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 46 毫秒
1.
Hyperpolarization-activated, cyclic nucleotide-gated cation (HCN) channels are critical regulators of neuronal excitability, but less is known about their possible roles in synaptic plasticity and memory circuits. Here, we characterized the HCN gene organization, channel properties, distribution, and involvement in associative and nonassociative forms of learning in Aplysia californica. Aplysia has only one HCN gene, which codes for a channel that has many similarities to the mammalian HCN channel. The cloned acHCN gene was expressed in Xenopus oocytes, which displayed a hyperpolarization-induced inward current that was enhanced by cGMP as well as cAMP. Similarly to its homologs in other animals, acHCN is permeable to K+ and Na+ ions, and is selectively blocked by Cs+ and ZD7288. We found that acHCN is predominantly expressed in inter- and motor neurons, including LFS siphon motor neurons, and therefore tested whether HCN channels are involved in simple forms of learning of the siphon-withdrawal reflex in a semiintact preparation. ZD7288 (100 μM) significantly reduced an associative form of learning (classical conditioning) but had no effect on two nonassociative forms of learning (intermediate-term sensitization and unpaired training) or baseline responses. The HCN current is enhanced by nitric oxide (NO), which may explain the postsynaptic role of NO during conditioning. HCN current in turn enhances the NMDA-like current in the motor neurons, suggesting that HCN channels contribute to conditioning through this pathway.Hyperpolarization-activated, cyclic nucleotide-gated (HCN), cation nonselective ion channels generate hyperpolarization-activated inward currents (Ih) and thus tend to stabilize membrane potential (13). In addition, binding of cyclic nucleotides (cAMP and cGMP) to the C-terminal cyclic nucleotide binding domain (CNBD) enhances Ih and thus couples membrane excitability with intracellular signaling pathways (2, 4). HCN channels are widely important for numerous systemic functions such as hormonal regulation, heart contractility, epilepsy, pain, central pattern generation, sensory perception (415), and learning and memory (1624).However, in previous studies it has been difficult to relate the cellular effects of HCN channels directly to their behavioral effects, because of the immense complexity of the mammalian brain. We have therefore investigated the role of HCN channels in Aplysia, which has a numerically simpler nervous system (25). We first identified and characterized an HCN gene in Aplysia, and showed that it codes for a channel that has many similarities to the mammalian HCN channel. We found that the Aplysia HCN channel is predominantly expressed in motor neurons including LFS neurons in the siphon withdrawal reflex circuit (26, 27). We therefore investigated simple forms of learning of that reflex in a semiintact preparation (2830) and found that HCN current is involved in classical conditioning and enhances the NMDA-like current in the motor neurons. These results provide a direct connection between HCN channels and behavioral learning and suggest a postsynaptic mechanism of that effect. HCN current in turn is enhanced by nitric oxide (NO), a transmitter of facilitatory interneurons, and thus may contribute to the postsynaptic role of NO during conditioning.  相似文献   

2.
What are the implications for the existence of subthreshold ion channels, their localization profiles, and plasticity on local field potentials (LFPs)? Here, we assessed the role of hyperpolarization-activated cyclic-nucleotide–gated (HCN) channels in altering hippocampal theta-frequency LFPs and the associated spike phase. We presented spatiotemporally randomized, balanced theta-modulated excitatory and inhibitory inputs to somatically aligned, morphologically realistic pyramidal neuron models spread across a cylindrical neuropil. We computed LFPs from seven electrode sites and found that the insertion of an experimentally constrained HCN-conductance gradient into these neurons introduced a location-dependent lead in the LFP phase without significantly altering its amplitude. Further, neurons fired action potentials at a specific theta phase of the LFP, and the insertion of HCN channels introduced large lags in this spike phase and a striking enhancement in neuronal spike-phase coherence. Importantly, graded changes in either HCN conductance or its half-maximal activation voltage resulted in graded changes in LFP and spike phases. Our conclusions on the impact of HCN channels on LFPs and spike phase were invariant to changes in neuropil size, to morphological heterogeneity, to excitatory or inhibitory synaptic scaling, and to shifts in the onset phase of inhibitory inputs. Finally, we selectively abolished the inductive lead in the impedance phase introduced by HCN channels without altering neuronal excitability and found that this inductive phase lead contributed significantly to changes in LFP and spike phase. Our results uncover specific roles for HCN channels and their plasticity in phase-coding schemas and in the formation and dynamic reconfiguration of neuronal cell assemblies.Local field potentials (LFPs) have been largely believed to be a reflection of the synaptic drive that impinges on a neuron. In recent experimental and modeling studies, there has been a lot of debate on the source and spatial extent of LFPs (19). However, most of these studies have used neurons with passive dendrites in their models and/or have largely focused on the contribution of spike-generating conductances to LFPs (7, 8, 10, 11). Despite the widely acknowledged regulatory roles of subthreshold-activated ion channels and their somatodendritic gradients in the physiology and pathophysiology of synapses and neurons (1217), the implications for their existence on LFPs and neuronal spike phase have surprisingly remained unexplored. This lacuna in LFP analysis is especially striking because local and widespread plasticity of these channels has been observed across several physiological and pathological conditions, translating to putative roles for these channels in neural coding, homeostasis, disease etiology and remedies, learning, and memory (16, 1823).In this study, we focus on the role of hyperpolarization-activated cyclic nucleotide-gated (HCN) channels that mediate the h current (Ih) in regulating LFPs and theta-frequency spike phase. From a single-neuron perspective, HCN channels in CA1 pyramidal neurons play a critical role in regulating neuronal integration and excitability (14, 2427) and importantly introduce an inductive phase lead in the voltage response to theta-frequency oscillatory inputs (28), thereby enabling intraneuronal synchrony of incoming theta-frequency inputs (29). Given these and their predominant dendritic expression (25), we hypothesized HCN channels as regulators of LFPs through their ability to alter the amplitude and phase of the intracellular voltage response, thereby altering several somatodendritic transmembrane currents that contribute to LFPs. The CA1 region of the hippocampus offers an ideal setup to test this hypothesis, given the regular, open-field organization (4, 6, 7) of the pyramidal neurons endowed with well-established somatodendritic gradients in ion channel densities (16). As this organization enables us to assess the role of location-dependent channel expression profiles on LFPs across different strata, we tested our hypothesis, using a computational scheme involving morphologically realistic, physiologically constrained conductance-based model neurons. Our results positively test our hypothesis and provide specific evidence for novel roles for HCN channels and their inductive component in regulating LFP and spike phases, apart from enhancing spike-phase coherence. These results identify definite roles for HCN channels in phase-coding schemas and in the formation and dynamic reconfiguration of neuronal cell assemblies and argue for the incorporation of subthreshold-activated ion channels, their gradients, and their plasticity into the computation of LFPs.  相似文献   

3.
Binding of 3′,5′-cyclic adenosine monophosphate (cAMP) to hyperpolarization-activated cyclic nucleotide-gated (HCN) ion channels regulates their gating. cAMP binds to a conserved intracellular cyclic nucleotide-binding domain (CNBD) in the channel, increasing the rate and extent of activation of the channel and shifting activation to less hyperpolarized voltages. The structural mechanism underlying this regulation, however, is unknown. We used double electron–electron resonance (DEER) spectroscopy to directly map the conformational ensembles of the CNBD in the absence and presence of cAMP. Site-directed, double-cysteine mutants in a soluble CNBD fragment were spin-labeled, and interspin label distance distributions were determined using DEER. We found motions of up to 10 Å induced by the binding of cAMP. In addition, the distributions were narrower in the presence of cAMP. Continuous-wave electron paramagnetic resonance studies revealed changes in mobility associated with cAMP binding, indicating less conformational heterogeneity in the cAMP-bound state. From the measured DEER distributions, we constructed a coarse-grained elastic-network structural model of the cAMP-induced conformational transition. We find that binding of cAMP triggers a reorientation of several helices within the CNBD, including the C-helix closest to the cAMP-binding site. These results provide a basis for understanding how the binding of cAMP is coupled to channel opening in HCN and related channels.Ion channels are allosteric membrane proteins that open selective pores in response to various physiological stimuli, including binding of ligands and changes in transmembrane voltage (1). They are important for diverse physiological functions ranging from neurotransmission to muscle contraction. One such channel, the hyperpolarization-activated cyclic nucleotide-gated (HCN) ion channel, underlies the current (termed Ih, If, or Iq) produced in response to hyperpolarization of cardiac pacemaker cells and neurons (2). In the heart, HCN channels are responsible for pace-making activity and may have a role in the autonomic regulation of the heart rate (35). In the brain, HCN channels are involved in repetitive firing of neurons and dendritic integration (68). Despite the important physiological roles of HCN channels, the structure of the channels and molecular mechanism of their function are not completely understood.HCN channels are part of the voltage-gated K+ channel superfamily (9). Like other members of this family, they are tetramers, with each subunit having a voltage-sensor domain of four transmembrane helices (S1–S4) and a pore-lining domain consisting of two transmembrane helices separated by a reentrant loop (S5-P-S6; Fig. 1A). However, HCN channels contain two key specializations that make them unique among the voltage-gated ion channels: (i) They are activated by membrane hyperpolarization instead of depolarization, and (ii) they are regulated by the direct binding of cyclic nucleotides, like the ubiquitous second messenger cAMP, to a cytoplasmic domain in the carboxyl-terminal region of the channel. The direct binding of the agonist cAMP to HCN channels increases the rate and extent of activation and shifts the voltage dependence of activation to more depolarizing voltages.Open in a separate windowFig. 1.Study of conformational changes in HCN2 using DEER. (A, Upper) Putative transmembrane topology of HCN2 channels highlighting the voltage sensor domain (S1–S4) and the pore domain (S5–S6). Only two subunits are shown. (A, Lower) Crystal structure [Protein Data Bank (PDB) ID code 3ETQ] of the cysteine-free cytoplasmic carboxyl-terminal domain of HCN2. One subunit of the tetramer is shown in color. (B) Schematic diagram showing the distance change between two cysteine-attached MTSL spin labels in a protein upon cAMP binding. In this example, the two positions are closer in the presence of cAMP. (C) Raw DEER time traces for HCN2cys-free V537C,A624C labeled with MTSL are shown in black in the absence or presence of cAMP, as indicated. The colored curves are distance-distribution fits to the data. The oscillation frequency is higher in the presence of cAMP, indicating that the two positions are closer together in the ligand-bound form.The crystal structure of the carboxyl-terminal region bound to cAMP has been solved for several HCN channels (1014). The nearly identical structures consist of fourfold symmetrical tetramers predicted to connect directly to the S6 segments that form the ion-conducting pore (Fig. 1A). Each of the subunits contains two domains: the cyclic nucleotide-binding domain (CNBD) and the C-linker domain. The CNBD exhibits strong structural similarity to the CNBDs of other cyclic nucleotide-binding proteins, including cAMP-dependent protein kinase (PKA), the guanine nucleotide exchange factor Epac, and the Escherichia coli catabolite gene activator protein (CAP) (1519). The CNBD consists of an eight-stranded antiparallel β-roll, followed by two α-helices (B-helix and C-helix). cAMP binds in the anticonformation between the β-roll and the C-helix. The C-linker is a unique domain found only in HCN channels and their close homologs, cyclic nucleotide-gated (CNG) channels, and KCNH family K+ channels (14, 20, 21). It is situated between the CNBD and membrane-spanning domains of the channel, and is the site of virtually all intersubunit interactions in the structure (Fig. 1A). The C-linker has been found to play a key role in coupling conformational changes in the CNBD to opening of the pore (9, 22, 23).The ligand-induced movement of the C-helix is widely thought to initiate the conformational changes that lead to opening of the channel pore, but the structural evidence in support of this hypothesis is equivocal (10, 2429). The crystal structure of the HCN2 carboxyl-terminal region in the absence of ligand shows little difference from the cyclic nucleotide-bound structure (12). The only significant differences between the two structures are observed in the F′-helix of the C-linker and in the C-helix. The proximal half of the C-helix is in the same position in the cAMP-bound and unbound structures, whereas the distal half is missing from the apo structure, indicating that it is disordered or can access multiple conformations. In contrast, studies on the soluble carboxyl-terminal fragment using transition metal ion FRET (tmFRET) demonstrate a relatively large movement (∼5 Å) at the proximal end of the C-helix upon binding of cAMP (12). The tmFRET studies also indicate a smaller movement at the distal end of the C-helix and increased disorder in the C-helix in the absence of cyclic nucleotides (12, 26).In this study, we examined the cAMP-induced conformational transition in the CNBD of HCN2 using double electron–electron resonance (DEER) spectroscopy. DEER is a pulse electron paramagnetic resonance (EPR) method that can determine distances and resolve distance distributions between pairs of sites within proteins separated by about 15–80 Å (3033). In a typical DEER experiment, two sites in a protein are mutated to cysteines and labeled with small magnetic spin labels (Fig. 1B). DEER measures the pair’s magnetic through-space coupling via excitation of one label and probing of the other with a series of short microwave pulses. This method yields an oscillating signal whose frequency falls off with the third power of the distance between the labels (Fig. 1C). Crucially, DEER measures full-distance distributions, rather than just an average distance, providing quantitative information on structural heterogeneity and variability that is not accessible from X-ray crystal structures or ensemble FRET experiments. Using DEER, we found that the binding of cAMP to the isolated C-linker/CNBD of HCN2 causes the C-helix to move substantially toward the β-roll and decreases the conformational heterogeneity of the protein. These observations are the first step in understanding the mechanisms of ligand gating of HCN channels and the activation of other CNBD-containing proteins.  相似文献   

4.
Protein toxins from tarantula venom alter the activity of diverse ion channel proteins, including voltage, stretch, and ligand-activated cation channels. Although tarantula toxins have been shown to partition into membranes, and the membrane is thought to play an important role in their activity, the structural interactions between these toxins and lipid membranes are poorly understood. Here, we use solid-state NMR and neutron diffraction to investigate the interactions between a voltage sensor toxin (VSTx1) and lipid membranes, with the goal of localizing the toxin in the membrane and determining its influence on membrane structure. Our results demonstrate that VSTx1 localizes to the headgroup region of lipid membranes and produces a thinning of the bilayer. The toxin orients such that many basic residues are in the aqueous phase, all three Trp residues adopt interfacial positions, and several hydrophobic residues are within the membrane interior. One remarkable feature of this preferred orientation is that the surface of the toxin that mediates binding to voltage sensors is ideally positioned within the lipid bilayer to favor complex formation between the toxin and the voltage sensor.Protein toxins from venomous organisms have been invaluable tools for studying the ion channel proteins they target. For example, in the case of voltage-activated potassium (Kv) channels, pore-blocking scorpion toxins were used to identify the pore-forming region of the channel (1, 2), and gating modifier tarantula toxins that bind to S1–S4 voltage-sensing domains have helped to identify structural motifs that move at the protein–lipid interface (35). In many instances, these toxin–channel interactions are highly specific, allowing them to be used in target validation and drug development (68).Tarantula toxins are a particularly interesting class of protein toxins that have been found to target all three families of voltage-activated cation channels (3, 912), stretch-activated cation channels (1315), as well as ligand-gated ion channels as diverse as acid-sensing ion channels (ASIC) (1621) and transient receptor potential (TRP) channels (22, 23). The tarantula toxins targeting these ion channels belong to the inhibitor cystine knot (ICK) family of venom toxins that are stabilized by three disulfide bonds at the core of the molecule (16, 17, 2431). Although conventional tarantula toxins vary in length from 30 to 40 aa and contain one ICK motif, the recently discovered double-knot toxin (DkTx) that specifically targets TRPV1 channels contains two separable lobes, each containing its own ICK motif (22, 23).One unifying feature of all tarantula toxins studied thus far is that they act on ion channels by modifying the gating properties of the channel. The best studied of these are the tarantula toxins targeting voltage-activated cation channels, where the toxins bind to the S3b–S4 voltage sensor paddle motif (5, 3236), a helix-turn-helix motif within S1–S4 voltage-sensing domains that moves in response to changes in membrane voltage (3741). Toxins binding to S3b–S4 motifs can influence voltage sensor activation, opening and closing of the pore, or the process of inactivation (4, 5, 36, 4246). The tarantula toxin PcTx1 can promote opening of ASIC channels at neutral pH (16, 18), and DkTx opens TRPV1 in the absence of other stimuli (22, 23), suggesting that these toxin stabilize open states of their target channels.For many of these tarantula toxins, the lipid membrane plays a key role in the mechanism of inhibition. Strong membrane partitioning has been demonstrated for a range of toxins targeting S1–S4 domains in voltage-activated channels (27, 44, 4750), and for GsMTx4 (14, 50), a tarantula toxin that inhibits opening of stretch-activated cation channels in astrocytes, as well as the cloned stretch-activated Piezo1 channel (13, 15). In experiments on stretch-activated channels, both the d- and l-enantiomers of GsMTx4 are active (14, 50), implying that the toxin may not bind directly to the channel. In addition, both forms of the toxin alter the conductance and lifetimes of gramicidin channels (14), suggesting that the toxin inhibits stretch-activated channels by perturbing the interface between the membrane and the channel. In the case of Kv channels, the S1–S4 domains are embedded in the lipid bilayer and interact intimately with lipids (48, 51, 52) and modification in the lipid composition can dramatically alter gating of the channel (48, 5356). In one study on the gating of the Kv2.1/Kv1.2 paddle chimera (53), the tarantula toxin VSTx1 was proposed to inhibit Kv channels by modifying the forces acting between the channel and the membrane. Although these studies implicate a key role for the membrane in the activity of Kv and stretch-activated channels, and for the action of tarantula toxins, the influence of the toxin on membrane structure and dynamics have not been directly examined. The goal of the present study was to localize a tarantula toxin in membranes using structural approaches and to investigate the influence of the toxin on the structure of the lipid bilayer.  相似文献   

5.
Rickettsiae are responsible for some of the most devastating human infections. A high infectivity and severe illness after inhalation make some rickettsiae bioterrorism threats. We report that deletion of the exchange protein directly activated by cAMP (Epac) gene, Epac1, in mice protects them from an ordinarily lethal dose of rickettsiae. Inhibition of Epac1 suppresses bacterial adhesion and invasion. Most importantly, pharmacological inhibition of Epac1 in vivo using an Epac-specific small-molecule inhibitor, ESI-09, completely recapitulates the Epac1 knockout phenotype. ESI-09 treatment dramatically decreases the morbidity and mortality associated with fatal spotted fever rickettsiosis. Our results demonstrate that Epac1-mediated signaling represents a mechanism for host–pathogen interactions and that Epac1 is a potential target for the prevention and treatment of fatal rickettsioses.Rickettsiae are responsible for some of the most devastating human infections (14). It has been forecasted that temperature increases attributable to global climate change will lead to more widespread distribution of rickettsioses (5). These tick-borne diseases are caused by obligately intracellular bacteria of the genus Rickettsia, including Rickettsia rickettsii, the causative agent of Rocky Mountain spotted fever (RMSF) in the United States and Latin America (2, 3), and Rickettsia conorii, the causative agent of Mediterranean spotted fever endemic to southern Europe, North Africa, and India (6). A high infectivity and severe illness after inhalation make some rickettsiae (including Rickettsia prowazekii, R. rickettsii, Rickettsia typhi, and R. conorii) bioterrorism threats (7). Although the majority of rickettsial infections can be controlled by appropriate broad-spectrum antibiotic therapy if diagnosed early, up to 20% of misdiagnosed or untreated (1, 3) and 5% of treated RMSF cases (8) result in a fatal outcome caused by acute disseminated vascular endothelial infection and damage (9). Fatality rates as high as 32% have been reported in hospitalized patients diagnosed with Mediterranean spotted fever (10). In addition, strains of R. prowazekii resistant to tetracycline and chloramphenicol have been developed in laboratories (11). Disseminated endothelial infection and endothelial barrier disruption with increased microvascular permeability are the central features of SFG rickettsioses (1, 2, 9). The molecular mechanisms involved in rickettsial infection remain incompletely elucidated (9, 12). A comprehensive understanding of rickettsial pathogenesis and the development of novel mechanism-based treatment are urgently needed.Living organisms use intricate signaling networks for sensing and responding to changes in the external environment. cAMP, a ubiquitous second messenger, is an important molecular switch that translates environmental signals into regulatory effects in cells (13). As such, a number of microbial pathogens have evolved a set of diverse virulence-enhancing strategies that exploit the cAMP-signaling pathways of their hosts (14). The intracellular functions of cAMP are predominantly mediated by the classic cAMP receptor, protein kinase A (PKA), and the more recently discovered exchange protein directly activated by cAMP (Epac) (15). Thus, far, two isoforms, Epac1 and Epac2, have been identified in humans (16, 17). Epac proteins function by responding to increased intracellular cAMP levels and activating the Ras superfamily small GTPases Ras-proximate 1 and 2 (Rap1 and Rap2). Accumulating evidence demonstrates that the cAMP/Epac1 signaling axis plays key regulatory roles in controlling various cellular functions in endothelial cells in vitro, including cell adhesion (1821), exocytosis (22), tissue plasminogen activator expression (23), suppressor of cytokine signaling 3 (SOCS-3) induction (2427), microtubule dynamics (28, 29), cell–cell junctions, and permeability and barrier functions (3037). Considering the critical importance of endothelial cells in rickettsioses, we examined the functional roles of Epac1 in rickettsial pathogenesis in vivo, taking advantage of the recently generated Epac1 knockout mouse (38) and Epac-specific inhibitors (39, 40) generated from our laboratory. Our studies demonstrate that Epac1 plays a key role in rickettsial infection and represents a therapeutic target for fatal rickettsioses.  相似文献   

6.
In flowering plants, pollen tubes are guided into ovules by multiple attractants from female gametophytes to release paired sperm cells for double fertilization. It has been well-established that Ca2+ gradients in the pollen tube tips are essential for pollen tube guidance and that plasma membrane Ca2+ channels in pollen tube tips are core components that regulate Ca2+ gradients by mediating and regulating external Ca2+ influx. Therefore, Ca2+ channels are the core components for pollen tube guidance. However, there is still no genetic evidence for the identification of the putative Ca2+ channels essential for pollen tube guidance. Here, we report that the point mutations R491Q or R578K in cyclic nucleotide-gated channel 18 (CNGC18) resulted in abnormal Ca2+ gradients and strong pollen tube guidance defects by impairing the activation of CNGC18 in Arabidopsis. The pollen tube guidance defects of cngc18-17 (R491Q) and of the transfer DNA (T-DNA) insertion mutant cngc18-1 (+/−) were completely rescued by CNGC18. Furthermore, domain-swapping experiments showed that CNGC18’s transmembrane domains are indispensable for pollen tube guidance. Additionally, we found that, among eight Ca2+ channels (including six CNGCs and two glutamate receptor-like channels), CNGC18 was the only one essential for pollen tube guidance. Thus, CNGC18 is the long-sought essential Ca2+ channel for pollen tube guidance in Arabidopsis.Pollen tubes deliver paired sperm cells into ovules for double fertilization, and signaling communication between pollen tubes and female reproductive tissues is required to ensure the delivery of sperm cells into the ovules (1). Pollen tube guidance is governed by both female sporophytic and gametophytic tissues (2, 3) and can be separated into two categories: preovular guidance and ovular guidance (1). For preovular guidance, diverse signaling molecules from female sporophytic tissues have been identified, including the transmitting tissue-specific (TTS) glycoprotein in tobacco (4), γ-amino butyric acid (GABA) in Arabidopsis (5), and chemocyanin and the lipid transfer protein SCA in Lilium longiflorum (6, 7). For ovular pollen tube guidance, female gametophytes secrete small peptides as attractants, including LUREs in Torenia fournieri (8) and Arabidopsis (9) and ZmEA1 in maize (10, 11). Synergid cells, central cells, egg cells, and egg apparatus are all involved in pollen tube guidance, probably by secreting different attractants (915). Additionally, nitric oxide (NO) and phytosulfokine peptides have also been implicated in both preovular and ovular pollen tube guidance (1618). Thus, pollen tubes could be guided by diverse attractants in a single plant species.Ca2+ gradients at pollen tube tips are essential for both tip growth and pollen tube guidance (1927). Spatial modification of the Ca2+ gradients leads to the reorientation of pollen tube growth in vitro (28, 29). The Ca2+ gradients were significantly increased in pollen tubes attracted to the micropyles by synergid cells in vivo, compared with those not attracted by ovules (30). Therefore, the Ca2+ gradients in pollen tube tips are essential for pollen tube guidance. The Ca2+ gradients result from external Ca2+ influx, which is mainly mediated by plasma membrane Ca2+ channels in pollen tube tips. Thus, the Ca2+ channels are the key components for regulating the Ca2+ gradients and are consequently essential for pollen tube guidance. Using electrophysiological techniques, inward Ca2+ currents were observed in both pollen grain and pollen tube protoplasts (3136), supporting the presence of plasma membrane Ca2+ channels in pollen tube tips. Recently, a number of candidate Ca2+ channels were identified in pollen tubes, including six cyclic nucleotide-gated channels (CNGCs) and two glutamate receptor-like channels (GLRs) in Arabidopsis (3740). Three of these eight channels, namely CNGC18, GLR1.2, and GLR3.7, were characterized as Ca2+-permeable channels (40, 41) whereas the ion selectivity of the other five CNGCs has not been characterized. We hypothesized that the Ca2+ channel essential for pollen tube guidance could be among these eight channels.In this research, we first characterized the remaining five CNGCs as Ca2+ channels. We further found that CNGC18, out of the eight Ca2+ channels, was the only one essential for pollen tube guidance in Arabidopsis and that its transmembrane domains were indispensable for pollen tube guidance.  相似文献   

7.
Halogenated inhaled general anesthetic agents modulate voltage-gated ion channels, but the underlying molecular mechanisms are not understood. Many general anesthetic agents regulate voltage-gated Na+ (NaV) channels, including the commonly used drug sevoflurane. Here, we investigated the putative binding sites and molecular mechanisms of sevoflurane action on the bacterial NaV channel NaChBac by using a combination of molecular dynamics simulation, electrophysiology, and kinetic analysis. Structural modeling revealed multiple sevoflurane interaction sites possibly associated with NaChBac modulation. Electrophysiologically, sevoflurane favors activation and inactivation at low concentrations (0.2 mM), and additionally accelerates current decay at high concentrations (2 mM). Explaining these observations, kinetic modeling suggests concurrent destabilization of closed states and low-affinity open channel block. We propose that the multiple effects of sevoflurane on NaChBac result from simultaneous interactions at multiple sites with distinct affinities. This multiple-site, multiple-mode hypothesis offers a framework to study the structural basis of general anesthetic action.General anesthetic agents have been in use for more than 160 y. However, we still understand relatively little about their mechanisms of action, which greatly limits our ability to design safer and more effective general anesthetic agents. Ion channels of the central nervous system are known to be key targets of general anesthetic agents, as their modulation can account for the endpoints and side effects of general anesthesia (14). Many families of ion channels are modulated by general anesthetic agents, including ligand-gated, voltage-gated, and nongated ion channels (2, 57). Mammalian voltage-gated Na+ (NaV) channels, which mediate the upstroke of the action potential, are regulated by numerous inhaled general anesthetic agents (814), which generally cause inhibition. Previous work showed that inhaled general anesthetic agents, including sevoflurane, isoflurane, desflurane, and halothane, mediate inhibition by increasing the rate of Na+ channel inactivation, hyperpolarizing steady-state inactivation, and slowing recovery from inactivation (11, 1518). Inhibition of presynaptic NaV channels in the spinal cord is proposed to lead to inhibition of neurotransmitter release, facilitating immobilization—one of the endpoints of general anesthesia (14, 19, 20). Despite the importance of NaV channels as general anesthetic targets, little is known about interaction sites or the mechanisms of action.What is known about anesthetic sites in NaV channels comes primarily from the local anesthetic field. Local anesthetic agent binding to NaV channels is well characterized. These amphiphilic drugs enter the channel pore from the intracellular side, causing open-channel block (21). Investigating molecular mechanisms of mammalian NaV channel modulation by general anesthetic agents has been complicated by the lack of high-resolution structures of these channels as a result of their large size and pseudotetrameric organization. However, the recent discovery of the smaller, tetrameric bacterial Na+ channel family has provided an invaluable tool to characterize the structural features of NaV channels and investigate their interactions with general anesthetic agents at the molecular level (22, 23). Several bacterial Na+ channels have been crystallized (2427). These channels have a classical domain structure in which helices S1–S4 form the voltage sensor domain (VSD), S5 and S6 form the pore, and the S4–S5 linker connects the voltage sensor to the pore domain. One notable structural feature is the presence of “fenestrations” or hydrophobic tunnels through the pore domain (24).Although crystal structures are not yet reported, the bacterial Na+ channel NaChBac has been extensively characterized by electrophysiology (22, 2836). Additionally NaChBac exhibits conserved slow open channel block in response to local and general anesthetic agents (15, 37). These anesthetic agents reduce peak current and accelerate current decay, making it conceivable that local and general anesthetic agents could share a site of action in NaChBac. The local anesthetic binding site identified in the central cavity of the mammalian NaV1.2 channel, which mediates open channel block, is partially conserved in NaChBac (37, 38). A recent molecular dynamics (MD) modeling study found that isoflurane, which inhibits NaChBac (15), interacts with multiple regions of this channel, including the pore, the selectivity filter, and the S4–S5 linker/S6 interface (39). Although the importance of these interactions on the modulation of mammalian NaV channels remains to be determined, the available data indicate that NaChBac is currently one of the best starting points to investigate the mechanisms of action of sevoflurane.Here, we investigated NaChBac to gain structural insight into the mechanisms of inhaled anesthetic modulation of NaV channels. The focus of this work is sevoflurane because this anesthetic is commonly used in clinical settings and is a known inhibitor of several mammalian NaV channels (NaV 1.4, 1.7, and 1.8) (11, 13). A three-pronged approach incorporating MD simulation, whole-cell patch-clamp electrophysiology, and kinetic modeling suggests that sevoflurane acts on multiple sites to alter gating and permeation. Whereas the effect on gating results from modulating activation and inactivation gating at low concentrations (0.2 mM), the permeation effect is apparent at high concentrations (2 mM) and results from open channel block (2 mM). Although the net inhibitory effect of these multisite interactions is consistent with anesthetic-induced reduction of neuronal firing, general anesthesia does not simply result from a global reduction in firing. General anesthesia depends on complex mechanisms throughout the brain, which include increases and decreases in firing (3). Thus, precisely how Na+ channel activation by sevoflurane fits into the global effects of anesthesia remains to be seen. The present work helps elucidate the molecular mechanism of sevoflurane action on NaV channels.  相似文献   

8.
We examined the origins and functional evolution of the Shaker and KCNQ families of voltage-gated K+ channels to better understand how neuronal excitability evolved. In bilaterians, the Shaker family consists of four functionally distinct gene families (Shaker, Shab, Shal, and Shaw) that share a subunit structure consisting of a voltage-gated K+ channel motif coupled to a cytoplasmic domain that mediates subfamily-exclusive assembly (T1). We traced the origin of this unique Shaker subunit structure to a common ancestor of ctenophores and parahoxozoans (cnidarians, bilaterians, and placozoans). Thus, the Shaker family is metazoan specific but is likely to have evolved in a basal metazoan. Phylogenetic analysis suggested that the Shaker subfamily could predate the divergence of ctenophores and parahoxozoans, but that the Shab, Shal, and Shaw subfamilies are parahoxozoan specific. In support of this, putative ctenophore Shaker subfamily channel subunits coassembled with cnidarian and mouse Shaker subunits, but not with cnidarian Shab, Shal, or Shaw subunits. The KCNQ family, which has a distinct subunit structure, also appears solely within the parahoxozoan lineage. Functional analysis indicated that the characteristic properties of Shaker, Shab, Shal, Shaw, and KCNQ currents evolved before the divergence of cnidarians and bilaterians. These results show that a major diversification of voltage-gated K+ channels occurred in ancestral parahoxozoans and imply that many fundamental mechanisms for the regulation of action potential propagation evolved at this time. Our results further suggest that there are likely to be substantial differences in the regulation of neuronal excitability between ctenophores and parahoxozoans.Voltage-gated K+ channels are highly conserved among bilaterian metazoans and play a central role in the regulation of excitation in neurons and muscle. Understanding the functional evolution of these channels may therefore provide important insights into how neuromuscular excitation evolved within the Metazoa. Three major gene families, Shaker, KCNQ, and Ether-a-go-go (EAG) encode all voltage-gated K+ channels in bilaterians (1, 2). In this study, we examine the functional evolution and origins of the Shaker and KCNQ gene families. Shaker family channels can be definitively identified by a unique subunit structure that includes both a voltage-gated K+ channel core and a family-specific cytoplasmic domain within the N terminus known as the T1 domain. T1 mediates assembly of Shaker family subunits into functional tetrameric channels (3, 4). KCNQ channels are also tetrameric but lack a T1 domain and use a distinct coiled-coil assembly domain in the C terminus (5, 6). KCNQ channels can be identified by the presence of this family-specific assembly motif and high amino acid conservation within the K+ channel core. Both channel families are found in cnidarians (1, 7) and thus predate the divergence of cnidarians and bilaterians, but their ultimate evolutionary origins have not yet been defined.Shaker family K+ channels serve diverse roles in the regulation of neuronal firing and can be divided into four gene subfamilies based on function and sequence homology: Shaker, Shab, Shal, and Shaw (8, 9). The T1 assembly domain is only compatible between subunits from the same gene subfamily (4, 10) and thus serves to keep the subfamilies functionally segregated. Shaker subfamily channels activate rapidly near action potential threshold and range from rapidly inactivating to noninactivating. Multiple roles for Shaker channels in neurons and muscles have been described, but their most unique and fundamental role may be that of axonal action potential repolarization. Shaker channels are clustered to the axon initial segment and nodes of Ranvier in vertebrate neurons (1113) and underlie the delayed rectifier in squid giant axons (14). The Shaker subfamily is diverse in cnidarians (15, 16), and the starlet sea anemone Nematostella vectensis has functional orthologs of most identified Shaker current types observed in bilaterians (16).The Shab and Shal gene subfamilies encode somatodendritic delayed rectifiers and A currents, respectively (1720). Shab channels are important for maintaining sustained firing (21, 22), whereas the Kv4-based A current modulates spike threshold and frequency (17). Shab and Shal channels are present in cnidarians, but cnidarian Shab channels have not been functionally characterized, and the only cnidarian Shal channels expressed to date display atypical voltage dependence and kinetics compared with bilaterian channels (23). Shaw channels are rapid, high-threshold channels specialized for sustaining fast firing in vertebrates (24, 25) but have a low activation threshold and may contribute to resting potential in Drosophila (19, 26, 27). A Caenorhabditis elegans Shaw has slow kinetics but a high activation threshold (28), and a single expressed cnidarian Shaw channel has the opposite: a low activation threshold but relatively fast kinetics (29). Thus, the ancestral properties and function of Shaw channels is not yet understood. Further functional characterization of cnidarian Shab, Shal, and Shaw channels would provide a better understanding of the evolutionary status of the Shaker family in early parahoxozoans.KCNQ family channels underlie the M current in vertebrate neurons (30) that regulates subthreshold excitability (31). The M current provides a fundamental mechanism for regulation of firing threshold through the Gq G-protein pathway because KCNQ channels require phosphatidylinositol 4,5-bisphosphate (PIP2) for activation (32, 33). PIP2 hydrolysis and subsequent KCNQ channel closure initiated by Gq-coupled receptors produces slow excitatory postsynaptic potentials, during which the probability of firing is greatly increased (32, 33). The key functional adaptations of KCNQ channels for this physiological role that can be observed in vitro are (i) a requirement for PIP2 to couple voltage-sensor activation to pore opening (34, 35), and (ii) a hyperpolarized voltage–activation curve that allows channels to open below typical action potential thresholds. Both key features are found in vertebrate (30, 34, 3638), Drosophila (39), and C. elegans (40) KCNQ channels, suggesting they may have been present in KCNQ channels in a bilaterian ancestor. Evolution of the M current likely represented a major advance in the ability to modulate the activity of neuronal circuits, but it is not yet clear when PIP2-dependent KCNQ channels first evolved.Here, we examine the origins and functional evolution of the Shaker and KCNQ gene families. If we assume the evolution of neuronal signaling provided a major selective pressure for the functional diversification of voltage-gated K+ channels, then we can hypothesize that the appearance of these gene families might accompany the emergence of the first nervous systems or a major event in nervous system evolution. Recent phylogenies that place the divergence of ctenophores near the root of the metazoan tree suggest that the first nervous systems, or at least the capacity to make neurons, may have been present in a basal metazoan ancestor (4143) (Fig. S1). One hypothesis then is that much of the diversity of metazoan voltage-gated channels should be shared between ctenophores and parahoxozoans [cnidarians, bilaterians, and placozoans (44)]. However, genome analysis indicates that many “typical” neuronal genes are missing in ctenophores and the sponges lack a nervous system, leading to the suggestion that extant nervous systems may have evolved independently in ctenophores and parahoxozoans (42, 45). Thus, a second hypothesis is that important steps in voltage-gated K+ channel evolution might have occurred separately in ctenophores and parahoxozoans. We tested these hypotheses by carefully examining the phylogenetic distribution and functional evolution of Shaker and KCNQ family K+ channels. Our results support a model in which major innovations in neuromuscular excitability occurred specifically within the parahoxozoan lineage.  相似文献   

9.
10.
The identity and forms of activating ligands for ion channels are fundamental to their physiological roles in rapid electrical signaling. P2X receptor channels are ATP-activated cation channels that serve important roles in sensory signaling and inflammation, yet the active forms of the nucleotide are unknown. In physiological solutions, ATP is ionized and primarily found in complex with Mg2+. Here we investigated the active forms of ATP and found that the action of MgATP2− and ATP4− differs between subtypes of P2X receptors. The slowly desensitizing P2X2 receptor can be activated by free ATP, but MgATP2− promotes opening with very low efficacy. In contrast, both free ATP and MgATP2− robustly open the rapidly desensitizing P2X3 subtype. A further distinction between these two subtypes is the ability of Mg2+ to regulate P2X3 through a distinct allosteric mechanism. Importantly, heteromeric P2X2/3 channels present in sensory neurons exhibit a hybrid phenotype, characterized by robust activation by MgATP2− and weak regulation by Mg2+. These results reveal the existence of two classes of homomeric P2X receptors with differential sensitivity to MgATP2− and regulation by Mg2+, and demonstrate that both restraining mechanisms can be disengaged in heteromeric channels to form fast and sensitive ATP signaling pathways in sensory neurons.Seven subtypes of P2X receptors have been identified in mammals that can form either homomeric (P2X1, P2X2, P2X3, P2X4, P2X5, P2X7) or heteromeric (P2X1/2, P2X1/4, P2X1/5, P2X2/3, P2X2/5, P2X2/6, P2X4/6, and possibly, P2X4/7) channels (18). These subtypes of P2X receptors have distinct gating properties, pharmacology, and cellular distributions. P2X1 and P2X3 receptors desensitize within a few hundred milliseconds when opened by ATP, and their distributions are restricted to either smooth muscle cells and platelets (P2X1) or a subset of sensory neurons (P2X3) (1, 914). P2X2 and P2X4 receptors exhibit slow desensitization during prolonged ATP application, and these receptors are the most abundant subtypes in the central nervous system (15). P2X2 subunits also express in a subset of sensory neurons; however, in these cells they only form heteromeric channels with P2X3 subunits (12, 16, 17). In sensory neurons, P2X3 homomeric channels together with P2X2/3 heteromeric channels play important roles in mediating the primary sensory effects of ATP, and knock-out animals with either P2X3 deletion or P2X2 and P2X3 double-deletions have revealed critical roles in taste, pain, oxygen sensing, and bladder filling (1720).A long-standing conundrum in P2X receptor-mediated signaling concerns the forms of ATP that activate these channels. In neutral solutions, ATP is ionized and exists mostly as free ATP (ATP4−), an efficient chelator of divalent cations such as Mg2+, and to a lesser extent Ca2+ (21). In extracellular biological compartments, such as the synaptic cleft, Ca2+ and Mg2+ are present in the millimolar range, and therefore only a relatively small fraction of ATP released from vesicles is present in the free form. Although a range of important studies have explored the regulatory effects of Ca2+ and Mg2+ on P2X receptor channels (2230), the essential question of which forms of ATP serve as agonists remains unresolved. Several previous studies have reported that P2X2, P2X7, and the native P2X receptors in cilia are activated by ATP in solutions containing low concentrations of divalent cations, and that the addition of divalent cations shifts the concentration dependence for activation of the channels to higher ATP concentrations, suggesting that either ATP4− is the most active form of ATP or that divalent cations regulate those subtypes through allosteric mechanisms (27, 3036). In the present study, we investigated the form(s) of ATP that serve as agonists for a range of subtypes of P2X receptor channels. Our primary focus was to determine whether ATP4− or MgATP2− are the principal agonists and to explore whether Mg2+ might serve specific regulatory roles. Our results demonstrate that the action of MgATP2− and ATP4− differ between subtypes of P2X receptors, and reveal that heteromeric channels can have unique hybrid phenotypes, findings that will be crucial for understanding the physiological functions of these channels in both the peripheral and central nervous systems.  相似文献   

11.
Modulation of P/Q-type Ca2+ currents through presynaptic voltage-gated calcium channels (CaV2.1) by binding of Ca2+/calmodulin contributes to short-term synaptic plasticity. Ca2+-binding protein-1 (CaBP1) and Visinin-like protein-2 (VILIP-2) are neurospecific calmodulin-like Ca2+ sensor proteins that differentially modulate CaV2.1 channels, but how they contribute to short-term synaptic plasticity is unknown. Here, we show that activity-dependent modulation of presynaptic CaV2.1 channels by CaBP1 and VILIP-2 has opposing effects on short-term synaptic plasticity in superior cervical ganglion neurons. Expression of CaBP1, which blocks Ca2+-dependent facilitation of P/Q-type Ca2+ current, markedly reduced facilitation of synaptic transmission. VILIP-2, which blocks Ca2+-dependent inactivation of P/Q-type Ca2+ current, reduced synaptic depression and increased facilitation under conditions of high release probability. These results demonstrate that activity-dependent regulation of presynaptic CaV2.1 channels by differentially expressed Ca2+ sensor proteins can fine-tune synaptic responses to trains of action potentials and thereby contribute to the diversity of short-term synaptic plasticity.Neurons fire repetitively in different frequencies and patterns, and activity-dependent alterations in synaptic strength result in diverse forms of short-term synaptic plasticity that are crucial for information processing in the nervous system (13). Short-term synaptic plasticity on the time scale of milliseconds to seconds leads to facilitation or depression of synaptic transmission through changes in neurotransmitter release. This form of plasticity is thought to result from residual Ca2+ that builds up in synapses during repetitive action potentials and binds to a Ca2+ sensor distinct from the one that evokes neurotransmitter release (1, 2, 4, 5). However, it remains unclear how changes in residual Ca2+ cause short-term synaptic plasticity and how neurotransmitter release is regulated to generate distinct patterns of short-term plasticity.In central neurons, voltage-gated calcium (CaV2.1) channels are localized in high density in presynaptic active zones where their P/Q-type Ca2+ current triggers neurotransmitter release (611). Because synaptic transmission is proportional to the third or fourth power of Ca2+ entry through presynaptic CaV2.1 channels, small changes in Ca2+ current have profound effects on synaptic transmission (2, 12). Studies at the calyx of Held synapse have provided important insights into the contribution of presynaptic Ca2+ current to short-term synaptic plasticity (1317). CaV2.1 channels are required for synaptic facilitation, and Ca2+-dependent facilitation and inactivation of the P/Q-type Ca2+ currents are correlated temporally with synaptic facilitation and rapid synaptic depression (1317).Molecular interactions between Ca2+/calmodulin (CaM) and CaV2.1 channels induce sequential Ca2+-dependent facilitation and inactivation of P/Q-type Ca2+ currents in nonneuronal cells (1821). Facilitation and inactivation of P/Q-type currents are dependent on Ca2+/CaM binding to the IQ-like motif (IM) and CaM-binding domain (CBD) of the CaV2.1 channel, respectively (20, 21). This bidirectional regulation serves to enhance channel activity in response to short bursts of depolarizations and then to decrease activity in response to long bursts. In synapses of superior cervical ganglion (SCG) neurons expressing exogenous CaV2.1 channels, synaptic facilitation is induced by repetitive action potentials, and mutation of the IM and CBD motifs prevents synaptic facilitation and inhibits the rapid phase of synaptic depression (22). Thus, in this model synapse, regulation of presynaptic CaV2.1 channels by binding of Ca2+/CaM can contribute substantially to the induction of short-term synaptic plasticity by residual Ca2+.CaM is expressed ubiquitously, but short-term plasticity has great diversity among synapses, and the potential sources of this diversity are unknown. How could activity-dependent regulation of presynaptic CaV2.1 channels contribute to the diversity of short-term synaptic plasticity? CaM is the founding member of a large family of Ca2+ sensor (CaS) proteins that are differentially expressed in central neurons (2325). Two CaS proteins, Ca2+-binding protein-1 (CaBP1) and Visinin-like protein-2(VILIP-2), modulate facilitation and inactivation of CaV2.1 channels in opposite directions through interaction with the bipartite regulatory site in the C-terminal domain (26, 27), and they have varied expression in different types of central neurons (23, 25, 28). CaBP1 strongly enhances inactivation and prevents facilitation of CaV2.1 channel currents, whereas VILIP-2 slows inactivation and enhances facilitation of CaV2.1 currents during trains of stimuli (26, 27). Molecular analyses show that the N-terminal myristoylation site and the properties of individual EF-hand motifs in CaBP1 and VILIP-2 determine their differential regulation of CaV2.1 channels (27, 2931). However, the role of CaBP1 and VILIP-2 in the diversity of short-term synaptic plasticity is unknown, and the high density of Ca2+ channels and unique Ca2+ dynamics at the presynaptic active zone make extrapolation of results from studies in nonneuronal cells uncertain. We addressed this important question directly by expressing CaBP1 and VILIP-2 in presynaptic SCG neurons and analyzing their effects on synaptic plasticity. Our results show that CaM-related CaS proteins can serve as sensitive bidirectional switches that fine-tune the input–output relationships of synapses depending on their profile of activity and thereby maintain the balance of facilitation versus depression by the regulation of presynaptic CaV2.1 channels.  相似文献   

12.
Stimulating presynaptic terminals can increase the proton concentration in synapses. Potential receptors for protons are acid-sensing ion channels (ASICs), Na+- and Ca2+-permeable channels that are activated by extracellular acidosis. Those observations suggest that protons might be a neurotransmitter. We found that presynaptic stimulation transiently reduced extracellular pH in the amygdala. The protons activated ASICs in lateral amygdala pyramidal neurons, generating excitatory postsynaptic currents. Moreover, both protons and ASICs were required for synaptic plasticity in lateral amygdala neurons. The results identify protons as a neurotransmitter, and they establish ASICs as the postsynaptic receptor. They also indicate that protons and ASICs are a neurotransmitter/receptor pair critical for amygdala-dependent learning and memory.Although homeostatic mechanisms generally maintain the brain’s extracellular pH within narrow limits, neural activity can induce transient and localized pH fluctuations. For example, acidification may occur when synaptic vesicles, which have a pH of ∼5.2–5.7 (13), release their contents into the synapse. Studies of mammalian cone photoreceptors showed that synaptic vesicle exocytosis rapidly reduced synaptic cleft pH by an estimated 0.2–0.6 units (46). Transient synaptic cleft acidification also occurred with GABAergic transmission (7). Some, but not all, studies also reported that high-frequency stimulation (HFS) transiently acidified hippocampal brain slices, likely as a result of the release of synaptic vesicle contents (8, 9). Neurotransmission also induces a slower, more prolonged alkalinization (10, 11). In addition to release of synaptic vesicle protons, neuronal and glial H+ and HCO3 transporters, channels, H+-ATPases, and metabolism might influence extracellular pH (1012).ASICs are potential targets of reduced extracellular pH. ASICs are Na+-permeable and, to a lesser extent, Ca2+-permeable channels that are activated by extracellular acidosis (1319). In the brain, ASICs consist of homotrimeric and heterotrimeric complexes of ASIC1a, ASIC2a, and ASIC2b. The ASIC1a subunit is required for acid-activation in the physiological range (>pH 5.0) (20, 21). Several observations indicate that ASIC are located postsynaptically. ASICs are located on dendritic spines. Although similar to glutamate receptors, they are also present on dendrites and cell bodies (20, 2224). ASIC subunits interact with postsynaptic scaffolding proteins, including postsynaptic density protein 95 and protein interacting with C-kinase-1 (20, 2429). In addition, ASICs are enriched in synaptosome-containing brain fractions (20, 24, 30).Although these observations raised the possibility that protons might be a neurotransmitter, postsynaptic ASIC currents have not been detected in cultured hippocampal neurons (31, 32), and whether localized pH transients might play a signaling role in neuronal communication remains unclear. In previous studies of hippocampal brain slices, extracellular field potential recordings suggested impaired hippocampal long-term potentiation (LTP) in ASIC1a−/− mice (20), although another study did not detect an effect of ASIC1a (33). Another study using microisland cultures of hippocampal neurons suggested that the probability of neurotransmitter release increased in ASIC1a−/− mice (32).Here, we tested the hypothesis that protons are a neurotransmitter and that ASICs are the receptor. Criteria to identify substances as neurotransmitters have been proposed (34). Beg and colleagues (35) used these criteria to conclude that protons are a transmitter released from Caenorhabditis elegans intestine to cause muscle contraction. Key questions about whether protons meet criteria for a neurotransmitter are: Does presynaptic stimulation increase the extracellular proton concentration? Do protons activate currents in postsynaptic cells? Can exogenously applied protons reproduce effects of endogenous protons? What is the postsynaptic proton receptor? We studied lateral amygdala brain slices because amygdala-dependent fear-related behavior depends on a pH reduction (36). In addition, ASICs are abundantly expressed there, and ASIC1a−/− mice have impaired fear-like behavior (3638).  相似文献   

13.
14.
Human Bestrophin 1 (hBest1) is a calcium-activated chloride channel that regulates neuronal excitability, synaptic activity, and retinal homeostasis. Mutations in hBest1 cause the autosomal-dominant Best macular dystrophy (BMD). Because hBest1 mutations cause BMD, but a knockout does not, we wondered if hBest1 mutants exert a dominant negative effect through interaction with other calcium-activated chloride channels, such as hBest2, 3, or 4, or transmembrane member 16A (TMEM16A), a member of another channel family. The subunit architecture of Best channels is debated, and their ability to form heteromeric channel assemblies is unclear. Using single-molecule subunit analysis, we find that each of hBest1, 2, 3, and 4 forms a homotetrameric channel. Despite considerable conservation among hBests, hBest1 has little or no interaction with other hBests or mTMEM16A. We identify the domain responsible for assembly specificity. This domain also plays a role in channel function. Our results indicate that Best channels preferentially self-assemble into homotetramers.Bestrophin 1 is calcium-activated chloride channel (CACC) and has been shown to express in a variety of tissues (1). In the brain, Best1 plays a crucial role in the regulation of neuronal excitability and synaptic activity by releasing gliotransmitters such as glutamate and GABA from astrocytes upon G-protein–coupled receptor (GPCR) activation (25). In retinal pigment epithelium (RPE) cells, Best1 plays an important role in retinal homeostasis (1), and mutations in human Best1 have been implicated in several retinal degenerative diseases including Best Macular Dystrophy (BMD) (612) and Retinitis Pigmentosa (13).The human bestrophin family includes three additional members; hBest2, 3, and 4 (14, 15). All four members function in heterologous cells (1518) as anion-selective channels, whose main physiological charge carrier is chloride (15, 17, 1922), but which also permeate glutamate and GABA (3, 4).hBest1 contains six hydrophobic segments (S1–S6), with both N and C termini residing inside the cell. Two topology models have been proposed for hBest1 (15, 23). In the first model, S1, S2, S4, and S6 traverse the membrane, whereas S3 is intracellular and S5 forms a reentrant loop from outside (15). A more recent model has S1, S2, S5, and S6 traversing the membrane and S3 and S4, although hydrophobic, remaining on the intracellular side (23).Best1 is activated by Ca2+ with a Kd of ∼150 nM (24). Several pieces of evidence suggest that this activation is due to direct binding of Ca2+ (25, 26) to an EF hand located immediately after S6 (24). It is unclear how Ca2+-binding gates the channel and whether the EF hand is part of the gate or communicates with it.Although much progress has been made on Best channels (15, 17, 1922, 27), several fundamental aspects of the structure and function of this channel family are not understood. First, previous biochemical analysis has indicated that Best1 is multimeric (22, 27) but led to conflicting assessments of the number of subunits in the channel, with experiments on human Best1 suggesting a tetramer or pentamer (22) but experiments on porcine Best1 suggesting a dimer (27). Second, although coimmunoprecipitation suggests that hBest1 interacts with hBest2 (22), it is unclear if this is direct interaction. Moreover, virtually nothing is known about the determinants of assembly.In this study, we used single-molecule subunit counting and colocalization to address four major questions about the subunit assembly and function of hBest channels: (i) What is the subunit stoichiometry of hBest channels? (ii) Does hBest1 coassemble with any other member of the hBest family or with a member of different CACC family, transmembrane member 16A (TMEM16A)? (iii) How is subunit assembly specified? (iv) Does the assembly determinant play any role in channel function?  相似文献   

15.
Calcium (Ca2+) released from the sarcoplasmic reticulum (SR) is crucial for excitation–contraction (E–C) coupling. Mitochondria, the major source of energy, in the form of ATP, required for cardiac contractility, are closely interconnected with the SR, and Ca2+ is essential for optimal function of these organelles. However, Ca2+ accumulation can impair mitochondrial function, leading to reduced ATP production and increased release of reactive oxygen species (ROS). Oxidative stress contributes to heart failure (HF), but whether mitochondrial Ca2+ plays a mechanistic role in HF remains unresolved. Here, we show for the first time, to our knowledge, that diastolic SR Ca2+ leak causes mitochondrial Ca2+ overload and dysfunction in a murine model of postmyocardial infarction HF. There are two forms of Ca2+ release channels on cardiac SR: type 2 ryanodine receptors (RyR2s) and type 2 inositol 1,4,5-trisphosphate receptors (IP3R2s). Using murine models harboring RyR2 mutations that either cause or inhibit SR Ca2+ leak, we found that leaky RyR2 channels result in mitochondrial Ca2+ overload, dysmorphology, and malfunction. In contrast, cardiac-specific deletion of IP3R2 had no major effect on mitochondrial fitness in HF. Moreover, genetic enhancement of mitochondrial antioxidant activity improved mitochondrial function and reduced posttranslational modifications of RyR2 macromolecular complex. Our data demonstrate that leaky RyR2, but not IP3R2, channels cause mitochondrial Ca2+ overload and dysfunction in HF.Type 2 ryanodine receptor/Ca2+ release channel (RyR2) and type 2 inositol 1,4,5-trisphosphate receptor (IP3R2) are the major intracellular Ca2+ release channels in the heart (13). RyR2 is essential for cardiac excitation–contraction (E–C) coupling (2), whereas the role of IP3R2 in cardiomyocytes is less well understood (3). E–C coupling requires energy in the form of ATP produced primarily by oxidative phosphorylation in mitochondria (48).Both increased and reduced mitochondrial Ca2+ levels have been implicated in mitochondrial dysfunction and increased reactive oxygen species (ROS) production in heart failure (HF) (6, 7, 917). Albeit Ca2+ is required for activation of key enzymes (i.e., pyruvate dehydrogenase phosphatase, isocitrate dehydrogenase, and α-ketoglutarate dehydrogenase) in the tricarboxylic acid (also known as Krebs) cycle (18, 19), excessive mitochondrial Ca2+ uptake has been associated with cellular dysfunction (14, 20). Furthermore, the exact source of mitochondrial Ca2+ has not been clearly established. Given the intimate anatomical and functional association between the sarcoplasmic reticulum (SR) and mitochondria (6, 21, 22), we hypothesized that SR Ca2+ release via RyR2 and/or IP3R2 channels in cardiomyocytes could lead to mitochondrial Ca2+ accumulation and dysfunction contributing to oxidative overload and energy depletion.  相似文献   

16.
Rod photoreceptors consist of an outer segment (OS) and an inner segment. Inside the OS a biochemical machinery transforms the rhodopsin photoisomerization into electrical signal. This machinery has been treated as and is thought to be homogenous with marginal inhomogeneities. To verify this assumption, we developed a methodology based on special tapered optical fibers (TOFs) to deliver highly localized light stimulations. By using these TOFs, specific regions of the rod OS could be stimulated with spots of light highly confined in space. As the TOF is moved from the OS base toward its tip, the amplitude of saturating and single photon responses decreases, demonstrating that the efficacy of the transduction machinery is not uniform and is 5–10 times higher at the base than at the tip. This gradient of efficacy of the transduction machinery is attributed to a progressive depletion of the phosphodiesterase along the rod OS. Moreover we demonstrate that, using restricted spots of light, the duration of the photoresponse along the OS does not increase linearly with the light intensity as with diffuse light.Vertebrate photoreceptors, rods and cones, are morphologically specialized light sensing neurons and consist of four parts: an outer segment (OS), an inner segment (IS), the nuclear region, and the synapse (1). The OS of rod photoreceptors is stacked with thousands of lipid discs containing rhodopsin molecules that absorb photons (24). They are surrounded by a plasma membrane and differ in their lipid and protein composition.It is known that within 1 s, each excited rhodopsin, densely packed in the disc membrane, activates tens of G proteins (named transducin), each of which activates one phosphodiesterase (PDE) molecule (57). Activated PDEs rapidly hydrolyze cytoplasmic cyclic guanosine monophosphate (cGMP) thereby closing cyclic nucleotide-gated (CNG) channels (8, 9). In darkness, a current carried by Na+, K+, and Ca2+ ions, which is known as the photocurrent, enters via the CNG channels into the OS and is pumped out by Na+/K+ ATPase, located in the IS (1). This current can be recorded using suction electrodes (3, 10).The existence of distinct compartments on photoreceptor OSs is a consequence of their function. The OS is a highly modified nonmotile cilium developed for the absorption of light that is translated into the electrical signal. The IS and the nuclear region—containing the organelles and the nucleus—are dedicated to metabolism, homeostasis, and synthesis of the membrane and transportation of proteins and lipids supplied to the OS by an extensive trafficking through the tight restriction connecting the OS to the IS (11). Considering that, unlike typical cilia, the OS is continuously renewed throughout its entire life and given the varying density and asymmetry of the molecules involved, it is not surprising to find an efficacy gradient of phototransduction between the base and the tip. Indeed, the cholesterol (12, 13), the phospholipids (1315), the CNG channels, the PDE (16), and the rhodopsin molecules (12) have different distributions between discs and plasma membranes; moreover, cholesterol (12, 13) and PDE (16) also have a different spatial distribution along the OS.The biochemical machinery underlying phototransduction has been thoroughly investigated and described in a quantitative way (1). Previous studies in the early 1980s have demonstrated a lower sensitivity and slower kinetics of flash responses (3, 17, 18) at the tip of the OSs and that the amplitude of the single photon response is smaller at the OS tip than at the OS base (18). Despite these experimental observations, most of the models of phototransduction, developed subsequently, assumed a homogeneous efficacy of the transduction machinery along the OS (1, 1922). In the present paper, by using highly confined spots of light obtained with apertureless tapered optical fibers (TOFs), we reveal that the biochemical machinery inside the rod OS has a gradient of efficacy much larger than previously thought (17, 18), demonstrating that the assumption of a homogeneous OS is not true. By using better technology, it is possible to obtain more confined spots of light necessary to uncover that OSs, not only have a specialized and different structure and function from that of ISs, but are also formed by inhomogeneous compartments (the internal discs) and that the efficiency of the transduction machinery is 5–10 times lower at the OS tip than at the base. A progressive depletion of PDE might represent a possible explanation of both this loss of efficacy and the role of PDE and cGMP diffusion in determining the peculiar kinetic features of the photoresponses we obtained by these highly confined flashes of light.  相似文献   

17.
Ca2+ influx triggers the fusion of synaptic vesicles at the presynaptic active zone (AZ). Here we demonstrate a role of Ras-related in brain 3 (Rab3)–interacting molecules 2α and β (RIM2α and RIM2β) in clustering voltage-gated CaV1.3 Ca2+ channels at the AZs of sensory inner hair cells (IHCs). We show that IHCs of hearing mice express mainly RIM2α, but also RIM2β and RIM3γ, which all localize to the AZs, as shown by immunofluorescence microscopy. Immunohistochemistry, patch-clamp, fluctuation analysis, and confocal Ca2+ imaging demonstrate that AZs of RIM2α-deficient IHCs cluster fewer synaptic CaV1.3 Ca2+ channels, resulting in reduced synaptic Ca2+ influx. Using superresolution microscopy, we found that Ca2+ channels remained clustered in stripes underneath anchored ribbons. Electron tomography of high-pressure frozen synapses revealed a reduced fraction of membrane-tethered vesicles, whereas the total number of membrane-proximal vesicles was unaltered. Membrane capacitance measurements revealed a reduction of exocytosis largely in proportion with the Ca2+ current, whereas the apparent Ca2+ dependence of exocytosis was unchanged. Hair cell-specific deletion of all RIM2 isoforms caused a stronger reduction of Ca2+ influx and exocytosis and significantly impaired the encoding of sound onset in the postsynaptic spiral ganglion neurons. Auditory brainstem responses indicated a mild hearing impairment on hair cell-specific deletion of all RIM2 isoforms or global inactivation of RIM2α. We conclude that RIM2α and RIM2β promote a large complement of synaptic Ca2+ channels at IHC AZs and are required for normal hearing.Tens of CaV1.3 Ca2+ channels are thought to cluster within the active zone (AZ) membrane underneath the presynaptic density of inner hair cells (IHCs) (14). They make up the key signaling element, coupling the sound-driven receptor potential to vesicular glutamate release (57). The mechanisms governing the number of Ca2+ channels at the AZ as well as their spatial organization relative to membrane-tethered vesicles are not well understood. Disrupting the presynaptic scaffold protein Bassoon diminishes the numbers of Ca2+ channels and membrane-tethered vesicles at the AZ (2, 8). However, the loss of Bassoon is accompanied by the loss of the entire synaptic ribbon, which makes it challenging to distinguish the direct effects of gene disruption from secondary effects (9).Among the constituents of the cytomatrix of the AZ, RIM1 and RIM2 proteins are prime candidates for the regulation of Ca2+ channel clustering and function (10, 11). The family of RIM proteins has seven identified members (RIM1α, RIM1β, RIM2α, RIM2β, RIM2γ, RIM3γ, and RIM4γ) encoded by four genes (RIM1–RIM4). All isoforms contain a C-terminal C2 domain but differ in the presence of additional domains. RIM1 and RIM2 interact with Ca2+ channels, most other proteins of the cytomatrix of the AZ, and synaptic vesicle proteins. They interact directly with the auxiliary β (CaVβ) subunits (12, 13) and pore-forming CaVα subunits (14, 15). In addition, RIMs are indirectly linked to Ca2+ channels via RIM-binding protein (14, 16, 17). A regulation of biophysical channel properties has been demonstrated in heterologous expression systems for RIM1 (12) and RIM2 (13).A role of RIM1 and RIM2 in clustering Ca2+ channels at the AZ was demonstrated by analysis of RIM1/2-deficient presynaptic terminals of cultured hippocampal neurons (14), auditory neurons in slices (18), and Drosophila neuromuscular junction (19). Because α-RIMs also bind the vesicle-associated protein Ras-related in brain 3 (Rab3) via the N-terminal zinc finger domain (20), they are also good candidates for molecular coupling of Ca2+ channels and vesicles (18, 21, 22). Finally, a role of RIMs in priming of vesicles for fusion is the subject of intense research (18, 2127). RIMs likely contribute to priming via disinhibiting Munc13 (26) and regulating vesicle tethering (27). Here, we studied the expression and function of RIM in IHCs. We combined molecular, morphologic, and physiologic approaches for the analysis of RIM2α knockout mice [RIM2α SKO (28); see Methods] and of hair cell-specific RIM1/2 knockout mice (RIM1/2 cDKO). We demonstrate that RIM2α and RIM2β are present at IHC AZs of hearing mice, positively regulate the number of synaptic CaV1.3 Ca2+ channels, and are required for normal hearing.  相似文献   

18.
Being activated by depolarizing voltages and increases in cytoplasmic Ca2+, voltage- and calcium-activated potassium (BK) channels and their modulatory β-subunits are able to dampen or stop excitatory stimuli in a wide range of cellular types, including both neuronal and nonneuronal tissues. Minimal alterations in BK channel function may contribute to the pathophysiology of several diseases, including hypertension, asthma, cancer, epilepsy, and diabetes. Several gating processes, allosterically coupled to each other, control BK channel activity and are potential targets for regulation by auxiliary β-subunits that are expressed together with the α (BK)-subunit in almost every tissue type where they are found. By measuring gating currents in BK channels coexpressed with chimeras between β1 and β3 or β2 auxiliary subunits, we were able to identify that the cytoplasmic regions of β1 are responsible for the modulation of the voltage sensors. In addition, we narrowed down the structural determinants to the N terminus of β1, which contains two lysine residues (i.e., K3 and K4), which upon substitution virtually abolished the effects of β1 on charge movement. The mechanism by which K3 and K4 stabilize the voltage sensor is not electrostatic but specific, and the α (BK)-residues involved remain to be identified. This is the first report, to our knowledge, where the regulatory effects of the β1-subunit have been clearly assigned to a particular segment, with two pivotal amino acids being responsible for this modulation.High-conductance voltage- and calcium-activated potassium (BK) channels are homotetrameric proteins of α-subunits encoded by the slo1 gene (1). These channels are expressed in virtually all mammalian tissues, where they detect and integrate membrane voltage and calcium concentration changes dampening the responsiveness of cells when confronted with excitatory stimuli. They are abundant in the CNS and nonneuronal tissues, such as smooth muscle or hair cells. This wide distribution is associated with an outstandingly large functional diversity, in which BK channel activity appears optimally adapted to the particular physiological demands of each cell type (2). On the other hand, small alterations in BK channel function may contribute to the pathophysiology of hypertension, asthma, cancer, epilepsy, diabetes, and other conditions in humans (38). Alternative splicing, posttranslational modifications, and regulation by auxiliary proteins have been proposed to contribute to this functional diversity (1, 2, 916).The BK channel α-subunit is formed by a single polypeptide of about 1,200 amino acids that contains all of the key structural elements for ion permeation, gating, and modulation by ions and other proteins. Tetramers of α-subunits form functional BK channels. Each subunit has seven hydrophobic transmembrane segments (S0–S6), where the voltage-sensor domain (VSD) and pore domain (PD) reside (2). The N terminus faces the extracellular side of the membrane, whereas the C terminus is intracellular. The latter contains four hydrophobic α-helices (S7–S10) and the main Ca2+ binding sites (2). VSDs formed by segments S1–S4 harbor a series of charged residues across the membrane that contributes to voltage sensing (2). Upon membrane depolarization, each VSD undergoes a rearrangement (17) that prompts the opening of a highly K+-selective pore formed by the four PDs that come together at the symmetry center of the tetramer.Although BK channel expression is ubiquitous, in most physiological scenarios their functioning is provided by their coassembly with auxiliary proteins, such as β-subunits. This coassembly brings channel activity into the proper cell/tissue context (11, 13). Four different β-subunits have been cloned (β1–β4) (1824), all of which have been observed to modify BK channel function. Albeit to a different extent, all β-subunits modify the Ca2+ sensitivity, voltage dependence, and gating properties of BK channels, hence modifying plasma membrane excitability balance. Regarding auxiliary β-subunits, β1- and β2-subunits increase apparent Ca2+ sensitivity and decelerate macroscopic current kinetics (14, 20, 21, 2530); β2 and β3 induce fast inactivation as well as an instantaneous outward rectification (20, 21, 24, 31, 32); and β4 slows down activation and deactivation kinetics (12, 23) and modifies Ca2+ sensitivity (12, 33, 34).It should be kept in mind that β-subunits are potential targets for different molecules that modulate channel function, such as alcohol (35), estrogens (15), hormones (36), and fatty acids (37, 38). Additionally, scorpion toxin affinity in BK channels would tend to increase when β1 is coexpressed with the α-subunit (22).To identify the molecular elements that give β1 the ability to modulate the voltage sensor of BK channels, we constructed chimeric proteins of β1/β2- and β1/β3-subunits by swapping their N and C termini, the transmembrane (TM) segments, and the extracellular loops and recorded their gating currents. Two lysine residues that are unique to the N terminus of β1 were identified to be sufficient for BK voltage-sensor modulation.  相似文献   

19.
A series of mono- and dinuclear alkynylplatinum(II) terpyridine complexes containing the hydrophilic oligo(para-phenylene ethynylene) with two 3,6,9-trioxadec-1-yloxy chains was designed and synthesized. The mononuclear alkynylplatinum(II) terpyridine complex was found to display a very strong tendency toward the formation of supramolecular structures. Interestingly, additional end-capping with another platinum(II) terpyridine moiety of various steric bulk at the terminal alkyne would lead to the formation of nanotubes or helical ribbons. These desirable nanostructures were found to be governed by the steric bulk on the platinum(II) terpyridine moieties, which modulates the directional metal−metal interactions and controls the formation of nanotubes or helical ribbons. Detailed analysis of temperature-dependent UV-visible absorption spectra of the nanostructured tubular aggregates also provided insights into the assembly mechanism and showed the role of metal−metal interactions in the cooperative supramolecular polymerization of the amphiphilic platinum(II) complexes.Square-planar d8 platinum(II) polypyridine complexes have long been known to exhibit intriguing spectroscopic and luminescence properties (154) as well as interesting solid-state polymorphism associated with metal−metal and π−π stacking interactions (114, 25). Earlier work by our group showed the first example, to our knowledge, of an alkynylplatinum(II) terpyridine system [Pt(tpy)(C ≡ CR)]+ that incorporates σ-donating and solubilizing alkynyl ligands together with the formation of Pt···Pt interactions to exhibit notable color changes and luminescence enhancements on solvent composition change (25) and polyelectrolyte addition (26). This approach has provided access to the alkynylplatinum(II) terpyridine and other related cyclometalated platinum(II) complexes, with functionalities that can self-assemble into metallogels (2731), liquid crystals (32, 33), and other different molecular architectures, such as hairpin conformation (34), helices (3538), nanostructures (3945), and molecular tweezers (46, 47), as well as having a wide range of applications in molecular recognition (4852), biomolecular labeling (4852), and materials science (53, 54). Recently, metal-containing amphiphiles have also emerged as a building block for supramolecular architectures (4244, 5559). Their self-assembly has always been found to yield different molecular architectures with unprecedented complexity through the multiple noncovalent interactions on the introduction of external stimuli (4244, 5559).Helical architecture is one of the most exciting self-assembled morphologies because of the uniqueness for the functional and topological properties (6069). Helical ribbons composed of amphiphiles, such as diacetylenic lipids, glutamates, and peptide-based amphiphiles, are often precursors for the growth of tubular structures on an increase in the width or the merging of the edges of ribbons (64, 65). Recently, the optimization of nanotube formation vs. helical nanostructures has aroused considerable interests and can be achieved through a fine interplay of the influence on the amphiphilic property of molecules (66), choice of counteranions (67, 68), or pH values of the media (69), which would govern the self-assembly of molecules into desirable aggregates of helical ribbons or nanotube scaffolds. However, a precise control of supramolecular morphology between helical ribbons and nanotubes remains challenging, particularly for the polycyclic aromatics in the field of molecular assembly (6469). Oligo(para-phenylene ethynylene)s (OPEs) with solely π−π stacking interactions are well-recognized to self-assemble into supramolecular system of various nanostructures but rarely result in the formation of tubular scaffolds (7073). In view of the rich photophysical properties of square-planar d8 platinum(II) systems and their propensity toward formation of directional Pt···Pt interactions in distinctive morphologies (2731, 3945), it is anticipated that such directional and noncovalent metal−metal interactions might be capable of directing or dictating molecular ordering and alignment to give desirable nanostructures of helical ribbons or nanotubes in a precise and controllable manner.Herein, we report the design and synthesis of mono- and dinuclear alkynylplatinum(II) terpyridine complexes containing hydrophilic OPEs with two 3,6,9-trioxadec-1-yloxy chains. The mononuclear alkynylplatinum(II) terpyridine complex with amphiphilic property is found to show a strong tendency toward the formation of supramolecular structures on diffusion of diethyl ether in dichloromethane or dimethyl sulfoxide (DMSO) solution. Interestingly, additional end-capping with another platinum(II) terpyridine moiety of various steric bulk at the terminal alkyne would result in nanotubes or helical ribbons in the self-assembly process. To the best of our knowledge, this finding represents the first example of the utilization of the steric bulk of the moieties, which modulates the formation of directional metal−metal interactions to precisely control the formation of nanotubes or helical ribbons in the self-assembly process. Application of the nucleation–elongation model into this assembly process by UV-visible (UV-vis) absorption spectroscopic studies has elucidated the nature of the molecular self-assembly, and more importantly, it has revealed the role of metal−metal interactions in the formation of these two types of nanostructures.  相似文献   

20.
Mechanosensitive ion channels underlie neuronal responses to physical forces in the sensation of touch, hearing, and other mechanical stimuli. The fundamental basis of force transduction in eukaryotic mechanosensitive ion channels is unknown. Are mechanical forces transmitted directly from membrane to channel as in prokaryotic mechanosensors or are they mediated through macromolecular tethers attached to the channel? Here we show in cells that the K+ channel TRAAK (K2P4.1) is responsive to mechanical forces similar to the ion channel Piezo1 and that mechanical activation of TRAAK can electrically counter Piezo1 activation. We then show that the biophysical origins of force transduction in TRAAK and TREK1 (K2P2.1) two-pore domain K+ (K2P) channels come from the lipid membrane, not from attached tethers. These findings extend the “force-from-lipid” principle established for prokaryotic mechanosensitive channels MscL and MscS to these eukaryotic mechanosensitive K+ channels.Mechanosensation encompasses the host of processes that cells have evolved to sense and respond to mechanical forces ubiquitous in biology. Mechanosensation underlies our sense of touch, hearing, and balance as well as our ability to regulate blood and osmotic pressures. Despite their broad importance, the molecules involved in mechanosensation have been largely difficult to identify and characterize. Mechanosensitive ion channels are cells’ fastest mechanosensors and translate mechanical forces into cellular electrical signals to produce rapid neuronal responses to mechanical stimuli. Whereas the list of eukaryotic ion channels implicated in mechanosensation continues to grow (17), a fundamental question remains: How do these channels sense force?Mechanical force gating of ion channels can in principle occur either directly through the lipid bilayer or through accessory tether-forming proteins (8). Lipid bilayer-mediated gating can occur if a force induces tension in the cellular membrane, which can provide a tension-dependent energy difference between closed and open conformations. Tether-mediated gating can occur if a mechanical stimulus is transmitted along accessory proteins or other macromolecular structures (cytoskeletal or extracellular matrix) that are attached to the channel (8, 9). Only the bacterial mechanosensitive channels MscL and MscS have been demonstrated rigorously to undergo lipid bilayer-mediated gating via membrane tension (10). Using a reconstituted system of purified channel protein in defined lipids, these channels were mechanically activated by membrane tension induced with pressure applied to the patch pipette (1114). Difficulties in high level expression, purification, and reconstitution have precluded such an analysis of mechanosensitivity in eukaryotic ion channels with the same rigor as applied to MscL and MscS (15, 16). Whereas mechanosensitivity of eukaryotic channels has been demonstrated by poking cell membranes under whole-cell voltage clamp and by pressure activation of channels in patches excised from cells or membrane blebs from cells (17), these experiments have not distinguished a direct membrane-mediated mechanism from other mechanisms that would rely on additional macromolecular components inescapably present in the cell-based assays.TRAAK (K2P4.1) and TREK channels are members of the two-pore domain K+ (K2P) ion channels. Their gating is regulated by mechanical perturbation of the cell membrane as well as polyunsaturated fatty acids, other lipids, and temperature (1823). TRAAK/TREK knockout mice exhibit mechanical and temperature allodynia (24). On this basis, these channels have been proposed to regulate the noxious input threshold for pressure and temperature sensitivity in mouse dorsal root ganglia (24). TRAAK is the only eukaryotic mechanosensitive ion channel for which crystal structures have been determined (25, 26). The biophysical mechanisms underlying TRAAK and TREK mechanosensitivity, however, are unknown. Here we address whether or not mechanical forces are transmitted to TRAAK and TREK channels directly from the membrane.  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