首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
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   

2.
Familial hemiplegic migraine (FHM) is a rare subtype of migraine with aura. Mutations causing FHM type 3 have been identified in SCN1A, the gene encoding the Nav1.1 Na+ channel, which is also a major target of epileptogenic mutations and is particularly important for the excitability of GABAergic neurons. However, functional studies of NaV1.1 FHM mutations have generated controversial results. In particular, it has been shown that the NaV1.1-L1649Q mutant is nonfunctional when expressed in a human cell line because of impaired plasma membrane expression, similarly to NaV1.1 mutants that cause severe epilepsy, but we have observed gain-of-function effects for other NaV1.1 FHM mutants. Here we show that NaV1.1-L1649Q is nonfunctional because of folding defects that are rescuable by incubation at lower temperatures or coexpression of interacting proteins, and that a partial rescue is sufficient for inducing an overall gain of function because of the modifications in gating properties. Strikingly, when expressed in neurons, the mutant was partially rescued and was a constitutive gain of function. A computational model showed that 35% rescue can be sufficient for inducing gain of function. Interestingly, previously described folding-defective epileptogenic NaV1.1 mutants show loss of function also when rescued. Our results are consistent with gain of function as the functional effect of NaV1.1 FHM mutations and hyperexcitability of GABAergic neurons as the pathomechanism of FHM type 3.Epilepsy and migraine are common neurologic disorders that may have pathophysiological links (13). Mutations have been identified for some rare types of epilepsy and migraine (1, 46), opening a window for investigating their pathogenic mechanisms, which may provide useful information also about more common forms. The Na+ channel α subunit NaV1.1, encoded by the SCN1A gene, is the target of hundreds of epileptogenic mutations (79), and of mutations causing familial hemiplegic migraine type 3 (FHM-3), a rare subtype of migraine with aura characterized by hemiplegia during the attacks, which can also be caused by mutations of CaV2.1 Ca2+ channels and the α2 subunit of the Na+/K+ ATPase (FHM types 1 and 2) (1, 6). The results of most studies suggest that epileptogenic NaV1.1 mutations cause variable degrees of loss of function of NaV1.1, leading to reduced Na+ current and excitability in GABAergic neurons, and resulting in decreased inhibition in neuronal networks (1014). The most severe phenotypes (e.g., Dravet syndrome, an extremely severe epileptic encephalopathy) are in general caused by mutations that induce complete NaV1.1 loss of function, leading to haploinsufficiency (15). Thus, it has been hypothesized that a more severe loss of function would cause more severe epilepsy (8). Functional studies of NaV1.1 FHM mutations have generated more confusing results (1). For instance, we have reported gain-of-function effects for the mutant Q1489K causing pure FHM (16), and modulable gain-/loss-of-function effects for the mutant T1174S associated with FHM or mild epilepsy in different branches of the family (17). Overall, our results are consistent with a gain of function of NaV1.1 as the cause of FHM, which might induce cortical spreading depression (CSD), a probable pathomechanism of migraine, because of hyperexcitability of GABAergic interneurons (16). However, a study has reported loss of function for FHM hNaV1.1 mutants expressed in the human cell line tsA-201—in particular, complete loss of function for the L1649Q mutant because of lack of cell surface expression (18). L1649Q has been identified in a four-generation family with eight members presenting with FHM, without epilepsy or other neurologic symptoms (19); this is a puzzling result more consistent with a phenotype of severe epilepsy (7, 8). We have found that NaV1.1 epileptogenic mutations can induce loss of function by causing folding defects (20), which can be partially rescued by incubation of the transfected cells at lower temperatures (≤30 °C) or by molecular interactions (21, 22), as recently confirmed also for other epileptogenic NaV1.1 mutants (23, 24). We report here that L1649Q is a folding-defective mutant that, when partially rescued, is characterized by an overall gain of function, consistent with our hypothesis of FHM type 3 pathomechanism (16).  相似文献   

3.
Excitation-evoked Ca2+ influx is the fastest and most ubiquitous chemical trigger for cellular processes, including neurotransmitter release, muscle contraction, and gene expression. The voltage dependence and timing of Ca2+ entry are thought to be functions of voltage-gated calcium (CaV) channels composed of a central pore regulated by four nonidentical voltage-sensing domains (VSDs I–IV). Currently, the individual voltage dependence and the contribution to pore opening of each VSD remain largely unknown. Using an optical approach (voltage-clamp fluorometry) to track the movement of the individual voltage sensors, we discovered that the four VSDs of CaV1.2 channels undergo voltage-evoked conformational rearrangements, each exhibiting distinct voltage- and time-dependent properties over a wide range of potentials and kinetics. The voltage dependence and fast kinetic components in the activation of VSDs II and III were compatible with the ionic current properties, suggesting that these voltage sensors are involved in CaV1.2 activation. This view is supported by an obligatory model, in which activation of VSDs II and III is necessary to open the pore. When these data were interpreted in view of an allosteric model, where pore opening is intrinsically independent but biased by VSD activation, VSDs II and III were each found to supply ∼50 meV (∼2 kT), amounting to ∼85% of the total energy, toward stabilizing the open state, with a smaller contribution from VSD I (∼16 meV). VSD IV did not appear to participate in channel opening.Voltage-gated Ca2+ (CaV) channels respond to membrane depolarization by catalyzing Ca2+ influx. CaV-mediated elevation of intracellular [Ca2+] regulates such critical physiological functions as neurotransmitter and hormone release, axonal outgrowth, muscle contraction, and gene expression (1). Their relevance to human physiology is evident from the broad phenotypic consequences of CaV channelopathies (2). The voltage dependence of CaV-driven Ca2+ entry relies on the modular organization of the channel-forming α1 subunit (Fig. 1), which consists of four repeated motifs (I–IV), each comprising six membrane-spanning helical segments (S1–S6) (Fig. 1A). Segments S1–S4 form a voltage-sensing domain (VSD), whereas segments S5 and S6 contribute to the Ca2+-conductive pore (1). The VSDs surround the central pore (Fig. 1B). VSDs are structurally and functionally conserved modules (35) capable of transducing a change in the cell membrane electrical potential into a change of ion-specific permeability or enzyme activity. VSDs sense depolarization by virtue of a signature motif of positively charged Arg or Lys at every third position of helix S4 (Fig. 1D), which rearranges in response to depolarization (4, 610). In contrast to voltage-gated K+ (KV) channels but similar to pseudotetrameric voltage-gated Na+ (NaV) channels, the amino acid sequences encoding each VSD have evolved independently (Fig. 1D). In addition to their distinct primary structure, the four CaV VSDs may also gain distinct functional properties from the asymmetrical association of auxiliary subunits, such as β, α2δ, and calmodulin (1, 1116) (Fig. 1C). The structural divergence among VSD-driven channels was foreseen by the classical Hodgkin–Huxley model (17), in which four independent “gating particles” control the opening of homotetrameric KV channels and only three seem sufficient to open NaV channels. An early study by Kostyuk et al. (18) suggested that only two gating particles are coupled to CaV channel opening. We recognize today that gating particles correspond to VSDs, and in NaV channels, VSDs I–III control Na+ influx, whereas VSD IV is associated with fast inactivation (1921).Open in a separate windowFig. 1.CaV membrane topology, putative structure, and S4 helix homology. (A) CaV channel-forming α1 subunits consist of four concatenated repeats, each encompassing one voltage sensor domain (VSD) and a quarter of the central pore domain (PD) (1). Stars indicate the positions of fluorophore labeling. (B) The atomic structure of an NaV channel (Protein Data Bank ID code 4EKW; top view) (56) shown as a structural representation for the CaV α1 subunit. (C) The α1 subunit asymmetrically associates with auxiliary β, α2δ, and calmodulin (CaM) subunits (1116). (D) Sequence alignment of VSD helix S4 from each of four CaV1.2 repeats and the archetypal homotetrameric Shaker K+ channel. Conserved, positively charged Arg or Lys is in blue. Residues substituted by Cys for fluorescent labeling are marked: F231 (VSD I), L614 (VSD II), V994 (VSD II), and S1324 (VSD IV).In this study, we used fluorometry to probe the properties of four individual VSDs in a human L-type calcium channel CaV1.2, which is a widely expressed regulator of physiological processes, such as cardiac and smooth muscle contractility (22). Although the collective transition of the CaV VSDs and the pore has been investigated in studies measuring total charge displacement (gating currents) (23, 24), the activation properties and functional roles of each VSD are unknown. Evidence for the role of each VSD in L-type CaV channel operation has been presented from charge neutralization studies, but a clear picture has yet to emerge. Work on a chimeric L-type channel suggests that VSDs I and III drive channel opening (25), whereas other studies on CaV1.2 favored the involvement of VSD II over VSD I, with the roles of VSDs III and IV remaining unclear (26, 27).The individual optical reports of four CaV1.2 VSDs revealed that each operates with distinct biophysical parameters. We found that VSDs II and III exhibit voltage- and time-dependent characteristics compatible with channel opening and that they can be considered rate-limiting for activation. We compared the voltage and time dependence of the fluorescent signals and ionic currents with the predictions of thermodynamic models relevant to CaV domain organization. We found that CaV1.2 activation is compatible with a model of allosteric VSD–pore coupling, where VSDs II and III are the primary drivers of channel opening with a smaller contribution by VSD I. We discuss the mechanism of CaV1.2 voltage sensitivity, which exhibits similarities to but also clear differences from the related pseudotetrameric NaV1.4 channels.  相似文献   

4.
Haploinsufficiency of the voltage-gated sodium channel NaV1.1 causes Dravet syndrome, an intractable developmental epilepsy syndrome with seizure onset in the first year of life. Specific heterozygous deletion of NaV1.1 in forebrain GABAergic-inhibitory neurons is sufficient to cause all the manifestations of Dravet syndrome in mice, but the physiological roles of specific subtypes of GABAergic interneurons in the cerebral cortex in this disease are unknown. Voltage-clamp studies of dissociated interneurons from cerebral cortex did not detect a significant effect of the Dravet syndrome mutation on sodium currents in cell bodies. However, current-clamp recordings of intact interneurons in layer V of neocortical slices from mice with haploinsufficiency in the gene encoding the NaV1.1 sodium channel, Scn1a, revealed substantial reduction of excitability in fast-spiking, parvalbumin-expressing interneurons and somatostatin-expressing interneurons. The threshold and rheobase for action potential generation were increased, the frequency of action potentials within trains was decreased, and action-potential firing within trains failed more frequently. Furthermore, the deficit in excitability of somatostatin-expressing interneurons caused significant reduction in frequency-dependent disynaptic inhibition between neighboring layer V pyramidal neurons mediated by somatostatin-expressing Martinotti cells, which would lead to substantial disinhibition of the output of cortical circuits. In contrast to these deficits in interneurons, pyramidal cells showed no differences in excitability. These results reveal that the two major subtypes of interneurons in layer V of the neocortex, parvalbumin-expressing and somatostatin-expressing, both have impaired excitability, resulting in disinhibition of the cortical network. These major functional deficits are likely to contribute synergistically to the pathophysiology of Dravet syndrome.Dravet syndrome (DS), also referred to as “severe myoclonic epilepsy in infancy,” is a rare genetic epileptic encephalopathy characterized by frequent intractable seizures, severe cognitive deficits, and premature death (13). DS is caused by loss-of-function mutations in SCN1A, the gene encoding type I voltage-gated sodium channel NaV1.1, which usually arise de novo in the affected individuals (47). Like DS patients, mice with heterozygous loss-of-function mutations in Scn1a exhibit ataxia, sleep disorder, cognitive deficit, autistic-like behavior, and premature death (814). Like DS patients, DS mice first become susceptible to seizures caused by elevation of body temperature and subsequently experience spontaneous myoclonic and generalized tonic-clonic seizures (11). Global deletion of NaV1.1 impairs Na+ currents and action potential (AP) firing in GABAergic-inhibitory interneurons (810), and specific deletion of NaV1.1 in forebrain interneurons is sufficient to cause DS in mice (13, 15). These data suggest that the loss of interneuron excitability and resulting disinhibition of neural circuits cause DS, but the functional role of different subtypes of interneurons in the cerebral cortex in DS remains unknown.Neocortical GABAergic interneurons shape cortical output and display great diversity in morphology and function (16, 17). The expression of parvalbumin (PV) and somatostatin (SST) defines two large, nonoverlapping groups of interneurons (16, 18, 19). In layer V of the cerebral cortex, PV-expressing fast-spiking interneurons and SST-expressing Martinotti cells each account for ∼40% of interneurons, and these interneurons are the major inhibitory regulators of cortical network activity (17, 20). Layer V PV interneurons make synapses on the soma and proximal dendrites of pyramidal neurons (18, 19), where they mediate fast and powerful inhibition (21, 22). Selective heterozygous deletion of Scn1a in neocortical PV interneurons increases susceptibility to chemically induced seizures (23), spontaneous seizures, and premature death (24), indicating that this cell type may contribute to Scn1a deficits. However, selective deletion of Scn1a in neocortical PV interneurons failed to reproduce the effects of DS fully, suggesting the involvement of other subtypes of interneurons in this disease (23, 24). Layer V Martinotti cells have ascending axons that arborize in layer I and spread horizontally to neighboring cortical columns, making synapses on apical dendrites of pyramidal neurons (17, 25, 26). They generate frequency-dependent disynaptic inhibition (FDDI) that dampens excitability of neighboring layer V pyramidal cells (2729), contributing to maintenance of the balance of excitation and inhibition in the neocortex. However, the functional roles of Martinotti cells and FDDI in DS are unknown.Because layer V forms the principal output pathway of the neocortex, reduction in inhibitory input to layer V pyramidal cells would have major functional consequences by increasing excitatory output from all cortical circuits. However, the effects of the DS mutation on interneurons and neural circuits that provide inhibitory input to layer V pyramidal cells have not been determined. Here we show that the intrinsic excitability of layer V fast-spiking PV interneurons and SST Martinotti cells and the FDDI mediated by Martinotti cells are reduced dramatically in DS mice, leading to an imbalance in the excitation/inhibition ratio. Our results suggest that loss of NaV1.1 in these two major types of interneurons may contribute synergistically to increased cortical excitability, epileptogenesis, and cognitive deficits in DS.  相似文献   

5.
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   

6.
Loss-of-function mutations in the human voltage-gated sodium channel NaV1.7 result in a congenital indifference to pain. Selective inhibitors of NaV1.7 are therefore likely to be powerful analgesics for treating a broad range of pain conditions. Herein we describe the identification of µ-SLPTX-Ssm6a, a unique 46-residue peptide from centipede venom that potently inhibits NaV1.7 with an IC50 of ∼25 nM. µ-SLPTX-Ssm6a has more than 150-fold selectivity for NaV1.7 over all other human NaV subtypes, with the exception of NaV1.2, for which the selectivity is 32-fold. µ-SLPTX-Ssm6a contains three disulfide bonds with a unique connectivity pattern, and it has no significant sequence homology with any previously characterized peptide or protein. µ-SLPTX-Ssm6a proved to be a more potent analgesic than morphine in a rodent model of chemical-induced pain, and it was equipotent with morphine in rodent models of thermal and acid-induced pain. This study establishes µ-SPTX-Ssm6a as a promising lead molecule for the development of novel analgesics targeting NaV1.7, which might be suitable for treating a wide range of human pain pathologies.Normal pain is a key adaptive response that serves to limit our exposure to potentially damaging or life-threatening events. In contrast, aberrant long-lasting pain transforms this adaptive response into a debilitating and often poorly managed disease. Chronic pain affects ∼20% of the population, with the incidence rising significantly in elderly cohorts (1). The economic burden of chronic pain in the United States was recently estimated to be ∼$600 billion per annum, which exceeds the combined annual cost of cancer, heart disease, and diabetes (2). There are few drugs available for treatment of chronic pain, and many of these have limited efficacy and dose-limiting side-effects.Voltage-gated sodium (NaV) channels are integral transmembrane proteins that provide a current pathway for the rapid depolarization of excitable cells (1, 3), and they play a key role in conveying nociceptor responses to synapses in the dorsal horn (4). Humans contain nine different NaV channel subtypes, denoted NaV1.1 to NaV1.9 (5, 6). In recent years, NaV1.7 has emerged as a promising analgesic target based on several remarkable human genetic studies. Gain-of-function mutations in the SNC9A gene encoding the pore-forming α-subunit of NaV1.7 cause severe episodic pain in inherited neuropathies, such as erythromelalgia and paroxysmal extreme pain disorder (7), whereas loss-of-function mutations in SCN9A result in a congenital indifference to pain (CIP) (8). The latter phenotype can be recapitulated in rodents via complete knockout of NaV1.7 in all sensory and sympathetic neurons (9). Moreover, certain polymorphisms in SCN9A correlate with sensitivity to nociceptive inputs (10). Remarkably, apart from their inability to sense pain, loss of smell (anosmia) is the only other sensory impairment in individuals with CIP (11, 12). Thus, the combined genetic data suggest that subtype-selective blockers of NaV1.7 are likely to be useful analgesics for treating a broad range of pain conditions.Centipedes are one of the oldest extant arthropods, with the fossil record dating back 430 million y (13). Centipedes were one of the first terrestrial taxa to use venom as a predation strategy, and they have adapted to capture a wide variety of prey, including insects, fish, molluscs, amphibians, reptiles, and even mammals (13, 14). The centipede venom apparatus, which is unique and bears little resemblance to that of other arthropods, evolved by modification of the first pair of walking legs into a set of pincer-like claws (forcipules) (13). Venom is secreted via a pore located near the tip of each forcipule. There are ∼3,300 extant species of centipedes, yet the venom of only a handful has been studied in any detail. We recently demonstrated that the venom of the Chinese red-headed centipede Scolopendra subspinipes mutilans is replete with unique, disulfide-rich peptides that potently modulate the activity of mammalian voltage-gated ion channels (14), and therefore we decided to explore this venom as a potential source of NaV1.7 inhibitors. We describe the purification from this venom of a highly selective inhibitor of NaV1.7 that is a more effective analgesic than morphine in rodent pain models.  相似文献   

7.
A constitutional isomeric library synthesized by a modular approach has been used to discover six amphiphilic Janus dendrimer primary structures, which self-assemble into uniform onion-like vesicles with predictable dimensions and number of internal bilayers. These vesicles, denoted onion-like dendrimersomes, are assembled by simple injection of a solution of Janus dendrimer in a water-miscible solvent into water or buffer. These dendrimersomes provide mimics of double-bilayer and multibilayer biological membranes with dimensions and number of bilayers predicted by the Janus compound concentration in water. The simple injection method of preparation is accessible without any special equipment, generating uniform vesicles, and thus provides a promising tool for fundamental studies as well as technological applications in nanomedicine and other fields.Most living organisms contain single-bilayer membranes composed of lipids, glycolipids, cholesterol, transmembrane proteins, and glycoproteins (1). Gram-negative bacteria (2, 3) and the cell nucleus (4), however, exhibit a strikingly special envelope that consists of a concentric double-bilayer membrane. More complex membranes are also encountered in cells and their various organelles, such as multivesicular structures of eukaryotic cells (5) and endosomes (6), and multibilayer structures of endoplasmic reticulum (7, 8), myelin (9, 10), and multilamellar bodies (11, 12). This diversity of biological membranes inspired corresponding biological mimics. Liposomes (Fig. 1) self-assembled from phospholipids are the first mimics of single-bilayer biological membranes (1316), but they are polydisperse, unstable, and permeable (14). Stealth liposomes coassembled from phospholipids, cholesterol, and phospholipids conjugated with poly(ethylene glycol) exhibit improved stability, permeability, and mechanical properties (1720). Polymersomes (2124) assembled from amphiphilic block copolymers exhibit better mechanical properties and permeability, but are not always biocompatible and are polydisperse. Dendrimersomes (2528) self-assembled from amphiphilic Janus dendrimers and minidendrimers (2628) have also been elaborated to mimic single-bilayer biological membranes. Amphiphilic Janus dendrimers take advantage of multivalency both in their hydrophobic and hydrophilic parts (23, 2932). Dendrimersomes are assembled by simple injection (33) of a solution of an amphiphilic Janus dendrimer (26) in a water-soluble solvent into water or buffer and produce uniform (34), impermeable, and stable vesicles with excellent mechanical properties. In addition, their size and properties can be predicted by their primary structure (27). Amphiphilic Janus glycodendrimers self-assemble into glycodendrimersomes that mimic the glycan ligands of biological membranes (35). They have been demonstrated to be bioactive toward biomedically relevant bacterial, plant, and human lectins, and could have numerous applications in nanomedicine (20).Open in a separate windowFig. 1.Strategies for the preparation of single-bilayer vesicles and multibilayer onion-like vesicles.More complex and functional cell mimics such as multivesicular vesicles (36, 37) and multibilayer onion-like vesicles (3840) have also been discovered. Multivesicular vesicles compartmentalize a larger vesicle (37) whereas multibilayer onion-like vesicles consist of concentric alternating bilayers (40). Currently multibilayer vesicles are obtained by very complex and time-consuming methods that do not control their size (39) and size distribution (40) in a precise way. Here we report the discovery of “single–single” (28) amphiphilic Janus dendrimer primary structures that self-assemble into uniform multibilayer onion-like dendrimersomes (Fig. 1) with predictable size and number of bilayers by simple injection of their solution into water or buffer.  相似文献   

8.
Voltage-dependent gating of ion channels is essential for electrical signaling in excitable cells, but the structural basis for voltage sensor function is unknown. We constructed high-resolution structural models of resting, intermediate, and activated states of the voltage-sensing domain of the bacterial sodium channel NaChBac using the Rosetta modeling method, crystal structures of related channels, and experimental data showing state-dependent interactions between the gating charge-carrying arginines in the S4 segment and negatively charged residues in neighboring transmembrane segments. The resulting structural models illustrate a network of ionic and hydrogen-bonding interactions that are made sequentially by the gating charges as they move out under the influence of the electric field. The S4 segment slides 6–8 Å outward through a narrow groove formed by the S1, S2, and S3 segments, rotates ∼30°, and tilts sideways at a pivot point formed by a highly conserved hydrophobic region near the middle of the voltage sensor. The S4 segment has a 310-helical conformation in the narrow inner gating pore, which allows linear movement of the gating charges across the inner one-half of the membrane. Conformational changes of the intracellular one-half of S4 during activation are rigidly coupled to lateral movement of the S4–S5 linker, which could induce movement of the S5 and S6 segments and open the intracellular gate of the pore. We confirmed the validity of these structural models by comparing with a high-resolution structure of a NaChBac homolog and showing predicted molecular interactions of hydrophobic residues in the S4 segment in disulfide-locking studies.Voltage-gated sodium (NaV) channels are responsible for initiation and propagation of action potentials in nerve, muscle, and endocrine cells (1, 2). They are members of the structurally homologous superfamily of voltage-gated ion channel proteins that also includes voltage-gated potassium (KV), voltage-gated calcium (CaV), and cyclic nucleotide-gated (CNG) channels (3). Mammalian NaV and CaV channels consist of four homologous domains (I through IV), each containing six transmembrane segments (S1 through S6) and a membrane-reentrant pore loop between the S5 and S6 segments (1, 3). Segments S1–S4 of the channel form the voltage-sensing domain (VSD), and segments S5 and S6 and the membrane-reentrant pore loop form the pore. The bacterial NaV channel NaChBac and its relatives consist of tetramers of four identical subunits, which closely resemble one domain of vertebrate NaV and CaV channels, but provide much simpler structures for studying the mechanism of voltage sensing (4, 5). The hallmark feature of the voltage-gated ion channels is the steep voltage dependence of activation, which derives from the voltage-driven outward movement of gating charges in response to the membrane depolarization (6, 7). The S4 transmembrane segment in the VSD has four to seven arginine residues spaced at 3-aa intervals, which serve as gating charges in the voltage-sensing mechanism (815). The intracellular S4–S5 linker that connects the VSD to the pore plays a key role in coupling voltage-dependent conformational changes in the VSD to opening and closing of the pore (16). The gating charges are pulled in by the internally negative transmembrane electric field and released to move out on depolarization. Their outward movement must be catalyzed by the voltage sensor to reduce the large thermodynamic barrier to movement of charged amino acid residues across the membrane. The molecular mechanism by which the gating charges are stabilized in the hydrophobic transmembrane environment and the catalytic mechanism through which they are transported across the membrane in response to changes in membrane potential are the subjects of intense research efforts.Progress has been made in determining high-resolution structures of voltage sensors of KV and NaV channels in activated states (1720). However, high-resolution structures of resting and intermediate states of voltage sensors are unknown. The majority of evidence supports a sliding helix model of the voltage-dependent gating in which the gating charge-carrying arginines in S4 are proposed to sequentially form ion pairs with negatively charged residues in S1–S3 segments during activation of the channel (911, 21). However, the structural basis for stabilization of the gating charges in the membrane and catalysis of their movement through the hydrophobic membrane environment remain uncertain. Here, we have integrated bioinformatics analysis of NaV and KV channel families using the HHPred homology detection server (2224), high-resolution structural modeling using the Rosetta Membrane (2527) and Rosetta Symmetry methods (28), the X-ray structures of the Kv1.2-Kv2.1 chimeric channel and NavAb with activated VSDs (19, 20) and the MlotiK1 CNG channel in the resting state (29), and experimental data showing sequential state-dependent interactions between gating charges in S4 and negatively charged residues in S1–S3 (this work and refs. 3033). Predictions of the resulting voltage-sensing model are confirmed in this work by disulfide-locking studies and mutant cycle analysis of the interactions of hydrophobic residues in the S4 segment. This model reveals structural details of the voltage-dependent conformational changes in the VSD that stabilize and catalyze gating charge movement and are coupled to opening and closing of the intracellular activation gate of the ion-conducting pore.  相似文献   

9.
It is not known which kinetic step in the acto-myosin ATPase cycle limits contraction speed in unloaded muscles (V0). Huxley’s 1957 model [Huxley AF (1957) Prog Biophys Biophys Chem 7:255–318] predicts that V0 is limited by the rate that myosin detaches from actin. However, this does not explain why, as observed by Bárány [Bárány M (1967) J Gen Physiol 50(6, Suppl):197–218], V0 is linearly correlated with the maximal actin-activated ATPase rate (vmax), which is limited by the rate that myosin attaches strongly to actin. We have observed smooth muscle myosin filaments of different length and head number (N) moving over surface-attached F-actin in vitro. Fitting filament velocities (V) vs. N to a detachment-limited model using the myosin step size d = 8 nm gave an ADP release rate 8.5-fold faster and ton (myosin’s attached time) and r (duty ratio) ∼10-fold lower than previously reported. In contrast, these data were accurately fit to an attachment-limited model, V = N·v·d, over the range of N found in all muscle types. At nonphysiologically high N, V = L/ton rather than d/ton, where L is related to the length of myosin’s subfragment 2. The attachment-limited model also fit well to the [ATP] dependence of V for myosin-rod cofilaments at three fixed N. Previously published V0 vs. vmax values for 24 different muscles were accurately fit to the attachment-limited model using widely accepted values for r and N, giving d = 11.1 nm. Therefore, in contrast with Huxley’s model, we conclude that V0 is limited by the actin–myosin attachment rate.An important parameter used to characterize and compare muscle types is the maximum shortening velocity defined as the velocity of intact (1) or skinned (2) muscle fibers and myofibrils (3) shortening under no load (V0). Any model of muscle contraction must be able to quantitatively predict V0 in terms of rate constants in the acto-myosin ATPase cycle (4). Huxley’s 1957 model of muscle contraction (5), which is still the most widely accepted model, predicts that V0 is limited by the rate of detachment of myosin from actin, which is limited by the rate of ADP release from acto-myosin ((k-AD) (6) or the rate of attachment of ATP (kT·[ATP]) to acto-myosin (7) (Fig. 1A). Stated simply, Huxley’s model says that muscle contracts only as fast as the myosins at the end of their working cycle can detach from actin.Open in a separate windowFig. 1.Kinetic scheme and various myosin structures. (A) Kinetic scheme for myosin (M) attachment to actin (A), D = ADP, T= ATP, Pi = phosphate. Kw is the equilibrium constant for weak binding and kws is the forward rate constant for the weak to strong transition. See the text for other rate constants. (B) Classic IVMA with actin (green), being moved by myosin heads (orange) attached to coverslip (blue). (C) Myosin II molecule. (D) Comparison of side-polar (smooth muscle) and bipolar myosin filaments (skeletal and cardiac muscle). (E) Inverted IVMA, myosin filament (orange) moves over biotinylated actin attached to the coverslip (light blue) via biotinylated PEG brush (blue) and streptavidin (not shown for clarity) (21).In a classic study, Bárány (8) found a linear correlation between V0 and maximal ATPase rates (vmax) among different muscle types from organisms of broad evolutionary origin suggesting that V0 and ATPase are limited by the same kinetic step. However, it is well established that the ATPase rate (4) in myofibrils (9) and in purified acto-myosin in solution (4, 10) is not detachment-limited, but rather is attachment-limited, controlled by Kw·kws = katt (Fig. 1A) (4). If V0 is attachment-limited, unloaded muscles will contract essentially as fast as N myosin heads with a given step size attach weakly to actin and undergo the powerstroke to the strongly bound state.The mechanism that limits the velocity of muscle contraction has been addressed using the classic in vitro motility assay (IVMA), where actin filaments move over surface-attached monomeric myosin (1114) (Fig. 1B). The maximal velocity (Vf) is primarily limited by k-AD (1517) and therefore Vf = d·k-AD. However, recent studies show that Vf exceeds the detachment limit (1821) and that factors that specifically inhibit attachment kinetics also slow Vf, suggesting that attachment kinetics influence Vf. The effects of attachment kinetics on V0 appear to be even more significant because V0 is considerably faster than Vf (16, 2224), which may indicate that the kinetic steps limiting Vf and V0 are not the same. However, it is difficult to draw firm conclusions because several aspects of the IVMA may lead to slower velocities: random myosin orientation on the surface (25), inactive myosin (26), low ionic strength necessary to keep actin bound to myosin (16, 22), the absence of regulatory proteins (27), and absence of the same geometry as the muscle lattice.Because myosin II has a low duty ratio (r) (20, 28, 29), many molecules are required to generate processive motion. Therefore, the functional form of myosin in vivo is the filament (Fig. 1D). We have recently introduced an inverted IVMA in which smooth muscle myosin (SMM) filaments move over surface-attached actin [Fig. 1E (21)], which more closely mimics the geometry of a muscle lattice and minimizes surface interactions.Here, we used this inverted assay to address which kinetic step limits V0. We prepared side-polar SMM filaments (21) that, unlike bipolar skeletal and cardiac filaments, have all heads on one side of the filament orientated in the same direction (Fig. 1D) (3035). We examined the relationship between V and N because the N dependence of V is a powerful test to discriminate between a detachment-limited and an attachment-limited model. N was reduced at a fixed filament length by making SMM-rod cofilaments and higher N was achieved by making SMM filaments with a wide range of longer lengths. We found that within the range of N seen in essentially all types of myosin filaments in muscle, V was linearly related to N, which fits the model where V is limited by the ATPase rate (v). As N was increased far beyond the physiological N, V approached an N-independent detachment limit with V = L·k-AD (not d·k-AD) where L is the “tether length” of the region of the myosin tail that is proximal to the heads known as subfragment 2 (S2, Fig. 1C). We confirmed our findings by measuring V at several fixed N (cofilaments) as a function of [ATP]. These data make sense in the context of the nonlinear stiffness of full-length myosin II when it is measured within a myosin filament (36, 37).The equations that fit our data are consistent with the relationship between vmax and V0 originally shown by Bárány (8), leading to the conclusion that unloaded muscles contract at velocities that are limited by their respective ATPase activities (attachment-limited).  相似文献   

10.
11.
Structural and dynamic features of RNA folding landscapes represent critical aspects of RNA function in the cell and are particularly central to riboswitch-mediated control of gene expression. Here, using single-molecule fluorescence energy transfer imaging, we explore the folding dynamics of the preQ1 class II riboswitch, an upstream mRNA element that regulates downstream encoded modification enzymes of queuosine biosynthesis. For reasons that are not presently understood, the classical pseudoknot fold of this system harbors an extra stem–loop structure within its 3′-terminal region immediately upstream of the Shine–Dalgarno sequence that contributes to formation of the ligand-bound state. By imaging ligand-dependent preQ1 riboswitch folding from multiple structural perspectives, we reveal that the extra stem–loop strongly influences pseudoknot dynamics in a manner that decreases its propensity to spontaneously fold and increases its responsiveness to ligand binding. We conclude that the extra stem–loop sensitizes this RNA to broaden the dynamic range of the ON/OFF regulatory switch.A variety of small metabolites have been found to regulate gene expression in bacteria, fungi, and plants via direct interactions with distinct mRNA folds (14). In this form of regulation, the target mRNA typically undergoes a structural change in response to metabolite binding (59). These mRNA elements have thus been termed “riboswitches” and generally include both a metabolite-sensitive aptamer subdomain and an expression platform. For riboswitches that regulate the process of translation, the expression platform minimally consists of a ribosomal recognition site [Shine–Dalgarno (SD)]. In the simplest form, the SD sequence overlaps with the metabolite-sensitive aptamer domain at its downstream end. Representative examples include the S-adenosylmethionine class II (SAM-II) (10) and the S-adenosylhomocysteine (SAH) riboswitches (11, 12), as well as prequeuosine class I (preQ1-I) and II (preQ1-II) riboswitches (13, 14). The secondary structures of these four short RNA families contain a pseudoknot fold that is central to their gene regulation capacity. Although the SAM-II and preQ1-I riboswitches fold into classical pseudoknots (15, 16), the conformations of the SAH (17) and preQ1-II counterparts are more complex and include a structural extension that contributes to the pseudoknot architecture (14). Importantly, the impact and evolutionary significance of these “extra” stem–loop elements on the function of the SAH and preQ1-II riboswitches remain unclear.PreQ1 riboswitches interact with the bacterial metabolite 7-aminomethyl-7-deazaguanine (preQ1), a precursor molecule in the biosynthetic pathway of queuosine, a modified base encountered at the wobble position of some transfer RNAs (14). The general biological significance of studying the preQ1-II system stems from the fact that this gene-regulatory element is found almost exclusively in the Streptococcaceae bacterial family. Moreover, the preQ1 metabolite is not generated in humans and has to be acquired from the environment (14). Correspondingly, the preQ1-II riboswitch represents a putative target for antibiotic intervention. Although preQ1 class I (preQ1-I) riboswitches have been extensively investigated (1828), preQ1 class II (preQ1-II) riboswitches have been largely overlooked despite the fact that a different mode of ligand binding has been postulated (14).The consensus sequence and the secondary structure model for the preQ1-II motif (COG4708 RNA) (Fig. 1A) comprise ∼80–100 nt (14). The minimal Streptococcus pneumoniae R6 aptamer domain sequence binds preQ1 with submicromolar affinity and consists of an RNA segment forming two stem–loops, P2 and P4, and a pseudoknot P3 (Fig. 1B). In-line probing studies suggest that the putative SD box (AGGAGA; Fig. 1) is sequestered by pseudoknot formation, which results in translational-dependent gene regulation of the downstream gene (14).Open in a separate windowFig. 1.PreQ1 class II riboswitch. (A) Chemical structure of 7-aminomethyl-7-deazaguanosine (preQ1); consensus sequence and secondary structure model for the COG4708 RNA motif (adapted from reference 14). Nucleoside presence and identity as indicated. (B) S. pneumoniae R6 preQ1-II RNA aptamer investigated in this study. (C) Schematics of an H-type pseudoknot with generally used nomenclature for comparison.Here, we investigated folding and ligand recognition of the S. pneumoniae R6 preQ1-II riboswitch, using complementary chemical, biochemical, and biophysical methods including selective 2′-hydroxyl acylation analyzed by primer extension (SHAPE), mutational analysis experiments, 2-aminopurine fluorescence, and single-molecule fluorescence resonance energy transfer (smFRET) imaging. In so doing, we explored the structural and functional impact of the additional stem–loop element in the context of its otherwise “classical” H-type pseudoknot fold (2932) (Fig. 1C). Our results reveal that the unique 3′-stem–loop element in the preQ1-II riboswitch contributes to the process of SD sequestration, and thus the regulation of gene expression, by modulating both its intrinsic dynamics and its responsiveness to ligand binding.  相似文献   

12.
13.
Alphavirus envelope proteins, organized as trimers of E2–E1 heterodimers on the surface of the pathogenic alphavirus, mediate the low pH-triggered fusion of viral and endosomal membranes in human cells. The lack of specific treatment for alphaviral infections motivates our exploration of potential antiviral approaches by inhibiting one or more fusion steps in the common endocytic viral entry pathway. In this work, we performed constant pH molecular dynamics based on an atomic model of the alphavirus envelope with icosahedral symmetry. We have identified pH-sensitive residues that cause the largest shifts in thermodynamic driving forces under neutral and acidic pH conditions for various fusion steps. A series of conserved interdomain His residues is identified to be responsible for the pH-dependent conformational changes in the fusion process, and ligand binding sites in their vicinity are anticipated to be potential drug targets aimed at inhibiting viral infections.Alphaviruses, mosquito-borne human pathogens causing severe inflammations and fatal fevers, have infected many millions of people in recent outbreaks worldwide since 2005 (13). The lack of a vaccine or specific treatment prompts investigations of the fundamental mechanisms of the alphaviral lifecycle to facilitate the development of effective antiviral therapies (4). Alphaviruses have been reported to enter the cell through receptor-mediated endocytosis. Here, alphaviruses are ferried toward the perinuclear space of the host cell inside vesicles towed by molecular motors and delivered to specific locations for productive replication (511). Even when direct entry into the cytoplasm is possible (1115), the endocytic entry pathway facilitates the transportation of viruses across the crowded cytoplasmic space and delays detection by the immune system without leaving empty capsid or envelope as obvious evidence of the viral infection exposed outside the host cell (10, 11). Before the delivery of its viral genome into the cytoplasm of a host cell, the alphavirus must undergo a critical step of low pH-triggered membrane fusion, which is a common mechanism in the endocytic viral entry pathway among many different viruses. Understanding the mechanism of the low pH-triggered alphaviral membrane fusion is essential for the development of therapies against alphavirus as well as other viruses using similar endocytic entry mechanisms.Recent studies of the lifecycle of alphavirus reveal that a precursor, p62, is first synthesized as a chaperon forming a heterodimer with E1, which is essential for viral budding (16); p62 protects the E1 protein in the low-pH environment of the secretory pathway before being cleaved by cellular furin to produce mature E2-E1 and a smaller fragment, E3 (1721). After the virus buds from the cytoplasmic membrane, E3 is released from the virus particle under neutral pH conditions outside the host cell (13, 2224).On the surface of a mature alphavirus, 80 (E2–E1)3 viral spikes, organized in T = 4 icosahedral symmetry on the viral lipid membrane, enclose the viral capsid and genome (2543). On internalization of the mature virus in the endosome of the host cell in a new round of infection cycle, the increasingly acidified endosomal environment triggers a series of conformational changes in the alphaviral spike (E2–E1)3 (38), including the dissociation of E2 (42, 44, 45), release of a fusion loop on E1 (46, 47), and trimerization of E1 (48). The fusion loop, roughly residues 83–100 on the cd loop of each E1 protein (13, 49, 50), in the newly formed E1 homotrimer (HT), inserts into the endosomal membrane. Then, the E1 proteins fold back, pulling the viral and endosomal membranes together and thus, promoting membrane fusion (13, 24).Recently solved high-resolution structures of the alphavirus envelope proteins E2–E1 fitted into cryo-EM data representing the intact virus under both acidic and neutral pH conditions (43, 51, 52) provide excellent atomic models for studies of the low pH-triggered fusion process. The structure of Chikungunya virus (CHIKV) obtained at pH 8.0 represents the initial mature state (M state) of the (E2–E1)3 viral spike before the fusion process (51). Under pH 5.6, domain B (DB) of E2, which protects the E1 fusion loop, is observed to be disordered in Sindbis virus (52). The rest of the domains of the (E2–E1)3 spike show moderate conformational differences with an rmsd = 4.0 Å among Cα atoms compared with the structures obtained at pH 8.0 for CHIKV (43, 51). The structure of the envelope proteins in acidic conditions most likely depicts a fusion intermediate (FI) state (52) before E2 dissociation during the low pH-triggered fusion process. In addition, the crystal structure of the folded-back E1 HT (53) is a good model to describe the postfusion state.Based on these atomic models of the E2 and E1 envelope proteins and our previously developed constant pH molecular dynamics (CPHMD) method (5458), we simulated the envelope proteins with icosahedral symmetry under various pH conditions covering pH 2.0–9.0. We used pH replica exchange in CPHMD and calculated pKa values using pH titration fitting, which has been shown as a reliable and accurate approach to capture pKa values of protein residues in various systems (5964). Through the CPHMD modeling, we calculated the pKa of the possible pH-sensitive residues (Asp, Glu, and His) in the M, FI, dissociated E2 (Dis), and HT states. We, therefore, derive the shifts in the thermodynamic stabilities originating from each titrating residue for the steps from the M to the FI state (M→FI) of (E2–E1)3, from the FI to the Dis state (FI→Dis) of E2 proteins, and from the FI to the HT state (FI→HT) of E1 proteins as shown in Fig. 1D. For these processes, we assume that the virus is in the endosomal environment, and we do not consider possible receptor-induced conformational changes. Our residue-level resolution simulations and analyses allow us to identify the critical functional residues with significant pKa shifts and changes in thermodynamic stability in the low pH-triggered fusion activation. Our results suggest that the most pH-sensitive residues are highly conserved among different alphaviral species and that these critical residues control the pH threshold of fusion activities, provide guidance to further mutagenesis experiments, and lead to more fundamental understanding of low pH-triggered alphaviral membrane fusion.Open in a separate windowFig. 1.Structure and organization of alphaviral envelope proteins. (A) The alphaviral envelope modeled in our simulations. (B) The alphaviral envelope proteins in an MAU. (C) The heterodimer of E2 (DA–DB–DC) and E1 (DI–DII–DIII). (D) Structures of a viral spike in different conformational states simulated for shifts in pKa values and thermodynamic stabilities. E1 proteins are shown in blue, cyan, and light blue. E2 proteins are shown in red, magenta, and pink.  相似文献   

14.
15.
16.
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   

17.
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   

18.
19.
Through phase transition-induced band edge engineering by dual doping with In and Mo, a new greenish BiVO4 (Bi1-XInXV1-XMoXO4) is developed that has a larger band gap energy than the usual yellow scheelite monoclinic BiVO4 as well as a higher (more negative) conduction band than H+/H2 potential [0 VRHE (reversible hydrogen electrode) at pH 7]. Hence, it can extract H2 from pure water by visible light-driven overall water splitting without using any sacrificial reagents. The density functional theory calculation indicates that In3+/Mo6+ dual doping triggers partial phase transformation from pure monoclinic BiVO4 to a mixture of monoclinic BiVO4 and tetragonal BiVO4, which sequentially leads to unit cell volume growth, compressive lattice strain increase, conduction band edge uplift, and band gap widening.Photocatalytic water splitting with a particulate semiconductor powered by sunlight is an ideal route to cost-effective, large-scale, and sustainable hydrogen production because of its extreme simplicity. However, it is challenging, because it requires a rare photocatalyst that carries a combination of suitable band gap energy, appropriate band positions, and photochemical stability (15). Thus, reproducible photocatalytic systems for visible light-driven overall water splitting (OWS) by one-step photoexcitation are also rare, although there were several reports of such systems (46). In the best-known successful case, Domen and coworkers (5) reported in 2005 that a solid solution of GaN and ZnO [(Ga1−xZnx)(N1−xOx)] was a stable photocatalyst that could split water into H2 and O2 under visible light when modified with a cocatalyst. This system remains the most active and reproducible one-step OWS photocatalyst responsive to visible light so far (4).Scheelite monoclinic (m-) BiVO4 is a well-documented photocatalyst having suitable band gap energy (Eg ∼ 2.4 eV) for absorbing visible light (710). Also, it is chemically stable in aqueous solution under light irradiation. Thus, it functions as an excellent photocatalyst for O2 evolution under visible light in the presence of an appropriate electron acceptor (e.g., AgNO3). However, because the bottom of its conduction band is located at a more positive potential than the potential of water reduction [0 VRHE (reversible hydrogen electrode) at pH 7], it is incapable of evolving H2. In addition, it shows poor charge transport characteristics (11) and weak surface adsorption properties (12), causing low photocatalytic activity. To overcome these weaknesses, a variety of strategies, such as heterojunction structure formation (11, 13, 14), loading cocatalysts (8, 1517), and impurity doping (1, 7, 12, 1823), has been attempted. These strategies were successful in improving BiVO4’s oxidation capability for photoelectrochemical water oxidation (1, 11, 13, 15, 19, 2428) as well as the Z scheme (two-photon excitation) water splitting system (29). Also, cocatalyzed BixY1-xVO4 (x ∼ 0.5) (7, 8) and cocatalyzed Bi0.5La0.5VO4 (23) promoted OWS by raising the conduction band edge (CBE) position, but OWS under visible light irradiation over BiVO4-based photocatalysts has not been fully shown.To meet this challenge, we developed greenish BiVO4 (GBVOx; x = atom ratio of In and Mo), Bi1-xInxV1-xMoxO4, by simultaneously substituting In3+ for Bi3+ and Mo6+ for V5+ in the host lattice of m-BiVO4. The new GBVOx photocatalyst has a slightly larger band gap energy than the usual yellow scheelite m-BiVO4 as supported by the unique color change to green and a higher (more negative) conduction band than H+/H2 potential (0 VRHE at pH 7). Consequently, as depicted in Fig. 1, GBVOx is able to split water into H2 and O2 under visible light irradiation without using any sacrificial reagents (e.g., CH3OH or AgNO3). Herein, we report the dual-metal doping effects on the optical absorption behavior, crystal structure, and electronic band structure of BiVO4, which led to one-photon OWS under visible light irradiation. We elucidate the physical origin of the augmented photoresponse behaviors of GBVOx through density functional theory (DFT) calculation of electronic structure as well as a variety of physical and electrochemical characterizations.Open in a separate windowFig. 1.OWS reaction mechanism by GBVOx.  相似文献   

20.
The in vivo application of aptamers as therapeutics could be improved by enhancing target-specific accumulation while minimizing off-target uptake. We designed a light-triggered system that permits spatiotemporal regulation of aptamer activity in vitro and in vivo. Cell binding by the aptamer was prevented by hybridizing the aptamer to a photo-labile complementary oligonucleotide. Upon irradiation at the tumor site, the aptamer was liberated, leading to prolonged intratumoral retention. The relative distribution of the aptamer to the liver and kidney was also significantly decreased, compared to that of the free aptamer.Aptamers are single-stranded nucleic acids that have emerged as a promising class of therapeutics owing to their relative ease of synthesis and high affinity and selectivity toward a range of targets including small molecules, proteins, viral particles, and living cells (16). Aptamers can fold into well-defined conformations and are more resistant to enzymatic degradation than other oligonucleotides (79). Aptamers have been suggested for imaging applications because their relatively small size and molecular mass (∼10 kDa) allow fast tissue penetration and clearance from blood (10, 11). The same characteristics make aptamers promising for effective delivery of diagnostic and therapeutic agents to tissues or organs. However, nonspecific accumulation of aptamers in normal tissues is undesirable (1215) because it diminishes the proportion of aptamer that targets the desired tissue. This can adversely affect the therapeutic index of the aptamer; this may be particularly true if the aptamer is conjugated to a drug or drug delivery device. Moreover, aptamers themselves can have nonspecific toxic effects (16, 17). Ideally aptamers would achieve a high concentration in a pathological tissue of interest while maintaining low levels elsewhere. The activity of aptamers can be modulated in vivo by binding to polymers or complementary oligonucleotide sequences (1820), but spatiotemporal regulation of aptamer activity in vivo has not been achieved, whereby activity would be enhanced in target tissues and not others. Here, we report a strategy to provide light-triggered control of aptamer function and distribution in vivo.Light is an excellent means of providing external spatiotemporal control of biological systems (2126). Many strategies have been developed to incorporate photosensitive groups in nucleotides that can control cellular function or affect biological pathways or gene expression by light (2429). Of particular interest, such approaches can be used to provide spatiotemporal control of gene activation (24). Here we hypothesized that light triggering can be used to achieve spatiotemporal control of binding of an aptamer injected systemically to its target tissue in vivo, which would have implications for control of delivery of therapeutic aptamers and/or conjugated drugs or drug delivery systems. We designed a photo-triggerable system whereby the aptamer of interest is inactivated by hybridization to a photo-labile complementary oligonucleotide. Upon irradiation, the complementary sequence breaks down, releasing the functional aptamer (Fig. 1). The aptamer of interest is the single-stranded DNA 26-mer aptamer AS1411 (A1411; sequence: 5′-GGT GGT GGT GGT TGT GGT GGT GGT GG-3′) that binds with high affinity and selectivity to nucleolin (3032), which is overexpressed on the cell membrane of several types of cancer cells, including the 4T1 breast cancer cells used here (3335). A1411 has been used for cancer targeting in vitro and in vivo (35, 36). A complementary photo-triggerable inhibitory oligonucleotide (OliP) was designed [sequence: 5′-CCA CCA//CCA CCA//CAA CCA C-3′, where // indicates photo-labile 1-(2-nitrophenyl)ethyl bonds (37); Scheme S1].Open in a separate windowFig. 1.The AS1411 aptamer (A1411, red DNA strand) is hybridized to a complementary oligonucleotide (OliP, green DNA strand) containing photo-cleavable bonds (black dots). The hybridized complex cannot bind cells. The A1411 can be released by light-triggered breakage of the OliP, allowing binding to cell surfaces.  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