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.
Changes in holding potential (Vh), affect both gating charge (the Q(Vh) curve) and peak ionic current (the F(Vh) curve) seen at positive test potentials. Careful comparison of the Q(Vh) and F(Vh) distributions indicates that these curves are similar, having two slopes (approximately 2.5e for Vh from -115 to -90 mV and approximately 4e for Vh from -90 to -65 mV) and very negative midpoints (approximately -86 mV). Thus, gating charge movement and channel availability appear closely coupled under fully-equilibrated conditions. The time course by which channels approach equilibration was explored using depolarizing prepulses of increasing duration. The high slope component seen in the F(Vh) and Q(Vh) curves is not evident following short depolarizing prepulses in which the prepulse duration approximately corresponds to the settling time for fast inactivation. Increasing the prepulse duration to 10 ms or longer reveals the high slope, and left-shifts the midpoint to more negative voltages, towards the F(Vh) and Q(Vh) distributions. These results indicate that a separate slow-moving voltage sensor affects the channels at prepulse durations greater than 10 ms. Charge movement and channel availability remain closely coupled as equilibrium is approached using depolarizing pulses of increasing durations. Both measures are 50% complete by 50 ms at a prepulse potential of -70 mV, with proportionately faster onset rates when the prepulse potential is more depolarized. By contrast, charge movement and channel availability dissociate during recovery from prolonged depolarizations. Recovery of gating charge is considerably faster than recovery of sodium ionic current after equilibration at depolarized potentials. Recovery of gating charge at -140 mV, is 65% complete within approximately 100 ms, whereas less than 30% of ionic current has recovered by this time. Thus, charge movement and channel availability appear to be uncoupled during recovery, although both rates remain voltage sensitive. These data suggest that channels remain inactivated due to a separate process operating in parallel with the fast gating charge. We demonstrate that this behavior can be simulated by a model in which the fast charge movement associated with channel activation is electrostatically-coupled to a separate slow voltage sensor responsible for the slow inactivation of channel conductance.  相似文献   

2.
Nonlinear or asymmetric charge movement was recorded from single ventricular myocytes cultured from 17-d-old embryonic chick hearts using the whole-cell patch clamp method. The myocytes were exposed to the appropriate intracellular and extracellular solutions designed to block Na+, Ca2+, and K+ ionic currents. The linear components of the capacity and leakage currents during test voltage steps were eliminated by adding summed, hyperpolarizing control step currents. Upon depolarization from negative holding potentials the nonlinear charge movement was composed of two distinct and separable kinetic components. An early rapidly decaying component (decay time constant range: 0.12-0.50 ms) was significant at test potentials positive to -70 mV and displayed saturation above 0 mV (midpoint -35 mV; apparent valence 1.6 e-). The early ON charge was partially immobilized during brief (5 ms) depolarizing test steps and was more completely immobilized by the application of less negative holding potentials. A second slower-decaying component (decay time constant range: 0.88-3.7 ms) was activated at test potentials positive to -60 mV and showed saturation above +20 mV (midpoint -13 mV, apparent valence 1.9 e-). The second component of charge movement was immobilized by long duration (5 s) holding potentials, applied over a more positive voltage range than those that reduced the early component. The voltage dependencies for activation and inactivation of the Na+ and Ca2+ ionic currents were determined for myocytes in which these currents were not blocked. There was a positive correlation between the voltage dependence of activation and inactivation of the Na+ and Ca2+ ionic currents and the activation and immobilization of the fast and slow components of charge movement. These complementary kinetic and steady-state properties lead to the conclusion that the two components of charge movement are associated with the voltage-sensitive conformational changes that precede Na+ and Ca2+ channel openings.  相似文献   

3.
Gating currents were measured by subtracting the linear component of the capacitative current recorded at very positive or very negative potentials. When the membrane is depolarized for a few minutes, repolarized to the usual holding potential (HP) of --70 mV for 1 ms, and then pulsed to 0 mV, the charge transferred in 2--4 ms is approximately 50% of that which was transferred during the same pulse holding at --70 mV. This charge decrease, called slow inactivation of the gating current, was found to be consistent with a shift of the charge vs. potential (Q-V) curve to more hyperpolarized potentials. When the HP is 0 mV, the total charge available to move is the same as the total charge available when the HP is --70 mV. The time constants of the fast component of the ON gating current are smaller at depolarized holding potentials than at --70 mV. When the HP is --70 mV and a prepulse of 50 ms duration is given to 0 mV, the Q-V curve is also shifted to more hyperpolarized potentials (charge immobilization), but the effect is not as pronounced as the one obtained by holding at 0 mV. When the HP is 0 mV, a prepulse to --70 mV for 50 ms partially shifts back the Q-V curve, indicating that fast inactivation of the gating charge may be recovered in the presence of slow inactivation. A physical model consisting of a gating particle that interacts with a fast inactivating particle, and a slow inactivating particle, reproduces most of the experimental results.  相似文献   

4.
When attached outside the voltage-sensing S4 segment of the Shaker potassium channel, the fluorescent probe tetramethylrhodamine (TMRM) undergoes voltage-dependent fluorescence changes (DeltaF) due to differential interaction with a pH-titratable external protein-lined vestibule (Cha, A., and F. Bezanilla. 1998. J. Gen. Physiol. 112:391-408.). We attached TMRM at the same sites [corresponding to M356C and A359C in the wild-type (wt) channel] in a deletion mutant of Shaker where all but the five amino acids closest to S4 had been removed from the S3-S4 linker. In the deletion mutant, the maximal DeltaF/F seen was diminished 10-fold, and the DeltaF at M356C became pH independent, suggesting that the protein-lined vestibule is made up in large part by the S3-S4 linker. The residual DeltaF showed that the probe still interacted with two putative quenching groups near the S4 segment. One group was detected by M356C-TMRM (located outside of S3 in the deletion mutant) and reported on deactivation gating charge movement when applying hyperpolarizing voltage steps from a holding potential of 0 mV. During activating voltage steps from a holding potential of -90 mV, the fluorescence lagged considerably behind the movement of gating charge over a range of potentials. Another putative quenching group was seen by probes attached closer to the S4 and caused a DeltaF at extreme hyperpolarizations (more negative than -90 mV) only. A signal from the interaction with this group in the wt S3-S4 linker channel (at L361C) correlated with gating charge moving in the hyperpolarized part of the Q-V curve. Probe attached at A359C in the deletion mutant and at L361C in wt channel showed a biphasic DeltaF as the probe oscillated between the two groups, revealing that there is a transient state of the voltage sensor in between, where the probe has maximal fluorescence. We conclude that the voltage sensor undergoes two distinct conformational changes as seen from probes attached outside the S4 segment.  相似文献   

5.
Sodium channel activations, measured as the fraction of channels open to peak conductance for different test potentials (F[V]), shows two statistically different slopes from holding potential more positive than -90 mV. A high valence of 4-6e is indicated a test potentials within 35 mV of the apparent threshold potential (circa -65 mV at -85 mV holding potential). However, for test potentials positive to -30 mV, the F(V) curve shows a 2e valence. The F(V) curve for crayfish axon sodium channels at these "depolarized" holding potentials thus closely resembles classic data obtained from other preparations at holding potentials between -80 and -60 mV. In contrast, at holding potentials more negative than -100 mV, the high slope essentially disappears and the F(V) curve follows a single Boltzmann distribution with a valence of approximately 2e at all potentials. Neither the slope of this simple distribution nor its midpoint (-20 mV) was significantly affected by removal of fast inactivation with pronase. The change in F(V) slope, when holding potential is increased from -85 to -120 mV, does not appear to be caused by the contribution of a second channel type. The simple voltage dependence of sodium current found at Vh -120 mV be used by to discriminate between models of sodium channel activation, and rules out models with three particles of equal valence.  相似文献   

6.
Intramembrane charge movement was recorded in rat and rabbit ventricular cells using the whole-cell voltage clamp technique. Na and K currents were eliminated by using tetraethylammonium as the main cation internally and externally, and Ca channel current was blocked by Cd and La. With steps in the range of -110 to -150 used to define linear capacitance, extra charge moves during steps positive to approximately -70 mV. With holding potentials near -100 mV, the extra charge moving outward on depolarization (ON charge) is roughly equal to the extra charge moving inward on repolarization (OFF charge) after 50-100 ms. Both ON and OFF charge saturate above approximately +20 mV; saturating charge movement is approximately 1,100 fC (approximately 11 nC/muF of linear capacitance). When the holding potential is depolarized to -50 mV, ON charge is reduced by approximately 40%, with little change in OFF charge. The reduction of ON charge by holding potential in this range matches inactivation of Na current measured in the same cells, suggesting that this component might arise from Na channel gating. The ON charge remaining at a holding potential of -50 mV has properties expected of Ca channel gating current: it is greatly reduced by application of 10 muM D600 when accompanied by long depolarizations and it is reduced at more positive holding potentials with a voltage dependence similar to that of Ca channel inactivation. However, the D600-sensitive charge movement is much larger than the Ca channel gating current that would be expected if the movement of channel gating charge were always accompanied by complete opening of the channel.  相似文献   

7.
Intramembrane charge movement was measured on skeletal muscle fibers of the frog in a single Vaseline-gap voltage clamp. Charge movements determined both under polarized conditions (holding potential, VH = -100 mV; Qmax = 30.4 +/- 4.7 nC/micro(F), V = -44.4 mV, k = 14.1 mV; charge 1) and in depolarized states (VH = 0 mV; Qmax = 50.0 +/- 6.7 nC/micro(F), V = -109.1 mV, k = 26.6 mV; charge 2) had properties as reported earlier. Linear capacitance (LC) of the polarized fibers was increased by 8.8 +/- 4.0% compared with that of the depolarized fibers. Using control pulses measured under depolarized conditions to calculate charge 1, a minor change in the voltage dependence (to V = -44.6 mV and k = 14.5 mV) and a small increase in the maximal charge (to Qmax = 31.4 +/- 5.5 nC/micro(F] were observed. While in most cases charge 1 transients seemed to decay with a single exponential time course, charge 2 currents showed a characteristic biexponential behavior at membrane potentials between -90 and -180 mV. The voltage dependence of the rate constant of the slower component was fitted with a simple constant field diffusion model (alpha m = 28.7 s-1, V = -124.0 mV, and k = 15.6 mV). The midpoint voltage (V) was similar to that obtained from the Q-V fit of charge 2, while the steepness factor (k) resembled that of charge 1. This slow component could also be isolated using a stepped OFF protocol; that is, by hyperpolarizing the membrane to -190 mV for 200 ms and then coming back to 0 mV in two steps. The faster component was identified as an ionic current insensitive to 20 mM Co2+ but blocked by large hyperpolarizing pulses. These findings are consistent with the model implying that charge 1 and the slower component of charge 2 interconvert when the holding potential is changed. They also explain the difference previously found when comparing the steepness factors of the voltage dependence of charge 1 and charge 2.  相似文献   

8.
Sodium channel gating currents in frog skeletal muscle   总被引:7,自引:5,他引:2       下载免费PDF全文
Charge movements similar to those attributed to the sodium channel gating mechanism in nerve have been measured in frog skeletal muscle using the vaseline-gap voltage-clamp technique. The time course of gating currents elicited by moderate to strong depolarizations could be well fitted by the sum of two exponentials. The gating charge exhibits immobilization: at a holding potential of -90 mV the proportion of charge that returns after a depolarizing prepulse (OFF charge) decreases with the duration of the prepulse with a time course similar to inactivation of sodium currents measured in the same fiber at the same potential. OFF charge movements elicited by a return to more negative holding potentials of -120 or -150 mV show distinct fast and slow phases. At these holding potentials the total charge moved during both phases of the gating current is equal to the ON charge moved during the preceding prepulse. It is suggested that the slow component of OFF charge movement represents the slower return of charge "immobilized" during the prepulse. A slow mechanism of charge immobilization is also evident: the maximum charge moved for a strong depolarization is approximately doubled by changing the holding potential from -90 to -150 mV. Although they are larger in magnitude for a -150-mV holding potential, the gating currents elicited by steps to a given potential have similar kinetics whether the holding potential is -90 or -150 mV.  相似文献   

9.
Charge movement was measured in frog cut twitch fibers with the double Vaseline-gap technique. Steady-state inactivation of charge movement was studied by changing the holding potential from -90 mV to a level ranging from -70 to -30 mV. Q beta and Q gamma at each holding potential were separated by fitting the Q-V plot with a sum of two Boltzmann distribution functions. At -70 mV Q beta and Q gamma were inactivated to 54.0% (SEM 2.2) and 82.7% (SEM 3.0) of the amounts at -90 mV. At holding potentials greater than or equal to -60 mV, more Q gamma was inactivated than Q beta, and at -30 mV Q gamma was completely inactivated but Q beta was not. There was no holding potential at which Q beta was unaffected and Q gamma was completely inactivated. The differences between the residual fractions of Q beta and Q gamma are significant at all holding potentials (P less than 0.001-0.05). The plot of the residual fraction of Q beta or Q gamma versus holding potential can be fitted well by an inverted sigmoidal curve that is a mirror image of the activation curve of the respective charge component. The pair of curves for Q gamma correlates well with those for tension generation or Ca release obtained by other investigators. The time courses of the inactivation of Q beta and Q gamma were studied by obtaining several Q-V plots with conditioning depolarizations lasting 1-20 s and separating each Q-V plot into Q beta and Q gamma components by fitting with a sum of two Boltzmann distribution functions. The inactivation time constant of Q beta was found to be 5-10 times as large as that of Q gamma. During repetitive stimulation, prominent I gamma humps could be observed in TEST-minus-CONTROL current traces and normal Q gamma components could be separated from the Q-V plots, whether 20 or 50 mM EGTA was present in the internal solution, whether 2 or 10 stimulations were used, and whether the stimuli were separated by 400 ms or 6 s. Repetitive stimulation slowed the kinetics of the I gamma hump and could shift the Q-V curve slightly in the depolarizing direction in some cases, resulting in an apparent suppression of charge at the potentials that fall on the steep part of the Q-V curve.  相似文献   

10.
Single sodium channel currents were analysed in cell attached patches from single ventricular cells of guinea pig hearts in the presence of a novel cardiotonic compound DPI 201-106. The mean single channel conductance of DPI-treated Na channels was not changed by DPI (20.8 +/- 4 pS, control, 3 patches; 21.3 +/- 1 pS with DPI, 5 mumol/1,3 patches). DPI voltage-dependently prolongs the cardiac sodium channel openings by removal of inactivation at potentials positive to -40 mV. At potentials negative to -40 mV a clustering of short openings at the very beginning of the depolarizing voltage steps can be observed causing a transient time course of the averaged currents. Long openings induced an extremely slow inactivation. Short openings, long openings and nulls appeared in groups referring to a modal gating behaviour of DPI-treated sodium channels. DPI-modified Na channels showed a monotonously prolonged mean open time with increased depolarizing voltage steps, e.g. the open state probability within a sweep was increased. However, the number of non-empty sweeps was decreased with the magnitude of the depolarizing steps, e.g. the probability of the channel being open as calculated from the averaged currents was voltage-dependently decreased by DPI (50% decrease at -50.7 +/- 9 9 mV, 3 patches). Short and long openings of DPI-modified channels could be separated by variation of the holding potential. The occurrence of long Na channel openings was much more suppressed by reducing the holding potential (half maximum inactivation at -112 +/- 8 mV, 4 patches) than that of short openings (half maximum inactivation at -88 +/- 8 mV, 4 patches). Otherwise, short living openings completely disappeared at potentials positive to -40 mV where the occurrence of long openings was favoured. The differential voltage dependence of blocking and activating effects of DPI on cardiac Na channels as well as the differential voltage dependence of the appearance of short and long openings refers to a modal gating behaviour of cardiac Na channels.  相似文献   

11.
The bacterial sodium channel, NaChBac, from Bacillus halodurans provides an excellent model to study structure-function relationships of voltage-gated ion channels. It can be expressed in mammalian cells for functional studies as well as in bacterial cultures as starting material for protein purification for fine biochemical and biophysical studies. Macroscopic functional properties of NaChBac have been described previously (Ren, D., B. Navarro, H. Xu, L. Yue, Q. Shi, and D.E. Clapham. 2001. Science. 294:2372-2375). In this study, we report gating current properties of NaChBac expressed in COS-1 cells. Upon depolarization of the membrane, gating currents appeared as upward inflections preceding the ionic currents. Gating currents were detectable at -90 mV while holding at -150 mV. Charge-voltage (Q-V) curves showed sigmoidal dependence on voltage with gating charge saturating at -10 mV. Charge movement was shifted by -22 mV relative to the conductance-voltage curve, indicating the presence of more than one closed state. Consistent with this was the Cole-Moore shift of 533 micros observed for a change in preconditioning voltage from -160 to -80 mV. The total gating charge was estimated to be 16 elementary charges per channel. Charge immobilization caused by prolonged depolarization was also observed; Q-V curves were shifted by approximately -60 mV to hyperpolarized potentials when cells were held at 0 mV. The kinetic properties of NaChBac were simulated by simultaneous fit of sodium currents at various voltages to a sequential kinetic model. Gating current kinetics predicted from ionic current experiments resembled the experimental data, indicating that gating currents are coupled to activation of NaChBac and confirming the assertion that this channel undergoes several transitions between closed states before channel opening. The results indicate that NaChBac has several closed states with voltage-dependent transitions between them realized by translocation of gating charge that causes activation of the channel.  相似文献   

12.
The steady-state gating of individual batrachotoxin-modified sodium channels in neutral phospholipid bilayers exhibits spontaneous, reversible changes in channel activation, such that the midpoint potential (Va) for the gating curves may change, by 30 mV or more, with or without a change in the apparent gating valence (za). Consequently, estimates for Va and, in particular, za from ensemble-averaged gating curves differ from the average values for Va and za from single-channel gating curves. In addition to these spontaneous variations, the average Va shifts systematically as a function of [NaCl] (being -109, -88, and -75 mV at 0.1, 0.5, and 1.0 M NaCl), with no systematic variation in the average za (approximately 3.7). The [NaCl]-dependent shifts in Va were interpreted in terms of screening of fixed charges near the channels' gating machinery. Estimates for the extracellular and intracellular apparent charge densities (sigma e = -0.7 and sigma i = -0.08 e/nm2) were obtained from experiments in symmetrical and asymmetrical NaCl solutions using the Gouy-Chapman theory. In 0.1 M NaCl the extracellular and intracellular surface potentials are estimated to be -94 and -17 mV, respectively. The intrinsic midpoint potential, corrected for the surface potentials, is thus about -30 mV, and the standard free energy of activation is approximately -12 kJ/mol. In symmetrical 0.1 M NaCl, addition of 0.005 M Ba2+ to the extracellular solution produced a 17-mV depolarizing shift in Va and a slight reduction in za. The shift is consistent with predictions using the Gouy-Chapman theory and the above estimate for sigma e. Subsequent addition of 0.005 M Ba2+ to the intracellular solution produced a approximately 5-mV hyperpolarizing shift in the ensemble-averaged gating curve and reduced za by approximately 1. This Ba(2+)-induced shift is threefold larger than predicted, which together with the reduction in za implies that Ba2+ may bind at the intracellular channel surface.  相似文献   

13.
Steady-state and kinetic properties of gating currents of the Shaker K+ channels were studied in channels expressed in Xenopus oocytes and recorded with the cut-open oocyte voltage clamp. The charge versus potential (Q-V) curve reveals at least two components of charge, the first moving in the hyperpolarized region (V1/2 = -63 mV) and the second, with a larger apparent valence, moving in the more depolarized region (V1/2 = -44 mV). The kinetic analysis of gating currents revealed also two exponential decaying components that corresponded in their voltage dependence with the charge components described in the steady-state. The first component was found to correlate with the effects of prepulses that produce the Cole-Moore shift of the ionic and gating currents and seems to be occurring completely within closed conformations of the channel. The second component seems to be related to the events occurring between the closed states just preceding, but not including, the transition to the open state. The ON and OFF gating currents exhibit a pronounced rising phase at potentials at which the second component becomes important, and this region corresponds to the potential range where the channel opens. The results could not be explained with simple parallel models, but the data can be fitted to a sequential model that could be related to a first rearrangement of the putative four subunits in cooperative fashion, followed by a concerted charge movement that leads to the open channel. The first series of charge movements are produced by transitions between several closed states carrying less than two electronic charges per step, while a step carrying about 3.5 electronic charges can explain the second component. This step is followed by the transition to the open state carrying less than 0.5 electronic charges. This model is able to reproduce all the kinetic and steady-state properties of the gating currents and predicts many of the properties of the ionic currents.  相似文献   

14.
The effects of extracellular saxitoxin (STX) and tetrodotoxin (TTX) on gating current (IgON) were studied in voltage clamped crayfish giant axons. At a holding potential (VH) of -90 mV, integrated gating charge (QON) was found to be 56% suppressed when 200 nM STX was added to the external solution, and 75% suppressed following the addition of 200 nM TTX. These concentrations of toxin are sufficiently high to block greater than 99% of sodium channels. A smaller suppression of IgON was observed when 1 nM STX was used (KD = 1-2 nM STX). The suppression of IgON by external toxin was found to be hold potential dependent, with only minimal suppression observed at the most hyperpolarized hold potentials, -140 to -120 mV. The maximal effect of these toxins on IgON was observed at hold potentials where the QON vs. VH plot was found to be steepest, -100 to -80 mV. The suppression of IgON induced by TTX is partially relieved following the removal of fast inactivation by intracellular treatment with N-bromoacetamide (NBA). The effect of STX and TTX on IgON is equivalent to a hyperpolarizing shift in the steady state inactivation curve, with 200 nM STX and 200 nM TTX inducing shifts of 4.9 +/- 1.7 mV and 10.0 +/- 2.1 mV, respectively. Our results are consistent with a model where the binding of toxin displaces a divalent cation from a negatively charged site near the external opening of the sodium channel, thereby producing a voltage offset sensed by the channel gating apparatus.  相似文献   

15.
The effect of acidosis on the transient outward K(+) current (I(to)) of rat ventricular myocytes has been investigated using the perforated patch-clamp technique. When the holding potential was -80 mV, depolarizing pulses to potentials positive to -20 mV activated I(to) in subepicardial cells but activated little I(to) in subendocardial cells. Exposure to an acid solution (pH 6.5) had no significant effect on I(to) activated from this holding potential in either subepicardial or subendocardial cells. When the holding potential was -40 mV, acidosis significantly increased I(to) at potentials positive to -20 mV in subepicardial cells but had little effect on I(to) in subendocardial cells. The increase in I(to) in subepicardial cells was inhibited by 10 mM 4-aminopyridine. In subepicardial cells, acidosis caused a +8.57-mV shift in the steady-state inactivation curve. It is concluded that in subepicardial rat ventricular myocytes acidosis increases the amplitude of I(to) as a consequence of a depolarizing shift in the voltage dependence of inactivation.  相似文献   

16.
Batrachotoxin-activated rat brain Na+ channels were reconstituted in neutral planar phospholipid bilayers in high ionic strength solutions (3 M NaCl). Under these conditions, diffuse surface charges present on the channel protein are screened. Nevertheless, the addition of extracellular and/or intracellular Ba2+ caused the following alterations in the gating of Na+ channels: (a) external (or internal) Ba2+ caused a depolarizing (or hyperpolarizing) voltage shift in the gating curve (open probability versus membrane potential curve) of the channels; (b) In the concentration range of 10-120 mM, extracellular Ba2+ caused a larger voltage shift in the gating curve of Na+ channels than intracellular Ba2+; (c) voltage shifts of the gating curve of Na+ channels as a function of external or internal Ba2+ were fitted with a simple binding isotherm with the following parameters: for internal Ba2+, delta V0.5,max (maximum voltage shift) = -11.5 mV, KD = 64.7 mM; for external Ba2+, delta V0.5,max = 13.5 mV, KD = 25.8 mM; (d) the change in the open probability of the channel caused by extracellular or intracellular Ba2+ is a consequence of alterations in both the opening and closing rate constants. Extracellular and intracellular divalent cations can modify the gating kinetics of Na+ channels by a specific modulatory effect that is independent of diffuse surface potentials. External or internal divalent cations probably bind to specific charges on the Na+ channel glycoprotein that modulate channel gating.  相似文献   

17.
The gating charge and voltage dependence of the open state to the inactivated state (O-->I) transition was measured for the voltage- dependent mammalian cardiac Na channel. Using the site 3 toxin, Anthopleurin-A (Ap-A), which selectively modifies the O-->I transition (see Hanck, D. A., and M. F. Sheets. 1995. Journal of General Physiology. 106:601-616), we studied Na channel gating currents (Ig) in voltage-clamped single canine cardiac Purkinje cells at approximately 12 degrees C. Comparison of Ig recorded in response to step depolarizations before and after modification by Ap-A toxin showed that toxin-modified gating currents decayed faster and had decreased initial amplitudes. The predominate change in the charge-voltage (Q-V) relationship was a reduction in gating charge at positive potentials such that Qmax was reduced by 33%, and the difference between charge measured in Ap-A toxin and in control represented the gating charge associated with Na channels undergoing inactivation by O-->I. By comparing the time course of channel activation (represented by the gating charge measured in Ap-A toxin) and gating charge associated with the O-->I transition (difference between control and Ap-A charge), the influence of activation on the time course of inactivation could be accounted for and the inherent voltage dependence of the O-->I transition determined. The O-->I transition for cardiac Na channels had a valence of 0.75 e-. The total charge of the cardiac voltage-gated Na channel was estimated to be 5 e-. Because charge is concentrated near the opening transition for this isoform of the channel, the time constant of the O-->I transition at 0 mV could also be estimated (0.53 ms, approximately 12 degrees C). Prediction of the mean channel open time-voltage relationship based upon the magnitude and valence of the O- ->C and O-->I rate constants from INa and Ig data matched data previously reported from single Na channel studies in heart at the same temperature.  相似文献   

18.
S Zhang  S J Kehl    D Fedida 《Biophysical journal》2001,81(1):125-136
Zinc ions are known to induce a variable depolarizing shift of the ionic current half-activation potential and substantially slow the activation kinetics of most K(+) channels. In Kv1.5, Zn(2+) also reduces ionic current, and this is relieved by increasing the external K(+) or Cs(+) concentration. Here we have investigated the actions of Zn(2+) on the gating currents of Kv1.5 channels expressed in HEK cells. Zn(2+) shifted the midpoint of the charge-voltage (Q-V) curve substantially more (approximately 2 times) than it shifted the V(1/2) of the g-V curve, and this amounted to +60 mV at 1 mM Zn(2+). Both Q1 and Q2 activation charge components were similarly affected by Zn(2+), which indicated free access of Zn(2+) to channel closed states. The maximal charge movement was also reduced by 1 mM Zn(2+) by approximately 15%, from 1.6 +/- 0.5 to 1.4 +/- 0.47 pC (n = 4). Addition of external K(+) or Cs(+), which relieved the Zn(2+)-induced ionic current reduction, decreased the extent of the Zn(2+)-induced Q-V shift. In 135 mM extracellular Cs(+), 200 microM Zn(2+) reduced ionic current by only 8 +/- 1%, compared with 71% reduction in 0 mM extracellular Cs(+), and caused a comparable shift in both the g-V and Q-V relations (17.9 +/- 0.6 mV vs. 20.8 +/- 2.1 mV, n = 6). Our results confirm the presence of two independent binding sites involved in the Zn(2+) actions. Whereas binding to one site accounts for reduction of current and binding to the other site accounts for the gating shift in ionic current recordings, both sites contribute to the Zn(2+)-induced Q-V shift.  相似文献   

19.
Measurements were made of the kinetics and steady-state properties of the sodium conductance changes in the giant axon of the crab Carcinus maenas. The conductance measurements were made in the presence of small concentrations of tetrodotoxin and as much electrical compensation as possible in order to minimize errors caused by the series resistance. After an initial delay of 10-150 microsec, the conductance increase during depolarizing voltage clamp pulses followed the Hodgkin-Huxley kinetics. Values of the time constant for the activation of the sodium conductance lay on a bell-shaped curve with a maximum under 180 microsec at -40 mV (at 18 degrees C). Values of the time constant for the inactivation of the sodium conductance were also fitted using a bell-shaped curve with a maximum under 7 msec at -70 mV. The effects of membrane potential on the fraction of Na channels available for activation studied using double pulse protocols suggest that hyperpolarizing potentials more negative than -100 mV lock a fraction of the Na channels in a closed conformation.  相似文献   

20.
The effect of low pH on the kinetics of Na channel ionic and gating currents was studied in frog skeletal muscle fibers. Lowering external pH from 7.4 to 5.0 slows the time course of Na current consistent with about a +25-mV shift in the voltage dependence of activation and inactivation time constants. Similar shifts in voltage dependence adequately describe the effects of low pH on the tail current time constant (+23.3 mV) and the gating charge vs. voltage relationship (+22.1 mV). A significantly smaller shift of +13.3 mV described the effect of pH 5.0 solution on the voltage dependence of steady state inactivation. Changes in the time course of gating current at low pH were complex and could not be described as a shift in voltage dependence. tau g, the time constant that describes the time course of the major component of gating charge movement, was slowed in pH 5.0 solution by a factor of approximately 3.5 for potentials from -60 to +45 mV. We conclude that the effects of low pH on Na channel gating cannot be attributed simply to a change in surface potential. Therefore, although it may be appropriate to describe the effect of low pH on some Na channel kinetic properties as a "shift" in voltage dependence, it is not appropriate to interpret such shifts as a measure of changes in surface potential. The maximum gating charge elicited from a holding potential of -150 mV was little affected by low pH.(ABSTRACT TRUNCATED AT 250 WORDS)  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