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.
Yang X  Chen H  Xu W  He Z  Ma M 《Plant cell reports》2007,26(10):1889-1897
The callus of Pteris vittata was induced from gametophytes generated from spores in vitro, and grew rapidly with periodical medium change. Arsenic tolerance and accumulation of P. vittata callus were compared with those of Arabidopsis thaliana callus. Cell death was not detected in P. vittata callus even at arsenate concentrations up to 2 mM; however, A. thaliana callus died at low (0.2 mM) arsenate concentrations. Meanwhile, P. vittata callus accumulated almost three times more As than A. thaliana callus when exposed to 0.2 mM arsenate. About 60% of the total As was removed when 7.5 g of P. vittata callus was cultured on 150 ml of half-strength MS liquid medium containing 450 μg As for 2 days. Furthermore, P. vittata callus, sporophytes, and gametophytes all grew well under 1 mM of arsenate and accumulated 1,250; 1,150 and 2,180 mg kg−1 dry weight As when grown on 2 mM arsenate for 15 or 30 days. The characteristics of non-differentiated cells, large biomass, ease of culture, good synchronization, and excellent As sequestering, make the callus of P. vittata a new ideal system to study the mechanisms of As hyperaccumulation and phytoremediation in As-contaminated groundwater.  相似文献   

2.
Rui  Li  Jiajun  Li  Chaoyuan  Wu 《Hydrobiologia》1990,(1):499-503
Kappaphycus alvarezii, cultured in a region deficient in nitrogen with intermittent application of 10 mM ammonium at three-day intervals for 1 hour, attained a daily growth rate of 4.6%, which was considered to be the optimum growth rate. Under such conditions, the C/N ratio was 29 with a carrageenan content of 58% and gel strength of 45–70 g cm–2. Gel strength of carrageenan was high when nitrogen content was high and vice versa.  相似文献   

3.
Lu CT  Mei XG 《Biotechnology letters》2003,25(17):1437-1439
When, on the 15th day of growth, an elicitor from Fusarium solani was added at 40 mg l–1 to Cistanche deserticola cell suspension cultures, the contents of echinacoside, acteoside and total phenylethanoid glycosides (PeGs) in cultured cells all increased over the next 27 d by over 100% to 15 mg g–1 dry wt, 9 mg g–1 dry wt and 57 mg g–1 dry wt, respectively. The final biomass (1.3 mg dry wt ml–1) was not affected.  相似文献   

4.
Summary Saccharomyces cerevisiae in the form of baker's yeast, cells cultivated on a yeast extract-peptone-glucose medium, as well as cells immobilized in 18% (w/v) polyacrylamide gel showed the ability to hydrolyze 1.727 mM sodium phytate solution at 45°C, pH 4.6, in a stirred tank reactor. Seventy percent yield of dephosphorylation was observed after 2 h using a baker's yeast concentration of 5.8 g dry matter per 100 ml. Hydrolytic activity at 1.8–2.0 M Pi min–1 was observed between 1st and 3rd h of the reaction in cells cultured 24 or 48 h. No inhibition by the substrate was found at sodium phytate concentrations of 0.587–1.727 mM. After 1.5 h of hydrolysis a single, well distinguished peak ofmyo-inositol-triphosphate was the main product found. By means of immobilization the stability of the biocatalyst was enhanced 3.3-fold and reached its half-life at 64 ninety-minute runs.  相似文献   

5.
Sulphate uptake by Amphidinium carterae, Amphidinium klebsii and Gymnodinium microadriaticum grown on artificial seawater medium with sulphate, cysteine, methionine or taurine as sulphur source occurred via an active transport system which conformed to Michaelis-Menten type saturation kinetics. Values for K m ranged from 0.18–2.13 mM and V max ranged from 0.2–24.2 nmol · 105 cells–1 · h–1. K m for symbiotic G. microadriaticum was 0.48 mM and V max was 0.2 nmol · 105 cells–1 · h–1. Sulphate uptake was slightly inhibited by chromate and selenate, but not by tungstate, molybdate, sulphite or thiosulphate. Cysteine and methionine (0.1 mM), but not taurine, inhibited sulphate uptake by symbiotic G. microadriaticum, but not by the two species of Amphidinium. Uptake was inhibited 45–97% under both light and dark conditions by carbonylcyanide 3-chlorophenylhydrazone (CCCP); under dark conditions sulphate uptake was 40–60% of that observed under light conditions and was little affected by 3-(3,4-dichlorophenyl) 1,1-dimethylurea (DCMU).The uptake of taurine, cysteine and methionine by A. carterae, A. klebsii, cultured and symbiotic G. microadriaticum conformed to Michaelis-Menten type saturation kinetics. K m values of taurine uptake ranged from 1.9–10 mM; for cysteine uptake from 0.6–3.2 mM and methionine from 0.001–0.021 mM. Cysteine induced a taurine uptake system with a K m of 0.3–0.7 mM. Cysteine and methionine uptake by all organisms was largely unaffected by darkness or by DCMU in light or darkness. CCCP significantly inhibited uptake of these amino acids. Thus energy for cysteine and methionine uptake was supplied mainly by respiration. Taurine uptake by A. carterae was independent of light but was inhibited by CCCP, whereas uptake by A. klebsii and symbiotic G. microadriaticum was partially dependent on photosynthetic energy. Taurine uptake by cultured G. microadriaticum was more dependent on photosynthetic energy and was more sensitive to CCCP. Cysteine inhibited uptake of methionine and taurine by cultured and symbiotic G. microadriaticum to a greater extent than in the Amphidinium species. Methionine did not greatly affect taurine uptake, but did inhibit cysteine uptake. Taurine did not affect the uptake of cysteine or methionine.  相似文献   

6.
Callus cultures were initiated from soybean (Glycine max (L.) Merr cv. Acme) cotyledons onMiller's basal medium supplemented with 2 mg L–1NAA and 0.5 mg L–1 kinetin. Growing cells wereexposed to increasing concentrations of NaCl in themedium. A concentration of 100 mM NaCl completelyinhibited callus growth. After incubation for 28 d,cells which could tolerate this concentration of NaClgrew to form cell colonies. A NaCl-tolerant line wasobtained through continuous subculturing on 100 mMNaCl. Salt tolerance in this culture was characterizedby an altered growth behavior, reduced cell volume, and accumulation of Na+, Cl, proline and sugars when grown under salt stress, as well as on normal media. These characteristics, which proved tobe stable after the culture was transferred to asalt-free medium, is commonly associated with halophytes. Presented data suggest that this salt tolerance is the result of a shift towards a halophytic behavior.  相似文献   

7.
Four from 18 strains of Erwinia herbicola tested had nitrogenase activity and grew with N2 as sole source of nitrogen under strict anaerobic conditions with a doubling time of 20–24 h. Nitrogenase activity started only 96–120 h after transfer to a special medium maintained under anaerobic conditions. A ten fold increase in protein per culture found after the maximum nitrogenase activity of 80–130 nmol C2H4. mg protein-1·min-1 was accompanied by a fall in pH of the medium (20 mM phosphate buffer and in 125 mM Tris-buffer) from pH 7.2 to 5.4 or less, but only to 6.8 in 100 mM phosphate buffer. In all cases we found a sharp curtailing of nitrogenase activity 48 h after the maximum. The bacteria utilized only 35–50% of the nitrogen fixed for growth. Erwinia herbicola strains differed from two strains of Enterobacter agglomerans in being unable to fix nitrogen on agar surfaces exposed to air. Specific nitrogenase activity in Erwinia herbicola is compared with data reported for other Enterobacteriaceae and is found to be higher than that reported for Klebsiella pneumoniae, Enterobacter cloacae or Citrobacter freundii.  相似文献   

8.
Biodegradation of toxic amides by immobilized Rhodococcus rhodochrous NHB-2 has been studied to generate data for future development of reactors for the treatment of simulated wastewater containing various toxic amides. The whole resting cells were immobilized in different matrices like agar, polyacrylamide and alginate. Agar gel beads were selected for the treatment of simulated wastewater containing 100mM each acetamide, propionamide, and 10mM of acrylamide and packed in a highly compact five-stage plug flow reactor. The immobilized bacterium worked well in a broad pH range from 5 to 10, with an optimum at 8.7. The apparent K m-value for the turnover of acetamide for the resting cells was determined to be around 40mM at pH 8.5 and 55°C, whereas the K m-value of the purified amidase was predicted to be about 20 mM. This organism exhibited greater turnover of aliphatic amides as compared to aromatic amides. Although these cells showed maximal amide-degrading activity at 55°C, simulated wastewater treatment was carried out at 45°C, because of the greater stability of the amidase activity at that temperature. Of note, indices for overall temperature stability, based on the temperature dependence of apparent first order kinetic temperature denaturation constants, were determined to be –7.9±1.1×10–4, and –13.7±1.3×10–4, –14.5±0.7×10–4, and –13.7±0.8×10–4°Cmin, for free cells and cells immobilized in alginate, agar and polyacrylamide respectively. After 250min the reactor showed maximum degradation of acetamide, propionamide and acrylamide of about 97, 100 and 90%, respectively by using 883 enzyme activity units per reactor stage. The results of this investigation showed that R. rhodochrous NHB-2 expressing thermostable amidase could be used for the efficient treatment of wastewater containing toxic amides. Therefore, we suggest that this microbe has a very high potential for the detoxification of toxic amides from industrial effluents and other wastewaters.  相似文献   

9.
Specific nitrogenase activity inAzospirillum brasilense ATCC 29145 in surface cultures under air is enhanced from about 50 nmol C2H4·mg protein-1·h-1 to 400 nmol C2H4 by the addition of 1 mM phenol. 0.5 and 2 mM phenol added increase the rate 5-fold and 4-fold. This enhancement effect is observed only between 2 and 3 days after inoculation, with only a small reduction of the growth of the cells by the phenol added. In surface cultures under 1% O2, nitrogenase activity is slightly reduced by the addition of 1–0.01 mM phenol. Utilization of succinate is enhanced during the period of maximum enhancement of nitrogenase activity by 60% by addition of 1 mM phenol. The cells did not produce14CO2 from [U-14C] phenol, neither in surface cultures nor in liquid cultures and less than 0.1% of the phenol was incorporated into the cells. A smaller but significant enhancement of nitrogenase activity by about 100% in surface cultures under air was found withKlebsiella pneumoniae K 11 after addition of 1 mM phenol. However, inRhizobium japonicum 61-A-101 all phenol concentrations above 0.01 mM reduced nitrogenase activity. With 1 mM phenol added activity was reduced to less than 10% with no effect on the growth in the same cultivation system. With thisRhizobium japonicum strain significant quantities of phenol (25 mol in 24 h by 2·1012 cells) were metabolized to14CO2, with phenol as sole carbon source. WithAzospirillum brasilense in liquid culture under 1% and 2% O2 in the gas phase, no enhancement of nitrogenase activity by phenol was noticed.  相似文献   

10.
The signal events of 1 mM Ce4+ (Ce(NH4)2(NO3)6)-induced apoptosis of cultured Taxus cuspidata cells were investigated. The percentage of apoptotic cells increased from 0.82% to 51.32% within 6 days. Caspase-3-like protease activity became notable during the second day of Ce4+-treatment, and the maximum activity was 5-fold higher than that of control cells at the fourth day. When the experiment system was pretreated with acetyl-Asp-Glu-Val-Asp-aldehyde (Ac-DEVD-CHO) at 100 M, caspase-3-like activity resulted in distinct inhibition by 70% and 77.3% after 3 and 4 days of induction. Furthermore, 100 M Ac-DEVD-CHO partially reduced the apoptotic cells by 58.6% and 60.8% at day 4 and 5 respectively. Ce4+ induced superoxide anions (O2·–) transient burst, and the first peak appeared at around 3.7–4 h, the second appeared at about 7 h. Both O2·– burst and cell apoptosis were effectively suppressed by application of diphenyl iodonium (NADPH oxidase inhibitor). Inhibition of O2·– production attenuated caspase-3-like activation by 49% and 53.6% during day 3 and 4 respectively. In addition, a total of 15 protein spots changed in response to caspase-3-like protease activation were identified by two-dimensional gel electrophoresis. These results suggest that Ce4+ of 1 mM induces apoptosis in suspension cultures of T. cuspidata through O2·– burst as well as caspase-3-like protease activation. The burst of O2·– exerts its activity as an upstream of caspase-3-like activation. Our results also implicate that other signal pathways independent of an O2·– burst possibly participate in mediating caspase-3-like protease activation.  相似文献   

11.
A simple enzyme mixture containing 2% Cellulase Onozuka R–10 and1% Macerozyme R–10 prepared in deionised water supplemented with 3% NaCland 1 mM CaCl2 was developed for isolating rapidlyprotoplasts from different species of Monostroma,Enteromorpha and Ulva. The yield fordifferent species of Monostroma ranged from 9.6 ×106 to 10.2 × 106 cells g–1f. wt thallus, and forEnteromorpha from 3.48 × 106 to 11.7× 106 cells g–1 f. wt and forUlva from 4.58 × 106 to 26.8 ×106 cells g–1 f. wt. The overallregeneration rate of the protoplasts isolated was usually > 90% and showednormal morphogenesis. The method yields rapid mass production of viableprotoplasts with high regeneration rates.  相似文献   

12.
Enargite (Cu3AsS4) was leached at 70°C by Sulfolobus BC in shake-flasks. The highest copper dissolution (52% after 550 h of leaching) was obtained with bacteria and 1 g l–1 ferric ion. In the absence of ferric ion, Sulfolobus BC catalyzes the bioleaching of enargite through a direct mechanism after adhesion onto the mineral surface. In ferric bioleaching, arsenic precipitated as ferric arsenate and arsenic remained associated to the solid residues, preventing the presence of a high dissolved arsenic concentration in the leaching solution. About 90% inhibition of bacterial growth rate and activity was observed for dissolved arsenic concentrations above 600 mg l–1 for As(III) and above 1000 mg l–1 for As(V). Arsenic-bearing copper ores and concentrates could be leached by Sulfolobus BC in the presence of ferric iron due to the favourable precipitation of arsenic ion as ferric arsenate, avoiding significant bacterial inhibition.  相似文献   

13.
The effects of concentration of amino acids, nitrate, and ammonium on the growth and taxol production in cultures of cell line TY-21 of Taxus yunnanensis were investigated. Addition of 20 different amino acids each at 15–20 mg l–1 to B5 medium significantly improved callus growth but inhibited taxol formation in the cultures. The optimum nitrate concentration was 20–30 mM for both growth and taxol production. Ammonium greatly suppressed growth but strongly promoted taxol formation in the cells when it was the sole inorganic nitrogen in the medium. Culturing the suspension cells in nitrate-containing medium for 15 days and then in a medium in which ammonium was the sole inorganic nitrogen for 7 days increased taxol yield by 104%, reaching up to 28.1 mg l–1.  相似文献   

14.
Two strains ofLeptospirillum-like bacteria, L6 and L8, have been isolated from a mixed inoculum, also containingThiobacillus ferrooxidans andT. thiooxidans, cultured for one year with a colbaltiferous pyrite as energy substrate in a 100 I continuous bioleaching laboratory unit. Several physiological properties of the strains are described. The vibrio-shaped microorganisms grew at pH values lower than 1.3. Their growth rate was maximum between 2.5 and 8.0 g l1 ferrous iron. The optimal growth temperature was 37.5° C. Ferric iron had a stimulative effect on bacterial development up to 8 g l–1, and growth was as rapid at 14 g l–1 ferric iron as at 8 g l–1. The negative influence of cobalt on the final cell concentration was observed at 0.5 g l–1, but the growth rate was not affected up to 2 g l–1. The G + C content of strains L8 is 55.6 mol%.  相似文献   

15.
Summary Whole cells and cell-free extracts ofAcinetobacter calcoaceticus containing benzoylformate decarboxylase efficiently condensed benzoylformate and acetaldehyde to produce the acyloin compoundS-(–)-2-hydroxypropiophenone. Optimal concentrations of acetaldehyde cosubstrate for this reaction were found to be 1600 and 800 mM when whole cells and cell-free extracts were used respectively as biocatalysts. In both cases, optimal benzoylformate concentration was 100 mM. Temperature and pH optima for the biotransformation reaction were 30°C and 6.0 respectively. Under optimised conditions, maximum production of 2-hydroxypropiophenone, amounting to 8.4 g L–1, occurred after a 2-h incubation. Product formation equivalent to 6.95 g in 1 h corresponded to a productivity of 267 mg acyloin per g dry cells per h.  相似文献   

16.
The role of carbon dioxide in glucose metabolism of Bacteroides fragilis   总被引:2,自引:0,他引:2  
The effect of CO2 concentration on growth and glucose fermentation of Bacteroides fragilis was studied in a defined mineral medium. Batch culture experiments were done in closed tubes containing CO2 concentrations ranging from 10% to 100% (with appropriate amounts of bicarbonate added to maintain the pH at 6.7). These experiments revealed that CO2 had no influence on growth rate or cell yield when the CO2 concentration was above 30% CO2 (minimum available CO2–HCO 3 - , 25.5 mM), whereas a slight decrease in these parameters was observed at 20% and 10% CO2 (available CO2–HCO 3 - , 17 and 8.5 mM, respectively). If CO2–HCO 3 - concentrations were below 10 mM, the lag phase lengthened and a decrease in maximal growth rate and cell yield were observed. The amount of acetate made decreased, while d-lactate concentration increased. A net production of CO2 allowed growth under conditions of extremely low concentrations of added CO2.When B. fragilis was grown in continuous culture with 100% CO2 or 100% N2, the dilution rate influenced the concentrations of acetate, succinate, propionate, d-lactate, l-malate and formate formed. Decreasing the dilution rate favored propionate and acetate production under both conditions. When the organism was grown with 100% N2, the amount of propionate formed was greater than the amount of succinate formed at all dilution rates. Except at slow dilution rates the reverse was true when 100% CO2 was used. B. fragilis was unable to grow at dilution rates faster than 0.154 h-1 when grown with 100% N2; the Y glc max was 67.9 g DW cells/mol glucose and m s was 0.064 mmol glucose/g DW·h. If the gas atmosphere was 100% CO2 the organism was washed out of the culture when the dilution rate exceeded 0.38 h-1; the Y glc max was 59.4 g DW cells/mol glucose and m s was 0.094 mmol glucose/g DW·h.Measurement of the phosphoenolpyruvate (PEP) carboxykinase (E.C. 4.1.1.49) with whole, permeabilized cells of B. fragilis showed an increase of specific enzyme activity with decreasing CO2 concentrations. The mechanisms used by B. fragilis to adjust to low levels of CO2 are discussed.  相似文献   

17.
The effects of nitrate and silicate levels, and carbon source on growth, biochemical composition and fatty acid composition ofNitzschia inconspicua were investigated using batch cultures. Within the range of silicate levels supplied (8.8–176 M), no marked variations in growth trend, biochemical composition or fatty acid composition were shown. Biomass at stationary phase, ranging from 64–66 mg ash-free dry weight (AFDW) L–1, and specific growth rate () based on chlorophylla (0.41–0.50 d–1) of the cultures grown within 0.3–3.0 mM NaNO3 were not significantly different. Cultures supplemented with glucose (0.1 % w/v), acetate (0.1 % w/v) or 5% CO2 attained higher biomass (85, 85, 97 mg AFDW L–1) than the control which was grown in synthetic seawater and agitated by magnetic stirring. Cells grown at <3.0 mM NaNO3 contained higher carbohydrate contents (14.8–21.5% AFDW) than those grown at 3.0 mM (4.0% AFDW). Lipid content increased at the expense of proteins in cells aerated with 5% CO2. The dominant fatty acids, 16:0 and 16:1, ranged from 35.7–45.0% and 36.4–45.4% total fatty acids (TFA), respectively, while the relative proportions of 20:4 (n-6) and 20:5 (n-3) ranged from 1.7–5.4% and 3.4–5.9% TFA respectively. Cultures aerated with 5% CO2 attained the highest biomass (97 mg AFDW L–1) and yield of 20:5 (n-3) (0.34 mg L–1).  相似文献   

18.
Callus cultures of Saussurea medusa were cultivated on solid culture medium supplemented with either Ce3+, La3+, Nd3+ or a mixture of rare earth elements. Ce3+, 0.05 mM, gave the highest biomass (0.53 g dry wt per flask) and total flavonoids (27.5 mg per flask), which were, 70% and 100% higher than those without Ce3+ addition, respectively. Ce3+, 0.01–0.1 mM, or La3+, 0.05 mM, or the mixture of rare earth elements, 0.025–0.1 mM, can substitute for 6-benzyladenine, and 0.025 mM Ce3+ can partly substitute for naphthaleneacetic acid in promoting cell growth and biosynthesis of total flavonoids in S. medusa.  相似文献   

19.
A Pseudomonas sp. strain NGK1 (NCIM 5120) capable of utilizing 2-methylnaphthalene (2-MN) was immobilized in various matrices namely, polyurethane foam (PUF), alginate, agar and polyvinyl alcohol (PVA) (1.5 × 1012 c.f.u. g–1 beads). The degradation rates of 25 and 50 mM 2-MN by freely suspended cells (2 × 1011 c.f.u. ml–1) and immobilized cells in batches, semi-continuous with shaken culture and continuous degradation in a packed-bed reactor were compared. The PUF-immobilized cells achieved higher degradation of 25 and 50 mM of 2-MN than freely suspended cells and the cells immobilized in alginate, agar or PVA. The PVA- and PUF-immobilized cells could be reused for more than 30 and 20 cycles respectively, without losing any degradation capacity. The effect of dilution rates on the rate of degradation of 25 and 50 mM 2-MN with freely suspended and immobilized cells were compared in the continuous system. Increase in dilution rate increased the degradation rate only up to 1 h–1 in free cells with 25 mM 2-MN and no significant increase was observed with 50 mM 2-MN. With immobilized cells, the degradation rate increased with increase in dilution rate up to 1.5 h–1 for 25 mM and 1 h–1 for 50 mM 2-MN. These results revealed that the immobilized cell systems are more efficient than freely suspended cells for biodegradation of 2-MN.  相似文献   

20.
The marine PrasinophyteTetraselmis may be cultured under both mixotrophic (photoheterotrophic) and heterotrophic conditions. The growth rate was slightly lower, and pigment levels and lipid composition were radically affected on heterotrophic culture in 1 L fermenters. Total chlorophyll levels of dark grown cultures were less than 1% of those observed in mixotrophically grown cells, the chlorophylla : b ratio also decreased as did the carotenoid content. In addition, the total amounts of lipids including polyunsaturated fatty-acids were also lower in heterotrophically cultured cells: 6.4 mg g–1 (dried alga) and 0.35 mg g–1 (dried alga); as compared to 37.1 mg g–1 (dried alga) and 18.5 mg g–1 (dried alga), for cells grown in the light. However, gross morphology and final yield (>16 g l–1) were relatively unaffected. The algae produced were spray-dried and tested for their suitability as an aquaculture feed.Address 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