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.
Voltage-dependent Ca(2+) channels play important roles in cerebellar functions including motor coordination and learning. Since abundant expression of Ca(V)2.3 Ca(2+) channel gene in the cerebellum was detected, we searched for possible deficits in the cerebellar functions in the Ca(V)2.3 mutant mice. Behavioral analysis detected in delayed motor learning in rotarod tests in mice heterozygous and homozygous for the Ca(V)2.3 gene disruption (Ca(V)2.3+/- and Ca(V)2.3-/-, respectively). Electrophysiological analysis of mutant mice revealed perplexing results: deficit in long-term depression (LTD) at the parallel fiber Purkinje cell synapse in Ca(V)2.3+/- mice but apparently normal LTD in Ca(V)2.3-/- mice. On the other hand, the number of spikes evoked by current injection in Purkinje cells under the current-clamp mode decreased in Ca(V)2.3 mutant mice in a gene dosage-dependent manner, suggesting that Ca(V)2.3 channel contributed to spike generation in Purkinje cells. Thus, Ca(V)2.3 channel seems to play some roles in cerebellar functions.  相似文献   

2.
As sperm prepare for fertilization, surface Ca(2+) channels must open to initiate required, Ca(2+)-mediated events. However, the molecular identity and functional properties of sperm Ca(2+) channels remain uncertain. Here, we use rapid local perfusion and single-cell photometry to examine the kinetics of calcium responses of mouse sperm to depolarizing stimuli. The linear rise of intracellular [Ca(2+)] evoked by approximately 10-s applications of an alkaline high [K(+)] medium directly reports activity of voltage-gated Ca(2+) channels. Little response occurs if external Ca(2+) is removed or if external or internal pH is elevated without depolarization. Responses are inhibited 30-40% by 30-100 micrometer Ni(2+) and more completely by 100-300 micrometer Cd(2+). They resist the dihydropyridines nitrendipine and PN200-110, but 1-10 micrometer mibefradil inhibits reversibly. They also resist the venom toxins calciseptine, omega-conotoxin MVIIC, and kurtoxin, but omega-conotoxin GVIA (5 micrometer) inhibits approximately 50%. GVIA also partially blocks transient, low voltage activated Ca(2+) currents of patch-clamped spermatids. Differential sensitivity of sperm responses to Ni(2+) and Cd(2+) and partial blockade by GVIA indicate that depolarization opens at least two types of voltage-gated Ca(2+) channels in epididymal sperm examined prior to capacitation. Involvement of a previously undetected Ca(V)2.2 (N-type) channel, suggested by the action of GVIA, is substantiated by immunodetection of Ca(2+) channel alpha(1B) subunits in sperm and sperm extracts. Resistance to dihydropyridines, calciseptine, MVIIC, and kurtoxin indicates that Ca(V)1, Ca(V)2.1, and Ca(V)3 (L-, P/Q-, and T-type) channels contribute little to this evoked response. Partial sensitivity to 1 micrometer mibefradil and an enhanced sensitivity of the GVIA-resistant component of response to Ni(2+) suggest participation of a Ca(V)2.3 (R-type) channel specified by previously found alpha(1E) subunits. Our examination of depolarization-evoked Ca(2+) entry indicates that mature sperm possess a larger palette of voltage-gated Ca(2+) channels than previously thought. Such diversity may permit specific responses to multiple cues encountered on the path to fertilization.  相似文献   

3.
4.
Ca2+和突触细胞融合   总被引:1,自引:0,他引:1  
神经突触传递对于神经系统功能的实现具有十分重要的意义,而神经突触传递涉及到突触囊泡膜和突触前膜的融合,3种膜蛋白SNARE特异性识别并形成复合物,从而介导了神经递质的释放。Ca^2 通过其感受器突触结合蛋白而调节了突触细胞的融合过程,也最终影响了神经元的胞吐作用。  相似文献   

5.
6.
7.
Ca2+ influx into presynaptic terminals via voltage-dependent Ca2+ channels triggers fast neurotransmitter release as well as different forms of synaptic plasticity. Using electrophysiological and genetic techniques we demonstrate that presynaptic Ca2+ entry through Cav2.3 subunits contributes to the induction of mossy fiber LTP and posttetanic potentiation by brief trains of presynaptic action potentials while they do not play a role in fast synaptic transmission, paired-pulse facilitation, or frequency facilitation. This functional specialization is most likely achieved by a localization remote from the release machinery and by a Cav2.3 channel-dependent facilitation of presynaptic Ca2+ influx. Thus, the presence of Cav2.3 channels boosts the accumulation of presynaptic Ca2+ triggering presynaptic LTP and posttetanic potentiation without affecting the low release probability that is a prerequisite for the enormous plasticity displayed by mossy fiber synapses.  相似文献   

8.
There is a growing body of evidence suggesting a functional relationship between Ca2+ signals generated in astroglia and the functioning of nearby excitatory synapses. Interference with endogenous Ca2+ homeostasis inside individual astrocytes has been shown to affect synaptic transmission and its use-dependent changes. However, establishing the causal link between source-specific, physiologically relevant intracellular Ca2+ signals, the astrocytic release machinery and the consequent effects on synaptic transmission has proved difficult. Improved methods of Ca2+ monitoring in situ will be essential for resolving the ambiguity in understanding the underlying Ca2+ signalling cascades.  相似文献   

9.
10.
CaVβ subunits are formed by a Src homology 3 domain and a guanylate kinase-like (GK) domain connected through a variable HOOK domain. Complete deletion of the Src homology 3 domain (75 residues) as well as deletion of the HOOK domain (47 residues) did not alter plasma membrane density of CaV2.3 nor its typical activation gating. In contrast, six-residue deletions in the GK domain disrupted cell surface trafficking and functional expression of CaV2.3. Mutations of residues known to carry nanomolar affinity binding in the GK domain of CaVβ (P175A, P179A, M195A, M196A, K198A, S295A, R302G, R307A, E339G, N340G, and A345G) did not significantly alter cell surface targeting or gating modulation of CaV2.3. Nonetheless, mutations of a quartet of leucine residues (either single or multiple mutants) in the α3, α6, β10, and α9 regions of the GK domain were found to significantly impair cell surface density of CaV2.3 channels. Furthermore, the normalized protein density of CaV2.3 was nearly abolished with the quadruple CaVβ3 Leu mutant L200G/L303G/L337G/L342G. Altogether, our observations suggest that the four leucine residues in CaVβ3 form a hydrophobic pocket surrounding key residues in the α-interacting domain of CaV2.3. This interaction appears to play an essential role in conferring CaVβ-induced modulation of the protein density of CaVα1 subunits in CaV2 channels.  相似文献   

11.
12.
Apamin: a specific toxin to study a class of Ca2+-dependent K+ channels   总被引:5,自引:0,他引:5  
Apamin is a bee venom neurotoxin of 18 amino-acids containing two disulfide bridges. Current clamp and voltage clamp experiments have shown that externally applied apamin blocks specifically at low concentration (0.1 microM) the Ca2+-dependent slow K+ conductance which mediates the long-lasting after-hyperpolarization in neuroblastoma cells and rat muscle cells in culture. The apamin-sensitive Ca2+-dependent slow K+ conductance is voltage-dependent and tetraethylammonium (TEA) insensitive. It is distinct from the high conductance Ca2+-dependent K+ channel revealed by patch clamp experiments. Biochemical characterization of the apamin receptor in rat striated muscle, neuroblastoma cells, rat synaptosomes, smooth muscles and hepatocytes was carried out with the use of a radiolabelled monoiodo-apamin derivative (125I-apamin) of high specific radioactivity (2 000 Ci/mmol). The dissociation constant of the apamin-receptor complex is between 15 and 60 pM for all tissue preparations. The density of binding sites is very low; it varied between 1 and 40 fmol/mg of protein. Radiation inactivation analysis indicates a molecular weight for the apamin receptor of 250 000 daltons whereas affinity labelling with 125I-apamin results in covalent labelling of a single polypeptide chain with a molecular weight of about 30 000 daltons. We conclude that the apamin-sensitive Ca2+-dependent K+ channel is probably a large oligomeric structure containing one subunit of 30 000 daltons.  相似文献   

13.
The cardiac L-type voltage-dependent calcium channel is responsible for initiating excitation-contraction coupling. Three sequences (amino acids 1609-1628, 1627-1652, and 1665-1685, designated A, C, and IQ, respectively) of its alpha(1) subunit contribute to calmodulin (CaM) binding and Ca(2+)-dependent inactivation. Peptides matching the A, C, and IQ sequences all bind Ca(2+)CaM. Longer peptides representing A plus C (A-C) or C plus IQ (C-IQ) bind only a single molecule of Ca(2+)CaM. Apocalmodulin (ApoCaM) binds with low affinity to the IQ peptide and with higher affinity to the C-IQ peptide. Binding to the IQ and C peptides increases the Ca(2+) affinity of the C-lobe of CaM, but only the IQ peptide alters the Ca(2+) affinity of the N-lobe. Conversion of the isoleucine and glutamine residues of the IQ motif to alanines in the channel destroys inactivation (Zühlke et al., 2000). The double mutation in the peptide reduces the interaction with apoCaM. A mutant CaM unable to bind Ca(2+) at sites 3 and 4 (which abolishes the ability of CaM to inactivate the channel) binds to the IQ, but not to the C or A peptide. Our data are consistent with a model in which apoCaM binding to the region around the IQ motif is necessary for the rapid binding of Ca(2+) to the C-lobe of CaM. Upon Ca(2+) binding, this lobe is likely to engage the A-C region.  相似文献   

14.
Homeostatic synaptic plasticity is a negative-feedback mechanism for compensating excessive excitation or inhibition of neuronal activity. When neuronal activity is chronically suppressed, neurons increase synaptic strength across all affected synapses via synaptic scaling. One mechanism for this change is alteration of synaptic AMPA receptor (AMPAR) accumulation. Although decreased intracellular Ca2+ levels caused by chronic inhibition of neuronal activity are believed to be an important trigger of synaptic scaling, the mechanism of Ca2+-mediated AMPAR-dependent synaptic scaling is not yet understood. Here, we use dissociated mouse cortical neurons and employ Ca2+ imaging, electrophysiological, cell biological, and biochemical approaches to describe a novel mechanism in which homeostasis of Ca2+ signaling modulates activity deprivation-induced synaptic scaling by three steps: (1) suppression of neuronal activity decreases somatic Ca2+ signals; (2) reduced activity of calcineurin, a Ca2+-dependent serine/threonine phosphatase, increases synaptic expression of Ca2+-permeable AMPARs (CPARs) by stabilizing GluA1 phosphorylation; and (3) Ca2+ influx via CPARs restores CREB phosphorylation as a homeostatic response by Ca2+-induced Ca2+ release from the ER. Therefore, we suggest that synaptic scaling not only maintains neuronal stability by increasing postsynaptic strength but also maintains nuclear Ca2+ signaling by synaptic expression of CPARs and ER Ca2+ propagation.  相似文献   

15.
L-type Ca2+ channels in mammalian brain neurons have either a CaV1.2 or CaV1.3 pore-forming subunit. Recently, it was shown that CaV1.3 Ca2+ channels underlie autonomous pacemaking in adult dopaminergic neurons in the substantia nigra pars compacta, and this reliance renders them sensitive to toxins used to create animal models of Parkinson’s disease. Antagonism of these channels with the dihydropyridine antihypertensive drug isradipine diminishes the reliance on Ca2+ and the sensitivity of these neurons to toxins, pointing to a potential neuroprotective strategy. However, for neuroprotection without an antihypertensive side effect, selective CaV1.3 channel antagonists are required. In an attempt to identify potent and selective antagonists of CaV1.3 channels, 124 dihydropyridines (4-substituted-1,4-dihydropyridine-3,5-dicarboxylic diesters) were synthesized. The antagonism of heterologously expressed CaV1.2 and CaV1.3 channels was then tested using electrophysiological approaches and the FLIPR Calcium 4 assay. Despite the large diversity in substitution on the dihydropyridine scaffold, the most CaV1.3 selectivity was only about twofold. These results support a highly similar dihydropyridine binding site at both CaV1.2 and CaV1.3 channels and suggests that other classes of compounds need to be identified for CaV1.3 selectivity.  相似文献   

16.
Calmodulin binds to IQ motifs in the α1 subunit of CaV1.1 and CaV1.2, but the affinities of calmodulin for the motif and for Ca2+ are higher when bound to CaV1.2 IQ. The CaV1.1 IQ and CaV1.2 IQ sequences differ by four amino acids. We determined the structure of calmodulin bound to CaV1.1 IQ and compared it with that of calmodulin bound to CaV1.2 IQ. Four methionines in Ca2+-calmodulin form a hydrophobic binding pocket for the peptide, but only one of the four nonconserved amino acids (His-1532 of CaV1.1 and Tyr-1675 of CaV1.2) contacts this calmodulin pocket. However, Tyr-1675 in CaV1.2 contributes only modestly to the higher affinity of this peptide for calmodulin; the other three amino acids in CaV1.2 contribute significantly to the difference in the Ca2+ affinity of the bound calmodulin despite having no direct contact with calmodulin. Those residues appear to allow an interaction with calmodulin with one lobe Ca2+-bound and one lobe Ca2+-free. Our data also provide evidence for lobe-lobe interactions in calmodulin bound to CaV1.2.The complexity of eukaryotic Ca2+ signaling arises from the ability of cells to respond differently to Ca2+ signals that vary in amplitude, duration, and location. A variety of mechanisms decode these signals to drive the appropriate physiological responses. The Ca2+ sensor for many of these physiological responses is the Ca2+-binding protein calmodulin (CaM).2 The primary sequence of CaM is tightly conserved in all eukaryotes, yet it binds and regulates a broad set of target proteins in response to Ca2+ binding. CaM has two domains that bind Ca2+ as follows: an amino-terminal domain (N-lobe) and a carboxyl-terminal domain (C-lobe) joined via a flexible α-helix. Each lobe of CaM binds two Ca2+ ions, and binding within each lobe is highly cooperative. The two lobes of CaM, however, have distinct Ca2+ binding properties; the C-lobe has higher Ca2+ affinity because of a slower rate of dissociation, whereas the N-lobe has weaker Ca2+ affinity and faster kinetics (1). CaM can also bind to some target proteins in both the presence and absence of Ca2+, and the preassociation of CaM in low Ca2+ modulates the apparent Ca2+ affinity of both the amino-terminal and carboxyl-terminal lobes. Differences in the Ca2+ binding properties of the lobes and in the interaction sites of the amino- and carboxyl-terminal lobes enable CaM to decode local versus global Ca2+ signals (2).Even though CaM is highly conserved, CaM target (or recognition) sites are quite heterogeneous. The ability of CaM to bind to very different targets is at least partially due to its flexibility, which allows it to assume different conformations when bound to different targets. CaM also binds to various targets in distinct Ca2+ saturation states as follows: Ca2+-free (3), Ca2+ bound to only one of the two lobes, or fully Ca2+-bound (47). In addition, CaM may bind with both lobes bound to a target (5, 6) or with only a single lobe engaged (8). If a target site can bind multiple conformers of CaM, CaM may undergo several transitions that depend on Ca2+ concentration, thereby tuning the functional response. Identification of stable intermediate states of CaM bound to individual targets will help to elucidate the steps involved in this fine-tuned control.Both CaV1.1 and CaV1.2 belong to the L-type family of voltage-dependent Ca2+ channels, which bind apoCaM and Ca2+-CaM at carboxyl-terminal recognition sites in their α1 subunits (914). Ca2+ binding to CaM, bound to CaV1.2 produces Ca2+-dependent facilitation (CDF) (14). Whether CaV1.1 undergoes CDF is not known. However, both CaV1.2 and CaV1.1 undergo Ca2+- and CaM-dependent inactivation (CDI) (14, 15). CaV1.1 CDI is slower and more sensitive to buffering by 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid than CaV1.2 CDI (15). Ca2+ buffers are thought to influence CDI and/or CDF in voltage-dependent Ca2+ channels by competing with CaM for Ca2+ (16).The conformation of the carboxyl terminus of the α1 subunit is critical for channel function and has been proposed to regulate the gating machinery of the channel (17, 18). Several interactions of this region include intramolecular contacts with the pore inactivation machinery and intermolecular contacts with CaM kinase II and ryanodine receptors (17, 1922). Ca2+ regulation of CaV1.2 may involve several motifs within this highly conserved region, including an EF hand motif and three contiguous CaM-binding sequences (10, 12). ApoCaM and Ca2+-CaM-binding sites appear to overlap at the site designated as the “IQ motif” (9, 12, 13), which are critical for channel function at the molecular and cellular level (14, 23).Differences in the rate at which 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid affects CDI of CaV1.1 and CaV1.2 could reflect differences in their interactions with CaM. In this study we describe the differences in CaM interactions with the IQ motifs of the CaV1.1 and the CaV1.2 channels in terms of crystal structure, CaM affinity, and Ca2+ binding to CaM. We find the structures of Ca2+-CaM-IQ complexes are similar except for a single amino acid change in the peptide that contributes to its affinity for CaM. We also find that the other three amino acids that differ in CaV1.2 and CaV1.1 contribute to the ability of CaV1.2 to bind a partially Ca2+-saturated form of CaM.  相似文献   

17.
During vertebrate locomotion, spinal neurons act as oscillators when initiated by glutamate release from descending systems. Activation of NMDA receptors initiates Ca2+-mediated intrinsic membrane potential oscillations in central pattern generator (CPG) neurons. NMDA receptor-dependent intrinsic oscillations require Ca2+-dependent K+ (KCa2) channels for burst termination. However, the location of Ca2+ entry mediating KCa2 channel activation, and type of Ca2+ channel – which includes NMDA receptors and voltage-gated Ca2+ channels (VGCCs) – remains elusive. NMDA receptor-dependent Ca2+ entry necessitates presynaptic release of glutamate, implying a location at active synapses within dendrites, whereas VGCC-dependent Ca2+ entry is not similarly constrained. Where Ca2+ enters relative to KCa2 channels is crucial to information processing of synaptic inputs necessary to coordinate locomotion. We demonstrate that Ca2+ permeating NMDA receptors is the dominant source of Ca2+ during NMDA-dependent oscillations in lamprey spinal neurons. This Ca2+ entry is synaptically located, NMDA receptor-dependent, and sufficient to activate KCa2 channels at excitatory interneuron synapses onto other CPG neurons. Selective blockade of VGCCs reduces whole-cell Ca2+ entry but leaves membrane potential and Ca2+ oscillations unaffected. Furthermore, repetitive oscillations are prevented by fast, but not slow, Ca2+ chelation. Taken together, these results demonstrate that KCa2 channels are closely located to NMDA receptor-dependent Ca2+ entry. The close spatial relationship between NMDA receptors and KCa2 channels provides an intrinsic mechanism whereby synaptic excitation both excites and subsequently inhibits ventral horn neurons of the spinal motor system. This places the components necessary for oscillation generation, and hence locomotion, at glutamatergic synapses.  相似文献   

18.
The sarcoplasmic reticulum (SR) of skeletal muscle contains K+, Cl, and H+ channels may facilitate charge neutralization during Ca2+ release. Our recent studies have identified trimeric intracellular cation (TRIC) channels on SR as an essential counter-ion permeability pathway associated with rapid Ca2+ release from intracellular stores. Skeletal muscle contains TRIC-A and TRIC-B isoforms as predominant and minor components, respectively. Here we test the physiological function of TRIC-A in skeletal muscle. Biochemical assay revealed abundant expression of TRIC-A relative to the skeletal muscle ryanodine receptor with a molar ratio of TRIC-A/ryanodine receptor ∼5:1. Electron microscopy with the tric-a−/− skeletal muscle showed Ca2+ overload inside the SR with frequent formation of Ca2+ deposits compared with the wild type muscle. This elevated SR Ca2+ pool in the tric-a−/− muscle could be released by caffeine, whereas the elemental Ca2+ release events, e.g. osmotic stress-induced Ca2+ spark activities, were significantly reduced likely reflecting compromised counter-ion movement across the SR. Ex vivo physiological test identified the appearance of “alternan” behavior with isolated tric-a−/− skeletal muscle, i.e. transient and drastic increase in contractile force appeared within the decreasing force profile during repetitive fatigue stimulation. Inhibition of SR/endoplasmic reticulum Ca2+ ATPase function could lead to aggravation of the stress-induced alternans in the tric-a−/− muscle. Our data suggests that absence of TRIC-A may lead to Ca2+ overload in SR, which in combination with the reduced counter-ion movement may lead to instability of Ca2+ movement across the SR membrane. The observed alternan behavior with the tric-a−/− muscle may reflect a skeletal muscle version of store overload-induced Ca2+ release that has been reported in the cardiac muscle under stress conditions.  相似文献   

19.
Hearing relies on faithful signal transmission by cochlear inner hair cells (IHCs) onto auditory fibres over a wide frequency and intensity range. Exocytosis at IHC ribbon synapses is triggered by Ca2+ inflow through CaV1.3 (L-type) Ca2+ channels. We investigated the macroscopic (whole-cell) and elementary (cell-attached) properties of Ca2+ currents in IHCs positioned at the middle turn (frequency ∼2 kHz) of the adult gerbil cochlea, which is their most sensitive hearing region. Using near physiological recordings conditions (body temperature and a Na+ based extracellular solution), we found that the macroscopic Ca2+ current activates and deactivates very rapidly (time constant below 1 ms) and inactivates slowly and only partially. Single-channel recordings showed an elementary conductance of 15 pS, a sub-ms latency to first opening, and a very low steady-state open probability (P o: 0.024 in response to 500-ms depolarizing steps at ∼−18 mV). The value of P o was significantly larger (0.06) in the first 40 ms of membrane depolarization, which corresponds to the time when most Ca2+ channel openings occurred clustered in bursts (mean burst duration: 19 ms). Both the P o and the mean burst duration were smaller than those previously reported in high-frequency basal IHCs. Finally, we found that middle turn IHCs are likely to express about 4 times more Ca2+ channels per ribbon than basal cells. We propose that middle-turn IHCs finely-tune CaV1.3 Ca2+ channel gating in order to provide reliable information upon timing and intensity of lower-frequency sounds.  相似文献   

20.
The L-type calcium channel (LTCC) CaV1.3 is regarded as a new potential therapeutic target for Parkinson’s disease. Calcium influx through CaV1.3 LTCC during autonomous pacemaking in adult dopaminergic neurons of the substantia nigra pars compacta is related to the generation of mitochondrial oxidative stress in animal models. Development of a CaV1.3 antagonist selective over CaV1.2 is essential because CaV1.2 pore-forming subunits are the predominant form of LTCCs and are abundant in the central nervous and cardiovascular systems. We have explored 1,4-dihydropyrimidines and 4H-pyrans to identify potent and selective antagonists of CaV1.3 relative to CaV1.2 LTCCs. A library of 36 dihydropyridine (DHP)-mimic 1,4-dihydropyrimidines and 4H-pyrans was synthesized, and promising chiral compounds were resolved. The antagonism studies of CaV1.3 and CaV1.2 LTCCs using DHP mimic compounds showed that dihydropyrimidines and 4H-pyrans are effective antagonists of DHPs for CaV1.3 LTCCs. Some 1,4-dihydropyrimidines are more selective than isradipine for CaV1.3 over CaV1.2, shown here by both calcium flux and patch-clamp electrophysiology experiments, where the ratio of antagonism is around 2–3. These results support the hypothesis that the modified hydrogen bonding donor/acceptors in DHP-mimic dihydropyrimidines and 4H-pyrans can interact differently with DHP binding sites, but, in addition, the data suggest that the binding sites of DHP in CaV1.3 and CaV1.2 LTCCs are very similar.  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