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.
CaV1.3 L-type channels control inner hair cell (IHC) sensory and sinoatrial node (SAN) function, and excitability in central neurons by means of their low-voltage activation and inactivation properties. In SAN cells CaV1.3 inward calcium current (ICa) inactivates rapidly whereas in IHCs inactivation is slow. A candidate suggested in slowing CaV1.3 channel inactivation is the presynaptically located ribbon-synapse protein RIM that is expressed in immature IHCs in presynaptic compartments also expressing CaV1.3 channels. CaV1.3 channel gating is also modulated by an intramolecular C-terminal mechanism. This mechanism was elicited during analysis of human C-terminal splice variants that differ in the length of their C-terminus and that modulates the channel's negative activation range and slows calcium-dependent inactivation.  相似文献   

2.
The gating modifier toxins are a large family of protein toxins that modify either activation or inactivation of voltage-gated ion channels. omega-Aga-IVA is a gating modifier toxin from spider venom that inhibits voltage-gated Ca(2+) channels by shifting activation to more depolarized voltages. We identified two Glu residues near the COOH-terminal edge of S3 in the alpha(1A) Ca(2+) channel (one in repeat I and the other in repeat IV) that align with Glu residues previously implicated in forming the binding sites for gating modifier toxins on K(+) and Na(+) channels. We found that mutation of the Glu residue in repeat I of the Ca(2+) channel had no significant effect on inhibition by omega-Aga-IVA, whereas the equivalent mutation of the Glu in repeat IV disrupted inhibition by the toxin. These results suggest that the COOH-terminal end of S3 within repeat IV contributes to forming a receptor for omega-Aga-IVA. The strong predictive value of previous mapping studies for K(+) and Na(+) channel toxins argues for a conserved binding motif for gating modifier toxins within the voltage-sensing domains of voltage-gated ion channels.  相似文献   

3.
4.
5.
Alternative splicing is a critical mechanism used extensively in the mammalian nervous system to increase the level of diversity that can be achieved by a set of genes. This review focuses on recent studies of voltage-gated calcium (Ca) channel Cavα1 subunit splice isoforms in neurons. Voltage-gated Ca channels couple changes in neuronal activity to rapid changes in intracellular Ca levels that in turn regulate an astounding range of cellular processes. Only ten genes have been identified that encode Cavα1 subunits, an insufficient number to account for the level of functional diversity among voltage-gated Ca channels. The consequences of regulated alternative splicing among the genes that comprise voltage-gated Ca channels permits specialization of channel function, optimizing Ca signaling in different regions of the brain and in different cellular compartments. Although the full extent of alternative splicing is not yet known for any of the major subtypes of voltage-gated Ca channels, it is already clear that it adds a rich layer of structural and functional diversity”.  相似文献   

6.
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   

7.
8.
Reactive oxygen species (ROS) are increasingly recognized as second messengers in many cellular processes. While high concentrations of oxidants damage proteins, lipids and DNA, ultimately resulting in cell death, selective and reversible oxidation of key residues in proteins is a physiological mechanism that can transiently alter their activity and function. Defects in ROS producing enzymes cause disturbed immune response and disease.Changes in the intracellular free Ca2+ concentration are key triggers for diverse cellular functions. Ca2+ homeostasis thus needs to be precisely tuned by channels, pumps, transporters and cellular buffering systems. Alterations of these key regulatory proteins by reversible or irreversible oxidation alter the physiological outcome following cell stimulation. It is therefore necessary to understand which proteins are regulated and if this regulation is relevant in a physiological- and/or pathophysiological context. Because ROS are inherently difficult to identify and to measure, we first review basic oxygen redox chemistry and methods of ROS detection with special emphasis on electron paramagnetic resonance (EPR) spectroscopy. We then focus on the present knowledge of redox regulation of Ca2+ permeable ion channels such as voltage-gated (CaV) Ca2+ channels, transient receptor potential (TRP) channels and Orai channels.  相似文献   

9.
L-type voltage gated calcium channels (VGCCs) interact with a variety of proteins that modulate both their function and localization. A-Kinase Anchoring Proteins (AKAPs) facilitate L-type calcium channel phosphorylation through β adrenergic stimulation. Our previous work indicated a role of neuronal AKAP79/150 in the membrane targeting of Ca(V)1.2 L-type calcium channels, which involved a proline rich domain (PRD) in the intracellular II-III loop of the channel.(1) Here, we show that mutation of proline 857 to alanine (P857A) into the PRD does not disrupt the AKAP79-induced increase in Ca(v)1.2 membrane expression. Furthermore, deletion of two other PRDs into the carboxy terminal domain of Ca(V)1.2 did not alter the targeting role of AKAP79. In contrast, the distal carboxy terminus region of the channel directly interacts with AKAP79. This protein-protein interaction competes with a direct association of the channel II-III linker on the carboxy terminal tail and modulates membrane targeting of Ca(V)1.2. Thus, our results suggest that the effects of AKAP79 occur through relief of an autoinhibitory mechanism mediated by intramolecular interactions of Ca(v)1.2 intracellular regions.  相似文献   

10.
L-type voltage gated calcium channels (VGCCs) interact with a variety of proteins that modulate both their function and localization. A-Kinase Anchoring Proteins (AKAPs) facilitate L-type calcium channel phosphorylation through β adrenergic stimulation. Our previous work indicated a role of neuronal AKAP79/150 in the membrane targeting of CaV1.2 L-type calcium channels, which involved a proline rich domain (PRD) in the intracellular II-III loop of the channel.1 Altier C, Dubel SJ, Barrère C, Jarvis SE, Stotz SC, Spaetgens RL, et al. Trafficking of L-type calcium channels mediated by the postsynaptic scaffolding protein AKAP79. J Biol Chem 2002; 277:33598 - 603; http://dx.doi.org/10.1074/jbc.M202476200; PMID: 12114507 [Crossref], [PubMed], [Web of Science ®] [Google Scholar] Here, we show that mutation of proline 857 to alanine (P857A) into the PRD does not disrupt the AKAP79-induced increase in Cav1.2 membrane expression. Furthermore, deletion of two other PRDs into the carboxy terminal domain of CaV1.2 did not alter the targeting role of AKAP79. In contrast, the distal carboxy terminus region of the channel directly interacts with AKAP79. This protein-protein interaction competes with a direct association of the channel II-III linker on the carboxy terminal tail and modulates membrane targeting of CaV1.2. Thus, our results suggest that the effects of AKAP79 occur through relief of an autoinhibitory mechanism mediated by intramolecular interactions of Cav1.2 intracellular regions.  相似文献   

11.
The expression of voltage-gated calcium channels (VGCCs) has not been reported previously in melanoma cells in spite of increasing evidence of a role of VGCCs in tumorigenesis and tumour progression. To address this issue we have performed an extensive RT-PCR analysis of VGCC expression in human melanocytes and a range of melanoma cell lines and biopsies. In addition, we have tested the functional expression of these channels using Ca(2+) imaging techniques and examined their relevance for the viability and proliferation of the melanoma cells. Our results show that control melanocytes and melanoma cells express channel isoforms belonging to the Ca(v) 1 and Ca(v) 2 gene families. Importantly, the expression of low voltage-activated Ca(v) 3 (T-type) channels is restricted to melanoma. We have confirmed the function of T-type channels as mediators of constitutive Ca(2+) influx in melanoma cells. Finally, pharmacological and gene silencing approaches demonstrate a role for T-type channels in melanoma viability and proliferation. These results encourage the analysis of T-type VGCCs as targets for therapeutic intervention in melanoma tumorigenesis and/or tumour progression.  相似文献   

12.
Voltage-gated sodium and calcium channels are responsible for inward movement of sodium and calcium during electrical signals in cell membranes. Their principal subunits are members of a gene family and can function as voltage-gated ion channels by themselves. They are expressed in association with one or more auxiliary subunits which increase functional expression and modify the functional properties of the principal subunits. Structural elements which are required for voltage-dependent activation, selective ion conductance, and inactivation have been identified, and their mechanisms of action are being explored through mutagenesis, expression in heterologous cells, and functional analysis. These experiments reveal that these two channels are built on a common structural theme with variations appropriate for functional specialization of each channel type.  相似文献   

13.
The skeletal and cardiac muscle dihydropyridine receptors (DHPRs) differ with respect to their rates of channel activation and in the means by which they control Ca2+ release from the sarcoplasmic reticulum (Adams, B.A., and K.G. Beam. 1990. FASEB J. 4:2809-2816). We have examined the functional properties of skeletal (SkEIIIK) and cardiac (CEIIIK) DHPRs in which a highly conserved glutamate residue in the pore region of repeat III was mutated to a positively charged lysine residue. Using expression in dysgenic myotubes, we have characterized macroscopic ionic currents, intramembrane gating currents, and intracellular Ca2+ transients attributable to these two mutant DHPRs. CEIIIK supported very small inward Ca2+ currents at a few potentials (from -20 to +20 mV) and large outward cesium currents at potentials greater than +20 mV. SkEIIIK failed to support inward Ca2+ flux at any potential. However, large, slowly activating outward cesium currents were observed at all potentials greater than + 20 mV. The difference in skeletal and cardiac Ca2+ channel activation kinetics was conserved for outward currents through CEIIIK and SkEIIIK, even at very depolarized potentials (at +100 mV; SkEIIIK: tau(act) = 30.7 +/- 1.9 ms, n = 11; CEIIIK: tau(act) = 2.9 +/- 0.5 ms, n = 7). Expression of SkEIIIK in dysgenic myotubes restored both evoked contractions and depolarization-dependent intracellular Ca(2+) transients with parameters of voltage dependence (V(0.5) = 6.5 +/- 3.2 mV and k = 9.3 +/- 0.7 mV, n = 5) similar to those for the wild-type DHPR (Garcia, J., T. Tanabe, and K.G. Beam. 1994. J. Gen. Physiol. 103:125-147). However, CEIIIK-expressing myotubes never contracted and failed to exhibit depolarization-dependent intracellular Ca2+ transients at any potential. Thus, high Ca2+ permeation is required for cardiac-type excitation-contraction coupling reconstituted in dysgenic myotubes, but not skeletal-type. The strong rectification of the EIIIK channels made it possible to obtain measurements of gating currents upon repolarization to -50 mV (Qoff) following either brief (20 ms) or long (200 ms) depolarizing pulses to various test potentials. For SkEIIIK, and not CEIIK, Qoff was significantly (P < 0.001) larger after longer depolarizations to +60 mV (121.4 +/- 2.0%, n = 6). The increase in Qoff for long depolarizations exhibited a voltage dependence similar to that of channel activation. Thus, the increase in Q(off) may reflect a voltage sensor movement required for activation of L-type Ca2+ current and suggests that most DHPRs in skeletal muscle undergo this voltage-dependent transition.  相似文献   

14.
15.
Voltage-dependent calcium channels are classified into low voltage-activated and high voltage-activated channels. We have investigated the molecular basis for this difference in voltage dependence of activation by constructing chimeras between a low voltage-activated channel (Ca(V)3.1) and a high voltage-activated channel (Ca(V)1.2), focusing on steady-state activation properties. Wild type and chimeras were expressed in oocytes, and two-electrode voltage clamp recordings were made of calcium channel currents. Replacement of domains I, III, or IV of the Ca 3.1 channel with the corresponding domain of Ca(V)1.2 led (V)to high voltage-activated channels; for these constructs the current/voltage (I/V) curves were similar to those for Ca(V)1.2 wild type. However, replacement of domain II gave only a small shift to the right of the I/V curve and modulation of the activation kinetics but did not lead to a high voltage-activating channel with an I/V curve like Ca 1.2. We also investigated the role of the voltage sensor (V)S4 by replacing the S4 segment of Ca(V)3.1 with that of Ca 1.2. For domain I, there was no shift in the I/V curve (V)as compared with Ca(V)3.1, and there were relatively small shifts to the right for domains III and IV. Taken together, these results suggest that domains I, III, and IV (rather than domain II) are apparently critical for channel opening and, therefore, contribute strongly to the difference in voltage dependence of activation between Ca 3.1 and Ca(V)1.2. However, the S4 segments in domains I, (V)III, and IV did not account for this difference in voltage dependence.  相似文献   

16.
Low voltage activation of Ca(V)1.3 L-type Ca(2+) channels controls excitability in sensory cells and central neurons as well as sinoatrial node pacemaking. Ca(V)1.3-mediated pacemaking determines neuronal vulnerability of dopaminergic striatal neurons affected in Parkinson disease. We have previously found that in Ca(V)1.4 L-type Ca(2+) channels, activation, voltage, and calcium-dependent inactivation are controlled by an intrinsic distal C-terminal modulator. Because alternative splicing in the Ca(V)1.3 alpha1 subunit C terminus gives rise to a long (Ca(V)1.3(42)) and a short form (Ca(V)1.3(42A)), we investigated if a C-terminal modulatory mechanism also controls Ca(V)1.3 gating. The biophysical properties of both splice variants were compared after heterologous expression together with beta3 and alpha2delta1 subunits in HEK-293 cells. Activation of calcium current through Ca(V)1.3(42A) channels was more pronounced at negative voltages, and inactivation was faster because of enhanced calcium-dependent inactivation. By investigating several Ca(V)1.3 channel truncations, we restricted the modulator activity to the last 116 amino acids of the C terminus. The resulting Ca(V)1.3(DeltaC116) channels showed gating properties similar to Ca(V)1.3(42A) that were reverted by co-expression of the corresponding C-terminal peptide C(116). Fluorescence resonance energy transfer experiments confirmed an intramolecular protein interaction in the C terminus of Ca(V)1.3 channels that also modulates calmodulin binding. These experiments revealed a novel mechanism of channel modulation enabling cells to tightly control Ca(V)1.3 channel activity by alternative splicing. The absence of the C-terminal modulator in short splice forms facilitates Ca(V)1.3 channel activation at lower voltages expected to favor Ca(V)1.3 activity at threshold voltages as required for modulation of neuronal firing behavior and sinoatrial node pacemaking.  相似文献   

17.
18.
The lipid diacylglycerol (DAG) analogue 1‐oleoyl‐2‐acetyl‐sn‐glycerol (OAG) was used to verify the existence of DAG‐sensitive channels in cortical neurons dissociated from E13 mouse embryos. Calcium imaging experiments showed that OAG increased the cytosolic concentration of Ca2+ ([Ca2+]i) in nearly 35% of the KCl‐responsive cells. These Ca2+ responses disappeared in a Ca2+‐free medium supplemented with EGTA. Mn2+ quench experiments showed that OAG activated Ca2+‐conducting channels that were also permeant to Ba2+. The OAG‐induced Ca2+ responses were unaffected by nifedipine or omega‐conotoxin GVIA (Sigma‐Aldrich, Saint‐Quentin Fallavier, France) but blocked by 1‐[β‐(3‐(4‐Methoxyphenyl)propoxy)‐4‐methoxyphenethyl]‐1H‐imidazole hydrochloride (SKF)‐96365 and Gd3+. Replacing Na+ ions with N‐methyl‐d ‐glucamine diminished the amplitude of the OAG‐induced Ca2+ responses showing that the Ca2+ entry was mediated via Na+‐dependent and Na+‐independent mechanisms. Experiments carried out with the fluorescent Na+ indicator CoroNa Green showed that OAG elevated [Na+]i. Like OAG, the DAG lipase inhibitor RHC80267 increased [Ca2+]i but not the protein kinase C activator phorbol 12‐myristate 13‐acetate. Moreover, the OAG‐induced Ca2+ responses were not regulated by protein kinase C activation or inhibition but they were augmented by flufenamic acid which increases currents through C‐type transient receptor potential protein family (TRPC) 6 channels. In addition, application of hyperforin, a specific activator of TRPC6 channels, elevated [Ca2+]i. Whole‐cell patch‐clamp recordings showed that hyperforin activated non‐selective cation channels. They were blocked by SKF‐96365 but potentiated by flufenamic acid. Altogether, our data show the presence of hyperforin‐ and OAG‐sensitive Ca2+‐permeable channels displaying TRPC6‐like properties. This is the first report revealing the existence of second messenger‐operated channels in cortical neurons.  相似文献   

19.
Transepithelial transport of calcium involves uptake at the apical membrane, movement across the cell, and extrusion at the basolateral membrane. Active vitamin D metabolites regulate the latter two processes by induction of calbindin D and the plasma membrane ATPase (calcium pump), respectively. The expression of calbindin D and the calcium pump declines with age in parallel with transepithelial calcium transport. The apical uptake of calcium is thought to be mediated by the recently cloned calcium channels-CaT1 (or ECaC2, TRPV6) and CaT2 (or ECaC1, TRPV5). The purpose of these studies was to determine whether there were age-related changes in intestinal calcium channel regulation and to identify the dietary factors responsible for their regulation. Young (2 months) and adult (12 months) rats were fed either a high calcium or low calcium diet for 4 weeks. The low calcium diet significantly increased duodenal CaT1 and CaT2 mRNA levels in both age groups, but the levels in the adult were less than half that of the young. The changes in calcium channel expression with age and diet were significantly correlated with duodenal calcium transport and with calbindin D levels. To elucidate the relative roles of serum 1,25(OH)2D3 and calcium in the regulation of calcium channel expression, young rats were fed diets containing varying amounts of calcium and vitamin D. Dietary vitamin D or exogenous 1,25(OH)2D3 more than doubled CaT1 mRNA levels, and this regulation was independent of dietary or serum calcium. These findings suggest that the apical calcium channels, along with calbindin and the calcium pump, may play a role in intestinal calcium transport and its modulation by age, dietary calcium, and 1,25(OH)2D3.  相似文献   

20.
We have previously demonstrated that formation of a complex between L-type calcium (Ca(2+)) channel alpha(1C) (Ca(V)1.2) and beta subunits was necessary to target the channels to the plasma membrane when expressed in tsA201 cells. In the present study, we identified a region in the C terminus of the alpha(1C) subunit that was required for membrane targeting. Using a series of C-terminal deletion mutants of the alpha(1C) subunit, a domain consisting of amino acid residues 1623-1666 ("targeting domain") in the C terminus of the alpha(1C) subunit has been identified to be important for correct targeting of L-type Ca(2+) channel complexes to the plasma membrane. Although cells expressing the wild-type alpha(1C) and beta(2a) subunits exhibited punctate clusters of channel complexes along the plasma membrane with little intracellular staining, co-expression of deletion mutants of the alpha(1C) subunit that lack the targeting domain with the beta(2a) subunit resulted in an intracellular localization of the channels. In addition, three other regions in the C terminus of the alpha(1C) subunit that were downstream of residues 1623-1666 were found to contribute to membrane targeting of the L-type channels. Deletion of these domains in the alpha(1C) subunit resulted in a reduction of plasma membrane-localized channels, and a concomitant increase in channels localized intracellularly. Taken together, these results have demonstrated that a targeting domain in the C terminus of the alpha(1C) subunit was required for proper plasma membrane localization of the L-type Ca(2+) channels.  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