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 109 毫秒
1.
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 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   

4.
5.
Physiologically, α-synuclein chaperones soluble NSF attachment protein receptor (SNARE) complex assembly and may also perform other functions; pathologically, in contrast, α-synuclein misfolds into neurotoxic aggregates that mediate neurodegeneration and propagate between neurons. In neurons, α-synuclein exists in an equilibrium between cytosolic and membrane-bound states. Cytosolic α-synuclein appears to be natively unfolded, whereas membrane-bound α-synuclein adopts an α-helical conformation. Although the majority of studies showed that cytosolic α-synuclein is monomeric, it is unknown whether membrane-bound α-synuclein is also monomeric, and whether chaperoning of SNARE complex assembly by α-synuclein involves its cytosolic or membrane-bound state. Here, we show using chemical cross-linking and fluorescence resonance energy transfer (FRET) that α-synuclein multimerizes into large homomeric complexes upon membrane binding. The FRET experiments indicated that the multimers of membrane-bound α-synuclein exhibit defined intermolecular contacts, suggesting an ordered array. Moreover, we demonstrate that α-synuclein promotes SNARE complex assembly at the presynaptic plasma membrane in its multimeric membrane-bound state, but not in its monomeric cytosolic state. Our data delineate a folding pathway for α-synuclein that ranges from a monomeric, natively unfolded form in cytosol to a physiologically functional, multimeric form upon membrane binding, and show that only the latter but not the former acts as a SNARE complex chaperone at the presynaptic terminal, and may protect against neurodegeneration.α-Synuclein is an abundant presynaptic protein that physiologically acts to promote soluble NSF attachment protein receptor (SNARE) complex assembly in vitro and in vivo (13). Point mutations in α-synuclein (A30P, E46K, H50Q, G51D, and A53T) as well as α-synuclein gene duplications and triplications produce early-onset Parkinson''s disease (PD) (410). Moreover, α-synuclein is a major component of intracellular protein aggregates called Lewy bodies, which are pathological hallmarks of neurodegenerative disorders such as PD, Lewy body dementia, and multiple system atrophy (1114). Strikingly, neurotoxic α-synuclein aggregates propagate between neurons during neurodegeneration, suggesting that such α-synuclein aggregates are not only intrinsically neurotoxic but also nucleate additional fibrillization (1518).α-Synuclein is highly concentrated in presynaptic terminals where α-synuclein exists in an equilibrium between a soluble and a membrane-bound state, and is associated with synaptic vesicles (1922). The labile association of α-synuclein with membranes (23, 24) suggests that binding of α-synuclein to synaptic vesicles, and its dissociation from these vesicles, may regulate its physiological function. Membrane-bound α-synuclein assumes an α-helical conformation (2532), whereas cytosolic α-synuclein is natively unfolded and monomeric (refs. 25, 26, 31, and 32; however, see refs. 33 and 34 and Discussion for a divergent view). Membrane binding by α-synuclein is likely physiologically important because in in vitro experiments, α-synuclein remodels membranes (35, 36), influences lipid packing (37, 38), and induces vesicle clustering (39). Moreover, membranes were found to be important for the neuropathological effects of α-synuclein (4044).However, the relation of membrane binding to the in vivo function of α-synuclein remains unexplored, and it is unknown whether α-synuclein binds to membranes as a monomer or oligomer. Thus, in the present study we have investigated the nature of the membrane-bound state of α-synuclein and its relation to its physiological function in SNARE complex assembly. We found that soluble monomeric α-synuclein assembles into higher-order multimers upon membrane binding and that membrane binding of α-synuclein is required for its physiological activity in promoting SNARE complex assembly at the synapse.  相似文献   

6.
Large glycosylating toxins are major virulence factors of various species of pathogenic Clostridia. Prototypes are Clostridium difficile toxins A and B, which cause antibiotics-associated diarrhea and pseudomembranous colitis. The current model of the toxins’ action suggests that receptor binding is mediated by a C-terminal domain of combined repetitive oligopeptides (CROP). This model is challenged by the glycosylating Clostridium perfringens large cytotoxin (TpeL toxin) that is devoid of the CROP domain but still intoxicates cells. Using a haploid genetic screen, we identified LDL receptor-related protein 1 (LRP1) as a host cell receptor for the TpeL toxin. LRP1-deficient cells are not able to take up TpeL and are not intoxicated. Expression of cluster IV of LRP1 is sufficient to rescue toxin uptake in these cells. By plasmon resonance spectroscopy, a KD value of 23 nM was determined for binding of TpeL to LRP1 cluster IV. The C terminus of TpeL (residues 1335–1779) represents the receptor-binding domain (RBD) of the toxin. RBD-like regions are conserved in all other clostridial glycosylating toxins preceding their CROP domain. CROP-deficient C. difficile toxin B is toxic to cells, depending on the RBD-like region (residues 1349–1811) but does not interact with LRP1. Our data indicate the presence of a second, CROP-independent receptor-binding domain in clostridial glycosylating toxins and suggest a two-receptor model for the cellular uptake of clostridial glycosylating toxins.Clostridial glycosylating toxins are major pathogenicity factors that are responsible for numerous severe diseases in humans and animals. Prototypes of these toxins are Clostridium difficile toxins A and B, the causative agents of antibiotics-associated diarrhea and pseudomembranous colitis (1, 2). During recent years, morbidity and mortality of C. difficile-induced infections (CDI) largely increased (3, 4). In fact, CDI advanced to one of the most important nosocomial infections in developed countries. Other members of the family of clostridial glycosylating toxins are Clostridium sordellii lethal and hemorrhagic toxin, and the α-toxin from Clostridium novyi, which cause gas gangrene syndromes (5). All these toxins have a very similar primary structure comprising an amino acid sequence identity of 40–90% (1, 5). Recently, an ABCD model has been proposed for these toxins with an N-terminally located glycosyltransferase domain (domain A), a subsequent cysteine protease domain for autoproteolytic cleavage (domain C), a putative pore-forming and delivery domain (domain D), and a C-terminal binding domain (domain B) (6). After binding to cell surface receptors, the toxins are endocytosed in a clathrin-dependent and dynamin-dependent manner (7). At a low pH of endosomes, the toxins insert into endosomal membranes and form pores, which probably allow translocation of the glycosyltransferase (GT) and cysteine protease domains into the cytosol where inositol hexakisphosphate activates the protease for autoproteolytic cleavage and release of the GT into the cytosol (810). In the cytosol, Rho and/or Ras proteins are glucosylated or modified by GlcNAcylation, resulting in inhibition of these switch proteins and, eventually, in inflammation and cell death (1113). Although autoproteolytic processing and toxin-induced glycosylation of Rho/Ras proteins are well characterized, the interaction of the toxins with cell surface receptors is still enigmatic. Cell surface-binding is suggested to be mediated by C-terminal polypeptide repeats (B domain) termed combined repetitive oligopeptides (CROP) that might recognize cell surface carbohydrate structures (1417). Recombinant fragments of this C-terminal toxin part blocks toxin binding and cytotoxicity (18). Moreover, monoclonal antibodies raised against this toxin portion prevent cell intoxication (19). Putative receptors have been described for C. difficile toxin A, including carbohydrates, glycophospholipids, and proteins (16, 17, 20, 21). The hypothesis that the C-terminal part of clostridial glycosylating toxins is solely responsible for receptor interaction has been challenged. For example, after removal of the CROP domain of toxin A or toxin B, the toxins were still cytotoxic (10, 22). Thus, we and others hypothesized that a receptor binding site different from the CROP domain might be involved in toxin binding.Recently, TpeL, the most recent member of the family of clostridial glycosylating toxins that is produced by Clostridium perfringens type A, B, and C strains, was described (23). TpeL exhibits 30–40% amino acid sequence identity with other clostridial glycosylating toxins and shares with them the glycosyltransferase domain, the cysteine protease domain, and the delivery domain. However, it does not possess a CROP domain. Nevertheless, TpeL intoxicates cells and kills mice (23). Toxic effects of TpeL are probably due to GlcNAcylation of Ras proteins at threonine-35 (24). In addition, TpeL-induced modification of Rac has been reported (25).Thus, TpeL is a model toxin to unravel the interaction of clostridial glycosylating toxins with target cells. Here, we identify the low-density lipoprotein receptor-related protein 1 (LRP1) as a target molecule for binding and cell entry of TpeL. We report that the C-terminal part of TpeL binds to LRP1 and represents the receptor-binding domain. Furthermore, we show that the respective part within C. difficile toxin B, which resembles to the receptor-binding domain of TpeL, binds to cells. Our study strongly supports a two-receptor model of clostridial glycosylating toxins and offers an additional perspective in the understanding of the pathogenicity of this group of clinically important toxins.  相似文献   

7.
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   

8.
To dissect the kinetics of structural transitions underlying the stepping cycle of kinesin-1 at physiological ATP, we used interferometric scattering microscopy to track the position of gold nanoparticles attached to individual motor domains in processively stepping dimers. Labeled heads resided stably at positions 16.4 nm apart, corresponding to a microtubule-bound state, and at a previously unseen intermediate position, corresponding to a tethered state. The chemical transitions underlying these structural transitions were identified by varying nucleotide conditions and carrying out parallel stopped-flow kinetics assays. At saturating ATP, kinesin-1 spends half of each stepping cycle with one head bound, specifying a structural state for each of two rate-limiting transitions. Analysis of stepping kinetics in varying nucleotides shows that ATP binding is required to properly enter the one-head–bound state, and hydrolysis is necessary to exit it at a physiological rate. These transitions differ from the standard model in which ATP binding drives full docking of the flexible neck linker domain of the motor. Thus, this work defines a consensus sequence of mechanochemical transitions that can be used to understand functional diversity across the kinesin superfamily.Kinesin-1 is a motor protein that steps processively toward microtubule plus-ends, tracking single protofilaments and hydrolyzing one ATP molecule per step (16). Step sizes corresponding to the tubulin dimer spacing of 8.2 nm are observed when the molecule is labeled by its C-terminal tail (710) and to a two-dimer spacing of 16.4 nm when a single motor domain is labeled (4, 11, 12), consistent with the motor walking in a hand-over-hand fashion. Kinesin has served as an important model system for advancing single-molecule techniques (710) and is clinically relevant for its role in neurodegenerative diseases (13), making dissection of its step a popular ongoing target of study.Despite decades of work, many essential components of the mechanochemical cycle remain disputed, including (i) how much time kinesin-1 spends in a one-head–bound (1HB) state when stepping at physiological ATP concentrations, (ii) whether the motor waits for ATP in a 1HB or two-heads–bound (2HB) state, and (iii) whether ATP hydrolysis occurs before or after tethered head attachment (4, 11, 1420). These questions are important because they are fundamental to the mechanism by which kinesins harness nucleotide-dependent structural changes to generate mechanical force in a manner optimized for their specific cellular tasks. Addressing these questions requires characterizing a transient 1HB state in the stepping cycle in which the unattached head is located between successive binding sites on the microtubule. This 1HB intermediate is associated with the force-generating powerstroke of the motor and underlies the detachment pathway that limits motor processivity. Optical trapping (7, 19, 21, 22) and single-molecule tracking studies (4, 811) have failed to detect this 1HB state during stepping. Single-molecule fluorescence approaches have detected a 1HB intermediate at limiting ATP concentrations (11, 12, 14, 15), but apart from one study that used autocorrelation analysis to detect a 3-ms intermediate (17), the 1HB state has been undetectable at physiological ATP concentrations.Single-molecule microscopy is a powerful tool for studying the kinetics of structural changes in macromolecules (23). Tracking steps and potential substeps for kinesin-1 at saturating ATP has until now been hampered by the high stepping rates of the motor (up to 100 s−1), which necessitates high frame rates, and the small step size (8.2 nm), which necessitates high spatial precision (7). Here, we apply interferometric scattering microscopy (iSCAT), a recently established single-molecule tool with high spatiotemporal resolution (2427) to directly visualize the structural changes underlying kinesin stepping. By labeling one motor domain in a dimeric motor, we detect a 1HB intermediate state in which the tethered head resides over the bound head for half the duration of the stepping cycle at saturating ATP. We further show that at physiological stepping rates, ATP binding is required to enter this 1HB state and that ATP hydrolysis is required to exit it. This work leads to a significant revision of the sequence and kinetics of mechanochemical transitions that make up the kinesin-1 stepping cycle and provides a framework for understanding functional diversity across the kinesin superfamily.  相似文献   

9.
Hyperpolarization-activated, cyclic nucleotide-gated cation (HCN) channels are critical regulators of neuronal excitability, but less is known about their possible roles in synaptic plasticity and memory circuits. Here, we characterized the HCN gene organization, channel properties, distribution, and involvement in associative and nonassociative forms of learning in Aplysia californica. Aplysia has only one HCN gene, which codes for a channel that has many similarities to the mammalian HCN channel. The cloned acHCN gene was expressed in Xenopus oocytes, which displayed a hyperpolarization-induced inward current that was enhanced by cGMP as well as cAMP. Similarly to its homologs in other animals, acHCN is permeable to K+ and Na+ ions, and is selectively blocked by Cs+ and ZD7288. We found that acHCN is predominantly expressed in inter- and motor neurons, including LFS siphon motor neurons, and therefore tested whether HCN channels are involved in simple forms of learning of the siphon-withdrawal reflex in a semiintact preparation. ZD7288 (100 μM) significantly reduced an associative form of learning (classical conditioning) but had no effect on two nonassociative forms of learning (intermediate-term sensitization and unpaired training) or baseline responses. The HCN current is enhanced by nitric oxide (NO), which may explain the postsynaptic role of NO during conditioning. HCN current in turn enhances the NMDA-like current in the motor neurons, suggesting that HCN channels contribute to conditioning through this pathway.Hyperpolarization-activated, cyclic nucleotide-gated (HCN), cation nonselective ion channels generate hyperpolarization-activated inward currents (Ih) and thus tend to stabilize membrane potential (13). In addition, binding of cyclic nucleotides (cAMP and cGMP) to the C-terminal cyclic nucleotide binding domain (CNBD) enhances Ih and thus couples membrane excitability with intracellular signaling pathways (2, 4). HCN channels are widely important for numerous systemic functions such as hormonal regulation, heart contractility, epilepsy, pain, central pattern generation, sensory perception (415), and learning and memory (1624).However, in previous studies it has been difficult to relate the cellular effects of HCN channels directly to their behavioral effects, because of the immense complexity of the mammalian brain. We have therefore investigated the role of HCN channels in Aplysia, which has a numerically simpler nervous system (25). We first identified and characterized an HCN gene in Aplysia, and showed that it codes for a channel that has many similarities to the mammalian HCN channel. We found that the Aplysia HCN channel is predominantly expressed in motor neurons including LFS neurons in the siphon withdrawal reflex circuit (26, 27). We therefore investigated simple forms of learning of that reflex in a semiintact preparation (2830) and found that HCN current is involved in classical conditioning and enhances the NMDA-like current in the motor neurons. These results provide a direct connection between HCN channels and behavioral learning and suggest a postsynaptic mechanism of that effect. HCN current in turn is enhanced by nitric oxide (NO), a transmitter of facilitatory interneurons, and thus may contribute to the postsynaptic role of NO during conditioning.  相似文献   

10.
Embryonic stem cell-based therapies exhibit great potential for the treatment of Parkinson’s disease (PD) because they can significantly rescue PD-like behaviors. However, whether the transplanted cells themselves release dopamine in vivo remains elusive. We and others have recently induced human embryonic stem cells into primitive neural stem cells (pNSCs) that are self-renewable for massive/transplantable production and can efficiently differentiate into dopamine-like neurons (pNSC–DAn) in culture. Here, we showed that after the striatal transplantation of pNSC–DAn, (i) pNSC–DAn retained tyrosine hydroxylase expression and reduced PD-like asymmetric rotation; (ii) depolarization-evoked dopamine release and reuptake were significantly rescued in the striatum both in vitro (brain slices) and in vivo, as determined jointly by microdialysis-based HPLC and electrochemical carbon fiber electrodes; and (iii) the rescued dopamine was released directly from the grafted pNSC–DAn (and not from injured original cells). Thus, pNSC–DAn grafts release and reuptake dopamine in the striatum in vivo and alleviate PD symptoms in rats, providing proof-of-concept for human clinical translation.Parkinson’s disease (PD) is a chronic progressive neurodegenerative disorder characterized by the specific loss of dopaminergic neurons in the substantia nigra pars compacta and their projecting axons, resulting in loss of dopamine (DA) release in the striatum (1). During the last two decades, cell-replacement therapy has proven, at least experimentally, to be a potential treatment for PD patients (27) and in animal models (815). The basic principle of cell therapy is to restore the DA release by transplanting new DA-like cells. Until recently, obtaining enough transplantable cells was a major bottleneck in the practicability of cell therapy for PD. One possible source is embryonic stem cells (ESCs), which can develop infinitely into self-renewable pluripotent cells with the potential to generate any type of cell, including DA neurons (DAns) (16, 17).Recently, several groups including us have introduced rapid and efficient ways to generate primitive neural stem cells (pNSCs) from human ESCs using small-molecule inhibitors under chemically defined conditions (12, 18, 19). These cells are nonpolarized neuroepithelia and retain plasticity upon treatment with neuronal developmental morphogens. Importantly, pNSCs differentiate into DAns (pNSC–DAn) with high efficiency (∼65%) after patterning by sonic hedgehog (SHH) and fibroblast growth factor 8 (FGF8) in vitro, providing an immediate and renewable source of DAns for PD treatment. Importantly, the striatal transplantation of human ESC-derived DA-like neurons, including pNSC–DAn, are able to relieve the motor defects in a PD rat model (1113, 15, 1923). Before attempting clinical translation of pNSC–DAn, however, there are two fundamental open questions. (i) Can pNSC–DAn functionally restore the striatal DA levels in vivo? (ii) What cells release the restored DA, pNSC–DAn themselves or resident neurons/cells repaired by the transplants?Regarding question 1, a recent study using nafion-coated carbon fiber electrodes (CFEs) reported that the amperometric current is rescued in vivo by ESC (pNSC–DAn-like) therapy (19). Both norepinephrine (NE) and serotonin are present in the striatum (24, 25). However, CFE amperometry/chronoamperometry alone cannot distinguish DA from other monoamines in vivo, such as NE and serotonin (Fig. S1) (see also refs. 2628). Considering that the compounds released from grafted ESC-derived cells are unknown, the work of Kirkeby et al. was unable to determine whether DA or other monoamines are responsible for the restored amperometric signal. Thus, the key question of whether pNSC–DAn can rescue DA release needs to be reexamined for the identity of the restored amperometric signal in vivo.Regarding question 2, many studies have proposed that DA is probably released from the grafted cells (8, 12, 13, 20), whereas others have proposed that the grafted stem cells might restore striatal DA levels by rescuing injured original cells (29, 30). Thus, whether the grafted cells are actually capable of synthesizing and releasing DA in vivo must be investigated to determine the future cellular targets (residual cells versus pNSC–DAn) of treatment.To address these two mechanistic questions, advanced in vivo methods of DA identification and DA recording at high spatiotemporal resolution are required. Currently, microdialysis-based HPLC (HPLC) (3133) and CFE amperometric recordings (34, 35) have been used independently by different laboratories to assess evoked DA release from the striatum in vivo. The major advantage of microdialysis-based HPLC is to identify the substances secreted in the cell-grafted striatum (33), but its spatiotemporal resolution is too low to distinguish the DA release site (residual cells or pNSC–DAn). In contrast, the major advantage of CFE-based amperometry is its very high temporal (ms) and spatial (μm) resolution, making it possible to distinguish the DA release site (residual cells or pNSC–DAn) in cultured cells, brain slices, and in vivo (3439), but it is unable to distinguish between low-level endogenous oxidizable substances (DA versus serotonin and NE) in vivo.In the present study, we developed a challenging experimental paradigm of combining the two in vivo methods, microdialysis-based HPLC and CFE amperometry, to identify the evoked substance as DA and its release site as pNSC–DAn in the striatum of PD rats.  相似文献   

11.
The diphthamide on human eukaryotic translation elongation factor 2 (eEF2) is the target of ADP ribosylating diphtheria toxin (DT) and Pseudomonas exotoxin A (PE). This modification is synthesized by seven dipthamide biosynthesis proteins (DPH1–DPH7) and is conserved among eukaryotes and archaea. We generated MCF7 breast cancer cell line-derived DPH gene knockout (ko) cells to assess the impact of complete or partial inactivation on diphthamide synthesis and toxin sensitivity, and to address the biological consequence of diphthamide deficiency. Cells with heterozygous gene inactivation still contained predominantly diphthamide-modified eEF2 and were as sensitive to PE and DT as parent cells. Thus, DPH gene copy number reduction does not affect overall diphthamide synthesis and toxin sensitivity. Complete inactivation of DPH1, DPH2, DPH4, and DPH5 generated viable cells without diphthamide. DPH1ko, DPH2ko, and DPH4ko harbored unmodified eEF2 and DPH5ko ACP- (diphthine-precursor) modified eEF2. Loss of diphthamide prevented ADP ribosylation of eEF2, rendered cells resistant to PE and DT, but does not affect sensitivity toward other protein synthesis inhibitors, such as saporin or cycloheximide. Surprisingly, cells without diphthamide (independent of which the DPH gene compromised) were presensitized toward nuclear factor of kappa light polypeptide gene enhancer in B cells (NF-κB) and death-receptor pathways without crossing lethal thresholds. In consequence, loss of diphthamide rendered cells hypersensitive toward TNF-mediated apoptosis. This finding suggests a role of diphthamide in modulating NF-κB, death receptor, or apoptosis pathways.Eukaryotic translation elongation factor 2 (eEF2) is a highly conserved protein and essential for protein biosynthesis. EEF2 enables peptide-chain elongation by translocating the peptide–tRNA complex from the A- to the P-site of the ribosome (1, 2). The diphthamide modification at His715 of human eEF2 (or at the corresponding position in other species) is conserved in all eukaryotes (3) and in archaeal counterparts. It is generated by proteins that are encoded by seven genes (4). Proteins encoded by dipthamide biosynthesis protein (DPH)1, DPH2, DPH3, and DPH4 (DNAJC24) attach a 3-amino-3-carboxypropyl (ACP) group to eEF2. This intermediate is converted by the methyltransferase DPH5 to diphthine, which is subsequently amidated to diphthamide by DPH6 and DPH7 (5).Diphthamide synthesis was previously described in yeast and other eukaryotes (46). However, the “complete picture” is (with the exception of the yeast pathway) to a large portion is composed of observations made in different cell types on single genes. Many reports related to diphthamide synthesis of mammalian cells describe “partial knockouts” and “partial phenotypes” (i.e., reduced levels but not complete loss of diphthamide modification or toxin sensitivities) (79). Because mammalian genomes are more complex than that of yeast, carrying extendend gene families, mammalian cells may compensate—at least to some degree—functional loss of genes that may be unique and essential in yeast. If and to what degree mammalian cells can compensate a partial or complete loss of DPH gene functionality (and with what consequences) is unknown to date.So far, the function of diphthamide on eEF2 also remained rather elusive. Reports indicate that it contributes to translation fidelity (1013). On the other hand, DPH genes or eEF2 can be mutated to prevent diphthamide attachment, yet cells carrying such mutations are viable (5, 11, 14, 15). Animals with heterozygous DPH knockouts (DPHko) can be generated, but homozygous DPH1ko, DPH3ko, and DPH4ko are embryonic lethal (13, 1618). Because these studies are based on inactivation of individual genes, it is difficult to discriminate between phenotypes caused by gene loss and phenotypes as a consequence of loss of diphthamide.Diphthamide-modified eEF2 is the target of ADP ribosylating toxins, including Pseudomonas exotoxin A (PE) and diphtheria toxin (DT) (19). These bacterial proteins enter cells and catalyze ADP ribosylation of diphthamide using nictotinamide adenine dinucleotide (NAD) as substrate (20, 21). This inactivates eEF2, arrests protein synthesis, and kills (14). Tumor-targeted PE and DT derivatives are applied in cancer therapies (2228) and their efficacy depends on toxin sensitivity of target cells. Therefore, information about factors (and their relative contributions) that influences cellular sensitivities toward diphthamide-modifying toxins may predict therapy responses. For example, alterations in OVCA1 (human DPH1) were described for ovarian cancers (16, 29), yet it is not known if and to what degree such alterations would affect sensitivities of tumor cells toward PE-derived drugs.Here we describe MCF7 breast cancer cell line derivatives with heterozygous or complete DPH gene inactivations. These cells are applied to analyze the contributions of individual DPHs not only to diphthamide synthesis and toxin sensitivity, but also to address gene dose effects. Because the set of knockout cell lines is derived from the same parent cell and provides loss of diphthamide as common consequence of inactivation of different genes, these cells can also shed light on the biological relevance of the diphthamide modification.  相似文献   

12.
Worldwide dissemination of antibiotic resistance in bacteria is facilitated by plasmids that encode postsegregational killing (PSK) systems. These produce a stable toxin (T) and a labile antitoxin (A) conditioning cell survival to plasmid maintenance, because only this ensures neutralization of toxicity. Shortage of antibiotic alternatives and the link of TA pairs to PSK have stimulated the opinion that premature toxin activation could be used to kill these recalcitrant organisms in the clinic. However, validation of TA pairs as therapeutic targets requires unambiguous understanding of their mode of action, consequences for cell viability, and function in plasmids. Conflicting with widespread notions concerning these issues, we had proposed that the TA pair kis-kid (killing suppressor-killing determinant) might function as a plasmid rescue system and not as a PSK system, but this remained to be validated. Here, we aimed to clarify unsettled mechanistic aspects of Kid activation, and of the effects of this for kis-kid–bearing plasmids and their host cells. We confirm that activation of Kid occurs in cells that are about to lose the toxin-encoding plasmid, and we show that this provokes highly selective restriction of protein outputs that inhibits cell division temporarily, avoiding plasmid loss, and stimulates DNA replication, promoting plasmid rescue. Kis and Kid are conserved in plasmids encoding multiple antibiotic resistance genes, including extended spectrum β-lactamases, for which therapeutic options are scarce, and our findings advise against the activation of this TA pair to fight pathogens carrying these extrachromosomal DNAs.Plasmids serve as extrachromosomal DNA platforms for the reassortment, mobilization, and maintenance of antibiotic resistance genes in bacteria, enabling host cells to colonize environments flooded with antimicrobials and to take advantage of resources freed by the extinction of nonresistant competitors. Fueled by these selective forces and aided by their itinerant nature, plasmids disseminate resistance genes worldwide shortly after new antibiotics are developed, which is a major clinical concern (13). However, in antibiotic-free environments, such genes are dispensable. There, the cost that plasmid carriage imposes on cells constitutes a disadvantage in the face of competition from other cells and, because plasmids depend on their hosts to survive, also a threat to their own existence.Many plasmids keep low copy numbers (CNs) to minimize the problem above, because it reduces burdens to host cells. However, this also decreases their chances to fix in descendant cells, a new survival challenge (4). To counteract this, plasmids have evolved stability functions. Partition systems pull replicated plasmid copies to opposite poles in host cells, facilitating their inheritance by daughter cells (5). Plasmids also bear postsegregational killing (PSK) systems, which encode a stable toxin and a labile antitoxin (TA) pair that eliminates plasmid-free cells produced by occasional replication or partition failures. Regular production of the labile antitoxin protects plasmid-containing cells from the toxin. However, antitoxin replenishment is not possible in cells losing the plasmid, and this triggers their elimination (5).TA pairs are common in plasmids disseminating antibiotic resistance in bacterial pathogens worldwide (2, 610). The link of these systems to PSK and the exiguous list of alternatives in the pipeline have led some to propose that chemicals activating these TA pairs may constitute a powerful antibiotic approach against these organisms (5, 1113). However, the appropriateness of these TA pairs as therapeutic targets requires unequivocal understanding of their function in plasmids. Although PSK systems encode TA pairs, not all TA pairs might function as PSK systems, as suggested by their abundance in bacterial chromosomes, where PSK seems unnecessary (1416). Moreover, the observation that many plasmids bear several TA pairs (610) raises the intriguing question of why they would need more than one PSK system, particularly when they increase the metabolic burden that plasmids impose on host cells (17). Because PSK functions are not infallible, their gathering may provide a mechanism for reciprocal failure compensation, minimizing the number of cells that escape killing upon plasmid loss (5). Alternatively, some TA pairs may stabilize plasmids by mechanisms different from PSK, and their grouping might not necessarily reflect functional redundancy (18).This may be the case in plasmid R1, which encodes TA pairs hok-sok (host killing-suppressor of killing) and kis(pemI)-kid(pemK) (1923). Inconsistent with PSK, we had noticed that activation of toxin Kid occurred in cells that still contained R1, and that this happened when CNs were insufficient to ensure plasmid transmission to descendant cells. We also found that Kid cleaved mRNA at UUACU sites, which appeared well suited to trigger a response that prevented plasmid loss and increased R1 CNs without killing cells, as suggested by our results. In view of all this, we argued that Kid and Kis functioned as a rescue system for plasmid R1, and not as a PSK system (24). This proposal cannot be supported by results elsewhere, suggesting that Kid may cleave mRNA at simpler UAH sites (with H being A, C, or U) (25, 26), a view that has prevailed in the literature (14, 16, 2729). Moreover, other observations indicate that our past experiments may have been inappropriate to conclude that Kid does not kill Escherichia coli cells (30, 31). Importantly, Kid, Kis, and other elements that we found essential for R1 rescue are conserved in plasmids conferring resistance to extended-spectrum β-lactamases, a worrying threat to human health (1, 610, 32). Therapeutic options to fight pathogens carrying these plasmids are limited, and activation of Kid may be perceived as a good antibiotic alternative. Because the potential involvement of this toxin in plasmid rescue advises against such approach, we aimed to ascertain here the mode of action; the effects on cells; and, ultimately, the function of Kid (and Kis) in R1.  相似文献   

13.
In humans, spontaneous movements are often preceded by early brain signals. One such signal is the readiness potential (RP) that gradually arises within the last second preceding a movement. An important question is whether people are able to cancel movements after the elicitation of such RPs, and if so until which point in time. Here, subjects played a game where they tried to press a button to earn points in a challenge with a brain–computer interface (BCI) that had been trained to detect their RPs in real time and to emit stop signals. Our data suggest that subjects can still veto a movement even after the onset of the RP. Cancellation of movements was possible if stop signals occurred earlier than 200 ms before movement onset, thus constituting a point of no return.It has been repeatedly shown that spontaneous movements are preceded by early brain signals (18). As early as a second before a simple voluntary movement, a so-called readiness potential (RP) is observed over motor-related brain regions (13, 5). The RP was found to precede the self-reported time of the “‘decision’ to act” (ref. 3, p. 623). Similar preparatory signals have been observed using invasive electrophysiology (8, 9) and functional MRI (7, 10), and have been demonstrated also for choices between multiple-response options (6, 7, 10), for abstract decisions (10), for perceptual choices (11), and for value-based decisions (12). To date, the exact nature and causal role of such early signals in decision making is debated (1220).One important question is whether a person can still exert a veto by inhibiting the movement after onset of the RP (13, 18, 21, 22). One possibility is that the onset of the RP triggers a causal chain of events that unfolds in time and cannot be cancelled. The onset of the RP in this case would be akin to tipping the first stone in a row of dominoes. If there is no chance of intervening, the dominoes will gradually fall one-by-one until the last one is reached. This has been coined a ballistic stage of processing (23, 24). A different possibility is that participants can still terminate the process, akin to taking out a domino at some later stage in the chain and thus preventing the process from completing. Here, we directly tested this in a real-time experiment that required subjects to terminate their decision to move once a RP had been detected by a brain–computer interface (BCI) (2531).  相似文献   

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

15.
In flowering plants, pollen tubes are guided into ovules by multiple attractants from female gametophytes to release paired sperm cells for double fertilization. It has been well-established that Ca2+ gradients in the pollen tube tips are essential for pollen tube guidance and that plasma membrane Ca2+ channels in pollen tube tips are core components that regulate Ca2+ gradients by mediating and regulating external Ca2+ influx. Therefore, Ca2+ channels are the core components for pollen tube guidance. However, there is still no genetic evidence for the identification of the putative Ca2+ channels essential for pollen tube guidance. Here, we report that the point mutations R491Q or R578K in cyclic nucleotide-gated channel 18 (CNGC18) resulted in abnormal Ca2+ gradients and strong pollen tube guidance defects by impairing the activation of CNGC18 in Arabidopsis. The pollen tube guidance defects of cngc18-17 (R491Q) and of the transfer DNA (T-DNA) insertion mutant cngc18-1 (+/−) were completely rescued by CNGC18. Furthermore, domain-swapping experiments showed that CNGC18’s transmembrane domains are indispensable for pollen tube guidance. Additionally, we found that, among eight Ca2+ channels (including six CNGCs and two glutamate receptor-like channels), CNGC18 was the only one essential for pollen tube guidance. Thus, CNGC18 is the long-sought essential Ca2+ channel for pollen tube guidance in Arabidopsis.Pollen tubes deliver paired sperm cells into ovules for double fertilization, and signaling communication between pollen tubes and female reproductive tissues is required to ensure the delivery of sperm cells into the ovules (1). Pollen tube guidance is governed by both female sporophytic and gametophytic tissues (2, 3) and can be separated into two categories: preovular guidance and ovular guidance (1). For preovular guidance, diverse signaling molecules from female sporophytic tissues have been identified, including the transmitting tissue-specific (TTS) glycoprotein in tobacco (4), γ-amino butyric acid (GABA) in Arabidopsis (5), and chemocyanin and the lipid transfer protein SCA in Lilium longiflorum (6, 7). For ovular pollen tube guidance, female gametophytes secrete small peptides as attractants, including LUREs in Torenia fournieri (8) and Arabidopsis (9) and ZmEA1 in maize (10, 11). Synergid cells, central cells, egg cells, and egg apparatus are all involved in pollen tube guidance, probably by secreting different attractants (915). Additionally, nitric oxide (NO) and phytosulfokine peptides have also been implicated in both preovular and ovular pollen tube guidance (1618). Thus, pollen tubes could be guided by diverse attractants in a single plant species.Ca2+ gradients at pollen tube tips are essential for both tip growth and pollen tube guidance (1927). Spatial modification of the Ca2+ gradients leads to the reorientation of pollen tube growth in vitro (28, 29). The Ca2+ gradients were significantly increased in pollen tubes attracted to the micropyles by synergid cells in vivo, compared with those not attracted by ovules (30). Therefore, the Ca2+ gradients in pollen tube tips are essential for pollen tube guidance. The Ca2+ gradients result from external Ca2+ influx, which is mainly mediated by plasma membrane Ca2+ channels in pollen tube tips. Thus, the Ca2+ channels are the key components for regulating the Ca2+ gradients and are consequently essential for pollen tube guidance. Using electrophysiological techniques, inward Ca2+ currents were observed in both pollen grain and pollen tube protoplasts (3136), supporting the presence of plasma membrane Ca2+ channels in pollen tube tips. Recently, a number of candidate Ca2+ channels were identified in pollen tubes, including six cyclic nucleotide-gated channels (CNGCs) and two glutamate receptor-like channels (GLRs) in Arabidopsis (3740). Three of these eight channels, namely CNGC18, GLR1.2, and GLR3.7, were characterized as Ca2+-permeable channels (40, 41) whereas the ion selectivity of the other five CNGCs has not been characterized. We hypothesized that the Ca2+ channel essential for pollen tube guidance could be among these eight channels.In this research, we first characterized the remaining five CNGCs as Ca2+ channels. We further found that CNGC18, out of the eight Ca2+ channels, was the only one essential for pollen tube guidance in Arabidopsis and that its transmembrane domains were indispensable for pollen tube guidance.  相似文献   

16.
17.
KCNE1 (E1) β-subunits assemble with KCNQ1 (Q1) voltage-gated K+ channel α-subunits to form IKslow (IKs) channels in the heart and ear. The number of E1 subunits in IKs channels has been an issue of ongoing debate. Here, we use single-molecule spectroscopy to demonstrate that surface IKs channels with human subunits contain two E1 and four Q1 subunits. This stoichiometry does not vary. Thus, IKs channels in cells with elevated levels of E1 carry no more than two E1 subunits. Cells with low levels of E1 produce IKs channels with two E1 subunits and Q1 channels with no E1 subunits—channels with one E1 do not appear to form or are restricted from surface expression. The plethora of models of cardiac function, transgenic animals, and drug screens based on variable E1 stoichiometry do not reflect physiology.Voltage-gated potassium (KV) channels include four α-subunits that form a single, central ion conduction pathway with four peripheral voltage sensors (13). Incorporation of accessory β-subunits modifies the function of KV channels to suit the diverse requirements of different tissues. KCNE genes encode minK-related peptides (MiRPs) (46), β-subunits with a single transmembrane span that assemble with a wide array of KV α-subunits (7, 8) to control surface expression, voltage dependence, and kinetics of gating transitions, unitary conductance, ion selectivity, and pharmacology of the resultant channel complexes (4, 915). IKslow (IKs) channels in the heart and inner ear are formed by the α-subunit encoded by KCNQ1 (called Q1, KVLQT1, KV7.1, or KCNQ1) and the β-subunit encoded by KCNE1 (called E1, mink, or KCNE1) (16, 17). Inherited mutations in Q1 and E1 are associated with cardiac arrhythmia and deafness.The number of E1 subunits in IKs channels has been a longstanding matter of disagreement. We first argued for two E1 subunits per channel based on the suppression of current by an E1 mutant (18). Subsequently, we reached the same conclusion by determining the total number of channels using radiolabeled charybdotoxin (CTX), a scorpion toxin that blocks channels when one molecule binds in the external conduction pore vestibule, and an antibody-based luminescence assay to tally E1 subunits (19). Morin and Kobertz (20) used iterative chemical linkage between CTX in the pore and E1, and they also assigned two accessory subunits to >95% of IKs channels without gathering evidence for variation in subunit valence. Furthermore, when we formed IKs channels from separate E1 and Q1 subunits and compared them with channels enforced via genetic encoding to contain two or four E1 subunits (19), we observed the natural IKs channels to have the same gating attributes, small-molecule pharmacology, and CTX on and off rates (a reflection of pore vestibule structure) as channels encoded with two E1 subunits but not those with four. These findings support the conclusion that two E1 subunits are necessary, sufficient, and the normal number in IKs channels.In contrast, others have argued that IKs channels have variable stoichiometry with one to four E1 subunits, or even more (2124). Recently, Nakajo et al. (25) applied single-particle spectroscopy to the question; this powerful “gold-standard” tool has been a valuable strategy to assess the subunit composition of ion channels (2628) and should be expected to improve on prior investigations conducted on populations of IKs channels and subject, therefore, to the simplifying assumptions that attend macroscopic studies (29). Nakajo et al. (25) reported a variable number of E1 subunits, from one to four, in IKs channels studied in Xenopus laevis oocytes. The impact of this result has been striking because it has engendered new models of cardiac physiology, altered models of IKs channel biosynthesis and function, stimulated the use of transgenic animals artificially enforced to express IKs channels with four E1 subunits (by expression of a fused E1–Q1 subunit), and prompted cardiac drug design based on the assumption that IKs channels can form with one E1 subunit (23, 3032).We were concerned that the conclusions of Nakajo et al. (25) were in error because they appraised only a limited fraction of particles that were immobile in the oocyte membrane; counted E1 and Q1 asynchronously rather than simultaneously (increasing the risk that particles moved into or out of the field of view); and studied Q1 and E1 appended not only with the fluorescent proteins (FP) required to count subunits by photobleaching but also with a common trafficking motif that suppressed channel mobility by interacting with an overexpressed anchoring protein, thereby risking nonnatural aggregation of subunits.Here, to resolve mobility problems and obviate the need for modification of subunits with targeting motifs, we describe and perform single-fluorescent-particle photobleaching at the surface of live mammalian cells, demonstrating three spectroscopic counting approaches: standard, asynchronous subunit counting; simultaneous, two-color subunit counting; and toxin-directed, simultaneous, two-color photobleaching. To analyze the data, we use two statistical approaches—one to assess the degree of colocalization of objects in dual-color images (33) and the other to infer stoichiometry from single-molecule photobleaching (34). These methods also allow determination of the surface density of assemblies of defined subunit composition and are therefore useful to assess the formation and life cycle of membrane protein complexes.We report that single IKs channels at the surface of mammalian cells contain two E1 subunits—no more and no less. This finding refutes the single-particle studies of Nakajo et al. (25) in oocytes and macroscopic studies (2124, 3032), arguing that forcing cells to express excess E1 produces IKs channels containing more than two E1 subunits and that low levels of E1 yields IKs channels with less than two E1 subunits. Not once did we observe an IKs channel with three or four E1 subunits. Moreover, simultaneous, two-color subunit counting revealed that low amounts of E1 relative to Q1 [ratios like those reported in human cardiac ventricle (35, 36)] produced two types of channels on the cell surface: IKs channels (with two E1 subunits) and Q1 channels (with no E1 subunits). Finally, E1 was shown to increase in IKs channel surface expression threefold, as we predicted based on assessment of IKs channel unitary conductance (11), whereas few E1 subunits were on the surface outside of IKs channels, even when E1 was expressed alone. This finding indicates that E1 does not travel to the surface readily on its own, that two E1 subunits facilitate IKs channel trafficking to the surface (or enhance surface residence time compared with Q1 channels), and that IKs channels with only one E1 subunit do not form, do not reach the surface, or are rapidly recycled.  相似文献   

18.
A liquid–liquid transition (LLT) in a single-component substance is an unconventional phase transition from one liquid to another. LLT has recently attracted considerable attention because of its fundamental importance in our understanding of the liquid state. To access the order parameter governing LLT from a microscopic viewpoint, here we follow the structural evolution during the LLT of an organic molecular liquid, triphenyl phosphite (TPP), by time-resolved small- and wide-angle X-ray scattering measurements. We find that locally favored clusters, whose characteristic size is a few nanometers, are spontaneously formed and their number density monotonically increases during LLT. This strongly suggests that the order parameter of LLT is the number density of locally favored structures and of nonconserved nature. We also show that the locally favored structures are distinct from the crystal structure and these two types of orderings compete with each other. Thus, our study not only experimentally identifies the structural order parameter governing LLT, but also may settle a long-standing debate on the nature of the transition in TPP, i.e., whether the transition is LLT or merely microcrystal formation.Liquid-liquid transition (LLT) is an intriguing phenomenon in which a liquid transforms into another one via a first-order transition. This means that there can be more than two liquid states for a single-component substance. Despite its counterintuitive nature, there have recently been many pieces of experimental and numerical evidence for the existence of LLT, for various liquids such as water (15), aqueous solutions (68), triphenyl phosphite (912), l-butanol (13), phosphorus (14), silicon (15, 16), germanium (17), and Y2O3–Al2O3 (18, 19). This suggests that the LLT may be rather universally observed for various types of liquids. However, none of the LLTs reported so far is free from criticisms (20, 21), mainly because these LLTs take place under experimentally difficult conditions [e.g., at high temperature and pressure (14, 15, 1719)] or in a supercooled state below the melting point (13, 57, 9, 10), where the transition is inevitably contaminated by microcrystal formation. The latter is not limited to experiments but arises in numerical simulations, often causing many controversies [LLT (2225) vs. crystallization (2628)]. For ST2 water, however, this issue has recently been settled by an extensive simulation study by Palmer et al. (4).One of the hottest and long-standing debates is on the nature of the transition found in a molecular liquid, triphenyl phosphite (TPP), by Kivelson and his coworkers (29). The transition is very easy to access experimentally, because it takes place at ambient pressure and at a temperature range between 230 and 210 K and the transformation speed is slow enough to follow the kinetics. Since the finding of this transition (29, 30), many researchers thus have been interested in this intriguing phenomenon and there have been hot discussions on the nature of the transition (20, 21). Some people interpreted this as a liquid-associated phenomenon (9, 10, 31, 32), but others interpret it differently. All of the controversies come from the fact that this transition accompanies microcrystal formation and thus the final state, which is called “glacial phase,” often contains microcrystallites. This led many researchers to explain the transition by non-LLT scenarios, which include a defect-ordered phase scenario predicted by a frustration limited domain theory (29, 30, 33, 34), a microcrystallization scenario (3538), and a liquid-crystal or plastic-crystal phase scenario (39). Each scenario captures a certain feature of the glacial phase, but fails in explaining all of the experimental results in a consistent manner. Similar situations are often seen in other candidates of LLTs, such as l-butanol [LLT (13) vs. microcrystallization (4043)], confined water [LLT (5) vs. other phenomena (4446)], and aqueous solutions [LLT (6, 7) vs. microcrystallization (8, 28, 47, 48)]. For TPP, however, some pieces of experimental evidence supportive of the LLT scenario rather than the microcrystallization scenario have recently been reported (11, 12).We propose a two-order-parameter (TOP) model of a liquid to explain LLT (20, 49). The main point of this model is that it is necessary to consider the spatiotemporal hierarchical nature of a liquid to understand LLT. More specifically, we argue that in addition to density order parameter ρ describing a gas–liquid transition, we need an additional scalar order parameter S, which is the number density of locally favored structures (LFS). In this model, LLT is a consequence of the cooperative ordering of the scalar nonconserved order parameter S, i.e., the cooperative formation of LFS. In other words, LLT is regarded as a gas–liquid-like transition of LFS: one liquid is a gas state of LFS (low-S state), and the other is its liquid state (high-S state). Recently, it was proposed by Anisimov and coworkers (50, 51) that the thermodynamic ordering field conjugate to the order parameter is the conversion equilibrium constant, which further characterizes the nature of LLT. We explained our experimental observation of LLT in TPP in terms of this model (9, 10). We also studied the phase transition dynamics and the physical and chemical properties of the second liquid state (liquid II), which were also explained by the model (20, 21).However, we have not had any direct experimental evidence for the formation of such LFS up to now; thus, an open question is, what is the relevant order parameter governing LLT, although the link of the order parameter to the enthalpy (9, 10), the refractive index (or, density) (9, 10, 29, 30), and the polarity associated with local molecular ordering (12) has been suggested for LLT in TPP. There have been structural studies on LLT by X-ray and neutron scattering measurements, focusing on local liquid structures at an inter- and intramolecular scale (36, 38, 5254) and mesoscopic structures (34, 55). However, there has been no experimental evidence for the presence of locally favored structures, which characterize the liquid state uniquely, or the order parameter has still not been identified from a microscopic viewpoint.Here we study the structural change of TPP during LLT by time-resolved small- and wide-angle X-ray scattering measurements, which cover a length scale from a single molecule size ( ~  1 nm) to more than tens of nanometers. We show, to our knowledge, the first direct evidence for the presence of LFS and the temporal increase upon the liquid I-to-liquid II transformation. Furthermore, we also find an indication of the formation of microcrystallites during LLT. However, we reveal that LFS and microcrystallites have different sizes and growth kinetics, indicating that although they sometimes appear simultaneously during the process of LLT, LLT itself is driven by the formation of LFS and not by that of microcrystallites. We also discover that LFS are destroyed upon crystallization, clearly indicating not only that these two types of orderings are competing with each other but also that LFS is a structure unique to the liquid state. Our findings provide a comprehensive view on the long-standing controversy on the origin of the glacial phase, which was discovered by Kivelson and his coworkers (29, 30), and show that the fraction of LFS may be the relevant order parameter of LLT. This suggests that a liquid can have a spatiotemporal hierarchical structure at a low temperature, contrary to the common picture of a high-temperature liquid where the structure is random and homogeneous beyond the molecular size.  相似文献   

19.
20.
The surface of a living cell provides a platform for receptor signaling, protein sorting, transport, and endocytosis, whose regulation requires the local control of membrane organization. Previous work has revealed a role for dynamic actomyosin in membrane protein and lipid organization, suggesting that the cell surface behaves as an active composite composed of a fluid bilayer and a thin film of active actomyosin. We reconstitute an analogous system in vitro that consists of a fluid lipid bilayer coupled via membrane-associated actin-binding proteins to dynamic actin filaments and myosin motors. Upon complete consumption of ATP, this system settles into distinct phases of actin organization, namely bundled filaments, linked apolar asters, and a lattice of polar asters. These depend on actin concentration, filament length, and actin/myosin ratio. During formation of the polar aster phase, advection of the self-organizing actomyosin network drives transient clustering of actin-associated membrane components. Regeneration of ATP supports a constitutively remodeling actomyosin state, which in turn drives active fluctuations of coupled membrane components, resembling those observed at the cell surface. In a multicomponent membrane bilayer, this remodeling actomyosin layer contributes to changes in the extent and dynamics of phase-segregating domains. These results show how local membrane composition can be driven by active processes arising from actomyosin, highlighting the fundamental basis of the active composite model of the cell surface, and indicate its relevance to the study of membrane organization.The cell surface mediates interactions between the cell and the outside world by serving as the site for signal transduction. It also facilitates the uptake and release of cargo and supports adhesion to substrates. These diverse roles require that the cell surface components involved in each function are spatially and temporally organized into domains spanning a few nanometers (nanoclusters) to several micrometers (microdomains). The cell surface itself may be considered as a fluid–lipid bilayer wherein proteins are embedded (1). In the living cell, this multicomponent system is supported by an actin cortex, composed of a branched network of actin and a collection of filaments (24).Current models of membrane organization fall into three categories: those invoking lipid–lipid and lipid–protein interactions in the plasma membrane [e.g., the fluid mosaic model (1, 5) and the lipid raft hypothesis (6)], or those that appeal to the membrane-associated actin cortex (e.g., the picket fence model) (7), or a combination of these (8, 9). Although these models based on thermodynamic equilibrium principles have successfully explained the organization and dynamics of a range of membrane components and molecules, there is a growing class of phenomena that appears inconsistent with chemical and thermal equilibrium, which might warrant a different explanation. These include aspects of the organization and dynamics of outer leaflet glycosyl-phosphatidylinositol-anchored proteins (GPI-anchored proteins) (1013), inner leaflet Ras proteins (14), and actin-binding transmembrane proteins (13, 15, 16).Recent experimental and theoretical work has shown that these features can be explained by taking into account that many cortical and membrane proteins are driven by ATP-consuming processes that drive the system out of equilibrium (13, 15, 17). The membrane models mentioned above have by-and-large neglected this active nature of the actin cortex where actin filaments are being continuously polymerized and depolymerized (1821), in addition to being persistently acted upon by a variety of myosin motors (2224) that consume ATP and exert contractile stresses on cortical actin filaments, continually remodeling the architecture of the cortex (4, 21, 25). These active processes in turn can generate tangential stresses and currents on the cell surface, which could drive the dynamics and local composition of membrane components at different scales (22, 2629).Actin polymerization is proposed to be driven at the membrane by two nucleators, the Arp2/3 complex, which creates a densely branched network, as well as formins that nucleate filaments (18, 21, 30). A number of myosin motors are also associated with the juxtamembranous actin cortex, of which nonmuscle myosin II is the major component in remodeling the cortex and creating actin flows (4, 23, 25, 26, 31, 32). Based on our observations that the clustering of cell surface components that couple directly or indirectly to cortical actin [e.g., GPI-anchored proteins, proteins of the Ezrin, Radaxin, or Moesin (ERM) family (13, 15)] depends on myosin activity, we proposed that this clustering arises from the coupling to contractile actomyosin platforms (called “actin asters”) produced at the cortex (15, 33).A coarse-grained theory describing this idea has been put forward and corroborated by the verification of its key predictions in live cells (15, 33), but a systematic identification of the underlying microscopic processes is lacking. Given the complexity of numerous processes acting at the membrane of a living cell, we use an in vitro approach to study the effect of an energy-consuming actomyosin network on the dynamics of membrane molecules that directly interact with filamentous actin.A series of in vitro studies have explored the organization of confined, dynamic filaments (both actin and microtubules) (3439) or the role of actin architecture on membrane organization (4046). Indeed, these studies have yielded insights into the nontrivial emergent configurations that mixtures of polar filaments and motors can adopt when fueled by ATP (3437), in particular constitutively remodeling steady states that display characteristics of active mechanics (38, 39, 47). However, the effect of linking these mechanics to the confining lipid bilayer and its organization has not been studied.The consequences of actin polymerization on membrane organization, in particular on giant unilamellar vesicles (GUVs), have been addressed in a number of studies on the propulsion of GUVs by an actin comet tail (40, 45, 46). In those experiments, the apparent advection of membrane bound ActA or WASP toward the site of actin polymerization is mainly due to the change in binding affinity of WASP to actin through Arp2/3 (44) and the spherical geometry resulting in the drag of actin to one pole of the vesicle after symmetry break of the actin shell. That this dynamic process changes the bulk properties of the bilayer, namely the critical temperature of a phase-separating lipid bilayer, was shown by Liu and Fletcher (40) when the actin nucleator N-WASP was connected to a lipid species (PIP2) that was capable of partitioning into one of the two phases.Besides these pioneering studies on the effects of active processes on membrane organization, little was done to directly test the effect of active lateral stresses as well as actomyosin remodeling at the membrane, particularly on the dynamics and organization of membrane-associated components.To this end, we build an active composite in vitro by stepwise addition of components: a supported lipid bilayer with an actin-binding component, actin filaments, and myosin motors. By systematically varying the concentrations of actin and myosin as well as the average actin filament length, we find distinct states of actomyosin organization at the membrane surface upon complete ATP consumption. More importantly, we find that the ATP-fueled contractile actomyosin currents induce the transient accumulation of actin-binding membrane components. As predicted, the active mechanics of actin and myosin at physiologically relevant ATP concentrations drives the system into a nonequilibrium steady state with anomalous density fluctuations and the transient clustering of actin-binding components of the lipid bilayer (15, 33). Finally, connection of this active layer of actomyosin to a phase-segregating bilayer, influences its phase behavior and coarsening dynamics.  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