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 0 毫秒
1.
The kinetics of Na+/d-glucose cotransport (SGLT) were reevaluated in rabbit renal brush border membrane vesicles isolated from the whole kidney cortex using a fast-sampling, rapid-filtration apparatus (FSRFA, US patent #5,330,717) for uptake measurements. Our results confirm SGLT heterogeneity in this preparation, and both high (HAG) and low (LAG) affinity glucose transport pathways can be separated over the 15–30°C range of temperatures. It is further shown that: (i) Na+ is an essential activator of both HAG and LAG; (ii) similar energies of activation can be estimated from the linear Arrhenius plots constructed from the V max data of HAG and LAG, thus suggesting that the lipid composition and/or the physical state of the membrane do not affect much the functioning of SGLT; (iii) similar V max values are observed for glucose and galactose transport through HAG and LAG, thus demonstrating that the two substrates share the same carrier agencies; and (iv) phlorizin inhibits both HAG and LAG competitively and with equal potency (K i = 15 μm). Individually, these data do not allow us to resolve conclusively whether the kinetic heterogeneity of SGLT results from the expression in the proximal tubule of either two independent transporters (rSGLT1 and rSGLT2) or from a unique transporter (rSGLT1) showing allosteric kinetics. Altogether and compared to the kinetic characteristics of the cloned SGLT1 and SGLT2 systems, they do point to a number of inconsistencies that lead us to conclude the latter possibility, although it is recognized that the two alternatives are not mutually exclusive. It is further suggested, from the differences in the K m values of HAG transport in the kidney as compared to the small intestine and SGLT1 cRNA-injected oocytes, that renal SGLT1 activity is somehow modulated, maybe through heteroassociation with (a) regulatory subunit(s) that might also contribute quite significantly to sugar transport heterogeneity in the kidney proximal tubule. Received: 25 October 1995/Revised: 10 June 1996  相似文献   

2.
In order to define potential interaction sites of SGLT1 with the transport inhibitor phlorizin, mutagenesis studies were performed in a hydrophobic region of loop 13 (aa 604–610), located extracellularly, close to the C-terminus. COS 7 cells were transiently transfected with the mutants and the kinetic parameters of α-methyl-d-glucopyranoside (AMG) uptake into the cells were investigated. Replacement of the respective amino acids with lysine reduced the maximal uptake rate: Y604K showed 2.2%, L606K 48.4%, F607K 15.1%, C608K 13.1%, G609K 14.1%, and L610K 17.2% of control. In all mutants the apparent K i for phlorizin increased at least by a factor of 5 compared to the wild-type K i of 4.6 ± 0.7 μmol/l; most striking changes were observed for Y604K (K i = 75.3 ± 19.0 μmol/l) and C608K (K i = 83.6 ± 13.9 μmol/l). Replacement of these amino acids with a nonpolar amino acid instead of lysine such as in Y604F, Y604G and C608A showed markedly higher affinities for phlorizin. In cells expressing the mutants the apparent affinity of AMG uptake for the sugar was not statistically different from that of the wild type (K m = 0.8 ± 0.2 mmol/l). These studies suggest that the region between amino acids 604 and 610 is involved in the interaction between SGLT1 and phlorizin, probably by providing a hydrophobic pocket for one of the aromatic rings of the aglucone moiety of the glycoside. Received: 29 March 2001/Revised: 15 June 2001  相似文献   

3.
Using cotransporters as drug delivery vehicles is a topic of continuing interest. We examined glucose derivatives containing conjugated aromatic rings using two isoforms of the Na+/glucose cotransporter: human SGLT1 (hSGLT1) and pig SGLT3 (pSGLT3, SAAT1). Our studies indicate that there is similarity between SGLT1 and SGLT3 in the overall architecture of the vestibule leading to the sugar-binding site but differences in translocation pathway interactions. Indican was transported by hSGLT1 with higher affinity (K0.5 0.06 mm) and 2-naphthylglucose with lower affinity (K0.5 0.5 mm) than α-methyl-d-glucopyranoside (αMDG, 0.2 mm). Both were poorly transported (maximal velocities, I max , 14% and 8% of αMDG). Other compounds were inhibitors (K i s 1–13 mm). In pSGLT3, indican and 2-naphthylglucose were transported with higher affinity than αMDG (K0.5s 0.9, 0.2 and 2.5 mm and relative I max s of 80, 25 and 100%). Phenylglucose and arbutin were transported with higher I max s (130 and 120%) and comparable K0.5s (8 and 1 mm). Increased affinity of indican relative to αMDG suggests that nitrogen in the pyrrole ring is favorable in both transporters. Higher affinity of 2-naphthylglucose for pSGLT3 than hSGLT1 suggests more extensive hydrophobic/aromatic interaction in pSGLT3 than in hSGLT1. Our results indicate that bulky hydrophobic glucosides can be transported by hSGLT1 and pSGLT3, and discrimination between them is based on steric factors and requirements for H-bonding. This provides information for design of glycosides with potential therapeutic value. Received: 18 February 2000/Revised: 13 April 2000  相似文献   

4.
The rabbit Na+/glucose cotransporter (SGLT1) exhibits a presteady-state current after step changes in membrane voltage in the absence of sugar. These currents reflect voltage-dependent processes involved in cotransport, and provide insight on the partial reactions of the transport cycle. SGLT1 presteady-state currents were studied as a function of external Na+, membrane voltage V m , phlorizin and temperature. Step changes in membrane voltage—from the holding V h to test values, elicited transient currents that rose rapidly to a peak (at 3–4 msec), before decaying to the steady state, with time constants τ≈4–20 msec, and were blocked by phlorizin (K i ≈30 μm). The total charge Q was equal for the application of the voltage pulse and the subsequent removal, and was a function of V m . The Q-V curves obeyed the Boltzmann relation: the maximal charge Q max was 4–120 nC; V 0.5, the voltage for 50% Q max was −5 to +30 mV; and z, the apparent valence of the moveable charge, was 1. Q max and z were independent of V h (between 0 and −100 mV) and temperature (20–30°C), while increasing temperature shifted V 0.5 towards more negative values. Decreasing [Na+] o decreased Q max, and shifted V 0.5 to more negative voltages 9by −100 mV per 10-fold decrease in [Na+] o ). The time constant τ was voltage dependent: the τ-V relations were bell-shaped, with maximal τmax 8–20 msec. Decreasing [Na+] o decreased τmax, and shifted the τ-V curves towards more negative voltages. Increasing temperature also shifted the τ-V curves, but did not affect τmax. The maximum temperature coefficient Q 10 for τ was 3–4, and corresponds to an activation energy of 25 kcal/mole. Simulations of a 6-state ordered kinetic model for rabbit Na+/glucose cotransport indicate that charge-movements are due to Na+-binding/dissociation and a conformational change of the empty transporter. The model predicts that (i) transient currents rise to a peak before decay to steady-state; (ii) the τ-V relations are bell-shaped, and shift towards more negative voltages as [Na+] o is reduced; (iii) τmax is decreased with decreasing [Na+] o ; and (iv) the Q-V relations are shifted towards negative voltages as [Na+] o is reduced. In general, the kinetic properties of the presteady-state currents are qualitatively predicted by the model. Received: 12 August 1996/Revised: 30 September 1996  相似文献   

5.
The present study deals with the phenotypic adaptation of tonoplast fluidity in the CAM plant Kalancho? daigremontiana to changes in growth temperature. Tonoplast fluidity was characterized by measuring fluorescence depolarization in membranes labeled with fluorescent fatty acid analogues and by following formation of eximeres in membranes labeled by eximere-forming fluorophores. With both techniques it was found that exposure of the plants to higher growth temperature compared with the control decreased the fluidity of the tonoplast while exposure to lower growth temperature caused the opposite. Three hours of high temperature treatment (raised from 25°C to 35°C; ``heat shock') were sufficient to decrease the tonoplast fluidity to roughly the same extent as growth under high temperature for 30 days. The phenotypic response of tonoplast fluidity to changes in growth temperature was found only in the complete membrane, not however in the lipid matrix deprived of the membrane proteins. Heat treatments of the plants decreased the lipid/protein ratio while exposure to low temperature (for 30 days) increased it. Heat treatments led to a decrease in the percentage of linolenic acid (C18:3) and linoleic acid (C18:2), heat shock and low temperature treatments induced an increase in the percentage of linoleic acid (C18:3), with concomitant decrease in the percentage of linoleic acid (C18:2). However, in the case of heat shock, increase in linolenic acid concerned mainly monogalactosyldiacylglycerol, while with low temperature treatment linoleic acid increased in phosphatidylcholine. Both treatment of the plants with high and low temperature led to a slight decrease in the contribution of phosphatidylcholine and phosphoethanolamine to the total phospholipid content of the tonoplast. High-temperature treatment of the plants not only decreased the phospholipid/protein ratio in the tonoplast, but also led to the occurrence of a 35 kDa polypeptide in the tonoplast which cross-reacted with an antiserum against the tonoplast H+-ATPase holoenzyme. The important role of membrane proteins in bringing about the phenotypic rigidization of the tonoplast was mimicked by reconstitution experiments showing that incorporation of the proteins isolated from the tonoplast into phosphatidylcholine vesicles decreased the fluidity of this membrane system. As to be expected from the analyses in the natural membrane, the degree of this effect depended on the phospholipid/protein ratio. Received: 4 March 1998/Revised: 28 July 1998  相似文献   

6.
7.
Rate and equilibrium measurements of ryanodine binding to terminal cysternae fractions of heavy sarcoplasmic reticulum vesicles demonstrate that its activation by high concentrations of monovalent salts is based on neither elevated osmolarity nor ionic strength. The effect of the ions specifically depends on their chemical nature following the Hofmeister ion series for cations (Li+ < NH+ 4 < K∼ Cs+≤ Na+) and anions (gluconate < Cl < NO3 ∼ ClO4 ∼ SCN) respectively, indicating that both are involved in the formation of the salt-protein complex that can react with ryanodine. Activation by rising salt concentrations exhibits saturation kinetics with different dissociation constants (25–11 m) and different degrees of cooperativity (n= 1.5–4.0) for the respective salts. Maximal second order binding rates between 40,000 and 80,000 (m −1· sec−1) were obtained for chlorides and nitrates of 1a group alkali ions with the exception of lithium supporting only rates of maximally 10,000 (M−1· sec−1). The nitrogen bases, NH+ 4 and Tris+, in combination with chloride or nitrate, behave divergently. High maximal binding rates were achieved only with NH4NO3. The dissociation constants for the ryanodine–protein complexes obtained by measurements at equilibrium proved to depend differently on salt concentration, yet, converging to 1–3 nm for the applied salts at saturating concentrations. The salts do not affect dissociation of the ryanodine protein complex proving that the effect of salts on the protein's affinity for ryanodine is determined by their effect on the on-rate of ryanodine binding. ATP and its analogues modify salt action resulting in elevated maximal binding rates and reduction or abolition of binding cooperativity. Linear relations have been obtained by comparing the rates of ryanodine binding at different salt concentrations with the rates or the initial amplitudes (15 sec) of salt induced calcium release from actively loaded heavy vesicles indicating that the various salts promote specifically and concentration dependently channel opening and its reaction with ryanodine. Received: 9 February 1998/Revised: 24 April 1998  相似文献   

8.
Analysis of the mechanistic basis by which sodium-coupled transport systems respond to changes in membrane potential is inherently complex. Algebraic expressions for the primary kinetic parameters (K m and V max ) consist of multiple terms that encompass most rate constants in the transport cycle. Even for a relatively simple cotransport system such as the Na+/alanine cotransporter in LLC-PK1 cells (1:1 Na+ to substrate coupling, and an ordered binding sequence), the algebraic expressions for K m for either substrate includes ten of the twelve rate constants necessary for modeling the full transport cycle. We show here that the expression of K m of the first-bound substrate (Na+) simplifies markedly if the second-bound substrate (alanine) is held at a low concentration so that its' binding becomes the rate limiting step. Under these conditions, the expression for the K Na m includes rate constants for only two steps in the full cycle: (i) binding/dissociation of Na+, and (ii) conformational `translocation' of the substrate-free protein. The influence of imposed changes in membrane potential on the apparent K Na m for the LLC-PK1 alanine cotransporter at low alanine thus provides insight to potential dependence at these sites. The data show no potential dependence for K Na m at 5 μm alanine, despite marked potential dependence at 2 mm alanine when the full algebraic expression applies. The results suggest that neither translocation of the substrate-free form of the transporter nor binding/dissociation of extracellular sodium are potential dependent events for this transport system. Received: 10 April 1998/Revised: 6 July 1998  相似文献   

9.
The general purpose of this theoretical work is to contribute to understand the physiological role of the electrogenic properties of the sodium pump, by studying a dynamic model that integrates diverse processes of ionic and water transport across the plasma membrane. For this purpose, we employ a mathematical model that describes the rate of change of the intracellular concentrations of Na+, K+ and Cl, of the cell volume, and of the plasma membrane potential (V m ). We consider the case of a nonexcitable, nonpolarized cell expressing the sodium pump; Na+, K+, Cl and water channels, and cotransporters of KCl and NaCl in its plasma membrane. We particularly analyze here the conditions under which the physiological V m can be generated in a predominantly electrogenic fashion, as a result of the activity of the sodium pump. A major conclusion of this study is that, for the cell model considered, a low potassium permeability is not a sufficient condition for a predominantly electrogenic generation of the V m by the sodium pump. The presence of an electroneutral exchange of Na+ and K+ represents a necessary additional requirement. Received: 8 September 1999/Revised: 21 March 2000  相似文献   

10.
The emergence of jawed vertebrates was predicated on the appearance of several innovations, including tooth formation. The development of teeth requires the participation of several specialized genes, in particular, those necessary for the formation of hard tissues—dentin, enamel, and cementum. Some vertebrates, most conspicuously birds, secondarily lost the tooth-forming ability. To determine the fate of some of the tooth-forming genes in the birds, we tested a domestic fowl cDNA library for the expression of the dentin matrix protein 1 (DMP1) gene. The library was prepared from the poly(A+) RNA isolated from the jaws of 11- to 13-day-old embryos and the testing was carried out by the polymerase chain reaction with degenerate primers designed on the basis of the available mammalian and reptile sequences. A chicken homologue of the DMP1 gene identified by this approach was shown to be expressed in the jaws and long bones, the same two tissues as in mammals. The chicken DMP1 gene has an exon/intron organization similar to that of its mammalian and reptile counterparts. The chicken gene contains three short highly conserved segments, the rest of the gene being poorly alignable or not alignable with its mammalian or reptilian homologues. The distribution of similarities and dissimilarities along the gene is indicative of a mode of evolution in which only short segments are kept constant, while the rest of the gene is relatively free to vary as long as the proportion of certain amino acid residues is retained in the encoded polypeptide. The DMP1 gene may have been retained in birds because of its involvement in bone formation. Received: 5 April 1999 / Accepted: 9 August 1999  相似文献   

11.
To investigate the causes and functional significance of rapid sex-determining protein evolution we compared three Caenorhabditis elegans genes encoding members of the protein phosphatase 2C (PP2C) family with their orthologs from another Caenorhabditis species (strain CB5161). One of the genes encodes FEM-2, a sex-determining protein, while the others have no known sex-determining role. FEM-2's PP2C domain was found to be more diverged than the other PP2C domains, supporting the notion that sex-determining proteins are subjected to selective pressures that allow for or cause rapid divergence. Comparison of the positions of amino acid substitutions in FEM-2 with a solved three-dimensional structure suggests that the catalytic face of the protein is highly conserved among C. elegans, CB5161, and another closely related species C. briggsae. However, the non-conserved regions of FEM-2 cannot be said to lack functional importance, since fem-2 transgenes from the other species were unable to rescue the germ-line defect caused by a C. elegans fem-2 mutation. To test whether fem-2 functions as a sex-determining gene in the other Caenorhabditis species we used RNA-mediated interference (RNAi). fem-2 (RNAi) in C. elegans and C. briggsae caused germ-line feminization, but had no noticeable effect in CB5161. Thus the function of fem-2 in CB5161 remains uncertain. Received: 11 April 2001 / Accepted: 6 August 2001  相似文献   

12.
Docosahexaenoic acid is found to be bound to three equivalent sites on albumin with the same affinities as palmitic acid at 0–38°C, which demonstrates that ethene-1,2-diyl- and methylene-groups contribute equally to the affinity. The equilibrium dissociation constants (K dm s) for red cell membrane binding sites of linoleic- and docosahexaenoic acid at pH 7.3 are determined at temperatures between 0 and 37°C. The temperature-independent capacities for binding are 12 ± 1 and 25.4 ± 3.0 nmoles g−1 ghosts respectively. Double isotope binding experiments reveal that the unsaturated fatty acids: arachidonic-, linoleic-, docosahexaenoic-, and oleic acid have partially shared capacities in ratios approximately 1:2:4:5, in contrast to the noncompetitive binding of palmitic acid. The observations suggest a two-tier binding limitation. One is the number of protein sites binding fatty acid anions electrostatically and the other is the number of suitable annular lipids adaptively selected among membrane lipids by the hydrocarbon chain. These competition conditions are confirmed by measurements of the tracer exchange efflux at near 0°C from albumin-free and albumin-filled ghosts of linoleic- and docosahexaenoic acid, either alone or in the presence of arachidonic- and palmitic acid. Under equilibrium conditions, the calculated ratios of inside to outside membrane binding is below 0.5 for four unsaturated fatty acids. The unidirectional rate constants of translocation between the inside and the outside correlate with the number of double bonds in these fatty acids, which are also correlated with the dissociation rate constants of the complexes with albumin. The membrane permeation occurs presumably by binding of the anionic unsaturated fatty acids to an integral protein followed by channeling of the neutral form between opposite binding sites of the protein through annular lipids encircling the protein. Received: 30 June 1997/Revised: 23 February 1998  相似文献   

13.
The effects of Ni2+ were evaluated on slowly-decaying, high-voltage-activated (HVA) Ca2+ currents expressed by pyramidal neurons acutely dissociated from guinea-pig piriform cortex. Whole-cell, patch-clamp recordings were performed with Ba2+ as the charge carrier. Ni2+ blocked HVA Ba2+ currents (I Bas) with an EC50 of approximately 60 μm. Additionally, after application of nonsaturating Ni2+ concentrations, residual currents activated with substantially slower kinetics than both total and Ni2+-sensitive I Bas. None of the pharmacological components of slowly decaying, HVA currents activated with kinetics significantly different from that of total currents, indicating that the effect of Ni2+ on I Bas kinetics cannot be attributed to the preferential inhibition of a fast-activating component. The effect of Ni2+ on I Ba amplitude was voltage-independent over the potential range normally explored in our experiments (−60 to +20 mV), hence the Ni2+-dependent decrease of I Ba activation rate is not due to a voltage- and time-dependent relief from block. Moreover, Ni2+ significantly reduced I Ba deactivation speed upon repolarization, which also is not compatible with a depolarization-dependent unblocking mechanism. The dependence on Ni2+ concentration of the I Ba activation-rate reduction was remarkably different from that found for I Ba block, with an EC50 of ∼20 μm and a Hill coefficient of ∼1.73 vs.∼1.10. These results demonstrate that Ni2+, besides inhibiting the I Bas under study probably by exerting a blocking action on the pore of the underlying Ca2+ channels, also interferes with Ca2+-channel gating kinetics, and strongly suggest that the two effects depend on Ni2+ occupancy of binding sites at least partly distinct. Received: 13 July 2000/Revised: 9 November 2000  相似文献   

14.
Nuclear magnetic resonance (NMR) microimaging and proton relaxation times were used to monitor differences between the hydration state of the nucleus and cytoplasm in the Rana pipiens oocyte. Individual isolated ovarian oocytes were imaged in a drop of Ringer's solution with an in-plane resolution of 80 μm. Proton spin echo images of oocytes arrested in prophase I indicated a marked difference in contrast between nucleoplasm and cytoplasm with additional intensity gradations between the yolk platelet-rich region of the cytoplasm and regions with little yolk. Neither shortening τe (spin echo time) to 9 msec (from 18 msec) nor lengthening τr (spin recovery time) to 2 sec (from 0.5 sec) reduced the observed contrast between nucleus and cytoplasm. Water proton T1 (spin-lattice) relaxation times of oocyte suspensions indicated three water compartments that corresponded to extracellular medium (T1= 3.0 sec), cytoplasm (T1= 0.8 sec) and nucleoplasm (T1= 1.6 sec). The 1.6 sec compartment disappeared at the time of nuclear breakdown. Measurements of plasma and nuclear membrane potentials with KCl-filled glass microelectrodes demonstrated that the prophase I oocyte nucleus was about 25 mV inside positive relative to the extracellular medium. A model for the prophase-arrested oocyte is proposed in which a high concentration of large impermeant ions together with small counter ions set up a Donnan-type equilibrium that results in an increased distribution of water within the nucleus in comparison with the cytosol. This study indicates: (i) a slow exchange between two or more intracellular water compartments on the NMR time-scale, (ii) an increased rotational correlation time for water molecules in both the cytoplasmic and nuclear compartments compared to bulk water, and (iii) a higher water content (per unit dry mass) of the nucleus compared to the cytoplasm, and (iv) the existence of a large (about 75 mV positive) electropotential difference between the nuclear and cytoplasmic compartments. Received: 18 January 1996/Revised: 29 April 1996  相似文献   

15.
The cloned intestinal peptide transporter is capable of electrogenic H+-coupled cotransport of neutral di- and tripeptides and selected peptide mimetics. Since the mechanism by which PepT1 transports substrates that carry a net negative or positive charge at neutral pH is poorly understood, we determined in Xenopus oocytes expressing PepT1 the characteristics of transport of differently charged glycylpeptides. Transport function of PepT1 was assessed by flux studies employing a radiolabeled dipeptide and by the two-electrode voltage-clamp-technique. Our studies show, that the transporter is capable of translocating all substrates by an electrogenic process that follows Michaelis Menten kinetics. Whereas the apparent K0.5 value of a zwitterionic substrate is only moderately affected by alterations in pH or membrane potential, K0.5 values of charged substrates are strongly dependent on both, pH and membrane potential. Whereas the affinity of the anionic dipeptide increased dramatically by lowering the pH, a cationic substrate shows only a weak affinity for PepT1 at all pH values (5.5–8.0). The driving force for uptake is provided mainly by the inside negative transmembrane electrical potential. In addition, affinity for proton interaction with PepT1 was found to depend on membrane potential and proton binding subsequently affects the substrate affinity. Furthermore, our studies suggest, that uptake of the zwitterionic form of a charged substrate contributes to overall transport and that consequently the stoichiometry of the flux-coupling ratios for peptide: H+/H3O+ cotransport may vary depending on pH. Received: 19 August 1996/Revised: 10 October 1996  相似文献   

16.
Spectrofluorimetric measurements were conducted to quantify, in real-time, membrane permeability changes resulting from the treatment of Sf9 insect cells (Spodoptera frugiperda, Lepidoptera) with different Bacillus thuringiensis Cry insecticidal proteins. Coumarin-derived CD222 and Merocyanin-540 probes were respectively used to monitor extracellular K+ and membrane potential variations upon Sf9 cells incubation with Cry toxins. Our results establish that Cry1C induces, after a delay, the depolarization of the cell membrane and the full depletion of intracellular K+. These changes were not observed upon Sf9 cells treated with Cry1A family toxins. Both the rate of the K+ efflux and the delay before its onset were dependent on toxin concentration. Both parameters were sensitive to temperature but only the delay was affected by pH. Cry1C-induced K+ efflux was inhibited by lanthanum ions in a dose-dependent manner. This study provides the first kinetic and quantitative characterization of the ion fluxes through the channels formed by a Cry toxin in the plasma membrane of a susceptible insect cell line. Received: 4 October 1999/Revised: 21 December 1999  相似文献   

17.
The Kedem-Katchalsky (KK) equations are often used to obtain information about the osmotic properties and conductance of channels to water. Using human red cell membranes, in which the osmotic flow is dominated by Aquaporin-1, we show here that compared to NaCl the reflexion coefficient of the channel for methylurea, when corrected for solute volume exchange and for the water permeability of the lipid membrane, is 0.54. The channels are impermeable to these two solutes which would seem to rule out flow interaction and require a reflexion coefficient close to 1.0 for both. Thus, two solutes can give very different osmotic flow rates through a semi-permeable pore, a result at variance with both classical theory and the KK formulation. The use of KK equations to analyze osmotic volume changes, which results in a single hybrid reflexion coefficient for each solute, may explain the discrepancy in the literature between such results and those where the equations have not been employed. Osmotic reflexion coefficients substantially different from 1.0 cannot be ascribed to the participation of other 'hidden' parallel aqueous channels consistently with known properties of the membrane. Furthermore, we show that this difference cannot be due to second-order effects, such as a solute-specific interaction with water in only part of the channel, because the osmosis is linear with driving force down to zero solute concentration, a finding which also rules out the involvement of unstirred-layer effects. Reflexion coefficients smaller than 1.0 do not necessitate water-solute flow interaction in permeable aqueous channels; rather, the osmotic behaviour of impermeable molecular-sized pores can be explained by differences in the fundamental nature of water flow in regions either accessible or inaccessible to solute, created by a varying cross-section of the channel.  相似文献   

18.
Band 3 (AE1), the anion exchanger of the human erythrocyte membrane, mediates not only fluxes of small hydrophilic anions (e.g., chloride, oxalate), but also the flip-flop of long-chain amphiphilic anions (e.g., dodecylsulfate). Treatment of erythrocytes with papain, long known to inhibit the transport of the former type of anions, accelerates the transport of the latter type. In an attempt to elucidate the basis of these opposite responses to papain, several small amphiphilic arylalkyl sulfonates and -sulfates were tested for the response of their transport, via AE1, to papain. Although all these probes are most likely transported by a flux and not by flip-flop, their transport was inhibited by papain only in some cases, but accelerated in others. Different responses to papain therefore most likely do not reflect differences between transport by flux or by flip. The transports of different species of anions also differed considerably in the changes of their sensitivity, to noncovalent and some covalent inhibitors, brought about by papain treatment. While oxalate transport remained as sensitive as in native cells, transports of small amphiphilic anions lost their sensitivity to a major extent, regardless of the inhibition or acceleration of their transport by papain. The results are discussed in the light of present concepts of the structural organisation of AE1, and interpreted in terms of a model of different transport subsites for different species of anions in this transporter. Received: 20 June 2000/Revised: 1 November 2000  相似文献   

19.
The maxi cation channel from the plasma membrane of rye (Secale cereale L.) roots was studied following its incorporation into planar phosphatidylethanolamine bilayers. Current recordings were made in the presence of 100-mm KCl containing quinine on both sides of the bilayer. Quinine produced voltage- and concentration-dependent blockade of the channel, reducing its apparent unitary current and open probability. The voltage-dependence suggested that blockade was effected from the cytoplasmic side by cationic quinine. Blockade was modelled using a kinetic scheme with two independent blocked states termed B1 and B2 (B1OB2). Rate constants promoting fast kinetics (k 1 and k −1 ) were found to be several orders of magnitude greater than those promoting slow kinetics (k 2 and k −2 ). Analysis of the fast kinetics indicated that the rate constants for blockade of the open channel at the first site (k 1 ) and its clearance (k −1 ) had voltage-dependencies (zδ p ) of 0.41 and −0.71, respectively, and that the equilibrium dissociation constant for the binding site (K d (0)) was about 1 mm. Analysis of the slow kinetics indicated that the rate constants for blockade of the open channel at the second site (k 2 ) and its clearance (k −2 ) had zδ p values of 0.12 and −1.27, respectively. The K d (0) value for the second binding site was about 10 mm. Received: 20 January 1998/Revised: 1 May 1998  相似文献   

20.
The permeation properties of adenosine 3′, 5′-cyclic monophosphate (cAMP)-activated recombinant rat olfactory cyclic nucleotide-gated channels (rOCNC1) in human embryonic kidney (HEK 293) cells were investigated using inside-out excised membrane patches. The relative permeability of these rOCNC1 channels to monovalent alkali cations and organic cations was determined from measurements of the changes in reversal potential upon replacing sodium in the bathing solution with different test cations. The permeability ratio of Cl relative to Na+ (P Cl /P Na ) was about 0.14, confirming that these channels are mainly permeable to cations. The sequence of relative permeabilities of monovalent alkali metal ions in these channels was P Na P K > P Li > P Cs P Rb , which closely corresponds to a high-strength field sequence as previously determined for native rat olfactory receptor neurons (ORNs). The permeability sequence for organic cations relative to sodium was P NH3OH > P NH4 > P Na > P Tris > P Choline > P TEA , again in good agreement with previous permeability ratios obtained in native rat ORNs. Single-channel conductance sequences agreed surprisingly well with permeability sequences. These conductance measurements also indicated that, even in asymmetric bi-ionic cation solutions, the conductance was somewhat independent of current direction and dependent on the composition of both solutions. These results indicate that the permeability properties of rOCNC1 channels are similar to those of native rat CNG channels, and provide a suitable reference point for exploring the molecular basis of ion selectivity in recombinant rOCNC1 channels using site-directed mutagenesis. Received: 3 July 2000/Revised: 29 August 2000  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