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.
Chlorisondamine (CHL) blocks behavioural responses to nicotine for several weeks or months in rats. Persistent blockade has also been demonstrated ex vivo, in assays of nicotine-evoked striatal dopamine release. Central administration of [3H]-CHL leads to long-term retention of radiolabel in nigrostriatal dopaminergic neurons and in few other cell groups. We investigated whether an analogous blockade also occurs in noradrenergic neurons in the brain and in cultured pheochromocytoma (PC12) cells, which have a similar noradrenergic phenotype. Administration of CHL (10 mg kg(-1) s.c. or 10 microg i.c.v.), 21 days prior, resulted in a near-total block of nicotine-evoked release of hippocampal [3H]-noradrenaline ([3H]-NA) from superfused rat synaptosomes; NMDA-evoked [3H]-NA release was unaffected. Three weeks after administration of [3H]-CHL (10 microg i.c.v.), preferential accumulation of radiolabel was observed in the locus coeruleus, which provides the entire noradrenergic innervation to hippocampus, as well as in previously noted structures. In rat pheochromocytoma (PC12) cells, nicotine evoked [3H]-NA release (EC50 approximately 30 microM). This effect was blocked by co-incubation with mecamylamine (10 microM) or CHL (1 microM) but was not affected by alpha-bungarotoxin. As in the hippocampus, the nicotinic agonist cytisine was at least as efficacious as nicotine. Acute exposure of PC12 cells to CHL 10 or 100 microM (but not 1 microM), followed by 90 min wash-out, almost completely blocked release evoked by 30 microM nicotine. More prolonged (24 h) exposure to CHL 100 microM (but not 1 or 10 microM), followed by 3 days of wash-out, partially inhibited release evoked by nicotine, leaving responses to high K+ unchanged. A significant (30%) reduction was also seen 5 days after exposure. We conclude that persistent nicotinic blockade by CHL is neither restricted to mesostriatal dopamine neurons, nor to the CNS, nor to neurons possessing the same nicotinic receptor pharmacology. In addition, the persistent blockade does not appear to result from an acute blocking action, but may be dependent upon intracellular accumulation of the antagonist.  相似文献   

2.
1. Isoarecolone was approximately 250 times less potent than nicotine as an inhibitor of [3H]-nicotine binding to rat brain membranes. Isoarecolone failed to inhibit the binding of the nicotinic ligand [125I]-alpha-bungarotoxin or of the muscarinic ligand [3H]-QNB. 2. Nicotine (0.01-30 microM) evoked the release of [3H]-dopamine from striatal and frontal cortex synaptosomes, with EC50 values of approximately 0.5 microM in each case. This release was largely mecamylamine-sensitive. 3. Isoarecolone (1-200 microM) evoked predominantly mecamylamine-sensitive dopamine release from both striatal and cortical synaptosomes, with a potency at least 20 times less than that of nicotine. The maximum effect of isoarecolone was less than that of nicotine, particularly in the frontal cortex preparation. 4. In control rats treated chronically with saline, neither nicotine nor isoarecolone had clear effects on locomotor activity at the doses tested. Chronic treatment with nicotine clearly sensitized rats to the locomotor activating effect of isoarecolone was seen at a dose about 40 times larger than that of nicotine. 5. The low potency and efficacy of isoarecolone in facilitating sensitized locomotor activity resembled its lower potency and efficacy, compared with nicotine, in evoking dopamine release in vitro. The agonist profile of the nicotinic receptor population mediating dopamine release may determine the pharmacological characteristics of consequent locomotor behaviour.  相似文献   

3.
The modulation of [3H]ACh release by nicotinic compounds was studied in superfused rat hippocampal synaptosomes loaded with [3H]choline, (-)-Nicotine (0.1-10 microM) evoked a dose-dependent increase in [3H]ACh release; higher concentrations were less effective. Nicotine-evoked release was Ca(2+)-dependent, and blocked by the nicotinic antagonists dihydro-beta-erythroidine, mecamylamine, and pempidine. The alpha 7-selective antagonist methyllycaconitine did not inhibit nicotine-evoked release when tested at 1 microM, although at 10 microM some attenuation of the response was observed. Six agonists tested were equally efficacious in stimulating [3H]ACh release, as judged by the maximum responses, and gave the following EC50 values: (+/-)-epibatidine 0.12 microM; (+)-anatoxin-a 0.14 microM; (-)-nicotine 0.99 microM; (-)-cytisine 1.06 microM; ABT-418 2.6 microM; isoarecolone 43 microM. Each agonist generated a "bell-shaped" dose response curve, suggesting desensitisation at higher concentrations. This is supported by analysis of repetitive stimulation with (-)-nicotine and (-)-cytisine: S2/S1 ratios declined sharply with increasing concentration, whereas subsequent KC1-evoked release remained constant. These results are discussed in terms of possible nicotinic receptor subtypes that might be present on hippocampal nerve terminals.  相似文献   

4.
To characterize the properties of nicotinic acetylcholine receptors (nAChRs) in autonomic ganglia, we examined specific L-[3H]nicotine binding to membrane fraction prepared from cultured bovine adrenal chromaffin cells, using a modified filtration method. Binding of L-[3H]nicotine to non-treated glass fiber filters interfered with detection of specific binding to the membrane fraction. Presoaking the glass fiber filters in 3% or greater concentrations of polyethyleneimine solution (sixty times higher than the reported concentration) for more than 5 h could reduce the binding of L-[3H]nicotine to the filters to background level. Furthermore, specific L-[3H]nicotine binding to the membrane fraction was detected only when the membrane fraction was prepared in buffer containing no divalent cations such as Ca2+ and Mg2+ (EDTA and EGTA were added) and protease inhibitors. Specific binding of L-[3H]nicotine was saturable and reversible. Scatchard analysis revealed a single class of high affinity binding sites with an average Kd of 8.7 nM and a Bmax of 43.2 fmol/mg protein. Specific binding was sensitive to cholinergic agonists (carbamylcholine and L-nicotine) and ganglionic stimulating agents (lobeline and DMPP), but was resistant to neuromuscular blocking agents (alpha-bungarotoxin, d-tubocurarine) and hexamethonium. These results suggest that high-affinity nicotine binding sites on adrenal chromaffin cells are nAChRs of ganglion-type, which have properties different from nAChRs on the neuromuscular junction but similar to nAChRs in the brain.  相似文献   

5.
To provide anatomical information on the complex effects of acetylcholine (ACh) in the olfactory bulb (OB), the distribution of different cholinergic muscarinic and nicotinic receptor sub-types was studied by quantitative in vitro autoradiography. The muscarinic M1-like and M2-like sub-types, as well as the nicotinic bungarotoxin-insensitive (alpha 4 beta 2-like) and bungarotoxin-sensitive (alpha 7-like) receptors were visualized using [3H]pirenzepine, [3H]AF-DX 384, [3H]cytisine and [125I] alpha-bungarotoxin (BTX), respectively. In parallel, labelling patterns of [3H]vesamicol (vesicular acetylcholine transport sites) and [3H]hemicholinium-3 (high-affinity choline uptake sites), two putative markers of cholinergic nerve terminals, were investigated. Specific labelling for each cholinergic radioligand is distributed according to a characteristic laminar and regional pattern within the OB revealing the lack of a clear overlap between cholinergic afferents and receptors. The presynaptic markers, [3H]vesamicol and [3H]hemicholinium-3, demonstrated similar laminar pattern of distribution with two strongly labelled bands corresponding to the glomerular layer and the area around the mitral cell layer. Muscarinic M1-like and M2-like receptor sub-types exhibited unique distribution with their highest levels seen in the external plexiform layer (EPL). Intermediate M1-like and M2-like binding densities were found throughout the deeper bulbar layers. In the glomerular layer, the levels of muscarinic receptor subtypes were low, the level of M2-like sites being higher than M1. Both types of nicotinic receptor sub-types displayed distinct distribution pattern. Whereas [125I] alpha-BTX binding sites were mostly concentrated in the superficial bulbar layers, [3H]cytisine binding was found in the glomerular layers, as well as the mitral cell layer and the underlying laminae. An interesting feature of the present study is the visualization of two distinct cholinoceptive glomerular subsets in the posterior OB. The first one exhibited high levels of both [3H]vesamicol and [3H]hemicholinium-3 sites. It corresponds to the previously identified atypical glomeruli and apparently failed to express any of the cholinergic receptors under study. In contrast, the second subset of glomeruli is not enriched with cholinergic nerve terminal markers but displayed high amounts of [3H]cytisine/nicotinic binding sites. Taken together, these results suggest that although muscarinic receptors have been hypothesized to be mostly involved in cholinergic olfactory processing and short-term memory in the OB, nicotinic receptors, especially of the cytisine/ alpha 4 beta 2 sub-type, may have important roles in mediating olfactory transmission of efferent neurons as well as in a subset of olfactory glomeruli.  相似文献   

6.
The presynaptic nicotinic modulation of dopamine release from striatal nerve terminals is well established, but the subtype(s) of neuronal nicotinic acetylcholine receptor (nAChR) underlying this response has not been identified. Recently, alpha-conotoxin-MII has been reported to inhibit potently and selectively the rat alpha3beta2 combination of nAChR subunits. Here we have synthesised the peptide, confirmed its specificity, and examined its effect on the (+/-)-anatoxin-a-evoked release of [3H]dopamine from rat striatal synaptosomes and slices. Alpha-conotoxin-MII (112 nM) completely blocked acetylcholine-evoked currents of alpha3beta2 nAChRs expressed in Xenopus oocytes (IC50 = 8.0 +/- 1.1 nM). Pairwise combinations of other nicotinic subunits were not blocked by 112 nM alpha-conotoxin-MII. On perfused striatal synaptosomes and slices, alpha-conotoxin-MII dose-dependently inhibited [3H]dopamine release evoked by 1 microM (+/-)-anatoxin-a with IC50 values of 24.3 +/- 2.9 and 17.3 +/- 0.1 nM, respectively. The dose-response curve was shifted to the right with increasing agonist concentrations. However, the maximal inhibition of responses achieved by alpha-conotoxin-MII (112 nM) was 44.9 +/- 5.4% for synaptosomes and 25.0 +/- 4.1% for slices, compared with an inhibition by 10 microM mecamylamine of 77.9 +/- 3.7 and 88.0 +/- 2.1%, respectively. These results suggest the presence of presynaptic alpha3beta2-like nAChRs on striatal dopaminergic terminals, but the incomplete block of (+/-)-anatoxin-a-evoked [3H]dopamine release by alpha-conotoxin-MII also supports the participation of nAChRs composed of other subunits. The lower inhibition found in slices is consistent with an additional indirect nicotinic stimulation of dopamine release via an alpha-conotoxin-MII-insensitive nAChR.  相似文献   

7.
Antagonists of alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropanoic acid (AMPA) receptors may have therapeutic potential as psychotropic agents. A series of mononitro- and dinitro-2- and 3-hydroxyphenylalanines was prepared, and their activity compared with willardiine, 5-nitrowillardiine, AMPA, and 2,4,5-trihydroxyphenylalanine (6-hydroxydopa) as inhibitors of specific [3H]AMPA and [3H]kainate binding in rat brain homogenates. The most active compounds were highly acidic (pKa 3-4), namely, 2-hydroxy-3,5-dinitro-DL-phenylalanine (13; [3H]AMPA IC50 approximately equal to 25 microM) and 3-hydroxy-2,4-dinitro-DL-phenylalanine (19; [3H]AMPA IC50 approximately equal to 5 microM). Two other dinitro-3-hydroxyphenylalanines, and 3,5-dinitro-DL-tyrosine, were considerably less active. Various mononitrohydroxyphenylalanines, which are less acidic, were also less active or inactive, and 2- and 3-hydroxyphenylalanine (o- and m-tyrosine) were inactive. Compounds 13 and 19, DL-willardiine (pKa 9.3, [3H]AMPA IC50 = 2 microM), and 5-nitro-DL-willardiine (pKa 6.4, [3H]AMPA IC50 = 0.2 microM) displayed AMPA > kainate selectivity in binding studies. Compound 19 was an AMPA-like agonist, but 13 was an antagonist in an AMPA-evoked norepinephrine release assay in rat hippocampal nerve endings. Also, compound 13 injected into the rat ventral pallidum antagonized the locomotor activity elicited by systemic amphetamine.  相似文献   

8.
1. We have demonstrated recently that exogenous prostaglandin E2 (PGE2) inhibits electrical field stimulation (EFS)-induced acetylcholine (ACh) release from parasympathetic nerve terminals innervating guinea-pig trachea. In the present study, we have attempted to characterize the pre-junctional prostanoid receptor(s) responsible for the inhibitory action of PGE2 and to assess whether other prostanoids modulate, at a prejunctional level, cholinergic neurotransmission in guinea-pig trachea. To this end, we have investigated the effect of a range of both natural and synthetic prostanoid agonists and antagonists on EFS-evoked [3H]-ACh release. 2. In epithelium-denuded tracheal strips pretreated with indomethacin (10 microM), PGE2 (0.1 nM-1 microM) inhibited EFS-evoked [3H]-ACh release in a concentration-dependent manner with an EC50 and maximal effect of 7.62 nM and 74% inhibition, respectively. Cicaprost, an IP-receptor agonist, PGF2alpha and the stable thromboxane mimetic, U46619 (each at 1 microM), also inhibited [3H]-ACh release by 48%, 41% and 35%, respectively. PGD2 (1 microM) had no significant effect on [3H]-ACh release. 3. The selective TP-receptor antagonist, ICI 192,605 (0.1 microM), completely reversed the inhibition of cholinergic neurotransmission induced by U-46619, but had no significant effect on similar responses effected by PGE2 and PGF2alpha. 4. A number of EP-receptor agonists mimicked the ability of PGE2 to inhibit [3H]-ACh release with a rank order of potency: GR63799X (EP3-selective) > PGE2 > M&B 28,767 (EP3 selective) > 17-phenyl-omega-trinor PGE2 (EP1-selective). The EP2-selective agonist, AH 13205 (1 microM), did not affect EFS-induced [3H]-ACh release. 5. AH6809 (10 microM), at a concentration 10 to 100 times greater than its pA2 at DP-, EP1- and EP2-receptors, failed to reverse the inhibitory effect of PGE2 or 17-phenyl-omega-trinor PGE2 on [3H]-ACh release. 6. These results suggest that PGE2 inhibits [3H]-ACh release from parasympathetic nerves supplying guinea-pig trachea via an interaction with prejunctional prostanoid receptors of the EP3-receptor subtype. Evidence for inhibitory prejunctional TP- and, possibly, IP-receptors was also obtained although these receptors may play only a minor role in suppressing [3H]-ACh release when compared to receptors of the EP3-subtype. However, the relative importance of the different receptors will depend not only on the sensitivity of guinea-pig trachea to prostanoids but on the nature of the endogenous ligands released locally that have activity on parasympathetic nerves.  相似文献   

9.
Ethanol disrupts signal transduction mediated by a variety of G-protein coupled receptors. We examined the effects of ethanol on arachidonic acid release mediated by muscarinic acetylcholine receptors. Chinese hamster ovary (CHO) cells transfected with the different subtypes of human muscarinic receptors (M1 to M5) were incubated with [3H]arachidonic acid ([3H]AA) for 18 hr, washed, and exposed to the cholinergic agonist carbamylcholine for 15 min. Carbamylcholine induced [3H]AA release from CHO cells expressing M1, M3, or M5, but not M2 or M4, muscarinic receptors. Dose response curves revealed that carbamylcholine stimulated [3H]AA release by up to 12-fold with an ECo of approximately 0.4 microM; maximal responses were obtained with 10 microM carbamylcholine. Exposure of M1-, M3-, or M5-expressing cells to ethanol for 5 min before stimulating with carbamylcholine reduced [3H]AA release by 40 to 65%; 50% of the maximal inhibition was obtained with an ethanol concentration of 30 to 50 mM. Ethanol did not affect basal [3H]AA release measured in the absence of carbamylcholine. Dose response curves suggest that ethanol acts as a noncompetitive inhibitor of muscarinic receptor-induced [3H]AA release insofar as maximal [3H]AA release was depressed in the presence of ethanol with no apparent change in the EC50 for stimulation by carbamylcholine. Exposure of CHO cells to 38 mM ethanol for 48 hr increased [3H]AA release induced by carbamylcholine without affecting basal [3H]AA release or altering the EC50 for carbamylcholine. These results indicate that ethanol acutely inhibits muscarinic receptor signaling through the arachidonic acid pathway in a noncompetitive manner, but chronically enhances muscarinic signaling through the same pathway.  相似文献   

10.
The effects of SR140333 and SR48968 (neurokinin1 and neurokinin2 tachykinin receptor antagonists, respectively) on the N-methyl-D-aspartate-evoked release of [3H]acetylcholine (previously formed from [3H]choline) were investigated in striosome-enriched areas and in the matrix of the rat striatum using an in vitro microsuperfusion method. In both striatal compartments, SR140333 and SR48968 did not modify the 50 microM N-methyl-D-aspartate-evoked release of [3H]acetylcholine. However, in low concentrations, both SR140333 (0.1 microM to 1 pM) and SR48968 (0.1 microM to 0.1 nM) markedly enhanced the 1 mM N-methyl-D-aspartate (+10 microM D-serine)-evoked release of [3H]acetylcholine in striosome-enriched areas. These responses were dopamine-dependent since they were not observed any more following the local blockade of D2 receptors by sulpiride or of dopamine synthesis by alpha-methyl-p-tyrosine. A dopamine-dependent disinhibitory effect (of lower amplitude) on the 1 mM N-methyl-D-aspartate (+10 microM D-serine)-evoked release of [3H]acetylcholine was also induced by SR48968 (0.1 microM to 0.1 nM) (but not by SR140333) in the matrix. In addition, in the matrix, as shown only in the presence of alpha-methyl-p-tyrosine, both SR140333 and SR48968 reduced the 1 mM N-methyl-D-aspartate (+10 microM D-serine)-evoked response and these non-dopamine-mediated inhibitory effects only occurred at the highest tested concentration (0.1 microM) of the antagonists. Indicating the specificity of these responses, the effects of SR140333 were reproduced by RP67580, another neurokinin1 receptor antagonist and, as expected from previous binding studies, corresponding SR140333 and SR48968 enantiomers were without effect. These results suggest that under potent stimulation of N-methyl-D-aspartate receptors, endogenously released substance P and neurokinin A (or related tachykinins) regulate differently the N-methyl-D-aspartate-evoked release of [3H]acetylcholine in striosomes and in the matrix. The inhibitory effects of these tachykinins on the evoked release of [3H]acetylcholine are mediated by dopamine. On the contrary, their facilitatory responses are only observed in the matrix under blockade of dopamine transmission.  相似文献   

11.
We have examined the effects of riluzole, a neuroprotective drug which stabilizes voltage-dependent sodium channels in their inactivated state and inhibits the release of glutamate in-vivo and in-vitro, on the release of newly taken up [3H]dopamine induced by ouabain, a potent and selective inhibitor of Na+/K+-ATPase in mouse striatal slices in-vitro. Riluzole potently (IC50 (concentration resulting in 50% inhibition) = 0.9+/-0.3 microM) and dose-dependently antagonized ouabain-stimulated [3H]dopamine release, the effect being observed at low concentrations. Tetrodotoxin (1 microM) and nomifensine (10 microM) also abolished ouabain-induced [3H]dopamine release. Blockade of glutamate receptors with dizocilpine (1 microM) and 6-(1H-imidazol-1-yl)-7-nitro-2,3(1H,4H)-quinoxalinedione (YM-90K; 10 microM), alone or in combination, was without effect. Incubation of striatal slices with 50 microM La3+, which blocks voltage-dependent calcium channels, did not inhibit [3H]dopamine release induced by ouabain. The potent effects of riluzole observed in this model are probably related to its ability to block voltage-dependent sodium channels. The consequences of this activity are critically discussed in relation to the protective action of riluzole previously reported in various models of Parkinson's disease and other neurodegenerative disorders.  相似文献   

12.
We studied the effects of cotinine, the major metabolite of nicotine, on nicotine-induced increase in [3H]phorbol dibutyrate binding, activation of protein kinase C and [3H]noradrenaline release in primary cultured bovine adrenal chromaffin cells. Cotinine (1 mM, 15 min.) and nicotine (10 microM, 5 min.) increased the [3H]phorbol binding by 100% and 150%, respectively. Both a short-term (10 min.) and a long-term (24 hr) pretreatment with cotinine inhibited the effect of nicotine. A 24 hr pretreatment with cotinine (1 mM) also reduced the nicotine-induced increase in membrane-bound protein kinase C activity. Cotinine pretreatment (10 min.) dose-dependently inhibited the release of [3H]noradrenaline induced by nicotine and dimethylphenylpiperazinium. Cotinine pretreatment did not reduce the [3H]noradrenaline release induced by high extracellular potassium (56 mM) or veratrine (10 mg l-1). The results indicate that cotinine inhibits activation of protein kinase C and noradrenaline release induced by nicotinic agonists in primary cultures of bovine adrenal chromaffin cells. The results suggest that pre-existing cotinine could modify responses to acute exposure to nicotine in neural systems.  相似文献   

13.
Acute nicotine administration stimulated [3H]norepinephrine ([3H]NE) release from cultured fetal locus coeruleus (LC) cells. The effect was concentration dependent, with an EC50 of 0.9 microM, and was abolished by removal of calcium from, or addition of tetrodotoxin (500 nM) to, the assay buffer. Other nicotinic receptor agonists stimulated [3H]NE release, with the rank order of potency being (+)-epibatidine > (-)-nicotine > 1,1-dimethyl-4-phenylpiperazinium (DMPP). Whereas (-)-nicotine and (+/-)-epibatidine exhibited equal maximal responses, DMPP was a partial agonist and (-)-cytisine had no agonist activity. Nicotine-stimulated release of [3H]NE was blocked by nicotinic receptor antagonists, with an order of potency of mecamylamine > lobeline > cytisine > methyllycaconitine > dihydro-beta-erythroidine. The pharmacological profile of this nicotinic receptor is largely consistent with that described previously for an alpha4beta2 subunit combination, although discrepancies in the efficacies of agonists were observed. No additivity in NMDA- and nicotine-stimulated [3H]NE release was observed, suggesting a common signal transduction mechanism. However, the pharmacological characteristics of MK-801 blockade of nicotine-induced responses were not consistent with those of an NMDA receptor. We therefore conclude that nicotine directly releases [3H]NE from LC cells and does not act indirectly via activation of glutamate release.  相似文献   

14.
1. The mechanisms underlying the postjunctional blockade induced by phenthonium [N-(4-phenyl) phenacyl 1-hyoscyamine] were investigated in mammalian and amphibian muscles. This muscarinic antagonist was previously shown to enhance specifically the spontaneous acetylcholine (ACh) release at concentrations that blocked neuromuscular transmission. 2. In both rat diaphragm and frog sartorius muscles, phenthonium (Phen, 1-100 microM) depressed the muscle twitches elicited by nerve stimulation (IC50: 23 microM and 5 microM, respectively), and blocked the nerve-evoked muscle action potential. The neuromuscular blockade was not reversed after incubation with neostigmine. 3. Equal concentrations of Phen decreased the rate of rise and prolonged the falling phase of the directly elicited action potential in frog sartorius muscle fibres, indicating that the drug also affects the sodium and potassium conductance. 4. Phen (50 and 100 microM) protected the ACh receptor against alpha-bungarotoxin (BUTX) blockade in the mouse diaphragm allowing recording of endplate potentials and action potentials after 5 h wash with physiological salt solution. 5. Phen (10-100 microM) produced a concentration- and voltage-dependent decrease of the endplate current (e.p.c.), and induced nonlinearity of the current-voltage relationship. At high concentrations Phen also shortened the decay time constant of e.p.c (tau(e.p.c.)) and reduced its voltage sensitivity. 6. At the same range of concentrations, Phen also reduced the initial rate of [125I]-BUTX binding to junctional ACh receptors of the rat diaphragm (apparent dissociation constant = 24 microM), the relationship between the degree of inhibition and antagonist concentration being that expected for a competitive mechanism. 7. It is concluded that Phen affects the electrical excitability of the muscle fibre membrane, and blocks neuromuscular transmission through a mechanism that affects the agonist binding to its recognition site and ionic channel conductance of the nicotinic ACh receptor.  相似文献   

15.
Inositol 1,4,5-trisphosphate (InsP3) is a second messenger responsible for the rapid and discontinuous release of Ca2+ from intracellular stores. In this study, the effects of the sulfhydryl reagent thimerosal were investigated on Ca2+ mobilization and on InsP3 binding. Thimerosal was shown to release Ca2+, in a dose-dependent manner, with an EC50 of 135.8 +/- 5.2 microM, from bovine adrenal cortex microsomes. Thimerosal-induced Ca2+ release was not prevented by heparin (250 micrograms/ml), ruling out a participation of InsP3 receptor in that effect. The slow rate of thimerosal-induced Ca2+ release rather suggested an inhibition of microsomal Ca2+ ATPase. At submaximal concentration, thimerosal (100 microM) was also shown to potentiate the release of Ca2+ induced by InsP3. Dose-response experiments revealed that thimerosal enhanced the apparent affinity of InsP3 by a factor 2.21 +/- 0.28, without modifying the maximal amount of Ca2+ released by InsP3. Thimerosal also enhanced, in a dose-dependent manner, [3H]InsP3 binding to adrenal cortex microsomes (EC50 = 43.3 +/- 7.6 microM). A similar effect was also observed on [3H]InsP3 binding to solubilized receptors, suggesting a direct modification of the receptor protein by thimerosal. The effects of thimerosal on Ca2+ release and [3H]InsP3 binding were abolished in the presence of the reducing agent dithiothreitol (1 mM), suggesting a modification by thimerosal of specific thiol groups on these microsomal proteins. Scatchard analysis revealed that thimerosal (100 microM) increased InsP3 receptor affinity by 1.87 +/- 0.26-fold. Kinetic analysis indicated that this increased affinity was due to an enhancement of InsP3 association rate constant. The concomitant increases of binding affinity and Ca2+ releasing potency suggest that the high affinity state of InsP3 receptor is a functional state.  相似文献   

16.
Clonal neurohybridoma NCB-20 cells express a valproate-insensitive succinic semialdehyde reductase activity that transforms succinic semialdehyde into gamma-hydroxybutyrate. This activity (1.14+/-0.16 nmol/min/mg protein) was similar to the lowest activity existing in adult rat brain. [3H]gamma-Hydroxybutyrate labels a homogeneous population of sites on NCB-20 cell membranes (Kd=250+/-44.4nM, Bmax=180+/-16.2fmol/mg protein) that apparently represents specific gamma-hydroxybutyrate binding sites characterized previously on brain cell membranes. Finally, an Na+-dependent uptake of [3H]gamma-hydroxybutyrate was expressed in NCB-20 cells with a Km of 35+21.1 microM and a Vmax of 80+/-14.2 pmol/min/mg protein. A three-day treatment with 1 mM dibutyryl-cyclic-AMP induced a three-fold increase in the cellular succinic semialdehyde reductase activity. In parallel, a K+-evoked release of [3H]gamma-hydroxybutyrate occurred. This release was Ca2+ dependent and was not present in undifferentiated cells. Cyclic-AMP treatment induced a decrease of [3H]gamma-hydroxybutyrate binding sites, which could be due to spontaneous gamma-hydroxybutyrate release. Patch-clamp experiments carried out on differentiated NCB-20 cells revealed the presence of Ca2+ conductances which were partially inhibited by 50 microM gamma-hydroxybutyrate. This gamma-hydroxybutyrate-induced effect was blocked by the gamma-hydroxybutyrate receptor antagonist NCS-382, but not by the GABA(B) antagonist CGP-55845. These results demonstrate the presence of an active gamma-hydroxybutyratergic system in NCB-20 cells which possesses the ability to release gamma-hydroxybutyrate. These cells express specific gamma-hydroxybutyrate receptors which modulate Ca2+ currents independently of GABA(B) receptors.  相似文献   

17.
The effect of 7-nitroindazole (7-NI), an inhibitor of neuronal nitric oxide synthase (nNOS) on the dimethylphenylpiperazinium(DMPP)-evoked release of [3H]noradrenaline ([3H]NA) from rat hippocampal slices was studied. The effect of DMPP (20 microM) to increase the basal release of [3H]NA was significantly potentiated by 7-NI (40 microM). In our previous study we showed that the response to DMPP has two components, a nicotinic receptor-mediated, [Ca2+]-dependent exocytosis followed by a [Ca2+]-independent, uptake blocker-sensitive carrier-mediated release. To clarify which part of the response was affected by the inhibition of nNOS, we investigated the effect of 7-NI on the nicotine-evoked NA release (nicotine has only receptor-mediated effect) and on the DMPP-evoked NA release in Ca(2+)-free medium where the receptor-mediated component is abolished. Nicotine (100 microM) significantly increased the basal release of [3H]NA but this release was not affected, whereas in Ca(2+)-free medium the response to DMPP (20 microM) was still potentiated by 7-NI (40 microM). In the presence of the NA uptake blocker desipramine (10 microM) DMPP (20 microM) was unable to provoke NA release independently from the presence or absence of 7-NI (40 microM). Our data show that 7-NI influences the carrier-mediated component of DMPP-evoked [3H]NA release, which indicates that nitric oxide produced by nNOS may play a role in the regulation of the NA uptake carrier.  相似文献   

18.
Pharmacological characterization of [3H]benzodiazepine binding to membrane preparations of adult rat hippocampus and neonatal rat brain have demonstrated, in addition to the omega 1 and omega 2 populations of central omega benzodiazepine binding sites associated with GABAA receptors, the existence of binding sites with microM affinity for the imidazopyridines zolpidem and alpidem. In the present study we have investigated their comparative autoradiographic distribution using [3H]flumazenil as a ligand. In the neonatal rat CNS, the imidazopyridine derivatives zolpidem and alpidem were found to discriminate two [3H]flumazenil binding site populations with an IC50 value ratio of more than 200-fold. In the different regions investigated (spinal cord, striatum, neocortex and inferior colliculus) the low affinity component had IC50 values of 20-40 microM (zolpidem) and 5-15 microM (alpidem) and accounted for ca. 50% of the total binding site population. In the adult rat, these imidazopyridine derivatives displayed a greater displacing potency in the cerebellum (IC50 = 6 and 36 nM, respectively) than in the hippocampus (IC50 = 37 and 403 nM, respectively). In the cerebellum, [3H]flumazenil binding was fully displaced by 1 microM of either compound and Hill coefficients of displacement curves were close to unity. In the hippocampus, 25% of [3H]flumazenil binding were resistant to 3 microM zolpidem or 1 microM alpidem, but were displaced by 100 microM of either compound. CL 218,872 also displayed a greater displacing potency in the cerebellum (IC50 = 83 nM) than in the hippocampus (IC50 = 711 nM) but [3H]flumazenil binding in the hippocampus was fully displaced by 10 microM of this compound. In adult rat hippocampus, zolpidem and alpidem were found to discriminate between three central omega site subtypes which display high (IC50 = 31 and 6.1 nM, for these imidazopyridine derivatives. In contrast, CL 218,872 discriminated between omega 1 and omega 2 sites but not between two omega 2 receptor subpopulations. omega 1 sites were mainly localized in layer IV of the sensorimotor cortex, cerebellum, substantia nigra, olfactory bulb and inferior colliculus. omega 2I sites were present in the cortical mantle (with higher levels in the cingulate and olfactory than in the sensorimotor cortex) and in subcortical (hippocampus, hypothalamus and nucleus accumbens) limbic structures. In the hippocampus, hypothalamus, spinal cord and nucleus accumbens, omega 2L sites accounted for more than 25% of the specific [3H]flumazenil binding; the density of these sites was minor in the cortex and in most pyramidal and extrapyramidal system structures.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
AIM: To investigate age related alterations in glutamate N-methyl-D-aspartate (NMDA) receptor binding produced by the modulatory compounds glutamate, glycine, and magnesium (Mg2+) sulphate. METHODS: The effects produced by glutamate plus glycine, and Mg2+ on the binding of [3H]MK-801, a ligand for the N-methyl-D-aspartate ion channel phencyclidine site, were measured in membrane preparations made from prefrontal cortex from human neonate (n = 5), infant (n = 6), and adult (n = 6) necropsy brains. RESULTS: Neonatal brains had the least [3H]MK-801 binding, suggesting either a low density of NMDA receptors or a more restricted access of [3H]MK-801 to cation channel sites. Infant brains had the most [3H]MK-801 binding which was stimulated to a greater extent by L-glutamate (100 microM) and glycine (10 microM) than in neonatal and adult brains. MG2+ invariably inhibited [3H]MK-801 binding. However, the Mg2+ IC50 value was higher in neonatal brain (3.6 mM) than infant (1.4 mM) and adult (0.87 mM) brains. CONCLUSION: Infant brain may have excess NMDA receptors which are hyper responsive to glutamate and glycine. The lower potency of Mg2+ to inhibit [3H]MK-801 binding in neonatal cortex may be because newborn babies have NMDA receptors without the normal complement of Mg2+ sites. The findings suggest that therapeutic NMDA receptor block in neonates requires higher concentrations of magnesium sulphate in brain tissue.  相似文献   

20.
The retina possesses cholinergic amacrine cells which release acetylcholine (ACh) in response to flickering light. Using an eye-cup preparation in anaesthetized rabbits we found that when the retina was exposed to nociceptin, the light-evoked release of ACh was reduced in a concentration-dependent manner (IC50 = 100 nM), the maximum effect being 60% inhibition. Opioid receptors were not involved in the inhibitory effect of nociceptin because its action was not blocked by naloxone (1 microM) and furthermore mu-opioids enhanced the light-evoked release of ACh. Using rabbit retina homogenates we found that the retina possessed a substantial number of high-affinity binding sites for [3H]-nociceptin indicating the presence of ORL1-receptors. Since [des-Phe1]-nociceptin, which has no affinity for the ORL1-receptor, had no effect on the light-evoked release of ACh it is unlikely that the action of nociceptin was simply non-specific. We conclude that the inhibitory effect of nociceptin on retinal ACh release involves activation of the ORL1 receptors.  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