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.
P A Rosenberg 《Glia》1991,4(1):91-100
The function of astrocytes in cerebral cortex may be studied by comparing the properties of conventional, astrocyte-rich cultures with astrocyte-poor cultures in which astrocyte proliferation has been stringently suppressed. Exposure of astrocyte-poor, but not astrocyte-rich, cultures to fresh medium containing 2 mM glutamine resulted in the death of most neurons within 24 h. This study was undertaken to understand the basis for the apparent toxicity of glutamine in astrocyte-poor cultures. The toxicity of glutamine was found to be mediated by glutamate, which demonstrated an LD50 as a neurotoxin in astrocyte-poor cultures of 2 microM. Exposure of astrocyte-poor (but not astrocyte-rich) cultures to fresh medium containing glutamine for 17.5-24 h resulted in the accumulation of substantial quantities of glutamate (255 +/- 158 microM; mean +/- standard deviation) coincident with the death of neurons in the cultures. Exposure of astrocyte-poor cultures to glutamate in the absence of glutamine did not result in the accumulation of extracellular glutamate. Both the neuronal death and the extracellular glutamate accumulation in astrocyte-poor cultures exposed to glutamine could be blocked by N-methyl-D-aspartate (NMDA) antagonists. These observations suggest that astrocytes as well as glutamine may play an important role in the pathogenesis of glutamate neurotoxicity in the central nervous system.  相似文献   

2.
The pharmacological properties of glutamate agonists were compared in astrocyte-rich and astrocyte-poor cultures derived from embryonic rat cerebral cortex. The object of this investigation was to determine the extent to which glutamate uptake might influence the receptor-mediated neurotoxic actions of these compounds. In astrocyte-rich cultures, using 30 min exposures, we observed that the potencies of the poorly transported agonists NMDA (35 microM) and D-glutamate (89 microM) were higher than that of L-glutamate (205 microM). In astrocyte-poor cultures, L-glutamate was much more potent, with an EC50 of 5 +/- 4 microM (3-12 microM), for a 30 min exposure, whereas the potencies of NMDA and D-glutamate were essentially unchanged. L- and D-aspartate were also more effective in astrocyte-poor cultures, again with EC50 values of approximately 6-10 microM, as compared with 130 and 108 microM, respectively, in astrocyte-rich cultures. In other experiments, blocking sodium-dependent glutamate uptake in astrocyte-rich cultures, by using a sodium-free medium, made glutamate as potent an agonist as in astrocyte-poor cultures. Finally, we directly assessed the glutamate uptake system in astrocyte-rich and astrocyte-poor cultures and found that uptake was reduced approximately 25-fold in the astrocyte-poor cultures. These results show that in the presence of abundant astrocytes the neurotoxic potencies of L-glutamate, L-aspartate, and D-aspartate are substantially under-estimated.  相似文献   

3.
Astrocytes are thought to control extracellular glutamate concentrations ([Glu]o) in the brain, thereby protecting neurons from excitotoxic injury. We investigated the effects of metabotropic glutamate receptor (mGluR) agonists on glutamate transport and [Glu]o in primary hippocampal astrocytic cultures. Acute or chronic exposure of astrocytes to the mGluR agonist trans‐1‐aminocyclopentane‐1,3‐dicarboxylic acid (trans‐ACPD) or its active isomer 1S,3R‐ACPD reduced [Glu]o in a time‐ and dose‐dependent manner (44.5 ± 3.6% reductions of [Glu]o in astrocytes from P0–P10 rats and 65.9 ± 4.1 % from rats P20 by 100 μM 1S,3R‐ACPD, EC50 ∼ 5 μM). 1S,3R‐ACPD effects developed slowly (median effective at ∼60 min) and persisted for several hours after agonist removal. ACPD‐pretreated astrocytes established lower steady‐state [Glu]o levels. ACPD effects persisted in the presence of the glutamate uptake inhibitors D ,L ‐threo‐β‐hydroxyaspartate (THA) and L ‐trans‐pyrrolidine‐2,4‐dicarboxylate (PDC) but were impaired by disruption of the transmembrane Na+, K+, or H+ gradients. In addition, 1S,3R‐ACPD had no effects on intracellular glutamate content and did not directly block glutamate transport. Furthermore, ACPD effects could be mimicked by glutamate per se and several other compounds presumed to be mGluR agonists, although (S)‐3,5‐dihydroxyphenylglycine (DHPG), (2S,2R,3R)‐2‐(2,3‐dicarboxycyclopropyl)glycine (DCG‐IV), and L ‐(+)‐2‐amino‐4‐phosphonobutyric acid (L ‐AP4) were without effect. These data suggest that glutamate and certain mGluR agonists may regulate [Glu]o by modulating the transmembrane equilibrium of glutamate transport, especially by attenuating glutamate release. GLIA 25:270–281, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

4.
Primary cultures of granule neurons derived from cerebella of postnatal rats are endowed with Glu receptors. Glu receptor agonists exert a trophic influence on differentiating granule cells but, with maturation, the cells become vulnerable to excitatory amino acids. Here we show that the P2 purinoceptor antagonist basilen blue abolishes in rat cerebellar granule neurons the cytotoxic action of glutamate with an IC50 in the 10–20 μM range. Within the same concentrations, basilen blue inhibits binding of [3H]ATP to cerebellar granule cells, glutamate-evoked release (but not uptake) of [3H] D-aspartate and Ca2+ uptake. Furthermore, the extracellular phosphorylation of a major 45-kDa endogenous ecto-protein substrate of cerebellar granule neurons is inhibited with an IC50 of about 1 μM. Similar effects are elicited by 5-adenylylimidodiphosphate, a P2 purinoceptor agonist, when supplied to the neurons for 8 days previously to the addition of glutamate. Our data point to the use of P2 purinoceptor modulators as novel elements for understanding and controlling glutamate-mediated excitatory neurotoxicity and neurotransmission. We suggest a possible involvement of P2 purinoceptors in these actions. © 1996 Wiley-Liss, Inc.  相似文献   

5.
Poly(ADP‐ribose) polymerase‐1 (PARP‐1) is a ubiquitous nuclear enzyme involved in genomic stability. Excessive oxidative DNA strand breaks lead to PARP‐1‐induced depletion of cellular NAD+, glycolytic rate, ATP levels, and eventual cell death. Glutamate neurotransmission is tightly controlled by ATP‐dependent astrocytic glutamate transporters, and thus we hypothesized that astrocytic PARP‐1 activation by DNA damage leads to bioenergetic depletion and compromised glutamate uptake. PARP‐1 activation by the DNA alkylating agent, N‐methyl‐N′‐nitro‐N‐nitrosoguanidine (MNNG), caused a significant reduction of cultured cortical astrocyte survival (EC50 = 78.2 ± 2.7 μM). HPLC revealed MNNG‐induced time‐dependent reductions in NAD+ (98%, 4 h), ATP (71%, 4 h), ADP (63%, 4 h), and AMP (66%, 4 h). The maximal [3H]glutamate uptake rate (Vmax) also declined in a manner that corresponded temporally with ATP depletion, falling from 19.3 ± 2.8 in control cells to 2.1 ± 0.8 nmol/min/mg protein 4 h post‐MNNG. Both bioenergetic depletion and loss of glutamate uptake capacity were attenuated by genetic deletion of PARP‐1, directly indicating PARP‐1 involvement, and by adding exogenous NAD+ (10 mM). In mixed neurons/astrocyte cultures, MNNG neurotoxicity was partially mediated by extracellular glutamate and was reduced by co‐culture with PARP‐1−/− astrocytes, suggesting that impairment of astrocytic glutamate uptake by PARP‐1 can raise glutamate levels sufficiently to have receptor‐mediated effects at neighboring neurons. Taken together, these experiments showed that PARP‐1 activation leads to depletion of the total adenine nucleotide pool in astrocytes and severe reduction in neuroprotective glutamate uptake capacity. © 2009 Wiley‐Liss, Inc.  相似文献   

6.
The contribution of ionotropic and metabotropic glutamate receptors to inositol polyphosphate accumulation in carp retinal slices was investigated using myo-[2-3H]inositol prelabelling. In the presence of the glutamate agonists quisqualate, (RS)-α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid (AMPA) and trans-(±)-1-amino-1, 3-cyclopentane-dicarboxylic acid (t-ACPD), formation of [3H]inositol phosphate was significantly increased in a dose-dependent manner, with EC50 values of 350 nM, 1.5 μM and 10 μM respectively. The complete AMPA-induced response and a large component of the quisqualate-induced response were inhibited in a competitive manner when the ionotropic antagonist 6-cyano-7-nitroquinoxaline-2, 3-dione (CNQX) was present. Furthermore, the remaining level of quisqualate-induced [3H]inositol phosphate formation closely matched that produced by ACPD alone, and coincubation of AMPA and ACPD showed additive effects, suggesting that the quisqualate-induced response resulted from coactivation of metabotropic and ionotropic glutamate receptors. The ionotropic component was partially reduced in the presence of cobalt, suggesting indirect effects resulting from synaptic interactions. We could exclude indirect effects through depolarization-induced release of other neurotransmitters. Only serotonin (EC50 1 μM) and carbachol (at a concentration of 1 mM) stimulated [3H]inositol phosphate formation, but their antagonists did not affect the quisqualate response and coactivation with quisqualate and serotonin or carbachol resulted in additive effects. The ionotropic component was completely suppressed when Ca2+ was omitted from the medium and cobalt was present. This makes it likely that the ionotropic component resulted from Ca2+ entry through AMPA-gated channels and subsequent Ca2+-dependent activation of phospholipase C.  相似文献   

7.
Primary cultures of gerbil mesencephalon were used for studying the modulation exerted by tachykinin NK3 receptor activation on the activity of dopamine (DA) neurons. Dopamine neurons were identified by their ability to take up [3H]DA in a nomifensine-dependent manner. Moreover, tyrosine hydroxylase immunohistochemistry revealed that these neurons accounted for 5–7% of the total cell population. The NK3 receptor agonists, senktide (EC50= 0.58 nM) and [MePhe7]neurokinin B (EC50= 3 nM), increased spontaneous [3H]DA release in a concentration-dependent manner. In contrast, tested at a supramaximal concentration (10-7 M), neither septide nor substance P were found to affect [3H]DA release. The senktide-evoked [3H]DA release was not observed when extracellular Ca2+ was chelated, but was unaffected by nomifensine. This indicates that this increase in [3H]DA outflow resulted more from an exocytotic process than from reversal of carrier-mediated DA uptake. Moreover, the senktide effect was unaffected by the Na+ channel blocker tetrodotoxin, a result suggesting a direct action of senktide on DA neurons. The non-peptide NK3 receptor antagonist, SR 142801, shifted or blocked (ICs0 = 0.89 nM) the senktide-evoked [3H]DA release, while its (-)-antipode, SR 142806, was 80-fold less potent, in agreement with binding data. Selective antagonists for NK, (SR 140333) or NK2 (SR 48968) receptors failed to reduce the senktide effect. Light scanning microscopic analysis of mesencephalic cells loaded with the Ca2+ sensitive dye, fluo-3, showed that senktide induced a rise in cytosolic Ca2+ in 8-10% of the cell population. The senktide-induced elevation in intracellular Ca2+ was rapid in onset and transient (at lo4 M) or more sustained with no further increase in fluorescence intensity (at 10-7 M). The proportion of senktide-responsive cells was not significantly modified when extracellular Ca2+ was chelated, but was reduced by 87% in the presence of SR 142801 and by 75% in cultures that were pre-treated with the DA neurotoxin l-methyl-4-phenylpyridinium. The present study shows that enhancement of spontaneous [3H]DA release and intracellular Ca2+ mobilization may be observed after NK3 receptor stimulation and that both biochemical events are likely to occur in DA neurons.  相似文献   

8.
Polyamines such as spermidine potentiate activation of theN-methyl-D-aspartate (NMDA)-type excitatory amino acid receptor. The goal of the present study was to investigate interactions between the putative polyamine binding site and previously described sites for glutamate and glycine. Binding of the high-potency PCP receptor ligand [3H]MK-801 to well-washed rat brain membranes was used as an in vitro probe of NMDA receptor activation. Spermidine concentration-response studies were performed in the absence and presence of both glutamate and glycine, with and withoutD-(−)-2-amino-5-phosphonovaleric acid (D(−)AP-5) or 7-chlorokynurenic acid (7Cl-KYN). Incubation in the presence of spermidine alone induced a 20.4-fold increase in [3H]MK-801 binding with an EC50 value of 13.3 μM. The mean concentration of spermidine which induced maximal stimulation of binding was 130 μM (n = 10,S.E.M.= 24.66,range= 25–250 μM). Glutamate (10 μM) decreased the EC50 value for spermidine-induced stimulation of [3H]MK-801 binding to 3.4 μM. Glycine (10 μM) did not significantly alter either maximum spermidine-induced [3H]MK-801 binding or the EC50 value for spermidine-induced stimulation of [3H]MK-801 binding. Incubation in the presence of the specific glutamate antagonistD(−)AP-5 attenuated [3H]MK-801 binding in a glutamate-reversible fashion. The competitive glycine antagonist 7Cl-KYN decreased maximum spermidine-induced [3H]MK-801 binding in a glycine-reversible fashion. In addition, 7Cl-KYN increased the EC50 value for spermidine-induced stimulation of [3H]MK-801 binding whileD(−)AP-5 was without effect. These findings suggest that glutamate and glycine regulate the polyamine binding site differentially. PCP-like agents induce a psychotomimetic state closely resembling schizophrenia by inhibiting NMDA receptor-mediated neurotransmission. The ability of polyamines to modulate NMDA receptor functioning suggests a potential site for pharmacological intervention.  相似文献   

9.
The effect of glutamate of [Ca2+]i and on [3H]γ-aminobutyric acid (GABA) release was studied on cultured chick embryonic retina cells. It was observed that glutamate (100 μM) increases the [Ca2+]i by Ca2+ influx through Ca2+ channels sensitive to nitrendipine, but not to ω-conotoxin GVIA (ω-Cg Tx) (50%), and by other channels insensitive to either Ca2+ channel blocker. Mobilization of Ca2+ by glutamate required the presence of external Na+, suggesting that Na+ mobilization through the ionotropic glutamate receptors is necessary for the Ca2+ channels to open. The increase in [Ca2+]i was not related to the release of [3H]GABA induced by glutamate, suggesting that the pathway for the entry of Ca2+ triggered by glutamate does not lead to exocytosis. In fact, the glutamate-induced release of [3H]GABA was significantly depressed by Cao2+, but it was dependent on Nao+, just as was observed for the [3H]GABA release induced by veratridine (50 μM). The veratridine-induced release could be fully inhibited by TTX, but this toxin had no effect on the glutamate-induced [3H]GABA release. Both veratridine- and glutamate-induced [3H]GABA release were inhibited by 1-(2-(((diphenylmethylene)amino)oxy)ethyl)-1,2,5,6-tetrahydro-3-pyridine-carboxylic acid (NNC-711), a blocker of the GABA carrier. Blockade of the NMDA and non-NMDA glutamate receptors with MK-801 and 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX), respectively, almost completely blocked the release of [3H]GABA evoked by glutamate. Continuous depolarization with 50 mM K+ induced maximal release of [3H]GABA of about 1.5%, which is much smaller than the release evoked by glutamate under the same conditions (6.0–6.5%). Glycine (3 μM) stimulated [3H]GABA release induced by 50 mM K+, and this effect was blocked by MK-801, suggesting that the effect of K+ on [3H]GABA release was partially mediated through the NMDA receptor which probably was stimulated by glutamate released by K+ depolarization. We conclude that glutamate induces Ca2+-independent release of [3H]GABA through reversal of the GABA carrier due to Na+ entry through the NMDA and non-NMDA, TTX-insensitive, channels. Furthermore the GABA carrier seems to be inhibited by Ca2+ entering by the pathways open by glutamate. This Ca2+ does not lead to exocytosis, probably because the Ca2+ channels used are located at sites far from the active zones.  相似文献   

10.
In the present work we investigated the mechanisms controlling the release of acetylcholine (ACh) and of γ-aminobutyric acid (GABA) from cultures of amacrine-like neurons, containing a subpopulation of cells which are simultaneously GABAergic and cholinergic. We found that 81.2 ± 2.8% of the cells present in the culture were stained immunocytochemically with an antibody against choline acetyltransferase, and 38.5 ± 4.8% of the cells were stained with an antibody against GABA. Most of the cells containing GABA (87.0 ± 2.9%) were cholinergic. The release of acetylcholine and GABA was mostly Ca2+-dependent, although a significant release of [3H]GABA occurred by reversal of its transporter. Potassium evoked the Ca2+-dependent release of [3H]GABA and [3H]acetylcholine, with EC50 of 31.0 ± 1.0 mm and 21.6 ± 1.1 mm , respectively. The Ca2+-dependent release of [3H]acetylcholine was significantly inhibited by 1 μm tetrodotoxin and by low (30 nm ) ω-conotoxin GVIA (ω-CgTx GVIA) concentrations, or by high (300 nm ) nitrendipine (Nit) concentrations. On the contrary, the release of [14C]GABA was reduced by 30 nm nitrendipine, or by 500 nm ω-CgTx GVIA, but not by this toxin at 30 nm . The release of either transmitters was unaffected by 200 nm ω-Agatoxin IVA (ω-Aga IVA), a toxin that blocks P/Q-type voltage-sensitive Ca2+ channels (VSCC). The results show that Ca2+-influx through ω-CgTx GVIA-sensitive N-type VSCC and through Nit-sensitive L-type VSCC induce the release of ACh and GABA. However, the significant differences observed regarding the Ca2+ channels involved in the release of each neurotransmitter suggest that in amacrine-like neurons containing simultaneously GABA and acetylcholine the two neurotransmitters may be released in distinct regions of the cells, endowed with different populations of VSCC.  相似文献   

11.
The effects of hexachlorocyclohexane (HCH) isomers and some GABAergic compounds on [3H]noradrenaline (NA) release from rat hippocampal slices prelabelled with 80 nM [3H]NA were determined. The convulsant γ-HCH isomer facilitated (EC50 = 21 μM) and the depressant δ-HCH isomer reduced (EC50 = 48 μM) the Ca2+-dependent K+-evoked release of [3H]NA, whereas α- and β-HCH isomers did not show any effect. Moreover, α- and δ-HCH isomers antagonized the facilitation of evoked [3H]NA release induced by the γ-HCH isomer. The GABAergic convulsant drugs, bicuculline, picrotoxin and pentylenetetrazol, did not cause any modification of the evoked [3H]NA release even at high concentrations. Neither bicuculline nor picrotoxin blocked the effects of HCH isomers on K+-evoked release of [3H]NA. Exposure of slices to diazepam reduced the K+-evoked release of [3H]NA (EC50 = 33 μM) in a manner similar to that of the δ-HCH isomer. In addition, diazepam (50 μM) blocked the γ-HCH effect and caused an additive inhibitory response with the δ-HCH isomer. On the other hand, diazepam and δ-HCH induced a time-dependent Ca2+-independent enhancement of basal [3H]NA release. The results suggest that modulation of [3H]NA release in the hippocampus by HCH isomers may be involved in the central actions of these compounds, and that sites other than the classic GABAA receptor may underlie their presynaptic mechanisms of action.  相似文献   

12.
We studied the early and late effects of L- trans -pyrrolidine-2,4-dicarboxylate (PDC), a competitive inhibitor of glutamate uptake with low affinity for glutamate receptors, in co-cultures of rat cortical neurons and glia expressing spontaneous excitatory amino acid (EAA) neurotransmission. At 100 or 200 μM, PDC induced different patterns of electrical changes: 100 μM prolonged tetrodotoxin-sensitive excitation triggered by synaptic glutamate release; 200 μM produced sustained, tetrodotoxin-insensitive and EAA-mediated neuronal depolarization, overwhelming synaptic activity. At 200 μM, but not at 100 μM, PDC caused rapid elevation of the glutamate concentration ([Glu]0) in the culture medium, resulting in NMDA receptor-mediated excitotoxic death of neurons 24 h later. The increase in [Glu]0 was largely insensitive to tetrodotoxin, independent of extracellular Ca2+, and present also in astrocyte-pure cultures. By the use of glutamate transporters functionally reconstituted in liposomes, we showed directly that PDC activates carrier-mediated release of glutamate via heteroexchange. Glutamate release and delayed neurotoxicity in our cultures were suppressed if PDC was applied in a Na+-free medium containing Li+. However, replacement of Na+ with choline instead of Li+ did not result in an identical effect, suggesting that Li+ does not act simply as an external Na+ substitute. In conclusion, our data indicate that alteration of glutamate transport by PDC has excitotoxic consequences and that active release of glutamate rather than just uptake inhibition is responsible for the generation of neuronal injury.  相似文献   

13.
We studied the effect of ionotropic glutamate receptor agonists on the release of endogenous glutamate or of [3H]D -aspartate from reaggregate cultures (retinospheroids) or from monolayer cultures of chick retinal cells, respectively. Kainate increased the fluorescence ratio of the Na+ indicator SBFI and stimulated a dose-dependent release of glutamate in low (0.1 mM) Ca2+ medium, as measured using a fluorometric assay. Under the same experimental conditions, the release evoked by N-methyl-D -aspartate (NMDA; 400 μM) was about half of that evoked by the same kainate concentration; α-amino-3-hydroxy-5-methyl-4-isoxasolepropionic acid (AMPA; 400 μM) did not trigger a significant response. In the presence of 1 mM CaCl2, all of the agonists increased the [Ca2+]i, as determined with the fluorescence dye Indo-1, but the glutamate release evoked by NMDA and kainate was significantly lower than that measured in 0.1 mM CaCl2 medium. Inhibition by Ca2+ of the kainate-stimulated release of glutamate was partially reversed by the phospholipase A2 inhibitor oleiloxyethyl phosphorylcholine (OPC), suggesting that the effect was mediated by the release of arachidonic acid, which inhibits the glutamate carrier. Accordingly, kainate, NMDA, and AMPA stimulated a Ca2+-dependent release of [3H]arachidonic acid, and the direct addition of the exogenous fatty acid to the medium decreased the release of glutamate evoked by kainate in low (0.1 mM) CaCl2 medium. In monolayer cultures, we showed that NMDA, kainate, and AMPA also stimulated the release of [3H]D -aspartate, but in this case release in the presence of 1 mM CaCl2 was significantly higher than that evoked in media with no added Ca2+. The ranking order of efficacy for stimulation of Ca2+-dependent release of [3H]D -aspartate was NMDA ≪ kainate < AMPA. © 1996 Wiley-Liss, Inc.  相似文献   

14.
Glutamate transport in nearly pure rat cortical neurons in culture (less than 0.2% astrocytes) is potently inhibited by dihydrokainate, l -serine-O-sulphate, but not by l -α-amino-adipate. This system allows for a test of the hypothesis that glutamate transport is important for protecting neurons against the toxicity of endogenous synaptically released glutamate. In support of this hypothesis, a 20–24 h exposure to 1 mm dihydrokainate reduced cell survival to only 14.8 ± 9.8% in neuronal cultures (P < 0.001;n = 3), although it had no effect on neuronal survival in astrocyte-rich cultures (P > 0.05;n = 3). Dihydrokainate also significantly caused accumulation of glutamate in the extracellular medium of cortical neuronal cultures (6.6 ± 4.9 μm , compared to 1.2 ± 0.3 μm in control, n = 14, P < 0.01). The neurotoxicity of dihydrokainate was blocked by 10 μm MK-801, 10 μm tetrodotoxin, and an enzyme system that degrades extracellular glutamate. The latter two also abolished the accumulation of glutamate in the extracellular medium. Dihydrokainate (1 mm ) inhibited the 45calcium uptake stimulated by 30 μm N-methyl-d -aspartate (NMDA), but not by higher concentrations consistent with a weak antagonist action of dihydrokainate at the NMDA receptor. Whole cell recordings showed that 1 mm dihydrokainate produced ≈ 25% inhibition of 30 μm NMDA-induced current in cortical neurons. Dihydrokainate (1 mm ) alone generated a small current (17% of the current produced by 30 μm NMDA) that was blocked by 30 μm 5,7-dichlorokynurenate and only weakly by 10 μm 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX). These results suggest that the toxicity of dihydrokainate in neuronal cultures is due to its ability to block glutamate transport in these cultures, and that dihydrokainate-sensitive neuronal glutamate transport may be important in protecting neurons against the toxicity of synaptically released glutamate.  相似文献   

15.
Non-synaptic release may be the major route of excitatory amino acid (EAA) efflux during cerebral ischemia. Possible routes of non-synaptic release include non-specific anion channels, reversal of Na+-, CI?-, or Ca2+-dependent uptake, and cell lysis. In the present study we employ a novel approach to show reversal of Na+-dependent uptake as a major route of EAA efflux from astrocyte cultures under conditions of energy failure. Primary rat astrocyte cultures were subjected to combined blockade of glycolytic and oxidative metabolism after incubation with [3H]-D-aspartate (D-ASP). Energy failure produced an efflux of D-ASP that was maximal by 90 minutes. The efflux over this period was reduced by more than 50% in cells that had been pre-loaded with PDC (L-transpyrrolidine-2,4-dicarboxylic acid) or TBHA (threo-β-hydroxyaspartic acid), compounds that are competitive inhibitors of Na+-dependent glutamate uptake. The effect of pre-loading with the inhibitors was concentration dependent. No effect was seen if the inhibitors were added after induction of energy failure, suggesting that the attenuation of D-ASP efflux resulted from binding of the inhibitors to an intracellular site. These results provide strong evidence that EAA efflux from astrocytes under conditions of energy failure occurs largely through reversal of Na+-dependent uptake. © 1995 Wiley-Liss, Inc.
  • 1 This article is a US Government work and, as such, is in the public domain in the United States of America.
  •   相似文献   

    16.
    Autoradiographic techniques were used to test if positive modulators of AMPA-type glutamate receptors have regionally differentiated effects on ligand binding. Cyclothiazide, a drug with ten fold greater effects on `flip' than `flop' splice variants of the receptors, had unequal effects across the subdivisions of hippocampus; i.e., it reduced [3H]AMPA binding in field CA3 with an EC50 of 24 μM and in field CA1 and dentate gyrus with EC50s between 60 and 100 μM. The EC50 for the drug's influence on binding was also significantly lower in the superficial than in the deeper layers of the neocortex, though these differences were not as pronounced as those in the hippocampus. The ampakine CX614, a compound with a modest preference for flop variants, had a slightly lower EC50 for its effects on [3H]AMPA binding in CA1 than in CA3. This result was confirmed with [3H]fluorowillardiine binding. The effects of the ampakine in neocortex tended to be greater in the deeper than superficial layers but this did not reach statistical significance. These results indicate that differential effects of modulators on AMPA receptor subunits are reflected in their relative potency across brain subdivisions. This raises the possibility that subclasses of positive modulators will exhibit a measurable degree of selectivity in their physiological and behavioral influences.  相似文献   

    17.
    18.
    Prior work indicated that serotonin transporter (SERT) inhibitors competitively inhibit substrate‐induced [3H]5‐HT release, producing rightward shifts in the substrate‐dose response curve and increasing the EC50 value without altering the Emax. We hypothesized that this finding would not generalize across a number of SERT inhibitors and substrates, and that the functional dissociation constant (Ke) of a given SERT inhibitor would not be the same for all tested substrates. To test this hypothesis, we utilized a well‐characterized [3H]5‐HT release assay that measures the ability of a SERT substrate to release preloaded [3H]5‐HT from rat brain synaptosomes. Dose‐response curves were generated for six substrates (PAL‐287 [naphthylisopropylamine], (+)‐fenfluramine, (+)‐norfenfluramine, mCPP [meta‐chlorophenylpiperazine], (±)‐MDMA, 5‐HT) in the absence and presence of a fixed concentration of three SERT inhibitors (indatraline, BW723C86, EG‐1‐149 [4‐(2‐(benzhydryloxy)ethyl)‐1‐(4‐bromobenzyl)piperidine oxalate]). Consistent with simple competitive inhibition, all SERT inhibitors increased the EC50 value of all substrates. However, in many cases a SERT inhibitor decreased the Emax value as well, indicating that in the presence of the SERT inhibitor the substrate became a partial releaser. Moreover, the Ke values of a given SERT inhibitor differed among the six SERT substrates, indicating that each inhibitor/substrate combination had a unique interaction with the transporter. Viewed collectively, these findings suggest that it may be possible to design SERT inhibitors that differentially regulate SERT function. Synapse 64:862–869, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

    19.
    Nitric oxide synthase (NOS) activity was determined in adult rat frontal cortex and hippocampus by measuring the conversion of L-[3H]arginine to L-[3H]citrulline. N-methyl-D-aspartate (NMDA), but not kainate or α-amino-3-hydroxy-5-methylisoxazole-4-propionate (AMPA), stimulated NOS activity. This effect was concentration dependent (EC50 ≈ 30μM) and was inhibited by tetrodotoxin, EGTA, Nω-nitro-L-arginine (NOARG), Mg2+, phencyclidine, and (cis)-4-phosphonomethyl-2-piperidine carboxylate (CGS 19755), but not by 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX). NOS activity was increased to an even greater extent by the calcium ionophores ionomycin and A23187 and by depolarization with 50 mM K+. Interestingly, neither caffeine nor 1-aminocyclopentane-1,3-dicarboxylic acid (1S,3R-ACPD), drugs that would be expected to increase intracellular Ca2+ concentration by release of Ca2+ from intracellular ryanodine- and inositol-1,4,5-trisphosphate-sensitive stores, respectively, had any significant effect on NOS activity. It is concluded that NOS can be activated by NMDA binding to a classic NMDA glutamate receptor subtype as well as by depolarization or other agents that increase the influx of extracellular Ca2+. The paradoxical lack of effect of caffeine, as well as the inhibitory effect of tetrodotoxin, are discussed. © 1994 Wiley-Liss, Inc.  相似文献   

    20.
    The glutamate agonist LY404,039 has been used to treat schizophrenia. Because all currently used antipsychotics act on dopamine receptors, it was decided to examine whether this glutamate agonist also had an affinity for dopamine D2 receptors in vitro. The present data show that LY404,039 inhibited the binding of [3H]domperidone and [3H](+)PHNO by 15.5 ± 1.5% to the high‐affinity state, D2High, of cloned dopamine D2Long receptors and rat striatal tissue with dissociation constants of between 8.2 and 12.6 nM. This high‐affinity component of LY404,039 on the binding of [3H]domperidone was inhibited by the presence of guanine nucleotide, indicating an agonist action of the drug at D2High. LY404,039 also stimulated the incorporation of [35S]GTP‐γ‐S into D2Long receptors (EC50% = 80 ± 15 nM) over the same range of concentrations as occurred for the inhibition of [3H]domperidone by LY404,039 at D2High (IC50%High = 50 ± 10 nM). A possible clinical antipsychotic action of LY404,039 may depend on the combined stimulation of glutamate receptors and a partial dopamine agonist action that would interfere with neurotransmission at D2High receptors. Synapse 63:935–939, 2009. © 2009 Wiley‐Liss, Inc.  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