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 93 毫秒
1.
Loading of mouse motor nerve terminals with EGTA-AM, but not with BAPTA-AM, inhibited the release of the neurotransmitter in response to stimulation of the nerve with rare (0.3 Hz) “single” pulses. During rhythmic stimulation with short (50 EPP) high-frequency (20 Hz) series, BAPTA-AM buffer modified burst pattern in a dose-dependent manner: it replaced the phase of initial facilitation by persistent depression of secretion and decreased its plateau level at the end of the burst. In contrast, loading of the nerve terminals with EGTA-AM buffer produced no effect on the phase of initial facilitation, but decreased the plateau level to the same degree as BAPTA-AM did. Probably, the different effects of both buffers on secretion of neurotransmitter reflect peculiarities of involvement of fast and slow Ca2+ signals of motor terminals in single and rhythmic release of the neurotransmitter.  相似文献   

2.
Protein kinase C blocker chelerythrine prevented the increase in quantal content of single and rhythmic evoked end-plate potentials after disinhibition of L-type Ca(2+)-channels with paxillin. Phorbol ester increased quantal content of single end-plate potentials and changed rhythmic activity of mouse motor synapses. The effects of phorbol ester were to a great extent neutralized by L-type Ca(2+)-channel blocker nitrendipine and were completely abolished by K(+)-channel blocker 4-aminopyridine. Thus, we discovered different facilitations of transmission after protein kinase C activation with calcium current through L-type channels and with phorbol ester.  相似文献   

3.
Toosendanin (TSN) is a triterpenoid extracted from Melia toosendan Sieb et Zucc, which was used as a digestive tract-parasiticide and agricultural insecticide in ancient China. TSN was demonstrated to be a selective presynaptic blocker and an effective antibotulismic agent. By interfering with neurotransmitter release through an initial facilitation followed by a subsequent depression, TSN eventually blocks synaptic transmission at both the neuro-muscular junction and central synapses. Despite sharing some similar actions with botulinum neurotoxin (BoNT), TSN has a marked antibotulismic effect in vivo and in vitro. Studies suggest that the antibotulismic effect of TSN is achieved by preventing BoNT from approaching its enzymatic substrate, the SNARE protein. It is also found that TSN can induce differentiation and apoptosis in several cell lines, and suppress proliferation of various human cancer cells. TSN inhibits various K(+)-channels, selectively facilitates Ca(2+)-influx via L-type Ca(2+) channels and increases intracellular Ca(2+) concentration ([Ca(2+)](i)). The TSN-induced [Ca(2+)](i) increase and overload could be responsible for the TSN-induced biphasic effect on transmitter release, cell differentiation, apoptosis as well as the cytoxicity of TSN.  相似文献   

4.
Electrophysiological and optical methods were used to study exo- and endocytosis of synaptic vesicles underlying secretion of the neurotransmitter from motor nerve terminals in frog sternocutaneous muscle. Increase in extracellular concentration of K+ or sucrose produced similar increase in the frequency of miniature endplate currents recorded by extracellular microelectrode. Fluorescent microscopy revealed bright spots in nerve terminal during stimulation of secretion with high-potassium solutions in the presence of endocytosis marker FM1-43. These spots corresponded to clusters of synaptic vesicles that passed through the cycles of exo- and endocytosis. Subsequent high-potassium stimulation of exocytosis in normal Ringer solution led to disappearance of marker spots, while in hyperosmotic saline the spots were preserved. No spots were seen after stimulation of neurotransmitter secretion with sucrose in the presence of FM1-43. It is concluded that quantal secretion of the neurotransmitter in frog motor nerve endings can be realized via both complete exocytosis of synaptic vesicles with subsequent endocytosis and kiss-and-run mechanism with the formation of a temporary pore.  相似文献   

5.
Precise localization of various ion channels into proper subcellular compartments is crucial for neuronal excitability and synaptic transmission. Axonal K(+) channels that are activated by depolarization of the membrane potential participate in the repolarizing phase of the action potential, and hence regulate action potential firing patterns, which encode output signals. Moreover, some of these channels can directly control neurotransmitter release at axonal terminals by constraining local membrane excitability and limiting Ca(2+) influx. K(+) channels differ not only in biophysical and pharmacological properties, but in expression and subcellular distribution as well. Importantly, proper targeting of channel proteins is a prerequisite for electrical and chemical functions of axons. In this review, we first highlight recent studies that demonstrate different roles of axonal K(+) channels in the local regulation of axonal excitability. Next, we focus on research progress in identifying axonal targeting motifs and machinery of several different types of K(+) channels present in axons. Regulation of K(+) channel targeting and activity may underlie a novel form of neuronal plasticity. This research field can contribute to generating novel therapeutic strategies through manipulating neuronal excitability in treating neurological diseases, such as multiple sclerosis, neuropathic pain, and Alzheimer's disease.  相似文献   

6.
Ca(2+) clearance in frog motor nerve terminals was studied by fluorometry of Ca(2+) indicators. Rises in intracellular Ca(2+) ([Ca(2+)](i)) in nerve terminals induced by tetanic nerve stimulation (100 Hz, 100 or 200 stimuli: Ca(2+) transient) reached a peak or plateau within 6-20 stimuli and decayed at least in three phases with the time constants of 82-87 ms (81-85%), a few seconds (11-12%), and several tens of seconds (less than a few percentage). Blocking both Na/Ca exchangers and Ca(2+) pumps at the cell membrane by external Li(+) and high external pH (9.0), respectively, increased the time constants of the initial and second decay components with no change in their magnitudes. By contrast, similar effects by Li(+) alone, but not by high alkaline alone, were seen only on 200 stimuli-induced Ca(2+) transients. Blocking Ca(2+) pumps at Ca(2+) stores by thapsigargin did not affect 100 stimuli-induced Ca(2+) transients but increased the initial decay time constant of 200 stimuli-induced Ca(2+) transients with no change in other parameters. Inhibiting mitochondrial Ca(2+) uptake by carbonyl cyanide m-chlorophenylhydrazone markedly increased the initial and second decay time constants of 100 stimuli-induced Ca(2+) transients and the amplitudes of the second and the slowest components. Plotting the slopes of the decay of 100 stimuli-induced Ca(2+) transients against [Ca(2+)](i) yielded the supralinear [Ca(2+)](i) dependence of Ca(2+) efflux out of the cytosol. Blocking Ca(2+) extrusion or mitochondrial Ca(2+) uptake significantly reduced this [Ca(2+)](i)-dependent Ca(2+) efflux. Thus Ca(2+)-dependent mitochondrial Ca(2+) uptake and plasmalemmal Ca(2+) extrusion clear out a small Ca(2+) load in frog motor nerve terminals, while thapsigargin-sensitive Ca(2+) pump boosts the clearance of a heavy Ca(2+) load. Furthermore, the activity of plasmalemmal Ca(2+) pump and Na/Ca exchanger is complementary to each other with the slight predominance of the latter.  相似文献   

7.
The Ca(2+)-dependent release of the neurotransmitter glutamate from purified nerve terminals (synaptosomes) of the rat hippocampus was studied in a rapid perfusion apparatus. The response of the terminals was investigated with respect to the kinetics and duration of the release of endogenous glutamate upon brief and sustained stimulation and upon repetitive stimulation. The synaptosomes were stimulated by sustained chemical depolarization (0.5-3 min 30 mM K+). The cellular levels of glutamate, free Ca2+ and ATP in the nerve terminals were measured. The Ca(2+)-dependent release of glutamate showed an immediate elevation upon K(+)-depolarization. When the stimulation was maintained, a prolonged phase of glutamate release was observed. After 3 min, the Ca(2+)-dependent release stopped, although K(+)-depolarization was still effective. When synaptosomes were stimulated again after a relatively short stimulation period (30 s), the second response was similar to the previous one. After a longer stimulation period, maintained until termination of release, the second response did not show the immediate initial elevation of Ca(2+)-dependent glutamate release. Only 30 s after stimulation the release developed with a time profile comparable to the first response. This initial lack of response was not due to low cytosolic levels of glutamate or ATP or to changes in cellular Ca(2+)-buffering. It can be concluded that the capacity to release glutamate after brief depolarizations is fully restored during the repolarization period. However, if stimulation periods are of long duration (until termination of release), this capacity is no longer fully restored, especially with respect to a fast component of release. New glutamate is recruited only during the subsequent depolarization and with a delay.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
O(2)-sensing in the carotid body occurs in neuroectoderm-derived type I glomus cells where hypoxia elicits a complex chemotransduction cascade involving membrane depolarization, Ca(2+) entry and the release of excitatory neurotransmitters. Efforts to understand the exquisite O(2)-sensitivity of these cells currently focus on the coupling between local P(O2) and the open-closed state of K(+)-channels. Amongst multiple competing hypotheses is the notion that K(+)-channel activity is mediated by a phagocytic-like multisubunit enzyme, NADPH oxidase, which produces reactive oxygen species (ROS) in proportion to the prevailing P(O2). In O(2)-sensitive cells of lung neuroepithelial bodies (NEB), multiple studies confirm that ROS levels decrease in hypoxia, and that E(M) and K(+)-channel activity are indeed controlled by ROS produced by NADPH oxidase. However, recent studies in our laboratories suggest that ROS generated by a non-phagocyte isoform of the oxidase are important contributors to chemotransduction, but that their role in type I cells differs fundamentally from the mechanism utilized by NEB chemoreceptors. Data indicate that in response to hypoxia, NADPH oxidase activity is increased in type I cells, and further, that increased ROS levels generated in response to low-O(2) facilitate cell repolarization via specific subsets of K(+)-channels.  相似文献   

9.
Rises in free [Ca2+]i in response to various tetanic stimuli (Ca2+ transient) in frog motor nerve terminals were measured by recording fluorescence changes of Ca2+ indicators and analyzed in relation to short-term synaptic plasticity. Ca2+ transients reached a plateau after 10-20 impulses at 100 Hz and decayed in a three-exponential manner, in which the fast component was predominant. The plateau and fast component of the Ca2+ transient were elevated infralinearly with an increase in tetanus frequency. Computer simulation showed that the Ca2+ transients estimated from fluorescence changes faithfully reflect the true changes in [Ca2+]i except for the initial 20 ms. A slow Ca2+ chelator, EGTA, loaded into the nerve terminal, decreased the magnitude of both the fast and slow components of facilitation of transmitter release and the time constant of the former. A fast Ca2+ chelator, BAPTA, decreased the magnitude of fast facilitation but slightly increased its time constant. These results suggest that Ca2+ transients in the frog motor nerve terminals are primarily caused by Ca2+ entry and are dissipated by three components, in which the rate of the fast component is equivalent to that of free Ca2+ diffusion. The residual Ca2+ in the nerve terminals after stimulation accounts for the fast component of facilitation.  相似文献   

10.
Allana TN  Lin JW 《Neuroscience》2008,154(4):1242-1254
We investigated the mechanism of facilitation at the crayfish inhibitory neuromuscular junction before and after blocking P-type Ca(2+) channels. P-type channels have been shown to be closer to releasable synaptic vesicles than non-P-type channels at this synapse. Prior to the block of P-type channels, facilitation evoked by a train of 10 action potentials at 100 Hz was increased by application of 40 mM [Mg(2+)](o), but decreased by pressure-injected EGTA. Blocking P-type channels with 5 nM omega-Aga IVA, which reduced total Ca(2+) influx and release to levels comparable to that recorded in 40 mM [Mg(2+)](o), did not change the magnitude of facilitation. We explored whether this observation could be attributed to the buffer saturation model of facilitation, since increasing the Ca(2+) channel-vesicle separation could potentially enhance the role of endogenous buffers. The characteristics of facilitation in synapses treated with omega-Aga IVA were probed with broad action potentials in the presence of K(+) channel blockers. After Ca(2+) channel-vesicle separation was increased by omega-Aga IVA, facilitation probed with broad action potential was still decreased by EGTA injection and increased by 40 mM [Mg(2+)](o). EGTA-AM perfusion was used to test the impact of EGTA over a range of concentration in omega-Aga IVA-poisoned preparations. The results showed a concentration dependent decrease in facilitation as EGTA concentration rose. Thus, probing facilitation with EGTA and reduced Ca(2+) influx showed that characteristics of facilitation are not changed after the role of endogenous buffer is enhanced by increasing Ca(2+) channel-vesicle separation. There is no clear indication that buffer saturation has become the dominant mechanism for facilitation after omega-Aga IVA poisoning. Finally, we sought correlation between residual Ca(2+) and the magnitude of facilitation. Using fluorescence transients of a low affinity Ca(2+) indicator, we calculated the ratio of fluorescence amplitude measured immediately before test pulse (residual Ca(2+)) to that evoked during action potential (local Ca(2+)). This ratio provides an estimate of relative changes between residual Ca(2+) and local Ca(2+) important for release. There is a significant increase in the ratio when Ca(2+) influx is reduced by 40 mM [Mg(2+)](o). The magnitude of facilitation exhibited a clear and positive correlation with the ratio, regardless of separation between Ca(2+) channels and releasable vesicles. This correlation suggests the importance of relative changes between residual and local Ca(2+) and lends support to the residual Ca(2+) hypothesis of facilitation.  相似文献   

11.
The translocation of synaptic Zn(2+) from nerve terminals into selectively vulnerable neurons may contribute to the death of these neurons after global ischemia. We hypothesized that cellular Zn(2+) overload might be lethal for reasons similar to cellular Ca(2+) overload and tested the hypothesis that Zn(2+) neurotoxicity might be mediated by the activation of nitric oxide synthase. Although Zn(2+) (30-300microM) altered nitric oxide synthase activity in cerebellar extracts in solution, it did not affect nitric oxide synthase activity in cultured murine neocortical neurons. Cultured neurons exposed to 300-500microM Zn(2+) for 5min under depolarizing conditions developed widespread degeneration over the next 24h that was unaffected by the concurrent addition of the nitric oxide synthase inhibitor N(G)-nitro-L-arginine. Furthermore, Zn(2+) neurotoxicity was attenuated when nitric oxide synthase activity in the cultures was induced by exposure to cytokines, exogenous nitric oxide was added or nitric oxide production was pharmacologically enhanced. The unexpected protective effect of nitric oxide against Zn(2+) toxicity may be explained, at least in part, by reduction of toxic Zn(2+) entry. Exposure to nitric oxide donors reduced Ba(2+) current through high-voltage activated calcium channels, as well as K(+)-stimulated neuronal uptake of 45Ca(2+) or 65Zn(2+). The oxidizing agents thimerosal and 2,2'-dithiodipyridine also reduced K(+)-stimulated cellular 45Ca(2+) uptake, while akylation of thiols by pretreatment with N-ethylmaleimide blocked the reduction of 45Ca(2+) uptake by a nitric oxide donor.The results suggest that Zn(2+)-induced neuronal death is not mediated by the activation of nitric oxide synthase; rather, available nitric oxide may attenuate Zn(2+) neurotoxicity by reducing Zn(2+) entry through voltage-gated Ca(2+) channels, perhaps by oxidizing key thiol groups.  相似文献   

12.
Because Ca(2+) plays diverse roles in intracellular signaling in neurons, several types of calcium channels are employed to control Ca(2+) influx in these cells. Our experiments focus on resolving the paradox of why whole-cell current has not been observed under typical recording conditions for one type of calcium channel that is highly expressed in frog sympathetic neurons. These channels, referred to as E(f)-channels, are present in the membrane at a density greater than the channels that carry approximately 90% of whole-cell current in low Ba(2+); but, E(f)-current has not been detected in low Ba(2+). Using Ca(2+) instead of Ba(2+) as the charge carrier, we recorded a possible E-type current in frog sympathetic neurons. The current was resistant to specific blockers of N-, L-, and P/Q-type calcium channels but was more sensitive to Ni(2+) block than was N- or L-current. Current amplitude in Ca(2+) is slightly greater than that in Ba(2+). In 3 mM Ca(2+), the current contributed approximately 12% of total current at peak voltage and increased at voltages more hyperpolarized to the peak, reaching approximately 40% at -30 mV, where whole-cell current starts to activate. The presence of E(f)-current in 3 mM Ca(2+) suggests a potential role for E(f)-channels in regulating calcium influx into sympathetic neurons.  相似文献   

13.
Rat Meynert neurons were acutely isolated using a dissociation technique that maintains functional GABAergic presynaptic boutons. Miniature inhibitory postsynaptic currents (mIPSCs) were recorded under voltage-clamp conditions using whole cell patch-clamp recordings. Using the frequency of mIPSCs as a measure of presynaptic terminal excitability, the existence of a Na(+)/Ca(2+) exchanger (NCX) in these GABAergic nerve terminals was clearly demonstrated. Both the frequency and the amplitude of mIPSCs were unaffected by replacement of extracellular Na(2+). However, in this Na(+)-free external solution, ouabain could now induce a transient increase of mIPSCs frequency, which was not inhibited by adding Cd(2+) or cyclopiazonic acid but was inhibited by removing external Ca(2+). This indicates that this transient potentiation was dependent on external Ca(2+), but that this Ca(2+) influx was not via voltage-dependent Ca(2+) channels. KB-R7943, an inhibitor of NCX, at a concentration of 3 x 10(-6) M, reduced this transient increase of mIPSCs frequency without affecting mIPSCs amplitude and the response to exogenous GABA. These results demonstrate the existence of NCX in these GABAergic nerve terminals. In zero external Na(+), ouabain causes an accumulation of intraterminal Na(+) and a resultant influx of Ca(2+) through the reversed mode operation of NCX. However, under more physiological conditions, NCX may also operate in a forward mode and serve to maintain low intracellular [Ca(2+)] in nerve terminals.  相似文献   

14.
Experiments on frog neuromuscular synapses using extracellular microelectrode recording of endplate currents (EPC) and nerve ending (NE) responses were performed to study the mechanisms of facilitation of quantum secretion of acetylcholine on replacement of extracellular Ca ions with Sr ions. Solutions with a Ca ion concentration of 0.5 mM (calcium solutions) or a Sr ion concentration of 1 mM (strontium solutions) were used; the basal levels of neurotransmitter secretion (in conditions of low-frequency stimulation) were essentially identical. In calcium solutions, the drop in EPC facilitation on paired-pulse stimulation as the interimpulse interval was increased from 5 to 500 msec was described by the sum of three exponential components – the early, the first, and the second. In strontium solutions, facilitation was decreased as compared with the level in calcium solutions predominantly because of decreases in the early and first components. At the same time, EPC facilitation in conditions of rhythmic stimulation (10 or 50 impulses/sec) in strontium solution was significantly increased as compared with the level in calcium solutions. In strontium solutions in conditions of high-frequency stimulation at 50 impulses/sec, there was also a marked decrease in the amplitude of the third phase of the NE response, reflecting NE potassium currents. These data lead to the conclusion that the facilitation sites underlying the first and early components had lower affinities for Sr ions than for Ca ions. Increases in facilitation in strontium solutions in conditions of high-frequency rhythmic activity resulted from two mechanisms: more marked widening of the NE action potential and an increase in the divalent cation influx current due to weak activation of the Ca2+-dependent potassium current in the presence of Sr ions, as well as the slow dynamics of the removal of Sr ions from the NE axoplasm as compared with that in the presence of Ca ions. Translated from Rossiiskii Fiziologicheskii Zhurnal imeni I. M. Sechenova, Vol. 94, No. 2, pp, 142–151, February, 2008.  相似文献   

15.
By perineurial recording, the effects of toosendanin (TSN), a presynaptic blocker, on nerve terminal calcium currents (I(Ca)) were observed in innervated triangularis sterni of the mouse and cutaneous pectoris of the frog. It was found that TSN blocked the slow component of I(Ca) insensitive to nifedipine and omega-conotoxin-GVIA, and increasing the extracellular Ca(2+) concentration partially antagonized the inhibitory effect in mouse motor nerve terminals. However, in the frog, TSN increased the slow component of I(Ca) and this effect disappeared in the presence of nifedipine in perfusion solution. Based on previous data showing that the slow component of I(Ca) were mediated by different subtypes of calcium channels in mouse and frog motor nerve terminals, we presume that TSN could exercise different effects on various subtypes of calcium channels.  相似文献   

16.
PURPOSE: This study examined the expression and function of inward rectifier K(+) channels in cultured rat hepatic stellate cells (HSC). MATERIALS AND METHODS: The expression of inward rectifier K(+) channels was measured using real-time RT-PCR, and electrophysiological properties were determined using the gramicidin-perforated patch-clamp technique. RESULTS: The dominant inward rectifier K(+) channel subtypes were K(ir)2.1 and K(ir)6.1. These dominant K(+) channel subtypes decreased significantly during the primary culture throughout activation process. HSC can be classified into two subgroups: one with an inward-rectifying K(+) current (type 1) and the other without (type 2). The inward current was blocked by Ba(2+) (100 microM) and enhanced by high K(+) (140 mM), more prominently in type 1 HSC. There was a correlation between the amplitude of the Ba(2+)-sensitive current and the membrane potential. In addition, Ba(2+) (300 microM) depolarized the membrane potential. After the culture period, the amplitude of the inward current decreased and the membrane potential became depolarized. CONCLUSION: HSC express inward rectifier K(+) channels, which physiologically regulate membrane potential and decrease during the activation process. These results will potentially help determine properties of the inward rectifier K(+) channels in HSC as well as their roles in the activation process.  相似文献   

17.
Adenosine is a neuromodulator that controls neurotransmitter release through inhibitory A1 and facilitatory A2A receptors. Although both adenosine receptor-mediated inhibition and facilitation of glutamate release have been observed, it is not clear whether both A1 and A2A receptors are located in the same glutamatergic nerve terminal or whether they are located on different populations of these terminals. Thus, we have tested if single pyramidal glutamatergic neurons from the hippocampus simultaneously expressed A1 and A2A receptor mRNA and if A1 and A2A receptors were co-localized in hippocampal glutamatergic nerve terminals. Single cell PCR analysis of visually identified pyramidal neurons revealed the simultaneous presence of A1 and A2A receptor mRNA in four out 16 pyramidal cells possessing glutamatergic markers but not GABAergic or astrocytic markers. Also, A1 and A2A receptor immunoreactivities were co-localized in 26 +/- 4% of nerve terminals labeled with antibodies against vesicular glutamate transporters type 1 or 2, i.e. glutamatergic nerve terminals. This indicates that glutamatergic neurons in the hippocampus co-express A1 and A2A receptors and that these two receptors are co-localized in a subset of glutamatergic nerve terminals.  相似文献   

18.
Propagation of odor-induced Ca(2+) transients from the cilia/knob to the soma in mammalian olfactory receptor neurons (ORNs) is thought to be mediated exclusively by high-voltage-activated Ca(2+) channels. However, using confocal Ca(2+) imaging and immunocytochemistry we identified functional T-type Ca(2+) channels in rat ORNs. Here we show that T-type Ca(2+) channels in ORNs also mediate propagation of odor-induced Ca(2+) transients from the knob to the soma. In the presence of the selective inhibitor of T-type Ca(2+) channels mibefradil (10-15 microM) or Ni(2+) (100 microM), odor- and forskolin/3-isobutyl-1-methyl-xanthine (IBMX)-induced Ca(2+) transients in the soma and dendrite were either strongly inhibited or abolished. The percentage of inhibition of the Ca(2+) transients in the knob, however, was 40-50% less than that in the soma. Ca(2+) transients induced by 30 mM K(+) were partially inhibited by mibefradil, but without a significant difference in the extent of inhibition between the knob and soma. Furthermore, an increase of as little as 2.5 mM in the extracellular K(+) concentration (7.5 mM K(+)) was found to induce Ca(2+) transients in ORNs, and such responses were completely inhibited by mibefradil or Ni(2+). Total replacement of extracellular Na(+) with N-methyl-d-glutamate inhibited none of the odor-, forskolin/IBMX- or 7.5 mM K(+)-induced Ca(2+) transients. Positive immunoreactivity to the Ca(v)3.1, Ca(v)3.2 and Ca(v)3.3 subunits of the T-type Ca(2+) channel was observed throughout the soma, dendrite and knob. These data suggest that involvement of T-type Ca(2+) channels in the propagation of odor-induced Ca(2+) transients in ORNs may contribute to signal transduction and odor sensitivity.  相似文献   

19.
Stimulation-induced facilitation of transmitter release was examined at spiny lobster (Palinurus japonicus) neuromuscular junctions by measuring excitatory junctional potentials (EJPs). We found three components of facilitation with the decay time constants about 16, 200 and 1000 ms, respectively. The decay time constants of the nerve terminal free Ca(2+) concentration after stimulation agreed well with the two slower time constants of facilitation. The relationship among the three components of facilitation and a yet slower component, S, was investigated on the basis of several models. The models that have a multiplicative relationship between any two components of facilitation, and the additive model between all components could not account for the results of experiments. Only the 'unified power model', which assumes that facilitation and S are described by the 3-4th power of the sum of underlying components, could account for both the growth process of EJPs during stimulation and the effects of Ca(2+)-chelators. Loading Ca(2+)-chelators, BAPTA and EGTA, into the presynaptic terminals resulted in reduction, but not elimination, of any component of the facilitation. These results suggest that in the spiny lobster neuromuscular junction, the 'unified power model' can describe the relationship between three components of facilitation and S, and that residual free Ca(2+) enhances all three components of facilitation, although it may not be essential for them.  相似文献   

20.
In the circular muscle of the guinea-pig gastric antrum, a decrease in the external Na(+) to less than 20 mM produced depolarization of the membrane with transient prolongation of the slow wave. This was followed by a high rhythmic activity. The activity was inhibited by reapplication of Na(+) before recovery. The depolarization in Na(+)-deficient solution was prevented and rhythmic activity continued at about 5/min for at least 6 min by simultaneous removal of K(+), Ca(2+), or Cl(-). After exposure to a Na(+)- and Cl(-)-deficient solution for a few minutes, reapplication of the Na(+) in Cl(-)-deficient solution inhibited generation of the slow wave until Cl(-) reapplication. Similar results were obtained when Na(+) and Cl(-) were reapplied in the absence of K(+) after exposure to a Na(+)-, K(+)-free, and Cl(-)-deficient solution, although the inhibition was weaker than Na(+) reapplication in a Cl(-)-deficient solution. In the presence of furosemide or bumetanide, a strong inhibition of activity was produced by the reapplication of Na(+) and Cl(-) after exposure to an Na(+)- and Cl(-)-deficient solution. A hypothesis is presented that intracellular Ca(2+) concentration ([Ca(2+)](i)) is the most important factor determining the generation and frequency of the slow wave and that [Ca(2+)](i) is regulated by the Na(+) concentration gradient across the plasma membrane. The recovery of the Na(+) concentration gradient by Na(+) reapplication after removal of Na(+) and Cl(-) is mainly controlled by a Na(+)-K(+)-Cl(-) co-transport.  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