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.
Calcium-dependent inactivation has been described as a negative feedback mechanism for regulating voltage-dependent calcium influx in cardiac cells. Most recent evidence points to the C-terminus of the alpha1C subunit, with its EF-hand binding motif, as being critical in this process. The EF-hand binding motif is mostly conserved between the C-termini of six of the seven alpha1 subunit Ca2+ channel genes. The role of E1537 in the C-terminus of the alpha1C calcium channel inactivation was investigated here after expression in Xenopus laevis oocytes. Whole-cell currents were measured in the presence of 10 mM Ba2+ or 10 mM Ca2+ after intracellular injection of 1,2-bis(2-aminophenoxy)ethane-N,N,N',N'-tetraacetic acid. Against all expectations, our results showed a significant reduction in the rate of voltage-dependent inactivation as measured in Ba2+ solutions for all E1537 mutants, whereas calcium-dependent inactivation appeared unscathed. Replacing the negatively charged glutamate residue by neutral glutamine, glycine, serine, or alanine significantly reduced the rate of Ba2+-dependent inactivation by 1.5-fold (glutamine) to 3.5-fold (alanine). The overall rate of macroscopic inactivation measured in Ca2+ solutions was also reduced, although a careful examination of the distribution of the fast and slow time constants suggests that only the slow time constant was significantly reduced in the mutant channels. The fast time constant, the hallmark of Ca2+-dependent inactivation, remained remarkably constant among wild-type and mutant channels. Moreover, inactivation of E1537A channels, in both Ca2+ and Ba2+ solutions, appeared to decrease with membrane depolarization, whereas inactivation of wild-type channels became faster with positive voltages. All together, our results showed that E1537 mutations impaired voltage-dependent inactivation and suggest that the proximal part of the C-terminus may play a role in voltage-dependent inactivation in L-type alpha1C channels.  相似文献   

2.
We have examined the kinetics of whole-cell T-current in HEK 293 cells stably expressing the alpha1G channel, with symmetrical Na(+)(i) and Na(+)(o) and 2 mM Ca(2+)(o). After brief strong depolarization to activate the channels (2 ms at +60 mV; holding potential -100 mV), currents relaxed exponentially at all voltages. The time constant of the relaxation was exponentially voltage dependent from -120 to -70 mV (e-fold for 31 mV; tau = 2.5 ms at -100 mV), but tau = 12-17 ms from-40 to +60 mV. This suggests a mixture of voltage-dependent deactivation (dominating at very negative voltages) and nearly voltage-independent inactivation. Inactivation measured by test pulses following that protocol was consistent with open-state inactivation. During depolarizations lasting 100-300 ms, inactivation was strong but incomplete (approximately 98%). Inactivation was also produced by long, weak depolarizations (tau = 220 ms at -80 mV; V(1/2) = -82 mV), which could not be explained by voltage-independent inactivation exclusively from the open state. Recovery from inactivation was exponential and fast (tau = 85 ms at -100 mV), but weakly voltage dependent. Recovery was similar after 60-ms steps to -20 mV or 600-ms steps to -70 mV, suggesting rapid equilibration of open- and closed-state inactivation. There was little current at -100 mV during recovery from inactivation, consistent with 相似文献   

3.
The possibility that guinea pig pancreatic alpha 2 cells are equipped with more than one type of Ca2+ channel was explored using the patch-electrode voltage-clamp technique. At a holding potential of -100 mV, a slowly developing (tau m approximately 5 ms at -40 mV assuming m2 kinetics) Ca2+ current appeared. This conductance first became detectable at potentials of about -60 mV and reached a maximum amplitude of 50-100 pA between -30 and -20 mV. During long depolarizations, it inactivated completely (tau h approximately 100 ms at -40 mV). Half-maximal steady state inactivation was observed at about -60 mV. A second, more rapidly developing (tau m approximately 2 ms at 0 mV) Ca2+ current was observed during pulses to -40 mV and above. It had a peak amplitude of 150-200 pA between 0 and 10 mV, was less dependent on the holding potential, and inactivated very little, even during long pulses. Both conductances were blocked by Co2+ but were unaffected by tetrodotoxin. The rapidly developing current differed from the slowly developing one in being sensitive to the antagonists D-600 and nifedipine, conducting Ba2+ better than Ca2+, increasing upon exposure to forskolin, and showing time-dependent decay (rundown). These findings indicate that the alpha 2 cells are equipped with two kinds of Ca2+ channels.  相似文献   

4.
In order to investigate the currently unknown cellular signaling pathways of T-type Ca(2+) channels, we decided to construct a new cell line which would stably express alpha(1G) and Kir2.1 subunits in HEK293 cells (HEK293/alpha(1G)/Kir2.1). Compared to cells which only expressed alpha(1G) (HEK293/alpha(1G)), HEK293/alpha(1G)/Kir2.1 cells produced an enormous inward rectifying current which was blocked by external Ba(2+) and Cs(+) in a concentration-dependent manner. The expression of Kir2.1 channels contributed significantly to the shift of membrane potential from -12.2+/-2.8 to -57.3+/-3.7mV. However, biophysical and pharmacological properties of alpha(1G)-mediated Ca(2+) channels remained unaffected by the expression of Kir2.1 subunits, except for the enlarging of the window current region. Biochemical activation of alpha(1G) channels using 150mM KCl brought about an increase in [Ca(2+)](i), which was blocked by mibefradil, the T-type Ca(2+) channel blocker. These data suggest that the HEK293/alpha(1G)/Kir2.1 cell line would have potential uses in the study of T-type Ca(2)(+) channel-mediated signaling pathways and possibly useful in the development of new therapeutic drugs associated with T-type Ca(2)(+) channels.  相似文献   

5.
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   

6.
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   

7.
Inactivation of currents carried by Ba2+ and Ca2+, as well as intramembrane charge movement from L-type Ca2+ channels were studied in guinea pig ventricular myocytes using the whole-cell patch clamp technique. Prolonged (2 s) conditioning depolarization caused substantial reduction of charge movement between -70 and 10 mV (charge 1, or charge from noninactivated channels). In parallel, the charge mobile between -70 and -150 mV (charge 2, or charge from inactivated channels) was increased. The availability of charge 2 depended on the conditioning pulse voltage as the sum of two Boltzmann components. One component had a central voltage of -75 mV and a magnitude of 1.7 nC/microF. It presumably is the charge movement (charge 2) from Na+ channels. The other component, with a central voltage of approximately - 30 mV and a magnitude of 3.5 nC/microF, is the charge 2 of L-type Ca2+ channels. The sum of charge 1 and charge 2 was conserved after different conditioning pulses. The difference between the voltage dependence of the activation of L-type Ca2+ channels (half-activation voltage, V, of approximately -20 mV) and that of charge 2 (V of -100 mV) made it possible to record the ionic currents through Ca2+ channels and charge 2 in the same solution. In an external solution with Ba2+ as sole metal the maximum available charge 2 of L-type Ca2+ channels was 10-15% greater than that in a Ca(2+)-containing solution. External Cd2+ caused 20-30% reduction of charge 2 both from Na+ and L-type Ca2+ channels. Voltage- and Ca(2+)-dependent inactivation phenomena were compared with a double pulse protocol in cells perfused with an internal solution of low calcium buffering capacity. As the conditioning pulse voltage increased, inactivation monitored with the second pulse went through a minimum at about 0 mV, the voltage at which conditioning current had its maximum. Charge 2, recorded in parallel, did not show any increase associated with calcium entry. Two alternative interpretations of these observations are: (a) that Ca(2+)- dependent inactivation does not alter the voltage sensor, and (b) that inactivation affects the voltage sensor, but only in the small fraction of channels that open, and the effect goes undetected. A model of channel gating that assumes the first possibility is shown to account fully for the experimental results. Thus, extracellular divalent cations modulate voltage-dependent inactivation of the Ca2+ channel. Intracellular Ca2+ instead, appears to cause inactivation of the channel without affecting its voltage sensor.  相似文献   

8.
Ca(v)2.1 mediates voltage-gated Ca2+ entry into neurons and the release of neurotransmitters at synapses of the central nervous system. An inactivation process that is modulated by the auxiliary beta-subunits regulates Ca2+ entry through Ca(v)2.1. However, the molecular mechanism of this alpha1-beta-subunit interaction remains unknown. Herein we report the identification of new determinants within segment IVS6 of the alpha(1)2.1-subunit that markedly influence channel inactivation. Systematic substitution of residues within IVS6 with amino acids of different size, charge, and polarity resulted in mutant channels with rates of fast inactivation (k(inact)) ranging from a 1.5-fold slowing in V1818I (k(inact) = 0.98 +/- 0.09 s(-1) compared with wild type alpha(1)2.1/alpha2-delta/beta1a k(inact) = 1.35 +/- 0.25 s(-1) to a 75-fold acceleration in mutant M1811Q (k(inact) = 102 +/- 3 s(-1). Coexpression of mutant alpha(1)2.1-subunits with beta(2a) resulted in two different phenotypes of current inactivation: 1) a pronounced reduction in the rate of channel inactivation or 2) an attenuation of a slow component in I(Ba) inactivation. Simulations revealed that these two distinct inactivation phenotypes arise from a beta2a-subunit-induced destabilization of the fast-inactivated state. The IVS6- and beta2a-subunit-mediated effects on Ca(v)2.1 inactivation are likely to occur via independent mechanisms.  相似文献   

9.
Voltage-dependent inactivation of CaV2.3 channels was investigated using point mutations in the beta-subunit-binding site (AID) of the I-II linker. The quintuple mutant alpha1E N381K + R384L + A385D + D388T + K389Q (NRADK-KLDTQ) inactivated like the wild-type alpha1E. In contrast, mutations of alpha1E at position R378 (position 5 of AID) into negatively charged residues Glu (E) or Asp (D) significantly slowed inactivation kinetics and shifted the voltage dependence of inactivation to more positive voltages. When co-injected with beta3, R378E inactivated with tau(inact) = 538 +/- 54 ms (n = 14) as compared with 74 +/- 4 ms (n = 21) for alpha1E (p < 0.001) with a mid-potential of inactivation E(0.5) = -44 +/- 2 mV (n = 10) for R378E as compared with E(0.5) = -64 +/- 3 mV (n = 9) for alpha1E. A series of mutations at position R378 suggest that positively charged residues could promote voltage-dependent inactivation. R378K behaved like the wild-type alpha1E whereas R378Q displayed intermediate inactivation kinetics. The reverse mutation E462R in the L-type alpha1C (CaV1.2) produced channels with inactivation properties comparable to alpha1E R378E. Hence, position 5 of the AID motif in the I-II linker could play a significant role in the inactivation of Ca(V)1.2 and CaV2.3 channels.  相似文献   

10.
Voltage-activated calcium channels can be divided into two subgroups based on their activation threshold, low-voltage-activated (LVA) and high-voltage-activated (HVA). Auxiliary subunits of the HVA calcium channels contribute significantly to biophysical properties of the channels. We have cloned and characterized members of two families of auxiliary subunits: alpha2delta and gamma. Two new alpha2delta subunits, alpha2delta-2 and alpha2delta-3, regulate all classes of HVA calcium channels. While the ubiquitous alpha2delta-2 modulates both neuronal and non-neuronal channels with similar efficiency, the alpha2delta-3 subunit regulates Ca(v)2.3 channels more effectively. Furthermore, alpha2delta-2 may modulate the LVA Ca(v)3.1 channel. Four new gamma subunits, gamma-2, gamma-3, gamma-4 and gamma-5, were characterized. The gamma-2 subunit modulated both the non-neuronal Ca(v)1.2 channel and the neuronal Ca(v)2.1 channel. The gamma-4 subunit affected only the Ca(v)2.1 channel. The gamma-5 subunit may be a regulatory subunit of the LVA Ca(v)3.1 channel. The Ca(v)1.2 channel is a major target for treatment of cardiovascular diseases. We have mapped the interaction site for clinically important channel blockers - dihydropyridines (DHPs) - and analysed the underlying inhibition mechanism. High-affinity inhibition is characterized by interaction with inactivated state of the channel. Its structural determinants are amino acids of the IVS6 segment, with smaller contribution of the IS6 segment, which contributes to voltage-dependence of DHP inhibition. Removal of amino acids responsible for the high-affinity inhibition revealed a low-affinity open channel block, in which amino acids of the IIIS5 and IIIS6 segments take part. Experiments with a permanently charged DHP suggested that there is another low-affinity interaction site on the alpha(1) subunit. We have cloned and characterized murine neuronal LVA Ca(v)3.1 channel. The channel has high sensitivity to the organic blocker mibefradil, moderate sensitivity to phenytoin, and low sensitivity to ethosuximide, amiloride and valproat. The channel is insensitive to tetrodotoxin and DHPs. The inorganic blockers Ni2+ and Cd2+ are moderately effective compared to La3+. The current through the Ca(v)3.1 channel inactivates faster with Ba2+ compared to Ca2+. Molecular determinants of fast inactivation are located in amino side of the intracellular carboxy terminus. The voltage dependence of charge movement is very shallow compared to the voltage dependence of current activation. Transfer of 30 % of charge correlates with activation of 70 % of measurable macroscopic current. Prolonged depolarization does not immobilize charge movement of the Ca(v)3.1 channel.  相似文献   

11.
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   

12.
Complimentary DNA clones encoding the alpha1C and beta2a subunits of guinea-pig cardiac L-type Ca2+ channels were isolated using the PCR method. The open reading frame encoded 2,169 amino acids for the alpha1C and 597 amino acids for the beta2a subunit. The proteins showed 94.2 and 94.8%, respectively, identity to the respective subunit of the rabbit protein. The message size of the guinea pig alpha1C and beta2a subunits was 8.0 and 3.5/4.0 kb, respectively. RT-PCR analysis revealed that the alpha1C subunit is expressed exclusively in the heart, while the beta2a subunit is expressed in the heart, cerebellum, whole brain, and stomach. The alpha1C and beta2a subunits are transiently expressed in BHK (baby hamster kidney) cells, and the channel currents were studied using the whole-cell patch clamp technique in medium containing 30 mM Ba2+. In cells expressing alpha1C alone, the Ba2+ current was activated at -30 mV and more positive potentials and peaked at about 10 mV. The co-expression of beta2a with alpha1C did not affect the voltage-dependence of the current, but increased the peak current and accelerated current decay. In cells transfected with guinea pig alpha1C and rabbit beta1+alpha2/delta, a Ba2+ current comparable to those in native myocytes was observed. The Ba2+ current can be blocked completely by nifedipine and is enhanced 3-fold by Bay K 8644. On the other hand, neither forskolin nor okadaic acid affects the Ba2+ current, suggesting that cAMP-mediated modulation is not easily reproduced in transfected cells, unlike that seen in native cardiac myocytes.  相似文献   

13.
It has been shown that beta auxiliary subunits increase current amplitude in voltage-dependent calcium channels. In this study, however, we found a novel inhibitory effect of beta3 subunit on macroscopic Ba(2+) currents through recombinant N- and R-type calcium channels expressed in Xenopus oocytes. Overexpressed beta3 (12.5 ng/cell cRNA) significantly suppressed N- and R-type, but not L-type, calcium channel currents at "physiological" holding potentials (HPs) of -60 and -80 mV. At a HP of -80 mV, coinjection of various concentrations (0-12.5 ng) of the beta3 with Ca(v)2.2alpha(1) and alpha(2)delta enhanced the maximum conductance of expressed channels at lower beta3 concentrations but at higher concentrations (>2.5 ng/cell) caused a marked inhibition. The beta3-induced current suppression was reversed at a HP of -120 mV, suggesting that the inhibition was voltage dependent. A high concentration of Ba(2+) (40 mM) as a charge carrier also largely diminished the effect of beta3 at -80 mV. Therefore, experimental conditions (HP, divalent cation concentration, and beta3 subunit concentration) approaching normal physiological conditions were critical to elucidate the full extent of this novel beta3 effect. Steady-state inactivation curves revealed that N-type channels exhibited "closed-state" inactivation without beta3, and that beta3 caused an approximately 40-mV negative shift of the inactivation, producing a second component with an inactivation midpoint of approximately -85 mV. The inactivation of N-type channels in the presence of a high concentration (12.5 ng/cell) of beta3 developed slowly and the time-dependent inactivation curve was best fit by the sum of two exponential functions with time constants of 14 s and 8.8 min at -80 mV. Similar "ultra-slow" inactivation was observed for N-type channels without beta3. Thus, beta3 can have a profound negative regulatory effect on N-type (and also R-type) calcium channels by causing a hyperpolarizing shift of the inactivation without affecting "ultra-slow" and "closed-state" inactivation properties.  相似文献   

14.
During sustained depolarization, voltage-gated Ca2+ channels progressively undergo a transition to a nonconducting, inactivated state, preventing Ca2+ overload of the cell. This transition can be triggered either by the membrane potential (voltage-dependent inactivation) or by the consecutive entry of Ca2+ (Ca2+-dependent inactivation), depending on the type of Ca2+ channel. These two types of inactivation are suspected to arise from distinct underlying mechanisms, relying on specific molecular sequences of the different pore-forming Ca2+ channel subunits. Here we report that the voltage-dependent inactivation (of the alpha1A Ca2+ channel) and the Ca2+-dependent inactivation (of the alpha1C Ca2+ channel) are similarly influenced by Ca2+ channel beta subunits. The same molecular determinants of the beta subunit, and therefore the same subunit interactions, influence both types of inactivation. These results strongly suggest that the voltage and the Ca2+-dependent transitions leading to channel inactivation use homologous structures of the different alpha1 subunits and occur through the same molecular process. A model of inactivation taking into account these new data is presented.  相似文献   

15.
The molecular basis for inactivation in Ca(V)2.3 (alpha 1E) channels was studied after expression of alpha 1E/alpha 1C (Ca(V)2.3/Ca(V)1.2) chimeras in Xenopus oocytes. In the presence of 10 mM Ba(2+), the CEEE chimera (Repeat I+part of the I-II linker from Ca(V)1.2) displayed inactivation properties similar to Ca(V)1.2 despite being more than 90% homologous to Ca(V)2.3. The transmembrane segments of Repeat I did not appear to be crucial as inactivation of EC(IS1-6)EEE was not significantly different than Ca(V)2.3. In contrast, EC(AID)EEE, with the beta-subunit binding domain from Ca(V)1.2, tended to behave like Ca(V)1.2 in terms of inactivation kinetics and voltage dependence. A detailed kinetic analysis revealed nonetheless that CEEE and EC(AID)EEE retained the fast inactivation time constant (tau(fast) approximately equal to 20-30 ms) that is a distinctive feature of Ca(V)2.3. Altogether, these data suggest that the region surrounding the AID binding site plays a pivotal albeit not exclusive role in determining the inactivation properties of Ca(V)2.3.  相似文献   

16.
In cochlea inner hair cells (IHCs), L-type Ca(2+) channels (LTCCs) formed by alpha1D subunits (D-LTCCs) possess biophysical and pharmacological properties distinct from those of alpha1C containing C-LTCCs. We investigated to which extent these differences are determined by alpha1D itself by analyzing the biophysical and pharmacological properties of cloned human alpha1D splice variants in tsA-201 cells. Variant alpha1D(8A,) containing exon 8A sequence in repeat I, yielded alpha1D protein and L-type currents, whereas no intact protein and currents were observed after expression with exon 8B. In whole cell patch-clamp recordings (charge carrier 15-20 mm Ba(2+)), alpha1D(8A) - mediated currents activated at more negative voltages (activation threshold, -45.7 versus -31.5 mV, p < 0.05) and more rapidly (tau(act) for maximal inward currents 0.8 versus 2.3 ms; p < 0.05) than currents mediated by rabbit alpha1C. Inactivation during depolarizing pulses was slower than for alpha1C (current inactivation after 5-s depolarizations by 90 versus 99%, p < 0.05) but faster than for LTCCs in IHCs. The sensitivity for the dihydropyridine (DHP) L-type channel blocker isradipine was 8.5-fold lower than for alpha1C. Radioligand binding experiments revealed that this was not due to a lower affinity for the DHP binding pocket, suggesting that differences in the voltage-dependence of DHP block account for decreased sensitivity of D-LTCCs. Our experiments show that alpha1D(8A) subunits can form slowly inactivating LTCCs activating at more negative voltages than alpha1C. These properties should allow D-LTCCs to control physiological processes, such as diastolic depolarization in sinoatrial node cells, neurotransmitter release in IHCs and neuronal excitability.  相似文献   

17.
Ba(2+) current through the L-type Ca(2+) channel inactivates essentially by voltage-dependent mechanisms with fast and slow kinetics. Here we found that slow inactivation is mediated by an annular determinant composed of hydrophobic amino acids located near the cytoplasmic ends of transmembrane segments S6 of each repeat of the alpha(1C) subunit. We have determined the molecular requirements that completely obstruct slow inactivation. Critical interventions include simultaneous substitution of A752T in IIS6, V1165T in IIIS6, and I1475T in IVS6, each preventing in additive manner a considerable fraction of Ba(2+) current from inactivation. In addition, it requires the S405I mutation in segment IS6. The fractional inhibition of slow inactivation in tested mutants caused an acceleration of fast inactivation, suggesting that fast and slow inactivation mechanisms are linked. The channel lacking slow inactivation showed approximately 45% of the sustained Ba(2+) or Ca(2+) current with no indication of decay. The remaining fraction of the current was inactivated with a single-exponential decay (pi(f) approximately 10 ms), completely recovered from inactivation within 100 ms and did not exhibit Ca(2+)-dependent inactivation properties. No voltage-dependent characteristics were significantly changed, consistent with the C-type inactivation model suggesting constriction of the pore as the main mechanism possibly targeted by Ca(2+) sensors of inactivation.  相似文献   

18.
We describe here several novel properties of the human alpha(1G) subunit that forms T-type calcium channels. The partial intron/exon structure of the corresponding gene CACNA1G was defined and several alpha(1G) isoforms were identified, especially two isoforms that exhibit a distinct III-IV loop: alpha(1G-a) and alpha(1G-b). Northern blot and dot blot analyses indicated that alpha(1G) mRNA is predominantly expressed in the brain, especially in thalamus, cerebellum, and substantia nigra. Additional experiments have also provided evidence that alpha(1G) mRNA is expressed at a higher level during fetal life in nonneuronal tissues (i.e. kidney, heart, and lung). Functional expression in HEK 293 cells of a full-length cDNA encoding the shortest alpha(1G) isoform identified to date, alpha(1G-b), resulted in transient, low threshold activated Ca(2+) currents with the expected permeability ratio (I(Sr) > I(Ca) >/= I(Ba)) and channel conductance ( approximately 7 pS). These properties, together with slowly deactivating tail currents, are typical of those of native T-type Ca(2+) channels. This alpha(1G)-related current was inhibited by mibefradil (IC(50) = 2 microM) and weakly blocked by Ni(2+) ions (IC(50) = 148 microM) and amiloride (IC(50) > 1 mM). We showed that steady state activation and inactivation properties of this current can generate a "window current" in the range of -65 to -55 mV. Using neuronal action potential waveforms, we show that alpha(1G) channels produce a massive and sustained Ca(2+) influx due to their slow deactivation properties. These latter properties would account for the specificity of Ca(2+) influx via T-type channels that occurs in the range of physiological resting membrane potentials, differing considerably from the behavior of other Ca(2+) channels.  相似文献   

19.
The origin of Ibetanull, the Ca2+ current of myotubes from mice lacking the skeletal dihydropyridine receptor (DHPR) beta1a subunit, was investigated. The density of Ibetanull was similar to that of Idys, the Ca2+ current of myotubes from dysgenic mice lacking the skeletal DHPR alpha1S subunit (-0.6 +/- 0.1 and -0.7 +/- 0.1 pA/pF, respectively). However, Ibetanull activated at significantly more positive potentials. The midpoints of the GCa-V curves were 16.3 +/- 1.1 mV and 11.7 +/- 1.0 mV for Ibetanull and Idys, respectively. Ibetanull activated significantly more slowly than Idys. At +30 mV, the activation time constant for Ibetanull was 26 +/- 3 ms, and that for Idys was 7 +/- 1 ms. The unitary current of normal L-type and beta1-null Ca2+ channels estimated from the mean variance relationship at +20 mV in 10 mM external Ca2+ was 22 +/- 4 fA and 43 +/- 7 fA, respectively. Both values were significantly smaller than the single-channel current estimated for dysgenic Ca2+ channels, which was 84 +/- 9 fA under the same conditions. Ibetanull and Idys have different gating and permeation characteristics, suggesting that the bulk of the DHPR alpha1 subunits underlying these currents are different. Ibetanull is suggested to originate primarily from Ca2+ channels with a DHPR alpha1S subunit. Dysgenic Ca2+ channels may be a minor component of this current. The expression of DHPR alpha1S in beta1-null myotubes and its absence in dysgenic myotubes was confirmed by immunofluorescence labeling of cells.  相似文献   

20.
Rapid inactivation of Ca2+ release-activated Ca2+ (CRAC) channels was studied in Jurkat leukemic T lymphocytes using whole-cell patch clamp recording and [Ca2+]i measurement techniques. In the presence of 22 mM extracellular Ca2+, the Ca2+ current declined with a biexponential time course (time constants of 8-30 ms and 50-150 ms) during hyperpolarizing pulses to potentials more negative than -40 mV. Several lines of evidence suggest that the fast inactivation process is Ca2+ but not voltage dependent. First, the speed and extent of inactivation are enhanced by conditions that increase the rate of Ca2+ entry through open channels. Second, inactivation is substantially reduced when Ba2+ is present as the charge carrier. Third, inactivation is slowed by intracellular dialysis with BAPTA (12 mM), a rapid Ca2+ buffer, but not by raising the cytoplasmic concentration of EGTA, a slower chelator, from 1.2 to 12 mM. Recovery from fast inactivation is complete within 200 ms after repolarization to -12 mV. Rapid inactivation is unaffected by changes in the number of open CRAC channels or global [Ca2+]i. These results demonstrate that rapid inactivation of ICRAC results from the action of Ca2+ in close proximity to the intracellular mouths of individual channels, and that Ca2+ entry through one CRAC channel does not affect neighboring channels. A simple model for Ca2+ diffusion in the presence of a mobile buffer predicts multiple Ca2+ inactivation sites situated 3-4 nm from the intracellular mouth of the pore, consistent with a location on the CRAC channel itself.  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