratory incubations, denitrification was always limited by C availability rather than by NO3. DO levels were above 0.5 mg L−1 at the inlet but did not limit NO3-N removal. pH increased steadily from about 6 to 7 along the length of the bed. Dissolved inorganic carbon (C-CO2) increased in average about 27.8 mg L−1, whereas DOC decreased slightly by about 0.2 mg L−1 along the length of the bed. The bed surface emitted on average 78.58 μg m−2 min−1 N2O-N (reflecting 1% of the removed NO3-N), 0.238 μg m−2 min−1 CH4 and 12.6 mg m−2 min−1 CO2. Dissolved N2O-N increased along the length of the bed and the bed released on average 362 g dissolved N2O-N per day coupled with N2O emission at the surface about 4.3% of the removed NO3-N as N2O. Mechanisms to reduce the production of this GHG need to be investigated if denitrification beds are commonly used. Dissolved CH4 concentrations showed no trends along the length of the bed, ranging from 5.28 μg L−1 to 34.24 μg L−1. Sulfate (SO42−) concentrations declined along the length of the bed on three of six samplings; however, declines in SO42− did not appear to be due to SO42− reduction because S2− concentrations were generally undetectable. Ammonium (NH4+) (range: <0.0007 mg L−1 to 2.12 mg L−1) and NO2 concentrations (range: 0.0018 mg L−1 to 0.95 mg L−1) were always very low suggesting that anammox was an unlikely mechanism for NO3 removal in the bed. C longevity was calculated from surface emission rates of CO2 and release of dissolved carbon (DC) and suggested that there would be ample C available to support denitrification for up to 39 years.This study showed that denitrification beds can be an efficient tool for reducing high NO3 concentrations in effluents but did produce some GHGs. Over the course of a year NO3 removal rates were always limited by C and temperature and not by NO3 or DO concentration.  相似文献   

17.
Abstract The production of nitrogen-containing gases by denitrification in three organisms was examined using membrane inlet mass spectrometry. The effects of O2 (during both growth and maintenance) and of pH, nitrate concentration and carbon source were tested in non-proliferating cell suspensions. Two strains of Pseudomonas aeruginosa were capable of co-respiration of NO3 and O2 and, under controlled O2 supply, gave oscillatory denitrification. Variations in culture and assay conditions affected both the rate of denitrification and the ratio of end products (N2O:N2). Higher rates were seen following anaerobic growth. Optimum values of pH and nitrate concentration for denitrification are given. Generally, the optimum pH was 7.0–7.5, approximately that of the growth medium. Optimum nitrate concentration was generally 20 mM.  相似文献   

18.
Produced water from the Coleville oil field in Saskatchewan, Canada was used to inoculate continuous up-flow packed-bed bioreactors. When 7.8 mM sulfate and 25 mM lactate were present in the in-flowing medium, H(2)S production (souring) by sulfate-reducing bacteria (SRB) was prevented by addition of 17.5 mM nitrate or 20 mM nitrite. Changing the sulfate or lactate concentration of the in-flowing medium indicated that the concentrations of nitrate or nitrite required for containment of souring decreased proportionally with a lowered concentration of the electron donor lactate, while the sulfate concentration of the medium had no effect. Microbial communities were dominated by SRB. Nitrate addition did not give rise to changes in community composition, indicating that lactate oxidation and H(2)S removal were caused by the combined action of SRB and nitrate-reducing, sulfide-oxidizing bacteria (NR-SOB). Apparently the nitrite concentrations formed by these NR-SOB did not inhibit the SRB sufficiently to cause community shifts. In contrast, significant community shifts were observed upon direct addition of high concentrations (20 mM) of nitrite. Strains NO3A and NO2B, two newly isolated, nitrate-reducing bacteria (NRB) emerged as major community members. These were found to belong to the epsilon-division of the Proteobacteria, to be most closely related to Campylobacter lari, and to oxidize lactate with nitrate or nitrite as the electron acceptor. Thus the mechanism of microbial H(2)S removal in up-flow packed-bed bioreactors depended on whether nitrate (SRB/NR-SOB) or nitrite (SRB/NR-SOB as well as NRB) was used. However, the amount of nitrate or nitrite needed to completely remove H(2)S was dictated by the electron donor (lactate) concentration, irrespective of mechanism.  相似文献   

19.
Abstract Denitrification in intact sediment cores was measured by the acetylene inhibition technique and compared with the nitrate flux between water and sediment. Less than half of the nitrate-N consumed by the sediment could be recovered as nitrous oxide-N. The low recovery rate of nitrous oxide from intact sediment cores indicated losses of nitrous oxide by diffusion down to nitrate-free sediment layers, where reduction of nitrous oxide may take place. In sediment slurries 100% of nitrate-N could be recovered as nitrous oxide-N as long as the nitrate concentration in the liquid phase was above 10 μM. Nitrous oxide added to nitrate-free sediment slurries was reduced regardless of whether acetylene was present or not. Therefore denitrification may be significantly underestimated by this method.  相似文献   

20.
Cells of the non-N2-fixing cyanobacteriumPhormidium laminosum were immobilized in polyurethane (PU) foams either by absorption or by entrapment in the PU prepolymer followed by polymerisation and by adsorption onto polyvinyl (PV) foams. Although entrapment caused toxicity problems which lead to rapid death of the immobilized cells, they were immobilized successfully by adsorption onto PU or PV foams and maintained their photosynthetic electron transport activities (PS I, II, I + II) for at least 7 weeks. Changes in the morphology resulting from immobilization, as revealed by scanning electron microscopy (SEM) and low temperature-SEM, were investigated. Batch cultures and a continuous-flow packed bed photobioreactor were used to study nitrate removal from water. The effects of light intensity and CO2 concentration on bioreactor performance were studied with respect to the nitrate uptake efficiency of the system. It was concluded thatP. laminosum immobilized on polymer foams is of potential value for biological nitrate removal in a continuous-flow system. author for correspondence  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