首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Alternative splicing of the skeletal muscle CaV1.1 voltage-gated calcium channel gives rise to two channel variants with very different gating properties. The currents of both channels activate slowly; however, insertion of exon 29 in the adult splice variant CaV1.1a causes an ∼30-mV right shift in the voltage dependence of activation. Existing evidence suggests that the S3–S4 linker in repeat IV (containing exon 29) regulates voltage sensitivity in this voltage-sensing domain (VSD) by modulating interactions between the adjacent transmembrane segments IVS3 and IVS4. However, activation kinetics are thought to be determined by corresponding structures in repeat I. Here, we use patch-clamp analysis of dysgenic (CaV1.1 null) myotubes reconstituted with CaV1.1 mutants and chimeras to identify the specific roles of these regions in regulating channel gating properties. Using site-directed mutagenesis, we demonstrate that the structure and/or hydrophobicity of the IVS3–S4 linker is critical for regulating voltage sensitivity in the IV VSD, but by itself cannot modulate voltage sensitivity in the I VSD. Swapping sequence domains between the I and the IV VSDs reveals that IVS4 plus the IVS3–S4 linker is sufficient to confer CaV1.1a-like voltage dependence to the I VSD and that the IS3–S4 linker plus IS4 is sufficient to transfer CaV1.1e-like voltage dependence to the IV VSD. Any mismatch of transmembrane helices S3 and S4 from the I and IV VSDs causes a right shift of voltage sensitivity, indicating that regulation of voltage sensitivity by the IVS3–S4 linker requires specific interaction of IVS4 with its corresponding IVS3 segment. In contrast, slow current kinetics are perturbed by any heterologous sequences inserted into the I VSD and cannot be transferred by moving VSD I sequences to VSD IV. Thus, CaV1.1 calcium channels are organized in a modular manner, and control of voltage sensitivity and activation kinetics is accomplished by specific molecular mechanisms within the IV and I VSDs, respectively.  相似文献   

2.
《Biophysical journal》2021,120(20):4429-4441
The voltage-gated calcium channel CaV1.1 belongs to the family of pseudo-heterotetrameric cation channels, which are built of four structurally and functionally distinct voltage-sensing domains (VSDs) arranged around a common channel pore. Upon depolarization, positive gating charges in the S4 helices of each VSD are moved across the membrane electric field, thus generating the conformational change that prompts channel opening. This sliding helix mechanism is aided by the transient formation of ion-pair interactions with countercharges located in the S2 and S3 helices within the VSDs. Recently, we identified a domain-specific ion-pair partner of R1 and R2 in VSD IV of CaV1.1 that stabilizes the activated state of this VSD and regulates the voltage dependence of current activation in a splicing-dependent manner. Structure modeling of the entire CaV1.1 in a membrane environment now revealed the participation in this process of an additional putative ion-pair partner (E216) located outside VSD IV, in the pore domain of the first repeat (IS5). This interdomain interaction is specific for CaV1.1 and CaV1.2 L-type calcium channels. Moreover, in CaV1.1 it is sensitive to insertion of the 19 amino acid peptide encoded by exon 29. Whole-cell patch-clamp recordings in dysgenic myotubes reconstituted with wild-type or E216 mutants of GFP-CaV1.1e (lacking exon 29) showed that charge neutralization (E216Q) or removal of the side chain (E216A) significantly shifted the voltage dependence of activation (V1/2) to more positive potentials, suggesting that E216 stabilizes the activated state. Insertion of exon 29 in the GFP-CaV1.1a splice variant strongly reduced the ionic interactions with R1 and R2 and caused a substantial right shift of V1/2, whereas no further shift of V1/2 was observed on substitution of E216 with A or Q. Together with our previous findings, these results demonstrate that inter- and intradomain ion-pair interactions cooperate in the molecular mechanism regulating VSD function and channel gating in CaV1.1.  相似文献   

3.
Study reveals how a slowly activating calcium channel is able to control rapid excitation–contraction coupling in skeletal muscle.

Skeletal muscle contraction is initiated by action potentials that depolarize the muscle fiber and trigger the rapid release of Ca2+ from the SR via RYR1 channels. This process of excitation–contraction coupling depends on voltage-gated CaV1.1 channels in the plasma membrane, or sarcolemma, of muscle fibers. But CaV1.1 channels are only slowly activated by changes in the sarcolemma membrane potential, and it is therefore unclear how they are able to trigger the much faster activation of RYR1 channels. In this issue of JGP, Savalli et al. reveal that this paradox can be explained by the fact that each of CaV1.1’s four voltage-sensing domains (VSDs) have distinct biophysical properties (1).Nicoletta Savalli (left), Riccardo Olcese (center), and colleagues reveal the distinct physical properties of the CaV1.1 channel’s four voltage-sensing domains (VSD I–IV, right). VSD-I shows slow activation kinetics and is the main contributor to the opening of CaV1.1. The other VSDs activate much faster and may therefore be coupled to RYR1 to mediate the rapid release of Ca2+ from the SR during skeletal muscle contraction.RYR1 channels have no voltage-sensing machinery of their own and therefore rely on a physical connection to CaV1.1 channels to release Ca2+ and initiate muscle contraction in response to muscle fiber depolarization. But RYR1 channels open ∼25 times faster than CaV1.1 channels. “So, how can these slowly activating CaV1.1 channels trigger the rapid release of Ca2+ from the SR?” asks Riccardo Olcese, a professor at the David Geffen School of Medicine, UCLA.Olcese and colleagues, including Assistant Project Scientist Nicoletta Savalli, suspected that the answer might lie in the fact that, like many other voltage-gated ion channels, CaV1.1 has four VSDs that alter their conformation in response to voltage changes. These domains are similar, but not identical, to each other, potentially enabling them to have distinct biophysical properties and perform distinct functions. Indeed, Olcese and colleagues previously demonstrated that, in the closely related channel CaV1.2, only VSDs II and III are involved in pore opening (2, 3).Savalli et al. used voltage-clamp fluorometry to compare the properties of CaV1.1’s VSDs, expressing the channel in Xenopus oocytes and labeling each of its VSDs in turn with an environmentally sensitive fluorophore to report voltage-dependent changes in their conformation (1). “We found that the four VSDs were very heterogenous in both their kinetics and voltage dependencies,” says Olcese. “VSD-I had very slow kinetics, compatible with the slow activation of the CaV1.1 pore. The other three VSDs had much faster kinetics and could, therefore, be good candidates to be the voltage sensors for RYR1 activation.”Olcese and colleagues confirmed the importance of VSD-I for CaV1.1 activation by analyzing a naturally occurring, charge-neutralizing mutation in this domain, R174W, that is linked to malignant hyperthermia (4). The team found that this mutation reduced the voltage-sensitivity of VSD-I and abolished the ability of CaV1.1 to conduct Ca2+ at physiological membrane potentials, but had no effect on the behavior of the other three VSDs.Finally, Savalli et al. applied their data on both the wild-type and mutant VSDs to an allosteric model of CaV activation (2, 3), which predicted that VSD-I contributes most of the energy required to stabilize the open state of CaV1.1, while the other VSDs contribute little to nothing.Thus, CaV1.1 activation is mainly driven by a single VSD—a mechanism that hasn’t been seen in any other voltage-gated ion channel—leaving the other VSDs free to perform other functions, such as the rapid activation of RYR1. Olcese and colleagues now want to pinpoint exactly which VSD(s) are coupled to RYR1 and determine how they trigger rapid Ca2+ release from the SR.  相似文献   

4.
Initiation of skeletal muscle contraction is triggered by rapid activation of RYR1 channels in response to sarcolemmal depolarization. RYR1 is intracellular and has no voltage-sensing structures, but it is coupled with the voltage-sensing apparatus of CaV1.1 channels to inherit voltage sensitivity. Using an opto-electrophysiological approach, we resolved the excitation-driven molecular events controlling both CaV1.1 and RYR1 activations, reported as fluorescence changes. We discovered that each of the four human CaV1.1 voltage-sensing domains (VSDs) exhibits unique biophysical properties: VSD-I time-dependent properties were similar to ionic current activation kinetics, suggesting a critical role of this voltage sensor in CaV1.1 activation; VSD-II, VSD-III, and VSD-IV displayed faster activation, compatible with kinetics of sarcoplasmic reticulum Ca2+ release. The prominent role of VSD-I in governing CaV1.1 activation was also confirmed using a naturally occurring, charge-neutralizing mutation in VSD-I (R174W). This mutation abolished CaV1.1 current at physiological membrane potentials by impairing VSD-I activation without affecting the other VSDs. Using a structurally relevant allosteric model of CaV activation, which accounted for both time- and voltage-dependent properties of CaV1.1, to predict VSD-pore coupling energies, we found that VSD-I contributed the most energy (~75 meV or ∼3 kT) toward the stabilization of the open states of the channel, with smaller (VSD-IV) or negligible (VSDs II and III) energetic contribution from the other voltage sensors (<25 meV or ∼1 kT). This study settles the longstanding question of how CaV1.1, a slowly activating channel, can trigger RYR1 rapid activation, and reveals a new mechanism for voltage-dependent activation in ion channels, whereby pore opening of human CaV1.1 channels is primarily driven by the activation of one voltage sensor, a mechanism distinct from that of all other voltage-gated channels.  相似文献   

5.
Extracellular S3-S4 linkers of domain IV (IVS3-S4) of L-type Ca2+ channels (CaV1) are subject to alternative splicing, resulting into distinct gating profiles serving for diverse physiological roles. However, it has remained elusive what would be the determining factor of IVS3-S4 effects on CaV1 channels. In this study, we systematically compared IVS3-S4 variants from CaV1.1-1.4, and discover that the flexibility of the linker plays a prominent role in gating characteristics. Chimeric analysis and mutagenesis demonstrated that changes in half activation voltage (V1/2) or activation time constant (τ) are positively correlated with the numbers of flexible glycine residues within the linker. Moreover, antibodies that reduce IVS3-S4 flexibility negatively shifted V1/2, emerging as a new category of CaV1 enhancers. In summary, our results suggest that the flexibility or rigidity of IVS3-S4 linker underlies its modulations on CaV1 activation (V1/2 and τ), paving the way to dissect the core mechanisms and to develop innovative perturbations pertaining to voltage-sensing S4 and its vicinities.  相似文献   

6.
Mutations in the voltage sensor domain (VSD) of CaV1.1, the α1S subunit of the L-type calcium channel in skeletal muscle, are an established cause of hypokalemic periodic paralysis (HypoPP). Of the 10 reported mutations, 9 are missense substitutions of outer arginine residues (R1 or R2) in the S4 transmembrane segments of the homologous domain II, III (DIII), or IV. The prevailing view is that R/X mutations create an anomalous ion conduction pathway in the VSD, and this so-called gating pore current is the basis for paradoxical depolarization of the resting potential and weakness in low potassium for HypoPP fibers. Gating pore currents have been observed for four of the five CaV1.1 HypoPP mutant channels studied to date, the one exception being the charge-conserving R897K in R1 of DIII. We tested whether gating pore currents are detectable for the other three HypoPP CaV1.1 mutations in DIII. For the less conserved R1 mutation, R897S, gating pore currents with exceptionally large amplitude were observed, correlating with the severe clinical phenotype of these patients. At the R2 residue, gating pore currents were detected for R900G but not R900S. These findings show that gating pore currents may occur with missense mutations at R1 or R2 in S4 of DIII and that the magnitude of this anomalous inward current is mutation specific.  相似文献   

7.
A channelopathy mutation in segment IIS6 of CaV1.4 (I745T) has been shown to cause severe visual impairment by shifting the activation and inactivation curves to more hyperpolarised voltages and slowing activation and inactivation kinetics. A similar gating phenotype is caused by the corresponding mutation, I781T, in CaV1.2 (midpoint of activation curve (V0.5) shifted to -37.7 ± 1.2 mV). We show here that wild type gating can partially be restored by a helix stabilising rescue mutation N785A. V0.5 of I781T/N785A (V0.5 = -21.5 ± 0.6 mV) was shifted back towards wild type (V0.5 = -9.9±1.1 mV). Homology models developed in our group (see accompanying article for details) were used to perform MD-simulations on wild-type and mutant channels. Systematic changes in segment IIIS6 (M1187 - F1194) and in helix IIS6 (N785-L786) were observed. The simulated structural changes in S6 segments of I781T/N785A were less pronounced than in I781T. A delicate balance between helix flexibility and stability enabling the formation of hydrophobic seals at the inner channel mouth appears to be important for wild type CaV1.2 gating. Our study illustrates that effects of mutations in the lower part of IIS6 may not be localized to the residue or even segment being mutated, but may affect conformations of interacting segments.  相似文献   

8.
Activity of voltage-gated Cav1.3 L-type Ca2+ channels is required for proper hearing as well as sinoatrial node and brain function. This critically depends on their negative activation voltage range, which is further fine-tuned by alternative splicing. Shorter variants miss a C-terminal regulatory domain (CTM), which allows them to activate at even more negative potentials than C-terminally long-splice variants. It is at present unclear whether this is due to an increased voltage sensitivity of the Cav1.3 voltage-sensing domain, or an enhanced coupling of voltage-sensor conformational changes to the subsequent opening of the activation gate. We studied the voltage-dependence of voltage-sensor charge movement (QON-V) and of current activation (ICa-V) of the long (Cav1.3L) and a short Cav1.3 splice variant (Cav1.342A) expressed in tsA-201 cells using whole cell patch-clamp. Charge movement (QON) of Cav1.3L displayed a much steeper voltage-dependence and a more negative half-maximal activation voltage than Cav1.2 and Cav3.1. However, a significantly higher fraction of the total charge had to move for activation of Cav1.3 half-maximal conductance (Cav1.3: 68%; Cav1.2: 52%; Cav3.1: 22%). This indicated a weaker coupling of Cav1.3 voltage-sensor charge movement to pore opening. However, the coupling efficiency was strengthened in the absence of the CTM in Cav1.342A, thereby shifting ICa-V by 7.2 mV to potentials that were more negative without changing QON-V. We independently show that the presence of intracellular organic cations (such as n-methyl-D-glucamine) induces a pronounced negative shift of QON-V and a more negative activation of ICa-V of all three channels. These findings illustrate that the voltage sensors of Cav1.3 channels respond more sensitively to depolarization than those of Cav1.2 or Cav3.1. Weak coupling of voltage sensing to pore opening is enhanced in the absence of the CTM, allowing short Cav1.342A splice variants to activate at lower voltages without affecting QON-V.  相似文献   

9.
Activity of voltage-gated Cav1.3 L-type Ca2+ channels is required for proper hearing as well as sinoatrial node and brain function. This critically depends on their negative activation voltage range, which is further fine-tuned by alternative splicing. Shorter variants miss a C-terminal regulatory domain (CTM), which allows them to activate at even more negative potentials than C-terminally long-splice variants. It is at present unclear whether this is due to an increased voltage sensitivity of the Cav1.3 voltage-sensing domain, or an enhanced coupling of voltage-sensor conformational changes to the subsequent opening of the activation gate. We studied the voltage-dependence of voltage-sensor charge movement (QON-V) and of current activation (ICa-V) of the long (Cav1.3L) and a short Cav1.3 splice variant (Cav1.342A) expressed in tsA-201 cells using whole cell patch-clamp. Charge movement (QON) of Cav1.3L displayed a much steeper voltage-dependence and a more negative half-maximal activation voltage than Cav1.2 and Cav3.1. However, a significantly higher fraction of the total charge had to move for activation of Cav1.3 half-maximal conductance (Cav1.3: 68%; Cav1.2: 52%; Cav3.1: 22%). This indicated a weaker coupling of Cav1.3 voltage-sensor charge movement to pore opening. However, the coupling efficiency was strengthened in the absence of the CTM in Cav1.342A, thereby shifting ICa-V by 7.2 mV to potentials that were more negative without changing QON-V. We independently show that the presence of intracellular organic cations (such as n-methyl-D-glucamine) induces a pronounced negative shift of QON-V and a more negative activation of ICa-V of all three channels. These findings illustrate that the voltage sensors of Cav1.3 channels respond more sensitively to depolarization than those of Cav1.2 or Cav3.1. Weak coupling of voltage sensing to pore opening is enhanced in the absence of the CTM, allowing short Cav1.342A splice variants to activate at lower voltages without affecting QON-V.  相似文献   

10.
Calcium entry through voltage-gated calcium channels (VGCC) initiates diverse cellular functions. VGCC pore-forming subunit (CaVα1) contains four homology repeats, each encompassing a voltage sensor and a pore domain. Three main classes of CaVα1 subunits have been described, CaV1, CaV2 and CaV3 that differ in their voltage-dependence of activation and in the extent in which this process is modulated by the auxiliary β-subunit (CaVβ). Association of CaVβ induces a coil-to-helix conformation of the I-II intracellular linker joining the first and second repeat of CaVα1 that is thought to be crucial for modulation of channel function. When expressed in Xenopus laevis oocytes in the absence of CaVβ the voltage to reach 50% activation (V0.5) for CaV1.2 and CaV2.3 differs by more than 60 mV and the channel current-carrying capacity by more than thirty-fold. Here we report that the difference in V0.5 is reduced to about 30 mV and the current-carrying capacity becomes virtually identical when the I-II linkers of CaV1.2 and CaV2.3 are swapped. Co-expression with CaVβ increases the current-carrying capacity of chimeric channels by the same extent, while the difference in V0.5 with respect to their corresponding parental channels vanishes. Our findings indicate that CaVβ modulatory potency is determined by both, the nature of the I-II linker and the pore-forming subunit background. Moreover, they demonstrate that the I-II linker encodes self-reliant molecular determinants for channel activation and suggest that besides to the secondary structure adopted by this segment upon CaVβ association, its chemical nature is as well relevant.  相似文献   

11.
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   

12.
There is growing evidence indicating that the pore structure of voltage-gated ion channels (VGICs) influences gating besides their conductance. Regarding low voltage-activated (LVA) Ca2+ channels, it has been demonstrated that substitutions of the pore aspartate (D) by a glutamate (D-to-E substitution) in domains III and IV alter channel gating properties such as a positive shift in the channel activation voltage dependence. In the present report, we evaluated the effects of E-to-D substitution in domains I and II on the CaV3.1 channel gating properties. Our results indicate that substitutions in these two domains differentially modify the gating properties of CaV3.1 channels. The channel with a single mutation in domain I (DEDD) presented slower activation and faster inactivation kinetics and a slower recovery from inactivation, as compared with the WT channel. In contrast, the single mutant in domain II (EDDD) presented a small but significant negative shift of activation voltage dependence with faster activation and slower inactivation kinetics. Finally, the double mutant channel (DDDD) presented somehow intermediate properties with respect to the two single mutants but with fastest deactivation kinetics. Overall, our results indicate that single amino acid modification of the selectivity filter of LVA Ca2+ channels in distinct domains differentially influence their gating properties, supporting a pore pseudo-symmetry.  相似文献   

13.
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   

14.
The delivery of Ca2+ into cells by CaV channels provides the trigger for many cellular actions, such as cardiac muscle contraction and neurotransmitter release. Thus, a full understanding of Ca2+ permeation through these channels is critical. Using whole-cell voltage-clamp recordings, we recently demonstrated that voltage modulates the apparent affinity of N-type (CaV2.2) channels for permeating Ca2+ and Ba2+ ions. While we took many steps to ensure the high fidelity of our recordings, problems can occur when CaV currents become large and fast, or when currents run down. Thus, we use here single channel recordings to further test the hypothesis that permeating ions interact with N-type channels in a voltage-dependent manner. We also examined L-type (CaV1.2) channels to determine if these channels also exhibit voltage-dependent permeation. Like our whole-cell data, we find that voltage modulates N-channel affinity for Ba2+ at voltages > 0 mV, but has little or no effect at voltages < 0 mV. Furthermore, we demonstrate that permeation through L-channel is also modulated by voltage. Thus, voltage-dependence may be a common feature of divalent cation permeation through CaV1 and CaV2 channels (i.e. high-voltage activated CaV channels). The voltage dependence of CaV1 channel permeation is likely a mechanism mediating sustained Ca2+ influx during the plateau phase of the cardiac action potential.  相似文献   

15.
Voltage dependence and kinetics of CaV1.2 activation are affected by structural changes in pore-lining S6 segments of the α1-subunit. Significant effects are induced by either proline or threonine substitutions in the lower third of segment IIS6 (“bundle crossing region”), where S6 segments are likely to seal the channel in the closed conformation (Hohaus, A., Beyl, S., Kudrnac, M., Berjukow, S., Timin, E. N., Marksteiner, R., Maw, M. A., and Hering, S. (2005) J. Biol. Chem. 280, 38471–38477). Here we report that S435P in IS6 results in a large shift of the activation curve (-25.9 ± 1.2 mV) and slower current kinetics. Threonine substitutions at positions Leu-429 and Leu-434 induced a similar kinetic phenotype with shifted activation curves (L429T by -6.6 ± 1.2 and L434T by -12.1 ± 1.7 mV). Inactivation curves of all mutants were shifted to comparable extents as the activation curves. Interdependence of IS6 and IIS6 mutations was analyzed by means of mutant cycle analysis. Double mutations in segments IS6 and IIS6 induce either additive (L429T/I781T, -34.1 ± 1.4 mV; L434T/I781T, -40.4 ± 1.3 mV; L429T/L779T, -12.6 ± 1.3 mV; and L434T/L779T, -22.4 ± 1.3 mV) or nonadditive shifts of the activation curves along the voltage axis (S435P/I781T, -33.8 ± 1.4 mV). Mutant cycle analysis revealed energetic coupling between residues Ser-435 and Ile-781, whereas other paired mutations in segments IS6 and IIS6 had independent effects on activation gating.Ca2+ current through CaV1.2 channels initiates muscle contraction, release of hormones and neurotransmitters, and affects physiological processes such as vision, hearing, and gene expression (1). Their pore-forming α1-subunit is composed of four homologous domains formed by six transmembrane segments (S1–S6) (2). The signal of the voltage-sensing machinery, consisting of multiple charged amino acids (located in segments S4 and adjacent structures of each domain), is transmitted to the pore region (3). Conformational changes in pore lining S6 and adjacent segments finally lead to pore openings (activation) and closures (inactivation).Our understanding of how CaV1.2 channels open and close is largely based on extrapolations of structural information from potassium channels. The crystal structures of the closed conformation of two bacterial potassium channels (KcsA and MlotiK) (4, 5) show a gate located at the intracellular channel mouth formed by tightly packed S6 helices. The crystal structure of the open conformation of Kv1.2 (6, 7) revealed a bent S6 with the highly conserved PXP motif apparently acting as a hinge (see 8). The activation mechanism proposed for MthK channels involves helix bending at a highly conserved glycine at position 83 (see Ref. 9, “glycine gating hinge” hypothesis).Compared with potassium channels, the pore of CaV is asymmetric, and none of the four S6 segments has a putative helix-bending PXP motif. Furthermore, the conserved glycine (corresponding to position 83 in MthK, see Ref. 10) is only present in segments IS6 and IIS6 (for review see Ref. 11). We have shown that substituting proline for this glycine in IIS6 of CaV1.2 does not significantly affect gating (12).Zhen et al. (13) investigated the pore lining S6 segments of CaV2.1 using the substituted cysteine accessibility method. The accessibility of cysteines was changed by opening and closing the channel, consistent with the gate being on the intracellular side. The general picture of a channel gate close to the inner channel mouth of CaV1.2 was recently supported by pharmacological studies (14).Substitution of hydrophilic residues in the lower third of segment IIS6 of CaV1.2 (LAIA motif, 779–784, see Ref. 12) induces pronounced changes in channel gating as follows: a shift in the voltage dependence of activation accompanied by a slowing of the activation kinetics near the footstep of the m(V) curve and a slowing of deactivation at all potentials. Interestingly, these changes in channel gating resemble the effects of proline substitution of Gly-219 in the bacterial sodium channel from Bacillus halodurans (“Gly-219 gating hinge,” see Ref. 15).The strongest shifts of the activation curve reported so far were observed for proline substitutions (12). As prolines in an α-helix cause a rigid kink with an angle of about 26° (16), we hypothesized that these mutants were causing a kink in helix IIS6 similar to a bend that would normally occur flexibly during the activation process (12).Here we extend our previous study by systematically substituting residues in segment IS6 of CaV1.2 by proline or the small and polar threonine. Several functional IS6 mutants with shifted activation and inactivation characteristics were identified (S435P, L429T, and L434T), and the interdependence of IS6 and IIS6 mutations was analyzed. Mutant cycle analysis revealed both mutually independent and energetically coupled contributions of IS6 and IIS6 residues on activation gating.  相似文献   

16.
Voltage-dependent calcium (CaV) 1.3 channels are involved in the control of cellular excitability and pacemaking in neuronal, cardiac, and sensory cells. Various proteins interact with the alternatively spliced channel C-terminus regulating gating of CaV1.3 channels. Binding of a regulatory calcium-binding protein calmodulin (CaM) to the proximal C-terminus leads to the boosting of channel activity and promotes calcium-dependent inactivation (CDI). The C-terminal modulator domain (CTM) of CaV1.3 channels can interfere with the CaM binding, thereby inhibiting channel activity and CDI. Here, we compared single-channel gating behavior of two natural CaV1.3 splice isoforms: the long CaV1.342 with the full-length CTM and the short CaV1.342A with the C-terminus truncated before the CTM. We found that CaM regulation of CaV1.3 channels is dynamic on a minute timescale. We observed that at equilibrium, single CaV1.342 channels occasionally switched from low to high open probability, which perhaps reflects occasional binding of CaM despite the presence of CTM. Similarly, when the amount of the available CaM in the cell was reduced, the short CaV1.342A isoform showed patterns of the low channel activity. CDI also underwent periodic changes with corresponding kinetics in both isoforms. Our results suggest that the competition between CTM and CaM is influenced by calcium, allowing further fine-tuning of CaV1.3 channel activity for particular cellular needs.  相似文献   

17.
The delivery of Ca2+ into cells by CaV channels provides the trigger for many cellular actions, such as cardiac muscle contraction and neurotransmitter release. Thus, a full understanding of Ca2+ permeation through these channels is critical. Using whole-cell voltage-clamp recordings, we recently demonstrated that voltage modulates the apparent affinity of N-type (CaV2.2) channels for permeating Ca2+ and Ba2+ ions. While we took many steps to ensure the high fidelity of our recordings, problems can occur when CaV currents become large and fast, or when currents run down. Thus, we use here single channel recordings to further test the hypothesis that permeating ions interact with N-type channels in a voltage-dependent manner. We also examined L-type (CaV1.2) channels to determine if these channels also exhibit voltage-dependent permeation. Like our whole-cell data, we find that voltage modulates N-channel affinity for Ba2+ at voltages > 0 mV, but has little or no effect at voltages < 0 mV. Furthermore, we demonstrate that permeation through L-channel is also modulated by voltage. Thus, voltage-dependence may be a common feature of divalent cation permeation through CaV1 and CaV2 channels (i.e. high-voltage activated CaV channels). The voltage dependence of CaV1 channel permeation is likely a mechanism mediating sustained Ca2+ influx during the plateau phase of the cardiac action potential.  相似文献   

18.
Point mutations in pore-lining S6 segments of CaV1.2 shift the voltage dependence of activation into the hyperpolarizing direction and significantly decelerate current activation and deactivation. Here, we analyze theses changes in channel gating in terms of a circular four-state model accounting for an activation R–A–O and a deactivation O–D–R pathway. Transitions between resting-closed (R) and activated-closed (A) states (rate constants x(V) and y(V)) and open (O) and deactivated-open (D) states (u(V) and w(V)) describe voltage-dependent sensor movements. Voltage-independent pore openings and closures during activation (A–O) and deactivation (D–R) are described by rate constants α and β, and γ and δ, respectively. Rate constants were determined for 16-channel constructs assuming that pore mutations in IIS6 do not affect the activating transition of the voltage-sensing machinery (x(V) and y(V)). Estimated model parameters of 15 CaV1.2 constructs well describe the activation and deactivation processes. Voltage dependence of the “pore-releasing” sensor movement ((x(V)) was much weaker than the voltage dependence of “pore-locking” sensor movement (y(V)). Our data suggest that changes in membrane voltage are more efficient in closing than in opening CaV1.2. The model failed to reproduce current kinetics of mutation A780P that was, however, accurately fitted with individually adjusted x(V) and y(V). We speculate that structural changes induced by a proline substitution in this position may disturb the voltage-sensing domain.  相似文献   

19.
Recently, we reported that homozygous deletion of alternative exon 33 of CaV1.2 calcium channel in the mouse resulted in ventricular arrhythmias arising from increased CaV1.2Δ33 ICaL current density in the cardiomyocytes. We wondered whether heterozygous deletion of exon 33 might produce cardiac phenotype in a dose-dependent manner, and whether the expression levels of RNA splicing factors known to regulate alternative splicing of exon 33 might change in human heart failure. Unexpectedly, we found that exon 33+/? cardiomyocytes showed similar CaV1.2 channel properties as wild-type cardiomyocyte, even though CaV1.2Δ33 channels exhibit a gain-in-function. In human hearts, we found that the mRNA level of splicing factor Rbfox1, but not Rbfox2, was downregulated in dilated cardiomyopathy, and CACNA1C mRNA level was dramatically decreased in the both of dilated and ischemic cardiomyopathy. These data imply Rbfox1 may be involved in the development of cardiomyopathies via regulating the alternative splicing of CaV1.2 exon 33. (149 words)  相似文献   

20.
The CaV1.2 L-type calcium channel is a key conduit for Ca2+ influx to initiate excitation-contraction coupling for contraction of the heart and vasoconstriction of the arteries and for altering membrane excitability in neurons. Its α1C pore-forming subunit is known to undergo extensive alternative splicing to produce many CaV1.2 isoforms that differ in their electrophysiological and pharmacological properties. Here, we examined the structure-function relationship of human CaV1.2 with respect to the inclusion or exclusion of mutually exclusive exons of the N-terminus exons 1/1a and IS6 segment exons 8/8a. These exons showed tissue selectivity in their expression patterns: heart variant 1a/8a, one smooth-muscle variant 1/8, and a brain isoform 1/8a. Overall, the 1/8a, when coexpressed with CaVβ2a, displayed a significant and distinct shift in voltage-dependent activation and inactivation and inactivation kinetics as compared to the other three splice variants. Further analysis showed a clear additive effect of the hyperpolarization shift in V1/2inact of CaV1.2 channels containing exon 1 in combination with 8a. However, this additive effect was less distinct for V1/2act. However, the measured effects were β-subunit-dependent when comparing CaVβ2a with CaVβ3 coexpression. Notably, calcium-dependent inactivation mediated by local Ca2+-sensing via the N-lobe of calmodulin was significantly enhanced in exon-1-containing CaV1.2 as compared to exon-1a-containing CaV1.2 channels. At the cellular level, the current densities of the 1/8a or 1/8 variants were significantly larger than the 1a/8a and 1a/8 variants when coexpressed either with CaVβ2a or CaVβ3 subunit. This finding correlated well with a higher channel surface expression for the exon 1-CaV1.2 isoform that we quantified by protein surface-expression levels or by gating currents. Our data also provided a deeper molecular understanding of the altered biophysical properties of alternatively spliced human CaV1.2 channels by directly comparing unitary single-channel events with macroscopic whole-cell currents.  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