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 477 毫秒
1.
A bench-scale experiment was conducted in a 701. tank of tap water to examine the effect of four design variables on oxygen transfer in a fine pore diffused aeration system. The experiment used non-steady state gas transfer methodology to examine the effect of air flow rate, air flow rate per diffuser, orifice diameter and reduced tank surface area on the overall oxygen transfer coefficient (KLa20, h−1); standard oxygen transfer rate (OT2, g O2 h−1); energy efficiency (Ep, g O2 kWh−1) and oxygen transfer efficiency (Eo, %). The experiments demonstrated that KLa20 and OTs increased with air flow rate (9.4–18.8 1 min−1) in the 40 and 140 μ diameter orifice range; however, Ep and E0 were not affected. Reducing the air flow rate per fine pore diffuser (40 and 140 μ diameter pore size) significantly increased KLa20, OTs, Ep and E0. A decrease in orifice diameter from 140 to 40 μ had no effect on KLa20, OTs, Ep and E0. A reduction in tank surface area had a marginally significant inverse effect on KLa20 and OTs, and no effect on Ep and Eo. The mean bubble size produced by the 40 and 140 μ diffusers was 4.0 and 4.2 mm, respectively. There was no consistent effect of air flow rate on bubble size within the range of air flow rates used in this experiment. In clean water aeration applications, the optimum system efficiency will be obtained using the largest number of fine pore diffusers operated at low air flow rates per diffuser. In wastewater treatment plants, higher air flow rates per diffuser should be used to prevent diffuser biofouling and keep biological solids in suspension. Wastewater systems are purposely operated at less than optimum transfer efficiencies in exchange for reduced diffuser maintenance and improved mixing. In either situation, changes in tank surface area and diffuser pore size (provided that pore diameter remains between 40 and 140 μ) are unlikely to have any significant effect on aeration system efficiency.  相似文献   

2.
Comparative study of three standard aeration systems: bubble diffuser; turbine aeration (Rushton type): Rushton turbine as surface aerator, with use of original correlation functions. This comparative study is carried out by relating: oxygen efficiency (kg O2 kW·h−1), to unitary power (Wm−3).  相似文献   

3.
Humic acid, which is a typical microbially refractory organic substance, was extracted from a landfill leachate. The humic acid solution (COD = 367 mg 1−1; TOC = 293 mg 1−1; BOD = 27 mg 1−1) was applied to a batch scale activated sludge treatment after the modification of its biodegradability by γ-ray irradiation. The BOD increased to 64 mg 1−1 by irradiation of 15 kGy (1.5 Mrad), while the COD and TOC decreased to 231 and 230 mg 1−1, respectively. When the irradiated sample was treated with an activated sludge, the BOD decreased rapidly in 2–3 h to about 15 mg 1−1 which was a similar value as the unirradiated sample was treated. The elimination efficiency of TOC by the sludge treatment was approximately equal to that obtained by irradiation of 15 kGy. These facts suggest a utility of applying microbial processes after radiation treatment of microbially refractory wastewaters.  相似文献   

4.
Pressure flotation of abattoir wastewater using carbon dioxide or a 1:3 mixture of carbon dioxide and nitrogen has been investigated in laboratory-scale tests. When the wastewater was not chemically pretreated, the pH fell to 4.9–5.8, but removal efficiencies for suspended solids, chemical oxygen demand and fat were no better than for pressure flotation using air. When the wastewater was pretreated with alum at dose rates up to 150 mgl−1, removal efficiencies using the carbon dioxide/nitrogen mixture were close to those obtained by dissolved air flotation of wastewater chemically pretreated with acid and alum.  相似文献   

5.
Denitrification studies with glycerol as a carbon source   总被引:1,自引:0,他引:1  
Based on the results of experimental work, the use of glycerol as a carbon source for denitrification is discussed. Investigations were carried out in modified UASB reactor using a mixed bacterial population in medium containing 600 mg NO3---N and essential biogens. It was found that at the most favourable C:N RATIO = 1.0 the efficiency of denitrification depended on nitrate load as well as on cell residence time. At nitrate loads in the range 220–670 mg NO3---Nl−1 day−1 (0.08–0.14 mg NO3---N mg−1 day−1) nitrogen removal was 0.6–0.12 mg N mg−1 day−1, respectively. Denitrification unit biocenosis was composed of bacteria, fungi and protozoa. The number of denitrifying bacteria per sludge weight unit within the studied range of nitrate loads was constant and averaged 23 × 107 cells mg−1.  相似文献   

6.
The aims of this study were to demonstrate the (1) feasibility of psychrophilic, or low-temperature, anaerobic digestion (PAD) of phenolic wastewaters at 10–15 °C; (2) economic attractiveness of PAD for the treatment of phenol as measured by daily biogas yields and (3) impact on bioreactor performance of phenol loading rates (PLRs) in excess of those previously documented (1.2 kg phenol m−3 d−1). Two expanded granular sludge bed (EGSB)-based bioreactors, R1 and R2, were employed to mineralise a volatile fatty acid-based wastewater. R2 influent wastewater was supplemented with phenol at an initial concentration of 500 mg l−1 (PLR, 1 kg m−3 d−1). Reactor performance was measured by chemical oxygen demand (COD) removal efficiency, CH4 composition of biogas and phenol removal (R2 only). Specific methanogenic activity, biodegradability and toxicity assays were employed to monitor the physiological capacity of reactor biomass samples. The applied PLR was increased to 2 kg m−3 d−1 on day 147 and phenol removal by day 415 was 99% efficient, with 4 mg l−1 present in R2 effluent. The operational temperature of R1 (control) and R2 was reduced by stepwise decrements from 15 °C through to a final operating temperature of 9.5 °C. COD removal efficiencies of c. 90% were recorded in both bioreactors at the conclusion of the trial (day 673), when the phenol concentration in R2 effluent was below 30 mg l−1. Daily biogas yields were determined during the final (9.5 °C) operating period, when typical daily R2 CH4 yields of c. 3.3 l CH4 g−1 CODremoved d−1 were recorded. The rate of phenol depletion and methanation by R2 biomass by day 673 were 68 mg phenol g VSS−1 d−1 and 12–20 ml CH4 g VSS−1 d−1, respectively.  相似文献   

7.
We have investigated electrical potential and acoustic emissions signals associated with rock deformation. Five rock types were studied; Clashach, Bentheim and Darley Dale sandstones (all quartz-rich) and a Seljadur basalt and Portland limestone (both quartz-free), both air dry and the rocks were tested in distilled water. Shallow crustal conditions were simulated in a triaxial rock deformation cell with a confining pressure simulating depth of 40 MPa, pore pressures ranging 5–35 MPa, and strain rates 10−7–10−4 s−1. Precursory electric potential signals prior to failure were observed in both saturated and dry samples of the quartz-rich sandstones, but only observed in the water saturated quartz-free rocks. Co-seismic electrical signals were obtained in all tests, providing strong evidence that two of the main sources for precursory and co-seismic signals are the piezoelectric and electrokinetic phenomena. Lowering the strain rate resulted in an increase in the number of acoustic emissions. The pore volume changes during compaction and dilatancy remained approximately constant for all strain rates. Streaming potential generated by fluid flow across the sample was also measured at different stages of deformation. The potential signals increased with the pore pressure gradient.  相似文献   

8.
Interstitial P levels in Lake Mendota and Lake Wingra were evaluated as a function of season and water column and sediment depth. Interstitial water was obtained by the centrifugation-filtration method. Temporal variations were observed over the entire 15 cm sediment depth interval examined in all four locations evaluated. Interstitial reactive P (IRP) levels in Lake Mendota ranged from 0.014–1.67 mg l−1 at the 5–6 m water column depth and from 1.20–5.75 mg l−1 at the 18–19.5 m depth. IRP levels in Lake Wingra ranged from 0.029–2.15 mg l−1 at 3.5 m and from 0.191–3.96 mg l−1 at 2 m. Variations in interstitial P were attributed to variations in oxidation state of Fe as influenced by oxygen transport and reduction rates.  相似文献   

9.
Zero net growth in a membrane bioreactor with complete sludge retention   总被引:5,自引:0,他引:5  
A bench-scale membrane bioreactor was operated with complete sludge retention in order to evaluate biological processes and biomass characteristics over the long term. The investigation was carried out by feeding a bench-scale plant with real sewage under constant volumetric loading rate (VLR=1.2 gCOD Lreact−1 h−1). Biological processes were monitored by measuring substrate removal efficiencies and biomass-related parameters. The latter included bacterial activity as determined through respirometric tests specifically aimed at investigating long term heterotrophic and nitrifying activity. After about 180 days under the imposed operating conditions, the system reached equilibrium conditions with constant VSS concentration of 16–18 g L−1, organic loading rate (OLR) below 0.1 gCOD gVSS−1 d−1 and specific respiration rates of 2–3 mgO2 gVSS−1 h−1. These conditions were maintained for more than 150 days, confirming that an equilibrium had been achieved between biomass growth, endogenous metabolism, and solubilization of inorganic materials.  相似文献   

10.
Utsumi H  Han YH  Ichikawa K 《Water research》2003,37(20):4924-4928
Hydroxyl (OH) radical is proposed as an important factor in the ozonation of water. In the present study, the enhancing effect of 3-chlorophenol on OH radical generation was mathematically evaluated using electron spin resonance (ESR)/spin-trapping technique. OH radical was trapped with a 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) as a stable adduct, DMPO–OH. The initial velocity of DMPO–OH generation in ozonated water containing 3-chlorophenol was quantitatively measured using a combined system of ESR spectroscopy with stopped-flow apparatus which was controlled by home-made software. The initial velocity of DMPO–OH generation increased as a function of the concentration of ozone and the more effectively of 3-chlorophenol concentration. The relation among ozone concentration, amount of 3-chlorophenol and the initial velocity (ν0) of DMPO–OH generation was mathematically analyzed and the following equation was obtained, ν0 (10−6M/s)={9.7×[3-chlorophenol (10−9M)]+0.0005} exp(57×[ozone (10−9M)]). The equation fitted very well with the experimental results, and the correlation coefficient was larger than 0.99. The equation for the enhancing effect by 3-chlorophenol should provide useful information to optimize the condition in ozone treatment process of water containing phenolic pollutants.  相似文献   

11.
Wet oxidation of phenol by air or oxygen over a Pt/TiO2 catalyst is studied in a batch reactor in the temperature range 150–200°C, pressure range 34–82 atm, and a catalyst loading range of 0–4 g catalyst L−1. The catalyst was powdered 4.45% Pt/TiO2 with a maximum particle size of 105 μm. Results show complete oxidation of phenol and almost complete total organic carbon (TOC) removal. Small amounts of stable organic acids are formed in side reactions of the phenol degradation pathway and are not readily degraded. Experimental results show that the reaction rate decreases by increased oxygen concentration. Theoretical rate expressions are derived, based on postulated oxidation and TOC reduction mechanisms.  相似文献   

12.
A residual soluble Pb ≤ 0.2 mg/l has been measured after hydrocerussite precipitation in the range of pH 9–10 and filtration with 0.45 μm filter, at a total carbonate concentration (CT) of about 1.5 × 10−4 M. Filterability and sedimentation were optimized by minimizing the “relative supersaturation” coefficient during precipitation. Pb abatement was implemented in the pH range 9–9.5, following a preliminary precipitation step consisting of a slow, gradual change of pH from 6.8 to 7.8 with 0.1 M NaOH solution. The crystals formed may settle within 24 h, allowing Pb to be recovered as Pb3(CO3)2(OH)2.  相似文献   

13.
A numerical and an analytical model were developed to predict the volatile organic compound (VOC) emission rate from dry building materials. Both models consider the mass diffusion process within the material and the mass convection and diffusion processes in the boundary layer. All the parameters, the mass diffusion coefficient of the material, the material/air partition coefficient, and the mass transfer coefficient of the air can be either found in the literature or calculated using known principles.

The predictions of the models were validated at two levels: with experimental results from the specially designed test and with predictions made by a CFD model. The results indicated that there was generally good agreement between the model predictions, the experimental results, and the CFD results. The analytical and numerical models then were used to investigate the impact of air velocity on emission rates from dry building materials. Results showed that the impact of air velocity on the VOC emission rate increased as the VOC diffusion coefficient of the material increased. For the material with a diffusion coefficient >10−10 m2/s, the VOC emission rate increased as the velocity increased; air velocity had significant effect on the VOC emission. For the material with a VOC diffusion coefficient <10−10 m2/s, the VOC emission rate increased as the velocity increased only in the short-term; <24 h. In the medium to long-term time range, the VOC emission rate decreased slightly as the air velocity increased; velocity did not have much impact on these materials. Furthermore, the study also found that the VOC concentration distribution within the material; the VOC emission rate and the VOC concentration in the air were linearly proportional to the initial concentration. However, the normalized emitted mass was not a function of the initial concentration: it was a function of the properties of the VOC and the material.  相似文献   


14.
Studies on marine biological filters: Model filters   总被引:1,自引:0,他引:1  
J.F. Wickins 《Water research》1983,17(12):1769-1780
Model systems of 401. capacity were used to study chemical changes which affected buffering in continuously recycled sea water and which restricted nitrification in percolating biological filters at 28°C.

Sustained hydrogen ion production during the microbial oxidation of ammonia to nitrite caused continuous carbon dioxide formation from carbonate and bicarbonate. The carbon dioxide was steadily lost to the air through vigorous aeration, leaving < 2 mg inorganic carbon 1−1 in the sea water. Oxidation of nitrate to nitrate did not significantly reduce pH nor deplete buffer capacity.

Ammonia oxidation was severely inhibited by the combination of low pH and dissolved inorganic carbon levels, but similar low levels of pH produced when carbon dioxide was bubbled through the water had only a moderate effect. Inhibition could be rapidly overcome or prevented by additions of inorganic carbon, sodium hydroxide or sodium dihydrogen phosphate.

Values recorded for the maximum specific growth rate of Nitrosomonas and Nitrobacter were 0.53 and 0.81 d−1 respectively. The corresponding generation or doubling times were calculated to be 31.4 and 20.5 h.

Some evidence was found for the uptake of phosphate and the formation of hydroxylamine during nitrification.  相似文献   


15.
A study was made of the effect of water hardness at different concentrations (viz. 0, 80, 120, 160, 240, 320, 400 and 480 mg l−1 as CaCO3) on the toxicity of cadmium metal (5 mg 1−1) as sulphate to saprophytic and nitrifying bacteria, with respect to the rate constant (K) and ultimate biochemical oxygen demand (L) which were calculated from BOD data (15 days) using the Thomas Graphical Method. Glucose was used as a source of carbon for micro-organisms. It was observed that the toxicity of cadmium to micro-organisms (both saprophytic and nitrifying) decreased with increasing hardness and reached a maximum at 320 mg 1−1 as CaCO3 for nitrifying and 400 mg l−1 as CaCO3 for saprophytic bacteria. After these hardness levels, the ultimate BOD (L) and rate constant (K) showed a decrease. Nitrifying bacteria were found to be more sensitive to the metal as well as to its complexation with calcium or with other ions as they retained their normal activity at a lower hardness level as compared to saprophytic bacteria.  相似文献   

16.
The effect of effluent composition involving the common anionic species Cl, SO42− and CO32− on the efficiency of nickel(II) precipitation, modelling lime (CaO) as the precipitant, has been investigated using the solubility domain approach. Solubility domains were based on the phases that were found to limit metal solubility for systems representing potential effluent composition limits. These phases were found to resemble their mineralized counterparts, but with a lower degree of structural order. At higher SO42− and CO32− concentrations both gypsum (CaSO4·2H2O) and calcite (CaCO3) were formed, but these had little effect on the observed residual nickel solubility. The calculated solubility domains were found to generally encompass the experimentally determined solubilities, thereby providing quality assurance ranges for hydroxide precipitation. The effect of the complexing anions tartrate and EDTA4− on residual Ni(II) in solution as well as the effects of the addition of Fe(III) on the removal of Ni(II) complexed by these species are described.  相似文献   

17.
A new model for compartment fires is proposed in which the new concept of combustion efficiency based on the mixing process of the fuel gas and air has been considered. This new concept was formulated by the mixing parameter, μ. It was defined as μ1 − exp(−τ*) and τ* was related to the residence time tR and mixing time tM, that is, τ* = tR/tM.

A simple one zone model was used in order to demonstrate the effect of the mixing process. Theoretical results were in good agreement with the experimental results of methanol and PMMA compartment fires, and especially the scale effect of the compartment was predicted successfully. Further, the similarity law for this scale effect was investigated, and the upper and lower limits of flashover were defined using a new number F. This F number was found to be the key parameter for the prediction of the compartment fire behavior.  相似文献   


18.
Wang GP  Liu JS  Tang J 《Water research》2004,38(20):1927-4474
Sediment cores, representing a range of watershed characteristics and anthropogenic impacts, were collected from two freshwater marshes at the Xianghai wetlands (Ramsar site no. 548) in order to trace the historical variation of nutrient accumulation. Cores were 210Pb- and 137Cs-dated, and these data were used to calculate sedimentation rates and sediment accumulation rates. Ranges of dry mass accumulation rates and sedimentation rates were 0.27–0.96 g m−2 yr−1 and 0.27–0.90 cm yr−1, respectively. The effect of human activities on increased sediment accumulation rates was observed. Nutrients (TOC, N, P, and S) in sediment were analyzed and nutrient concentration and accumulation were compared in two marshes with different hydrologic regime: an “open” marsh (E-0) and a partly “closed” marsh (F-0). Differences in physical and chemical characteristics between sediments of “open” and partly “closed” marsh were also observed. The “open” marsh sequestered much higher amounts of TOC (1.82%), N (981.1 mg kg−1), P (212.17 mg kg−1), and S (759.32 mg kg−1) than partly “closed” marsh (TOC: 0.32%, N: 415.35 mg kg−1, P: 139.64 mg kg−1, and S: 624.45 mg kg−1), and the “open” marsh indicated a rather large historical variability of TOC, N, P, and S inputs from alluvial deposits. Nutrient inputs (2.16–251.80 g TOC m−2 yr−1, 0.43–20.12 g N m−2 yr−1, 0.39–3.03 g P m−2 yr−1, 1.60–15.13 g S m−2 yr−1) into the Xianghai wetlands of China are in the high range compared with reported nutrient accumulation rates for freshwater marshes in USA. The vertical variation, particularly for N, P, and S indicated the input history of the nutrients of the Xianghai wetlands developed in three periods—before 1950s, 1950–1980s, and after 1980s. The ratios between anthropogenic and natural inputs showed that the relative anthropogenic inputs of TOC, N, P, and S have been severalfold (TOC: 1.68–11.21, N: 0.47–3.67, P: 0.24–1.36, and S: 1.46–2.96) greater than values of their natural inputs after 1980s. The result is probably attributable, in part, to two decades of surface coal mining activities, urban sewage, and agriculture runoff within the upstream region of the Huolin River. Our findings suggest that the degree of anthropogenic disturbance within the surrounding watershed regulates wetland sediment, TOC, N, P, and S accumulation.  相似文献   

19.
We report simultaneous laboratory measurements of seismic velocities and fluid permeability on lava flow basalt from Etna (Italy) and columnar basalt from Seljadur (Iceland). Measurements were made in a servo-controlled steady-state-flow permeameter at effective pressures from 5–80 MPa, during both increasing and decreasing pressure cycles. Selected samples were thermally stressed at temperatures up to 900 °C to induce thermal crack damage. Acoustic emission output was recorded throughout each thermal stressing experiment.

At low pressure (0–10 MPa), the P-wave velocity of the columnar Seljadur basalt was 5.4 km/s, while for the Etnean lava flow basalt it was only 3.0–3.5 km/s. On increasing the pressure to 80 MPa, the velocity of Etnean basalt increased by 45%–60%, whereas that of Seljadur basalt increased by less than 2%. Furthermore, the velocity of Seljadur basalt thermally stressed to 900 °C fell by about 2.0 km/s, whereas the decrease for Etnean basalt was negligible. A similar pattern was observed in the permeability data. Permeability of Etnean basalt fell from about 7.5×10−16 m2 to about 1.5×10−16 m2 over the pressure range 5–80 MPa, while that for Seljadur basalt varied little from its initial low value of 9×10−21 m2. Again, thermal stressing significantly increased the permeability of Seljadur basalt, whilst having a negligible effect on the Etnean basalt. These results clearly indicate that the Etnean basalt contains a much higher level of crack damage than the Seljadur basalt, and hence can explain the low velocities (3–4 km/s) generally inferred from seismic tomography for the Mt. Etna volcanic edifice.  相似文献   


20.
Two large circular enclosures, each containing approx. 550 m3 of water, 14 m deep, and open to the mud-water interface, were used to monitor the effects of hypolimnetic aeration. One enclosure was held as a control, the other aerated every 3 or 4 days for a period long enough (usually < 2 h) to maintain hypolimnetic O2 levels at > 4 mg 1−1. Nutrient additions (10 g of 90% H3 PO4 and 250 g NaNO3 per week) to each enclosure were controlled from the commencement of the experiment (17 June 1980) until its completion (2 November 1981). Temperatures in both enclosures were identical. Hypolimnetic O2 levels in the control fell to zero during both summers, but remained at > 4 mg 1−1 in the aerated enclosure. Free N2 concentrations in the hypolimnion of the aerated enclosure was higher than in the control. Concentrations of H2S in the control hypolimnion increased to > 5 mg 1−1 and concentrations of CH4 increased to > 18 mg 1−1. Both remained at or near zero in the aerated enclosure. Tests of aerator efficiency suggested that the full air-lift design that was employed had an average O2 exchange efficiency of 42% which is higher than the values reported for most other designs.  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