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.
Cholinergic nerve terminals were affinity purified from rat caudate nucleus. On stimulation with both 22.6 mM KCl and 50 microM veratridine, ATP was released in a Ca2+-dependent manner. The molar ratio of released acetylcholine to ATP (9:1) was closer to that found in isolated cholinergic vesicles (7:1) than whole terminals (3:1). Extracellular [14C]ATP was rapidly metabolized by these terminals to adenosine and inosine via ectonucleotidases. The terminals had a saturable, high-affinity uptake mechanism for adenosine (Km = 16.6 microM). Veratridine stimulation also caused the Ca2+-dependent release of nucleosides in a dipyridamole-sensitive manner. Both theophylline treatment and inhibition of extracellular ATP breakdown resulted in increased ATP and nucleoside release. Extracellular adenosine was shown to inhibit acetylcholine release, probably via the A1 receptor. The role of extracellular purines at the cholinergic nerve terminal is discussed.  相似文献   

2.
The release of 3H-labeled purines at rest and during electrical stimulation was investigated in slices of rat cortex prelabeled with [3H]adenine and perfused with Krebs solution. A linear relationship was found between radioactivity efflux and stimulation frequency from 2.5 to 20 Hz. At frequencies of less than 2.5 Hz, no increase in radioactivity efflux was detected. The amount of tritium released per pulse increased with stimulation frequency up to 10 Hz and declined at 20 Hz. The tritium efflux from the slices at rest and at a stimulation frequency of 10 Hz, analyzed by HPLC with ultraviolet absorbance detection at 254 nm, consisted mostly of adenosine, inosine, and hypoxanthine. The 3H-labeled purine release evoked by 10-Hz stimulation increased with current intensity from 15 to 100 mA/cm2. At 20 mA/cm2, addition of 0.5 microM tetrodotoxin to the superfusing Krebs solution brought about a 98% decrease of 3H-labeled purine release. At higher current strength, the percentage of tetrodotoxin-sensitive-evoked tritium efflux was smaller. At 30 mA/cm2, 86% of the evoked release was tetrodotoxin sensitive. Under these stimulation conditions, tritium efflux showed a 69% decrease when the slices were superfused with calcium-free Krebs solution containing 0.5 mM EGTA. The muscarinic agonist oxotremorine (30 microM) significantly enhanced the 10-Hz-stimulated 3H-labeled purine release. The effect of oxotremorine was partially prevented by tetrodotoxin, was antagonized by atropine (1.5 microM), and was mimicked by addition of physostigmine (3.8 microM) to the superfusion fluid. Atropine alone did not affect the evoked release, and none of the drugs modified the basal tritium efflux.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
The purines ATP and adenosine can act as a coordinated team of transmitters. As extracellular adenosine is frequently derived from the enzymatic dephosphorylation of released ATP, the distinct actions of the two purines can be synchronized. In retinal ganglion cells (RGCs), stimulation of the P2X7 receptor for ATP leads to increased intracellular Ca2+ and death. Here we define the contrasting effects of adenosine and identify protective actions mediated by the A3 receptor. Adenosine attenuated the rise in Ca2+ produced by the P2X7 agonist 3'-O-(4-benzoylbenzoyl)ATP (BzATP). Adenosine was also neuroprotective, increasing the survival of ganglion cells exposed to BzATP. The A3 adenosine receptor agonist 2-chloro-N6-(3-iodobenzyl)-adenosine-5'-N-methyluronimide (Cl-IB-MECA) mimicked the inhibition of the Ca2+ rise, whereas the A3 antagonist 3-Ethyl-5-benzyl-2-methyl-4-phenylethynyl-6-phenyl-1,4-(+/-)-dihydropyridine-3,5-dicarboxylate (MRS-1191) reduced the protective effects of adenosine. Both Cl-IB-MECA and a second A3 receptor agonist IB-MECA reduced the cell loss triggered by BzATP. The actions of BzATP were mimicked by ATPgammaS, but not by ATP. In summary, adenosine can stop the rise in Ca2+ and cell death resulting from stimulation of the P2X7 receptor on RGCs, with the A3 adenosine receptor contributing to this protection. Hydrolysis of ATP into adenosine and perhaps inosine shifts the balance of purinergic action from that of death to the preservation of life.  相似文献   

4.
Abstract: Propentofylline is a novel neuroprotective agent that has been shown to act as an adenosine transport inhibitor as well as an adenosine receptor antagonist. In the present series of experiments we have compared the effects of propentofylline with those of known adenosine transport inhibitors and receptor antagonists on the formation of adenosine in rat hippocampal slices. The ATP stores were labeled by incubating the slices with [3H]-adenine. The total 3H overflow and the overflow of endogenous and 3H-labeled adenosine, inosine, and hypoxanthine were measured. Adenosine release, secondary to ATP breakdown, was induced both by hypoxia/hypoglycemia and by electrical field stimulation. Propentofylline (20–500 µM) increased the release of endogenous and radiolabeled adenosine, without increasing the total release of purines. Thus, the drug altered the pattern of released purines, i.e., increasing adenosine and decreasing inosine and hypoxanthine. This pattern, which was observed when purine release was induced both by electrical field stimulation and by hypoxia/hypoglycemia, was shared by the nucleoside transport inhibitor dipyridamole (1 µM) and by mioflazine (1 µM) and nitrobenzylthioinosine (1 µM). By contrast, other xanthines, including theophylline (100 µM) and 8-cyclopentyltheophylline (10 µM), enprofylline (100 µM), or torbafylline (300 µM), if anything, increased the total release of purines without alterations of the pattern of release. These results indicate that nucleoside transport inhibitors can decrease the release of purines from cells and at the same time increase the concentration of extracellular adenosine, possibly by preventing its uptake and subsequent metabolism. This change in purine metabolism may be beneficial with regard to cell damage after ischemia. The results also indicate that propentofylline behaves in such a potentially beneficial manner.  相似文献   

5.
ATP promoted biphasic effects on both basal and fMLP-stimulated arachidonic acid (AA) release in neutrophil-like HL60 cells: stimulation in the micromolar range (EC50 = 3.2 +/- 0.9 microM) and inhibition at higher concentrations (EC50 = 90 +/- 11 microM). ATP also inhibited UTP- and platelet activating factor-stimulated AA release. Only stimulatory effects of ATP on basal or fMLP-stimulated phospholipase C were observed. The inhibitory effect of ATP on AA release was not due to reacylation of released AA, chelation of extracellular Ca2+, cell permeabilization, or changes in the rise of [Ca2+]i induced by agonist. The inhibition was rapid, being detected within 5-15 s. The inhibitory effect of ATP on fMLP-stimulated AA release could be desensitized by pretreatment of the cells with 2 mM ATP, but not 20 microM ATP, the concentration that resulted in maximal release of AA and inositol phosphates. The inhibition by ATP was neither dependent on generation of adenosine by ATP hydrolysis nor the result of direct interaction of ATP with P1 purinergic receptors. Among other nucleotides tested (CTP, GTP, ITP, TTP, XTP, adenosine 5'-(beta,gamma-methylene)triphosphate (AMP-PCP), adenyl-5'-yl imidodiphosphate (AMP-P(NH)P), ADP, adenosine 5'-O-(3-thiotriphosphate) (ATP gamma S), and UTP), only UTP and ATP gamma S displayed biphasic effects with potencies and efficacies almost identical to those of ATP. The other nucleotides only exhibited stimulatory effects (EC50 = 60-300 microM). The results are consistent with a model of dual regulation of AA release by two distinct subtypes of P2U receptors in HL60 cells.  相似文献   

6.
ATP and adenosine are important extracellular regulators of glomerular functions. In this study, ATP release from glomeruli suspension and its extracellular metabolism were investigated. Basal extraglomerular ATP concentration (1nM) increased several fold during inhibition of ecto-ATPase activity, reflecting the basal ATP release rate. Mechanical perturbation increased the amounts of ATP released from glomeruli. ATP added to glomeruli was almost completely degraded within 20 minutes. In that time, AMP was the main product of extracellular ATP metabolism. Significant accumulation of AMP was observed after 5 min (194 +/-16 microM) and 20 min (271 +/-11 microM), whereas at the same time concentration of adenosine was only 10 muM. A competitive inhibitor of ecto-5-nucleotidase alpha-beta-methylene-ADP (AOPCP), decreased extraglomerular ATP and adenosine concentration by 80% and 50%, respectively. Similarly, AMP (100 microM) also markedly reduced extraglomerular ATP accumulation, whereas IMP, its deamination product, was not effective. P1, P5-diadenosine pentaphosphate (Ap5A) - an inhibitor of ecto-adenylate kinase prevented significantly the disappearance of ATP from extraglomerular media caused by AMP. These findings demonstrate that the decrease in extracellular ATP concentration observed after addition of AOPCP or AMP is caused by the presence of ecto-adenylate kinase activity in the glomeruli. The enzyme catalyses reversible reaction 2ADP<->ATP+AMP, and a rise in the AMP concentration can lead to fall in ATP level. The present study provides evidence the extraglomerular accumulation of ATP reflects both release of ATP from glomeruli cells and its metabolism by ecto-enzymes. Our data suggest that AMP, produced from ATP in the Bowman's capsular space, might plays a dual role as a substrate for ecto-adenylate kinase and ecto-nucleotidase reactions being responsible for the regulation of intracapsular ATP and adenosine concentration. We conclude that AMP degrading and converting ecto-enzymes effectively determine the balance between ATP and adenosine concentration and thus the activation of P2 and/or adenosine receptors.  相似文献   

7.
Human adipocytes are of limited viability (7 +/- 2% release of lactate dehydrogenase/h) and contain active ectophosphatases which are capable of sequentially degrading ATP to adenosine. At densities of 30,000-40,000 cells/ml, human fat cell suspensions accumulated adenosine, inosine, and hypoxanthine, and their concentrations were 38 +/- 8, 120 +/- 10, and 31 +/- 7 nmol/liter after 3 h of incubation. Dipyridamole (10 mumol/liter), an inhibitor of nucleoside transport, caused a 5-7-fold increase in adenosine accumulation which was reduced by 85% on inhibition of ectophosphatases by beta-glycerophosphate and antibodies against ecto-5'-nucleotidase or alpha, beta-methylene 5'-adenosine diphosphate (10 mumol/liter), respectively, indicating that most of the adenosine is produced in the extracellular compartment. Accordingly, the spontaneous accumulation of adenosine was reduced beyond 5 nmol/liter on inhibition of ectophosphatase activities or removal of extracellular AMP by AMP deaminase (4 units/ml). Added adenosine (30 nmol/liter) disappeared until its concentration approached 5 nmol/liter. Isoproterenol (1 mumol/liter) had no effect on adenosine accumulation regardless whether purine production from extracellular sources was minimized or not. In contrast to adenosine, the concentrations of inosine and hypoxanthine displayed only a modest decrease (30-50%) on inhibition of ectophosphatase activities. In addition, isoproterenol caused a 2-3-fold increase in inosine and hypoxanthine production which was concentration-dependent and could be inhibited by propranolol. It is concluded that the adenosine that accumulates in human adipocyte suspensions is almost exclusively derived from adenine nucleotides which are released by leaking cells. By contrast, inosine and hypoxanthine are produced inside the cells, and the release of these latter purines appears to be linked to ATP turnover via adenylate cyclase.  相似文献   

8.
The release of radioactive metabolites from isolated rat superior cervical ganglia was measured under various conditions following preloading with 3H-adenosine. The 3H label was recovered primarily in the adenosine metabolites, ATP, ADP, AMP, IMP, and inosine, rather than in adenosine itself. Increased release was evoked by preganglionic stimulation or by exposure to a high-K+ medium, whereas in a low-Ca2+-high-Mg2+ medium, both spontaneous release and evoked release of most metabolites were inhibited. Exposure of the ganglion to an atmosphere of N2 also increased the release of most labeled metabolites, but this release was not substantially affected by a low-Ca2+ medium. The fluorescent derivatives of the endogenous adenine-containing compounds present in the ganglion were prepared from homogenates and separated by high-performance liquid chromatography (HPLC). By the end of the testing period (6 hr), levels of ATP in the isolated ganglia had dropped to 10-20% of the initial values, while levels of ADP, AMP, and adenosine increased. There was little difference in these values between nonstimulated ganglia and those exposed to N2 or to a high-K+ medium.  相似文献   

9.
Adenosine is well known to be released during cerebral metabolic stress and is believed to be neuroprotective. ATP release under similar circumstances has been much less studied. We have now used biosensors to measure and compare in real time the release of ATP and adenosine during in vitro ischaemia in hippocampal slices. ATP release only occurred following the anoxic depolarisation, whereas adenosine release was apparent almost immediately after the onset of ischaemia. ATP release required extracellular Ca2+. By contrast adenosine release was enhanced by removal of extracellular Ca2+, whilst TTX had no effect on either ATP release or adenosine release. Blockade of ionotropic glutamate receptors substantially enhanced ATP release, but had only a modest effect on adenosine release. Carbenoxolone, an inhibitor of gap junction hemichannels, also greatly enhanced ischaemic ATP release, but had little effect on adenosine release. The ecto-ATPase inhibitor ARL 67156, whilst modestly enhancing the ATP signal detected during ischaemia, had no effect on adenosine release. Adenosine release during ischaemia was reduced by pretreatment with homosysteine thiolactone suggesting an intracellular origin. Adenosine transport inhibitors did not inhibit adenosine release, but instead they caused a twofold increase of release. Our data suggest that ATP and adenosine release during ischaemia are for the most part independent processes with distinct underlying mechanisms. These two purines will consequently confer temporally distinct influences on neuronal and glial function in the ischaemic brain.  相似文献   

10.
The contribution of neuronal ATP to interstitial adenosine levels was investigated in isolated perfused rat hearts. Ventricular surface transudates, representing interstitial fluid, were analyzed for norepinephrine, ATP, and adenosine. Exocytotic release of norepinephrine was induced by electrical stimulation of cardiac efferents emanating from the stellate ganglion. Ganglion stimulation increased contractility, interstitial norepinephrine, ATP, and adenosine. Interstitial adenosine was 11- to 27-fold higher than interstitial ATP, suggesting that the released ATP is unlikely the only source of adenosine. In the presence of AOPCP (alpha,beta-methyleneadenosine 5'-diphosphate), an ecto-5'-nucleotidase inhibitor, the ganglion-stimulated increase in interstitial ATP and adenosine reached levels similar to those in the absence of AOPCP, also suggesting that adenosine does not derive from extracellular ATP. The perfusate Ca2+ was raised from 1 to 4 mM to determine the importance of the enhanced contractile function on the levels of norepinephrine, ATP, and adenosine. The results were increases in contractility and interstitial norepinephrine, ATP, and adenosine, which were not suppressed with atenolol, indicating a norepinephrine-independent release of ATP and adenosine. Reserpine treatment and administration of guanethidine depleted the catecholamine stores and diminished the catecholamine release, respectively. However, neither agent altered Ca2+-induced increases in ATP and adenosine. It is concluded that the amount of neuronal-derived ATP is low and most likely does not contribute significantly to interstitial levels of adenosine. Furthermore, elevations in interstitial norepinephrine, ATP, and adenosine are associated with neuronal-independent increases in contractile function.  相似文献   

11.
The presence of 5 or 20 microM choline in the eserinized medium superfusing striatal slices enhanced the spontaneous release of acetylcholine (ACh) at both concentrations and, at 20 microM, the release of transmitter evoked by electrical field stimulation. Neither the electrical stimulation nor the addition of choline altered choline acetyltransferase activity. These results show that ACh release is dependent on the availability of extracellular choline. The rate of choline efflux was 7 times higher than the rate of ACh release, was not affected by stimulation, and was increased by 40% when hemicholinium-3 (HC-3), an inhibition of choline uptake, was present. The muscarinic antagonist atropine (1 microM) increased the evoked release of ACh into both the choline-free medium and that containing 20 microM choline. An adenosine receptor antagonist, 1,3-diethyl-8-phenyl xanthine (10 microM), failed to affect ACh release or the enhancement of release produced by atropine. In medium containing HC-3, stimulation of the slices elicited ACh release for the first 20 min of the 30 min stimulation period (15 Hz); thereafter, although stimulation was continued, the rate of release decreased to that associated with spontaneous release. Tissue ACh contents were not modified by the addition of choline or atropine to the medium, but were depressed by HC-3. Neither atropine nor HC-3 altered tissue choline content. The total amount of ACh + choline released during an experiment was 5-15 times higher than the decrease in tissue levels of these two compounds during the same period of time.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Abstract: The release of adenosine and ATP evoked by electrical field stimulation in rat hippocampal slices was investigated with the following two patterns of stimulation: (1) a brief, high-frequency burst stimulation (trains of stimuli at 100 Hz for 50 ms applied every 2 s for 1 min), to mimic a long-term potentiation (LTP) stimulation paradigm, and (2) a more prolonged (3 min) and low-frequency (5 Hz) train stimulation, to mimic a long-term depression (LTD) stimulation paradigm. The release of ATP was greater at a brief, high-frequency burst stimulation, whereas the release of [3H]adenosine was slightly greater at a more prolonged and low-frequency stimulation. To investigate the source of extracellular adenosine, the following two pharmacological tools were used; α,β-methylene ADP (AOPCP), an inhibitor of ecto-5′-nucleotidase, to assess the contribution of the catabolism of released adenine nucleotides as a source of extracellular adenosine, and S-(4-nitrobenzyl)-6-thioinosine (NBTI), an inhibitor of adenosine transporters, to assess the contribution of the release of adenosine, as such, as a source of extracellular adenosine. At low-frequency stimulation, NBTI inhibited by nearly 50% the evoked outflow of [3H]adenosine, whereas AOPCP inhibited [3H]adenosine outflow only marginally. In contrast, at high-frequency stimulation, AOPCP inhibited by 30% the evoked release of [3H]adenosine, whereas NBTI produced a 40% inhibition of [3H]adenosine outflow. At both frequencies, the kinetics of evoked [3H]adenosine outflow was affected in different manners by AOPCP and NBTI; NBTI mainly depressed the rate of evoked [3H]adenosine outflow, whereas AOPCP mainly inhibited the later phase of evoked [3H]adenosine accumulation. These results show that there is a simultaneous, but quantitatively different, release of ATP and adenosine from rat hippocampal slices stimulated at frequencies that can induce plasticity phenomena such as LTP (100 Hz) or LTD (5 Hz). The source of extracellular adenosine is also different according to the frequency of stimulation; i.e., at a brief, high-frequency stimulation there is a greater contribution of released adenine nucleotides for the formation of extracellular adenosine than at a low frequency with a more prolonged stimulation.  相似文献   

13.
We studied the effect of adenosine nucleotides on several aspects of the functional activation of human peripheral blood polymorphonuclear leukocytes (PMN). Radiolabeled ATP bound to PMN in a manner suggesting the existence of specific binding sites because: 1) binding was reversed (92 +/- 6%) by 100-fold excess concentrations of unlabeled ATP but minimally by either ADP (43 +/- 12%) or GTP (37 +/- 8%); and 2) binding saturation was achieved (i.e., specific binding did not increase) above 250 microM ATP. Binding studies revealed that significant ATP hydrolysis occurred, even at low temperatures and in the presence of phosphatase inhibitors. Adenosine nucleotides activated signal transduction mechanisms in PMN because: 1) 1 to 100 microM ATP and 5'-adenylylimidodiphosphate (AMP-PNP) stimulated increased production of 1,2-diacylglycerols; 2) ATP (0.5 to 500 microM) and ADP (0.1 to 10 mM) induced increased insoluble protein kinase (PKC) activity in a dose-dependent manner when used at concentrations greater than 50 microM; 3) ATP (greater than or equal to 50 microM) induced a shift in the solubility of phorbol receptors from mostly soluble (89% in untreated cells) to mostly insoluble (68%), whereas ADP, GTP, and GDP were effective at higher concentrations; and 4) greater than or equal to 50 microM ATP stimulated increased phosphorylation of endogenous PMN proteins. AMP-PNP induced PKC activity and phosphoprotein changes that were qualitatively similar to those observed when PMN were treated with ATP, suggesting that extracellular ATP hydrolysis is not required for signal transduction to activate PKC. Functionally, ATP stimulated the secretion of specific (but not azurophil) granules because vitamin B12-binding protein and low levels of lysozyme, but not beta-glucuronidase, were released; qualitatively similar results were obtained by using AMP-PNP. These results suggest that certain adenosine nucleotides employed at physiologically relevant concentrations stimulate increased 1,2-diacylglycerol production, PKC activity, granule secretion, and endogenous phosphoprotein formation in a manner that is independent of extracellular ATP hydrolysis.  相似文献   

14.
The relative roles of mitochondrial (mito) ATP-sensitive K(+) (mitoK(ATP)) channels, protein kinase C (PKC), and adenosine kinase (AK) in adenosine-mediated protection were assessed in Langendorff-perfused mouse hearts subjected to 20-min ischemia and 45-min reperfusion. Control hearts recovered 72 +/- 3 mmHg of ventricular pressure (50% preischemia) and released 23 +/- 2 IU/g lactate dehydrogenase (LDH). Adenosine (50 microM) during ischemia-reperfusion improved recovery (149 +/- 8 mmHg) and reduced LDH efflux (5 +/- 1 IU/g). Treatment during ischemia alone was less effective. Treatment with 50 microM diazoxide (mitoK(ATP) opener) during ischemia and reperfusion enhanced recovery and was equally effective during ischemia alone. A(3) agonism [100 nM 2-chloro-N(6)-(3-iodobenzyl)-adenosine-5'-N-methyluronamide], A(1) agonism (N(6)-cyclohexyladenosine), and AK inhibition (10 microM iodotubercidin) all reduced necrosis to the same extent as adenosine, but less effectively reduced contractile dysfunction. These responses were abolished by 100 microM 5-hydroxydecanoate (5-HD, mitoK(ATP) channel blocker) or 3 microM chelerythrine (PKC inhibitor). However, the protective effects of adenosine during ischemia-reperfusion were resistant to 5-HD and chelerythrine and only abolished when inhibitors were coinfused with iodotubercidin. Data indicate adenosine-mediated protection via A(1)/A(3) adenosine receptors is mitoK(ATP) channel and PKC dependent, with evidence for a downstream location of PKC. Adenosine provides additional and substantial protection via phosphorylation to 5'-AMP, primarily during reperfusion.  相似文献   

15.
We have previously reported that adenosine formed in response to reduced arteriolar and/or tissue PO(2) preserves endothelial nitric oxide (NO) synthesis during sympathetic vasoconstriction in the rat intestine. To more precisely identify the site and mechanism of adenosine formation under these conditions, we tested the hypothesis that ATP released in response to reduced O(2) levels serves as a source of adenosine. Direct application of ATP to the wall of first-order arterioles elicited dose-dependent dilations of 15-33% above resting diameter that were reduced by 71-80% by the 5'-ectonucleotidase inhibitor alpha,beta-methyleneadenosine 5'-diphosphate (AOPCP, 4.5 x 10(-5) M) and completely abolished by N(G)-monomethyl-L-arginine (L-NMMA, 10(-4) M). Under control conditions, sympathetic nerve stimulation at 3 and 8 Hz induced arteriolar constrictions of 11 +/- 1 and 19 +/- 1 microm, respectively. These responses were enhanced by 58-69% in the presence of L-NMMA or when local PO(2) was maintained at resting levels. However, in the presence of AOPCP, the enhancing effects of L-NMMA and the high O(2) superfusate on sympathetic constriction were preserved. These results suggest that, although exogenously applied ATP can stimulate arteriolar NO release in the intestine largely through its sequential extracellular hydrolysis to adenosine, this process does not contribute to adenosine formation and sustained NO release during sympathetic constriction in this vascular bed.  相似文献   

16.
Calcium-dependent release of purines was previously demonstrated in cultures of chick retinal cells stimulated with high potassium concentrations but there is no evidence for an exocytotic mechanism of adenosine release from presynaptic terminals. Here we show that activation of NMDA or alpha-amino-3-hydroxy-5-methylisoxazole-4-propionate (AMPA)/kainate glutamate ionotropic receptors promotes a two- to three-fold increase in the release of purines from these cultures. Approximately 96% of intracellular radioactivity is found as nucleotides after incubation with [(3)H]adenosine, but more than 85% of glutamate-stimulated released material is found as inosine (60%), hypoxanthine (19.9%) and adenosine (7.8%). The release is prevented by removal of extracellular calcium, by the transporter blocker nitrobenzylthioinosine, or inhibitors of calcium/calmodulin-dependent protein kinase II (CAMK II). The uptake of [(3)H]adenosine, but not of [(3)H]GABA or [(3)H]choline, is also blocked by 1-[N,O-bis(5-isoquinolinesulfonyl)-N-methyl-l-tyrosyl]-4-phenylpiperazine (KN62), N-[2-(N-(4-chlorocinnamyl)-N-methylaminomethyl)phenyl-N-[2-hydroxiethyl]-4-methoxybenzenesulfonamide (KN93) or the myristoylated autocamtide-2-related inhibitory peptide, suggesting that the enzyme modulates the nucleoside transporter. The distribution of intracellular purines was not affected by KN62. These results indicate that activation of glutamate receptors triggers the release of purines from retinal cells by a mechanism involving calcium influx, CAMK II and the nitrobenzylthioinosine-sensitive nucleoside transporter. The regulation of adenosine release by glutamate receptors and CAMK II could have important consequences in the presynaptic control of glutamate release.  相似文献   

17.
The metabolic pathway of inositol phospholipids represents a series of synthetic and hydrolytic reactions with inositol as a by-product. Hence, the rate of [3H]inositol release from prelabeled phospholipids can be used as a reflection of activity of this pathway. In the frog sympathetic ganglion prelabeled with [3H]inositol, we studied the effect of synaptic activity (orthodromic stimulation) on release of 3H-label into the medium. This release was interpreted as [3H]inositol release. The value was low at rest and increased significantly by 32% during orthodromic stimulation (20 Hz for 5 min). However, on cessation of the stimulation, [3H]inositol release increased rapidly by 148% and remained elevated for at least 45 min. This increase in [3H]inositol release during and after the stimulation period was reduced by suffusion of the ganglia with adenosine. We hypothesized that synaptic activation releases a long-lasting stimulatory agonist and a short-lasting inhibitory (adenosine) agonist or agonists affecting [3H]inositol release. To demonstrate the presence of a stimulatory agonist, two sympathetic ganglia were used. One was prelabeled with [3H]inositol, and the other was not. The two ganglia were placed together in a 5-microliter droplet of Ringer's solution containing atropine. Orthodromic stimuli applied to the nonlabeled ganglion elicited release of [3H]inositol from the nonstimulated ganglion. To test whether the adenosine formed during orthodromic stimulation inhibits [3H]inositol release, we destroyed endogenous adenosine by suffusion of the ganglia with adenosine deaminase during the stimulation period. We found that adenosine deaminase induced large increases in [3H]inositol release during the stimulation period, in contrast to an increase seen only during the poststimulation period when adenosine deaminase was omitted. Because [3H]inositol release is assumed to parallel changes in content of inositol phosphates, we anticipated no changes of the levels of these compounds during orthodromic stimulation. However, measurements showed that levels of inositol phosphates and inositol phospholipids were all elevated except for phosphatidylinositol 4-phosphate. On termination of the stimulus, they remained elevated, with a further increase in levels of inositol trisphosphate and phosphatidylinositol 4-phosphate. We conclude that endogenous adenosine inhibits [3H]inositol release, possibly by modulating several of the steps of the inositol phospholipid pathway.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
ATP is released into extracellular space as an autocrine/paracrine molecule by mechanical stress and pharmacological-receptor activation. Released ATP is partly metabolized by ectoenzymes to adenosine. In the present study, we found that adenosine causes ATP release in Madin-Darby canine kidney cells. This release was completely inhibited by CPT (an A1 receptor antagonist), U-73122 (a phospholipase C inhibitor), 2-APB (an inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) receptor blocker), thapsigargin (a Ca2+-ATPase inhibitor), and BAPTA/AM (an intracellular Ca2+ chelator), but not by DMPX (an A2 receptor antagonist). However, forskolin, epinephrine, and isoproterenol, inducers of cAMP accumulation, failed to release ATP. Adenosine increased intracellular Ca2+ concentrations that were strongly blocked by CPT, U-73122, 2-APB, and thapsigargin. Moreover, adenosine enhanced accumulations of Ins(1,4,5)P3 that were significantly reduced by U-73122 and CPT. These data suggest that adenosine induces the release of ATP by activating an Ins(1,4,5)P3 sensitive-Ca2+ pathway through the stimulation of A1 receptors.  相似文献   

19.
The aim of the study was to elucidate the role of nucleoside transport systems in the postischemic release of nucleosides and nucleobases accumulated by the rat liver during cold storage. Livers were preserved for 24 h in Euro-Collins (EC) or in a lactobionate-based solution (LBS) without exogenous adenosine. The rates of release of uric acid, xanthine, hypoxanthine, inosine, adenosine, uridine, and cytidine were monitored during early reperfusion. The greater part of the purines and pyrimidines (up to 80%) was lost in the first 2 min of reperfusion. After storage in EC, uric acid and xanthine formed more than 90% of the total purines released; nucleosides did not exceed 5% of the total. After storage in LBS, hypoxanthine formed more than 80% of purine efflux and the release of inosine and uridine was increased 5-10 times. These changes were shown to be due to the presence of allopurinol in LBS. Dipyridamole (an inhibitor of equilibrative nucleoside transporters) decreased the efflux of uric acid after storage in EC but residual release remained high. Dipyridamole exerted the most pronounced effect on the release of nucleosides (inosine and uridine) from livers stored in LBS. The use of sodium-free media for liver preservation and reperfusion did not alter the rates of purine and pyrimidine release. We conclude that equilibrative nucleoside transporters mediate the postischemic release of nucleosides and also, but to a less degree, of uric acid. Simple diffusion is an important factor in the release of nucleobases. Active Na(+)/nucleoside cotransport does not play an important role in early reperfusion.  相似文献   

20.
Activation of glycolysis by insulin in cultured rat hepatocytes is preceded by an activation of phosphofructokinase 2 (PFK 2) and subsequent rise of the fructose 2,6-bisphosphate [Fru(2,6)P2] level. Extracellular addition of ATP or puromycin prevented the hormonal effect on glycolysis. The mechanism through which the purines abolished glycolytic stimulation was investigated. 1. 50 microM ATP completely prevented the 3-5-fold insulin-dependent increase of glycolysis, irrespective of whether the cells initially possessed a low or a high Fru(2,6)P2 content. 50 microM puromycin prevented the stimulation of glycolysis by insulin only in cells whose initial Fru(2,6)P2 levels were low and had to be increased by insulin prior to the increase in glycolysis. It did not antagonize the action of insulin cells with initial high Fru(2,6)P2 content. 2. ATP exerted effects on its own; it decreased initially high Fru(2,6)P2 levels by 95% within 10 min and decreased the basal glycolytic rate by 60%. Half-maximal effects on the Fru(2,6)P2 level were obtained with about 25 microM ATP or 15 microM adenosine 5'[beta, gamma-methylene]triphosphate. ADP and adenosine-5-[gamma-thio]triphosphate were as effective as ATP, whereas 100 microM adenosine 5'[alpha, beta-methylene]triphosphate elicited no effect. Puromycin neither decreased high Fru(2,6)P2 levels nor inhibited basal glycolysis. 3. Extracellular ATP (100 microM) led to inhibition of the active form of PFK 2. Intracellular levels of Glc6P, citrate, ATP, ADP and AMP were increased by extracellular ATP, the phosphoenolpyruvate content was decreased, Fru6P and glycerol 3-phosphate levels stayed constant. Puromycin did not inhibit PFK 2. 4. Both puromycin and ATP prevented the insulin-dependent rise of the Fru(2,6)P2 level, they abolished the activation of PFK 2 by the hormone. Puromycin did not block the accumulation of Fru(2,6)P2 provoked by glucose addition; ATP also antagonized the glucose-dependent increase. 5. 100 microM ATP elevated the cAMP-dependent protein kinase activity ratio from 0.1 to 0.38 and increased the level of inositol trisphosphate by 16-fold within 5 min, whereas puromycin was without effect on either level. It is concluded that the two purines block the insulin effect on glycolysis by preventing the hormone increasing the Fru(2,6)P2 level. The mode of action, however, seems to be different: ATP antagonizes insulin action in that it leads to increased inhibition of PFK 2 whereas puromycin prevents the activation of PFK 2 by insulin.  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