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.
We have measured the time course of release of 42K and 86Rb from an occluded state of the Na,K-pump using a rapid filtration apparatus. We have found that at 20 degrees C and in the presence of ATP, 42K is released with a rate constant of approximately 45 s-1 and 86Rb with a rate constant of approximately 20 s-1; both ATP and ADP are effective at a low affinity site (Kd approximately 0.3 and 1 mM, respectively) with the rate of deocclusion being only half as great in ADP as in ATP. Mg2+ stimulates 2-fold at low concentrations probably by forming MgATP, and free Mg2+ is strongly inhibitory at high concentrations (Kd approximately 10 mM). Mg2+ also decreases the affinity for ATP, and the data are consistent with mixed type inhibition; from the analysis the dissociation constant is approximately 1 mM for the inhibitory Mg2+ and the Rb+-occluded form without ATP. The rate of 42K or 86Rb release increases monotonically with pH while ATPase activity decreases above pH 8, so that deocclusion is not rate-limiting in the overall cycle at high pH. This is reflected by a convergence of the rate of Na,K-ATPase and Na,Rb-ATPase activities at high pH and by a decrease in the observed steady-state level of the occluded 86Rb intermediate at high pH. K+, Rb+, Na+, and Cs+, but not Li+, increase the rate of 42K and 86Rb release at constant ionic strength, presumably at sites other than the transport sites. The spontaneous rate of deocclusion is only approximately 0.1 s-1 at low ionic strength in the absence of nucleotides, and it is increased markedly by all cations tested except Li+. Overall the data are consistent with deocclusion as a rate-limiting step in the Na,K-pump cycle.  相似文献   

2.
The rate of 86Rb or 42K release from an occluded form of the phosphorylated Na+ pump has been studied using a rapid filtration apparatus described previously. The rate constant of release is 5-15 s-1, and 42K and 86Rb dissociate at approximately the same rate. Mg2+ is required for deocclusion in the presence of Pi at a site which has the same affinity as the site involved in stabilization of E2(K) with ATP; we propose that Na,K-ATPase has only one site for Mg2+ (apart from Mg2+ complexed with ATP), that the affinity of this site for Mg2+ is increased by Pi binding and decreased by ATP binding, and that Mg2+ is bound and released in the normal transport cycle. In the presence of K+, Cs+, Rb+, or Tl+, the release of two distinct 86Rb ions can be observed, the slow release from one site ("s" site) being blocked by occupancy of the site vacated by the other ("f", fast site). By a sequence of incubations, labeled 86Rb can be placed at either site, and the rate of dissociation monitored individually; in the absence of K+, dissociation from the s site proceeds after a lag in which the f site is vacated. The results are consistent with a "flickering-gate" model of deocclusion to the extracellular pump face, in which the site is exposed to the medium only long enough for a single ion to be released. When deocclusion to the intracellular face is promoted with ATP, ions are released from both sites at the same rate, presumably because the E2----E1 conformational change is rate-limiting. Unlabeled ions co-occluded with 86Rb increase the ATP-stimulated rate of release in the order Rb+ less than Tl+ less than Cs+ less than K+; since the same rank order is observed when dissociation from the s site is monitored in the presence of these ions and MgPi we propose that the latter process proceeds toward the intracellular pump face. 86Rb release from the vanadate-inhibited enzyme has the characteristics of Pi-stimulated release but is approximately 25-fold slower. ATP binds to both the phosphorylated and vanadate-inhibited forms of Na,K-ATPase and increases the rate of deocclusion, apparently to both the intracellular and extracellular faces of the pump.  相似文献   

3.
We have studied the effect of various amines on the rate of release of 86Rb from the occluded state of dog kidney Na,K-ATPase formed by pre-incubation of the enzyme with 86Rb. In the presence of MgPi, various amines act like K+ or Rb+ in blocking the release of 86Rb from one of two sites for occlusion (the "s" site). Of 38 amines tested, tetrapropylamine and various benzyl amines exhibit the highest affinity; the K1/2 for these compounds is 2-5 mM. In the presence of ATP, when 86Rb is presumably released towards the intracellular face of the pump in the normal mode of operation, 86Rb release is blocked by the presence of amine, but only if the amine is also included in a preincubation with MgPi. The data are consistent with a model in which the interaction of amine with one of the transport sites (the "f" site) prevents the E2----E1 transformation that is stimulated by ATP. When 86Rb deocclusion from the f site has occurred in the presence of amine, the lone 86Rb at the s site can be released in the presence of ATP if the amine is removed from the medium. This suggests that a single 86Rb ion at the s site can be released to the intracellular face of the membrane, and therefore that transport can occur with only one K+ site occupied. The amine that blocks release of one 86Rb ion does not itself become occluded: (a) The interaction of amine and ATP is only seen when both ligands are present in the medium; (b) the effects of amines are not "remembered" after a brief exposure to a rinse medium; (c) with the vanadate-inhibited enzyme, benzyltriethylamine and tetrapropylamine are only weakly effective in blocking 86Rb release from the s site; and (d) organic cations exhibit very low affinity in competition with 86Rb for occlusion at equilibrium. Thus the results are consistent with the idea that monofunctional amines block by binding to the f site but that, unlike K+ and Rb+, they do not become occluded. In contrast, at equilibrium ethylenediamine prevents 86Rb occlusion in a competitive manner, suggesting the possibility of occlusion of the bifunctional amine.  相似文献   

4.
45Ca2+-40Ca2+ exchangeability of 45Ca bound to the calcium transport sites of unphosphorylated sarcoplasmic reticulum Ca2+-ATPase at equilibrium has been found to be heterogeneous: Half of the bound calcium is [Ca2+]-dependent in a slowly exchangeable (k less than 0.3 s-1), "occluded" state in the Ca2+-ATPase, and the other calcium is [Ca2+]-independent in a rapidly exchangeable (k approximately 0.3 s-1), "unoccluded" state (Nakamura, J. (1986) Biochim. Biophys. Acta 870, 495-501). In this paper, the two different forms of exchangeable calcium were studied after phosphorylation of the enzyme by ATP without added Mg2+ at pH 7.0 and 0 degree C. By the phosphorylation, the degree of the occlusion became higher (k less than 0.03 s-1). The unoccluded calcium was, however, not significantly affected. The more highly occluded calcium exchanged at the same rate as the decay rate of the phosphoenzyme (EP) in the steady state at a ratio of about 1:1. The occluded calcium was relieved by dephosphorylation of EP by ADP. These results suggest that 1 mol of ADP-sensitive EP more highly occluded 1 mol of calcium, already occluded before phosphorylation. After transformation of ADP-sensitive EP to its ADP-insensitive form by the addition of 20 mM Mg2+ at pH 8.8, the unoccluded calcium was rapidly (k = 0.1-0.3 s-1) released from the transformed EP. However, the occluded calcium was maintained in an occluded state in which the calcium was slowly (k approximately 0.01 s-1) released from the EP without exchange. The results suggest that calcium occlusion in the ADP-sensitive EP is not relieved by the loss of ADP sensitivity of the EP itself.  相似文献   

5.
In experiments performed at 37 degrees C, Ca2+ reversibly inhibits the Na+-and (Na+ + K+)-ATPase activities and the K+-dependent phosphatase activity of (Na+ + K+)-ATPase. With 3 mM ATP, the Na+-ATPase was less sensitive to CaCl2 than the (Na+ + K+)-ATPase activity. With 0.02 mM ATP, the Na+-ATPase and the (Na+ + K+)-ATPase activities were similarly inhibited by CaCl2. The K0.5 for Ca2+ as (Na+ + K+)-ATPase inhibitor depended on the total MgCl2 and ATP concentrations. This Ca2+ inhibition could be a consequence of Ca2+-Mg2+ competition, Ca . ATP-Mg . ATP competition or a combination of both mechanisms. In the presence of Na+ and Mg2+, Ca2+ inhibited the K+-dependent dephosphorylation of the phosphoenzyme formed from ATP, had no effect on the dephosphorylation in the absence of K+ and inhibited the rephosphorylation of the enzyme. In addition, the steady-state levels of phosphoenzyme were reduced in the presence both of NaCl and of NaCl plus KCl. With 3 mM ATP, Ca2+ alone sustained no more than 2% of the (Na+ + K+)-ATPase activity and about 23% of the Na+-ATPase activity observed with Mg2+ and no Ca2+. With 0.003 mM ATP, Ca2+ was able to maintain about 40% of the (Na+ + K+)-ATPase activity and 27% of the Na+-ATPase activity seen in the presence of Mg2+ alone. However, the E2(K)-E1K conformational change did not seem to be affected. Ca2+ inhibition of the K+-dependent rho-nitrophenylphosphatase activity of the (Na+ + K+)-ATPase followed competition kinetics between Ca2+ and Mg2+. In the presence of 10 mM NaCl and 0.75 mM KCl, the fractional inhibition of the K+-dependent rho-nitrophenylphosphatase activity as a function of Ca2+ concentration was the same with and without ATP, suggesting that Ca2+ indeed plays the important role in this process. In the absence of Mg2+, Ca2+ was unable to sustain any detectable ouabain-sensitive phosphatase activity, either with rho-nitrophenylphosphate or with acetyl phosphate as substrate.  相似文献   

6.
Effects of various cations on the dephosphorylation of (Na+ + K+)-ATPase, phosphorylated by ATP in 50 mM imidazole buffer (pH 7.0) at 22 degrees C without added Na+, have been studied. The dephosphorylation in imidazole buffer without added K+ is extremely sensitive to K+-activation (Km K+ = 1 microM), less sensitive to Mg2+-activation (Km Mg2+ = 0.1 mM) and Na+-activation (Km Na+ = 63 mM). Imidazole and Na+ effectively inhibit K+-activated dephosphorylation in linear competitive fashion (Ki imidazole 7.5 mM, Ki Na+ 4.6 mM). The Ki for Na+ is independent of the imidazole concentration, indicating different and non-interacting inhibitory sites for Na+ and imidazole. Imidazole inhibits Mg2+-activated dephosphorylation just as effective as K+-activated dephosphorylation, as judged from the Ki values for imidazole in the two processes. Tris buffer and choline chloride, like imidazole, inhibit dephosphorylation in the presence of residual K+ (less than 1 microM), but less effectively in terms of I50 values and extent of inhibition. Tris inhibits to the same extent as choline. This indicates different inhibitory sites for Tris or choline and for imidazole. These findings indicate that high steady-state phosphorylation levels in Na+-free imidazole buffer are due to the induction of a phosphorylating enzyme conformation and to the inhibition of (K+ + Mg2+)-stimulated dephosphorylation.  相似文献   

7.
The hydrolysis of ATP catalyzed by purified (Na,K)-ATPase from pig kidney was more sensitive to Mg2+ inhibition when measured in the presence of saturating Na+ and K+ concentrations [(Na,K)-ATPase] than in the presence of Na+ alone, either at saturating [(Na,Na)-ATPase] or limiting [(Na,0)-ATPase] Na+ concentrations. This was observed at two extreme concentrations of ATP (3 mM where the low-affinity site is involved and 3 microM where only the catalytic site is relevant), although Mg2+ inhibition was higher at low ATP concentration. In the case of (Na,Na)-ATPase activity, inhibition was barely observed even at 10 mM free Mg2+ when ATP was 3 mM. When (Na,K)-ATPase activity was measured at different fixed K+ concentrations the apparent Ki for Mg2+ inhibition was lower at higher monovalent cation concentration. When K+ was replaced by its congeners (Rb+, NH+4, Li+), Mg2+ inhibition was more pronounced in those cases in which the dephosphorylating cation forms a tighter enzyme-cation complex after dephosphorylation. This effect was independent of the ATP concentration, although inhibition was more marked at lower ATP for all the dephosphorylating cations. The K0.5 for ATP activation at its low-affinity site, when measured in the presence of different dephosphorylating cations, increased following the sequence Rb+ greater than K+ greater than NH+4 greater than Li+ greater than none. The K0.5 values were lower with 0.05 mM than with 10 mM free Mg2+ but the order was not modified. The trypsin inactivation pattern of (Na,K)-ATPase indicated that Mg2+ kept the enzyme in an E1 state. Addition of K+ changed the inactivation into that observed with the E2 enzyme form. On the other hand, K+ kept the enzyme in an E2 state and addition of Mg2+ changed it to an E1 form. The K0.5 for KCl-induced E1-to-E2 transformation (observed by trypsin inactivation profile) in the presence of 3 mM MgCl2 was about 0.9 mM. These results concur with two mechanisms for free Mg2+ inhibition of (Na,K)-ATPase: "product" and dead-end. The first would result from Mg2+ interaction with the enzyme in the E2(K) occluded state whereas the second would be brought about by a Mg2+-enzyme complex with the enzyme in an E1 state.  相似文献   

8.
The binding and conformational properties of the divalent cation site required for H+,K(+)-ATPase catalysis have been explored by using Ca2+ as a substitute for Mg2+. 45Ca2+ binding was measured with either a filtration assay or by passage over Dowex cation exchange columns on ice. In the absence of ATP, Ca2+ was bound in a saturating fashion with a stoichiometry of 0.9 mol of Ca2+ per active site and an apparent Kd for free Ca2+ of 332 +/- 39 microM. At ATP concentrations sufficient for maximal phosphorylation (10 microM), 1.2 mol of Ca2+ was bound per active site with an apparent Kd for free Ca2+ of 110 +/- 22 microM. At ATP concentrations greater than or equal to 100 microM, 2.2 mol of Ca2+ were bound per active site, suggesting that an additional mole of Ca2+ bound in association with low affinity nucleotide binding. At concentrations sufficient for maximal phosphorylation by ATP (less than or equal to 10 microM), APD, ADP + Pi, beta,gamma-methylene-ATP, CTP, and GTP were unable to substitute for ATP. Active site ligands such as acetyl phosphate, phosphate, and p-nitrophenyl phosphate were also ineffective at increasing the Ca2+ affinity. However, vanadate, a transition state analog of the phosphoenzyme, gave a binding capacity of 1.0 mol/active site and the apparent Kd for free Ca2+ was less than or equal to 18 microM. Mg2+ displaced bound Ca2+ in the absence and presence of ATP but Ca2+ was bound about 10-20 times more tightly than Mg2+. The free Mg2+ affinity, like Ca2+, increased in the presence of ATP. Monovalent cations had no effect on Ca2+ binding in the absence of ATP but dit reduce Ca2+ binding in the presence of ATP (K+ = Rb+ = NH4 + greater than Na+ greater than Li+ greater than Cs+ greater than TMA+, where TMA is tetramethylammonium chloride) by reducing phosphorylation. These results indicate that the Ca2+ and Mg2+ bound more tightly to the phosphoenzyme conformation. Eosin fluorescence changes showed that both Ca2+ and Mg2+ stabilized E1 conformations (i.e. cytosolic conformations of the monovalent cation site(s)) (Ca.E1 and Mg.E1). Addition of the substrate acetyl phosphate to either Ca.E1 or Mg.E1 produced identical eosin fluorescence showing that Ca2+ and Mg2+ gave similar E2 (extracytosolic) conformations at the eosin (nucleotide) site. In the presence of acetyl phosphate and K+, the conformations with Ca2+ or Mg2+ were also similar. Comparison of the kinetics of the phosphoenzyme and Ca2+ binding showed that Ca2+ bound prior to phosphorylation and dissociated after dephosphorylation.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
In order to characterize low affinity ATP-binding sites of renal (Na+,K+) ATPase and sarcoplasmic reticulum (Ca2+)ATPase, the effects of ATP on the splitting of the K+-sensitive phosphoenzymes were compared. ATP inactivated the dephosphorylation in the case of (Na+,K+)ATPase at relatively high concentrations, while activating it in the case of (Ca2+)ATPase. When various nucleotides were tested in place of ATP, inactivators of (Na+,K+)ATPase were found to be activators in (Ca2+)ATPase, with a few exceptions. In the absence of Mg2+, the half-maximum concentration of ATP for the inhibition or for the activation was about 0.35 mM or 0.25 mM, respectively. These values are comparable to the previously reported Km or the dissociation constant of the low affinity ATP site estimated from the steady-state kinetics of the stimulation of ATP hydrolysis or from binding measurements. By increasing the concentration of Mg2+, but not Na+, the effect of ATP on the phosphoenzyme of (Na+,K+)ATPase was reduced. On the other hand, Mg2+ did not modify the effect of ATP on the phosphoenzyme of (Ca2+)ATPase. During (Na+,K+)ATPase turnover, the low affinity ATP site appeared to be exposed in the phosphorylated form of the enzyme, but the magnesium-complexed ATP interacted poorly with the reactive K+-sensitive phosphoenzyme, which has a tightly bound magnesium, probably because of interaction between the divalent cations. In the presence of physiological levels of Mg2+ and K+, ATP appeared to bind to the (Na+,K+)ATPase only after the dephosphorylation, while it binds to the (Ca2+)-ATPase before the dephosphorylation to activate the turnover.  相似文献   

10.
The effects of anoxia were studied in freshly isolated rat hepatocytes maintained in agarose gel threads and perfused with Krebs-Henseleit bicarbonate buffer (KHB). Cytosolic free calcium (Ca2+i) was measured with aequorin, intracellular sodium (Na+i) with SBFI, intracellular pH (pHi) with BCECF, lactic dehydrogenase (LDH) by the increase in NADH absorbance during lactate oxidation to pyruvate, ATP by 31P NMR spectroscopy in real time, and intracellular free Mg2+ (Mg2+i) from the chemical shift of beta-ATP relative to alpha-ATP in the NMR spectra. Anoxia was induced by perfusing the cells with KHB saturated with 95% N2, 5% CO2. After 1 h of anoxia, beta-ATP fell 66%, and 85% after 2 h, while the Pi/ATP ratio increased 10-fold from 2.75 to 28.3. Under control conditions, the resting cytosolic free calcium was 127 +/- 6 nM. Anoxia increased Ca2+i in two distinct phases: a first rise occurred within 15 min and reached a mean value of 389 +/- 35 nM (p less than 0.001). A second peak reached a maximum value of 1.45 +/- 0.12 microM (p less than 0.001) after 1 h. During the first hour of anoxia, Na+i increased from 15.9 +/- 2.4 mM to 32.2 +/- 1.2 mM (p less than 0.001), Mg2+i doubled from 0.51 +/- 0.05 to 1.12 +/- 0.01 mM (p less than 0.001), and pHi decreased from 7.41 +/- 0.03 to 7.06 +/- 0.1 (p less than 0.001). LDH release doubled during the first hour and increased 6-fold during the second hour of anoxia. Upon reoxygenation, ATP, Ca2+i, Mg2+i, Na+i, and LDH returned near the control levels within 45 min. To determine whether the increased LDH release was related to the rise in Ca2+i, and whether the increased Ca2+i was caused by Ca2+ influx, the cells were perfused with Ca(2+)-free KHB (+ 0.1 mM EGTA) during the anoxic period. After 2 h of anoxia in Ca(2+)-free medium, beta-ATP again fell 90%, but Ca2+i, after the first initial peak, fell below control levels, and LDH release increased only 2.7-fold. During reoxygenation, Ca2+i, ATP, Na+i, and LDH returned near the control levels within 45 min. These results suggest that the rise in Ca2+i induced by anoxia is caused by an influx of Ca2+ from the extracellular fluid, and that LDH release and cell injury may be related to the resulting rise in Ca2+i.  相似文献   

11.
B Vilsen 《Biochemistry》1999,38(35):11389-11400
Mutant Phe788 --> Leu of the rat kidney Na+,K(+)-ATPase was expressed in COS cells to active-site concentrations between 40 and 60 pmol/mg of membrane protein. Analysis of the functional properties showed that the discrimination between Na+ and K+ on the two sides of the system is severely impaired in the mutant. Micromolar concentrations of K+ inhibited ATP hydrolysis (K(0.5) for inhibition 107 microM for the mutant versus 76 mM for the wild-type at 20 mM Na+), and at 20 mM K+, the molecular turnover number for Na+,K(+)-ATPase activity was reduced to 11% that of the wild-type. This inhibition was counteracted by Na+ in high concentrations, and in the total absence of K+, the mutant catalyzed Na(+)-activated ATP hydrolysis ("Na(+)-ATPase activity") at an extraordinary high rate corresponding to 86% of the maximal Na+,K(+)-ATPase activity. The high Na(+)-ATPase activity was accounted for by an increased rate of K(+)-independent dephosphorylation. Already at 2 mM Na+, the dephosphorylation rate of the mutant was 8-fold higher than that of the wild-type, and the maximal rate of Na(+)-induced dephosphorylation amounted to 61% of the rate of K(+)-induced dephosphorylation. The cause of the inhibitory effect of K+ on ATP hydrolysis in the mutant was an unusual stability of the K(+)-occluded E2(K2) form. Hence, when E2(K2) was formed by K+ binding to unphosphorylated enzyme, the K(0.5) for K+ occlusion was close to 1 microM in the mutant versus 100 microM in the wild-type. In the presence of 100 mM Na+ to compete with K+ binding, the K(0.5) for K+ occlusion was still 100-fold lower in the mutant than in the wild-type. Moreover, relative to the wild-type, the mutant exhibited a 6-7-fold reduced rate of release of occluded K+, a 3-4-fold increased apparent K+ affinity in activation of the pNPPase reaction, a 10-11-fold lower apparent ATP affinity in the Na+,K(+)-ATPase assay with 250 microM K+ present (increased K(+)-ATP antagonism), and an 8-fold reduced apparent ouabain affinity (increased K(+)-ouabain antagonism).  相似文献   

12.
The present work compares the effects of several ligands (phosphatase substrates, MgCl2, RbCl and inorganic phosphate) and temperature on the phosphatase activity and the E2(Rb) occluded conformation of Na+/K+-ATPase. Cooling from 37 degrees C to 20 degrees C and 0 degrees C (hydrolysis experiments) or from 20 degrees C to 0 degrees C (occlusion experiments) had the following consequences: (i) dramatically reduced the Vmax for p-nitrophenyl phosphate and acetyl phosphate hydrolysis but it produced little or no changes in the Km for the substrates; (ii) led to a 5-fold drop in the Km for the inorganic phosphate-induced di-occlusion of E2(Rb); (iii) reduced the K0.5 and curve sigmoidicity of the Rb-stimulated hydrolysis of p-nitrophenyl phosphate and acetyl phosphate and the Rb-promoted E2(Rb) formation. At 20 degrees C, in the presence of 1 mM RbCl and no Mg2+, acetyl phosphate did not affect E2(Rb); with 3 mM MgCl2, acetyl phosphate stimulated a release of Rb from E2(Rb) both in the presence and absence of RbCl in the incubation mixture. As a function of acetyl phosphate concentration the Km for iRb release was indistinguishable from the Km found for stimulation of hydrolysis and enzyme phosphorylation under identical experimental conditions; in addition, the extrapolated di-occluded fraction corresponding to maximal hydrolysis was not different from 100%. These results indicate that although E2(K) might be an intermediary in the phosphatase reaction, the most abundant enzyme conformation during phosphatase turnover is E2 which has no K+ occluded in it. The ligand interactions associated to phosphatase activity do not support an equivalence of this reaction with the dephosphorylation step in the Na+ + K+-dependent ATP hydrolysis; on the other hand, there are similarities with the reversible binding of inorganic phosphate in the presence of Mg2+ and K+ ions.  相似文献   

13.
The rate of phosphorylation of sodium and potassium ion-transport adenosine triphosphatase by 10 microM [gamma-32P]ATP was much slower with Ca2+ than with Mg2+ (0.13-10 mM) in the presence of 16 to 960 mM Na+ at 0 degrees C and pH 7.4. In the presence of a fixed concentration of Mg2+ or Ca2+, the rate became slower with increasing Na+ concentration. When the Na+ concentration was fixed, the rate became slower with decreasing divalent cation concentration. Sodium ions appear to antagonize the divalent cation in the phosphorylation to slow its rate. In the presence of 1 mM Ca2+ and 126 or 270 mM Na+, the rate was slow enough to permit the manual addition of a chasing solution at various times before the phosphorylation reached the steady state. Therefore, we studied the time-dependent change of the sensitivity to ADP or to K+ of the phosphoenzyme by a chase with unlabeled ATP containing ADP or K+ during the time range from the transient to the steady state of the phosphorylation. The ADP sensitivity decreased and the K+ sensitivity increased with the progress of the phosphorylation. With 270 mM Na+, the phosphoenzyme found at 1 s, when its amount was 5.5% of the maximum level, was virtually completely sensitive to ADP. Under these conditions, it was concluded that the form of the phosphoenzyme initially produced from the enzyme.ATP complex has ADP sensitivity and that the phosphoenzyme acquires K+ sensitivity later. The initially produced ADP-sensitive phosphoenzyme partially lost its normal instability and sensitivity upon adding a chelating agent, probably because of dissociation of a divalent cation from the phosphoenzyme.  相似文献   

14.
Oligomycin reduced the fluorescence intensity of an N-(p-(2-benzimidazoly)phenyl) maleimide (BIPM) probe at Cys-964 of the alpha-chain of pig kidney Na+,K(+)-ATPase with increase in the concentration of Na+ with a Hill coefficient of nh = 0.77 with Kh = 231 mM. The maximum fluorescence decrease was around 80% of the value observed after accumulation of ADP-sensitive phosphoenzyme (E1P) in the presence of 2 M Na+. The addition of Mg2+ and ATP with Na+ or choline chloride to give the same final ligand concentration to the Na(+)-enzyme-oligomycin complex formed with 16 mM Na+ + 1,984 mM choline chloride or 2 M Na+ induced rapid phosphorylation (20 or 21/s) and slower fluorescence decrease (12.1 +/- 1.2 or 10.1 +/- 3.2/s). These additions to the Na(+)-enzyme complex formed under the former or the latter conditions induced slow phosphorylation (13/s) prior to a much slower fluorescence decrease (3.4 +/- 0.3 or 8.6 +/- 0.7/s). The addition of Ca2+ and ATP to these enzyme complexes induced rapid fluorescence changes (21-11/s) followed by one order of magnitude slower rates of phosphorylation (1.5-1.3 s). These data suggest that the decrease in BIPM fluorescence induced by ATP with Ca2+ or with Mg2+, reflects the change of the Na+ binding state before or after the formation of E1P, respectively.  相似文献   

15.
Vanadate inhibition of the Ca2+-ATPase from human red cell membranes   总被引:4,自引:0,他引:4  
(1) VO3(-) combines with high affinity to the Ca2+-ATPase and fully inhibits Ca2+-ATPase and Ca2+-phosphatase activities. Inhibition is associated with a parallel decrease in the steady-state of the Ca2+-dependent phosphoenzyme. (2) VO3(-) blocks hydrolysis of ATP at the catalytic site. The sites for VO3(-) also exhibit negative interactions in affinity with the regulatory sites for ATP of the Ca2+-ATPase. (3) The sites for VO39-) show positive interaactions in affinity with sites for Mg2+ and K+. This accounts for the dependence on Mg2+ and K+ of the inhibition by VO3(-). Although, with less effectiveness, Na2+ and K+ substitutes for K+ whereas Li+ does not. The apparent affinites for Mg24 and K+ for inhibiton by VO3(-) seem to be less than those for activation of the Ca2+-ATPase. (4) Inhibition by VO3(-) is independent of Ca2+ at concentrations up to 50 microM. Higher concentrations of Ca2+ lead to a progressive release of the inhibitiory effect of VO3(-).  相似文献   

16.
D M Chipman  A Lev 《Biochemistry》1983,22(19):4450-4459
Glutaraldehyde treatment of electroplax membrane preparations of Na,K-ATPase leads to irreversible changes in the enzymic behavior of the protein, which are not due to modification of the active site. When the glutaraldehyde treatment is carried out in a medium containing K+ and without Na+, the "K+-modified enzyme" so produced shows the following changes in enzymic properties: The steady-state phosphorylation by ATP and the rate of ATP-ADP exchange are decreased to approximately 40% of control, while Na,K-ATPase activity decreases to approximately 15% of control. Phosphatase activity is decreased very little, but the potassium activation parameters of the reaction are changed, from K0.5 approximately equal to 5 mM and nH = 1.9 in control to K0.5 approximately equal to 0.5 mM and nH = 1 in K+-modified enzyme. KI(app) for nucleotide inhibition of phosphatase activity is increased significantly. Changes in the cation dependence of the ATPase reaction are also observed. All of these effects can be explained by assuming that the cross-linking of surface groups in protein subunits when they are in conformation E2 shifts the intrinsic conformational equilibrium of the enzyme toward E2. We considered the simplest mathematical model for the coupling between K+ binding and the conformational equilibrium, with equivalent potassium sites that must be simultaneously in the same state. If one assumes that the potassium activation of phosphatase activity in the K+-modified enzyme reflects the affinity for K+ of E2, the behavior of the phosphatase activity in the native enzyme can be fit if there are only two potassium sites, whose affinity is 80-fold higher in E2 than in E1, and the equilibrium constant for E2 in equilibrium E1 is about 250. The same sites can explain the activation of dephosphorylation during ATP hydrolysis. Independent of the model chosen, potassium ions must be required for the catalytic action of form E2 and cannot be merely "allosteric activators". The enzyme modified with glutaraldehyde in a medium containing Na+ also has interesting properties, but their rationalization is less straightforward. The Na,K-ATPase activity is inhibited more than the "partial reactions", as in the K+-modified enzyme. We suggest that this is a generally expected result of modifications of the enzyme.  相似文献   

17.
We have previously demonstrated that Na+,K(+)-ATPase can be phosphorylated by 100 microM ATP and 5 mM Mg2+ and in the absence of Na+, provided that 40% dimethylsulfoxide (Me2SO) is present. Phosphorylation was stimulated by K+ up to a steady-state level of about 50% of Etot (Barrabin et al. (1990) Biochim. Biophys. Acta 1023, 266-273). Here we describe the time-course of phosphointermediate (EP) formation and of dephosphorylation of EP at concentrations of Mg2+ from 0.1 to 5000 microM and of K+ from 0.01 to 100 mM. The results were simulated by a simplified version of the commonly accepted Albers-Post model, i.e. a 3-step reaction scheme with a phosphorylation, a dephosphorylation and an isomerization/deocclusion step. Furthermore it was necessary to include an a priori, Mg(2+)- and K(+)-independent, equilibration between two enzyme forms, only one of which (constituting 14% of Etot) reacted directly with ATP. The role of Mg(2+) was two-fold: At low Mg2+, phosphorylation was stimulated by Mg2+ due to formation of the substrate MgATP, whereas at higher concentrations it acted as an inhibitor at all three steps. The affinity for the inhibitory Mg(2+)-binding was increased several-fold, relative to that in aqueous media, by dimethylsulfoxide. K+ stimulated dephosphorylation at all Mg(2+)-concentrations, but at high, inhibitory [Mg2+], K+ also stimulated the phosphorylation reaction, increasing both the rate coefficient and the steady-state level of EP. Generally, the presence of Me2SO seems to inhibit the dephosphorylation step, the isomerization/deocclusion step, and to a lesser extent (if at all) the phosphorylation reaction, and we discuss whether this reflects that Me2SO stabilizes occluded conformations of the enzyme even in the absence of monovalent cations. The results confirm and elucidate the stimulating effect of K+ on EP formation from ATP in the absence of Na+, but they leave open the question of the molecular mechanism by which Me2SO, inhibitory Mg2+ and stimulating K+ interact with the Na+,K(+)-ATPase.  相似文献   

18.
Lowering the extracellular K+ content from 6 to 0.6 mM causes a rise, and elevation from 6 to 8.5 mM a fall of 45Ca++ efflux from the vascular smooth muscle cells of the arteria carotis communis of cattle. In contrast, a level of 17 mM K+ has no influence. Removal of extracellular calcium does not block these effects. 10(-4) M ouabain also induces a rise in Ca++ efflux, additional potassium reduction then being without effect; 10(-9) M ouabain is of no influence. The 45Ca++ efflux kinetics correlates with the activity of the isolated Na,K-ATPase. Tonus increases of the vascular strips by 10(-4) M ouabain and potassium deficiency cannot be blocked by 4 mM lanthanum or removal of extracellular calcium. Unlike sodium, potassium stimulates the active Ca++ binding and the activity of the Ca-ATPase of the microsomal fraction. The ative Ca++ binding of the mitochondria is stimulated by both ions. It is postulated that the activity of the plasma membrane Na,K-pump is able to regulate the tonus of big arteries through alteration of Ca++ storage processes.  相似文献   

19.
We have recently shown that inactivation of renal Na,K-ATPase by 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide occurs via an intramolecular cross-link formed between an activated carboxyl group and an endogenous nucleophile (Pedemonte, C.H., and Kaplan, J.H. (1986) J. Biol. Chem. 261, 3632-3639). The modified enzyme shows the same level of Rb+ binding as untreated enzyme: 3.16 and 2.93 ATP-sensitive mumol of Rb+ binding/mumol of phosphoenzyme, respectively. Thus, the Rb+ binding site and the transition accomplished by low affinity nucleotide binding which accelerates de-occlusion are not greatly affected by the carbodiimide inactivation. 1 mM K+ reduces the ADP binding to the high affinity nucleotide binding site to the same extent in normal and 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide-treated enzyme and Na+ counteracts this effect. Thus, the competition between Na+ and K+ ions for binding to the free enzyme are also largely unaltered by the modification. Phosphorylation from ATP (microM) in the presence of Na+ and Mg2+ ions and from inorganic phosphate in the presence of Mg2+ ions (in the absence or presence of ouabain) is greatly inhibited (85%) following carbodiimide treatment. The extent of inhibition of phosphorylation quantitatively correlates with the residual Na,K-ATPase activity (15%). Consequently, the rate of inactivation by carbodiimide is reduced when a greater proportion of the enzyme is in the phosphorylated form. Fluoroscein isothiocyanate, which inhibits the Na,K-ATPase by covalently modifying a lysine residue close to the high affinity binding site for ATP in the alpha-subunit does not bind to the carbodiimide-inactivated enzyme. Since high affinity nucleotide binding is only partially inhibited by the modification produced by the carbodiimide this suggests that the lysine residue to which fluoroscein isothiocyanate binds is not specifically required for competent nucleotide binding.  相似文献   

20.
Previously, we proposed the following reaction machanism for the transport ATPase (EC 3.6.1.3) reaction in the presence of high concentrations of Mg2+ and Na+:(see article). Some kinetic and thermodynamic properties of steps 3 and 4 were investigated, and the following results were obtained. 1. When the reaction was started by adding ATP to the enzyme in the presence of 50 mM Na+ and 0.5 mM K+ or in the presence of 50mM Na+ and 0.5mM Rb+, the amount of E ADP P increased with time and maintained a constant level after reaching a maximum. We could not observe the initial burst of EP formation, which was observed by Post er al. in the presence of 8 mM Na+ and 0.01 mM Rb+. 2. The existence of quasi-equilibrium between E2ATP and E ADP P in the presence of low concentrations of Na+ was suggested by the fact that the values of the reciprocal of the equilibrium constant, K3 of step 3 obtained by the following three methods were almost the same. a) The value of 1+K3 was estimated from the ratio of vo/[EP] to kd, where vo is the rate of ATP hydrolysis in the steady state, [EP] the concentration of EP, and kd the first-order rate constant of EP disappearance after stopping EP formation. b) This value was also calculated from the ratio of the amount of P1 liberated to that of decrease in EP after stopping EP formation. c) The value of K3 was also calculated from the initial rapid decrease in EP on adding K+ and EDTA, assuming that the rapid decrease was due to a shift of the equilibrium toward E2ATP on adding K+. For example, the value of K3 with 10mM NaCL and 0.5mM KCL was 7--11. Although ATP formation due to a shift of the equilibrium toward E2ATP by a K+ jump in the presence of a low concentration of Na+ was observed at 0 degrees, the amount of ATP formed by a K+ jump at 15 degrees was less than the value expected from the shift of the equilibrium. 3. The values of delta H degrees and delta S degrees of step 3 were estimated in the presence of a sufficient amount of Na+ and in the absence of K+. They were +4--+5 kcal mole minus 1 and +15--+16 entropy units mole minus1, respectively. On the basis of kinetic studies of the elementary steps and the overall reaction of Na+-K+-dependent ATPase [EC 3.6.1.3], we (1--4) showed that a phosphorylated intermediate, EP, is formed via two kinds of enzyme-substrate complex, E1ATP and E2ATP, that the EP is in K+-dependent quasi-equilibrium with E2ATP, and that in the presence of high concentration of Mg2+, EP is in a high-energy state and contains bound ADP, E ADP P.(see article).  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