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 effects of amino acids (aa) and N-(diisopropyloxyphosphoryl)-amino acids (DIPP-aa) on cell membranes were investigated by evaluating water and methyl urea permeability. Permeability coefficients Pf and Ps were determined by standard osmotic methods for cells ofPisum sativum stem base epidermis after 20 min exposure to a 5 mM solution of each aa and DIPP-aa. The Pf value ofP. sativum epidermal cells (untreated controls) was 1.3 ± 0.4 × 10-3 μm s-1. Treat ments with the diisopropyl-oxyphosphoryl derivatives of three one charged and three polar amino acids (serine, threonine, asparagine, and aspartic acid) and unsubstituted (free) serine and threonine increased water permeability up to about two fold of the control value. Serine and threonine and their DIPP-derivatives increased methyl urea permeability (controls 1.03 ± 0.09 × 10-3 μm s-1) 30 to 80 percent Other amino acids and their DIPP-derivatives caused small or insignificant changes of water permeability. Only certain polar amino acids and their DIPP-derivatives increased the osmotic water and methyl urea permeation through the plasma membrane. The specificity of these molecules on plasma membranes suggests that the active amino acids (serine and threonine) and their DIPP-derivatives interact with charged membrane molecules. The relatively small changes in water and methyl urea permeability may indicate that the effective aa’s and their DIPP-derivatives interact with phospholipids rather than aquaporin. A concurring alteration of water channel proteins, however, cannot excluded.  相似文献   

2.
Water and solute relations of young roots of Phaseolus coccineus have been measured using the root pressure probe. Biphasic root pressure relaxations were obtained when roots were treated with solutions containing different osmotic test solutes. From the relaxations, the hydraulic conductivity (Lpr), the permeability coefficient (Psr), and the reflection coefficient (σsr) of the roots could be evaluated. Lpr was 1.8 to 8.4 . 10?8 m . s?1 . MPa?1 and Psr (in 10?10 m . s?1): methanol, 27–62; ethanol, 44–73; urea, 5–11; mannitol, 1.5; KCl, 7.1–9.2; NaCl, 2.1; NaNO3, 3.7. The hydraulic conductivity was similar when using osmotic and hydrostatic pressure gradients as driving forces. The hydraulic conductivity of individual root cortex cells (Lp) was by two orders of magnitude larger than Lpr (Lp = 0.3 to 4.7 . 10?6 m . s?1 . MPa?1) which indicated a predominant cell-to-cell rather than an apoplasmic transport of water in the Phaseolus root. Except for distances shorter than 20 mm from the root apex, the hydraulic resistance of the roots was limited by the radial movement of water across the root cylinder and not by the hydraulic resistance within the xylem. Reflection coefficients were low: methanol: 0.16–0.34; ethanol: 0.15–0.47; urea: 0.41–0.51; mannitol: 0.68; KCl: 0.43–0.54; NaCl: 0.59; NaNO3: 0.54. The transport coefficients (Lpr, Psr, σsr) have been critically examined for influences of unstirred layers and active transport. The low σsr suggests that the common treatment of the root as a rather perfect osmometer (σsr = 1) analogous to plant cells should be treated cautiously. The reasons for the low σsr and the possible implications of the absolute values of the transport parameters for the absorption of water and nutrients are discussed.  相似文献   

3.
The reflection coefficient, σj, which measures the coupling between the jth solute and water transport across a semipermeable membrane, varies between 0 and 1.0. Values of σj significantly less than 1.0 provide irreversible thermodynamic proof that there is coupling between the transport of solute and solvent and thus that they share a common pathway. We have developed an improved method for measuring σ and have used it to determine that σethylene glycol = 0.71 ± 0.03 and σurea = 0.65 ± 0.03, in agreement with many, but not all, previous determinations. Since both of these values are significantly lower than 1.0, they show that there is a common ethylene glycol/water pathway and a common urea/water pathway. Addition of first one and then two methyl groups to urea increases σ to 0.89 ± 0.04 for methylurea and 0.98 ± 0.4 for 1,3-dimethylurea, consistent with passage through an aqueous pore with a sharp cutoff in the 6–7 Å region.  相似文献   

4.
5.
Measurement of the transport parameters that govern the passage of urea and amides across the red cell membrane leads to important questions about transport of water. It had initially been thought that small protein channels, permeable to water and small solutes, traversed the membrane (see Solomon, 1987). Recently, however, very strong evidence has been presented that the 28 kDa protein, CHIP28, found in the red cell membrane, is the locus of the water channel (see Agre et al., 1993). CHIP28 transports water very rapidly but does not transport small nonelectrolytes such as urea. The irreversible thermodynamic parameter, σ i , the reflection coefficient, is a measure of the relationship between the permeability of the solute and that of water. If a solute permeates by dissolution in the membrane, σ i = 1.0; if it permeates by passage through an aqueous channel, σ i < 1.0. For urea, Goldstein and Solomon (1960) found that σurea= 0.62 ± 0.03 which meant that urea crosses the red cell membrane in a water-filled channel. This result and many subsequent observations that showed that σurea < 1.0 are at variance with the observation that CHIP28 is impermeable to urea. In view of this problem, we have made a new series of measurements of σ i for urea and other small solutes by a different method, which obviates many of the criticisms Macey and Karan (1993) have made of our earlier method. The new method (Chen et al., 1988), which relies upon fluorescence of the intracellular dye, fluorescein sulfonate, leads to the corrected value, σurea,corr= 0.64 ± 0.03 for ghosts, in good agreement with earlier data for red cells. Thus, the conclusion on irreversible thermodynamic and other grounds that urea and water share a common channel is in disagreement with the view that CHIP28 provides the sole channel for water entrance into the cell. Received: 6 February 1996/Revised: 20 May 1996  相似文献   

6.
Annett Hertel  Ernst Steudle 《Planta》1997,202(3):324-335
Using the cell pressure probe, the effects of temperature on hydraulic conductivity (Lp; osmotic water permeability), solute permeability (permeability coefficient, Ps), and reflection coefficients (σs) were measured on internodes of Chara corallina, Klein ex Willd., em R.D.W.. For the first time, complete sets of transport coefficients were obtained in the range between 10 and 35 °C which provided evidence about pathways of water and solutes as they move across the plasma membrane (water channel and bilayer arrays). Test solutes used to check for the selectivity of water channels were monohydric alcohols of different molecular size and shape (ethanol, n-propanol, iso-propanol, and tert-butanol) and heavy water (HDO). Within the limits of accuracy, Q10 values for Lp and for the diffusive water permeability (Pd) were identical (Q10 for Lp = 1.29 ± 0.17 (± SD; n = 15 cells) and Q10 for Pd = 1.25 ± 0.16 (n = 5 cells)). The Q10 values were equivalent to activation energies of Ea = 16.8 ± 6.4 and 16.6 ± 10.0 kJ · mol−1, respectively, which is similar to that of self-diffusion or of viscous flow of water. The Q10 values and activation energies for Ps of the alcohols were significantly larger (ethanol: Q10 = 1.68 ± 0.16, Ea = 37.1 ± 5.9 kJ · mol−1; n-propanol: Q10 =  1.75 ± 0.40, Ea = 43.1 ± 15.3 kJ · mol−1; iso-propanol: Q10 = 2.12 ± 0.42, Ea =  52.2 ± 14.6 kJ · mol−1; tert-butanol: Q10 = 2.13 ± 0.56, Ea = 51.6 ± 17.1 kJ · mol−1; ±SD; n = 5 to 6 cells). Effects of temperature on reflection coefficients were most pronounced. With increasing temperature, σs values of the alcohols decreased and those of HDO increased. The data indicate that water and solutes use different pathways when crossing the membrane. Ordinary and isotopic water use water channels and the other test solutes use the bilayer array (composite transport model of membrane). Changes in σs values with temperature were found to be a sensitive measure for the open/closed state of water channels. The decrease of σs with temperature was theoretically predicted from the temperature dependence of Ps and Lp. Differences between predicted and measured values of σs allowed estimation of the bypass flow (slippage) of solutes through water channels which did not completely exclude test solutes. The permeability of channels depended on the structure and size of test solutes. It is concluded that water channels are much less selective than is usually thought. Since water channels represent single-file or no-pass pores, solutes drag along considerable amounts of water as they diffuse across channels. This results in low overall values of σs. The σs of HDO was extremely low. Its response to temperature was opposite to that for the σs of the alcohols. This suggested a stronger effect of temperature on the hydraulic (osmotic) than on the diffusive water flow across individual water channels, i.e. a differential sensitivity of different mechanisms to temperature. Received: 10 October 1996 / Accepted: 2 December 1996  相似文献   

7.
K Okita  A Teramoto  H Fujita 《Biopolymers》1970,9(6):717-738
A new procedure for evaluating u and σ characterizing σ-helix-forming polypeptides in solution was derived from Nagai's theory for the helix–coil transition of such polymers. Here u is the activity for helix formation from random coil, and σ is the helix initiation parameter. The necessary data are the helical content fN at fixed solvent and temperature as a function of N, where N is the degree of polymerization of the polypeptide sample. Such data were obtained from ORD measurements on a number of fractionated samples of poly-N5-(3-hydroxypropyl)-L -glutamine (PHPG) in mixtures of water and methanol covering the complete range of composition and at various termperatures (5–40°C). When analyzed in terms of the proposed procedure, they yielded values of σ which were in the range (3.2 ± 0.6) × 10?4, substantially independent of solvent composition and temperature. These values were much larger than those obtained recently for σ of poly(β-benzyl-L -aspartate) in m-cresol and in a mixture of chloroform and DCA. The data for [η] and s0 (limiting sedimentation coefficient) as functions of molecular weight indicated that the molecular shape of PHPG in pure methanol is essentially rodlike, whereas that in pure water is not entirely randomly coiled but rather may be regarded as an interrupted helix. These indications were consistent with the results from ORD measurements. When plotted against the corresponding values of fN, the values of [η] and [s0] for PHPG in mixtures of water and methanol of various compositions and temperatures formed smooth composite curves, and we attributed these phenomena to the fact that σ of PHPG was nearly constant under these solvent conditions. Here [s0] stands for a reduced limiting sedimentation coefficient which is equal to the inverse friction factor of the solute molecule.  相似文献   

8.
F.W. Kleinhans 《Cryobiology》1998,37(4):271-289
The analysis of experiments for the purpose of determining cell membrane permeability parameters is often done using the Kedem–Katchalsky (KK) formalism (1958). In this formalism, three parameters, the hydraulic conductivity (Lp), the solute permeability (Ps), and a reflection coefficient (ς), are used to characterize the membrane. Sigma was introduced to characterize flux interactions when water and solute (cryoprotectant) cross the membrane through a common channel. However, the recent discovery and characterization of water channels (aquaporins) in biological membranes reveals that aquaporins are highly selective for water and do not typically cotransport cryoprotectants. In this circumstance, sigma is a superfluous parameter, as pointed out by Kedem and Katchalsky. When sigma is unneeded, a two-parameter model (2P) utilizing onlyLpandPsis sufficient, simpler to implement, and less prone to spurious results. In this paper we demonstrate that the 2P and KK formalism yield essentially the same result (LpandPs) when cotransporting channels are absent. This demonstration is accomplished using simulation techniques to compare the transport response of a model cell using a KK or 2P formalism. Sigma is often misunderstood, even when its use is appropriate. It is discussed extensively here and several simulations are used to illustrate and clarify its meaning. We also discuss the phenomenological nature of transport parameters in many experiments, especially when both bilayer and channel transport are present.  相似文献   

9.
This study was undertaken to quantify the effect of aluminum and calcium on membrane permeability. The influence of Ca2+ (0.2-3.7 millimolar) and Al3+ (0-3.7 millimolar) on the permeability of root cortical cells of Quercus rubra was measured using three nonelectrolytes (urea, methyl urea, and ethyl urea) as permeators of progressively increasing lipid solubility. Water permeability was also measured. Al3+ (a) increased membrane permeability to the nonelectrolytes, (b) decreased the membrane's partiality for lipid permeators, and (c) decreased membrane permeability to water. Ca2+ had effects on permeability that were opposite to those of Al3+. When Al3+ and Ca2+ were tested in combination, these opposite effects counteracted each other. The results suggest that Al3+ altered the architecture of membrane lipids.  相似文献   

10.
The effect of 0.5 ppm ozone for 0.5-1 hr on plant cell membrane permeability was ascertained. Permeabilities to both water and solutes were estimated by measuring leaf disc weight changes and following tritiated water and 86Rb fluxes. Measurements were made immediately after ozone exposure and 24 hr after exposure. The reflection coefficient, σ, an index of solute permeability, decreased in ozone-treated primary leaves of pinto bean (Phaseolus vulgaris). The latter indicates an increase in membrane solute permeability or internal solute leakage. Water and THO flux estimates both indicated a decrease in membrane permeability to water; both the hydraulic conductivity (Lp) and the water diffusional coefficient (LD) apparently decreased, an anomaly which is discussed. These data indicate that ozone has a direct effect on membrane function by altering permeability characteristics. We assume from these data that cell membranes are primary target sites for ozone injury.  相似文献   

11.
J. Schönherr 《Planta》1976,128(2):113-126
Summary The upper astomatous cuticle of Citrus aurantium L. leaves was isolated enzymatically or chemically, extracted with lipid solvents and used for the determination of water diffusion (P d ) and osmotic water permeability (P f ). The water permeability was strongly dependent on the pH value and the cations of the buffer solutions. In presence of monovalent alkali metal ions P d increased almost five fold between pH 3 and 11. The shape of the plot P d vs. pH suggests the presence of 3 different dissociable groups fixed to the membrane matrix. They are tentatively identified as two carboxyl groups dissociating between pH 3 to 6 and 6 to 9, respectively, and as phenolic hydroxyl groups dissociating above pH 9. The carboxyl group dissociating between pH 6 and 9 discriminated between alkali metal ions according to their ionic radius. Water permeability was lowest in the Li+ from and increased in the order Li++++. The water permeability of membranes in Ca2+ form was only slightly higher than that of membranes in H+ form and little dependent on pH. The energy of activation which amounted to 13 kcal mol–1 was constant over the temperature range of 5 to 40°C and pH independent. Since P f was greater than P d it was concluded that the cutin matrix contained polar pores and that water transport caused by a chemical potential gradient was both by diffusion and by viscous flow. The porous nature of the membranes was also confirmed by the fact that they are permselective according to size of the permeating molecule. Using the empirical equations of Paganelli and Solomon (1957) and Nevis (1958) the equivalent radius of the pores was estimated to be 0.46 and 0.45 nm, respectively. This estimate is in good agreement with the observations that (a) [14C]urea (molecular radius r s =0.264 nm) and [3H]glucose (r s =0.444 nm) penetrated the membranes and (b) the reflection coefficient was equal to one for raffinose (r s =0.654 nm) and sucrose (r s =0.555 nm) but 0.95 for glucose and 0.78 for urea. Both, the reflection coefficient and the pore radius estimates were pH independent, hence the increase in water permeability with increasing pH was due to an increase in the number of pores per unit area (1 cm2) from 5x1010 at pH 3 to 15.8x1010 at pH 9.Abbreviations THO tritiated water - HEPES N-2-hydroxyethylpiperazine-N-2-ethane sulfonic acid - MES (N-morpholino) ethane sulphic acid - SADH succinic acid 2,2-dimethyl hydrazide  相似文献   

12.
Summary A mathematical analysis of the relationship between the flux of water, f H2O, and the flux of potassium f K, to the xylem of exuding root systems of Ricinus communis, is presented. Previous analyses (Baker and Weatherley, 1969; Minchin and Baker, 1969) have indicated the presence of a water dependent and a water independent f K both of which vary with the external concentration of potassium, Cm, supplied as potassium nitrate.The present analysis reveals that whereas at Cm values<1 mM both components of f K contribute ions to the osmotically active solutions within the osmotic barrier, at Cm values>1 mM only the water dependent f K is responsible for the osmotic work. This suggests that the ions are released within different regions of the stele. It is proposed that at cm values<1 mM both components are released from the inner stelar tissues whilst at higher Cm values the water dependent f K is released from the outer stelar tissues. This requires that the solute permeability of the plasmalemma of the outer stelar tissues increases markedly at or about Cm values of 1 mM.It is postulated that the required separation of the two f K components within the stelar symplasm at Cm values>1 mM is due to the water independent f K being in a bound state, possibly being transported along a chain of binding sites whilst the water dependent f K is in a free state within the aqueous phase of the cytoplasm.  相似文献   

13.
Osmotic responses of maize roots   总被引:16,自引:0,他引:16  
Water and solute relations of excised seminal roots of young maize (Zea mays L) plants, have been measured using the root pressure probe. Upon addition of osmotic solutes to the root medium, biphasic root pressure relaxations were obtained as theoretically expected. The relaxations yielded the hydraulic conductivity Lp r) the permeability coefficient (P sr), and the reflection coefficient (σ sr) of the root. Values of Lp r in these experiments were by nearly an order of magnitude smaller than Lp r values obtained from experiments where hydrostatic pressure gradients were used to induce water flows. The value of P sr was determined for nine different osmotica (electrolytes and nonelectrolytes) which resulted in rather variable values (0.1·10-8–1.7·10-8m·s-1). The reflection coefficient σ sr of the same solutes ranged between 0.3 and 0.6, i.e. σ sr was low even for solutes for which cell membranes exhibit a σ s≈1. Deviations from the theoretically expected biphasic responses occured which may have reflected changes of either P sr or of active pumping induced by the osmotic change. The absolute values of Lp r, P sr, and σ sr have been critically examined for an underestimation by unstirred layer effecs. The data indicate a considerable apoplasmic component for radial movement of water in the presence of hydrostatic gradients and also some solute flow byppassing root protoplasts. In the presence of osmotic gradients, however, there was a substantial cell-to-cell transport of water. Cutting experiments demonstrated that the hydraulic resistance for the longitudinal movement of water was much smaller than for radial transport except for the apical ends of the segments (length=5 to 20 mm). The differences in Lp r as well as the low σ sr values suggest that the simple osmometer model of the root with a single osmotic barrier exhibiting nearly semipermeable properties should be extended for a composite membrane model with hydraulic and osmotic barriers arranged in series and in parallel.  相似文献   

14.
Summary Comparison of transintegumental membrane permeability and partition coefficients of selected nonelectrolytes was attempted to correlate the parameters of lipid solubility, hydrophilicity, and membrane permeation in male and female schistosomes (parasites of the portal venous tributaries of man). Surface permation (measured by the triple isotope technique) and octanol/water partition coefficients were determined for 17 compounds (acetamide, aminopyrine, antipyrine, benzyl alcohol, butanol, caffeine, ethanol, ethylene glycol, glycerol, inosine, mannitol, methanol, polyethylene glycol, propylene glycol, sucrose, thiourea, and urea).Linear regression analyses comparing the logarithm of the partition coefficient to transintegumental uptakes indicate a positive correlation in both sexes:R=0.76 (P<0.001) for males, andR=0.77 (P<0.001) for females. Similarly, linear regression analyses comparing hydrogen bond number with the logarithm of tissue uptake index demonstrate a high (negative) correlation in both males (R=–0.85,P<0.001) and females (R=–0.90,P<0.001). The male and female schistosomes showed no statistically significant differences in correlation of these parameters. Surface permeation was the same in male and female schistosomes, suggesting that male-female variations in integumental uptake rates previously observed may be restricted to metabolites which enter by way of a selective carrier system.  相似文献   

15.
In general, erythrocytes are highly permeable to water, urea and glycerol. However, expression of aquaporin isoforms in erythrocytes appears to be species characteristic. In the present study, human (hRBC) and bovine (bRBC) erythrocytes were chosen for comparative studies due to their significant difference in membrane glycerol permeability.Osmotic water permeability (Pf) at 23 °C was (2.89 ± 0.37) × 10−2 and (5.12 ± 0.61) × 10−2 cm s−1 for human and bovine cells, respectively, with similar activation energies for water transport. Glycerol permeability (Pgly) for human ((1.37 ± 0.26) × 10−5 cm s−1) differed in three orders of magnitude from bovine erythrocytes ((5.82 ± 0.37) × 10−8 cm s−1) that also showed higher activation energy for glycerol transport. When compared to human, bovine erythrocytes showed a similar expression pattern of AQP1 glycosylated forms on immunoblot analysis, though in slight higher levels, which could be correlated with the 1.5-fold larger Pf found. However, AQP3 expression was not detectable. Immunofluorescence analysis confirmed the absence of AQP3 expression in bovine erythrocyte membranes.In conclusion, lack of AQP3 in bovine erythrocytes points to the lipid pathway as responsible for glycerol permeation and explains the low glycerol permeability and high Ea for transport observed in ruminants.  相似文献   

16.
Abstract An artificial osmotic cell has been constructed using reverse osmosis membranes. The cell consisted of a thin film of an osmotic solution (thickness: 100 to 200 μm) containing a non-permeating solute and was bounded between the membrane and the front plate of a pressure transducer which continuously recorded cell turgor. The membrane was supported by metal grids to withstand positive and negative pressures (P). At maximum, negative pressures of up to –0.7 MPa (absolute) could be created within the film on short-term and pressures of up to –0.3 MPa could be maintained without cavitation for several hours. As with living plant cells, the application of osmotic solutions of a non-permeating solute resulted in monophasic relaxations of turgor pressure from which the hydraulic conductivity of the membrane (Lp) and the elastic modulus of the cell (?) could be estimated. The application of solutions with permeating solutes resulted in biphasic pressure relaxation curves (as for living cells) from which the permeability (Ps) and reflection (σs) coefficients could be evaluated for the given membrane. Lp, Ps, and σs were independent of P and did not change upon transition from the positive to the negative range of pressure. It is concluded that the artificial cell could be used to simulate certain transport properties of living cells and to study phenomena of negative pressure as they occur in the xylem and, perhaps, also in living cells of higher plants.  相似文献   

17.
Intra- and transcellular water movements in plants are regulated by the water permeability of the plasma membrane (PM) and vacuolar membrane (VM) in plant cells. In the present study, we investigated the osmotic water permeability of both PM (P f1) and VM (P f2), as well as the bulk osmotic water permeability of a protoplast (P f(bulk)) isolated from radish (Raphanus sativus) roots. The values of P f(bulk) and P f2 were determined from the swelling/shrinking rate of protoplasts and isolated vacuoles under hypo- or hypertonic conditions. In order to minimize the effect of unstirred layer, we monitored dropping or rising protoplasts (vacuoles) in sorbitol solutions as they swelled or shrunk. P f1 was calculated from P f(bulk) and P f2 by using the ‘three-compartment model’, which describes the theoretical relationship between P f1, P f2 and P f(bulk) (Kuwagata and Murai-Hatano in J Plant Res, 2007). The time-dependent changes in the volume of protoplasts and isolated vacuoles fitted well to the theoretical curves, and solute permeation of PM and VM was able to be neglected for measuring the osmotic water permeability. High osmotic water permeability of more than 500 μm s−1, indicating high activity of aquaporins (water channels), was observed in both PM and VM in radish root cells. This method has the advantage that P f1 and P f2 can be measured accurately in individual higher plant cells. Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorized users. It includes four appendices, four tables and two figures. Mari Murai-Hatano and Tsuneo Kuwagata contributed equally to the paper. An erratum to this article is available at .  相似文献   

18.
19.
Methanogenesis from ethanol was studied in batch cocultures of a proton-reducing acetogenic Desulfovibrio sp. together with one of eight methanogenic bacteria representing five genera. A mathematical model of mutualistic cocultures predicts an equalisation in the specific growth rates of the two species which defines a specific growth rate for the coculture. At non-limiting ethanol concentrations the model predicts that the specific growth rate of the coculture is dependent upon the K s (H2) of the methanogen and its maximum specific growth rate in axenic culture when utilising H2 as the energy source. We demonstrate experimentally that with methanogens known to have similar K s (H2) values, the specific growth rates of methanogenic mutualistic cocultures are dependent upon the maximum specific growth rates of the methanogens.  相似文献   

20.
Summary Osmotic responses of slices of dogfish rectal gland to hypotonic (urea-free) and hypertonic media were studied. Transfer of tissue from isotonic (890 mosM) to hypotonic (550 mosM) saline produced an osmotic swelling associated with a slow net uptake of cell K+ (and Cl) and a slow, two-component efflux of urea. Media made hypertonic (1180 mosM) by addition of urea or mannitol produced osmotic shrinkage with a net loss of KCl. The cell osmotic responses in hypotonic media were lower than predicted for an ideal osmometer. No volume regulatory responses were seen subsequent to the initial osmotic effects. The cation influx in hypotonic media lacked specificity: in the presence of 0.5 mM ouabain or in K+-free media a net influx of Na+ was found. At steady state, the cell membrane potential evaluated from the Nernst potentials of K+ and triphenylmethyl phosphonium+, was independent of medium tonicity, suggesting the membrane potential as a determinant in the cellular osmotic response. Zero-time86Rb+ fluxes were measured:86Rb+ influx was not affected by hypotonicity, implying an unchanged operation of the Na+–K+-ATPase. On the other hand,86Rb+ efflux was significantly reduced at hypotonicity; this effect was transient, the efflux returning to the control value once the new steady state of cell volume had been reached. A controlled efflux system is therefore involved in the cell osmotic response. The absence of the volume regulatory phenomenon suggests that the cells are not equipped with a volume-sensing mechanism.Abbreviations and symbols DW dry weight - E extracellular (polyethylene glycol) space - E Nernst potential - H2Oe H2Oi tissue water, extra- and intracellular - TPMP + triphenyl methyl phosphonium salt - WW wet weight  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