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 343 毫秒
1.
Adsorption of hydrophobic contaminants at the particle/water interface is one of the key processes controlling their fate in the aquatic environment. The sorption of the natural female hormones oestrone and 17-oestradiol has been studied under simulated riverine conditions. Both the kinetics and the effects of varying fundamental environmental parameters (e.g. sediment properties) on the thermodynamic equilibrium partition coefficient (K p) have been studied in continuous and batch sorption experiments, respectively. Results showed that the sorption of oestrone and 17-oestradiol by sediment was relatively slow, reaching equilibrium in 50 days. In addition, relatively small adsorption of both oestrone and 17-oestradiol onto the sediment was observed, with K p values between 200 and 250 mL g–1. The comparable K p values of the two compounds reflect their structural similarity. It can be concluded that the two endocrine disruptors, oestrone and 17-oestradiol remain primarily in association with the aqueous phase.  相似文献   

2.
The photosynthesis–irradiance response of Ecklonia radiata (C. Agardh) J. Agardh, a common kelp in the temperate southern hemisphere, was investigated in situ throughout the year and across a depth profile at West Island, South Australia. Temperature and irradiance environment altered throughout the year, varying at 3 m between 14–20°C and 279–705 mol photons m–2 s–1. Photosynthetic capacity (Pm) varied throughout the year between 177–278 mol O2 g–1 dry wt h–1 at 3 m and 133–348 mol O2 g–1 dry wt h–1 at 10 m. The irradiance required for sub-saturation of photosynthesis (Ek) varied between 97–152 and 81–142 mol photons m–2 s–1 for 3 m and 10 m respectively, and the respiration rate varied between 15–36 and 13–20 mol O2 g–1 dry wt h–1 for 3 m and 10 m. A clear seasonal change in photokinetic parameters was detected and provided strong evidence for a seasonal acclimation response. During winter an increase in the efficiency of light utilisation at low irradiance () was accompanied by a decrease in both Ek and that required for photosynthetic compensation. Pm also increased during the winter and autumn months and respiratory requirements decreased. These changes enable E. radiata to display an optimal photosynthetic performance throughout the year despite significant changes in the surrounding environment.Communicated by P.W. Sammarco, Chauvin  相似文献   

3.
There is increasing evidence that suspension feeders play a significant role in plankton–benthos coupling. However, to date, active suspension feeders have been the main focus of research, while passive suspension feeders have received less attention. To increase our understanding of energy fluxes in temperate marine ecosystems, we have examined the temporal variability in zooplankton prey capture of the ubiquitous Mediterranean gorgonian Leptogorgia sarmentosa. Prey capture was assessed on the basis of gut content from colonies collected every 2 weeks over a year. The digestion time of zooplankton prey was examined over the temperature range of the species at the study site. The main prey items captured were small (80–200 µm), low-motile zooplankton (i.e. eggs and invertebrate larvae). The digestion time of zooplankton prey increased when temperature decreased (about 150% from 21°C to 13°C; 15 h at 13°C, 9 h at 17°C, and 6 h at 21°C), a pattern which has not previously been documented in anthozoans. Zooplankton capture rate (prey polyp–1 h–1) varied among seasons, with the greatest rates observed in spring (0.16±0.02 prey polyp–1 h–1). Ingestion rate in terms of biomass (g C polyp–1 h–1) showed a similar trend, but the differences among the seasons were attenuated by seasonal differences in prey size. Therefore, ingestion rate did not significantly vary over the annual cycle and averaged 0.019±0.002 g C polyp–1 h–1. At the estimated ingestion rates, the population of L. sarmentosa removed between 2.3 and 16.8 mg C m–2 day–1 from the adjacent water column. This observation indicates that predation by macroinvertebrates on seston should be considered in energy transfer processes in littoral areas, since even species with a low abundance may have a detectable impact.Communicated by S.A. Poulet, Roscoff  相似文献   

4.
The life-history of the crown-of thorns starfish (Acanthaster planci) includes a planktotrophic larva that is capable of feeding on particulate food. It has been proposed, however, that particulate food (e.g. microalgae) is scarce in tropical water columns relative to the nutritional requirements of the larvae of A. planci, and that periodic shortages of food play an important role in the biology of this species. It has also been proposed that non-particulate sources of nutrition (e.g. dissolved organic matter, DOM) may fuel part of the nutritional requirements of the larval development of A. planci as well. The present study addresses the ability of A. planci larvae to take up several DOM species and compares rates of DOM uptake to the energy requirements of the larvae. Substrates transported in this study have been previously reported to be transported by larval asteroids from temperate and antarctic waters. Transport rates (per larval A. planci) increased steadily during larval development and some substrates had among the highest mass-specific transport rates ever reported for invertebrate larvae. Maximum transport rates (J max in) for alanine increased from 15.5 pmol larva–1 h–1 (13.2 pmol g–1 h–1) for gastrulas (J max in=38.7 pmol larva–1 h–1 or 47.4 pmol g–1 h–1) to 35.0 pmol larva–1 h–1 (13.1 pmol g–1 h–1) for early brachiolaria (J max in just prior to settlement=350.0 pmol larva–1 h–1 or 161.1 pmol g–1 h–1) at 1 M substrate concentrations. The instantaneous metabolic demand for substrates by gastrula, bipinnaria and brachiolaria stage larvae could be completely satisfied by alanine concentrations of 11, 1.6 and 0.8 M, respectively. Similar rates were measured in this study for the essential amino acid leucine, with rates increasing from 11.0 pmol larva–1 h–1 (or 9.4 pmol g–1 h–1) for gastrulas (J max in=110.5 pmol larva–1 h–1 or 94.4 pmol g–1 h–1) to 34.0 pmol larva–1 h–1 (or 13.0 pmol g–1 h–1) for late brachiolaria (J max in=288.9 pmol larva–1 h–1 or 110.3 pmol g–1 h–1) at 1 M substrate concentrations. The essential amino acid histidine was transported at lower rates (1.6 pmol g–1 h–1 at 1 M for late brachiolaria). Calculation of the energy contribution of the transported species revealed that larvae of A. planci can potentially satisfy 0.6, 18.7, 29.9 and 3.3% of their total energy requirements (instantaneous energy demand plus energy added to larvae as biomass) during embryonic and larval development from external concentrations of 1 M of glucose, alanine, leucine and histidine, respectively. These data demonstrate that a relatively minor component of the DOM pool in seawater (dissolved free amino acids, DFAA) can potentially provide significant amounts of energy for the growth and development of A. planci during larval development.  相似文献   

5.
Concentrations of thallium in phytoplankton (0.02 to 0.8 g g–1), zooplankton (0.03 to 0.5 g g–1) and ichthyoplankton (0.1 g g–1) from the central Pacific were comparable , as were the atomic ratios of thallium to calcium (3x10–6) and to potassium (1x10–6) in those organisms. These relatively constant ratios, plus the biounlimited ocean profile of thallium, indicate that it is rapidly cycled through plankton in the same manner as potassium, its principal biogeochemical analogue. The higher atomic ratios of thallium to potassium in pelagic clays (6x10–6) and ferromanganese nodules (4x10–3) suggest that both biological transport processes and abiotic transport processes influence this trace element's oceanic cycle.  相似文献   

6.
Beryllium and aluminium contents in uncontaminated soils from six countries are reported. The means and ranges of beryllium in the surface soils were as follows: 1.43(0.20–5.50)g g–1 in Thailand (n=28), 0.7 (0.31–1.03) g g–1 in Indonesia (n=12), 0.99(0.82–1.32) g g–1 in New Zealand (n=3), 0.58(0.08-1.68)g g–1 in Brazil (n=16), 3.52(2.49–4.97)g g–1 in the former Yugoslavia (n=10), and 1.56(1.01–2.73) g g–1 in the former USSR (n=8). The mean and range of beryllium contents of the surface soils in Japan (1.17(0.27–1.95)g g–1 n=27) are situated within the values of the soils from these countries except for the Yugoslav soils derived from limestones. The mean of the mean beryllium contents of the surface soils in all these countries is 1.42 g g–1 which will be used as a tentative average content of beryllium in uncontaminated surface soils, except for the soils derived from parent materials high in beryllium content. The beryllium contents of the subsoils were higher than those of the surface soils in New Zealand and Yugoslavia as is the case with Japan. The correlation coefficient between the contents of beryllium and aluminium in all the soil samples (n=113) including surface soils and subsoils was 0.505 (p < 0.001).  相似文献   

7.
The vertical distribution, diel gut pigment content and oxygen consumption of Calanus euxinus were studied in April and September 1995 in the Black Sea. Gut pigment content of C. euxinus females was associated with diel vertical migration of the individuals, and it varied with depth and time. Highest gut pigment content was observed during the nighttime, when females were in the chlorophyll a (chl a) rich surface waters, but significant feeding also occurred in the deep layer. Gut pigment content throughout the water column varied from 0.8 to 22.0 ng pigment female–1 in April and from 0.2 to 21 ng pigment female–1 in September 1995. From the diel vertical migration pattern, it was estimated that female C. euxinus spend 7.5 h day–1 in April and 10.5 h day–1 in September in the chl a rich surface waters. Daily consumption by female C. euxinus in chl a rich surface waters was estimated by taking into account the feeding duration and gut pigment concentrations. Daily carbon rations of female C. euxinus, derived from herbivorous feeding in the euphotic zone, ranged from 6% to 11% of their body carbon weight in April and from 15% to 35% in September. Oxygen consumption rates of female and copepodite stage V (CV) C. euxinus were measured at different temperatures and at different oxygen concentrations. Oxygen consumption rates at oxygen-saturated concentration ranged from an average of 0.67 g O2 mg–1 dry weight (DW) h–1 at 5°C to 2.1 g O2 mg–1 DW h–1 at 23°C for females, and ranged from 0.48 g O2 mg–1 DW h–1 at 5°C to 1.5 g O2 mg–1 DW h–1 at 23°C for CVs. The rate of oxygen consumption at 16°C varied from 0.62 g O2 mg–1 DW h–1 at 0.65 mg O2 l–1 to 1.57 g O2 mg–1 DW h–1 at 4.35 mg O2 l–1 for CVs, and from 0.74 g O2 mg–1 DW h–1 at 0.57 mg O2 l–1 to 2.24 g O2 mg–1 DW h–1 at 4.37 mg O2 l–1 for females. From the oxygen consumption rates, daily requirements for the routine metabolism of females were estimated, and our results indicate that the herbivorous daily ration was sufficient to meet the routine metabolic requirements of female C. euxinus in April and September in the Black Sea.Communicated by O. Kinne, Oldendorf/Luhe  相似文献   

8.
Gas-liquid interface measurements were conducted in a strongly turbulent free-surface flow (i.e., stepped cascade). Local void fractions, bubble count rates, bubble size distributions and gas-liquid interface areas were measured simultaneously in the air-water flow region using resistivity probes. The results highlight the air-water mass transfer potential of a stepped cascade with measured specific interface area over 650 m–1 and depth-average specific area up to 310 m–1. A comparison between single-tip and double-tip resistivity probes suggests that simple robust single-tip probes may provide accurate, although conservative, gas-liquid interfacial properties. The latter device may be used in the field and in prototype plants. Notation a = specific interface area (m–1); a mean = depth-average specific interface area (m–1): a mean=frac1Y 90limits sup> Y 90 sup 0(1–C)dy; C = local void fraction; C gas = dissolved gas concentration (kg m–3); C mean = depth-average mean air concentration defined as: C mean=1–d/Y 90; C s = saturation concentration (kg m–3); D = dimensionless air bubble diffusivity (defined by [1]); d = equivalent clear-water flow depth (m): d=limits sup> Y 90 sup 0(1–C) dy; dab = air bubble diameter (m); dc = critical flow depth (m); for a rectangular channel: d c=sqrt[3]q w 2/g; F = air bubble count rate (Hz); F max = maximum bubble count rate (Hz), often observed for C=50%; g = gravity acceleration (m s–2); h = step height (m); K L = liquid film coefficient (m s–1); K = integration constant defined as: K=tanh –1 sqrt0.1)+(2D)–1 [1]; L = chute length (m); N = velocity distribution exponent; ———– *Corresponding author, E-mail: h.chanson@mailbox.uq.edu.au Q w = water discharge (m3 s–1); q w = water discharge per unit width m2 s–1); t = time (s); V = local velocity (m s–1); V c = critical flow velocity (m s–1); for a rectangular channel: V c=sqrt[3]q w g V max = maximum air-water velocity (m s–1); V 90 = characteristic air-water velocity (m s–1) where C = 90%; W = channel width (m); x = longitudinal distance (m) measured along the flow direction (i.e., parallel to the pseudo-bottom formed by the step edges); y = distance (m) normal to the pseudo-bottom formed by the step edges; Y90 = characteristic distance (m) where C=0.90; Y 98 = characteristic distance (m) where C=0.98; = slope of pseudo-bottom by the step edges; = diameter (m).  相似文献   

9.
We studied Na+/K+ ATPase activity and ultrastructure in gills of the hyper-hypo-regulating crab Chasmagnathus granulatus Dana, 1851 acclimated to different salinities: 10, 30 and 45, known to be hypo-, iso-, and hyper-osmotic to the hemolymph, respectively. After centrifugation of homogenates at 11,000 g, Na+/K+–ATPase activity was almost entirely found in the pellets from the posterior (6–8) and anterior (3–5) gills, whereas very little was detected in the supernatant liquid. Specific activity of gill 6 was 41.3, 30.2, and 28.2 µmol Pi h–1 mg prot–1 for crabs acclimated to 10, 30, and 45, respectively, the result for 10 being significantly higher than those at 30 and 45. Although the concentration of sodium at which the reaction rate is half-maximal (K M) was similar in the three acclimation salinities, only the enzyme from crabs acclimated to 10 was inhibited by high sodium concentration. Specific activity of gill 5 increased with the increment in external salinity (10.1, 15, and 18.1 µmol Pi h–1 mg prot–1 for 10, 30, and 45, respectively), the only significant difference being that between the extreme salinities. The epithelium thickness of the dorsal portion of gill 6 showed a variation among salinities: 21.7, 15.8 and 17.2 µm for 10, 30 and 45, respectively. There were significant differences in epithelium thickness between the 10 and the other salinities. In all three salinities, the ultrastructure of gill 6 epithelium showed a high density of mitochondria, estimated by their volume fraction (Vv m=0.307–0.355). These mitochondria were packed between extensive basolateral membrane interdigitations in ionocytes and pillar cells. Gill 5 showed three cell types: pillars which possess mitochondria packed between membrane folds only in their interdigitations with neighbouring cells; type-I cells 8.0 µm thick with low density of mitochondria (Vv m=0.088), and type-II cells, 9.9 µm thick and rich in mitochondria (Vv m=0.423), but lacking basolateral interdigitations. Vv m of type-I cells of gill 5 was significantly lower than those of type-II cells of the same gill and the ionocytes of gill 6. No significant difference in Vv m was detected between the latter cell types.Communicated by P.W. Sammarco, Chauvin  相似文献   

10.
G. R. Parsons 《Marine Biology》1990,104(3):363-367
Routine metabolic rates of bonnethead sharks,Sphyrna tiburo, of 95 to 4 650 g, ranged from 70.4 to 15.0 kcal kg–1 d–1. Over the size range 34 to 95 cm total length, shark swimming-velocities varied from about 29 to 67 cm s–1. Swimming velocities predicted using Weih's cost-optimization model were similar to observed velocities. The total cost of transport (the energetic cost of transporting 1 unit of body mass 1 km distance) for 1 to 8 kg sharks varied from 0.67 to 0.40 cal g–1 km–1. The energetic range (an estimation of the distance traveled after a 25% reduction in body weight) indicates that a 1 kg bonnethead shark would travel 500 km distance in 17 d before displaying a 25% reduction in weight. An 8 kg individual would travel 830 km in 23 d. Although the bonnethead shark is a continuously active species, its routine metabolic rate and the efficiency of its locomotory system may be similar to that of typical bony fishes.  相似文献   

11.
The pattern of growth (biomass accumulation) in Ecklonia radiata throughout the year and across a depth profile was investigated using the traditional hole-punch method, and the information presented in context with concurrently measured in situ net productivity rates. The rate of net daily productivity showed a lack of consistent seasonal variability, remaining constant throughout the year at two of the four depths measured (3 m and 12 m), and becoming higher during winter at another (5 m). Throughout the year, rates of net daily productivity differed significantly across the depth profile. Net daily productivity rates averaged 0.017 g C g–1 dwt day–1 and 0.005 g C g–1 dwt day–1 at a depth of 3 m (1,394 mol O2 g–1 dwt day–1) and 10 m (382 mol O2 g–1 dwt day–1) respectively. In contrast, the biomass accumulation rate of E. radiata was highly seasonal, with low rates of growth occurring in autumn (0.002 g dwt g–1 dwt day–1 at both 3 and 10 m) and summer (0.007 and 0.004 g dwt g–1 dwt day–1 at 3 and 10 m respectively) and higher rates in spring (0.016 and 0.007 g dwt g–1 dwt day–1 at 3 and 10 m respectively) and winter (0.015 and 0.008 g dwt g–1 dwt day–1 at 3 and 10 m respectively). The proportion of assimilated carbon used for biomass accumulation varied throughout the year, between 5% and 41% at 3 m and between 28% and 128% at 10 m. The rates of biomass accumulation at all depths represented only a small proportion of the amount of carbon assimilated annually.Communicated by P.W. Sammarco, Chauvin  相似文献   

12.
Eelgrass (Zostera marina L.) has access to nutrient pools in both the water column and sediments. We investigated the potential for eelgrass to utilize nitrate nitrogen by measuring nitrate reductase (NR) activity with an in vivo tissue assay. Optimal incubation media contained 60 mM nitrate, 100 mM phosphate, and 0.5% 1-propanol at pH 7.0. Leaves had significantly higher NR activity than roots (350 vs 50 nmoles NO 2 produced g FW–1 h–1). The effects of growing depth (0.8 m MLW, 1.2 m, 3.0 m, 5.0 m) and location within the eelgrass meadow (patch edge vs middle) on NR activity were examined using plants collected from three locations in the Woods Hole area, Massachusetts, USA, in July 1987. Neither depth nor position within the meadow appear to affect NR activity. Nitrate enrichment experiments (200 M NO 3 for 6 d) were conducted in the laboratory to determine if NR activity could be induced. Certain plants from shallow depth (1.2 m) showed a significant response to enrichment, with NR activity increasing from >100 up to 950 nmoles NO 2 g FW–1 h–1 over 6 d. It appears that Z. marina growing in very shallow water (0.8 m) near a shoreline may be affected by ground water or surface run-off enrichments, since plants from this area exhibited rates up to 1 600 nmol NO 2 g FW–1 h–1. Water samples from this location consistently had slightly higher NO 3 concentrations (1.4 M) than all other collection sites (0.7 M). Thus, it is possible that chronic run-off or localized groundwater inputs can create sufficient NO 3 enrichment in the water column to induce nitrate reductase activity in Zostera leaves.  相似文献   

13.
The effect of cholinergic antagonists on the bradycardia induced by waterborne copper in the Mediterranean limpet Patella caerulea was investigated by using non-invasive recording of cardiac activity of whole animals. Preliminary tests were conducted to check the role of cholinergic and serotoninergic systems in the control of heart rate of P. caerulea. Superfusing the whole limpets with carbachol (cholinergic agonist) at 5×10–5 M produced a negative inotropic and chronotropic effect (bradycardia), while superfusion with 5-hydroxytryptamine produced a positive inotropic and chronotropic effect (tachycardia). Exposure of limpets to a solution of copper in artificial seawater (0.25 mg l–1, 3 h) reduced their heart rate to about 80% the value recorded in copper-free water. This bradycardia was inhibited by injecting the limpets with atropine (cholinergic muscarinic antagonist) at 21 g g–1 wet flesh weight and with benzoquinonium [cholinergic nicotinic antagonist blocking the K+ mediated acetylcholine (ACh) response] at 10 and 100 g g–1 prior to copper exposure. In contrast, D-tubocurarine (cholinergic nicotinic antagonist blocking Na+ mediated ACh response) had no effect at 85 g g–1. These results agree with the involvement of the cholinergic system in the bradycardic response of limpets to copper, and support the view that gastropod ACh receptors do not fit the vertebrate nicotinic–muscarinic classification.Communicated by R. Cattaneo-Vietti, Genova  相似文献   

14.
Silicon and carbon uptake rates were studied over a 24 h light/dark cycle in a synchronised culture of the marine diatom Cylindrotheca fusiformis (Reimann et Lewin) using 32Si and 14C. The silicic acid uptake rate per cell (cSi) varied between 1.2 and 20.0 fmol Si cell–1 h–1 and was closely correlated to the G2+M phase of the cell cycle. A linear and significant relationship was determined between the percentage of cells present in G2+M and cSi. Evolution of the soluble free-silicon pool was studied simultaneously. The concentration of the total soluble free pool of silicon (QPSi) varied from 1% to 7% of the total silicon content. A significant difference of 1.5 fmol Si cell–1 between QPSi and the labelled free pool (QnpSi) was measured, indicating the presence of an unlabelled fraction of the pool. The concentration of QnpSi was around 1.0 fmol Si cell–1 prior to cell division and did not change as a function of cSi, which indicated a feedback mechanism coupling uptake into the free pool and incorporation into the frustule. In parallel, 14C uptake variation (cC) was measured during the division of the population. The value of cC varied between 0.44 and 0.78 pmol C cell–1 h–1 and appeared to be maximal when cells were in the G1 phase. This variation of cC marginally affected the total carbon content of the cells (QTC) in comparison with the light/dark cycle. The variations in the Si/C ratio, from 0.021 to 0.046, demonstrated the different control mechanisms of Si and C metabolisms during the course of the cell- and photocycle.Communicated by S.A. Poulet, Roscoff  相似文献   

15.
Photosynthetic performance in the kelp Laminaria solidungula J. Agardh was examined from photosynthesis irradiance (P-I) parameters calculated from in situ 14C uptake experiments, using whole plants in the Stefansson Sound Boulder Patch, Alaskan Beaufort Sea, in August 1986. Rates of carbon fixation were determined from meristematic, basal blade, and second blade tissue in young and adult sporophytes. Differences in saturating irradiance (I k, measured as photosynthetically active radiation, PAR), photosynthetic capacity (P max), and relative quantum efficiency () were observed both between young and adult plants and between different tissue types. I k was lowest in meristematic tissue (20 to 30 E m–2 s–1) for both young and adult plants, but consistently 8 to 10 E m–2 s–1 higher in young plants compared to adults in all three tissues. Average I k for non-meristematic tissue in adult plants was 38 E m–2 s–1. Under saturating irradiances, young and adult plants exhibited similar rates of carbon fixation on an area basis, but under light limitation, fixation rates were highest in adult plants for all tissues. P max was generally highest in the basal blade and lowest in meristematic tissue. Photosynthetic efficiency () ranged between 0.016 and 0.027 mol C cm–2 h–1/E m–2 s–1, and was highest in meristematic tissue. The relatively lower I k and higher exhibited by L. solidungula in comparison to other kelp species are distinct adaptations to the near absence of light during the eight-month ice-covered period and in summer when water turbidity is high. Continuous measurement of in situ quantum irradiance made in summer showed that maximum PAR can be less than 12 E m–2 s–1 for several days when high wind velocities increase water turbulence and decrease water transparency.The Univeristy of Texas Marine Science Institute Contribution No. 695  相似文献   

16.
Ammonium excretion of a dense population (~1 500 individuals m–2) of the ophiuridOphiothrix fragilis (Abildgaard) was measured in the Dover Straits (French coast) between May 1989 and March 1990: the excretion rate varied from 4.8 µg N g–1 dry wt h–1 in November to 12.8 µg N g–1 dry wt h–1 in June. Mean individual ammonium excretion,E, wasE=0.019t +1.26 (whereE=µg N individual–1 andt=time in min;r=0.80;N=81). Variations in the ammonium excretion rate during a tidal cycle appeared to arise from variations in the duration of the suspension-feeding activity ofO. fragilis, which was governed by the strength of the tidal current. During short-term starvation, excretion was low (E=0.009t+1.47;r=0.91;N=17), increasing with increasing length of starvation [E=4.62lnt–2.5;r=0.95;N=17], as observed for other echinoderms; this could be due to catabolism of tissue. The daily ammonia flux from thisO. fragilis population to the water column was estimated at 41 mg N m–2 d–1.  相似文献   

17.
The activity of Na–K-ATPase was measured in crude homogenates prepared from various organs (leg muscle, pincer muscle, heart, testes, digestive gland, hypodermis, gills 1–9) of shore crabs, Carcinus maenas L., acclimated to salinities ranging between 10 and 50 S (in steps of 10 S). In all salinities tested, Na–K-ATPase activity was highest in posterior gills 7–9 (10–12 mol Pi mg protein-1 h-1), followed by anterior gills 1–6 (ca. 2.5 mol Pi mg protein-1 h-1) and the other organs (in most cases far below 2mol Pi mg protein-1 h-1). In gills only, Na–K-ATPase activity was salinity-dependent, with the highest values in the lowest salinities and vice versa. In gills 7–9, Na–K-ATPase activity was increased more than threefold following a reduction in salinity from 50 to 10 S. Na–K-ATPase activity, expressed as percentage of total ATPase activity, amounted to 60–80% in gills, about 60% in hypodermis and 20–40% in the other organs. Ouabain, a specific inhibitor of Na–K-ATPase activity, reduced serum osmolalities in crabs kept at 9–10 S only when injected into the hemolymph (1 and 5 · 10-5 M), but had no effect when dissolved in ambient water (10-4 M). The results obtained underline that crustacean gills are the main organs for ionic regulation, and confirm the hypothesis of the central role of the Na–K-ATPase in active Na uptake as the basic mechanism of hyperregulation in dilute media. Reduction of serum osmolalities following injection of ouabain into the hemolymph confirms previous reports on localization of the sodium pump in the basolateral parts of epithelial cells.  相似文献   

18.
Soil studies, conducted in Maryland, Minnesota and Louisiana, have described the urban pattern of lead contamination. They have shown that the highest amounts of lead cluster within the interior of the largest cities. The results of the New Orleans urban patterns of distribution of soil lead provided the basis for further study. The hypothesis was tested that elementary school properties have the same pattern of soil lead contamination as their neighbouring residential communities. Thirty New Orleans Public Elementary Schools were selected for this study. Surface samples (2.5cm or 1 inch depth) were collected from playgrounds and next to entrances of each school. Results showed that soil lead on school properties follows the same relative contamination patterns (pvalue10–5) as soil lead on residential properties of neighbouring communities. Schools however, have significantly lower lead contamination than the neighbouring residential properties. Innercity school properties present a higher risk of soil lead exposure than mid and outercity schools. Soils next to innercity school entrances showed the highest lead, with 18.5% having concentrations over 400gg–1. Systematic landscaping around the school entrances would significantly reduce the hazard from lead dust contaminated soils.  相似文献   

19.
During two expeditions of the R.V. Polarstern to the Arctic Ocean, pack ice and under-ice water samples were collected during two different seasons: late summer (September 2002) and late winter (March/April 2003). Physical and biological properties of the ice were investigated to explain seasonal differences in species composition, abundance and distribution patterns of sympagic meiofauna (in this case: heterotrophs >20 µm). In winter, the ice near the surface was characterized by extreme physical conditions (minimum ice temperature: –22°C, maximum brine salinity: 223, brine volume: 5%) and more moderate conditions in summer (minimum ice temperature: –5.6°C, maximum brine salinity: 94, most brine volumes: 5%). Conditions in the lowermost part of the ice did not differ to a high degree between summer and winter. Chlorophyll a concentrations (chl a) showed significant differences between summer and winter: during winter, concentrations were mostly <1.0 µg chl a l–1, while chl a concentrations of up to 67.4 µmol l–1 were measured during summer. The median of depth-integrated chl a concentration in summer was significantly higher than in winter. Integrated abundances of sympagic meiofauna were within the same range for both seasons and varied between 0.6 and 34.1×103 organisms m–2 in summer and between 3.7 and 24.8×103 organisms m–2 in winter. With regard to species composition, a comparison between the two seasons showed distinct differences: while copepods (42.7%) and rotifers (33.4%) were the most abundant sea-ice meiofaunal taxa during summer, copepod nauplii dominated the community, comprising 92.9% of the fauna, in winter. Low species abundances were found in the under-ice water, indicating that overwintering of the other sympagic organisms did not take place there, either. Therefore, their survival strategy over the polar winter remains unclear.Communicated by O. Kinne, Oldendorf/Luhe  相似文献   

20.
R. Beiras  J. Widdows 《Marine Biology》1995,122(4):597-603
The acute and long-term effects of neurotransmitters dopamine (DA), serotonin (SE) and norepinephrine (NE) on the feeding rates of Mytilus edulis veliger larvae were investigated through concentration-response curves. Increasing DA concentrations increasingly inhibited food uptake. Acute exposure to high levels of DA caused long-term inhibitory effects on feeding rates (10–5 MDA) and growth rates (3x10–4 MDA). Feeding activity was also inversely related to NE concentration. SE concentrations between 10–8–3x10–7 M supported enhanced feeding rates. Neither NE nor SE showed long-term inhibitory effects on feeding at concentrations <10–4 M. These results were consistent with the observed effects of the different neurotransmitters on the swimming pattern of the larvae. The experimental evidence supports the model of ciliary control in adult mussels, involving dual innervation of the ciliated cells of the velum, with excitatory serotonergic and inhibitory dopaminergic fibers.  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