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.
Abstract: [3H] γ -Aminobutyric acid ([3H]GABA) binding to purified lipids was examined in an organic solvent-aqueous partition system. In addition, the [3H]GABA binding capacity in the partition system was compared with the capacity of lipids to alter sodium-dependent [3H]GABA uptake into synaptosomes isolated from rat whole brains. [3H]GABA was found to bind to all of the lipids studied in the organic solvent-aqueous partition system [phosphatidic acid (PA), phosphatidylethanolamine (PE), phosphatidylinositol (PI), phosphatidylserine (PS), gangliosides, and sulfatide], although PS exhibited the greatest binding capacity. [3H]GABA uptake into synaptosomes was enhanced by PS (48.0%) but was not altered by any other lipid. PS enhancement of [3H]GABA uptake required the presence of sodium and was blocked by nipecotic acid (10 μ m ). These results suggest that PS may play a role in the sodium-dependent GABA reuptake process in the presynaptic nerve end.  相似文献   

2.
The interaction of the negatively-charged phosphatidylserine (PS) and gamma-Aminobutyric acid (GABA) is examined in black lipid membranes (BLM) and inverse micelles. GABA does not permeate through PS membranes and, in concentrations of 10(-5)-10(-4) M, it reduces the negative potential at the membrane-aqueous solution interface. The effect is owing to the adsorption of the GABA cationic species and the consequent decrease of the negative surface charge density of the membrane. When the intrinsic pH of the membrane-solution interface is considered, the Gouy-Chapman-Stern theory describes the GABA screening effect and makes it possible to calculate the GABA-PS binding constant. This value is compared with that obtained measuring the partition of 14C-GABA between an organic phase containing PS and the aqueous solution. The results presented strongly suggest that the electrostatic force plays a major role in GABA-PS interaction.  相似文献   

3.
1. 3H-gamma-Aminobutyric acid (GABA) release elicited by a depolarizing K+ stimulus or by noradrenergic transmitter was examined in rat pineals in vitro. 2. The release of 3H-GABA was detectable at a 20 mM K+ concentration in medium and increased steadily up to 80 mM K+. 3. In a Ca2+-free medium 3H-GABA release elicited by 30 mM K+, but not that elicited by 50 mM K+, became blunted. 4. Norepinephrine (NE; 10(-6)-10(-4) M) stimulated 3H-GABA release from rat pineal explants in a dose-dependent manner. 5. The activity of 10(-5) M NE on pineal GABA release was suppressed by equimolecular amounts of prazosin or phentolamine (alpha 1- and alpha 1/alpha 2-adrenoceptor blockers, respectively) and was unaffected by propranolol (beta-adrenoceptor blocker). 6. The alpha 1-adrenoceptor agonist phenylephrine (10(-7)-10(-5) M) and the beta-adrenoceptor agonist isoproterenol (10(-5) M) mimicked the GABA releasing activity of NE, while 10(-7) M isoproterenol failed to affect it; the alpha 2-adrenoceptor agonist clonidine (10(-7)-10(-5) M) did not modify 3H-GABA release. 7. The addition of 10(-4) M GABA or of the GABA transaminase inhibitor gamma-acetylenic GABA or aminooxyacetic acid inhibited the melatonin content and/or release to the medium in rat pineal organotypic cultures. 8. GABA at concentrations of 10(-5) M or greater partially inhibited the NE-induced increase in melatonin production by pineal explants. 9. The depressant effect of GABA on melatonin production was inhibited by the GABA type A receptor antagonist bicuculline; bicuculline alone increased the pineal melatonin content. Baclofen, a GABA type B receptor agonist, did not affect the pineal melatonin content or release. 10. The decrease in serotonin (5-HT) content of rat pineal explants brought about by NE was not modified by GABA; GABA by itself increased 5-HT levels. 11. These results indicate that (a) GABA is released from rat pineals by a depolarizing stimulus of K+ through a mechanism which is partially Ca2+ dependent; (b) NE releases rat pineal GABA via interaction with alpha 1-adrenoceptors; (c) GABA inhibits melatonin production in vitro via interaction with GABA type A receptor sites; and (d) GABA's effect on NE-induced melatonin release does not correlate with the lack of effect on the NE-induced decrease in pineal 5-HT content.  相似文献   

4.
In the present study, existence of (3)H-GABA uptake mechanism in bovine spermatozoa and the modulation of (3)H-GABA transport by GABA itself were evaluated. The hypothesis was tyrosine phosphorylation affects transporter (GAT) function. (3)H-GABA uptake assays were performed on bovine spermatozoa and it resulted to be temperature- and time-dependent and K(m) was 1.48muM. Uptake was inhibited by the metabolic inhibitor ouabain and different blockers of GAT-1 (beta-alanine, l-DABA, nipecotic acid, tiagabine). Extracellular GABA up-regulated GABA transport, while the addition of SKF89976A, a high affinity inhibitor of the rat brain GABA transporter, reduced GABA uptake. Tyrosine phosphorylation affects transporter function since genistein, a broad-spectrum tyrosine kinase inhibitor, decreased (3)H-GABA uptake. Reduction in uptake did not occur in the presence of daidzein, an inactive genistein analogue. Furthermore, the genistein-mediated reduction in transport could be prevented by the tyrosine phosphatase inhibitor pervanadate. The action of these drugs on GABA transport is likely mediated through the GABA transporter GAT-1 since SKF89976A blocked a majority of GABA uptake. Wash-out experiments indicated that the genistein effect was reversible. When the experiments were conducted using "in vitro" capacitated spermatozoa there was no detectable uptake. Present results demonstrate that the carrier-mediated GABA uptake system in bovine spermatozoa modulates its function in response to extracellular GABA, that changes in lipid distribution and membrane composition which occur during capacitation eliminates GABA uptake and suggest the involvement of tyrosine phosphorylation in GABA transport.  相似文献   

5.
The efflux of 45Ca2+ and 3H-GABA from cat cerebral cortex has been measured in vivo by a superfusion technique. Under control conditions, the efflux of both these isotopes is sensitive to increments in the calcium concentration of the superfusion medium but insensitive to the addition of 25 mM GABA. After 40 min of contact with medium containing 25 mM glutamate, however, neither 45Ca2+ nor 3H-GABA efflux is affected by added Ca2+ but 3H-GABA efflux becomes considerably enhanced by 25 mM GABA. The mean membrane potentials at this time were found to be less than controls. Depolarization generally continued past this time, recovery beginning 150 min after superfusion with glutamate. Peaks in 45Ca2+ efflux were also correlated with seizure activity which was sometimes elicited by topical glutamate. It is suggested that the transport systems for Ca2+ and GABA may be sensitive to states of neural activity, such as spreading depression.  相似文献   

6.
There are now increasing evidences suggesting that GABA is able of direct interaction with certain endocrine cells. In the present study, highly specific anti-GABA-glutaraldehyde antibodies and 3H-GABA uptake were used at the light and electron microscope levels to investigate the occurrence of cells containing endogenous GABA or taking up exogenous GABA in the mucosal antrum and corpus of the rat stomach. Only certain endocrine cell types of both regions were immunostained or grain-labelled. However, the morphology of their secretory granules did not allow to identify the nature of their hormone with certainty but suggested that somatostatin-like cells could interact with GABA. The combination of gastrin and somatostatin immunodetection with 3H-GABA uptake autoradiography at the light microscope level, revealed that a subpopulation of somatostatin-like cells and other still unidentified endocrine cells are able to take up GABA, while the gastrin-like cells are not. These results reinforce the hypothesis that certain endocrine cell types of the diffuse endocrine system of the digestive tract are able to directly interact with GABA.  相似文献   

7.
The aim of the present paper was to determine whether the release of glutamate from putative "glutamergic" terminals in the cerebellum is influenced by gamma-aminobutyric acid (GABA). In a group of preliminary experiments, we present biochemical evidence in favour of a neurotransmitter role of glutamate in the cerebellum: (1) endogenous glutamate was released from depolarized cerebellar synaptosomal preparations in a Ca2+-dependent away; (2) [14C]glutamate was synthesized from [14C]glutamine in cerebellar synaptosomes, and the newly synthesized [14C]glutamate was released released in a Ca2+-dependent way; (3) the elevation of cyclic GMP elicited by depolarization of cerebellar slices in the presence of Ca2+ was partly reversed by the glutamate antagonist glutamic acid diethyl ester, which probably prevented the interaction of endogenously released glutamate with postsynaptic receptors. GABA and muscimol at low concentrations (2--20 micrometers) potentiated the depolarization-induced release of D-[3H]aspartate (a glutamate analogue which labels the glutamate "reuptake pool") from cerebellar synaptosomes. The effect was concentration dependent and was largely prevented by two GABA antagonists, bicuculline and picrotoxin. The stimulation of D-[3H]aspartate release evoked by muscimol was linearly related to the logarithm of K+ concentration in the depolarizing medium. GABA did not affect the overall release of endogenous glutamate, but potentiated, in a picrotoxin-sensitive manner, the depolarization-evoked release of [14C]glutamate previously synthesized from [14C]glutamine. Since nerve endings are the major site of glutamate synthesis from glutamine, GABA and muscimol appear to exert their stimulatory effect at the level of "glutamergic" nerve terminals, probably after interacting with presynaptic GABA receptors. The possible functional significance of these findings is briefly discussed.  相似文献   

8.
1. Cerebral-cortex slices prelabelled with gamma-amino[1-(14)C]butyrate (GABA) were incubated in a glucose-saline medium. After the initial rapid uptake there was no appreciable re-entry of (14)C into the GABA pool, either from the medium or from labelled metabolites formed in the tissue. The kinetic constants of GABA metabolism were determined by computer simulation of the experimental results by using mathematical procedures. The GABA flux was estimated to be 0.03mumol per min/g, or about 8% of the total flux through the tricarboxylic acid cycle. It was found that the assumption of compartmentation did not greatly affect the estimates of the GABA flux. 2. The time-course of incorporation of (14)C into amino acids associated with the tricarboxylic acid cycle was followed with [1-(14)C]GABA and [U-(14)C]-glucose as labelled substrates. The results were consistent with the utilization of GABA via succinate. This was confirmed by determining the position of (14)C in the carbon skeletons of aspartate and glutamate formed after the oxidation of [1-(14)C]GABA. These results also indicated that under the experimental conditions the reversal of reactions catalysed by alpha-oxoglutarate dehydrogenase and glutamate decarboxylase respectively was negligible. The conversion of [(14)C]GABA into gamma-hydroxybutyrate was probably also of minor importance, but decarboxylation of oxaloacetate did occur at a relatively slow rate. 3. When [1-(14)C]GABA was the labelled substrate there was evidence of a metabolic compartmentation of glutamate since, even before the peak of the incorporation of (14)C into glutamate had been reached, the glutamine/glutamate specific-radioactivity ratio was greater than unity. When [U-(14)C]glucose was oxidized this ratio was less than unity. The heterogeneity of the glutamate pool was indicated also by the relatively high specific radioactivity of GABA, which was comparable with that of aspartate during the whole incubation time (40min). The rates of equilibration of labelled amino acids between slice and medium gave evidence that the permeability properties of the glutamate compartments labelled as a result of oxidation of [1-(14)C]GABA were different from those labelled by the metabolism of [(14)C]glucose. The results showed therefore that in brain tissue incubated under the conditions used, the organization underlying metabolic compartmentation was preserved. The observed concentration ratios of amino acids between tissue and medium were also similar to those obtaining in vivo. These ratios decreased in the order: GABA>acidic acids>neutral amino acids>glutamine. 4. The approximate pool sizes of the amino acids in the different metabolic compartments were calculated. The glutamate content of the pool responsible for most of the labelling of glutamine during oxidation of [1-(14)C]GABA was estimated to be not more than 30% of the total tissue glutamate. The GABA content of the ;transmitter pool' was estimated to be 25-30% of the total GABA in the tissue. The structural correlates of metabolic compartmentation were considered.  相似文献   

9.
Summary There are now increasing evidences suggesting that GABA is able of direct interaction with certain endocrine cells. In the present study, highly specific anti-GABA-glutaraldehyde antibodies and 3H-GABA uptake were used at the light and electron microscope levels to investigate the occurrence of cells containing endogenous GABA or taking up exogenous GABA in the mucosal antrum and corpus of the rat stomach. Only certain endocrine cell types of both regions were immunostained or grain-labelled. However, the morphology of their secretory granules did not allow to identify the nature of their hormone with certainty but suggested that somatostatin-like cells could interact with GABA. The combination of gastrin and somatostatin immunodetection with 3H-GABA uptake autoradiography at the light microscope level, revealed that a subpopulation of somatostatin-like cells and other still unidentified endocrine cells are able to take up GABA, while the gastrin-like cells are not. These results reinforce the hypothesis that certain endocrine cell types of the diffuse endocrine system of the digestive tract are able to directly interact with GABA.  相似文献   

10.
Abstract—
  • 1 Metabolism of [2-14C]pyruvate, [1-14C]acetate and [5-14C]citrate in the rat cerebral cortex slices was studied in the presence of halothane. Metabolites assayed include acetylcholine (ACh), citrate, glutamate, glutamine, γ-aminobutyrate (GABA) and aspartate. The trichloroacetic acid soluble extract, the trichloroacetic acid insoluble precipitate and its lipid extract were also studied.
  • 2 In control experiments, pyruvate preferentially labelled ACh, citrate, glutamate, GABA and aspartate. Acetate labeled ACh, but to a lesser extent than pyruvate. Acetate also labeled lipids and glutamine. Citrate labeled lipids but not ACh and served as a preferential precursor for glutamine. These data support a three-compartment model for cerebral tricarboxylic acid cycle metabolism.
  • 3 Halothane caused increases in GABA and aspartate contents and a decrease in ACh content. It has no effect on the contents of citrate, glutamate and glutamine.
  • 4 Halothane preferentially inhibited the metabolic transfer of radioactivity from pyruvate into almost all metabolites, an effect probably not related to pyruvate permeability. This is interpreted as halothane depression of the‘large metabolic compartment’ which includes the nerve endings.
  • 5 Halothane increased the metabolic transfer of radioactivity from acetate into lipids but did not alter such a transfer into the trichloracetic acid extract.
  • 6 Halothane increased the metabolic transfer of radioactivity from citrate into the trichloroacetic acid precipitate, lipids and especially glutamine. Transfer of citrate radioactivity into GABA was somewhat decreased.
  • 7 The differential effects of halothane on acetate and citrate utilization suggest that the ‘small metabolic compartment’ should be subdivided. Therefore, at least three metabolic compartments are demonstrated.
  • 8 Halothane did not interfere with the dicarboxylic acid portion of the tricarboxylic acid cycle.
  相似文献   

11.
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   

12.
—The uptake of [U-14C]glutamate into the amphibian brain was studied in vitro using brains from toads (Bufo boreas) adapted either to a fresh water (FWA) or an hyperosmotic saline (HOA) environment. Initial rates of 14C-glutamate uptake showed a single apparent Km of about 0·2 mm . Uptake by HOA brains was slower than that by FWA brains, reflecting perhaps a non-competitive type of inhibition by the higher content of glutamate in the HOA brains. Although the glutamate content of HOA brains was maintained during prolonged incubation at twice the level found in FWA toads, other metabolic parameters measured in the two types of brain preparations were surprisingly similar. Tissue to medium concentration ratios of greater than 3000:1 were generated by both FWA and HOA brains. In both brain systems the clearance of glutamate from the medium was accompanied by a rapid conversion of the amino acid to glutamine and its release into the medium. In both the FWA and HOA toad brain systems some [U-14C]glutamate was metabolized to aspartate and GABA; in both systems the specific radioactivity (SA) of glutamine in the tissue was from two to four times greater than that of glutamate; also the SA of glutamine released into the medium was higher by several orders of magnitude than the SA of glutamine in brain tissues. These and other findings support the concept that, in both the FWA and HOA toad brains, transport processes are instrumental in preserving low extracellular levels of glutamate but that mechanisms other than transport are responsible for the maintenance of different levels of glutamate in the FWA and HOA toad brains.  相似文献   

13.
Isolated synaptosomes were used to study the problem of net accumulation of neurotransmitters. The time-course and the kinetics of exogenous and endogenous GABA transport were studied by liquid-scintillation counting and HPLC-amino acid analysis respectively. Different pools of GABA were suggested by a 6-fold difference in tissue-to-medium-ratio of endogenous vs. exogenous GABA. Net accumulation, exchange and net efflux of GABA was found to be a function of the GABA concentration in the incubation medium. The Kms for net accumulation and for 3H-GABA accumulation were 2.68 +/- 1.16 and 6.19 +/- 1.26 microM respectively, whereas the Vmaxs were 5.9 +/- 4.9 and 134 +/- 13 pmol/mg w.w. min respectively. This means that the transport studies which use exogenous substances (e.g. 3H-GABA) considerably overestimate the transport by overlooking the magnitude of the counter transport.  相似文献   

14.
The relationship between neuronal glutamate turnover, the glutamate/glutamine cycle and de novo glutamate synthesis was examined using two different model systems, freshly dissected rat retinas ex vivo and in vivo perfused rat brains. In the ex vivo rat retina, dual kinetic control of de novo glutamate synthesis by pyruvate carboxylation and transamination of alpha-ketoglutarate to glutamate was demonstrated. Rate limitation at the transaminase step is likely imposed by the limited supply of amino acids which provide the alpha-amino group to glutamate. Measurements of synthesis of (14)C-glutamate and of (14)C-glutamine from H(14)CO(3) have shown that (14)C-amino acid synthesis increased 70% by raising medium pyruvate from 0.2 to 5 mM. The specific radioactivity of (14)C-glutamine indicated that approximately 30% of glutamine was derived from (14)CO(2) fixation. Using gabapentin, an inhibitor of the cytosolic branched-chain aminotransferase, synthesis of (14)C-glutamate and (14)C-glutamine from H(14)CO(3)(-) was inhibited by 31%. These results suggest that transamination of alpha-ketoglutarate to glutamate in Müller cells is slow, the supply of branched-chain amino acids may limit flux, and that branched-chain amino acids are an obligatory source of the nitrogen required for optimal rates of de novo glutamate synthesis. Kinetic analysis suggests that the glutamate/glutamine cycle accounts for 15% of total neuronal glutamate turnover in the ex vivo retina. To examine the contribution of the glutamate/glutamine cycle to glutamate turnover in the whole brain in vivo, rats were infused intravenously with H(14)CO(3)(-). (14)C-metabolites in brain extracts were measured to determine net incorporation of (14)CO(2) and specific radioactivity of glutamate and glutamine. The results indicate that 23% of glutamine in the brain in vivo is derived from (14)CO(2) fixation. Using published values for whole brain neuronal glutamate turnover, we calculated that the glutamate/glutamine cycle accounts for approximately 60% of total neuronal turnover. Finally, differences between glutamine/glutamate cycle rates in these two model systems suggest that the cycle is closely linked to neuronal activity.  相似文献   

15.
It has been shown in experiments on rat cortex slices preincubated with 3H-GABA that chlorodiazepoxide (10(-6), 3.10(-5) M) does not change basal and electric stimulation-induced release of the label. It has been also shown that it does not eliminate the autoinhibitory effect of GABA on electric stimulation-induced release of 3H-GABA. However, harmane and some other (but not all) derivatives given at the same concentrations increase 3H-GABA release induced by electric stimulation and abolish the inhibitory effect of GABA without changing or slightly raising spontaneous release of 3H-GABA. It is concluded that harmane enhances the electrically stimulated release of the transmitter by GABAergic axons whatever the effect on benzodiazepine-binding areas of GABA receptors.  相似文献   

16.
M S Starr 《Life sciences》1985,37(24):2249-2255
Slices of rat substantia nigra were preloaded with tritiated gamma-aminobutyrate (GABA) or dopamine (DA) and perfused with Krebs solution containing 5 microM aminooxyacetic acid or 10 microM nialamide to inhibit the catabolism of GABA and DA respectively. Repeated brief exposures to high potassium medium (+ 30 mM K+ for 1 min) evoked a consistent pattern of calcium-dependent 3H efflux against which the effects of opiates (10-400 microM) were assessed. Opiate agonists inhibited K+-induced 3H-GABA efflux in the following decreasing order of potency: bremazocine greater than D-Ala2-Met5-enkephalinamide (ENK) greater than SKF 10047 much greater than morphine, consistent with the participation of kappa, delta, sigma and to a lesser extent mu opiate receptors respectively. Naloxone (1 microM) partially antagonised the response to morphine and ENK, while ICI 154129 attenuated ENK only. Save for a GABA-releasing action of SKF 10047 at high doses, none of the compounds altered basal outflow of 3H-GABA. Naloxone, in the dose range 10-400 microM, also significantly inhibited depolarisation-induced release of 3H-GABA. In parallel experiments none of the compounds tested were found to influence 3H-DA release in concentrations up to 40 microM, but thereafter suppressed K+-induced 3H-DA outflow indiscriminately. The results are discussed with reference to the possible mechanism(s) via which injected and endogenous opiates may affect motor performance by attenuating GABA transmission in the nigra.  相似文献   

17.
Conversion of γ-Hydroxybutyrate to γ-Aminobutyrate In Vitro   总被引:3,自引:3,他引:0  
[3H]gamma-Hydroxybutyric acid [( 3H]GHB) at physiological concentration incubated with brain slices in Krebs-Ringer medium produced [3H]gamma-aminobutyric acid [( 3H]GABA). This compound was identified by its Rf values on thin-layer chromatograms and by analysis of the dansyl derivatives of the free amino acid fraction. No labelled glutamate could be detected. Brain slices incubated with labelled glutamate and nonradioactive GHB generated labelled 2-oxoglutarate, suggesting that gamma-aminobutyrate-2-oxoglutarate transaminase (GABA-T) is involved in catalyzing this reaction. Furthermore, specific inhibitors of GABA-T blocked the production of labelled GABA from labelled GHB and of labelled 2-oxoglutarate from labelled glutamate. Transformation of [3H]GHB into [3H]GABA was not inhibited by malonate, demonstrating that the succinate-linked pathway is not involved in the generation of GABA. The kinetic characteristics of the multienzyme system involved in GHB degradation studied in vitro are compatible with the production of GABA in vivo.  相似文献   

18.
Abstract— The amino acid and carbohydrate metabolism of confluent cultures of C-6 glioma cells has been investigated. It was observed that the presence of glutamine in the incubation fluid was essential to maintain high glutamine levels in the cells during a 2 h incubation. When cells were incubated in a cerebrospinal fluid-like medium glutamate, glutamine, aspartate and γ-aminobutyrate (GABA) levels were comparable to those occurring in whole forebrain of adult rat in vivo. Glucose uptake was high, approx 1 μmol/mg protein/2 h, 50% of which was accounted for by lactate production. Of the remaining glucose uptake a substantial proportion was unaccounted for by known oxygen-coupled citric acid cycle flux, or glycogen or amino acid synthesis. Interestingly, the cells released into the medium significant amounts of the neuroinhibitory amino acids, GABA and glycine, and rapidly cleared the medium of the neuroexcitatory amino acids glutamate and aspartate. Metabolism of [2-14C]glucose and [3H]acetate by the cells indicated rapid labelling of the glutamate and aspartate pools of the cells by glucose in 1 h, but the relative specific activities of glutamine and GABA were much lower. The metabolism of tracer concentrations of [3H]acetate to glutamate by the cells indicated greater dilution of this isotope compared to that of labelled glucose. However, the ratio of 3H to 14C radioactivity in glutamate and other amino acids was similar to that in the mixture of glucose and acetate added to the medium. Therefore, some active route of acetate metabolism which communicates metabolically with the route of glucose metabolism to glutamate appears to exist in the cells. Significant acetate activation and fatty acid turnover would explain the present results. Some of the amino acid labelling patterns observed in these studies are not consistent with these glial-like cells behaving as models for the small compartment of amino acid metabolism in brain. Enzyme measurements corroborated the metabolic studies. Glutamate decarboxylase activity was 3–10% of the level found in whole brain. GABA transaminase was also low compared to brain as was glutamine synthetase. Glutamate dehydrogenase was present at levels equal to or higher than those of whole brain.  相似文献   

19.
Saffrey  M. J.  Marcus  N.  Jessen  K. R.  Burnstock  G. 《Cell and tissue research》1983,234(1):231-235
The occurrence of neurons possessing high-affinity uptake sites for GABA was studied in the myenteric plexus of the guinea-pig ileum, caecum, and proximal and distal colon, the rat proximal colon, and the chicken gizzard with the use of 3H-GABA and autoradiography. Experiments were carried out on plexuses that had been freshly isolated from the gut wall or on isolated plexuses that had been maintained as explant cultures for 7 to 14 days. Scattered neurons selectively labelled with 3H-GABA were found in the myenteric plexuses from all the areas examined. The results suggest that GABAergic neurons are widely distributed in the enteric nervous system.  相似文献   

20.
Cyclo[VKLdKVdYPLKVKLdYP] (GS14dK(4)), a synthetic tetradecameric ring-size analogue of the naturally occurring antimicrobial peptide gramicidin S (GS), retains the strong antimicrobial activity of GS but is 15-20 times less hemolytic. To characterize its interaction with lipid membranes and to understand the molecular basis of its capacity to lyse bacterial cells, in preference to erythrocytes, we have investigated the interactions of GS14dK(4) with detergent micelles and with lipid bilayer model membranes by Fourier transform infrared spectroscopy and compared our results with those of a similar study of GS [Lewis, R. N. A. H., et al. (1999) Biochemistry 38, 15193-15203]. In both aqueous and organic solvent solutions, GS14dK(4) adopts a beta-sheet conformation that is somewhat distorted and more sensitive to the polarity of its environment than GS. Like GS, GS14dK(4) is completely or partially excluded from gel-state lipid bilayers but interacts strongly with liquid-crystalline lipid bilayers and detergent micelle, and interacts more strongly with more fluid liquid-crystalline lipid systems. However, its interactions are more strongly influenced by membrane lipid order and fluidity, and unlike GS, it is essentially excluded from cholesterol-containing phospholipid bilayers. Also, GS14dK(4) is excluded from cationic lipid bilayers, but partitions more strongly and/or penetrates more deeply into anionic lipid bilayers than into those composed of either zwitterionic or nonionic lipids. Anionic lipids also facilitate GS14dK(4) interactions with multicomponent lipid bilayers which are predominantly zwitterionic or nonionic. Although GS14dK(4) generally penetrates and/or partitions into zwitterionic or uncharged lipid bilayers less strongly than does GS, its greater size and altered distribution of positive charges make it intrinsically more perturbing with regard to membrane organization once associated with lipid bilayers. This fact, combined with its relatively strong interactions with anionic phospholipids, may explain why GS14dK(4) retains relatively high antimicrobial activity. However, its low hemolytic activity is probably largely attributable to its low propensity to penetrate and/or partition into cholesterol-containing zwitterionic lipid membranes.  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