首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The β-subunit of the voltage-sensitive K+ channels shares 15–30% amino acid identity with the sequences of aldo–keto reductases (AKR) genes. However, the AKR properties of the protein remain unknown. To begin to understand its oxidoreductase properties, we examine the pyridine coenzyme binding activity of the protein in vitro. The cDNA of Kvβ2.1 from rat brain was subcloned into a prokaryotic expression vector and overexpressed in Escherichia coli. The purified protein was tetrameric in solution as determined by size exclusion chromatography. The protein displayed high affinity binding to NADPH as determined by fluorometric titration. The KD values for NADPH of the full-length wild-type protein and the N-terminus deleted protein were 0.1±0.007 and 0.05±0.006 M, respectively — indicating that the cofactor binding domain is restricted to the C-terminus, and is not drastically affected by the absence of the N-terminus amino acids, which form the ball and chain regulating voltage-dependent inactivation of the α-subunit. The protein displayed poor affinity for other coenzymes and the corresponding values of the KD for NADH and NAD were between 1–3 μM whereas the KD for FAD was >10 μM. However, relatively high affinity binding was observed with 3-acetyl pyridine NADP, indicating selective recognition of the 2′ phosphate at the binding site. The selectivity of Kvβ2.1 for NADPH over NADP may be significant in regulating the K+ channels as a function of the cellular redox state.  相似文献   

2.
Connexin 26 (Cx26) and connexin 30 (Cx30) form hemichannels that release ATP from the endolymphatic surface of cochlear supporting and epithelial cells and also form gap junction (GJ) channels that allow the concomitant intercellular diffusion of Ca2+ mobilizing second messengers. Released ATP in turn activates G-protein coupled P2Y2 and P2Y4 receptors, PLC-dependent generation of IP3, release of Ca2+ from intracellular stores, instigating the regenerative propagation of intercellular Ca2+ signals (ICS). The range of ICS propagation is sensitive to the concentration of extracellular divalent cations and activity of ectonucleotidases. Here, the expression patterns of Cx26 and Cx30 were characterized in postnatal cochlear tissues obtained from mice aged between P5 and P6. The expression gradient along the longitudinal axis of the cochlea, decreasing from the basal to the apical cochlear turn (CT), was more pronounced in outer sulcus (OS) cells than in inner sulcus (IS) cells. GJ-mediated dye coupling was maximal in OS cells of the basal CT, inhibited by the nonselective connexin channel blocker carbenoxolone (CBX) and absent in hair cells. Photostimulating OS cells with caged inositol (3,4,5) tri-phosphate (IP3) resulted in transfer of ICS in the lateral direction, from OS cells to IS cells across the hair cell region (HCR) of medial and basal CTs. ICS transfer in the opposite (medial) direction, from IS cells photostimulated with caged IP3 to OS cells, occurred mostly in the basal CT. In addition, OS cells displayed impressive rhythmic activity with oscillations of cytosolic free Ca2+ concentration ([Ca2+]i) coordinated by the propagation of Ca2+ wavefronts sweeping repeatedly through the same tissue area along the coiling axis of the cochlea. Oscillations evoked by uncaging IP3 or by applying ATP differed greatly, by as much as one order of magnitude, in frequency and waveform rise time. ICS evoked by direct application of ATP propagated along convoluted cellular paths in the OS, which often branched and changed dynamically over time. Potential implications of these findings are discussed in the context of developmental regulation and cochlear pathophysiology.  相似文献   

3.
The dependence of the efficacy of the influence of a kava-pyrone (±)-kavain (330 µM) on the frequency of activation of Na+ channels and voltage dependence of the effects of (±)-kavain on the rate of inactivation of these channels were studied in experiments on isolated neurons from the rat hippocampus. In all series of experiments, the holding potential equalled –100 mV. The efficacy of (±)-kavain-induced blockade of Na+ channels was independent of the frequency of stimulation within the range up to 10/sec. In the control experiments, the rate of inactivation increased with the rise of depolarization from –40 mV to +30 mV, and then the saturation effect was observed. At the membrane potential of –40 mV, the rate of (±)-kavain-evoked inactivation increased approximately by a factor of 2.5. At the more positive shifts of the membrane potential, the efficacy of the effects of (±)-kavain on the rate of inactivation became noticeably reduced, and at +30 mV (±)-kavain exerted no distinct influence on this parameter.Neirofiziologiya/Neurophysiology, Vol. 28, No. 6, pp. 312–315, November–December, 1996.  相似文献   

4.
σ-Receptors are integral membrane proteins that have been implicated in a number of biological functions, many of which involve the modulation of ion channels. A wide range of synthetic ligands activate σ-receptors, but endogenous σ-receptor ligands have proven elusive. One endogenous ligand, dimethyltryptamine (DMT), has been shown to act as a σ-receptor agonist. Progesterone and other steroids bind σ-receptors, but the functional consequences of these interactions are unclear. Here we investigated progesterone binding to σ(1)- and σ(2)-receptors and evaluated its effect on σ-receptor-mediated modulation of voltage-gated Na(+) channels. Progesterone binds both σ-receptor subtypes in liver membranes with comparable affinities and blocks photolabeling of both subtypes in human embryonic kidney 293 cells that stably express the human cardiac Na(+) channel Na(v)1.5. Patch-clamp recording in this cell line tested Na(+) current modulation by the σ-receptor ligands ditolylguanidine, PB28, (+)SKF10047, and DMT. Progesterone inhibited the action of these ligands to varying degrees, and some of these actions were reduced by σ(1)-receptor knockdown with small interfering RNA. Progesterone inhibition of channel modulation by drugs was consistent with stronger antagonism of σ(2)-receptors. By contrast, progesterone inhibition of channel modulation by DMT was consistent with stronger antagonism of σ(1)-receptors. Progesterone binding to σ-receptors blocks σ-receptor-mediated modulation of a voltage-gated ion channel, and this novel membrane action of progesterone may be relevant to changes in brain and cardiovascular function during endocrine transitions.  相似文献   

5.
6.
Ion channels exert a vital role in the dialogue between male and female gametes and thus in the generation of new individuals in many species. Intracellular Ca2+ is possibly the key messenger between gametes. Different Ca2+-permeable channels have been detected in the plasma membrane and in the organelle-like acrosome membrane of sperm, which play vital roles in determining sperm fertilizing ability. Recent reports from several laboratories have adequately documented that the Ca2+-permeable channels of a sperm control a variety of functions ranging from motility to the acrosome reaction. In this article, we have reviewed the data from our and other laboratories, and have documented the mechanisms of different Ca2+-permeable channels involved in the fertilization event.  相似文献   

7.
Voltage-gated K+ and Na+ channels are involved in diverse physiological processes including excitability of heart, muscular and neuronal cells, as well as release of hormones and neurotransmitters. These channels have modular structure and contain five membrane domains: four voltage-sensing domains (VSDs) and one pore domain. VSDs of different channels contain unique ligand-binding sites and are considered as potential pharmacological targets. Modular organization of ion channels points to the possibility of NMR structural studies of isolated VSDs apart from the pore. Here, the feasibility of such studies is considered by the example of VSD of human Kv2.1 channel and VSD-I of human Nav1.4 channel. Cell-free protein expression systems based on the S30 bacterial extract from E. coli, which allow us to produce milligram quantities of VSD samples, including their analogues labeled with stable isotopes, were developed. The choice of membrane- mimicking media that provide long-term stability of the native structure of the membrane protein and high-quality of NMR spectra is a crucial step in NMR studies. Screening of various environments showed that the domains of the Kv2.1 and Nav1.4 channels are unstable in media containing phospholipids: micelles of short-chain lipid DC7PC and lipid-detergent bicelles based on zwitterionic or anionic saturated lipids (DMPC and DMPG). It was demonstrated that the optimal media for NMR studies are the mixtures of zwitterionic and weakly cationic detergents (FOS-12/LDAO). The VSD sample of the Nav1.4 channel in FOS- 12/LDAO environment aggregated irreversibly within a few days despite the high-quality spectra. It is likely that VSDs of human K+ and Na+ channels are not completely autonomous membrane domains and the contacts with other domains of the channel are required for their stabilization.  相似文献   

8.
The macroscopic algae Characeae are usually assumed to occur in waterbodies supplied by groundwater with low phosphate content, but the indicative value of the species is seldom defined in bibliography. Former braided channels of the Rhône river are supplied with groundwater originating from the main channel (seepage) or from hillslope aquifer. The aim of the present paper was to determine if it possible to use the Characeae as indicators of physicochemical characteristies of water in order to assess the origin of groundwater supplying former river channels. Four former braided channels of the Rhône River colonized by Characeae were investigated, and the physico-chemical characteristics of i) the channels, ii) the groundwater and iii) the river were measured over a period of several months. Species are arranged along a gradient of conductivity, alkalinity, ammonium and phosphate content of the water. Charophyte species can indicate the origin of groundwater, either seepage or hillslope nutrient-poor aquifer, and integrate both the average value of the chemical parameter, and their variations. C. hispida occurs in a nutrient-poor channel mainly supplied by highly calcareous groundwater coming from hillslope aquifer. Chara major has requirements close to those of C. hispida, but is more tolerant to periodic inputs of nutrients. C. vulgaris and N. syncarpa both tolerate mesotrophic waters originating from both hillslope aquifer and seepage, and C. globularis is associated to a channel mainly supplied by mesotrophic to eutrophic river seepage.  相似文献   

9.
In excitable cells, the initiation of the action potential results from the opening of voltage-gated sodium channels. These channels undergo a series of conformational changes between open, closed, and inactivated states. Many models have been proposed for the structural transitions that result in these different functional states. Here, we compare the crystal structures of prokaryotic sodium channels captured in the different conformational forms and use them as the basis for examining molecular models for the activation, slow inactivation, and recovery processes. We compare structural similarities and differences in the pore domains, specifically in the transmembrane helices, the constrictions within the pore cavity, the activation gate at the cytoplasmic end of the last transmembrane helix, the C-terminal domain, and the selectivity filter. We discuss the observed differences in the context of previous models for opening, closing, and inactivation, and present a new structure-based model for the functional transitions. Our proposed prokaryotic channel activation mechanism is then compared with the activation transition in eukaryotic sodium channels.Eukaryotic sodium channels are complex, multidomain proteins that assemble into pseudotetrameric structures. Sodium channels are also present in some prokaryotes (Ren et al., 2001), where they are homologous but simpler, single-domain proteins that assemble to form functional homotetramers (Nurani et al., 2008). Both eukaryotic and prokaryotic sodium channels consist of voltage-sensor (VS) and pore regions that are physically separated in three dimensions. Pore-only channels can be constructed, which are capable of undergoing opening and closing transitions and support ion flux in a manner similar to that of intact channels (McCusker et al., 2011; Shaya et al., 2011). Although the kinetics of their processes of conversion between states are different, both eukaryotic and prokaryotic channels exhibit activation, recovery, and slow inactivation transitions (Koishi et al., 2004; Charalambous and Wallace, 2011); in addition, eukaryotic channels undergo a fast inactivation process that is not observed in prokaryotic channels (Pavlov et al., 2005). Prokaryotic sodium channels have also been shown to be blocked by eukaryotic sodium channel antagonists (Bagnéris et al., 2014).Many models have been proposed for the structural changes that result in the open, closed, and slow inactivated states of prokaryotic sodium channels (e.g., Kuzmenkin et al., 2004; Zhao et al., 2004; Pavlov et al., 2005; Shafrir et al., 2008). Although there are, as yet, no crystal structures of eukaryotic sodium channels, the structures of several prokaryotic sodium channel orthologues have recently been determined by x-ray crystallography (Payandeh et al., 2011, 2012; McCusker et al., 2012; Zhang et al., 2012; Bagnéris et al., 2013; Shaya et al., 2014). These include the voltage-gated sodium channels from Magnetococcus marinus, formerly known as Magnetococcus spirillium (NavMs); from Arcobacter butzler (NavAb); from Rickettsiales sp. HIMB114 (NavRh); and from Alkalilimnicola ehrlichei (NavAe). These structures now make it possible to observe various states of prokaryotic sodium channels because they have apparently captured the pore regions in partially and fully open (NavMs), closed (NavAb and NavAe), and inactivated (NavAb and NavRh) conformations, thereby enabling structural comparisons that provide new insights into the transition processes.Because the transmembrane pore regions of the NavMs and NavAb pores have very high sequence identities (∼69%; McCusker et al., 2012; Fig. 1), our comparisons will primarily be confined to these structures because with this level of similarity, differences seen are likely to be state dependent rather than a consequence of sequence dissimilarities. They enable comparisons of structures that we designate as the following: partially (pOpenMs; Protein Data Bank [PDB] accession no. 4F4L; 3.5-Å resolution; McCusker et al., 2012) and fully open (OpenMs; PDB accession no. 3ZJZ, 2.9-Å resolution; Bagnéris et al., 2013), closed (ClosedAb; PBD accession no. 3RVY, an I217C mutant; 2.7-Å resolution; Payandeh et al., 2011), and inactivated (InactivatedAb; PDB accession no. 4EKW, 3.2-Å resolution; Payandeh et al., 2012) forms. The other inactivated form for which there is a crystal structure, NavRh (PDB accession no. 4DXW; 3.0-Å resolution; Zhang et al., 2012), and the other closed form, NavAe (PDB accession no. 4LTO; 3.5-Å resolution; Shaya et al., 2014), are more distant homologues (only 42 and 45% identity, respectively, in the pore region; Fig. 1). In addition, although both the NavAb and NavMs homologues have been shown to support sodium flux (Payandeh et al., 2011; D’Avanzo et al., 2013), neither NavRh nor NavAe has yet been shown to exhibit functional activity; therefore, we considered them to be less suitable for detailed comparisons. Although the resolutions of the structures are modest, all are sufficient to define the types and magnitudes of the conformational differences described in the comparisons made in this Review. In addition to these crystal structures, however, two structures of another orthologue, NavCt (PDB accession no. 4BGN, ∼9-Å resolution; Tsai et al., 2013), have been determined by electron crystallography, but those structures are at too low a resolution to make viable comparisons.Open in a separate windowFigure 1.Sequence alignment, using Clustal Omega (Sievers et al., 2011), of the prokaryotic sodium channel orthologues: NavMs from Magnetococcus marinus MC-1 (UniProt accession no. A0L5S6), NavAb from Arcobacter butzleri RM4018 (UniProt accession no. A8EVM5), NavAe from Alkalilimnicola ehrlichii MLHE-1 (UniProt accession no. Q0ABW0), NavCt from Caldalkalibacillus thermarum TA2.A1 (UniProt accession no. F5L478), NavRh from Rickettsiales sp. HIMB114 (UniProt accession no. D0RMU8), and domain IV (DIV) of the human Nav1.4 sodium channel (UniProt accession no. P35499). The identities of the helical regions (transmembrane helices S1–S6), the N-terminal intracellular helix S1N, the S4–S5 linker helix, the P1 and P2 pore helices, and the intracellular C-terminal coiled-coil (CC) helix are indicated in the horizontal colored tubes above the sequences and are based on the crystal structures of NavAb for the VS region (cyan), on the crystal structure of NavMs for the pore region (green), and on the site-directed electron paramagnetic resonance spectroscopy of NavMs and circular dichroism truncation studies of NaChBac for the CTD (gray). The vertical magenta bar indicates the extracellular negatively charged (ENC) region formed by D49 (in S2), and the red bars are the intracellular negatively charged (INC) region formed by E59 (in S2) and D81 (in S3) involved in the gating charge transfer across the membrane through sequential interactions with four arginine residues in S4 (indicated by the cyan vertical bars); in the “up” conformation, which corresponds to the activated state, residues E59 and D81 form the salt-bridged pairs. The residues comprising the SF are highlighted in light green vertical bars. The residues in purple are proposed to be the start of the twist in S6 that is implicated in activation gate opening (light purple from the partially open structure and dark purple in the fully open structure). The residue in yellow indicates the location of the final hydrophobic constriction region (HC3) in the closed structure, which effectively corresponds to the location of the activation gate.This Review focuses on comparisons of the ClosedAb–pOpenMs–OpenMs–InactivatedAb crystal structures, developing a structure-based model for the functional transitions. Although these crystal structures provide an important new structural context for understanding the nature of the transitions, it must be remembered that any models developed from them will ultimately require verification by functional studies under voltage-clamp conditions that probe native structures in real time and in the context of cell membranes.

Defining the conformational states represented in the crystal structures

The full-length NavAb channel has its VS in the “up” conformation (with residues E59 and D81 forming salt-bridged pairs to arginines in S4; Fig. 1) (Payandeh et al., 2011). This conformation of the VS is considered to be associated with the activated or fully open state as defined by disulfide cross-linking experiments (DeCaen et al., 2008, 2009). In addition, because in crystals there is no transmembrane potential, we expected that the structure would include an open form of the pore, as would be seen when the channel gate opens in response to membrane depolarization. However, in these crystals, the activation gate, and thus the pore region, appears to be in the closed conformation: the pore has no continuous transmembrane pathway that would enable the passage of sodium ions from the extracellular to the intracellular surface (McCusker et al., 2012; Fig. 2, A and B). Although the selectivity filter (SF) is sufficiently open to enable ions to enter the pore, the bottom of its cavity (at the activation gate) is blocked (Payandeh et al., 2011) so ions cannot exit. This structure has thus been described as being in a “pre-open” state, in other words, with a VS conformation that has primed the gate for opening but in which the gate is actually closed. This demonstrates that the structures of the VS and pore regions can be uncoupled, and may indicate the presence of an additional conformation intermediate between the “open” and “closed” states. In the comparisons in this Review, which focuses on the pore region responsible for the opening and closing of the transmembrane pathway, we have designated this structure as “closed” (ClosedAb) based on the state of the pore region. It is very similar in dimensions and features to the NavAe closed pore (Shaya et al., 2014), a structure that does not contain a VS region, another indication that the pore region state is not solely determined by the VS state.Open in a separate windowFigure 2.Accessible surface differences between the open and closed structures. (A; middle) Accessible surface plots (made using CAVER software; Chovancova et al., 2012) showing radius versus distance through the central axis of the pore for ClosedAb-I217C (slate blue), ClosedAe (red), InactivatedAb (gray), OpenMs (light green), and pOpenMs (turquoise). (Left and right) Slab surface mode/cartoon depictions of the pores, sliced along the transmembrane direction for the ClosedAb (left, slate blue) and OpenMs (right, light green) pores, highlighting the sites of the first minor constriction (blue underlay: present for both forms with the responsible residues, V213 in the closed form and I215 in the open form), the second minor construction (orange underlay: present only in the closed form, residue I217C), and the third major constriction (yellow underlay: present only for the closed form, residue M221) at the intracellular end of the cavity. I217C is the mutation in the closed structure that enabled crystallization at higher resolution. (B) Detailed view of the three “hydrophobic constriction” (HC) regions noted above shown in cartoon and stick mode for the ClosedAb (slated blue) and OpenMs (light green) structures. The distances shown were measured between two diagonally opposite residues. It is clear that the narrow constrictions at HC2 and HC3 (6.62 and 4.81 Å) seen in the closed structure are not present in the open structure (where the equivalent distances are 14.84 and 17.18 Å).One of the two NavMs pore structures (Bagnéris et al., 2013) has been designated to be in the fully open state (OpenMs) by virtue of its transmembrane pathway being of sufficient diameter along the full length of the pore, from the extracellular surface to the intracellular surface (Fig. 2, A and B), to enable the translocation of sodium ions across the cell membrane (Ulmschneider et al., 2013). It has all four monomers with a splayed conformation at their intracellular ends. In contrast, in the partially open (pOpenMs) structure (McCusker et al., 2012) (which lacks the last 37 residues of its C-terminal domain [CTD]), only one of the four monomers of the tetramer adopts an open splayed conformation, with the result that the activation gate is only partially open; it is not sufficient to permit ion translocation but is more widely open than the closed structure (Fig. 2 A). The other three monomers in this structure are effectively equivalent to the closed-state pore conformations. A model “open” structure was constructed (McCusker et al., 2012) based on using the most open of the four monomers to produce a symmetric tetramer, which was very similar to the structure later determined of the fully open state (Bagnéris et al., 2013). In the partially open structure, the distal end of the CTD, which has been proposed to form a stabilizing four-helix coiled-coil (Powl et al., 2010; Irie et al., 2012; Bagnéris et al., 2013; Shaya et al., 2014), is absent, although it is present in the fully open structure; this may have enabled the uncoupled opening of the four S6 helices, and this asymmetry suggests there may be an asynchrony in the functional transition between states.The two inactivated forms of NavAb (InactivatedAb; Payandeh et al., 2012) and NavRh (InactivatedRh; Zhang et al., 2012) have been so designated based on their closed pores and collapsed SFs (although the authors of the NavAb structure were careful to describe it as being in a “potentially” inactivated state, largely because there was no confirming functional data). Both of these structures have closed activation gates, which are very similar to the ClosedAb structure, but their SFs are unlike any of the other structures, and would be too narrow to enable ions to enter the pore (Figs. 2 A, middle, and 4, C and D, and Table S1).Open in a separate windowFigure 4.Comparisons of the SF regions. (Left and middle columns) Views from the extracellular surface showing the size of the SF central holes, defined as the white areas in the middle of each structure. (Right column) Side view superpositions of the SFs in cartoon and stick depictions, in each case comparing the OpenMs structure (light green) with the corresponding structure in that row. The identities of the SF residues (TLES) are indicated. Only two monomers are shown for clarity. (A; left) OpenMs (pale green) and (middle) pOpenMs (turquoise) structures. (B; middle) ClosedAb (slate) and overlay (right) comparison of ClosedAb and OpenMs. (C) As in B for InactivatedAb (AB tetramer; gray). (D) As in B for InactivatedAb (CD tetramer; gray). (E) As in B for ClosedAe (raspberry). The distances between the narrowest parts of the SFs in each case are given in Table S1.

Structural transitions associated with opening and closing

The extracellular turret and vestibule surfaces (comprised of the S5–S6 linker, including the P1, P2, and SF regions; Fig. 1) are very similar in the OpenMs and ClosedAb pores, suggesting that these are not substantially involved in the gating transition. In addition, the N-terminal ends of their S6 helices (Fig. 3 B) are virtually superimposable, as are the C-terminal ends of the S5 helices. However, the intracellular surface features differ substantially (Fig. 3 A) as a result of differences in the relative orientations and splaying of the C-terminal ends of their S6 helices away from the pore axis. Also, the N-terminal ends of the S5 helices appear to have moved slightly with respect to each other. Had the open pore structure had its VS attached, this latter movement would have impinged on the S4–S5 linker (Fig. 3 A, circle) of an adjacent monomer. We speculate that the reason the NavAb pore is closed is that the linker region is in the closed conformation, and that if the linker had been more closely coupled to the VS in its open conformation in the crystal, the NavAb pore structure would also have been of an open conformation.Open in a separate windowFigure 3.Differences between the open and closed structures. (A) Comparison of the OpenMs (green) pore and the ClosedAb (slate blue) crystal structures, depicted in cylindrical mode viewed from the intracellular surface. The equivalent residue numbers for the pore domain were G129 to M221 in NavMs and G130 to M222 in NavAb. Three-dimensional alignments (in all cases the least-squares superpositions were done using residues 145–198 [or their sequence equivalents] at the top of helices S5 and S6) and figures were made using PyMOL software (Schrödinger, LLC). The motions associated with the S5 and S6 helices are indicated by the small and large arrows, respectively. One of the ClosedAb monomers is shown in gray, so that it can be seen that the region of the S4–S5 linker that the S5 helix in the open state would impinge on (magenta circle) is in the adjacent, not the same, monomer. (B) The Cα carbons of the S6 helixes (in stick motif) showing that the ClosedAb (slate blue), InactivatedAb (gray), and ClosedAe (red) structures overlay closely, but that the OpenMs (green) deviates from the other structures starting at residue T206. (C) Plot of the delta phi (blue) and delta psi (red) angles in the S6 helix as a function of residue number. Values are those of OpenMs structure (PDB accession no. 3ZJZ-A chain) minus those of the ClosedAb structure (3RVY-A chain), demonstrating that the differences start after residue T206 in NavMs and continue to the end of the S6 helix. The single peak at around residue 155 is not related to the transition but simply arises from different interactions of the two proteins, with the different crystallization detergent molecules present adjacent to this site. (D) Secondary structure alignments compared using the 2Struc server (Klose et al., 2010). The position corresponding to the T206 residue in helix S6 is indicated by the black box in both parts. (Top) OpenMs versus ClosedAb. The locations of the S5 and S6 helices are indicated by the horizontal green bars. Both structures have essentially identical secondary structures, even around T206. (Bottom) OpenMs versus InactivatedAb. The biggest differences are at the top of S5 (purple box) and in the turret loop (cyan box), not in helix S6 nor the region around T206.When the structures of the S6 pore helices of all the conformations are examined in detail (Fig. 3 B), the differences at the quaternary structural level between the open and closed (and inactivated) structures appear to arise from a twist in the middle of the S6 helix structures beginning at residue T206. The extracellular ends of all of the S6 helices in all of the closed and inactivated structures superimpose closely (Fig. 3 B), but the OpenMs structure is an outlier. This results from a change in the backbone Ramachandran angles between the open and closed forms, which perpetrates along the C-terminal end of S6 starting at T206 (Fig. 3 C). The consequence of this deviation is that the C-terminal end of the S6 helix in the OpenMs structure moves away from the axis of the top part of the helix, thereby opening the activation gate. In the lower resolution pOpenMs structure, residue T209 appeared to be the focal point of the bend, but in the higher resolution fully OpenMs structure, it is clear that the beginning of the bend occurs one turn earlier in S6, at T206. Notably, the changes between open and closed structures are subtle enough so that they do not change the secondary structure (Fig. 3 D) of the S6 helix. What starts as a relatively small change in the middle of the helix is translated to a motion of >5 Å at the end of S6 that forms the activation gate. This suggests a simple (energetically inexpensive) mechanism for opening and closing: a twist in the backbone that does not disrupt the hydrogen bonding pattern is sufficient to open the gate. It does not require major rearrangements of the rest of the structure and is thus compatible with a rapid opening, such as that which gives rise to the initial phase of the action potential in excitable cells. The asynchrony of the motions of the four monomers, as suggested by the partially open structure, suggests that there may also be an asynchrony of the process of opening.The open and closed structures have very similar cavity regions (Fig. 2 A), with a minor hydrophobic constriction (designated HC1) near the bottom of the cavity (Fig. 2 A) near residue I215 in NavMs (equivalent to V213 in NavAb) that is not sufficient to prevent ions continuing their passage. However, the ClosedAb form has two additional major hydrophobic constrictions (HC2 and HC3) further down toward the extracellular surface (Fig. 2 A) at approximately the level of residues M221 and I217 (numbering according to NavAb sequence; notably, the latter corresponds to the site of the mutation to a cysteine in the NavAb construct). Neither of these constrictions is seen in the open NavMs structure. HC3 (which obscures the exit) would exclude passage of even totally dehydrated sodium ions in the closed structure.Because the SFs of NavMs and NavAb have identical sequences (TLESWS; Fig. 1), they can be directly compared. The SFs of the OpenMs (Fig. 4 A) and ClosedAb (Fig. 4 B) structures are superimposable (Fig. 4 B, right), thus indicating that the closing of the activation gate, which prevents ion translocation, does not affect the potential entry of ions into the pore. In contrast, the SFs of the inactivated forms (Fig. 4, C and D) are very different from those of both the open and closed states, suggesting that they are altered during the inactivation process.

The structure and role of the CTD in opening and closing

All of the initial crystal structures determined (NavAb, NavMs, and NavRh) have elucidated the nature of the transmembrane domains (effectively to the end of the S6 helix) of the channels, but the structures of their CTDs (beyond the end of helix S6) were not interpretable, although these regions were present in the proteins used to produce the crystals. The structure of the CTD of the open conformation was first determined not by crystallography but by a combination of spectroscopic methods (Powl et al., 2010; Bagnéris et al., 2013), and shown to have a disordered region adjacent to the activation gate and a distal coiled-coil region (Figs. 1 and and5);5); molecular dynamics calculations (Bagnéris et al., 2013) suggested that flexibility of the region adjacent to the activation gate in the open state could account for the lack of defined structure in the crystal. The CTD in the closed NavAe pore (Shaya et al., 2014) is the only other such domain that has been defined structurally, this time by crystallographic means. Like the open CTD, its distal end formed a coiled-coil, but unlike the open CTD, the proximal end of the structure, nearest to the activation gate, was helical and relatively well ordered. These observations suggested that although the C-terminal end of the CTD remains unchanged in the two states, the end adjacent to the activation gate undergoes an ordered-to-disordered transition when the gate is opened (Fig. 5), thereby accommodating the separation movement at the end of the S6 helix that forms the gate, without uncoiling the distal coiled-coil that has been proposed to have a structural role in stabilizing the tetrameric structure in the membrane (Mio et al., 2010; Powl et al., 2010). Functional studies on the NavMs channel (Bagnéris et al., 2013), as well as on another orthologue (NavSulP from Sulfitobacer pontiacus; Irie et al., 2012), have indicated that this region (especially the EEE motif near the top of the CTD) plays a role in inactivation, recovery, and closing, perhaps by promotion of the conformational change of the S6 helix. In addition, in the NavAe pore structure, there is some density in the extracellular region beyond the end of the transmembrane segments that could be a calcium or other cation. Domain DIV of the human Nav1.4 sodium channel also possesses a similar negatively charged region just after its activation gate (Fig. 1), which appears to be a common feature in the prokaryotic sodium channels.Open in a separate windowFigure 5.Schematic diagram of (1) closed, (2) open, and (3) inactivated states of prokaryotic sodium channels based on crystallographic studies. Only two monomers are shown in each figure for clarity. The color scheme is as in Fig. 1. The VS S1–S4 helices are depicted as cyan bars, the S4–S5 linker is an orange bar, the S5–S6 helices of the pore region are green bars, and the coiled-coil region of the CTD is gray. The extracellular negatively charged region in S2 is in a magenta circle, and the intracellular negatively charged regions in S2 and S3 are in red circles. The four arginines in S4 involved in the gating charge transfer across the membrane through sequential interactions, with the extracellular negatively charged region and the intracellular negatively charged region represented by a “+.” The residues forming the SF are indicated by red boxes. The residue in purple is the start of the S6 twist that results in the open gate; the residue in yellow is the third hydrophobic constriction site in the closed form, which is not constricted in the open form, and indicates the position of the activation gate. The white bars represent the helical region present in the CTD in the closed and inactivated structures, and the corresponding dotted gray lines are the disordered CTD linker region in the open conformation.

What produces the open and closed states in the crystals?

An obvious suggestion as to what causes NavMs pore structure to be open is that it is missing the constraints provided by the VS and S4–S5 linker regions present in the full-length channel structures. However, this cannot be the only reason, as the two structures of pore-only constructs, NavMs and NavAe, are in different states (NavMs is open, whereas NavAe is closed). It is also not because of the presence of CTD structures because, again, the CTDs are present in both of these two structures, albeit in different conformations. Furthermore, the partially open NavMs construct is missing the distal end of its CTD, yet one of its monomers is in the open state. The CTDs are present but not visible in any of the other structures. It is possible that the different CTD structures have a role in driving the transition (for example because of the different conformations at their N-terminal ends), but this cannot be determined from the structures in hand. It could also be because of the differences in crystal packing that enable the NavMs pores to open: in the NavMs crystals, the transmembrane regions of different pores are aligned next to each other, as if in a membrane bilayer, with a large enough gap to fit the CTD between two separate bilayers and thus allow sufficient room for the bottom of the pore to open. In the NavAe crystals, the CTDs are sandwiched between other CTDs and pore regions of symmetrically related molecules, rigidifying the whole structure and potentially preventing the opening.The different structures in the crystals may be caused by (or alternatively, result in) the presence or absences of ions within the pore, as sodium ions are visible in the SF of only the NavMs open-state crystal forms. However, in a version of the NavAb orthologue engineered to be calcium selective (Tang et al., 2014), the presence of calcium ions in that pore did not produce an open state, so the mere presence of ions appears not to be sufficient to produce an open pore. The most likely candidate in vivo is the VS-pore linker between helices S4 and S5, which could act as a lever, either pushing or pulling the pore domain when the VS is activated to its “up” position. But the linker is not present in either the NavMs open or partially open structures, nor in the NavAe closed structures. And although it is present in the NavAb structures, it is in the “wrong” position to influence the opening of the pore, so it may be a contributing, but not the only, driving force. This is an important issue and will need to be resolved in the future to enable a full understanding of the opening and closing processes.

Structural transitions associated with inactivation

Comparisons of the structure of the InactivatedAb channel with both the ClosedAb and OpenMs structures indicate areas of the protein primarily involved in the inactivation transition. The InactivatedAb structure has a very similar upper cavity region to both the open and closed structures, and an asymmetric inactivation gate conformation that, like the symmetrical gate of the closed state, is occluded. The activation gate structures of the S6 helices of both the InactivatedAb and InactivatedRh channels closely overlay the ClosedAb (Fig. 3 B) but not the OpenMs structures, suggesting that the hinge necessary to open the gate has not been activated. Although their S6 conformations appear to be the same, the structures of the two types of InactivatedAb tetramers (“AB” and “CD”) and the ClosedAb form do exhibit differences at their C-terminal ends: the inactivated forms have between two and eight fewer ordered residues visible in the crystals, suggesting that they exhibit more mobility or flexibility in this region. Because the CTD has not been resolved in any of the inactivated form crystals, we cannot make any conclusions about whether the regions adjacent to the transmembrane helices are disordered (as in the open state) or helical (as in the closed state). Nor can we discern even if they exhibit the same coiled-coil structures at their distal ends that are present in both the open and closed forms.In contrast, the SF region of the InactivatedAb structure (as well as that of the inactivated NavRh structure) is very different from either the open or closed states (Fig. 4). It has been described as “collapsed,” and indeed it is of insufficient diameter to enable a full dehydrated sodium ion to enter into the pore, much less exit it. Because in this case comparisons can be made for two crystal structures of the same orthologue (NavAb), the differences in the inactivated and closed structures cannot be attributed to SF sequence differences. The changes involve both the P1 and P2 helices as well as the SF itself, and involve modest differences in the secondary structures of the turret and top of S5 (Fig. 3 D). The observation that structural changes associated with the inactivated structure primarily involve the SF region corresponds well with earlier functional studies suggesting that the region close to the SF is responsible for inactivation (Pavlov et al., 2005). Interestingly, too, is the correspondence with studies done many years ago using circular dichroism spectroscopy (Cronin et al., 2003) on eukaryotic (electric eel) sodium channels that were induced to adopt open, closed, and inactivated states through the use of toxins and drugs. Those functional studies indicated that the secondary structures of the open and closed channels were surprisingly very similar but that the secondary structures of the inactivated form was considerably different. This also corresponds with the structural observations for the prokaryotic channels, where there are no apparent changes in secondary structure between open and closed states but significant differences between the open and the inactivated state.The structural comparisons thus suggest that inactivation closes not only the extracellular but also the intracellular ends of the transmembrane passageway, although the extracellular changes appear to be the defining ones for the inactivation state.

Structural models for sodium channel opening–closing–inactivation mechanisms

Early models (e.g., Armstrong and Bezanilla, 1973; Hille, 1975; Lehmann-Horn and Jurkat-Rott, 1999; see also Hille, 2001, and Armstrong, 2007) proposed for eukaryotic sodium channel charge movement and gating were based primarily on functional observations and were produced, for the most part, before either any sodium channel sequences were determined or any crystal structures of any members of the voltage-gated cation channel superfamily were available. Structural models included suggestions such as the constriction of the SF or the shutting of two flaps (Lehmann-Horn and Jurkat-Rott, 1999), one at the extracellular surface and one at the intracellular surface, as being responsible for inactivation and closing, respectively. With the sequencing of the first eukaryotic sodium channel (Noda et al., 1984), three-dimensional models were developed that proposed the involvement of specific regions for the activation and inactivation processes (Guy and Seetharamulu, 1986; Yu et al., 2005). Many other approaches, including mutational effects, interactions with toxins, drug and ligand binding, and various spectroscopic studies, have since contributed to structure–function models of sodium channels (Catterall, 2012). More recently, the crystal structures of potassium channels (Doyle et al., 1998; Long et al., 2007) provided structural templates and information for understanding the electromechanical coupling and mechanism of their opening and closing. The sequence homology of sodium channels to these other members of the voltage-gated ion channel superfamily suggested that they would share a common architecture, albeit with very different SF regions, and led to more detailed models for sodium channel structures and functions (Yarov-Yarovoy et al., 2001; Shafrir et al., 2008; Zarrabi et al., 2010).Soon after the prokaryotic sodium channels were identified as simpler target molecules for biophysical studies (Ren et al., 2001), and well before any of the prokaryotic orthologue crystal structures were determined, a hinge model for NaChBac (the first sodium channel orthologue whose sequence was determined) was proposed. This model for opening of the activation gate, like the models for the MthK potassium channel (Jiang et al., 2002), was based on the presence of a glycine residue in the middle of the S6 helix. Because glycines lack side chains, this type of residue could act as a flexible pivot for the top and bottom of the S6 helix, which could then move to open the gate. This model was supported by functional studies that showed that mutations of the glycine altered the rate of inactivation and recovery (Ito et al., 2004; Zhao et al., 2004), and spectroscopic studies that showed that changing the glycine to a conformationally less flexible serine produced a more thermally stable protein (O’Reilly et al., 2008); later molecular dynamics calculations (Barber et al., 2012) also supported a hinge-bending model involving this glycine. This model for the mechanism became less favored, however, after the determination of the sequences of other orthologues, as they did not have a glycine in the equivalent position.The solution of the crystal structure of the prokaryotic sodium channel NavAb, with sequence and structural homology (especially in the VS regions) to single-domain tetrameric prokaryotic and eukaryotic potassium channel structures, prompted the suggestion (Payandeh et al., 2011) that the mechanism behind the coupling and pore opening might be similar to another model proposed for K+ channels (Long et al., 2007). Indeed, the VS of the ClosedAb structure aligns well with the VS of the open Kv1.2 structure, although they diverge in the pore, specifically at the beginning of the base of the S5 helix, and their SFs are completely different both in sequence and in structure (one using side chains and backbone carbonyls, and the other narrower one using only backbone carbonyls to create the ion-binding sites). That model had the VS and pore regions moving as modular units, with the S4–S5 linker movement causing the conformational change. In the sodium channel case, the proposed change was a rotation-like motion of the S5 and S6 helices relative to each other, opening the pore in an iris-like dilation motion. This bears semblance to the twisting mechanism proposed for the opening and closing of the nonspecific ion channel in nicotinic acetylcholine receptors (Unwin, 2003), where the transmembrane helices of adjacent subunits change positions relative to each other, primarily on one surface, producing a rotational squeezing and release mechanism that changes the size of the pore along its length.

A new structure-based model for the opening–closing–inactivation mechanisms

The observed structural differences described above between OpenMs and ClosedAb do not support a model based on a modular S5–S6 iris-like motion. In the prokaryotic sodium channel crystal structures, the C-terminal end of the S5 helix exhibits very little change between the two different states, and nothing suggests its movement is coupled with that of the S6 helix in either the same or an adjacent subunit within the tetramer. Instead, the opening and closing appears to involve a twist/bend in the middle of the S6 helix, which moves the activation gate to occlude or open the intracellular surface, enabling ion egress from the cavity (McCusker et al., 2012; Bagnéris et al., 2013; Video 1). This is remarkably similar to the earlier hinge model, except that there is no glycine at the hinge position. Instead, a threonine (T206 in the NavMs numbering scheme) is located at the juncture, in an equivalent position in the sequence to the glycine found in the NaChBac orthologue. The structure of NavMs shows that there would be sufficient space to enable the backbone motion even with the threonine side chain, as the Cβ substituent does not impinge on the region that moves. However, the enhanced flexibility of a glycine residue in NaChBac could be responsible for its faster rate of opening (Ren et al., 2001) than NavMs (Ulmschneider et al., 2013).Because the S1–S4 helices of the VS structure are already in the open conformation in the ClosedAb and InactivatedAb structures, it initially seemed likely that the S4–S5 linker in these structures would be in a conformation similar to that which it would adopt in the open pore state. However, comparisons of the OpenMs and ClosedAb structures indicate that the positions of especially the S5 helices in the open pore would mitigate against the linker in the ClosedAb structure being completely in the activated state because the end of the open state S5 helix would impinge on the linker. It is therefore suggested that the linker is in a “primed” state but that it moves in concert with the S5 and S6 helices as the activation gate is opened.The structure-based model for the slow inactivation process, like several of the early models (Lehmann-Horn and Jurkat-Rott, 1999), appears to primarily involve a closing of the SF to ion entry on the extracellular surface (Fig. 5). Rather than a flap folding over the entry, however, it appears that the SF itself folds in toward the center of the ion pathway, blocking entry of any sodium ions into the pore. It also appears, contrary to the double-flap mechanism, that the activation gate in the inactivated form is closed rather than open, although this could be a consequence of the crystallization and packing conditions (Video 2).Flexibility at the C-terminal region of S6 thus appears to be the defining feature of the prokaryotic voltage-gated sodium channel. It may be at least partially responsible for channel opening and for the cascade of events that occur at the SF and P loops during inactivation. The presence of a flexible linking region to the CTD would enable the opening and closing to occur without uncoiling the distal coiled-coil region, which may play an important role in stabilizing the tetrameric quaternary structure of the prokaryotic sodium channels (Video 3). This feature is not present (nor necessary) in eukaryotic sodium channels, which are single polypeptide chains. Instead, they seem to have a CTD that plays an important role in fast inactivation, a function that prokaryotic sodium channels do not exhibit.

Similarities and differences in the activation mechanisms of prokaryotic and eukaryotic sodium channels

Prokaryotic and eukaryotic sodium channels exhibit considerable homologies in the sequences of their transmembrane regions (Fig. 1), especially in the C-terminal ends of their S6 helices, although their SFs and CTDs are entirely different. Likewise, both exhibit transitions between activation, recovery, and slow inactivation functional states, albeit with different kinetics (as well as the absence of fast inactivation in prokaryotic channels). Therefore, it is of interest to consider if they might involve similar structural features in a common mechanism for opening and closing.New studies (Bagnéris et al., 2014) on the structure and electrophysiological consequences of binding channel blocker drugs to prokaryotic and human Nav1.1 channels have shown the parallel nature of their block by a range of ligands that bind in the hydrophobic cavity and prevent passage of ions from the extracellular to intracellular surfaces. This apparently occurs not by closing the activation gate but by the hydrophobic ligands creating a barrier to exit at the end of the SF region, and is entirely consistent with the mechanisms described above.Furthermore, a recent study was done to test whether a similar mechanism of opening and closing could be operating in human Nav1.4 and prokaryotic sodium channels (Oelstrom et al., 2014). Channels with their fast inactivation mechanism disabled could be functionally held in either the open or closed state. An accessibility labeling technique was used to probe sites in S6 of domain IV (Fig. 1). Residues C terminal to the proposed activation gate were accessible to labeling from the intracellular side in both the open and closed states, whereas residues above this were only accessible in the open state. In the open state, the channel could be labeled up to the ring of hydrophobic residues immediately preceding the equivalent to residue T206 in the prokaryotic channels, again consistent with a mechanism involving a hinge mechanism in the middle of S6. These results strongly support the type of mechanism for opening and closing presented above, with a flexible hinge region that opens the pore up to the intracellular compartment in the open state, but not in the closed state.However, it is important to note that significant differences in the activation/inactivation and selectivity mechanisms do exist between prokaryotic and eukaryotic channels (see, for example, Ahern, 2013; Goldschen-Ohm et al., 2013; Finol-Urdaneta et al., 2014, for detailed discussions). These arise in large part because of the asymmetry of the pseudotetrameric eukaryotic structures as opposed to the simpler homotetramers that comprise the prokaryotic channels. They result in asynchronous movements of VSs, asymmetric ion-binding sites, and novel loops between the different domains that have specialized roles, for example in fast inactivation, which are not seen in the prokaryotic channels. Nevertheless, the availability of several crystal forms of prokaryotic voltage-gated sodium channels in different conformational states has vastly increased our knowledge of the structure–function relationships for these channels, and provides new insight into their mechanisms of opening, closing, and inactivation (Fig. 5). These, in turn, may inform our understanding of both the structure and function of sodium channels in general. The new model proposed, based on the detailed structures of prokaryotic channels in crystal environments, will, however, await confirmation by further functional experiments on channels in their biological context, as well as structural studies of eukaryotic channels.  相似文献   

10.
11.
Ca2+-activated Cl? currents have been implicated in many cellular processes in different cells, but for many years, their molecular identity remained unknown. Particularly intriguing are Ca2+-activated Cl? currents in olfactory transduction, first described in the early 90s. Well characterized electrophysiologically, they carry most of the odorant-induced receptor current in the cilia of olfactory sensory neurons (OSNs). After many attempts to determine their molecular identity, TMEM16B was found to be abundantly expressed in the cilia of OSNs in 2009 and having biophysical properties like those of the native olfactory channel. A TMEM16B knockout mouse confirmed that TMEM16B was indeed the olfactory Cl? channel but also suggested a limited role in olfactory physiology and behavior.

The question then arises of what the precise role of TMEM16b in olfaction is. Here we review the long story of this channel and its possible roles.  相似文献   


12.
Summary Properties of the single Cl channels were studied in excised patches of surface membrane from molluscan neurones using single-channel recording technique. These channels are controlled by Ca2+ and K+ acting on cytoplasmic and outer membrane surfaces, respectively, and by the membrane potential. The channels display about 16 intermediate conductance sublevels, each of them being multiples of 12.5 pS. The upper level of the channel conductance is about 200 pS. The channel behavior is consistent with an aggregation of channel-forming subunits into a cluster.  相似文献   

13.
Summary Squid giant axon could be excited in concentrated glycerol solutions containing normal concentrations of electrolytes, when osmolalities of solutions inside and outside the axon were matched. These glycerol solutions did not freeze at the temperature as low as –19°C. The nerve excitation in these solutions were observed at this low temperature. The excitation process at this low temperature was slowed down and time constants of the excitation kinetics were several hundredfold larger than those in normal seawater at 10°C, under which temperature the squid habituated. The temperature coefficients for the electrophysiological membrane parameters under this condition were larger than those in normal seawater above 0°C. The Q10 value for the conduction velocity was 2.0 and that of the duration of the action potential was around 8.5. The time course of the membrane currents was also slowed with the Q10 value of around 5 and the magnitude decreased with the Q10 value of around 2 as the temperature was lowered. The Q10 values for the kinetics of the on process of the Na-channel were around 4.5 and were almost the same as those of the off process of the Na-channel in the wide range of the temperature below 0°C. The Q10 value of the on process of K-channel was around 6.5 and was larger than those for Na-channel. The Q10 values increased gradually as the temperature was lowered.  相似文献   

14.
Testosterone induces vasorelaxation through non-genomic mechanisms in several isolated blood vessels, but no study has reported its effects on the canine basilar artery, an important artery implicated in cerebral vasospasm. Hence, this study has investigated the mechanisms involved in testosterone-induced relaxation of the canine basilar artery. For this purpose, the vasorelaxant effects of testosterone were evaluated in KCl- and/or PGF-precontracted arterial rings in vitro in the absence or presence of several antagonists/inhibitors/blockers; the effect of testosterone on the contractile responses to CaCl2 was also determined. Testosterone (10-180 μM) produced concentration-dependent relaxations of KCl- or PGF-precontracted arterial rings which were: (i) unaffected by flutamide (10 μM), dl-aminoglutethimide (10 μM), actinomycin D (10 μM), cycloheximide (10 μM), SQ 22,536 (100 μM) or ODQ (30 μM); and (ii) significantly attenuated by the blockers 4-aminopyridine (KV; 1 mM), BaCl2 (KIR; 30 μM), iberiotoxin (BKCa2+; 20 nM), but not by glybenclamide (KATP; 10 μM). In addition, testosterone (31, 56 and 180 μM) and nifedipine (0.01-1 μM) produced a concentration-dependent blockade of the contraction to CaCl2 (10 μM to 10 mM) in arterial rings depolarized by 60 mM KCl. These results, taken together, show that testosterone relaxes the canine basilar artery mainly by blockade of voltage-dependent Ca2+ channels and, to a lesser extent, by activation of K+ channels (KIR, KV and BKCa2+). This effect does not involve genomic mechanisms, production of cAMP/cGMP or the conversion of testosterone to 17β-estradiol.  相似文献   

15.
Jarvis SE  Zamponi GW 《Cell calcium》2005,37(5):483-488
Calcium entry through presynaptic voltage-gated calcium channels is essential for neurotransmitter release. The two major types of presynaptic calcium channels contain a synaptic protein interaction site that physically interacts with synaptic vesicle release proteins. This is thought to tighten the coupling between the sources of calcium entry and the neurotransmitter release machinery. Conversely, the binding of synaptic proteins to presynaptic calcium channels regulates calcium channel activity. Hence, presynaptic calcium channels act not only as the masters of the synaptic release process, but also as key targets for feedback inhibition.  相似文献   

16.
In the microcirculation, longitudinal conduction of vasomotor responses provides an essential means of coordinating flow distribution among vessels in a complex network. Spread of current along the vessel axis can display a regenerative component, which leads to propagation of vasomotor signals over many millimeters; the ionic basis for the regenerative response is unknown. We examined the responses to 10 s of focal electrical stimulation (30 Hz, 2 ms, 30 V) of mouse cremaster arterioles to test the hypothesis that voltage-dependent Na(+) (Na(v)) and Ca(2+) channels might be activated in long-distance signaling in microvessels. Electrical stimulation evoked a vasoconstriction at the site of stimulation and a spreading, nondecremental conducted dilation. Endothelial damage (air bubble) blocked conduction of the vasodilation, indicating an involvement of the endothelium. The Na(v) channel blocker bupivacaine also blocked conduction, and TTX attenuated it. The Na(v) channel activator veratridine induced an endothelium-dependent dilation. The Na(v) channel isoforms Na(v)1.2, Na(v)1.6, and Na(v)1.9 were detected in the endothelial cells of cremaster arterioles by immunocytochemistry. These findings are consistent with the involvement of Na(v) channels in the conducted response. BAPTA buffering of endothelial cell Ca(2+) delayed and reduced the conducted dilation, which was almost eliminated by Ni(2+), amiloride, or deletion of alpha(1H) T-type Ca(2+) (Ca(v)3.2) channels. Blockade of endothelial nitric oxide synthase or Ca(2+)-activated K(+) channels also inhibited the conducted vasodilation. Our findings indicate that an electrically induced signal can propagate along the vessel axis via the endothelium and can induce sequential activation of Na(v) and Ca(v)3.2 channels. The resultant Ca(2+) influx activates endothelial nitric oxide synthase and Ca(2+)-activated K(+) channels, triggering vasodilation.  相似文献   

17.
Voltage-gated sodium (Na(v)) channels are the molecular targets of β-scorpion toxins, which shift the voltage dependence of activation to more negative membrane potentials by a voltage sensor-trapping mechanism. Molecular determinants of β-scorpion toxin (CssIV) binding and action on rat brain sodium channels are located in the S1-S2 (IIS1-S2) and S3-S4 (IIS3-S4) extracellular linkers of the voltage-sensing module in domain II. In IIS1-S2, mutations of two amino acid residues (Glu(779) and Pro(782)) significantly altered the toxin effect by reducing binding affinity. In IIS3-S4, six positions surrounding the key binding determinant, Gly(845), define a hot spot of high-impact residues. Two of these substitutions (A841N and L846A) reduced voltage sensor trapping. The other three substitutions (N842R, V843A, and E844N) increased voltage sensor trapping. These bidirectional effects suggest that the IIS3-S4 loop plays a primary role in determining both toxin affinity and efficacy. A high resolution molecular model constructed with the Rosetta-Membrane modeling system reveals interactions of amino acid residues in sodium channels that are crucial for toxin action with residues in CssIV that are required for its effects. In this model, the wedge-shaped CssIV inserts between the IIS1-S2 and IIS3-S4 loops of the voltage sensor, placing key amino acid residues in position to interact with binding partners in these extracellular loops. These results provide new molecular insights into the voltage sensor-trapping model of toxin action and further define the molecular requirements for the development of antagonists that can prevent or reverse toxicity of scorpion toxins.  相似文献   

18.
Atrial-selective inhibition of cardiac Na(+) channel current (I(Na)) and I(Na)-dependent parameters has been shown to contribute to the safe and effective management of atrial fibrillation. The present study examined the basis for the atrial-selective actions of ranolazine. Whole cell I(Na) was recorded at 15°C in canine atrial and ventricular myocytes and in human embryonic kidney (HEK)-293 cells expressing SCN5A. Tonic block was negligible at holding potentials from -140 to -100 mV, suggesting minimal drug interactions with the closed state. Trains of 40 pulses were elicited over a range of holding potentials to determine use-dependent block. Guarded receptor formalism was used to analyze the development of block during pulse trains. Use-dependent block by ranolazine increased at more depolarized holding potentials, consistent with an interaction of the drug with either preopen or inactivated states, but was unaffected by longer pulse durations between 5 and 200 ms, suggesting a weak interaction with the inactivated state. Block was significantly increased at shorter diastolic intervals between 20 and 200 ms. Responses in atrial and ventricular myocytes and in HEK-293 cells displayed a similar pattern. Ranolazine is an open state blocker that unbinds from closed Na(+) channels unusually fast but is trapped in the inactivated state. Kinetic rates of ranolazine interactions with different states of atrial and ventricular Na(+) channels were similar. Our data suggest that the atrial selectivity of ranolazine is due to a more negative steady-state inactivation curve, less negative resting membrane potential, and shorter diastolic intervals in atrial cells compared with ventricular cells at rapid rates.  相似文献   

19.
Connexinchannels are gated by transjunctional voltage(Vj)or CO2 via distinct mechanisms.The cytoplasmic loop (CL) and arginines of a COOH-terminal domain(CT1) of connexin32 (Cx32) wereshown to determine CO2sensitivity, and a gating mechanism involvingCL-CT1 association-dissociationwas proposed. This study reports that Cx32 mutants, tandem, 5R/E, and5R/N, designed to weaken CL-CT1interactions, display atypicalVjand CO2 sensitivities when testedheterotypically with Cx32 wild-type channels inXenopus oocytes. In tandems, two Cx32monomers are linked NH2-to-COOH terminus. In 5R/E and 5R/N mutants, glutamates or asparagines replaceCT1 arginines. On the basis of theintriguing sensitivity of the mutant-32 channel toVjpolarity, the existence of a "slow gate" distinct from theconventionalVjgate is proposed. To a lesser extent the slow gatemanifests itself also in homotypic Cx32 channels. Mutant-32 channelsare more CO2 sensitive than homotypic Cx32 channels, andCO2-induced chemical gating isreversed with relative depolarization of the mutant oocyte, suggesting Vjsensitivity of chemical gating. A hypothetical pore-plugging modelinvolving an acidic cytosolic protein (possibly calmodulin) is discussed.

  相似文献   

20.
High voltage-activated Ca2+ (CaV) channels are protein complexes containing pore-forming α1 and auxiliary β and α2δ subunits. The subcellular localization and membrane interactions of the β subunits play a crucial role in regulating CaV channel inactivation and its lipid sensitivity. Here, we investigated the effects of membrane phosphoinositide (PI) turnover on CaV2.2 channel function. The β2 isoform β2e associates with the membrane through electrostatic and hydrophobic interactions. Using chimeric β subunits and liposome-binding assays, we determined that interaction between the N-terminal 23 amino acids of β2e and anionic phospholipids was sufficient for β2e membrane targeting. Binding of the β2e subunit N terminus to liposomes was significantly increased by inclusion of 1% phosphatidylinositol 4,5-bisphosphate (PIP2) in the liposomes, suggesting that, in addition to phosphatidylserine, PIs are responsible for β2e targeting to the plasma membrane. Membrane binding of the β2e subunit slowed CaV2.2 current inactivation. When membrane phosphatidylinositol 4-phosphate and PIP2 were depleted by rapamycin-induced translocation of pseudojanin to the membrane, however, channel opening was decreased and fast inactivation of CaV2.2(β2e) currents was enhanced. Activation of the M1 muscarinic receptor elicited transient and reversible translocation of β2e subunits from membrane to cytosol, but not that of β2a or β3, resulting in fast inactivation of CaV2.2 channels with β2e. These results suggest that membrane targeting of the β2e subunit, which is mediated by nonspecific electrostatic insertion, is dynamically regulated by receptor stimulation, and that the reversible association of β2e with membrane PIs results in functional changes in CaV channel gating. The phospholipid–protein interaction observed here provides structural insight into mechanisms of membrane–protein association and the role of phospholipids in ion channel regulation.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号