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 15 毫秒
1.
A modified Rotating Biological Contactor (RBC) was used for the treatability studies of synthetic tapioca wastewaters. The RBC used was a four stage laboratory model and the discs were modified by attaching porous nechlon sheets to enhance biofilm area. Synthetic tapioca wastewaters were prepared with influent concentrations from 927 to 3600 mg/l of COD. Three hydraulic loads were used in the range of 0.03 to 0.09 m3·m–2·d–1 and the organic loads used were in the range of 28 to 306 g COD· m–2·d–1. The percentage COD removal were in the range from 97.4 to 68. RBC was operated at a rotating speed of 18 rpm which was found to be the optimal rotating speed. Biokinetic coefficients based on Kornegay and Hudson models were obtained using linear analysis. Also, a mathematical model was proposed using regression analysis.List of Symbols A m2 total surface area of discs - d m active depth of microbial film onany rotating disc - K s mg ·l–1 saturation constant - P mg·m–2·–1 area capacity - Q l·d–1 hydraulic flow rate - q m3·m–2·d–1 hydraulic loading rate - S 0 mg·l–1 influent substrate concentration - S e mg·l–1 effluent substrate concentration - w rpm rotational speed - V m3 volume of the reactor - X f mg·l–1 active biomass per unit volume ofattached growth - X s mg·l–1 active biomass per unit volume ofsuspended growth - X mg·l–1 active biomass per unit volume - Y s yield coefficient for attachedgrowth - Y A yield coefficient for suspendedgrowth - Y yield coefficient, mass of biomass/mass of substrate removed Greek Symbols hr mean hydraulic detention time - (max)A d–1 maximum specific growth rate forattached growth - (max)s d–1 maximum specific growth rate forsuspended growth - max d–1 maximum specific growth rate - d–1 specific growth rate - v mg·l–1·hr–1 maximum volumetric substrateutilization rate coefficient  相似文献   

2.
In this work, a new kinetic approach was proposed to describe the microbial growth, substrate consumption, and formation and utilization of the intracellular storage products (X STO) in activated sludge. It was found that the formation of X STO was coupled with energy generation and respiration and that the X STO formation rate was proportional to the substrate utilization rate. A high amount of external substrate resulted in a relatively rapid storage process with a large fraction of substrate electrons for X STO formation. The maximum growth rate of active biomass on X STO and the yield coefficient for growth on the storage polymers were estimated as 0.12 h−1 and 0.60 g chemical oxygen demand (COD) X g−1 CODSTO, respectively. This established model was verified with the experimental results from two different case studies with pure and mixed cultures. Results showed that this kinetic model was able to accurately and mechanistically describe the microbial storage processes.  相似文献   

3.
Summary  Growth of Gibberella fujikuroi in submerged cultures occurs as micelles or filamentous hyphae dispersed in fluid and pellets or stable, spherical agglomerations. Gibberella fujikuroi growth, substrate consumption and bikaverin production kinetics obtained from submerged batch fermentation were fitted to three different sigmoid models: two and three-parameter Gompertz models and one Logistic model. Growth fitting was used to compare between models and select the best one by means of an F test. The best model for describing growth was the two-parameter Gompertz model and was used for glucose consumption and bikaverin production fitting. Data from eight different schemes of fermentations were analysed and parameter estimation was carried out by means of minimization of residual sum of squares. Some characteristic values obtained with the two-parameter Gompertz model fit are: μ=0.028 h−1, Yx/s=0.1089 g substrate/g biomass, α =0.1384 g product/g biomass.  相似文献   

4.
Integrative and replicative plasmids for the expression driven by the P43 promoter and secretion of recombinant proteins in Bacillus subtilis were constructed. The plasmids named pInt and pRep respectively were tested for the production of recombinant human interferon gamma (rhIFN-γ). A synthetic hIFN-γ gene employing the optimized B. subtilis codon usage was fused with the Bacillus licheniformis α-amylase signal peptide (sp-amyL) encoding sequence. The integrative construct produced 2.5 ± 0.2 mg l−1 and the replicative system produced 20.3 ± 0.8 mg l−1 of total recombinant rhIFN-γ. The results showed that secretion of hIFN-γ was the bottleneck for the overexpression of mature rhIFN-γ by B. subtilis.  相似文献   

5.
The kinetic and general growth features of Bacillus thuringiensis var. israelensis were evaluated. Initial glucose concentration (S 0) in fermentation media varied from 10 to 152 g/l. The results afforded to characterize four morphologically and physiologically well-defined culture phases, independent of S 0 values: Phase I, vegetative growth; Phase II, transition to sporulation; Phase III, sporulation; and Phase IV, spores maturation and cell lysis. Important process parameters were also determined. The maximum specific growth rates (μ X,m) were not affected with S 0 up to 75 g/l (1.0–1.1 per hour), but higher glucose concentrations resulted in growth inhibition by substrate, revealed by a reduction in μ X,m values. These higher S 0 values led to longer Phases III and IV and delayed sporulation. Similar biomass concentrations (X m = 15.2–15.9 g/l) were achieved with S 0 over 30.8 g/l, with increasing residual substrate, suggesting a limitation in some other nutrients and the use of glucose to form other metabolites. In this case, with S 0 from 30.8 to 152 g/l, cell yield (Y X/S ) decreased from 0.58 to 0.41 g/g. On the other hand, with S 0 = 10 g/l growth was limited by substrate, and Y X/S has shown its maximum value (0.83 g/g).  相似文献   

6.
A novel short-chain dehydrogenases/reductases superfamily (SDRs) reductase (PsCR) from Pichia stipitis that produced ethyl (S)-4-chloro-3-hydroxybutanoate with greater than 99% enantiomeric excess, was purified to homogeneity using fractional ammonium sulfate precipitation followed by DEAE-Sepharose chromatography. The enzyme purified from recombinant Escherichia coli had a molecular mass of about 35 kDa on SDS–PAGE and only required NADPH as an electron donor. The Km value of PsCR for ethyl 4-chloro-3-oxobutanoate was 4.9 mg/mL and the corresponding Vmax was 337 μmol/mg protein/min. The catalytic efficiency value was the highest ever reported for reductases from yeasts. Moreover, PsCR exhibited a medium-range substrate spectrum toward various keto and aldehyde compounds, i.e., ethyl-3-oxobutanoate with a chlorine substitution at the 2 or 4-position, or α,β-diketones. In addition, the activity of the enzyme was strongly inhibited by SDS and β-mercaptoethanol, but not by ethylene diamine tetra acetic acid.  相似文献   

7.
Thirty-six 2.5-year-old wether Inner Mongolian White Cashmere Goats (IMWG) (BW = 42.7 ± 3.44 kg) were used to determine the effects of dietary copper (Cu) concentration on growth performance, nutrient digestibility and fiber characteristics during the cashmere slow-growing period. Wethers were stratified by weight and randomly assigned to four dietary treatments, which included a control diet containing 5.60 mg Cu/kg DM, the control diet supplied, respectively, with 10, 20 and 30 mg Cu/kg DM (total dietary Cu level of 5.60, 15.6, 25.6 and 35.6 mg/kg DM). The experiment lasted 50 days including a 10-day preliminary trial and 10-day metabolism trial. Average daily feed intake (ADFI) did not differ among treatment groups (P > 0.05), except that the supplement providing 30 mg Cu/kg DM decreased average daily gain and gain efficiency (P < 0.05). Copper supplementation had no influence on digestibility of DM, CP and ADF (P > 0.05), however, NDF digestibility of the treatment group supplemented with 30 mg Cu/kg DM was lower compared with that of other groups (P < 0.05). Length and growth rate of cashmere fiber were higher in the treatment group supplemented with 20 mg Cu/kg DM compared with other groups (P < 0.05), but cashmere diameter was not affected by Cu supplementation (P > 0.05). In conclusion, supplementation of Cu at the levels of 10, 20 and 30 mg/kg DM to the basal diet containing 5.60 mg Cu/kg DM had no influence on ADFI or nutrient digestibility of DM, CP and ADF in cashmere goats, while 30 mg Cu/kg DM supplementation had a negative effect on growth performance and NDF digestibility. However, 20 mg Cu/kg DM supplementation of the basal diet enhanced cashmere growth. Hence, the appropriate supplemental level during the cashmere slow-growing period is deemed to be 20 mg Cu/kg DM (total dietary Cu level of 25.6 mg/kg DM).  相似文献   

8.
This study evaluates a two-stage bioprocess for recovering bioenergy in the forms of hydrogen and methane while treating organic residues of ethanol fermentation from tapioca starch. A maximum hydrogen production rate of 0.77 mmol H2/g VSS/h can be achieved at volumetric loading rate (VLR) of 56 kg COD/m3/day. Batch results indicate that controlling conditions at S0/X0 = 12 with X0 = 4000 mg VSS/L and pH 5.5-6 are important for efficient hydrogen production from fermentation residues. Hydrogen-producing bacteria enriched in the hydrogen bioreactor are likely utilizing lactate and acetate for biohydrogen production from ethanol-fermentation residues. Organic residues remained in the effluent of hydrogen bioreactor can be effectively converted to methane with a rate of 0.37 mmol CH4/g VSS/h at VLR of 8 kg COD/m3/day. Approximately 90% of COD in ethanol-fermentation residues can be removed and among that 2% and 85.1% of COD can be recovered in the forms of hydrogen and methane, respectively.  相似文献   

9.
This study evaluated the biological treatability of produced water (PW), the water separated from oil at the wellhead which contains both dispersed oil and low levels of heavy metals, using waste stabilisation ponds (WSPs). We examined both chemical oxygen demand (COD) and oil and grease (O&G) removal using different process configurations (hydraulic retention time (HRT), aerobic and anaerobic conditions, oil skimming, effluent recycle) in a small (10 L) reactor being fed a synthetic PW (COD = 1050–1350 mg L−1, O&G = 400–500 μL L−1, 6 gNaCl/L). The reactor was operated for 6 months, and at a HRT of 6 days (8 with evaporation) COD removals were greater than 85%, and improved over time to >90%, while O&G removals (measured with a newly developed method) were greater than 82% and also improved with time. Operating with an anaerobic section, oil skimming and 300% recycling were all found to enhance COD removal.  相似文献   

10.
The kinetic properties of a microsomal gill (Na+,K+)-ATPase from the freshwater shrimp, Macrobrachium olfersii, acclimated to 21‰ salinity for 10 days were investigated using the substrate p-nitrophenylphosphate. The enzyme hydrolyzed this substrate obeying cooperative kinetics at a rate of 123.6 ± 4.9 U mg− 1 and K0.5 = 1.31 ± 0.05 mmol L− 1. Stimulation of K+-phosphatase activity by magnesium (Vmax = 125.3 ± 7.5 U mg− 1; K0.5 = 2.09 ± 0.06 mmol L− 1), potassium (Vmax = 134.2 ± 6.7 U mg− 1; K0.5 = 1.33 ± 0.06 mmol L− 1) and ammonium ions (Vmax = 130.1 ± 5.9 U mg− 1; K0.5 = 11.4 ± 0.5 mmol L− 1) was also cooperative. While orthovanadate abolished p-nitrophenylphosphatase activity, ouabain inhibition reached 80% (KI = 304.9 ± 18.3 μmol L− 1). The kinetic parameters estimated differ significantly from those for freshwater-acclimated shrimps, suggesting expression of different isoenzymes during salinity adaptation. Despite the ≈2-fold reduction in K+-phosphatase specific activity, Western blotting analysis revealed similar α-subunit expression in gill tissue from shrimps acclimated to 21‰ salinity or fresh water, although expression of phosphate-hydrolyzing enzymes other than (Na+,K+)-ATPase was stimulated by high salinity acclimation.  相似文献   

11.
Summary Cell growth and phenol degradation kinetics were studied at 10°C for a psychrotrophic bacterium, Pseudomonas putida Q5. The batch studies were conducted for initial phenol concentrations, So, ranging from 14 to 1000 mg/1. The experimental data for 14<=So<=200 mg/1 were fitted by non-linear regression to the integrated Haldane substrate inhibition growth rate model. The values of the kinetic parameters were found to be: m=0.119 h–1, K S=5.27 mg/1 and K I=377 mg/1. The yield factor of dry biomass from substrate consumed was Y=0.55. Compared to mesophilic pseudomonads previously studied, the psychrotrophic strain grows on and degrades phenol at rates that are ca. 65–80% lower. However, use of the psychrotrophic microorganism may still be economically advantageous for waste-water treatment processes installed in cold climatic regions, and in cases where influent waste-water temperatures exhibit seasonal variation in the range 10–30°C.Nomenclature K S saturation constant (mg/l) - K I substrate inhibition constant (mg/l) - specific growth rate (h–1) - m maximum specific growth rate without substrate inhibition (h–1) - max maximum achievable specific growth rate with substrate inhibition (h–1) - S substrate (phenol) concentration (mg/l) - So initial substrate concentration (mg/l) - Smax substrate concentration corresponding to max (mg/l) - t time (h) - X cell concentration, dry basis (mg DW/l) - Xf final cell concentration, dry basis (mg DW/l) - Xo initial cell concentration, dry basis (mg DW/l) - Y yield factor (mg DW cell produced/mg substrate consumed)  相似文献   

12.
Mathematical model parameters for the methanogenic degradation of propylene glycol were estimated in a sequential manner by means of an optimization technique. Model parameters determined from an initial experimental data set using one bioreactor were then verified with the results from a second bioreactor. The proposed methodology is a useful tool to obtain model parameters for continuous flow reactors with completely mixed regime. Abbrevations: S – substrate concentration (mg COD l–1); S in – influent substrate concentration (mg COD l–1); D L – dilution rate (day–1); – stoichiometric coefficients (ND); nx – number of microbial species (ND); X S – fixed biomass concentration (mg biomass l–1); X L – suspended biomass concentration of (mg biomass l–1); k d – decay rate of biomass (day–1); b S – specific detachment rate of biofilm (day–1); – specific growth rate of biomass (day–1); m – maximum specific growth rate of biomass (day–1); K S – half saturation constant (mg COD l–1); K I – inhibition constant (mg COD l–1).  相似文献   

13.
The seasonal variability of specific growth rate and the carbon stable isotope ratio (δ13C) of leaf blades (δ13Cleaf) of a temperate seagrass, Zostera marina (within 10 days old) were measured simultaneously, together with the δ13C of dissolved inorganic carbon (δ13CDIC) at three sites in the semi-closed Akkeshi estuary system, northeastern Japan, in June, September, and November 2004. The δ13Cleaf ranged from −16.2 to −6.3‰ and decreased from summer to winter. The simultaneous measurement of the δ13Cleaf, growth rate, and morphological parameters (mean leaf length and width, mean number of leaves per shoot, and sheath length) of the seagrass and δ13CDIC in the surrounding water allowed us to compare directly the δ13Cleaf and specific growth rate of seagrass. The difference in the δ13C of seagrass leaves relative to the source DIC (Δδ13Cleaf − DIC) was the least negative (−11 to −7‰) in June at all three sites and became more negative (−17 to −8‰) as the specific growth rate decreased. This positive correlation between Δδ13Cleaf − DIC and specific growth rate can be used to diagnose the growth of seagrasses. Δδ13Cleaf − DIC changed by −1.7 ± 0.2‰ when the leaf specific growth rate decreased by 1% d−1.  相似文献   

14.
Acarbose-fructoside (acarbose-Fru) was newly synthesized via the acceptor reaction of a levansucrase from Leuconostoc mesenteroides B-512 FMC with acarbose and sucrose. The resultant product was separated with 10.5% purification yield via Bio-gel P-2 column chromatography and HPLC. Its structure was determined to be 1I-β-d-fructofuranosyl α-acarbose, according to the results of 1H, 13C, HSQC, and HMBC analyses. Acarbose-Fru was inhibited competitively on α-glucosidase (A. niger and baker's yeast) but mixed noncompetitively on α-amylases (A. oryzae and porcine pancreatic). Compared to acarbose, acarbose-Fru exhibited inhibition potency of 1.12 or 1.52 on A. niger α-glucosidase or A. oryzae α-amylase, respectively. Additionally, acarbose-Fru was identified as a novel substrate for dextransucrase with Km and Vmax values of 189.0 mM and 8.51 μmol/(mg min), respectively. Therefore, acarbose-Fru as a substrate might be synthesized novel acarbose derivatives by using dextransucrase.  相似文献   

15.

To interpret the biological nutrient removal in a cyclic activated sludge system (CAS), a modified model was developed by combining the process of simultaneous storage and growth, and the kinetics of soluble microbial product (S SMP) and extracellular polymeric substance (X EPS) with activated sludge model no. 3 (ASM3). These most sensitive parameters were initially selected whilst parameters with low sensitivity were given values from literature. The selected parameters were then calibrated on an oxygen uptake rate test and a batch CAS reactor on an operational cycle. The calibrated model was validated using a combination of the measurements from a batch CAS reactor operated for 1 month and the average deviation method. The simulations demonstrated that the modified model was capable of predicting higher effluent concentrations compared to outputs of the ASM3 model. Additionally, it was also shown that the average deviation of effluent S COD, S NH, S SMP and X EPS simulated with the modified model was all less than 1 mg L−1. In summary, the model could effectively describe biological processes in a CAS reactor and provide a wonderful tool for operation.

  相似文献   

16.
As a part of the investigations on the microbial lipid production using the yeast Rhodotorula gracilis, CFR-1, kinetics of the biomass synthesis has been studied using shake flask experiments. Using a medium containing a carbon to nitrogen ratio of 701, the rates of biomass production were followed at different initial substrate concentrations in the range of 20–100 kg/m3. A logistic model was found to be reasonably adequate to describe the kinetics of the growth of biomass; the maximum specific growth rate of 0.105 h–1 was applicable for substrate concentrations less than 60 kg/m3, which gave reasonable agreement between predicted and actual biomass concentration values.List of Symbols S 0, X 0 kg/m3 Initial concentrations of sugar, non lipid biomass respectively - X, X(t) kg/m3 Concentrations of non lipid biomass at any time t - dX/dt kg/(m3 · h) Rate of biomass growth - h–1 Specific growth rate - max h–1 Maximum specific growth rate - K s mol/dm3 Monods constant - X max kg/m3 Maximum biomass reached in a run  相似文献   

17.
The aggregation of proteins is believed to be intimately connected to many neurodegenerative disorders. We recently reported an “Ockham's razor”/minimalistic approach to analyze the kinetic data of protein aggregation using the Finke–Watzky (F–W) 2-step model of nucleation (A → B, rate constant k1) and autocatalytic growth (A + B → 2B, rate constant k2). With that kinetic model we have analyzed 41 representative protein aggregation data sets in two recent publications, including amyloid β, α-synuclein, polyglutamine, and prion proteins (Morris, A. M., et al. (2008) Biochemistry 47, 2413-2427; Watzky, M. A., et al. (2008) Biochemistry 47, 10790–10800). Herein we use the F–W model to reanalyze protein aggregation kinetic data obtained under the experimental conditions of variable temperature or pH 2.0 to 8.5. We provide the average nucleation (k1) and growth (k2) rate constants and correlations with variable temperature or varying pH for the protein α-synuclein. From the variable temperature data, activation parameters ΔG, ΔH, and ΔS are provided for nucleation and growth, and those values are compared to the available parameters reported in the previous literature determined using an empirical method. Our activation parameters suggest that nucleation and growth are energetically similar for α-synuclein aggregation (ΔGnucleation = 23(3) kcal/mol; ΔGgrowth = 22(1) kcal/mol at 37 °C). From the variable pH data, the F–W analyses show a maximal k1 value at pH ~ 3, as well as minimal k1 near the isoelectric point (pI) of α-synuclein. Since solubility and net charge are minimized at the pI, either or both of these factors may be important in determining the kinetics of the nucleation step. On the other hand, the k2 values increase with decreasing pH (i.e., do not appear to have a minimum or maximum near the pI) which, when combined with the k1 vs. pH (and pI) data, suggest that solubility and charge are less important factors for growth, and that charge is important in the k1, nucleation step of α-synuclein. The chemically well-defined nucleation (k1) rate constants obtained from the F–W analysis are, as expected, different than the 1/lag-time empirical constants previously obtained. However, k2 × [A]0 (where k2 is the rate constant for autocatalytic growth and [A]0 is the initial protein concentration) is related to the empirical constant, kapp obtained previously. Overall, the average nucleation and average growth rate constants for α-synuclein aggregation as a function of pH and variable temperature have been quantitated. Those values support the previously suggested formation of a partially folded intermediate that promotes aggregation under high temperature or acidic conditions.  相似文献   

18.
Song Y  Wang C  Wang C  Lv L  Chen Y  Zuo Z 《Animal reproduction science》2009,110(3-4):306-318
The present study was undertaken to examine the effect of administered recombinant mouse leptin on the recovery of regressed ovary in fasted ducks. Twenty-eight ducks were divided into five groups: fed ad libitum (control; n = 5), fasted control (FC; n = 5), fasted + low dose of leptin (F + L; n = 5), fasted + medium dose of leptin (F + M; n = 5) and fasted + high dose of leptin (F + H; n = 3). All four fasted groups were fasted for 2 days and then ad libitum and the ducks were treated with leptin at doses of 0 (control and FC), 50 (F + L), 250 (F + M) and 1000 (F + H) μg/kg body weight/day on day 3–5. Results showed that a moderate dose of leptin (250 μg/kg body weight/day) injected during the re-feeding period: (i) promoted the recovery of the regressed ovary as evidenced by an increase in ovary weight and recovery of yellow hierarchical follicles; (ii) elevated the plasma 17β-estradiol (E2) level; (iii) increased the mRNA levels of ovary follicle-stimulating hormone receptor (FSHR), luteinizing hormone receptor (LHR) and estrogen receptor-β (ER-β). Furthermore, the results also showed that a high dose of leptin (1000 μg/kg body weight/day) may have a negative effect on the recovery of the regressed ovary. In conclusion, this study indicates that, in ducks, leptin may be involved in the recovery of the regressed ovary caused by 2 days of fasting. This effect may be related to increased plasma E2 levels and stimulation of the mRNA levels of ovarian FSHR, LHR and especially ER-β.  相似文献   

19.
A novel method for the determination of microbial growth kinetics on hydrophobic volatile organic compounds (VOC) has been developed. A stirred tank reactor was operated as a fed-batch system to which the VOC was continuously fed via the gas phase, assuring a constant VOC concentration in the mineral medium. A flow of air was saturated with the VOC, and then mixed with a further flow of air, to obtain a predetermined VOC concentration. Thus, different VOC concentrations in the mineral medium could be obtained by altering the VOC concentration in the feed gas. The growth kinetics of Xanthobacter autotrophicus GJ10 on 1,2-dichloroethane (DCE) and of Pseudomonas sp. strain JS150 on MonoChloroBenzene (MCB) were assessed using this method. The growth of strain JS150 was strongly inhibited at MCB concentrations higher than 160 mg l−1, and the results were fitted using a piecewise function. The growth kinetics of strain GJ10 were described by the Luong model where maximum growth rate μmax = 0.12 h−1, substrate saturation constant K S = 7.8 mg l−1, and maximum substrate concentration S m (above which growth is completely inhibited) = 1080 mg l−1. Varying nitrogen and oxygen flows enabled the effect of oxygen concentration on the growth kinetics of Pseudomonas JS150 to be determined. Received: 30 November 1998 / Received revision: 19 March 1999 / Accepted: 20 March 1999  相似文献   

20.
We investigated Fe plaque formation and Ca, Cu, Mn, Zn, and P uptake capacities of fifteen kinds of wetland plants. The test plants were cultured in 3 l nutrient solutions for 8 days. Fe plaque was induced by adding 200 mg l−1 Fe2+ as FeSO4·7H2O for 4 days in one set of experiment and 8 days in another. This plaque ranged from 2.38 to 8.67 mg g−1 of plant root after 4 days and from 4.56 to 15.71 mg g−1 of plant root after 8-day treatment. In both experimental durations, the plaque was significantly correlated with root surface area (r = 0.904 and 0.878, P < 0.01). Thus, Canna generalis, Typha latifolia and Thalia dealbata, with their larger root surface areas (>1,400 cm2), formed relatively greater Fe plaque amounts. The amounts of Ca, Cu, Zn and P in the Fe plaques were significantly correlated with Fe plaque amount, (r = 0.819, 0.742, 0.693, 0.917, respectively, for these four elements for the 4-day treatment; and r = 0.917, 0.768, 0.949, 0.872, respectively, for 8-day treatment, P < 0.01). Plants varied widely in accumulating Ca, Cu, Mn, Zn, and P in their tissues. The amounts accumulated on root were significantly correlated with Fe plaque amount in both for 4- and 8-day exposure treatments with Fe (r = 0.973, 0.847, 0.709, 0.837, 0.892, respectively, for 4-day treatment; and r = 0.943, 0.691, 0.843, 0.957, 0.983, respectively, for 8-day treatment, P < 0.01). No such significant correlations were found for the Fe plaque in shoot. Canna generalis, Typha latifolia and Thalia dealbata were superior in Ca, P and Zn uptake, while Canna generalis and Thalia dealbata accumulated Cu and Mn well in case of concentrated wastewater treatment.  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