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.
The fine structure and motility of spermatozoa and the composition of the seminal plasma of the perch Perca fluviatilis are investigated by electron microscopy, computer assisted cell motility analysis (CMA) and biochemical methods. The spermatozoon is asymmetrical as the flagellum inserts mediolateral on the nucleus. It lacks an acrosome, has an ovoid head and a small midpiece with one mitochondrion. Sperm motility–initiated in distilled water (10° C)–is characterized as follows: 85·0 ± 2·7% of the spermatozoa are motile, the main swimming type (10 ± 1 s after motility initiation) is the linear motion (61·4 ± 24·4%) and the average swimming velocity is 122·4 ± 21·9 μm s–1. When motility is initiated with NaCl, glucose or sucrose solutions of 100 mosmol kg–1 the percentage of motile spermatozoa and the swimming types are similar as in water, but the swimming velocity (174·0 ± 22·3 μm s–1) is significantly higher. Motility is inhibited by high osmolality of the diluent: when increasing the osmolality of the saline solutions to 350 mosmol kg–1 sperm motility is totally suppressed while potassium (10–40 mmol 1–1) does not affect motility parameters. pH optimum for sperm motility is between pH 7·0 and 8·5. The seminal fluid contains 124·01 ± 21·68 mmol 1–1 sodium, 10·22 ± 1·11 mmol 1–1 potassium and 0·72 ± 0·26 mmol 1–1 calcium. pH is 8·25 ± 0·09, and osmolality 283·90 ± 37·19 mosmol kg–1. The following organic components were determined: monosaccharides (glucose 63 ± 19 μmol 1–1, fructose 54 ± 28 μmol 1–1, galactose 59 ± 25 μmol 1–1), lipids (cholesterol 5·51 ± 6·42 μmol 1–1, triglycerides 72 ± l00 μmol l–1, cholesteryloleate 15–150 μmol 1–1, phosphatidylcholine 26 · 31 μmol 1–1, glycolipids 1–10 mg 100 m1–1), lactate 108 ± 99 μmol 1–1, hydroxybutyrate 102 ± 99 nmol 1–1, choline 59 ± 159 μmol 1–1, protein 344·75 ± 59·06 mg 100m1–1, enzymes (β-d -glucuronidase l.4 ± 0.7 μmol h–1 100 ml–1, protease (caseolytic activity) 1·0 ± 0·6 μmol h–1 100 ml–1, alkaline phosphatase 2520·0 ± 861·0 μmol h–1 100 ml–1, acid phosphatase 44.0 ± 16.0 μmol h–1 100 ml–1, glucose-6-phosphate dehydrogenase 38·9 ± 86·9 μmol h–1 100 ml–1, lactate dehydrogenase 134·4 ± 69·6 μmol h–1 100 ml–1, butyrylcholine esterase 0·014 ± 0·010 μmol h–1 100 ml–1, adenosine triphosphatase 562·8 ± 665·4 μmol h –1 100 ml–1).  相似文献   

2.
The ecophysiology of the hypotonic response was studied in the charophyte alga, Lamprothamnium papulosum, which was grown in a marine (SW; 1072 mosmol kg–1) and a brackish (1/2 SW; 536 mosmol kg–1) environment. The cells produced an extracellular mucilage identified by histochemical staining as a mixture of sulphated and carboxylated polysaccharides. The thickness and chemical composition of the mucilage layer was a function of environmental salinity and cell age. Mucilage progressively increased in thickness from the apex (9 SW cells: 12·6 ± 1·8 μm; 15 1/2 SW cells: 4·8 ± 0·7 μm) to the base of the plants (15 SW cells: 44·8 ± 3·3 μm; nine 1/2 SW cells: 23·8 ± 2·5 μm); with a corresponding increase in the sulphated proportion. The mucilage was significantly thicker in SW plants. Hydraulic conductivity (Lp) at the apex of SW plants, measured by transcellular osmosis, was 8·3 × 10–13 m s–1 Pa–1. This was close to Lp of freshwater Chara (8·5 × 10–13 m s–1 Pa–1) which lacked mucilage. Basal SW cells with thicker mucilage had a smaller apparent Lp of 3·5 × 10–13 m s–1 Pa–1. The electrophysiology of the resting state and hypotonic response was compared in cells from the two environments based on current/voltage (I/V) analysis. The resting potential difference (PD) and conductance differed (11 SW cells: – 102·4 ± 10·1 mV, eight SW cells: 18·6 ± 2·4 S m–2; 19 1/2 SW cells: –125·7 ± 5·9 mV, 8·3 ± 0·8 S m–2). The type of cellular response to a hypotonic shock (decrease of 268 mosmol kg–1) also differed. In 1/2 SW plants, only the apical cells with thin mucilage responded classically with depolarization, conductance increase, Ca2+ influx, cessation of cytoplasmic streaming, and K+ and Cl effluxes. Older cells making up the bulk of the plants responded with depolarization, but continued cytoplasmic streaming, and had only a small increase in conductance; or depolarized transiently without altering the I/V profile, conductance or streaming speed. Most cells remained depolarized and in the K+ state 1 h post-shock. Cells treated with the K+ channel blocker tetraethylammonium chloride also depolarized and remained depolarized. The SW cells depolarized but otherwise responded minimally to a 268 mosmol kg–1 drop in osmolarity and required a further 268 mosmol kg–1 down-step to elicit a change in the conductance. A spectrum of responses was measured in successively older and more mucilaginous cells from the same marine plant. We discuss the ecophysiological significance of the mucilage layer which modulates the cellular response to osmotic shock and which can be secreted to different degrees by plants inhabiting environments of different salinity.  相似文献   

3.
Two populations of Phyllariopsis purpurascens from the Straits of Gibraltar area were investigated during 1990 and 1991. This area has very strong currents (~2·5 m s-1 at spring tide). The first sporophytes (less than 0·5 cm in length) were observed in April; they disappeared after the first storm period in October. The estimated number of spores produced in an unilocular sporangium was 64. The fertility period started in May and continued until October, but the maximum sorus surface area (27%) was reached in September. The maximum sporangial density was 1476±236 mm-2. The slow development of fertility and its correlation with the seasonal increase in temperature suggest that sporangial development is controlled by temperature and not by photoperiod. The drag effect produced by the strong currents (at spring tides and storm periods) is sufficient to break adult plants (with 0·01 m2 in surface intercepting the current). At the beginning of the sporophytic production cycle (210 days), the meristematic zone produces more tissue for blade growth than for stipe growth. Holdfast growth occurs in the first 3 months. The mortality was 90% for sporophytes 0–10 cm in length, and decreased to ~50% in adult plants. One cohort in the cycle has been identified. The net primary production (calculated by the graphic model of Allen) was 0·04 g dry weight m-2·day-1 from April to October (210 days). The production is among the lowest values estimated for seaweeds.  相似文献   

4.
Limnological studies on some lakes in the Netherlands   总被引:6,自引:0,他引:6  
Most Dutch lakes are small and shallow, resulting from peat dredging since the late 18th century. However, deep lakes have appeared recently owing to sand digging. Limnological features of one such lake, Wijde Biik (N. Holland), were studied during 1968–70. The lake with an average depth of 11 4 m (maximum depth 31 m) and area of 2·65 × 106m2 is one of the deepest and biggest in the Loosdrecht-lakes area. The lake is 125 cm below sea level, and underground water-movements play an important part in the lake's hydrology. The lake exhibits thermal stratification on warm and calm days; since the lake surroundings are open and flat, wind and nocturnal cooling destroy such a stratification. There is continuous circulation from autumn through spring. The O2 saturation (%) in the upper water varies from 70 to 120%. Bottom waters were never anaerobic (lowest values 10% O2 saturation). CI? (2·8m-equiv.) and HCO?3 (1·9 m-equiv.) were the dominant anions just as Ca++ (2·77 m-equiv.) and Na+ (2·5 m-equiv.) formed the main cations. Chlorides have increased 2·5-fold in the 40 years as a geochemical consequence of deepening. The surface drainage has minor influence on lake's water chemistry. Part-P (10–140 μg/I) and PO4-P (2–40 μg/1) recorded maximum and minimum respectively, and NO3-N (0·05–1·15 mg/1) its minimum, during Microcystis abundance in August 1968. The SiO2-Si decreased from February 1969 (400 μg/l) to June 1969 as Diatotna elongatum increased. The Si-decrease to <30% of the 1932 values is due to removal of Si-rich clay and silt, due to sand digging. Chlorococcales were the important lake algae. Desmids were poor. Microcystis dominated as a rule from July-September, achieving from 15 to 31 colonies/ml. Poor light transmission rather than nutrients limits plankton growth as also the primary production in the lake. Copepoda were the dominant zooplankton. Bosmina coregoni recorded between 2 and 44 individuaIs/1 in summer 1968 and was the main cladoceran. The average primary production during summers of 1969 and 1970 was 380 and 497 mg C m?2 day?1 respectively. Light limited production below 1 m—1 % light in 1969 and 10% in 1970 penetrated down to 4 m. About 70% of the production took place in the upper 2 m. Calculation of production according to theoretical models under-estimated the observed values by 12% because Z0·5Ik lay much above (0·8–2·8 m) the expected value of 3·5 m. It is suggested that in turbid lakes like Wijde Blik in situ incubations should be done at 0·5 m intervals in the upper 2 or 3 m.  相似文献   

5.
Cladophora glomerata grown in continuous-flow culture was found to have optimal specific growth rate (μ) at, or near, 20°C. Specific growth rate increased linearly with increased duration of illumination per day up to 24h, and increased light intensity up to 6000 lx. Undissociated ammoniacal nitrogen (0·185 mg 1-1) reduced μ to 50% of that at 0·010 mg 1-1: 0·077–1·057 mg NO2-N1-1 and 7·2–15·2 mg NO3-N1-1 had no significant effect on μ. At 4·9 mg PO4-P1-1, μ was 48% of that at 1·9 mg1-1. The critical medium PO4-P concentration was less than 0·098 mg1-1. Specific growth rate was reduced to 50% of that in the based medium by 0·036 mg Cu1-1, 0·070 mg Zn1-1 and 1·03 mg Pb1-1. Results are discussed in the context of the natural distribution of the alga in the field situation.  相似文献   

6.
Zinc and salinity effects on membrane transport in Chara connivens   总被引:1,自引:1,他引:0  
Pressure-probe measurements showed that the pressure relaxation of internodal cells of the freshwater alga Chara connivens slowed considerably when 1–5 mol m?3 Zn2+, or more especially Zn2+ and 75 mol m?3 NaCl, were present in the medium for periods of 1 h or longer. These results indicate that the water permeability of the Chara membrane is decreased by Zn2+, and that this effect is enhanced by 75 mol m?3 NaCl. Specific values taken after 375 min exposure were: 5 mol m?3 Zn2+ and 75 mol m?3 NaCl caused the half-time for bulk water movement to increase from 7·8±2·3 to 79·5±5·4s, corresponding to a decrease in the hydraulic conductivity (Lp) from (13·0±3·3) × 10?7 m s?1 mPa?1 to (1·25±0·23) × 10?7 m s?1 MPa?1 (mean±S.D., n= 10). These changes are not seen in the presence of NaCl alone, and to a reduced extent in the presence of 5 mol m?3Zn2+ alone (after 375 min, Lp was (2·4±0·1) × 10?7 m s?1 MPa?1, mean±S.D., n = 6). Ca2+ cannot substitute for Zn2+, but seems to competitively inhibit Zn2+. There was another, kinetically distinct effect of Zn2+: the ingress of Na+ within 15 min of exposure to 75 mol m?3 NaCl is halved by the presence of 1–5 mol m?3 Zn2+, although internal osmolality is little changed by Zn2+. In spite of this, Zn2+ does not exert the long-term protection against NaCl that has been reported for Ca2+. Depending on the concentration of Zn2+ and the duration of the exposure, the effects on water permeability were fully or partly reversible within 24–48 h. The mechanism of these changes is difficult to identify. One possibility is a zinc-induced restriction of trans-membrane channels to give single-file channels which can be blocked by salt.  相似文献   

7.
It is generally accepted that hypertension and other vascular pathologies increase in diabetes mellitus (DM) patients as a result of the renin–angiotensin–aldosterone (RAA) system. In this study, changes in the renin‐angiotensin‐aldosterone (RAA) system level was determined in Streptozotocin (STZ)‐injected rats. A total of 46 female Wistar albino rats (180–220 g body weight) was utilized in these experiments. STZ was given intraperitoneally to induce diabetes in rats. Streptozotocin (60 mg kg−1 body weight) was dissolved in 0·1 m citrate–‐phosphate buffer (pH 4–5). The non‐diabetic rats were injected with sterilized buffer alone to act as a control group. Blood glucose levels were 398±8·2 mg dl−1, 488±11·75 mg dl−1 and 658±29·6 mg dl−1 at days 3, 12 and 30 respectively. The level of plasma renin activity (PRA) was measured as 7·69±1·07 ng ml−1 h−1; 1·82±0·22 ng ml−1 h−1 and 0·67±0·12 ng ml−1 h−1 at days 3, 12 and 30, respectively. These values showed that the PRA levels are decreased with increased time period. Serum angiotensin converting enzyme (ACE, E.C. 3.4.15.1) levels were increased at days 12 and 30 (p<0·05 and p<0·005), whereas serum aldosterone levels were increased at days 3 and 12 (p<0·05). The level of urea and creatinine increased at days 12 and 30 (p<0·05 and p<0·005, respectively) when compared to the control group. The data from these experiments indicate that the PRA level decreased whereas ACE activity level increased in diabetic rats compared with the control. Aldosterone levels increased at the first stage of the experiment, but then decreased by the end of the experiment as a result of changes in renin and ACE levels. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

8.
SODIUM-DEPENDENT EFFLUX AND EXCHANGE OF GABA IN SYNAPTOSOMES   总被引:12,自引:10,他引:2  
Abstract— The influx and efflux of [3H]GABA were investigated in synaptosomes. Two efflux components were detected. The first, termed spontaneous efflux, was not affected by the external sodium chloride concentration. The second, termed GABA-stimulated efflux, was observed when low levels of GABA were added to the incubation medium and was found to require external sodium chloride. The rate of spontaneous efflux at 0°C was about 37 per cent of the rate at 27°C but both GABA-stimulated efflux and GABA influx were completely inhibited at 0°C. The stimulation of efflux by external GABA followed simple Michaelis–Menten kinetics with respect to external GABA. The concentration of external GABA required for half-maximal stimulation was 4·9 ± 1·4 μm and the Vmax for efflux was 1·0 ± 0·6 nmol. min-1.mg-1 of protein. A similar stimulation of efflux was observed with GABA analogue l -2,4-diamino-butyric acid which is a competitive inhibitor of influx. The concentration of external l -2,4-diaminobutyric acid required for half-maximal stimulation of efflux was 51 ± 12 μm and the Vmax for efflux was 0·8 ± 0·5 nmol.min-1.mg-1 of protein. Since the sodium-dependency, temperature sensitivity, and kinetic properties of the GABA-stimulated efflux system were similar to the influx system, GABA-stimulated efflux was attributed to carrier-mediated exchange diffusion. Measurement of efflux and influx in the same preparation showed there was a net efflux when total fluxes were considered and that the exchange ratio (influx to GABA-stimulated efflux) was 0·9 when carrier-mediated fluxes were considered. The effect of the temperature of the fluid used to rinse synaptosomes collected on filters in influx experiments was investigated. There was no detectable difference in measured values of influx between samples rinsed with cold fluid (0°C) and warm fluid (27°C). The endogenous GABA content of synaptosomes was found to be 20·3 ± 2·5 nmol GABA per mg of protein. From this value, the cytoplasmic concentration of GABA in synaptosomes was estimated to be a maximum of 40 mm . About 5 per cent of total cerebral cortical GABA was found in the synaptosomal fraction.  相似文献   

9.
Ultrasonic telemetry was used to compare post‐release survival and movements of Atlantic sharpnose sharks Rhizoprionodon terraenovae in a coastal area of the north‐east Gulf of Mexico. Ten fish were caught with standardized hook‐and‐line gear during June to October 1999. Atlantic sharpnose sharks were continuously tracked after release for periods of 0·75 to 5·90 h and their positions recorded at a median interval of 9 min. Individual rate of movement was the mean of all distance and time measurements for each fish. Mean ± s.e . individual rate of movement was 0·45 ± 0·06 total lengths per second (LT s?1) and ranged from 0·28 to 0·92 LT s?1 over all fish. Movement patterns did not differ between jaw and internally hooked Atlantic sharpnose sharks. Individual rate of movement was inversely correlated with bottom water temperature at capture (r2 = 0·52, P ≤ 0·05). No consistent direction in movement was detected for Atlantic sharpnose sharks after release, except that they avoided movement towards shallower areas. Capture‐release survival was high (90%), with only one fish not surviving, i.e. this particular fish stopped movement for a period of 10 min. Total rate of movement was total distance over total time (m min?1) for each Atlantic sharpnose shark. Mean total rate of movement was significantly higher immediately after release at 21·5 m min?1 over the first 1·5 h of tracking, then decreased to 11·2 m min?1 over 1·5–6 h, and 7·7 m min?1 over 3–6 h (P ≤ 0·002), which suggested initial post‐release stress but quick recovery from capture. Thus, high survival (90%) and quick recovery indicate that the practice of catch‐and‐release would be a viable method to reduce capture mortality for R. terraenovae.  相似文献   

10.
Aims: To compare physical properties of spores that were produced in broth sporulation media at greater than 108 spores ml−1. Methods and Results: Bacillus atrophaeus reproducibly sporulated in nutrient broth (NB) and sporulation salts. Microscopy measurements showed that the spores were 0·68 ± 0·11 μm wide and 1·21 ± 0·18 μm long. Coulter Multisizer (CM3) measurements revealed the spore volumes and volume-equivalent spherical diameters, which were 0·48 ± 0·38 μm3 and 0·97 ± 0·07 μm, respectively. Bacillus cereus reproducibly sporulated in NB, sporulation salts, 200 mmol l−1 glutamate and antifoam. Spores were 0·95 ± 0·11 μm wide and 1·31 ± 0·17 μm long. Spore volumes were 0·78 ± 0·61 μm3 and volume-equivalent spherical diameters were 1·14 ± 0·11 μm. Bacillus atrophaeus spores were hydrophilic and B. cereus spores were hydrophobic. However, spore hydrophobicity was significantly altered after treatment with pH-adjusted bleach. Conclusions: The utility of a CM3 for both quantifying Bacillus spores and measuring spore sizes was demonstrated, although the volume between spore exosporium and spore coat was not measured. This study showed fundamental differences between spores from a Bacillus subtilis- and B. cereus-group species. Significance and Impact of the Study: This is useful for developing standard methods for broth spore production and physical characterization of both living and decontaminated spores.  相似文献   

11.
N-2′-Acetoxybenzoyl (aspirin) derivatives (degree of substitution 0·35–1·00) of chitosan, N-desulphated heparin and 2-amino-2-deoxy-d-glucose were prepared by methods that gave yields in the range 65–86%. The salicylate of chitosan was isolated with a 98% yeild. Aspirin or salicylic acid was released much more slowly from N-(2′-acetoxybenzoyl)-chitosan than from the salicylate of chitosan, and much faster at 37°C in 0·1 m NaOH solution than in 2% aqueous acetic acid solution. Salicylic acid was isolated from the dialysate (0·1 m NaOH solution) of N-(2′-acetoxybenzoyl)-chitosan.  相似文献   

12.
Petioles of water‐sufficient intact Vicia faba L. plants were infused with 1 µm abscisic acid (ABA) to simulate the import of root‐source ABA. This protocol permitted quantitative ABA delivery, up to 300 pmol ABA over 60 min, to the leaf without ambiguities associated with perturbations in plant–water status. The ABA concentrations in whole‐leaf samples and in apoplastic sap increased with the amount infused; ABA degradation was not detected. The ABA concentration in apoplastic sap was consistent with uptake of imported ABA into the leaf symplast, but this interpretation is qualified. Our focus was quantitative cellular compartmentation of imported ABA in guard cells. Unlike when leaves are stressed, the guard‐cell symplast ABA content did not increase because of ABA infusion (P = 0·48; 3·0 ± 0·5 versus 4·0 ± 1·2 fg guard‐cell‐pair?1). However, the guard‐cell apoplast ABA content increased linearly (R2 = 0·98) from ?0·2 ± 0·5 to 3·1 ± 1·3 fg guard‐cell‐pair?1 (≈ 3·1 µm ) and was inversely related to leaf conductance (R2 = 0·82). Apparently, xylem ABA accumulates in the guard‐cell wall as a result of evaporation of the apoplast solution. This mechanism provides for integrating transpiration rate and ABA concentration in the xylem solution.  相似文献   

13.
Wild adult specimens of the Peruvian anchovy Engraulis ringens were captured and reared to validate the daily periodicity of otolith microincrement formation. The postcapture stress generated spontaneous spawning, making it possible to conduct a rearing trial on larvae first in an artificial nutrient‐enriched system (ANES) for 52 days followed by an artificial feeding regime in a culture tank until day 115 post‐hatch. Microincrements of the sagittal otoliths of sacrificed juveniles [mean ± s.d. total length (LT) = 5·13 ± 0·37 cm, range 5–6 cm; c.v. = 7·5%] showed very distinct light and dark zones. The slope of the relationship between the total number of increments after the hatch check and days elapsed after hatching was not significantly different from 1. The transfer from ANES to the artificial feeding regime induced a mark in the sagittal otoliths. The number of microincrements after this induced mark coincided with the number of days elapsed after the transfer date. In parallel experiments, adult E. ringens (mean ± s.d. LT = 14·92 ± 0·55 cm, range 13–16 cm) were exposed to one of two fluorescent marking immersion treatments with either alizarin red S (ARS; 25 mg l?1 per 6 h) or oxytetracycline hydrochloride (OTC; 200 mg l?1 per 10 h). The microincrements between fluorescent bands were distinct, ranging from 0·89 to 2·75 µm (mean ± s.d. =1·43 ± 0·28 µm; c.v. = 32%) and from 0·71 to 2·89 µm (1·53 ± 0·27 µm; c.v. = 35%) for ARS and OTC, respectively. The relationship between the number of microincrements between marks and the number of elapsed days for ARS and OCT treatments indicated that there was a significant correspondence between the number of increases observed and the number of days. Hence, daily microincrements of otoliths of E. ringens are likely to be formed in juveniles and adults under natural conditions.  相似文献   

14.
The penetration of 137Cs by contact exchange through cuticular membranes from the adaxial surface of leaves of Pyrus communis and Prunus cerasus has been investigated. The resistance of the cuticles to the caesium penetration was dependent on the counter-ions associated with the fixed negative ion exchange sites in the membrane. The mobility of hydrated potassium ions and their tenuous connection to -COOH?-groups in the membrane encouraged caesium permeation in contrast to cuticular membranes with predominantly protonized ion exchange sites. Divalent calcium ions caused a strong reduction (4–20 times) of the caesium permeability which is decisive for the calculation of the caesium uptake by the intact leaf. Under these conditions, a penetration rate of the deposited caesium of 0·11±0·05% h?1 for pear cuticles and of 0·036±0·025% h?1 for cherry cuticles was measured after the adjustment to steady state conditions. Approximately 12–24% and 4·5–7·5% of initially retained caesium could be absorbed by the leaves of pear and cherry, respectively, in a rain-free period of 7 d in the area of Munich after wet deposition of fallout from the Chernobyl reactor accident. Furthermore, the caesium penetration from the physiological inside to the outside of the membrane was found to be smaller by a factor of 100–150 compared with that of the opposite direction.  相似文献   

15.
Striped trumpeter larvae reared in algal cell‐induced turbid water (greenwater) fed equally well in clearwater in a light intensity range of 1–10 μmol s‐1 m‐2, when evaluated in terms of both the proportion of larvae feeding and larval feeding intensity. An ontogenetic improvement in photopic visual sensitivity of larvae was indicated by improved feeding at 0·1 μmol s‐1 m‐2, from 26±5% of larvae feeding and 0·027±0·005 rotifers consumed per feeding larva min‐1 on day 8, to 96±2% and 0·221±0·007 rotifers consumed larva‐1 min‐1 on day 23 post‐hatching. Algal cell‐induced turbidity was shown to reduce incident irradiance with depth, indicated by increasing coefficients of attenuation (1·4–33·1) with increasing cell densities (0–2×106 cells ml‐1), though light intensities in the feeding experiment test chambers, at the algal cell densities tested, were within the optimal range for feeding (1–10 μmol s‐1 m‐2). Algae‐induced turbidity had different effects on larval feeding response dependent upon the previous visual environment of the larvae. Young larvae (day 9 post‐hatching) reared in clearwater showed decreased feeding capabilities with increasing turbidity, from 98±1% feeding and 0·153±0·022 rotifers consumed larva‐1 min‐1 in clearwater to 61±10% feeding and 0·042±0·004 rotifers consumed larva‐1 min‐1 at 56 NTU, while older clearwater reared larvae fed well at all turbidities tested. Likewise, greenwater reared larvae had increased feeding capabilities in the highest algal cell densities tested (32 and 66 NTU) compared with those in low algal cell density (6 NTU), and clearwater (0·7 NTU) to which they were naïve.  相似文献   

16.
The kinetic parameters of the inhibition of pigeon brain acetylchlolinesterase (AChE) by procaine hydrochloride were investigated. Procaine (0·083–1·67 mM) reversibly inhibited AChE activity (15–83 percent) in a concentration dependent manner, the IC50 being about 0·38 mM. The Michaelis-Menten constant (Km) for the hydrolysis of acetylthiocholine iodide was found to be 1·53 × 10?4 M and the Vmax was 1·06 μmol min?1 mg?1 protein. Dixon as well as Lineweaver-Burk plots and their secondary replots indicated that the nature of the inhibition is of the linear mixed type which is considered to be a mixture of partial competitive and pure non-competitive. The values of Ki(slope) and Ki (intercepts) were estimated as 0·14 mM and 0·22 mM respectively by the primary Dixon and by the secondary replots of the Lineweaver-Burk plot. The Ki′/Ki ratio shows that procaine has a greater affinity of binding for the peripheral than for the active site.  相似文献   

17.
A series of experiments were carried out to construct an energy budget for juvenile thick lipped mullet, Crenimugil labrosus Risso. A partial factorial experimental design was used to examine the effects of temperature, fish size and meal size on growth. The maximum ration that the fish were able to ingest completely per day was found to be 0·8, 1·4 and 2·3% wet body weight (b.w.) at 13,18 and 23°C, respectively. Ingested maintenance requirements (M.R.) were estimated to be 137, 205 and 288 cal fish-1 day-1 at 13, 18 and 23°C, respectively. At 18deg; C, M.R. varied as 25 W1.04 cal day-1, where W= fish weight (g). Growth rate increased with increasing temperature. Maximal conversion efficiency was 21–24% and was achieved closer to the maximum ingested ration with increasing temperature. The relationship between respiration rate and W at 18deg; C for 3-20 g fish is described by: respiration rate (ml O2 h-1) = 0·128 W0.976 The energy cost of apparent specific dynamic action at 18deg; C was found to vary between 5·1% and 23·6% of the calorific value of the ingested meal (1% wet b.w.) , mean (± S.E.)=10·2 ± 2·0%. Post mortem analyses of groups of fish fed 0·2, 0·8 and 1·5% wet b.w. meals showed a significant increase in total lipid and a significant decrease in water content with increasing ratio size. A negative correlation was found between body water content and total lipid (and calories). The mean assimilation efficiency (±s.e.) for 5–10 g mullet at 18deg; C was 73·9 ± 3·6%. The observations reported in this study were brought together to construct an energy budget for juvenile C. labrosus which was found to give a reliable prediction (within 10%) of energy demand and growth under the prevailing experimental conditions. Both gross (K1) and net (K2) growth efficiencies, based on energy values, increased with increasing ratio size up to satiation and were independent of temperature. The maximum values of K1 and K2 observed were 0·33 and 0·46, respectively. The third order efficiency (K3) appeared to be independent of temperature and ration size; mean values ranged between 0·66 and 0·84.  相似文献   

18.
Although vanadium is found abundantly in animal and plant kingdoms its biological effects are not clear. Vanadate compounds have been shown to normalize blood glucose levels in streptozotocin treated rats, enhance glucose oxidation and improve the sensitivity to insulin by enhanced receptor binding in rat adipocytes. The aim of the present study was to investigate the effect of vanadate, at high (0–8 mmol l?1) and low (0–1·0 mmol l?1) physiological concentrations, on [125I]-insulin binding in the placenta of three groups of pateints, namely from normal (N) controls, gestational diabetics (GDM) and women with risk factors in their medical history for developing diabetes mellitus (RF). Vanadate at low concentrations (0·2–0·6 mmol l?1) enhanced the maximal binding 2-fold in GDM placenta but only increased (up to 1·2-fold) the binding slightly at high cncentrations (5 mmol l?1). However with placenta from normal or women at risk, vanadate increased the [125I]-insulin binding up to 1·2-fold both at low and high concentrations. Thus it appears that vanadate augements insulin binding in the placenta from women with gestational diabetes mellitus.  相似文献   

19.
Crustose coralline algae occupied ~1%–2% (occasionally up to 7%) of the sea floor within their depth range of 15–50 m, and they were the dominant encrusting organisms and macroalgae beyond 20 m depth in Young Sound, NE Greenland. In the laboratory, oxygen microelectrodes were used to measure net photosynthesis (P) versus downwelling irradiance (Ed) and season for the two dominant corallines [Phymatolithon foecundum (Kjellman) Düwel et Wegeberg 1996 and Phymatolithon tenue (Rosenvinge) Düwel et Wegeberg 1996] representing> 90% of coralline cover. Differences in P‐Ed curves between the two species, the ice‐covered and open‐water seasons, or between specimens from 17 and 36 m depth were insignificant. The corallines were low light adapted, with compensation irradiances (Ec) averaging 0.7–1.8 μmol photons·m ? 2·s ? 1 and light adaptation (Ek) indices averaging 7–17 μmol photons·m ? 2·s ? 1. Slight photoinhibition was evident in most plants at irradiances up to 160 μmol photons·m ? 2·s ? 1. Photosynthetic capacity (Pm) was low, averaging 43–67 mmol O2·m ? 2 thallus·d ? 1 (~250–400 g C·m ? 2 thallus·yr ? 1). Dark respiration rates averaged ~5 mmol O2·m ? 2 thallus·d ? 1. In ice covered periods, Ed at 20 m depth averaged ~1 μmol photons·m ? 2·s ? 1, with daily maxima of 2–3 μmol photons·m ? 2·s ? 1. During the open water season, Ed at 20 m depth averaged ~7 μmol photons·m ? 2·s ? 1 with daily maxima of ~30 μmol photons·m ? 2·s ? 1. Significant net primary production of corallines was apparently limited to the 2–3 months with open water, and the small contribution of corallines to primary production seems due to low Pm values, low in situ irradiance, and their relatively low abundance in Young Sound.  相似文献   

20.
The kinetics of the chlorophyll fluorescence rise induced by adding 20 mM MgCl2 to a suspension of isolated pea chloroplasts treated with 3-(3,4-dichlorophenyl)-1, 1-dimethylurea (DCMU) have been examined experimentally and theoretically as a function of temperature. The application of similarity arguments and particle aggregation theory to the experimental results suggests that at the first approximation, the salt-induced time-dependent fluorescence changes may be described by the diffusion-controlled lateral movement of Photosystem II pigment-protein complexes. From an analysis of the temperature dependence of the fluorescence changes, estimates obtained for the lateral diffusion coefficients were 1.85 · 10?12–3.08 · 10?11 cm2/s over the temperature range 10°C ? T?30°C.  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