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.
Aerobic granulation is a process in which suspended biomass aggregate and form discrete well-defined granules in aerobic systems. To investigate the properties and kinetics of aerobic granular sludge, aerobic granules were cultivated with glucose synthetic wastewater in a series of sequencing batch reactors (SBR). The spherical shaped granules were observed on 8th day with the mean diameter of 0.1 mm. With the organic loading rate (OLR) being increased to 4.0 g COD L−1 d−1, aerobic granules grew matured with spherical shape. The size of granules ranged from 1.2 to 1.8 mm, and the corresponding settling velocity of individual granule was 24.2–36.4 m h−1. The oxygen utilization rate (OUR) of mature granules was 41.90 g O2 kg MLSS−1 h−1, which was two times higher than that of activated sludge (18.32 g O2 kg MLSS−1 h−1). The experimental data indicated that the substrate utilization and biomass growth kinetics generally followed Monod's kinetics model. The corresponding kinetic coefficients of k (maximum specific substrate utilization rate), Ks (half velocity coefficient), Y (growth yield coefficient) and Kd (decay coefficient) were determined as follows, kc = 23.65 d−1, Kc = 3367.05 mg L−1, KN = 0.038 d−1, KN = 29.65 mg L−1, Y = 0.1927–0.2022 mg MMLS (mg COD)−1 and Kd = 0.00845–0.0135 d−1, respectively. Those properties of aerobic granules made aerobic granules system had a short setup period, high substrate utilization rate and low sludge production.  相似文献   

2.
The entomogenous fungus Cordyceps taii, a traditional Chinese medicinal mushroom, exhibits potent important pharmacological effects and it has great potential for health foods and medicine. In this work, the effects of oxygen supply on production of biomass and bioactive helvolic acid were studied in shake-flask fermentation of C. taii mycelia. The value of initial volumetric oxygen transfer coefficient (KLa) within 10.1–33.8 h−1 affected the cell growth, helvolic acid production and expression levels of biosynthetic genes. The highest cell concentration of 17.2 g/L was obtained at 14.3 h−1 of initial KLa. The highest helvolic acid production was 9.6 mg/L at 10.1 h−1 of initial KLa. The expression levels of three genes encoding hydroxymethylglutaryl-CoA synthase, hydroxymethylglutaryl-CoA reductase and squalene synthase were down-regulated on day 2 and day 8 but up-regulated on day 14 at an initial KLa value of 10.1 h−1 vs. 33.8 h−1, which well corresponded to the helvolic acid biosynthesis in those conditions. The information obtained would be helpful for improving the biomass and helvolic acid production in large-scale fermentation of C. taii.  相似文献   

3.
《Aquatic Botany》2005,81(2):157-173
The main photosynthesis and respiration parameters (dark respiration rate, light saturated production rate, saturation irradiance, photosynthetic efficiency) were measured on a total of 23 macrophytes of the Thau lagoon (2 Phanerogams, 5 Chlorophyceae, 10 Rhodophyceae and 6 Phaeophyceae). Those measurements were performed in vitro under controlled conditions, close to the natural ones, and at several seasons. Concomitantly, measurements of pigment concentrations, carbon, phosphorous and nitrogen contents in tissues were performed. Seasonal intra-specific variability of photosynthetic parameters was found very high, enlightening an important acclimatation capacity. The highest photosynthetic capacities were found for Chlorophyceae (e.g. Monostroma obscurum thalli at 17 °C, 982 μmol O2 g−1 dw h−1 and 9.1 μmol O2 g−1 dw h−1/μmol photons m−2 s−1, respectively for light saturated net production rate and photosynthetic efficiency) and Phanerogams (e.g. Nanozostera noltii leaves at 25 °C, 583 μmol O2 g−1 dw h−1 and 2.6 μmol O2 g−1 dw h−1/μmol photons m−2 s−1 respectively for light saturated net production rate and photosynthetic efficiency). As expected, species with a high surface/volume ratio were found to be more productive than coarsely branched thalli and thick blades shaped species. Contrary to Rd (ranging 6.7–794 μmol O2 g−1 dw h−1, respectively for Rytiphlaea tinctoria at 7 °C and for Dasya sessilis at 25 °C) for which a positive relationship with water temperature was found whatever the species studied, the evolution of P/I curves with temperature exhibited different responses amongst the species. The results allowed to show summer nitrogen limitation for some species (Gracilaria bursa-pastoris and Ulva spp.) and to propose temperature preferences based on the photosynthetic parameters for some others (N. noltii, Zostera marina, Chaetomorpha linum).  相似文献   

4.
《Aquatic Botany》2005,81(4):326-342
The effects of NH4+ or NO3 on growth, resource allocation and nitrogen (N) uptake kinetics of two common helophytes Phragmites australis (Cav.) Trin. ex Steudel and Glyceria maxima (Hartm.) Holmb. were studied in semi steady-state hydroponic cultures. At a steady-state nitrogen availability of 34 μM the growth rate of Phragmites was not affected by the N form (mean RGR = 35.4 mg g−1 d−1), whereas the growth rate of Glyceria was 16% higher in NH4+-N cultures than in NO3-N cultures (mean = 66.7 and 57.4 mg g−1 d−1 of NH4+ and NO3 treated plants, respectively). Phragmites and Glyceria had higher S/R ratio in NH4+ cultures than in NO3 cultures, 123.5 and 129.7%, respectively.Species differed in the nitrogen utilisation. In Glyceria, the relative tissue N content was higher than in Phragmites and was increased in NH4+ treated plants by 16%. The tissue NH4+ concentration (mean = 1.6 μmol g fresh wt−1) was not affected by N treatment, whereas NO3 contents were higher in NO3 (mean = 1.5 μmol g fresh wt−1) than in NH4+ (mean = 0.4 μmol g fresh wt−1) treated plants. In Phragmites, NH4+ (mean = 1.6 μmol g fresh wt−1) and NO3 (mean = 0.2 μmol g fresh wt−1) contents were not affected by the N regime. Species did not differ in NH4+ (mean = 56.5 μmol g−1 root dry wt h−1) and NO3 (mean = 34.5 μmol g−1 root dry wt h−1) maximum uptake rates (Vmax), and Vmax for NH4+ uptake was not affected by N treatment. The uptake rate of NO3 was low in NH4+ treated plants, and an induction phase for NO3 was observed in NH4+ treated Phragmites but not in Glyceria. Phragmites had low Km (mean = 4.5 μM) and high affinity (10.3 l g−1 root dry wt h−1) for both ions compared to Glyceria (Km = 6.3 μM, affinity = 8.0 l g−1 root dry wt h−1). The results showed different plasticity of Phragmites and Glyceria toward N source. The positive response to NH4+-N source may participates in the observed success of Glyceria at NH4+ rich sites, although other factors have to be considered. Higher plasticity of Phragmites toward low nutrient availability may favour this species at oligotrophic sites.  相似文献   

5.
α-Glucuronidase (EC 3.2.1.139) of family GH 115 from Scheffersomyces stipitis is a valuable enzyme for the modification of water-soluble xylan into insoluble biopolymers, due to its unique ability to act on polymeric xylans. The influence of growth rate on the production of α-glucuronidase by recombinant Saccharomyces cerevisiae MH1000pbk10D-glu in glucose-limited fed-batch culture was studied at 14 and 100 L scale. At and below the critical specific growth rate (μcrit) of 0.12 h−1 at 14 L scale, the biomass yield coefficient (Yx/s) remained constant at 0.4 g g−1 with no ethanol production, whereas ethanol yields relative to biomass (keth/x) of up to 0.54 g g−1 and a steady decrease in Yx/s were observed at μ > 0.12 h−1. Production of α-glucuronidase was growth associated at a product yield (kα-glu/x) of 0.45 mg g−1, with the highest biomass (37.35 g/L) and α-glucuronidase (14.03 mg/L) concentrations, were recorded during fed-batch culture at or near to μcrit. Scale-up with constant kLa from 14 to 100 L resulted in ethanol concentrations of up to 2.5 g/L at μ = 0.12 h−1. At this scale, α-glucuronidase yield could be maximised at growth rates below μcrit, to prevent localised high glucose concentration pockets at the feed entry zone that would induce oxido-reductive metabolism. This is the first report where recombinant production of α-glucuronidase (EC 3.2.1.139) by S. cerevisiae was optimised for application at pilot scale.  相似文献   

6.
In the current work nanoparticles (NPs) of α-amylase were generated in an aqueous solution using high-intensity ultrasound, and were subsequently immobilized on polyethylene (PE) films, or polycarbonate (PC) plates, or on microscope glass slides. The α-amylase NPs coated on the solid surfaces have been characterized by ESEM, TEM, FTIR, XPS and AFM. The substrates immobilized with α-amylase were used for hydrolyzing soluble potato starch to maltose. The amount of enzyme introduced in the substrates, leaching properties, and the catalytic activity of the immobilized enzyme were compared. The catalytic activity of the amylase deposited on the three solid surfaces was compared to that of the same amount of free enzyme at different pHs and temperatures. α-Amylase coated on PE showed the best catalytic activity in all the examined parameters when compared to native amylase, especially at high temperatures. When immobilized on glass, α-amylase showed better activity than the native enzyme over all pH and temperature values studied. However, the immobilization on PC did not improve the enzyme activity at any pH and any temperature compared to the free amylase. The kinetic parameters, Km and Vmax were also calculated. The amylase coated PE showed the most favorable kinetic parameters (Km = 5 g L−1 and Vmax = 5E−07 mol mL−1 min−1). In contrast, the anchored enzyme-PC exhibited unfavorable kinetic parameters (Km = 16 g L−1, Vmax = 4.2E−07 mol mL−1 min−1). The corresponding values for amylase-glass were Km = 7 g L−1, Vmax = 1.8E−07 mol mL−1 min−1, relative to those obtained for the free enzyme (Km = 6.6 g L−1, Vmax = 3.3E−07 mol mL−1 min−1).  相似文献   

7.
《Process Biochemistry》2010,45(7):1036-1042
A recombinant strain of Escherichia coli with CYP102A1 gene was developed for the demethylation of colchicine into their derivatives. The CYP102A1 gene responsible for demethylation was isolated from Bacillus megaterium ACBT03 and amplified using suitable primers. The amplified product was cloned into pET28a+ expression vector using host E. coli BL21(DE3) cells. The CYP3A4 (product of CYP102A1 gene) protein expression and other parameters like substrate toxicity, product toxicity and enzyme activity were optimized in shake flasks; and further scaled-up to 5 l bioreactor with 3 l working volume. In 5 l bioreactor, dissolved oxygen (DO) was optimized for maximum specific growth and enhanced 3-demethylated colchicine (3-DMC) production. The optimized conditions from shake flasks were scaled-up to 70 l bioreactor and resulted into ∼80% conversion of 20 mM colchicine in 48 h with a volumetric productivity of 6.62 mg l−1 h−1. Scale-up factors were measured as volumetric oxygen transfer coefficient (kLa) i.e., 56 h−1 and impeller tip velocity (Vtip) i.e., 7.065 m s−1, respectively. The kinetic parameters Km, kcat, and kcat/Km of the CYP3A4 enzyme using colchicine as the substrate were determined to be 271 ± 30 μM, 8533 ± 25 min−1, and 31.49 μM min−1, respectively, when IPTG induced recombinant E. coli culture was used.  相似文献   

8.
Aerobic production-scale processes are constrained by the technical limitations of maximum oxygen transfer and heat removal. Consequently, microbial activity is often controlled via limited nutrient feeding to maintain it within technical operability. Here, we present an alternative approach based on a newly engineered Escherichia coli strain. This E. coli HGT (high glucose throughput) strain was engineered by modulating the stringent response regulation program and decreasing the activity of pyruvate dehydrogenase. The strain offers about three-fold higher rates of cell-specific glucose uptake under nitrogen-limitation (0.6 gGlc gCDW−1 h−1) compared to that of wild type, with a maximum glucose uptake rate of about 1.8 gGlc gCDW−1 h−1 already at a 0.3 h−1 specific growth rate. The surplus of imported glucose is almost completely available via pyruvate and is used to fuel pyruvate and lactate formation. Thus, E. coli HGT represents a novel chassis as a host for pyruvate-derived products.  相似文献   

9.
《Harmful algae》2011,10(6):531-539
Temporal and spatial variability in the kinetic parameters of uptake of nitrate (NO3), ammonium (NH4+), urea, and glycine was measured during dinoflagellate blooms in Changjiang River estuary and East China Sea coast, 2005. Karenia mikimotoi was the dominant species in the early stage of the blooms and was succeeded by Prorocentrum donghaiense. The uptake of nitrogen (N) was determined using 15N tracer techniques. The results of comparison kinetic parameters with ambient nutrients confirmed that different N forms were preferentially taken up during different stages of the bloom. NO3 (Vmax 0.044 h−1; Ks 60.8 μM-N) was an important N source before it was depleted. NH4+ (Vmax 0.049 h−1; Ks 2.15 μM-N) was generally the preferred N. Between the 2 organic N sources, urea was more preferred when K. mikimotoi dominated the bloom (Vmax 0.020 h−1; Ks 1.35 μM-N) and glycine, considered as a dominant amino acid, was more preferred when P. donghaiense dominated the bloom (Vmax 0.025 h−1; Ks 1.76 μM-N). The change of N uptake preference by the bloom-forming algae was also related to the variation in ambient N concentrations.  相似文献   

10.
《Aquatic Botany》2002,72(3-4):219-233
We studied the potential role of dissolved inorganic carbon (DIC) in determining vegetation dominance of Potamogeton pectinatus L. and Chara aspera Deth. ex Willd. by monitoring the seasonal dynamics of DIC in a shallow lake and comparing the use of DIC of the two species. The HCO3-concentration in summer dropped from 2.5 to <0.5 mM with seasonally increasing Chara biomass, whereas outside the vegetation concentrations remained at 2.5 mM. Inside Potamogeton spp. vegetation DIC decreased from 2.5 to ca. 0.75 mM HCO3. A growth experiment showed ash-free biomass for P. pectinatus was nearly two times as high as for C. aspera at 3 mM HCO3, but almost two times lower at 0.5 mM than at 3.0. In a separate experiment, P. pectinatus precultured at a relatively low HCO3-level had a lower net photosynthetic rate (Pmax, 0.1 mmol O2 g−1 DW h−1) than C. aspera (Pmax, 0.1 mmol O2 g−1 DW h−1) over the range of HCO3-concentrations tested (Pmax, 0.14 mmol O2 g−1 DW h−1). In response to CO2 no significant differences between the compensation points (P. pectinatus, 28 mM; C. aspera 66 mM), were observed, but the photosynthetic rate increased faster than for C. aspera than for P. pectinatus. Under field conditions, the use of CO2 is not important since inside vegetation CO2-concentrations were below 10 μM, and thus, not available for photosynthesis of either species during the main part of the growth season. It is suggested that C. aspera may be a better competitor for HCO3 than P. pectinatus in conditions with a low HCO3 supply. As HCO3 is a strong limiting factor for growth inside the vegetation and probably the only carbon source available, the superior ability of C. aspera to use HCO3 may be an important factor explaining its present dominance in Veluwemeer.  相似文献   

11.
β-Glucosidase catalyzes the sequential breakdown of cyanogenic glycosides in cyanogenic plants. The β-glucosidase from Prunus armeniaca L. was purified to 8-fold, and 20% yield was obtained, with a specific activity of 281 U/mg protein. The enzyme showed maximum activity in 0.15 M sodium citrate buffer, pH 6, at 35 °C with p-nitrophenylglucopyranoside as substrate. The β-glucosidase from wild apricot was used successfully for the saccharification of cellobiose into D-glucose. This enzyme has a Vmax of 131.6 μmol min−1 mg−1 protein, Km of 0.158 mM, Kcat of 144.8 s−1, Kcat/Km of 917.4 mM−1 s−1, and Km/Vmax of 0.0012 mM min mg μmole−1, using cellobiose as substrate. The half-life, deactivation rate coefficient, and activation energy of this β-glucosidase were 12.76 h, 1.509 × 10−5 s−1, and 37.55 kJ/mol, respectively. These results showed that P. armeniaca is a potential source of β-glucosidase, with high affinity and catalytic capability for the saccharification of cellulosic material.  相似文献   

12.
Aerobic granular sludge was cultivated in a glass sequencing batch reactor (SBR) with glucose synthetic wastewater. The spherical shaped granules were observed on 4th day with the mean diameter of 0.1 mm. With the increase of chemical oxygen demand (COD) concentration of the influent, aerobic granules grew matured, the size of which ranged from 1.2 to 1.9 mm. The aerobic granular sludge could sustain high organic loading rate (about 4.0 g COD L−1 d−1), with good settling ability (settling velocity 36 m/h) and high biomass concentration (MLSS 6.7 ±0.2 g/L). Experimental data indicated that the substrate utilization and biomass growth kinetics followed Monod's kinetics model approximately. The corresponding kinetic coefficients of maximum specific substrate utilization rate (k), half velocity coefficient (Ks), growth yield coefficient (Y) and decay coefficient (Kd) were 13.2 d−1, 275.8 mg/L, 0.183–0.250 mg MLSS/mg COD and 0.023–0.075 d−1, respectively, which made aerobic granules have short setup period, high rate of substrate utilization and little surplus sludge.  相似文献   

13.
The aim of this study was to develop a bioprocess for l- and d-lactic acid production from raw sweet potato through simultaneous saccharification and fermentation by Lactobacillus paracasei and Lactobacillus coryniformis, respectively. The effects of enzyme and nitrogen source concentrations as well as of the ratio of raw material to medium were investigated. At dried material concentrations of 136.36–219.51 g L−1, yields of 90.13–91.17% (w/w) and productivities of 3.41–3.83 g L−1 h−1 were obtained with lactic acid concentrations as high as 198.32 g L−1 for l-lactic acid production. In addition, d-lactic acid was produced with yields of 90.11–84.92% (w/w) and productivities of 2.55–3.11 g L−1 h−1 with a maximum concentration of 186.40 g L−1 at the same concentrations of dried material. The simple and efficient process described in this study will benefit the tuber and root-based lactic acid industries without requiring alterations in plant equipment.  相似文献   

14.
《Process Biochemistry》2007,42(4):686-692
Pseudomonas putida 33 wild strain, subjected to gamma ray mutagenesis and designated as P. putida 300-B mutant was used as microbial rhamnolipid-producer by using distant carbon sources (viz. hydrocarbons, waste frying oils ‘WFOs’, vegetable oil refinery wastes and molasses) in the minimal media under shake flask conditions. The behavior of glucose as co-substrate and growth initiator was examined. The 300-B mutant strain showed its ability to grow on all the substrates tested and produced rhamnolipid surfactants to different extents however; soybean and corn WFOs were observed to be preferred carbon sources followed by kerosene and paraffin oils, respectively. The best cell biomass (3.5 g l−1) and rhamnolipids yield (4.1 g l−1) were obtained with soybean WFO as carbon source and glucose as growth initiator under fed-batch cultivation showing an optimum specific growth rate (μ) of 0.272 h−1, specific product yield (qp) of 0.318 g g−1 h and volumetric productivity (PV) of 0.024 g l−1 h. The critical micelle concentration of its culture supernatant was observed to be 91 mg rhamnolipids l−1 and surface tension as 31.2 mN m−1.  相似文献   

15.
In wild-type Escherichia coli, 1 mol of CO2 was fixated in 1 mol of succinic acid generation anaerobically. The key reaction in this sequence, catalyzed by phosphoenolpyruvate carboxylase (PPC), is carboxylation of phosphoenolpyruvate to oxaloacetate. Although inactivation of pyruvate formate-lyase and lactate dehydrogenase is found to enhance the PPC pathway for succinic acid production, it results in excessive pyruvic acid accumulation and limits regeneration of NAD+ from NADH formed in glycolysis. In other organisms, oxaloacetate is synthesized by carboxylation of pyruvic acid by pyruvate carboxylase (PYC) during glucose metabolism, and in E. coli, nicotinic acid phosphoribosyltransferase (NAPRTase) is a rate-limiting enzyme of the NAD(H) synthesis system. To achieve the NADH/NAD+ ratio decrease as well as carbon flux redistribution, co-expression of NAPRTase and PYC in a pflB, ldhA, and ppc deletion strain resulted in a significant increase in cell mass and succinic acid production under anaerobic conditions. After 72 h, 14.5 g L−1 of glucose was consumed to generate 12.08 g L−1 of succinic acid. Furthermore, under optimized condition of CO2 supply, the succinic acid productivity and the CO2 fixation rate reached 223.88 mg L−1 h−1 and 83.48 mg L−1 h−1, respectively.  相似文献   

16.
Cheese whey powder (CWP) solution with different CWP or sugar concentrations was fermented to ethanol in a continuous fermenter using pure culture of Kluyveromyces marxianus (DSMZ 7239). Sugar concentration of the feed CWP solution varied between 55 and 200 g l−1 while the hydraulic residence time (HRT) was kept constant at 54 h. Ethanol formation, sugar utilization and biomass formation were investigated as functions of the feed sugar concentration. Percent sugar utilization and biomass concentrations decreased and the effluent sugar concentration increased with increasing feed sugar concentrations especially for the feed sugar contents above 100 g l−1. Ethanol concentration and productivity (DP) increased with increasing feed sugar up to 100 g l−1 and then decreased with further increases in the feed sugar content. The highest ethanol concentration (3.7%, v v−1) and productivity (0.54 gE l−1 h−1) were obtained with the feed sugar content of 100 g l−1 or 125 g l−1. The ethanol yield coefficient (YP/S) was also maximum (0.49 gE gS−1) when the feed sugar was between 100 and 125 g l−1. The growth yield coefficient (YX/S) decreased steadily from 0.123 to 0.063 gX gS−1 when the feed sugar increased from 55 to 200 g l−1 due to adverse effects of high sugar contents on yeast growth. The optimal feed sugar concentration maximizing the ethanol productivity and sugar utilization was between 100 and 125 g l−1 under the specified experimental conditions.  相似文献   

17.
The anaerobic accumulation of several organic pollutants from industrial wastewaters, as storage substrates, and their subsequent aerobic biodegradation using a wastewater treatment mixed microbial culture for biological nutrient removal has been studied. The amount and the kinetics of substrate accumulation in the anaerobic stage depended on the characteristics of the wastewater fed to the anaerobic stage. Depending on the substrate used, levels of between 27 and 86% of storage polymers were accumulated with respect to the level obtained on feeding with acetate. The biodegradation kinetics were studied by modelling respirometry results. During the aerobic stage, oxygen-consumption data obtained in the respirometric tests were fitted to a model using a non-linear fitting estimation method. The simulation data obtained correlated well with the experimental oxygen-consumption data. The estimated kinetic parameters obtained indicate that each storage polymer was degraded at a different rate. However, the values obtained for the storage polymer half-saturation coefficient, KS: 16 mg COD l−1, and for the coefficient for endogenous respiration, b: 0.008 h−1, were similar in all the experiments. The results indicate that each substrate produces the synthesis of a specific storage polymer that is degraded at a different rate.  相似文献   

18.
A study was conducted on H2S removal in a biotrickling filter packed with open-pore polyurethane foam. Thiobacillus denitrificans was used as inoculum and a mixed culture population was developed during the process. The inhibitory effect of sulphate concentration (1.8–16.8 g L−1), pH (6.9–8.6), trickling liquid velocity (TLV, 9.1–22.8 m h−1), H2S inlet concentration (20–157 ppmv) and the empty bed residence time (EBRT, 9–57 s) on the H2S removal efficiency (RE) were thoroughly investigated. An increase in pH from 6.9 to 8.5 led to a corresponding increase in H2S removal. In addition, an inhibitory effect of sulphate concentration was observed from 16.8 g L−1 and the maximum elimination capacity was found to be 22 gS m−3 h−1 (RE 98%). The RE was constant (98.8 ± 0.30%) for EBRT  16 s, but a decrease in the EBRT from 16 to 9 s led to a corresponding decrease in RE from 98.2 to 89.6% for a TLV of 9.1 m h−1 and from 97.9 to 94.9% for a TLV of 22.8 m h−1 (inlet load of 11.0 ± 0.2 gS m−3 h−1). The sulphur oxidation capacity in the biotrickling filter was not diminished by the presence of other bacteria.  相似文献   

19.
In this study, the soap stock as a sole carbon source was used for growing a carotenoid producing yeast (Rhodotorula rubra). The application of soap stock resulted in increase of carotenoids yield up to 5.36 folds when compared with the grown cultures on glucose. On the best Monod equation fitted on the specific growth rate (μ) data, the maximum specific growth rate (μm) and half-saturation concentration (KS) were respectively determined at 0.064 h−1 and 3.26 g L−1 for total fatty acids presented in soap stock. Further tests on the carotenogenesis process were carried out in a cell-immobilized airlift photobioreactor where the natural loofa sponge was used for immobilization of the cells. The performance of the bioreactor was statistically studied by the response surface methodology (RSM) where aeration rate of 0.11 vvm and light irradiation intensity of 2517 Lx provided an optimum condition for producing β-carotene with a specific production rate of 22.65 mg gcell−1 day−1.  相似文献   

20.
《Aquatic Botany》2005,82(3):181-192
The influence of nitrogen and phosphorus pulses on Chaetomorpha linum (Muller) Kutzing growth and photosynthesis was studied in laboratory experiments. Photosynthesis and growth of C. linum from Tancada lagoon seems limited by both nitrogen and phosphorus, as indicated by the high rate (4.7–11.6 mg O2 g−1 dry weight h−1) of light-saturated photosynthesis (Pm) and growth rates observed under nitrogen plus phosphorus enrichment in relation to enrichment by nitrogen alone (2.9–7.6 mg O2 g−1 dry weight h−1). Significant increase in nitrogen and phosphorus content as percentage of dry weight was observed in C. linum fertilized with a single nutrient or with nitrogen plus phosphorus. In Tancada lagoon, when availability of nitrogen to primary producers is by pulses, an increase of nitrate concentration in the water column (from 6 to 100 μM) has a greater effect on growth of C. linum (growth rate: 0.13 day−1) than an increase in ammonium concentration (from 20 to 100 μM and growth rate: 0.11 day−1). For a given thallus nitrogen content (0.6–1.4% N), both Pm and the photosynthetic efficiency (α) normalized to dry weight were correlated (r2 = 0.73, p < 0.005) indicating that variations in electron transport were coupled to variations in C-fixation capacity. Optimizing both α and Pm may be a general characteristic of thin-structured opportunistic algae in more variable estuarine environments.  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