当前位置: 首页 > 医学版 > 期刊论文 > 基础医学 > 生理学进展 > 2005年 > 第4期 > 正文
编号:11256185
Sodium Channel Inactivation: Molecular Determinants and Modulation
http://www.100md.com 生理学进展 2005年第4期
     Physiologisches Institut, University of Kiel, Kiel, Germany

    ABSTRACT

    Voltage-gated sodium channels open (activate) when the membrane is depolarized and close on repolarization (deactivate) but also on continuing depolarization by a process termed inactivation, which leaves the channel refractory, i.e., unable to open again for a period of time. In the "classical" fast inactivation, this time is of the millisecond range, but it can last much longer (up to seconds) in a different slow type of inactivation. These two types of inactivation have different mechanisms located in different parts of the channel molecule: the fast inactivation at the cytoplasmic pore opening which can be closed by a hinged lid, the slow inactivation in other parts involving conformational changes of the pore. Fast inactivation is highly vulnerable and affected by many chemical agents, toxins, and proteolytic enzymes but also by the presence of -subunits of the channel molecule. Systematic studies of these modulating factors and of the effects of point mutations (experimental and in hereditary diseases) in the channel molecule have yielded a fairly consistent picture of the molecular background of fast inactivation, which for the slow inactivation is still lacking.

    I. INTRODUCTION

    A. Scope

    The subject of this article is the voltage-gated sodium channel that plays a key role in membrane excitation and is dealt with in an enormous number of papers. Inactivation (defined in sect. IIA) of this channel appears to be its most vulnerable kinetic feature as it is influenced, mostly slowed or abolished, by all kinds of chemical agents such as drugs, toxins, or mutations, often of only a single amino acid residue of the channel molecule. The amino acid sequence of the sodium channel molecule and its transmembrane topology have been known for quite some time. Nevertheless, relating function to structural details is still unsatisfactory, but site-directed mutations are increasingly employed to solve such questions. The molecular exploration of hereditary diseases has helped much to identify relevant regions.

    Papers dealing with sodium channels, even those that are limited to inactivation, are so numerous that often only summaries, in particular of older work, can be quoted. Reviews of more recent papers on structure and function of sodium channels have been written by Catterall (67, 68), Denac et al. (106), Fozzard and Hanck (130), Marban et al. (272), and Ogata and Ohishi (331); extensive reviews of kinetic aspects of inactivation by Patlak (341) and of mechanisms by Goldin (146) have appeared. Highly readable accounts are found in books on ion channels in general (5, 181). Other reviews will be quoted in several other sections.

    In this paper inactivation is first described as observed in electrophysiological experiments including patch-clamp studies. These are followed by interpretations by kinetic models and gating current experiments. Then the structure will be dealt with on a molecular level as derived from experimental point mutations. Next chemical modulation will be described and finally genetic modulation, observed in hereditary diseases (channelopathies).

    B. Terminology

    There exist many isoforms of the sodium channel (144) that have been differently termed. To eliminate the confusion a panel of leading researchers has agreed on a uniform nomenclature following that for potassium channels (147). These terms will be given, at least once, in parentheses to those used in the original papers.

    II. PHENOMENA

    A. Time Course of (Fast) Inactivation

    The typical voltage-gated sodium channel opens on depolarization and closes rapidly on repolarization or, more slowly, on sustained depolarization. The latter process is termed inactivation and leaves the channel refractory for some time after repolarization. In the classical study of the squid giant axon, Hodgkin and Huxley (184, 185) described the sodium conductance (channel gating) being proportional to m3 · h, where m and h are variables of time and potential that can assume values between 0 and 1: m increases and h decreases on depolarization. This leads to a nonmonotonic time course with 1 eC h expressing inactivation whereby h develops exponentially as h(t) = h eC (h eC ho) exp (-t/h) with ho and h the starting and final value of h, respectively, and h the inactivation time constant. For large depolarizations h becomes 0, meaning that inactivation reaches completion.

    Changes in methods led to results that necessitated different formalisms. Thus in squid axons internally perfused with NaF solution, Chandler and Meves (75) found inactivation to be incomplete (h = 0.1) and at very positive membrane potentials h increasing to 0.3 or more. These authors explained their results by assuming that h is the sum of two components, h = h1 + h2, with the components connected through an inactive (closed) state. Only during very long depolarizing pulses lasting for several seconds did the "maintained" current inactivate, which Chandler and Meves (76) described by an additional variable s so that h = (h1 + h2)s. Such slow inactivation of varying time constants is observed in many preparations as will be shown below; in KF-perfused squid axons, it gives rise to action potentials lasting for seconds. On blocking IK, a noninactivating fraction of INa was later found also in nonperfused squid axons (412). It was further studied at very positive membrane potentials and with an inverse Na+ gradient (92). The voltage dependence of this fraction has again been examined in more detail recently (359), but the conclusions have been questioned (88).

    In amphibian myelinated nerve fibers, inactivation was originally studied with short impulses and could be described by a monoexponential process (131). When in this preparation the potassium current was eliminated with tetraethylammonium ions (TEA), or by other means, the sodium current could be followed for any length of time. Inactivation then turned out to be diphasic (see Fig. 4, trace "toxin free") and so was recovery from inactivation (84) which was interpreted by a second inactivating state in series. Several authors observed diphasic inactivation (see Ref. 230), but some interpreted it differently, recognizing that recovery from inactivation posed the more difficult part of description. Ochs et al. (330) tested several three-state models and found their results to be best fit with two open states connected through a closed one. To account for biphasic tail currents on repolarization, Elinder and rhem (119) suggested another model with two open states leading to two different inactivated states. Schmidtmayer (394) formulated a cyclic three-state model with one open and two closed inactivated states. Whereas these models were based on the idea of a single population of sodium channels, Benoit et al. (35) assumed that there are two, however interconvertible, types of sodium channels that differ not only in their inactivation kinetics but also in their susceptibility to blocking agents.

    In the papers mentioned so far, it was tacitly assumed, as implied by the m · h formalism, that inactivation proceeds independently from activation. This was questioned on the grounds of experiments on Myxicola axons. If in this preparation steady-state inactivation was determined with the classical protocol of a conditioning pulse of varying amplitude followed by a test pulse of constant amplitude (184), the position of the observed h(V) curve was clearly shifted to more negative potentials for weaker test pulses (151). Such shift had been predicted by Hoyt (197) for activation-inactivation coupling. Another point in favor of coupling was seen in a delayed onset of inactivation as determined with a constant test pulse following a conditioning pulse of constant amplitude but varying duration. Such delay was reported for Myxicola axons (151) and squid axons (45), but in the latter preparation, the delay could be minimized by leaving a gap between conditioning and test pulse to allow activation to subside (140). This does not seem to apply to Myxicola axons in which introduction of a gap did not eliminate the delay (148, 149, 150) nor did it in neuronal and cardiac channels (42). Kniffki et al. (230) showed that in toad nodes of Ranvier the initial delay is due to the activation during the conditioning pulse and implicit in the classical m · h description. Thus the situation is complicated, moreover since the kinetic experiments are prone to flaws in method, some of which were reviewed by Meves (282).

    The question of whether or not inactivation is coupled to activation is of course essential for understanding the mechanism of inactivation. A great number of kinetic models have been proposed that attempt to reconcile results on bulk sodium currents with so-called gating currents (see sect. IIIB) or single-channel recordings (see sect. IID).

    B. Differences Between Channel Isoforms

    Activation and inactivation characteristics like Vm, the potential at which activation reaches half-maximal values, and Vh, the potential at which h = 0.5, differ in different isoforms, but also among species and, with cloned channels, depend on the cells in which they are expressed (269; see below). Thus, in Na+ channels of human heart (hH1 = Nav1.5), skeletal muscle (hSkM1 = Nav1.4), and rat brain (rIIA = Nav1.2), expressed in mammalian tsA201 cells, Vm = eC48, eC28, and eC22 mV and Vh = eC92, eC72, and eC61 mV, respectively. The kinetics of activation and inactivation are also quite different. Thus t1/2, the time to reach half of their maximal current amplitude, values are 0.82, 0.48, and 0.40 ms for the three isoforms at 20°C (Ref. 333, containing ample references). The time constants of the rapid component of inactivation decrease with depolarization and reach asymptotic values at V >20 mV of 0.41 ms (hH1), 0.26 ms (hSkM1), and 0.27 ms (rIIA). Clearly at positive potentials, where channels inactivate mostly from the open state, heart channels do so more slowly. Recovery from inactivation is also quite different: rec, measured at V = eC100 mV, was 44.7 ms (hH1), 4.7 ms (hSkM1), and 7.6 ms (rIIA). Extensive descriptions of these isoforms (70) and their evolution (142) have recently been compiled.

    C. Slow Inactivation

    After prolonged depolarization, often achieved by changes of holding potential in the voltage clamp, recovery from inactivation may proceed very slowly with time constants in the second to minute range and in several phases. Such changes have been termed "slow" or "ultraslow" inactivation and have been reported for amphibian myelinated nerve fibers (52, 129, 238, 318, 458), frog muscle (12), frog ventricular myocytes (132), and rat muscle (382, 415). Slow changes in INa following changes in membrane potential have also been observed in lobster, crayfish (378), Myxicola (392), and squid (1, 75, 76, 287) axons, in the latter preparation even after block of fast inactivation by intra-axonal application of pronase (379).

    Slow inactivation was also observed in mammalian neuronal cells, such as rat hippocampal neurons (288), tetrodotoxin-resistant cells of dorsal root ganglia of rat (332), or cultivated neuroblastoma cells (353). In human cardiac muscle (471), slow inactivation is present but is only 40% complete compared with 80% observed in channels of human skeletal muscle (367).

    Fast inactivation is important for action potential repolarization, and in mammalian nodes of Ranvier, which almost lack phasic potassium channels, it is the only repolarizing force besides the leakage current (85). Slow inactivation, on the other hand, may play a role in regulating excitability (383), such as by modulating burst discharges as can be demonstrated by computations (120). In reality, however, this modulation appears to be complicated since slow inactivation not only depends on resting potential but also on previous history of action potential firing (288, 448). Also, persistent INa of tetrodotoxin (TTX)-resistant channels (100) may affect the resting potential as suggested by computer simulations (175).

    D. Single-Channel Results

    Single-channel records of normal sodium channels were preferentially obtained from cell-attached neuroblastoma and myocardial cells. They show one or a few short openings at the first 10eC20 ms (at room temperature) after the start of a depolarizing impulse which, if kept on, leads the channels in an inactivated state. An early thorough analysis of single neuroblastoma channels revealed a fast "microscopic" inactivation from the open state even in the case of slow "macroscopic" inactivation as expressed by h at smaller depolarizations where a slower activation is rate limiting (6, 7). The underlying scheme predicts that removal of inactivation would unmask the slow activation reaching its peak much later than INa of untreated cells. This was indeed observed in neuroblastoma cells intracellularly treated with papain (153) but seems to be the exception rather than the rule, since other channel subtypes show a fast activation combined with relatively slow inactivation.

    In some heart cells, openings lasting clearly longer than in neuronal cells were observed that correspond to a slow inactivating state as illustrated by ensemble-averaged records; even a transition from fast to slowly inactivating state was observed as well as a persistent state (48, 49). In rat skeletal muscle, slow inactivation manifests itself as a decreasing number of channels of unchanged open time and conductance (380).

    An extensive single-channel study has also been done on cells of the rat entorhinal cortex which generate a persistent current caused by prolonged and often delayed bursts (267). Open times within these bursts (but not interburst closed times) were strongly voltage dependent. In neuroblastoma cells, multiple, longer, and late openings were seen only in channels modified by batrachotoxin (354), sea anemone toxin, scorpion toxin, or chloramine-T (226, 303, 308, 323). Sea anemone toxin prolonged the open state (including repeated openings) in cardiomyocytes (121, 402) as well as in cloned myocardial channels (hH1 = Nav1.5; Ref. 74). In a comparative study on single cardiac and neuronal channels, modified by sea anemone toxin, initial clusters of multiple openings and periodic subsequent reopenings were observed in either preparation. However, toxin induced a larger persistent current in neuronal than in cardiac channels (42).

    Similar single-channel phenomena have been observed in mutant channels with slowed inactivation (280) whereby particularly delayed inactivation is accompanied with very long mean open times (244, 417). Mutant channels completely deficient of inactivation (IFM/QQQ; see sect. IVB) showed repeated openings and closings throughout long depolarizing pulses (159). In almost all these experiments, single-channel conductance was unaffected by modifications of kinetics. Analysis, however, may be complicated by the existence of conductance substrates and changes following excisions of membrane patches (see Ref. 309).

    E. Temperature Effects

    In their classical description of ion channel kinetics in squid axons, Hodgkin and Huxley (185) assumed a temperature coefficient, Q10, of inactivation of 3. Similar results have since been obtained with many other preparations (181). Hence, in comparing time constants it is necessary to consider the temperature at which they were measured. Usually the Q10 values of the time constant(s) increase at lower temperatures as one would expect if kinetics are described by a simple Arrhenius plot of log h vs. TeC1. However, in some preparations such as skeletal muscle, a break in this plot is observed which may point to a transition in the lipid membrane phase (85, 405), complicating the interpretation in terms of activation enthalpy. In rat nerve, the steady-state inactivation h(V) was found to be shifted in the hyperpolarizing direction on cooling (404), so was h(V) in Xenopus nerve (207), whereas in rat muscle only the steady-state curve of slow inactivation s(V) was similarly shifted; this may explain the reduced availability of mammalian sodium channels if tested at room rather than at body temperature (381).

    III. INTERPRETATIONS

    A. Kinetic Models

    As already mentioned in section IIA, the classical description of sodium channel gating by Hodgkin and Huxley (184, 185) implied activation and inactivation to be independent processes, linked only by their dependence on membrane potential. However, later results such as the existence of more than one inactivation component, single-channel results, and the discovery of four nonidentical domains to form the sodium channel -subunit required extended models, competently reviewed by Hille (181). Most of these models propose that inactivation derives its voltage dependence from that of activation. However, often amino acid substitutions in the voltage sensors shift h(V) in the opposite direction compared with shifts in m(V), arguing against a strict activation-inactivation coupling (236).

    The general concept is that on depolarization a channel moves from (several) closed (C) resting state(s) through an open (O) state to one (or several) inactivated (I) state(s). Gating is assumed to be a Markov process meaning that the rate constants of transitions between these states are "oblivious" of how a given state is reached. A critical feature to model is the recovery from inactivation that does not seem to pass through the open state as no ionic current is usually seen during this period. An exception is observed in cerebellar neurons that produce a "resurgent" current on repolarization (2, 360). This current is assumed to reflect unblocking of an open-channel block in a portion of channels by a hypothetical particle during strong depolarization, constituting an additional mechanism of inactivation. It thus enhances recovery from inactivation, which enables these neurons to fire at high frequency (360). In most other channels the details of recovery require CI and IC steps leading to cyclic connections of states. Such cycles underlie the restraint of microscopic reversibility so that the product of rate constants going clockwise must equal the product going counterclockwise. In the following scheme 1 this would mean k34kOIkIC = k43kCIkIO.

    (1)

    In a study of various five-state models, that of the kind of scheme 1 yielded the best fit of single-channel results (see sect. IID) as tested with the maximum likelihood method (193), allowing inactivation both from closed and open states. Block by a synthetic inactivation gate peptide (see sect. IVB) could be described by such a model (280). Considering also gating current results (see sect. IIIB) led to extended models such as in scheme 2 with two inactivated states or even more complicated models. Details are found in the excellent review of Patlak (342).

    (2)

    Another model based on single-channel results has been proposed by Correa and Bezanilla (93) and allows for a second open state that is reached through an inactivated state; this scheme may underlie the classical description of Chandler and Meves (75) mentioned in section IIA.

    It should be noted that the classical Hodgkin-Huxley formulation m3 · h can be restated as a consecutive movement through eight states as in scheme 3 with k12 = 3m, k21 = m, k23 = 2m, k32 = 2m, k34 = m and k43 = 3m (180); to describe a number of single-channel results, it appears not to be inferior to a model of the type of scheme 1 (78).

    (3)

    A similar model involving five closed states has been proposed for hippocampal neurons (241) and was later supplemented with another open state to describe inactivation in heart channels and its temperature dependence (200).

    While the kinetic models described so far are based on voltage-clamp experiments that employ square voltage pulses, a different method called "nonequilibrium response spectroscopy" uses rapidly fluctuating (up to 14 kHz) potentials to extract more kinetic details (290).

    B. Gating Currents and Inactivation

    The potential-dependent gating of channels requires a voltage sensor bearing charges that move within the membrane as its potential changes, generating a tiny "gating current." This became measurable on channel opening and closing after suppression of the much larger ionic current (16, 221), whereas movement of the inactivation gate did not seem to contribute a specific fraction of gating current. Instead, inactivation causes a substantial fraction (2/3) to become immobilized, and immobilization proceeds and recovers with the same time course as does inactivation (17, 286, 287; for reviews, see Refs. 11, 15). This indicates that inactivation is coupled to activation from which it derives most of its voltage dependence (181). Nevertheless, some genuine voltage dependence of inactivation exists (194, 219, 220). The advantage of gating current measurements is to reveal kinetic steps that are not accompanied by changes of ionic currents.

    A great number of papers have been published dealing, one way or the other, with the relationship between inactivation and gating current, many of them in K+ channels. As for Na+ channels, more recent studies tried to reconcile gating currents with details of channel structure, in particular the existence of four domains, D1eCD4, each with six segments of which S4 acts as voltage sensor (see sect. IVA). Although segments S4 of all domains seem to be involved in activation (237), D4S4 clearly has the largest effect on inactivation (236), and mutation of central arginines has specific effects on inactivation and gating charge immobilization in rNav1.2 (239). S4 immobilization in D3 and D4 (but not in D1 or D2) has also been observed by site-directed fluorescence labeling (73).

    Another approach is to compare effects of modulating agents on ionic currents with those on gating currents (223). As one would expect, irreversibly removing inactivation in squid axons by chloramine-T (CT) or pronase also stops charge immobilization but with other similarly acting agents like batrachotoxin (BTX) immobilization remains (439). In amphibian nodes of Ranvier, comparable results were obtained with CT (111) but not with BTX, which eliminated immobilization (114) as did internally applied iodate (113). Comparative studies on nodes of Ranvier, also with site 3 toxins, have been summarized by Meves (284). In mammalian muscle channels (rNav1.4) and cardiac channels (hNav1.5), site 3 toxins of Anthopleura, ApA and ApB, reduce the maximal gating charge by about one-third, although muscle channels inactivate clearly faster than heart channels (407). Studies of ApA effects on gating charge in mutant hNav1.5 revealed the importance of the outermost arginine in D4S4 for charge immobilization (409) Charge immobilization has also been achieved in squid axons by intracellular treatment with the positively charged sulfhydryl reagents MTSET (2-trimethylammonioethylmethane thiosulfonate) and MTS-PTrEA {[3(triethylammonium)propyl]methanesulfonate}. These reagents irreversibly modify native cysteine(s) in a potential-dependent manner and promote inactivation from closed states, whereas they do not affect activation; MTS-PTrEA shifts the steady-state inactivation curve to more negative potential values and renders it much shallower (222).

    IV. MOLECULAR MECHANISMS

    A. Channel Structure and How to Relate It to Function

    1. Structure

    During the last 20 years the molecular analysis of the sodium channel has made enormous progress through experiments most recently described by Catterall (68). Here it may suffice to summarize our present knowledge. The channel purified from mammalian brain consists of the large -subunit (260 kDa) with the pore, the 1- (36 kDa) and 2-subunits (33 kDa) that contain extracellular immunoglobulin-like folds as illustrated by Figure 1 in section IVB. The -subunit consists of four domains (D1-D4) each with six transmembrane -helical segments (S1-S6) of which S4 bears several positive charges originating from arginine or lysine residues. The proteins of the domains wrap around a central pore such that the P-loops (SS1-SS2) between S5 and S6 form part of the pore lining.

    The positively charged residues of S4 are interspaced by two hydrophobic residues forming a spiral ribbon of positive charges on the -helix. The "classical" view has it that negatively charged residues in adjacent segments form ion pairs that stabilize S4 against the pull exerted at rest by the (inside) negative membrane potential. Depolarization releases S4, the voltage sensor, to move outward initiating the structural changes that open the pore (66, 166). This general concept has been varied and adapted to comply with new evidence, in particular gained from the three-dimensional structure of the bacterial potassium channel KcsA as reviewed by Bezanilla (43).

    Comparable results of sodium channels are soon to be expected as a recently described bacterial channel, NaChBac, may provide the protein in quantities needed for structural studies (365). A mutation, G219P, in S6 of this channel, dramatically reduces the rate of the already very slow deactivation and inactivation and shifts membrane potential (Vm) by eC51 mV (521). It seems that this mutation strongly favors bending of the S6 helix and that G219 serves as a gating hinge. A low-resolution three-dimensional structure study of a sodium channel (employing cryo-electron microscopy and image reconstruction) revealed, in addition to the ion-conducting central pore, four peripherally located transmembrane "gating pores" in which the S4 movement is thought to take place (69, 391).

    More recently, X-ray crystallography of another bacterial channel, KvAP, yielded results that led to a quite different interpretation of voltage sensor movement. S4 and part of S3 are supposed to form a "paddle" at the periphery of the channel which, on depolarization, moves like a lever through the surrounding membrane lipid from the intracellular to the extracellular side, thereby opening the pore (205, 206). This unconventional view, summarized by MacKinnon (265), is at present controversially discussed on the basis of new experimental results (4, 29, 53, 133, 247; see also comment in Ref. 190). Whatever the outcome, it certainly will affect our interpretation of sodium channel gating.

    2. Methods

    To derive function from structure, many methods have been employed. Early experiments revealed the sidedness of certain modulating agents or procedures giving valuable information as to where certain functions may be located. Thus, for instance, scorpion toxins act only from outside, pronase treatment only from inside the excitable cell to slow or delete inactivation. Comparable information was obtained with specific antibodies against portions of the channel (see sect. IVB). Decisive progress, however, was achieved by molecular genetics enabling site-directed mutagenesis. Mutants are expressed in Xenopus oocytes but most frequently in mammalian cells such as HEK293 (human embryonic kidney cells) and their derivative tsA201, also in Chinese hamster ovary cells (CHO) or in the Escherichia coli strain BL21. Very successful was replacing single amino acid residues by cysteine to which access could then be tested with sulfhydryl reagents ("substituted-cysteine-accessibility method", SCAM; Ref. 213), whereby access could change with the functional state. Experimental point mutations were employed in most of the papers cited in section IV, BeCD, whereas natural mutants found in hereditary diseases ("channelopathies") are described in section VI. In Figure 1, the localization of the most important mutations is marked by black numbers on white, those of channelopathies as white numbers on black background.

    B. Localization of the Gate Mediating Fast Inactivation

    The intracellular linker between domains D3 and D4 plays an important role as derived from studies with antibodies directed against the linker which completely blocks fast inactivation (39, 461, 462); also, inactivation is lost or slowed if this linker is cut (432) or bears deletions or mutations (217, 344). In the following description, mutants of Nav1.2 are given if not otherwise mentioned. In these Nav1.2 channels, the critical motif is the hydrophobic triad I1488, F1489, M1490 (IFM; "h" in Fig. 1). Inactivation is completely blocked in I1488Q-F1489Q-M1490Q (IFM/QQQ, corresponds to I1303Q-F1304Q-M1305Q in Nav1.4), but mutant F1489Q alone considerably slows inactivation (502) as does the equivalent F1304Q in muscle channels (Nav1.4) for which kinetic modeling led to the assumption of three inactivated states (328). Inactivation is restored by adding short peptides containing IFM (KIFMK-amide) to the intracellular side (117, 118) or in channels in which inactivation is slowed by external application of -toxin of the scorpion Leiurus quinquestriatus (116). SCAM experiments with the mutant F1489C show that F becomes inaccessible to MTSET on inactivation, suggesting that IFM serves as a hydrophobic latch for a hinged lid formed by the D3-D4 linker (216). The NMR solution structure of the isolated inactivation gate (the linker peptide) has been identified as a stably folded core consisting of a -helix capped by an NH2-terminal turn (373). It is supposed that on gate closing the core (the latch) pivots on a more flexible hinge region. Molecular dynamic simulation predicts an additional helical segment in the D3-D4 linker, on the other side of IFM, which possibly helps to guide the inactivation particle towards its receptor (416). Another solution structure study has been done in a medium of low dielectric constant (293).

    More recent experiments on mutant muscle channels, Nav1.4, revealed that the charge of residues beyond the IFM motif also plays an important role: reduction of positive charge (K1317N,K1318N) accelerates inactivation kinetics in the absence of the -subunit (289), whereas charge reversal in the presence of the -subunit (E1314R,E1315R) slows open-state inactivation but accelerates closed-state inactivation (165). It is concluded that clusters of negatively and positively charged residues in the D3-D4 linker differentially regulate the kinetics of fast inactivation.

    The D3-D4 linker of Nav1.2 (brain) channels also contains a serine, S1506, distal of IFM, whose phosphorylation by protein kinase C (PKC) slows inactivation without an effect on the steady-state curve h(V) (500, 501). Phosphorylation of other sites in the D1-D2 linker reduces INa, an effect that is also observed on phosphorylation by protein kinase A (99, 401). In cardiac Na+ channels (Nav1.5) phosphorylation by PKC at the equivalent site S1505 causes a strong negative shift of h(V) pointing to stabilized inactivation from closed states; the S1505A mutant is almost unaffected by PKC (352). In muscle e?-channels (Nav1.4), the equivalent serine S1321 does not seem to play a key role in shifting h(V) by PKC, since it is also observed in a S1321A mutant (30).

    Fast inactivation of Nav1.2 is also interrupted in mutants F1764A/I1765A/L1766A of D4S6 (corresponding to F1579/80/81 in Nav1.4, in Fig. 1), originally thought to form part of the hydrophobic latch receptor (278). Later experiments with F1764A/V1774A excluded a direct interaction with the IFM motif, since application of KIFMK-amide nevertheless restored inactivation (279). KIMFK restores fast inactivation of open but not of closed channels in the F1651A/L1660A mutant in the D4S4-S5 intracellular loop (approximately at in Fig. 1). This suggests that the IFM motif interacts with the loop during inactivation of closed channels (280). In Nav1.4 cysteine substitution, F1579C, inhibits both fast and slow inactivation, whereas Y1586C and I1575C enhance them (21). Interestingly, these residues belong to the LA binding site, which hence appears to be involved in both types of inactivation.

    The D3-D4 linker, however, is not the only determinant of fast inactivation. Thus if in the more slowly inactivating heart channel (Nav1.5) the linker is replaced by that of the faster brain channel (Nav1.2), its properties are not transferred (171). On the other hand, replacing the COOH terminals accelerates inactivation but also magnifies the differences in voltage dependence of the steady-state inactivation (270). Similar results have been obtained with Nav1.4/Nav1.5 chimeras (108, 109). In a more detailed study of the heart channel COOH terminal, it was shown that truncation of its distal part only reduces current density. Truncation of the proximal part, consisting of six helices, additionally shifts the inactivation curve and clearly increases the fraction of noninactivating channels (91). Further experiments with mutants, in particular truncation of the highly charged sixth helix which increases the persistent INa, suggest that the COOH terminal stabilizes inactivation and minimizes channel reopening (301). The COOH terminal contains a binding site of calmodulin, which is important for the functional expression of channels. In Nav1.6 calmodulin also slows inactivation in a calcium-dependent manner (176).

    Important insights have also been gained from experimental changes of the voltage sensor segment, S4. Point mutations of charged amino acids in S4 to cysteine changed the kinetics of deactivation from open and inactivated states in a domain-specific fashion (164). In heart channels (Nav1.5), charge-neutralizing or -reversing substitutions shifted the activation curve m(V) and decreased its slope as had already been demonstrated for D1S4 in the pioneering work of Ste筯mer et al. (432). Inactivation time constants were markedly decreased only in D4S4 mutations (Ref. 82; see also sect. IVB). In rat brain channels (rNav1.2), addition of a positive charge just external of the outermost positively charged residue of D4S4 (F1625R or F1625K) led to a split of h(V) into two components and a large shift towards hyperpolarization but only to a shift in neutral mutants. These and other findings add to the idea that the D4S4 movement controls the inactivation gate, whereby the countercharges may play an important role in the D4S4 position (512).

    The still poorly understood connection between sensor movement and gating has recently been subject of several "perspective" papers (44, 134, 189, 243). It should be mentioned that the "paddle" hypothesis leaves the problem of the sensor-gate coupling unresolved.

    C. Localization of Site(s) Responsible for Slow Inactivation

    The mechanism of slow inactivation is more complicated, located in structures distinct from those of fast inactivation, and it is less well understood (472). Since it bears similarities to the C-type inactivation of potassium channels, which is connected to the external pore lining, corresponding residues in sodium channels have been the target of site-directed mutagenesis. Thus, in the absence of 1, mutant W402C (P-region of D1, ; Ref. 22) eliminates slow inactivation, whereas mutation of the adjacent residue, E403C or E403R, seems to favor entry into the slow inactivation state (520) as does W434A (492). E403 together with E758, D1214, and D1532 at comparable positions in domains 2, 3, and 4, respectively, form an outer ring of charges whose structural rearrangement was found to be associated with slow inactivation (507). Also, slow inactivation alters the accessibility by the positively charged MTSEA (2-aminoethylmethanethiosulfonate) to the outer pore cysteine of mutant F1236C (P-region of D3,), pointing to a structural rearrangement which, incidentally, may be linked to use-dependent LA action (334). A mutant of the adjacent K1237S or E leads to a very slow ("ultra-slow") type of inactivation (447). Such very slow inactivation is also enhanced by mutant A1529D in the P-loop of D4, part of the putative selectivity filter (177).

    As already mentioned, slow inactivation of human cardiac sodium channels (hNav1.5) develops more slowly and is only 40% complete versus 80% in skeletal muscle channels (Nav1.4) (367). Chimeras of domains of these two channels show that slow inactivation can be modulated by all four domains, with D1 and D2 being more prominent (335). Interestingly, a single residue in D2S5-S6 (P-region), V754 of Nav1.4 and the corresponding I891 of Nav1.5, confers their parental properties to the chimera; however, considering the other experimental evidence, it seems unlikely that the P-region is directly involved in slow inactivation (470). Also, mutant V787K (D2S6; ) markedly enhanced slow inactivation, whereas with V787C it was further slowed, incomplete, and less voltage dependent than in wild type (336). Mutants Y401C and G1530C in the P-region of D1 and D4, respectively, were modified by MTSET at the same rate during slow inactivation as in the noninactivated state, which suggests that the outer mouth of the pore remains open in either state (430).

    In mutant channels devoid of fast inactivation, slow inactivation remains intact (127, 463). In such mutants, substitution of positive charges in S4 of D1 or D2 shifts the slow inactivation to more positive potentials, whereas with intact fast inactivation, S4 mutations in D2 and D3 cause a negative shift of slow inactivation; obviously fast and slow inactivation interact (236), but the extent is differently interpreted (127, 450, 463). Coupling between fast and slow inactivation has also been proposed from analysis of point mutation F1403Q in Nav1.4 (328). Also, mutation L1482C in the D4S4-S5 loop of human muscle Na+ channels disrupts fast but enhances slow inactivation (8). Mutants of charged residues around the fast-inactivation IFM particle, D1309Q or R and EE1314,15RR, right-shift the steady-state relation between slow inactivation and membrane potential, suggesting that these residues may interact with the structures that control slow inactivation (275).

    Further studies of the interaction between fast and slow inactivation were done on mutations in segment D4S4 (which is likely to be involved in the coupling) with the method of cysteine modification. Fast inactivation of mutant R1454C (near ) is similar to that of wild type but becomes much slower on modification by MTSET and MTSES (2-sulfonatoethylmethanethiosulfonate) (510). However, only treatment with the latter (negatively charged) reagent also affects slow inactivation, whereby peak current is reduced ("use dependent") during trains of depolarizing pulses (291). Interestingly, this effect is partially counteracted by a point mutation in the P-region of D1 (W408A) which, in the presence of the 1-subunit, shifts activation to more negative potentials, hastens fast inactivation (449), and attenuates recovery from slow inactivation (211).

    D. Importance of the -Subunits

    In Xenopus oocytes, coexpression of the -subunit of Nav1.2 with 1-subunit of neuronal and skeletal muscle channels increases the current density, accelerates inactivation, and shifts the steady-state inactivation curve in the hyperpolarizing direction (202, 343) as illustrated by Figure 2. Comparable results have been obtained with Nav1.8 and, to a lesser extent, with Nav1.7, channels found in dorsal root ganglia (466). Coexpression also accelerates recovery from inactivation (32, 62). It seems that the -subunit has a slow and a fast gating mode leading to diphasic inactivation. The latter mode is favored on binding of 1 whereby the inactivation kinetics of either component does not change (296). An opposite effect, slowing of inactivation and inducing a persistent INa, is observed by coexpression of G protein -subunits with Nav1.2 (264). -Subunits expressed alone in mammalian cells inactivate almost as fast as the native preparation (203), which has been attributed to an endogenous splice variant, 1A, present in these cells (299). This hypothesis was subsequently rejected on the grounds of experiments with antisense oligonucleotides (297). An additional subunit, 3, predominant during development (406), has been found that accelerates inactivation but less than 1 with which it is closely related (300). The distribution of 3 in human tissues and comparative amino acid sequences of 1, 2, and 3 is described in Stevens et al. (426). Most recently, a novel disulfide-linked subunit, 4, has been identified which shows similarities with 2; on coexpression it shifts the activation curve in the hyperpolarizing direction without affecting inactivation (516).

    The 2-subunit does not modulate hH1 (=Nav1.5) or IIA (=Nav1.2) and 1/2 chimeras, expressed in Xenopus oocytes, have served to determine the regions of 1 necessary for modulation: hH1 channels via the transmembrane portion ("membrane anchor") plus additional regions, IIA channels via the extracellular region only (276, 277, 522) as has been suggested before (79).

    These effects are not yet fully understood, but the functional domains for the interaction seem to be highly conserved (62, 343). Further experiments revealed that only the extracellular domain of 1 is essential for the interaction (79, 277), whereby segment D4SS2-S6 of the -subunit plays an important role (see broken-line connection in Fig. 1; Ref. 351). Earlier studies pointed to this region: inactivation of human heart channels (hH1 = Nav1.5), in contrast to skeletal muscle channels (hSkM1 = Nav1.4), is not much accelerated on coexpression (in Xenopus oocytes, Ref. 329) with 1, but chimeras containing only D4S5-S6 of hSkM1 show the typical acceleration (268). As for the increased channel density, it is interesting to note that in HEK293 cells the hH1-1 complex forms already in the endoplasmic reticulum which may facilitate trafficking to the plasma membrane (523).

    V. CHEMICAL MODULATION

    A. Modulation by Toxins and Local Anesthetics

    Various toxins and other chemical agents slow or even abolish inactivation. Several groups of toxins have been characterized by binding studies that originally led to the definition of five binding sites (47, 67, 495) that were eventually extended to nine sites (524). Experiments on mutant channels (see sect. IV) reveal that toxins within one binding group often do not occupy identical but rather overlapping sites. Also, there exists considerable allosteric interaction between the sites. In the following section, toxin binding to sites 2, 3, 5, and 6 will be discussed as well as other chemicals that affect Na+ channel inactivation.

    1. Site 2 toxins: veratridine, batrachotoxin, grayanotoxin, and aconitine

    The toxins of this group are lipid soluble, which enables them to access binding sites embedded in the membrane. The most important of these compounds are veratridine (VT), an alkaloid from lilaceous plants, BTX, which is secreted by the skin of Colombian arrow-poison frogs, aconitine (AC) from plants of the buttercup family, and grayanotoxin (GTX) from plants of the heather family. Although these toxins differ widely in their structure, their common effect is to keep sodium channels open, hence they are termed "agonists." The underlying mechanism is a large shift of activation in the hyperpolarizing direction and a slowed or even abolished inactivation. Also, the toxins clearly bind to open channels. Details are found in reviews (64, 106, 224, 294, 455) and in the book of Hille (181).

    A typical VT (60 e蘉) effect is illustrated by Figure 3, which was obtained with long depolarizing impulses on a frog node of Ranvier. It not only shows the slowly developing Na+ inward current followed by a large slowly decaying current tail (cut off) at the end of the pulse but also the slow current reduction on adding benzocaine during the impulse. In contrast, if benzocaine was applied to Na+ channels of a Ranvier node kept open by CT (see sect. VA5), block was very fast (half-time 60 ms). From this and other results it is hypothesized that channels kept open by VT cannot be directly blocked, and the current reduction shown in trace 2 in Figure 3 is determined by the rate with which VT-modified channels close during the pulse and become susceptible to the local anesthetic (457).

    Site 2 toxins show "use dependence" in that their effects increase on repeated membrane depolarization; measures that open sodium channels (site 3 toxins, mutations) facilitate their action. In contrast to toxins of other groups, they also affect ion permeation, impressively demonstrated with VT by a sudden reduction of single-channel conductance on binding (27, 413, 477) which explains the much reduced macroscopic INa in the presence of VT (453). GTX-modified channels also have a reduced conductance (508). Likewise sodium channels from eel electroplax inserted in planar bilayers have reduced conductances in the sequence unmodified > BTX > GTX > VT (115, 364).

    Obviously these toxins considerably distort the channel since also selectivity for Na+ is reduced. Thus permeability of ammonium ions increases considerably in the presence of VT (251) and in the presence of AC ammonium ions may even become more permeable than Na+ and the relative permeabilities of Cs+, Rb+, and K+ increase (161, 304). Comparable results have been reported under the influence of BTX and VT (135) and AC (362). On the other hand, Li+ that passes native sodium channels readily becomes clearly less permeable under VT treatment (135). Hence, in Li+ Ringer solution, the typical VT-induced afterpotentials are no longer observed (453). Details of altered selectivity caused by VT including ion flux measurements are found in a review (455), that of other site 2 toxins also elsewhere (181).

    2. Site 3 toxins: scorpion, sea anemone, and spider toxins

    Another binding site has been characterized and termed site 3 to which -scorpion toxins and sea anemone toxins bind (for a review, see Ref. 67). Site 3 toxins have been the subject of many studies so that the reader is referred to a selection of reviews treating toxin structure (156, 218, 326, 347, 370, 498) and/or electrophysiological effects (245, 246, 283, 414, 454). These toxins consist of a polypeptide chain held together by several disulfide bonds; they act from outside the membrane and inhibit inactivation and render it incomplete, but their binding sites seem to be overlapping rather than identical (372). An example of the effect of ATX II, a toxin of the sea anemone Anemonia sulcata, is given in Figure 4, which was obtained in a voltage-clamped frog node of Ranvier at a saturating (with respect to the late current) concentration of 5 e蘉.

    There exists a large sequence homology among the many -scorpion toxins (347) and sea anemone toxins (326) but little between the two. Not only do the toxins of various scorpion species differ with respect to their effects, but also to the target channel isoforms as in amphibian nodes of Ranvier (285, 305, 485), rat skeletal muscle (80) versus heart muscle (81). TTX-resistant sodium channels of dorsal root ganglion cells are also resistant to sea anemone toxins (and other similarly acting toxins) in contrast to TTX-sensitive channels of this preparation (Ref. 389; see below and sect. V, A3, A5, and B).

    As for the toxin molecule, subtle changes may yield largely different effects. Thus in neuroblastoma cells, sea anemone toxins APE 1eC1 and 1eC2 of Anthopleura elegantissima, differing by four amino acids, cause noninactivating currents of 20 and 40%, respectively (54). Likewise, two toxins from another sea anemone (Anthopleura xantogrammatica), ApA and ApB, show great differences in affinity for mammalian (tsA201 cells) channels despite a strong sequence homology, arising solely from different rates, koff, of toxin-channel dissociation (41). In frog nerve ATX I, which differs from ATX II by several amino acids, is ineffective and does not antagonize ATX II (395). Affinity may also be affected: a 20-fold change in EC50 of slowing inactivation of cloned hH1 (=Nav1.5) by Bunodosoma granulifera toxins II and III which differ by only one amino acid (157). The exclusive targets of some -scorpion toxins are insect sodium channels, which makes them potential selective insecticides (155, 524).

    Another type of scorpion toxins, termed -like, is toxic to both mammals and insects. Typical representatives are LqhIII of Leiurus quinquestriatus hebraeus and BomIII and BomIV of Buthus occitanus mardochei (72, 80, 137, 138); LqhIII also affects frog axons where, however, it acts like the classical -toxin (37). The -like toxins seem to bind to a site that is differentially related to site 3 (72). The venom of the scorpion Tityus serrulatus contains an interesting toxin, TiTx, which binds to muscle surface channels with a very high affinity (26). In neuroblastoma cells, TiTx reduces peak INa, causes an inward current to flow near the resting potential, but also slows inactivation (468). The toxin thus shows effects of both - and -toxins. Similar results have been observed in Xenopus nodes of Ranvier (208).

    Binding of the scorpion -toxins is weaker on depolarization (36, 71, 285, 403, 429, 496), but the authors do not completely agree as to whether the open and/or inactivated state confers the reduced affinity. Toxin II of the sea anemone Anemonia sulcata (ATX II) binding is not reduced on depolarization (285, 473), whereas actions of ATX III and IV on crayfish axons were clearly reduced (497). More recently, voltage dependence of scorpion toxin binding was found in a voltage range where activation and inactivation saturates so that it may originate from other sources (80, 81). Moreover, in steady-state binding experiments to rat brain synaptosomes, depolarization, achieved by high K+ concentration, yielded different kinetic results than short applications.

    The toxins not only slow inactivation but render it incomplete (reviewed in Refs. 106, 454), inducing a persistent INa component (36) which is more prominent in neuronal than in cardiac Na+ channels. Also, the toxins enhance the rate of recovery from inactivation through closed states (42).

    Slowed and incomplete inactivation is also induced by funnel-web spider toxins that have been shown to bind to site 3 and compete with scorpion -toxins (258, 259), some of them also act on insect Na+ channels (163). These toxins are without effect on the inactivation of TTX-resistant channels of rat dorsal root ganglia (321, 322, 434). The spider toxins -atracotoxins form a new family of polypeptides with no similarity to scorpion -toxins (128). -ACTX-Hv1a (formerly versutoxin) is contained in the venom of the funnel-web spider Hadronyche versuta, which also produces the less effective -atracotoxin-Hv1b of no insecticidal effect (434). -ACTX-Ar1 (formerly robustoxin) from Atrax robustus is dangerous to humans; it exhibits a 83% amino acid sequence homology with -ACTX-Hv1a (321). Another spider, Paracoelotes luctuosus, produces the insecticidal -palutoxins with effects similar to those of -scorpion toxins (95). Reviews of spider toxins have been published (94, 162).

    3. Site 5 toxins and persistent sodium current

    In many preparations a small persistent INa is observed already in normal saline; although small, it may be important in regulating excitability (97, 436, 443). Different channel isoforms with comparable inactivation kinetics nevertheless have persistent components of distinctly different size (83, 104, 267, 525). Some such results have been interpreted as "window" currents flowing in the potential range where steady-state activation and inactivation curves overlap (19, 317, 384), but this interpretation does not fit results in ventricular myocytes (388) and especially in mammalian neurons (13, 97, 214, 340).

    In addition to site 3 toxins, another group of toxins, binding to site 5, causes persistent currents including ciguatoxin, which is responsible for the ciguatera fish poisoning. Site 5 toxins are lipid-soluble polyethers produced by dinoflagellates (20, 136). At least in frog nodes of Ranvier, ciguatoxin merely induces a late (persistent) component without affecting inactivation kinetics. Activation of this component is shifted by 30 mV towards hyperpolarization. The size of this component depends on the holding potential (VH) being three times larger at VH = eC70 than eC120 mV (38). In TTX-sensitive channels of rat dorsal root ganglia neurons, pacific ciguatoxin-1 shifts the activation curve and the steady-state inactivation curve in the hyperpolarizing direction. In TTX-resistant channels, toxin mainly increases the rate of recovery from inactivation (428).

    Another group of dinoflagellate toxins that bind to site 5 are the brevetoxins (also polyether molecules), the cause of paralytic or neurotoxic shellfish poisoning. Brevetoxin-3 (PbTx-3), for example, produces a hyperpolarizing shift of the activation curve in TTX-sensitive channels of rat sensory neurons accompanied by an inhibition of inactivation (204). In TTX-sensitive Na+ channels of rat brain, brevetoxin PbTx-3 causes a similar shift of the activation curve which, together with a slowing of inactivation, leads to hyperexcitability (349). Actions of site 5 toxins on ion channels have been reviewed (20, 102, 274, 505). Studies of brevetoxin derivatives to elucidate active centers of the toxin have appeared (136, 204); some of these derivatives act as antagonists (350).

    4. Site 6 toxins: -conotoxins

    Marine snails of the genus Conus produce a great variety of toxins that act on different ion channels including sodium channels (reviewed in Refs. 125, 445). Many block these channels, but a group of polypeptides, consisting of 26eC32 amino acids with 3 disulfide bridges, termed -conotoxins, slows inactivation like -toxins of scorpions but does not bind to site 3. Hence, a new site 6, located at the extracellular side of the membrane, was defined (122; summarized in Ref. 524). The main target of the toxins TxVIA (C. textile), NgVIA (C. nigropunctatis), and GmVIA (C. gloriamaris) is the sodium channel of mollusk neurons (122, 123, 172, 173, 411). Mammalian cells, as contained in rat brain synaptosomes, generally do not respond (but see below), but 22Na influx stimulated by veratridine is further increased (123). Also, TxVIA, lacking electrophysiologically detectable effects on insect axons or frog muscle, nevertheless bind to "silent" receptors on these preparations (410).

    Toxins of the fish-hunting Conus striatus retard inactivation in mouse neuroblastoma cells (154), in frog sympathetic neurons (55), and in frog nodes of Ranvier (as does -EVIA) and shift activation in the hyperpolarizing direction (170). Binding seems to be potential dependent with an optimum between eC100 and eC60 mV (154).

    Recently, inhibition of inactivation was also reported for mammalian channels; Am 2766, a toxin from C. amadis, was found to affect rat brain channels (433) and -EVIA, a toxin from C. ermineus, mammalian neuronal and amphibian Na+ channels but not muscle or cardiac channels (24). Table 1 recapitulates the neurotoxins mentioned, their electrophysiological effects, and the putative location of their binding sites.

    5. Other toxins and agents affecting inactivation

    A) OTHER POLYPEPTIDE TOXINS. There are several other toxins slowing inactivation that do not seem to bind to site 3 and are chemically quite different such as Goniopora coral toxin (9.7 kDa; Refs. 152, 307) and pompilidotoxin from solitary wasps (1.5 kDa; Refs. 226, 387). Interestingly, rat heart cells are much less affected by pompilidotoxin (226), and TTX-resistant channels of rat trigeminal ganglion cells are not affected at all (387). Binding of these toxins is reduced by depolarization. A recently studied toxin, PnTx2eC6, isolated from the spider Phoneutria nigriventer venom, slows inactivation kinetics in frog muscle channels and does not seem to bind to site 3 (273). A review of spider and wasp toxins has recently appeared (107).

    B) INSECTICIDES. Insect sodium channels are the target of many insecticides, especially dichlorodiphenyltrichloroethane (DDT) analogs and pyrethroids, synthetic analogs of the natural pyrethrin of Chrysanthemum flowers (311). These agents inhibit inactivation (346, 467, 524), but they also affect squid axons (262), lobster axons (312), as well as amphibian (14, 112, 178, 469) and mammalian channels (141, 142), however at higher concentrations and to a varying degree (155, 310, 418). Pyrethroids bind to a separate site allosterically coupled to sites 2 (345) and 3 (451). TTX-resistant channels of rat dorsal root ganglia are more sensitive to pyrethroids than TTX-sensitive channels of the same preparation (141, 435, 441). The differential sensitivity of sodium channel isoforms to pyrethroids has recently been reviewed by Soderlund et al. (418).

    Pyrethroids shift the relation between activation and membrane potential in the direction of hyperpolarization and cause a similar shift of inactivation; in their presence, large and prolonged tail currents are observed after repolarization. These tails are drastically further prolonged on lowering the temperature so that the depolarizing afterpotentials increase and with it the deadly repetitive discharges (141, 302, 419). Open channels of squid axons have a higher affinity for pyrethroids (263), and in frog muscle fibers, opening is a prerequisite to the action of the pyrethroid deltamethrin (252). Although many features remind of those caused by veratridine, pyrethroids do not reduce single-channel conductance nor do they change selectivity (509).

    C) CHLORAMINE-T. Chloramine-T (N-chloro-p-toluenesulfonamide; CT) is a mild oxidant (mol wt 228) that irreversibly slows inactivation and renders it incomplete without affecting activation so that, on repolarization, it leaves the tail currents almost unchanged (169, 454, 479). In cloned muscle channels, elimination of fast inactivation by CT even accelerated development of slow inactivation without changing recovery or steady-state slow inactivation (491). CT acts on either side of the membrane in crayfish axons (198) but on rat brain channels, inserted in planar lipid bilayers, only if applied to the inside (98), possibly since the agent cannot penetrate the bilayer in contrast to natural membranes (327). CT is commonly thought to act by modifying methionine groups (479), but this explanation has been questioned (356). Besides, the action of CT cannot be highly specific as it also delays the inactivation of some types of potassium channels (377, 424). CT is preferably applied for only a few minutes, since longer treatments destroy the preparations with the possible exception of squid axons (480).

    If shortly applied (to frog nerve fibers) while the membrane is depolarized, CT only reduces peak INa, whereas application during a slight hyperpolarization affects only inactivation, leaving the peak current unchanged as shown in Figure 5, A and B (393). It appears that CT immobilizes the inactivation gate in a portion of channels: in the closed state in Figure 5A, in the open state in Figure 5B. A very similar result has been obtained with an IFM/CFM mutant labeled with a photoactivable cross-linking group with slowed inactivation; on ultraviolet irradiation of the hyperpolarized membrane, a large persistent INa (see Fig. 5D) was observed, whereas irradiation during a lasting depolarization only reduced the peak current (see Fig. 5C), comparable to Figure 5A (192).

    A state-independent effect on inactivation, however, has been observed in crayfish axons in which the blocking action turned out to be reversible on washing (198). CT pretreatment considerably enhances the action of agents that preferably bind to open channels such as batrachotoxin (440), veratridine (454), grayanotoxin (518), or pyrethroids (419). In frog nodes of Ranvier, cooling further slows the modified inactivation and increases the persistent INa fraction accordingly; this argues against a separate channel population to be responsible for this current component (394).

    D) N-BROMOACETAMIDE. N-bromoacetamide (NBA; mol wt 138) is another agent that inhibits inactivation if applied to the inside of squid axons (337, 339) or crayfish axons (390), GH3 cells (194), or cardiac channels (231, 292). External application succeeded in other preparations such as neuroblastoma cells (153), amphibian muscle (325), or myelinated nerve fibers (478), possibly because the membranes of these preparations can be more easily penetrated. Removal of inactivation by NBA is faster at eC100 mV than at eC30 mV as if the inactivation gate is partially protected in its closed configuration (390). NBA was also found to slow (reversibly) inactivation of certain K+ channels but possibly by a different mechanism (338).

    E) GLUTARALDEHYDE. Glutaraldehyde (mol wt 100), a protein cross-linking agent, acts like NBA on frog nerve (393) and muscle (89), where its possible effect of lowering the internal pH (which inhibits inactivation) could not be excluded (325). Glutaraldehyde inhibits inactivation (and clearly reduces peak INa) of squid axons (191) and similarly affects rat cardiac myocytes as studied in single-channel experiments (232, 234). Glutaraldehyde also affects K+ channels (191).

    F) IODATE. Iodate, internally applied, inhibits inactivation in amphibian (90, 320, 393, 396, 421) as well as rat myelinated nerve fibers (319), frog muscle (325), rabbit Schwann cells (196), and rat cardiac myocytes (28, 39, 90, 232).

    G) POSITIVE INOTROPIC AGENTS. A group of synthetic agents exerts a positive inotropic effect by inhibiting inactivation of cardiac sodium channels and thus prolonging the open state of these channels leading to an increased Na+ influx (186): racemic DPI 201eC106 and its congener BDF 9148 with their active components S-DPI (=DPI 205eC430) and S-BDF (=BDF 9196), respectively. The effects of S-DPI were mostly studied in cardiomyocytes but are also observed in mouse skeletal muscle (341) or neuroblastoma cells (376). Electrophysiological experiments suggest that the three agents bind to different but allosterically coupled sites (399), although DPI 201eC106 prolongs the open state of sodium channels, as do ATX II or VT. The effects of S-DPI have been studied in single-channel experiments (233, 234, 324) that have also revealed conductance substates (341, 403). Many other results with this cardiotonic are found in a review (106).

    BDF 9148 increases action potential duration by delaying inactivation of cardiac sodium channels of guinea pigs (363) and humans (186, 306). Another congener, BDF 9198, delays the slow component of inactivation in guinea pig heart and induces a persistent INa component (517). A similar effect has been reported for still another congener, the S-enantiomer of RWJ 24517: casatrin (452). In cultured ventricular myocytes, BDF 9145 proved to interact synergistically with site 2 toxin VT (477).

    H) PROTEOLYTIC ENZYMES. In preparation for internal perfusion of squid axons, pronase, a mixture of proteolytic enzymes, was employed to facilitate extrusion of the axoplasm. Prolonged enzymatic treatment caused inactivation, but not activation, to be destroyed, which suggested that the inactivation gate is accessible from the axoplasmic side (18, 374). This evidence helped to formulate the ball-and-chain model of fast inactivation (17). Later the active pronase component alkaline proteinase b was identified (375). Removal of squid axon inactivation was employed in studying activation and deactivation (337), steady-state properties of activation (427), and use-dependent block by many local anesthetics (reviewed in Ref. 310). In crayfish axons, pronase treatment revealed that the blocking agent methylene blue, internally applied, binds to open channels (422) and that the rate of removal of inactivation is voltage dependent (390).

    The proteolytic enzyme trypsin, which cleaves peptides at lysine and arginine residues, removes inactivation if applied to the interior of excitable cells. This has been observed in GH3 cells (460), N18 neuroblastoma cells (153), or cardiac myocytes in which, however, -chymotrypsin is more effective (87). These and other experiments employing agents to remove inactivation were used to study the relation of activation and inactivation. Also, means to prolong the open state improve measurement of currents through single channels inserted in lipid bilayers (3).

    I) FREE FATTY ACIDS AND PHOSPHOLIPIDS. Dietary polyunsaturated fatty acids (PUFA), in particular of the n-3 class as contained in fish oil, are antiarrhythmic agents. They exert their effect on cardiac (Nav1.5) Na+ channels by reducing INa and shifting the steady-state inactivation curve in the hyperpolarizing direction. Since other, structurally unrelated, heart channels are also inhibited, an indirect effect via membrane phospholipids was discussed. However, channels with a mutation in D1S6 (N406K) are significantly less sensitive to PUFAs, and this inhibitory effect is increased on coexpression of 1 (506). This points to a direct effect on the -subunit rather than to an effect on the membrane phospholipids such as their packing (348) or fluidity (253).

    6. Local anesthetics preferentially bind to inactivated channels

    Block by local anesthetics (LA) varies with structure and physical properties of drug (56, 181) and of the type of channel (106, 400) and depends on potential history of the membrane: relatively weakly if resting, but strongly if depolarized. Especially effective are frequent pulses that may lead to increasing block until a new steady level is reached. This phasic block has been termed "use dependent" (96) and has been the subject of numerous studies and several interpretations of which the "modulated receptor hypothesis" has gained most interest. The hypothesis assumes that affinity for LA depends on the conformational state of the channel with the inactivated state showing the highest affinity (179, 188). This energetically favored conformation, then, is stabilized in the presence of LA, resulting in a shift of the steady-state inactivation curve to more negative potentials. A different interpretation ("guarded receptor hypothesis") assumes that drug binds to a constant-affinity receptor whose access is regulated by channel gates leading to apparently variable affinities (423). Either hypothesis has been used to describe LA results, mostly obtained with antiarrhythmics on cardiac myocytes, however with no clear decision. This has been acknowledged in a critical assessment by one of the proponents of the guarded receptor hypothesis (158) who in a later study showed its limitations, at least in preparations with chemically slowed gating (160). On the other hand, a recent study of mutated cardiac channels offers arguments in favor of this hypothesis (361). It should be mentioned that a more peripheral role of the inactivation gate in LA action has been suggested (Ref. 464; see sect. VB3).

    Because block of inactivated channels apparently plays a role in the understanding of LA action, experiments on channels with impaired inactivation have repeatedly been done, however with varying results. Thus in squid axons treated with CT, use dependence persisted, but after pronase treatment it was almost absent (57, 481). CT-treated frog nerve also showed use dependence (519). In this preparation, channels kept open by CT (during 14-s depolarizing pulses) were blocked on sudden application of benzocaine within a fraction of a second. This shows that the LA receptor remains readily accessible (457). If tested with 15-ms impulses after equilibration in benzocaine, the late INa was much more reduced than the peak current. This reflects a weaker resting block, and affinity for resting channels was estimated to be 17 times lower than for inactivated channels (281). In this preparation benzocaine uniformly reduced INa during a test pulse if inactivation was completely abolished by batrachotoxin (398). Benzocaine enhanced the residual steady-state inactivation of batrachotoxin-modified brain channels, rNav1.2 (486).

    Many LA are racemates whose constituents often differ pharmacologically, in particular in binding to the receptor (182, 515) but also in unbinding leading to differences in use dependence (86, 105, 459). No stereoselective action on TTX-resistant channels was observed (51). LA stereoisomers have also been tested on mutant sodium channels to elucidate the structure of the receptor as shown in section VB3.

    It may be mentioned here that partial block of TTX-resistant channels of dorsal root ganglia by La3+ leads to a slowed inactivation of the remaining inward INa, whereas block and slowing are much less if INaflows outward. The kinetic effect appears to reflect that blocked channels cannot inactivate, which becomes visible as the blocking-unblocking reaction at a site near the selectivity filter is significantly faster than the inactivation rate (242).

    B. Toxins and Receptor Sites

    1. Molecular determinants of toxin binding site 2

    Site 2, to which the lipid-soluble toxins VT, BTX, AC, and GTX bind (64, 65; for a review, see Ref. 495), turned out to very likely consist of overlapping receptors. Attempts to localize the binding site(s) employed point mutations. Thus it was found that if residues in the middle of S6 of several domains were changed to the positively charged lysine, the channel, expressed in a mammalian cell, became resistant to BTX: N433K, N434K and N437K in D1 (493), N784K and L788K in D2 (489), S1276K and L1280K in D3 (490), and F1579K and N1584K in D4 (494); other mutants in D4, F1764A and I1760A, caused affinity to be reduced 60- and 4-fold, respectively (257). These mutants showed little change in gating. In mutant N434K on the other hand, VT binding was not abolished but seemed to reduce channel conductance (as in WT, Ref. 252) and affect gating (493). Similar observations were also made with F1579K and N1584K, which led to the hypothesis of VT action that in a first step peak INa was reduced and then VT trapped within the D1S6 and D4S6 domain interface (482). More recently, the assumed close alignment of the VT receptor and the LA receptor (see also sect. VB3), already suggested earlier (see Fig. 3; Ref. 457), was studied by cysteine substitutions of three residues constituting the LA receptor: N434C/L1280C/F1579C (S6 of D1, D3, and D4; rNav1.4). This mutant remained susceptible to LA block, whereas VT not only failed to keep the channel open but progressively blocked it on repetitive pulsing. This block was prevented by simultaneous application of LA (488). The authors suggest that the VT and LA receptors overlap extensively with that of VT being situated in the inner channel vestibule. VT blocks also wild-type channels, although imperfectly, which may lead to the drastically reduced unitary conductance, in contrast to the interpretation of Barnes and Hille (27).

    The moderately BTX-resistant mutant F1710C in D4S6 of Nav1.3 channels (expressed in Xenopus oocytes) allows rapid dissociation of toxin (which in wild type binds almost irreversibly) from its receptor, but only in open channels. This led to the conclusion that BTX reaches its receptor from the cytoplasmic side only when the activation gate is open (256), but access and dissociation occur probably only during the initial steps of depolarization (103).

    The BTX-resistant mutants I433K, N434K, and L437K of D1S4 and their matches in rat heart muscle, V406K and L410K, turned out to be resistant to GTX as well (201). Other completely GTX-resistant mutants were found in D4S6: I1575A and Y1586K, whereas Y1586A and F1579K were partially resistant. It is suggested that GTX and BTX have overlapping but not identical binding sites. In mutant F1579A, on the other hand, GTX was more effective than in wild type (225). Recent binding studies on mutant Nav1.4 channels revealed two more GTX binding sites on D2S6 (N784) and D3S6 (S1276). Also, systematic substitutions of F1579 (D4S6) showed that the smaller the substituent residue, the more both rates kon and koff were increased, leading to almost unchanged values of KD, whereas substitution of Y1586 (D4S6) selectively increased koff, thus increasing KD (266). The authors conclude that all four S6 segments contribute to the binding site but F1579 appears to control access and Y1586 binding. Less is known about AC action on mutant channels. In human heart muscle, the related alkaloid lappaconitine which, in contrast to the agonist AC, blocks (irreversibly) sodium channels, fails to bind to mutants F1760K or N1765K of D4S6, corresponding to the BTX-resistant mutants F1579K and N1584K in skeletal muscle (503).

    2. Molecular determinants of toxin binding site 3

    Site 3 toxins act only if applied to the external side of the membrane, and earlier experiments employing photoaffinity labeling placed the receptor site on the extracellular loop S5-S6 of D1 (444). This was confirmed by antibody mapping that revealed another spot on the equivalent loop of D4 (446). Experiments with modified toxins (see summary in Ref. 372) suggested acidic residues to be required for toxin binding. In experiments on mutant -subunits (Nav1.2 expressed in mammalian cells) several residues in D4S3-S4 could be identified, with E1613R and E1613H showing considerably reduced binding of scorpion (Leiurus quinquestriatus) toxin and sea anemone (Anemonia sulcata) toxin ATX II, however to a somewhat different extent for the two groups of toxins. Extensive studies have been done testing three types of Leiurus quinquestriatus hebraeus toxins on rNav1.4 channels into which D4S3-S4 linkers of various isoforms had been incorporated. These studies show that recombining a few amino acid residues in site 3 leads to phenotypes of great toxicological variety (254). Wild-type cardiac channels have a much reduced affinity for scorpion toxin, and substitution of cardiac D4S3 into skeletal muscle -subunits rendered them rather insensitive to toxin (372). TTX-resistant Nav1.8 is resistant to Leiurus quinquestriatus hebraeus toxin as well and so is a chimera of the sensitive muscle channel Nav1.4 containing, in the D4S3-S4 linker, four additional amino acids (SLEN) existing in Nav1.8 (385).

    Although binding affinities of wild-type and mutant (E1613R) -subunits differ largely, the voltage dependence of scorpion toxin binding at equilibrium, KD(V), is quite similar. It is suggested that the affinity changes are due to a conformational change required for fast inactivation (see sect. IVB) that follows the voltage-dependent transitions among closed states (372). It was found that the increase in KD is predominantly due to an increased dissociation rate (koff) of scorpion toxin, whereas association (kon) appears to be independent of potential, at least between eC120 and eC80 mV. Comparable results have been obtained with different scorpion toxins (80). However, the situation is more complicated since in steady-state binding studies on synaptosomes with toxin II of Leiurus quinqestriatus hebraeus (LqhII), kon turned out to be far more affected by depolarization than koff. Fast depolarization led to a biphasic unbinding whose first phase revealed a fast koff as observed in electrophysiological experiments on rBII sodium channels (rNav1.2) expressed in mammalian cells; the second slower phase resembled the almost voltage-independent off-rate observed in the steady-state binding experiments attributed to binding to the slow-inactivated state (139). Strong depolarizing pulses also weaken the effect of Ts3, a toxin from the scorpion Tityus serrulatus, assumedly due to dissociation from its binding site (58).

    Point mutations in Nav1.2, away from the putative receptor in the central part of D4S4, that slow inactivation by slowing S4 movement, also reduce the effect of ATX II without changing its affinity. This has been interpreted as an electrostatic interaction of the positively charged toxin with the outermost S4 residue (240); it helps to understand how a toxin that binds to the external side eventually influences the inactivation "lid" at the internal side. Inhibition of D4S4 movement by a site 3 toxin has also been observed in Nav1.5 channels (408).

    3. Molecular determinants of the binding sites of local anesthetics

    The importance of inactivation for phasic block by local anesthetics (LA) prompted experiments with inactivation-deficient mutants with varying results depending on the LA. Thus in IFM/QQQ mutants (in D3-D4 linker, "h" in Fig. 1) use dependence of lidocaine block was absent (Nav1.5, ventricular myocytes; Ref. 331) as with disopyramide, whereas it was retained with the lidocaine derivative RAD-243 or flecainide (159). In F/Q mutants of rat muscle channels, rNav1.4 (coexpressed with 1 in Xenopus oocytes), use-dependent action of lidocaine was observed, at least at higher concentrations, and lidocaine blocked the plateau INa much more efficiently than the peak current; this was not observed in IFM/QQQ mutants (23). In their attempt to describe these effects by a kinetic model, the authors concluded that lidocaine acts as an allosteric effector to enhance inactivation. A mutant of this F, F1304C, reacts with MTSET applied to the inside in noninactivated (hyperpolarized) channels thereby blocking inactivation. However, if the channels are depolarized, the cysteine becomes unreactive and may thus indicate the gate position also during depolarization-dependent block by lidocaine. After a depolarizing pulse, the gate reopens with the same kinetics independent of whether or not lidocaine is present, but in the latter case, INa recovers very slowly as the anesthetic slowly unbinds (464). This unexpected result suggests that the inactivation gate may play a more marginal role in local anesthesia.

    Other experiments to determine the LA binding site were done by site-directed mutations in the middle of segment S6 of domains D1 to D4. Mutations F1764A and Y1771A in D4S6 of rat brain IIA channels, rNav1.2, expressed in Xenopus oocytes, facing the pore ( in Fig. 1), drastically decreased the affinity of open and inactivated channels for etidocaine, leading to a nearly complete loss of use and voltage dependence (357). The authors concluded that the hydrophobic residues F1764 and Y1771 are determinants of the LA binding site, and substitution with alanine destabilizes drug binding. Mutation N1769A, oriented away from the pore, considerably increased the affinity only of the resting channel, possibly by an indirect effect. Mutant I1760A, closer to the extracellular side of D4S6, enabled the quaternary LA QX314 to block on external application, which would be ineffective in the wild type. In subsequent experiments, these authors found comparable results with lidocaine, the anticonvulsant phenytoin and, to a lesser extent, with quinidine and flecainide, suggesting an overlapping common receptor site (358). In rat muscle channels, rNav1.4, expressed in HEK293 cells, etidocaine, and the neutral LA benzocaine were tested on mutants N1584A and F1579A (equivalent to N1769A and F1764A of Nav1.2), revealing an increased block of the former and a decreased block of the latter by both LA (483). This suggests a common receptor for the two types of LA, already deduced from kinetic experiments on wild-type channels (397). Also, mutant I1575A (equivalent to I1760A in Nav1.2) allowed the channel to be blocked by externally applied QX314. Substitution by hydrophilic lysine did not change resting block by benzocaine at F1579K but decreased it at N1584K and increased it at Y1586K (504).

    BTX-resistant mutants N434K and L437K (D1S6, near in Fig. 1) bind LA significantly weaker than wild type, whereas N434D restored or even enhanced binding (484). Cysteine substitutions N434C enhanced inactivation (shift of Vh to more negative potentials) and increased sensitivity to lidocaine, whereas I436C shifted Vh in the opposite direction (235). The authors also observed a reduction of use-dependent block in D2S6 mutants like I782C and V786C and concluded that determinants of LA binding in D1S6 and D2S6 are subsidiary to those in D4S6. This confirms the extensive study on alanine-substituted residues in D1S6 and D2S6 (514) in Nav1.2 channels and those in D3S6 (513). The results of these two papers, together with earlier findings, suggest that the LA receptor site is formed by residues in D3S6 and D4S6 with the contribution of a single amino acid in D1S6.

    The BTX-resistant mutants S1276K and L1280K (D3S6) showed much reduced bupivacaine binding to the inactivated state, but clearly less to the resting state. It appears that LAs interact with these S6 segments when the channel is in its inactivated state (490). BTX-resistant mutants N784K and L788K (in D2S6) remain sensitive to bupivacaine (489). Incidentally, the BTX-resistant N434R shows a clear stereoselectivity, decreasing S(eC)-bupivacaine potency definitely more than of its enantiomer; similar results were obtained with cocaine but not with RAC 109 (314). Stereoselective bupivacaine block of inactivated Nav1.4 channels was also found with mutant L1280 (D3S6), leading to the assumption that this residue and N434 interact directly with LA, facing each other in the pore (316). Weak stereoselectivity for bupivacaine was also observed in heart channel (Nav1.5) mutants F1760K, Y1765K, and N406K, corresponding to F1575K, Y1584K, and N434K in Nav1.4 channels (315).

    Other experiments on mutants of heart sodium channels, Nav1.5, such as F1760K (expressed in HEK293 cells) revealed that block of resting and inactivated states by the antidepressant amitriptyline (a use-dependent blocker of sodium channels) is greatly reduced and use dependence eliminated (313).

    Finally, experiments have been done on mutants causing hereditary muscle diseases (see sect. VIB) to test whether LA could restore normal inactivation behavior. This was not completely achieved in equine hyperkalemic periodic paralysis (386) or paramyotonia congenita (124, 438). Incidentally, two agents of possible clinical significance, the preservative 4-chloro-m-cresol and the bacteriostatic diluent benzylalcohol, have been tested on these mutants, showing typical LA behavior (167, 168).

    4. Binding sites for agents of low molecular weight

    A) CT. Little is known of the precise location of the CT target, generally assumed to be a methionine group (see sect. VA5). The M of the IFM triad (in the D3-D4 linker) forming the inactivation lid could be a likely candidate as it stabilizes the inactivation state via hydrophobic interactions (373); oxidation of M would disrupt these interactions (195). However, this interpretation may be questionable, since an IFI mutant remains sensitive to CT (487). Quionez et al. (355), working on muscle fibers, postulate at least two critical methionine groups.

    B) NBA. The incomplete inactivation of cardiac Na+ channels after treatment with NBA and iodate was indistinguishable from that on applying an antibody targeting a portion of the D3-D4 linker, which provided a substrate for NBA but not for the action of iodate (39). In myelinated frog nerve fibers, application of near-saturating concentrations of ATX II after iodate pretreatment yielded additive effects, suggesting different mechanisms by which these two agents act (396).

    VI. GENETIC MODULATION: CHANNELOPATHIES

    Mutations of single amino acid residues in the channel molecule are not only produced in the laboratory but are recognized as the cause of many hereditary diseases called "channelopathies." The first channelopathies were identified in skeletal muscle by electrophysiological experiments that later led to genetic and functional expression studies. Defects of sodium channel inactivation, mostly leading to hyperexcitability, are also found in heart muscle and in the central nervous system leading to a great variety of clinical symptoms. An extensive review of genetics, chemistry, and symptoms is available (248), also in a shorter version (249).

    A. Heart Muscle

    The most serious disorder in heart muscle caused by ion channel defects is the long Q-T syndrome (LQT) due to disturbed myocardial repolarization that may lead to ventricular arrhythmias or even sudden death (215, 227, 271). However, only one (rare) type, LQT3, is linked to a sodium channel gene, SCN5A, encoding the -subunit. One defect in hH1 (hNav1.5) channels is the deletion KPQ of three consecutive amino acids (K1505, P1506, Q1507) in the D3-D4 linker (not far from the inactivation "lid") accompanied by a persistent current (34, 77, 476). A persistent INa component was also observed in the point mutations N1325S (in D3S4-S5) and R1644H (in D4S4; Ref. 476); slowed inactivation was seen in R1623Q (in D4S4, Ref. 212). N1325 and R1644 are located near the docking sites of the inactivation gate, whereas R1623 seems to be involved in activation-inactivation coupling.

    In section IVB, it was mentioned that the COOH terminal stabilizes inactivation. It is therefore to be expected that inherited mutations in the COOH terminal of cardiac sodium channels may cause arrhythmias. For example, E1784K in the acidic domain within the early portion of the COOH terminal is accompanied by a small persistent INa (499). Y1795C ( in Fig. 1) produces LQT3 with the onset of inactivation slowed, whereas Y1795H speeds inactivation leading to the Brugada syndrome, an ST elevation in ECG leads V1 through V3 (369). It is suggested that in Y1795C a disulfide bond is formed with a partner cysteine in the channel (442). Another mutant, 1795insD, in which an aspartate is inserted, may lead to both LQT3 and Brugada syndrome (46), possibly because fast inactivation is disrupted, but slow inactivation is augmented which could reduce INa at rapid heart rates (465).

    B. Skeletal Muscle

    A great number of hereditary skeletal muscle diseases are due to mutations in the sodium channel -subunit (gene SCN4A) that lead to slowed inactivation. Many reviews have appeared (25, 59, 146, 187, 199, 209, 229, 248, 250). The inactivation defects induce trains of action potentials or even sustained depolarization of the sarcolemma, with the former causing myotonic symptoms and the latter paralyses due to persistent inactivation. The most common diseases are paramyotonia congenita (PMC), hyperkalemic periodic paralysis (HyperPP), and potassium-aggravated myotonia (PAM). PMC is a paradoxical myotonia with cold-induced muscle stiffening during exercise followed by weakness or paralysis. HyperPP shows episodic attacks of generalized weakness triggered by rest after body exertion or K+ intake, accompanied by hyperkalemia. Hypokalemic periodic paralysis (HypoPP-2) is a rare disorder characterized by intermittent attacks of weakness accompanied by a decrease in serum K+ concentration. PAM is a generalized myotonia aggravated by K+ intake, showing no weakness.

    The underlying mutations are found in expected locations such as the D3-D4 linker containing the "lid" (PMC, PAM) or D4S6 (HyperPP, PAM) but also in other locations as S4-S5 linker of D2 ( in Fig. 1; PMC) or in D4S3 and D4S4 for PMC ( ) and the intracellular linker S4-S5 in D3 ( ; HyperPP, PAM); the mechanisms are not always understood. Lists of mutations are found in several reviews (60, 61, 229, 248). Often the mutations were studied in heterologously expressed -subunits (hNav1.4) rather than in native muscle fibers or even on rat channels (rNav1.4; Ref. 298). More recent papers and some containing particularly interesting mutations are mentioned below.

    Thus cold aggravation of PMC has recently been studied again; that caused by mutation R1448H (in D4S4) showed a clearly increased "window" current on cooling (295), R1448C a disproportionally slowed deactivation on repolarization (110). Mutation R1441P in rat (corresponding to R1448 in humans) also slowed deactivation, which opposes action potential repolarization and thus exacerbates PMC; this double cause could contribute to cold aggravation as was shown by computer simulation (126). Mutation T1313A/M ( in D3-D4 linker, distal to "h" in Fig. 1), another cause of PMC, does not change temperature sensitivity per se but further slowing of inactivation on cooling may push it over the threshold of PMC (50).

    The most common HyperPP mutations are T704M in D2S5 and M1592V in D4S6; they cause a small persistent INa with T704M also affecting activation (511). Shifts of both activation and inactivation in the hyperpolarizing direction have recently been observed in a nearby mutation, I1495F, which also enhances slow inactivation (31). Slow inactivation is also affected in rat muscle T698M (corresponding to T704M in human channels; in Fig. 1) leading to a very quick recovery from slow inactivation which, combined with a shift in Vm of activation, is thought to result in the persistent INa mentioned (101). Other authors assign the defective slow inactivation caused by mutation M1585V (human: M1592V) only a supporting role (174). A rare form of painful myotonia without weakness has recently been observed due to mutation V445M in D1S6 ( ) in which slow inactivation was enhanced (437, 475).

    Whereas serum K+ concentration is usually high during attacks of HyperPP, another periodic paralysis, hypokalemic periodic paralysis (HypoPP) is accompanied by reduced K+ concentration. Only some cases (HypoPP-2) of this disease are due to Na+ channel mutations such as R672G/H/S (in D2S4) leading to enhanced inactivation and reduced INa (210, 425). In a study of the first mutation in human Na+ channels, R669H, that has been associated with HypoPP-2, no effects on fast inactivation but an enhanced slow inactivation was found (431).

    Potassium-aggravated myotonia (PAM) has been observed in mutations with impaired fast inactivation such as S804F (in D2S6), I1160V (in D3S4-S5), G1306A/E/V ( in D3-D4 linker), and V1589M (in D4S6), extensively described in the review of Kleopa and Barchi (229). Possibly K+ aggravates indirectly by depolarizing the membrane (187, 368).

    C. Central Nervous System

    Mutations in the -subunit of brain sodium channels (Nav1.1 and others) may cause some forms of childhood-onset epilepsy: generalized epilepsy with febrile seizures plus (GEFS+2) and severe myoclonic epilepsy of infancy (SMEI) with GEFS+ being considered "benign" but SMEI "intractable." Mutations in the 1-subunit have been identified as cause of GEFS+1, rendering this subunit ineffective in accelerating inactivation (see sect. IVD), which may thus lead to hyperexcitability (474). Such "gain-of-function" changes, for instance due to persistent INa, may be associated with either GEFS+ or SMEI as no simple correlation exists between clinical symptoms and electrophysiological behavior (366). Even "loss-of-function" mutations, associated with reduced current densities, may cause GEFS+2 or SMEI (260), an unexpected effect that could be indirect by eliminating inhibition.

    An increasing number of mutations causing GEFS+2 have been detected (146), and some have been heterologously expressed in human cells to study their electrophysiology (9, 255, 260, 261). For instance, R1648H (in D4S4) induces a sizable, T875M (in D2S4) and W1204R (cytoplasmic D2-D3 linker, close to D3S1) a small persistent current (261). Also, T875M enhances slow inactivation, another example of reduced excitability resulting in seizures (10, 420).

    VII. CONCLUSION

    Fast inactivation is a highly important feature of sodium channel kinetics as it helps to repolarize the excitable membrane during an action potential, in some preparations such as mammalian nodes of Ranvier being the only repolarizing force. Inactivation modulates the affinity of the channels for various chemical agents, in particular for local anesthetics and similarly acting drugs, leading to increasing block during trains of action potentials that may lead to a beneficial antiarrhythmic effect. The molecular mechanism of fast inactivation as an intrinsic block of the intra-axonal pore entrance is fairly well understood but, less well, why some subtle changes away from this location have drastic effects on inactivation. Although the secondary structure of the channel molecule is known, much is still to be learned about the tertiary structure and higher orders, in particular the exact conformational changes underlying gating. Such information is now increasingly becoming available through X-ray and NMR studies that eventually will also clarify the more complicated mechanism of slow inactivation.

    REFERENCES

    Adelman WJ and Palti Y. The effects of external potassium and long duration voltage conditioning on the amplitude of sodium currents in the giant axon of the squid. J Gen Physiol 54: 589eC606, 1969.

    Afshari FS, Ptak K, Khaliq ZM, Grieco TM, Slater NT, McCrimmon DR, and Raman IM. Resurgent Na currents in four classes of neurons of the cerebellum. J Neurophysiol 92: 2831eC2843, 2004.

    Agnew WS, Cooper EC, Shenkel S, Correa AM, James WM, Ukomadu C, and Tomiko SA. Voltage-sensitive sodium channels: agents that perturb inactivation gating. Ann NY Acad Sci 625: 200eC223, 1991.

    Ahern CA and Horn R. Specificity of charge-carrying residues in the voltage sensor of potassium channels. J Gen Physiol 123: 205eC216, 2004.

    Aidley WJ and Stanfield PR. Ion Channels: Molecules in Action. Cambridge, UK: Cambridge Univ. Press, 1996.

    Aldrich RW, Corey DP, and Stevens CF. A reinterpretation of mammalian sodium channel gating based on single channel recording. Nature 306: 436eC441, 1983.

    Aldrich RW and Stevens CF. Voltage-dependent gating of single sodium channels from mammalian neuroblastoma cells. J Neurosci 7: 418eC431, 1987.

    Alekov AK, Peter W, Mitrovic N, Lehmann-Horn F, and Lerche H. Two mutations in the IV/S4eCS5 segment of the human skeletal muscle Na+ channel disrupt fast and enhance slow inactivation. Neurosci Lett 306: 173eC176, 2001.

    Alekov AK, Rahman MM, Mitrovic N, Lehmann-Horn F, and Lerche H. A sodium channel mutation causing epilepsy in man exhibits subtle defects in fast inactivation and activation in vitro. J Physiol 529: 533eC539, 2000.

    Alekov AK, Rahman MM, Mitrovic N, Lehmann-Horn F, and Lerche H. Enhanced inactivation and acceleration of activation of the sodium channel associated with epilepsy in man. Eur J Neurosci 13: 2171eC2176, 2001.

    Almers W. Gating currents and charge movement in excitable membranes. Rev Physiol Biochem Pharmacol 82: 96eC190, 1978.

    Almers W, Stanfield PR, and Ste筯mer W. Slow changes in currents through sodium channels in frog muscle membrane. J Physiol 339: 253eC271, 1983.

    Alzheimer C, Schwindt PC, and Crill WE. Modal gating of Na+ channels as a mechanism of persistent Na+ current in pyramidal neurons from rat and cat sensomotor cortex. J Neurosci 13: 660eC673, 1993.

    rhem P and Frankenhaeuser B. DDT and related substances: effects on permeability properties of myelinated Xenopus nerve fibre. Potential clamp analysis. Acta Physiol Scand 91: 502eC511, 1974.

    Armstrong CM. Sodium channels and gating currents. Physiol Rev 61: 644eC683, 1981.

    Armstrong CM and Bezanilla F. Charge movement associated with the opening and closing of the activation gate of the Na channel. J Gen Physiol 63: 533eC552, 1974.

    Armstrong CM and Bezanilla F. Inactivation of the sodium channel. II. Gating current experiments. J Gen Physiol 70: 567eC590, 1977.

    Armstrong CM, Bezanilla F, and Rojas E. Destruction of sodium conductance inactivation in squid axons perfused with pronase. J Gen Physiol 62: 375eC391, 1973.

    Attwell D, Cohen I, Eisner D, Ohba M, and Ojeda C. The steady state TTX-sensitive ("window") sodium current in cardiac Purkinje fibres. Pfle筭ers Arch 379: 137eC142, 1979.

    Baden DG. Brevetoxins: unique polyether dinoflagellate toxins. FASEB J 3: 1807eC1817, 1989.

    Bai CX, Glaaser IW, Sawanobori T, and Sunami A. Involvement of local anesthetic binding sites on IVS6 of sodium channels in fast and slow inactivation. Neurosci Lett 337: 41eC45, 2003.

    Balser JR, Nuss HB, Chiamvimonvat N, Peerez-Garce猘 MT, Marban E, and Tomaselli GF. External pore residue mediates slow inactivation in e? rat skeletal muscle sodium channels. J Physiol 494: 431eC442, 1996.

    Balser JR, Nuss HB, Orias DW, Johns DC, Marban E, and Tomaselli GF. Local anesthetics as effectors of allosteric gating. Lidocaine effects on inactivation-deficient rat skeletal muscle Na channels. J Clin Invest 98: 2874eC2886, 1996.

    Barbier J, Lamthanh H, Le Gall F, Favreau P, Benoit E, Chen H, Gilles N, Ilan N, Heinemann SH, Gordon D, Menez A, and Molge?J. A -conotoxin from Conus ermineus venom inhibits inactivation in vertebrate neuronal Na+ channels but not in skeletal and cardiac muscles. J Biol Chem 279: 4680eC4685, 2004.

    Barchi RL. Ion channel mutations and diseases of skeletal muscle. Neurobiol Dis 4: 254eC264, 1997.

    Barhanin J, Iledefonse M, Rougier O, Sampaio SV, Giglio JR, and Lazdunski M. Tityus toxin, a high affinity effector of the Na+ channel in muscle, with a selectivity for channels in the surface membrane. Pfle筭ers Arch 400: 22eC27, 1984.

    Barnes S and Hille B. Veratridine modifies open sodium channels. J Gen Physiol 91: 421eC443, 1988.

    Beck W, Benz I, Bessler W, Jung G, and Kohlhardt M. Responsiveness of cardiac Na+ channels to a site-directed antiserum against the cytosolic linker between domains III and IV and their sensitivity to other modifying agents. J Membr Biol 134: 213eC239, 1993.

    Bell DC, Yao H, Saenger RC, Riley JH, and Siegelbaum SA. Changes in local S4 environment provide a voltage-sensing mechanism for mammalian hyperpolarization-activated HCN channels. J Gen Physiol 123: 5eC20, 2004.

    Bendahhou S, Cummins TR, Potts JF, Tong J, and Agnew WS. Serine-1321-independent regulation of the e? adult skeletal muscle Na+ channel by protein kinase C. Proc Natl Acad Sci USA 92: 12003eC12007, 1995.

    Bendahhou S, Cummins TR, Tawil R, Waxman SG, and Ptacek LJ. Activation and inactivation of the voltage-gated sodium channel: role of segment S5 revealed by a novel hyperkalaemic periodic paralysis mutation. J Neurosci 19: 4762eC4771, 1999.

    Bennett PB, Makita N, and George A. A molecular basis for gating mode transitions in human skeletal muscle Na+ channels. FEBS Lett 326: 21eC24, 1993.

    Bennett PB, Valenzuela C, Chen LQ, and Kallen RG. On the molecular nature of the lidocaine receptor of cardiac Na+ channels. Modification of block by alterations in the -subunit III-IV interdomain. Circ Res 77: 584eC592, 1995.

    Bennett PB, Yazawa K, Makita N, and George AL. Molecular mechanism for an inherited cardiac arrhythmia. Nature 376: 683eC685, 1995.

    Benoit E, Corbier A, and Dubois JM. Evidence for two transient sodium currents in the frog node of Ranvier. J Physiol 361: 339eC360, 1985.

    Benoit E and Dubois JM. Properties of maintained sodium current induced by a toxin from Androctonus scorpion in frog node of Ranvier. J Physiol 383: 93eC114, 1987.

    Benoit E and Gordon D. The scorpion -like toxin LqhIII specifically alters sodium channel inactivation in frog myelinated axons. Neuroscience 104: 551eC559, 2001.

    Benoit E, Legrand AM, and Dubois JM. Effects of ciguatoxin on current and voltage clamped frog myelinated nerve fibre. Toxicon 24: 357eC364, 1986.

    Benz I, Beck W, Kraas W, Stoll D, Jung G, and Kohlhardt M. Two types of modified cardiac Na+ channels after cytosolic interventions at the -subunit capable of removing Na+ inactivation. Eur Biophys J 25: 189eC200, 1997.

    Benz I and Kohlhardt M. Chemically modified cardiac Na+ channels and their sensitivity to antiarrhythmics: is there a hidden drug receptor J Membr Biol 139: 191eC201, 1994.

    Benzinger GR, Drum CL, Chen LQ, Kallen RG, and Hanck DA. Differences in the binding sites of two site-3 sodium channel toxins. Pfle筭ers Arch 434: 742eC749, 1997.

    Benzinger GR, Tonkovich GS, and Hanck DA. Augmentation of recovery from inactivation by site-3 Na channel toxins. A single-channel and whole-cell study of persistent currents. J Gen Physiol 113: 333eC346, 1999.

    Bezanilla F. The voltage sensor in voltage-dependent ion channels. Physiol Rev 80: 555eC592, 2000.

    Bezanilla F. Voltage sensor movements. J Gen Physiol 120: 465eC473, 2002.

    Bezanilla F and Armstrong CM. Inactivation of the sodium channel. I. Sodium current experiments. J Gen Physiol 70: 549eC566, 1977.

    Bezzina C, Veldkamp MV, van den Berg MP, Postma AV, Rook MB, Viersma JW, van Langen IM, Tan-Sindhunata G, Bink-Boelkens MTE, van der Hout AH, Mannens MMAM, and Wilde AAM. A single Na+ channel mutation causing both long-QT and Brugada syndromes. Circ Res 85: 1206eC1213, 1999.

    Blumenthal KM and Seibert AL. Voltage-gated sodium channel toxins. Poisons, probes, and future promise. Cell Biochem Biophys 38: 215eC238, 2003.

    Bhle T and Benndorf K. Multimodal action of single Na+ channels in myocardiac mouse cells. Biophys J 68: 121eC130, 1995.

    Bhle T, Steinbis M, Biskup C, Koopmann R, and Benndorf K. Inactivation of single cardiac Na+ channels in three different gating modes. Biophys J 75: 1740eC1748, 1998.

    Bouhours M, Sternberg D, Davoine CS, Ferrer X, Willer JC, Fontaine B, and Tabti N. Functional characterization and cold sensitivity of T1313A, a new mutation of the skeletal muscle sodium channel causing paramyotonia congenita in humans. J Physiol 554: 635eC647, 2003.

    Bru ME, Branitzki P, Olschewski A, Vogel W, and Hempelmann G. Block of neuronal tetrodotoxin-resistant Na+ currents by stereoisomers of piperidine local anesthetics. Anesth Analg 91: 1499eC1505, 2000.

    Brismar T. Slow mechanism for sodium permeability inactivation in myelinated nerve fibre of Xenopus laevis. J Physiol 270: 283eC297, 1977.

    Broomand A, Mnnikk R, Larsson HP, and Elinder F. Molecular movements of the voltage sensor in a K channel. J Gen Physiol 122: 741eC748, 2003.

    Bruhn T, Schaller R, Schulze C, Sanchez-Rodriguez J, Dannmeier C, Ravens U, Heubach JF, Eckhardt K, Schmidtmayer J, Schmidt H, Aneiros A, Wachter E, and Beeress L. Isolation and characterization of five neurotoxic and cardiotoxic polypeptides from the sea anemone Anthopleura elegantissima. Toxicon 39: 693eC702, 2001.

    Bulaj G, De la Cruz R, Azimi-Zonooz A, West P, Watkins M, Yoshikami D, and Olivera BM. -Conotoxin structure/function through cladistic analysis. Biochemistry 40: 13201eC13208, 2001.

    Butterworth JF and Strichartz GR. Molecular mechanisms of local anesthesia: a review. Anesthesiology 72: 711eC734, 1990.

    Cahalan MD. Local anesthetic block of sodium channels in normal and pronase-treated squid giant axons. Biophys J 23: 285eC311, 1978.

    Campos FV, Coronas FIV, and Beiro PSL. Voltage-dependent displacement of the scorpion toxin Ts3 from sodium channels and its implication on the control of inactivation. Br J Pharmacol 142: 1115eC1122, 2004.

    Cannon SC. Sodium channel defects in myotonia and periodic paralysis. Annu Rev Neurosci 19: 141eC164, 1996.

    Cannon SC. Ion channel defects in the hereditary myotonias and periodic paralyses. In: Scientific American Molecular Neurology, edited by Martin JB. New York: Sci Am, 1998, p. 257eC277.

    Cannon SC. Spectrum of sodium channel disturbances in the nondystrophic myotonias and periodic paralyses. Kidney Int 57: 772eC779, 2000.

    Cannon SC, McClatchey A, and Gusella J. Modification of Na+ current conducted by the rat skeletal muscle alpha subunit by coexpression with a human brain subunit. Pfle筭ers Arch 423: 155eC157, 1993.

    Catterall WA. Binding of scorpion toxin to receptor sites associated with sodium channels in frog muscle. Correlation of voltage-dependent binding with activation. J Gen Physiol 74: 375eC391, 1979.

    Catterall WA. Neurotoxins that act on voltage-sensitive sodium channels in excitable membranes. Annu Rev Pharmacol Toxicol 20: 15eC43, 1980.

    Catterall WA. The molecular basis of neuronal excitability. Science 223: 653eC661, 1984.

    Catterall WA. Voltage-dependent gating of sodium channels: correlating structure and function. Trends Neurosci 9: 7eC10, 1986.

    Catterall WA. Cellular and molecular biology of voltage-gated sodium channels. Physiol Rev 72 Suppl: S15eCS48, 1992.

    Catterall WA. From ionic currents to molecular mechanisms: the structure and function of voltage-gated sodium channels. Neuron 26: 13eC25, 2000.

    Catterall WA. A 3D view of sodium channels. Nature 409: 988eC991, 2001.

    Catterall WA, Chandy KG, and Gutman GA (Editors). The IUPHAR Compendium of Voltage-Gated Ion Channels. Leeds, UK: IUPHAR Media, 2002.

    Catterall WA, Ray R, and Morrow CS. Membrane potential dependent binding of scorpion toxin to action potential Na+ ionophore. Proc Natl Acad Sci USA 73: 2682eC2686, 1976.

    Cesteele S, Stankiewicz M, Mansuelle P, De Waard M, Dargent B, Gilles N, Pelhate M, Rochat H, Martin-Eauclaire MF, and Gordon D. Scorpion -like toxins toxic to both mammals and insects differentially interact with receptor site 3 on voltage-gated sodium channels in mammals and insects. Eur J Neurosci 11: 975eC985, 1999.

    Cha A, Ruben PC, George AL, Fujimoto E, and Bezanilla F. Voltage sensors in domains III and IV, but not I and II, are immobilized by Na+ channel fast inactivation. Neuron 22: 73eC87, 1999.

    Chahine M, Plante E, and Kallen RG. Sea anemone toxin (ATX II) modulation of heart and skeletal muscle sodium channel -subunits expressed in tsA201 cells. J Membr Biol 152: 39eC48, 1996.

    Chandler WK and Meves H. Evidence for two types of sodium conductance in axons perfused with sodium fluoride. J Physiol 211: 653eC678, 1970.

    Chandler WK and Meves H. Slow changes in membrane permeability and long-lasting action potentials in axons perfused with fluoride solutions. J Physiol 211: 707eC728, 1970.

    Chandra A, Starmer CF, and Grant AO. Multiple effects if KPQ deletion mutation on gating of human cardiac Na+ channels expressed in mammalian cells. Am J Physiol Heart Circ Physiol 274: H1643eCH1654, 1998.

    Chay TR. The Hodgkin-Huxley Na+ channel model versus the five-state Markovian model. Biopolymers 31: 1483eC1502, 1991.

    Chen C and Cannon SC. Modulation of Na+ channel inactivation by the 1 subunit: a deletion analysis. Pfle筭ers Arch 431: 186eC195, 1995.

    Chen H, Gordon D, and Heinemann SH. Modulation of cloned skeletal muscle sodium channels by scorpion toxins LqhII, LqhIII, and LqhIT. Pfle筭ers Arch 439: 423eC432, 2000.

    Chen H and Heinemann SH. Interaction of scorpion -toxin with cardiac sodium channels: binding properties and enhancement of slow inactivation. J Gen Physiol 117: 505eC518, 2001.

    Chen LQ, Santarelli V, Horn R, and Kallen RG. A unique role for the S4 segment in domain 4 in the inactivation of sodium channels. J Gen Physiol 118: 549eC556, 1996.

    Chen YH, Dale TJ, Romanos MA, Whittaker WRJ, Xie XM, and Clare JJ. Cloning, distribution and functional analysis of the type III sodium channel from human brain. Eur J Neurosci 12: 4281eC4289, 2000.

    Chiu SY. Inactivation of sodium channels: second order kinetics in myelinated nerve. J Physiol 273: 573eC596, 1977.

    Chiu SY, Ritchie JM, Rogart RB, and Stagg D. A quantitative description of membrane currents in rabbit myelinated nerve. J Physiol 292: 149eC166, 1979.

    Clarkson CW. Stereoselective block of cardiac sodium channels by RAC109 in single guinea pig ventricular myocytes. Circ Res 65: 1306eC1323, 1989.

    Clarkson CW. Modification of Na channel inactivation by -chymotrypsin in single cardiac myocytes. Pfle筭ers Arch 417: 48eC57, 1990.

    Clay JR. On the persistent sodium current in squid giant axons. J Neurophysiol 89: 640eC644, 2003.

    Collins CA, Arispe N, and Rojas E. Slow Na+ conductance inactivation following modification of fast inactivation gating in frog muscle fibres. Biomed Res 4: 363eC373, 1983.

    Conti F, Hille B, Neumcke B, Nonner W, and Stmpfli R. Conductance of the sodium channel in myelinated nerve fibres with modified sodium inactivation. J Physiol 262: 729eC742, 1976.

    Cormier JW, Rivolta I, Tateyama M, Yang AS, and Kass R. Secondary structure of the human cardiac Na+ channel C terminus. Evidence for a role of helical structures in modulation of channel inactivation. J Biol Chem 277: 9233eC9241, 2002.

    Correa AM and Bezanilla F. Gating of the squid sodium channel at positive potentials. I. Macroscopic ionic and gating currents. Biophys J 66: 1853eC1863, 1994.

    Correa AM and Bezanilla F. Gating of the squid sodium channel at positive potentials. II. Single channels reveal two open states. Biophys J 66: 1864eC1878, 1994.

    Corzo G and Escoubas P. Pharmacologically active spider peptide toxins. Cell Mol Life Sci 60: 2409eC2426, 2003.

    Corzo G, Escoubas P, Stankiewicz M, Pelhate M, Kristensen CP, and Nakajima T. Isolation, synthesis and pharmacological characterization of -palutoxins IT, novel insecticidal toxins from the spider Paracoelotes luctuosus (Amaurobiidae). Eur J Biochem 267: 5783eC5795, 2000.

    Courtney KR. Mechanism of frequency-dependent inhibition of sodium currents in frog myelinated nerve by the lidocaine derivative GEA 968. J Pharmacol Exp Ther 195: 225eC236, 1975.

    Crill WE. Persistent sodium current in mammalian central neurons. Annu Rev Physiol 58: 349eC362, 1996.

    Cukierman S. Inactivation modifiers of Na+ currents and the gating of rat brain Na+ channels in planar lipid bilayers. Pfle筭ers Arch 419: 514eC521, 1991.

    Cukierman S. Regulation of voltage-dependent sodium channels. J Membr Biol 151: 203eC214, 1996.

    Cummins TR, Dib-Hajj SD, Black JA, Akopian AN, Wood JN, and Waxman SG. A novel persistent tetrodotoxin-resistant sodium current in SNS-null and wild type small primary sensory neurons. J Neurosci 19: 1eC6, 1999.

    Cummins TR and Sigworth FJ. Impaired slow inctivation in mutant sodium channels. Biophys J 71: 227eC236, 1996.

    Dechraoui MY, Naar J, Pauillac S, and Legrand AM. Ciguatoxins and brevetoxins, neurotoxic polyether compounds active on sodium channels. Toxicon 37: 125eC143, 1999.

    De Leon L and Ragsdale DS. State-dependent access to the batrachotoxin receptor on the sodium channel. Neuroreport 14: 1353eC1356, 2003.

    Del Negro CA, Koshiya N, Butera RJ, and Smith JC. Persistent sodium current, membrane properties and bursting behavior of pre-Btzinger complex inspiratory neurons in vitro. J Neurophysiol 88: 2242eC2250, 2002.

    De Luca A, Natuzzi F, Desaphy JF, Loni G, Lentini G, Franchini C, Tortorella V, and Conte Camerino D. Molecular determinants of mexiletine structure for potent and use-dependent block of skeletal muscle sodium channels. Mol Pharmacol 57: 268eC277, 2002.

    Denac H, Meuysssen M, and Scholtysik G. Structure, function and pharmacology of voltage-gated sodium channels. Naunyn-Schmiedbergs Arch Pharmacol 362: 453eC479, 2000.

    De Olivera Beleboni R, Pizzo R, Fontana ACK, De Carolino R OG, Coutinho-Netto J, and Dos Santos WF. Spider and wasp neurotoxins: pharmacological and biochemical aspects. Eur J Pharmacol 493: 1eC17, 2004.

    Desche簄es I, Chen LQ, Kallen RC, and Chahine M. Electrophysiological study of chimeric sodium channels from heart and skeletal muscle. J Membr Biol 164: 25eC34, 1998.

    Desche簄es I, Trottier E, and Chahine M. Implication of the C-terminal region of the -subunit of voltage-gated sodium channels in fast inactivation. J Membr Biol 183: 103eC114, 2001.

    Dice MS, Abbruzzese JL, Wheeler JT, Groome JR, Fujimoto E, and Ruben PC. Temperature-sensitive defects in paramyotonia congenita mutants R1448C and T1313M. Muscle Nerve 30: 277eC288, 2004.

    Drews G. Effects of chloramine-T on charge movement and fraction of open channels in frog nodes of Ranvier. Pfle筭ers Arch 409: 251eC257, 1987.

    Dubois JM and Bergman C. Asymmetrical currents and sodium currents in Ranvier nodes exposed to DDT. Nature 266: 741eC742, 1977.

    Dubois JM and Schneider MF. Kinetics of intramembrane charge movement and sodium current in frog node of Ranvier. J Gen Physiol 79: 571eC602, 1982.

    Dubois JM, Schneider MF, and Khodorov BI. Voltage dependence of intramembrane charge movement and conductance activation of batrachotoxin-modified sodium channels of frog node of Ranvier. J Gen Physiol 81: 829eC844, 1983.

    Duch DS, Hernandez A, Levinson SR, and Urban B. Grayanotoxin-I-modified eel electroplax sodium channels. Correlation with batrachotoxin and veratridine modifications. J Gen Physiol 100: 623eC645, 1992.

    Eaholtz G, Colvin A, Leonard D, Taylor C, and Catterall WA. Block of brain sodium channels by peptide mimetics of the isoleucine, phenylalanine, and methionine (IFM) motif from the inactivation gate. J Gen Physiol 113: 279eC203, 1999.

    Eaholtz G, Scheuer T, and Catterall WA. Restoration of inactivation and block of open sodium channels by an inactivation gate peptide. Neuron 12: 1041eC1048, 1994.

    Eaholtz G, Zagotta WN, and Catterall WA. Kinetic analysis of block of open sodium channels by a peptide containing the isoleucine, phenylalanine, methionine (IFM) motif from the inactivation gate. J Gen Physiol 111: 75eC82, 1998.

    Elinder F and rhem P. Tail currents in the myelinated axon of Xenopus laevis suggest a two-open-state Na channel. Biophys J 73: 179eC185, 1997.

    Elliott JR. Slow Na+ channel inactivation and bursting discharge in a simple model axon: implications for neuropathic pain. Brain Res 754: 221eC226, 1997.

    El-Sherif N, Fozzard HA, and Hanck DA. Dose-dependent modulation of the cardiac sodium channel by sea anemone toxin ATX II. Circ Res 70: 285eC301, 1992.

    Fainzilber M, Kofman O, Zlotkin E, and Gordon D. A new neurotoxin receptor site on sodium channels is identified by a conotoxin that affects sodium channel inactivation in molluscs and acts as an antagonist in rat brain. J Biol Chem 269: 2574eC2580, 1994.

    Fainzilber M, Lodder JC, Kits KS, Kofman O, Vinnitsky I, Van Rietschoten J, Zlotkin E, and Gordon D. A new conotoxin affecting sodium current inactivation interacts with the -conotoxin receptor site. J Biol Chem 270: 1123eC1129, 1995.

    Fan Z, George AL, Kyle JW, and Makielski JC. Two human paramyotonia congenita mutations have opposite effects on lidocaine block of Na+ channels expressed in a mammalian cell line. J Physiol 296: 275eC286, 1996.

    Favreau P, Le Gall F, Benoit E, and Molge?J. A review on conotoxins targeting ion channels and acetylcholine receptors in vertebrate neuromuscular junction. Acta Physiol Pharmacol Ther Latinoam 49: 257eC267, 1999.

    Featherstone DE, Fujimoto E, and Ruben PC. A defect in skeletal muscle sodium channel deactivation exacerbates hyperexcitability in human paramyotonia congenita. J Physiol 506: 627eC638, 1998.

    Featherstone DE, Richmond JE, and Ruben PC. Interaction between fast and slow inactivation in Skm1 sodium channels. Biophys J 71: 3098eC3109, 1996.

    Fletcher JI, Chapman BE, Mackay JP, Howden MEH, and King GF. The structure of versutoxin (-atracotoxin-Hv1) provides insights into the binding of site 3 neurotoxins to the voltage-gated sodium channel. Structure 5: 1525eC1535, 1997.

    Fox JM. Ultra-slow inactivation of the ionic currents through the membrane of myelinated nerve. Biochim Biophys Acta 426: 232eC244, 1976.

    Fozzard HA and Hanck DA. Structure and function of voltage-dependent sodium channels: comparison of brain II and cardiac isoforms. Physiol Rev 76: 887eC926, 1996.

    Frankenhaeuser B. Quantitative description of sodium current in myelinated nerve fibres of Xenopus laevis. J Physiol 151: 491eC501, 1960.

    Furue T, Yakehiro M, Yamaoka K, Sumii K, and Seyama I. Characteristics of two slow inactivation mechanisms and their influence on the sodium channel activity of frog ventricular myocytes. Pfle筭ers Arch 436: 631eC638, 1998.

    Gandhi CS, Clark E, Loots E, Pralle A, and Isacoff EY. The orientation and molecular movement of a K+ channel voltage-sensing domain. Neuron 40: 515eC525, 2003.

    Gandhi CS and Isacoff EY. Molecular models of voltage sensing. J Gen Physiol 120: 455eC463, 2002.

    Garber SS and Miller V. Single Na+ channels activated by veratridine and batrochotoxin. J Gen Physiol 89: 459eC480, 1987.

    Gawley RE, Rein KS, Jeglitsch G, Adams DJ, Theodorakis EA, Tiebes J, Nicolaou KC, and Baden DG. The relationship of brevetoxin "length" and A-ring functionality to binding and activity in neuronal sodium channels. Chem Biol 2: 533eC541, 1995.

    Gilles N, Blanchet C, Shichor I, Zaninetti M, Lotan I, Bertrand D, and Gordon D. A scorpion -like toxin that is active on insects and mammals reveals an unexpected specificity and distribution of sodium channel subtypes in rat brain neurons. J Neurosci 19: 8730eC8739, 1999.

    Gilles N, Chen H, Wilson H, Le Gall F, Montoya G, Molge?J, Schnherr R, Nicholson G, Heinemann SH, and Gordon D. Scorpion and -like toxins differentially interact with sodium channels in mammalian CNS and periphery. Eur J Neurosci 12: 2823eC2832, 2000.

    Gilles N, Leipold E, Chen H, Heinemann SH, and Gordon D. Effect of depolarization on binding kinetics of scorpion -toxin highlights conformational changes of rat brain sodium channels. Biochemistry 40: 14576eC14584, 2001.

    Gillespie JI and Meves H. The time course of sodium inactivation in squid giant axons. J Physiol 299: 289eC307, 1980.

    Ginsburg KS and Narahashi T. Differential sensitivity of tetrodotoxin-sensitive sodium channels to the insecticide allethrin in rat dorsal root ganglion neurons. Brain Res 627: 239eC248, 1993.

    Ginsburg KS and Narahashi T. Time course and temperature dependence of allethrin modulation of sodium channels in rat dorsal root ganglion cells. Brain Res 847: 38eC49, 1999.

    Goldin AL. Accessory subunits and sodium channel inactivation. Curr Opin Neurobiol 3: 272eC277, 1993.

    Goldin AL. Resurgence of sodium channel research. Annu Rev Physiol 63: 871eC894, 2001.

    Goldin AL. Evolution of voltage-gated Na+ channels. J Exp Biol 205: 575eC584, 2002.

    Goldin AL. Mechanisms of sodium channel inactivation. Curr Opin Neurobiol 13: 284eC290, 2003.

    Goldin AL, Barchi RL, Caldwell JH, Hofmann F, Howe JR, Hunter JC, Kallen RG, Mandel G, Meisler MH, Berwald-Netter Y, Noda M, Tamkun MM, Waxman SG, Wood JN, and Catterall WA. Nomenclature of voltage-gated sodium channels. Neuron 28: 365eC368, 2000.

    Goldman L. Kinetics of channel gating in excitable membranes. Q Rev Biophys 9: 491eC526, 1976.

    Goldman L. Sodium channel opening as a precursor to inactivation. A route to the inactivated state. Eur Biophys J 16: 321eC325, 1989.

    Goldman L and Kenyon JL. Delays in inactivation development and activation kinetics in Myxicola giant axons. J Gen Physiol 80: 83eC102, 1982.

    Goldman L and Schauf CL. Inactivation of the sodium current in Myxicola giant axons. Evidence for coupling to the activation process. J Gen Physiol 59: 659eC675, 1972.

    Gonoi T, Ashida K, Feller D, Schmidt J, Fujiwara M, and Catterall WA. Mechanism of action of a polypeptide neurotoxin from the coral Goniopora on sodium channels in mouse neuroblastoma cells. Mol Pharmacol 29: 347eC354, 1986.

    Gonoi T and Hille B. Gating of Na channels. Inactivation modifiers discriminate among models. J Gen Physiol 89: 253eC274, 1987.

    Gonoi T, Ohizumi Y, Kobayashi J, Nakamura H, and Catterall WA. Actions of a polypeptide toxin from the marine snail Conus striatus on voltage-sensitive sodium channels. Mol Pharmacol 32: 691eC698, 1987.

    Gordon D. A new approach to insect-pest control: combination of neurotoxins with voltage sensitive sodium channels to increase selectivity and specificity. Invert Neurosci 3: 103eC116, 1997.

    Gordon D, Savarin P, Gurevitz M, and Zinn-Justin S. Functional anatomy of scorpion toxins affecting sodium channels. J Toxicol Toxin Rev 17: 131eC159, 1998.

    Goudet C, Ferrer T, Gale L, Artiles A, Batista CF, Possani LD, Alvarez J, Aneiros A, and Tytgat J. Characterization of two Bunodosoma granulifera toxins active on cardiac sodium channels. Br J Pharmacol 134: 1195eC1206, 2001.

    Grant AO. Models of drug interaction with the sodium channel. Clin Invest Med 14: 447eC457, 1991.

    Grant AO, Chandra R, Keller C, Carboni M, and Starmer CF. Block of wild-type and inactivation-deficient cardiac sodium channels IFM/QQQ stably expressed in mammalian cells. Biophys J 79: 3019eC3035, 2000.

    Grant AO, Wendt DJ, Zilberter Y, and Starmer CF. Kinetics of interaction of disopyramide with the cardiac sodium channel: fast dissociation from open channels at normal rest potential. J Membr Biol 136: 199eC214, 1993.

    Grishchenko II, Naumov AP, and Zubov AN. Gating and selectivity of aconitine-modified sodium channels in neuroblastoma cells. Neuroscience 9: 549eC554, 1983.

    Grishin E. Polypeptide neurotoxins from spider venoms. Eur J Biochem 264: 276eC280, 1999.

    Grolleau F, Stankiewicz M, Birinyi-Strachan L, Wang XH, Nicholson GM, Pelhate M, and Lapied B. Electrophysiological analysis of the neurotoxic action of a funnel-web spider toxin, -atracotoxin-Hv1a, on insect voltage-gated Na+ channels. J Exp Biol 204: 711eC721, 2001.

    Groome JR, Fujimoto E, George AL, and Ruben PC. Differential effects of homologous S4 mutations in human skeletal muscle sodium channels on deactivation gating from open and inactivated states. J Physiol 516: 687eC698, 1999.

    Groome JR, Fujimoto E, and Ruben PC. Negative charges in the DIII-DIV linker of human skeletal muscle Na+ channels regulate deactivation gating. J Physiol 548: 85eC96, 2003.

    Guy HR and Seetharamulu P. Molecular model of the action potential sodium channel. Proc Natl Acad Sci USA 83: 508eC512, 1986.

    Haeseler G, Leuwer M, Kavan J, We箁z A, Dengler R, and Piepenbrock S. Voltage-dependent block of normal and mutant muscle sodium channels by 4-Chloro-m-Cresol. Br J Pharmacol 128: 1259eC1267, 1999.

    Haeseler G, Mamarvar M, Bufler J, Dengler R, Hecker H, Aronson JK, Piepenbrock S, and Leuwer M. Voltage-dependent blockade of normal and mutant sodium channels by benzylalcohol. Br J Pharmacol 130: 1321eC1330, 2000.

    Hahin R. Removal of inactivation causes time-invariant sodium current delays. J Gen Physiol 92: 331eC350, 1988.

    Hahin R, Wang GK, Shapiro BI, and Strichartz G. Alterations in sodium channel gating produced by the venom of the marine mollusc Conus striatus. Toxicon 29: 245eC259, 1991.

    Hartmann HA, Tiedeman AA, Chen SF, Brown AM, and Kirsch GE. Effects of III-IV linker mutations on human heart Na+ channel inactivation gating. Circ Res 75: 114eC122, 1994.

    Hasson A, Fainzilber M, Gordon D, Zlotkin E, and Spira ME. Alteration of sodium currents by new peptide toxins from the venom of a molluscivorous Conus snail. Eur J Neurosci 5: 56eC64, 1993.

    Hasson A, Shon KJ, Olivera BM, and Spira ME. Alterations of voltage-activated sodium current by a novel conotoxin from the venom of Conus gloriamaris. J Neurophysiol 73: 1295eC1301, 1995.

    Hayward LJ, Brown RH, and Cannon SC. Slow inactivation differs among mutant Na channels associated with myotonia and periodic paralysis. Biophys J 72: 1204eC1219, 1997.

    Herzog RI, Cummins TR, and Waxman SG. Persistent TTX-resistant Na+ current affects resting potential and response to depolarization in simulated spinal sensory neurons. J Neurophysiol 86: 1351eC1364, 2001.

    Herzog RI, Liu C, Waxman SG, and Cummins TR. Calmodulin binds to the C terminus of sodium channels Nav1.4 and Nav16 and differentially modulates their functional properties. J Neurosci 23: 8261eC8270, 2003.

    Hilber K, Sandtner W, Kudlacek O, Glaaser IW, Kyle JW, French RJ, Fozzard HA, Dudley SC, and Todt H. The selectivity filter of the voltage-gated sodium channel is involved in channel activation. J Biol Chem 276: 27831eC27839, 2001.

    Hille B. Pharmacological modifications of the sodium channels of frog nerve. J Gen Physiol 51: 199eC219, 1968.

    Hille B. Local anesthetics: hydrophilic and hydrophobic pathways for the drug-receptor reaction. J Gen Physiol 69: 497eC515, 1977.

    Hille B. Ionic channels in excitable membranes. Current problems and biophysical approaches. Biophys J 22: 283eC294, 1978.

    Hille B. Ion Channels of Excitable Membranes (3rd ed.). Sunderland, MA: Sinauer, 2001.

    Hille B, Courtney K, and Dum R. Rate and site of action of local anesthetics in myelinated nerve fibers. In: Molecular Mechanisms of Anesthesia. Progress in Anesthesiology, edited by B. R. Fink. New York: Raven, 1975, vol. 1, p. 13eC20.

    Hille B, Leibowitz MD, Sutro JB, Schwarz JR, and Holan G. State-dependent modification of sodium channels by lipid-soluble agonists. In: Proteins of Excitable Membranes, edited by B. Hille and D. M. Fambrough. New York: Wiley, 1987, p. 109eC123.

    Hodgkin AL and Huxley AF. The dual effect of membrane potential on sodium conductance in the giant axon of Loligo. J Physiol 116: 496eC506, 1952.

    Hodgkin AL and Huxley AF. A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol 117: 500eC544, 1952.

    Hoey A, Amos GJ, and Ravens U. Comparison of the action potential prolonging and positive inotropic activity of DPI 201eC106 and BDF 9148 in human ventricular myocardium. J Mol Cell Cardiol 26: 985eC994, 1994.

    Hoffman EP, Lehmann-Horn F, and Re筪el R. Overexcited or inactive: ion channels in muscle disease. Cell 80: 681eC686, 1995.

    Hondeghem LM and Katzung BG. Time- and voltage-dependent interactions of antiarrhythmic drugs with cardiac sodium channels. Biochim Biophys Acta 472: 373eC398, 1977.

    Horn R. Coupled movements in voltage-gated ion channels. J Gen Physiol 120: 449eC453, 2002.

    Horn R. How S4 segments move charge. Let me count the ways. J Gen Physiol 123: 1eC4, 2004.

    Horn R, Brodwick MS, and Eaton DC. Effect of protein cross-linking reagents on membrane currents of squid axon. Am J Physiol Cell Physiol 238: C127eCC132, 1980.

    Horn R, Ding S, and Gruber HJ. Immobilizing the moving parts of voltage-gated ion channels. J Gen Physiol 116: 461eC475, 2000.

    Horn R and Vandenberg CA. Statistical properties of single sodium channels. J Gen Physiol 84: 505eC534, 1984.

    Horn R, Vandenberg CA, and Lange K. Statistical analysis of single sodium channels. Effects of N-bromoacetamide. Biophys J 45: 323eC335, 1984.

    Hoshi T and Heinmann SH. Regulation of cell function by methionine oxidation and reduction. J Physiol 531: 1eC11, 2001.

    Howe JR and Ritchie JM. Multiple kinetic components of sodium channel inactivation in rabbit Schwann cells. J Physiol 455: 529eC566, 1992.

    Hoyt RC. Sodium inactivation in nerve fibers. Biophys J 8: 1074eC1097, 1968.

    Huang JMC, Tanguy J, and Yeh JZ. Removal of sodium inactivation and block of sodium channels by chloramine-T in crayfish and squid giant axons. Biophys J 52: 155eC163, 1987.

    Hudson AJ, Ebers GC, and Bulman DE. The skeletal muscle sodium and chloride channel diseases. Brain 118: 547eC563, 1995.

    Irvine LA, Jafri S, and Winslow RC. Cardiac sodium channel Markov model with temperature dependence and recovery from inactivation. Biophys J 76: 1868eC1885, 1999.

    Ishii H, Kinoshita E, Kimura T, Yakehiro M, Yamaoka K, Imoto K, Mori Y, and Seyama I. Point-mutations related to the loss of batrachotoxin binding abolish the grayanotoxin effect in Na+ channel isoforms. Jpn J Physiol 49: 457eC461, 1999.

    Isom LL, De Jongh KS, Patton DE, Reber BFX, Offord J, Charbonneau H, Walsh K, Goldin AL, and Catterall WA. Primary structure and functional expression of the 1 subunit of the rat brain sodium channel. Science 256: 839eC842, 1992.

    Isom LL, Scheuer T, Brownstein AB, Ragsdale DS, Murphy BJ, and Catterall WA. Functional co-expression of the 1 and type IIA subunits of sodium channels in a mammalian cell line. J Biol Chem 270: 3306eC3312, 1995.

    Jeglitsch G, Rein K, Baden DG, and Adams DJ. Brevetoxin-3 (PbTx-3) and its derivatives modulate single tetrodotoxin-sensitive sodium channels in rat sensory neurons. J Pharmacol Exp Ther 284: 516eC525, 1998.

    Jiang Y, Lee A, Chen J, Ruta V, Cadene M, Chait BT, and MacKinnon R. X-ray structure of a voltage-dependent K+ channel. Nature 423: 33eC41, 2003.

    Jiang Y, Ruta V, Chen J, Lee A, and MacKinnon R. The principle of gating charge movement in a voltage-dependent K+ channel. Nature 423: 42eC48, 2003.

    Jonas P. Temperature dependence of gating current in myelinated nerve fibers. J Membr Biol 112: 277eC289, 1989.

    Jonas P, Vogel W, Arantes EC, and Giglio JR. Toxin of the scorpion Tityus serrulatus modifies both activation and inactivation of sodium permeability of nerve membrane. Pfle筭ers Arch 407: 92eC99, 1986.

    Jurkat-Rott K, Lerche H, and Lehmann-Horn F. Skeletal muscle channelopathies. J Neurol 249: 1493eC1502, 2002.

    Jurkat-Rott K, Mitrovic N, Hang C, Kouzmekine A, Iaizzo P, Herzog J, Lerche H, Nicole S, Vale-Santos J, Chauveau D, Fontaine B, and Lehmann-Horn F. Voltage-sensor sodium channel mutations cause hypokalemic periodic paralysis type 2 by enhanced inactivation and reduced current. Proc Natl Acad Sci USA 97: 9549eC9554, 2000.

    Kambouris NG, Hastings LA, Stepanovic S, Marban E, Tomaselli GF, and Balser JR. Mechanistic link between lidocaine block and inactivation probed by outer pore mutations in the rat e? skeletal muscle sodium channel. J Physiol 512: 693eC705, 1998.

    Kambouris NG, Nuss HB, Johns DC, Tomaselli GF, Marban E, and Balser JR. Phenotypic characterization of a novel long-QT syndrome mutation (R1623) in the cardiac sodium channel. Circulation 97: 640eC644, 1998.

    Karlin A and Akabas MH. Substituted-cysteine accessibility method. In: Methods in Enzymology. Ion Channels, edited by P. M. Conn. San Diego, CA: Academic, 1998, vol. 293, pt. B, p. 123eC145.

    Kay AR, Sugimori M, and Lline R. Kinetic and stochastic properties of a persistent sodium current in mature guinea pig cerebellar Purkinje cells. J Neurophysiol 80: 1167eC1179, 1998.

    Keating MT and Sanguinetti MC. Molecular and cellular mechanisms of cardiac arrhythmias. Cell 104: 569eC580, 2001.

    Kellenberger S, Scheuer T, and Catterall WA. Movement of the Na+ channel inactivation gate during inactivation. J Biol Chem 271: 30971eC30979, 1996.

    Kellenberger S, West JW, Scheuer T, and Catterall WA. Molecular analysis of the putative inactivation particle in the inactivation gate of brain type IIA Na+ channels. J Gen Physiol 109: 589eC605, 1997.

    Kem WR. Sea anemone toxins: structure and action. In: The Biology of Nematocysts, edited by D. A. Hessenger and H. N. Lehoff. San Diego, CA: Academic, 1988, p. 375eC405.

    Keynes RD. The kinetics of voltage-gated ion channels. Q Rev Biophys 27: 339eC434, 1994.

    Keynes RD and Elinder F. The screw-helical voltage gating of ion channels. Proc R Soc Lond B Biol Sci 266: 843eC852, 1999.

    Keynes RD and Rojas E. Kinetics and steady-state properties of the charged system controlling sodium conductance in the squid giant axon. J Physiol 239: 393eC434, 1974.

    Khodakhah K, Melishchu A, and Armstrong CM. Charge immobilization caused by modification of internal cysteines in squid Na channels. Biophys J 75: 2821eC2829, 1998.

    Khodorov BI. Sodium inactivation and drug-induced immobilization of the gating charge in nerve membrane. Prog Biophys Mol Biol 37: 49eC89, 1981.

    Khodorov BI. Batrachotoxin as a tool to study voltage-sensitive sodium channels of excitable membranes. Prog Biophys Mol Biol 45: 57eC148, 1985.

    Kimura T, Kinoshita E, Yamaoka K, Yuki T, Takehira M, and Seyama I. On site of action of grayanotoxin in domain 4 segment 6 of rat skeletal muscle sodium channel. FEBS Lett 465: 18eC22, 2000.

    Kinoshita E, Maejima H, Yamaoka K, Konno K, Kawai N, Shimuzu E, Yokote S, Nakayama H, and Seyama S. Novel wasp toxin discriminates between neuronal and cardiac sodium channels. Mol Pharmacol 59: 1457eC1463, 2001.

    Kirsch GE. Ion channel defects in cardiac arrhythmias. J Membr Biol 170: 181eC190, 1999.

    Kirsch GE, Skattebl A, Possani LD, and Brown AM. Modification of Na channel gating by an scorpion toxin from Tityus serrulatus. J Gen Physiol 93: 67eC83, 1989.

    Kleopa KA and Barchi RL. Genetic disorders of neuromuscular ion channels. Muscle Nerve 26: 299eC325, 2002.

    Kniffki KD, Siemen D, and Vogel W. Development of sodium permeability inactivation in nodal membranes. J Physiol 313: 37eC48, 1981.

    Kohlhardt M, Fichtner H, and Frbe U. Differences in open state of NBA-modified Na+ channels. Eur Biophys J 15: 289eC292, 1988.

    Kohlhardt M, Fichtner H, and Frbe U. Gating in iodate-modified single cardiac Na+ channels. J Membr Biol 112: 67eC78, 1989.

    Kohlhardt M, Frbe U, and Herzig JW. Modification of single cardiac Na+ channels by DPI 201eC106. J Membr Biol 89: 163eC172, 1986.

    Kohlhardt M, Frbe U, and Herzig JW. Properties of normal and non-inactivating single cardiac Na+ channels. Proc R Soc Lond B Biol Sci 232: 71eC93, 1987.

    Kondratiev A and Tomaselli GF. Altered gating and local anesthetic block mediated by residues in the I-S6 and II-S6 transmembrane segments of voltage-dependent Na+ channels. Mol Pharmacol 64: 741eC752, 2003.

    Kontis KJ and Goldin AL. Sodium channel inactivation is altered by substitution of voltage sensor positive charges. J Gen Physiol 110: 403eC413, 1997.

    Kontis KJ, Rounaghi A, and Goldin AL. Sodium channel activation gating is affected by substitutions of voltage sensor positive charges in all four domains. J Gen Physiol 110: 391eC401, 1997.

    Koppenhfer E and Vogel W. Wirkung von Tetrodotoxin und Tetrathylammoniumchlorid an der Innenseite der Schne箁ringsmembran von Xenopus laevis. Pfle筭ers Arch 313: 361eC380, 1969.

    Ke筯n FJP and Greeff NG. Movement of voltage sensor S4 in domain 4 is tightly coupled to sodium channel fast inactivation and gating charge immobilization. J Gen Physiol 114: 167eC183, 1999.

    Ke筯n FJP and Greeff NC. Gating properties of a sodium channel with three arginines substituted by histidines in the central part of voltage sensor S4D4. J Membr Biol 193: 23eC34, 2003.

    Kuo CC and Bean BP. Na+ channels must deactivate to recover from inactivation. Neuron 12: 819eC829, 1994.

    Kuo CC, Chen WY, and Yang YC. Block of tetrodotoxin-resistant Na+ channel pore by multivalent cations: gating modification and Na+ flow dependence. J Gen Physiol 124: 27eC42, 2004.

    Larsson HP. The search is on for the voltage sensor-to-gate coupling. J Gen Physiol 120: 475eC481, 2002.

    Lawrence JH, Orias DW, Balser JR, Nuss HB, Tomaselli GF, O'Rourke B, and Marban E. Single-channel analysis of inactivation-defective rat skeletal muscle sodium channels containing the F1304Q mutation. Biophys J 71: 1285eC1294, 1996.

    Lazdunski M, Balerna M, Chicheportiche R, Fosset M, Jacques Y, Lombet A, Romey G, Schweitz H, and Vincent JP. Marine neurotoxins to study the voltage-dependent sodium channel in excitable membranes. In: Neurotoxins: Fundamental and Clinical Advances, edited by I. W. Chubb and L. G. Geffen. Adelaide, Australia: Flinders Univ. of South Australia, 1979, p. 111eC119.

    Lazdunski M, Frelin C, Barhanin J, Lombet A, Meiri H, Pauron D, Romey G, Schmidt A, Schweitz H, Vigne O, and Vijverberg HPM. Polypeptide toxins as tools to study voltage-sensitive Na+ channels. Ann NY Acad Sci 479: 204eC220, 1986.

    Lee HC, Wang JM, and Swartz KJ. Interaction between extracellular hanatoxin and the resting conformation of the voltage-sensor paddle in Kv channels. Neuron 40: 527eC536, 2003.

    Lehmann-Horn F and Jurkat-Rott K. Voltage-gated ion channels and hereditary disease. Physiol Rev 79: 1217eC1372, 1999.

    Lehmann-Horn F and Jurkat-Rott K. Channelopathies: voltage-gated sodium channels. Pharmaceut News 8: 29eC36, 2001.

    Lehmann-Horn F and Re筪el R. Molecular pathophysiology of voltage-gated ion channels. Rev Physiol Biochem Pharmacol 128: 195eC268, 1996.

    Leibowitz MD, Schwarz JR, Holan G, and Hille B. Electrophysiological comparison of insecticide and alkaloid agonists of Na channel. J Gen Physiol 90: 75eC93, 1987.

    Leibowitz MD, Sutro SB, and Hille B. Voltage-dependent gating of veratridine-modified Na channels. J Gen Physiol 87: 25eC46, 1986.

    Leifert WR, McMurphie EJ, and Saint DA. Inhibition of cardiac sodium currents in adult rat myocytes by n-3 polyunsaturated fatty acids. J Physiol 520: 671eC679, 1999.

    Leipold E, Lu S, Gordon D, Hansel A, and Heinemann SH. Combinatorial interaction of scorpion toxins Lqh-2, Lqh-3, and LqIT with sodium channel receptor sites-3. Mol Pharmacol 65: 685eC691, 2004.

    Lerche H, Jurkat-Rott K, and Lehmann-Horn F. Ion channels and epilepsy. Am J Med Genet 106: 146eC159, 2001.

    Li HL, Hadid D, and Ragsdale DS. The batrachotoxin receptor on the voltage-gated sodium channel is guarded by the channel activation gate. Mol Pharmacol 61: 905eC912, 2002.

    Linford NJ, Cantrell AC, Qu Y, Scheuer T, and Catterall WA. Interaction of batrachotoxin with the local anesthetic receptor site in transmembrane segment IVS6 of the voltage-gated sodium channel. Proc Natl Acad Sci USA 95: 13947eC13952, 1998.

    Little MJ, Wilson H, Zappia C, Cesteele S, Tyler MI, Martin-Eauclaire MF, Gordon D, and Nicholson GM. -Atracotoxins from Australian funnel web spiders compete with scorpion -toxins on both rat brain and insect sodium channels. FEBS Lett 439: 246eC252, 1998.

    Little MJ, Zappia C, Gilles N, Connor M, Tyler MI, Martin-Eauclaire MF, Gordon D, and Nicholson GM. -Atracotoxins from the Australian funnel-web spiders compete with scorpion -toxin binding but differentially modulate alkaloid toxin activation of voltage-gated sodium channels. J Biol Chem 273: 27076eC27083, 1998.

    Lossin C, Rhodes TH, Desai RR, Vanoye CG, Wang D, Carniciu S, Devinsky O, and George AL. Epilepsy-associated dysfunction in the voltage-gated neuronal sodium channel SCN1A. J Neurosci 23: 11289eC11295, 2003.

    Lossin C, Wang DW, Rhodes TH, Vanoye CG, and George AL. Molecular basis of an inherited epilepsy. Neuron 34: 877eC884, 2002.

    Lund AE and Narahashi T. Interaction of DDT with sodium channels in squid giant axon membrane. Neuroscience 6: 2253eC2258, 1981.

    Lund AE and Narahashi T. Kinetics of sodium channel modification by the insecticide tetramethrin in squid giant axons. J Pharmacol Exp Ther 219: 464eC473, 1981.

    Ma JY, Catterall WA, and Scheuer T. Perisistent sodium current through brain sodium channels by G protein subunits. Neuron 19: 443eC452, 1997.

    MacKinnon R. Potassium channels. FEBS Lett 555: 62eC65, 2003.

    Maejima H, Kinoshita E, Seyama I, and Yamaoka K. Distinct sites regulating grayanotoxin binding and unbinding to D4S6 of Nav1.4 sodium channels as revealed by improved estimation of toxin sensitivity. J Biol Chem 278: 9464eC9471, 2003.

    Magistretti J and Alonso A. Fine gating properties of channels responsible for persistent sodium current generation in entorhinal cortex neurons. J Gen Physiol 120: 855eC973, 2002.

    Makita N, Bennett PB, and George AL. Molecular determinants of 1 subunit-induced gating modulation in voltage-dependent Na+ channels. J Neurosci 16: 7117eC7127, 1996.

    Mandel G. Tissue-specific expression of the voltage-sensitive sodium channel. J Membr Biol 125: 193eC205, 1992.

    Mantegazza M, Yu FH, Catterall WA, and Scheuer T. Role of the C-terminal domain in inactivation of brain and cardiac sodium channels. Proc Natl Acad Sci USA 98: 15348eC15353, 2001.

    Marbe E. Cardiac channelopathies. Nature 415: 213eC218, 2002.

    Marban E, Yamagishi T, and Tomaselli GF. Structure and function of voltage-gated sodium channels. J Physiol 508: 647eC657, 1998.

    Matavel A, Cruz JS, Penaforte CL, Arae瞛o DAM, Kalapothakis E, Prado VF, Diniz CR, Cordeiro MN, and Beiro PSL. Electrophysiological characterization and molecular identification of the Phoneutria nigriventer peptide toxin PnTx2eC6. FEBS Lett 523: 219eC223, 2002.

    Mattei C, Molge?J, Legrand AM, and Benoit E. Ciguatoxines et breeveetoxines: dissection de leur actions neurobiologiques. J Soc Biol 193: 329eC344, 1999.

    McCollum EJ, Vilin YY, Spackman E, Fujimoto E, and Ruben PC. Negatively charged residues adjacent to IFM motif in DIII-DIV linker of hNav1.4 differentially affect slow inactivation. FEBS Lett 552: 163eC169, 2003.

    McCormick KA, Isom LL, Ragsdale D, Scheuer T, and Catterall WA. Molecular determinants of Na+ channel function in the extracellular domain of the 1 subunit. J Biol Chem 273: 3954eC3962, 1998.

    McCormick KA, Srinivasan J, White K, Scheuer T, and Catterall WA. The extracellular domain of the 1 subunit is both necessary and sufficient for 1-like modulation of sodium channel gating. J Biol Chem 274: 32638eC32646, 1999.

    McPhee JC, Ragsdale DS, Scheuer T, and Catterall WA. A mutation in segment IVS6 disrupts fast inactivation of sodium channels. Proc Natl Acad Sci USA 91: 12346eC12350, 1994.

    McPhee JC, Ragsdale DS, Scheuer T, and Catterall WA. A critical role for transmembrane segment IVS6 of the sodium channel subunit in fast inactivation. J Biol Chem 270: 12025eC12034, 1995.

    McPhee JC, Ragsdale DS, Scheuer T, and Catterall WA. A critical role for the S4eCS5 intracellular loop in domain IV of the sodium channel -subunit in fast inactivation. J Biol Chem 273: 1121eC1129, 1998.

    Meeder T and Ulbricht W. Action of benzocaine on sodium channels of frog node of Ranvier treated with chloramine-T. Pfle筭ers Arch 409: 265eC273, 1987.

    Meves H. Inactivation of the sodium permeabiltity in squid giant nerve fibres. Prog Biophys Mol Biol 33: 207eC230, 1978.

    Meves H. Effect of scorpion toxin on sodium channels. In: Structure and Function in Excitable Cells, edited by D. C. Chang, I. Tasaki, W. J. Adelman, and H. R. Leuchtag. New York: Plenum, 1983, p. 273eC280.

    Meves H. The gating current of the node of Ranvier. In: Ion Channels, edited by T. Narahashi. New York: Plenum, 1990, vol. 2, p. 65eC121.

    Meves H, Rubly N, and Watt DD. Voltage-dependent effects of a scorpion toxin on sodium current inactivation. Pfle筭ers Arch 402: 24eC33, 1984.

    Meves H and Vogel W. Inactivation of the asymmetrical displacement current in giant axons of Loligo forbesi. J Physiol 267: 377eC393, 1977.

    Meves H and Vogel W. Slow recovery of sodium current and gating current from inactivation. J Physiol 267: 395eC410, 1977.

    Mickus T, Jung HY, and Spruston N. Properties of slow, cumulative sodium channel inactivation in rat hippocampal CA1 pyramidal neurons. Biophys J 76: 846eC860, 1999.

    Miller JR, Patel MK, John JE, Mounsey JP, and Moorman JR. Contributions of charged residues in a cytoplasmic linking region to Na channel gating. Biophys Biochim Acta 1509: 275eC291, 2000.

    Millonas MM and Hanck DA. Nonequilbrium response spectroscopy of voltage-sensitive ion channel gating. Biophys J 74: 210eC229, 1998.

    Mitrovic N, George AL, and Horn R. Role of domain 4 in sodium channel slow inactivation. J Gen Physiol 115: 707eC717, 2000.

    Mitsuiye T and Noma A. Quantification of exponential Na+ current activation in N-bromoacetamide-treated cardiac myocytes of guinea-pig. J Physiol 465: 245eC263, 1993.

    Miyamoto K, Nakagawa T, and Kuroda Y. Solution structure of the cytoplasmic linker between domain III-S6 and domain IV-S1 (III-IV linker) of the rat brain sodium channel in SDS micelles. Biopolymers 59: 380eC393, 2001.

    Moczydlowski E and Schild L. Unitary properties of the batrachotoxin-trapped state of voltage-sensitive sodium channels. In: Handbook of Membrane Channels, Molecular and Cellular Physiology, edited by C. Peracchia. San Diego, CA: Academic, 1994, p. 137eC160.

    Mohammadi B, Mitrovic N, Lehmann-Horn F, Dengler R, and Bufler J. Mechanisms of cold sensitivity of paramyotonia congenita mutation R1448H and overlap syndrome mutation M1360V. J Physiol 547: 691eC698, 2003.

    Moran O. Modulation of the rSkM1 sodium channel -subunit by the 1-subunit does not modify the inactivation kinetics. Neurosci Res Commun 26: 17eC26, 2000.

    Moran O, Conti F, and Tammaro P. Sodium channel heterologous expression in mammalian cells and the role of the endogenous 1-subunits. Neurosci Lett 336: 175eC179, 2003.

    Moran O, Nizzari M, and Conti F. Myopathic mutations affect differently the inactivation of the two gating modes of sodium channels. J Bioerg Biomembr 31: 591eC608, 1999.

    Moran O, Nizzari M, and Conti F. Endogenous expression of the 1A sodium channel subunit in HEK-293 cells. FEBS Lett 473: 132eC134, 2000.

    Morgan K, Stevens EB, Shah B, Cox PJ, Dixon AK, Lee K, Pinnock RD, Hughes J, Richardson PJ, Mizuguchi K, and Jackson AP. 3: an additional auxiliary subunit of the voltage-sensitive sodium channel that modulates channel gating with distinct kinetics. Proc Natl Acad Sci USA 97: 2308eC2313, 2000.

    Motoike HK, Liu H, Glaaser IW, Yang AS, Tateyama M, and Kass RS. The Na+ channel inactivation gate is a molecular complex: a novel role of the COOH-terminal domain. J Gen Physiol 123: 155eC165, 2004.

    Motomura H and Narahashi T. Temperature dependence of pyrethroid modification of single sodium channels in rat hippocampal neurons. J Membr Biol 177: 23eC39, 2000.

    Mozhayeva GN, Naumov AP, Kuryshev YA, and Nosyreva ED. Some properties of sodium channels in neuroblastoma cells modified with scorpion toxin and chloramine-T. Single channel measurements. Gen Physiol Biophys 9: 3eC18, 1990.

    Mozhayeva GN, Naumov AP, Negulyaev YA, and Nosyreva ED. The permeability of aconitine-modified sodium channels to univalent cations in myelinated nerve. Biochim Biophys Acta 466: 461eC473, 1977.

    Mozhayeva GN, Naumov AP, Nosyreva ED, and Grishin EV. Potential-dependent interaction of toxin from venom of the scorpion Buthus eupeus with sodium channels in myelinated fibre. Voltage clamp experiments. Biochim Biophys Acta 597: 587eC602, 1980.

    Me筶ler-Ehmsen J, Nbauer M, and Schwinger RHG. Na+ channel modulating effect of the inotropic compound (S-)BDF 9196 in human myocardium. Naunyn-Schmiedebergs Arch Pharmacol 359: 60eC64, 1999.

    Muramatsu I, Fujiwara M, Miura A, and Narahashi T. Effects of goniopora toxin on crayfish giant axon. J Pharmacol Exp Ther 234: 307eC315, 1985.

    Nagy K. Mechanism of inactivation of single channels after modification by chloramine-T, sea anemone toxin and scorpion toxin. J Membr Biol 106 :29eC40, 1988.

    Nagy K, Kiss T, and Hof D. Single Na channels in mouse neuroblastoma cell membrane. Indications for two open states. Pfle筭ers Arch 399: 302eC308, 1983.

    Narahashi T. Neuroreceptors and ion channels as the basis for drug action: past, present, and future. J Pharmacol Exp Ther 294: 1eC26, 2000.

    Narahashi T, Ginsburg KS, Nagata K, Song JH, and Tabebayashi H. Ion channels as targets for insecticides. Neurotoxicology 19: 581eC590, 1998.

    Narahashi T and Haas HG. Interaction of DDT with the components of lobster nerve membrane conductance. J Gen Physiol 51: 177eC198, 1968.

    Nau C, Seaver M, Wang SY, and Wang GK. Block of human heart hH1 sodium channels by amitryptiline. J Pharmacol Exp Ther 292: 1015eC1023, 2000.

    Nau C, Wang SY, Strichartz GR, and Wang GK. Point mutations at N434 in D1eCS6 of e? Na+ channels modulate binding affinity and stereoselectivity of local anesthetic enantiomers. Mol Pharmacol 56: 404eC413, 1999.

    Nau C, Wang SY, Strichartz GR, and Wang GK. Block of human heart hH1 sodium channnels by the enantiomers of bupivacine. Anesthesiology 93: 1022eC1033, 2000.

    Nau C, Wang SY, and Wang GK. Point mutations at L1280 in Nav1.4 channel D3eCS6 modulate binding affinity and stereoselectivity of bupivacaine enantiomers. Mol Pharmacol 63: 1398eC1406, 2003.

    Neumcke B. Diversity of sodium channels in adult and cultured cells, in oocytes and in lipid bilayers. Rev Physiol Biochem Pharmacol 115: 1eC49, 1990.

    Neumcke B, Fox JM, Drouin H, and Schwarz W. Kinetics of the slow variation of peak sodium current in the membrane of myelinated nerve following changes of holding potential or extracellular pH. Biochim Biophys Acta 426: 245eC257, 1976.

    Neumcke B, Schwarz JR, and Stmpfli R. A comparison of sodium currents in rat and frog myelinated nerve: normal and modified sodium inactivation. J Physiol 382: 175eC191, 1987.

    Neumcke B, Schwarz W, and Stmpfli R. Modification of sodium inactivation in myelinated nerve by Anemonia toxin II and iodate. Analysis of current fluctuations and current relaxations. Biochim Biophys Acta 600: 456eC466, 1980.

    Nicholson GM, Walsh R, Little MJ, and Tyler MI. Characterisation of the effects of robustoxin, a lethal neurotoxin from the Sydney funnel-web spider Atrax robustus, on sodium channel activation and inactivation. Pfle筭ers Arch 436: 117eC126, 1998.

    Nicholson GM, Willow M, Howden MEH, and Narahashi T. Modification of sodium channel gating and kinetics by versutoxin from the Australian funnel-web spider Hadronyche versuta. Pfle筭ers Arch 428: 400eC409, 1994.

    Niemann P, Schmidtmayer J, and Ulbricht W. Chloramine-T effect on sodium conductance of neuroblastoma cells as studied by whole-cell and single-channel analysis. Pfle筭ers Arch 418: 129eC136, 1991.

    Nilius B, Benndorf K, Markwardt F, and Franke T. Modulation of single cardiac sodium channels by DPI 201eC106. Gen Physiol Biophys 6: 409eC424, 1987.

    Nonner W, Spalding BC, and Hille B. Low intracellular pH and chemical agents slow inactivation gating in sodium channels of muscle. Nature 284: 360eC363, 1980.

    Norton RS. Structure and structure-function relationships of sea anemone proteins that interact with the sodium channel. Toxicon 29: 1051eC1084, 1991.

    Nosyreva ED, Grishchenko II, and Negulyaev YA. Effects of chloramine-T on activation and inactivation of sodium channels in neuroblastoma cells. Neurophysiology 19: 574eC579, 1987.

    Nuss HB, Balser JR, Orias DW, Lawrence JH, Tomaselli GF, and Marban E. Coupling between fast and slow inactivation revealed by analysis of a point mutation (F1403Q) in e? rat skeletal muscle sodium channels. J Physiol 494: 411eC429, 1996.

    Nuss HB, Chiamvimonat N, Peerez-Garce猘 MT, Tomaselli GF, and Marban E. Functional association of 1 subunit with human cardiac (hH1) and rat skeletal muscle (e?) sodium channel subunits expressed in Xenopus oocytes. J Gen Physiol 106: 1171eC1191, 1995.

    Ochs G, Bromm B, and Schwarz JR. A three-state model for inactivation of sodium permeability. Biochim Biophys Acta 645: 243eC252, 1981.

    Ogata N and Ohishi Y. Molecular diversity of structure and function of the voltage-gated Na+ channels. Jpn J Pharmacol 88: 365eC377, 2002.

    Ogata N and Tetebayashi H. Slow inactivation of tetrodotoxin-insensitive Na+ channels in neurons of rat dorsal root ganglia. J Membr Biol 129: 71eC80, 1992.

    O'Leary ME. Characterization of the isoform-specific differences in the gating of neuronal and muscle sodium channels. Can J Physiol Pharmacol 76: 1041eC1050, 1998.

    Ong BH, Tomaselli GF, and Balser JP. A structural rearrangement in the sodium channel pore linked to slow inactivation and use dependence. J Gen Physiol 116: 653eC661, 2000.

    O'Reilly JP, Wang SY, Kallen RG, and Wang GK. Comparison of slow inactivation in human heart and skeletal muscle Na+ channel chimaeras. J Physiol 515: 61eC73, 1999.

    O'Reilly JP, Wang SY, and Wang GK. Residue-specific effects on slow inactivation at V787 in D2eCS6 of Nav1.4 sodium channels. Biophys J 81: 2100eC2111, 2001.

    Oxford GS. Some kinetic and steady-state properties of sodium channels after removal of inactivation. J Gen Physiol 77: 1eC22, 1981.

    Oxford GS and Wagoner PK. The inactivating K+ current in GH3 pituitary cells and its modification by chemical reagents. J Physiol 410: 587eC612, 1989.

    Oxford GS, Wu CH, and Narahashi T. Removal of sodium channel inactivation in squid giant axons by N-bromoacetamide. J Gen Physiol 71: 227eC247, 1978.

    Parri HR and Crunelli V. Sodium current in rat and cat thalamocortical neurons: role of a non-inactivating component in tonic and burst firing. J Neurosci 18: 854eC867, 1998.

    Patlak JB. Sodium channel subconductance levels measured with a new variance-mean analysis. J Gen Physiol 92: 413eC430, 1988.

    Patlak JB. Molecular kinetics of voltage-dependent Na+ channels. Physiol Rev 71: 1047eC1080, 1991.

    Patton DE, Isom LL, Catterall WA, and Goldin AL. The adult rat brain 1 subunit modifies activation and inactivation gating of multiple sodium channel subunits. J Biol Chem 269: 17649eC17655, 1994.

    Patton DE, West JW, Catterall WA, and Goldin AL. Amino acid residues required for fast sodium channel inactivation. Charge neutralizations and deletions in the III-IV linker. Proc Natl Acad Sci USA 89: 10905eC10909, 1992.

    Pauron D, Barhanin J, Amichot M, Pralavorio M, Berge JB, and Lazdunski M. Pyrethroid receptor in the insect Na+ channel: alteration of its properties in pyrethroid-resistant flies. Biochemistry 28: 1673eC1677, 1989.

    Pichon Y, Guillet JC, Heilig U, and Pelhate M. Recent studies of the effects of DDT and pyrethroid insecticides on nervous activity in cockroaches. Pestic Sci 16: 627eC640, 1985.

    Possani LD, Becerril B, Delepierre M, and Tytgat J. Scorpion toxins specific for Na+ channels. Eur J Biochem 264: 287eC300, 1999.

    Pound EM, Kang JX, and Leaf A. Partitioning of polyunsaturated fatty acids, which prevent cardiac arrhythmias, into phospholipid cell membranes. J Lipid Res 42: 346eC351, 2001.

    Purkerson SL, Baden DG, and Fieber LA. Brevetoxin modulates neuronal sodium channels in two cell lines derived from rat brain. Neurotoxicology 20: 909eC920, 1999.

    Purkerson-Parker SL, Fieber LA, Rein KS, Podona T, and Baden DG. Brevetoxin derivatives that inhibit toxin activity. Chem Biol 7: 385eC393, 2000.

    Qu Y, Rogers JC, Chen SF, McCormick KA, Scheuer T, and Catterall WA. Functional roles of the extracellular segments of the sodium channel subunit in voltage-dependent gating and modulation by 1 subunits. J Biol Chem 274: 32647eC32654, 1999.

    Qu Y, Rogers JC, Tanada TN, Catterall WA, and Scheuer T. Phosphorylation of S1505 in the cardiac Na+ channel inactivation gate is required for modulation by protein kinase C. J Gen Physiol 108: 375eC379, 1996.

    Quandt FN. Modification of slow inactivation of single sodium channels by phenytoin in neuroblastoma cells. Mol Pharmacol 34: 557eC565, 1988.

    Quandt FN and Narahashi T. Modification of single Na+ channels by batrachotoxin. Proc Natl Acad Sci USA 79: 6732eC6736, 1982.

    Quionez M, DiFranco M, and Gonzeez F. Involvement of methionine residues in the fast inactivation mechanism of sodium current from toad skeletal muscle fibers. J Membr Biol 169: 83eC90, 1999.

    Rack M, Rubly N, and Waschow C. Effects of some chemical reagents on sodium current inactivation in myelinated nerve fibers of the frog. Biophys J 50: 557eC564, 1986.

    Ragsdale DS, McPhee JC, Scheuer T, and Catterall WA. Molecular determinants of state-dependent block of Na+ channels by local anesthetics. Science 265: 1724eC1728, 1994.

    Ragsdale DS, McPHee JC, Scheuer T, and Catterall WA. Common molecular determinants of local anesthetic, antiarrhythmic and anticonvulsant block of voltage-gated Na+ channels. Proc Natl Acd Sci USA 93: 9270eC9275, 1996.

    Rakowski RF, Gadsby DC, and De Weer P. Single ion occupancy and steady-state gating of Na channels in squid giant axon. J Gen Physiol 119: 235eC249, 2002.

    Raman IM and Bean BP. Inactivation and recovery of sodium currents in cerebellar Purkinje neurons: evidence for two mechanisms. Biophys J 80: 729eC737, 2001.

    Ramos E and O'Leary ME. State-dependent trapping of flecainide in the cardiac sodium channel. J Physiol 560: 37eC49, 2004.

    Rao S and Sikdar SK. Modification of a subunit of RIIA sodium channels by aconitine. Pfle筭ers Arch 439: 349eC355, 2000.

    Ravens U, Wettwer E, Pfeifer T, Himmel H, and Armah B. Characterization of the effects of the new inotropic agent BDF 9148 in isolated papillary muscles and myocytes of the guinea-pig heart. Br Pharmacol 104: 1019eC1023, 1991.

    Recio-Pinto E, Duch SD, Levinson SR, and Urban B. Purified and unpurified sodium channels from eel electroplax in planar lipid bilayers. J Gen Physiol 90: 375eC395, 1987.

    Ren D, Navarro B, Xu H, Yue L, Shi Q, and Clapham DE. A prokaryotic voltage-gated sodium channel. Science 294: 2372eC2375, 2001.

    Rhodes TH, Lossin C, Vanoye CG, Wang DW, and George AL. Noninactivating voltage-gated sodium channels in severe myoclonic epilepsy of infancy. Proc Natl Acad Sci USA 101: 11147eC11152, 2004.

    Richmond JE, Featherstone DE, Hartman HA, and Ruben PC. Slow inactivation in human cardiac sodium channels. Biophys J 74: 2945eC2952, 1998.

    Richmond JE, VanDeCarr D, Featherstone DE, George AL, and Ruben PC. Defective fast inactivation recovery and deactivation account for sodium channel myotonia in I1169V mutant. Biophys J 73: 1896eC1903, 1997.

    Rivolta I, Abriel H, Tateyama M, Liu H, Memmi M, Vardas P, Napolitano C, Priori SG, and Kass RS. Inherited Brugada and long QT-3 syndrome mutations of a single residue of the cardiac sodium channel confer distinct channel and clinical phenotypes. J Biol Chem 276: 30623eC30630, 2001.

    Rochat H, Bernard P, and Couraud F. Scorpion toxins: chemistry and mode of action. Adv Cytopharmacol 3: 325eC334, 1979.

    Roden DM, Balser JR, George AL, and Anderson ME. Cardiac ion channels. Annu Rev Physiol 64: 431eC475, 2002.

    Rogers JC, Qu Y, Tanada TN, Scheuer T, and Catterall WA. Molecular determinants of high affinity binding of -scorpion toxin and sea anemone toxin in the S3eCS4 extracellular loop in domain IV of the Na+ channel subunit. J Biol Chem 271: 15950eC15962, 1996.

    Rohl CA, Boeckman FA, Baker C, Scheuer T, Catterall WA, and Klevit RE. Solution structure of the sodium channel inactivation gate. Biochemistry 38: 855eC861, 1999.

    Rojas E and Armstrong CM. Sodium conductance activation without inactivation in pronase-perfused axons. Nature New Biol 229: 177eC178, 1971.

    Rojas E and Rudy B. Destruction of sodium conductance inactivation by a specific protease in perfused nerve fibres from Loligo. J Physiol 262: 501eC531, 1976.

    Romey G, Quast U, Pauron D, Frelin C, Renaud JF, and Lazdunski M. Na+ channels as sites of action of the cardiactive agent DPI 201eC106 with agonist and antagonist enantiomers. Proc Natl Acad Sci USA 84: 896eC900, 1987.

    Rouzaire-Dubois B and Dubois JM. Modification of electrophysiological and pharmacological properties of K channels in neuroblastoma cells induced by chloramine-T. Pfle筭ers Arch 416: 393eC397, 1990.

    Ruben PC, Starkus JG, and Rayner MD. Steady-state availability of sodium channels. Interactions between activation and slow inactivation. Biophys J 61: 941eC955, 1992.

    Rudy B. Slow inactivation of the sodium conductance in squid giant axons. Pronase resistance. J Physiol 283: 1eC21, 1978.

    Ruff RL. Single-channel basis of slow inactivation of Na+ channels in rat skeletal muscle. Am J Physiol Cell Physiol 271: C971eCC981, 1996.

    Ruff RL. Effects of temperature on slow and fast inactivation of rat skeletal muscle Na+ channels. Am J Physiol Cell Physiol 277: C937eCC947, 1999.

    Ruff RL, Simoncini L, and Ste筯mer W. Comparison between slow sodium channel inactivation in rat slow- and fast-twitch muscle. J Physiol 383: 339eC348, 1987.

    Ruff RL, Simoncini L, and Ste筯mer W. Slow sodium channel inactivation in mammalian muscle: a possible role in regulating excitability. Muscle Nerve 11: 502eC510, 1988.

    Rugiero F, Mistry M, Sage D, Black JA, Waxman SG, Crest M, Clerc N, Delmas P, and Gola M. Selective expression of a persistent tetrodotoxin-resistant Na+ current and Nav1.9 subunit in myenteric sensory neurons. J Neurosci 23: 2715eC2725, 2003.

    Saab CY, Cummins TR, Dib-Hajj SD, and Waxman SG. Molecular determinant of Nav1.8 sodium channel resistance to the venom from scorpion Leiurus quinquestriatus hebraeus. Neurosci Lett 331: 79eC82, 2002.

    Sah RL, Tsushima RG, and Backx PH. Effects of local anesthetics on Na+ channels containing equine hyperkalemic periodic paralysis mutation. Am J Physiol Cell Physiol 275: C389eCC400, 1998.

    Sahara Y, Gotoh M, Konno K, Miwa A, Tsubokawa H, Robinson HPC, and Kawai N. A new class of neurotoxin from wasp venom slows inactivation of sodium current. Eur J Neurosci 12: 1961eC1970, 2000.

    Saint DA, Ju YK, and Gage PW. A persistent sodium current in rat ventricular myocytes. J Physiol 453: 219eC231, 1992.

    Salceda E, Garateix A, and Soto E. The sea anemone toxins BgII and BgIII prolong the inactivation time course of the tetrodotoxin-sensitive sodium current in rat dorsal root ganglion neurons. J Pharmacol Exp Ther 303: 1067eC1074, 2002.

    Salgado VL, Yeh JZ, and Narahashi T. Voltage-dependent removal of sodium inactivation by N-bromoacetamide and pronase. Biophys J 47: 567eC571, 1985.

    Sato C, Ueno Y, Asal K, Takahashi K, Sato M, Engel A, and Fujiyoshi Y. The voltage-sensitive sodium channel is a bell-shaped molecule with several cavities. Nature 409: 1047eC1051, 2001.

    Schauf CL. Zonisamide enhances slow sodium inactivation in Myxicola. Brain Res 413: 185eC188, 1987.

    Schmidtmayer J. Behaviour of chemically modified sodium channels in frog nerve support a three-state model of inactivation. Pfle筭ers Arch 404: 21eC28, 1985.

    Schmidtmayer J. Voltage and temperature dependence of normal and chemically modified inactivation of sodium channels. Quantitative description by a cyclic three-state model. Pfle筭ers Arch 414: 273eC281, 1989.

    Schmidtmayer J, Stoye-Herzog M, and Ulbricht W. Rate of action of Anemonia sulcata toxin II on sodium channels in myelinated nerve fibres. Pfle筭ers Arch 394: 313eC319, 1982.

    Schmidtmayer J, Stoye-Herzog M, and Ulbricht W. Combined action of intraaxonal iodate and external sea anemone toxin ATX II on sodium channel inactivation of frog nerve fibres. Pfle筭ers Arch 398: 204eC209, 1983.

    Schmidtmayer J and Ulbricht W. Interaction of lidocaine and benzocaine in blocking sodium channels. Pfle筭ers Arch 387: 47eC54, 1980.

    Schneider MF and Dubois JM. Effects of benzocaine on the kinetics of normal and batrachotoxin-modified Na channels in frog node of Ranvier. Biophys J 50: 523eC530, 1986.

    Scholtysik G, Quast U, and Schaad A. Evidence for different receptor sites for the novel cardiotonic S-DPI 201eC106, ATX II and veratridine on the cardiac sodium channel. Eur J Pharmacol 125: 111eC118, 1986.

    Scholz A. Mechanisms of (local) anaesthetics on voltage-gated sodium and other ion channels. Br J Anaesth 89: 52eC61, 2002.

    Schreibmayer W. Isoform diversity and modulation of sodium channels by protein kinases. Cell Physiol Biochem 9: 187eC200, 1999.

    Schreibmayer W, Kazerani H, and Tritthart HA. A mechanistic interpretation of the action of toxin II from Anemonia sulcata on the cardiac sodium channel. Biochim Biophys Acta 901: 273eC282, 1987.

    Schreibmayer W, Tritthart HA, and Schindler H. The cardiac sodium channel shows a regular substate pattern indicating synchronized activity of several pathways instead of one. Biochim Biophys Acta 986: 172eC186, 1989.

    Schwarz JR. The effect of temperature on Na currents in rat myelinated nerve fibres. Pfle筭ers Arch 406: 397eC404, 1986.

    Schwarz W. Temperature experiments on nerve and muscle membranes of frogs. Pfle筭ers Arch 382: 27eC34, 1979.

    Shah BS, Stevens EB, Pinnock RD, Dixon AK, and Lee K. Developmental expression of the novel voltage-gated sodium channel auxiliary subunit 3, in rat CNS. J Physiol 534: 763eC776, 2001.

    Sheets MF and Hanck DA. Gating of skeletal and cardiac muscle sodium channels in mammalian cells. J Physiol 514: 425eC436, 1999.

    Sheets MF, Kyle JW, and Hanck DA. The role of the putative inactivation lid in sodium channel gating current immobilization. J Gen Physiol 115: 609eC619, 2000.

    Sheets MF, Kyle JW, Kallen RG, and Hanck DA. The Na channel voltage sensor associated with inactivation is localized to the external charged residues of domain IV, S4. Biophys J 77: 747eC757, 1999.

    Shichor I, Fainzilber M, Pelhate M, Malecot CO, Zlotkin E, and Gordon D. Interactions of -conotoxins with alkaloid neurotoxins reveal differences between silent and effective binding sites on voltage-sensitive sodium channels. J Neurochem 67: 2451eC2460, 1996.

    Shon KJ, Hasson A, Spira ME, Cruz LJ, Gray WR, and Olivera BM. -Conotoxin GmVIA, a novel peptide from the venom of Conus gloriamaris. Biochemistry 33: 11420eC11425, 1994.

    Shoukimas JJ and French RF. Incomplete inactivation of sodium currents in nonperfused squid axon. Biophys J 32: 857eC862, 1980.

    Sigel E. Effects of veratridine on single neuronal sodium channels expressed in Xenopus oocytes. Pfle筭ers Arch 410: 112eC120, 1987.

    Simard JM, Meves H, and Watt DD. Neurotoxins in venom from the North American scorpion, Centruroides sculpturatus Ewing. In: Natural Toxins: Toxicology, Chemistry and Safety, edited by Keeler R. F., N. B. Mandava, and A. T. Tu. Fort Collins, CO: Alaken, 1992, p. 236eC263.

    Simoncini L and Ste筯mer W. Slow sodium channel inactivation in rat fast-twitch muscle. J Physiol 383: 327eC337, 1987.

    Sirota FL, Pascutti PG, and Anteneodo C. Molecular modeling and dynamics of the sodium channel inactivation gate. Biophys J 82: 1207eC1215, 2002.

    Smith MR and Goldin AL. Interaction between the sodium channel inactivation linker and domain III S4eCS5. Biophys J 73: 1885eC1895, 1997.

    Soderlund DM, Smith TJ, and Lee SH. Differential sensitivity of sodium channel isoforms and sequence variants to pyrethroid insecticides. Neurotoxicology 21: 127eC138, 2000.

    Song JH and Narahashi T. Modulation of sodium channels of rat cerebellar Purkinje neurons by the pyrethroid tetramethrin. J Pharmacol Exp Ther 277: 445eC453, 1996.

    Spampanato J, Escayg A, Meisler MH, and Goldin AL. Functional effects of two voltage-gated sodium channel mutations that cause generalized epilepsy with febrile seizures plus type 2. J Neurosci 21: 7481eC7490, 2001.

    Stmpfli R. Intra-axonal iodate inhibits sodium inactivation. Experientia 30: 505eC508, 1974.

    Starkus JG, Heggeness ST, and Rayner MD. Kinetic analysis of sodium channel block by internal methylene blue in pronased crayfish giant axons. Biophys J 46: 205eC218, 1984.

    Starmer CF, Grant AO, and Strauss AC. Mechanisms of use-dependent block of sodium channels in excitable membranes by local anesthetics. Biophys J 46: 15eC27, 1984.

    Stephens GJ and Robertson B. Inactivation of the cloned potassium channel mouse Kv1.1 by the human Kv34 "ball" peptide and its chemical modification. J Physiol 484: 1eC13, 1995.

    Sternberg D, Maisonobe T, Jurkat-Rott K, Nicole S, Launay E, Chauveau D, Tabti N, Lehmann-Horn F, Hainque B, and Fontaine B. Hypokalemic periodic paralsysis type 2 caused by mutations at codon 672 in the muscle sodium channel gene SCN4A. Brain 124: 1091eC1099, 2001.

    Stevens EB, Cox PJ, Shah BS, Dixon AK, Richardson PJ, Pinnock RD, and Lee K. Tissue distribution and functional expression of the human voltage-gated sodium channel 3 subunit. Pfle筭ers Arch 441: 481eC488, 2001.

    Stimers JR, Bezanilla F, and Taylor RE. Sodium channel activation in the squid giant axon. Steady state properties. J Gen Physiol 85: 65eC82, 1985.

    Strachan C, Lewis RJ, and Nicholson GM. Differential actions of pacific ciguatoxin-1 on sodium channel subtypes in mammalian sensory neurons. J Pharmacol Exp Ther 288: 379eC388, 1999.

    Strichartz GR and Wang GK. Rapid voltage-dependent dissociation of scorpion -toxins coupled to Na channel inactivation in amphibian myelinated nerve. J Gen Physiol 88: 413eC435, 1986.

    Struyk AF and Cannon SC. Slow inactivation does not block the aqueous accessibility to the outer pore of voltage-gated Na channels. J Gen Physiol 120: 509eC516, 2002.

    Struyk AF, Scoggan KA, Bulman DE, and Cannon CS. The human skeletal muscle Na channel mutation R669H associated with hypokalemic periodic paralysis enhances slow inactivation. J Neurosci 20: 8610eC8617, 2000.

    Ste筯mer W, Conti F, Suzuki H, Wang X, Noda M, Yahagi N, Kubo H, and Numa S. Structural parts involved in activation and inactivation of the sodium channel. Nature 339: 597eC603, 1989.

    Sudarslal S, Majumdar S, Ramasamy P, Dhawan R, Pal PP, Ramaswami M, Lala AK, Sikdar SK, Sarma SP, Krishnan KS, and Balaram P. Sodium channel modulating activity in a -conotoxin from an Indian marine snail. FEBS Lett 553: 209eC212, 2003.

    Szeto TH, Birinyi-Strachan LC, Smith R, Connor M, Christie MJ, King GF, and Nicholson GM. Isolation and pharmacological characterization of -atracotoxin-Hv1b, a vertebrate-selective sodium channel toxin. FEBS Lett 470: 293eC299, 2000.

    Tabarean IV and Narahashi T. Potent modulation of tetrodotoxin-sensitive and tetrodotoxin-resistant sodium channels by the type II pyrethroid deltamethrin. J Pharmacol Exp Ther 284: 958eC965, 1998.

    Taddese A and Bean BP. Subthreshold sodium current from rapidly inactivating sodium channels drives spontaneous firing of tuberomammillary neurons. Neuron 33: 587eC600, 2002.

    Takahashi MP and Cannon SC. Enhanced slow inactivation by V445M: a sodium channel mutation associated with myotonia. Biophys J 76: 861eC868, 1999.

    Takahashi MP and Cannon SC. Mexiletine block of disease-associated mutations in S6 segments of the human skeletal muscle Na+ channel. J Physiol 537: 701eC714, 2001.

    Tanguy J and Yeh JZ. Batrachotoxin uncouples gating charge immobilization from fast Na inactivation in squid giant axons. Biophys J 54: 719eC730, 1988.

    Tanguy J and Yeh JZ. BTX modification of Na channels in squid axons. I. State dependence of BTX action. J Gen Physiol 97: 499eC519, 1991.

    Tatebayashi H and Narahashi T. Differential mechanism of action of the pyrethroid tetramethrin on tetrodotoxin-sensitive and tetrodotoxin-resistant sodium channels. J Pharmacol Exp Ther 270: 595eC603, 1994.

    Tateyama M, Liu H, Yang AS, Cormier JW, and Kass RS. Structural effects of an LQT-3 mutation on heart Na+ channel gating. Biophys J 86: 1843eC1851, 2004.

    Taylor CP. Na+ currents that fail to inactivate. Trends Neurosci 16: 455eC460, 1993.

    Tejedor FJ and Catterall WA. Site of covalent attachment of -scorpion toxin derivatives in domain I of the sodium channel subunit. Proc Natl Acad Sci USA 85: 8742eC8746, 1988.

    Terlau H and Olivera B. Conus venoms: a rich source of novel ion channel-targeted peptides. Physiol Rev 84: 41eC68, 2004.

    Thomsen WJ and Catterall WA. Localization of the receptor site for -scorpion toxins by antibody mapping: implications for sodium channel topology. Proc Natl Acad Sci USA 86: 10161eC10165, 1989.

    Todt H, Dudley SC, Kyle JW, French RJ, and Fozzard HA. Ultra-slow inactivation in e? Na+ channels is produced by a structural rearrangement of the outer vestibule. Biophys J 76: 1335eC1345, 1999.

    Toib A, Lyakhov V, and Marom S. Interaction between duration of activity and time course of recovery from slow inactivation in mammalian brain Na+ channels. J Neurosci 18: 1893eC1903, 1998.

    Tomaselli GF, Chiamvimonvat N, Nuss HB, Balser JR, Perez-Garce猘 MT, Xu RH, Orias DW, Backx PH, and Marban E. A mutation in the pore of the sodium channel alters gating. Biophys J 68: 1814eC1827, 1995.

    Townsend C and Horn R. Effect of alkali metal cations on slow inactivation of cardiac Na+ channels. J Gen Physiol 110: 23eC33, 1997.

    Trainer VL, McPhee JC, Boutelef-Bochan H, Baker C, Scheuer T, Babin D, Demoute JP, Guedin D, and Catterall WA. High affinity binding of pyrethroids to the subunit of brain sodium channels. Mol Pharmacol 51: 651eC657, 1997.

    Tsushima RG, Kelly JE, Salata JJ, Liberty KN, and Wasserstrom JA. Modification of cardiac Na+ current by RWJ 24517 and its enantiomers in guinea pig ventricular myocytes. J Pharmacol Exp Ther 291: 845eC855, 1999.

    Ulbricht W. The effect of veratridine on excitable membranes of nerve and muscle. Ergeb Physiol 61: 18eC71, 1969.

    Ulbricht W. The inactivation of sodium channels in the node of Ranvier and its chemical modification. In: Ion Channels, edited by Narahashi T. New York: Plenum, 1990, vol. 2, p 123eC268.

    Ulbricht W. Effects of veratridine on sodium currents and fluxes. Rev Physiol Biochem Pharmacol 133: 1eC54, 1998.

    Ulbricht W and Schmidtmayer J. Modification of sodium channels in myelinated nerve by Anemonia sulcata toxin II. J Physiol 77: 1103eC1111, 1981.

    Ulbricht W and Stoye-Herzog M. Distinctly different rates of benzocaine action on sodium channels of Ranvier nodes kept open by chloramine-T and veratridine. Pfle筭ers Arch 402: 439eC445, 1984.

    Ulbricht W and Wagner HH. The influence of pH on equilibrium effects of tetrodotoxin on myelinated nerve fibres of Rana esculenta. J Physiol 252: 159eC184, 1975.

    Valenzuela C, Snyders DJ, Bennett PB, Tamargo J, and Hondeghem LM. Stereoselective block of cardiac sodium channels by bupivacaine in guinea pig ventricular myocytes. Circulation 92: 3014eC3024, 1995.

    Vandenberg CA and Horn R. Inactivation viewed through single sodium channels. J Gen Physiol 84: 535eC561, 1984.

    Vassilev PM, Scheuer T, and Catterall WA. Identification of an intracellular peptide segment involved in sodium channel inactivation. Science 241: 1658eC1661, 1988.

    Vassilev PM, Scheuer T, and Catterall WA. Inhibition of inactivation of single sodium channels by a site-directed antibody. Proc Natl Acad Sci USA 86: 8147eC8151, 1989.

    Vedantham V and Cannon SC. Slow inactivation does not affect movement of the fast inactivation gate in voltage-gated Na+ channels. J Gen Physiol 111: 83eC93, 1998.

    Vedantham V and Cannon SC. The position of the fast inactivation gate during lidocaine block of voltage-gated Na+ channels. J Gen Physiol 113: 7eC16, 1999.

    Veldkamp MV, Viswanathan PC, Bezzina C, Baartscheer A, Wilde AAM, and Balser JR. Two distinct congenital arrhythmias evoked by a multidysfunctional Na+ channel. Circ Res 86: e91eCe97, 2000.

    Vijayaragavan K, O'Leary ME, and Chahine M. Gating properties of Nav1.7 and Nav18 peripheral nerve sodium channels. J Neurosci 21: 7909eC7918, 2001.

    Vijverberg HPM and De Weille JR. The interaction of pyrethroids with voltage-dependent Na channels. Neurotoxicology 6: 23eC34, 1985.

    Vijverberg HPM, Puaron D, and Lazdunski M. The effect of Tityus serrulatus scorpion toxin on Na channels in neuroblastoma cells. Pfle筭ers Arch 401: 297eC303, 1984.

    Vijverberg HPM, Van der Salm JM, and Van den Bercken J. Similar mode of action of pyrthroids and DDT on sodium channel gating in myelinated nerves. Nature 295: 601eC603, 1982.

    Vilin YY, Fujimoto E, and Ruben PC. A single residue differentiates between human cardiac and skeletal muscle Na+ channel slow inactivation. Biophys J 80: 2221eC2230, 2001.

    Vilin YY, Makita M, George AL, and Ruben PC. Structural determinants of slow inactivation in human cardiac and skeletal muscle sodium channels. Biophys J 77: 1384eC1393, 1999.

    Vilin YY and Ruben PC. Slow inactivation in voltage-gated sodium channels. Molecular substrates and contributions to channelopathies. Cell Biochem Biophys 35: 171eC190, 2001.

    Vincent JP, Balerna M, Barhanin J, Fosset M, and Lazdunski M. Binding of sea anemone toxin to receptor sites associated with gating system of sodium channels in synaptic nerve endings in vitro. Proc Natl Acad Sci USA 77: 1646eC1650, 1980.

    Wallace RH, Wang DW, Singh R, Scheffer IE, George AL, Phillips HA, Saar K, Reis A, Johnson EW, Sutherland GR, Berkovic SF, and Mulley JC. Febrile seizures and generalized epilepsy associated with a mutation in the Na+-channel 1 subunit gene SCN1B. Nat Genet 19: 366eC370, 1998.

    Wang DW, VanDeCarr D, Ruben PC, George AL, and Bennett PB. Functional consequences of a domain 1/S6 segment sodium channel mutation associated with painful congenital myotonia. FEBS Lett 448: 231eC234, 1999.

    Wang DW, Yazawa K, George AL, and Bennett PB. Characterization of human cardiac Na+ channel mutations in the congenital long QT syndrome. Proc Natl Acad Sci USA 93: 13200eC13205, 1996.

    Wang G, Dugas M, Armah BI, and Honerjger P. Sodium channel comodification with full activator reveals veratridine reaction dynamics. Mol Pharmacol 37: 144eC148, 1990.

    Wang GK. Modification of sodium channel inactivation in single myelinated nerve fibers by methionine-reactive chemicals. Biophys J 46: 121eC124, 1984.

    Wang GK. Irreversible modification of sodium channel inactivation in toad myelinated nerve fibres by the oxidant chloramine-T. J Physiol 346: 127eC141, 1984.

    Wang GK, Brodwick MS, and Eaton DC. Removal of sodium channel inactivation in squid axon by the oxidant chloramine-T. J Gen Physiol 86: 289eC302, 1985.

    Wang GK, Brodwick MS, Eaton DC, and Strichartz GR. Inhibition of sodium currents by local anesthetics in chloramine-T-treated squid axons. The role of channel activation. J Gen Physiol 89: 645eC667, 1987.

    Wang GK, Quan C, Seaver M, and Wang SY. Modification of wild-type and batrachotoxin-resistant muscle e? Na+ channels by veratridine. Pfle筭ers Arch 439: 705eC713, 2000.

    Wang GK, Quan C, and Wang SY. A common local anesthetic receptor for benzocaine and etidocaine in voltage-gated e? Na+ channels. Pfle筭ers Arch 435: 293eC302, 1998.

    Wang GK, Quan C, and Wang SY. Local anesthetic block of batrachotoxin-resistant muscle Na+ channels. Mol Pharmacol 54: 389eC396, 1998.

    Wang GK and Strichartz G. Kinetic analysis of the action of Leiurus scorpion -toxin on ionic currents in myelinated nerve. J Gen Physiol 86: 739eC762, 1985.

    Wang GK and Wang SY. Modification of cloned brain Na+ channels by batrachotoxin. Pfle筭ers Arch 427: 309eC316, 1994.

    Wang GK and Wang SY. Modification of human cardiac sodium channels gating by UVA light. J Membr Biol 189: 153eC165, 2002.

    Wang GK and Wang SY. Veratridine block of rat skeletal muscle Nav1.4 sodium channels in the inner vestibule. J Physiol 548: 667eC673, 2003.

    Wang SY, Barile M, and Wang GK. Disparate role of Na+ channel D2eCS6 residues in batrachotoxin and local anesthetic action. Mol Pharmacol 59: 1100eC1107, 2001.

    Wang SY, Nau C, and Wang GK. Residues in Na+ channel D3eCS6 segment modulate both batrachotoxin and local anesthetic affinities. Biophys J 79: 1379eC1387, 2000.

    Wang SY and Wang GK. Slow inactivation of muscle e? Na+ channels in permanently transfected mammalian cells. Pfle筭ers Arch 432: 692eC699, 1996.

    Wang SY and Wang GK. A mutation in segment I-S6 alters slow inactivation of sodium channels. Biophys J 72: 1633eC1640, 1997.

    Wang SY and Wang GK. Point mutations in segment I-S6 render voltage-gated Na+ channels resistant to batrachotoxin. Proc Natl Acad Sci USA 95: 2653eC2658, 1998.

    Wang SY and Wang GK. Batrachotoxin-resistant Na+ channels derived from point mutations in transmembrane segment D4eCS6. Biophys J 76: 3141eC3149, 1999.

    Wang SY and Wang GK. Voltage-gated sodium channels as primary targets of diverse lipid-soluble neurotoxins. Cell Signal 15: 151eC159, 2003.

    Warashina A and Fujita S. Effects of sea anemone toxins on the sodium inactivation process in crayfish axons. J Gen Physiol 81: 305eC323, 1983.

    Warashina A, Jiang ZY, and Ogura T. Potential-dependent action of Anemonia sulcata toxins III and IV on sodium channels in crayfish giant axons. Pfle筭ers Arch 411: 88eC93, 1988.

    Watt DD and Simard JM. Neurotoxic proteins in scorpion venoms. J Toxicol Toxin Rev 3: 181eC221, 1984.

    Wei J, Wang DW, Alings M, Fish F, Wathen M, Roden DM, and George AL. Congenital long-QT syndrome caused by a novel mutation in a conserved acidic domain of the cardiac Na+ channel. Circulation 99: 3165eC3171, 1999.

    West JW, Numann R, Murphy BJ, Scheuer T, and Catterall WA. A phosphorylation site in the Na+ channel required for modulation by protein kinase C. Science 254: 866eC868, 1991.

    West JW, Numann R, Murphy BJ, Scheuer T, and Catterall WA. Phosphorylation of a conserved protein kinase C site is required for modulation of Na+ currents in transfected Chinese hamster ovary cells. Biophys J 62: 31eC33, 1992.

    West JW, Patton DE, Scheuer T, Wang Y, Goldin AL, and Catterall WA. A cluster of hydrophobic amino acid residues required for fast Na+ channel inactivation. Proc Natl Acad Sci USA 89: 10910eC10914, 1992.

    Wright SN. Irreversible block of human heart (hH1) sodium channels by the plant alkaloid lappaconitine. Mol Pharmacol 59: 183eC192, 2001.

    Wright SN, Wang SY, and Wang GK. Lysine point mutations in Na+ channel D4eCS6 reduce inactivated channel block by local anesthetics. Mol Pharmacol 54: 737eC739, 1998.

    Wu CH and Narahashi T. Mechanism of action of novel marine neurotoxins on ion channels. Annu Rev Pharmacol Toxicol 28: 141eC161, 1988.

    Xiao YF, Ke Q, Wang SY, Auktor K, Yang Y, Wang GK, Morgan JP, and Leaf A. Single point mutations affect fatty acid block of human myocardial sodium channel subunit Na+ channels. Proc Natl Acad Sci USA 98: 3606eC3611, 2001.

    Xiong W, Li RA, Tian Y, and Tomaselli GF. Molecular motions of the outer ring of charge of the sodium channel: do they couple to slow inactivation J Gen Physiol 122: 323eC332, 2003.

    Yakehiro M, Yuki T, Yamaoka K, Furue T, Mori Y, Imoto K, and Seyama I. An analysis of the variations in potency of grayanotoxin analogs in modifying frog sodium channels of differing subtypes. Mol Pharmacol 58: 692eC700, 2000.

    Yamamoto D, Yeh JZ, and Narahashi T. Ion permeation and selectivity of squid axon sodium channels modified by tetramethrin. Brain Res 272: 193eC197, 1986.

    Yang N, George AL, and Horn R. Probing the outer vestibule of a sodium channel voltage sensor. Biophys J 73: 2260eC2268, 1997.

    Yang N, Ji S, Zhou M, Pteek LJ, Barchi RL, Horn R, and George AL. Sodium channel mutations in paramyotonia congenita exhibit similar biophysical phenotypes in vitro. Proc Natl Acad Sci USA 91: 12785eC12789, 1994.

    Yang YC and Kuo CC. The position of the fourth segment of domain 4 determines status of the inactivation gate in Na+ channels. J Neurosci 23: 4922eC4930, 2003.

    Yarov-Yarovoy V, Brown J, Sharp EM, Clare JJ, Scheuer T, and Catterall WA. Molecular determinants of voltage-dependent gating and binding of pore-blocking drugs in transmembrane segment IIIS6 of the Na+ channel subunit. J Biol Chem 276: 20eC27, 2001.

    Yarov-Yarovoy V, McPhee JC, Idsvoog D, Pate C, Scheuer T, and Catterall WA. Role of amino acid residues in transmembrane segments IS6 and IIS6 of the Na+ channel subunit in voltage-dependent gating and drug block. J Biol Chem 277: 35393eC35401, 2002.

    Yeh JZ. Blockage of sodium channels by stereo-isomers of local anesthetics. Prog Anesthesiol 2: 35eC44, 1980.

    Yu FH, Westenbroek RE, Silos-Santiago I, McCormick KA, Lawson D, Ge P, Ferriera H, Lilly J, DiStefano PS, Catterall WA, Scheuer T, and Curtis R. Sodium channel 4, a new disulfide-linked auxiliary subunit with similarity to 2. J Neurosci 23: 7577eC7585, 2003.

    Yuill KH, Convery MK, Dooley PC, Doggrell SA, and Hancox JC. Effects of BDF 9198 on action potentials and ionic currents from guinea-pig isolated ventricular myocytes. Br J Pharmacol 130: 1753eC1766, 2000.

    Yuki T, Yamaoka K, Yakehiro M, and Seyama I. State-dependent action of grayanotoxin I on Na+ channels in frog ventricular myocytes. J Physiol 534: 777eC790, 2001.

    Zaborovskaya LS and Khodorov BI. The role of inactivation in the cumulative blockage of voltage-dependent sodium channels by local anesthetics and antiarrhythmics. Gen Physiol Biophys 3: 517eC520, 1984.

    Zhang Z, Xu Y, Dong PH, Sharma D, and Chiamvimonvat N. A negatively charged residue in the outer mouth of rat sodium channel determines the gating kinetics of the channel. Am J Physiol Cell Physiol 284: C1247eCC1254, 2003.

    Zhao Y, Yarov-Yarovoy V, Scheuer T, and Catterall WA. A gating hinge in Na+ channels: a molecular switch for electrical signaling. Neuron 41: 859eC865, 2004.

    Zimmer T and Benndorf K. The human heart and rat brain IIA Na+ channels interact with different molecular regions. J Gen Physiol 120: 887eC895, 2002.

    Zimmer T, Biskup C, Bollensdorff C, and Benndorf K. The 1 subunit but not the 2 subunit colocalizes with the human heart Na+ channel (hH1) already within the endoplasmic reticulum. J Membr Biol 186: 13eC21, 2002.

    Zlotkin E. The insect voltage-gated sodium channel as target of insecticides. Annu Rev Entomol 44: 429eC455, 1999.

    Zygmunt AC, Eddlestone GT, Thomas GP, Nesterenko VV, and Antzelevitch D. Larger late conductance in M cells contributes to electrical heterogeneity in canine ventricle. Am J Physiol Heart Circ Physiol 281: H689eCH697, 2001.(Werner Ulbricht)