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 31 毫秒
1.
Levic S  Lv P  Yamoah EN 《PloS one》2011,6(12):e29005
Spontaneous action potentials have been described in developing sensory systems. These rhythmic activities may have instructional roles for the functional development of synaptic connections. The importance of spontaneous action potentials in the developing auditory system is underpinned by the stark correlation between the time of auditory system functional maturity, and the cessation of spontaneous action potentials. A prominent K(+) current that regulates patterning of action potentials is I(A). This current undergoes marked changes in expression during chicken hair cell development. Although the properties of I(A) are not normally classified as Ca(2+)-dependent, we demonstrate that throughout the development of chicken hair cells, I(A) is greatly reduced by acute alterations of intracellular Ca(2+). As determinants of spike timing and firing frequency, intracellular Ca(2+) buffers shift the activation and inactivation properties of the current to more positive potentials. Our findings provide evidence to demonstrate that the kinetics and functional expression of I(A) are tightly regulated by intracellular Ca(2+). Such feedback mechanism between the functional expression of I(A) and intracellular Ca(2+) may shape the activity of spontaneous action potentials, thus potentially sculpting synaptic connections in an activity-dependent manner in the developing cochlea.  相似文献   

2.
Regulation of TRPM2 by extra- and intracellular calcium   总被引:2,自引:0,他引:2       下载免费PDF全文
TRPM2 is a calcium-permeable nonselective cation channel that is opened by the binding of ADP-ribose (ADPR) to a C-terminal nudix domain. Channel activity is further regulated by several cytosolic factors, including cyclic ADPR (cADPR), nicotinamide adenine dinucleotide phosphate (NAADP), Ca(2+) and calmodulin (CaM), and adenosine monophosphate (AMP). In addition, intracellular ions typically used in patch-clamp experiments such as Cs(+) or Na(+) can alter ADPR sensitivity and voltage dependence, complicating the evaluation of the roles of the various modulators in a physiological context. We investigated the roles of extra- and intracellular Ca(2+) as well as CaM as modulators of ADPR-induced TRPM2 currents under more physiological conditions, using K(+)-based internal saline in patch-clamp experiments performed on human TRPM2 expressed in HEK293 cells. Our results show that in the absence of Ca(2+), both internally and externally, ADPR alone cannot induce cation currents. In the absence of extracellular Ca(2+), a minimum of 30 nM internal Ca(2+) is required to cause partial TRPM2 activation with ADPR. However, 200 microM external Ca(2+) is as efficient as 1 mM Ca(2+) in TRPM2 activation, indicating an external Ca(2+) binding site important for proper channel function. Ca(2+) facilitates ADPR gating with a half-maximal effective concentration of 50 nM and this is independent of extracellular Ca(2+). Furthermore, TRPM2 currents inactivate if intracellular Ca(2+) levels fall below 100 nM irrespective of extracellular Ca(2+). The facilitatory effect of intracellular Ca(2+) is not mimicked by Mg(2+), Ba(2+), or Zn(2+). Only Sr(2+) facilitates TRPM2 as effectively as Ca(2+), but this is due to Sr(2+)-induced Ca(2+) release from internal stores rather than a direct effect of Sr(2+) itself. Together, these data demonstrate that cytosolic Ca(2+) regulates TRPM2 channel activation. Its facilitatory action likely occurs via CaM, since the addition of 100 microM CaM to the patch pipette significantly enhances ADPR-induced TRPM2 currents at fixed [Ca(2+)](i) and this can be counteracted by calmidazolium. We conclude that ADPR is responsible for TRPM2 gating and Ca(2+) facilitates activation via calmodulin.  相似文献   

3.
Campos-Toimil M  Orallo F 《Life sciences》2007,80(23):2147-2153
The effects of (-)-epigallocatechin-3-gallate (EGCG), the most abundant catechin of tea, on Ca(2+)-permeable non-selective cation currents (NSCC) and voltage-operated Ca(2+) channels (VOCC) have been investigated in cultured rat aortic smooth muscle cells using the whole-cell voltage-clamp technique. Under the Cs(+)/tetraethylammonium (TEA)-containing internal solution, and in the presence of nifedipine (1 microM), EGCG (30 microM) activated a long-lasting inward current, with a reversal potential (E(rev)) of approximately 0 mV. This current was not significantly altered by the replacement of [Cl(-)](i) or [Cl(-)](o), implying that the inward current was not a chloride channel, but a NSCC. SKF 96365 (30 microM) and Cd(2+) (500 microM) almost completely abolished the EGCG-induced NSCC. A higher dose of EGCG (100 microM) additionally activated a nifedipine-sensitive inward current in the absence of depolarization protocol. EGCG (100 microM) also potentiated a nifedipine-sensitive voltage-dependent Ba(2+)-current during the first 5 min of incubation. However, after > 10 min of incubation with EGCG, this current was significantly inhibited. Our results suggest that EGCG caused a Ca(2+) influx into smooth muscle cells via VOCC (probably L-type) and other SKF-96365- and Cd(2+)-sensitive Ca(2+)-permeable channels. The action described here may be responsible for the contraction induced by EGCG in rat aortic rings and for the rise of the intracellular concentration of Ca(2+) in rat aortic smooth muscle cells evoked by this catechin. On the other hand, the inhibition of VOCC after > 10 min of incubation may be, in part, responsible for the relaxation of rat aorta induced by EGCG.  相似文献   

4.
The actions of crude venom from Anemesia species of spider were investigated in cultured dorsal root ganglion neurones from neonatal rats and hippocampal slices. Using mass spectrometry (MALDI-TOF MS), 10-12 distinct peptides with masses between about 3 and 10kDa were identified in the crude spider venom. At a concentration of 5 microg/ml crude Anemesia venom transiently enhanced the mean peak whole cell voltage-activated Ca(2+) current in a voltage-dependent manner and potentiated transient increases in intracellular Ca(2+) triggered by 30mM KCI as measured using Fura-2 fluorescence imaging. Additionally, 5-8 microg/ml Anemesia venom increased the amplitude of glutamatergic excitatory postsynaptic currents evoked in hippocampal slices. Omega-Conotoxin GVIA (1 microM) prevented the increase in voltage-activated Ca(2+) currents produced by Anemesia venom. This attenuation occurred when the cone shell toxin was applied before or after the spider venom. Anemesia venom (5 microg/ml) created no significant change in evoked action potentials but produced modest but significant inhibition of voltage-activated K(+) currents. At a concentration of 50 microg/ml Anemesia venom only produced reversible inhibitory effects, decreasing voltage-activated Ca(2+) currents. However, no significant effects on Ca(2+) currents were observed with a concentration of 0.5 microg/ml. The toxin(s) in the venom that enhanced Ca(2+) influx into sensory neurones was heat-sensitive and was made inactive by boiling or repetitive freeze-thawing. Boiled venom (5 microg/ml) produced significant inhibition of voltage-activated Ca(2+) currents and freeze-thawed venom inhibited Ca(2+) transients measured using Fura-2 fluorescence. Our data suggest that crude Anemesia venom contains components, which increased neuronal excitability and neurotransmission, at least in part this was mediated by enhancing Ca(2+) influx through N-type voltage-activated Ca(2+) channels.  相似文献   

5.
[Arg(8)]-vasopressin (AVP), at low concentrations (10-500 pM), stimulates oscillations in intracellular Ca(2+) concentration (Ca(2+) spikes) in A7r5 rat aortic smooth muscle cells. Our previous studies provided biochemical evidence that protein kinase C (PKC) activation and phosphorylation of voltage-sensitive K(+) (K(v)) channels are crucial steps in this process. In the present study, K(v) currents (I(Kv)) and membrane potential were measured using patch clamp techniques. Treatment of A7r5 cells with 100 pM AVP resulted in significant inhibition of I(Kv). This effect was associated with gradual membrane depolarization, increased membrane resistance, and action potential (AP) generation in the same cells. The AVP-sensitive I(Kv) was resistant to 4-aminopyridine, iberiotoxin, and glibenclamide but was fully inhibited by the selective KCNQ channel blockers linopirdine (10 microM) and XE-991 (10 microM) and enhanced by the KCNQ channel activator flupirtine (10 microM). BaCl(2) (100 microM) or linopirdine (5 microM) mimicked the effects of AVP on K(+) currents, AP generation, and Ca(2+) spiking. Expression of KCNQ5 was detected by RT-PCR in A7r5 cells and freshly isolated rat aortic smooth muscle. RNA interference directed toward KCNQ5 reduced KCNQ5 protein expression and resulted in a significant decrease in I(Kv) in A7r5 cells. I(Kv) was also inhibited in response to the PKC activator 4beta-phorbol 12-myristate 13-acetate (10 nM), and the inhibition of I(Kv) by AVP was prevented by the PKC inhibitor calphostin C (250 nM). These results suggest that the stimulation of Ca(2+) spiking by physiological concentrations of AVP involves PKC-dependent inhibition of KCNQ5 channels and increased AP firing in A7r5 cells.  相似文献   

6.
Snake vomeronasal receptor neurons in slice preparations were studied using the patch-clamp technique in the conventional and nystatin-perforated whole-cell configurations. The mean resting potential was approximately -70 mV; the average input resistance was 3 GOmega. Neurons required current injection of only 1-10 pA to display a variety of spiking patterns. Intracellular dialysis of 100 microM inositol 1,4,5-trisphosphate (IP(3)) evoked an inward current in 38% of neurons, with an average peak amplitude of 16.4 +/- 2.8 pA at a holding potential of -70mV. Application of 100 microM 3-deoxy-3-fluoro-D-myo-inositol 1,4,5-trisphosphate (F-IP(3)), a derivative of IP(3), also evoked an inward current in 4/8 (50%) neurons (32.6 +/- 58 pA at -70 mV, n = 4). The reversal potentials of the induced components were estimated to be -14 +/- 5 mV for IP(3) and -17 +/- 3 mV for F-IP(3). Bathing the neurons in 10 microM ruthenium red solution greatly reduced the IP(3)-evoked inward current to 1.6 +/- 1.1 pA at -70 mV (n = 6). With Cs(+)-containing internal solution, neither the Ca(2+)-ATPase inhibitor thapsigargin (1-50 microM) nor the Ca(2+)-ionophore ionomycin (10 microM) evoked a significant current response, suggesting that IP(3) can elicit current response in the neurons without mediation by intracellular Ca(2+) stores. Intracellular application of 1 mM cAMP evoked no detectable current response. Extracellular application of chemoattractant for snakes evoked a very large inward current. The reversal potential of the chemoattractant-induced current was similar to that of the IP(3)-induced current. The present results suggest that IP(3) may act as a second messenger in the transduction of chemoattractants in the garter snake vomeronasal organ.  相似文献   

7.
Two potential mechanisms by which the intracellular Ca(2 stores might modulate catecholamine release from bovine adrenal chromaffin cells were investigated: (i) that the cytosolic Ca(2+)transient caused by Ca(2+)release from the intracellular stores recruits additional chromaffin granules to a readily releasable pool that results in augmented catecholamine release when this is subsequently evoked, and (ii) that the Ca(2+)influx that follows depletion of intracellular stores (i.e. store-operated Ca(2+)entry) triggers release per se thereby augmenting evoked catecholamine release. When histamine or caffeine were applied in Ca(2+)-free perfusion media, a transient elevation of intracellular free Ca(2+)occurred owing to mobilization of Ca(2+)from the stores. When Ca(2+)was later readmitted to the perfusing fluid there followed a prompt and maintained rise in intracellular Ca(2+)concentrations of magnitude related to the degree of store mobilization. In parallel experiments, increased catecholamine secretion was measured under the conditions when Ca(2+)influx following store-mobilization occurred. Furthermore, the size of the catecholamine release increment correlated with the degree of Ca(2+)influx. Store-operated Ca(2+)entry evoked by mobilization with histamine and/or caffeine did not augment nicotine-evoked secretion per se; that is, it augmented evoked catecholamine release only to the extent that it increased basal catecholamine release. The nicotine-evoked catecholamine release was sensitive to cytosolic BAPTA, which, at the concentration used (50 microM BAPTA-AM), reduced release by approximately 25%. However, the increment in basal catecholamine release which followed Ca(2+)influx triggered by Ca(2+)store mobilization was not reduced by intracellular BAPTA. This finding is inconsistent with the hypothesis that the elevated cytosolic Ca(2+)from store mobilization recruits additional vesicles of catecholamine to the sub-plasmalemmal release sites to augment subsequently evoked secretion. This position is supported by the observation that histamine (10 microM) in Ca(2+)-free medium caused a pronounced elevation of cytosolic free Ca(2+), but this caused no greater catecholamine release when Ca(2+)was re-introduced than did prior exposure to Ca(2+)-free medium alone, which caused no elevation of cytosolic free Ca(2+). It is concluded that intracellular Ca(2+)stores can modulate secretion of catecholamines from bovine chromaffin cells by permitting Ca(2+)influx through a store-operated entry pathway. The results do not support the notion that the Ca(2+)released from intracellular stores plays a significant role in the recruitment of vesicles into the ready-release pool under the experimental conditions reported here.  相似文献   

8.
We examined the properties of voltage-dependent Ca(2+) channels (VDCCs) mediating 1-methyl-4-phenylpyridinium (MPP(+))-evoked [3H]DA release from rat striatal slices. In some cases, the Ca(2+)-independent efflux of neurotransmitters is mediated by the high-affinity neurotransmitter-uptake systems. To determine whether such a mechanism might be involved in MPP(+)-evoked [3H]DA release. MPP(+) (1,10 and 100 microM) evoked the release of [3H]DA from rat striatal slices in a concentration-dependent manner. In the absence of Ca(2+), MPP(+) (10 and 100 microM)-evoked [3H]DA release was significantly decreased to approximately 50% of control (a physiological concentration of Ca(2+)). In the presence of Ca(2+), nomifensine (0.1,1 and 10 microM) dose-dependently and significantly inhibited the MPP(+)-evoked release of [3H]DA. Nomifensine (1 and 10 microM) also dose-dependently and significantly inhibited the MPP(+)-evoked release of [3H]DA under Ca(2+)-free conditions. MPP(+)-evoked [3H]DA release was partly inhibited by nicardipine (1 and 10 microM), an L-type Ca(2+) channel blocker. On the other hand, the N-type Ca(2+) channel blocker omega-conotoxin-GVIA (omega-CTx-GVIA) (1 and 3 microM) did not affect this release. omega-agatoxin-IVA (omega-Aga-IVA) at low concentrations (0.1 microM), which are sufficient to block P-type Ca(2+) channels alone, also had no effect. On the other hand, MPP(+)-evoked [3H]DA release was significantly decreased by high concentrations of omega-Aga-IVA (0.3 microM) that would inhibit Q-type Ca(2+) channels. In addition, application of the Q-type Ca(2+) channel blocker omega-conotoxin-MVIIC (omega-CTx-MVIIC) (0.3 and 1 microM) also significantly inhibited MPP(+)-evoked [3H]DA release. These results suggest that MPP(+)-evoked [3H]DA release from rat striatal slices is largely mediated by Q-type Ca(2+) channels, and the Ca(2+)-independent component is mediated by reversal of the DA transport system.  相似文献   

9.
Microinjection of human Jurkat T-lymphocytes with nicotinic acid adenine dinucleotide phosphate (NAADP(+)) dose-dependently stimulated intracellular Ca(2+)-signaling. At a concentration of 10 nM NAADP(+) evoked repetitive and long-lasting Ca(2+)-oscillations of low amplitude, whereas at 50 and 100 nM, a rapid and high initial Ca(2+)-peak followed by trains of smaller Ca(2+)-oscillations was observed. Higher concentrations of NAADP(+) (1 and 10 microM) gradually reduced the initial Ca(2+)-peak, and a complete self-inactivation of Ca(2+)-signals was seen at 100 microM. The effect of NAADP(+) was specific as it was not observed with nicotinamide adenine dinucleotide phosphate. Both inositol 1,4, 5-trisphosphate- and cyclic adenosine diphosphoribose-mediated Ca(2+)-signaling were efficiently inhibited by coinjection of a self-inactivating concentration of NAADP(+). Most importantly, microinjection of a self-inactivating concentration of NAADP(+) completely abolished subsequent stimulation of Ca(2+)-signaling via the T cell receptor/CD3 complex, indicating that a functional NAADP(+) Ca(2+)-release system is essential for T-lymphocyte Ca(2+)-signaling.  相似文献   

10.
Nicotinic acid adenine dinucleotide phosphate (NAADP) has been shown to release intracellular Ca(2+) in several types of cells. We have used Ca(2+)-sensitive fluorescent dyes (Fura-2, Fluo-4) to measure intracellular Ca(2+) in astrocytes in culture and in situ. Bath-applied NAADP elicited a reversible and concentration-dependent Ca(2+) rise in up to 90% of astrocytes in culture (EC(50)=7 microM). The NAADP-evoked Ca(2+) rise was maintained in the absence of extracellular Ca(2+), but was suppressed after depleting the Ca(2+) stores of the ER with ATP (20 microM), with cyclopiazonic acid (10 microM) or with ionomycin (5 microM). P(2) receptor antagonist pyridoxalphosphate-6-azophenyl-2'4'-disulfonic acid (PPADS, 100 microM), IP(3) receptor blocker 2-aminoethoxydiphenyl borate (2-APB, 100 microM) and PLC inhibitor U73122 (10 microM) also reduced or suppressed the NAADP-evoked Ca(2+) rise. NAADP still evoked a Ca(2+) response after application of glycyl-l-phenylalanine-beta-naphthylamide (GPN, 200 microM), which permeabilizes lysosomes, or preincubation with H(+)-ATPase inhibitor bafilomycin A1 (4 microM) and of p-trifluoromethoxy carbonyl cyanide phenylhydrazone (FCCP, 2 microM), that impairs mitochondrial Ca(2+) handling. In acute brain slices, NAADP (10 microM) evoked Ca(2+) transients in cerebellar Bergmann glial cells and in hippocampal astrocytes. Our results suggest that NAADP recruits Ca(2+) from inositol 1,4,5-trisphosphate-sensitive Ca(2+) stores in mammalian astrocytes, at least partly by activating metabotropic P(2)Y receptors.  相似文献   

11.
We developed an improved mathematical model for a single primary pacemaker cell of the rabbit sinoatrial node. Original features of our model include 1) incorporation of the sustained inward current (I(st)) recently identified in primary pacemaker cells, 2) reformulation of voltage- and Ca(2+)-dependent inactivation of the L-type Ca(2+) channel current (I(Ca,L)), 3) new expressions for activation kinetics of the rapidly activating delayed rectifier K(+) channel current (I(Kr)), and 4) incorporation of the subsarcolemmal space as a diffusion barrier for Ca(2+). We compared the simulated dynamics of our model with those of previous models, as well as with experimental data, and examined whether the models could accurately simulate the effects of modulating sarcolemmal ionic currents or intracellular Ca(2+) dynamics on pacemaker activity. Our model represents significant improvements over the previous models, because it can 1) simulate whole cell voltage-clamp data for I(Ca,L), I(Kr), and I(st); 2) reproduce the waveshapes of spontaneous action potentials and ionic currents during action potential clamp recordings; and 3) mimic the effects of channel blockers or Ca(2+) buffers on pacemaker activity more accurately than the previous models.  相似文献   

12.
Modulation of the cloned human intermediate-conductance Ca(2+)-activated K(+) channel (hIK) by the compound 1-ethyl-2-benzimidazolinone (EBIO) was studied by patch-clamp technique using human embryonic kidney cells (HEK 293) stably expressing the hIK channels. In whole-cell studies, intracellular concentrations of free Ca(2+) were systematically varied, by buffering the pipette solutions. In voltage-clamp, the hIK specific currents increased gradually from 0 to approximately 300 pA/pF without reaching saturation even at the highest Ca(2+) concentration tested (300 nM). In the presence of EBIO (100 microM), the Ca(2+)-activation curve was shifted leftwards, and maximal currents were attained at 100 nM Ca(2+). In current-clamp, steeply Ca(2+)-dependent membrane potentials were recorded and the cells gradually hyperpolarised from -20 to -85 mV when Ca(2+) was augmented from 0 to 300 nM. EBIO strongly hyperpolarised cells buffered at intermediate Ca(2+) concentrations. In contrast, no effects were detected either below 10 nM (no basic channel activation) or at 300 nM Ca(2+) (V(m) close to E(K)). Without Ca(2+), EBIO-induced hyperpolarisations were not obtainable, indicating an obligatory Ca(2+)-dependent mechanism of action. When applied to inside-out patches, EBIO exerted a Ca(2+)-dependent increase in the single-channel open-state probability, showing that the compound modulates hIK channels by a direct action on the alpha-subunit or on a closely associated protein. In conclusion, EBIO activates hIK channels in whole-cell and inside-out patches by a direct mechanism, which requires the presence of internal Ca(2+).  相似文献   

13.
To clarify the mechanisms underlying the pancreatic β-cell response to varying glucose concentrations ([G]), electrophysiological findings were integrated into a mathematical cell model. The Ca(2+) dynamics of the endoplasmic reticulum (ER) were also improved. The model was validated by demonstrating quiescent potential, burst-interburst electrical events accompanied by Ca(2+) transients, and continuous firing of action potentials over [G] ranges of 0-6, 7-18, and >19 mM, respectively. These responses to glucose were completely reversible. The action potential, input impedance, and Ca(2+) transients were in good agreement with experimental measurements. The ionic mechanisms underlying the burst-interburst rhythm were investigated by lead potential analysis, which quantified the contributions of individual current components. This analysis demonstrated that slow potential changes during the interburst period were attributable to modifications of ion channels or transporters by intracellular ions and/or metabolites to different degrees depending on [G]. The predominant role of adenosine triphosphate-sensitive K(+) current in switching on and off the repetitive firing of action potentials at 8 mM [G] was taken over at a higher [G] by Ca(2+)- or Na(+)-dependent currents, which were generated by the plasma membrane Ca(2+) pump, Na(+)/K(+) pump, Na(+)/Ca(2+) exchanger, and TRPM channel. Accumulation and release of Ca(2+) by the ER also had a strong influence on the slow electrical rhythm. We conclude that the present mathematical model is useful for quantifying the role of individual functional components in the whole cell responses based on experimental findings.  相似文献   

14.
The goal of this study was to determine whether the protein kinase A (PKA) responsiveness of the cardiac L-type Ca(2+) current (ICa) is affected during transient increases in intracellular Ca(2+) concentration. Ventricular myocytes were isolated from 3- to 4-day-old neonatal rats and cultured on aligned collagen thin gels. When measured in 1 or 2 mM Ca(2+) external solution, the aligned myocytes displayed a large ICa that was weakly regulated (20% increase) during stimulation of PKA by 2 microM forskolin. In contrast, application of forskolin caused a 100% increase in ICa when the external Ca(2+) concentration was reduced to 0.5 mM or replaced with Ba(2+). This Ca(2+)-dependent inhibition was also observed when the cells were treated with 1 microM isoproterenol, 100 microM 3-isobutyl-1-methylxanthine, or 500 microM 8-bromo-cAMP. The responsiveness of ICa to PKA was restored during intracellular dialysis with a calmodulin (CaM) inhibitory peptide but not during treatment with inhibitors of protein kinase C, Ca(2+)/CaM-dependent protein kinase, or calcineurin. Adenoviral-mediated expression of a CaM molecule with mutations in all four Ca(2+)-binding sites also increased the PKA sensitivity of ICa. Finally, adult mouse ventricular myocytes displayed a greater response to forskolin and cAMP in external Ba(2+). Thus Ca(2+) entering the myocyte through the voltage-gated Ca(2+) channel regulates the PKA responsiveness of ICa.  相似文献   

15.
Quercetin is known to protect the cells suffering from oxidative stress. The oxidative stress elevates intracellular Ca(2+) concentration, one of the phenomena responsible for cell death. Therefore, we hypothesized that quercetin would protect the cells suffering from overload of intracellular Ca(2+). To test the hypothesis, the effects of quercetin on the cells suffering from oxidative stress and intracellular Ca(2+) overload were examined by using a flow cytometer with appropriate fluorescence probes (propidium iodide, fluo-3-AM, and annexin V-FITC) and rat thymocytes. The concentrations (1-30 microM) of quercetin to protect the cells suffering from intracellular Ca(2+) overload by A23187, a calcium ionophore, were similar to those for the cells suffering from oxidative stress by H(2)O(2). The cell death respectively induced by H(2)O(2) and A23187 was significantly suppressed by removal of external Ca(2+). Furthermore, quercetin greatly delayed the process of Ca(2+)-dependent cell death although it did not significantly affect the elevation of intracellular Ca(2+) concentration by H(2)O(2) and A23187, respectively. It is concluded that quercetin can protect the cells from oxidative injury in spite of increased concentration of intracellular Ca(2+). Results suggest that quercetin is also used for protection of cells suffering from overload of intracellular Ca(2+).  相似文献   

16.
Whole cell patch clamp and intracellular Ca(2+) transients in trout atrial cardiomyocytes were used to quantify calcium release from the sarcoplasmic reticulum (SR) and examine its dependency on the Ca(2+) trigger source. Short depolarization pulses (2-20 ms) elicited large caffeine-sensitive tail currents. The Ca(2+) carried by the caffeine-sensitive tail current after a 2-ms depolarization was 0.56 amol Ca(2+)/pF, giving an SR Ca(2+) release rate of 279 amol Ca(2+). pF(-1). s(-1) or 4.3 mM/s. Depolarizing cells for 10 ms to different membrane potentials resulted in a local maximum of SR Ca(2+) release, intracellular Ca(2+) transient, and cell shortening at 10 mV. Although 100 microM CdCl(2) abolished this local maximum, it had no effect on SR Ca(2+) release elicited by a depolarization to 110 or 150 mV, and the SR Ca(2+) release was proportional to the membrane potential in the range -50 to 150 mV with 100 microM CdCl(2). Increasing the intracellular Na(+) concentration ([Na(+)]) from 10 to 16 mM enhanced SR Ca(2+) release but reduced cell shortening at all membrane potentials examined. In the absence of TTX, SR Ca(2+) release was potentiated with 16 mM but not 10 mM pipette [Na(+)]. Comparison of the total sarcolemmal Ca(2+) entry and the Ca(2+) released from the SR gave a gain factor of 18.6 +/- 7.7. Nifedipine (Nif) at 10 microM inhibited L-type Ca(2+) current (I(Ca)) and reduced the time integral of the tail current by 61%. The gain of the Nif-sensitive SR Ca(2+) release was 16.0 +/- 4.7. A 2-ms depolarization still elicited a contraction in the presence of Nif that was abolished by addition of 10 mM NiCl(2). The gain of the Nif-insensitive but NiCl(2)-sensitive SR Ca(2+) release was 14.8 +/- 7.1. Thus both reverse-mode Na(+)/Ca(2+) exchange (NCX) and I(Ca) can elicit Ca(2+) release from the SR, but I(Ca) is more efficient than reverse-mode NCX in activating contraction. This difference may be due to extrusion of a larger fraction of the Ca(2+) released from the SR by reverse-mode NCX rather than a smaller gain for NCX-induced Ca(2+) release.  相似文献   

17.
Wang W  Hu GY  Wang YP 《Life sciences》2006,78(26):2989-2997
Magnesium lithospermate B (MLB) is the main water-soluble principle of Salviae Miltiorrhizae Radix (also called as 'Danshen' in the traditional Chinese medicine) for the treatment of cardiovascular diseases. MLB was found to possess a variety of pharmacological actions. However, it is unclear whether and how MLB affects the cardiac ion channels. In the present study, the effects of MLB on the voltage-activated ionic currents were investigated in single ventricular myocytes of adult guinea pigs. MLB reversibly inhibited L-type Ca(2+) current (I(Ca,L)). The inhibition was use-dependent and voltage-dependent (the IC(50) value of MLB was 30 microM and 393 microM, respectively, at the holding potential of -50 mV and -100 mV). In the presence of 100 microM MLB, both the activation and steady-state inactivation curves of I(Ca,L) were markedly shifted to hyperpolarizing membrane potentials, whereas the time course of recovery of I(Ca,L) from inactivation was not altered. MLB up to 300 microM had no significant effect on the fast-inactivating Na(+) current (I(Na)), delayed rectifier K(+) current (I(K)) and inward rectifier K(+) current (I(K1)). The results suggest that the voltage-dependent Ca(2+) antagonistic effect of MLB work in concert with its antioxidant action for attenuating heart ischemic injury.  相似文献   

18.
We investigated the actions of dantrolene Ca(2+)-induced on Ca(2+)-release (CICR) evoked by action potentials in cultured rat sensory neurons. The effect of dantrolene on action potential after-depolarization and voltage-activated calcium currents was studied in cultured neonatal rat dorsal root ganglion cells (DRG) using the whole-cell patch-clamp technique. Depolarizing current injection evoked action potentials and depolarizing after-potentials, which are activated as a result of CICR following a single action potential in some cells. The type of after-potentials was determined by inducing action potentials from the resting membrane potential. Extracellular application of dantrolene (10 microM) abolished after-depolarizations without affecting action potential properties. Furthermore, dantrolene significantly reduced repetitive action potentials after depolarizing current injection into these neurons, but had no significant effect on the steady-state current voltage relationship of calcium currents in these neurons. We conclude that dantrolene inhibits the induction of action potential after depolarizations by inhibiting CICR in cultured rat sensory neurons.  相似文献   

19.
The activity of the cardiac Na(+)/Ca(2+) exchanger (NCX1.1) undergoes continuous modulation during the contraction-relaxation cycle because of the accompanying changes in the electrochemical gradients for Na(+) and Ca(2+). In addition, NCX1.1 activity is also modulated via secondary, ionic regulatory mechanisms mediated by Na(+) and Ca(2+). In an effort to evaluate how ionic regulation influences exchange activity under pulsatile conditions, we studied the behavior of the cloned NCX1.1 during frequency-controlled changes in intracellular Na(+) and Ca(+) (Na(i)(+) and Ca(i)(2+)). Na(+)/Ca(2+) exchange activity was measured by the giant excised patch-clamp technique with conditions chosen to maximize the extent of Na(+)- and Ca(2+)-dependent ionic regulation so that the effects of variables such as pulse frequency and duration could be optimally discerned. We demonstrate that increasing the frequency or duration of solution pulses leads to a progressive decline in pure outward, but not pure inward, Na(+)/Ca(2+) exchange current. However, when the exchanger is permitted to alternate between inward and outward transport modes, both current modes exhibit substantial levels of inactivation. Changes in regulatory Ca(2+), or exposure of patches to limited proteolysis by alpha-chymotrypsin, reveal that this "coupling" is due to Na(+)-dependent inactivation originating from the outward current mode. Under physiological ionic conditions, however, evidence for modulation of exchange currents by Na(i)(+)-dependent inactivation was not apparent. The current approach provides a novel means for assessment of Na(+)/Ca(2+) exchange ionic regulation that may ultimately prove useful in understanding its role under physiological and pathophysiological conditions.  相似文献   

20.
We have studied the role of the actin cytoskeleton in bombesin-induced inositol 1,4,5-trisphosphate (IP(3))-production and Ca(2+)release in the pancreatic acinar tumour cell line AR4-2J. Intracellular and extracellular free Ca(2+)concentrations were measured in cell suspensions, using Fura-2. Disruption of the actin cytoskeleton by pretreatment of the cells with latrunculin B (10 microM), cytochalasin D (10 microM) or toxin B from Clostridium difficile (20 ng/ml) for 5-29 h led to inhibition of both, bombesin-stimulated IP(3)-production and Ca(2+)release. The toxins had no effect on binding of bombesin to its receptor, on Ca(2+)uptake into intracellular stores and on resting Ca(2+)levels. Ca(2+)mobilization from intracellular stores, induced by thapsigargin (100 nM) or IP(3)(1 microM) was not impaired by latrunculin B. In latrunculin B-pretreated cells inhibition of both, bombesin-stimulated IP(3)- production and Ca(2+)release was partly suspended in the presence of aluminum fluoride, an activator of G-proteins. Aluminum fluoride had no effect on basal IP(3)and Ca(2+)levels of control and toxin-pretreated cells. We conclude that disruption of the actin cytoskeleton impairs coupling of the bombesin receptor to its G-protein, resulting in inhibition of phospholipase C-activity with subsequent decreases in IP(3)-production and Ca(2+)release.  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