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.
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   

2.
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   

3.
The physiological functions of hydrogen sulfide (H2S) include vasorelaxation, stimulation of cellular bioenergetics, and promotion of angiogenesis. Analysis of human colon cancer biopsies and patient-matched normal margin mucosa revealed the selective up-regulation of the H2S-producing enzyme cystathionine-β-synthase (CBS) in colon cancer, resulting in an increased rate of H2S production. Similarly, colon cancer-derived epithelial cell lines (HCT116, HT-29, LoVo) exhibited selective CBS up-regulation and increased H2S production, compared with the nonmalignant colonic mucosa cells, NCM356. CBS localized to the cytosol, as well as the mitochondrial outer membrane. ShRNA-mediated silencing of CBS or its pharmacological inhibition with aminooxyacetic acid reduced HCT116 cell proliferation, migration, and invasion; reduced endothelial cell migration in tumor/endothelial cell cocultures; and suppressed mitochondrial function (oxygen consumption, ATP turnover, and respiratory reserve capacity), as well as glycolysis. Treatment of nude mice with aminooxyacetic acid attenuated the growth of patient-derived colon cancer xenografts and reduced tumor blood flow. Similarly, CBS silencing of the tumor cells decreased xenograft growth and suppressed neovessel density, suggesting a role for endogenous H2S in tumor angiogenesis. In contrast to CBS, silencing of cystathionine-γ-lyase (the expression of which was unchanged in colon cancer) did not affect tumor growth or bioenergetics. In conclusion, H2S produced from CBS serves to (i) maintain colon cancer cellular bioenergetics, thereby supporting tumor growth and proliferation, and (ii) promote angiogenesis and vasorelaxation, consequently providing the tumor with blood and nutritients. The current findings identify CBS-derived H2S as a tumor growth factor and anticancer drug target.The endogenous gasotransmitter hydrogen sulfide (H2S) is a stimulator of vasorelaxation (13), angiogenesis (35), and cellular bioenergetics (6, 7). H2S is generated from l-cysteine by two pyridoxal-5′-phospate–dependent enzymes, cystathionine-β-synthase (CBS) and cystathionine-γ-lyase (CSE), and by the combined action of cysteine aminotransferase and 3-mercaptopyruvate sulfurtransferase (3-MST) (810). H2S exerts its cellular actions via multiple mechanisms (115), including activation of potassium channels (13), stimulation of kinase pathways (4, 11, 12), and inhibition of phosphodiesterases (3, 15).Both ATP generation and angiogenesis are vital factors for the growth and proliferation of tumors (1619). Using human colon cancer tissues and cancer-derived cell lines, we have now conducted a series of in vitro and in vivo studies to explore whether endogenous, tumor cell-derived H2S plays a role as a tumor-derived survival factor. The results show that CBS is selectively overexpressed in colon cancer, and that H2S produced by it serves to maintain the tumor''s cellular bioenergetics and to promote tumor angiogenesis.  相似文献   

4.
A series of discrete decanuclear gold(I) μ3-sulfido complexes with alkyl chains of various lengths on the aminodiphosphine ligands, [Au10{Ph2PN(CnH2n+1)PPh2}43-S)4](ClO4)2, has been synthesized and characterized. These complexes have been shown to form supramolecular nanoaggregate assemblies upon solvent modulation. The photoluminescence (PL) colors of the nanoaggregates can be switched from green to yellow to red by varying the solvent systems from which they are formed. The PL color variation was investigated and correlated with the nanostructured morphological transformation from the spherical shape to the cube as observed by transmission electron microscopy and scanning electron microscopy. Such variations in PL colors have not been observed in their analogous complexes with short alkyl chains, suggesting that the long alkyl chains would play a key role in governing the supramolecular nanoaggregate assembly and the emission properties of the decanuclear gold(I) sulfido complexes. The long hydrophobic alkyl chains are believed to induce the formation of supramolecular nanoaggregate assemblies with different morphologies and packing densities under different solvent systems, leading to a change in the extent of Au(I)–Au(I) interactions, rigidity, and emission properties.Gold(I) complexes are one of the fascinating classes of complexes that reveal photophysical properties that are highly sensitive to the nuclearity of the metal centers and the metal–metal distances (159). In a certain sense, they bear an analogy or resemblance to the interesting classes of metal nanoparticles (NPs) (6069) and quantum dots (QDs) (7076) in that the properties of the nanostructured materials also show a strong dependence on their sizes and shapes. Interestingly, while the optical and spectroscopic properties of metal NPs and QDs show a strong dependence on the interparticle distances, those of polynuclear gold(I) complexes are known to mainly depend on the nuclearity and the internuclear separations of gold(I) centers within the individual molecular complexes or clusters, with influence of the intermolecular interactions between discrete polynuclear molecular complexes relatively less explored (3438), and those of polynuclear gold(I) clusters not reported. Moreover, while studies on polynuclear gold(I) complexes or clusters are known (3454), less is explored of their hierarchical assembly and nanostructures as well as the influence of intercluster aggregation on the optical properties (3438). Among the gold(I) complexes, polynuclear gold(I) chalcogenido complexes represent an important and interesting class (4451). While directed supramolecular assembly of discrete Au12 (52), Au16 (53), Au18 (51), and Au36 (54) metallomacrocycles as well as trinuclear gold(I) columnar stacks (3438) have been reported, there have been no corresponding studies on the supramolecular hierarchical assembly of polynuclear gold(I) chalcogenido clusters.Based on our interests and experience in the study of gold(I) chalcogenido clusters (4446, 51), it is believed that nanoaggegrates with interesting luminescence properties and morphology could be prepared by the judicious design of the gold(I) chalcogenido clusters. As demonstrated by our previous studies on the aggregation behavior of square-planar platinum(II) complexes (7780) where an enhancement of the solubility of the metal complexes via introduction of solubilizing groups on the ligands and the fine control between solvophobicity and solvophilicity of the complexes would have a crucial influence on the factors governing supramolecular assembly and the formation of aggregates (80), introduction of long alkyl chains as solubilizing groups in the gold(I) sulfido clusters may serve as an effective way to enhance the solubility of the gold(I) clusters for the construction of supramolecular assemblies of novel luminescent nanoaggegrates.Herein, we report the preparation and tunable spectroscopic properties of a series of decanuclear gold(I) μ3-sulfido complexes with alkyl chains of different lengths on the aminophosphine ligands, [Au10{Ph2PN(CnH2n+1)PPh2}43-S)4](ClO4)2 [n = 8 (1), 12 (2), 14 (3), 18 (4)] and their supramolecular assembly to form nanoaggregates. The emission colors of the nanoaggregates of 2−4 can be switched from green to yellow to red by varying the solvent systems from which they are formed. These results have been compared with their short alkyl chain-containing counterparts, 1 and a related [Au10{Ph2PN(C3H7)PPh2}43-S)4](ClO4)2 (45). The present work demonstrates that polynuclear gold(I) chalcogenides, with the introduction of appropriate functional groups, can serve as building blocks for the construction of novel hierarchical nanostructured materials with environment-responsive properties, and it represents a rare example in which nanoaggregates have been assembled with the use of discrete molecular metal clusters as building blocks.  相似文献   

5.
Protein misfolding and aggregation cause serious degenerative conditions such as Alzheimer’s, Parkinson, and prion diseases. Damage to membranes is thought to be one of the mechanisms underlying cellular toxicity of a range of amyloid assemblies. Previous studies have indicated that amyloid fibrils can cause membrane leakage and elicit cellular damage, and these effects are enhanced by fragmentation of the fibrils. Here we report direct 3D visualization of membrane damage by specific interactions of a lipid bilayer with amyloid-like fibrils formed in vitro from β2-microglobulin (β2m). Using cryoelectron tomography, we demonstrate that fragmented β2m amyloid fibrils interact strongly with liposomes and cause distortions to the membranes. The normally spherical liposomes form pointed teardrop-like shapes with the fibril ends seen in proximity to the pointed regions on the membranes. Moreover, the tomograms indicated that the fibrils extract lipid from the membranes at these points of distortion by removal or blebbing of the outer membrane leaflet. Tiny (15–25 nm) vesicles, presumably formed from the extracted lipids, were observed to be decorating the fibrils. The findings highlight a potential role of fibrils, and particularly fibril ends, in amyloid pathology, and report a previously undescribed class of lipid–protein interactions in membrane remodelling.The failure of molecular chaperones to prevent the accumulation of misfolded proteins results in protein aggregation and amyloid formation, processes associated with severe human degenerative diseases (1, 2). Despite the attention focused on these problems during the century since these disorders were first identified (35) and advances in understanding the structure of the cross-β conformation of amyloid fibrils in atomic detail (6, 7), the basic pathological mechanisms of amyloidosis remain poorly understood and therapeutic intervention is lacking. The identity of the toxic species and the mechanisms of cytotoxicity remain major unsolved problems. In some systems, there is evidence suggesting that prefibrillar oligomers, rather than the fully formed fibrils, are the source of toxicity (8, 9). In these cases, cytotoxicity is thought to result from the formation of specific membrane pores (10, 11) although alternative models including membrane destabilization or membrane thinning have also been proposed (1215). In other cases, toxicity may reside with the amyloid fibrils themselves. Evidence that toxicity correlates with fibrillar assemblies has been reported for yeast and mammalian prion proteins (16, 17), human lysozyme (18), Huntingtin exon 1, α-synuclein (19), and Amyloid-β (Aβ) (20, 21). Furthermore, Aβ plaques have been shown to form rapidly in vivo and to precede neuropathological changes in a mouse model (22). The end surfaces of fibrils (herein termed “fibril ends”) are unusually reactive entities: they play a key role in catalyzing recruitment and conformational conversion of amyloid-forming proteins (23, 24) and provide the sites for templated elongation of amyloid fibril growth (25, 26). Recently, Xue et al. (27) showed that short fibrils of β2-microglobulin (β2m), α-synuclein, and hen lysozyme, each prepared by fragmentation of longer fibrils, cause increased damage to membranes and disruption to cellular function compared with the initial long fibrils. Short and long fibril preparations differ in the number of fibril ends at a given protein concentration. Because these are known to be reactive sites, the above observations suggest a role for fibril ends in amyloid–lipid interaction and possibly in amyloid pathogenesis (23, 24, 27). Fragmented fibrils of all three proteins were also found to induce dye leakage from negatively charged liposomes, the most susceptible of which contain a mixture of the cellular lipids phosphatidylcholine (PC) and phosphatidylglycerol (PG), but liposomes with a variety of compositions were damaged by the fibrils in all cases (27).Here, we use β2m amyloid fibrils formed in vitro as a model system to investigate the structural basis of membrane damage by amyloid fibrils (27, 28). Previous studies have shown that these fibrils possess a parallel in register cross-β structure (29, 30) assembled into multidomain filaments coiled together, described by cryo-EM (28). These fibrils bind amyloid-specific ligands such as serum amyloid P component with a similar affinity to their ex vivo counterparts (31). Using the conditions under which β2m amyloid fibrils induce dye release from liposomes (pH 7.4), we examined the effects of both long (1,400 nm) and fragmented (400 nm) β2m fibrils (27), as well as various control preparations, on the 3D structures of the liposomes by confocal microscopy, cryo-EM, and tomography. We found pronounced distortions in the liposomes, interruptions to the bilayer structure, and extraction of lipids that were induced by the presence of amyloid fibrils. The most severe distortions were seen in proximity to the fibril ends, which are enriched in the fragmented fibril samples. This type of membrane remodelling appears distinct from the actions of other previously described proteins that induce membrane breakage, as in the action of membrane pore-forming proteins (32). The results suggest a role of fibrils in membrane damage that could contribute to the cellular dysfunction associated with amyloid disease.  相似文献   

6.
7.
8.
Kinesin-1 is a dimeric motor protein, central to intracellular transport, that steps hand-over-hand toward the microtubule (MT) plus-end, hydrolyzing one ATP molecule per step. Its remarkable processivity is critical for ferrying cargo within the cell: over 100 successive steps are taken, on average, before dissociation from the MT. Despite considerable work, it is not understood which features coordinate, or “gate,” the mechanochemical cycles of the two motor heads. Here, we show that kinesin dissociation occurs subsequent to, or concomitant with, phosphate (Pi) release following ATP hydrolysis. In optical trapping experiments, we found that increasing the steady-state population of the posthydrolysis ADP·Pi state (by adding free Pi) nearly doubled the kinesin run length, whereas reducing either the ATP binding rate or hydrolysis rate had no effect. The data suggest that, during processive movement, tethered-head binding occurs subsequent to hydrolysis, rather than immediately after ATP binding, as commonly suggested. The structural change driving motility, thought to be neck linker docking, is therefore completed only upon hydrolysis, and not ATP binding. Our results offer additional insights into gating mechanisms and suggest revisions to prevailing models of the kinesin reaction cycle.Since its discovery nearly 30 years ago (1), kinesin-1—the founding member of the kinesin protein superfamily—has emerged as an important model system for studying biological motors (2, 3). During “hand-over-hand” stepping, kinesin dimers alternate between a two–heads-bound (2-HB) state, with both heads attached to the microtubule (MT), and a one–head-bound (1-HB) state, where a single head, termed the tethered head, remains free of the MT (4, 5). The catalytic cycles of the two heads are maintained out of phase by a series of gating mechanisms, thereby enabling the dimer to complete, on average, over 100 steps before dissociating from the MT (68). A key structural element for this coordination is the neck linker (NL), a ∼14-aa segment that connects each catalytic head to a common stalk (9). In the 1-HB state, nucleotide binding is thought to induce a structural reconfiguration of the NL, immobilizing it against the MT-bound catalytic domain (2, 3, 1017). This transition, called “NL docking,” is believed to promote unidirectional motility by biasing the position of the tethered head toward the next MT binding site (2, 3, 1017). The completion of an 8.2-nm step (18) entails the binding of this tethered head to the MT, ATP hydrolysis, and detachment of the trailing head, thereby returning the motor to the ATP-waiting state (2, 3, 1017). Prevailing models of the kinesin mechanochemical cycle (2, 3, 10, 14, 15, 17), which invoke NL docking upon ATP binding, explain the highly directional nature of kinesin motility and offer a compelling outline of the sequence of events following ATP binding. Nevertheless, these abstractions do not speak directly to the branching transitions that determine whether kinesin dissociates from the MT (off-pathway) or continues its processive reaction cycle (on-pathway). The distance moved by an individual motor before dissociating—the run length—is limited by unbinding from the MT. The propensity for a dimer to unbind involves a competition among multiple, force-dependent transitions in the two heads, which are not readily characterized by traditional structural or bulk biochemical approaches. Here, we implemented high-resolution single-molecule optical trapping techniques to determine transitions in the kinesin cycle that govern processivity.  相似文献   

9.
Mechanosensitive ion channels are sensors probing membrane tension in all species; despite their importance and vital role in many cell functions, their gating mechanism remains to be elucidated. Here, we determined the conditions for releasing intact mechanosensitive channel of large conductance (MscL) proteins from their detergents in the gas phase using native ion mobility–mass spectrometry (IM-MS). By using IM-MS, we could detect the native mass of MscL from Escherichia coli, determine various global structural changes during its gating by measuring the rotationally averaged collision cross-sections, and show that it can function in the absence of a lipid bilayer. We could detect global conformational changes during MscL gating as small as 3%. Our findings will allow studying native structure of many other membrane proteins.One of the best candidates to explore the gating of mechanosensitive channels is the mechanosensitive channel of large conductance (MscL) from Escherichia coli. The crystal structure of MscL in its closed/nearly closed state from Mycobacterium tuberculosis revealed this channel as a homopentamer (1). Each subunit has a cytoplasmic N- and C-terminal domain as well as two α-helical transmembrane (TM) domains, TM1 and TM2, which are connected by a periplasmic loop. The five TM1 helices form the pore and the more peripheral TM2 helices interact with the lipid bilayer.MscL detects changes in membrane tension invoked by a hypoosmotic shock and couples the tension sensing directly to large conformational changes (1, 2). On the basis of a large body of structural and theoretical data, numerous gating models of MscL have been proposed (39). These models agree upon (i) the hydrophobic pore constriction of the channel and (ii) the channel opens by an iris-like rotation—i.e., a tilting and outward movement of transmembrane helices that make the channel wider and shorter (5). This mechanism is supported by patch-clamp (10), disulfide cross-linking (11), FRET spectroscopy (12), and site-directed spin labeling EPR experiments (6, 7), as well as computational studies (1315). So far, direct experimental results have only been observed for short-range local structural changes, and no measure of the overall global structural changes during channel gating have been reported. Because there is no crystal structure available for the open MscL channel, elucidating overall global structural changes from the onset of channel activation is of utmost importance for our understanding of the gating mechanism of mechanosensitive channels. Here, we provide direct experimental evidence for the key areal changes occurring during channel gating by combining our ability to activate MscL in a controlled manner to different subopen states (16) with a native ion mobility-mass spectrometry (IM-MS) approach.  相似文献   

10.
A dynamic transmembrane voltage field has been suggested as an intrinsic element in voltage sensor (VS) domains. Here, the dynamic field contribution to the VS energetics was analyzed via electrostatic calculations applied to a number of atomistic structures made available recently. We find that the field is largely static along with the molecular motions of the domain, and more importantly, it is minimally modified across VS variants. This finding implies that sensor domains transfer approximately the same amount of gating charges when moving the electrically charged S4 helix between fixed microscopic configurations. Remarkably, the result means that the observed operational diversity of the domain, including the extension, rate, and voltage dependence of the S4 motion, as dictated by the free energy landscape theory, must be rationalized in terms of dominant variations of its chemical free energy.Voltage sensor (VS) domains are electrically charged membrane proteins made of four packed helices (1). The fourth segment (S4) contains four highly conserved positively charged amino acids, R1 through R4. By interchanging its conformation between two main states (resting and activated) in response to voltage variations, VS domains displace the S4 charges across the membrane capacitance, giving rise to ΔQ, the so-called gating charge (2). As a result of their function of converting voltage variations into molecular motions, VS domains are ubiquitous in a number of electrically mediated processes, either as domain components of phosphatases (2) or proton (3) and ion channels (48).Despite the conservation of S4 sequences, nature has designed a variety of constructs (2) that present a wide range of voltage dependence and absolute rates of activation. For instance, the VS kinetics is markedly distinct between voltage-gated Na+ and K+ channels (9), a feature that complies with their respective role in the fast and slow phases of the action potential. Drastic kinetic shifts can even be observed in VS differing by point mutations (10, 11). In all of these constructs, the S4 operation results from the fine balance between the chemical and the electrical components of the relative free energy of the segment. Whereas the former depends on the S4 energy in the absence of an electrical driving force, the latter arises essentially from ?(r) (12), a dimensionless scalar field that reports the fraction of the membrane voltage coupled to every S4 charge qi. As mostly embodied in the transporter model (13), the reshaping of ?(r) along with S4 displacements appears as one potential mechanism impacting the sensing process. Facing the VS diversity, it has been unknown to which extent the field reshaping may impact the S4 operation in distinct constructs and account for its energetic differences.Here, by benefiting from an increasing number of atomistic structures of VS-containing channels or enzymes made available recently, we use all-atom molecular dynamics (MD) simulations (1416) in combination with electrostatic calculations (17, 18) to investigate voltage-coupling properties of sensor domains. Anticipating our results, we find that, although primarily conformation independent, ?(r) is minimally modified over the VS variants. This finding points to marginal dynamic contributions of the membrane voltage field to the energetics and diversity of VS proteins.  相似文献   

11.
Broadly neutralizing HIV antibodies (bNAbs) can recognize carbohydrate-dependent epitopes on gp120. In contrast to previously characterized glycan-dependent bNAbs that recognize high-mannose N-glycans, PGT121 binds complex-type N-glycans in glycan microarrays. We isolated the B-cell clone encoding PGT121, which segregates into PGT121-like and 10-1074–like groups distinguished by sequence, binding affinity, carbohydrate recognition, and neutralizing activity. Group 10-1074 exhibits remarkable potency and breadth but no detectable binding to protein-free glycans. Crystal structures of unliganded PGT121, 10-1074, and their likely germ-line precursor reveal that differential carbohydrate recognition maps to a cleft between complementarity determining region (CDR)H2 and CDRH3. This cleft was occupied by a complex-type N-glycan in a “liganded” PGT121 structure. Swapping glycan contact residues between PGT121 and 10-1074 confirmed their importance for neutralization. Although PGT121 binds complex-type N-glycans, PGT121 recognized high-mannose-only HIV envelopes in isolation and on virions. As HIV envelopes exhibit varying proportions of high-mannose- and complex-type N-glycans, these results suggest promiscuous carbohydrate interactions, an advantageous adaptation ensuring neutralization of all viruses within a given strain.Antibodies are essential for the success of most vaccines (1), and antibodies against HIV appear to be the only correlate of protection in the recent RV144 anti-HIV vaccine trial (2). Some HIV-1–infected patients develop broadly neutralizing serologic activity against the gp160 viral spike 2–4 y after infection (310), but these antibodies do not generally protect infected humans because autologous viruses escape through mutation (1113). Nevertheless, broadly neutralizing activity puts selective pressure on the virus (13) and passive transfer of broadly neutralizing antibodies (bNAbs) to macaques protects against simian/human immunodeficiency virus (SHIV) infection (1424). It has therefore been proposed that vaccines that elicit such antibodies may be protective against HIV infection in humans (10, 2528).The development of single-cell antibody cloning techniques revealed that bNAbs target several different epitopes on the HIV-1 gp160 spike (2935). The most potent HIV-1 bNAbs recognize the CD4 binding site (CD4bs) (31, 34, 36) and carbohydrate-dependent epitopes associated with the variable loops (32, 33, 37, 38), including the V1/V2 (antibodies PG9/PG16) (33) and V3 loops (PGTs) (32). Less is known about carbohydrate-dependent epitopes because the antibodies studied to date are either unique examples or members of small clonal families.To better understand the neutralizing antibody response to HIV-1 and the epitope targeted by PGT antibodies, we isolated members of a large clonal family dominating the gp160-specific IgG memory response from the clade A-infected patient who produced PGT121. We report that PGT121 antibodies segregate into two groups, a PGT121-like and a 10-1074–like group, according to sequence, binding affinity, neutralizing activity, and recognition of carbohydrates and the V3 loop. The 10-1074 antibody and related family members exhibit unusual potent neutralization, including broad reactivity against newly transmitted viruses. Unlike previously characterized carbohydrate-dependent bNAbs, PGT121 binds to complex-type, rather than high-mannose, N-glycans in glycan microarray experiments. Crystal structures of PGT121 and 10-1074 compared with structures of their germ-line precursor and a structure of PGT121 bound to a complex-type N-glycan rationalize their distinct properties.  相似文献   

12.
Chronic infection with hepatitis C virus (HCV) is an important cause of end stage liver disease worldwide. In the United States, most HCV-related disease is associated with genotype 1 infection, which remains difficult to treat. Drug and vaccine development was hampered by inability to culture patient isolates representing HCV genotypes 1–7 and subtypes; only a recombinant 2a genome (strain JFH1) spontaneously replicated in vitro. Recently, we identified three mutations F1464L/A1672S/D2979G (LSG) in the nonstructural (NS) proteins, essential for development of full-length HCV 2a (J6) and 2b (J8) culture systems in Huh7.5 cells. Here, we developed a highly efficient genotype 1a (strain TN) full-length culture system. We initially found that the LSG substitutions conferred viability to an intergenotypic recombinant composed of TN 5′ untranslated region (5′UTR)-NS5A and JFH1 NS5B-3′UTR; recovered viruses acquired two adaptive mutations located in NS3 and NS4B. Introduction of these changes into a replication-deficient TN full-length genome, harboring LSG, permitted efficient HCV production. Additional identified NS4B and NS5B mutations fully adapted the TN full-length virus. Thus, a TN genome with 8 changes (designated TN cell-culture derived, TNcc) replicated efficiently and released infectious particles of ∼5 log10 focus-forming units per mL; passaged TNcc did not require additional changes. IFN-α and directly acting antivirals targeting the HCV protease, NS5A, and NS5B, each inhibited full-length TN infection dose-dependently. Given the unique importance of genotype 1 for pathogenesis, this infectious 1a culture system represents an important advance in HCV research. The approach used and the mutations identified might permit culture development for other HCV isolates, thus facilitating vaccine development and personalized treatment.Hepatitis C virus (HCV) chronically infects an estimated 130–170 million people worldwide. The infection increases the risk of developing liver cirrhosis and liver cancer and results in more than 350,000 deaths annually. No HCV vaccine is available. Current standard treatment is based on IFN-α/ribavirin, which, however, has low efficacy against the most prevalent HCV variants (1). Incorporation of directly acting antivirals (DAAs) in treatment regimens improves sustained viral response rate, but a favorable outcome is challenged by fast emergence of drug resistance and differential responses of the different HCV genotypes (2). Thus, HCV infection continues to be a huge health and economic burden to the world population, and improved in vitro experimental systems would be important to permit additional studies of new antivirals and associated resistance patterns.HCV is a small enveloped virus belonging to the genus Hepacivirus in the family Flaviviridae. The HCV genome is a positive-sense single-strand RNA (∼9.6 kb), consisting of a single ORF flanked by 5′ and 3′ untranslated regions (UTRs). The ORF encodes virus structural proteins (core, E1, and E2), p7, and six nonstructural (NS) proteins NS2, NS3, NS4A, NS4B, NS5A, and NS5B (1). HCV isolates are classified into seven major variants (genotypes 1–7) and numerous subtypes (a, b, etc.) differing by ∼30% and ∼20%, respectively, in nucleotide and amino acid sequences (3). Genotype 1 is the most prevalent in the world. In the United States, Japan, China, and southern Europe over 70% of HCV patients are infected with genotype 1 (47). In northern Europe, genotype 1 accounts for ∼50% of HCV infections (8). Genotype 1 infection is more resistant to IFN-based treatment than infection with other genotypes (1).In vitro HCV culture systems are needed to study treatment regimens and associated viral resistance. However, drug and vaccine development have been greatly hampered by inability to culture patient isolates in vitro. Although a number of HCV full-length clones were shown to be infectious in chimpanzees (ref. 9 and cited references therein), only JFH1 (genotype 2a) spontaneously replicated in human hepatoma cells (Huh7 and its derivatives) and released infectious virus particles (10, 11). Efficient growth of JFH1 required culture adaptive mutations (12). Recently, we reported efficient J6cc (2a) and J8cc (2b) full-length culture systems (13); subsequently Date et al. reported on a full-length system for 2a strain JFH2 (14). A single full-length genotype 1a genome, H77-S, carrying mutations identified in the subgenomic replicon of the same strain, has been reported to release relatively small amounts of virus particles (15). The Con1 (1b) full-length culture system was reported, but a very low level of replication has limited its utility (16). Thus, efficient HCV full-length culture systems remained limited to genotype 2 isolates. The clinical importance of the marked genetic differences between HCV genotype isolates poses a critical need for development of robust full-length culture systems for HCV genotype 1 isolates.The unique replication capacity of JFH1 has permitted the development of JFH1-based HCV recombinants (17); we and others have reported different inter- and intragenotypic recombinants including core-NS2 (1724), 5′UTR-NS2 (25), NS3 protease/NS4A (26, 27), NS5A (28), and core-NS3 protease plus NS4A-NS5A (29) of various genotypes. These JFH1-based culture systems have been used for genotype-specific studies of HCV genes (19, 20, 22, 2528, 30), for testing of HCV DAAs (2628) and neutralizing antibodies (19, 20, 22, 30), and for studying aspects of virus–host interaction (25, 31, 32). Through studies of J6 (2a) recombinants with the entire or partial NS5B and 3′UTR from JFH1, we recently identified adaptive mutations F1464L in NS3 and A1672S in NS4A, designated LS (13). Combination of LS with selected mutations in NS5B and the 3′UTR allowed replication of the full-length J6 genome, which led to the identification of an additional unique mutation D2979G in NS5B, designated G (13). The LSG substitutions permitted the development of robust HCV full-length culture systems for J6 and for the prototype genotype 2b J8 strain (13). In this study, we used LSG as a critical component and a unique approach to develop a highly efficient full-length genotype 1a (strain TN) infectious culture system, named TN cell-culture derived (TNcc). The TNcc replicated efficiently in culture, released viral particles of ∼105 focus-forming units (FFU)/mL and did not require additional mutations after viral passage. We demonstrated that full-length TN responded dose dependently to IFN-α, as well as HCV DAAs already used in the clinic and being tested in clinical trials.  相似文献   

13.
The current epidemic of infections caused by antibiotic-resistant Gram-positive bacteria requires the discovery of new drug targets and the development of new therapeutics. Lipoteichoic acid (LTA), a cell wall polymer of Gram-positive bacteria, consists of 1,3-polyglycerol-phosphate linked to glycolipid. LTA synthase (LtaS) polymerizes polyglycerol-phosphate from phosphatidylglycerol, a reaction that is essential for the growth of Gram-positive bacteria. We screened small molecule libraries for compounds inhibiting growth of Staphylococcus aureus but not of Gram-negative bacteria. Compound 1771 [2-oxo-2-(5-phenyl-1,3,4-oxadiazol-2-ylamino)ethyl 2-naphtho[2,1-b]furan-1-ylacetate] blocked phosphatidylglycerol binding to LtaS and inhibited LTA synthesis in S. aureus and in Escherichia coli expressing ltaS. Compound 1771 inhibited the growth of antibiotic-resistant Gram-positive bacteria and prolonged the survival of mice with lethal S. aureus challenge, validating LtaS as a target for the development of antibiotics.Methicillin-resistant Staphylococcus aureus (MRSA), glycopeptide-resistant enterococci (GRE), and antibiotic-resistant Clostridium difficile are examples for newly emerging or reemerging drug-resistant Gram-positive pathogens with significant human morbidity (1). As commensals of the human skin or intestine, these microbes are continuously exposed to antibiotics and have evolved resistance traits against commonly used therapeutics (2). To develop novel antibiotics, a suitable molecular target must be identified (3). Ideal targets are found only in bacteria, not in humans, and are positioned in the bacterial envelope, accessible for small molecule inhibitors yet out of reach of multi–drug-resistance transporters protecting cytoplasmic targets (3, 4). Such attributes are found in penicillin-binding proteins and their inhibitors, β-lactam or glycopeptide antibiotics; however, efforts to identify other extracellular targets in Gram-positive bacteria have met with little success (5).Gram-positive bacteria incorporate lipoteichoic acids (LTAs) into their envelope to scavenge magnesium ions, direct autolysins to subcellular cell wall locations, and enable bacterial cell division (6, 7). LTA from S. aureus and other bacteria is composed of 1,3-polyglycerol-phosphate linked to glycolipid, which provides for LTA anchoring in membranes (8). The glycolipid moiety is composed of β-gentiobiosyldiacylglycerol [glucosyl-(1→6)-glucosyl-(1→3)-diacylglycerol (Glc2-DAG)] (8, 9). PgcA (α-phosphoglucomutase), GtaB (UTP:α-glucose-1-phosphate uridyl transferase), and YpfP (glycosyl-transferase) catalyze the three steps of glycolipid synthesis (1012), whereas LtaA transports Glc2-DAG across the lipid bilayer (10). S. aureus ltaA, ypfP, pgcA, or gtaB mutants cannot assemble Glc2-DAG but continue to synthesize polyglycerol-phosphate (10, 12). Glycolipid synthesis mutants are viable; however, the variants display an increase in size and aberrant cell shapes (10, 12).LTA synthesis involves the polymerization of polyglycerol-phosphate and its transfer to Glc2-DAG (13). This reaction is catalyzed by LTA synthase (LtaS), a protein with five transmembrane domains in its N-terminal domain and an extracellular sulfatase-like domain (pfam00884) at the C-terminal end (14). Genetic depletion or loss of the ltaS gene in S. aureus results in severe cell division defects, a phenotype that is exacerbated when staphylococci are grown at >30 °C (14, 15). Characterization of LtaS in Bacillus anthracis, Bacillus subtilis, and Listeria monocytogenes confirmed that membrane proteins with pfam00884 domains are indeed responsible for the synthesis of polyglycerol-phosphate LTA (7). Where examined, ltaS mutants exhibited diminished viability, increased cell size, and altered morphology (1619). These results suggested that LtaS of Gram-positive bacteria may represent an extracellular target for the development of antibiotics against drug-resistant Gram-positive bacteria. We here provide proof for this hypothesis by isolating a small molecule inhibitor of LTA synthesis.  相似文献   

14.
Voltage-gated sodium (Nav) channels play a fundamental role in the generation and propagation of electrical impulses in excitable cells. Here we describe two unique structurally related nanomolar potent small molecule Nav channel inhibitors that exhibit up to 1,000-fold selectivity for human Nav1.3/Nav1.1 (ICA-121431, IC50, 19 nM) or Nav1.7 (PF-04856264, IC50, 28 nM) vs. other TTX-sensitive or resistant (i.e., Nav1.5) sodium channels. Using both chimeras and single point mutations, we demonstrate that this unique class of sodium channel inhibitor interacts with the S1–S4 voltage sensor segment of homologous Domain 4. Amino acid residues in the “extracellular” facing regions of the S2 and S3 transmembrane segments of Nav1.3 and Nav1.7 seem to be major determinants of Nav subtype selectivity and to confer differences in species sensitivity to these inhibitors. The unique interaction region on the Domain 4 voltage sensor segment is distinct from the structural domains forming the channel pore, as well as previously characterized interaction sites for other small molecule inhibitors, including local anesthetics and TTX. However, this interaction region does include at least one amino acid residue [E1559 (Nav1.3)/D1586 (Nav1.7)] that is important for Site 3 α-scorpion and anemone polypeptide toxin modulators of Nav channel inactivation. The present study provides a potential framework for identifying subtype selective small molecule sodium channel inhibitors targeting interaction sites away from the pore region.Voltage-gated sodium (Nav) channels play an important role in the generation and propagation of electrical signals in excitable cells (13). Eukaryotic Nav channels are heteromeric membrane proteins composed of a pore-forming α-subunit and auxiliary β-subunits (3, 4). The mammalian genome encodes nine distinct α (Nav1.1–1.9) and four β subunits (3). The α-subunit comprises four homologous domains (D1–D4), each of which contains six transmembrane segments (S1–S6) (35). The S5 and S6 segments form the central pore separated by the SS1 and SS2 segments, which form the ion selectivity filter at its extracellular end. The S1 to S4 segments form the voltage sensor (35).Both naturally occurring and synthetic pharmacological modulators of sodium channel have been identified (611), and for many, their site of interaction has been defined. For example, the marine toxin TTX inhibits several Nav subtypes by interacting with amino acid residues within the SS1–SS2 segments that define the outer pore of the channel (12, 13). In contrast, the polypeptide α- and β-scorpion venom toxins, which enhance sodium channel activation or delay inactivation, and spider venom toxin sodium channel inhibitors like Protx II interact with specific residues on the S1–S4 voltage sensor regions within homologous Domain 2 (i.e., β-scorpion toxins, Protx II) or Domain 4 (i.e., α-scorpion toxins, anemone toxins) of the channel (8, 1418). Many synthetic small molecule inhibitors of Nav channels, including local anesthetic, antiepileptic, and antiarrhythmic agents, are believed to interact with amino acid residues within the S6 segment in Domain 4, which forms part of the pore lining and is structurally highly conserved across subclasses of mammalian Nav channels (11, 1922). This structural homology probably accounts for many clinically used local anesthetics and related antiepileptic and antiarrhythmic inhibitors, exhibiting little or no selectivity across the nine subtypes of mammalian Nav channels (23). In the clinic this absence of subtype selectivity can result in toxicities associated with unwanted interactions with off-target Nav channels (e.g., cardiac toxicity due to inhibition of cardiac Nav1.5 channels) (24, 25). Therefore, because of their importance in normal physiology and pathophysiology, identification of selective pharmacological modulators of Nav channels is of considerable interest to the scientific and medical communities (9, 23, 2629). For example, in addition to the therapeutic utility of sodium channel inhibitors described above, there has been recent interest in potentially targeting inhibition of specific Nav channel subtypes (i.e., Nav1.7, Nav1.8, and Nav1.3) for the treatment of pain (9, 3032).The present study describes the characterization of a class of subtype selective sodium channel inhibitor that interacts with a unique site on the voltage sensor region of homologous Domain 4. This inhibitory interaction site differs from previously reported inhibitor binding sites for TTX and local anesthetic-like modulators (11, 12).  相似文献   

15.
Membrane recruitment of cytohesin family Arf guanine nucleotide exchange factors depends on interactions with phosphoinositides and active Arf GTPases that, in turn, relieve autoinhibition of the catalytic Sec7 domain through an unknown structural mechanism. Here, we show that Arf6-GTP relieves autoinhibition by binding to an allosteric site that includes the autoinhibitory elements in addition to the PH domain. The crystal structure of a cytohesin-3 construct encompassing the allosteric site in complex with the head group of phosphatidyl inositol 3,4,5-trisphosphate and N-terminally truncated Arf6-GTP reveals a large conformational rearrangement, whereby autoinhibition can be relieved by competitive sequestration of the autoinhibitory elements in grooves at the Arf6/PH domain interface. Disposition of the known membrane targeting determinants on a common surface is compatible with multivalent membrane docking and subsequent activation of Arf substrates, suggesting a plausible model through which membrane recruitment and allosteric activation could be structurally integrated.Guanine nucleotide exchange factors (GEFs) activate GTPases by catalyzing exchange of GDP for GTP (1). Because many GEFs are recruited to membranes through interactions with phospholipids, active GTPases, or other membrane-associated proteins (15), GTPase activation can be restricted or amplified by spatial–temporal overlap of GEFs with binding partners. GEF activity can also be controlled by autoregulatory mechanisms, which may depend on membrane recruitment (611). Structural relationships between these mechanisms are poorly understood.Arf GTPases function in trafficking and cytoskeletal dynamics (5, 12, 13). Membrane partitioning of a myristoylated (myr) N-terminal amphipathic helix primes Arfs for activation by Sec7 domain GEFs (1417). Cytohesins comprise a metazoan Arf GEF family that includes the mammalian proteins cytohesin-1 (Cyth1), ARNO (Cyth2), and Grp1 (Cyth3). The Drosophila homolog steppke functions in insulin-like growth factor signaling, whereas Cyth1 and Grp1 have been implicated in insulin signaling and Glut4 trafficking, respectively (1820). Cytohesins share a modular architecture consisting of heptad repeats, a Sec7 domain with exchange activity for Arf1 and Arf6, a PH domain that binds phosphatidyl inositol (PI) polyphosphates, and a C-terminal helix (CtH) that overlaps with a polybasic region (PBR) (2128). The overlapping CtH and PBR will be referred to as the CtH/PBR. The phosphoinositide specificity of the PH domain is influenced by alternative splicing, which generates diglycine (2G) and triglycine (3G) variants differing by insertion of a glycine residue in the β1/β2 loop (29). Despite similar PI(4,5)P2 (PIP2) affinities, the 2G variant has 30-fold higher affinity for PI(3,4,5)P3 (PIP3) (30). In both cases, PIP3 is required for plasma membrane (PM) recruitment (23, 26, 3133), which is promoted by expression of constitutively active Arf6 or Arl4d and impaired by PH domain mutations that disrupt PIP3 or Arf6 binding, or by CtH/PBR mutations (8, 3436).Cytohesins are autoinhibited by the Sec7-PH linker and CtH/PBR, which obstruct substrate binding (8). Autoinhibition can be relieved by Arf6-GTP binding in the presence of the PIP3 head group (8). Active myr-Arf1 and myr-Arf6 also stimulate exchange activity on PIP2-containing liposomes (37). Whether this effect is due to relief of autoinhibition per se or enhanced membrane recruitment is not yet clear. Phosphoinositide recognition by PH domains, catalysis of nucleotide exchange by Sec7 domains, and autoinhibition in cytohesins are well characterized (8, 16, 17, 30, 3843). How Arf-GTP binding relieves autoinhibition and promotes membrane recruitment is unknown. Here, we determine the structural basis for relief of autoinhibition and investigate potential mechanistic relationships between allosteric regulation, phosphoinositide binding, and membrane targeting.  相似文献   

16.
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   

17.
The S4 segment and the S4–S5 linker of voltage-gated potassium (Kv) channels are crucial for voltage sensing. Previous studies on the Shaker and Kv1.2 channels have shown that phosphatidylinositol-4,5-bisphosphate (PIP2) exerts opposing effects on Kv channels, up-regulating the current amplitude, while decreasing the voltage sensitivity. Interactions between PIP2 and the S4 segment or the S4–S5 linker in the closed state have been highlighted to explain the effects of PIP2 on voltage sensitivity. Here, we show that PIP2 preferentially interacts with the S4–S5 linker in the open-state KCNQ2 (Kv7.2) channel, whereas it contacts the S2–S3 loop in the closed state. These interactions are different from the PIP2–Shaker and PIP2–Kv1.2 interactions. Consistently, PIP2 exerts different effects on KCNQ2 relative to the Shaker and Kv1.2 channels; PIP2 up-regulates both the current amplitude and voltage sensitivity of the KCNQ2 channel. Disruption of the interaction of PIP2 with the S4–S5 linker by a single mutation decreases the voltage sensitivity and current amplitude, whereas disruption of the interaction with the S2–S3 loop does not alter voltage sensitivity. These results provide insight into the mechanism of PIP2 action on KCNQ channels. In the closed state, PIP2 is anchored at the S2–S3 loop; upon channel activation, PIP2 interacts with the S4–S5 linker and is involved in channel gating.A series of ion channels, such as inward rectifier K+ (Kir) channels, transient receptor potential channels, and voltage-gated channels, are sensitive to the presence of phosphatidylinositol-4,5-bisphosphate (PIP2) in membranes (14). Structural studies on Kir channels (1, 2, 5) demonstrated that PIP2 directly interacts with the channels. Subsequent studies supported that PIP2 also interacts directly with voltage-gated potassium (Kv) channels (619). Several positive residues that may be critical for PIP2 activity have been identified (7, 11, 18, 2024). Previous studies on Kv1.2 and Shaker channels showed that PIP2 exerts opposing effects on Kv channels, up-regulating the current amplitude, while leading to a decrease in voltage sensitivity (7, 18). The S4 segment and the S4–S5 linker of Kv channels are crucial for voltage sensing. The interactions of PIP2 with the S4 segments and the S4–S5 linkers of the closed-state Shaker and Kv1.2 channels underlie the loss-of-function effect of PIP2 on voltage sensitivity (7, 18).The KCNQ (Kv7) family of slowly activated outwardly rectifying potassium channels is one of the Kv channel families that are sensitive to the presence of PIP2 in the membrane. KCNQ channels have been widely studied because of their important biological and pharmacological functions. Retigabine, a first-in-class K+ channel opener used for the treatment of epilepsy, adopts a unique mechanism to enhance the activity of KCNQ channels (25). PIP2 is important for the functions of KCNQ channels. Reduction of PIP2 affinity caused by congenic mutations of KCNQ channels is associated with long QT syndrome, suggesting critical physiological implications of PIP2 on KCNQ channels (23, 26). We reported that PIP2 also alters the pharmacological selectivity of KCNQ potassium channels (6). Zaydman et al. (27) showed that the coupling of voltage sensing and pore opening in the KCNQ1 channel requires PIP2 and suggested there is a PIP2 interaction site at the interface between the voltage-sensing domain (VSD) and the central pore domain (PD). However, the effects and interactions of PIP2 on KCNQ channels are not well understood.Here, by combining molecular dynamics (MD) simulations, mutagenesis, and electrophysiological determinations, we observed that the effects and interactions of PIP2 on KCNQ2 are different relative to the Shaker and Kv1.2 channels. PIP2 up-regulates both the current amplitude and voltage sensitivity of the KCNQ2 channel. PIP2 preferentially interacts with the S4–S5 linker of the open-state KCNQ2 channel and does not interact with the S4 segment or S4-S5 linker of the closed state. In the closed state, PIP2 only interacts with the S2–S3 loop. Furthermore, our electrophysiological experiments suggest that disruption of the interaction of PIP2 with the S4–S5 linker may decrease the voltage sensitivity and current amplitude, whereas disruption of the interaction with the S2–S3 loop only alters the current amplitude of the channel. These results provide insights into the mechanism of PIP2 action on Kv channels.  相似文献   

18.
The inositol 1,4,5-trisphosphate receptor (IP3R) in the endoplasmic reticulum mediates calcium signaling that impinges on intracellular processes. IP3Rs are allosteric proteins comprising four subunits that form an ion channel activated by binding of IP3 at a distance. Defective allostery in IP3R is considered crucial to cellular dysfunction, but the specific mechanism remains unknown. Here we demonstrate that a pleiotropic enzyme transglutaminase type 2 targets the allosteric coupling domain of IP3R type 1 (IP3R1) and negatively regulates IP3R1-mediated calcium signaling and autophagy by locking the subunit configurations. The control point of this regulation is the covalent posttranslational modification of the Gln2746 residue that transglutaminase type 2 tethers to the adjacent subunit. Modification of Gln2746 and IP3R1 function was observed in Huntington disease models, suggesting a pathological role of this modification in the neurodegenerative disease. Our study reveals that cellular signaling is regulated by a new mode of posttranslational modification that chronically and enzymatically blocks allosteric changes in the ligand-gated channels that relate to disease states.Ligand-gated ion channels function by allostery that is the regulation at a distance; the allosteric coupling of ligand binding with channel gating requires reversible changes in subunit configurations and conformations (1). Inositol 1,4,5-trisphosphate receptors (IP3Rs) are ligand-gated ion channels that release calcium ions (Ca2+) from the endoplasmic reticulum (ER) (2, 3). IP3Rs are allosteric proteins comprising four subunits that assemble a calcium channel with fourfold symmetry about an axis perpendicular to the ER membrane. The subunits of three IP3R isoforms (IP3R1, IP3R2, and IP3R3) are structurally divided into three domains: the IP3-binding domain (IBD), the regulatory domain, and the channel domain (36). Fitting of the IBD X-ray structures (7, 8) to a cryo-EM map (9) indicates that the IBD activates a remote Ca2+ channel by allostery (8); however, the current X-ray structure only spans 5% of each tetramer, such that the mechanism underlying allosteric coupling of the IBD to channel gating remains unknown.The IP3R in the ER mediates intracellular calcium signaling that impinges on homeostatic control in various subsequent intracellular processes. Deletion of the genes encoding the type 1 IP3R (IP3R1) leads to perturbations in long-term potentiation/depression (3, 10, 11) and spinogenesis (12), and the human genetic disease spinocerebellar ataxia 15 is caused by haploinsufficiency of the IP3R1 gene (1315). Dysregulation of IP3R1 is also implicated in neurodegenerative diseases including Huntington disease (HD) (1618) and Alzheimer’s disease (AD) (1922). IP3Rs also control fundamental cellular processes—for example, mitochondrial energy production (23, 24), autophagy regulation (2427), ER stress (28), hepatic gluconeogenesis (29), pancreatic exocytosis (30), and macrophage inflammasomes (31). On the other hand, excessive IP3R function promotes cell death processes including apoptosis by activating mitochondrial or calpain pathways (2, 17). Considering these versatile roles of IP3Rs, appropriate IP3R structure and function are essential for living systems, and aberrant regulation of IP3R closely relates to various diseases.Several factors such as cytosolic molecules, interacting proteins, and posttranslational modifications control the IP3-induced Ca2+ release (IICR) through allosteric sites in IP3Rs. Cytosolic Ca2+ concentrations strictly control IICR in a biphasic manner with activation at low concentrations and inhibition at higher concentrations. The critical Ca2+ sensor for activation is conserved among the three isoforms of IP3 and ryanodine receptors, and this sensor is located in the regulatory domain outside the IBD and the channel domain (32). A putative ATP regulatory region is deleted in opisthotonos mice, and IICR is also regulated by this mutation in the regulatory domain (33). Various interacting proteins, such as cytochrome c, Bcl-2-family proteins, ataxin-3, huntingtin (Htt) protein, Htt-associated protein 1A (HAP1A), and G-protein–coupled receptor kinase-interacting protein 1 (GIT1), target allosteric sites in the carboxyl-terminal tail (35). The regulatory domain and the carboxyl-terminal tail also undergo phosphorylation by the protein kinases A/G and B/Akt and contain the apoptotic cleavage sites for the protease caspase-3 (4, 5). These factors allosterically regulate IP3R structure and function to control cellular fates; therefore, understanding the allosteric coupling of the IBD to channel gating will elucidate the regulatory mechanism of these factors.Transglutaminase (TG) catalyses protein cross-linking between a glutamine (Gln) residue and a lysine (Lys) residue via an Nε-(γ-glutamyl)lysine isopeptide bond (34, 35). TG type 2 (TG2) is a Ca2+-dependent enzyme with widespread distribution and is highly inducible by various stimulations such as oxidative stress, cytokines, growth factors, and retinoic acid (RA) (34, 35). TG2 is considered a significant disease-modifying factor in neurodegenerative diseases including HD, AD, and Parkinson’s diseases (PD) (34, 3645) because TG2 might enzymatically stabilize aberrant aggregates of proteins implicated in these diseases—that is, mutant Htt, β-amyloid, and α-synuclein; however, the causal role of TG2 in Ca2+ signaling in brain pathogenesis has been unclear. Ablation of TG2 in HD mouse models is associated with increased lifespan and improved motor function (46, 47). However, TG2 knockout mice do not show impaired Htt aggregation, suggesting that TG2 may play a causal role in these disorders rather than TG2-dependent cross-links in aberrant protein aggregates (47, 48).In this study, we discovered a new mode of chronic and irreversible allosteric regulation in IP3R1 in which covalent modification of the receptor at Gln2746 is catalyzed by TG2. We demonstrate that up-regulation of TG2 modifies IP3R1 structure and function in HD models and propose an etiologic role of this modification in the reduction of neuronal signaling and subsequent processes during the prodromal state of the neurodegenerative disease.  相似文献   

19.
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   

20.
Archaebacterial halophiles (Haloarchaea) are oxygen-respiring heterotrophs that derive from methanogens—strictly anaerobic, hydrogen-dependent autotrophs. Haloarchaeal genomes are known to have acquired, via lateral gene transfer (LGT), several genes from eubacteria, but it is yet unknown how many genes the Haloarchaea acquired in total and, more importantly, whether independent haloarchaeal lineages acquired their genes in parallel, or as a single acquisition at the origin of the group. Here we have studied 10 haloarchaeal and 1,143 reference genomes and have identified 1,089 haloarchaeal gene families that were acquired by a methanogenic recipient from eubacteria. The data suggest that these genes were acquired in the haloarchaeal common ancestor, not in parallel in independent haloarchaeal lineages, nor in the common ancestor of haloarchaeans and methanosarcinales. The 1,089 acquisitions include genes for catabolic carbon metabolism, membrane transporters, menaquinone biosynthesis, and complexes I–IV of the eubacterial respiratory chain that functions in the haloarchaeal membrane consisting of diphytanyl isoprene ether lipids. LGT on a massive scale transformed a strictly anaerobic, chemolithoautotrophic methanogen into the heterotrophic, oxygen-respiring, and bacteriorhodopsin-photosynthetic haloarchaeal common ancestor.Halophilic archaebacteria (Haloarchaea) require concentrated salt solutions for survival and can inhabit saturated brine environments such as salt lakes, the Dead Sea, and salterns (1). In rRNA and phylogenomic analyses of informational genes, Haloarchaea always branch well within the methanogens (24). Haloarchaea can thus be seen as deriving from methanogen ancestors, but the physiology of methanogens and halophiles could hardly be more different. Methanogens are strict anaerobes, most species are lithoautotrophs that use electrons from H2 to reduce CO2 to methane (obligate hydrogenotrophic methanogens), thereby generating a chemiosmotic ion gradient for ATP synthesis in their energy metabolism, although some species can generate methane from reduced C1 compounds, or acetate in the case of aceticlastic forms (57). Their carbon metabolism involves the Wood–Ljungdahl (acetyl-CoA) pathway of CO2 fixation (57). In contrast, Haloarchaea are obligate heterotrophs that typically use O2 as the terminal acceptor of their electron transport chain, although many can also use alternative electron acceptors such as nitrate in addition to light harnessing via a bacteriorhodopsin-based proton pumping system (8). The evolutionary nature of that radical physiological transformation from anaerobic chemolithoautotroph to aerobic heterotroph is of interest.Many individual reports document that lateral gene transfer (LGT) from eubacteria was involved in the origin of at least some components of haloarchaeal metabolism. These include the operon for gas vesicle formation, which allows Haloarchaea to remain in surface waters (9), the newly identified methylaspartate cycle of acetyl-CoA oxidation (10), various components of the haloarchaeal aerobic respiratory chain (1118), and proteins involved in the assembly of FeS clusters (19). The sequencing of the first haloarchaeal genome over a decade ago identified some eubacterial genes that possibly could have been acquired by lateral gene transfer (11, 20), and whereas substantial data that would illuminate the origin of haloarchaeal physiology have accumulated since then, those data have not been subjected to comparative evolutionary analysis. Investigating the role of the environment in haloarchaeal genome evolution, Rhodes et al. (21) recently showed that Haloarchaea are indeed far more likely to acquire genes from other halophiles, but they did not address the issues at the focus of our present investigation, namely: How many eubacterial acquisitions are present in haloarchaeal genomes? How was the physiological transformation of methanogens to Haloarchaea affected by LGT? Do those acquisitions trace to the haloarchaeal common ancestor as a single acquisition or not?To discern whether the eubacterial genes in haloarchaeal genomes are the result of multiple independent transfers in individual lineages or the result of a single ancient mass acquisition, here we have analyzed 10 sequenced haloarchaeal genomes—Haloarcula marismortui (22), Halobacterium salinarum (23), Halobacterium sp. (20), Halomicrobium mukohataei (24), Haloquadratum walsbyi (25), Halorhabdus utahensis (26), Halorubrum lacusprofundi (27), Natrialba magadii (28), Natronomonas pharaonis (29), and Haloterrigena turkmmenica (30)—in the context of 65 other archaebacterial and >1,000 eubacterial reference genomes.  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