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.
2.
3.
At our body surface, the epidermis absorbs UV radiation. UV overexposure leads to sunburn with tissue injury and pain. To understand how, we focus on TRPV4, a nonselective cation channel highly expressed in epithelial skin cells and known to function in sensory transduction, a property shared with other transient receptor potential channels. We show that following UVB exposure mice with induced Trpv4 deletions, specifically in keratinocytes, are less sensitive to noxious thermal and mechanical stimuli than control animals. Exploring the mechanism, we find that epidermal TRPV4 orchestrates UVB-evoked skin tissue damage and increased expression of the proalgesic/algogenic mediator endothelin-1. In culture, UVB causes a direct, TRPV4-dependent Ca2+ response in keratinocytes. In mice, topical treatment with a TRPV4-selective inhibitor decreases UVB-evoked pain behavior, epidermal tissue damage, and endothelin-1 expression. In humans, sunburn enhances epidermal expression of TRPV4 and endothelin-1, underscoring the potential of keratinocyte-derived TRPV4 as a therapeutic target for UVB-induced sunburn, in particular pain.The surface epithelium (epidermis) of skin provides barrier protection against dehydration and the potentially harmful external environment (1). Accordingly, skin is the site of first interaction between ambient environment and immunologically competent organismal structures, and also the site for sentient responses (2). Sensory neurons in the dorsal root ganglia (DRG) and trigeminal ganglia (TG) are endowed with sensory transduction capacity for heat, cold, mechanical cues, itch, and pain, and their axons directly interface with skin epithelium (24).Against a background of suggestive findings (2, 57), we wondered whether the epidermis as a “forefront” of sensory signaling may function in sensitizing pain transduction in response to naturally occurring irritating cues. To elucidate mechanisms, we used a mouse sunburn model and induced a state of lowered sensory thresholds associated with tissue injury caused by UV radiation (810). UV-sunburn-evoked lowering of sensory thresholds shares major hallmarks of pathological pain, a valuable feature of this model. Skin tissue injury caused by UVB has been elucidated to be mediated by cytokines and chemokines, known from immunological responses, such as IL-1β and IL-6, which are also known to cause and facilitate pain (1119). Another more recent study identified a proinflammatory chemokine, CXCL5, as proalgesic in response to UVB overexposure of rat and human skin (20). An exciting new arena pertaining to molecular mechanisms of the skin’s response to noxious UV was recently opened by an elegant study that reported the role of UVB-mediated damage to noncoding RNA molecules in the skin (21). Unraveling a molecular mechanism, the Toll-like receptor 3 gene was found critical in signaling the proinflammatory actions of the UVB-damaged noncoding RNA molecules. However, this study focused on molecular mechanisms of acute inflammation in the skin.We intended to identify pain mechanisms that mediate the pain associated with UVB-mediated tissue injury. Pain in response to external environmental cues has been understood better because of scientific progress in the field of transient receptor potential (TRP) ion channels that have been found responsive to such cues, and which were found expressed in DRG and TG peripheral sensory neurons, which are the cells believed to be the primary transducers. Indeed, TRPV1, one of the founding members of the TRPV channel subfamily, has been identified as relevant for pain, including pathological pain, response to thermal cues, and most recently for itch (2231). However, TRPA1 (transient receptor potential ion channel, ankyrin subfamily, family member #1) and TRPM8 seem to be involved in transduction of pain-inducing stimuli as well (3236).Also a family member of the TRPV subfamily, TRPV4 is a multimodally activated, nonselective cation channel that is involved in physiological pain evoked by osmotic and mechanical, but not thermal, cues (3740). For pathological pain, it is relevant for inflammation- and nerve-damage-induced pain sensitization (4143). Of note, Trpv4−/− mice exhibit impaired skin-barrier function (44, 45). That said, TRPV4 is expressed in a number of different cell types, including robust expression in epidermal keratinocytes and also is detectable in skin-innervating sensory neurons. This “dual-location expression” of TRPV4 leaves the cellular mechanisms involved in the channel’s function and the functional contribution of environment-exposed keratinocytes vs. skin-innervating sensory neurons unclear.Against this background of dual-location TRPV4 expression and the role of TRPV4 in inflammatory and neuropathic pain, we now address whether epidermally derived TRPV4 is pathophysiologically relevant in sunburn pain and tissue damage. Using Trpv4 gene-targeted mice, selectively inducing targeting in postnatal keratinocytes, and topically applying selective TRPV4 inhibitors, we demonstrate that epidermal TRPV4 plays a prominent, hitherto unrecognized role in UVB-evoked skin tissue damage and pain of sunburn.  相似文献   

4.
Defective mitochondrial distribution in neurons is proposed to cause ATP depletion and calcium-buffering deficiencies that compromise cell function. However, it is unclear whether aberrant mitochondrial motility and distribution alone are sufficient to cause neurological disease. Calcium-binding mitochondrial Rho (Miro) GTPases attach mitochondria to motor proteins for anterograde and retrograde transport in neurons. Using two new KO mouse models, we demonstrate that Miro1 is essential for development of cranial motor nuclei required for respiratory control and maintenance of upper motor neurons required for ambulation. Neuron-specific loss of Miro1 causes depletion of mitochondria from corticospinal tract axons and progressive neurological deficits mirroring human upper motor neuron disease. Although Miro1-deficient neurons exhibit defects in retrograde axonal mitochondrial transport, mitochondrial respiratory function continues. Moreover, Miro1 is not essential for calcium-mediated inhibition of mitochondrial movement or mitochondrial calcium buffering. Our findings indicate that defects in mitochondrial motility and distribution are sufficient to cause neurological disease.Motor neuron diseases (MNDs), including ALS and spastic paraplegia (SP), are characterized by the progressive, length-dependent degeneration of motor neurons, leading to muscle atrophy, paralysis, and, in some cases, premature death. There are both inherited and sporadic forms of MNDs, which can affect upper motor neurons, lower motor neurons, or both. Although the molecular and cellular causes of most MNDs are unknown, many are associated with defects in axonal transport of cellular components required for neuron function and maintenance (16).A subset of MNDs is associated with impaired mitochondrial respiration and mitochondrial distribution. This observation has led to the hypothesis that neurodegeneration results from defects in mitochondrial motility and distribution, which, in turn, cause subcellular ATP depletion and interfere with mitochondrial calcium ([Ca2+]m) buffering at sites of high synaptic activity (reviewed in ref. 7). It is not known, however, whether mitochondrial motility defects are a primary cause or a secondary consequence of MND progression. In addition, it has been difficult to isolate the primary effect of mitochondrial motility defects in MNDs because most mutations that impair mitochondrial motility in neurons also affect transport of other organelles and vesicles (1, 811).In mammals, the movement of neuronal mitochondria between the cell body and the synapse is controlled by adaptors called trafficking kinesin proteins (Trak1 and Trak2) and molecular motors (kinesin heavy chain and dynein), which transport the organelle in the anterograde or retrograde direction along axonal microtubule tracks (7, 1224). Mitochondrial Rho (Miro) GTPase proteins are critical for transport because they are the only known surface receptors that attach mitochondria to these adaptors and motors (1215, 18, 25, 26). Miro proteins are tail-anchored in the outer mitochondrial membrane with two GTPase domains and two predicted calcium-binding embryonic fibroblast (EF) hand motifs facing the cytoplasm (12, 13, 25, 27, 28). A recent Miro structure revealed two additional EF hands that were not predicted from the primary sequence (29). Studies in cultured cells suggest that Miro proteins also function as calcium sensors (via their EF hands) to regulate kinesin-mediated mitochondrial “stopping” in axons (15, 16, 26). Miro-mediated movement appears to be inhibited when cytoplasmic calcium is elevated in active synapses, effectively recruiting mitochondria to regions where calcium buffering and energy are needed. Despite this progress, the physiological relevance of these findings has not yet been tested in a mammalian animal model. In addition, mammals ubiquitously express two Miro orthologs, Miro1 and Miro2, which are 60% identical (12, 13). However, the individual roles of Miro1 and Miro2 in neuronal development, maintenance, and survival have no been evaluated.We describe two new mouse models that establish the importance of Miro1-mediated mitochondrial motility and distribution in mammalian neuronal function and maintenance. We show that Miro1 is essential for development/maintenance of specific cranial neurons, function of postmitotic motor neurons, and retrograde mitochondrial motility in axons. Loss of Miro1-directed retrograde mitochondrial transport is sufficient to cause MND phenotypes in mice without abrogating mitochondrial respiratory function. Furthermore, Miro1 is not essential for calcium-mediated inhibition of mitochondrial movement or [Ca2+]m buffering. These findings have an impact on current models for Miro1 function and introduce a specific and rapidly progressing mouse model for MND.  相似文献   

5.
Recent studies have identified molecular pathways driving forgetting and supported the notion that forgetting is a biologically active process. The circuit mechanisms of forgetting, however, remain largely unknown. Here we report two sets of Drosophila neurons that account for the rapid forgetting of early olfactory aversive memory. We show that inactivating these neurons inhibits memory decay without altering learning, whereas activating them promotes forgetting. These neurons, including a cluster of dopaminergic neurons (PAM-β′1) and a pair of glutamatergic neurons (MBON-γ4>γ1γ2), terminate in distinct subdomains in the mushroom body and represent parallel neural pathways for regulating forgetting. Interestingly, although activity of these neurons is required for memory decay over time, they are not required for acute forgetting during reversal learning. Our results thus not only establish the presence of multiple neural pathways for forgetting in Drosophila but also suggest the existence of diverse circuit mechanisms of forgetting in different contexts.Although forgetting commonly has a negative connotation, it is a functional process that shapes memory and cognition (14). Recent studies, including work in relatively simple invertebrate models, have started to reveal basic biological mechanisms underlying forgetting (515). In Drosophila, single-session Pavlovian conditioning by pairing an odor (conditioned stimulus, CS) with electric shock (unconditioned stimulus, US) induces aversive memories that are short-lasting (16). The memory performance of fruit flies is observed to drop to a negligible level within 24 h, decaying rapidly early after training and slowing down thereafter (17). Memory decay or forgetting requires the activation of the small G protein Rac, a signaling protein involved in actin remodeling, in the mushroom body (MB) intrinsic neurons (6). These so-called Kenyon cells (KCs) are the neurons that integrate CS–US information (18, 19) and support aversive memory formation and retrieval (2022). In addition to Rac, forgetting also requires the DAMB dopamine receptor (7), which has highly enriched expression in the MB (23). Evidence suggests that the dopamine-mediated forgetting signal is conveyed to the MB by dopamine neurons (DANs) in the protocerebral posterior lateral 1 (PPL1) cluster (7, 24). Therefore, forgetting of olfactory aversive memory in Drosophila depends on a particular set of intracellular molecular pathways within KCs, involving Rac, DAMB, and possibly others (25), and also receives modulation from extrinsic neurons. Although important cellular evidence supporting the hypothesis that memory traces are erased under these circumstances is still lacking, these findings lend support to the notion that forgetting is an active, biologically regulated process (17, 26).Although existing studies point to the MB circuit as essential for forgetting, several questions remain to be answered. First, whereas the molecular pathways for learning and forgetting of olfactory aversive memory are distinct and separable (6, 7), the neural circuits seem to overlap. Rac-mediated forgetting has been localized to a large population of KCs (6), including the γ-subset, which is also critical for initial memory formation (21, 27). The site of action of DAMB for forgetting has yet to be established; however, the subgroups of PPL1-DANs implicated in forgetting are the same as those that signal aversive reinforcement and are required for learning (2830). It leaves open the question of whether the brain circuitry underlying forgetting and learning is dissociable, or whether forgetting and learning share the same circuit but are driven by distinct activity patterns and molecular machinery (26). Second, shock reinforcement elicits multiple memory traces through at least three dopamine pathways to different subdomains in the MB lobes (28, 29). Functional imaging studies have also revealed Ca2+-based memory traces in different KC populations (31). It is poorly understood how forgetting of these memory traces differs, and it remains unknown whether there are multiple regulatory neural pathways. Notably, when PPL1-DANs are inactivated, forgetting still occurs, albeit at a lower rate (7). This incomplete block suggests the existence of an additional pathway(s) that conveys forgetting signals to the MB. Third, other than memory decay over time, forgetting is also observed through interference (32, 33), when new learning or reversal learning is introduced after training (6, 34, 35). Time-based and interference-based forgetting shares a similar dependence on Rac and DAMB (6, 7). However, it is not known whether distinct circuits underlie forgetting in these different contexts.In the current study, we focus on the diverse set of MB extrinsic neurons (MBENs) that interconnect the MB lobes with other brain regions, which include 34 MB output neurons (MBONs) of 21 types and ∼130 dopaminergic neurons of 20 types in the PPL1 and protocerebral anterior medial (PAM) clusters (36, 37). These neurons have been intensively studied in olfactory memory formation, consolidation, and retrieval in recent years (e.g., 24, 2830, 3848); however, their roles in forgetting have not been characterized except for the aforementioned PPL1-DANs. In a functional screen, we unexpectedly found that several Gal4 driver lines of MBENs showed significantly better 3-h memory retention when the Gal4-expressing cells were inactivated. The screen has thus led us to identify two types of MBENs that are not involved in initial learning but play important and additive roles in mediating memory decay. Furthermore, neither of these MBEN types is required for reversal learning, supporting the notion that there is a diversity of neural circuits that drive different forms of forgetting.  相似文献   

6.
Treatments with two-small molecule tropomyosin receptor kinase B (trkB) ligands, 7,8 dihydroxyflavone (7,8 DHF) and deoxygedunin, were evaluated for their ability to promote the regeneration of cut axons in injured peripheral nerves in mice in which sensory and motor axons are marked by YFP. Peripheral nerves were cut and repaired with grafts from strain-matched, nonfluorescent donors and secured in place with fibrin glue. Lengths of profiles of regenerating YFP+ axons were measured 2 wk later from confocal images. Axon regeneration was enhanced when the fibrin glue contained dilutions of 500-nM solution of either small-molecule trkB agonist. In mice in which the neurotrophin receptor trkB is knocked out selectively in neurons, axon regeneration is very weak, and topical treatment with 7,8 DHF had no effect on axon regeneration. Similar treatments with deoxygedunin had only a modest effect. In conditional BDNF knockout mice, topical treatments with either 7,8 DHF or deoxygedunin resulted in a reversal of the poor regeneration found in controls and produced significant enhancement of regeneration. In WT mice treated with 2 wk of daily i.p. injections of either 7,8 DHF or deoxygedunin (5 mg/kg), regenerating axon profiles were nearly twice as long as in controls. Restoration of direct muscle responses evoked by sciatic nerve stimulation to pretransection levels over an 8-wk survival period was found only in the treated mice. Treatments with either small-molecule trkB agonist enhanced axon regeneration and muscle reinnervation after peripheral nerve injuries.Despite the capacity for axons in injured peripheral nerves to regenerate and reinnervate targets in the periphery, functional recovery after peripheral nerve injury is poor (1, 2). A major factor contributing to these poor functional outcomes is the need for regenerating axons to elongate in a pathway in the distal segment of a cut nerve to reach their targets (3). The regenerating axons enter this pathway over a protracted period (4), and once in the pathway, their elongation is dependent upon a favorable balance of growth-promoting and growth-inhibiting molecules (5). The availability of growth-promoting molecules can be limiting. Their production by supporting (Schwann) cells in the distal nerve segment becomes compromised over time, which has a negative impact on functional recovery when repair of damaged nerves is delayed or when regeneration has to take place over long distances, as it often does in humans (6, 7).One of the key growth-promoting molecules in the pathway used by regenerating axons in peripheral nerves is the neurotrophin brain derived neurotrophic factor (BDNF). In the pathway, it is produced by transformed Schwann cells and stimulates neurite elongation by binding to tropomyosin receptor kinase B (trkB) receptors on the regenerating axons (8, 9). Axon regeneration in cut nerves is severely compromised if the availability of BDNF to regenerating axons is reduced, such as by treatments with function-blocking antibodies (10) or by genetic manipulation of BDNF expression in Schwann cells (9). Local treatment of repaired nerves with recombinant human BDNF (rhBDNF) promotes enhanced early regeneration in mice (9, 11).On the basis of these findings, one might expect that administration of rhBDNF might form an effective treatment strategy to promote the regeneration of axons in cut nerves. Unfortunately, rhBDNF is a relatively large molecule (13 Kd), and it does not cross the blood–brain or blood–nerve barriers effectively (12). In addition, its biological half-life is relatively short (1–3 h) (13, 14). When it has been used experimentally, it has been given in very large doses (15), which is both expensive and runs the risk of untoward side effects or toxicities (16).Because of these limitations, small molecules that mimic the action of BDNF have been sought. Two molecules have emerged from a screen for their cell survival promotion and cell-signaling properties. Both 7,8 dihydroxyflavone (7,8 DHF) (17) and deoxygedunin (18) have been shown to function as trkB agonists in vitro. Both molecules have been used to promote BDNF-dependent activities in vivo (1820). Because they are small enough to pass the blood–brain (or blood–nerve) barrier, they have been used systemically, even by oral administration (18). One or both of these molecules have been applied to the study of a wide range of different neurological model systems, including learning (21), memory loss (22), excitotoxic stress (23), depression (24, 25), fear conditioning (19), neuromuscular synaptic transmission (26), and age-related synaptic plasticity (27). If treatments with these molecules, especially systemically, proved effective in enhancing axon regeneration in injured peripheral nerves, such a result could form the basis for the development of a novel treatment strategy for peripheral nerve injuries. The objective of this study was to evaluate the effectiveness of treating injured peripheral nerves with these small molecules on axon regeneration. We show here that treatments with 7,8 DHF and deoxygedunin promote enhanced axon regeneration in cut peripheral nerves, and systemic treatments result in more extensive muscle reinnervation. We also show that this enhancement is produced by signaling through neuronal trkB receptors. A preliminary report of some of these data has been made.  相似文献   

7.
Drosophila melanogaster can acquire a stable appetitive olfactory memory when the presentation of a sugar reward and an odor are paired. However, the neuronal mechanisms by which a single training induces long-term memory are poorly understood. Here we show that two distinct subsets of dopamine neurons in the fly brain signal reward for short-term (STM) and long-term memories (LTM). One subset induces memory that decays within several hours, whereas the other induces memory that gradually develops after training. They convey reward signals to spatially segregated synaptic domains of the mushroom body (MB), a potential site for convergence. Furthermore, we identified a single type of dopamine neuron that conveys the reward signal to restricted subdomains of the mushroom body lobes and induces long-term memory. Constant appetitive memory retention after a single training session thus comprises two memory components triggered by distinct dopamine neurons.Memory of a momentous event persists for a long time. Whereas some forms of long-term memory (LTM) require repetitive training (13), a highly relevant stimulus such as food or poison is sufficient to induce LTM in a single training session (47). Recent studies have revealed aspects of the molecular and cellular mechanisms of LTM formation induced by repetitive training (811), but how a single training induces a stable LTM is poorly understood (12).Appetitive olfactory learning in fruit flies is suited to address the question, as a presentation of a sugar reward paired with odor induces robust short-term memory (STM) and LTM (6, 7). Odor is represented by a sparse ensemble of the 2,000 intrinsic neurons, the Kenyon cells (13). A current working model suggests that concomitant reward signals from sugar ingestion cause associative plasticity in Kenyon cells that might underlie memory formation (1420). A single activation session of a specific cluster of dopamine neurons (PAM neurons) by sugar ingestion can induce appetitive memory that is stable over 24 h (19), underscoring the importance of sugar reward to the fly.The mushroom body (MB) is composed of the three different cell types, α/β, α′/β′, and γ, which have distinct roles in different phases of appetitive memories (11, 2125). Similar to midbrain dopamine neurons in mammals (26, 27), the structure and function of PAM cluster neurons are heterogeneous, and distinct dopamine neurons intersect unique segments of the MB lobes (19, 2834). Further circuit dissection is thus crucial to identify candidate synapses that undergo associative modulation.By activating distinct subsets of PAM neurons for reward signaling, we found that short- and long-term memories are independently formed by two complementary subsets of PAM cluster dopamine neurons. Conditioning flies with nutritious and nonnutritious sugars revealed that the two subsets could represent different reinforcing properties: sweet taste and nutritional value of sugar. Constant appetitive memory retention after a single training session thus comprises two memory components triggered by distinct reward signals.  相似文献   

8.
The ability to acquire large-scale recordings of neuronal activity in awake and unrestrained animals is needed to provide new insights into how populations of neurons generate animal behavior. We present an instrument capable of recording intracellular calcium transients from the majority of neurons in the head of a freely behaving Caenorhabditis elegans with cellular resolution while simultaneously recording the animal’s position, posture, and locomotion. This instrument provides whole-brain imaging with cellular resolution in an unrestrained and behaving animal. We use spinning-disk confocal microscopy to capture 3D volumetric fluorescent images of neurons expressing the calcium indicator GCaMP6s at 6 head-volumes/s. A suite of three cameras monitor neuronal fluorescence and the animal’s position and orientation. Custom software tracks the 3D position of the animal’s head in real time and two feedback loops adjust a motorized stage and objective to keep the animal’s head within the field of view as the animal roams freely. We observe calcium transients from up to 77 neurons for over 4 min and correlate this activity with the animal’s behavior. We characterize noise in the system due to animal motion and show that, across worms, multiple neurons show significant correlations with modes of behavior corresponding to forward, backward, and turning locomotion.How do patterns of neural activity generate an animal’s behavior? To answer this question, it is important to develop new methods for recording from large populations of neurons in animals as they move and behave freely. The collective activity of many individual neurons appears to be critical for generating behaviors including arm reach in primates (1), song production in zebrafinch (2), the choice between swimming or crawling in leech (3), and decision-making in mice during navigation (4). New methods for recording from larger populations of neurons in unrestrained animals are needed to better understand neural coding of these behaviors and neural control of behavior more generally.Calcium imaging has emerged as a promising technique for recording dynamics from populations of neurons. Calcium-sensitive proteins are used to visualize changes in intracellular calcium levels in neurons in vivo which serve as a proxy for neural activity (5). To resolve the often weak fluorescent signal of an individual neuron in a dense forest of other labeled cells requires a high magnification objective with a large numerical aperture that, consequently, can image only a small field of view. Calcium imaging has traditionally been performed on animals that are stationary from anesthetization or immobilization to avoid imaging artifacts induced by animal motion. As a result, calcium imaging studies have historically focused on small brain regions in immobile animals that exhibit little or no behavior (6).No previous neurophysiological study has attained whole-brain imaging with cellular resolution in a freely behaving unrestrained animal. Previous whole-brain cellular resolution calcium imaging studies of populations of neurons that involve behavior investigate either fictive locomotion (3, 7), or behaviors that can be performed in restrained animals, such as eye movements (8) or navigation of a virtual environment (9). One exception has been the development of fluorescence endoscopy, which allows recording from rodents during unrestrained behavior, although imaging is restricted to even smaller subbrain regions (10).Investigating neural activity in small transparent organisms allows one to move beyond studying subbrain regions to record dynamics from entire brains with cellular resolution. Whole-brain imaging was performed first in larval zebrafish using two-photon microscopy (7). More recently, whole-brain imaging was performed in Caenorhabditis elegans using two-photon (11) and light-field microscopy (12). Animals in these studies were immobilized, anesthetized, or both and thus exhibited no gross behavior.C. elegans’ compact nervous system of only 302 neurons and small size of only 1 mm make it ideally suited for the development of new whole-brain imaging techniques for studying behavior. There is a long and rich history of studying and quantifying the behavior of freely moving C. elegans dating back to the mid-1970s (13, 14). Many of these works involved quantifying animal body posture as the worm moved, for example as in ref. 15. To facilitate higher-throughput recordings of behavior, a number of tracking microscopes (1618) or multiworm imagers were developed (19, 20) along with sophisticated behavioral analysis software and analytical tools (21, 22). Motivated in part to understand these behaviors, calcium imaging systems were also developed that could probe neural activity in at first partially immobilized (23) and then freely moving animals, beginning with ref. 24 and and then developing rapidly (17, 18, 2529). One limitation of these freely moving calcium imaging systems is that they are limited to imaging only a very small subset of neurons and lack the ability to distinguish neurons that lie atop one another in the axial direction of the microscope. Despite this limitation, these studies, combined with laser-ablation experiments, have identified a number of neurons that correlate or affect changes in particular behaviors including the AVB neuron pair and VB-type motor neurons for forward locomotion; the AVA, AIB, and AVE neuron pairs and VA-type motor neurons for backward locomotion; and the RIV, RIB, and SMD neurons and the DD-type motor neurons for turning behaviors (17, 18, 25, 26, 28, 30, 31). To move beyond these largely single-cell studies, we sought to record simultaneously from the entire brain of C. elegans with cellular resolution and record its behavior as it moved about unrestrained.  相似文献   

9.
High-frequency deep brain stimulation (HFS) is clinically recognized to treat parkinsonian movement disorders, but its mechanisms remain elusive. Current hypotheses suggest that the therapeutic merit of HFS stems from increasing the regularity of the firing patterns in the basal ganglia (BG). Although this is consistent with experiments in humans and animal models of Parkinsonism, it is unclear how the pattern regularization would originate from HFS. To address this question, we built a computational model of the cortico-BG-thalamo-cortical loop in normal and parkinsonian conditions. We simulated the effects of subthalamic deep brain stimulation both proximally to the stimulation site and distally through orthodromic and antidromic mechanisms for several stimulation frequencies (20–180 Hz) and, correspondingly, we studied the evolution of the firing patterns in the loop. The model closely reproduced experimental evidence for each structure in the loop and showed that neither the proximal effects nor the distal effects individually account for the observed pattern changes, whereas the combined impact of these effects increases with the stimulation frequency and becomes significant for HFS. Perturbations evoked proximally and distally propagate along the loop, rendezvous in the striatum, and, for HFS, positively overlap (reinforcement), thus causing larger poststimulus activation and more regular patterns in striatum. Reinforcement is maximal for the clinically relevant 130-Hz stimulation and restores a more normal activity in the nuclei downstream. These results suggest that reinforcement may be pivotal to achieve pattern regularization and restore the neural activity in the nuclei downstream and may stem from frequency-selective resonant properties of the loop.High-frequency (i.e., above 100 Hz) deep brain stimulation (HFS) of the basal ganglia (BG) and thalamus is clinically recognized to treat movement disorders in Parkinson’s disease (PD) (14), but its therapeutic mechanisms remain unclear (5, 6).Early hypotheses about HFS were derived from the rate-based model of the BG function (7, 8) and postulated the disruption of the output of the BG-thalamic system via either the inactivation of neurons in the stimulated site (target) (915), which would provide an effect similar to a surgical lesion, or the abnormal excitation of axons projecting out of the target (1619), which would disrupt the neuronal activity in the structures downstream, including any pathophysiological activity (20).More recently, an ever-growing number of experiments in PD humans and animal models of Parkinsonism has indicated that HFS affects the firing patterns of the neurons rather than the mean firing rate both in the target and the structures downstream (18, 19, 2131) and it replaces repetitive low-frequency (i.e., ≤50 Hz) bursting patterns with regularized (i.e., more tonic) patterns at higher frequencies (25, 26). It has been proposed that increased pattern regularity of neurons in the target may be therapeutic (5, 3237), but it is still unknown how this regularity comes about with HFS.It has been suggested that an increased pattern regularity can deplete the information content of the target output and this lack of information would act as an “information lesion” (33) and prevent the pathological activity from being transmitted within the BG-thalamic system (22, 33, 36). As a result, an information lesion in the target [typically, one among the subthalamic nucleus (STN), internal globus pallidus (GPi), or thalamus] would have effects similar to those of a destructive lesion in the same site, which has been reported to alleviate the movement disorders (38).Instead, studies (32, 34, 35, 37) have suggested that an increased pattern regularity of the BG output partly compensates the PD-evoked impairment of the information-processing capabilities of the thalamo-cortical system, and this restores a more faithful thalamic relay of the sensorimotor information (35, 39).Although intriguing, these hypotheses remain elusive on (i) the neuronal mechanisms that would elicit pattern regularization (e.g., why regularization would be relevant only for HFS) and (ii) the effects that increased regularity would have on the cortico-BG-thalamo-cortical loop.It has been hypothesized that pattern regularization occurs because axons projecting out of the target follow the pattern of the stimulus pulses (40, 41) and, given the segregated organization of the BG-thalamic connections (42), it has been assumed that pattern regularization percolates straightforward from the target to the structures immediately downstream (34, 36). However, this representation of the pattern regularization as a “local” effect can hardly be reconciled with the fact that HFS of any structure of the cortico-BG-thalamo-cortical loop is therapeutic for at least some movement disorders (14, 4347), nor does it explain why stimulation at frequencies above 160–180 Hz is not necessarily therapeutic despite the fact that the regularity of the axonal patterns may increase (48, 49). Moreover, coherence in the 8–30-Hz band among neurons across different structures may decrease under HFS but not for lower frequencies (26, 5052), which suggests the emergence of diffused changes in neuronal activity that would be hardly accounted for with purely local effects.There is emerging evidence, instead, that HFS affects multiple structures simultaneously. First, it has been shown that deep brain stimulation (DBS) may antidromically activate afferent axons and fibers of passage (5359), thus reaching structures not immediately downstream. Second, studies (57, 58) observed in 6-hydroxydopamine (6-OHDA)-intoxicated rats that the antidromic effects increase with the stimulation frequency and peak around 110–130 Hz. Third, it has been shown in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-intoxicated nonhuman primates (NHPs) that STN DBS may evoke similar poststimulus responses in different BG structures, both downstream from and upstream to the STN (5, 27, 28, 30, 60). Finally, it has been reported that the cortico-BG-thalamo-cortical system consists of multiple sets of reentrant, interconnected, and partially overlapping neuronal loops (5, 42, 61, 62), which means that the structures upstream to the target (e.g., the striatum) may play an important role in the therapeutic mechanisms of HFS.Altogether, these results suggest that (A) pattern regularization is a global effect that exploits the closed-loop nature of the cortico-BG-thalamo-cortical system and selectively emerges only for specific HFS values, and that (B) the therapeutic merit of pattern regularization has to deal with the restoration of a more normal functionality of the entire cortico-BG-thalamo-cortical loop rather than with variations in the information content of one specific structure.We explored hypotheses (A) and (B) and assessed the system-wide effects of DBS by constructing a computational model of the cortico-BG-thalamo-cortical loop in both normal and parkinsonian conditions and by simulating the effects of STN DBS both at low (20–80 Hz) and high (100–180 Hz) frequencies. The model includes populations of single-compartment neurons and interneurons from motor cortex, striatum, GPi, and thalamus according to a network topology derived from the NHP anatomy, and it simulates both the orthodromic and antidromic effects of DBS. As a result, this model reproduced both average activity and discharge patterns of single units in NHP and rats under normal and parkinsonian conditions, with and without DBS, for all modeled structures.We show through numerical simulation that hypothesis (A) is significantly contributed by reinforcement mechanisms in the striatum. These mechanisms are selectively elicited by HFS, facilitate the percolation of regularized discharge patterns from the striatum to the GPi, and have a primary role in (B), because the percolated striato-pallidal input combines with the local effects of STN DBS to restore the thalamic relay function (63).  相似文献   

10.
11.
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   

12.
13.
The activation of cochlear progenitor cells is a promising approach for hair cell (HC) regeneration and hearing recovery. The mechanisms underlying the initiation of proliferation of postnatal cochlear progenitor cells and their transdifferentiation to HCs remain to be determined. We show that Notch inhibition initiates proliferation of supporting cells (SCs) and mitotic regeneration of HCs in neonatal mouse cochlea in vivo and in vitro. Through lineage tracing, we identify that a majority of the proliferating SCs and mitotic-generated HCs induced by Notch inhibition are derived from the Wnt-responsive leucine-rich repeat-containing G protein-coupled receptor 5 (Lgr5+) progenitor cells. We demonstrate that Notch inhibition removes the brakes on the canonical Wnt signaling and promotes Lgr5+ progenitor cells to mitotically generate new HCs. Our study reveals a new function of Notch signaling in limiting proliferation and regeneration potential of postnatal cochlear progenitor cells, and provides a new route to regenerate HCs from progenitor cells by interrupting the interaction between the Notch and Wnt pathways.Sensory hair cell (HC) loss is the major cause of hearing loss and balance disorder. In nonmammalian vertebrates, HCs are regenerated in both auditory and vestibular systems after HC loss, leading to functional recovery of hearing and balance function (13). In mammals, limited spontaneous HC regeneration occurs in the vestibular system (48). In the adult mammalian vestibular sensory epithelium, inner ear stem cells were isolated with the capacity to differentiate into HCs and other inner ear cell types (9). In contrast, only neonatal mammalian cochleae have limited HC regeneration capacity in vivo, and harbor stem cells or progenitor cells that could proliferate and regenerate new HCs (1012); however, no spontaneous HC regeneration has been observed in the mature cochlea (13).Recent studies reported that in neonatal mouse cochlea, Wnt-responsive leucine-rich repeat-containing G protein-coupled receptor 5 (Lgr5+) cells are the progenitors with the capacity to regenerate HCs under certain condition (11, 12, 1417). However, endogenous Lgr5+ progenitors maintain mitotic quiescence in neonatal mouse cochlea, suggesting the existence of negative regulators that inhibit the proliferation of those progenitors. Overexpressing β-catenin in the Lgr5+ or Sox2+ [SRY (sex determining region Y)-box 2] cells initiates proliferation by forming BrdU+ foci adjacent to HCs (14, 16), serving as a potential approach to overcome the nonproliferative barrier of progenitors in the cochlear sensory epithelium. Alternatively, the identification and removal of the negative regulators could provide a new route to activate cochlear progenitor proliferation to achieve HC generation.Inner ear sensory epithelium consists of a mosaic of HCs and supporting cells (SCs), generated from the same precursor pool in the prosensory domain during development (18, 19). The formation of the mosaic HC and SC pattern is mediated by lateral inhibition through the Notch signaling pathway (20, 21). Evidence from birds and mice suggests that Notch signaling negatively regulates the formation of HCs and the loss of Notch signaling generates supernumerary ectopic HCs at the expense of SCs (2226). During early embryonic development, it has been shown that inhibition of Notch/JAG2 and DLL1 may prolong the proliferation process of the prosensory cells in the inner ear (27, 28). Notch signaling may play an important role maintaining the homeostasis of cochlear sensory epithelium on cell number and structures.Although Wnt and Notch signaling are two fundamental pathways that regulate progenitor cell proliferation and determine the cell fate in the inner ear, their relationship remains largely unclear in the postnatal mouse cochlea. Here, by inhibiting the Notch signaling using Notch1 conditional knockout (KO) mice in vivo and by γ-secretase inhibitor IX (DAPT) treatment in vitro, we found that both inhibitions led to the proliferation of SCs and mitotic generation of HCs in the postnatal cochlear sensory epithelium. Lineage tracing demonstrated that a majority of the proliferating SCs and mitotically generated HCs were of Lgr5+ lineage. In addition, we showed that Notch inhibition resulted in β-catenin up-regulation in the Sox2+ SCs, whereas inhibition of Wnt signaling significantly decreased SCs proliferation and mitotic generation of HCs induced by Notch inhibition.  相似文献   

14.
15.
How activity of sensory neurons leads to perceptual decisions remains a challenge to understand. Correlations between choices and single neuron firing rates have been found early in vestibular processing, in the brainstem and cerebellum. To investigate the origins of choice-related activity, we have recorded from otolith afferent fibers while animals performed a fine heading discrimination task. We find that afferent fibers have similar discrimination thresholds as central cells, and the most sensitive fibers have thresholds that are only twofold or threefold greater than perceptual thresholds. Unlike brainstem and cerebellar nuclei neurons, spike counts from afferent fibers do not exhibit trial-by-trial correlations with perceptual decisions. This finding may reflect the fact that otolith afferent responses are poorly suited for driving heading perception because they fail to discriminate self-motion from changes in orientation relative to gravity. Alternatively, if choice probabilities reflect top-down inference signals, they are not relayed to the vestibular periphery.The neural basis of perception holds a long-standing fascination for neuroscientists. How do the properties of single neurons and populations of neurons relate to, and account for, sensory perception? In all sensory systems, information about the world is first translated into neural activity by peripheral receptor neurons, and then transformed by multiple stages of subcortical and cortical processing into a perceptual decision. How and where does perception emerge from multiple neural representations that appear to be at least partially redundant? These questions have been addressed often in sensory and multisensory cortex (e.g., in refs. 13 for vestibular perception), but much less is known about how the activity of sensory afferents relates to perceptual sensitivity and perceptual decisions.One way to assess a potential role of sensory neurons in a perceptual task is to compare neuronal and perceptual sensitivity, measured simultaneously in the same subject (4). Although this comparison has been done many times for cortical neurons (1, 57), little is known about how the sensitivity of peripheral afferents compares with behavior apart from microneurography studies of tactile afferents in humans (8, 9). To our knowledge, the present study provides the first direct comparison of afferent neuronal sensitivity and perceptual sensitivity, measured simultaneously in experimental animals. Results of such comparisons have important implications for understanding how population encoding and decoding may constrain and shape the information that guides behavior.Another way that neuroscientists have explored the functional links between sensory neurons and perception is by measuring the trial-by-trial correlations between neural activity and perceptual decisions, which are typically quantified as “choice probabilities” (CPs) (10). The “bottom-up” explanation of CPs is that trial-to-trial variability in the activity of sensory neurons drives variability in decisions; in contrast, the “top-down” explanation is that higher level signals related to decisions or featural attention are fed back to modulate the responses of sensory neurons (reviewed in ref. 11). To our knowledge, CPs have never been measured for primary afferents in any sensory system. Such measurements are potentially of interest because they may help clarify whether trial-to-trial response variability at the sensory periphery can be correlated with decisions, and whether top-down signals may propagate all of the way back to afferents.For some sensory systems and perceptual tasks, the relationships between afferent activity, behavioral sensitivity, and perceptual decisions cannot be examined because afferents do not exhibit the same forms of stimulus selectivity seen in the cortex (e.g., binocular disparity or direction tuning in primate visual neurons). In this regard, the vestibular system may provide an alternative model to probe the origin of CPs because similar basic forms of spatiotemporal directional selectivity are seen at many levels of processing, from afferents to cortex (12). Recently, we provided the first demonstration (13), to our knowledge, that subcortical neurons in the vestibular nuclei (VN) and cerebellar nuclei (CN) could exhibit robust CPs, and these effects were even larger than those effects measured in some cortical areas under identical stimulus conditions (e.g., ref. 1). Interestingly, we also found that the CPs of these subcortical neurons were correlated with the degree to which the neurons represented translation (heading) without being confounded by head orientation relative to gravity. This result suggested that CPs might emerge in the vestibular pathways after resolution of the tilt-translation ambiguity (14). If so, this logic predicts that otolith afferents, which encode the net gravitoinertial acceleration, and therefore confound tilt and translation (15), would not exhibit CPs.We simultaneously measured heading discrimination performance while recording from otolith afferents, the primary sensory neurons that carry linear acceleration information to the brain. We addressed the following questions. First, how does the heading sensitivity of otolith afferents compare with central vestibular neurons? Are central vestibular neurons more sensitive than afferents because they have greater average modulation amplitudes than afferents (1623), or do changes in response variability counteract these differences in response gain? Second, do otolith afferents show choice-related activity, and how do CPs of afferents compare with CPs measured in central neurons and cortical areas?  相似文献   

16.
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   

17.
During development, a neuron transitions from a state of rapid growth to a stable morphology, and neurons within the adult mammalian CNS lose their ability to effectively regenerate in response to injury. Here, we identify a novel form of neuronal regeneration, which is remarkably independent of DLK-1/DLK, KGB-1/JNK, and other MAPK signaling factors known to mediate regeneration in Caenorhabditis elegans, Drosophila, and mammals. This DLK-independent regeneration in C. elegans has direct genetic and molecular links to a well-studied form of endogenous activity-dependent ectopic axon outgrowth in the same neuron type. Both neuron outgrowth types are triggered by physical lesion of the sensory dendrite or mutations disrupting sensory activity, calcium signaling, or genes that restrict outgrowth during neuronal maturation, such as SAX-1/NDR kinase or UNC-43/CaMKII. These connections suggest that ectopic outgrowth represents a powerful platform for gene discovery in neuronal regeneration. Moreover, we note numerous similarities between C. elegans DLK-independent regeneration and lesion conditioning, a phenomenon producing robust regeneration in the mammalian CNS. Both regeneration types are triggered by lesion of a sensory neurite via reduction of neuronal activity and enhanced by disrupting L-type calcium channels or elevating cAMP. Taken as a whole, our study unites disparate forms of neuronal outgrowth to uncover fresh molecular insights into activity-dependent control of the adult nervous system’s intrinsic regenerative capacity.One of the principal goals of modern neuroscience is the comprehensive understanding and therapeutic application of neuronal regeneration (1). This goal is particularly relevant in the case of the mammalian central nervous system (CNS), which regenerates poorly. Efforts to enhance axon regeneration generally fall into two broad categories: eliminating or blocking nonpermissive extrinsic inhibitors or promoting a neuron’s intrinsic regenerative capacity (2). Although much research in previous decades focused on the extrinsic angle, recent encouraging developments, particularly in invertebrate models, increasingly examine the cell intrinsic control of regeneration. Studies demonstrate that axon regeneration recruits or recapitulates mechanisms involved in a diverse range of biological processes, including synapse formation, stress response, apoptosis, and development.During development, neuronal electrical activity acts as a common intracellular feedback mechanism to establish appropriate connections and modulate outgrowth (3). Subsequently, a neuron transitions from a state of rapid growth to a stable morphology, and neurons within the adult mammalian CNS lose their ability to effectively regenerate in response to injury. A striking exception to this paradigm is lesion conditioning, a phenomenon exemplified by the dorsal root ganglion (DRG), where peripheral axon damage enables robust central axon regeneration (4). The peripheral lesion is thought to revert the neuron to a growth-permissive state in an activity-dependent manner (5). These observations suggest that activity-dependent inhibitors of neurite outgrowth may represent a potent therapeutic target for enhancing neuronal regeneration in the face of injury or disease.The nematode Caenorhabditis elegans is a well-established and powerful model system (6) for the genetic and molecular study of both activity-dependent neuronal development and regeneration. Specifically, several studies (79) have observed spontaneous ectopic axon outgrowths from four bipolar sensory neurons of mutant animals raised under stressed conditions (high temperature). An extensive genetic study indicated that the mutations disrupt sensory activity, which triggers cell-specific ectopic outgrowth after the neurons establish their initial pattern of innervations (10). More recently, the application of advanced subcellular laser ablation techniques (11) to C. elegans has permitted quantitative in vivo study of single-neuron regeneration in this genetically tractable system. Several studies have revealed the role of conserved mitogen-activated protein kinase (MAPK) pathways, including the dual leucine zipper kinase (DLK) and the parallel but coordinated c-Jun N-terminal kinase (JNK) pathways (12, 13, 14). Other studies have clarified the beneficial role of caspase activity (15) and the role of calcium signaling (16, 17). Thus, C. elegans, by virtue of its facile genetics and amenability to precise laser surgery, is an excellent system for studying axon outgrowth and regeneration.In a previous study, we sought to dissect the subcellular components of the C. elegans ASJ sensory neurons by femtosecond laser surgery to establish their role in development and behavior (18). ASJ axons rapidly regenerate to the vicinity of their original synaptic targets, but regeneration is prevented by mutation of dlk-1, which plays a crucial role in neuronal regeneration in C. elegans (12), Drosophila (19), and mammals (20, 21). Remarkably, although dlk-1(ju476) completely prevents regeneration after axotomy, we discovered that strong DLK-independent regeneration returns under two conditions, which we explore in this study. First, regeneration occurs after severing both the axon and sensory dendrite, indicating that sensory dendrite cuts trigger robust DLK-independent regeneration. Second, even without a dendrite cut, regeneration occurs under mutations that also trigger ectopic axon outgrowth. Here, we demonstrate that this novel type of regeneration is largely independent of both the DLK-1/DLK and the KGB-1/JNK pathways. We find numerous direct links between this DLK-independent regeneration and ectopic outgrowth in the same neurons. Finally, we note similarities between DLK-independent regeneration in C. elegans and regeneration after lesion conditioning in mammals, including shared phenotypes, role of L-type voltage-gated calcium channel (L-VGCC), and mediation by cyclic adenosine monophosphate (cAMP). As such, our study establishes a novel regeneration pathway in C. elegans, where physical, genetic, or pharmacological lesion of a sensory neurite or sensory signaling removes activity-dependent developmental inhibitors to permit robust regeneration.  相似文献   

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

19.
High-density accumulation of voltage-gated sodium (Nav) channels at nodes of Ranvier ensures rapid saltatory conduction along myelinated axons. To gain insight into mechanisms of node assembly in the CNS, we focused on early steps of nodal protein clustering. We show in hippocampal cultures that prenodes (i.e., clusters of Nav channels colocalizing with the scaffold protein ankyrinG and nodal cell adhesion molecules) are detected before myelin deposition along axons. These clusters can be induced on purified neurons by addition of oligodendroglial-secreted factor(s), whereas ankyrinG silencing prevents their formation. The Nav isoforms Nav1.1, Nav1.2, and Nav1.6 are detected at prenodes, with Nav1.6 progressively replacing Nav1.2 over time in hippocampal neurons cultured with oligodendrocytes and astrocytes. However, the oligodendrocyte-secreted factor(s) can induce the clustering of Nav1.1 and Nav1.2 but not of Nav1.6 on purified neurons. We observed that prenodes are restricted to GABAergic neurons, whereas clustering of nodal proteins only occurs concomitantly with myelin ensheathment on pyramidal neurons, implying separate mechanisms of assembly among different neuronal subpopulations. To address the functional significance of these early clusters, we used single-axon electrophysiological recordings in vitro and showed that prenode formation is sufficient to accelerate the speed of axonal conduction before myelination. Finally, we provide evidence that prenodal clusters are also detected in vivo before myelination, further strengthening their physiological relevance.Voltage-gated sodium (Nav) channels are highly enriched at the axon initial segment (AIS) and the node of Ranvier, allowing generation and rapid propagation of action potentials by saltatory conduction in myelinated fibers. These axonal domains also contain cell adhesion molecules [e.g., neurofascin 186 (Nfasc186)] and the scaffolding proteins ankyrinG (AnkG) and βIV spectrin, which provide a potential link with the actin cytoskeleton (1). Flanking the nodes are the paranodes, where axoglial junctions between paranodal myelin loops and the axon are formed through interactions between axonal contactin-associated protein (Caspr)/contactin and glial Nfasc155 (2, 3). Although the mechanisms of nodal assembly are best characterized in the peripheral nervous system (49), less is known about the cellular and molecular mechanisms underlying node assembly in the CNS. ECM proteins, adhesion molecules, such as Nfasc186, and also, axoglial paranodal junctions have been shown to trigger CNS nodal clustering, although their respective roles remain uncertain (1019). Moreover, axonal clustering of Nav channels before myelin deposition and oligodendroglial contact has been shown to occur in retinal ganglion cell (RGC) cultures, where these clusters were induced by oligodendroglial-secreted factor(s) (20, 21).Here, we have investigated the cellular and molecular mechanisms underlying nodal protein assembly in hippocampal neuron cultures. We first showed that evenly spaced clusters of Nav channels, colocalizing with Nfasc186 and AnkG, are detected along axons before myelination. Strikingly, this prenode assembly is restricted to GABAergic interneurons, suggesting the existence of different mechanisms of nodal assembly. The prenodal clustering can be induced on purified neurons by the addition of oligodendroglial-secreted factor(s) and also depends on intrinsic cues, such as AnkG. Furthermore, we also provide evidence that these clusters are detected in vivo before myelination on hippocampal tissue sections. Finally, to gain insight into their functional significance, we performed in vitro simultaneous somatic and axonal recordings and showed that the presence of prenodes increases the speed of action potential propagation along axons before myelination.  相似文献   

20.
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