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 31 毫秒
1.
Wen  Zhao  Shuang-Lin  Dong 《Hydrobiologia》2003,492(1-3):181-190
Primary productivity, biomass and chlorophyll-a of size fractionated phytoplankton (<0.22 m, <3 m, <8 m, <10 m, <40 m, <64 m, <112 m and <200 m) were estimated in 6 ponds and 5 experimental enclosures. The results showed that the planktonic algae less than 10 m are important in the biomass and production of phytoplankton in saline–alkaline ponds. The production of size fractionated phytoplankton corresponding to <112 m, <10 m and <3 m in saline–alkaline ponds were 10.5 ± 6.6 , 8.6 ± 5.4 and 0.33 ± 0.1 mgC l–1 d–1, respectively. Mean community respiration rate was 1.80 ± 0.73, 1.69 ± 0.90 and 1.38 ± 1.12 mgC l–1 d–1, respectively. The average production of phytoplankton corresponding to micro- (10–112 m), nano- (3–10 m) and pico- (<3 m) were 1.61, 8.30 and 0.33 mgC l–1 d–1, respectively. The ratio of those to the total phytoplankton production was 15%, 79% and 3%, respectively. The mean respiration rate of the different size groups was 0.11, 0.31 and 1.38 mgC l–1 d–1; the ratio of those to total respiration of phytoplankton was 6%, 17% and 77%, respectively. The production of size-fractionated phytoplankton corresponding to <200 m, <10 m and <3 m in enclosures was 2.19 ± 1.63, 2.08 ± 1.75 and 0.22 ± 0.08 mgC l–1 d-1, respectively. Mean community respiration rates were 1.25 ± 1.55, 1.17 ± 1.42 and 0.47 ± 0.32 mgC l–1 d–1, respectively. The average production of phytoplankton corresponding to micro- (10–200 m), nano- (3–10 m) and pico- (<3 m) plankton was 0.11, 1.86 and 0.22 mgC l–1 d–1, respectively. The ratio of those to the total production of phytoplankton was 5%, 85% and 10%, respectively. The mean respiration rate of different size groups were 0.08, 0.72 and 0.46 mgC l–1 d–1, the ratio of those to total respiration of phytoplankton was 6%, 57% and 37%, respectively. The concentrations of chlorophyll-a of the phytoplankton in the corresponding size of micro- (10–112 m), nano- (3–10 m) and pico- (<3 m) plankton in the experimental ponds were 19.3, 98.2 and 11. 9 g l–1, respectively. The ratio of those to the total chlorophyll-a was 15%, 76% and 9%, respectively. The concentrations of chlorophyll-a of phytoplankton micro- (10–200 m), nano- (3–10 m) and pico- (<3 m) plankton in enclosures were 1.7, 34.3 and 3.0 g l–1, respectively. The ratio of those to the total chlorophyll-a was 4%, 88% and 8%, respectively.  相似文献   

2.
Callus was initiated from immature leaf and stem segments of rose (Rosa hybrida cv. Landora) and subcultured every four weeks on a basal medium of half-strength Murashige & Skoog (1962) salts plus 30 g l-1 sucrose (1/2 MS) and supplemented with 2.2 M BA, 5.4 M NAA and 2.2–9.0 M 2,4-D. Embryogenic callus and subsequently somatic embryos were obtained from 8-week-old callus culture on 1/2 MS+2.2 M BA+0.05 M NAA+0.3 M GA3+200–800 mg l-1 L-proline. Long-term cultures were established and maintained for up to 16 months by repeated subculture of embryogenic callus on L-proline deficient medium. About 12% of cotyledonary stage embryos taken from cultures cold-stored at 8±1°C for 4 days germinated on 1/2 MS+2.2 M BA+0.3 M GA3+24.7 M adenine sulphate.Abbreviations BA benzyladenine - NAA -naphthaleneacetic acid - 2,4-D 2,4-dichlorophenoxyacetic acid - GA3 gibberellic acid  相似文献   

3.
Three-year-old spruce (Picea abies) saplings were planted and cultivated for 2 years in pots with 3 1 substrate, consisting of a homogenized mixture of sand, peat and forest soil with a high organic content (volume ratio 11.52). This substrate was amended with 10–180 mol Cd [kg soil dry weight (DW)]–1, 50–7500 mol Zn (kg soil DW)–1 (determined with 1 M ammonium acetate extracts) or combinations of both elements. Annual xylem growth rings in stems of plants treated with 50 mol Cd (kg soil DW)–1 or 7500 mol Zn (kg soil DW)–1 were significantly narrower than in control plants. Growth reductions were more pronounced in the second year of the experiment. The contents of Cd and Zn in stem wood and needles were positively correlated with the substrate concentrations. The Mg contents of the spruce needles were inversely correlated with soil concentrations of Cd and Zn. Root development was impeded at moderate concentrations of Cd (50 mol kg–1) or Zn (1000 mol kg–1) in the substrate. The adverse effects of potentially toxic trace elements, like Cd or Zn, on xylem growth of spruce plants are discussed with regard to possible growth reductions in forest trees under field conditions.  相似文献   

4.
In order to maintain axenic seedstock cultures axenically of thecommercially important red seaweed, Porphyra yezoensis, aprocedure was developed for axenic isolation and culture of conchocelis andmonospores. For axenic isolation of the conchocelis, contaminated microalgaewere most effectively removed by filtering contaminated samples through a100-m mesh after sonication. Removal of bacteria and otheralgaewas accomplished using a mixture of 5 agents (0.02% chitosan, 100 gml–1 GeO2, 10 gml–1 ampicillin, 40 gml–1 kanamycin and 200 gml–1 streptomycin). Axenic single colonies wereisolatedfrom a semi-solid medium prepared from 1% transfer gel. After collectingmonospores from the 40–50% density layer on a percoll-gradient, removalofbacteria and fungi from the monospores was accomplished using a mixture of 5antibiotics (3.5 g ml–1 nystatin, 2 mgml–1 ampicillin, 400 gml–1 kanamycin, 50 gml–1 neomycin and 800 gml–1 streptomycin). Axenic single juvenile blades wereisolated from a semi-solid medium prepared from 0.5% transfer gel.  相似文献   

5.
The application of abscisic acid (ABA), either as a racemic mixture or as optically resolved isomers, increases freezing tolerance in a bromegrass (Bromus inermis Leyss) cell culture and induces the accumulation of several heat-stable proteins. Two stereoisomers of an ABA analog, 23 dihydroacetylenic abscisyl alcohol (DHA), were used to study the role of ABA-induced processes in the acquisition of freezing tolerance in these cells. Freezing tolerance was unchanged in the presence of (–) DHA (LT50 -9°C), and no increase in heat-stable protein accumulation was detected; however, the (+) enantiomer increased the freezing tolerance (LT50 -13°C) and induced the accumulation of these polypeptides. All three forms of ABA increased freezing tolerance in the bromegrass cells, although (–) ABA was less effective than either (+) or (±) ABA when added at equal concentrations. Cells pretreated with 20 or 50 M (–) DHA displayed lower levels of freezing tolerance following the addition of 2.5, 7.5 or 25 M (±) ABA. Full freezing tolerance could be restored by increasing the concentration of (±) ABA to > 25 M. Pretreatment of cells with (–) DHA (20 or 50 M) had no effect on freezing tolerance when 25 M (+) ABA was added. The induction of freezing tolerance by 25 M (–) ABA was completely inhibited by the presence of 20 M (–) DHA. The accumulation of ABA-responsive heat-stable proteins was inhibited by pretreatment with 20 M (–) DHA in cells treated with 2.5 or 7.5M (+) ABA, and in cells treated with 25 M (–) ABA. The accumulation of these polypeptides was restored when (±) or (+) ABA was added at a concentration of 25 M. The analysis of proteins which cross-reacted with a dehydrin antibody revealed a similar inhibitory pattern as seen with the other ABA-responsive proteins. The effects of the various isomers of ABA and DHA on cell osmolarity and sucrose uptake was also investigated. In both cases, (±) and (+) ABA had pronounced effects on the parameters measured, whereas (–) ABA treated cells gave substantially different results. In both sucrose uptake and cell osmolarity, DHA had no significant effect on the results obtained following (±) or (+) ABA treatment. Maximum freezing tolerance was only observed in cells when both heat-stable protein accumulation and sucrose uptake were observed.Abbreviations ABA abscisic acid - DHA 2,3 dihydroacetylenicabscisyl alcohols - DMSO dimethyl sulfoxide - LT50 temperature at which 50% of cells are killed The authors would like to acknowledge the technical assistance of Angela Bollman, Bruce Ewan and Angela Shaw. This work was supported by grants from the Natural Science and Engineering Research Council of Canada to L.V.G. and N.H.L., and a grant from the University of Saskatchewan to R.W.W.  相似文献   

6.
Summary Isolated and homogenised Deiters' neurons from the lateral vestibular nucleus of rabbit in a Krebs-Ringer solution containing no added Mg++, 1.3 moles/ ml and 5 moles/ml Mg++, broke down ATP at the maximal rate of 0.29+-0.20, 2.40+–0.20, and 5.95+–0.63 moles/cell/hr. In 1.3 mM Mg++ solution the single cell homogenates took up phosphate at the mean rate of 2.6+–0.2 moles/cell/hr. If the rabbits were injected 1 hour before with 20 mg/kg body weight of 2-amino-1-propene-1,1,3, tricarbonitrile (triap), the breakdown of ATP in these latter media was 0.82+–0.44, 2,5+–0.60, and 6.7+– 1.1 moles/cell/hr, respectively, and the quantity of inorganic liberated did not decrease.  相似文献   

7.
Pesticides and heavy metals in Danish streambed sediment   总被引:2,自引:0,他引:2  
Kronvang  B.  Laubel  A.  Larsen  S. E.  Friberg  N. 《Hydrobiologia》2003,494(1-3):93-101
The role of streambed sediment as a sink for pesticides and heavy metals was investigated in 30 Danish lowland streams. The investigated streams drain catchments varying in hydrology, topography, soil type and land use. The <250 m newly accumulated fraction of the uppermost 1–2 cm layer of streambed sediment was analysed for 19 old and modern pesticides and 9 heavy metals. DDE was present in the sediment of all the streams. Of the herbicides, fungicides and insecticides currently in use, the most frequently detected was diuron (50.0%), fenpropimorph (66.7%) and lambda-cyhalothrin (6.7%), respectively. The pesticides detected in the highest concentration were fenpropimorph (1700 ng g–1), propiconazole (130 ng g–1) and isoproturon (110 ng g–1). The heavy metals are listed in order of increasing median concentration: Cd (0.80 g g–1), Co (9.1 g g–1), As (12.0 g g–1), Ni (19.0 g g–1), Cr (19.2 g g–1), Pb (19.7 g g–1), Cu (20.1 g g–1), V (28.5 g g–1), Zn (103 g g–1). The average number of pesticides detected in the 27 streams draining predominantly agricultural catchments was (3.7±2.0) being higher (p=0.077) than in the three streams draining non-agricultural catchments (1.7±0.6). Pesticides were significantly related to catchment size, soil type and hydrological regime. Several heavy metals (Cr, Cu, Pb, V and Zn) were related to urban activity and soil type.  相似文献   

8.
Armengol  X.  Boronat  L.  Camacho  A.  Wurtsbaugh  W. A. 《Hydrobiologia》2001,(1):107-114
Grazing rates of zooplankton were analysed in the summer of 1999 in Yellow Belly Lake, an oligotrophic system in the Sawtooth Mountains of Idaho (U.S.A.). The colonial rotifer Conochilus unicornis was a dominant species in the epilimnion, with densities reaching 20 colonies l–1 (ca. 400 ind. l–1). Clearance rates were measured with an in situ Haney Grazing chamber and synthetic microspheres 5, 9 and 23m in diameter. At epilimnetic temperatures of around 14 °C, mean clearance rates for 5m particles ranged from 30 to 65 l ind.–1 h –1. Clearance rates were 2–9 times higher on the 5m spheres than on the 9 m spheres, and C. unicornis almost never fed on the 23 m spheres. Grazing rates did not change over the diel cycle. Clearance rates declined more than 10-fold as temperatures declined from 14 °C in the epilimnion to 7 °C in the metalimnion. In the epilimnion, grazing by C. unicornis was more important than grazing by crustaceans in the community, at least on particles 9m. The results show the importance of grazing by rotifers in lakes, and the significance of spatial variations that influence grazing rates.  相似文献   

9.
This study addresses the temporal distribution of forms of phosphorus in the soil of a temporarily flooded riparian forest of the valley of the river Garonne (Southwest of France). A sequential extraction for forms of phosphorus of increasing chemical stability was used. During the study period (13 months), the forest was flooded a few days during March and May. In winter, resin-Pi concentration was high (26 g g–1) in comparison to spring values (<9 g g–1). NaHCO3-Po, NaHCO3-Pi or NaOH-Pi concentrations increased during winter (up to 74, 124 and 78 g g–1 respectively) and decreased significantly during spring (32, 44 and 32 g g–1 respectively). This pattern was attributed to simultaneous mineralization and plant uptake during the growing season and to the flood events (erosional processes and P-release). During summer and fall, resin-Pi concentration increased significantly (up to 26 g g–1 in October). NaHCO3-Po concentrations remained low during spring and summer (<33 g g–1), and increased significantly in fall (>45 g g–1 NaHCO3-Pi or NaOH-Pi increased in late spring or summer (90 g g–1 and 68 g g–1 respectively). Increasing concentrations of the labile forms during late spring or summer were ascribed to the warm temperature and soil dryness that limited plant growth. HCl-Pi increased regularly after the floods (174 g g–1 before the flood events to 254 g g–1 after the floods). Residual P presented a similar pattern i.e. 214 g g–1 and 279 g g–1 respectively before and after the flood events. This pattern was attributed to a progressive incorporation of flood deposits to the soil.  相似文献   

10.
Development of an L6 myoblast in vitro model of moniliformin toxicosis   总被引:1,自引:0,他引:1  
L6 myoblasts were used as an in vitro model to investigate the role of moniliformin and its interaction with monensin in turkey knockdown syndrome and sudden death syndromes in poultry. Cell viability and microscopic and ultrastructural alterations noted in L6 myoblasts cultured in the presence of moniliformin (0.0–0.3 g/l) were compared to those observed in parallel cultures also containing one of the following compounds: selenium (0–0.004 ng/l), thiamine (0–0.3 g/l), or pyruvate (0–0.46 g/l). Marked dilation of the RER, membranous whorls, glycogen deposition, membrane-bound cytoplasmic inclusions and necrosis were observed in myoblasts exposed to 0.03/2-0.30 g moniliformin/l medium. Supplementation of medium with thiamine and pyruvate, or selenium, provided significant protection to cells exposed to 0.0–0.3 g/l or 0.0–0.15 g moniliformin/l, respectively. Dose-dependent differences in protein and ATP production were not detected. Myoblasts grown in medium containing 0–0.15 g moniliformin/l and 7.5–50.0 M A23187, beauvericin or monensin had degrees of cytotoxicity similar to parallel cultures receiving only an ionophore. L6 myoblasts were a useful model of moniliformin toxicosis. The findings of this study suggest cytotoxicity due to moniliformin in L6 myoblasts may be due in part to oxidative damage and altered pyruvate metabolism, and that moniliformin does not predispose myoblasts to ionophore toxicosis. This study supports the results of in vivo investigations in poultry that moniliformin and monensin do not act synergistically to induce knockdown or monensin toxicosis.  相似文献   

11.
Summary Observations of aperture changes as sucrose is added to the solution bathing epidermal strips ofCommelina communis L. allow calculation of the osmotic changes required to open or close the stomatal pore, for comparison with changes in potassium content. With isolated guard cells, in strips in which all cells other than guard cells have been killed, the internal osmotic changes required are 83 mosmol kg–1 m–1 below 10m aperture, 129 mosmol kg–1 m–1 in the range 10–15 m, and 180 mosmol kg–1 m–1 above 15 m. For opening against subsidiary cell turgor in addition to guard cell turgor, in intact strips with live subsidiary and epidermal cells, these figures should each be increased by about 33 mosmol kg–1 m–1. A change in subsidiary cell turgor is magnified in its effects on the water relations of the guard cell by a factor greater than 3.7 for equal changes in the water potential of the two cells, or greater than 4.7 at constant volume of the guard cell.  相似文献   

12.
Synopsis Arsenic persists in Chautauqua Lake, New York waters 13 years after cessation of herbicide (sodium arsenite) application and continues to cycle within the lake. Arsenic concentrations in lake water ranged from 22.4–114.81 g l–1, = 49.0 ag l–1. Well water samples generally contained less than 10 g l–1 arsenic. Arsenic concentrations in lake water exceeded U.S. Public Health Service recommended maximum concentrations (10 g l–1) and many samples exceeded the maximum permissible limit (50 g l–1). Fish accumulated arsenic from water but did not magnify it. Fish to water arsenic ratios ranged from 0.4–41.6. Black crappie (Pomoxis nigromaculatus) contained the highest arsenic concentrations (0.14–2.04 g g–1 ), X = 0.7 g g–1) while perch (Perca flavescens), muskellunge (Esox masquinongy) and largemouth bass (Micropterus salmoides) contained the lowest concentrations (0.02–0.13 g g–1). Arsenic concentrations in fish do not appear to pose a health hazard for human consumers.  相似文献   

13.
A study of the effects of two selected organophosphorus insecticides methylpyrimifos and chlorpyrifos on soil microflora in an agricultural loam was made. The insecticides had concentrations of 10 to 300 g g–1. The presence of methylpyrimifos at concentrations of 100 to 300 g g–1 or chlorpyrifos at concentrations from 10 to 300 g g–1 significantly decreased aerobic dinitrogen fixing bacteria and dinitrogen fixation. Nitrifying bacteria decreased at concentrations of 200 and 300 g g–1 of methylpyrimifos. The presence of 10 to 300 g g–1 of chlorpyrifos decreased the total number of bacteria. However, fungal populations and denitrifying bacteria were not affected as a consequence of the addition of the organophosphorus insecticides to the agricultural soil, showing that these microorganisms can tolerate high amounts of those insecticides.  相似文献   

14.
Summary In separated outer medullary collecting duct (MCD) cells, the time course of binding of the fluorescent stilbene anion exchange inhibitor, DBDS (4,4-dibenzamido-2,2-stilbene disulfonate), to the MCD cell analog of band 3, the red blood cell (rbc) anion exchange protein, can be measured by the stopped-flow method and the reaction time constant, DBDS, can be used to report on the conformational state of the band 3 analog. In order to validate the method we have now shown that the ID50,DBDS,MCD (0.5±0.1 m) for the H2-DIDS (4,4-diisothiocyano-2,2-dihydrostilbene disulfonate) inhibition of DBDS is in agreement with the ID50,Cl ,MCD (0.94±0.07 m) for H2-DIDS inhibition of MCD cell Cl flux, thus relating DBDS directly to anion exchange. The specific cardiac glycoside cation transport inhibitor, ouabain, not only modulates DBDS binding kinetics, but also increases the time constant for Cl exchange by a factor of two, from Cl=0.30±0.02 sec to 0.56±0.06 sec (30mm NaHCO3). The ID50,DBDS,MCD for the ouabain effect on DBDS binding kinetics is 0.003±0.001 m, so that binding is about an order of magnitude tighter than that for inhibition of rbc K+ flux (K I,K +,rbc=0.017 m). These experiments indicate that the Na+,K-ATPase, required to maintain cation gradients across the MCD cell membrane, is close enough to the band 3 analog that conformational information can be exchanged. Cytochalasin E (CE), which binds to the spectrin/actin complex in rbc and other cells, modulates DBDS binding kinetics with a physiological ID50,DBDS,MCD (0.076±0.005 m); 2 m CE also more than doubles the Cl exchange time constant from 0.20±0.04 sec to 0.50±0.08 sec (30mm NaHCO3). These experiments indicate that conformational information can also be exchanged between the MCD cell band 3 analog and the MCD cell cytoskeleton.  相似文献   

15.
Isolated embryos ofKarwinskia humboldtiana were cultured in vitro. The growth of embryos and development to plantlets on woody plant medium supplemented with indole-3-acetic acid 6.10-2 mol l–1, gibberellic acid (GA3) 3.10-2 mol l–1, and 6-benzylaminopurine (BA) 2 mol l–1 was obtained. Multiplication of shoots and rooting of excised shoots has been achieved. Callus formation on modified Murashige-Skoog medium supplemented with 1-naphthaleneacetic acid 10 mol l–1, GA3 14 mol l–1, and kinetin 5 mol l–1 on hypocotyls, or on root cultures on medium supplemented with 2.4-dichlorophenoxyacetic acid 10 mol l–1 and BA 10 mol l–1 was induced.Abbreviations BA 6-benzylaminopurine - 2,4-d 2,4-dichlorophenoxyacetic acid - GA3 gibberellic acid - IAA indole-3-acetic acid - NAA 1-naphthaleneacetic acid - TEM transmission electron microscopy  相似文献   

16.
Conversion of methanol to CH4 has a large isotope effect so that a small contribution of methanol-dependent CH4 production may decrease the 13CH4 of total CH4 production. Therefore, we investigated the role of methanol for CH4 production. Methanol was not detectable above 10 M in anoxic methanogenic rice field soil. Nevertheless, addition of 13C-labeled methanol (99% enriched) resulted in immediate accumulation of 13CH4. Addition of 0.1 M 13C-methanol resulted in increase of the 13CH4 from –47 to –6 within 2 h, followed by a slow decrease. Addition of 1 M 13C-methanol increased 13CH4 to +500 within 4 h, whereas 10 M increased 13CH4 to +2500 and continued to increase. These results indicate that the methanol concentrations in situ, which diluted the 13C-methanol added, were 0.1 M and that the turnover of methanol contributed only about 2% to total CH4 production at 0.1 M. However, contribution increased up to 5 and 17% when 1 and 10 M methanol were added, respectively. Anoxic rice soil that was incubated at different temperatures between 10 and 37 °C exhibited maximally 2–6% methanol-dependent methanogenesis about 1–2 h after addition of 1 M 13C-methanol. Only at 50 °C, contribution of methanol to CH4 production reached a maximum of 10%. After longer (7–10 h) incubation, however, contribution generally was only 2–4%. Methanol accumulated in the soil when CH4 production was inhibited by chloroform. However, the accumulated methanol accounted for only up to 0.7 and 1.2% of total CH4 production at 37 and 50 °C, respectively. Collectively, our results show that methanol-dependent methanogenesis was operating in anoxic rice field soil but contributed only marginally to total CH4 production and the isotope effect observed at both low and high temperature.  相似文献   

17.
Three mannose-binding lectins were assayed in artificial diets for their toxic and growth-inhibitory effects on nymphal development of the peach-potato aphid Myzus persicae. The snowdrop (Galanthus nivalis) lectin GNA was the most toxic, with an induced nymphal mortality of 42% at 1500 g ml–1 (30 M) and an IC50 (50% growth inhibition) of 630 g ml–1 (13 M). The daffodil (Narcissus pseudonarcissus) lectin NPA and a garlic (Allium sativum) lectin ASA induced no significant mortality in the range 10–1500 g ml–1, but did result in growth inhibition of 59% (NPA) and 26% (ASA) at 1500 g ml–1 (40 M for NPA, 63 M for ASA). All three lectins were responsible for a slight but significant growth stimulation when ingested at 10 g ml–1, reaching +26%, +18% and +11% over the control values for the garlic lectin, the daffodil lectin and the snowdrop lectin, respectively. GNA, as well as the glucose/mannose binding lectin Concanavalin A, were also provided at sublethal doses throughout the life cycle of the aphids, and effects on adult performance were monitored. Adult survival was not significantly altered, but both lectins adversely affected total fecundity and the dynamics of reproduction, resulting in significant reduction in calculated r ms (population intrinsic rate of natural increase) on lectin-containing diets. These effects are discussed in relation to the use of transgenic plants expressing these toxic lectins for potential control of aphid populations.  相似文献   

18.
Summary The rate of fluid transport across rabbit corneal endothelium has been measured with an automatic volumetric method. The present resolution of the procedure is 1–3 nanoliters, and intervals of measurement can be made as small as seconds. In the presence of glucose, oxidized glutathione (GSSG), and adenosine, the maximal rates were 6.2±1.0 l/hr cm2, and 8.2±0.8 l/hr cm2 if a large portion of the stroma was dissected away. In the presence of glucose and GSSG only, the rates were lower, namely 3.7±0.5 l/hr cm2. The rates consistently increased or decreased when adenosine was added or deleted, respectively, during given experiments. The stimulation of fluid transport by adenosine was in the order of 40–50%. The results raise the possibility that this transport mechanism might be subject to metabolic control.  相似文献   

19.
Protistan community grazing rates upon both bacterioplankton and autotrophic picoplankton were estimated using fluorescently-labeled prey and by measurement of extracellular hydrolysis of 4-methylumbelliferyl (MUF) -N-acetylglucosaminide in a eutrophic reservoir and an oligo-mesotrophic lake during phytoplankton blooms. In addition, enzyme methods were optimized in bacterivorous flagellate cultures by two enzyme assays, based on fluorometric detection of protistan digestive activity, which were compared and calibrated independently against flagellate bacterivory. Enzymatic hydrolyses of MUF -N,N,N-triacetylchitotriose and MUF -N-acetylglucosaminide were measured in cell-free (sonicated) and whole-cell (unsonicated) samples. The hydrolysis of both substrates, using the whole-cell enzyme assay at in situ pH, was correlated significantly with total grazing rate of Bodo saltans. Thus the whole-cell enzyme assay with MUF -N-acetylglucosaminide was used for freshwater samples. High-affinity (K m < 1 mol 1–1) and low-affinity (K m > 100 mol 1–1) enzymes were distinguished kinetically in most samples from both systems studied. Activities (V max ) of the high-affinity enzyme varied from 0.24 to 1.43 nmol 1–1 h–1. Protistan community grazing on bacterioplankton was in the range of 0.15–1.36 g C 1–1 h–1. both for lake and reservoir, the differences being observed in grazing on picocyanobacteria (lake, 0.03-0.22 g C 1–1 h–1. reservoir, 0.35–1.56 g C 1–1) h–1. The enzyme activities were correlated significantly with the protistan grazing both on bacterioplakton (r s = 0.62, P < 0.001) and total procaryotic picoplankton (the sum of organic carbon grazed from bacteria and picocyanobacteria, r s = 0.73, P < 0.001) in the eutrophic reservoir. Weaker relationships (r s = 0.42) with a lower slope were found for the oligo-mesotrophic lake. Ingestion rate studies are time-consuming and the digestive enzyme assay with MUF -N-acetylglucosaminide presents a rapid alternative for estimating total protistan prokaryotic picoplanktivory in freshwaters.  相似文献   

20.
Jia  Yinsuo  Gray  V.M. 《Photosynthetica》2003,41(4):605-610
We determined for Vicia faba L the influence of nitrogen uptake and accumulation on the values of photon saturated net photosynthetic rate (P Nmax), quantum yield efficiency (), intercellular CO2 concentration (C i), and carboxylation efficiency (C e). As leaf nitrogen content (NL) increased, the converged onto a maximum asymptotic value of 0.0664±0.0049 mol(CO2) mol(quantum)–1. Also, as NL increased the C i value fell to an asymptotic minimum of 115.80±1.59 mol mol–1, and C e converged onto a maximum asymptotic value of 1.645±0.054 mol(CO2) m–2 s–1 Pa–1 and declined to zero at a NL-intercept equal to 0.596±0.096 g(N) m–2. fell to zero for an NL-intercept of 0.660±0.052 g(N) m–2. As NL increased, the value of P Nmax converged onto a maximum asymptotic value of 33.400±2.563 mol(CO2) m–2 s–1. P N fell to zero for an NL-intercept of 0.710±0.035 g(N) m–2. Under variable daily meteorological conditions the values for NL, specific leaf area (L), root mass fraction (Rf), P Nmax, and remained constant for a given N supply. A monotonic decline in the steady-state value of Rf occurred with increasing N supply. L increased with increasing N supply or with increasing NL.  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