首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The present study addresses the effect of intracellular Na(+) and membrane potential on the binding of dopamine (DA) to the dopamine transporter (DAT). Perforation of plasma membranes of DAT-expressing cells with gramicidin diminished DA uptake and decreased the potency (increases K(i)) of DA in inhibiting the binding of cocaine analog [(3)H]2beta-carbomethoxy-3beta-(4-fluorophenyl)tropane (CFT). It also compromised the ability of external Na(+) to reduce DA K(i). No substantial effect on DA K(i) was observed upon gramicidin treatment in Na(+)-free buffer, membrane depolarization with high [K(+)](o), or elevation of [Na(+)](i) with monensin under non-depolarizing conditions. Elevation of DA K(i) was greater at more positive potentials when [Na(+)](i) was raised to a similar level, or at higher [Na(+)](i) when the membrane was depolarized to a similar level. In cells expressing D313N DAT, DA K(i) was significantly higher but less sensitive to gramicidin than that in wild-type (WT) cells. In contrast, DA K(i) in cell-free membranes was insensitive to Na(+), gramicidin, and D313N mutation. The data suggest that (i) intracellular Na(+) plays a role in affecting the external access to DA binding sites at DAT on depolarized plasma membranes of cells, and (ii) access to DA binding sites in cell-free membranes may occur from the intracellular side of the membrane. Unlike DA binding, CFT binding to both cells and membranes was sensitive to Na(+) and D313N mutation but insensitive to gramicidin, consistent with exclusively external access to sites that are different from but conformationally linked to those for DA.  相似文献   

2.
Phospholemman (PLM), when phosphorylated at serine 68, relieves its inhibition on Na(+)-K(+)-ATPase but inhibits Na(+)/Ca(2+) exchanger 1 (NCX1) in cardiac myocytes. Under stress when catecholamine levels are high, enhanced Na(+)-K(+)-ATPase activity by phosphorylated PLM attenuates intracellular Na(+) concentration ([Na(+)](i)) overload. To evaluate the effects of PLM on NCX1 on in vivo cardiac contractility, we injected recombinant adeno-associated virus (serotype 9) expressing either the phosphomimetic PLM S68E mutant or green fluorescent protein (GFP) directly into left ventricles (LVs) of PLM-knockout (KO) mice. Five weeks after virus injection, ~40% of isolated LV myocytes exhibited GFP fluorescence. Expression of S68E mutant was confirmed with PLM antibody. There were no differences in protein levels of α(1)- and α(2)-subunits of Na(+)-K(+)-ATPase, NCX1, and sarco(endo)plasmic reticulum Ca(2+)-ATPase between KO-GFP and KO-S68E LV homogenates. Compared with KO-GFP myocytes, Na(+)/Ca(2+) exchange current was suppressed, but resting [Na(+)](i), Na(+)-K(+)-ATPase current, and action potential amplitudes were similar in KO-S68E myocytes. Resting membrane potential was slightly lower and action potential duration at 90% repolarization (APD(90)) was shortened in KO-S68E myocytes. Isoproterenol (Iso; 1 μM) increased APD(90) in both groups of myocytes. After Iso, [Na(+)](i) increased monotonically in paced (2 Hz) KO-GFP but reached a plateau in KO-S68E myocytes. Both systolic and diastolic [Ca(2+)](i) were higher in Iso-stimulated KO-S68E myocytes paced at 2 Hz. Echocardiography demonstrated similar resting heart rate, ejection fraction, and LV mass between KO-GFP and KO-S68E mice. In vivo closed-chest catheterization demonstrated enhanced contractility in KO-S68E compared with KO-GFP hearts stimulated with Iso. We conclude that under catecholamine stress when [Na(+)](i) is high, PLM minimizes [Na(+)](i) overload by relieving its inhibition of Na(+)-K(+)-ATPase and preserves inotropy by simultaneously inhibiting Na(+)/Ca(2+) exchanger.  相似文献   

3.
Intracellular Na(+)-concentration, [Na(+)](i) modulates excitation-contraction coupling of cardiac myocytes via the Na(+)/Ca(2+) exchanger (NCX). In cardiomyocytes from rainbow trout (Oncorhyncus mykiss), whole cell patch-clamp studies have shown that Ca(2+) influx via reverse-mode NCX contributes significantly to contraction when [Na(+)](i) is 16 mM but not 10 mM. However, physiological [Na(+)](i) has never been measured. We recorded [Na(+)](i) using the fluorescent indicator sodium-binding benzofuran isophthalate in freshly isolated atrial and ventricular myocytes from rainbow trout. We examined [Na(+)](i) at rest and during increases in contraction frequency across three temperatures that span those trout experience in nature (7, 14, and 21 degrees C). Surprisingly, we found that [Na(+)](i) was not different between atrial and ventricular cells. Furthermore, acute temperature changes did not affect [Na(+)](i) in resting cells. Thus, we report a resting in vivo [Na(+)](i) of 13.4 mM for rainbow trout cardiomyocytes. [Na(+)](i) increased from rest with increases in contraction frequency by 3.2, 4.7, and 6.5% at 0.2, 0.5, and 0.8 Hz, respectively. This corresponds to an increase of 0.4, 0.6, and 0.9 mM at 0.2, 0.5, and 0.8 Hz, respectively. Acute temperature change did not significantly affect the contraction-induced increase in [Na(+)](i). Our results provide the first measurement of [Na(+)](i) in rainbow trout cardiomyocytes. This surprisingly high [Na(+)](i) is likely to result in physiologically significant Ca(2+) influx via reverse-mode NCX during excitation-contraction coupling. We calculate that this Ca(2+)-source will decrease with the action potential duration as temperature and contraction frequency increases.  相似文献   

4.
We have shown previously that partial inhibition of the cardiac myocyte Na(+)/K(+)-ATPase activates signal pathways that regulate myocyte growth and growth-related genes and that increases in intracellular Ca(2+) concentration ([Ca(2+)](i)) and reactive oxygen species (ROS) are two essential second messengers within these pathways. The aim of this work was to explore the relation between [Ca(2+)](i) and ROS. When myocytes were in a Ca(2+)-free medium, ouabain caused no change in [Ca(2+)](i), but it increased ROS as it did when the cells were in a Ca(2+)-containing medium. Ouabain-induced increase in ROS also occurred under conditions where there was little or no change in [Na(+)](i). Exposure of myocytes in Ca(2+)-free medium to monensin did not increase ROS. Increase in protein tyrosine phosphorylation, an early event induced by ouabain, was also independent of changes in [Ca(2+)](i) and [Na(+)](i). Ouabain-induced generation of ROS in myocytes was antagonized by genistein, a dominant negative Ras, and myxothiazol/diphenyleneiodonium, indicating a mitochondrial origin for the Ras-dependent ROS generation. These findings, along with our previous data, indicate that increases in [Ca(2+)](i) and ROS in cardiac myocytes are induced by two parallel pathways initiated at the plasma membrane: One being the ouabain-altered transient interactions of a fraction of the Na(+)/K(+)-ATPase with neighboring proteins (Src, growth factor receptors, adaptor proteins, and Ras) leading to ROS generation, and the other, inhibition of the transport function of another fraction of the Na(+)/K(+)-ATPase leading to rise in [Ca(2+)](i). Evidently, the gene regulatory effects of ouabain in cardiac myocytes require the downstream collaborations of ROS and [Ca(2+)](i).  相似文献   

5.
Recently, we reported that ouabain kills renal epithelial and vascular endothelial cells independently of elevation of the [Na(+)](i)/[K(+)](i) ratio. These observations raised the possibility of finding cardiotonic steroids (CTS) that inhibit the Na(+),K(+) pump without attenuating cell survival and vice versa. To test this hypothesis, we compared CTS action on Na(+),K(+) pump, [Na(+)](i) content, and survival of Madin-Darby canine kidney cells. At a concentration of 1 microM, ouabain and other tested cardenolides, as well as bufadienolides such as bufalin, cinobufagin, cinobufotalin, and telobufotoxin, led to approximately 10-fold inhibition of the Na(+),K(+) pump, a 2-3-fold decrease in staining with dimethylthiazol-diphenyltetrazolium (MTT), and massive death indicated by detachment of approximately 80% of cells and caspase-3 activation. In contrast, Na(+),K(+) pump inhibition and elevation of [Na(+)](i) seen in the presence of 3 microM marinobufagenin (MBG) and marinobufotoxin did not affect MTT staining and cell survival. Inhibition of the Na(+),Rb(+) pump in K(+)-free medium was not accompanied by a decline of MTT staining and cell detachment but increased sensitivity to CTS. In K(+)-free medium, half-maximal inhibition of (86)Rb influx was observed in the presence of 0.04 microM ouabain and 0.1 microM MBG, whereas half-maximal detachment and decline of MTT staining were detected at 0.03 and 0.004 microM of ouabain versus 10 and 3 microM of MBG, respectively. Both ouabain binding and ouabain-induced [Na(+)](i),[K(+)](i)-independent signaling were suppressed in the presence of MBG. Thus, our results show that CTS exhibit distinctly different potency in Na(+),K(+) pump inhibition and triggering of [Na(+)](i)/[K(+)](i)-independent signaling, including cell death.  相似文献   

6.
In whole-cell recordings from HEK293 cells stably transfected with the delayed rectifier K(+) channel Kv2.1, long depolarizations produce current-dependent changes in [K(+)](i) that mimic inactivation and changes in ion selectivity. With 10 mM K(o)(+) or K(i)(+), and 140-160 mM Na(i,o)(+), long depolarizations shifted the reversal potential (V(R)) toward E(Na). However, similar shifts in V(R) were observed when Na(i,o)(+) was replaced with N-methyl-D-glucamine (NMG(+))(i, o). In that condition, [K(+)](o) did not change significantly, but the results could be quantitatively explained by changes in [K(+)](i). For example, a mean outward K(+) current of 1 nA for 2 s could decrease [K(+)](i) from 10 mM to 3 mM in a 10 pF cell. Dialysis by the recording pipette reduced but did not fully prevent changes in [K(+)](i). With 10 mM K(i,o)(+), 150 mM Na(i)(+), and 140 mM NMG(o)(+), steps to +20 mV produced a positive shift in V(R), as expected from depletion of K(i)(+), but opposite to the shift expected from a decreased K(+)/Na(+) selectivity. Long steps to V(R) caused inactivation, but no change in V(R). We conclude that current-dependent changes in [K(+)](i) need to be carefully evaluated when studying large K(+) currents in small cells.  相似文献   

7.
Hypertrophy and failure (H/F) in humans and large mammals are characterized by a change from a positive developed force-frequency relationship (+FFR) in normal myocardium to a flattened or negative developed force-frequency relationship (-FFR) in disease. Altered Ca(2+) homeostasis underlies this process, but the role of intracellular Na(+) concentration ([Na(+)](i)) in H/F and frequency-dependent contractility reserve is unclear. We hypothesized that altered [Na(+)](i) is central to the -FFR response in H/F feline myocytes. Aortic constriction caused left ventricular hypertrophy (LVH). We found that as pacing rate was increased, contraction magnitude was maintained in isolated control myocytes (CM) but decreased in LVH myocytes (LVH-M). Quiescent LVH-M had higher [Na(+)](i) than CM (LVH-M 13.3 +/- 0.3 vs. CM 8.9 +/- 0.2 mmol/l; P < 0.001) with 0.5-Hz pacing (LVH-M 14.9 +/- 0.5 vs. CM 10.8 +/- 0.4 mmol/l; P < 0.001) but were not different at 2.5 Hz (17.0 +/- 0.7 vs. control 16.0 +/- 0.7 mmol/l; not significant). [Na(+)](i) was altered by patch pipette dialysis to define the effect of [Na(+)](i) on contraction magnitude and action potential (AP) wave shape at slow and fast pacing rates. Using AP clamp, we showed that LVH-M require increased [Na(+)](i) and long diastolic intervals to maintain normal shortening. Finally, we determined the voltage dependence of contraction for Ca(2+) current (I(Ca))-triggered and Na(+)/Ca(2+) exchanger-mediated contractions and showed that there is a greater [Na(+)](i) dependence of contractility in LVH-M. These data show that increased [Na(+)](i) is essential for maintaining contractility at slow heart rates but contributes to small contractions at fast rates unless rate-dependent AP shortening is prevented, suggesting that altered [Na(+)](i) regulation is a critical contributor to abnormal contractility in disease.  相似文献   

8.
The aim of this study was to investigate whether or not the activity of the cardiac Na(+)-Ca(2+) exchanger might be directly sensitive to external K(+) concentration ([K(+)](e)). Measurements of whole-cell exchanger current (I(NaCa)) were made at 37 degrees C from guinea-pig isolated ventricular myocytes, using whole-cell patch clamp recording with major interfering conductances blocked. Changing [K(+)](e) from 0 to 5mM significantly reduced both outward and inward exchange currents in a time-dependent manner. Various [K(+)](e) between 1 and 15 mM were tested and the inhibitory effect was observed to be concentration-dependent. At steady-state, 5mM [K(+)](e) decreased the density of Ni(2+)-sensitive current by 52.8+/-4.3% (mean+/-S.E.M., n=6) and of 0Na0Ca-sensitive current by 39.0+/-4.4% (n=5). The possibility that the inhibitory effect of external K(+) on I(NaCa) might wholly or in part be secondary to activation of the sarcolemmal Na(+)-K(+) pump was investigated by testing the effect of K(+) addition in the presence of a high concentration of strophanthidin (500 microM). Ni(2+)-sensitive I(NaCa) was still observed to be sensitive to external K(+) (I(NaCa) decreased by 39.4+/-9.4%, n=4), suggesting that the inhibitory effect could occur independently of activation of the Na(+)-K(+) pump. The effect of external K(+) on I(NaCa) was verified using a baby hamster kidney (BHK) cell line stably expressing the cardiac Na(+)-Ca(2+) exchanger isoform, NCX1. Similar to native I(NaCa), NCX1 current was also suppressed by [K(+)](e). However, [K(+)](e) did not alter current amplitude in untransfected BHK cells. The effect of [K(+)](e) on I(NaCa) could not be attributed to simply adding any monovalent cation back to the external solution, since it was not reproduced by application of equimolar Li(+), Cs(+) and TEA(+). Rb(+), however, could mimic the effect of K(+). Collectively, these data suggest that external K(+) at physiologically and pathologically relevant concentrations might be able to modulate directly the activity of the cardiac Na(+)-Ca(2+) exchanger.  相似文献   

9.
Hille C  Walz B 《Cell calcium》2006,39(4):305-311
Stimulation with the neurotransmitter dopamine causes an amplitude-modulated increase in the intracellular Ca(2+) concentration ([Ca(2+)](i)) in epithelial cells of the ducts of cockroach salivary glands. This is completely attributable to a Ca(2+) influx from the extracellular space. Additionally, dopamine induces a massive [Na(+)](i) elevation via the Na(+)K(+)2Cl(-) cotransporter (NKCC). We have reasoned that Ca(2+)-entry is mediated by the Na(+)Ca(2+) exchanger (NCE) operating in the Ca(2+)-entry mode. To test this hypothesis, [Ca(2+)](i) and [Na(+)](i) were measured by using the fluorescent dyes Fura-2, Fluo-3, and SBFI. Inhibition of Na(+)-entry from the extracellular space by removal of extracellular Na(+) or inhibition of the NKCC by 10 microM bumetanide did not influence resting [Ca(2+)](i) but completely abolished the dopamine-induced [Ca(2+)](i) elevation. Simultaneous recordings of [Ca(2+)](i) and [Na(+)](i) revealed that the dopamine-induced [Na(+)](i) elevation preceded the [Ca(2+)](i) elevation. During dopamine stimulation, the generation of an outward Na(+) concentration gradient by removal of extracellular Na(+) boosted the [Ca(2+)](i) elevation. Furthermore, prolonging the dopamine-induced [Na(+)](i) rise by blocking the Na(+)/K(+)-ATPase reduced the recovery from [Ca(2+)](i) elevation. These results indicate that dopamine induces a massive NKCC-mediated elevation in [Na(+)](i), which reverses the NCE activity into the reverse mode causing a graded [Ca(2+)](i) elevation in the duct cells.  相似文献   

10.
Myocellular sodium homeostasis is commonly disrupted during critical illness for unknown reasons. Recent data suggest that changes in intracellular sodium content and the amount of ATP provided by glycolysis are closely related. The role of glycolysis and oxidative phosphorylation in providing fuel to the Na(+)-K(+) pump was investigated in resting rat extensor digitorum longus muscles incubated at 30 degrees C for 1 h. Oxidative inhibition with carbonyl cyanide m-chlorophenylhydrazone, known as CCCP (0.2 microM), or by hypooxygenation did not alter myocellular sodium or potassium content ([Na(+)](i), [K(+)](i), respectively), whereas treatment with iodoacetic acid (0.3 mM), which effectively blocked glycolysis, dramatically increased [Na(+)](i) and the [Na(+)](i)/[K(+)](i) ratio. Experiments using ouabain and measurements of myocellular high-energy phosphates indicate that Na(+)-K(+)-ATPase activity is only impaired when glycolysis is inhibited. The data suggest that normal glycolysis is required to regulate intracellular sodium in fast-twitch skeletal muscles, because it is the predominant source of the fuel for the Na(+)-K(+)-ATPase.  相似文献   

11.
The role of Na(+), K(+), Cl(-)-cotransport (NKCC) in apoptosis of HepG2 human hepatoblastoma cells was investigated. Pinacidil (Pin), an activator of ATP-sensitive K(+) (K(ATP)) channels, induced apoptosis in a dose- and time-dependent manner in HepG2 cells. Pin increased intracellular K(+) concentration ([K(+)](i)). Bumetanide and furosemide, NKCC inhibitors, significantly inhibited the Pin-induced increased [K(+)](i) and apoptosis, whereas K(ATP) inhibitors (glibenclamide and tolbutamide) had no effects. The Pin-induced [K(+)](i) increase was significantly prevented by reducing extracellular Cl(-) concentration, and Pin also increased intracellular Na(+) concentration ([Na(+)](i)), further indicating that these effects of Pin may be due to NKCC activation. In addition, Pin induced a rapid and sustained increase in intracellular Ca(2+) concentration ([Ca(2+)](i)), which was completely prevented by the NKCC inhibitors. Treatment with EGTA or BAPTA/AM markedly inhibited the Pin-induced apoptosis. Inhibitors of Na(+), Ca(2+)-exchanger, bepridil, and benzamil significantly prevented both [Ca(2+)](i) increase and apoptosis induced by Pin. Taken together, these results suggest that Pin increases [Na(+)](i) through NKCC activation, which leads to stimulation of reverse-mode of Na(+), Ca(2+) exchanger, resulting in [Ca(2+)](i) increase, and in turn, apoptosis. These results further suggest that NKCC may be a good target for induction of apoptosis in human hepatoma cells.  相似文献   

12.
13.
The antibiotic Beauvericin (BEA) was previously shown to express ionophoric properties under simple experimental systems. Its channel-forming activity was examined in inside-out patches of ventricular myocytes and synthetic membranes with the patch clamp and fluorescence imaging techniques. Current transitions to several open state levels were evident after wash-in. The BEA channel is cation-selective. Conductance and kinetics are presented for K+ and Na+ substates and main states. The pore was blocked by La3+. In myocytes, the [K+]i was reduced, while [Na+]i and [Ca2+]i increased, leading to cytolysis. These results indicate that BEA forms cation-selective channels in lipid membranes, which can affect the ionic homeostasis.  相似文献   

14.
In patients with hyperkalemic periodic paralysis (HyperKPP), attacks of muscle weakness or paralysis are triggered by K(+) ingestion or rest after exercise. Force can be restored by muscle work or treatment with β(2)-adrenoceptor agonists. A missense substitution corresponding to a mutation in the skeletal muscle voltage-gated Na(+) channel (Na(v)1.4, Met1592Val) causing human HyperKPP was targeted into the mouse SCN4A gene (mutants). In soleus muscles prepared from these mutant mice, twitch, tetanic force, and endurance were markedly reduced compared with soleus from wild type (WT), reflecting impaired excitability. In mutant soleus, contractility was considerably more sensitive than WT soleus to inhibition by elevated [K(+)](o). In resting mutant soleus, tetrodotoxin (TTX)-suppressible (22)Na uptake and [Na(+)](i) were increased by 470 and 58%, respectively, and membrane potential was depolarized (by 16 mV, P < 0.0001) and repolarized by TTX. Na(+),K(+) pump-mediated (86)Rb uptake was 83% larger than in WT. Salbutamol stimulated (86)Rb uptake and reduced [Na(+)](i) both in mutant and WT soleus. Stimulating Na(+),K(+) pumps with salbutamol restored force in mutant soleus and extensor digitorum longus (EDL). Increasing [Na(+)](i) with monensin also restored force in soleus. In soleus, EDL, and tibialis anterior muscles of mutant mice, the content of Na(+),K(+) pumps was 28, 62, and 33% higher than in WT, respectively, possibly reflecting the stimulating effect of elevated [Na(+)](i) on the synthesis of Na(+),K(+) pumps. The results confirm that the functional disorders of skeletal muscles in HyperKPP are secondary to increased Na(+) influx and show that contractility can be restored by acute stimulation of the Na(+),K(+) pumps. Calcitonin gene-related peptide (CGRP) restored force in mutant soleus but caused no detectable increase in (86)Rb uptake. Repeated excitation and capsaicin also restored contractility, possibly because of the release of endogenous CGRP from nerve endings in the isolated muscles. These observations may explain how mild exercise helps locally to prevent severe weakness during an attack of HyperKPP.  相似文献   

15.
Previous work demonstrated that Na(+)-K(+) pump activity within skeletal muscle is attenuated in myocardial infarction (MI). This may lead to enhanced interstitial K(+) concentration ([K(+)](o)) in the muscle. We tested the hypothesis that [K(+)](o) rises with muscle contraction and that, in rats with MI, the rate of rise in [K(+)](o) is greater than it is in control animals. Microdialysis probes were inserted in the skeletal muscle of six healthy control and six MI rats. The ends of the probes were then attached to the K(+) electrodes, and [K(+)](o) was continuously measured. Muscle contraction was induced by electrical stimulation of the sciatic nerves for 1 min. Stimulation at 1 and 3 Hz increased muscle [K(+)](o) by 14.2% and 44.7% in controls and by 22.9% and 62.8% in MI rats (P < 0.05 vs. controls), respectively. When ouabain, an inhibitor of Na(+)-K(+) pump, was added to the perfusate, muscle [K(+)](o) rose significantly. This effect of ouabain was significantly attenuated in MI animals. In conclusion, when compared with that in control animals, an increase of [K(+)](o) in exercising muscle is augmented in MI rats, likely due to an attenuation of Na(+)-K(+) pump activity.  相似文献   

16.
There are well-documented differences in ion channel activity and action potential shape between epicardial (EPI), midmyocardial (MID), and endocardial (ENDO) ventricular myocytes. The purpose of this study was to determine if differences exist in Na/K pump activity. The whole cell patch-clamp was used to measure Na/K pump current (I(P)) and inward background Na(+)-current (I(inb)) in cells isolated from canine left ventricle. All currents were normalized to membrane capacitance. I(P) was measured as the current blocked by a saturating concentration of dihydro-ouabain. [Na(+)](i) was measured using SBFI-AM. I(P)(ENDO) (0.34 +/- 0.04 pA/pF, n = 17) was smaller than I(P)(EPI) (0.68 +/- 0.09 pA/pF, n = 38); the ratio was 0.50 with I(P)(MID) being intermediate (0.53 +/- 0.13 pA/pF, n = 19). The dependence of I(P) on [Na(+)](i) or voltage was essentially identical in EPI and ENDO (half-maximal activation at 9-10 mM [Na(+)](i) or approximately -90 mV). Increasing [K(+)](o) from 5.4 to 15 mM caused both I(P)(ENDO) and I(P)(EPI) to increase, but the ratio remained approximately 0.5. I(inb) in EPI and ENDO were nearly identical ( approximately 0.6 pA/pF). Physiological [Na(+)](i) was lower in EPI (7 +/- 2 mM, n = 31) than ENDO (12 +/- 3 mM, n = 29), with MID being intermediate (9 +/- 3 mM, n = 22). When cells were paced at 2 Hz, [Na(+)](i) increased but the differences persisted (ENDO 14 +/- 3 mM, n = 10; EPI 9 +/- 2 mM, n = 10; and MID intermediate, 11 +/- 2 mM, n = 9). Based on these results, the larger I(P) in EPI appears to reflect a higher maximum turnover rate, which implies either a larger number of active pumps or a higher turnover rate per pump protein. The transmural gradient in [Na(+)](i) means physiological I(P) is approximately uniform across the ventricular wall, whereas transporters that utilize the transmembrane electrochemical gradient for Na(+), such as Na/Ca exchange, have a larger driving force in EPI than ENDO.  相似文献   

17.
The effect of eccentric contractions on intracellular Na(+) concentration ([Na(+)](i)) and its distribution were examined in isolated rat and mouse muscle fiber bundles. [Na(+)](i) was measured with either Na(+)-binding benzofuran isophthalate or sodium green. Ten isometric contractions had no significant effect on force (measured after 5 min of recovery) and caused no significant change in the resting [Na(+)](i) (7.2 +/- 0.5 mM). In contrast 10 eccentric contractions (40% stretch at 4 muscle lengths/s) reduced developed force at 100 Hz to 45 +/- 3% of control and increased [Na(+)](i) to 16.3 +/- 1.6 mM (n = 6; P < 0.001). The rise of [Na(+)](i) occurred over 1-2 min and showed only minimal recovery after 30 min. Confocal images of the distribution of [Na(+)](i) showed a spatially uniform distribution both at rest and after eccentric contractions. Gd(3+) (20 microM) had no effect on resting [Na(+)](i) or control tetanic force but prevented the rise of [Na(+)](i) and reduced the force deficit after eccentric damage. These data suggest that Na(+) entry after eccentric contractions may occur principally through stretch-sensitive channels.  相似文献   

18.
In cardiomyocytes, a major decrease in the level of sarco/endoplasmic reticulum Ca(2+) ATPase (SERCA) can severely impair systolic and diastolic functions. In mice with cardiomyocyte-specific conditional excision of the Serca2 gene (SERCA2 KO), end-stage heart failure developed between four and seven weeks after gene deletion combined with [Na(+)](i) elevation and intracellular acidosis. In this study, to investigate the underpinning changes in Ca(2+) dynamics and metabolic homeostasis, we developed data-driven mathematical models of Ca(2+) dynamics in the ventricular myocytes of the control, four-week, and seven-week SERCA2 knockout (KO) mice. The seven-week KO model showed that elevated [Na(+)](i) was due to increased Na(+) influxes through the Na(+)/Ca(2+) exchanger (NCX) and the Na(+)/H(+) exchanger, with the latter exacerbated by intracellular acidosis. Furthermore, NCX upregulation in the seven-week KO model resulted in increased ATP consumption for ion transport. Na(+) accumulation in the SERCA KO due to NCX upregulation and intracellular acidosis potentially play a role in the development of heart failure, by initiating a reinforcing cycle involving: a mismatch between ATP demand and supply; an increasingly compromised metabolism; a decreased pH(i); and, finally, an even greater [Na(+)](i) elevation.  相似文献   

19.
Immunological stimulation of rat mucosal-type mast cells (RBL-2H3 line) by clustering of their Fcepsilon receptors (FcepsilonRI) causes a rapid and transient increase in free cytoplasmic Ca(2+) ion concentration ([Ca(2+)](i)) because of its release from intracellular stores. This is followed by a sustained elevated [Ca(2+)](i), which is attained by Ca(2+) influx. Because an FcepsilonRI-induced increase in the membrane permeability for Na(+) ions has also been observed, and secretion is at least partially inhibited by lowering of extracellular sodium ion concentrations ([Na(+)](o)), the operation of a Na(+)/Ca(2+) exchanger has been considered. We found significant coupling between the Ca(2+) and Na(+) ion gradients across plasma membranes of RBL-2H3 cells, which we investigated employing (23)Na-NMR, (45)Ca(2+), (85)Sr(2+), and the Ca(2+)-sensitive fluorescent probe indo-1. The reduction in extracellular Ca(2+) concentrations ([Ca(2+)](o)) provoked a [Na(+)](i) increase, and a decrease in [Na(+)](o) results in a Ca(2+) influx as well as an increase in [Ca(2+)](i). Mediator secretion assays, monitoring the released beta-hexosaminidase activity, showed in the presence of extracellular sodium a sigmoidal dependence on [Ca(2+)](o). However, the secretion was not affected by varying [Ca(2+)](o) as [Na(+)](o) was lowered to 0.4 mM, while it was almost completely inhibited at [Na(+)](o) = 136 mM and [Ca(2+)](o) < 0.05 mM. Increasing [Na(+)](o) caused the secretion to reach a minimum at [Na(+)](o) = 20 mM, followed by a steady increase to its maximum value at 136 mM. A parallel [Na(+)](o) dependence of the Ca(2+) fluxes was observed: Antigen stimulation at [Na(+)](o) = 136 mM caused a pronounced Ca(2+) influx. At [Na(+)](o) = 17 mM only a slight Ca(2+) efflux was detected, whereas at [Na(+)](o) = 0.4 mM no Ca(2+) transport across the cell membrane could be observed. Our results clearly indicate that the [Na(+)](o) dependence of the secretory response to FcepsilonRI stimulation is due to its influence on the [Ca(2+)](i), which is mediated by a Na(+)-dependent Ca(2+) transport.  相似文献   

20.
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   

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