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 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   

2.
The kinetics of Na(+)-Ca2+ exchange current after a cytoplasmic Ca2+ concentration jump (achieved by photolysis of DM-nitrophen) was measured in excised giant membrane patches from guinea pig or rat heart. Increasing the cytoplasmic Ca2+ concentration from 0.5 microM in the presence of 100 mM extracellular Na+ elicits an inward current that rises with a time constant tau 1 < 50 microseconds and decays to a plateau with a time constant tau 2 = 0.65 +/- 0.18 ms (n = 101) at 21 degrees C. These current signals are suppressed by Ni2+ and dichlorobenzamil. No stationary current, but a transient inward current that rises with tau 1 < 50 microseconds and decays with tau 2 = 0.28 +/- 0.06 ms (n = 53, T = 21 degrees C) is observed if the Ca2+ concentration jump is performed under conditions that promote Ca(2+)-Ca2+ exchange (i.e., no extracellular Na+, 5 mM extracellular Ca2+). The transient and stationary inward current is not observed in the absence of extracellular Ca2+ and Na+. The application of alpha-chymotrypsin reveals the influence of the cytoplasmic regulatory Ca2+ binding site on Ca(2+)-Ca2+ and forward Na(+)-Ca2+ exchange and shows that this site regulates both the transient and stationary current. The temperature dependence of the stationary current exhibits an activation energy of 70 kj/mol for temperatures between 21 degrees C and 38 degrees C, and 138 kj/mol between 10 degrees C and 21 degrees C. For the decay time constant an activation energy of 70 kj/mol is observed in the Na(+)-Ca2+ and the Ca(2+)-Ca2+ exchange mode between 13 degrees C and 35 degrees C. The data indicate that partial reactions of the Na(+)-Ca2+ exchanger associated with Ca2+ binding and translocation are very fast at 35 degrees C, with relaxation time constants of about 6700 s-1 in the forward Na(+)-Ca2+ exchange and about 12,500 s-1 in the Ca(2+)-Ca2+ exchange mode and that net negative charge is moved during Ca2+ translocation. According to model calculations, the turnover number, however, has to be at least 2-4 times smaller than the decay rate of the transient current, and Na+ inward translocation appears to be slower than Ca2+ outward movement.  相似文献   

3.
The Na(+)-Ca2+ exchanger from Drosophila was expressed in Xenopus and characterized electrophysiologically using the giant excised patch technique. This protein, termed Calx, shares 49% amino acid identity to the canine cardiac Na(+)-Ca2+ exchanger, NCX1. Calx exhibits properties similar to previously characterized Na(+)-Ca2+ exchangers including intracellular Na+ affinities, current-voltage relationships, and sensitivity to the peptide inhibitor, XIP. However, the Drosophila Na(+)-Ca2+ exchanger shows a completely opposite response to cytoplasmic Ca2+. Previously cloned Na(+)-Ca2+ exchangers (NCX1 and NCX2) are stimulated by cytoplasmic Ca2+ in the micromolar range (0.1- 10 microM). This stimulation of exchange current is mediated by occupancy of a regulatory Ca2+ binding site separate from the Ca2+ transport site. In contrast, Calx is inhibited by cytoplasmic Ca2+ over this same concentration range. The inhibition of exchange current is evident for both forward and reverse modes of transport. The characteristics of the inhibition are consistent with the binding of Ca2+ at a regulatory site distinct from the transport site. These data provide a rational basis for subsequent structure-function studies targeting the intracellular Ca2+ regulatory mechanism.  相似文献   

4.
The deduced amino acid sequence of the cardiac sarcolemmal Na(+)-Ca2+ exchanger has a region which could represent a calmodulin binding site. As calmodulin binding regions of proteins often have an autoinhibitory role, a synthetic peptide with this sequence was tested for functional effects on Na(+)-Ca2+ exchange activity. The peptide inhibits the Na(+)-dependent Ca2+ uptake (KI approximately 1.5 microM) and the Nao(+)-dependent Ca2+ efflux of sarcolemmal vesicles in a noncompetitive manner with respect to both Na+ and Ca2+. The peptide is also a potent inhibitor (KI approximately 0.1 microM) of the Na(+)-Ca2+ exchange current of excised sarcolemmal patches. The binding site for the peptide on the exchanger is on the cytoplasmic surface of the membrane. The exchanger inhibitory peptide binds calmodulin with a moderately high affinity. From the characteristics of the inhibition of the exchange of sarcolemmal vesicles, we deduce that only inside-out sarcolemmal vesicles participate in the usual Na(+)-Ca2+ exchange assay. This contrasts with the common assumption that both inside-out and right-side-out vesicles exhibit exchange activity.  相似文献   

5.
Zhou HY  Han CY  Wang XL 《生理学报》2006,58(2):136-140
心肌缺血损伤过程中,胞内Na^+、ATP及pH都出现明显变化。钠/钙交换对心肌细胞的钙平衡起重要的调节作用。本实验采用膜片钳全细胞记录豚鼠心室肌细胞钠/钙交换电流,研究温度和胞内Na^+、ATP及pH对钠/钙交换双向电流的影响。结果表明,温度从22℃升至34℃,钠/钙交换电流增大约4倍,而pH值的改变对钠/钙交换双向电流没有明显的影响。在22~24℃时,同时耗竭胞内ATP和胞内酸化对钠/钙交换双向转运功能影响程度小;而在34—37℃时,同时耗竭胞内ATP和胞内酸化能抑制钠/钙交换双向电流的外向和内向成分,且内向成分抑制程度高于外向成分抑制程度。表明同时耗竭胞内ATP和胞内酸化对钠/钙交换的作用具有温度依赖性。胞内Na^+超载能使钠/钙交换电流的外向成分增加,但不增加或减少内向电流(即正向转运)成分。因此,胞内酸化及耗竭胞内ATP损伤细胞排钙机制和胞内钠超载通过钠/钙反向交换引起钙内流是引起心肌细胞钙超载的两个独立的重要因素。  相似文献   

6.
Increases in intracellular free Ca(2+)+ concentration (Ca(2+)+ oscillations) occur during meiotic maturation and fertilization of mammalian oocytes but little is known about the mechanisms of Ca(2+) homeostasis in these cells. Cells extrude Ca(2+) from the cytosol using two main transport processes, the Ca(2+)-ATPase and the Na(+)-Ca(2+) exchanger. The aim of this study was to determine whether Na(+)-Ca(2+) exchange activity is present in immature and mature mouse oocytes. Na(+)-Ca(2+) exchange can be revealed by altering the Na(+) concentration gradient across the plasma membrane and recording intracellular free Ca(2+) concentrations using Ca(2+)-sensitive fluorescent dyes. Depletion of extracellular Na(+) caused an immediate increase in Ca(2+) concentration in immature oocytes and a delayed increase in mature oocytes. The Na(+) ionophore, monensin, caused an increase in intracellular Ca(2+) in immature oocytes similar to that induced by Na(+)-depleted medium. In mature oocytes, monensin had no effect on intracellular Ca(2+) but the time taken for Ca(2+) to reach a peak value on removal of extracellular Na(+) was significantly decreased. Finally, addition of Ca(2+) to immature oocytes incubated in Ca(2+)-free medium caused an increase in the concentration of intracellular Ca(2+) that was dependent upon the presence of extracellular Na(+). This effect was not seen in mature oocytes. The data show that Na(+)-Ca(2+) exchange occurs in immature and mature mouse oocytes and that Ca(2+) homeostasis in immature oocytes is more sensitive to manipulations that activate Na(+)-Ca(2+) exchange.  相似文献   

7.
本研究探讨低氧和AlF-4等药物经G-蛋白敏感的跨膜信号通路对心血管肌源性张力的调控作用。在含有稳定表达Na+-Ca2+交换蛋白的CK1.4细胞中,用fura-2荧光影像确定细胞低氧对胞浆游离Ca2+[Ca2+]i的影响。在离体犬心乳头肌、颈动脉、主动脉及肺动脉恒温灌流样本中,用张力-电换能器测量低氧灌流和AlF-4等药物对心血管肌源性张力的影响。在整体犬体内,按拉丁方设计,用125Isod-1获得不同剂量VISA高效剂细胞内分布等药代动力学参数。结果表明:①在CK1.4细胞中,低氧抑制Na+-Ca2+交换蛋白,产生Ca2+内流,升高[Ca2+]i;②低氧灌流削弱AlF-4所致的血管收缩而明显易化Ca2+内流所致的心乳头肌收缩,与结果1)吻合;③VISA高效剂分布至细胞内,协同AlF-4,模拟并激活G-蛋白敏感的跨膜信号通路,显著改善低氧所致的Na+-Ca2+交换蛋白等跨膜大分子和心血管收缩蛋白氧化损害。  相似文献   

8.
We have investigated temperature dependence of Ca2+ uptake by the cardiac sarcolemmal Na(+)-Ca2+ exchanger from dog, rabbit and bullfrog. In native rabbit sarcolemmal vesicles, Ca2+ affinity of the Na(+)-Ca2+ exchanger is unchanged from 7 to 37 degrees C; however, the initial velocity of Ca2+ uptake declines much more steeply below 22 degrees C than above 22 degrees C. In native dog sarcolemma, the temperature dependence of Na(+)-Ca2+ exchange velocity is similar to that of native rabbit. However, in frog heart the velocity of Na(+)-Ca2+ exchange declines much more slowly with decreasing temperature at both temperature ranges. Reconstitution of the Na(+)-Ca2+ exchanger into artificial lipid vesicles consisting of either asolectin or phosphatidylserine, phosphatidylcholine, and cholesterol has little effect on temperature dependence of Na(+)-Ca2+ exchange velocity in any of the three species. We conclude that the lesser temperature sensitivity of the cardiac sarcolemmal Na(+)-Ca2+ exchanger of a poikilothermic species is at least partly an intrinsic property of the transport protein.  相似文献   

9.
Ciliates possess diverse Ca2+ homeostasis systems, but little is known about the occurrence of a Na(+)-Ca2+ exchanger. We studied Na(+)-Ca2+ exchange in the ciliate Euplotes crassus by digital imaging. Cells were loaded with fura-2/AM or SBF1/AM for fluorescence measurements of cytosolic Ca2+ and Na+ respectively. Ouabain pre-treatment and Na+o substitution in fura-2/AM-loaded cells elicited a bepridil-sensitive [Ca2+]i rise followed by partial recovery, indicating the occurrence of Na(+)-Ca2+ exchanger working in reverse mode. In experiments on prolonged effects, ouabain, Na+o substitution, and bepridil all caused Ca2+o-dependent [Ca2+]i increase, showing a role for Na(+)-Ca2+ exchange in Ca2+ homeostasis. In addition, by comparing the effect of orthovanadate (affecting not only Ca2+ ATPase, but also Na(+)-K+ ATPase and, hence, Na(+)-Ca2+ exchange) to that of bepridil on [Ca2+]i, it was shown that Na(+)-Ca2+ exchange contributes to Ca2+ homeostasis. In electrophysiological experiments, no membrane potential variation was observed after bepridil treatment suggesting compensatory mechanisms for ion effects on cell membrane voltage, which also agrees with membrane potential stability after ouabain treatment. In conclusion, data indicate the presence of a Na(+)-Ca2+ exchanger in the plasma membrane of E. crassus, which is essential for Ca2+ homeostasis, but could also promote Ca2+ entry under specific conditions.  相似文献   

10.
The sarcolemmal Na(+)-Ca2+ exchanger is regulated by intracellular Ca2+ at a high affinity Ca2+ binding site separate from the Ca2+ transport site. Previous data have suggested that the Ca2+ regulatory site is located on the large intracellular loop of the Na(+)-Ca2+ exchange protein, and we have identified a high-affinity 45Ca2+ binding domain on this loop (Levitsky, D. O., D. A. Nicoll, and K. D. Philipson. 1994. Journal of Biological Chemistry. 269:22847-22852). We now use electrophysiological and mutational analyses to further define the Ca2+ regulatory site. Wild-type and mutant exchangers were expressed in Xenopus oocytes, and the exchange current was measured using the inside- out giant membrane patch technique. Ca2+ regulation was measured as the stimulation of reverse Na(+)-Ca2+ exchange (intracellular Na+ exchanging for extracellular Ca2+) by intracellular Ca2+. Single-site mutations within two acidic clusters of the Ca2+ binding domain lowered the apparent Ca2+ affinity at the regulatory site from 0.4 to 1.1-1.8 microM. Mutations had parallel effects on the affinity of the exchanger loop for 45Ca2+ binding (Levitsky et al., 1994) and for functional Ca2+ regulation. We conclude that we have identified the functionally important Ca2+ binding domain. All mutant exchangers with decreased apparent affinities at the regulatory Ca2+ binding site also have a complex pattern of altered kinetic properties. The outward current of the wild-type Na(+)-Ca2+ exchanger declines with a half time (th) of 10.8 +/- 3.2 s upon Ca2+ removal, whereas the exchange currents of several mutants decline with th values of 0.7-4.3 s. Likewise, Ca2+ regulation mutants respond more rapidly to Ca2+ application. Study of Ca2+ regulation has previously been possible only with the exchanger operating in the reverse mode as the regulatory Ca2+ and the transported Ca2+ are then on opposite sides of the membrane. The use of exchange mutants with low affinity for Ca2+ at regulatory sites also allows demonstration of secondary Ca2+ regulation with the exchanger in the forward or Ca2+ efflux mode. In addition, we find that the affinity of wild-type and mutant Na(+)-Ca2+ exchangers for intracellular Na+ decreases at low regulatory Ca2+. This suggests that Ca2+ regulation modifies transport properties and does not only control the fraction of exchangers in an active state.  相似文献   

11.
The possible contribution of Na(+)-Ca(2+) exchange to the triggering of Ca(2+) release from the sarcoplasmic reticulum in ventricular cells remains unresolved. To gain insight into this issue, we measured the "trigger flux" of Ca(2+) crossing the cell membrane in rabbit ventricular myocytes with Ca(2+) release disabled pharmacologically. Under conditions that promote Ca(2+) entry via Na(+)-Ca(2+) exchange, internal [Na(+)] (10 mM), and positive membrane potential, the Ca(2+) trigger flux (measured using a fluorescent Ca(2+) indicator) was much greater than the Ca(2+) flux through the L-type Ca(2+) channel, indicating a significant contribution from Na(+)-Ca(2+) exchange to the trigger flux. The difference between total trigger flux and flux through L-type Ca(2+) channels was assessed by whole-cell patch-clamp recordings of Ca(2+) current and complementary experiments in which internal [Na(+)] was reduced. However, Ca(2+) entry via Na(+)-Ca(2+) exchange measured in the absence of L-type Ca(2+) current was considerably smaller than the amount inferred from the trigger flux measurements. From these results, we surmise that openings of L-type Ca(2+) channels increase [Ca(2+)] near Na(+)-Ca(2+) exchanger molecules and activate this protein. These results help to resolve seemingly contradictory results obtained previously and have implications for our understanding of the triggering of Ca(2+) release in heart cells under various conditions.  相似文献   

12.
Experiments are described demonstrating that Na(+)-Ca2+ exchange of retinal rod disc membrane is highly sensitive to light. The Na(+)-Ca2+ exchanger was shown to possess two types of binding sites with different affinities for calcium. The low affinity binding sites (KCaD = 5.8 mumol/l) are light-insensitive. After bleaching, KD of the high affinity Ca2(+)-binding sites an Ki for Na+ changed from 0.2 to 0.3 mumol/l and from 3.2 to 0.7 nmol/l, respectively. Light inhibits the steady-state Ca2+ uptake by a factor of 1.5. Photocontrol of the Na(+)-Ca2+ exchanger affinity is observed at the physiological level of rhodopsin bleaching.  相似文献   

13.
The possible structural changes in the major isotype of parvalbumin from the toad (Bufo bufo japonicus) skeletal muscle caused by Ca2+ and Mg2+ binding have been analyzed by microcalorimetric titrations. Parvalbumin was titrated with Ca2+ in both the absence and presence of Mg2+ and with Mg2+ in the absence of Ca2+, at pH 7.0, and at 5 degrees, 15 degrees, and 25 degrees C. The two sites in a molecule were equivalent on Mg2(+)-Ca2+ exchange, but distinguishable on Ca2+ and Mg2+ binding. The reactions of parvalbumin with Ca2+ are exothermic at every temperature in both the absence and presence of Mg2+, but those with Mg2+ are always endothermic except for the binding to site 1 at 25 degrees C. The magnitudes of the hydrophobic and internal vibrational contributions to the heat capacity and entropy changes of parvalbumin on Ca2+ and Mg2+ binding and Mg2(+)-Ca2+ exchange have been estimated by the empirical method of Sturtevant [Sturtevant, J. M. (1977) Proc. Natl Acad. Sci. USA 74, 2236-2240]. Although no major conformational changes were noted between Ca2(+)- and Mg2(+)-bound forms of toad parvalbumin, the conformational difference was larger in Ca2+ (or Mg2+) binding to site 1 than site 2. This may indicate that the metal-free form is much less stable than any form with Ca2+ (or Mg2+) bound at one site at least. On Mg2(+)-Ca2+ exchange, the vibrational as well as hydrophobic entropy is only slightly increased in a parallel manner. In contrast, on Ca2+ (or Mg2+) binding, the hydrophobic entropy increases but the vibrational entropy decreases; the former indicates the sequestering of nonpolar groups from the surface to the interior of a molecule, and the latter suggests that the overall structures are tightened on Ca2+ (or Mg2+) binding but loosened on Mg2(+)-Ca2+ exchange. Despite the clear distinctions in the thermodynamic features, the conformational changes of toad parvalbumin are essentially the same as those of the two isotypes of bullfrog parvalbumins on Ca2+ binding and Mg2(+)-Ca2+ exchange.  相似文献   

14.
In order to compare the importance of Na(+)-Ca2+ exchange in the regulation of cytosolic Ca2+ concentration (Ca2+i), acini obtained from rat pancreas and submandibular glands as well as cardiac myocytes were loaded with Na+ by inhibition of Na(+)-K+ ATPase activity then loaded with fura-2. In the exocrine tissues, incubation in K(+)-free buffer or with ouabain had no substantial effect on resting Ca2+i or on the changes in Ca2+i following exposure to carbachol as compared with acini incubated under control conditions. In contrast, rat cardiac myocytes, treated identically, showed marked changes in Ca2+i under resting and stimulated conditions as compared with controls. We conclude that the Na(+)-Ca2+ exchange systems of rat pancreatic and submandibular gland acini contribute little to the overall regulation of Ca2+i at rest during cholinergic stimulation.  相似文献   

15.
The Na+-Ca2+ exchange mechanism in cardiac sarcolemmal vesicles can catalyze the exchange of Ca2+ on either side of the sarcolemmal membrane for Na+ on the opposing side. Little is known regarding the relative affinities of Na+ and Ca2+ for exchanger binding sites on the intra- and extracellular membrane surfaces. We have previously reported (Philipson, K.D. and Nishimoto, A.Y. (1982) J. Biol. Chem. 257, 5111-5117) a method for measuring the Na+-Ca2+ exchange of only the inside-out vesicles in a mixed population of sarcolemmal vesicles (predominantly right-side-out). We concluded that the apparent Km(Ca2+) for Na+i-dependent Ca2+ uptake was similar for inside-out and right-side-out vesicles. In the present study, we examine in detail Na+o-dependent Ca2+ efflux from both the inside-out and the total population of vesicles. To load vesicles with Ca2+ prior to measurement of Ca2+ efflux, four methods are used: 1, Na+-Ca2+ exchange; 2, passive Ca2+ diffusion; 3, ATP-dependent Ca2+ uptake; 4, exchange of Ca2+ for Na+ which has been actively transported into vesicles by the Na+ pump. The first two methods load all sarcolemmal vesicles with Ca2+, while the latter two methods selectively load inside-out vesicles with Ca2+. We are able to conclude that the dependence of Ca2+ efflux on the external Na+ concentration is similar in inside-out and right-side-out vesicles. Thus the apparent Km(Na+) values (approximately equal to 30 mM) of the Na+-Ca2+ exchanger are similar on the two surfaces of the sarcolemmal membrane. In other experiments, external Na+ inhibited the Na+i-dependent Ca2+ uptake of the total population of vesicles much more potently than that of the inside-out vesicles. Apparently Na+ can compete for the Ca2+ binding site more effectively on the external surface of right-side-out than on the external surface of inside-out vesicles. Thus, although affinities for Na+ or Ca2+ (in the absence of the other ion) appear symmetrical, the interactions between Na+ and Ca2+ at the two sides of the exchanger are not the same. The Na+-Ca2+ exchanger is not a completely symmetrical transport protein.  相似文献   

16.
Treatment of canine cardiac sarcolemmal vesicles with phospholipase D resulted in a large stimulation (up to 400%) of Na+-Ca2+ exchange activity. The phospholipase D treatment decreased the apparent Km (Ca2+) for the initial rate of Nai+-dependent Ca2+ uptake from 18.2 +/- 2.6 to 6.3 +/- 0.3 microM. The Vmax increased from 18.0 +/- 3.6 to 31.5 +/- 3.6 nmol of Ca2+/mg of protein/s. The effect was specific for Na+-Ca2+ exchange; other sarcolemmal transport enzymes ((Na+, K+)-ATPase; ATP-dependent Ca2+ transport) are inhibited by incubation with phospholipase D. Phospholipase D had little effect on the passive Ca2+ permeability of the sarcolemmal vesicles. After treatment with 0.4 unit/ml of phospholipase D (20 min, 37 degrees C), the sarcolemmal content of phosphatidic acid rose from 0.9 +/- 0.2 to 8.9 +/- 0.4%; simultaneously, Na+-Ca2+ exchange activity increased 327 +/- 87%. It is probable that the elevated phosphatidic acid level is responsible for the enhanced Na+-Ca2+ exchange activity. In a previous study (Philipson, K. D., Frank, J. S., and Nishimoto, A. Y. (1983) J. Biol. Chem. 258, 5905-5910), we hypothesized that negatively charged phospholipids were important in Na+-Ca2+ exchange, and the present results are consistent with this hypothesis. Stimulation of Na+-Ca2+ exchange by phosphatidic acid may be important in explaining the Ca2+ influx which accompanies the phosphatidylinositol turnover response which occurs in a wide variety of tissues.  相似文献   

17.
The modulation of rat brain Na(+)-Ca2+ exchange by K+   总被引:1,自引:0,他引:1  
The involvement of potassium ions in the Na(+)-Ca2+ exchange process was studied in rat brain synaptic plasma membrane (SPM) vesicles. Addition of equimolar [K+] to the intravesicular and the extravesicular medium led to a stimulation of the Na+ gradient-dependent Ca2+ influx; this stimulation was noticeable already at 0.5 mM and reached its maximum at 2 mM K+. The magnitude of the K+ stimulation was between 1.3-2.5-fold in different SPM preparations. K+ ions also stimulated the Na(+)-dependent Ca2+ efflux. K+ stimulation of Na(+)-Ca2+ exchange is of considerable specificity, since it is not mimicked by either Li+ or H+. The following lines of evidence suggest that K+ modulation of Na(+)-Ca2+ exchange involves the catalytic moiety of the transporter itself and not an unrelated K+ channel which modulates the membrane potential. 1) K+ stimulation of the transport process was conserved following reconstitution of the transporter into phospholipid-rich liposomes, an experimental condition which presumably separates the native membrane proteins among different vesicular structures. 2) K+ stimulation of Na+ gradient-dependent Ca2+ influx persists also when the build up of negative inside membrane potential is prevented by addition of carbonyl cyanide p-trifluoromethoxy phenylhydrazone which renders the membrane highly permeable to protons both in the native and the reconstituted preparation. 3) K+ stimulation of Na+ gradient-dependent Ca2+ influx is obtained also when tetraethylammonium chloride, 2,3-diaminopyridine and Cs+ are added to the Ca2+ uptake medium. Reconstituted SPM vesicles take up 86Rb+ in response to activation of Na+ gradient-dependent Ca2+ influx. The ratio of Ca2+ taken up by SPM vesicles in a Na+ gradient-dependent manner to the corresponding amounts of Rb+ taken up varies between 8-5 in different SPM preparations. If the stoichiometry of the process is 1 Rb+/1 Ca2+, then Rb+ cotransport is mediated by 10-20% of the transporters present in the preparation.  相似文献   

18.
A mechanism is developed for Na(+)-Ca2+ exchange using a new approach made possible by the availability of computer software that allows the systematic search of a large parameter space for optimum sets of parameters to fit multiple sets of experimental data. The approach was to make the experimental data dictate the form of the mechanism: the qualitative features of the data dictating the number and nature of the states of the exchanger and their interrelationship, and the quantitative aspects of the data dictating the values of the rate constants that govern the amount of each state relative to the total amount of exchanger. A single set of experimental data served this initial purpose, namely, observations of equilibrium Ca(2+)-Ca2+ exchange in cardiac sarcolemmal vesicles (Slaughter et al., 1983, J. biol. Chem. 258, 3183-3190). From this data a minimum mechanism was induced having 56 states (SYM56), which gave satisfactory quantitative fits to the experimental data. With this set of parameters additional experimental data were fitted, from the same preparation, the single cardiac cell and the squid giant axon, with some changes in parameters, but none dramatic. In spite of the symmetric nature of the mechanism, i.e. binding constants for Na+ and Ca2+ do not depend on the orientation of the binding sites, the mechanism exhibits marked asymmetric behavior similar to that observed experimentally. Finally, in accounting for Ca(2+)-Ca2+ exchange in the absence of monovalent cations, Ca2+ influx becomes dependent on intracellular Ca(2+)--an unexpected outcome--exactly in keeping with the "essential activator" role of intracellular Ca2+ observed by DiPolo & Beaugé (1987, J. gen. Physiol. 90, 505-525). Observations of Na(+)-Ca2+ exchange in the retinal rod outer segment are well fitted with a simplified version of SYM56 comprising 25 states (namely, SYM25), supporting the notion that the exchanger in the retinal rod outer segment differs from that in cardiac sarcolemma and squid axon. Maximum turnover rate of 840 sec-1 for SYM56 and 20 sec-1 for SYM25 are comparable to those reported for the exchanger in cardiac muscle and retinal rod outer segment, respectively.  相似文献   

19.
Ion transport and regulation of Na(+)-Ca(2+) exchange were examined for two alternatively spliced isoforms of the canine cardiac Na(+)-Ca(2+) exchanger, NCX1.1, to assess the role(s) of the mutually exclusive A and B exons. The exchangers examined, NCX1.3 and NCX1.4, are commonly referred to as the kidney and brain splice variants and differ only in the expression of the BD or AD exons, respectively. Outward Na(+)-Ca(2+) exchange activity was assessed in giant, excised membrane patches from Xenopus laevis oocytes expressing the cloned exchangers, and the characteristics of Na(+)(i)- (i.e., I(1)) and Ca(2+)(i)- (i.e., I(2)) dependent regulation of exchange currents were examined using a variety of experimental protocols. No remarkable differences were observed in the current-voltage relationships of NCX1.3 and NCX1.4, whereas these isoforms differed appreciably in terms of their I(1) and I(2) regulatory properties. Sodium-dependent inactivation of NCX1.3 was considerably more pronounced than that of NCX1.4 and resulted in nearly complete inhibition of steady state currents. This novel feature could be abolished by proteolysis with alpha-chymotrypsin. It appears that expression of the B exon in NCX1.3 imparts a substantially more stable I(1) inactive state of the exchanger than does the A exon of NCX1.4. With respect to I(2) regulation, significant differences were also found between NCX1.3 and NCX1.4. While both exchangers were stimulated by low concentrations of regulatory Ca(2+)(i), NCX1.3 showed a prominent decrease at higher concentrations (>1 microM). This does not appear to be due solely to competition between Ca(2+)(i) and Na(+)(i) at the transport site, as the Ca(2+)(i) affinities of inward currents were nearly identical between the two exchangers. Furthermore, regulatory Ca(2+)(i) had only modest effects on Na(+)(i)-dependent inactivation of NCX1.3, whereas I(1) inactivation of NCX1.4 could be completely eliminated by Ca(2+)(i). Our results establish an important role for the mutually exclusive A and B exons of NCX1 in modulating the characteristics of ionic regulation and provide insight into how alternative splicing tailors the regulatory properties of Na(+)-Ca(2+) exchange to fulfill tissue-specific requirements of Ca(2+) homeostasis.  相似文献   

20.
Saponins can both permeabilize cell plasma membranes and cause positive inotropic effects in isolated cardiac muscles. Different saponins vary in their relative abilities to cause each effect suggesting that different mechanisms of action may be involved. To investigate this possibility, we have compared the effects of seven different saponins on the passive Ca2+ permeability and Na+-Ca2+ exchange activity of isolated canine cardiac sarcolemmal membranes. Saponins having hemolytic activity reversibly increased the passive efflux of Ca2+ from sarcolemmal vesicles preloaded with 45Ca2+ with the following order of potency: echinoside-A greater than echinoside-B greater than holothurin-A greater than holothurin-B greater than sakuraso-saponin. Ginsenoside-Rd and desacyl-jego-saponin, which lack hemolytic activity, had no significant effect on this variable. The saponins also stimulated Na+-Ca2+ exchange activity measured as Na+-dependent Ca2+ uptake by sarcolemmal vesicles. Ginsenoside-Rd and desacyl-jego-seponin, which did not affect passive Ca2+ permeability, stimulated the uptake, while in contrast, echinoside-A and -B only slightly increased or decreased this latter variable. Thus, the abilities of these compounds to enhance Na+-Ca2+ exchange activity seem to be inversely related to their abilities to increase the Ca2+ permeability. Effects by the echinosides on Na+-Ca2+ exchange may be masked by the loss of Ca2+ from the vesicles due to the increased permeability. These results suggest that the saponins interact with membrane constituent(s) that can influence the passive Ca2+ permeability and the Na+-Ca2+ exchange activity of cardiac sarcolemmal membranes.  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