首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The dihydropyridine (DHP)-binding site has been identified within L-type Ca(2+) channel alpha(1C) subunit. However, the molecular mechanism underlying modulation of Ca(2+) channel gating by DHPs has not been clarified. To search for novel determinants of high affinity DHP binding, we introduced point mutations in the rat brain Ca(2+) channel alpha(1C) subunit (rbCII or Ca(v)1.2c) based on the comparison of amino acid sequences between rbCII and the ascidian L-type Ca(2+) channel alpha(1) subunit, which is insensitive to DHPs. The alpha(1C) mutants (S1115A, S1146A, and A1420S) and rbCII were transiently expressed in BHK6 cells with beta(1a) and alpha(2)/delta subunits. The mutation did not affect the electrophysiological properties of the Ca(2+) channel, or the voltage- and concentration-dependent block of Ca(2+) channel currents produced by diltiazem and verapamil. However, the S1115A channel was significantly less sensitive to DHP antagonists. Interestingly, in the S1115A channel, DHP agonists failed to enhance whole-cell Ca(2+) channel currents and the prolongation of mean open time, as well as the increment of NP(o). Responsiveness to the non-DHP agonist FPL-64176 was also markedly reduced in the S1115A channel. When S1115 was replaced by other amino acids (S1115D, S1115T, or S1115V), only S1115T was slightly sensitive to S-(-)-Bay K 8644. These results indicate that the hydroxyl group of Ser(1115) in IIIS5-S6 linker of the L-type Ca(2+) channel alpha(1C) subunit plays a critical role in DHP binding and in the action of DHP Ca(2+) channel agonists.  相似文献   

2.
Cloning of three members of low-voltage-activated (LVA) calcium channel family, predominantly neuronal alpha1G and alpha1I, and ubiquitous alpha1H, enabled to investigate directly their electrophysiological and pharmacological profile as well as their putative subunit composition. All the three channels are half-activated at membrane potential about -40 mV and half-inactivated at about -70 mV. Kinetics of alpha1G and alpha1H channels activation and inactivation are similar and faster than that of alpha1I channel. All the three channels are blocked with high affinity by the organic blocker mibefradil. Another high affinity blocker is kurtoxin. Cloned LVA channels are relatively insensitive to antiepileptics, dihydropyridines and omega-conotoxins. Ni2+ is high affinity blocker of alpha1H channel only. Amiloride inhibits the alpha1H channel. The subunit composition of LVA channel remains unclear. Out of known high-voltage-activated calcium channel subunits, alpha2delta-2 and gamma-5 subunits significantly and systematically modified activation and/or inactivation of the current. In contrast, alpha2delta-1, alpha2delta-3, gamma-2 and gamma-4 subunits failed to modulate the current or had only minor effects.  相似文献   

3.
The intracellular loops that interlink the four transmembrane domains of Ca(2+)- and Na(+)-channels (Ca(v), Na(v)) have critical roles in numerous forms of channel regulation. In particular, the intracellular loop that joins repeats I and II (I-II loop) in high voltage-activated (HVA) Ca(2+) channels possesses the binding site for Ca(v)beta subunits and plays significant roles in channel function, including trafficking the alpha(1) subunits of HVA channels to the plasma membrane and channel gating. Although there is considerable divergence in the primary sequence of the I-II loop of Ca(v)1/Ca(v)2 HVA channels and Ca(v)3 LVA/T-type channels, evidence for a regulatory role of the I-II loop in T-channel function has recently emerged for Ca(v)3.2 channels. In order to provide a comprehensive view of the role this intracellular region may play in the gating and surface expression in Ca(v)3 channels, we have performed a structure-function analysis of the I-II loop in Ca(v)3.1 and Ca(v)3.3 channels using selective deletion mutants. Here we show the first 60 amino acids of the loop (post IS6) are involved in Ca(v)3.1 and Ca(v)3.3 channel gating and kinetics, which establishes a conserved property of this locus for all Ca(v)3 channels. In contrast to findings in Ca(v)3.2, deletion of the central region of the I-II loop in Ca(v)3.1 and Ca(v)3.3 yielded a modest increase (+30%) and a reduction (-30%) in current density and surface expression, respectively. These experiments enrich our understanding of the structural determinants involved in Ca(v)3 function by highlighting the unique role played by the intracellular I-II loop in Ca(v)3.2 channel trafficking, and illustrating the prominent role of the gating brake in setting the slow and distinctive slow activation kinetics of Ca(v)3.3.  相似文献   

4.
Mutations of eight to nine amino acids of IIIS5, IIIS6 and IVS6 segments were shown to reconstruct the dihydropyridine (DHP) interaction site in the non-L-type alpha1E or alpha1A calcium channels. The reconstructed site enabled enantiomer-selective inhibition and activation of the expressed chimeras by DHPs but failed to transfer voltage dependence of the current inhibition. Here we show that transfer of four non-conserved amino acids from the IS6 segment to the DHP-sensitive alpha1E chimera increased the inhibition by (+)isradipine at the hyperpolarized membrane potential of -100 mV and enhanced the voltage-dependent block.  相似文献   

5.
Replacement of L-type Ca(2+) channel alpha(1) subunit residue Thr-1066 in segment IIIS5 by a tyrosine residue conserved in the corresponding positions of non-L-type Ca(2+) channels eliminates high dihydropyridine sensitivity through a steric mechanism. To determine the effects of this mutation on phenylalkylamine interaction, we exploited the availability of Ca(v)1.2DHP(-/-) mice containing the T1066Y mutation. In contrast to dihydropyridines, increased protein-dependent binding of the phenylalkylamine (-)-[(3)H]devapamil occurred to Ca(v)1.2DHP(-/-) mouse brain microsomes. This effect could be attributed to an at least 2-fold increase in affinity as determined by saturation analysis and binding inhibition experiments. The latter also revealed a higher affinity for (-)-verapamil but not for (-)-gallopamil. The mutation caused a pronounced slowing of (-)-[(3)H]devapamil dissociation, indicating a stabilization of the drug-channel complex. The increased affinity of mutant channels was also evident in functional studies after heterologous expression of wild type and T1066Y channels in Xenopus laevis oocytes. 100 mum (-)-verapamil inhibited a significantly larger fraction of Ba(2+) inward current through mutant than through WT channels. Our results provide evidence that phenylalkylamines also interact with the IIIS5 helix and that the geometry of the IIIS5 helix affects the access and/or binding of different chemical classes of Ca(2+) channel blockers to their overlapping binding domains. Mutation of Thr-1066 to a non-L-type tyrosine residue can be exploited to differentially affect phenylalkylamine and dihydropyridine binding to L-type Ca(2+) channels.  相似文献   

6.
7.
It has been suggested that the auxiliary subunits of high voltage-activated (HVA) calcium channels modulate T-type, low voltage-activated (LVA) calcium channels. Such a regulation has yet to be documented, especially because there has been no biochemical characterization of T-channels. To monitor total protein levels and plasma membrane expression of T-channels in living cells, external epitopes (hemagglutinin, FLAG) were introduced into human recombinant Ca(V)3 channels that were also N-terminally fused to green fluorescent protein. Utilizing Western blot techniques, fluorescence flow cytometry, immunofluorescence, luminometry, and electrophysiology, we describe here that beta(1b) and alpha(2)-delta(1) subunits enhance the level of Ca(V)3 proteins as well as their plasma membrane expression in various expression systems. We also report that, in both Xenopus oocytes and mammalian cells, the alpha(2)-delta(1) subunits increase by at least and beta(1b) 2-fold the current density of Ca(V)3 channels with no change in the electrophysiological properties. Altogether, these data indicate that HVA auxiliary subunits modulate Ca(V)3 channel surface expression, suggesting that the membrane targeting of HVA and LVA alpha(1) subunits is regulated dynamically through the expression of a common set of regulatory subunits.  相似文献   

8.
9.
The structural determinant of the permeation and selectivity properties of high voltage-activated (HVA) Ca(2+) channels is a locus formed by four glutamate residues (EEEE), one in each P-region of the domains I-IV of the alpha(1) subunit. We tested whether the divergent aspartate residues of the EEDD locus of low voltage-activated (LVA or T-type) Ca(2+) channels account for the distinctive permeation and selectivity features of these channels. Using the whole-cell patch-clamp technique in the HEK293 expression system, we studied the properties of the alpha(1G) T-type, the alpha(1C) L-type Ca(2+) channel subunits, and alpha(1G) pore mutants, containing aspartate-to-glutamate conversions in domain III, domain IV, or both. Three characteristic features of HVA Ca(2+) channel permeation, i.e. (a) Ba(2+) over Ca(2+) permeability, (b) Ca(2+)/Ba(2+) anomalous mole fraction effect (AMFE), and (c) high Cd(2+) sensitivity, were conferred on the domain III mutant (EEED) of alpha(1G). In contrast, the relative Ca(2+)/Ba(2+) permeability and the lack of AMFE of the alpha(1G) wild type channel were retained in the domain IV mutant (EEDE). The double mutant (EEEE) displayed AMFE and a Cd(2+) sensitivity similar to that of alpha(1C), but currents were larger in Ca(2+)- than in Ba(2+)-containing solutions. The mutation in domain III, but not that in domain IV, consistently displayed outward fluxes of monovalent cations. H(+) blocked Ca(2+) currents in all mutants more efficiently than in alpha(1G). In addition, activation curves of all mutants were displaced to more positive voltages and had a larger slope factor than in alpha(1G) wild type. We conclude that the aspartate residues of the EEDD locus of the alpha(1G) Ca(2+) channel subunit not only control its permeation properties, but also affect its activation curve. The mutation of both divergent aspartates only partially confers HVA channel permeation properties to the alpha(1G) Ca(2+) channel subunit.  相似文献   

10.
We have positionally cloned and characterized a new calcium channel auxiliary subunit, alpha(2)delta-2 (CACNA2D2), which shares 56% amino acid identity with the known alpha(2)delta-1 subunit. The gene maps to the critical human tumor suppressor gene region in chromosome 3p21.3, showing very frequent allele loss and occasional homozygous deletions in lung, breast, and other cancers. The tissue distribution of alpha(2)delta-2 expression is different from alpha(2)delta-1, and alpha(2)delta-2 mRNA is most abundantly expressed in lung and testis and well expressed in brain, heart, and pancreas. In contrast, alpha(2)delta-1 is expressed predominantly in brain, heart, and skeletal muscle. When co-expressed (via cRNA injections) with alpha(1B) and beta(3) subunits in Xenopus oocytes, alpha(2)delta-2 increased peak size of the N-type Ca(2+) currents 9-fold, and when co-expressed with alpha(1C) or alpha(1G) subunits in Xenopus oocytes increased peak size of L-type channels 2-fold and T-type channels 1.8-fold, respectively. Anti-peptide antibodies detect the expression of a 129-kDa alpha(2)delta-2 polypeptide in some but not all lung tumor cells. We conclude that the alpha(2)delta-2 gene encodes a functional auxiliary subunit of voltage-gated Ca(2+) channels. Because of its chromosomal location and expression patterns, CACNA2D2 needs to be explored as a potential tumor suppressor gene linking Ca(2+) signaling and lung, breast, and other cancer pathogenesis. The homologous location on mouse chromosome 9 is also the site of the mouse neurologic mutant ducky (du), and thus, CACNA2D2 is also a candidate gene for this inherited idiopathic generalized epilepsy syndrome.  相似文献   

11.
12.
The molecular basis of the Ca2+ channel block by (+)-cis-diltiazem was studied in class A/L-type chimeras and mutant alpha1C-a Ca2+ channels. Chimeras consisted of either rabbit heart (alpha1C-a) or carp skeletal muscle (alpha1S) sequence in transmembrane segments IIIS6, IVS6, and adjacent S5-S6 linkers. Only chimeras containing sequences from alpha1C-a were efficiently blocked by (+)-cis-diltiazem, whereas the phenylalkylamine (-)-gallopamil efficiently blocked both constructs. Carp skeletal muscle and rabbit heart Ca2+ channel alpha1 subunits differ with respect to two nonconserved amino acids in segments IVS6. Transfer of a single leucine (Leu1383, located at the extracellular mouth of the pore) from IVS6 alpha1C-a to IVS6 of alpha1S significantly increased the (+)-cis-diltiazem sensitivity of the corresponding mutant L1383I. An analysis of the role of the two heterologous amino acids in a L-type alpha1 subunit revealed that corresponding amino acids in position 1487 (outer channel mouth) determine recovery of resting Ca2+ channels from block by (+)-cis-diltiazem. The second heterologous amino acid in position 1504 of segment IVS6 (inner channel mouth) was identified as crucial inactivation determinant of L-type Ca2+ channels. This residue simultaneously modulates drug binding during membrane depolarization. Our study provides the first evidence for a guarded and modulated benzothiazepine receptor on L-type channels.  相似文献   

13.
Molecular cloning and expression studies established the existence of three T-type Ca(2+) channel (Ca(v)3) alpha(1) subunits: Ca(v)3.1 (alpha(1G)), Ca(v)3.2 (alpha(1H)), and Ca(v)3.3 (alpha(1I)). Although all three channels are low voltage-activated, they display considerable differences in their kinetics, with Ca(v)3.1 and Ca(v)3.2 channels activating and inactivating much faster than Ca(v)3.3 channels. The goal of the present study was to determine the structural elements that confer the distinctively slow kinetics of Ca(v)3.3 channels. To address this question, a series of chimeric channels between Ca(v)3.1 and Ca(v)3.3 channels were constructed and expressed in Xenopus oocytes. Kinetic analysis showed that the slow activation and inactivation kinetics of the Ca(v)3.3 channel were not completely abolished by substitution with any one portion of the Ca(v)3.1 channel. Likewise, the Ca(v)3.1 channel failed to acquire the slow kinetics by simply adopting one portion of the Ca(v)3.3 channel. These findings suggest that multiple structural elements contribute to the slow kinetics of Ca(v)3.3 channels.  相似文献   

14.
Native smooth muscle L-type Ca(v)1.2 calcium channels have been shown to support a fraction of Ca(2+) currents with a window current that is close to resting potential. The smooth muscle L-type Ca(2+) channels are also more susceptible to inhibition by dihydropyridines (DHPs) than the cardiac channels. It was hypothesized that smooth muscle Ca(v)1.2 channels exhibiting hyperpolarized shift in steady-state inactivation would contribute to larger inhibition by DHP, in addition to structural differences of the channels generated by alternative splicing that modulate DHP sensitivities. In addition, it has also been shown that alternative splicing modulates DHP sensitivities by generating structural differences in the Ca(v)1.2 channels. Here, we report a smooth muscle L-type Ca(v)1.2 calcium channel splice variant, Ca(v)1.2SM (1/8/9(*)/32/Delta33), that when expressed in HEK 293 cells display hyperpolarized shifts for steady-state inactivation and activation potentials when compared with the established Ca(v)1.2b clone (1/8/9(*)/32/33). This variant activates from more negative potentials and generates a window current closer to resting membrane potential. We also identified the predominant cardiac isoform Ca(v)1.2CM clone (1a/8a/Delta9(*)/32/33) that is different from the established Ca(v)1.2a (1a/8a/Delta9(*)/31/33). Importantly, Ca(v)1.2SM channels were shown to be more sensitive to nifedipine blockade than Ca(v)1.2b and cardiac Ca(v)1.2CM channels when currents were recorded in either 5 mM Ba(2+) or 1.8 mM Ca(2+) external solutions. This is the first time that a smooth muscle Ca(v)1.2 splice variant has been identified functionally to possess biophysical property that can be linked to enhanced state-dependent block by DHP.  相似文献   

15.
16.
Pan ZH  Hu HJ  Perring P  Andrade R 《Neuron》2001,32(1):89-98
Transmitter release in neurons is thought to be mediated exclusively by high-voltage-activated (HVA) Ca(2+) channels. However, we now report that, in retinal bipolar cells, low-voltage-activated (LVA) Ca(2+) channels also mediate neurotransmitter release. Bipolar cells are specialized neurons that release neurotransmitter in response to graded depolarizations. Here we show that these cells express T-type Ca(2+) channel subunits and functional LVA Ca(2+) currents sensitive to mibefradil. Activation of these currents results in Ca(2+) influx into presynaptic terminals and exocytosis, which we detected as a capacitance increase in isolated terminals and the appearance of reciprocal currents in retinal slices. The involvement of T-type Ca(2+) channels in bipolar cell transmitter release may contribute to retinal information processing.  相似文献   

17.
L-type Ca(2+) channels in native tissues have been found to contain a pore-forming alpha(1) subunit that is often truncated at the C terminus. However, the C terminus contains many important domains that regulate channel function. To test the hypothesis that C-terminal fragments may associate with and regulate C-terminal-truncated alpha(1C) (Ca(V)1.2) subunits, we performed electrophysiological and biochemical experiments. In tsA201 cells expressing either wild type or C-terminal-truncated alpha(1C) subunits in combination with a beta(2a) subunit, truncation of the alpha(1C) subunit by as little as 147 amino acids led to a 10-15-fold increase in currents compared with those obtained from control, full-length alpha(1C) subunits. Dialysis of cells expressing the truncated alpha(1C) subunits with C-terminal fragments applied through the patch pipette reconstituted the inhibition of the channels seen with full-length alpha(1C) subunits. In addition, C-terminal deletion mutants containing a tethered C terminus also exhibited the C-terminal-induced inhibition. Immunoprecipitation assays demonstrated the association of the C-terminal fragments with truncated alpha(1C) subunits. In addition, glutathione S-transferase pull-down assays demonstrated that the C-terminal inhibitory fragment could associate with at least two domains within the C terminus. The results support the hypothesis the C- terminal fragments of the alpha(1C) subunit can associate with C-terminal-truncated alpha(1C) subunits and inhibit the currents through L-type Ca(2+) channels.  相似文献   

18.
1,4‐Dihydropyridines (DHPs) have been developed to treat hypertension, angina, and nerve system disease. They are thought to mainly target the L‐type calcium channels, but low selectivity prompts them to block Cav1.2 and Cav3.1 channels simultaneously. Recently, some novel DHPs with different hydrophobic groups have been synthesized and among them M12 has a higher selectivity for Cav3.1. However, the structural information about Cav3.1‐DHPs complexes is not available in the experiment. Thus, we combined homology modeling, molecular docking, molecular dynamics simulations, and binding free energy calculations to quantitatively elucidate the inhibition mechanism of DHPs. The calculated results indicate that our model is in excellent agreement with experimental results. On the basis of conformational analysis, we identify the main interactions between DHPs and calcium channels and further elaborate on the different selectivity of ligands from the micro perspective. In conjunction with energy distribution, we propose that the binding sites of Cav3.1‐DHPs is characterized by several interspersed hydrophobic amino acid residues on the IIIS6 and IVS6 segments. We also speculate the favorable function groups on prospective DHPs. Besides, our model provides important information for further mutagenesis experiments.  相似文献   

19.
Ikeda M  Matsumoto S 《Life sciences》2003,73(9):1175-1187
We examined the subtypes and characteristics of the Ca(2+) channel in small (diameter < 30 microm) trigeminal ganglion (TG) neurons from neonatal rats by means of whole cell patch clamp techniques. There were two current components, low-voltage activated (LVA) and high-voltage activated (HVA) I(Ba), with different activation ranges and waveforms. LVA I(Ba) elicited from a depolarizing step pulse at a holding potential (HP) of -80 mV was inhibited by 0.25 mM amiloride (62%), which did not produce any significant inhibition of the peak amplitude of HVA I(Ba). The application of 0.5 mM amiloride inhibited 10% of the HVA I(Ba). The LVA I(Ba) was also reduced by changing the HP from -80 to -60 mV (61%), and under these conditions the peak amplitude of HVA I(Ba) did not change significantly. In addition, HVA I(Ba) and LVA I(Ba) showed marked differences in their inactivation properties. Experiments with several Ca(2+) channel blockers revealed that on average, 26% of the HVA I(Ba) was nifedipine (10 microM) sensitive, 55% was sensitive to omega-conotoxinGVIA (1 microM), 4% was blocked by omega-agatoxinIVA (1 microM), and the remainder of the current that was resistant to the co-application of all three Ca(2+) channel blockers was 15% of the total current. These results suggest that the application of amiloride and the alteration of the holding potential level can discriminate between HVA and LVA Ba(2+) currents in TG neurons, and that TG neurons expressed T-, L-, N-, P-/Q- and R-type Ca(2+) channels.  相似文献   

20.
Voltage-gated sodium channels consist of a pore-forming alpha subunit associated with beta1 subunits and, for brain sodium channels, beta2 subunits. Although much is known about the structure and function of the alpha subunit, there is little information on the functional role of the 16 extracellular loops. To search for potential functional activities of these extracellular segments, chimeras were studied in which an individual extracellular loop of the rat heart (rH1) alpha subunit was substituted for the corresponding segment of the rat brain type IIA (rIIA) alpha subunit. In comparison with rH1, wild-type rIIA alpha subunits are characterized by more positive voltage-dependent activation and inactivation, a more prominent slow gating mode, and a more substantial shift to the fast gating mode upon coexpression of beta1 subunits in Xenopus oocytes. When alpha subunits were expressed alone, chimeras with substitutions from rH1 in five extracellular loops (IIS5-SS1, IISS2-S6, IIIS1-S2, IIISS2-S6, and IVS3-S4) had negatively shifted activation, and chimeras with substitutions in three of these (IISS2-S6, IIIS1-S2, and IVS3-S4) also had negatively shifted steady-state inactivation. rIIA alpha subunit chimeras with substitutions from rH1 in five extracellular loops (IS5-SS1, ISS2-S6, IISS2-S6, IIIS1-S2, and IVS3-S4) favored the fast gating mode. Like wild-type rIIA alpha subunits, all of the chimeric rIIA alpha subunits except chimera IVSS2-S6 were shifted almost entirely to the fast gating mode when coexpressed with beta1 subunits. In contrast, substitution of extracellular loop IVSS2-S6 substantially reduced the effectiveness of beta1 subunits in shifting rIIA alpha subunits to the fast gating mode. Our results show that multiple extracellular loops influence voltage-dependent activation and inactivation and gating mode of sodium channels, whereas segment IVSS2-S6 plays a dominant role in modulation of gating by beta1 subunits. Evidently, several extracellular loops are important determinants of sodium channel gating and modulation.  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