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.
Autoradiographic studies of [3H]aldosterone [( 3H-A] and [3H]dexamethasone binding sites in intact target cells (isolated collecting tubules of rabbit and rat kidney) revealed an almost exclusive nuclear localization of the hormone-receptor complexes. In the present work we compared the nucleo-cytoplasmic repartition of [3H]A-receptor complexes studied in parallel by biochemical and autoradiographic methods. In addition, the thermo-dependency of the nuclear translocation was examined. Kidney pyramids were incubated in vitro with [3H]A (2 X 10(-9) M) in the presence or absence of a 100-fold excess unlabelled A, at 30 degrees C for 1 h or 4 degrees C for 2 h. Then tissue was processed for isolation of nuclear and cytoplasmic fractions, on the one hand, or for obtention of microdissected tubular segments on which autoradiographs on dry films were performed. Autoradiographs showed that the specific labelling was almost exclusively nuclear without significant cytoplasmic labelling, at both 30 or 4 degrees C. This indicates that almost all binding sites migrated rapidly into nuclei, and that this translocation did not depend on temperature. In contrast, parallel biochemical experiments yielded classical results, that is, at 30 degrees C, the presence of specific binding sites in both cytoplasm and nuclei with a predominance in cytoplasm. At 4 degrees C, the cytoplasmic binding was unchanged, but nuclear binding was drastically reduced, indicating thermodependency of nuclear translocation, when studied by biochemical methods including cell disruption. Autoradiographic results thus questioned the classical notion of thermo-dependent nuclear translocation of aldosterone-receptor complexes, based on results obtained by biochemical methods.  相似文献   

2.
Lifetable demography and reproductive traits of a Kenyan strain of the rotifer Brachionus angularis were investigated using individual and small batch culture approaches. The rotifer was identified morphologically before conducting studies at 20, 25 and 30 °C, using Chlorella vulgaris at 2.5 × 105 to 2.5 × 107 cells ml–1. The rotifers were highly fecund, producing 2.11 ± 0.07 offspring female–1 day–1 and reproductive, producing 8.43 ± 0.24 offspring female–1 at 25 °C with 2.5 × 106 algal cells ml–1. The highest intrinsic rate of natural increase (0.74 ± 0.02 d–1), specific population growth rate (0.49 ± 0.01), longest life expectancy at hatching (12.41 ± 0.28 d) and shortest generation time (2.87 ± 0.03 d) also occurred at 25 °C with 2.5 × 106 algal cells ml–1. The duration of hatching to first spawning was shortest (2.86 ± 0.21 h) at 30 °C with 2.5 × 107 algal cells ml–1 and longest (8.83 ± 0.39 h) at 20 °C with 2.5 × 105 algal cells ml–1. The highest population density (255.7 ± 12.6 ind. ml–1) was realised at 25 °C with 2.5 × 106 cells ml–1 on Day 8, whereas the lowest population density (122.0 ± 3.6 ind. ml–1) was realised at 20 °C with 2.5 × 105 cells ml–1 on Day 8. The lorica length and width of the Kenyan strain of B. angularis are 85.6 ± 3.1 µm and 75.4 ± 3.6 µm, respectively. The rotifer optimally reproduces at 25 °C when fed with 2.5 × 106 algal cells ml–1.  相似文献   

3.
Eucalypts are major emitters of biogenic volatile organic compounds (BVOCs), especially volatile isoprenoids. Emissions and incorporation of 13C in BVOCs were measured in Eucalyptus camaldulensis branches exposed to rapid heat stress or progressive temperature increases, in order to detect both metabolic processes and their dynamics. Isoprene emission increased and photosynthesis decreased with temperatures rising from 30°C to 45°C, and an increasing percentage of unlabelled carbon was incorporated into isoprene in heat‐stressed leaves. Intramolecular labelling was also incomplete in isoprene emitted by heat‐stressed leaves, suggesting increasing contribution of respiratory (and possibly also photorespiratory) carbon. At temperature above 45°C, a drop of isoprene emission was mirrored by the appearance of unlabelled monoterpenes, green leaf volatiles, methanol, and ethanol, indicating that the emission of stored volatiles was mainly induced by cellular damage. Emission of partially labelled acetaldehyde was also observed at very high temperatures, suggesting a double source of carbon, with a large unlabelled component likely transported from roots and associated to the surge of transpiration at very high temperatures. Eucalypt plantations cover large areas worldwide, and our findings may dramatically change forecast and modelling of future BVOC emissions at planetary level, especially considering climate warming and frequent heat waves.  相似文献   

4.
As part of a programme of comparative measurements of P d (diffusional water permeability) the RBCs (red blood cells) from an aquatic monotreme, platypus (Ornithorhynchus anatinus), and an aquatic reptile, saltwater crocodile (Crocodylus porosus) were studied. The mean diameter of platypus RBCs was estimated by light microscopy and found to be ~6.3 μm. P d was measured by using an Mn2+‐doping 1H NMR (nuclear magnetic resonance) technique. The P d (cm/s) values were relatively low: ~2.1×10?3 at 25°C, 2.5×10?3 at 30°C, 3.4×10?3 at 37°C and 4.5 at 42°C for the platypus RBCs and ~2.8×10?3 at 25°C, 3.2×10?3 at 30°C, 4.5×10?3 at 37°C and 5.7×10?3 at 42°C for the crocodile RBCs. In parallel with the low water permeability, the E a,d (activation energy of water diffusion) was relatively high, ~35 kJ/mol. These results suggest that “conventional” WCPs (water channel proteins), or AQPs (aquaporins), are probably absent from the plasma membranes of RBCs from both the platypus and the saltwater crocodile.  相似文献   

5.
A new method of labelling human RBCs with 99mTc by the use of Sn-α-d-glucose 1-phosphate (GP) is presented. It was tested for spleen imaging in 16 normal volunteers. The labelling was carried out during the heating (30 min at 49.5 °C) to damage the cells and cooling (10 min at R.T.) steps. The labelling efficiency was 95.5 ± 1.5% with Sn-GP and was better than Sn-PYP (61.5 ± 25.0%). The radioactivity retention of RBCs was > 96% after 6 washings. The spleen was delineated very well in all the subjects. The spleen activity reached a plateau at 20 min post-injection. Spleen-to-liver and spleen-to-cardiac blood pool concentration ratios were high (10.4 ± 2.2 and 10.1 ± 3.2, respectively) calculated at 30 min. The method is simple, rapid and efficient.  相似文献   

6.
Corneal cryopreservation requires that endothelial cells remain viable and intercellular structure be preserved. High viability levels for cryopreserved endothelial cells have been achieved, but preserving intercellular structure, especially endothelial attachment to Descemet's membrane, has proved difficult. Cell detachment apparently is not caused by ice, suggesting osmotic or chemical mechanisms. Knowledge of the permeation kinetics of cryoprotectants (CPAs) into endothelial cells and stroma is essential for controlling osmotic and chemical activity and achieving adequate tissue permeation prior to cooling. Proton nuclear magnetic resonance (NMR) spectroscopy was used to assess the permeation of dimethyl sulfoxide (DMSO) into isolated rabbit corneas. Corneas with intact epithelia were exposed to isotonic medium or 2.0 mol/L DMSO for 60 min and subsequently transferred to 2.0 or 4.0 mol/L DMSO, respectively, at 22, 0, or −10°C. DMSO concentration in the cornea was measured vs time. The Kedem-Katchalsky model was fitted to the data. Hydraulic permeability (m3/N·s) is 7.1×10−13+216%-11% at 22°C, 8.2×10−13+235%−21% at 0°C, and 1.7×10−14+19% −16% at −10°C. The reflection coefficient is 1.0+2%−1% at 22°C and 0°C, and 0.9±5% at −10°C. Solute mobility (cm/s) is 5.9×10−6+6%–11% at 22°C, 3.1×10−6+12%−11% at 0°C, and 5.0×10−8 cm/s+59%−40% at −10°C.  相似文献   

7.
《Insect Biochemistry》1989,19(8):809-814
The interaction of locust high density lipophorin (HDLp) with pieces of fat body tissue was studied at 33°C using a radiolabelled ligand binding assay. Under the assay conditions, binding of tritium-labelled HDLp ([3H]HDLp) was demonstrated to correlate linearly with tissue concentration up to ∼ 7 mg of fat body protein per ml of incubation medium. The [3H]HDLp binding that was displaceable by a 20-fold excess of unlabelled HDLp (which is an approximation of the specific binding) reached equilibrium after ∼ 2 h, whereas low levels of non-displaceable binding increased linearly during this time interval. Analysis of the concentration dependent total binding of [3H]HDLp revealed the presence of a specific binding site with an equilibrium dissociation constant of Kd = 3.1 (±0.5) × 10−7 M and a maximal binding capacity of 9.8 (±0.5) ng μg−1 tissue protein. Competition experiments demonstrated that the affinity of unlabelled HDLp for the binding site is similar to the affinity of [3H]HDLp. Unlabelled low density lipophorin (LDLp), however, was shown to have an approx. 20-fold lower affinity for the binding site.  相似文献   

8.
Papaverine inhibited the basal renin secretion of rat kidney slices incubated in a physiological salt solution at 37°C. Inhibition was concentration-dependent; secretion was 99 ± 0.2 % inhibited by 5 × 10?4 M papaverine, and 8 × 10?5 M was the estimated ED50. In contrast, 2 × 10?4 M IBMx (3-isobutyl-1-methyl-xanthine) increased the rate of secretion from 215 ± 17 to 366 ± 30 ng hr?1mg?1/20 min (p < 0.001). Isoprotenol (4 × 10?7, 8 × 10?7, and 5 × 10?6 M) stimulated renin secretion in a concentration-dependent manner; the stimulatory effects were antagonized by papaverine but unaffected by IBMx. Thus, two known inhibitors of phosphodiesterase--IBMx and papaverine--produce sharply contrasting effects on basal and on isoproterenol-stimulated renin secretion from rat kidney slices.  相似文献   

9.
The relationship between elongation growth and the incorporation of [3H]gibberellin A1 ([3H]GA1) into a 2,000g pelletable (2KP) fraction from lettuce (Lactuca sativa L., cv. Arctic) hypocotyl sections has been examined. Sections were loaded with incremental amounts of GA1 under conditions where growth was arrested (5° C) or permitted (30° C) and, after 16 h, all were transferred to a GA-free medium at 30° C. Growth and 2KP radioactivity were measured at this point and after a further 24 h in the chase medium. Uptake was reduced by 80% at 5° C, as compared to 30° C, but 2KP labelling and protein synthesis were only reduced by half. The growth rate of the 5° C pretreated sections during the chase period was comparable to that observed during the pulse in the 30° C material but the dose/response relationship was flatter. Low temperature sections incorporated a much higher percentage of GA1 uptake into the 2KP fraction (27% at maximum) but the absolute levels of labelling at this temperature were lower than those measured at 30° C. The data are interpreted as showing that 2KP labelling is not a consequence of growth. It must either precede response or be an unconnected concurrent process.  相似文献   

10.
In suspensions of epididymal spermatozoa in vitro at +10°C and +37°C, all nuclei-containing and mitochondria-containing structures (normal spermatozoa, spermatozoa with the bent and coiled tails, complexes of head and neck) are with propidium iodide and rhodamine 123, respectively. Intracellular ATP concentration determined by a bioluminescent method in mitochondria-containing elements of suspension decreases (significantly faster at 37°C than at 10°C) up to a certain unchangeable level (2.5 × 10?8 M/l at 37°C and to 1.6 × 10?8 M/l at 10°C per 106 of mitochondria-containing elements). Mechanisms of spermatozoa destruction are discussed.  相似文献   

11.
The biological control of Aonidiella aurantii (Maskell) on citrus can be achieved with periodic releases of the parasitoid Aphytis melinus DeBach. A. melinus is normally reared on parthenogenetic strains of the scale Aspidiotus nerii Bouché. The developmental rate, crawler production and survival of A. nerii were studied at two temperatures (25±1 and 30±1°C) and two levels of relative humidity (RH; 55±5 and 85±5%) on squash. Temperature had an important effect on developmental rate, but RH did not. At 30°C, no growth was observed and no crawlers were produced. At 25°C, development of the L1, L2 and young females’ stages required 24.4±0.7, 11.1±0.8 and 13.2±0.3 days, respectively. At 25°C, times until the appearance of mature females, the start of crawler production and the peak of crawler production were 48.7±0.1, 50.1±0.7 and 63.3±1.0 days, respectively. Crawler production lasted 42.6±1.9 days, and the 50% production level was reached on day 13. Initially, crawler survival was higher at 30°C (65.9±2.8%) than at 25°C (51.7±4.8%), but by the end of development survival was much higher at 25 than 30°C (88.0±2.1 vs. 22.9±2.7%). RH had no effect on either initial or final survival. Average progeny production was of 28.0±2.8 crawlers per female at 25°C. The relationship between the weight of groups of crawlers and their number was: Number=1031+835,500×weight of group (g) (R 2=0.826), which can assist in colony management. The importance of selecting the correct strain of A. nerii is emphasised.  相似文献   

12.
Nuclear import of plasmid DNA mediated by a nuclear localization signal (NLS) derived from SV40 T antigen was investigated in a cell-free extract. In vitro assembled sea urchin male pronuclei were incubated in a 100,000g supernatant of a zebrafish fertilized egg lysate, together with fluorescently labeled plasmid DNA bound to NLS or nuclear import deficient reverse NLS (revNLS) peptides. After 3 hr, DNA-NLS, but not DNA-revNLS, complexes were bound around the nuclear periphery. We demonstrate that nuclear import of DNA-NLS complexes is a two-step process involving binding to, and translocation across, the nuclear envelope. Binding is ATP-independent, occurs at 0°C and is Ca2+-independent. By contrast, translocation requires ATP hydrolysis, Ca2+, is temperature dependent and is blocked by the lectin wheat germ agglutinin. Both binding and translocation are competitively inhibited by albumin-NLS conjugates, require heat-labile cytosolic factors, and are inhibited by N-ethylmaleimide treatment of the cytosol. Binding and translocation are differentially affected by cytosol dilutions, suggesting that at least two distinct soluble fractions are required for nuclear import. The requirements for NLS-mediated nuclear import of plasmid DNA are similar to those for nuclear import of protein-NLS conjugates in permeabilized cells. © 1996 Wiley-Liss, Inc.  相似文献   

13.
We report for the first time the presence of a sex steroid-binding protein in the plasma of green sea turtles Chelonia mydas, which provides an insight into reproductive status. A high affinity, low capacity sex hormone steroid-binding protein was identified in nesting C. mydas and its thermal profile was established. In nesting C. mydas testosterone and oestradiol bind at 4°C with high affinity (K a = 1.49 ± 0.09 × 109 M−1; 0.17 ± 0.02 × 107 M−1) and low binding capacity (B max = 3.24 ± 0.84 × 10−5 M; 0.33 ± 0.06 × 10−4 M). The binding affinity and capacity of testosterone at 23 and 36°C, respectively were similar to those determined at 4°C. However, oestradiol showed no binding activity at 36°C. With competition studies we showed that oestradiol and oestrone do not compete for binding sites. Furthermore, in nesting C. mydas plasma no high-affinity binding was observed for adrenocortical steroids (cortisol and corticosterone) and progesterone. Our results indicate that in nesting C. mydas plasma temperature has a minimal effect on the high-affinity binding of testosterone to sex steroid-binding protein, however, the high affinity binding of oestradiol to sex steroid-binding protein is abolished at a hypothetically high (36°C) sea/ambient/body temperature. This suggests that at high core body temperatures most of the oestradiol becomes biologically available to the tissues rather than remaining bound to a high-affinity carrier.  相似文献   

14.
Infrared laser traps (optical tweezers) were used to study laser-induced organelle movements in the marine alga Pyrocystis noctiluca (Dinophyta). These cells are highly suitable for optical micromanipulation due to their large size and extensive vacuole. Experiments were done with plastids held by optical tweezers and moved from the nuclear area into the vacuole. The subsequent retraction movement was analysed for speed. The displaced organelles remained connected to their original position by a thin cytoplasmic strand, often less than 1 μm in diameter. When the organelles were released they rapidly returned at an initial rate of 81.7 ± 7.8 μm . s?1 (overall displacement 50 μm, measured distance 20 μm, 25 °C ± 1 °C, number of cells 22), slowing down with progressive retraction of the connecting strand. The return movement was reduced to 4.2 ± 0.2 μ .s?1 (n = 10) when the organelles were displaced and held for 1 min. Displacement to a longer distance increased the rate of return movement. A change from a high to a low environmental temperature significantly reduced movement from 94.5 ± 9.0 . s?1 (30 °C ± 1 °C, n = 22) to 34.5 ± 2.7 μm .s?1 (5°C ± 1 °C, n = 22). Nocodazole and N-ethylmaleimide (NEM), inhibitors of microtubules and acto-myosin, respectively, did not affect the retraction of the connecting strand, but at high concentrations of NEM it became increasingly difficult to move organelles away from the nuclear area. We suggest that the return movement of organelles within laser-induced artificial strands mainly depends on the viscoelastic properties of the tonoplast. The quantification of these properties by optical tweezers allows determination of reactions of plant cells to temperature changes.  相似文献   

15.
Kinetic comparisons of mesophilic and thermophilic aerobic biomass   总被引:1,自引:0,他引:1  
Kinetic parameters describing growth and decay of mesophilic (30°C) and thermophilic (55°C) aerobic biomass were determined in continuous and batch experiments by using oxygen uptake rate measurements. Biomass was cultivated on a single soluble substrate (acetate) in a mineral medium. The intrinsic maximum growth rate (μ max) at 55°C was 0.71±0.09 h−1, which is 1.5 times higher than the μ max at 30°C (0.48±0.11 h−1). The biomass decay rates increased from 0.004 h−1 at 30°C to 0.017 h−1 at 55°C. Monod constants were very low for both types of biomass: 9±2 mg chemical oxygen demand (COD) l−1at 30°C and 3±2 mg COD l−1at 55°C. Theoretical biomass yields were similar at 30 and 55°C: 0.5 g biomass COD (g acetate COD)−1. The observed biomass yields decreased under both temperature conditions as a function of the cell residence time. Under thermophilic conditions, this effect was more pronounced due to the higher decay rates, resulting in lower biomass production at 55°C compared to 30°C. Electronic Publication  相似文献   

16.
Quantification of the diffusion of small molecules and large lipid transporting lipoproteins across arterial tissues could be useful in elucidating the mechanism(s) of atherosclerosis. Optical coherence tomography (OCT) was used to determine the effect of temperature on the rate of diffusion of glucose and low‐density lipoproteins (LDL) in human carotid endarterectomy tissue in vitro. The permeability rate for glucose was calculated to be (3.51 ± 0.27) × 10–5 cm/s (n = 13) at 20 °C, and (3.70 ± 0.44) × 10–5 cm/s (n = 5) at 37 °C; for LDL the rate was (2.42 ± 0.33) × 10–5 cm/s (n = 5) at 20 °C and (4.77 ± 0.48) × 10–5 cm/s (n = 7) at 37 °C, where n is the number of samples. These results demonstrate that temperature does not significantly influence the permeation of small molecules (e.g. glucose), however, raising the temperature does significantly increase the permeation of LDL. These results provide new information about the capacity of an atherogenic lipoprotein to traverse the intimal layer of the artery. These results also demonstrate the potential of OCT for elucidating the dynamics of lipoprotein perfusion across the arterial wall. (© 2009 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim)  相似文献   

17.
Alkaline hydrolysis and subcritical water degradation were investigated as ex-situ remediation processes to treat explosive-contaminated soils from military training sites in South Korea. The addition of NaOH solution to the contaminated soils resulted in rapid degradation of the explosives. The degradation of explosives via alkaline hydrolysis was greatly enhanced at pH ≥12. Estimated pseudo-first-order rate constants for the alkaline hydrolysis of 2,4-dinitrotoluene (DNT), 2,4,6-trinitrotoluene (TNT) and hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) in contaminated soil at pH 13 were (9.6?±?0.1)×10?2, (2.2?±?0.1)×10?1, and (1.7?±?0.2)×10?2 min?1, respectively. In the case of subcritical water degradation, the three explosives were completely removed at 200–300°C due to oxidation at high temperatures and pressures. The degradation rate increased as temperature increased. The pseudo-first-order rate constants for DNT, TNT, and RDX at 300°C were (9.4?±?0.8)×10?2, (22.8?±?0.3)×10?2, and (16.4?±?1.0)×10?2, respectively. When the soil-to-water ratio was more than 1:5, the extent of alkaline hydrolysis and subcritical water degradation was significantly inhibited.  相似文献   

18.
The speckled peacock bass Cichla temensis is a popular sport and food fish that generates substantial angling tourism and utilitarian harvest within its range. Its popularity and value make this species important for management and a potential aquaculture candidate for both fisheries enhancement and food fish production. However, little is known of optimal physiochemical conditions in natural habitats, which also are important for the development of hatchery protocols for handling, spawning and grow-out. Speckled peacock bass have been documented to have high sensitivity to extreme temperatures, but the metabolic underpinnings have not been evaluated. In this study, the effects of temperature (25, 30 and 35°C) on the standard metabolic rate (SMR) and lower dissolved oxygen tolerance (LDOT) of juvenile speckled peacock bass (mean ± standard error total length 153 ± 2 mm and wet weight 39.09 ± 1.37 g) were evaluated using intermittent respirometers after an acclimation period of 2 weeks. Speckled peacock bass had the highest SMR at 35°C (345.56 ± 19.89 mgO2 kg−1 h−1), followed by 30°C (208.16 ± 12.45 mgO2 kg−1 h−1) and 25°C (144.09 ± 10.43 mgO2 kg−1 h−1). Correspondingly, the Q10, or rate of increase in aerobic metabolic rate (MO2) relative to 10°C, for 30–35°C was also greater (2.76) than from 25 to 30°C (2.08). Similarly, speckled peacock bass were the most sensitive to hypoxia at the warmest temperature, with an LDOT at pO2 of 90 mmHg (4.13 mg l−1) at 35°C compared to pO2 values of 45 mmHg (2.22 mg l−1) and 30 mmHg (1.61 mg l−1) at 30 and 25°C, respectively. These results indicate that speckled peacock bass are sensitive to temperatures near 35°C, therefore we recommend managing and rearing this species at 25–30°C.  相似文献   

19.
  • 1.1. The diffusional water permeability (Pd) of rabbit red blood cell (RBC) membrane has been monitored by a doping nuclear magnetic resonance (NMR) technique on control cells and following inhibition with p-chloromercuribenzene sulfonate (PCMBS).
  • 2.2. The values of Pd were around 6.3 × 10−3 cm/sec at 15°C, 7.0 × 10−3cm/sec at 20°C, 8.0 × 10−3 cm/sec at 25°C, 9.1 × 10−3 cm/sec at 30°C and10.7 × 10−3 cm/sec at 37°C.
  • 3.3. Systematic studies on the effects of PCMBS on water diffusion indicated that the maximal inhibition was reached in 15 min at 37°C with 0.5 mM PCMBS.
  • 4.4. The values of maximal inhibition were around 71–74% at all temperatures.
  • 5.5. The basal permeability to water was estimated as 1.6 × 10−3cm/sec at 15°C, 2.0 × 10−3cm/sec at 20°C, 2.4 × 10−3cm/sec at 25°C, 2.6 × 10−3cm/sec at 30°C, and 3.1× 10−3 cm/secat 37°C.
  • 6.6. The activation energy of water diffusion was around 18 kJ/mol and increased to 27 kcal/mol after incubation with PCMBS in conditions of maximal inhibition of water diffusion.
  • 7.7. The membrane polypeptide electrophoretic pattern of rabbit RBCs has been compared with its human counterpart.
  • 8.8. The rabbit membrane contained a higher amount of spectrin (bands 1 and 2), while the band 6 (glyceraldehyde-3-phosphate dehydrogenase) was markedly less intense.
  • 9.9. Considerable differences in the electrophoretic patterns of the two sources of RBC membranes appeared in the bands migrating in the band 4.5 region and in front of band 7, where some polypeptides were apparent in higher amounts in the rabbit RBC membrane.
  相似文献   

20.
The aim of this study was to determine acute toxicity in the post larvae of the white shrimp Litopenaeus vannamei after 96 h of exposure to dissolved arsenic under three different temperatures and salinity conditions. Recent reports have shown an increase in the presence of this metalloid in coastal waters, estuaries, and lagoons along the Mexican coast. The white shrimp stands out for its adaptability to temperature and salinity changes and for being the main product for many commercial fisheries; it has the highest volume of oceanic capture and production in Mexican shrimp farms. Lethal concentrations (LC50–96 h) were obtained at nine different combinations (3?×?3 combinations in total) of temperature (20, 25, and 30 °C) and salinity (17, 25, and 33) showing mean LC50–96 h values (±standard error) of 9.13?±?0.76, 9.17?±?0.56, and 6.23?±?0.57 mgAs?L?1(at 20 °C and 17, 25, and 33 salinity); 12.29?±?2.09, 8.70?±?0.82, and 8.03?±?0.59 mgAs?L?1 (at 25 °C and 17, 25, and 33 salinity); and 7.84?±?1.30, 8.49?±?1.40, and 7.54?±?0.51 mgAs?L?1 (at 30 °C and 17, 25, and 33 salinity), respectively. No significant differences were observed for the optimal temperature and isosmotic point of maintenance (25 °C–S 25) for the species, with respect to the other experimental conditions tested, except for at 20 °C–S 33, which was the most toxic. Toxicity under 20 °C–S 33 conditions was also higher than 25 °C–S 17 and 20 °C (S 17 or 25). The least toxic condition was 25 °C–S 17. All this suggests that the toxic effect of arsenic is not affected by temperature changes; it depends on the osmoregulatory pattern developed by the shrimp, either hyperosmotic at low salinity or hiposmotic at high salinity, as observed at least on the extreme salinity conditions here tested (17 and 33). However, further studies testing salinities near the isosmotic point (between 20 and 30 salinities) are needed to clarify these mechanisms.  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