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 968 毫秒
1.
The ability of mebendazole and fenbendazole to bind to tubulin in cytosolic fractions from 8-day Ascaris suum embryos was determined by inhibition studies with [3H]colchicine. Colchicine binding in the presence of 1·10?6 M mebendazole was completely inhibited during a 6 h incubation period at 37°C. Inhibition of colchicine binding to A. suum embryonic tubulin by mebendazole and fenbendazole appeared to be noncompetative. The inhibition constants of mebendazole and fenbendazole for A. suum embryonic tubulin were 1.9·10?8 M and 6.5·10?8 M, respectively. Mebendazole and fenbendazole appeared to be competitive inhibitors of colchicine binding to bovine brain tubulin. The inhibition constants of mebendazole and fenbendazole for bovine brain tubulin were 7.3·10?6 M and 1.7·10?5 M, respectively. These values are 250–400 times greater than the inhibition constants of fenbendazole and mebendazole for A. suum embryonic tubulin. Differential binding affinities between nematode tubulin and mammalian tubulin for benzimidazoles may explain the selective toxicity. The importance of tubulin as a receptor for anthelmintic benzimidazoles in animal parasitic nematodes is discussed.  相似文献   

2.
The osmotic permeability coefficient (Pf) for water movement across Novikoff hepatoma cells was found to be 82 ± 3 (S.E.) · 10?5 cm · s?1 at 20°C. The corresponding diffusional permeability coefficient for 3HHO (Pd) was 97 ± 10 (S.E.) · 10?5 cm · s?1, therefore the ratio PfPd is close to unity. The apparent activation energy for water filtration was 10.4 ± 0.4 (S.E.) kcal · mol?1. This value is significantly greater than the activation energy for the self diffusion of water. The product of the hydraulic permeability coefficient and the viscosity coefficient for water was temperature-dependent. However, the product of the hydraulic permeability coefficient and the viscosity coefficient for membrane lipid did not vary with temperature. These data are interpreted as evidence for water movement across a lipid membrane barrier rather than through aqueous channels.  相似文献   

3.
The transport of sterols incorporated into the lecithin bilayer of small unilamellar liposomes through a model membrane was studied. A two-chamber diffusion cell containing liposomes with incorporated [4-14C]cholesterol or β-[4-14C]sitosterol in the donor chamber and liposomes with unlabeled cholesterol in the receiver chamber was used. The permeability coefficients of the sterols through silastic rubber membranes which served as a model membrane were measured. The permeability for cholesterol incorporated into liposomes in a phosphatidyl choline/cholesterol molar ratio of 1 : 1, produced by sonication for 1 h, and subsequent centrifugation at 100000 × g for 1 h, was 1.6 · 10?8 cm sec?1. Dilution of the liposome suspension did not change the permeability coefficient significantly. The permeability coefficient of sitosterol incorporated into liposomes was about 4-times smaller than that of cholesterol. These results suggest that the sterols were delivered to the silastic membrane by the intact liposomes and that free solute was not involved in the transport to the membrane to a significant degree. The large differences in the permeability coefficients between cholesterol and sitosterol indicate that an aqueous interfacial barrier was crossed by the sterol during the delivery to the membrane.  相似文献   

4.
Leakage of the entrapped anionic fluorophore carboxyfluorescein was used as a measure of the permeability of liposomes to several different acids. Carboxyfluorescein leakage increased with increasing buffer concentration at a given pH and depended on its chemical nature: apolar weak acids such as acetic or pyruvic acids induced fast leakage at relatively high pH (4 to 5), while glycine, aspartic, citric and hydrochloric acids induced leakage only at lower pH. Fluorescence leakage measurements reflected the acidification of the liposomes' aqueous spaces, which was primarily caused by the diffusion of undissociated acid molecules across the lipid bilayer. A simple mathematical model in accord with this hypothesis and assuming that carboxyfluorescein leakage was directly related to the proportion of its neutral lactone form, described satisfactorily the carboxyfluorescein leakage kinetics and allowed rough estimation of permeability coefficients for carboxyfluorescein (neutral lactone form; 9 · 10?9 cm · s?1), acetic acid (>1 · 10?7cm · s?1) and glycine (cation: 6 · 10?9 cm · s?1). These results are consistent with low effective proton permeability of liposomes (<5 · 10?12cm · s?1) and with the permeability coefficient of HCl (3 · 10?3 cm · s?1) reported by Nozaki and Tanford ((1981) Proc. Natl. Acad. Sci. U.S.A. 78, 4324–4328). Diffusion of weak acid molecules across lipid membranes has implications for drug encapsulation and delivery, and may be of biological significance.  相似文献   

5.
6.
Three photosynthetic parameters of 7 species of marine diatoms were studied using Na214CO3 at 5–8 C using log phase axenic cultures. The cell volumes of the different species varied from 70 μm3 to 40 × 105μm3. The present experiment is consistent with the interpretation that the initial slope α (mg C · [mg chl a]?1· h?1· w?1· m2) of photosynthesis vs. light curves is controlled by self-shading of chlorophyll a in the cell. Pm, the rate of photosynthesis at light saturation (mg C · [mg cell, C]?1· h?1) and R, the intercept at zero light intensity (mg C · [mg cell C]?1· H?1) are both dependent on the ratio of surface area to volume of cell.  相似文献   

7.
To establish the capacity of the leaf mesophyll plasmalemma of Phaseolus vulgaris L. to supply ascorbate (ASC) into the cell wall by simple diffusion, a method for calculating plasmalemma diffusional conductivity to ascorbic acid (AA) in intact leaves was evaluated. The core of the approach is that in the presence of a sink for ascorbate in the cell wall, cell wall total ascorbic acid concentration [TAA]cw (=[ASC]cw+[AA]cw) reaches zero at some positive whole‐leaf total ascorbic acid concentration [TAA]l. It is shown that [TAA]l at [TAA]cw=0 is proportional to the sink for ASC in the cell wall and the reciprocal of plasmalemma conductivity. The predicted proportional relationship between [TAA]cw and [TAA]l was confirmed by decreasing TAA levels in leaves through predarkening. Furthermore, increasing the sink intensity for ASC in the cell wall by the acute exposure of leaves to 450 nmol ozone mol?1 during re‐illumination, [TAA]cw reached zero at 2.7‐fold higher [TAA]l than without ozone, and the slope of the relationship increased twofold. Plasmalemma diffusional conductivities to AA of 2.9×10?6 and 1.8×10?6 m s?1, needed to maintain [TAA]cw at the observed level, were calculated from the increase in [TAA]l at [TAA]cw=0 and from the two different estimates of the sink for ASC. A value of 1.3×10?6 m s?1 was calculated on the basis of the oil‐water distribution coefficient for TAA. It is concluded that the demand for ASC in the mesophyll cell wall of the investigated leaves could be met by simple diffusion of AA through the plasmalemma. From the measured increase in the slope of the relationship [TAA]cw versus [TAA]l, an increase in the cell wall pH of 0.3 units was estimated under the influence of ozone.  相似文献   

8.
Mice infected with Echinococcus granulosus were given one dose of 15 mg kg?1 of 14C-mebendazole by gavage. Blood, hydatid cyst fluid and membranes were collected and counted at varying intervals thereafter. Radioactivity in blood peaked by 16 h and declined rapidly thereafter. Activity in hydatid cyst fluid paralleled that in blood but in amounts of only 5–10%. While levels of radioactivity in hydatid cyst membranes for the most part paralleled those of blood, in several samples they remained stable or increased from 16 to 48 h while those in blood had decreased to baseline. Protoscolices lost all signs of viability by 48 h after treatment.  相似文献   

9.
The transport of 3-O-methylglucose in white fat cells was measured under equilibrium exchange conditions at 3-O-methylglucose concentrations up to 50 mM with a previously described method (Vinten, J., Gliemann, J. and Østerlind, K. (1976) J. Biol. Chem. 251, 794–800). Under these conditions the main part of the transport was inhibitable by cytochalasin B. The inhibition was found to be of competitive type with an inhibition constant of about 2.5 · 10?7 M, both in the absence and in the presence of insulin (1μM). The cytochalasin B-insensitive part of the 3-O-methylglucose permeability was about 2 · 10?9 cm · s?1, and was not affected by insulin. As calculated from the maximum transport capacity, the half saturation constant and the volume/ surface ratio, the maximum permeability of the fat cell membrane to 3-O-methylglucose at 37°C and in the presence of insulin was 4.3 · 10?6 cm · s?1. From the temperature dependence of the maximum transport capacity in the interval 18–37°C and in the presence of insulin, an Arrhenius activation energy of 14.8 ± 0.44 kcal/mol was found. The corresponding value was 13.9 ± 0.89 in the absence of insulin. The half saturating concentration of 3-O-methylglucose was about 6 mM in the temperature interval used, and it was not affected by insulin, although this hormone increased the maximum transport capacity about ten-fold to 1.7 mmol · s?1 per 1 intracellular water at 37°C.  相似文献   

10.
In slow mainstream flows (<4–6 cm · s?1), the transport of dissolved nutrients to seaweed blade surfaces is reduced due to the formation of thicker diffusion boundary layers (DBLs). The blade morphology of Macrocystis pyrifera (L.) C. Agardh varies with the hydrodynamic environment in which it grows; wave‐exposed blades are narrow and thick with small surface corrugations (1 mm tall), whereas wave‐sheltered blades are wider and thinner with large (2–5 cm) edge undulations. Within the surface corrugations of wave‐exposed blades, the DBL thickness, measured using an O2 micro‐optode, ranged from 0.67 to 0.80 mm and did not vary with mainstream velocities between 0.8 and 4.5 cm · s?1. At the corrugation apex, DBL thickness decreased with increasing seawater velocity, from 0.4 mm at 0.8 cm · s?1 to being undetectable at 4.5 cm · s?1. Results show how the wave‐exposed blades trap fluid within the corrugations at their surface. For wave‐sheltered blades at 0.8 cm · s?1, a DBL thickness of 0.73 ± 0.31 mm within the edge undulation was 10‐fold greater than at the undulation apex, while at 2.1 cm · s?1, DBL thicknesses were similar at <0.07 mm. Relative turbulence intensity was measured using an acoustic Doppler velocimeter (ADV), and overall, there was little evidence to support our hypothesis that the edge undulations of wave‐sheltered blades increased turbulence intensity compared to wave‐exposed blades. We discuss the positive and negative effects of thick DBLs at seaweed surfaces.  相似文献   

11.
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   

12.
The permeability of rye leaf protoplasts to glycerol was determined using 1,3-14C glycerol and liquid scintillation spectrometry. Estimates were 1.0×10−8 m s−1 at 0°C and 4.1×10−8 m s−1 at 22 and 31°C. The activation energy for glycerol permeability was 32.8 kJ/mol. The effect of electroporation on glycerol uptake was also explored. Treatments were performed with a field strength of 100 V/cm and an exponential decay constant of 5.8 ms. At 22 °C, electroporation affected the rate and extent of glycerol permeation, causing an increase in the intercept of the glycerol uptake curve and a decrease in the slope. Electroporation had no significant effect on glycerol uptake when performed at 0°C, when the cells were electroporated at 0°C then warmed to 31 °C, or when the cells were electroporated at 22 °C then cooled to 0°C. The results at 22°C were consistent with an influx of glycerol during electroporation. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

13.
Three rice varieties, cv. Norin 36, cv. Norin 37 and cv. Yubae, were grown in a loam with a 20 cm water-table which gave aerobic conditions to a depth of not less than 15 to 17 cm. Under these conditions Norin 36 grew more vigorously and tillered more frequently than the other two varieties. The rates of oxygen diffusion at 23°C from roots up to 11 cm in length were however appreciably lower for Norin 36 (4.3 × 10?8g · cm?2 of root surface · min?1) than for Norin 37 or Yubae (c. 7.8 × 10?8g). A considerable increase (up to 200 %) in the oxygen diffusion rate (ODR) from the roots occurred if they were cooled to 3°C, and at this temperature differences in ODR between the varieties were not significant. For a purely physical system, because of the decrease in the diffusion coefficient of oxygen in water, and, the increase in oxygen solubility, a drop of c. 20 % in ODR should accompany the above 20°C drop in temperature. A 16 % drop was recorded for artificial ‘roots’ under these conditions. It was concluded that respiratory activity at the higher temperature must have been responsible for the low readings and intervarietal differences observed at 23°C. By increasing the 3°C values by 25 % a mean value of 14.2 × 10?8g · cm?2 of root surface · min?1 was recorded for the three varieties, being the probable ODR at 23°C in the absence of a respiratory factor. Calculations show that respiratory activity removed enough oxygen to reduce the ODR for Norin 36 by more than 9 × 10?8g, and for Norin 37 and Yubae by c. 6.7 × 10?8g · cm?2 of root surface · min?1. Anatomical investigations showed that cortical breakdown was always extensive at 4 to 4.5 cm from the apex of the roots. In some cases however breakdown had not occurred in the basal segment of the root. No opinion could be formed as to whether differences in the amount of cortical breakdown between the varieties might have occasioned the respiratory differences observed. An interesting feature of the root anatomy was the failure of breakdown in those regions of the roots through which lateral roots emerged.  相似文献   

14.
Transport properties of cuticular waxes from 40 different plant species were investigated by measuring desorption rates of 14C-labelled octadecanoic acid from isolated and subsequently reconstituted wax. Diffusion coefficients (D) of octadecanoic acid in reconstituted waxes, calculated from the slopes of the regression lines fitted to the linearized portions of desorption kinetics, ranged from 1.2 × 10?19 m2 s?1 (Senecio kleinia leaf) to 2.9 × 10?17 m2 s?1 (Malus cf. domestica fruit). Cuticular water permeabilities (cuticular transpiration) measured with intact cuticular membranes isolated from 24 different species varied from 1.7 × 10?11 m s?1 (Vanilla planifolia leaf) up to 2.1 × 10?9 m s?1 (Malus cf. domestica fruit), thus covering a range of more than 2 orders of magnitude. Cuticular water permeabilities were highly correlated with diffusion coefficients of octadecanoic acid in isolated cuticular wax of the same species. It is therefore possible to estimate cuticular barrier properties of stomatous leaf surfaces or of leaves where isolation of the cuticle is impossible by measuring D of octadecanoic acid in isolated waxes of these leaves.  相似文献   

15.
The characteristics of water diffusional permeability (P) of human red blood cells were studied on isolated erythrocytes by a doping nuclear magnetic resonance technique. In order to estimate the basal permeability the maximal inhibition of water diffusion was induced by exposure of red blood cells to p-chloromercuribenzene sulfonate (PCMBS) under various conditions (concentration, duration, temperature). The lowest values of P were around 0.7×10–3 cm s–1 at 10°C, 1.2×10–3 cm s–1 at 15°C, 1.4×10–3 cm s–1 at 20°C, 1.8×10–3 cm s–1 at 25°C, 2.1×10–3 cm s–1 at 30°C and 3.5×10–3 cm s–1 at 37°C. The mean value of the activation energy of water diffusion (Ea,d) was 25 kJ/mol for control and 43.7 kJ/mol for PCMBS-inhibited erythrocytes. The values of P and Ea,d obtained after induction of maximal inhibition of water diffusion by PCMBS can be taken as references for the basal permeability to water of the human red blood cell membrane.  相似文献   

16.
Critical (<30 min) and prolonged (>60 min) swimming speeds in laboratory chambers were determined for larvae of six species of Australian freshwater fishes: trout cod Maccullochella macquariensis, Murray cod Maccullochella peelii, golden perch Macquaria ambigua, silver perch Bidyanus bidyanus, carp gudgeon Hypseleotris spp. and Murray River rainbowfish Melanotaenia fluviatilis. Developmental stage (preflexion, flexion, postflexion and metalarva) better explained swimming ability than did length, size or age (days after hatch). Critical speed increased with larval development, and metalarvae were the fastest swimmers for all species. Maccullochella macquariensis larvae had the highest critical [maximum absolute 46·4 cm s?1 and 44·6 relative body lengths (LB) s?1] and prolonged (maximum 15·4 cm s?1, 15·6 LB s?1) swimming speeds and B. bidyanus larvae the lowest critical (minimum 0·1 cm s?1, 0·3 LB s?1) and prolonged swimming speeds (minimum 1·1 cm s?1, 1·0 LB s?1). Prolonged swimming trials determined that the larvae of some species could not swim for 60 min at any speed, whereas the larvae of the best swimming species, M. macquariensis, could swim for 60 min at 44% of the critical speed. The swimming performance of species with precocial life‐history strategies, with well‐developed larvae at hatch, was comparatively better and potentially had greater ability to influence their dispersal by actively swimming than species with altricial life‐history strategies, with poorly developed larvae at hatch.  相似文献   

17.
Colonies of the stream-inhabiting cyanobacterium Nostoc parmelioides Kützing often contain a single endosymbiotic dipteran larva Cricotopus nostocicola (Wirth), which induces a morphological change from small, spherical colonies to larger, ear-shaped colonies. At a current velocity of 0 cm · s?1, whole colonies containing the midge showed overall rates of 14CO2 uptake and nitrogenase activity that were higher than those when the midge was absent (sphere-shaped colonies). Spherical colonies incubated at current velocities of 5-10 cm · s?1did not show higher rates of 14CO2 or 15N2 incorporation than those with the larvae (ear-shaped colonies). Ear-shaped colonies extended well into regions of higher current velocity, whereas spherical colonies did not. Photosynthesis of ear-shaped colonies was stimulated by increased current velocity, increased inorganic C and decreased O2 concentrations. Moreover, levels of O2 at the surface of midge-inhabited colonies decreased with increased current velocity. The morphological change induced by the larva is detrimental (lowers photosynthesis and N2 fixation) in quiescent water but not at current velocities above 10 cm · s?1. This is probably a result of higher diffusion of O2 and CO2 associated with the midge-induced morphology.  相似文献   

18.
The action of anticonvulsant drugs, phenytoin, diazepam, clonazepam and phenobarbitone, was tested on the release of [14C]-GABA from tissue slices of rat cerebral cortex. All drugs caused a significant dose-dependent depression of the 33mM-K+-evoked release of [14C]-GABA but had little effect on the resting release of [14C]-GABA, except at high concentrations. The IC50 values for inhibition of K+-evoked release of [14C]-GABA were 4.7 × 10?5, 7 × 10?5, 28 × 10?5 and 7.9 × 10?4M for diazepam, clonazepam, phenytoin and phenobarbitone respectively. Trifluoperazine also caused a similar and complete inhibition of [14C]-GABA release with an IC50 of 1 × 10?5M. The effect of diazepam and trifluoperazine were additive. The inhibition by trifluoperazine could be overcome by addition of exogenous calmodulin, whereas that of diazepam, phenytoin or phenobarbitone was not overcome. It is proposed that the anticonvulsants tested inhibit calcium-dependent transmitter release at a site distal to the formation of a calcium-calmodulin complex, which is presumably activated by this complex. Trifluoperazine, on the other hand, acts by reducing the availability of calmodulin.  相似文献   

19.
Diffusion of histamine, theophylline and tryptamine through planar lipid bilayer membranes was studied as a function of pH. Membranes were made of egg phosphatidylcholine plus cholesterol (1 : 1 mol ratio) in tetradecane. Tracer fluxes and electrical conductances were used to estimate the permeabilities to nonionic and ionic species. Only the nonionic forms crossed the membrane at a significant rate. The membrane permeabilities to the nonionic species were: histamine, 3.5 · 10?5cm · s?1; theophylline, 2.9 · 10?4cm · s?1; and tryptamine, 1.8 · 10?1cm · s?1. Chemical reactions in the unstirred layers are important in the transport of tryptamine and theophylline, but not histamine. For example, as pH decreased from 10.0 to 7.5 the ratio of nonionic (B) to ionic (BH+) tryptamine decreased by 300-fold, but the total tryptamine permeability decreased only 3-fold. The relative insensitivity of the total tryptamine permeability to the ratio, [B]/[BH+], is due to the rapid interconversion of B and BH+ in the instirred layers. Our model describing diffusion and reaction in the unstirred layers can explain some ‘anomalous’ relationships between pH and weak acid/base transport through lipid bilayer and biological membranes.  相似文献   

20.
Maximum sustained swimming speeds, swimming energetics and swimming kinematics were measured in the green jack Caranx caballus (Teleostei: Carangidae) using a 41 l temperature‐controlled, Brett‐type swimming‐tunnel respirometer. In individual C. caballus [mean ±s.d. of 22·1 ± 2·2 cm fork length (LF), 190 ± 61 g, n = 11] at 27·2 ± 0·7° C, mean critical speed (Ucrit) was 102·5 ± 13·7 cm s?1 or 4·6 ± 0·9 LF s?1. The maximum speed that was maintained for a 30 min period while swimming steadily using the slow, oxidative locomotor muscle (Umax,c) was 99·4 ± 14·4 cm s?1 or 4·5 ± 0·9 LF s?1. Oxygen consumption rate (M in mg O2 min?1) increased with swimming speed and with fish mass, but mass‐specific M (mg O2 kg?1 h?1) as a function of relative speed (LF s?1) did not vary significantly with fish size. Mean standard metabolic rate (RS) was 170 ± 38 mg O2 kg?1 h?1, and the mean ratio of M at Umax,c to RS, an estimate of factorial aerobic scope, was 3·6 ± 1·0. The optimal speed (Uopt), at which the gross cost of transport was a minimum of 2·14 J kg?1 m?1, was 3·8 LF s?1. In a subset of the fish studied (19·7–22·7 cm LF, 106–164 g, n = 5), the swimming kinematic variables of tailbeat frequency, yaw and stride length all increased significantly with swimming speed but not fish size, whereas tailbeat amplitude varied significantly with speed, fish mass and LF. The mean propulsive wavelength was 86·7 ± 5·6 %LF or 73·7 ± 5·2 %LT. Mean ±s.d . yaw and tailbeat amplitude values, calculated from lateral displacement of each intervertebral joint during a complete tailbeat cycle in three C. caballus (19·7, 21·6 and 22·7 cm LF; 23·4, 25·3 and 26·4 cm LT), were 4·6 ± 0·1 and 17·1 ± 2·2 %LT, respectively. Overall, the sustained swimming performance, energetics, kinematics, lateral displacement and intervertebral bending angles measured in C. caballus were similar to those of other active ectothermic fishes that have been studied, and C. caballus was more similar to the chub mackerel Scomber japonicus than to the kawakawa tuna Euthynnus affinis.  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