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.
There is controversy over whether Ca(2+) binds to the BK(Ca) channel's intracellular domain or its integral-membrane domain and over whether or not mutations that reduce the channel's Ca(2+) sensitivity act at the point of Ca(2+) coordination. One region in the intracellular domain that has been implicated in Ca(2+) sensing is the "Ca(2+) bowl". This region contains many acidic residues, and large Ca(2+)-bowl mutations eliminate Ca(2+) sensing through what appears to be one type of high-affinity Ca(2+)-binding site. Here, through site-directed mutagenesis we have mapped the residues in the Ca(2+) bowl that are most important for Ca(2+) sensing. We find acidic residues, D898 and D900, to be essential, and we find them essential as well for Ca(2+) binding to a fusion protein that contains a portion of the BK(Ca) channel's intracellular domain. Thus, much of our data supports the conclusion that Ca(2+) binds to the BK(Ca) channel's intracellular domain, and they define the Ca(2+) bowl's essential Ca(2+)-sensing motif. Overall, however, we have found that the relationship between mutations that disrupt Ca(2+) sensing and those that disrupt Ca(2+) binding is not as strong as we had expected, a result that raises the possibility that, when examined by gel-overlay, the Ca(2+) bowl may be in a nonnative conformation.  相似文献   

2.
Selective permeability in voltage-gated Ca(2+) channels is dependent upon a quartet of pore-localized glutamate residues (EEEE locus). The EEEE locus is widely believed to comprise the sole high-affinity Ca(2+) binding site in the pore, which represents an overturning of earlier models that had postulated two high-affinity Ca(2+) binding sites. The current view is based on site-directed mutagenesis work in which Ca(2+) binding affinity was attenuated by single and double substitutions in the EEEE locus, and eliminated by quadruple alanine (AAAA), glutamine (QQQQ), or aspartate (DDDD) substitutions. However, interpretation of the mutagenesis work can be criticized on the grounds that EEEE locus mutations may have additionally disrupted the integrity of a second, non-EEEE locus high-affinity site, and that such a second site may have remained undetected because the mutated pore was probed only from the extracellular pore entrance. Here, we describe the results of experiments designed to test the strength of these criticisms of the single high-affinity locus model of selective permeability in Ca(2+) channels. First, substituted-cysteine accessibility experiments indicate that pore structure in the vicinity of the EEEE locus is not extensively disrupted as a consequence of the quadruple AAAA mutations, suggesting in turn that the quadruple mutations do not distort pore structure to such an extent that a second high affinity site would likely be destroyed. Second, the postulated second high-affinity site was not detected by probing from the intracellularly oriented pore entrance of AAAA and QQQQ mutants. Using inside-out patches, we found that, whereas micromolar Ca(2+) produced substantial block of outward Li(+) current in wild-type channels, internal Ca(2+) concentrations up to 1 mM did not produce detectable block of outward Li(+) current in the AAAA or QQQQ mutants. These results indicate that the EEEE locus is indeed the sole high-affinity Ca(2+) binding locus in the pore of voltage-gated Ca(2+) channels.  相似文献   

3.
Isolated sarcoplasmic reticulum vesicles in the presence of Mg(2+) and absence of Ca(2+) retain significant ATP hydrolytic activity that can be attributed to the Ca(2+)-ATPase protein. At neutral pH and the presence of 5 mM Mg(2+), the dependence of the hydrolysis rate on a linear ATP concentration scale can be fitted by a single hyperbolic function. MgATP hydrolysis is inhibited by either free Mg(2+) or free ATP. The rate of ATP hydrolysis is not perturbed by vanadate, whereas the rate of p-nitrophenyl phosphate hydrolysis is not altered by a nonhydrolyzable ATP analog. ATP binding affinity at neutral pH and in a Ca(2+)-free medium is increased by Mg(2+) but decreased by vanadate when Mg(2+) is present. It is suggested that MgATP hydrolysis in the absence of Ca(2+) requires some optimal adjustment of the enzyme cytoplasmic domains. The Ca(2+)-independent activity is operative at basal levels of cytoplasmic Ca(2+) or when the Ca(2+) binding transition is impeded.  相似文献   

4.
Fluorescent ryanodine revealed the distribution of ryanodine receptors in the submembrane cytoplasm (less than a few micrometers) of cultured bullfrog sympathetic ganglion cells. Rises in cytosolic Ca(2+) ([Ca(2+)](i)) elicited by single or repetitive action potentials (APs) propagated at a high speed (150 microm/s) in constant amplitude and rate of rise in the cytoplasm bearing ryanodine receptors, and then in the slower, waning manner in the deeper region. Ryanodine (10 microM), a ryanodine receptor blocker (and/or a half opener), or thapsigargin (1-2 microM), a Ca(2+)-pump blocker, or omega-conotoxin GVIA (omega-CgTx, 1 microM), a N-type Ca(2+) channel blocker, blocked the fast propagation, but did not affect the slower spread. Ca(2+) entry thus triggered the regenerative activation of Ca(2+)-induced Ca(2+) release (CICR) in the submembrane region, followed by buffered Ca(2+) diffusion in the deeper cytoplasm. Computer simulation assuming Ca(2+) release in the submembrane region reproduced the Ca(2+) dynamics. Ryanodine or thapsigargin decreased the rate of spike repolarization of an AP to 80%, but not in the presence of iberiotoxin (IbTx, 100 nM), a BK-type Ca(2+)-activated K(+) channel blocker, or omega-CgTx, both of which decreased the rate to 50%. The spike repolarization rate and the amplitude of a single AP-induced rise in [Ca(2+)](i) gradually decreased to a plateau during repetition of APs at 50 Hz, but reduced less in the presence of ryanodine or thapsigargin. The amplitude of each of the [Ca(2+)](i) rise correlated well with the reduction in the IbTx-sensitive component of spike repolarization. The apamin-sensitive SK-type Ca(2+)-activated K(+) current, underlying the afterhyperpolarization of APs, increased during repetitive APs, decayed faster than the accompanying rise in [Ca(2+)](i), and was suppressed by CICR blockers. Thus, ryanodine receptors form a functional triad with N-type Ca(2+) channels and BK channels, and a loose coupling with SK channels in bullfrog sympathetic neurons, plastically modulating AP.  相似文献   

5.
Functional large-conductance Ca(2+)- and voltage-activated K(+) (BK) channels can be assembled from four alpha subunits (Slo1) alone, or together with four auxiliary beta1 subunits to greatly increase the apparent Ca(2+) sensitivity of the channel. We examined the structural features involved in this modulation with two types of experiments. In the first, the tail domain of the alpha subunit, which includes the RCK2 (regulator of K(+) conductance) domain and Ca(2+) bowl, was replaced with the tail domain of Slo3, a BK-related channel that lacks both a Ca(2+) bowl and high affinity Ca(2+) sensitivity. In the second, the Ca(2+) bowl was disrupted by mutations that greatly reduce the apparent Ca(2+) sensitivity. We found that the beta1 subunit increased the apparent Ca(2+) sensitivity of Slo1 channels, independently of whether the alpha subunits were expressed as separate cores (S0-S8) and tails (S9-S10) or full length, and this increase was still observed after the Ca(2+) bowl was mutated. In contrast, beta1 subunits no longer increased Ca(2+) sensitivity when Slo1 tails were replaced by Slo3 tails. The beta1 subunits were still functionally coupled to channels with Slo3 tails, as DHS-I and 17 beta-estradiol activated these channels in the presence of beta1 subunits, but not in their absence. These findings indicate that the increase in apparent Ca(2+) sensitivity induced by the beta1 subunit does not require either the Ca(2+) bowl or the linker between the RCK1 and RCK2 domains, and that Slo3 tails cannot substitute for Slo1 tails. The beta1 subunit also induced a decrease in voltage sensitivity that occurred with either Slo1 or Slo3 tails. In contrast, the beta1 subunit-induced increase in apparent Ca(2+) sensitivity required Slo1 tails. This suggests that the allosteric activation pathways for these two types of actions of the beta1 subunit may be different.  相似文献   

6.
Activation of the cardiac ryanodine receptor (RyR2) by Ca(2)+ is an essential step in excitation-contraction coupling in heart muscle. However, little is known about the molecular basis of activation of RyR2 by Ca(2)+. In this study, we investigated the role in Ca(2)+ sensing of the conserved glutamate 3987 located in the predicted transmembrane segment M2 of the mouse RyR2. Single point mutation of this conserved glutamate to alanine (E3987A) reduced markedly the sensitivity of the channel to activation by Ca(2)+, as measured by using single-channel recordings in planar lipid bilayers and by [(3)H]ryanodine binding assay. However, this mutation did not alter the affinity of [(3)H]ryanodine binding and the single-channel conductance. In addition, the E3987A mutant channel was activated by caffeine and ATP, was inhibited by Mg(2)+, and was modified by ryanodine in a fashion similar to that of the wild-type channel. Coexpression of the wild-type and mutant E3987A RyR2 proteins in HEK293 cells produced individual single channels with intermediate sensitivities to activating Ca(2)+. These results are consistent with the view that glutamate 3987 is a major determinant of Ca(2)+ sensitivity to activation of the mouse RyR2 channel, and that Ca(2)+ sensing by RyR2 involves the cooperative action between ryanodine receptor monomers. The results of this study also provide initial insights into the structural and functional properties of the mouse RyR2, which should be useful for studying RyR2 function and regulation in genetically modified mouse models.  相似文献   

7.
Mutational analyses have suggested that BK channels are regulated by three distinct divalent cation-dependent regulatory mechanisms arising from the cytosolic COOH terminus of the pore-forming alpha subunit. Two mechanisms account for physiological regulation of BK channels by microM Ca2+. The third may mediate physiological regulation by mM Mg2+. Mutation of five aspartate residues (5D5N) within the so-called Ca2+ bowl removes a portion of a higher affinity Ca2+ dependence, while mutation of D362A/D367A in the first RCK domain also removes some higher affinity Ca2+ dependence. Together, 5D5N and D362A/D367A remove all effects of Ca2+ up through 1 mM while E399A removes a portion of low affinity regulation by Ca2+/Mg2+. If each proposed regulatory effect involves a distinct divalent cation binding site, the divalent cation selectivity of the actual site that defines each mechanism might differ. By examination of the ability of various divalent cations to activate currents in constructs with mutationally altered regulatory mechanisms, here we show that each putative regulatory mechanism exhibits a unique sensitivity to divalent cations. Regulation mediated by the Ca2+ bowl can be activated by Ca2+ and Sr2+, while regulation defined by D362/D367 can be activated by Ca2+, Sr2+, and Cd2+. Mn2+, Co2+, and Ni2+ produce little observable effect through the high affinity regulatory mechanisms, while all six divalent cations enhance activation through the low affinity mechanism defined by residue E399. Furthermore, each type of mutation affects kinetic properties of BK channels in distinct ways. The Ca2+ bowl mainly accelerates activation of BK channels at low [Ca2+], while the D362/D367-related high affinity site influences both activation and deactivation over the range of 10-300 microM Ca2+. The major kinetic effect of the E399-related low affinity mechanism is to slow deactivation at mM Mg2+ or Ca2+. The results support the view that three distinct divalent-cation binding sites mediate regulation of BK channels.  相似文献   

8.
In skeletal muscle, Mg(2+) exerts a dual inhibitory effect on RyR1, by competing with Ca(2+) at the activation site and binding to a low affinity Ca(2+)/Mg(2+) inhibitory site. Pharmacological activators of RyR1 must overcome the inhibitory action of Mg(2+) before Ca(2+) efflux can occur. In normal muscle, where the free [Mg(2+)](i) is approximately 1mM, even prolonged exposure to millimolar levels of volatile anesthetics does not initiate SR Ca(2+) release. However, when the cytosolic [Mg(2+)] is reduced below the physiological range, low levels of volatile anesthetic within the clinically relevant range (1mM) can initiate SR Ca(2+) release, in the form of a propagating Ca(2+) wave. In human muscle fibers from malignant hyperthermia susceptible patients, such Ca(2+) waves occur when 1mM halothane is applied at physiological [Mg(2+)](i). There is increasing evidence to suggest that defective Mg(2+) regulation of RyR1 confers susceptibility to malignant hyperthermia. At the molecular level, interactions between critical RyR1 subdomains may explain the clustering of RyR1 mutations and associated effects on Mg(2+) regulation.  相似文献   

9.
Cui J  Aldrich RW 《Biochemistry》2000,39(50):15612-15619
The activation of BK type Ca(2+)-activated K(+) channels depends on both voltage and Ca(2+). We studied three point mutations in the putative voltage sensor S4 or S4-S5 linker regions in the mslo1 BK channels to explore the relationship between voltage and Ca(2+) in activating the channel. These mutations reduced the steepness of the open probability - voltage (P(o) - V) relation and increased the shift of the P(o) - V relations on the voltage axis in response to increases in the calcium concentration. It is striking that these two effects were reciprocally related for all three mutations, despite different effects of the mutations on other aspects of the voltage dependence of channel gating. This reciprocal relationship suggests strongly that the free energy contributions to channel activation provided by voltage and by calcium binding are simply additive. We conclude that the Ca(2+) binding sites and the voltage sensors do not directly interact. Rather they both affect the mslo1 channel opening through an allosteric mechanism, by influencing the conformational change between the closed and open conformations. The mutations changed the channel's voltage dependence with little effect on its Ca(2+) affinitiy.  相似文献   

10.
We examined the effects of the mitochondrial Ca(2+)-activated K(+) (mitoBK(Ca)) channel activator NS 1619 on L-type Ca(2+) channels in rat ventricular myocytes. NS 1619 inhibited the Ca(2+) current in a dose-dependent manner. NS 1619 shifted the activation curve to more positive potentials, but did not have a significant effect on the inactivation curve. Pretreatment with inhibitors of membrane BK(Ca) channel, mitoBK(Ca) channel, protein kinase C, protein kinase A, and protein kinase G had little effect on the Ca(2+) current and did not alter the inhibitory effect of NS 1619 significantly. The application of additional NS 1619 in the presence of isoproterenol, a selective beta-adrenoreceptor agonist, reduced the Ca(2+) current to approximately the same level as a single application of NS 1619. In conclusion, our results suggest that NS 1619 inhibits the Ca(2+) current independent of the mitoBK(Ca) channel and protein kinases. Since NS 1619 is widely used to study mitoBK(Ca) channel function, it is essential to verify these unexpected effects of NS 1619 before experimental data can be interpreted accurately.  相似文献   

11.
The binding of Ca(2+) to two adjacent Ca(2+)-binding domains, CBD1 and CBD2, regulates ion transport in the Na(+)/Ca(2+) exchanger. As sensors for intracellular Ca(2+), the CBDs form electrostatic switches that induce the conformational changes required to initiate and sustain Na(+)/Ca(2+) exchange. Depending on the presence of a few key residues in the Ca(2+)-binding sites, zero to four Ca(2+) ions can bind with affinities between 0.1 to 20 μm. Importantly, variability in CBD2 as a consequence of alternative splicing modulates not only the number and affinities of the Ca(2+)-binding sites in CBD2 but also the Ca(2+) affinities in CBD1.  相似文献   

12.
Large-conductance Ca(2+)-activated K(+) channels (BK(Ca) channels) are regulated by the tissue-specific expression of auxiliary beta subunits. Beta1 is predominantly expressed in smooth muscle, where it greatly enhances the BK(Ca) channel's Ca(2+) sensitivity, an effect that is required for proper regulation of smooth muscle tone. Here, using gating current recordings, macroscopic ionic current recordings, and unitary ionic current recordings at very low open probabilities, we have investigated the mechanism that underlies this effect. Our results may be summarized as follows. The beta1 subunit has little or no effect on the equilibrium constant of the conformational change by which the BK(Ca) channel opens, and it does not affect the gating charge on the channel's voltage sensors, but it does stabilize voltage sensor activation, both when the channel is open and when it is closed, such that voltage sensor activation occurs at more negative voltages with beta1 present. Furthermore, beta1 stabilizes the active voltage sensor more when the channel is closed than when it is open, and this reduces the factor D by which voltage sensor activation promotes opening by approximately 24% (16.8-->12.8). The effects of beta1 on voltage sensing enhance the BK(Ca) channel's Ca(2+) sensitivity by decreasing at most voltages the work that Ca(2+) binding must do to open the channel. In addition, however, in order to fully account for the increase in efficacy and apparent Ca(2+) affinity brought about by beta1 at negative voltages, our studies suggest that beta1 also decreases the true Ca(2+) affinity of the closed channel, increasing its Ca(2+) dissociation constant from approximately 3.7 microM to between 4.7 and 7.1 microM, depending on how many binding sites are affected.  相似文献   

13.
To investigate the metal-binding properties of KChIP1, the interaction of KChIP1 and mutated KChIP1 with divalent cations (Mg(2+), Ca(2+), Sr(2+), and Ba(2+)) was explored by 8-anilinonaphthalene-1-sulfonate (ANS) fluorescence. It showed that KChIP1 possessed two types of Ca(2+)-binding sites, high-affinity and low-affinity Ca(2+)-binding sites. However, only low-affinity-binding site for Mg(2+), Sr(2+), and Ba(2+) was observed. The metal-binding properties of KChIP1 are not appreciably affected after removal of the N-terminal portion and EF-hand 1. Deleting the EF-hand 4 of KChIP1 abolishes its high-affinity Ca(2+)-binding site, but retains the intact low-affinity-binding site for metal ions. A decrease in the nonpolarity of ANS-binding site occurs with all mutants. However, the binding of ANS with KChIP1 is no longer observed after removal of EF-hands 3 and 4. Intermolecular interaction assessed by chemical cross-linking suggested that KChIP1 had a propensity to form dimer in the absence of metal ions, and a KChIP1 tetramer was pronouncedly produced in the presence of metal ions. Noticeably, the oligomerization state depends on the integrity of EF-hand 4. Taken together, our data suggest that EF-hand 4 is of structural importance as well as functional importance for fulfilling the physiological function of KChIP1.  相似文献   

14.
Activation of large conductance Ca(2+)-activated K(+) channels is controlled by both cytoplasmic Ca(2+) and membrane potential. To study the mechanism of voltage-dependent gating, we examined mSlo Ca(2+)-activated K(+) currents in excised macropatches from Xenopus oocytes in the virtual absence of Ca(2+) (<1 nM). In response to a voltage step, I(K) activates with an exponential time course, following a brief delay. The delay suggests that rapid transitions precede channel opening. The later exponential time course suggests that activation also involves a slower rate-limiting step. However, the time constant of I(K) relaxation [tau(I(K))] exhibits a complex voltage dependence that is inconsistent with models that contain a single rate limiting step. tau(I(K)) increases weakly with voltage from -500 to -20 mV, with an equivalent charge (z) of only 0.14 e, and displays a stronger voltage dependence from +30 to +140 mV (z = 0.49 e), which then decreases from +180 to +240 mV (z = -0.29 e). Similarly, the steady state G(K)-V relationship exhibits a maximum voltage dependence (z = 2 e) from 0 to +100 mV, and is weakly voltage dependent (z congruent with 0.4 e) at more negative voltages, where P(o) = 10(-5)-10(-6). These results can be understood in terms of a gating scheme where a central transition between a closed and an open conformation is allosterically regulated by the state of four independent and identical voltage sensors. In the absence of Ca(2+), this allosteric mechanism results in a gating scheme with five closed (C) and five open (O) states, where the majority of the channel's voltage dependence results from rapid C-C and O-O transitions, whereas the C-O transitions are rate limiting and weakly voltage dependent. These conclusions not only provide a framework for interpreting studies of large conductance Ca(2+)-activated K(+) channel voltage gating, but also have important implications for understanding the mechanism of Ca(2+) sensitivity.  相似文献   

15.
BK channels modulate neurotransmitter release due to their activation by voltage and Ca(2+). Intracellular Mg(2+) also modulates BK channels in multiple ways with opposite effects on channel function. Previous single-channel studies have shown that Mg(2+) blocks the pore of BK channels in a voltage-dependent manner. We have confirmed this result by studying macroscopic currents of the mslo1 channel. We find that Mg(2+) activates mslo1 BK channels independently of Ca(2+) and voltage by preferentially binding to their open conformation. The mslo3 channel, which lacks Ca(2+) binding sites in the tail, is not activated by Mg(2+). However, coexpression of the mslo1 core and mslo3 tail produces channels with Mg(2+) sensitivity similar to mslo1 channels, indicating that Mg(2+) sites differ from Ca(2+) sites. We discovered that Mg(2+) also binds to Ca(2+) sites and competitively inhibits Ca(2+)-dependent activation. Quantitative computation of these effects reveals that the overall effect of Mg(2+) under physiological conditions is to enhance BK channel function.  相似文献   

16.
The universal signal for egg activation at fertilization is a rise in cytoplasmic Ca(2+) with defined spatial and temporal kinetics. Mammalian and amphibian eggs acquire the ability to produce such Ca(2+) signals during a maturation period that precedes fertilization and encompasses resumption of meiosis and progression to metaphase II. In Xenopus, immature oocytes produce fast, saltatory Ca(2+) waves that can be oscillatory in nature in response to IP(3). In contrast, mature eggs produce a single continuous, sweeping Ca(2+) wave in response to IP(3) or sperm fusion. The mechanisms mediating the differentiation of Ca(2+) signaling during oocyte maturation are not well understood. Here, I characterized elementary Ca(2+) release events (Ca(2+) puffs) in oocytes and eggs and show that the sensitivity of IP(3)-dependent Ca(2+) release is greatly enhanced during oocyte maturation. Furthermore, Ca(2+) puffs in eggs have a larger spatial fingerprint, yet are short lived compared to oocyte puffs. Most interestingly, Ca(2+) puffs cluster during oocyte maturation resulting in a continuum of Ca(2+) release sites over space in eggs. These changes in the spatial distribution of elementary Ca(2+) release events during oocyte maturation explain the continuous nature and slower speed of the fertilization Ca(2+) wave.  相似文献   

17.
The voltage- and Ca2+-dependent gating mechanism of large-conductance Ca2+-activated K+ (BK) channels from cultured rat skeletal muscle was studied using single-channel analysis. Channel open probability (Po) increased with depolarization, as determined by limiting slope measurements (11 mV per e-fold change in Po; effective gating charge, q(eff), of 2.3 +/- 0.6 e(o)). Estimates of q(eff) were little changed for intracellular Ca2+ (Ca2+(i)) ranging from 0.0003 to 1,024 microM. Increasing Ca2+(i) from 0.03 to 1,024 microM shifted the voltage for half maximal activation (V(1/2)) 175 mV in the hyperpolarizing direction. V(1/2) was independent of Ca2+(i) for Ca2+(i) < or = 0.03 microM, indicating that the channel can be activated in the absence of Ca2+(i). Open and closed dwell-time distributions for data obtained at different Ca2+(i) and voltage, but at the same Po, were different, indicating that the major action of voltage is not through concentrating Ca2+ at the binding sites. The voltage dependence of Po arose from a decrease in the mean closing rate with depolarization (q(eff) = -0.5 e(o)) and an increase in the mean opening rate (q(eff) = 1.8 e(o)), consistent with voltage-dependent steps in both the activation and deactivation pathways. A 50-state two-tiered model with separate voltage- and Ca2+-dependent steps was consistent with the major features of the voltage and Ca2+ dependence of the single-channel kinetics over wide ranges of Ca2+(i) (approximately 0 through 1,024 microM), voltage (+80 to -80 mV), and Po (10(-4) to 0.96). In the model, the voltage dependence of the gating arises mainly from voltage-dependent transitions between closed (C-C) and open (O-O) states, with less voltage dependence for transitions between open and closed states (C-O), and with no voltage dependence for Ca2+-binding and unbinding. The two-tiered model can serve as a working hypothesis for the Ca2+- and voltage-dependent gating of the BK channel.  相似文献   

18.
The COOH-terminal S9-S10 tail domain of large conductance Ca(2+)-activated K(+) (BK) channels is a major determinant of Ca(2+) sensitivity (Schreiber, M., A. Wei, A. Yuan, J. Gaut, M. Saito, and L. Salkoff. 1999. Nat. Neurosci. 2:416-421). To investigate whether the tail domain also modulates Ca(2+)-independent properties of BK channels, we explored the functional differences between the BK channel mSlo1 and another member of the Slo family, mSlo3 (Schreiber, M., A. Yuan, and L. Salkoff. 1998. J. Biol. Chem. 273:3509-3516). Compared with mSlo1 channels, mSlo3 channels showed little Ca(2+) sensitivity, and the mean open time, burst duration, gaps between bursts, and single-channel conductance of mSlo3 channels were only 32, 22, 41, and 37% of that for mSlo1 channels, respectively. To examine which channel properties arise from the tail domain, we coexpressed the core of mSlo1 with either the tail domain of mSlo1 or the tail domain of mSlo3 channels, and studied the single-channel currents. Replacing the mSlo1 tail with the mSlo3 tail resulted in the following: increased open probability in the absence of Ca(2+); reduced the Ca(2+) sensitivity greatly by allowing only partial activation by Ca(2+) and by reducing the Hill coefficient for Ca(2+) activation; decreased the voltage dependence approximately 28%; decreased the mean open time two- to threefold; decreased the mean burst duration three- to ninefold; decreased the single-channel conductance approximately 14%; decreased the K(d) for block by TEA(i) approximately 30%; did not change the minimal numbers of three to four open and five to seven closed states entered during gating; and did not change the major features of the dependency between adjacent interval durations. These observations support a modular construction of the BK channel in which the tail domain modulates the gating kinetics and conductance properties of the voltage-dependent core domain, in addition to determining most of the high affinity Ca(2+) sensitivity.  相似文献   

19.
Ca(2+) channel inactivation in the neurons of the freshwater snail, Lymnaea stagnalis, was studied using patch-clamp techniques. In the presence of a high concentration of intracellular Ca(2+) buffer (5 mM EGTA), the inactivation of these Ca(2+) channels is entirely voltage dependent; it is not influenced by the identity of the permeant divalent ions or the amount of extracellular Ca(2+) influx, or reduced by higher levels of intracellular Ca(2+) buffering. Inactivation measured under these conditions, despite being independent of Ca(2+) influx, has a bell-shaped voltage dependence, which has often been considered a hallmark of Ca(2+)-dependent inactivation. Ca(2+)-dependent inactivation does occur in Lymnaea neurons, when the concentration of the intracellular Ca(2+) buffer is lowered to 0.1 mM EGTA. However, the magnitude of Ca(2+)-dependent inactivation does not increase linearly with Ca(2+) influx, but saturates for relatively small amounts of Ca(2+) influx. Recovery from inactivation at negative potentials is biexponential and has the same time constants in the presence of different intracellular concentrations of EGTA. However, the amplitude of the slow component is selectively enhanced by a decrease in intracellular EGTA, thus slowing the overall rate of recovery. The ability of 5 mM EGTA to completely suppress Ca(2+)-dependent inactivation suggests that the Ca(2+) binding site is at some distance from the channel protein itself. No evidence was found of a role for serine/threonine phosphorylation in Ca(2+) channel inactivation. Cytochalasin B, a microfilament disrupter, was found to greatly enhance the amount of Ca(2+) channel inactivation, but the involvement of actin filaments in this effect of cytochalasin B on Ca(2+) channel inactivation could not be verified using other pharmacological compounds. Thus, the mechanism of Ca(2+)-dependent inactivation in these neurons remains unknown, but appears to differ from those proposed for mammalian L-type Ca(2+) channels.  相似文献   

20.
Depletion of intracellular Ca(2+) stores evokes Ca(2+) entry across the plasma membrane by inducing Ca(2+) release-activated Ca(2+) (CRAC) currents in many cell types. Recently, Orai and STIM proteins were identified as the molecular identities of the CRAC channel subunit and the endoplasmic reticulum Ca(2+) sensor, respectively. Here, extensive database searching and phylogenetic analysis revealed several lineage-specific duplication events in the Orai protein family, which may account for the evolutionary origins of distinct functional properties among mammalian Orai proteins. Based on similarity to key structural domains and essential residues for channel functions in Orai proteins, database searching also identifies a putative primordial Orai sequence in hyperthermophilic archaeons. Furthermore, modern Orai appears to acquire new structural domains as early as Urochodata, before divergence into vertebrates. The evolutionary patterns of structural domains might be related to distinct functional properties of Drosophila and mammalian CRAC currents. Interestingly, Orai proteins display two conserved internal repeats located at transmembrane segments 1 and 3, both of which contain key amino acids essential for channel function. These findings demonstrate biochemical and physiological relevance of Orai proteins in light of different evolutionary origins and will provide novel insights into future structural and functional studies of Orai proteins.  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