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.
Photoreceptors need the support of pigment epithelial (PE) and Müller glial cells in order to maintain visual sensitivity and neurotransmitter resynthesis. In rod outer segments (ROS), all-trans-retinal is transformed to all-trans-retinol by retinol dehydrogenase using NADPH. NADPH is restored in ROS by the pentose phosphate pathway utilizing high amounts of glucose supplied by choriocapillaries. The retinal formed is transported to PE cells where regeneration of 11-cis-retinal occurs. Müller cells take up and metabolize glucose predominantly to lactate which is massively released into the extracellular space (ES). Lactate is taken up by photoreceptors, where it is transformed to pyruvate which, in turn, enters the Krebs cycle in mitochondria of the inner segment. Stimulation of neurotransmitter release by darkness induces 130% rise in the amount of glutamate released into ES. Glutamate is transported into Müller cells where it is predominantly transformed to glutamine. Stimulation of photoreceptors induces an eightfold increase in glutamine formation. It appears, therefore, that there is a signaling function in the transfer of amino acids from Müller cells to photoreceptors. Work on the model-system of the honeybee retina demonstrated that photoreceptors release NH4+ and glutamate in a stimulus-dependent manner which, in turn, contribute to the biosynthesis of alanine in glia. Alanine released into the extracellular space is taken up and used by photoreceptors. Glial cells take glutamate by high-affinity transporters. This uptake induces a transient change in glial cell metabolism. The transformation of glutamate to glutamine is possibly also controlled by the uptake of NH4+ which directly affects cellular metabolism.  相似文献   

2.
Mammalian astrocytes convert glutamate to glutamine and bee retinal glial cells convert pyruvate to alanine. To maintain such amination reactions these glial cells may take up NH4+/NH3. We have studied the entry of NH4+/NH3 into bundles of glial cells isolated from bee retina by using the fluorescent dye BCECF to measure pH. Ammonium caused intracellular pH to decrease by a saturable process: the rate of change of pH was maximal for an ammonium concentration of about 5 mM. This acidifying response to ammonium was abolished by the loop diuretic bumetanide (100 microM) and by removal of extracellular Cl-. These results strongly suggest that ammonium enters the cell by contransport of NH4+ with Cl-. Removal of extracellular Na+ did not abolish the NH(4+)-induced acidification. The NH(4+)-induced pH change was unaffected when nearly all K+ conductance was blocked with 5 mM Ba2+ showing that NH4+ did not enter through Ba(2+)-sensitive ion channels. Application of 2 mM NH4+ led to a large increase in total intracellular proton concentration estimated to exceed 13.5 mEq/L. As the cell membrane appeared to be permeable to NH3, we suggest that when NH4+ entered the cells, NH3 left, so that protons were shuttled into the cell. This shuttle, which was strongly dependent on internal and external pH, was quantitatively modelled. In retinal slices, 2 mM NH4+ alkalinized the extracellular space: this alkalinization was reduced in the absence of bath Cl-. We conclude that NH4+ enters the glial cells in bee retina on a cotransporter with functional similarities to the NH4+(K+)-Cl- cotransporter described in kidney cells.  相似文献   

3.
To gain insight into neuronal-glial signaling in brain, cerebellar Bergmann glia and granule neurons were studied in acutely isolated slices with the aid of laser scanning confocal microscopy. Both Bergmann glia and granule neurons responded to N-methyl-D-aspartate (NMDA) with a rise in [Ca2+]i. However, the glial NMDA response was frequently inhibited by tetrodotoxin, suggesting that the response depended on neuronal action potentials, rather than on direct activation of NMDA receptors on the Bergmann glia. Further experiments demonstrated that the NMDA response in Bergmann glia was not inhibited by a combination of non-NMDA glutamate receptor blockers 6-cyano-7-nitroquinoxaline-2,3-dione and alpha-methyl-4-carboxyphenylglycine. Bergmann glia also responded to norepinephrine and high K+, and the responses were not inhibited by tetrodotoxin. The glial norepinephrine response was blocked by phentolamine but not by the removal of external Ca2+, indicating a direct activation of alpha1-adrenergic receptors that mediated release of Ca2+ from intracellular stores. The KCl-induced response in both neurons and glia was dependent on external Ca2+ and was blocked by verapamil or nifedipine. In summary, our data indicate that Bergmann glia in situ recognize a signal(s) released from neurons during neuronal activity.  相似文献   

4.
Glutamate is the most abundant excitatory amino acid in the central nervous system. It has also been described as a potent toxin when present in high concentrations because excessive stimulation of its receptors leads to neuronal death. Glial influence on neuronal survival has already been shown in the central nervous system, but the mechanisms underlying glial neuroprotection are only partly known. When cells isolated from newborn rat retina were maintained in culture as enriched neuronal populations, 80% of the cells were destroyed by application of excitotoxic concentrations of glutamate. Massive neuronal death was also observed in newborn retinal cultures containing large numbers of glia, or when neurons were seeded onto feeder layers of purified cells prepared from immature (postnatal 8 day) rat retina. When newborn retinal neurons were seeded onto feeder layers of purified glial cells prepared from adult retinas, application of excitotoxic amino acids no longer led to neuronal death. Furthermore, neuronal death was not observed in mixed neuron/glial cultures prepared from adult retina. However, in all cases (newborn and adult) application of kainate led to amacrine cell-specific death. Activity of glutamine synthetase, a key glial enzyme involved in glutamate detoxification, was assayed in these cultures in the presence or absence of exogenous glutamate. Whereas pure glial cultures alone (from young or adult retina) showed low activity that was not stimulated by glutamate addition, mixed or co-cultured neurons and adult glia exhibited up to threefold higher levels of activity following glutamate treatment. These data indicate that two conditions must be satisfied to observe glial neuroprotection: maturation of glutamine synthetase expression, and neuron-glial signalling through glutamate-elicited responses.  相似文献   

5.
6.
Glial-neuronal communication was studied by monitoring the effect of intercellular glial Ca2+ waves on the electrical activity of neighboring neurons in the eyecup preparation of the rat. Calcium waves in astrocytes and Müller cells were initiated with a mechanical stimulus applied to the retinal surface. Changes in the light-evoked spike activity of neurons within the ganglion cell layer occurred when, and only when, these Ca2+ waves reached the neurons. Inhibition of activity was observed in 25 of 53 neurons (mean decrease in spike frequency, 28 +/- 2%). Excitation occurred in another five neurons (mean increase, 27 +/- 5%). Larger amplitude Ca2+ waves were associated with greater modulation of neuronal activity. Thapsigargin, which reduced the amplitude of the glial Ca2+ increases, also reduced the magnitude of neuronal modulation. Bicuculline and strychnine, inhibitory neurotransmitter antagonists, as well as 6-Nitro-7-sulphamoylbenzo[f]quinoxaline-2,3-dione (NBQX) and D(-)-2-amino-7-phosphonoheptanoic acid (D-AP7), glutamate antagonists, reduced the inhibition of neuronal activity associated with glial Ca2+ waves, suggesting that inhibition is mediated by inhibitory interneurons stimulated by glutamate release from glial cells. The results suggest that glial cells are capable of modulating the electrical activity of neurons within the retina and thus, may directly participate in information processing in the CNS.  相似文献   

7.
Excessive activation of N-methyl-D-aspartate (NMDA) receptor channels (NRs) is a major cause of neuronal death associated with stroke and ischemia. Cerebellar granule neurons in vivo, but not in culture, are relatively resistant to toxicity, possibly owing to protective effects of glia. To evaluate whether NR-mediated signaling is modulated when developing neurons are cocultured with glia, the neurotoxic responses of rat cerebellar granule cells to applied NMDA or glutamate were compared in astrocyte-rich and astrocyte-poor cultures. In astrocyte-poor cultures, significant neurotoxicity was observed in response to NMDA or glutamate and was inhibited by an NR antagonist. Astrocyte-rich neuronal cultures demonstrated three significant differences, compared with astrocyte-poor cultures: (a) Neuronal viability was increased; (b) glutamate-mediated neurotoxicity was decreased, consistent with the presence of a sodium-coupled glutamate transport system in astrocytes; and (c) NMDA- but not kainate-mediated neurotoxicity was decreased, in a manner that depended on the relative abundance of glia in the culture. Because glia do not express NRs or an NMDA transport system, the mechanism of protection is distinct from that observed in response to glutamate. No differences in NR subunit composition (evaluated using RT-PCR assays for NR1 and NR2 subunit mRNAs), NR sensitivity (evaluated by measuring NR-mediated changes in intracellular Ca2+ levels), or glycine availability as a coagonist (evaluated in the presence and absence of exogenous glycine) were observed between astrocyte-rich and astrocyte-poor cultures, suggesting that glia do not directly modulate NR composition or function. Nordihydroguaiaretic acid, a lipoxygenase inhibitor, blocked NMDA-mediated toxicity in astrocyte-poor cultures, raising the possibility that glia effectively reduce the accumulation of highly diffusible and toxic arachidonic acid metabolites in neurons. Alternatively, glia may alter neuronal development/phenotype in a manner that selectively reduces susceptibility to NR-mediated toxicity.  相似文献   

8.
9.
Cells from major types of gliomas, i.e. oligodendrogliomas and glioblastomas, are able to generate action potentials upon a current injection similar to neurons (Patt et al. (1996) Neuroscience, 71, 601-611; Labrakakis et al. (1997b) J. Neuropath. Exp. Neurol., 56, 243-254. Here, we report that activation of ionotropic glutamate receptors by the selective agonist, kainate, or by glutamate itself, depolarized the tumour cells in culture and living slices from tumour tissue, and can elicit volleys of action potentials, as recorded with the patch-clamp technique. Sixty-six percent of the glioblastoma cells, 44% of the astocytoma and 86% of the oligodendroglioma cells responded to glutamate and the specific agonist of AMPA/kainate receptors, kainate. The involvement of non-NMDA (N-methyl-D-aspartate) receptors is further supported by the observation that both kainate and glutamate currents were blocked by CNQX (6-cyano-7-nitroquinoxaline-2,3-dione). The receptor activation was accompanied by an increase in cytosolic Ca2+, as recorded with a fura-2 microfluorometric system. The Ca2+ elevation was mediated by the activation of Ca2+ channels due to membrane depolarization. The presence of voltage-gated Ca2+ channels was confirmed by patch-clamp experiments. Taken together, these findings imply that the electrophysiological properties of glioma cells are more reminiscent of those of neurons than of glial cells.  相似文献   

10.
Previous studies have shown that complete blockade of metabolism in embryonic chick retina causes a time-dependent increase in the release of glutamate into the extracellular space. The present study examined the cellular source of this glutamate, i.e., neuronal and/or glial. Pure cultures of retinal neurons or glia were labeled for 10 min at 37 degrees C with [3H]acetate. Retinal glia, but not retinal neurons, were found to selectively and preferentially metabolize acetate, thus producing 3H-labeled amino acids in the glial compartment. This finding provides direct evidence to substantiate findings from several other laboratories that have indirectly determined the preferential metabolism of acetate by glia by using mixed neuronal/glial populations. To study the cellular source of glutamate released during total metabolic blockade, whole retina were prelabeled with [3H]acetate plus [U-14C]glucose (to label the neuronal compartment). Total metabolic blockade was instituted with a combination of iodoacetate (IOA) plus KCN, and the release of glutamate into the medium was followed at 5, 15, and 30 min. During total energy blockade, net extracellular glutamate was not elevated at 5 min [0.17 +/- 0.02 vs. 0.12 +/- 0.01 microM for treated vs. control retina (means +/- SEM), respectively], but was increased significantly at 15 (1.2 +/- 0.26 microM) and 30 min (2.6 +/- 0.22 microM). Total [3H]glutamate in the medium during IOA/KCN treatment was unchanged at 5 min, but was increased 1.5- and threefold above basal levels at 15 and 30 min, respectively. During the time when extracellular glutamate increased, the specific activity of [3H]glutamate remained fairly constant, 731 +/- 134 and 517 +/- 82 dpm/nmol (means +/- SEM) at 15 and 30 min, respectively. In contrast, 14C-labeled glutamate in the medium did not increase during IOA/KCN treatment and paralleled basal levels. Thus, the specific activity of 14C-labeled extracellular glutamate decreased from 309 +/- 87 dpm/nmol at 15 min to 42 +/- 8 dpm/nmol at 30 min. Prior loading of the tissue with 0.5 mM trans-pyrrolidine-2,4-dicarboxylate (t-PDC), a glutamate transport inhibitor, blocked 57% of the glutamate released at 30 min of IOA/KCN exposure, suggesting that reversal of an Na+-dependent glutamate transporter was a key contributor to the appearance of extracellular glutamate during energy deprivation. The increase in extracellular [3H]glutamate, constancy of the specific activity of extracellular [3H]glutamate, decrease in the specific activity of extracellular [14C]glutamate, and attenuation of release by prior loading with t-PDC indicate that glial pools of glutamate released via reversal of the transporter contribute significantly to the rise in extracellular glutamate after metabolic inhibition in this preparation.  相似文献   

11.
In the last few years, there has been an important increase in interest in nitric oxide (NO) as an intercellular messenger, and its putative role in numerous CNS functions is being continually updated. Arginine, the nitric oxide precursor, has been found in our laboratory to be released following stimulation of the white matter in the cerebellum and of sensory afferents in the thalamus. Since arginine is localized in glial cells while the nitric oxide synthesizing enzyme is localized in different cells (predominantly in neurons), these findings may represent a transfer of arginine from glia to neurons in order to supply the nitric oxide synthase with its substrate. The mechanism underlying this glial-neuronal interaction seems to involve the activation of excitatory amino acid receptors present on glial cells. Our results speak for an intense crosstalk between neurons and glia (activation of glial receptors by neurotransmitter released from neurons) and between glia and neurons (supply of the nitric precursor arginine from glia to neurons). The form in which NO is released from cells has been much debated. The chemical identity of the endothelial-derived relaxing factor in particular is still a matter of dispute, the major contender being NO. and a S-nitrosothiol compound. Based on the strong reactivity of NO for thiols and on the presence of cysteine and glutathione at the mM level intracellularly and microM level extracellularly, we have investigated whether S-nitrosothiols, i.e. S-nitrosoglutathione, may be the potential "package" form in which NO could be stored. We demonstrated, with HPLC coupled to mass spectrometry techniques, the presence of endogenous nitrosoglutathione in rat brain tissue. This packaging of NO in the form of nitrosothiols might serve to facilitate its transfer, prolong its life, and target its delivery to specific effectors. That could confer a specificity of action to the widely diffusable messenger NO, may determine the range of effectiveness of NO and mitigate its adverse cytotoxic effects.  相似文献   

12.
In order to investigate the role of neuron-glia interactions in the response of astroglial to a non-invasive cerebellar cortex injury, we have used two cases of the ataxic form of Creutzfeldt-Jakob disease (CJD) with distinct neuronal loss and diffuse astrogliosis. The quantitative study showed no changes in cell density of either Purkinje or Bergmann glial cells in CJ-1, whereas in the more affected CJ-2 a loss of Purkinje cells and an increase of Bergmann glial cells was found. The granular layer in both CJD cases showed a similar loss of granule cells (about 60%) in parallel with the significant increase in GFAP+ reactive astrocytes. GFAP immunostaining revealed greater reactivity of Bergmann glia in CJ-2 than in CJ-1, as indicated by the thicker glial processes and the higher optical density. Granular layer reactive astrocytes were regularly spaced. In both CJD cases there was strict preservation of the spatial arrangement of all astroglial subtypes--Fa?anas cells, Bergmann glia and granular layer astrocytes. Reactive Fa?anas and Bergmann glial cells and microglia/macrophages expressed vimentin, while only a few vimentin+ reactive astrocytes were detected in the granular layer. Karyometric analysis showed that the increase in nuclear volume in reactive astroglia was directly related with the level of glial hypertrophy. The number of nucleoli per nuclear section was constant in astroglial cells of human controls and CJD, suggesting an absence of polyploidy in reactive astroglia. Ultrastructural analysis revealed junctional complexes formed by the association of macula adherens and gap junctions. In the molecular layer numerous vacant dendritic spines were ensheathed by lamellar processes of reactive Bergmann glia. Our results suggest that quantitative (neuron/astroglia ratio) and qualitative changes in the interaction of neurons with their region-specific astroglial partners play a central role in the astroglial response pattern to the pathogenic agent of CJD.  相似文献   

13.
Increased extracellular glutamate levels are related to glial and neuronal damage. Glutamate-mediated toxicity is limited by glial uptake and metabolic transformation of glutamate to glutamine and the energetic compounds alanine and lactate which are utilized by surrounding neurons. Under in vitro conditions, barbiturates have been shown to reduce glutamate uptake and its further metabolism, possibly impeding metabolic coupling between astrocytes and neurons. The aims were to investigate if under clinical conditions, the barbiturate thiopental reduces important detoxification of glutamate, resulting in lower CSF glutamine, alanine and lactate levels as opposed to patients receiving midazolam. During long-term administration of thiopental and midazolam, pathologically elevated ventricular CSF glutamate levels were associated with significantly increased glutamine and alanine levels up to 14 days after trauma. CSF lactate, however, remained normal. These data suggest that long-term administration of thiopental and midazolam under clinical conditions does not impede enzymatic activities responsible for detoxification and metabolism of glutamate.  相似文献   

14.
Because glial cells have been characterized by electrophysiological studies as being silent and inactive, neuroscientists have overlooked the roles of these cells in the dynamic function of the central nervous system. Recent measurements of intracellular Ca2+ concentration, however, revealed the dynamic and active features of the glial cells. Some populations of these cells gave rise to increase in intracellular Ca2+ concentration by stimulation with neurotransmitters such as glutamate, acetylcholine, serotonin, noradrenaline and histamine through the activation of specific neurotransmitter receptors distributed on the glial cells. Although the roles of the increased intracellular Ca2+ concentration in glial cells have not been elucidated, these properties suggested the functional participation of glial cells in synaptic modulation and plasticity.  相似文献   

15.
Previous studies showed that in cultured chick ciliary ganglion neurons and CNS glia, adenosine can be synthesized by hydrolysis of 5'-AMP and that the accumulation of the adenosine degradative products inosine and hypoxanthine was significantly greater in glial than in neuronal cultures. Furthermore, previous immunochemical and histochemical studies in brain showed that adenosine deaminase and nucleoside phosphorylase are localized in endothelial and glial cells but are absent in neurons; however, adenosine deaminase may be found in a few neurons in discrete brain regions. These results suggested that adenosine degradative pathways may be more active in glia. Thus, we have determined if there is a differential distribution of adenosine deaminase, nucleoside phosphorylase, and xanthine oxidase enzyme fluxes in glia, comparing primary cultures of central and ciliary ganglion neurons and glial cells from chick embryos. Hypoxanthine-guanine phosphoribosyltransferase and production of adenosine by S-adenosylhomocysteine hydrolase activity were also examined. Our results show that there is a distinct profile of purine metabolizing enzymes for glia and neurons in culture. Both cell types have an S-adenosylhomocysteine hydrolase, but it was more active in neurons than in glia. In contrast, in glia the enzymatic activities of xanthine oxidase (443 +/- 61 pmol/min/10(7) cells), nucleoside phosphorylase (187 +/- 8 pmol/min/10(7) cells), and adenosine deaminase (233 +/- 32 pmol/min/10(7) cells) were more active at least 100, 20, and five times, respectively, than in ciliary ganglion neurons and 100, 100, and nine times, respectively, than in central neurons.  相似文献   

16.
Radial glial cells, present in many parts of the embryonic vertebrate central nervous system (CNS), have been implicated in the guidance of neuroblasts from the ventricular zone to their laminar destinations. Moreover, radial glial cells may be progenitors of some CNS neurons and glia. To gain new insight into the structure and development of these cells, we have generated and characterized a panel of monoclonal antibodies that recognize radial glial cells of the chick optic tectum. Mice were immunized with homogenates of embryonic day (E) 10 tectum, and antibodies were analyzed by immunofluorescence and immunoblotting. We describe here three pairs of antibodies. 1) H5 and a previously generated antibody, R5 (Dr?ger et al., J. Neurosci. 4:2025, 1984), stain the whole extent of the radial glial cell from E7 to E20. In cultures prepared from E10 tecta, both stain a filamentous meshwork in glial cells but not in neurons. On immunoblots, both recognize a protein of approximately 52 kD that is closely related (or identical) to vimentin. 2) H28 and H29 stain radial glia between E7 and E14, but not later. Moreover, H28 and H29 staining is markedly more intense in the ventricular and intermediate zones than in the laminae of the tectal plate. Both of these antibodies recognize an intracellular epitope in cultured glial cells and a protein of approximately 35 kD on immunoblots. 3) H2 and H27 recognize antigens concentrated in the most superficial processes and endfeet of radial glia in late (E16-E20) embryos. They stain distinct structures in cultured glia, suggesting that they recognize distinct antigens. H27 recognizes a protein of approximately 29 kD on immunoblots. Thus antibodies H5 and R5 are good markers of radial glial cells at all stages, whereas the others define antigens that are developmentally regulated and localized to discrete domains. Together, these antibodies can be used to study temporal and spatial specializations of radial glia.  相似文献   

17.
As a first step in understanding how the supply of the neurotransmitter histamine is maintained in a photoreceptor, we followed the uptake and metabolism of the immediate precursor of histamine, histidine. [3H]Histidine taken up into photoreceptors and glia was detected using autoradiography, and synthesis of [3H]histamine from [3H]histidine was assayed with thin-layer chromatography. Photoreceptors from barnacles were pulsed (15 min) with [3H]histidine (0.2-200 microM), then maintained in normal saline for up to 24 hr. Autoradiography showed that photoreceptor somata, axons, and presynaptic arbors were labeled, but only weakly, like (nonhistaminergic) ganglion cells. Label instead was concentrated over surrounding glia. Stimulating preparations with light did not increase photoreceptor labeling. Grain counts from photoreceptor axons showed uptake of [3H]histidine into these neurons by a Na+-dependent mechanism with a Km of approximately 50 microM. Over 24 hr only 1% of the [3H]histidine taken up by preparations was converted to [3H]histamine either in the dark or in the light. Injections of [3H]histidine directly into photoreceptors established that synthesis takes place within the photoreceptors and confirmed that stimulation with light did not measurably affect the rate of conversion of [3H]histidine to [3H]histamine. These results suggest that de novo synthesis of transmitter is unlikely to be as important as its reuptake in maintaining neurotransmitter supply in these photoreceptor terminals. In support of this conclusion, photoreceptors accumulated more label when transmitter release was stimulated with high K+ and histamine uptake was antagonized with chlorpromazine.  相似文献   

18.
Alanine dehydrogenase [EC 1. 4. 1. 1] was purified to homogeneity from a crude extract of Enterobacter aerogenes ICR 0220. The enzyme had a molecular mass of about 245 kDa and consisted of six identical subunits. The enzyme showed maximal activity at about pH 10.9 for the deamination of L-alanine and at about pH 8.7 for the amination of pyruvate. The enzyme required NAD+ as a coenzyme. Analogs of NAD+, deamino-NAD+ and nicotinamide guanine dinucleotide served as coenzymes. Initial-velocity and product inhibition studies suggested that the deamination of L-alanine proceeded through a sequential ordered binary-ternary mechanism. NAD+ bound first to the enzyme, followed by L-alanine, and the products were released in the order of ammonia, pyruvate, and NADH. The Km were 0.47 mM for L-alanine, 0.16 mM for NAD+, 0.22 mM for pyruvate, 0.067 mM for NADH, and 66.7 mM for ammonia. The Km for L-alanine was the smallest in the alanine dehydrogenases studied so far. The enzyme gene was cloned into Escherichia coli JM109 cells and the nucleotides were sequenced. The deduced amino acid sequence was very similar to that of the alanine dehydrogenase from Bacillus subtilis. However, the Enterobacter enzyme has no cysteine residue. In this respect, the Enterobacter enzyme is different from other alanine dehydrogenases.  相似文献   

19.
No enteric neurons or glia develop in the gut below the rostral foregut in mice lacking glial cell line-derived neurotrophic factor (GDNF) or Ret. We analyzed the nature and age dependence of the effects of GDNF and, for comparison, those of NT-3, on the in vitro development of the precursors of enteric neurons and glia. Positive and negative immunoselection with antibodies to p75(NTR) were used to isolate crest-derived and crest-depleted populations of cells from the fetal rat bowel at E12, 14, and 16. Cells were typed immunocytochemically. GDNF stimulated the proliferation of nestin-expressing precursor cells isolated at E12, but not at E14-16. GDNF promoted the development of peripherin-expressing neurons (E12 > E14-16) and expression of TrkC. GDNF inhibited expression of S-100-expressing glia at E14-16. NT-3 did not affect cells isolated at E12, never stimulated precursors to proliferate, and promoted glial as well as neuronal development at E14-16. GFRalpha-1 was expressed both by crest- and non-crest-derived cells, although only crest-derived cells anchored GFRalpha-1 and GFRalpha-2 (GFRalpha-1 > GFRalpha-2). GDNF increased the number of neurons anchoring GFRalpha-1. GFRalpha-1 is immunocytochemically detectable in neurons of the E13 intestine and persists in adult neurons of both plexuses. We suggest that GDNF stimulates the proliferation of an early (E12) NT-3-insensitive precursor common to enteric neurons and glia; by E14, this common precursor is replaced by specified NT-3-responsive neuronal and glial progenitors. GDNF exerts a neurotrophic, but not a mitogenic, effect on the neuronal progenitor. The glial progenitor is not maintained by GDNF.  相似文献   

20.
Sodium-dependent transport into astrocytes is critical for maintaining the extracellular concentrations of glutamate below toxic levels in the central nervous system. In this study, the expression of the glial glutamate transporters GLT-1 and GLAST was studied in primary cultures derived from cortical tissue. In primary astrocytes, GLAST protein levels were approximately one half of those observed in cortical tissue, but GLT-1 protein was present at very low levels compared with cortical tissue. Maintenance of these astrocytes in medium supplemented with dibutyryl-cAMP (dbcAMP) caused a dramatic change in cell morphology, increased GLT-1 and GLAST mRNA levels approximately 5-fold, increased GLAST protein approximately 2-fold, and increased GLT-1 protein >/=8-20-fold. These increases in protein expression were accompanied by 2-fold increases in the Vmax and Km values for Na+-dependent L-[3H]glutamate transport activity. Although GLT-1 is sensitive to inhibition by dihydrokainate in heterologous expression systems, no dihydrokainate sensitivity was observed in astrocyte cultures that expressed GLT-1. Biotinylation with a membrane-impermeant reagent, separation of the biotinylated/cell surface proteins, and subsequent Western blotting demonstrated that both GLT-1 and GLAST were present at the cell surface. Coculturing of astrocytes with neurons also induced expression of GLT-1, which colocalized with the glial specific marker, glial fibrillary acidic protein. Neurons induced a small increase in GLAST protein. Several studies were performed to examine the mechanism by which neurons regulate expression of the glial transporters. Three different protein kinase A (PKA) antagonists did not block the effect of neurons on glial expression of GLT-1 protein, but the addition of dbcAMP to mixed cultures of neurons and astrocytes did not cause GLT-1 protein to increase further. This suggests that neurons do not regulate GLT-1 by activation of PKA but that neurons and dbcAMP regulate GLT-1 protein through convergent pathways. As was observed with GLT-1, the increases in GLAST protein observed in cocultures were not blocked by PKA antagonists, but unlike GLT-1, the addition of dbcAMP to mixed cultures of neurons and astrocytes caused GLAST protein to increase approximately 2-fold. Neurons separated from astrocytes with a semipermeable membrane increased GLT-1 protein, indicating that the effect of neurons was mediated by a diffusible molecule. Treatment of cocultures with high concentrations of either N-methyl-D-aspartate or glutamate killed the neurons, caused GLT-1 protein to decrease, and caused GLAST protein to increase. These studies suggest that GLT-1 and GLAST protein are regulated independently in astrocyte cultures and that a diffusible molecule secreted by neurons induces expression of GLT-1 in astrocytes.  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