首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Thrombin effect increasing swelling-induced glutamate efflux was examined in cultured cortical astrocytes, cerebellar granule neurons (CGN), hippocampal and cortical neurons. Hypotonic glutamate efflux (monitored by d-[3H]aspartate) from cortical astrocytes was increased by thrombin (5 U/mL) to reach 16 % of the cell pool in 5 min. Thrombin had lower effects in CGN, and marginal effects in hippocampal and cortical neurons. These differences were related to the magnitude of thrombin-evoked cytosolic calcium rise. The protease-activated receptor 1 is expressed in astrocytes and neurons. In astrocytes exposed to chemical ischemia (sodium iodoacetate plus sodium azide) d-[3H]aspartate release showed a first phase (20–40 min) of initial low efflux which progressively increases; and a second phase (40–60 min) of larger efflux coincident with cell swelling. Efflux at the first phase was 52 % inhibited by the glutamate transporter blocker DL-threo-β-benzyloxyaspartic-acid (TBOA) and 11 % by the volume-sensitive pathway blocker phloretin. At the second phase, efflux was reduced 52 and 38 % respectively, by these blockers. In CGN d-[3H]aspartate efflux increased sharply and then decreased. This efflux was 32 % reduced by calcium omission, 46 % by TBOA and 32 % by 4-[(2-butyl-6,7dichloro-2-cyclopentyl-2,3-dihydro-1oxo-1H-inden-5-yl)oxy] butanoic-acid. Thrombin enhanced this release by 32 %. Ischemic treatment increased astrocyte mortality from 4 % in controls to 39 and 61 % in ischemia and ischemia plus thrombin, respectively. Cell death was prevented by phloretin. CGN viability was unaffected by the treatment. These results suggest that coincidence of swelling and thrombin during ischemia elevates extracellular glutamate prominently from astrocyte efflux, which may endanger neurons in vivo.  相似文献   

2.
Microglia are the resident immune cells of the CNS, which are important for preserving neural tissue functions, but may also contribute to neurodegeneration. Activation of these cells in infection, inflammation, or trauma leads to the release of various toxic molecules, including reactive oxygen species (ROS) and the excitatory amino acid glutamate. In this study, we used an electrophysiologic approach and a d‐[ 3 H] aspartate (glutamate) release assay to explore the ROS‐dependent regulation of glutamate‐permeable volume‐regulated anion channels (VRACs). Exposure of rat microglia to hypo‐osmotic media stimulated Cl? currents and d ‐[3H]aspartate release, both of which were inhibited by the selective VRAC blocker, DCPIB. Exogenously applied H2O2 potently increased swelling‐activated glutamate release. Stimulation of microglia with zymosan triggered production of endogenous ROS and strongly enhanced glutamate release via VRAC in swollen cells. The effects of zymosan were attenuated by the ROS scavenger, MnTMPyP, and by two inhibitors of NADPH oxidase (NOX), diphenyliodonium and thioridazine. However, zymosan‐stimulated glutamate release was insensitive to other NOX blockers, apocynin and HEBSF. This pharmacologic profile pointed to the potential involvement of apocynin‐insensitive NOX4. Using RT‐PCR we confirmed that NOX4 is expressed in rat microglial cells along with NOX1 and NOX2. To check for potential involvement of phagocytic NOX2, we stimulated this isoform using protein kinase C (PKC) activator, phorbol 12‐myristate 13‐acetate or inhibited it with the broad spectrum PKC blocker, Gö6983. Both agents potently modulated endogenous ROS production by NOX2 but not VRAC activity. Taken together, these data suggest that the anion channel VRAC may contribute to microglial glutamate release and that its activity is regulated by endogenous ROS originating from NOX4.  相似文献   

3.
Our study was aimed at investigating whether complement, a complex of soluble and membrane‐associated serum proteins, could, in addition to its well‐documented post‐synaptic activity, also pre‐synaptically affect the release of classic neurotransmitters in central nervous system (CNS). Complement (dilution 1 : 10 to 1 : 10000) elicited the release of preloaded [3H]‐d ‐aspartate ([3H]d ‐ASP) and endogenous glutamate from mouse cortical synaptosomes in a dilution‐dependent manner. It also evoked [3H]d ‐ASP release from mouse hippocampal, cerebellar, and spinal cord synaptosomes, as well as from rat and human cortical nerve endings, but left unaltered the release of GABA, [3H]noradrenaline or [3H]acetylcholine. Lowering external Na+ (from 140 to 40 mM) or Ca2+ (from 1.2 to 0.1 mM) ions prevented the 1 : 300 complement‐evoked [3H]d ‐ASP release from mouse cortical synaptosomes. Complement‐induced releasing effect was unaltered in synaptosomes entrapped with the Ca2+ ions chelator 1,2‐bis‐(2‐aminophenoxy) ethane‐N,N,N’,N’, tetra‐acetic acid or with pertussis toxin. Nifedipine,/ω‐conotoxin GVIA/ω‐conotoxin MVIIC mixture as well as the vesicular ATPase blocker bafilomycin A1 were also inefficacious. The excitatory amino acid transporter blocker DL‐threo‐ß‐benzyloxyaspartic acid, on the contrary, reduced the complement‐evoked releasing effect in a concentration‐dependent manner. We concluded that complement‐induced releasing activity is restricted to glutamatergic nerve endings, where it was accounted for by carrier‐mediated release. Our observations afford new insights into the molecular events accounting for immune and CNS crosstalk.

  相似文献   


4.
Vitamin C is transported in the brain by sodium vitamin C co‐transporter 2 (SVCT‐2) for ascorbate and glucose transporters for dehydroascorbate. Here we have studied the expression of SVCT‐2 and the uptake and release of [14C] ascorbate in chick retinal cells. SVCT‐2 immunoreactivity was detected in rat and chick retina, specially in amacrine cells and in cells in the ganglion cell layer. Accordingly, SVCT‐2 was expressed in cultured retinal neurons, but not in glial cells. [14C] ascorbate uptake was saturable and inhibited by sulfinpyrazone or sodium‐free medium, but not by treatments that inhibit dehydroascorbate transport. Glutamate‐stimulated vitamin C release was not inhibited by the glutamate transport inhibitor l ‐β‐threo‐benzylaspartate, indicating that vitamin C release was not mediated by glutamate uptake. Also, ascorbate had no effect on [3H] d ‐aspartate release, ruling out a glutamate/ascorbate exchange mechanism. 2‐Carboxy‐3‐carboxymethyl‐4‐isopropenylpyrrolidine (Kainate) or NMDA stimulated the release, effects blocked by their respective antagonists 6,7‐initroquinoxaline‐2,3‐dione (DNQX) or (5R,2S)‐(1)‐5‐methyl‐10,11‐dihydro‐5H‐dibenzo[a,d]cyclohepten‐5,10‐imine hydrogen maleate (MK‐801). However, DNQX, but not MK‐801 or 2‐amino‐5‐phosphonopentanoic acid (APV), blocked the stimulation by glutamate. Interestingly, DNQX prevented the stimulation by NMDA, suggesting that the effect of NMDA was mediated by glutamate release and stimulation of non‐NMDA receptors. The effect of glutamate was neither dependent on external calcium nor inhibited by 1,2‐bis (2‐aminophenoxy) ethane‐N′,N′,N′,N′,‐tetraacetic acid tetrakis (acetoxy‐methyl ester) (BAPTA‐AM), an internal calcium chelator, but was inhibited by sulfinpyrazone or by the absence of sodium. In conclusion, retinal cells take up and release vitamin C, probably through SVCT‐2, and the release can be stimulated by NMDA or non‐NMDA glutamate receptors.  相似文献   

5.
In this study, the functional consequences of the pharmacological modulation of the M‐current (IKM) on cytoplasmic Ca2+ intracellular Ca2+concentration ([Ca2+]i) changes and excitatory neurotransmitter release triggered by various stimuli from isolated rat cortical synaptosomes have been investigated. Kv7.2 immunoreactivity was identified in pre‐synaptic elements in cortical slices and isolated glutamatergic cortical synaptosomes. In cerebrocortical synaptosomes exposed to 20 mM [K+]e, the IKM activator retigabine (RT, 10 μM) inhibited [3H]d ‐aspartate ([3H]d ‐Asp) release and caused membrane hyperpolarization; both these effects were prevented by the IKM blocker XE‐991 (20 μM). The IKM activators RT (0.1–30 μM), flupirtine (10 μM) and BMS‐204352 (10 μM) inhibited 20 mM [K+]e‐induced synaptosomal [Ca2+]i increases; XE‐991 (20 μM) abolished RT‐induced inhibition of depolarization‐triggered [Ca2+]i transients. The P/Q‐type voltage‐sensitive Ca2+channel (VSCC) blocker ω‐agatoxin IVA prevented RT‐induced inhibition of depolarization‐induced [Ca2+]i increase and [3H]d ‐Asp release, whereas the N‐type blocker ω‐conotoxin GVIA failed to do so. Finally, 10 μM RT did not modify the increase of [Ca2+]i and the resulting enhancement of [3H]d ‐Asp release induced by [Ca2+]i mobilization from intracellular stores, or by store‐operated Ca2+channel activation. Collectively, the present data reveal that the pharmacological activation of IKM regulates depolarization‐induced [3H]d ‐Asp release from cerebrocortical synaptosomes by selectively controlling the changes of [Ca2+]i occurring through P/Q‐type VSCCs.  相似文献   

6.
Microdialysis in neostriatum of anaesthetized rats was performed to study effects on amino acid efflux of the glutamate uptake-inhibitor dihydrokainate (DHK). Both basal and K+-evoked (100 mM) efflux of glutamate increased in the presence of DHK. The increase in the basal glutamate efflux occurred at lower DHK concentrations than during K+-depolarization (when the extracellular glutamate concentration was several-fold higher), confirming that DHK is a competitive inhibitor. The increase in basal efflux caused by DHK did not exhibit Ca2+-dependency, whereas ∼50% of the increase in glutamate efflux during K+-depolarization was Ca2+-dependent. The Ca2+-dependent efflux is related to transmitter release, whereas the Ca2+-independent efflux is probably due to metabolic events and/or transport of DHK into cells in exchange for glutamate. Taurine efflux in response to DHK increased both during basal conditions and K+-depolarization, probably secondary to the increase in glutamate concentration, whereas aspartate, GABA, glutamine and alanine effluxes did not change.  相似文献   

7.
This study was undertaken in order to characterize the role of the glutamate/aspartate transporter (GLAST) in the glutathione (GSH) efflux induced by glutamate. Our results demonstrated that retinal cell cultures exhibit two mechanisms of GSH release, one Na+-independent and other Na+-dependent. Glutamate and aspartate induced GSH efflux only in presence of Na+. Treatment with PCD (L-trans-Pyrrolidine-2,4-dicarboxylate), a transportable glutamate uptake blocker, increased GSH release indicating that GSH can be carried by glutamate transporters in retinal cell cultures. Added to this, treatment with zinc ion cultures, a recognized inhibitor of GLAST blocked GSH efflux evoked by glutamate. Treatment with NMDA antagonist (MK-801) did not have any effect on the GSH release induced by glutamate. These results suggest that glutamate induces GLAST-mediated release of GSH from retinal cell cultures and this could represent an important mechanism of cellular protection against glutamate toxicity in the CNS.  相似文献   

8.
Abstract: In rat hippocampal synaptosomes, adenosine decreased the K+ (15 mM) or the kainate (1 mM) evoked release of glutamate and aspartate. An even more pronounced effect was observed in the presence of the stable adenosine analogue, R-phenylisopropyladenosine. All these effects were reversed by the selective adenosine A1 receptor antagonist 8-cyclo-pentyltheophylline. In the same synaptosomal preparation, K+ (30 mM) strongly stimulated the release of the preloaded [3H]adenosine in a partially Ca2+-dependent and tetrodotoxin (TTX)-sensitive manner. Moreover, in the same experimental conditions, both l -glutamate and l -aspartate enhanced the release of [3H]adenosine derivatives ([3H]ADD). The gluta-mate-evoked release was dose dependent and appeared to be Ca2+ independent and tetrodotoxin insensitive. This effect was not due to metabolism because even the nonmetabolizable isomers d -glutamate and d -aspartate were able to stimulate [3H]ADD release. In contrast, the specific glutamate agonists N-methyl-d -aspartate, kainate, and quisqualate failed to stimulate [3H]ADD release, suggesting that glutamate and aspartate effects were not mediated by known excitatory amino acid receptors. Moreover, NMDA was also ineffective in the absence of Mg2+ and l -glutamate-evoked release was not inhibited by adding the specific antagonists 2-amino-5-phosphonovaleric acid or 6–7-dinitroquinoxaline-2, 3-dione. The stimulatory effect did not appear specific for only excitatory amino acids, as γ-anunobutyric acid stimulated [3H]ADD release in a dose-related manner. These results suggest that, at least in synaptosomal preparations from rat hippocampus, adenosine and glutamate modulate each other's release. The exact mechanism of such interplay, although still, unknown, could help in the understanding of excitatory amino acid neurotoxicity.  相似文献   

9.
Abstract— Hemisections of toad brains, when incubated in a physiological medium containing no glutamine. released considerable amounts of this amino acid into the medium. When glutamine was included in the medium at a concentration of 0.2 mm the net efflux from the tissue was reduced but not totally prevented. Although there was no net uptake of glutamine, the tissue did accumulate [U-14C]glu-tamine and some of this labelled glutamine was rapidly metabolized to glutamate, GABA and aspartate. The precursor-product relationship for the metabolism of glutamine to glutamate differed from the classic single compartment model in that the specific radioactivity of glutamate rose very quickly to approx one-tenth that of glutamine, but increased slowly thereafter. These data suggest that the [14C]glutamine was taken up into two metabolically distinct compartments and/or that some of the [14C]glutamine was converted to [14C]glutamate during the uptake process. The uptake of [14C]glutamine was diminished when the tissue was incubated in a non-oxygenated medium or when Na+ was omitted (substituted with sucrose) and K+ was concomitantly elevated. However, on a relative basis, the incorporation of radioactivity into glutamate and GABA was increased by these incubation conditions. The metabolism of glutamine to aspartate was greatly depressed when the tissue was not oxygenated. The glutamate formed from [U-14C]glutamine taken up by the tissue was converted to GABA at a faster rate than was glutamate derived from [U-14C]glucose. [U-14C]gly-cerol or exogenous [U-14C]glutamate. This suggests that glutamine was metabolized to GABA selectively; i.e. on a relative basis, glutamine served as a better source of carbon for the synthesis of GABA than did glucose, glycerol or exogenous glutamate. When the brain hemisections were incubated in the normal physiological medium with or without glutamine. there was very little efflux of glutamate, GABA or aspartate from the tissue. However when NaCl was omitted from the medium (substituted with sucrose) and K+ was elevated to 29 miu. a marked efflux of these three amino acids into the medium did occur, and over a period of 160min, the content of each amino acid in the tissue was depleted considerably. When glutamine (0.2 mm ) was included in the Na+ deficient-high K.+ medium, the average amount of glutamate, GABA and aspartate in the tissue plus the medium was greater than when glutamine was not included in the medium. Such data indicate that CNS tissues can utilize glutamine for a net synthesis of glutamate, GABA and aspartate. The results of this study provide further evidence in support of the concept that the functional (transmitter) pools of glutamate and GABA are maintained and regulated in part via biosynthesis from glutamine. One specific mechanism instrumental in regulating the content of glutamate in nerve terminals may be a process of glutamine uptake coupled to deamidation.  相似文献   

10.
Using sodium (NaN3)-induced anoxia plus aglycaemia as a model of chemically-induced ischemia, we have characterized the endogenous release of excitatory and inhibitory amino acids from superfused hippocampal slices. Chemical ischemia produced an azide (1–30 mM) dose-dependent increase in the efflux of glutamate, aspartate and GABA. These increases were attenuated to varying degrees by removal of Ca2+, or the addition of the voltage dependent Na+-channel blocker tetrodotoxin (TTX), the selective Ca2+ channel blockers conotoxin MVIIA, MVIIC, and nifedipine, the NMDA antagonist MK801, the AMPA antagonist GYKI-52466. Similarly, addition of the GLT-1 glutamate transport inhibitor dihydrokainate (DHK) and the anti-estrogen/anion channel blocker tamoxifen also attenuated the efflux of glutamate and GABA. It would therefore appear that the increases in amino acid efflux induced by chemical ischemia originates from both the neuronal pool, via conventional exocytotic release, and glial sources via reversal of the GLT-1 transporter and anion channel regulated cell swelling.  相似文献   

11.
Summary A number of important experimental data do not support the widespread hypothesis that Na+-channels block is cerebroprotective, essentially because it reduces presynaptic glutamate release: (i) the inhibition of exocytosis by these compounds is not specific to glutamate; (ii) aspartate efflux produced by various stimuli was also reduced, but aspartate cannot be released by exocytosis because it is not concentrated within presynaptic vesicles; and (iii) glutamate accumulated extracellularly during ischaemic or traumatic insult to the CNS is mainly of cytosolic origin. As an alternative, we propose that use-dependent Na+-channel blockers enhance the resistance of nerve cells to insults, primarily by decreasing their energy demand, and that reduced efflux of glutamate and other compounds is aconsequence of attenuated cellular stress.  相似文献   

12.
In order to study the mechanism of the glutamate-aspartate translocator, rat liver mitochondria were loaded with either glutamate or aspartate. In the presence of ascorbate plus tetramethyl-p-phenylenediamine as an electron donor at the third energy conservation site, exchange of external glutamate for matrix aspartate is highly favored over the reverse exchange. In the absence of an energy source, although the asymmetry of the exchange rates is much smaller, it is still observable. Further studies have shown that the proton uptake accompanying influx of glutamate in exchange for aspartate efflux occurs by protonation of a group on the carrier (pK = 7.9) at the external side of the inner mitochondrial membrane, followed by deprotonation at the matrix surface. It is postulated that glutamate binds to the protonated form of the carrier and aspartate to the deprotonated form. Because of the alkaline pK, aspartate efflux is inhibited with increased matrix [H+] due to limitation of the availability of deprotonated carrier for aspartate binding. For the reverse exchange, aspartate uptake is inhibited by increasing external [H+]. Thus the rate of aspartate uptake by mitochondria is apparently impeded both by a proton motive force (Δp) unfavorable to entry of ions with net negative charge as well as by the small proportion of deprotonated carrier at the outer surface of the membrane. This conclusion is further illustrated by inhibition of the aspartate-aspartate exchange with increased [H+] and by addition of an energy source. The glutamate-glutamate exchange, however, showed a slight stimulation by increased [H+] and was unaffected by the energy state.The model initially proposed for the carrier, in which a neutral glutamate-carrier complex exchanges for a negatively charged aspartate-carrier complex, is tested further. Glutamate uptake was noncompetitively inhibited by external aspartate, which indicates that aspartate and glutamate bind to separate forms of the carrier. Intramitochrondrial glutamate at a concentration of 18 mm, however, had no effect on aspartate efflux. Arrhenius plots for the glutamate-aspartate and aspartate-glutamate exchanges were linear over the range of temperatures tested (1–35 °C and 5–25 °C, respectively) and provided an average value of 14.3 kcal/mol for the energy of activation. In addition, there appear to be two pools, exchangeable and nonexchangeable, of matrix aspartate available to the translocator, since extramitochondrial radiolabeled aspartate can equilibrate only with unlabeled matrix aspartate at low aspartate loading (1–2 nmol of aspartate/mg of protein). The physiological significance of the data is discussed.  相似文献   

13.
Glutamate, the major excitatory transmitter in the vertebrate brain, is removed from the synaptic cleft by a family of sodium‐dependent glutamate transporters profusely expressed in glial cells. Once internalized, it is metabolized by glutamine synthetase to glutamine and released to the synaptic space through sodium‐dependent neutral amino acid carriers of the N System (SNAT3/slc38a3/SN1, SNAT5/slc38a5/SN2). Glutamine is then taken up by neurons completing the so‐called glutamate/glutamine shuttle. Despite of the fact that this coupling was described decades ago, it is only recently that the biochemical framework of this shuttle has begun to be elucidated. Using the established model of cultured cerebellar Bergmann glia cells, we sought to characterize the functional and physical coupling of glutamate uptake and glutamine release. A time‐dependent Na+‐dependent glutamate/aspartate transporter/EAAT1‐induced System N‐mediated glutamine release could be demonstrated. Furthermore, D‐aspartate, a specific glutamate transporter ligand, was capable of enhancing the co‐immunoprecipitation of Na+‐dependent glutamate/aspartate transporter and Na+‐dependent neutral amino acid transporter 3, whereas glutamine tended to reduce this association. Our results suggest that glial cells surrounding glutamatergic synapses may act as sensors of neuron‐derived glutamate through their contribution to the neurotransmitter turnover.  相似文献   

14.
Abstract: Rat pinealocytes accumulate glutamate in microvesicles and secrete it through exocytosis so as to transmit signals intercellularly. Glutamate is involved in the negative regulation of norepinephrine-stimulated melatonin production. In this study, we found that aspartate is also released from cultured rat pinealocytes during the exocytosis of glutamate. The release of aspartate was triggered by addition of KCI or A23187 (a Ca2+ ionophore) in the presence of Ca2+ and was proportional to the amount of l -glutamate released. Furthermore, the release of aspartate was inhibited by both botulinum neurotoxin type E and L- or N-type voltage-gated Ca2+ channel blockers. Bay K 8644, an agonist for the L-type Ca2+ channel, stimulated the release of aspartate 2.1-fold. Immunohistochemical analyses with antibodies against aspartate and synaptophysin revealed that aspartate is colocalized with synaptophysin in a cultured pinealocyte. HPLC with fluorometric detection indicated that the released aspartate is of the l form, although pinealocytes also contain the d form in their cytoplasm, corresponding to ~30% of the total free aspartate. Radiolabeled l -aspartate was taken up by the microsomal fraction from bovine pineal glands in a Na+-dependent manner. The Na+-dependent uptake of l -aspartate was strongly inhibited by l -cysteine sulfinate, β-hydroxyaspartate, and l -serine-O-sulfate, inhibitors for the Na+-dependent glutamate/aspartate transporter on the plasma membrane. Na+-dependent sequestration of l -aspartate was also observed in cultured rat pinealocytes, which was inhibited similarly by these transporter inhibitors. These results strongly suggest that l -aspartate is released through microvesicle-mediated exocytosis from pinealocytes and is taken up again through the Na+-dependent transporter at the plasma membrane. The possible role of l -aspartate as an intercellular chemical transmitter in the pineal gland is discussed.  相似文献   

15.
The release of endogenous amino acids from depolarized rat hippocampal mossy fiber synaptosomes was investigated to assess the possible role(s) of glutamate and aspartate in mediating the excitatory mossy fiber synaptic input. The relative proportions of prodynorphin-derived peptides concomitantly released with amino acids were also determined to further characterize the biochemical basis for mossy fiber synaptic transmission. Of the 18 amino acids shown to be present in superfusate fractions by liquid chromatographic analysis, only glutamate was released at a significantly enhanced rate from K+-stimulated (35 mM KCl) mossy fiber nerve endings. The rates of glutamate and aspartate release were increased by 360±27% and 54±12% over baseline respectively. However, the K+-evoked release of glutamate was substantially more Ca2+-dependent (80%) than was the release of aspartate (49%). The veratridine (45 M)-evoked release of both acidic amino acids was entirely blocked by 1 M tetrodotoxin. Depolarization (45 mM KCl) also stimulated the release of the four prodynorphin (Dyn) products examined, in a rank order of Dyn B >> Dyn A(1–17) > Dyn A(1–8) >> Dyn A(1–13), with Dyn B efflux increasing by more than 5-fold over baseline values. These results suggest that the predominant excitatory amino acid in hippocampal mossy fiber synaptic transmission may be glutamate and that this synaptic input may be modulated by at least four different products of prodynorphin processing.The animals involved in this study were procured, maintained and used in accordance with the Animal Welfare Act and the Guide for the Care and Use of Laboratory Animals prepared by the Institute of Laboratory Animal Resources—National Research Council.  相似文献   

16.
Abstract— Superfused slices of the rat dentate gyrus were employed to study the release of GABA, glutamate and aspartate, which are considered strong neurotransmitter candidates in this region. The introduction of Ca2+ to a Ca2+-free superfusion medium containing a depolarizing agent augmented the efflux of all three amino acids. The response to application of Ca2+ nearly always occurred within 30 s, the shortest interval tested in these studies. The efflux rate reached a peak within 90 s and then declined to a level slightly greater than the prestimulation baseline. The failure to maintain the maximal rate with continued exposure to Ca2+ and depolarizing influences appeared not to result from a reduction in Ca2+ permeability caused by continuous depolarization. Ca2+ also stimulated the efflux of exogenously loaded radiolabeled GABA, glutamate and aspartate, but not proline. Exogenously loaded GABA was more readily released than endogenous GABA. Otherwise the effects of various treatments on their efflux rates were qualitatively similar. Mg2+ inhibited Ca2+-dependent efflux. Ba2+, but not Mg2+, stimulated amino acid efflux in the absence of Ca2+. Extracellular Na+ was not required to support Ca2+-dependent efflux. Addition of Ca2+ to a Ca2+-free medium in the absence of a depolarizing agent released GABA from the slices, but not glutamate or aspartate. K+-enriched medium and the depolarizing alkaloid, veratridine, stimulated both Ca2+-dependent and Ca2+-independent release processes. Na+-free medium enhanced the Ca2+-independent releasing action of elevated K+. Ca2+-independent release was inhibited by raising the Mg2+ concentration by 15 or 30 mM and appeared to be inhibited by Ca2+ as well. Amino acid output in the absence of Ca2+ is probably not directly related to transmission and is considered to result partially from a general increase in membrane permeability induced by depolarization in a Ca2+-free medium and partially from stimulation of carrier-mediated amino acid efflux. These results support previously suggested transmitter roles for GABA, glutamate and aspartate in the rat dentate gyrus.  相似文献   

17.
Efflux of various amino acids from slices of rat cerebellar hemispheres was determined under resting or depolarizing conditions. It was increased under high K+(50 mM) as compared to low K+ (5 mM) conditions by 1258 pmol/mg protein for aspartate, 478 for gamma-aminobutyric acid (GABA), 44,693 for glutamate, and 615 for glycine. These were significantly higher than the corresponding values obtained under low-Ca2+ (0.1 mM), high-Mg2+ (12 mM) conditions, whereas for 11 other amino acids the K+-induced efflux was similar under normal and low-Ca2+ concentrations. The K+-induced efflux of exogenously accumulated L-[3H]aspartate, D-[3H]aspartate, and L-[3H]glutamate was higher by factors of 2, 5.8, and 6.3, respectively, under normal Ca2+ conditions, as compared with low-Ca2+, high-Mg2+ conditions. After climbing fibre degeneration induced by destruction of the inferior olive with 3-acetylpyridine, release of endogenous aspartate and exogenous L-[3H]glutamate and D-[3H]aspartate was significantly reduced, by 26%, 38%, and 27%, respectively. These results support the hypothesis that climbing fibres may use aspartate or a related compound as a neurotransmitter. In rat cerebellar tissue, L-[3H]glutamate and L-[3H]aspartate differ in several aspects: (1) L-[3H]glutamate uptake was 4 times higher than that of L-[3H]aspartate; (2) fractional rate constant of K+- evoked release of L-[3H]aspartate was 7% X 2.5 min-1, and of L-[3H]glutamate 36% X 2.5 min-1; and (3) specific activity of L-[3H]glutamate in the eluate collected during K+ stimulation was 3.5 times the value in the tissue, whereas for L-[3H]aspartate, specific activities in the eluate and tissue were similar.  相似文献   

18.
The liver is the organ responsible for bisphenol A (BPA) metabolism, an environmental chemical agent. Exposure to this toxin is associated with liver abnormalities and dysfunction. An important role played by excitatory amino acid transporters (EAATs) of the slc1 gene family has been reported in liver injuries. To gain insight into a plausible effect of BPA exposure in the liver glutamate/aspartate transport, using the human hepatoblastoma cell line HepG2, we report a BPA‐dependent dynamic regulation of SLC1A3 and SLC1A2. Through the use of radioactive [3H]‐ d ‐aspartate uptake experiments and immunochemical approaches, we characterized time and dose‐dependent regulation of the protein levels and function of these transporters after acute exposure to BPA. An increase in nuclear Yin Yang 1 was found. These results suggest an important involvement of the EAATs in liver physiology and its disruption after acute BPA exposure.  相似文献   

19.
Ascidia callosa sperm are triggered to undergo initiation of the sperm reaction (mitochondrial swelling) by increasing the pH or lowering the Na+ concentration of the medium. The optimal [Na+] for acid release is 20 mM with excellent correlation between acid release and initiation of morphological changes. Increasing the [K+] to around 20 mM inhibits acid release when applied up to 1 min after triggering the sperm but with less inhibition at 2 and 4 min, suggesting that K+ inhibits initiation of acid release rather than acid release itself. Acid release and the sperm reaction can also be triggered by Cl?-free (NO?3 or glutamate substituted) seawater (SW). Cl? efflux accompanies H+ efflux with twice as many Cl? being released as H+. Both H+ and Cl? release in Cl?-free SW are dependent upon CO2 being present in HCO?3-free medium, suggesting that H+ efflux is in part Cl? and HCO?3-mediated. However, the chloride channel blocking agent SITS has no effect on H+ release and augments Cl? release. Acid release results in a substantial increase in internal pH as determined by partitioning of 9-amino acridine. We envision acid release from ascidian sperm as involving two systems, the Na+-dependent acidification system of unreacted sperm and the Cl?- and HCO?3-mediated H+ release at activation. The mechanism controlling acid release would then involve inactivation of the internal acidification process and activation of the chloride-bicarbonate-mediated alkalinization process.  相似文献   

20.
Aims: This study focused on the cloning, expression and characterization of recombinant α‐l ‐arabinosidases from Bifidobacterium longum H‐1. Methods and Results: α‐l ‐Arabinofuranosidase (AfuB‐H1) and bifunctional α‐l ‐arabinopyranosidase/β‐d ‐galactosidase (Apy‐H1) from B. longum H‐1 were identified by Southern blotting, and their recombinant enzymes were overexpressed in Escherichia coli BL21 (DE3). Recombinant AfuB‐H1 (rAfuB‐H1) was purified by single‐step Ni2+‐affinity column chromatography, whereas recombinant Apy‐H1 (rApy‐H1) was purified by serial Q‐HP and Ni2+‐affinity column chromatography. Enzymatic properties and substrate specificities of the two enzymes were assessed, and their kinetic constants were calculated. According to the results, rAfuB‐H1 hydrolysed p‐nitrophenyl‐α‐l ‐arabinofuranoside (pNP‐αL‐Af) and ginsenoside Rc, but did not hydrolyse p‐nitrophenyl‐α‐l ‐arabinopyranoside (pNP‐αL‐Ap). On the other hand, rApy‐H1 hydrolysed pNP‐αL‐Ap, p‐nitrophenyl‐β‐d ‐galactopyranoside (pNP‐βD‐Ga) and ginsenoside Rb2. Conclusions: Ginsenoside‐metabolizing bifidobacterial rAfuB‐H1 and rApy‐H1 were successfully cloned, expressed, and characterized. rAfuB‐H1 specifically recognized the α‐l ‐arabinofuranoside, whereas rApy‐H1 had dual functions, that is, it could hydrolyse both β‐d ‐galactopyranoside and α‐l ‐arabinopyranoside. Significance and Impact of the Study: These findings suggest that the biochemical properties and substrate specificities of these recombinant enzymes differ from those of previously identified α‐l ‐arabinosidases from Bifidobacterium breve K‐110 and Clostridium cellulovorans.  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