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 46 毫秒
1.
The effects of organic quaternary amines, tetraethylammonium (TEA) chloride and benzyltriethylammonium (BTEA) chloride, on Na,K pump current were examined in rat cardiac myocytes superfused in extracellular Na(+)-free solutions and whole-cell voltage-clamped with patch electrodes containing a high Na(+)-salt solution. Extracellular application of these quaternary amines competitively inhibited extracellular K(+) (K(+)(o)) activation of Na,K pump current; however, the concentration for half maximal inhibition of Na,K pump current at 0 mV (K(0)(Q)) by BTEA, 4.0 +/- 0.3 mM, was much lower than the K(0)(Q) for TEA, 26.6 +/- 0.7 mM. Even so, the fraction of the membrane electric field dissipated during K(+)(o) activation of Na,K pump current (lambda(K)), 39 +/- 1%, was similar to lambda(K) determined in the presence of TEA (37 +/- 2%) and BTEA (35 +/- 2%), an indication that the membrane potential (V(M)) dependence for K(+)(o) activation of the Na,K pump current was unaffected by TEA and BTEA. TEA was found to inhibit the Na,K pump current in a V(M)-independent manner, i.e., inhibition of current dissipated 4 +/- 2% of the membrane electric field. In contrast, BTEA dissipated 40 +/- 5% of the membrane electric field during inhibition of Na,K pump current. Thus, BTEA inhibition of the Na,K-ATPase is V(M)-dependent. The competitive nature of inhibition as well as the similar fractions of the membrane electric field dissipated during K(+)(o)-dependent activation and BTEA-dependent inhibition of Na,K pump current suggest that BTEA inhibits the Na,K-ATPase at or very near the enzyme's K(+)(o) binding site(s) located in the membrane electric field. Given previous findings that organic quaternary amines are not occluded by the Na,K-ATPase, these data clearly demonstrate that an ion channel-like structure provides access to K(+)(o) binding sites in the enzyme.  相似文献   

2.
The steady-state voltage and [Na(+)](o) dependence of the electrogenic sodium pump was investigated in voltage-clamped internally dialyzed giant axons of the squid, Loligo pealei, under conditions that promote the backward-running mode (K(+)-free seawater; ATP- and Na(+)-free internal solution containing ADP and orthophosphate). The ratio of pump-mediated (42)K(+) efflux to reverse pump current, I(pump) (both defined by sensitivity to dihydrodigitoxigenin, H(2)DTG), scaled by Faraday's constant, was -1.5 +/- 0.4 (n = 5; expected ratio for 2 K(+)/3 Na(+) stoichiometry is -2.0). Steady-state reverse pump current-voltage (I(pump)-V) relationships were obtained either from the shifts in holding current after repeated exposures of an axon clamped at various V(m) to H(2)DTG or from the difference between membrane I-V relationships obtained by imposing V(m) staircases in the presence or absence of H(2)DTG. With the second method, we also investigated the influence of [Na(+)](o) (up to 800 mM, for which hypertonic solutions were used) on the steady-state reverse I(pump)-V relationship. The reverse I(pump)-V relationship is sigmoid, I(pump) saturating at large negative V(m), and each doubling of [Na(+)](o) causes a fixed (29 mV) rightward parallel shift along the voltage axis of this Boltzmann partition function (apparent valence z = 0.80). These characteristics mirror those of steady-state (22)Na(+) efflux during electroneutral Na(+)/Na(+) exchange, and follow without additional postulates from the same simple high field access channel model (Gadsby, D.C., R.F. Rakowski, and P. De Weer, 1993. Science. 260:100-103). This model predicts valence z = nlambda, where n (1.33 +/- 0.05) is the Hill coefficient of Na binding, and lambda (0.61 +/- 0.03) is the fraction of the membrane electric field traversed by Na ions reaching their binding site. More elaborate alternative models can accommodate all the steady-state features of the reverse pumping and electroneutral Na(+)/Na(+) exchange modes only with additional assumptions that render them less likely.  相似文献   

3.
The transport of Na+ and Ca2+ ions in the cardiac Na(+)-Ca2+ exchanger can be described as separate events (Khananshvili, D. (1990) Biochemistry 29, 2437-2442). Thus, the Na(+)-Na+ and Ca(2+)-Ca2+ exchange reactions reflect reversible partial reactions of the transport cycle. The effect of diffusion potentials (K(+)-valinomycin) on different modes of the Na(+)-Ca2+ exchanger (Na(+)-Ca2+, Ca(2+)-Ca2+, and Na(+)-Na+ exchanges) were tested in reconstituted proteoliposomes, obtained from the Triton X-100 extracts of the cardiac sarcolemmal membranes. The initial rates of the Nai-dependent 45Ca-uptake (t = 1 s) were measured in EGTA-entrapped proteoliposomes at different voltages. At the fixed values of voltage [45 Ca]o was varied from 4 to 122 microM, and [Na]i was saturating (150 mM). Upon varying delta psi from -94 to +91 mV, the Vmax values were increased from 9.5 +/- 0.5 to 26.5 +/- 1.5 nmol.mg-1.s-1 and the Km from 17.8 +/- 2.5 to 39.1 +/- 5.2 microM, while the Vmax/Km values ranged from only 0.53 +/- 0.08 to 0.73 +/- 0.17 nmol.mg-1.s-1.microM-1. The equilibrium Ca(2+)-Ca2+ exchange was voltage sensitive at very low [Ca]o = [Ca]i = 2 microM, while at saturating [Ca]o = [Ca]i = 200 microM the Ca(2+)-Ca2+ exchange became voltage-insensitive. The rates of the equilibrium Na(+)-Na+ exchange appears to be voltage insensitive at saturating [Na]o = [Na]i = 160 mM. Under the saturating ionic conditions, the rates of the Na(+)-Na+ exchange were at least 2-3-fold slower than the Ca(2+)-Ca2+ exchange. The following conclusions can be drawn. (a) The near constancy of the Vmax/Km for Na(+)-Ca2+ exchange at different voltages is compatible with the ping-pong model proposed previously. (b) The effects of voltage on Vmax of Na(+)-Ca2+ exchange are consistent with the existence of a single charge carrying transport step. (c) It is not yet possible to clearly assign this step to the Na+ or Ca2+ transport half of the cycle although it is more likely that 3Na(+)-transport is a charge carrying step. Thus, the unloaded ion-binding domain contains either -2 or -3 charges (presumably carboxyl groups). (d) The binding of Na+ and Ca2+ appears to be weakly voltage-sensitive. The Ca(2+)-binding site may form a small ion-well (less than 2-3 A).  相似文献   

4.
Na,K-ATPase containing the amino acid substitution glutamate to alanine at position 779 of the alpha subunit (Glu779Ala) supports a high level of Na-ATPase and electrogenic Na+-Na+ exchange activity in the absence of K+. In microsomal preparations of Glu779Ala enzyme, the Na+ concentration for half maximal activation of Na-ATPase activity was 161 +/- 14 mM (n = 3). Furthermore, enzyme activity with 800 mM Na+ was found to be similar in the presence and absence of 20 mM K+. These results showed that Na+, with low affinity, could stimulate enzyme turnover as effectively as K+. To gain further insight into the mechanism of this enzyme activity, HeLa cells expressing Glu779Ala enzyme were voltage clamped with patch electrodes containing 115 mM Na+ during superfusion in K+-free solutions. Electrogenic Na+-Na+ exchange was observed as an ouabain-inhibitable outward current whose amplitude was proportional to extracellular Na+ (Na+(o)) concentration. At all Na+(o) concentrations tested (3-148 mM), exchange current was maximal at negative membrane potentials (V(M)), but decreased as V(M) became more positive. Analyzing this current at each V(M) with a Hill equation showed that Na+-Na+ exchange had a high-affinity, low-capacity component with an apparent Na+(o) affinity at 0 mV (K0(0.5)) of 13.4 +/- 0.6 mM and a low-affinity, high-capacity component with a K0(0.5) of 120 +/- 13 mM (n = 17). Both high- and low-affinity exchange components were V(M) dependent, dissipating 30 +/- 3% and 82 +/- 6% (n = 17) of the membrane dielectric, respectively. The low-affinity, but not the high-affinity exchange component was inhibited with 2 mM free ADP in the patch electrode solution. These results suggest that the high-affinity component of electrogenic Na+-Na+ exchange could be explained by Na+(o) acting as a low-affinity K+ congener; however, the low-affinity component of electrogenic exchange appeared to be due to forward enzyme cycling activated by Na+(o) binding at a Na+-specific site deep in the membrane dielectric. A pseudo six-state model for the Na,K-ATPase was developed to simulate these data and the results of the accompanying paper (Peluffo, R.D., J.M. Argüello, and J.R. Berlin. 2000. J. Gen. Physiol. 116:47-59). This model showed that alterations in the kinetics of extracellular ion-dependent reactions alone could explain the effects of Glu779Ala substitution on the Na,K-ATPase.  相似文献   

5.
We have used admittance analysis together with the black lipid membrane technique to analyze electrogenic reactions within the Na(+) branch of the reaction cycle of the Na(+)/K(+)-ATPase. ATP release by flash photolysis of caged ATP induced changes in the admittance of the compound membrane system that are associated with partial reactions of the Na(+)/K(+)-ATPase. Frequency spectra and the Na(+) dependence of the capacitive signal are consistent with an electrogenic or electroneutral E(1)P <--> E(2)P conformational transition which is rate limiting for a faster electrogenic Na(+) dissociation reaction. We determine the relaxation rate of the rate-limiting reaction and the equilibrium constants for both reactions at pH 6.2-8.5. The relaxation rate has a maximum value at pH 7.4 (approximately 320 s(-1)), which drops to acidic (approximately 190 s(-1)) and basic (approximately 110 s(-1)) pH. The E(1)P <--> E(2)P equilibrium is approximately at a midpoint position at pH 6.2 (equilibrium constant approximately 0.8) but moves more to the E(1)P side at basic pH 8.5 (equilibrium constant approximately 0.4). The Na(+) affinity at the extracellular binding site decreases from approximately 900 mM at pH 6.2 to approximately 200 mM at pH 8.5. The results suggest that during Na(+) transport the free energy supplied by the hydrolysis of ATP is mainly used for the generation of a low-affinity extracellular Na(+) discharge site. Ionic strength and lyotropic anions both decrease the relaxation rate. However, while ionic strength does not change the position of the conformational equilibrium E(1)P <--> E(2)P, lyotropic anions shift it to E(1)P.  相似文献   

6.
Time constants of slow inactivation were investigated in NH(2)-terminal deleted Shaker potassium channels using macro-patch recordings from Xenopus oocytes. Slow inactivation is voltage insensitive in physiological solutions or in simple experimental solutions such as K(+)(o)//K(+)(i) or Na(+)(o)//K(+)(i). However, when [Na(+)](i) is increased while [K(+)](i) is reduced, voltage sensitivity appears in the slow inactivation rates at positive potentials. In such solutions, the I-V curves show a region of negative slope conductance between approximately 0 and +60 mV, with strongly increased outward current at more positive voltages, yielding an N-shaped curvature. These changes in peak outward currents are associated with marked changes in the dominant slow inactivation time constant from approximately 1.5 s at potentials less than approximately +60 mV to approximately 30 ms at more than +150 mV. Since slow inactivation in Shaker channels is extremely sensitive to the concentrations and species of permeant ions, more rapid entry into slow inactivated state(s) might indicate decreased K(+) permeation and increased Na(+) permeation at positive potentials. However, the N-shaped I-V curve becomes fully developed before the onset of significant slow inactivation, indicating that this N-shaped I-V does not arise from permeability changes associated with entry into slow inactivated states. Thus, changes in the relative contributions of K(+) and Na(+) ions to outward currents could arise either: (a) from depletions of [K(+)](i) sufficient to permit increased Na(+) permeation, or (b) from voltage-dependent changes in K(+) and Na(+) permeabilities. Our results rule out the first of these mechanisms. Furthermore, effects of changing [K(+)](i) and [K(+)](o) on ramp I-V waveforms suggest that applied potential directly affects relative permeation by K(+) and Na(+) ions. Therefore, we conclude that the voltage sensitivity of slow inactivation rates arises indirectly as a result of voltage-dependent changes in the ion occupancy of these channels, and demonstrate that simple barrier models can predict such voltage-dependent changes in relative permeabilities.  相似文献   

7.
This study examines the conformations of the Na(+)/glucose cotransporter (SGLT1) during sugar transport using charge and fluorescence measurements on the human SGLT1 mutant G507C expressed in Xenopus oocytes. The mutant exhibited similar steady-state and presteady-state kinetics as wild-type SGLT1, and labeling of Cys507 by tetramethylrhodamine-6-maleimide had no effect on kinetics. Our strategy was to record changes in charge and fluorescence in response to rapid jumps in membrane potential in the presence and absence of sugar or the competitive inhibitor phlorizin. In Na(+) buffer, step jumps in membrane voltage elicited presteady-state currents (charge movements) that decay to the steady state with time constants tau(med) (3-20 ms, medium) and tau(slow) (15-70 ms, slow). Concurrently, SGLT1 rhodamine fluorescence intensity increased with depolarizing and decreased with hyperpolarizing voltages (DeltaF). The charge vs. voltage (Q-V) and fluorescence vs. voltage (DeltaF-V) relations (for medium and slow components) obeyed Boltzmann relations with similar parameters: zdelta (apparent valence of voltage sensor) approximately 1; and V(0.5) (midpoint voltage) between -15 and -40 mV. Sugar induced an inward current (Na(+)/glucose cotransport), and reduced maximal charge (Q(max)) and fluorescence (DeltaF(max)) with half-maximal concentrations (K(0.5)) of 1 mM. Increasing [alphaMDG](o) also shifted the V(0.5) for Q and DeltaF to more positive values, with K(0.5)'s approximately 1 mM. The major difference between Q and DeltaF was that at saturating [alphaMDG](o), the presteady-state current (and Q(max)) was totally abolished, whereas DeltaF(max) was only reduced 50%. Phlorizin reduced both Q(max) and DeltaF(max) (K(i) approximately 0.4 microM), with no changes in V(0.5)'s or relaxation time constants. Simulations using an eight-state kinetic model indicate that external sugar increases the occupancy probability of inward-facing conformations at the expense of outward-facing conformations. The simulations predict, and we have observed experimentally, that presteady-state currents are blocked by saturating sugar, but not the changes in fluorescence. Thus we have isolated an electroneutral conformational change that has not been previously described. This rate-limiting step at maximal inward Na(+)/sugar cotransport (saturating voltage and external Na(+) and sugar concentrations) is the slow release of Na(+) from the internal surface of SGLT1. The high affinity blocker phlorizin locks the cotransporter in an inactive conformation.  相似文献   

8.
Phosphorylation of shark rectal Na,K-ATPase by ATP in the presence of Na(+) was characterized by chemical quench experiments and by stopped-flow RH421 fluorescence. The appearance of acid-stable phosphoenzyme was faster than the rate of fluorescence increase, suggesting that of the two acid-stable phosphoenzymes formed, RH421 exclusively detects formation of E(2)-P, which follows formation of E(1)-P. The stopped-flow RH421 fluorescence response to ATP phosphorylation was biphasic, with a major fast phase with k(obs) approximately 90 s(-1) and a minor slow phase with a k(obs) of approximately 9 s(-1) (20 degrees C, pH 7.4). The observed rate constants for both the slow and the fast phase could be fitted with identical second-degree functions of the ATP concentration with apparent binding constants of approximately 3.1 x 10(7) M(-1) and 1. 8 x 10(5) M(-1), respectively. Increasing [ADP] decreased k(obs) for the rate of the RH421 fluorescence response to ATP phosphorylation. This could be accounted for by the reaction of ADP with the initially formed E(1)-P followed by a conformational change to E(2)-P. The biphasic stopped-flow RH421 responses to ATP phosphorylation could be simulated, assuming that in the absence of K(+) the highly fluorescent E(2)-P is slowly transformed into the "K(+)-insensitive" E'(2)-P subconformation forming a side branch of the main cycle.  相似文献   

9.
The current (I(p)) generated by the wild-type or the glutamate (E) 779 alanine (A) mutant of the rat Na(+)/K(+) pump alpha1-subunit expressed in HEK 293 cells was studied at 35 degrees C by means of whole-cell recording in Na(+)-free and Na(+)-containing solution. Glutamate 779 is located in the fifth transmembrane domain of the alpha-subunit of the Na(+)/K(+)-ATPase. Compared with the wild-type, the E779A mutant exhibited an apparent K(+)(o)-affinity decreased by a factor of 3-4 both in Na(+)-free and in Na(+)-containing media. The competition of Na(+)(o) and K(+)(o) for cation binding sites of the pump remained unchanged. Similarly, in Na(+)-free solution the shape of the I(p)-V curves for various external K(+)-concentrations ([K(+)](o)) was essentially the same. However, in Na(+)-containing solutions the shape of I(p)-V curves from cells expressing the mutant of the rat alpha1-subunit clearly differed from the shape observed in cells expressing the wild-type, but voltage dependence of the pump current persisted. A prominent Na(+)(o)-activated, electrogenic Na(+)-transport mediated by the pump, displaying little voltage dependence in the potential range tested (-80 to +60 mV), was present in the cells expressing the E779A mutant pump. The data suggest that exchanging E779 for A in the rat Na(+)/K(+) pump alpha1-subunit causes a modest decrease in the apparent K(+)(o) affinity and a profound, Na(+)(o)-dependent alteration in the electrogenicity of the mutant pump expressed in HEK 293 cells.  相似文献   

10.
Fedosova NU  Champeil P  Esmann M 《Biochemistry》2003,42(12):3536-3543
Transient kinetic analysis of nucleotide binding to pig kidney Na,K-ATPase using a rapid filtration technique shows that the interaction between nucleotide and enzyme apparently follows simple first-order kinetics both for ATP in the absence of Mg(2+) and for ADP in the presence or absence of Mg(2+). Rapid filtration experiments with Na,K-ATPase membrane sheets may nevertheless suffer from a problem of accessibility for a fraction of the ATPase binding sites. Accordingly, we estimate from these data that for ATP binding in the absence of Mg(2+) and the presence of 35 mM Na(+) at pH 7.0 at 20 degrees C, the bimolecular binding rate constant k(on) is about 30 microM(-1) x s(-1) and the dissociation rate constant k(off) is about 8 s(-1). In the presence of 10 mM Mg(2+), the binding rate constant is the same as that in the absence of Mg(2+). For ADP or MgADP the binding rate constant is about 20 microM(-1) x s(-1) and the dissociation rate constant is about 12 s(-1). Results of rapid-mixing stopped-flow experiments with the fluorescent dye eosin are also consistent with a one-step mechanism of binding of eosin to the ATPase nucleotide site. The implication of these results is that nucleotide binding to Na,K-ATPase both in the absence and presence of Mg(2+) appears to be a single-step event, at least on the time scale accessible in these experiments.  相似文献   

11.
Ward DG  Taylor M  Lilley KS  Cavieres JD 《Biochemistry》2006,45(10):3460-3471
ATP has high- and low-affinity effects on the sodium pump and other P-type ATPases. We have approached this question by using 2',3'-O-(trinitrophenyl)-8-azidoadenosine 5'-diphosphate (TNP-8N(3)-ADP) to photoinactivate and label Na,K-ATPase, both in its native state and after covalent FITC block of its high-affinity ATP site. With the native enzyme, the photoinactivation rate constant increases hyperbolically with a K(D(TNP-8N)3(-)(ADP)) of 0.11 microM; TNP-ATP and ATP protect the site with high affinities. The inactivation does not require Na(+), but K(+) inhibits with a K(K)' of 12 microM; Na(+) reverses this effect, with a K(Na) of 0.17 mM. This pattern suggests that Na(+) and K(+) are binding at sites in their "intracellular" conformation. It was known that FITC did not abolish the reverse phosphorylation by P(i), or the K(+)-phosphatase activity, and that TNP-8N(3)-ADP could subsequently photoinactivate the latter with >100-fold lower affinity; in that case, the cation sites acted as if facing outward [Ward, D. G., and Cavieres, J. D. (1998) J. Biol. Chem. 273, 14277-14284, 33759-33765]. Native and FITC-modified enzymes have now been photolabeled with TNP-8N(3)-[alpha-(32)P]ADP and alpha-chain soluble tryptic peptides separated by reverse-phase HPLC. With native Na,K-ATPase, three labeled peaks lead to the unique sequence alpha-(470)Ile-Val-Glu-Ile-Pro-Phe-Asn-Ser-Thr-Asn-X-Tyr-Gln-Leu-Ser-Ile-His-Lys(487), the dropped residue being alphaLys480. With the FITC enzyme, instead, two independent labeling and purification cycles return the sequence alpha-(721)Ala-Asp-Ile-Gly-Val-Ala-Met-Gly-Ile-Ala-Gly-Ser-Asp-Val-Ser-Lys(736). These results suggest that Na,K-ATPase also has a low-affinity nucleotide binding region, one that is under distinctive allosteric control by Na(+) and K(+). Moreover, the cation effects seem compatible with a slow, passive Na(+)/K(+) carrier behavior of the FITC-modified sodium pump.  相似文献   

12.
The kinetics of Na(+)-dependent partial reactions of the Na+,K(+)-ATPase from rabbit kidney were investigated via the stopped-flow technique, using the fluorescent labels N-(4-sulfobutyl)-4-(4-(p-(dipentylamino)phenyl)butadienyl)py ridinium inner salt (RH421) and 5-iodoacetamidofluorescein (5-IAF). When covalently labeled 5-IAF enzyme is mixed with ATP, the two labels give almost identical kinetic responses. Under the chosen experimental conditions two exponential time functions are necessary to fit the data. The dominant fast phase, 1/tau 1 approximately 155 s-1 for 5-IAF-labeled enzyme and 1/tau 1 approximately 200 s-1 for native enzyme (saturating [ATP] and [Na+], pH 7.4 and 24 degrees C), is attributed to phosphorylation of the enzyme and a subsequent conformational change (E1ATP(Na+)3-->E2P(Na+)3 + ADP). The smaller amplitude slow phase, 1/tau 2 = 30-45 s-1, is attributed to the relaxation of the dephosphorylation/rephosphorylation equilibrium in the absence of K+ ions (E2P<==>E2). The Na+ concentration dependence of 1/tau 1 showed half-saturation at a Na+ concentration of 6-8 mM, with positive cooperatively involved in the occupation of the Na+ binding sites. The apparent dissociation constant of the high-affinity ATP-binding site determined from the ATP concentration dependence of 1/tau 1 was 8.0 (+/- 0.7) microM. It was found that P3-1-(2-nitrophenyl)ethyl ATP, tripropylammonium salt (NPE-caged ATP), at concentrations in the hundreds of micromolar range, significantly decreases the value of 1/tau 1, observed. This, as well as the biexponential nature of the kinetic traces, can account for previously reported discrepancies in the rates of the reactions investigated.  相似文献   

13.
The contribution of electrostatic forces to the interaction of Na,K-ATPase with adenine nucleotides was investigated by studying the effect of ionic strength on nucleotide binding. At pH 7.0 and 20 degrees C, there was a qualitative correlation between the equilibrium dissociation constant (K(d)) values for ATP, ADP, and MgADP and their total charges. All K(d) values increased with increasing ionic strength. According to the Debye-Hückel theory, this suggests that the nucleotide binding site and its ligands have "effective" charges of opposite signs. However, quantitative analysis of the dependence on ionic strength shows that the product of the effective electrostatic charges on the ligand and the binding site is the same for all nucleotides, and is therefore independent of the total charge of the nucleotide. The data suggest that association of nucleotides with Na,K-ATPase is governed by a partial charge rather than the total charge of the nucleotide. This charge, interacting with positive charges on the protein, is probably the one corresponding to the alpha-phosphate of the nucleotide. Dissociation rate constants measured in complementary transient kinetic experiments were 13 s(-1) for ATP and 27 s(-1) for ADP, independent of the ionic strength in the range 0.1-0.5 M. This implies similar association rate constants for the two nucleotides (about 40 x 10(6) M(-1) s(-1) at I = 0.1 M). The results suggest that long-range Coulombic forces, affecting association rates, are not the main contributors to the observed differences in affinities, and that local interactions, affecting dissociation rates, may play an even greater role.  相似文献   

14.
A method for determining individual rate constants for nucleotide binding to and dissociation from membrane bound pig kidney Na,K-ATPase is presented. The method involves determination of the rate of relaxation when Na,K-ATPase in the presence of eosin is mixed with ADP or ATP in a stopped-flow fluorescence apparatus. It is shown that the nucleotide dependence of this rate of relaxation--taken together with measured equilibrium binding values for eosin and ADP--makes possible a reasonably reliable determination of the rate constant for dissociation of nucleotide, i.e., determination of the rate constant k-1 in the following model (where E denotes Na,K-ATPase): [formula: see text] All experiments are carried out at about 4 degrees C in a buffer containing 200 mM sucrose, 10 mM EDTA, 25 mM Tris and 73 mM NaCl (pH 7.4). Values obtained for the rate constants for dissociation are about 6 s-1 for ADP and 2-3 s-1 for ATP.  相似文献   

15.
Fedosova NU  Esmann M 《Biochemistry》2004,43(14):4212-4218
Correlation between the Na,K-ATPase affinity to ADP and the cation (its nature and concentration) present in the medium was investigated. In buffer with low ionic strength (I approximately 1 mM) high-affinity ADP binding was not observed, while a stepwise increase in the concentrations of added cation (Na(+), Tris(+), imidazole(+), N-methylglucamine(+), choline(+)) induced an increase in the ADP affinity. The effect was fully saturated at 30-50 mM for all of the cations tested. The maximal affinity for ADP was slightly higher in the presence of Na(+), Tris(+), or imidazole(+) than in the presence of N-methylglucamine(+) or choline(+) (equilibrium dissociation constant K(d) 0.2-0.3 vs 0.7 microM). The ADP dissociation rates from its complex with enzyme in the presence of Na(+) or Tris(+) were similar, implying identity of the nucleotide-binding enzyme conformations, which therefore are assigned to E(1). The ability to compete with K(+) clearly distinguished Na(+) from other cations, which speaks against the sole involvement of the transport sites in the induction of the ADP-binding E(1) conformation. Since the cations are similar in their mode of induction of the high ADP affinity but they demonstrate a pronounced difference in ability to compete with K(+), their effects cannot be combined within any scheme with only one type of cation-binding sites. We suggest that the high affinity toward nucleotide is induced by cation interactions within the protein or lipid and that these nucleotide-domain-related sites coexist with the transport sites, which bind only Na(+) or K(+).  相似文献   

16.
HCO3- exit across the basolateral membrane of the kidney proximal tubule cell is mediated via an electrogenic Na+:HCO3- cotransporter. We have studied the effect of pH on the activity of this cotransport system in basolateral membrane vesicles isolated from rabbit renal cortex. At constant internal pH 6.0, increasing the external pH and [HCO3-] increased the rate of 4,4'-diisothiocyanostilbene-2,2'-disulfonic acid-sensitive 22Na+ influx into the vesicles. To determine the role of internal pH on the activity of the Na+:HCO3- cotransport system, the influx of 22Na+ via HCO3-dependent Na(+)-Na+ exchange was measured in the absence of an initial pH and [HCO3-] gradient (pH(i) = pH(o), 5% CO2). Increasing the pH from 6.8 to 7.2 increased whereas, increasing the pH from 7.4 to 8.0 decreased the rate of 22Na+ influx via this exchange. Increasing pH at constant [HCO3-] (pH(i) = pH(o) = 8.0, 1.5% CO2 versus pH(i) = pH(o) = 7.2, 10% CO2) reduced the influx of 22Na+ via HCO3-dependent Na(+)-Na+ exchange. Increasing pH at constant [CO3(2-)](pH(i) = pH(o) = 8.0, 1.5% CO2 versus pH(i) = pH(o) = 7.2, 60% CO2) was associated with reduced 22Na+ uptake. Decreasing the pH (pH(i) = pH(o) = 6.3, 60% CO2 versus pH(i) = pH(o) = 7.2, 5% CO2) was associated with a reduced rate of HCO3(-)-dependent Na(+)-Na+ exchange. We conclude that the Na+:HCO3- cotransporter displays a significant pH sensitivity profile with the cotransporter being more functional at pH 7.0-7.4 and less active at more acid or alkaline pH. In addition, the results suggest that the pH sensitivity arises at the inner surface of the basolateral membrane.  相似文献   

17.
Microsomal Na, K-ATPase is activated by acetylcholine (5 x 10(-6)--10(-5) M) in a cell-free system including neuronal nuclei and the microsomal--cytoplasmic fraction. No enzyme activation by acetylcholine occurs in the presence of puromycin, actinomycin D and ribonuclease or upon removal of the nuclear or microsomal--cytoplasmic fraction from the system. After preincubation with acetylcholine the membranes reveal a better capacity for phosphorylation by [gamma-32P]ATP and dephosphorylation in the presence of ADP and Na+. The ATP binding by the membranes preincubated in a system with acetylcholine is also increased thereby. It was assumed that acetylcholine induces the synthesis of Na, K-ATPase or its protein activator.  相似文献   

18.
The resting membrane potential (V(m)) of isolated somatic longitudinal muscles of the earthworm Lumbricus terrestris was studied by glass microelectrodes. The inhibition of chloride permeability by low pH did not affect V(m) of the muscle fibers in isolated somatic longitudinal muscles of the earthworm Lumbricus terrestris which was -48.7 mV (inside negative) at pH 7.3 and -49.1 at pH 5.6. On the other hand, bathing the muscles in Cl(-) and Na(+)-free solutions, or application of the chloride transporter inhibitor furosemide and Na(+)-K(+)-ATPase inhibitor ouabain depolarized the V(m) by 3-5 mV. The effects of a Cl(-) -free solution and ouabain were not additive. This demonstrates relatively small contribution of equilibrium potential for Cl(-) to the resting membrane potential and electrogenic effect of Na(+)K(+)-ATPase which is dependent on the supply of Na(+)(i) ions by furosemide-sensitive and Cl(-)(e)- and Na(+)(e)-dependent electroneutral transport (most probably Na(+)K(+)Cl(-) cotransport).  相似文献   

19.
In immature neurones, the steady-state intracellular Cl- concentration [Cl-](i) is generally higher than expected for passive distribution, and this is believed to be due to Na(+)-K(+)-2Cl(-) co-transport. Here, we show that N2a neuroblastoma cells, incubated in HEPES-buffered NaCl medium maintain a [Cl-](i) around 60 mm, two- to threefold higher than expected for passive distribution at a membrane potential of - 49 mV. When the cells were transferred to a Cl(-) -free medium, [Cl-](i) decreased quickly (t(1/2) < 5 min), suggesting a high Cl- permeability. When the intracellular ATP concentration was reduced to less than 1 mm by metabolic inhibitors, the initial rate of (36) Cl- uptake was strongly inhibited (60-65%) while steady-state [Cl-](i) decreased to 24 mm, close to the value predicted from the Nernst equilibrium. Moreover, after reduction of [ATP](i) and [Cl-](i) by rotenone, the subsequent addition of glucose led to a reaccumulation of Cl-, in parallel with ATP recovery. Internal bicarbonate did not affect Cl- pumping, suggesting that Cl-/HCO(3)(-) exchange does not significantly contribute to active transport. Likewise, Na(+) -K(+) -2Cl(-) co-transport also appeared to play a minor role: although mRNA for the NKCC1 form of the co-transporter was detected in N2a cells, neither the initial rate of (36)Cl- uptake nor steady-state [Cl-](i) were appreciably decreased by 10 microm bumetanide or replacement of external Na(+) by choline. These results suggest that a highly active ATP-dependent mechanism, distinct from Na(+) -K(+) -2Cl(-) co-transport, is responsible for most of the inward Cl- pumping in N2a cells.  相似文献   

20.
[(3)H]noradrenaline ([(3)H]NA) released from sympathetic nerves in the isolated main pulmonary artery of the rabbit was measured in response to field stimulation (2Hz, 1ms, 60V for 3min) in the presence of uptake blockers (cocaine, 3 x10(-5)M and corticosterone, 5 x10(-5)M). The [(3)H]NA-release was fully blocked by the combined application of the selective and irreversible 'N-type' voltage-sensitive Ca(2+)-channel (VSCC)-blocker omega-conotoxin (omega-CgTx) GVIA (10(-8)M) and the 'non-selective' VSCC-blocker aminoglycoside antibiotic neomycin (3x10(-3)M). Na(+)-loading (Na(+)-pump inhibition by K(+)-free perfusion) was required to elicit further NA-release after blockade of VSCCs (omega-CgTx GVIA+neomycin). In K(+)-free solution, in the absence of functioning VSCCs (omega-CgTx GVIA+neomycin), the fast Na(+)-channel activator veratridine (10(-5)M) further potentiated the nerve-evoked release of [(3)H]NA. This NA-release was significantly inhibited by KB-R7943, and fully blocked by Ca(o)(2+)-removal. However, Li(+)-substitution was surprisingly ineffective. The non-selective K(+)-channel blocker 4-aminopyridine (4-AP, 10(-4)M) also further potentiated the nerve-evoked release of NA in K(+)-free solution. This potentiated release was concentration-dependently inhibited by KB-R7943, significantly inhibited by Li(+)-substitution and abolished by Ca(o)(2+)-removal. It is concluded that in Na(+)-loaded sympathetic nerves, in which the VSCCs are blocked, the reverse Na(+)/Ca(2+)-exchange-mediated Ca(2+)-entry is responsible for transmitter release on nerve-stimulation. Theoretically we suppose that the fast Na(+)-channel and the exchanger proteins are close to the vesicle docking sites.  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