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 125 毫秒
1.
To investigate the role of N-methyl-D-aspartate (NMDA) receptor activity in the stability of the presynaptic axon arbor and postsynaptic dendritic arbors in vivo, we took time-lapse confocal images of single DiI-labeled Xenopus retinotectal axons and optic tectal neurons in the presence and absence of the NMDA receptor antagonist, APV. Retinotectal axons or tectal neurons were imaged at 30-min intervals over 2 h, or twice over a 24-h period. Retinal axons in animals exposed to DL-APV (100 microM) showed an increase in rates of branch additions and a decrease in branch lifetimes over 2 h compared to untreated axons. Under the same experimental conditions, tectal neurons showed a decreased rate of branch tip additions and retractions. APV treatment over 24 h had no apparent effect on axon arbor morphology, but did decrease tectal cell dendritic arbor elaboration. These observations demonstrate that NMDA receptor activity in postsynaptic neurons stabilizes pre- and postsynaptic neuronal morphology in vivo.. However, when NMDA receptor activity is blocked, presynaptic retinal axons respond with increased arbor dynamics while postsynaptic tectal cell dendrites decrease arbor dynamics. Such differential responses of pre- and postsynaptic partners might increase the probability of coactive afferents converging onto a common target under conditions of lower NMDA receptor activity.  相似文献   

2.
In the developing retinotectal projection, retinal axon arbor structure changes rapidly within the target tectal neuropil at stages when the visual system functions to process visual information. In vivo imaging of single retinotectal axon arbors shows that up to 50% of the arbor branch length can be restructured within 8 h and short branchtips have average lifetimes of 10 min. To determine if presynaptic sites are restricted to the relatively stable part of the arbor or if they are also located on the more dynamic portions of the arbor, punctate staining of synaptic vesicle proteins (SVP) synapsin 1 and synaptophysin was mapped within individual retinal axons using double-label confocal immunocytochemistry. We report that SVP puncta were distributed throughout the retinotectal axon arbor. Notably, short branchtips, which are known to be extremely dynamic, contain the presynaptic machinery necessary for synaptic transmission. These data support a model in which activity-dependent mechanisms can influence presynaptic axon arbor morphology by modifying the rate of dynamic rearrangements of axonal branchtips. © 1998 John Wiley & Sons, Inc. J Neurobiol 35: 426–434, 1998.  相似文献   

3.
Diverse modes of axon elaboration in the developing neocortex   总被引:2,自引:2,他引:0       下载免费PDF全文
The development of axonal arbors is a critical step in the establishment of precise neural circuits, but relatively little is known about the mechanisms of axonal elaboration in the neocortex. We used in vivo two-photon time-lapse microscopy to image axons in the neocortex of green fluorescent protein-transgenic mice over the first 3 wk of postnatal development. This period spans the elaboration of thalamocortical (TC) and Cajal-Retzius (CR) axons and cortical synaptogenesis. Layer 1 collaterals of TC and CR axons were imaged repeatedly over time scales ranging from minutes up to days, and their growth and pruning were analyzed. The structure and dynamics of TC and CR axons differed profoundly. Branches of TC axons terminated in small, bulbous growth cones, while CR axon branch tips had large growth cones with numerous long filopodia. TC axons grew rapidly in straight paths, with frequent interstitial branch additions, while CR axons grew more slowly along tortuous paths. For both types of axon, new branches appeared at interstitial sites along the axon shaft and did not involve growth cone splitting. Pruning occurred via retraction of small axon branches (tens of microns, at both CR and TC axons) or degeneration of large portions of the arbor (hundreds of microns, for TC axons only). The balance between growth and retraction favored overall growth, but only by a slight margin. Given the identical layer 1 territory upon which CR and TC axons grow, the differences in their structure and dynamics likely reflect distinct intrinsic growth programs for axons of long projection neurons versus local interneurons.  相似文献   

4.
Javaherian A  Cline HT 《Neuron》2005,45(4):505-512
We have used in vivo time-lapse two-photon imaging of single motor neuron axons labeled with GFP combined with labeling of presynaptic vesicle clusters and postsynaptic acetylcholine receptors in Xenopus laevis tadpoles to determine the dynamic rearrangement of individual axon branches and synaptogenesis during motor axon arbor development. Control GFP-labeled axons are highly dynamic during the period when axon arbors are elaborating. Axon branches emerge from sites of synaptic vesicle clusters. These data indicate that motor neuron axon elaboration and synaptogenesis are concurrent and iterative. We tested the role of Candidate Plasticity Gene 15 (CPG15, also known as Neuritin), an activity-regulated gene that is expressed in the developing motor neurons in this process. CPG15 expression enhances the development of motor neuron axon arbors by promoting neuromuscular synaptogenesis and by increasing the addition of new axon branches.  相似文献   

5.
Fragments of injured axons that detach from their cell body break down by the molecularly regulated process of Wallerian degeneration (WD). Although WD resembles local axon degeneration, a common mechanism for refining neuronal structure, several previously examined instances of developmental pruning were unaffected by WD pathways. We used laser axotomy and time-lapse confocal imaging to characterize and compare peripheral sensory axon WD and developmental pruning in live zebrafish larvae. Detached fragments of single injured axon arbors underwent three stereotyped phases of WD: a lag phase, a fragmentation phase and clearance. The lag phase was developmentally regulated, becoming shorter as embryos aged, while the length of the clearance phase increased with the amount of axon debris. Both cell-specific inhibition of ubiquitylation and overexpression of the Wallerian degeneration slow protein (Wld(S)) lengthened the lag phase dramatically, but neither affected fragmentation. Persistent Wld(S)-expressing axon fragments directly repelled regenerating axon branches of their parent arbor, similar to self-repulsion among sister branches of intact arbors. Expression of Wld(S) also disrupted naturally occurring local axon pruning and axon degeneration in spontaneously dying trigeminal neurons: although pieces of Wld(S)-expressing axons were pruned, and some Wld(S)-expressing cells still died during development, in both cases detached axon fragments failed to degenerate. We propose that spontaneously pruned fragments of peripheral sensory axons must be removed by a WD-like mechanism to permit efficient innervation of the epidermis.  相似文献   

6.
Nitric oxide modulates retinal ganglion cell axon arbor remodeling in vivo   总被引:3,自引:0,他引:3  
Nitric oxide (NO) has been postulated to act as an activity-dependent retrograde signal that can mediate multiple aspects of synaptic plasticity during development. In the visual system, a role for NO in activity-dependent structural modification of presynaptic arbors has been proposed based on NO's ability to prune inappropriate projections and segregate axon terminals. However, evidence demonstrating that altered NO signaling does not perturb ocular dominance map formation leaves unsettled the role of NO during the in vivo refinement of visual connections. To determine whether NO modulates the structural remodeling of individual presynaptic terminal arbors in vivo we have: 1. Used NADPH-diaphorase histochemistry to determine the onset of NO synthase (NOS) expression in the Xenopus visual system. 2. Used in vivo time-lapse imaging to examine the role of NO during retinal ganglion cell (RGC) axon arborization. We show that NOS expression in the target optic tectum is developmentally regulated and localized to neurons that reside in close proximity to arborizing RGC axons. Moreover, we demonstrate that perturbations in tectal NO levels rapidly and significantly alter the dynamic branching of RGC arbors in vivo. Tectal injection of NO donors increased the addition of new branches, but not their stabilization in the long term. Tectal injection of NOS inhibitors increased the dynamic remodeling of axonal arbors by increasing branch addition and elimination and by lengthening pre-existing branches. Thus, these results indicate that altering NO signaling significantly modifies axon branch dynamics in a manner similar to altering neuronal activity levels (Cohen-Cory, 1999). Consequently, our results support a role for NO during the dynamic remodeling of axon arbors in vivo, and suggest that NO functions as an activity-dependent retrograde signal during the refinement of visual connections.  相似文献   

7.
Visual activity refines the retinotopic map formed on tectum during regeneration and development in goldfish through an N-methyl-D-aspartate (NMDA) receptor-mediated mechanism. Retinal arbors are enlarged in fish with unrefined maps. Here, we examined the effect of NMDA receptor blockers on the development of retinotectal arbors in zebrafish. Since visual behaviors begin 68-79 h postfertilization, we blocked NMDA receptors by immersion of larvae in MK801, AP5, or CPP starting at either 48 or 72 h. We then labeled axons with DiI at 72 or 96 h and examined them 5-9 h later. Arbors at 101-105 h (31 cases) were larger than at 77-79 h (11 cases): The average number of branches increased from 4.0 to 7.6 and the area (convex polygon method) increased by 42%. Blocking NMDA receptors with MK801 from 72 to 101-105 h significantly enlarged arbor size, but the number of branches remained roughly the same. The length and area of the arbors were both significantly increased (21% and 36%), whereas the width increased by a smaller amount (6%). This increase was reflected in longer distances between branches within the arbor (interbranch segments, +13%) as well as in the summed length of all branches (+28%). This selective effect on the extent but not number of branches is in agreement with our previous report of strobe effects in both developing and regenerating projections in goldfish, and supports the role of NMDA receptors in the first 24 h of synaptic transmission. We also used DiO to label arbors in time-lapse images taken at hourly intervals from 77 to 112 h. These sequences confirmed that individual arbors grew during this time, but showed that rates of branch addition and deletion and branch lifetimes were unaltered by the MK801 treatment. This is consistent with a simple model of random insertion of new branches and selective activity-driven elimination of those at the periphery to keep the normal arbor focused. Blocking NMDA receptors is postulated to randomize the elimination allowing the periphery to expand, thus accounting for the enlarged areas, without change in branch numbers or branch dynamics.  相似文献   

8.
Brain-derived neurotrophic factor (BDNF) modulates synaptic connectivity by increasing synapse number and by promoting activity-dependent axon arbor growth. Patterned neuronal activity is also thought to influence the morphological maturation of axonal arbors by directly influencing the stability of developing synapses. Here, we used in vivo time-lapse imaging to examine the relationship between synapse stabilization and axon branch stabilization, and to better understand the participation of BDNF in synaptogenesis. Green fluorescent protein (GFP)-tagged synaptobrevin II was used to visualize presynaptic specializations in individual DsRed2-labeled Xenopus retinal axons arborizing in the optic tectum. Neutralizing endogenous tectal BDNF with function-blocking antibodies significantly enhanced GFP-synaptobrevin cluster elimination, a response that was paralleled by enhanced branch elimination. Thus, synapse dismantling was associated with axon branch pruning when endogenous BDNF levels were reduced. To obtain a second measure of the role of BDNF during synapse stabilization, we injected recombinant BDNF in tadpoles with altered glutamate receptor transmission in the optic tectum. Tectal injection of the NMDA receptor antagonists APV or MK801 transiently induced GFP-synaptobrevin cluster dismantling, but did not significantly influence axon branch addition or elimination. BDNF treatment rescued synapses affected by NMDA receptor blockade: BDNF maintained GFP-synaptobrevin cluster density by maintaining their addition rate and rapidly inducing their stabilization. Consequently, BDNF influences synaptic connectivity in multiple ways, promoting not only the morphological maturation of axonal arbors, but also their stabilization, by a mechanism that influences both synapses and axon branches.  相似文献   

9.
Visual activity refines developing retinotectal maps and shapes individual retinal arbors via an NMDA receptor-dependent mechanism. As retinal axons grow into tectum, they slow markedly and emit many transient side branches behind the tip, assuming a "bottlebrush" morphology. Some branches are stabilized and branch further, giving rise to a compact arbor. The dynamic rate of branch addition and deletion is increased twofold when MK801 is used to block NMDA receptors, as if this prevents release of a stabilizing signal such as arachidonic acid (AA) from the postsynaptic neuron. In optic tract, AA mediates NCAM and L1 stimulation of axon growth by activating presynaptic protein kinase C (PKC) to phosphorylate GAP-43 and stabilize F-actin, and, if present in tectum, this growth control pathway could be modulated by postsynaptic activation. To test for the effects on arbor morphology of blocking PKC or AA release, we examined DiO-labeled retinal axons of larval zebrafish with time-lapse videomicroscopy. Bath application of the selective PKC inhibitor bisindolylmaleimide from 2 or 3 days onward doubled the rate at which side branches were added and deleted, as seen with MK801, and also prevented maturation of the arbor so that it retained a "bottlebrush" morphology. In order to selectively block the PKC being transported to retinal terminals, we injected the irreversible inhibitor calphostin C into the eye from which the ganglion cells were labeled, and this produced both effects seen with bath application. In contrast, there were no effects of control injections, which included Ringers into the same eye and the same dose into the opposite eye (actually much closer to the tectum of interest), to rule out the possibility that the inhibitor leaked from the eye to act on tectal cells. For comparison, we examined arbors treated with the NMDA blocker MK801 at half-hour time-lapse intervals, and detected the twofold rise in rates of branch addition and deletion previously reported in Xenopus larvae, but not the structural effect seen with the PKC inhibitors. In addition, we could produce both effects seen with PKC inhibitors by using RHC80267 to block AA release from DAG lipase, indicating that AA is the main drive for PKC activation. Thus, the results show a distinct role of AA and presynaptic PKC in both maturation of arbor structure and in the dynamic control of branching. The effects on branch dynamics were present regardless of the level of maturity of arbor structure. The fact that they mimicked those of MK801 suggests that presynaptic PKC may be involved in the NMDA receptor-driven stabilization of developing retinal arbors.  相似文献   

10.
Using in vivo rapid and long-interval two-photon time-lapse imaging of brain neuronal growth within the intact and unanesthetized Xenopus laevis tadpole, we characterize dynamic dendritic growth behaviors of filopodia, branches, and dendritic growth cones (DGCs), and analyze their contribution to persistent arbor morphology. The maturational progression of dynamic dendritogenesis was captured by short-term, 5 min interval, imaging for 1 h every day for 5 days, and the contribution of short-term growth to persistent structure was captured by imaging at 5 min intervals for 5 h, and at 2 h intervals for 10 h during the height of arbor growth. We find that filopodia and branch stability increases with neuronal maturation, and while the majority of dendritic filopodia rapidly retract, 3% to 7% of interstitial filopodia transition into persistent branches with lifetimes greater than 90 min. Here, we provide the first characterization of DGC dynamics, including morphology and behavior, in the intact and awake developing vertebrate brain. We find that DGCs occur on all growing branches indicating an essential role in branch elongation, and that DGC morphology correlates with dendritic branch growth behavior and varies with maturation. These results demonstrate that dendritogenesis involves a remarkable amount of continuous remodeling, with distinct roles for filopodia and DGCs across neuronal maturation. ? 2011 Wiley Periodicals, Inc. Develop Neurobiol 72: 615-627, 2012.  相似文献   

11.
During development, layer 2/3 neurons in the neocortex extend their axons horizontally, within the same layers, and stop growing at appropriate locations to form branches and synaptic connections. Firing and synaptic activity are thought to be involved in this process, but how neuronal activity regulates axonal growth is not clear. Here, we studied axonal growth of layer 2/3 neurons by exciting cell bodies or axonal processes in organotypic slice cultures of the rat cortex. For neuronal stimulation and morphological observation, plasmids encoding channelrhodopsin-2 (ChR2) and DsRed were coelectroporated into a small number of layer 2/3 cells. Firing activity induced by photostimulation (475 nm) was confirmed by whole-cell patch recording. Axonal growth was observed by time-lapse confocal microscopy, using a different excitation wavelength (560 nm), at 10–20-min intervals for several hours. During the first week in vitro, when spontaneous neuronal activity is low, DsRed- and ChR2-expressing axons grew at a constant rate. When high-frequency photostimulation (4 or 10 Hz) for 1 min was applied to the soma or axon, most axons paused in their growth. In contrast, lower-frequency stimulation did not elicit this pause behavior. Moreover, in the presence of tetrodotoxin, even high-frequency stimulation did not cause axonal growth to pause. These results indicate that increasing firing activity during development suppresses axon growth, suggesting the importance of neuronal activity for the formation of horizontal connections.  相似文献   

12.
The brain contains a complex network of axons rapidly communicating information between billions of synaptically connected neurons. The morphology of individual axons, therefore, defines the course of information flow within the brain. More than a century ago, Ramón y Cajal proposed that conservation laws to save material (wire) length and limit conduction delay regulate the design of individual axon arbors in cerebral cortex. Yet the spatial and temporal communication costs of single neocortical axons remain undefined. Here, using reconstructions of in vivo labelled excitatory spiny cell and inhibitory basket cell intracortical axons combined with a variety of graph optimization algorithms, we empirically investigated Cajal''s conservation laws in cerebral cortex for whole three-dimensional (3D) axon arbors, to our knowledge the first study of its kind. We found intracortical axons were significantly longer than optimal. The temporal cost of cortical axons was also suboptimal though far superior to wire-minimized arbors. We discovered that cortical axon branching appears to promote a low temporal dispersion of axonal latencies and a tight relationship between cortical distance and axonal latency. In addition, inhibitory basket cell axonal latencies may occur within a much narrower temporal window than excitatory spiny cell axons, which may help boost signal detection. Thus, to optimize neuronal network communication we find that a modest excess of axonal wire is traded-off to enhance arbor temporal economy and precision. Our results offer insight into the principles of brain organization and communication in and development of grey matter, where temporal precision is a crucial prerequisite for coincidence detection, synchronization and rapid network oscillations.  相似文献   

13.
The arbors of single axons terminating predominantly in layer IV of the representation of the hand in area 3b of owl monkeys were reconstructed from serial brain sections after axons beneath the cortex were severed and horseradish peroxidase was injected into the white matter. In addition to dense terminations in layer IV, these labeled axons generally had branches extending into deeper layer III, and a few had very sparse terminations in layer VI. Terminal arbors ranged from 100 to 900 μm in diameter, and fine branches with synaptic boutons were unevenly distributed, typically grouped in a large central cluster and one or more smaller side clusters. The results are consistent with three broad conclusions: (1) Since the arbors are large relative to the details of the somatotopic map in area 3b, all regions within a single arbor may not be equally effective in activating cortical cells. (2) Spatially separate branches of single axons may relate to spatially separate modules of neurons of the same class in a manner that allows them to receive the same inputs. (3) Many of the somatotopic changes that have been reported in the hand representation as a result of nerve manipulations in adults could result from alterations in synaptic effectiveness within the arbors of single axons.  相似文献   

14.
Li J  Erisir A  Cline H 《Neuron》2011,69(2):273-286
Dendrites, axons, and synapses are dynamic during circuit development; however, changes in microcircuit connections as branches stabilize have not been directly demonstrated. By combining in?vivo time-lapse imaging of Xenopus tectal neurons with electron microscope reconstructions of imaged neurons, we report the distribution and ultrastructure of synapses on individual vertebrate neurons and relate these synaptic properties to dynamics in dendritic and axonal arbor structure over hours or?days of imaging. Dynamic dendrites have a high density of immature synapses, whereas stable dendrites have sparser, mature synapses. Axons initiate contacts from multisynapse boutons on stable branches. Connections are refined by decreasing convergence from multiple inputs to postsynaptic dendrites and by decreasing divergence from multisynapse boutons to postsynaptic sites. Visual deprivation or NMDAR antagonists decreased synapse maturation and elimination, suggesting that coactive input activity promotes microcircuit development by concurrently regulating synapse elimination and maturation of remaining contacts.  相似文献   

15.
Visual activity refines developing retinotectal maps and shapes individual retinal arbors via an NMDA receptor‐dependent mechanism. As retinal axons grow into tectum, they slow markedly and emit many transient side branches behind the tip, assuming a “bottlebrush” morphology. Some branches are stabilized and branch further, giving rise to a compact arbor. The dynamic rate of branch addition and deletion is increased twofold when MK801 is used to block NMDA receptors, as if this prevents release of a stabilizing signal such as arachidonic acid (AA) from the postsynaptic neuron. In optic tract, AA mediates NCAM and L1 stimulation of axon growth by activating presynaptic protein kinase C (PKC) to phosphorylate GAP‐43 and stabilize F‐actin, and, if present in tectum, this growth control pathway could be modulated by postsynaptic activation. To test for the effects on arbor morphology of blocking PKC or AA release, we examined DiO‐labeled retinal axons of larval zebrafish with time‐lapse videomicroscopy. Bath application of the selective PKC inhibitor bisindolylmaleimide from 2 or 3 days onward doubled the rate at which side branches were added and deleted, as seen with MK801, and also prevented maturation of the arbor so that it retained a “bottlebrush” morphology. In order to selectively block the PKC being transported to retinal terminals, we injected the irreversible inhibitor calphostin C into the eye from which the ganglion cells were labeled, and this produced both effects seen with bath application. In contrast, there were no effects of control injections, which included Ringers into the same eye and the same dose into the opposite eye (actually much closer to the tectum of interest), to rule out the possibility that the inhibitor leaked from the eye to act on tectal cells. For comparison, we examined arbors treated with the NMDA blocker MK801 at half‐hour time‐lapse intervals, and detected the twofold rise in rates of branch addition and deletion previously reported in Xenopus larvae, but not the structural effect seen with the PKC inhibitors. In addition, we could produce both effects seen with PKC inhibitors by using RHC80267 to block AA release from DAG lipase, indicating that AA is the main drive for PKC activation. Thus, the results show a distinct role of AA and presynaptic PKC in both maturation of arbor structure and in the dynamic control of branching. The effects on branch dynamics were present regardless of the level of maturity of arbor structure. The fact that they mimicked those of MK801 suggests that presynaptic PKC may be involved in the NMDA receptor‐driven stabilization of developing retinal arbors. © 2003 Wiley Periodicals, Inc. J Neurobiol 58: 328–340, 2004  相似文献   

16.
Both blocking activity with intraocular tetrodotoxin (TTX) and synchronizing activity with a xenon strobe light (1 Hz) prevent retinotopic sharpening of regenerating optic projection in goldfish (Meyer, 1983; Schmidt, 1985; Cook and Rankin, 1986). In this study, we tested, in both normal and regenerating projections, the effects of these two treatments on individual optic arbors. Arbors were stained via anterograde transport of HRP, drawn in camera lucida from tectal whole mounts, and analyzed for spatial extent in the plane of the retinotopic map, order of branching, number of branch endings, depth of termination, and the caliber of the parent axon. In normal tectum, fine, medium, and coarse caliber axons gave rise to small, medium, and large arbors, which averaged 127 μm, 211 μm and 275 μm in horizontal extent, and terminated at characteristic depths. All three classes averaged roughly 21 branch endings. Optic arbors that regenerated with normal patterns of activity returned to a roughly normal appearance by 6–11 weeks postcrush: the same three calibers of axons gave rise to the same three sizes of arbors at the same depths, but they were much less stratified and were on average about 16% larger in horizontal extent. At this time point, arbors regenerated under TTX or strobe were on the average 71 and 119% larger, respectively, than the control-regenerated arbors (larger in all classes), although they had approximately the same number of branch endings and were equally poorly stratified. Synapses formed under strobe were also normal in appearance. Thus the only significant effect of both strobe and TTX treatment was to enlarge the spatial extent of arbor branches. Arbors that were not regenerating were very slightly (but significantly) enlarged by TTX block of activity or strobe illumination. As previous staining showed that regenerating axons initially make widespread branches and later retract many of those branches (Schmidt, Turcotte, Buzzard, and Tieman, 1988; Stuermer, 1988), the present findings support the idea that blocking activity or synchronizing activity prevents retinotopic sharpening by interfering with the elimination of some of the errant branches.  相似文献   

17.
The developing zebrafish retinotectal arbors make many trial branches with synapses but most are retracted. With NMDA blockers, branches are withdrawn at a higher rate, and a synapse on a branch not only stabilizes that branch, but biases new branches to form nearby. Here, we tested whether new branch formation requires the polarity complex, which is essential for organizing the cytoskeleton in initial axon formation. The complex (PAR3, PAR6, and atypical protein kinase C [aPKC]) is downstream of phosphatidyl‐inositol‐3‐kinase (PI3K), and its aPKC could be activated by retrograde arachidonic acid synaptic signaling. DiO‐labeled arbors in zebrafish were imaged on day 3 (before treatment) and 1–2 days after treatment to suppress or inhibit PAR3, PAR6, or PI3K. Intraocular antisense (AS) oligos to PAR3 or PAR6 both severely limited branch addition, which was most evident in arbors with few branches before treatment. As a result of the inability to branch, arbor segments grew longer than in controls. Both PI3K inhibition (LY294002) and AS suppression of PI3Kα and PI3Kδ isoforms likewise limited branch addition but also decreased growth, as the sum of segment lengths was below normal after 2 days. Both the results support the idea that the polarity complex and PI3K participate in arbor branch formation. The PKC inhibitor Go6983 also severely restricted branch addition and growth, as did bisindolyl‐maleimide and calphostin C reported previously, consistent with PKCζ, but not PKCµ, participation. These experiments suggest a mechanism whereby activity signaling could affect the branching of retinotectal arbors. © 2013 Wiley Periodicals, Inc. Develop Neurobiol 74: 591–601, 2014  相似文献   

18.
BACKGROUND: Trigeminal sensory neurons detect thermal and mechanical stimuli in the skin through their elaborately arborized peripheral axons. We investigated the developmental mechanisms that determine the size and shape of individual trigeminal arbors in zebrafish and analyzed how these interactions affect the functional organization of the peripheral sensory system. RESULTS: Time-lapse imaging indicated that direct repulsion between growing axons restricts arbor territories. Removal of one trigeminal ganglion allowed axons of the contralateral ganglion to cross the midline, and removal of both resulted in the expansion of spinal cord sensory neuron arbors. Generation of embryos with single, isolated sensory neurons resulted in axon arbors that possessed a vast capacity for growth and expanded to encompass the entire head. Embryos in which arbors were allowed to aberrantly cross the midline were unable to respond in a spatially appropriate way to mechanical stimuli. CONCLUSIONS: Direct repulsive interactions between developing trigeminal and spinal cord sensory axon arbors determine sensory neuron organization and control the shapes and sizes of individual arbors. This spatial organization is crucial for sensing the location of objects in the environment. Thus, a combination of undirected growth and mutual repulsion results in the formation of a functionally organized system of peripheral sensory arbors.  相似文献   

19.
Targeting of axons and dendrites to particular synaptic laminae is an important mechanism by which precise patterns of neuronal connectivity are established. Although axons target specific laminae during development, dendritic lamination has been thought to occur largely by pruning of inappropriately placed arbors. We discovered by in vivo time-lapse imaging that retinal ganglion cell (RGC) dendrites in zebrafish show growth patterns implicating dendritic targeting as a mechanism for contacting appropriate synaptic partners. Populations of RGCs labeled in transgenic animals establish distinct dendritic strata sequentially, predominantly from the inner to outer retina. Imaging individual cells over successive days confirmed that multistratified RGCs generate strata sequentially, each arbor elaborating within a specific lamina. Simultaneous imaging of RGCs and subpopulations of presynaptic amacrine interneurons revealed that RGC dendrites appear to target amacrine plexuses that had already laminated. Dendritic targeting of prepatterned afferents may thus be a novel mechanism for establishing proper synaptic connectivity.  相似文献   

20.
In the embryo of the leech Hirudo medicinalis, afferent projections of peripheral sensory neurons travel along common nerve tracts to the CNS, where they defasciculate, branch, and arborize into separate, modality-specific synaptic laminae. Previous studies have shown that this process requires, at least in part, the constitutive and then modality-specific glycosylations of tractin, a leech L1 homologue. We report here on the dynamics of growth of these projections as obtained by examining the morphology of single growing dye-filled sensory afferents as a function of time. Using 2-photon laser-scanning microscopy of the intact developing embryo, we obtained images of individual sensory projections at 3 to 30 min intervals, over several hours of growth, and at different stages of development. The time-lapse series of images revealed a highly dynamic and maturation-state-dependent pattern of growth. Upon entering the CNS, the growth cone-tipped primary axon sprouted numerous long filopodial processes, many of which appeared to undergo repeated cycles of extension and retraction. The growth cone was transformed into a sensory arbor through the formation of secondary branches that extended within the ganglionic neuropil along the anterior-posterior axis of the CNS. Numerous tertiary and quaternary processes grew from these branches and also displayed cycles of extension and retraction. The motility of these higher-order branches changed with age, with younger afferents displaying higher densities and greater motility than older, more mature sensory arbors. Finally, coincident with a reduction in higher order projections was the appearance of concavolar structures on the secondary processes. Rows of these indentations suggest the formation of presynaptic en-passant specializations accompanying the developmental onset of synapse formation.  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