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.
This study assessed the effect of two precursors (l-phenylalanine and p-amino benzoic acid) used alone or in combination with methyl jasmonate, on the growth and accumulation of paclitaxel, baccatin III and 10-deacetylbaccatin III in hairy root cultures of Taxus x media var. Hicksii. The greatest increase in dry biomass was observed after 4 weeks of culturing hairy roots in medium supplemented with 1 μM of l-phenylalanine (6.2 g L−1). Addition of 1 μM of l-phenylalanine to the medium also resulted in the greatest 10-deacetylbaccatin III accumulation (422.7 μg L−1), which was not detected in the untreated control culture. Supplementation with 100 μM of l-phenylalanine together with 100 μM of methyl jasmonate resulted in the enhancement of paclitaxel production from 40.3 μg L−1 (control untreated culture) to 568.2 μg L−1, the highest paclitaxel content detected in the study. The effect of p-amino benzoic acid on taxane production was less pronounced, and the highest yield of paclitaxel (221.8 μg L−1) was observed when the medium was supplemented with 100 μM of the precursor in combination with methyl jasmonate.Baccatin III was not detected under the conditions used in this experiment and the investigated taxanes were not excreted into the medium.  相似文献   

2.
Medium-chain-length polyhydroxyalkanoates (MCL-PHAs) were produced in carbon-limited, single-stage, fed-batch fermentations of Pseudomonas putida KT2440 by co-feeding nonanoic acid (NA) and glucose (G) to enhance the yield of PHA from NA. An exponential (μ = 0.25 h−1) followed by a linear feeding strategy at a NA:G ratio of 1:1 (w/w) achieved 71 g l−1 biomass containing 56% PHA. Although the same overall PHA productivity (1.44 g l−1 h−1) was obtained when NA alone was fed at the same specific growth rate, the overall yield of PHA from NA increased by 25% (0.66 g PHA g NA−1 versus 0.53 g g−1) with glucose co-feeding. Further increasing glucose in the feed (NA:G = 1:1.5) resulted in a slightly higher yield (0.69 g PHA g NA−1) but lower PHA content (48%) and productivity (1.16 g l−1 h−1). There was very little change in the PHA composition.  相似文献   

3.
Field and laboratory experiments were designed to determine the differential growth and toxin response to inorganic and organic nitrogen additions in Pseudo-nitzschia spp. Nitrogen enrichments of 50 μM nitrate (KNO3), 10 μM ammonium (NH4Cl), 20 μM urea and a control (no addition) were carried out in separate carboys with seawater collected from the mouth of the San Francisco Bay (Bolinas Bay), an area characterized by high concentrations of macronutrients and iron. All treatments showed significant increases in biomass, with chlorophyll a peaking on days 4–5 for all treatments except urea, which maintained exponential growth through the termination of the experiment. Pseudo-nitzschia australis Frenguelli abundance was 103 cells l−1 at the start of the experiment and increased by an order of magnitude by day 2. Particulate domoic acid (pDA) was initially low but detectable (0.15 μg l−1), and increased throughout exponential and stationary phases across all treatments. At the termination of the experiment, the urea treatment produced more than double the amount of pDA (9.39 μg l−1) than that produced by the nitrate treatment (4.26 μg l−1) and triple that of the control and ammonium treatments (1.36 μg l−1 and 2.64 μg l−1, respectively). The mean specific growth rates, calculated from increases in chlorophyll a and from cellular abundance of P. australis, were statistically similar across all treatments.These field results confirmed laboratory experiments conducted with a P. australis strain isolated from Monterey Bay, CA (isolate AU221-a) grown in artificial seawater enriched with 50 μM nitrate, 50 μM ammonium or 25 μM of urea as the sole nitrogen source. The exponential growth rate of P. australis was significantly slower for cells grown on urea (ca. 0.5 day−1) compared to the cells grown on either nitrate or ammonium (ca. 0.9 day−1). However the urea-grown cells produced more particulate and dissolved domoic acid (DA) than the ammonium- or nitrate-grown cells. The field and laboratory experiments demonstrate that P. australis is able to grow effectively on urea as the primary source of nitrogen and produced more pDA when grown on urea in both natural assemblages and unialgal cultures. These results suggest that the influence of urea from coastal runoff may prove to be more important in the development or maintenance of toxic blooms than previously thought, and that the source of nitrogen may be a determining factor in the relative toxicity of west coast blooms of P. australis.  相似文献   

4.
Can soil Chytridiomycota survive and grow in different osmotic potentials?   总被引:1,自引:1,他引:0  
Twenty isolates from soil in the orders Spizellomycetales, Blastocladiales and Chytridiales (Chytridiomycota) grew on complex solid media supplemented with 10 g l−1 sodium chloride. In a synthetic liquid medium, 4.4 g l−1 sodium chloride strongly inhibited growth in three of the five isolates, possibly because of the effect of the ions or osmolarity of the solution. The maximum concentration for growth in synthetic liquid medium with different osmotic potentials using polyethylene glycol (PEG) varied considerably amongst the isolates. Three patterns of growth with increasing concentrations of PEG were evident among isolates within the genus Rhizophydium. Up to the concentration where growth ceased, the dry weight of each isolate either decreased, remained constant, or in one case, increased. Most of the fungi survived when incubated at room temperature for 7 d in complex liquid media supplemented with 35 g l−1 sodium chloride or 300 g l−1 PEG. These data indicate that soil Chytridiomycota can survive various osmotic potentials that may occur during the wetting and drying phases in soils.  相似文献   

5.
The phytoplankton communities and the production of cyanobacterial toxins were investigated in two alkaline Kenyan crater lakes, Lake Sonachi and Lake Simbi. Lake Sonachi was mainly dominated by the cyanobacterium Arthrospira fusiformis, Lake Simbi by A. fusiformis and Anabaenopsis abijatae. The phytoplankton biomasses measured were high, reaching up to 3159 mg l−1 in L. Sonachi and up to 348 mg l−1 in L. Simbi. Using HPLC techniques, one structural variant of the hepatotoxin microcystin (microcystin-RR) was found in L. Sonachi and four variants (microcystin-LR, -RR, -LA and -YR) were identified in L. Simbi. The neurotoxin anatoxin-a was found in both lakes. To our knowledge this is the first evidence of cyanobacterial toxins in L. Sonachi and L. Simbi. Total microcystin concentrations varied from 1.6 to 12.0 μg microcystin-LR equivalents g−1 DW in L. Sonachi and from 19.7 to 39.0 μg microcystin-LR equivalents g−1 DW in L. Simbi. Anatoxin-a concentrations ranged from 0.5 to 2.0 μg g−1 DW in L. Sonachi and from 0 to 1.4 μg g−1 DW in L. Simbi. In a monocyanobacterial strain of A. fusiformis, isolated from L. Sonachi, microcystin-YR and anatoxin-a were produced. The concentrations found were 2.2 μg microcystin g−1 DW and 0.3 μg anatoxin-a g−1 DW. This is the first study showing A. fusiformis as producer of microcystins and anatoxin-a. Since A. fusiformis occurs in mass developments in both lakes, a health risk for wildlife can be expected.  相似文献   

6.
Effects of carbon concentration and carbon to nitrogen (C:N) ratio on six biocontrol fungal strains are reported in this paper. All fungal strains had extensive growth on the media supplemented with 6–12 g l−1 carbon and C:N ratios from 10:1 to 80:1, and differed in nutrient requirements for sporulation. Except for the two strains of Paecilomyces lilacinus, all selected fungi attained the highest spore yields at a C:N ratio of 160:1 when the carbon concentration was 12 g l−1 for Metarhizium anisopliae SQZ-1-21, 6 g l−1 for M. anisopliae RS-4-1 and Trichoderma viride TV-1, and 8 g l−1 for Lecanicillium lecanii CA-1-G. The optimal conditions for P. lilacinus sporulation were 8 g l−1 carbon with a C:N ratio of 10:1 for M-14 and 12 g l−1 carbon with a C:N ratio of 20:1 for IPC-P, respectively. The results indicated that the influence of carbon concentration and C:N ratio on fungal growth and sporulation is strain dependent; therefore, consideration for the complexity of nutrient requirements is essential for improving yields of fungal biocontrol agents.  相似文献   

7.
The dimorphic fungus Mucor indicus was grown in different forms classified as purely filamentous, mostly filamentous, mostly yeast-like and purely yeast-like, and the relationship between morphology and metabolite production, inhibitor tolerance and the cell wall composition was investigated. Low concentrations of spores in the inoculum with subsequent aeration promoted filamentous growth, whereas higher spore concentrations and anaerobic conditions promoted yeast-like growth. Ethanol was the main metabolite with glycerol next under all conditions tested. The yields of ethanol from glucose were between 0.39 and 0.42 g g−1 with productivities of 3.2–5.0 g l−1 h−1. The ethanol productivity of mostly filamentous cells was increased from 3.9 to 5.0 g l−1 h−1 by the presence of oxygen, whereas aeration of purely yeast-like cells showed no such effect. All growth forms were able to tolerate 4.6 g l−1 furfural and 10 g l−1 acetic acid and assimilate the sugars, although with different consumption rates. The cell wall content of the fungus measured as alkali insoluble materials (AIM) of the purely yeast-like cells was 26% of the biomass, compared to 8% of the pure filaments. However, the chitosan concentration of the filaments was 29% of the AIM, compared to 6% of the yeast-like cells.  相似文献   

8.
Forty species of fungi, representing a range of ecological and taxonomic groups, were tested for their ability to grow on agar media amended with lithium chloride (LiCl) at 1.5, 3 and 6 g l−1. Species of Trichoderma varied considerably in their sensitivity to LiCl; at one week on 6 g l−1 LiCl medium, the growth of seven species of Trichoderma was considerably inhibited; however, by three weeks at this level, four of the species tested were able to attain ≥30 % of control growth. Of the seven species tested, an isolate of T. viride was the most sensitive to LiCl in agar. Eleven other imperfect fungi also showed a range of ability to grow on agar amended with LiCl, from total inhibition to complete lack of inhibition. Six ascomycete fungi were greatly inhibited by LiCl at all levels; however, an isolate of Chaetomium globosum was highly tolerant of LiCl. Seven basidiomycete wood-decay fungi were quite sensitive to LiCl in agar, showing total to nearly total inhibition even at the lowest level; however, after three weeks, an isolate of Postia placenta was nearly uninhibited except at 6 g l−1. Five ectomycorrhizal basidiomycete fungi were totally inhibited by all levels of LiCl; however, one ectomycorrhizal imperfect fungus (Cenococcum graniforme) was able to grow at 3 g l−1 and was uninhibited at 1.5 g l−1. Four zygomycete fungus isolates were nearly unaffected in their growth by all levels of LiCl.  相似文献   

9.
Large-scale collections of Zygnemataceae in the continental United States of America were made between March and August in 1982, 1983, and 1984. Collections were made on a 31000-km transect through 35 states. Zygnemataceae were found at 318 sites was inspected. Temperature average 19°C and pH averaged 6.1 over all sites. Algal strains in collections were identified to genus, characterized for filament width, chloroplast number, and end wall type, then photographed and isolated into unialgal culture. Spirogyra was the most common genus collected(632 strains), followed in abundance by Zygnema (174 Strains) and Mougeotia (135 strains). These three genera contained 95% of the strains collected and were equally widely distributed. Strains of the three genera frequently occurred together; no genus displayed evidence of habitat specialization among the three habitat types: flowing water, permanent ponds or lakes, and temporary pools. In Spirogyra, strains with plane (flat) end walls were four times more abundant than those with replicate (interlocking) end walls. Spirogyra with plane end walls showed more variation in filament width than Zygnema, Mougeotia, or Spirogyra with replicate end walls. In Spirogyra with plane end walls, filament width was correlted with nuclear DNA content and number of strains found per collection site was twice that of other genera or Spirogyra, with replicate end walls. Spirogyra strains wider than 70 μm were more frequent on the northern part of the transect. It is proposed that polyploidy may be of widespread occurrence in Spirogyra with plane end walls and that associated morphological plasticity may account for the high apparent specied diversity and survival of the genus in a wider variety of microhabitats than occupied by other Zygnemataceae.  相似文献   

10.
Ergosterol contents of six wood-rotting basidiomycetes were analyzed under different cultivation conditions. Four white-rot and two brown-rot fungi were cultivated in liquid synthetic medium with low nutrient nitrogen (2 mM) and 0.1% glucose, and ergosterol in mycelial biomasses were measured weekly for 35 days. The highest ergosterol content per fungal dry mass in the white-rot fungi was found in Phanerochaete chrysosporium being 2100 μg g−1, while in Ceriporiopsis subvermispora it was 1700 μg g−1, Phlebia radiata 700 μg g−1, and Physisporinus rivulosus 560 μg g−1. In brown-rot fungi the ergosterol content was in Poria placenta 2868 μg g−1 and in Gloeophyllum trabeum 3915 μg g−1. On agar media, P. chrysosporium and P. radiata reached the highest ergosterol value in 7 days, while in wood block cultures the ergosterol contents were quite stable. The conversion factors for ergosterol-to-fungal biomass varied from 48 and 243, which were lower than values for ascomycetous soil fungi reported in the literature.  相似文献   

11.
The effect of elicitation with linoleic (C18:2) and α-linolenic (C18:3) fatty acids (LLA and α-LNA) was investigated in Panax ginseng C.A. Meyer adventitious roots cultured in 5 l balloon-type bioreactors. Fatty acids were added in culture medium at 0.0, 1.0, 2.5, 5.0, 10.0, and 20.0 μmol l−1 at day 40, at the end of exponential growth phase. Roots were harvested and assayed at day 47. Elicitation with both LLA and α-LNA enhanced accumulation of total polyphenolics and flavonoids in roots compared with control without elicitation. The highest accumulation of flavonoids was observed at 5.0 μmol l−1 for both elicitors. Total phenolics production reached its highest value of about 4.0 mg g−1 DW under the elicitation with 5.0 μmol l−1 LLA and 5.0–20.0 μmol l−1 α-LNA. Meanwhile, α-LNA was more effective than LLA for increasing biomass and ginsenoside production. The biomass of 11.1 g DW l−1 and maximal total ginsenoside content of 7.9 mg g−1 DW were achieved at 5 μmol l−1 α-linolenic acid. The essential polyunsaturated linoleic (C18:2) and α-linolenic (C18:3) fatty acids were accumulated in roots in response to elicitation while content of palmitic (C16:0) and oleic (C18:1) acids declined. The activities of SOD, G-POD and CAT were enhanced by two elicitors to similar extent while APX activity was preferably stimulated by α-LNA. Our results demonstrate that elicitation with α-linolenic acid stimulates production of biomass and secondary metabolites in bioreactor-cultured P. ginseng adventitious roots.  相似文献   

12.
In vitro grown inflorescences of Bambusa edulis were used to investigate the process of vegetative shoot growth in detail. The findings revealed that auxins and ACC could be significant growth regulators in this process. Overall, auxins [NAA, indolebutyric acid (IBA), and 2,4-dichlorophenoxyacetic acid (2,4-D)] induced inflorescences to grow vegetative shoots. However, the efficiency of shoot regeneration varied. A greater percentage (27.3–34.5) of inflorescences in the 5 mg l−1 NAA, 10 mg l−1 NAA, and 1 mg l−1 2,4-D treatments formed more vegetative shoots than those exposed to other treatments. IBA promoted shoot regeneration less effectively than NAA and 2,4-D. Fifty percent of regenerated vegetative shoots flowered after 2 months when the medium was supplemented with 5 mg l−1 NAA. All shoots that received 1 mg l−1 1-amino-cyclopropane-1-carboxylic acid (ACC) flowered in 5 mg l−1 NAA medium. Rooted plantlets were used to examine their survival following in vitro flowering. All plantlets with vegetative shoots, even those with inflorescences, survived and grew.  相似文献   

13.
The aim of this study in the field was to investigate whether there are differences between the outer archipelago (Gullmar Fjord) and a semi-enclosed fjord system (Koljö Fjord) in occurrences of D. acuta and D. acuminata as well as in their content of diarrheic shellfish toxin (DST) per cell. When all data pairs of cell toxicity of D. acuminata and the corresponding number of cells l−1 from the two sites were tested in a regression analysis, a statistically significant negative correlation became evident and was apparent as a straight line on a log–log plot (p < 0.0001). Obviously, there was an overall inverse relationship between the population density of D. acuminata and the toxin content per cell. Plotted on a linear scale, all data-pairs of cell toxicity and cell number made up a parabolic curve. On this curve the data-pairs could be separated into three groups: (i) D. acuminata occurring in numbers of fewer than approximately 100 cells l−1, and with a toxin content per cell above 5 ρg cell−1; (ii) cell numbers between 100 and approximately 250 cells l−1 with a cell toxin content from 5 to 2 ρg cell−1; (iii) when the population became greater than 250 cells l−1, the toxicity, with few exceptions, was less than 2 ρg cell−1. By applying this subdivision, some clear patterns of the distribution of the differently toxic D. acuminata became evident. When comparing the cell toxicity of the two sites, it was obvious that the D. acuminata cells from all depths from the Gullmar Fjord as a mean were significantly more toxic compared to the Koljö Fjord samples. The results have demonstrated that approximately 100 high-toxicity cells in a low-density population at surface may lead to the same accumulation of DST in a mussel as the ingestion of 1500 low-toxicity cells from a high-density pycnocline population.  相似文献   

14.
Substantial mortalities of Atlantic salmon (Salmo salar) at two aquaculture sites in Long Island Sound, off Grand Manan Island, Bay of Fundy (BoF) (New Brunswick, Canada) in September 2003, were associated with a bloom of Alexandrium fundyense (>3 × 105 cells L−1), a dinoflagellate alga that produces toxins which cause paralytic shellfish poisoning (PSP). Cells of A. fundyense collected from surface waters while fish were dying had total paralytic shellfish (PS) toxin concentrations of 70.6 pg STX equiv. (saxitoxin equivalents) cell−1 and PS toxin profiles rich in carbamate toxins (78.2%). The zooplankton sampled contained PS toxins (63.1 pg STX equiv. g−1 wet wt) and the toxin profile matched that of A. fundyense cells.Mean PS toxin levels were low (<4 μg STX equiv. 100 g−1 wet wt) in stomach, gill and muscle tissues of moribund salmon, suggesting that PS toxins are very lethal to salmon.The PS toxin concentrations in blue mussels (Mytilus edulis) growing on the salmon cages (37; 526 μg STX equiv. 100 g−1 wet wt) were the highest recorded to date from this region. Their PS toxin profiles showed enhanced carbamate contents (85.5%) compared with that found in A. fundyense. Blue mussels collected from an adjacent Canadian Food Inspection Agency (CFIA) monitoring site in Grand Manan had PS toxin concentrations of 4214 and 150 μg STX equiv. 100 g−1 wet wt in late September and December, respectively, well above the regulatory limit (RL), and horse mussels (Modiolus modiolus) collected in late September had PS toxin concentrations of 2357 μg STX equiv. 100 g−1 wet wt. Detoxification under laboratory conditions suggested that blue mussels may require up to 19 weeks for elimination below RL when they accumulate these high concentrations of PS toxins. This depuration period may be shorter in the field.PS toxin levels above RL were detected in hepatopancreatic tissues of lobster (Homarus americanus), with lower levels (<16 μg STX equiv. 100 g−1 wet wt) in tail muscle and gills.These results illustrate the movement of PS toxins through the marine food chain following an A. fundyense bloom in the BoF, and support earlier studies suggesting that kills from the region of zooplanktivorous fish, such as herring (Clupea harengus harengus), can be attributed to blooms of A. fundyense. This is the first reported incident of PSP associated with mortalities of caged Atlantic salmon in the BoF. Analyses of muscle tissues and viscera from the affected salmon indicated that any portion would not be a health hazard if consumed.  相似文献   

15.
Biomass, primary production and nutrient budgets associated to Sarcocornia perennis subspecies (ssp.) alpini were studied in the Palmones River estuary salt marsh (Southern Spain) to evaluate the nutrient sequestration capacity of the low marsh. Above- and belowground living and dead biomass, as well as carbon, nitrogen and phosphorus content were monitored during 1 year. Additionally, the fate of aboveground detritus was evaluated in an experiment on litter decomposition. The detritus production of S. perennis ssp. alpini was almost equivalent to its annual primary production indicating a rapid turnover of biomass. We calculated that only 12% of the aboveground detritus was exported out of the low marsh while the rest was decomposed in the sediment with a rate of 0.8 year−1. Changes in concentrations of total carbon, nitrogen and phosphorus in the sediment showed patterns related to S. perennis ssp. alpini belowground biomass. Our results suggested that the sediment functions as a net sink for nutrients accumulating 550 g C m−2 year−1, 55 g N m−2 year−1, and 13 g P m−2 year−1.  相似文献   

16.
Removal of phosphorus (P) by Ceratophyllum demersum L. and associated epiphytic periphyton was quantified by measuring the disappearance of soluble reactive P (SRP) from microcosms during 1-h in situ incubations conducted over a 1-year period. Initial P concentrations in these incubations ranged from 30 to >10,000 μg P L−1. Phosphorus removal was proportional to initial P concentrations and was weakly correlated with solar irradiance and water temperature. Removal rates (0.6–32.8 mg P m−2 d−1) and kv coefficients (0.68–1.93 h−1) from experiments run at low initial P concentrations (up to 200 μg P L−1) were comparable to results reported for other macrophytes. Removal rates from experiments run at the highest (>10,000 μg P L−1) initial P concentrations (5300 and 11,100 mg P m−2 d−1) most likely represented luxury nutrient consumption and were not thought to be sustainable long term. We were unable to determine a Vmax for P removal, suggesting that the nutrient-storage capability of the C. demersum/periphyton complex was not saturated during our short-term incubations. Based on N:P molar ratios, the marsh was P limited, while the C. demersum/periphyton complex was either N limited or in balance for N and P throughout this study. However, despite its tissue stoichiometry, the C. demersum/periphyton complex always exhibited an affinity for P. It appeared that the biochemical mechanisms, which mediate P removal, at least on a short-term basis, were more influenced by increases in ambient P levels than by tissue nutrient stoichiometry.  相似文献   

17.
Laccase-catalyzed oxidation of phenolic compounds in organic media   总被引:1,自引:0,他引:1  
Rhus vernificera laccase-catalyzed oxidation of phenolic compounds, i.e., (+)-catechin, (−)-epicatechin and catechol, was carried out in selected organic solvents to search for the favorable reaction medium. The investigation on reaction parameters showed that optimal laccase activity was obtained in hexane at 30 °C, pH 7.75 for the oxidation of (+)-catechin as well as for (−)-epicatechin, and in toluene at 35 °C, pH 7.25 for the oxidation of catechol. Ea and Q10 values of the biocatalysis in the reaction media of the larger log p solvents like isooctane and hexane were relatively higher than those in the reaction media of lower log p solvents like toluene and dichloromethane. Maximum laccase activity in the organic media was found with 6.5% of buffer as co-solvent. A wider range of 0–28 μg protein/ml in hexane than that of 0–16.7 μg protein/ml in aqueous medium was observed for the linear increasing conversion of (+)-catechin. The kinetic studies revealed that in the presence of isooctane, hexane, toluene and dichloromethane, the Km values were 0.77, 0.97, 0.53 and 2.9 mmol/L for the substrate of (+)-catechin; 0.43, 0.34, 0.14 and 3.4 mmol/L for (−)-epicatechin; 2.9, 1.8, 0.61 and 1.1 mmol/L for catechol, respectively, while the corresponding Vmax values were 2.1 × 10−2, 2.3 × 10−2, 0.65 × 10−2 and 0.71 × 10−2 δA/μg protein min); 1.8 × 10−2, 0.88 × 10−2, 0.19 × 10−2 and 1.0 × 10−2 δA/μg protein min); 0.48 × 10−2, 0.59 × 10−2, 0.67 × 10−2 and 0.54 × 10−2 δA/μg protein min), respectively. FT-IR indicated the formation of probable dimer from (+)-catechin in organic solvent. These results suggest that this laccase has higher catalytic oxidation capacity of phenolic compounds in suitable organic media and favorite oligomers could be obtained.  相似文献   

18.
A massive outbreak of Karenia brevis that had been ongoing for several months along the southwestern coast of Florida was sampled in early September 2005 off Sanibel Island to assess the utility of bio-optical features and ataxonomic analysis (quantification of eukaryotic and cyanobacterial picoplankton) by flow cytometry in monitoring red tide blooms. Sea-surface sampling followed aircraft visual location of discolored water. Within the most concentrated area of the bloom, chlorophyll a values exceeded 500 μg l−1, and concentrations of nitrate (0.3 μM ± 0.0) and ammonium (<0.2 μM) were depleted compared to high concentrations of total dissolved nitrogen, total dissolved phosphorus, and soluble reactive phosphorus (141 ± 34 μM, 16.5 ± 2.5 μM, and 6.44 ± 0.57 μM, respectively). Low water clarity in the bloom (Secchi depth transparency 0.3 m, Kd estimated at 4.83 m−1) was strongly influenced by attenuation from dinoflagellates as well as chromophoric dissolved organic matter (CDOM). The fact that the K. brevis bloom occurred in lower-salinity (30 psu), high-nutrient waters implicates riverine transport of land-based nutrients as a source of nutrient supplies that fueled or sustained the bloom. Throughout ongoing efforts to advance modeling and technological capabilities that presently lack reliable predictive capability, bio-optical remote sensing via aerial flyovers along with in-water sensor data can continue to provide accurate coverage of relatively large temporal and spatial features. Flow cytometry can provide conservative (because of some cell lysis), rapid, near-real-time validation of bloom components. The concentration and position of the organisms, along with water mass scalars, can also help to diagnose factors promoting K. brevis bloom development and dispersion.  相似文献   

19.
Red tide blooms of Cochlodinium polykrikoides in a coastal cove   总被引:1,自引:0,他引:1  
Successive blooms of the dinoflagellate Cochlodinium polykrikoides occurred in Pettaquamscutt Cove, RI, persisting from September through December 1980 and again from April through October 1981. Cell densities varied from <100 cells L−1 at the onset of the bloom and reached a maximum density exceeding 3.4 × 106 cells L−1 during the summer of 1981. The bloom was mainly restricted to the mid to inner region of this shallow cove with greatest concentrations localized in surface waters of the southwestern region during summer/fall periods of both years. Highly motile cells consisting of single, double and multiple cell zooids were found as chains of 4 and 8 cells restricted to the late August/September periods. The highest cell densities occurred during periods when annual temperatures were between 19 and 28 °C and salinities between 25 and 30. A major nutrient source for the cove was Crying Brook, located at the innermost region at the head of the cove. Inorganic nitrogen (NH3 and NO2 + NO3) from the brook was continually detectable throughout the study with maximum values of 57.5 and 82.5 μmol L−1, respectively. Phosphate (PO4-P) was always present in the source waters and rarely <0.5 μmol L−1; silicate always exceeded 30 μmol L−1 with maximum concentrations reaching 226 μmol L−1. Chlorophyll a and ATP concentrations during the blooms varied directly with cell densities. Maximum Chl a levels were 218 mg m−3 and ATP-carbon was >20 g C m−3. Primary production by the dinoflagellate-dominated community during the bloom varied between 4.3 and 0.07 g C m−3 d−1. Percent carbon turnover calculated from primary production values and ATP-carbon varied from 6 to 129% d−1. The dinoflagellates dominated the entire summer period; other flagellates and diatoms were present in lesser amounts. A combination of low washout rate due to the cove dynamics, active growth, and life cycles involving cysts allowed C. polykrikoides to maintain recurrent bloom populations in this area.  相似文献   

20.
Different induction strategies for fed-batch recombinant protein production under the control of the strong T5 promoter in Escherichia coli have been investigated. Since the production of recombinant rhamnulose 1-phosphate aldolase is growth-related, the productivity of the process can be strongly reduced due to the negative effect of protein expression on cell growth. IPTG pulse induction as well as inducer dosage have been applied and their advantages and drawbacks highlighted. Both strategies led to high levels of the recombinant protein, 1000 AU g DCW−1. Inducer concentration and inducer to biomass ratio were identified as the parameters influencing the rate of protein production and final enzymatic activity per gram of biomass. In pulse induction, the maximum enzymatic activity was found at inducer concentration of 70 μM. In continuous induction experiments, inducer concentrations between 4 and 12 μM were identified as the working range in which cell growth and recombinant protein accumulation occurred simultaneously. On the other hand, the amount of IPTG per gram of biomass required was 1.6 μmol IPTG gDCW−1 in pulse induction and between 0.3 and 0.5 μmol IPTG g DCW−1 in continuous induction.  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