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 15 毫秒
1.
Members of the ADP-ribosylation factor (ARF) family of small guanosine triphosphate-binding proteins play an essential role in membrane trafficking which subserves constitutive protein transport along exocytic and endocytic pathways within eukaryotic cell bodies. In growing neurons, membrane trafficking within motile growth cones distant from the cell body underlies the rapid plasmalemmal expansion which subserves axon elongation. We report here that ARF is a constituent of axonal growth cones, and that application of brefeldin A to neurons in culture produces a rapid arrest of axon extension that can be ascribed to inhibition of ARF function in growth cones. Our findings demonstrate a role for ARF in growth cones that is coupled tightly to the rapid growth of neuronal processes characteristic of developmental and regenerative axon elongation, and indicate that ARF participates not only in constitutive membrane traffic within the cell body, but also in membrane dynamics within growing axon endings.  相似文献   

2.
Members of the ADP‐ribosylation factor (ARF) family of small guanosine triphosphate–binding proteins play an essential role in membrane trafficking which subserves constitutive protein transport along exocytic and endocytic pathways within eukaryotic cell bodies. In growing neurons, membrane trafficking within motile growth cones distant from the cell body underlies the rapid plasmalemmal expansion which subserves axon elongation. We report here that ARF is a constituent of axonal growth cones, and that application of brefeldin A to neurons in culture produces a rapid arrest of axon extension that can be ascribed to inhibition of ARF function in growth cones. Our findings demonstrate a role for ARF in growth cones that is coupled tightly to the rapid growth of neuronal processes characteristic of developmental and regenerative axon elongation, and indicate that ARF participates not only in constitutive membrane traffic within the cell body, but also in membrane dynamics within growing axon endings. © 1999 John Wiley & Sons, Inc. J Neurobiol 38: 105–115, 1999  相似文献   

3.
Sytnyk  V. N.  Korogod  S. M.  Dityatev  A. E. 《Neurophysiology》2001,33(3):140-147
In our study, we showed that distribution of NCAM along the surface of actively growing in vitro neurites of murine hippocampal neurons comprises at least two components. The first component is reflected in exponential decline of the NCAM immunoreactivity from the soma and growth cone to a central part of the neurite. This component can be described by a model assuming diffusive redistribution of NCAM from the sites of its preferential insertion (from the neuron's soma and growth cones). The second component manifests itself as clusters of NCAM immunoreactivity irregularly distributed along the neurites. Some of these NCAM clusters, which were immunolabeled in a living neuron, can intermittently move back and forth along the neurites with a velocity up to 0.5 m/sec. Our data demonstrate that, besides passive NCAM diffusion, an active mechanism of NCAM redistribution exists in the central neurite part.  相似文献   

4.
Identified neurons of the buccal ganglion of the snail Helisoma when isolated from their ganglionic environment and plated in cell culture grow new neurites that are tipped with motile growth cones. Addition of the neurotransmitter serotonin to the culture medium surrounding actively growing neurons causes an immediate, premature cessation of neurite elongation in specific identified neurons. Serotonin selectively inhibits neurite extension of neurons B19 and P5 while having no effect on the extension of neuron B5. Coincident with the serotonin evoked inhibition of neurite elongation is an inhibition of growth cone motile activities and a retraction of growth cone filopodia and lamellipodia. One site of serotonin's growth inhibitory actions is directly at the growth cone rather than at the neurites or cell body. A second area of this study concerns connectivity. In Helisoma neurons the formation of electrical synaptic connections critically relies on both potential partner neurons having a mutual interaction of actively growing neurites. Neurons in a nongrowing state do not form electrical synapses (Hadley et al., 1983). As a result of inhibiting neurite extension, serotonin is able to affect synaptogenesis by preventing certain neurons (neurons B19) from forming electrical synaptic connections with other neurons (neurons B5) that are themselves competent to interconnect. Thus, by inhibiting neurite extension, serotonin is capable of regulating both the development of arborizations and the formation of connectivity.  相似文献   

5.
Axonogenesis involves a shift from uniform delivery of materials to all neurites to preferential delivery to the putative axon, supporting its more rapid extension. Waves, growth cone‐like structures that propagate down the length of neurites, were shown previously to correlate with neurite growth in dissociated cultured hippocampal neurons. Waves are similar to growth cones in their structure, composition and dynamics. Here, we report that waves form in all undifferentiated neurites, but occur more frequently in the future axon during initial neuronal polarization. Moreover, wave frequency and their impact on neurite growth are altered in neurons treated with stimuli that enhance axonogenesis. Coincident with wave arrival, growth cones enlarge and undergo a marked increase in dynamics. Through their engorgement of filopodia along the neurite shaft, waves can induce de novo neurite branching. Actin in waves maintains much of its cohesiveness during transport whereas actin in nonwave regions of the neurite rapidly diffuses as measured by live cell imaging of photoactivated GFP‐actin and photoconversion of Dendra‐actin. Thus, waves represent an alternative axonal transport mechanism for actin. Waves also occur in neurons in organotypic hippocampal slices where they propagate along neurites in the dentate gyrus and the CA regions and induce branching. Taken together, our results indicate that waves are physiologically relevant and contribute to axon growth and branching via the transport of actin and by increasing growth cone dynamics. © 2009 Wiley Periodicals, Inc. Develop Neurobiol 2009  相似文献   

6.
Growth cones of sympathetic neurons from the superior cervical ganglia of neonatal rats were studied using video-microscopy to determine events following contact between growth cones and other cell surfaces, including other growth cones and neurites. A variety of behaviors were observed to occur upon contact between growth cones. Most commonly, one growth cone would collapse and subsequently retract upon establishing filopodial contact with the growth cone of another sympathetic neuron. Contacts resulting in collapse and retraction were often accompanied by a rapid and transient burst of lamellipodial activity along the neurite 30-50 microns proximal to the retracting growth cone. In no instances did interactions between growth cones and either fibroblasts or red blood cells result in the growth cone collapsing, suggesting that a specific recognition event was involved. On several occasions, growth cones were seen to track other growth cones, although fasciculation was rare. In some cases, there was no obvious response between contacting growth cones. Growth cone-growth cone contact was almost four times more likely to result in collapse and retraction than was growth cone-neurite contact (45% vs 12%, respectively). These observations suggest that the superior cervical ganglion may be composed of neurons with different cell surface determinants and that these determinants are more concentrated on the surface of growth cones than on neurites. These results further suggest that contact-mediated inhibition of growth cone locomotion may play an important role in growth cone guidance.  相似文献   

7.
The right upper quadrant (RUQ) cells (R3-R13) of Aplysia regenerating in dissociated cell culture form unusually large growth cones. The movement of these growth cones was observed by time-lapse phase microscopy and their ultrastructure was examined by transmission electron microscopy. Their behavior and ultrastructure have features that are typical of growth cones in vitro. Additionally, they contain neurosecretory granules similar to those found in these cells in vivo. Because RUQ growth cones are large, they can be isolated by manual dissection. RUQ cells were grown in the presence of [35S]methionine and the labeled proteins transported to the growth cones were analyzed by SDS-PAGE. These proteins were compared to those in RUQ cell bodies, RUQ neurites, and to those in the neurites and cell bodies of other identified neurons grown in vitro. Most proteins synthesized by RUQ cells in vitro are transported to their growth cones, including several glycoproteins and the precursor to the R3-R14 neuropeptide. Neuropeptides are also synthesized by a number of other Aplysia neurons growing in vitro. We examined R2, LPL1, R15, and left upper quadrant neurons and found that their precursor peptides, like those of R3-R14, are readily recognized as major cell-specific radiolabeled bands on SDS gels. The presence in regenerating growth cones of neuropeptides, neurosecretory granules, and glycoproteins known to be rapidly transported toward synapses in vivo supports the emerging view that the growth cone in vitro contains not only a motility apparatus but also a macromolecular assembly capable of forming an active synapse immediately upon or shortly after contacting targets.  相似文献   

8.
During development, growth cones direct growing axons into appropriate targets. However, in some cortical pathways target innervation occurs through the development of collateral branches that extend interstitially from the axon shaft. How do such branches form? Direct observations of living cortical brain slices revealed that growth cones of callosal axons pause for many hours beneath their cortical targets prior to the development of interstitial branches. High resolution imaging of dissociated living cortical neurons for many hours revealed that the growth cone demarcates sites of future axon branching by lengthy pausing behaviors and enlargement of the growth cone. After a new growth cone forms and resumes forward advance, filopodial and lamellipodial remnants of the large paused growth cone are left behind on the axon shaft from which interstitial branches later emerge. To investigate how the cytoskeleton reorganizes at axon branch points, we fluorescently labeled microtubules in living cortical neurons and imaged the behaviors of microtubules during new growth from the axon shaft and the growth cone. In both regions microtubules reorganize into a more plastic form by splaying apart and fragmenting. These shorter microtubules then invade newly developing branches with anterograde and retrograde movements. Although axon branching of dissociated cortical neurons occurs in the absence of targets, application of a target-derived growth factor, FGF-2, greatly enhances branching. Taken together, these results demonstrate that growth cone pausing is closely related to axon branching and suggest that common mechanisms underlie directed axon growth from the terminal growth cone and the axon shaft.  相似文献   

9.
Sensory neurons were dissociated from lumbar dorsal root ganglia of embryonic chick and put into culture, either directly or after removing non-neuronal cells by density gradient centrifugation. The cells were grown on culture substrata of various kinds in medium containing nerve growth factor (NGF). After 24 h the cultures were fixed, mounted and analysed. Lengths of neurites were measured, and the numbers of primary processes formed at the cell body and of growth cones were counted. From these values, the rates of growth cone advance and frequency of growth cone branching were calculated. Neuronal outgrowths increased strikingly in length and complexity with embryonic age; there was a 3.5-fold increase in total neurite length and a 3-fold increase in the number of growth cones when neurons from 15-day embryos (E15) were compared with those from 8-day embryos (E8) grown on the same substratum (glass). Growth was markedly greater on surfaces prepared with laminin or conditioned medium compared with plain glass or air-dried collagen. When E15 neurons grown on glass were compared with those grown on laminin, for example, a 2.5-fold increase in total neurite length and a 3-fold increase in the number of growth cones was observed. Calculations showed that a major factor in these changes was an increase in the frequency of growth cone branching. The number of initial processes emanating from the cell body changed with age, but not with the different substrata tested. Non-neuronal cells when present in low numbers and in contact with neurons did not appear to influence neuronal geometry in a systematic way. Our results document the fact that both external factors (in this case, the nature of the culture substratum) and intrinsic factors (stage of development of the neuron) can influence the geometry of neurite outgrowth.  相似文献   

10.
The molecular layer of the cerebellum contains parallel fibers, the axons of granule neurons. We have examined the morphology and behavior of parallel fiber growth cones in the early postnatal rat cerebellum using the fluorescent tracer DiI. Parallel fiber growth cones distributed into three categories based on size and shape: short torpedo-like, long torpedo-like, and lamellopodial in form. The torpedo-like growth cones were modified by the addition of lamellopodia and/or filopodia, and the lamellopodial growth cones were often decorated with a filopodium. These three different growth cone morphologies were found throughout the growing region of the molecular layer. The nascent axons elaborated by premigratory granule neurons differed from the longer axons of more developed neurons in that they often had forked growth cones and extensive lamellopodial decoration along the axon shaft. Growth cones in living slices closely resembled those observed in the fixed preparations. The living growth cones exhibited frequent lamellopodial rearrangement and a side-to-side head-waving movement. The axon proximal to the growth cone was also dynamic. The axons curved and undulated, and mobile swellings formed along the axon shaft. These observations show that the growth cones of parallel fibers are similar to growth cones described for axons in other developing systems in terms of size, morphological characteristics, and dynamic behavior. © 1998 John Wiley & Sons, Inc. J Neurobiol 36: 91–104, 1998  相似文献   

11.
To determine the relationship between growth cone structure and motility, we compared the neurite extension rate, the form of individual growth cones, and the organization of f-actin in embryonic (E21) and postnatal (P30) sympathetic neurons in culture. Neurites extended faster on laminin than on collagen, but the P30 nerites were less than half as long as E21 neurites on both substrata. Growth cone shape was classified into one of five categories, ranging from fully lamellipodial to blunt endings. The leading margins of lamellipodia advanced smoothly across the substratum ahead of any filopodial activity and contained meshworks of actin filaments with no linear f-actin bundles, indicating that filopodia need not underlie lamellipodia. Rapid translocation (averaging 0.9-1.4 microns/min) was correlated with the presence of lamellipodia; translocation associated with filopodia averaged only 0.3-0.5 microns/min. This relationship extended to growth cones on a branched neurite where the translocation of each growth cone was dependent on its shape. Growth cones with both filopodial and lamellipodial components moved at intermediate rates. The prevalence of lamellipodial growth cones depended on age of the neurites; early in culture, 70% of E21 growth cones were primarily lamellipodial compared to 38% of P30 growth cones. A high percentage of E21 lamellipodial growth cones were associated with rapid neurite elongation (1.2 mm/day), whereas a week later, only 16% were lamellipodial, and neurites extended at 0.5 mm/day. Age-related differences in neurite extension thus reflected the proportion of lamellipodial growth cones present rather than disparities in basic structure or in the rates at which growth cones of a given type moved at different ages. Filopodia and lamellipodia are each sufficient to advance the neurite margin; however, rapid extension of superior cervical ganglion neurites was supported by lamellipodia independent of filopodial activity.  相似文献   

12.
The complex process of axon guidance is largely driven by the growth cone, which is the dynamic motile structure at the tip of the growing axon. During axon outgrowth, the growth cone must integrate multiple sources of guidance cue information to modulate its cytoskeleton in order to propel the growth cone forward and accurately navigate to find its specific targets1. How this integration occurs at the cytoskeletal level is still emerging, and examination of cytoskeletal protein and effector dynamics within the growth cone can allow the elucidation of these mechanisms. Xenopus laevis growth cones are large enough (10-30 microns in diameter) to perform high-resolution live imaging of cytoskeletal dynamics (e.g.2-4 ) and are easy to isolate and manipulate in a lab setting compared to other vertebrates. The frog is a classic model system for developmental neurobiology studies, and important early insights into growth cone microtubule dynamics were initially found using this system5-7 . In this method8, eggs are collected and fertilized in vitro, injected with RNA encoding fluorescently tagged cytoskeletal fusion proteins or other constructs to manipulate gene expression, and then allowed to develop to the neural tube stage. Neural tubes are isolated by dissection and then are cultured, and growth cones on outgrowing neurites are imaged. In this article, we describe how to perform this method, the goal of which is to culture Xenopus laevis growth cones for subsequent high-resolution image analysis. While we provide the example of +TIP fusion protein EB1-GFP, this method can be applied to any number of proteins to elucidate their behaviors within the growth cone.  相似文献   

13.
Laminin, an extracellular matrix molecule, is known to promote neurite growth. In the present study, the effects of soluble laminin on organelle transport and their relation to neurite growth were investigated in cultured dissociated mouse dorsal root ganglion (DRG) neurons. Laminin added into the extracellular medium was deposited on the surface of DRG neurons. DRG neurons incubated with soluble laminin exhibited branched, long, and thin neurites. Time-lapse study demonstrated that many small-diameter branches were newly formed after the addition of laminin. Thus, the growths of large-diameter primary neuritis, arising from cell bodies and branches extended from growth cones of primary neuritis, were analyzed separately. Laminin decreased the growth rate of primary neurites but increased that of branches. In primary neurites, acute addition of laminin rapidly decreased organelle movement in the neurite shaft and growth cone, accompanied by slowing of the growth cone advance. Branching of primary neurites occurred in response to laminin in some growth cones. In these growth cones, organelles protruded into nascent branches. In branches, soluble laminin increased organelle movement in the growth cone and the distal portion of the shaft. These results suggest that laminin inhibits the elongation of primary neurites but promotes branching and elongation of branches, all of which seem to be closely related to organelle transport.  相似文献   

14.
To understand the role of microtubules in growth cone turning, we observed fluorescently labeled microtubules in neurons as they encountered a substrate boundary. Neurons growing on a laminin-rich substrate avoided growing onto collagen type IV. Turning growth cones assumed heterogeneous morphologies and behaviors that depended primarily in their extent of adhesion to the substrate. We grouped these behaviors into three categories-sidestepping, motility, and growth-mediated reorientation. In sidestepping and motility-mediated reorientation, the growth cone and parts of the axon were not well attached to the substrate so the acquisition of an adherent lamella caused the entire growth cone to move away from the border and consequently reoriented the axon. In these cases, since the motility of the growth cone dominates its reorientation, the microtubules were passive, and reorientation occurred without significant axon growth. In growth-mediated reorientation, the growth cone and axon were attached to the substrate. In this case, microtubules reoriented within the growth cone to stabilize a lamella. Bundling of the reoriented microtubules was followed by growth cone collapse to form new axon, and further, polarized lamellipodial extension. These observations indicate that when the growth cone remains adherent to the substrate during turning, the reorientation and bundling of microtubules is an important, early step in growth cone turning.  相似文献   

15.
Neuronal regeneration after damage to an axon tract requires the rapid sealing of the injured plasma membrane and the subsequent formation of growth cones that can lead regenerating processes to their appropriate target. Membrane sealing and growth cone formation are Ca(2+)-dependent processes, but the signaling pathways activated by Ca(2+) to bring about these effects remain poorly understood. An in vitro injury model was employed in which neurites from identified snail neurons (Helisoma trivolvis) were transected with a glass microknife, and the formation of new growth cones from the distal portions of transected neurites was recorded at defined times after transection. This study presents three main results. First, phospholipase A(2) (PLA(2)), a calcium-activated enzyme, is necessary for membrane sealing in vitro. Second, PLA(2) activity is also required for the formation of a new growth cone after the membrane has sealed successfully. Thus, PLA(2) plays a dual role by affecting both growth cone formation and membrane sealing. Third, the injury-induced activation of PLA(2) by Ca(2+) controls growth cone formation through the production of leukotrienes, secondary metabolites of PLA(2) activity. Taken together, these results suggest that the injury-induced Ca(2+) influx acts via PLA(2) and leukotriene production to assure growth cone formation. These findings indicate that events that cause an inhibition of PLA(2) or lipoxygenases, enzymes that produce leukotrienes, could result in the inability of neurites to regenerate.  相似文献   

16.
In this study, we examined the subcellular distribution and functions of LIMK1 in developing neurons. Confocal microscopy, subcellular fractionation, and expression of several epitope-tagged LIMK1 constructs revealed that LIMK1 is enriched in the Golgi apparatus and growth cones, with the LIM domain required for Golgi localization and the PDZ domain for its presence at neuritic tips. Overexpression of wild-type LIMK1 suppresses the formation of trans-Golgi derived tubules, and prevents cytochalasin D-induced Golgi fragmentation, whereas that of a kinase-defective mutant has the opposite effect. Transfection of wild-type LIMK1 accelerates axon formation and enhances the accumulation of Par3/Par6, insulin-like growth factor (IGF)1 receptors, and neural cell adhesion molecule (NCAM) at growth cones, while inhibiting the Golgi export of synaptophysin-containing vesicles. These effects were dependent on the Golgi localization of LIMK1, paralleled by an increase in cofilin phosphorylation and phalloidin staining in the region of the Golgi apparatus, and prevented by coexpression of constitutive active cofilin. The long-term overexpression of LIMK1 produces growth cone collapse and axon retraction, an effect that is dependent on its growth cone localization. Together, our results suggest an important role for LIMK1 in axon formation that is related with its ability to regulate Golgi dynamics, membrane traffic, and actin cytoskeletal organization.  相似文献   

17.
Collapsin response mediator proteins (CRMPs) are involved in signal transduction after exposure of neural cells to the axon guidance molecule Semaphorin 3A/collapsin. All five known CRMPs are expressed in the developing cerebral cortex and neocortical neurons are responsive to Semaphorin 3A. Here, we examine the expression and subcellular localization of CRMPs in neocortical neurons and in neonatal rat brain. In neocortical neurons CRMP-4 was detected in the perikaryon with a diffuse cytosolic distribution. In neurites and at growth cones punctate staining patterns were observed. Extraction of neuron cultures with methyl-beta-cyclodextrin to deplete cholesterol caused rapid redistribution of the punctate CRMP-4 staining into larger patches and abundant growth cone collapse. Western blotting of brain extracts demonstrated for all CRMPs the existence of soluble, detergent-extractable, and Triton X-100-resistant forms. Furthermore, sucrose density gradient centrifugation after solubilization of brain membranes with Triton X-100 revealed that CRMP-1, -3, -5, and to a lower extent CRMP-4 are associated with a detergent-resistant fraction with low buoyant density, but CRMP-2 was not detectable in this fraction. Thus, we propose that lipid rafts form sites for the compartmentalization of signaling events involving specific CRMPs and that the integrity of these membrane microdomains is essential for the maintenance of growth cones.  相似文献   

18.
Growth cone behavior and production of traction force   总被引:11,自引:9,他引:2       下载免费PDF全文
《The Journal of cell biology》1990,111(5):1949-1957
The growth cone must push its substrate rearward via some traction force in order to propel itself forward. To determine which growth cone behaviors produce traction force, we observed chick sensory growth cones under conditions in which force production was accommodated by movement of obstacles in the environment, namely, neurites of other sensory neurons or glass fibers. The movements of these obstacles occurred via three, different, stereotyped growth cone behaviors: (a) filopodial contractions, (b) smooth rearward movement on the dorsal surface of the growth cone, and (c) interactions with ruffling lamellipodia. More than 70% of the obstacle movements were caused by filopodial contractions in which the obstacle attached at the extreme distal end of a filopodium and moved only as the filopodium changed its extension. Filopodial contractions were characterized by frequent changes of obstacle velocity and direction. Contraction of a single filopodium is estimated to exert 50-90 microdyn of force, which can account for the pull exerted by chick sensory growth cones. Importantly, all five cases of growth cones growing over the top of obstacle neurites (i.e., geometry that mimics the usual growth cone/substrate interaction), were of the filopodial contraction type. Some 25% of obstacle movements occurred by a smooth backward movement along the top surface of growth cones. Both the appearance and rate of movements were similar to that reported for retrograde flow of cortical actin near the dorsal growth cone surface. Although these retrograde flow movements also exerted enough force to account for growth cone pulling, we did not observe such movements on ventral growth cone surfaces. Occasionally obstacles were moved by interaction with ruffling lamellipodia. However, we obtained no evidence for attachment of the obstacles to ruffling lamellipodia or for directed obstacle movements by this mechanism. These data suggest that chick sensory growth cones move forward by contractile activity of filopodia, i.e., isometric contraction on a rigid substrate. Our data argue against retrograde flow of actin producing traction force.  相似文献   

19.
Growth cones are specialized sensorimotor structures at the tips of neurites implicated in pathfinding decisions and axonal outgrowth during neuronal development. We generated a mouse monoclonal antibody (mAb 2G13) against chick tectum and found that the antibody exclusively labelled axonal growth cones, particularly their filopodia and lamellipodia, in developing rat CNS and in embryonic neurons in culture. The high fidelity of the staining of growth cones by mAb 2G13 means that the antibody will be a useful marker for identifying growth cones. In growth cones of cultured neurons, mAb 2G13 labelling is intracellular and mainly associated with the filamentous actin cytoskeleton. Experiments with cytochalasins, which depolymerise filamentous actin, showed that 2G13p (the protein recognised by mAb 2G13) is physically associated with filamentous actin in growth cones. These properties of 2G13p suggest a role in growth cone motility.  相似文献   

20.
The netrin-1 receptor Deleted in Colorectal Cancer (DCC) is required for the formation of major axonal projections by embryonic cortical neurons, including the corpus callosum, hippocampal commissure, and cortico-thalamic tracts. The presentation of DCC by axonal growth cones is tightly regulated, but the mechanisms regulating DCC trafficking within neurons are not well understood. Here, we investigated the mechanisms regulating DCC recruitment to the plasma membrane of embryonic cortical neurons. In embryonic spinal commissural neurons, protein kinase A (PKA) activation recruits DCC to the plasma membrane and enhances axon chemoattraction to netrin-1. We demonstrate that PKA activation similarly recruits DCC and increases embryonic cortical neuron axon extension, which, like spinal commissural neurons, respond to netrin-1 as a chemoattractant. We then determined if depolarization might recruit DCC to the plasma membrane. Neither netrin-1 induced axon extension, nor levels of plasma membrane DCC, were altered by depolarizing embryonic spinal commissural neurons with elevated levels of KCl. In contrast, depolarizing embryonic cortical neurons increased the amount of plasma membrane DCC, including at the growth cone, and increased axon outgrowth evoked by netrin-1. Inhibition of PKA, phosphatidylinositol-3-kinase, protein kinase C, or exocytosis blocked the depolarization-induced recruitment of DCC and suppressed axon outgrowth. Inhibiting protein synthesis did not affect DCC recruitment, nor were the distributions of trkB or neural cell adhesion molecule (NCAM) influenced by depolarization, consistent with selective mobilization of DCC. These findings identify a role for membrane depolarization modulating the response of axons to netrin-1 by regulating DCC recruitment to the plasma membrane.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号