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.
Shaker B potassium channels undergo rapid N-type and slow C-type inactivation. While N-type inactivation is supposed to be mediated by occlusion of the pore by the N-terminal protein structure, the molecular mechanisms leading to C-type inactivation are less well understood. Considering N-type inactivation as a model for a protein conformational transition, we investigated inactivation of heterologously expressed Shaker B potassium channels and mutants thereof, showing various degrees of C-type inactivation, under high hydrostatic (oil) pressure. In addition to the derived apparent activation and reaction volumes (ΔV), experiments at various temperatures yielded estimates for enthalpic (ΔH) and entropic (TΔS) contributions. N-type inactivation was accelerated by increasing temperature and slowed by high hydrostatic pressure yielding at equilibrium ΔH = 76 kJ/mole, TΔS = 82 kJ/mole, and ΔV = 0.18 nm3 indicating that the transition to the N-type inactivated state is accompanied by an increase in volume and a decrease in order. N-terminally deleted ShΔ6–46 constructs with additional mutations at either position 449 or 463 were used to investigate C-type inactivation. In particular at high temperatures, inactivation occurred in two phases indicating more than one process. At equilibrium the following values were estimated for the major inactivation component of mutant ShΔ6–46 T449A: ΔH = –64 kJ/mole, TΔS = –60 kJ/mole, and ΔV = –0.25 nm3, indicating that the C-type inactivated state occupies a smaller volume and is more ordered than the noninactivated state. Thus, hydrostatic pressure affects N- and C-type inactivation in opposite ways. Received: 17 May 1997 / Accepted: 18 July 1997  相似文献   

2.
In Kv1.5, protonation of histidine 463 in the S5-P linker (turret) increases the rate of depolarization-induced inactivation and decreases the peak current amplitude. In this study, we examined how amino acid substitutions that altered the physico-chemical properties of the side chain at position 463 affected slow inactivation and then used the substituted cysteine accessibility method (SCAM) to probe the turret region (E456-P468) to determine whether residue 463 was unique in its ability to modulate the macroscopic current. Substitutions at position 463 of small, neutral (H463G and H463A) or large, charged (H463R, H463K, and H463E) side groups accelerated inactivation and induced a dependency of the current amplitude on the external potassium concentration. When cysteine substitutions were made in the distal turret (T462C-P468C), modification with either the positively charged [2-(trimethylammonium)ethyl] methanethiosulfonate bromide (MTSET) or negatively charged sodium (2-sulfonatoethyl) methanethiosulfonate reagent irreversibly inhibited current. This inhibition could be antagonized either by the R487V mutation (homologous to T449V in Shaker) or by raising the external potassium concentration, suggesting that current inhibition by MTS reagents resulted from an enhancement of inactivation. These results imply that protonation of residue 463 does not modulate inactivation solely by an electrostatic interaction with residues near the pore mouth, as proposed by others, and that residue 463 is part of a group of residues within the Kv1.5 turret that can modulate P/C-type inactivation. electrophysiology; voltage-gated potassium channels; substituted cysteine accessibility method  相似文献   

3.
Several voltage-gated channels share a proline-valine-proline (proline hinge) sequence motif at the intracellular side of S6. We studied the proline hinge in Kv1.4 channels, which inactivate via two mechanisms: N- and C-type. We mutated the second proline to glycine or alanine: P558A, P558G. These mutations were studied in the presence/absence of the N-terminal to separate the effects of the interaction between the proline hinge and N- and C-type inactivation. Both S6 mutations slowed or removed N- and C-type inactivation, and altered recovery from inactivation. P558G slowed activation and N- and C-type inactivation by nearly an order of magnitude. Sensitivity to extracellular acidosis and intracellular quinidine binding remained, suggesting that transmembrane communication in N- and C-type inactivation was preserved, consistent with our previous findings of major structural rearrangements involving S6 during C-type inactivation. P558A was very disruptive: activation was slowed by more than an order of magnitude, and no inactivation was observed. These results are consistent with our hypothesis that the proline hinge and intracellular S6 movement play a significant role in inactivation and recovery. Computer modeling suggests that both P558G and P558A mutations modify early voltage-dependent steps and make a final voltage-insensitive step that is rate limiting at positive potentials.  相似文献   

4.
We studied the effect of H(2)O(2) on the gating behavior of large-conductance Ca(2+)-sensitive voltage-dependent K(+) (K(V,Ca)) channels. We recorded potassium currents from single skeletal muscle channels incorporated into bilayers or using macropatches of Xenopus laevis oocytes membranes expressing the human Slowpoke (hSlo) alpha-subunit. Exposure of the intracellular side of K(V,Ca) channels to H(2)O(2) (4-23 mM) leads to a time-dependent decrease of the open probability (P(o)) without affecting the unitary conductance. H(2)O(2) did not affect channel activity when added to the extracellular side. These results provide evidence for an intracellular site(s) of H(2)O(2) action. Desferrioxamine (60 microM) and cysteine (1 mM) completely inhibited the effect of H(2)O(2), indicating that the decrease in P(o) was mediated by hydroxyl radicals. The reducing agent dithiothreitol (DTT) could not fully reverse the effect of H(2)O(2). However, DTT did completely reverse the decrease in P(o) induced by the oxidizing agent 5,5'-dithio-bis-(2-nitrobenzoic acid). The incomplete recovery of K(V,Ca) channel activity promoted by DTT suggests that H(2)O(2) treatment must be modifying other amino acid residues, e.g., as methionine or tryptophan, besides cysteine. Noise analysis of macroscopic currents in Xenopus oocytes expressing hSlo channels showed that H(2)O(2) induced a decrease in current mediated by a decrease both in the number of active channels and P(o).  相似文献   

5.
In K(+) channels, rearrangements of the pore outer vestibule have been associated with C-type inactivation gating. Paradoxically, the crystal structure of Open/C-type inactivated KcsA suggests these movements to be modest in magnitude. In this study, we show that under physiological conditions, the KcsA outer vestibule undergoes relatively large dynamic rearrangements upon inactivation. External Cd(2+) enhances the rate of C-type inactivation in an cysteine mutant (Y82C) via metal-bridge formation. This effect is not present in a non-inactivating mutant (E71A/Y82C). Tandem dimer and tandem tetramer constructs of equivalent cysteine mutants in KcsA and Shaker K(+) channels demonstrate that these Cd(2+) metal bridges are formed only between adjacent subunits. This is well supported by molecular dynamics simulations. Based on the crystal structure of Cd(2+)-bound Y82C-KcsA in the closed state, together with electron paramagnetic resonance distance measurements in the KcsA outer vestibule, we suggest that subunits must dynamically come in close proximity as the channels undergo inactivation.  相似文献   

6.
Kv4 channels mediate the somatodendritic A-type K+ current (I(SA)) in neurons. The availability of functional Kv4 channels is dynamically regulated by the membrane potential such that subthreshold depolarizations render Kv4 channels unavailable. The underlying process involves inactivation from closed states along the main activation pathway. Although classical inactivation mechanisms such as N- and P/C-type inactivation have been excluded, a clear understanding of closed-state inactivation in Kv4 channels has remained elusive. This is in part due to the lack of crucial information about the interactions between gating charge (Q) movement, activation, and inactivation. To overcome this limitation, we engineered a charybdotoxin (CTX)-sensitive Kv4.2 channel, which enabled us to obtain the first measurements of Kv4.2 gating currents after blocking K+ conduction with CTX (Dougherty and Covarrubias. 2006J. Gen. Physiol. 128:745-753). Here, we exploited this approach further to investigate the mechanism that links closed-state inactivation to slow Q-immobilization in Kv4 channels. The main observations revealed profound Q-immobilization at steady-state over a range of hyperpolarized voltages (-110 to -75 mV). Depolarization in this range moves <5% of the observable Q associated with activation and is insufficient to open the channels significantly. The kinetics and voltage dependence of Q-immobilization and ionic current inactivation between -153 and -47 mV are similar and independent of the channel's proximal N-terminal region (residues 2-40). A coupled state diagram of closed-state inactivation with a quasi-absorbing inactivated state explained the results from ionic and gating current experiments globally. We conclude that Q-immobilization and closed-state inactivation at hyperpolarized voltages are two manifestations of the same process in Kv4.2 channels, and propose that inactivation in the absence of N- and P/C-type mechanisms involves desensitization to voltage resulting from a slow conformational change of the voltage sensors, which renders the channel's main activation gate reluctant to open.  相似文献   

7.
In response to a prolonged membrane depolarization, inactivation autoregulates the activity of voltage-gated ion channels. Slow inactivation involving a localized constriction of the selectivity filter (P/C-type mechanism) is prevalent in many voltage-gated K+ channels of the Kv1 subfamily. However, the generalization of this mechanism to other Kv channel subfamilies has remained uncertain and controversial. In agreement with a “foot-in-the-door” mechanism and the presence of ion-ion interactions in the pore, elevated external K+ slows the development of P/C-type inactivation and accelerates its recovery. In sharp contrast and resembling the regulation of the hippocampal A-type K+ current, we found that Kv4.x channels associated with KChIP-1 (an auxiliary subunit) exhibit accelerated inactivation and unaffected recovery from inactivation when exposed to elevated external K+. This regulation depends on the ability of a permeant ion to enter the selectivity filter (K+ = Rb+ = NH4+ > Cs+ > Na+); and the apparent equilibrium dissociation constant of a single regulatory site is 8 mM for K+. By applying a robust quantitative global kinetic modeling approach to all macroscopic properties over a 210-mV range of membrane potentials, we determined that elevated external K+ inhibits unstable closed states outside the main activation pathway and thereby promotes preferential closed-state inactivation. These results suggest the presence of a vestigial and unstable P/C-type mechanism of inactivation in Kv4 channels and strengthen the concept of novel mechanisms of closed-state inactivation. Regulation of Kv4 channel inactivation by hyperkalemia may help to explain the pathophysiology of electrolyte imbalances in excitable tissues.  相似文献   

8.
pH-dependent modulation of Kv1.3 inactivation: role of His399   总被引:2,自引:0,他引:2  
The Kv1.3 K+ channel lacks N-type inactivation, but during prolonged depolarized periods it inactivates via the slow (P/C type) mechanism. It bears a titratable histidine residue in position 399 (equivalent of Shaker 449), a site known to influence the rate of slow inactivation. As opposed to several other voltage-gated K+ channels, slow inactivation of Kv1.3 is slowed when extracellular pH (pHo) is lowered under physiological conditions. Our findings are as follows. First, when His399 was mutated to a lysine, arginine, leucine, valine or tyrosine, extracellular acidification (pH 5.5) accelerated inactivation reminiscent of other Kv channels. Second, inactivation of the wild-type channel was accelerated by low pHo when the ionic strength of the external solution was raised. Inactivation of the H399K mutant was also accelerated by high ionic strength at pH 7.35 but not the inactivation of H399L. Third, after the external application of blocking barium ions, recovery of the wild-type current during washout was slower in low pHo. Fourth, the dissociation rate of Ba2+ was pH insensitive for both H399K and H399L. Furthermore, Ba2+ dissociation rates were equal for H399K and the wild type at pH 5.5 and were equal for H399L and the wild type at pH 7.35. These observations support a model in which the electric field of the protonated histidines creates a potential barrier for potassium ions just outside the external mouth of the pore that hinders their exit from the binding site controlling inactivation. In Kv1.3, this effect overrides the generally observed speeding of slow inactivation when pHo is reduced. extracellular pH; potassium channel; histidine; barium; high ionic strength  相似文献   

9.
Kv1.4 channels are Shaker-related voltage-gated potassium channels with two distinct inactivation mechanisms. Fast N-type inactivation operates by a ball-and-chain mechanism. Slower C-type inactivation is not so well defined, but involves intracellular and extracellular conformational changes of the channel. We studied the interaction between inactivation mechanisms using two-electrode voltage-clamp of Kv1.4 and Kv1.4ΔN (amino acids 2–146 deleted to remove N-type inactivation) heterologously expressed in Xenopus oocytes. We manipulated C-type inactivation by introducing a lysine-tyrosine point mutation (K532Y, equivalent to Shaker T449Y) that diminishes C-type inactivation. We used experimental data to develop a comprehensive computer model of Kv1.4 channels to determine the interaction between activation and N- and C-type inactivation mechanisms needed to replicate the experimental data. C-type inactivation began at lower voltage preactivated states, whereas N-type inactivation was coupled directly to the open state. A model with distinct N- and C-type inactivated states was not able to reproduce experimental data, and direct transitions between N- and C-type inactivated states were required, i.e., there is coupling between N- and C-type inactivated states. C-type inactivation is the rate-limiting step determining recovery from inactivation, so understanding C-type inactivation, and how it is coupled to N-type inactivation, is critical in understanding how channels act to repetitive stimulation.  相似文献   

10.
Kv3.3 K+ channels are believed to incorporate an NH2-terminal domain to produce an intermediate rate of inactivation relative to the fast inactivating K+ channels Kv3.4 and Kv1.4. The rate of Kv3.3 inactivation has, however, been difficult to establish given problems in obtaining consistent rates of inactivation in expression systems. This study characterized the properties of AptKv3.3, the teleost homologue of Kv3.3, when expressed in Chinese hamster ovary (CHO) or human embryonic kidney (HEK) cells. We show that the properties of AptKv3.3 differ significantly between CHO and HEK cells, with the largest difference occurring in the rate and voltage dependence of inactivation. While AptKv3.3 in CHO cells showed a fast and voltage-dependent rate of inactivation consistent with N-type inactivation, currents in HEK cells showed rates of inactivation that were voltage-independent and more consistent with a slower C-type inactivation. Examination of the mRNA sequence revealed that the first methionine start site had a weak Kozak consensus sequence, suggesting that the lack of inactivation in HEK cells could be due to translation at a second methionine start site downstream of the NH2-terminal coding region. Mutating the nucleotide sequence surrounding the first methionine start site to one more closely resembling a Kozak consensus sequence produced currents that inactivated with a fast and voltage-dependent rate of inactivation in both CHO and HEK cells. These results indicate that under the appropriate conditions Kv3.3 channels can exhibit fast and reliable inactivation that approaches that more typically expected of "A"-type K+ currents.  相似文献   

11.
Previous studies demonstrated that slow inactivation of the Shaker potassium channel can be made ~100-fold faster or slower by point mutations at a site in the outer pore (T449). However, the discovery that two forms of slow inactivation coexist in Shaker raises the question of which inactivation process is affected by mutation. Equivalent mutations in KV2.1, a channel exhibiting only U-type inactivation, have minimal effects on inactivation, suggesting that mutation of Shaker T449 acts on C-type inactivation alone, a widely held yet untested hypothesis. This study reexamines mutations at Shaker T449, confirming that T449A speeds inactivation and T449Y/V slow it. T449Y and T449V exhibit U-type inactivation that is enhanced by high extracellular potassium, in contrast to C-type inactivation in T449A which is inhibited by high potassium. Automated parameter estimation for a 12-state Markov model suggests that U-type inactivation occurs mainly from closed states upon weak depolarization, but primarily from the open state at positive voltages. The model also suggests that WT channels, which in this study exhibit mostly C-type inactivation, recover from inactivation through closed-inactivated states, producing voltage-dependent recovery. This suggests that both C-type and U-type inactivation involve both open-inactivated and closed-inactivated states.  相似文献   

12.
KCNQ1 inactivation bears electrophysiological characteristics different from classical N- and C-type inactivation in Shaker-like potassium channels. However, the molecular site of KCNQ1 inactivation has not yet been determined. KCNQ2 channels do not exert a fast inactivation in contrast to KCNQ1 channels. By expressing functional chimeras between KCNQ1 and KCNQ2 in Xenopus oocytes, we mapped the region of this inactivation to transmembrane domain S5 and the pore loop H5 and finally narrowed down the site to positions Gly(272) and Val(307) in KCNQ1. Exchanging these two amino acids individually with the analogous KCNQ2 residue abolished inactivation. Furthermore, a KCNQ1-like inactivation was introduced into KCNQ2 by mutagenesis in the corresponding region, confirming its relevance for the inactivation process. As KCNQ1 inactivation involves the regions S5 and H5, it exhibits a geography distinct from N- or C-type inactivation. Native cardiac I(Ks) channels comprising KCNQ1 and accessory MinK subunits do not inactivate because of the functional interaction of KCNQ1 with MinK. Mutations in KCNQ1 can lead to long QT1 syndrome, an inherited form of arrhythmia. The long QT1 mutant KCNQ1(L273F) displays a pronounced KCNQ1 inactivation. Here we show that when expressing mutant I(Ks) channels formed from KCNQ1(L273F) and MinK, MinK association no longer eliminates KCNQ1 inactivation. This results in smaller repolarizing currents in the heart and therefore represents a novel mechanism leading to long QT syndrome.  相似文献   

13.
Lowering external pH reduces peak current and enhances current decay in Kv and Shaker-IR channels. Using voltage-clamp fluorimetry we directly determined the fate of Shaker-IR channels at low pH by measuring fluorescence emission from tetramethylrhodamine-5-maleimide attached to substituted cysteine residues in the voltage sensor domain (M356C to R362C) or S5-P linker (S424C). One aspect of the distal S3-S4 linker α-helix (A359C and R362C) reported a pH-induced acceleration of the slow phase of fluorescence quenching that represents P/C-type inactivation, but neither site reported a change in the total charge movement at low pH. Shaker S424C fluorescence demonstrated slow unquenching that also reflects channel inactivation and this too was accelerated at low pH. In addition, however, acidic pH caused a reversible loss of the fluorescence signal (pKa = 5.1) that paralleled the reduction of peak current amplitude (pKa = 5.2). Protons decreased single channel open probability, suggesting that the loss of fluorescence at low pH reflects a decreased channel availability that is responsible for the reduced macroscopic conductance. Inhibition of inactivation in Shaker S424C (by raising external K+ or the mutation T449V) prevented fluorescence loss at low pH, and the fluorescence report from closed Shaker ILT S424C channels implied that protons stabilized a W434F-like inactivated state. Furthermore, acidic pH changed the fluorescence amplitude (pKa = 5.9) in channels held continuously at −80 mV. This suggests that low pH stabilizes closed-inactivated states. Thus, fluorescence experiments suggest the major mechanism of pH-induced peak current reduction is inactivation of channels from closed states from which they can activate, but not open; this occurs in addition to acceleration of P/C-type inactivation from the open state.  相似文献   

14.
Ether-à-go-go (EAG) and EAG-related gene (ERG) K(+) channels are close homologues but differ markedly in their gating properties. ERG1 channels are characterized by rapid and extensive C-type inactivation, whereas mammalian EAG1 channels were previously considered noninactivating. Here, we show that human EAG1 channels exhibit an intrinsic voltage-dependent slow inactivation that is markedly enhanced in rate and extent by 1-10 μM 3-nitro-N-(4-phenoxyphenyl) benzamide, or ICA105574 (ICA). This compound was previously reported to have the opposite effect on ERG1 channels, causing an increase in current magnitude by inhibition of C-type inactivation. The voltage dependence of 2 μM ICA-induced inhibition of EAG1 current was half-maximal at -73 mV, 62 mV negative to the half-point for channel activation. This finding suggests that current inhibition by the drug is mediated by enhanced inactivation and not open-channel block, where the voltage half-points for current inhibition and channel activation are predicted to overlap, as we demonstrate for clofilium and astemizole. The mutation Y464A in the S6 segment also induced inactivation of EAG1, with a time course and voltage dependence similar to that caused by 2 μM ICA. Several Markov models were investigated to describe gating effects induced by multiple concentrations of the drug and the Y464A mutation. Models with the smallest fit error required both closed- and open-state inactivation. Unlike typical C-type inactivation, the rate of Y464A- and ICA-induced inactivation was not decreased by external tetraethylammonium or elevated [K(+)](e). EAG1 channel inactivation introduced by Y464A was prevented by additional mutation of a nearby residue located in the S5 segment (F359A) or pore helix (L434A), suggesting a tripartite molecular model where interactions between single residues in S5, S6, and the pore helix modulate inactivation of EAG1 channels.  相似文献   

15.
Mechanisms underlying Kv4 channel inactivation and recovery are presently unclear, although there is general consensus that the basic characteristics of these processes are not consistent with Shaker (Kv1) N- and P/C-type mechanisms. Kv4 channels also differ from Shaker in that they can undergo significant inactivation from pre-activated closed-states (closed-state inactivation, CSI), and that inactivation and recovery kinetics can be regulated by intracellular KChIP2 isoforms. To gain insight into the mechanisms regulating Kv4.3 CSI and recovery, we have analyzed the effects of increasing [K(+)](o) from 2 mM to 98 mM in the absence and in the presence of KChIP2b, the major KChIP2 isoform expressed in the mammalian ventricle. In the absence of KChIP2b, high [K(+)](o) promoted Kv4.3 inactivated closed-states and significantly slowed the kinetics of recovery from both macroscopic and closed-state inactivation. Coexpression of KChIP2b in 2 mM [K(+)](o) promoted non-inactivated closed-states and accelerated the kinetics of recovery from both macroscopic and CSI. In high [K(+)](o), KChIP2b eliminated or significantly reduced the slowing effects on recovery. Attenuation of CSI by the S4 charge-deletion mutant R302A, which produced significant stabilization of non-inactivated closed-states, effectively eliminated the opposing effects of high [K(+)](o) and KChIP2b on macroscopic recovery kinetics, confirming that these results were due to alterations of CSI. Elevated [K(+)](o) therefore slows Kv4.3 recovery by stabilizing inactivated closed-states, while KChIP2b accelerates recovery by destabilizing inactivated closed-states. Our results challenge underlying assumptions of presently popular Kv4 gating models and suggest that Kv4.3 possesses novel allosteric mechanisms, which are absent in Shaker, for coupling interactions between intracellular KChIP2b binding motifs and extracellular K(+)-sensitive regulatory sites.  相似文献   

16.
The fast-inactivation process in the hERG channel can be affected by mutations in the pore or S6 domain, similar to the C-type inactivation in the Shaker channel. However, differences in the kinetics and voltage dependence of inactivation between these two channels suggest that different structural determinants may be involved. To explore this possibility, we mutated a serine in the outer mouth region of hERG (S631) to residues of different physicochemical properties and compared the resulting changes in the channel's inactivation process with those resulting from mutations of an equivalent position in the Shaker channel (T449). The most dramatic differences are seen when this position is occupied by a charged residue: S631K and S631E disrupted C-type inactivation in hERG, whereas T449K and T449E facilitate C-type inactivation in Shaker. S631K and S631E also disrupted the K selectivity of hERG pore, a change not seen in T449K or T449E of Shaker. To further study why there are such differences, we replaced S631 with cysteine. This allowed us to manipulate the properties of thiol groups at position 631 and correlate side-chain properties here with changes in channel function. S631C behaved like the wild-type channel when the thiol groups were in the reduced state. Oxidizing thiol groups with H2O2 or modifying them with MTSET or MTSES disrupted C-type inactivation and K selectivity, similar to the phenotype of S631K and S631E. The same thiol-modifying maneuvers did not affect the wild-type channel function. Our results suggest differences in the outer mouth structure between hERG and Shaker, and we propose a "molecular spring" hypothesis to explain these differences.  相似文献   

17.
Di-substituted cyclohexyl (DSC) derivatives inhibit the voltage-gated potassium channel, K(v)1.3, and have immunosuppressant activity (Schmalhofer et al. (2002) Biochemistry 41, 7781-7794). This class of inhibitors displays Hill coefficients of near 2 in functional assays, and trans DSC analogues appear to selectively interact with K(v)1.3 channel conformations related to C-type inactivation. To further understand the details of the DSC inhibitor interaction with potassium channels, trans-1-(N-n-propylcarbamoyloxy)-4-phenyl-4-(3-(2-methoxyphenyl)-3-oxo-2-azaprop-1-yl)cyclo-hexane (trans-NPCO-DSC) was radiolabeled with tritium, and its binding characteristics to K(v)1.3 channels were determined. Specific binding of [(3)H]-trans-NPCO-DSC to K(v)1.3 channels is a saturable, time-dependent, and fully reversible process. Saturation binding isotherms and competition binding experiments are consistent with the presence of two receptor sites for DSC derivatives on the K(v)1.3 channel that display positive allosteric cooperativity. The high affinity interaction of [(3)H]-trans-NPCO-DSC with K(v)1.3 channels appears to correlate with the rates of C-type inactivation of the channel. These data, taken together, mark the first demonstration of the existence of multiple binding sites for an inhibitor of an ion channel and suggest that the high affinity interaction of trans-NPCO-DSC and similar inhibitors with K(v)1.3 channels could be exploited for the development of selective molecules that target this protein.  相似文献   

18.
Kv1.4 encodes a slowly recovering transient outward current (I(to)), which inactivates by a fast N-type (intracellular ball and chain) mechanism but has slow recovery due to C-type inactivation. C-type inactivation of the NH(2)-terminal deletion mutant (fKv1.4DeltaN) was inhibited by 98 mM extracellular K(+) concentration ([K(+)](o)), whereas N-type was unaffected. In 98 mM [K(+)](o), removal of intracellular K(+) concentration ([K(+)](i)) speeded C-type inactivation but had no effect on N-type inactivation, suggesting that C-type inactivation is sensitive to K(+) binding to intracellular sites. C-type inactivation is thought to involve closure of the extracellular pore mouth. However, a valine to alanine mutation on the intracellular side of S6 (V561A) of fKv1.4DeltaN alters recovery and results in anomalous speeding of C-type inactivation with increasing [K(+)](o). Extracellular pH (pH(o)) modulated both N- and C-type inactivation through an S5-H5 linker histidine (H508) with acidosis speeding both N- and C-type inactivation. Mutation of an extracellular lysine to a tyrosine (K532Y) slowed C-type inactivation and inhibited the pH dependence of both N- and C-type inactivation. These results suggest that mutations, [K(+)], and pH modulate inactivation through membrane-spanning mechanisms involving S6.  相似文献   

19.
Mechanisms underlying Kv4 channel inactivation and recovery are presently unclear, although there is general consensus that the basic characteristics of these processes are not consistent with Shaker (Kv1) N- and P/C-type mechanisms. Kv4 channels also differ from Shaker in that they can undergo significant inactivation from pre-activated closed-states (closed-state inactivation, CSI), and that inactivation and recovery kinetics can be regulated by intracellular KChIP2 isoforms. To gain insight into the mechanisms regulating Kv4.3 CSI and recovery, we have analyzed the effects of increasing [K+]o from 2 mM to 98mM in the absence and in the presence of KChIP2b, the major KChIP2 isoform expressed in the mammalian ventricle. In the absence of KChIP2b, high [K+]o promoted Kv4.3 inactivated closed-states and significantly slowed the kinetics of recovery from both macroscopic and closed-state inactivation. Coexpression of KChIP2b in 2 mM [K+]o promoted non-inactivated closed-states and accelerated the kinetics of recovery from both macroscopic and CSI. In high [K+]o, KChIP2b eliminated or significantly reduced the slowing effects on recovery. Attenuation of CSI by the S4 charge-deletion mutant R302A, which produced significant stabilization of non-inactivated closed-states, effectively eliminated the opposing effects of high [K+]o and KChiP2b on macroscopic recovery kinetics, confirming that these results were due to alterations of CSI. Elevated [K+]o therefore slows Kv4.3 recovery by stabilizing inactivated closed-states, while KChIP2b accelerates recovery by destabilizing inactivated closed-states. Our results challenge underlying assumptions of presently popular Kv4 gating models and suggest that Kv4.3 possesses novel allosteric mechanisms, which are absent in Shaker, for coupling interactions between intracellular KChIP2b binding motifs and extracellular K+-sensitive regulatory sites.  相似文献   

20.
N R Matheson  J Travis 《Biochemistry》1985,24(8):1941-1945
Human alpha 1-proteinase inhibitor is easily susceptible to inactivation because of the presence of a methionyl residue at its reactive site. Thus, oxidizing species derived from the myeloperoxidase system (enzyme, H2O2, and C1-), as well as hypochlorous acid, can inactivate this inhibitor, although H2O2 alone has no effect. Butylated hydroxytoluene, a radical scavenger, partially protects alpha 1-proteinase inhibitor from the myeloperoxidase system and completely protects it from hypochlorous acid. Each oxidant also reacts differently with the inhibitor, in that the myeloperoxidase system and hypochlorous acid can each oxidize as many as six methionyl residues, but hypochlorous acid can also oxidize a single tyrosine residue. Myeloperoxidase can be inactivated by hypochlorous acid, by autoxidation in the presence of H2O2 and C1-, as well as by H2O2 alone. Butylated hydroxytoluene completely protects this enzyme from hypochlorous acid inactivation, does not affect the action of H2O2, and enhances autoinactivation. As many as six methionyl residues and two tyrosine residues could be oxidized during autoxidation and six methionine residues by H2O2 alone. Eight methionine residues and one tyrosine residue could be oxidized by hypochlorous acid. The tyrosine residue in myeloperoxidase was oxidized only at a relatively high concentration (600 microM) of hypochlorous acid at which point the enzyme simultaneously and completely lost its enzymatic activity. Loss of activity of myeloperoxidase could also be correlated with the loss of the heme groups present in the enzyme when a relatively high concentration of hypochlorous acid (600 microM) was used and also during autoxidation. It appears that once there is sufficient oxidant to modify one of the tyrosine residues, the heme group itself becomes susceptible.  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