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.
  1. It is unclear whether GABAA receptor-mediated hyperpolarizing and depolarizing synaptic potentials (IPSPAs and DPSPAs, respectively) are evoked by (a) the same populations of GABAergic interneurones and (b) exhibit similar regulation by allosteric modulators of GABAA receptor function. We have attempted to address these questions by investigating the effects of (a) known agonists for presynaptic receptors on GABAergic terminals, and (b) a range of GABAA receptor ligands, on each response.
  2. The GABA uptake inhibitor NNC 05-711 (10 μM) enhanced whereas bicuculline (10 μM) inhibited both IPSPAs and DPSPAs.
  3. (−)-Baclofen (5 μM), [D-Ala2,N-Me-Phe4,Gly5-ol]-enkephalin (DAGO; 0.5 μM), and carbachol (10 μM) caused substantial depressions (up to 99%) of DPSPAs that were reversed by CGP 55845A (1 μM), naloxone (10 μM) and atropine (5 μM), respectively. In contrast, 2-chloroadenosine (CADO; 10 μM) only slightly depressed DPSPAs. Quantitatively, the effect of each agonist was similar to that reported for IPSPAs.
  4. The neurosteroid ORG 21465 (1–10 μM), the anaesthetic propofol (50–500 μM), the barbiturate pentobarbitone (100–300 μM) and zinc (50 μM) all enhanced DPSPAs and IPSPAs.
  5. The benzodiazepine (BZ) agonist flunitrazepam (10–50 μM) and inverse agonist DMCM (1 μM) caused a respective enhancement and inhibition of both IPSPAs and DPSPAs. The BZω1 site agonist zolpidem (10–30 μM) produced similar effects to flunitrazepam.
  6. The anticonvulsant loreclezole (1–100 μM) did not affect either response.
  7. These data demonstrate that similar populations of inhibitory interneurones can generate both IPSPAs and DPSPAs by activating GABAA receptors that are subject to similar allosteric modulation.
  相似文献   

2.
  1. Radioligand binding and patch-clamp techniques were used to study the actions of γ-aminobutyric acid (GABA) and the general anaesthetics propofol (2,6-diisopropylphenol), pentobarbitone and 5α-pregnan-3α-ol-20-one on rat α1 and β3 GABAA receptor subunits, expressed either alone or in combination.
  2. Membranes from HEK293 cells after transfection with α1 cDNA did not bind significant levels of [35S]-tert-butyl bicyclophosphorothionate ([35S]-TBPS) (<0.03 pmol mg−1 protein). GABA (100 μM) applied to whole-cells transfected with α1 cDNA and clamped at −60 mV, also failed to activate discernible currents.
  3. The membranes of cells expressing β3 cDNAs bound [35S]-TBPS (∼1 pmol mg−1 protein). However, the binding was not influenced by GABA (10 nM–100 μM). Neither GABA (100 μM) nor picrotoxin (10 μM) affected currents recorded from cells expressing β3 cDNA, suggesting that β3 subunits do not form functional GABAA receptors or spontaneously active ion channels.
  4. GABA (10 nM–100 μM) modulated [35S]-TBPS binding to the membranes of cells transfected with both α1 and β3 cDNAs. GABA (0.1 μM–1 mM) also dose-dependently activated inward currents with an EC50 of 9 μM recorded from cells transfected with α1 and β3 cDNAs, clamped at −60 mV.
  5. Propofol (10 nM–100 μM), pentobarbitone (10 nM–100 μM) and 5α-pregnan-3α-ol-20-one (1 nM–30 μM) modulated [35S]-TBPS binding to the membranes of cells expressing either α1β3 or β3 receptors. Propofol (100 μM), pentobarbitone (1 mM) and 5α-pregnan-3α-ol-20-one (10 μM) also activated currents recorded from cells expressing α1β3 receptors.
  6. Propofol (1 μM–1 mM) and pentobarbitone (1 mM) both activated currents recorded from cells expressing β3 homomers. In contrast, application of 5α-pregnan-3α-ol-20-one (10 μM) failed to activate detectable currents.
  7. Propofol (100 μM)-activated currents recorded from cells expressing either α1β3 or β3 receptors reversed at the C1 equilibrium potential and were inhibited to 34±13% and 39±10% of control, respectively, by picrotoxin (10 μM). 5α-Pregnan-3α-ol-20-one (100 nM) enhanced propofol (100 μM)-evoked currents mediated by α1β3 receptors to 1101±299% of control. In contrast, even at high concentration 5α-pregnan-3α-ol-20-one (10 μM) caused only a modest facilitation (to 128±12% of control) of propofol (100 μM)-evoked currents mediated by β3 homomers.
  8. Propofol (3–100 μM) activated α1β3 and β3 receptors in a concentration-dependent manner. For both receptor combinations, higher concentrations of propofol (300 μM and 1 mM) caused a decline in current amplitude. This inhibition of receptor function reversed rapidly during washout resulting in a ‘surge'' current on cessation of propofol (300 μM and 1 mM) application. Surge currents were also evident following pentobarbitone (1 mM) application to cells expressing either receptor combination. By contrast, this phenomenon was not apparent following applications of 5α-pregnan-3α-ol-20-one (10 μM) to cells expressing α1β3 receptors.
  9. These observations demonstrate that rat β3 subunits form homomeric receptors that are not spontaneously active, are insensitive to GABA and can be activated by some general anaesthetics. Taken together, these data also suggest similar sites on GABAA receptors for propofol and barbiturates, and a separate site for the anaesthetic steroids.
  相似文献   

3.
  1. The site(s) at which P2-receptor agonists act to evoke contractions of the rat isolated tail artery was studied by use of P2-receptor antagonists and the extracellular ATPase inhibitor 6-N,N-diethyl-D-β,γ-dibromomethyleneATP (ARL 67156).
  2. Suramin (1 μM–1 mM) and pyridoxalphosphate-6-azophenyl-2′,4′-disulphonic acid (PPADS) (0.3–300 μM) inhibited contractions evoked by equi-effective concentrations of α,β-methyleneATP (α,β-meATP) (5 μM), 2-methylthioATP (2-meSATP) (100 μM) and adenosine 5′-triphosphate (ATP) (1 mM) in a concentration-dependent manner. Responses to α,β-meATP and 2-meSATP were abolished, but approximately one third of the peak response to ATP was resistant to suramin and PPADS.
  3. Contractions evoked by uridine 5′-triphosphate (UTP) (1 mM) were slightly inhibited by suramin (100 and 300 μM) and potentiated by PPADS (300 μM).
  4. Desensitization of the P2X1-receptor by α,β-meATP abolished contractions evoked by 2-meSATP (100 μM) and reduced those to ATP (1 mM) and UTP (1 mM) to 15±3% and 68±4% of control.
  5. Responses to α,β-meATP (5 μM) and 2-meSATP (100 μM) were abolished when tissues were bathed in nominally calcium-free solution, while the peak contractions to ATP (1 mM) and UTP (1 mM) were reduced to 24±6% and 61±13%, respectively, of their control response.
  6. ARL 67156 (3–100 μM) potentiated contractions elicited by UTP (1 mM), but inhibited responses to α,β-meATP (5 μM), 2-meSATP (100 μM) and ATP (1 mM) in a concentration-dependent manner.
  7. These results suggest that two populations of P2-receptors are present in the rat tail artery; ligand-gated P2X1-receptors and G-protein-coupled P2Y-receptors.
  相似文献   

4.
  1. Extracellular single unit recording techniques were used to compare the effects of (-)-3-amino-1-hydroxypyrrolidin-2-one ((–)-HA-966) and (±)-baclofen on the activity of dopamine-containing neurones in 300 μm slices of rat substantia nigra. Electrophysiological data were compared with the outcome of in vitro binding experiments designed to assess the affinity of (–)-HA-966 for γ-aminobutyric acid (GABAB) receptors.
  2. Bath application of (–)-HA-966 produced a concentration-dependent inhibition of dopaminergic neuronal firing (EC50=444.0 μM; 95% confidence interval: 277.6 μM–710.1 μM, n=27) which was fully reversible upon washout from the recording chamber. Although similar effects were observed in response to (±)-baclofen, the direct-acting GABAB receptor agonist proved to be considerably more potent than (–)-HA-966 (EC50=0.54 μM; 95% confidence interval: 0.44 μM–0.66 μM, n=29) in vitro.
  3. Low concentrations of chloral hydrate (10 μM) were without effect on the basal firing rate of nigral dopaminergic neurones but significantly increased the inhibitory effects produced by concomitant application of (–)-HA-966.
  4. The inhibitory effects of (–)-HA-966 were completely reversed in the presence of the GABAB receptor antagonists, CGP-35348 (100 μM) and 2-hydroxysaclofen (500 μM). Bath application of CGP-35348 alone increased basal firing rate. However, the magnitude of the excitation (9.2±0.3%) was not sufficient to account for the ability of the antagonist to reverse fully the inhibitory effects of (–)-HA-966.
  5. (–)-HA-966 (0.1–1.0 mM) produced a concentration-dependent displacement of [H]-GABA from synaptic membranes in the presence of isoguvacine (40 μM). However, the affinity of the drug for GABAB binding sites was significantly less than that of GABA (0.0005 potency ratio) and showed no apparent stereoselectivity.
  6. These results indicate that while (–)-HA-966 appears to act as a direct GABAB receptor agonist in vitro, its affinity for this receptor site is substantially less than that of GABA or baclofen and unlikely to account for the depressant actions of this drug which occur at levels approximately ten fold lower in vivo.
  相似文献   

5.
  1. The effect of protein tyrosine kinase inhibitors on human adenosine A1 receptor-mediated [3H]-inositol phosphate ([3H]-IP) accumulation has been studied in transfected Chinese hamster ovary cells (CHO-A1) cells.
  2. In agreement with our previous studies the selective adenosine A1 receptor agonist N6-cyclopentyladenosine (CPA) stimulated the accumulation of [3H]-IPs in CHO-A1 cells. Pre-treatment with the broad spectrum tyrosine kinase inhibitor genistein (100 μM; 30 min) potentiated the responses elicited by 1 μM (199±17% of control CPA response) and 10 μM CPA (234±15%). Similarly, tyrphostin A47 (100 μM) potentiated the accumulation of [3H]-IPs elicited by 1 μM CPA (280±32%).
  3. Genistein (EC50=13.7±1.2 μM) and tyrphostin A47 (EC50=10.4±3.9 μM) potentiated the [3H]-IP response to 1 μM CPA in a concentration-dependent manner.
  4. Pre-incubation with the inactive analogues of genistein and tyrphostin A47, daidzein (100 μM; 30 min) and tyrphostin A1 (100 μM; 30 min), respectively, had no significant effect on the accumulation of [3H]-IPs elicited by 1 μM CPA.
  5. Genistein (100 μM) had no significant effect on the accumulation of [3H]-IPs produced by the endogenous thrombin receptor (1 u ml−1; 100±10% of control response). In contrast, tyrphostin A47 produced a small augmentation of the thrombin [3H]-IP response (148±13%).
  6. Genistein (100 μM) had no effect on the [3H]-IP response produced by activation of the endogenous Gq-protein coupled CCKA receptor with the sulphated C-terminal octapeptide of cholecystokinin (1 μM CCK-8; 96±6% of control). In contrast, tyrphostin A47 (100 μM) caused a small but significant increase in the response to 1 μM CCK-8 (113±3% of control).
  7. The phosphatidylinositol 3-kinase inhibitor LY 294002 (30 μM) and the MAP kinase kinase inhibitor PD 98059 (50 μM) had no significant effect on the [3H]-IP responses produced by 1 μM CPA and 1 μM CCK-8.
  8. These observations suggest that a tyrosine kinase-dependent pathway may be involved in the regulation of human adenosine A1 receptor mediated [3H]-IP responses in CHO-A1 cells.
  相似文献   

6.
  1. The influence of L-NG-nitro-arginine (L-NOARG, 30 μM) on contractile responses to exogenous noradrenaline was studied in the rat anococcygeus muscle.
  2. Noradrenaline (0.1–100 μM) contracted the muscle in a concentration-dependent manner. L-NOARG (30 μM) had no effect on noradrenaline responses.
  3. Phenoxybenzamine (Pbz 0.1 μM) depressed by 46% (P<0.001) the maximum response and shifted to the right (P<0.001) the E/[A] curve to noradrenaline (pEC50 control: 6.92±0.09; pEC50 Pbz: 5.30±0.10; n=20).
  4. The nested hyperbolic null method of analysing noradrenaline responses after phenoxybenzamine showed that only 0.61% of the receptors need to be occupied to elicit 50% of the maximum response, indicating a very high functional receptor reserve.
  5. Contractile responses to noradrenaline after partial α1-adrenoceptor alkylation with phenoxybenzamine (0.1 μM) were clearly enhanced by L-NOARG.
  6. The potentiating effect of L-NOARG on noradrenaline responses after phenoxybenzamine was reversed by (100 μM) L-arginine but not by (100 μM) D-arginine.
  7. These results indicate that spontaneous release of NO by nitrergic nerves can influence the α1-adrenoceptor-mediated response to exogenous noradrenaline.
  相似文献   

7.
  1. In vitro studies were performed to examine the mechanisms underlying substance P-induced enhancement of constriction rate in guinea-pig mesenteric lymphatic vessels.
  2. Substance P caused an endothelium-dependent increase in lymphatic constriction frequency which was first significant at a concentration of 1 nM (115±3% of control, n=11) with 1 μM, the highest concentration tested, increasing the rate to 153±4% of control (n=9).
  3. Repetitive 5 min applications of substance P (1 μM) caused tachyphylaxis with tissue responsiveness tending to decrease (by an average of 23%) and significantly decreasing (by 72%) for application at intervals of 30 and 10 min, respectively.
  4. The competitive antagonist of tachykinin receptors, spantide (5 μM) and the specific NK1 receptor antagonist, WIN51708 (10 μM) both prevented the enhancement of constriction rate induced by 1 μM substance P.
  5. Endothelial cells loaded with the Ca2+ sensing fluophore, fluo 3/AM did not display a detectable change in [Ca2+]i upon application of 1 μM substance P.
  6. Inhibition of nitric oxide synthase by NG nitro-L-arginine (L-NOARG; 100 μM) had no significant effect on the response induced by 1 μM substance P.
  7. The enhancement of constriction rate induced by 1 μM substance P was prevented by the cyclo-oxygenase inhibitor, indomethacin (3 μM), the thromboxane A2 synthase inhibitor, imidazole (50 μM), and the thromboxane A2 receptor antagonist, SQ29548 (0.3 μM).
  8. The stable analogue of thromboxane A2, U46619 (0.1 μM) significantly increased the constriction rate of lymphangions with or without endothelium, an effect which was prevented by SQ29548 (0.3 μM).
  9. Treatment with pertussis toxin (PTx; 100 ng ml−1) completely abolished the response to 1 μM substance P without inhibiting either the perfusion-induced constriction or the U46619-induced enhancement of constriction rate.
  10. Application of the phospholipase A2 inhibitor, antiflammin-1 (1 nM) prevented the enhancement of lymphatic pumping induced by substance P (1 μM), without inhibiting the response to either U46619 (0.1 μM) or acetylcholine (10 μM).
  11. The data support the hypothesis that the substance P-induced increase in pumping rate is mediated via the endothelium through NK1 receptors coupled by a PTx sensitive G-protein to phospholipase A2 and resulting in generation of the arachidonic acid metabolite, thromboxane A2, this serving as the diffusible activator.
  相似文献   

8.
  1. The release of endogenous γ-aminobutyric acid (GABA) and glutamic acid in the human brain has been investigated in synaptosomal preparations from fresh neocortical samples obtained from patients undergoing neurosurgery to reach deeply located tumours.
  2. The basal outflows of GABA and glutamate from superfused synaptosomes were largely increased during depolarization with 15 mM KCl. The K+-evoked overflows of both amino acids were almost totally dependent on the presence of Ca2+ in the superfusion medium.
  3. The GABAB receptor agonist (−)-baclofen (1, 3 or 10 μM) inhibited the overflows of GABA and glutamate in a concentration-dependent manner. The inhibition caused by 10 μM of the agonist ranged from 45–50%.
  4. The effect of three selective GABAB receptor antagonists on the inhibition of the K+-evoked GABA and glutamate overflows elicited by 10 μM (−)-baclofen was investigated. Phaclofen antagonized (by about 50% at 100 μM; almost totally at 300 μM) the effect of (−)-baclofen on GABA overflow but did not modify the inhibition of glutamate release. The effect of (−)-baclofen on the K+-evoked GABA overflow was unaffected by 3-amino-propyl (diethoxymethyl)phosphinic acid (CGP 35348; 10 or 100 μM); however, CGP 35348 (10 or 100 μM) antagonized (−)-baclofen (complete blockade at 100 μM) at the heteroreceptors on glutamatergic terminals. Finally, [3-[[(3,4-dichlorophenyl) methyl]amino]propyl] (diethoxymethyl) phosphinic aid (CGP 52432), 1 μM, blocked the GABAB autoreceptor, but was ineffective at the heteroreceptors. The selectivity of CGP 52432 was lost at 30 μM, as the compound, at this concentration, inhibited completely the (−)-baclofen effect both on GABA and glutamate release.
  5. It is concluded that GABA and glutamate release evoked by depolarization of human neocortex nerve terminals can be affected differentially through pharmacologically distinct GABAB receptors.
  相似文献   

9.
  1. The interaction of melatonin (N-acetyl-5-methoxytryptamine) with 5-hydroxytryptamine4 (5-HT4) receptors and/or with melatonin receptors (ML1, ML2 sites) has been assessed in isolated strips of the guinea-pig proximal colon. In the same preparation, the pharmacological profile of a series of melatonin agonists (2-iodomelatonin, 6-chloromelatonin, N-acetyl-5-hydroxytryptamine (N-acetyl-5-HT), 5-methoxycarbonylamino-N-acetyltryptamine (5-MCA-NAT)) was investigated.
  2. In the presence of 5-HT1/2/3 receptor blockade with methysergide (1 μM) and ondansetron (10 μM), melatonin (0.1 nM–10 μM), 5-HT (1 nM–1 μM) and the 5-HT4 receptor agonist, 5-methoxytryptamine (5-MeOT: 1 nM–1 μM) caused concentration-dependent contractile responses. 5-HT and 5-MeOT acted as full agonists with a potency (−log EC50) of 7.8 and 8.0, respectively. The potency value for melatonin was 8.7, but its maximum effect was only 58% of that elicited by 5-HT.
  3. Melatonin responses were resistant to atropine (0.1 μM), tetrodotoxin (0.3 μM), and to blockade of 5-HT4 receptors by SDZ 205,557 (0.3 μM) and GR 125487 (3, 30 and 300 nM). The latter antagonist (3 nM) inhibited 5-HT-induced contractions with an apparent pA2 value of 9.6. GR 125487 antagonism was associated with 30% reduction of the 5-HT response maximum. Contractions elicited by 5-HT were not modified when melatonin (1 and 10 nM) was used as an antagonist.
  4. Like melatonin, the four melatonin analogues concentration-dependently contracted colonic strips. The rank order of agonist potency was: 2-iodomelatonin (10.8) >6-chloromelatonin (9.9) ⩾ N-acetyl-5-HT (9.8) ⩾5-MCA-NAT (9.6) >melatonin (8.7), an order typical for ML2 sites. In comparison with the other agonists, 5-MCA-NAT had the highest intrinsic activity.
  5. The melatonin ML1B receptor antagonist luzindole (0.3, 1 and 3 μM) had no effect on the concentration-response curve to melatonin. Prazosin, an α-adrenoceptor antagonist possessing moderate/high affinity for melatonin ML2 sites did not affect melatonin-induced contractions at 0.1 μM. Higher prazosin concentrations (0.3 and 1 μM) caused a non-concentration-dependent depression of the maximal response to melatonin without changing its potency. Prazosin (0.1 and 1 μM) showed a similar depressant behaviour towards the contractile responses to 5-MCA-NAT.
  6. In the guinea-pig proximal colon, melatonin despite some structural similarity with the 5-HT4 receptor agonist 5-MeOT, does not interact with 5-HT4 receptors (or with 5-HT1/2/3 receptors). As indicated by the rank order of agonist potencies and by the inefficacy of luzindole, the most likely sites of action of melatonin are postjunctional ML2 receptors. However, this assumption could not be corroborated with the use of prazosin as this ‘ML2 receptor antagonist'' showed only a non-concentration-dependent depression of the maximal contractile response to both melatonin and 5-MCA-NAT. Further investigation with the use of truly selective antagonists at melatonin ML2 receptors is required to clarify this issue.
  相似文献   

10.
  1. 5-Hydroxytryptamine (5-HT; 1 nM–100 μM) concentration-dependently inhibited the amplitude and frequency of spontaneous contractions in longitudinal and circular muscles of the porcine myometrium. The circular muscle (EC50; 68–84 nM) was more sensitive than the longitudinal muscle (EC50; 1.3–1.44 μM) to 5-HT. To characterize the 5-HT receptor subtype responsible for inhibition of myometrial contractility, the effects of 5-HT receptor agonists on spontaneous contractions and of 5-HT receptor antagonists on inhibition by 5-HT were examined in circular muscle preparations.
  2. Pretreatment with tetrodotoxin (1 μM), propranolol (1 μM), atropine (1 μM), guanethidine (10 μM) or L-NAME (100 μM) failed to change the inhibition by 5-HT, indicating that the inhibition was due to a direct action of 5-HT on the smooth muscle cells.
  3. 5-CT, 5-MeOT and 8-OH-DPAT mimicked the inhibitory response of 5-HT, and the rank order of the potency was 5-CT>5-HT>5-MeOT>8-OH-DPAT. On the other hand, oxymethazoline, α-methyl-5-HT, 2-methyl-5-HT, cisapride, BIMU-1, BIMU-8, ergotamine and dihydroergotamine had almost no effect on spontaneous contractions, even at 10–100 μM.
  4. Inhibition by 5-HT was not decreased by either pindolol (1 μM), ketanserin (1 μM), tropisetron (10 μM), MDL72222 (1 μM) or GR113808 (10 μM), but was antagonized by the following compounds in a competitive manner (with pA2 values in parentheses): methiothepin (8.05), methysergide (7.92), metergoline (7.4), mianserin (7.08), clozapine (7.06) and spiperone (6.86).
  5. Ro 20-1724 (20 μM) and rolipram (10 μM) significantly enhanced the inhibitory response of 5-HT, but neither zaprinast (10 μM) nor dipyridamole (10 μM) altered the response of 5-HT.
  6. 5-HT (1 nM–1 μM) caused a concentration-dependent accumulation of intracellular cyclic AMP in the circular muscle.
  7. From the present results, the 5-HT receptor, which is functionally correlated with the 5-HT7 receptor, mediates the inhibitory effect of 5-HT on porcine myometrial contractility. This inhibitory response is probably due to an increase in intracellular cyclic AMP through the activation of adenylate cyclase that is positively coupled to 5-HT7 receptors.
  相似文献   

11.
  1. In the guinea-pig isolated vas deferens preparation bathed in Tyrode''s solution, the prostacyclin analogues, cicaprost, TEI-9063, iloprost, taprostene and benzodioxane-prostacyclin, enhanced twitch responses to submaximal electrical field stimulation (20%-EFS). The high potency of cicaprost (EC150=1.3 nM) and the relative potencies of the analogues (equi-effective molar ratios=1.0, 0.85, 1.6, 17 and 82, respectively) suggest the involvement of a prostacyclin (IP-) receptor.
  2. Maximum enhancement induced by cicaprost in 2.5 mM K+ Krebs-Henseleit solution was similar to that in Tyrode solution (2.7 mM K+), but was progressively reduced as the K+ concentration was increased to 3.9, 5.9 and 11.9 mM. There was also a greater tendency for the other prostacyclin analogues to inhibit EFS responses in 5.9 mM standard K+ Krebs-Henseleit solution; this may be attributed to their agonist actions on presynaptic EP3-receptors resulting in inhibition of transmitter release.
  3. The EFS enhancing action of cicaprost was not affected by the α1-adrenoceptor antagonist prazosin (100 and 1000 nM). Cicaprost (20 and 200 nM) did not affect contractile responses of the vas deferens to either ATP (5 μM) or α,β-methylene ATP (1 μM) in the presence of tetrodotoxin (TTX, 100 nM). In addition, enhancement by cicaprost of responses to higher concentrations of ATP (30 and 300 μM) in the absence of TTX, as shown previously by others, was not seen. Prostaglandin E2 (PGE2, 10 nM) and another prostacyclin analogue TEI-3356 (20 nM) enhanced purinoceptor agonist responses. Unexpectedly, TTX (0.1 and 1 μM) partially inhibited contractions elicited by 10–1000 μM ATP; contractions elicited by 1–3 μM ATP were unaffected. Further studies are required to establish whether a pre- or post-synaptic mechanism is involved.
  4. In a separate series of experiments, cicaprost (5–250 nM), TEI-9063 (3–300 nM), 4-aminopyridine (10–100 μM) and tetraethylammonium (100–1000 μM) enhanced both 20%-EFS responses and the accompanying overflow of noradrenaline to a similar extent. In further experiments with the EP1-receptor antagonist AH 6809, TEI-3356 (1.0–100 nM) and the EP3-receptor agonist, sulprostone (0.1–1.0 nM) inhibited both maximal EFS responses and noradrenaline overflow, thus confirming previous reports of the high activity of TEI-3356 at the EP3-receptor. Cicaprost had no significant effect on noradrenaline overflow at 10 and 100 nM, but produced a modest inhibition at 640 nM.
  5. In conclusion, our studies show that prostacyclin analogues (particularly TEI-3356) can inhibit EFS responses of the guinea-pig vas deferens by acting as agonists at presynaptic EP3-receptors. Prostacyclin analogues (particularly cicaprost and TEI-9063) can also enhance EFS responses through activation of IP-receptors. The mechanism of the enhancement has not been rigorously established but from our results we favour a presynaptic action to increase transmitter release.
  相似文献   

12.
  1. The effects of the antidiabetic agent englitazone and the anorectic drug ciclazindol on ATP-sensitive K+ (KATP) channels activated by diazoxide and leptin were examined in the CRI-G1 insulin-secreting cell line using whole cell and single channel recording techniques.
  2. In whole cell current clamp mode, the hyperglycaemic agent diazoxide (200 μM) and the ob gene product leptin (10 nM) hyperpolarised CRI-G1 cells by activation of KATP currents. KATP currents activated by either agent were inhibited by tolbutamide, with an IC50 for leptin-activated currents of 9.0 μM.
  3. Application of englitazone produced a concentration-dependent inhibition of KATP currents activated by diazoxide (200 μM) with an IC50 value of 7.7 μM and a Hill coefficient of 0.87. In inside-out patches englitazone (30 μM) also inhibited KATP channel currents activated by diazoxide by 90.8±4.1%.
  4. In contrast, englitazone (1–30 μM) failed to inhibit KATP channels activated by leptin, although higher concentrations (>30 μM) did inhibit leptin actions. The englitazone concentration inhibition curve in the presence of leptin resulted in an IC50 value and Hill coefficient of 52 μM and 3.2, respectively. Similarly, in inside-out patches englitazone (30 μM) failed to inhibit the activity of KATP channels in the presence of leptin.
  5. Ciclazindol also inhibited KATP currents activated by diazoxide (200 μM) in a concentration-dependent manner, with an IC50 and Hill coefficient of 127 nM and 0.33, respectively. Furthermore, application of ciclazindol (1 μM) to the intracellular surface of inside-out patches inhibited KATP channel currents activated by diazoxide (200 μM) by 86.6±8.1%.
  6. However, ciclazindol was much less effective at inhibiting KATP currents activated by leptin (10 nM). Ciclazindol (0.1–10 μM) had no effect on KATP currents activated by leptin, whereas higher concentrations (>10 μM) did cause inhibition with an IC50 value of 40 μM and an associated Hill coefficient of 2.7. Similarly, ciclazindol (1 μM) had no significant effect on KATP channel activity following leptin addition in excised inside-out patches.
  7. In conclusion, KATP currents activated by diazoxide and leptin show different sensitivity to englitazone and ciclazindol. This may be due to differences in the mechanism of activation of KATP channels by diazoxide and leptin.
  相似文献   

13.
  1. Interations were investigated between loreclezole, chlormethiazole and pentobarbitone as potentiators of depolarization responses mediated by γ-aminobutyric acidA (GABAA) receptors on afferent nerve terminals in the rat cuneate nucleus in vitro. These drugs were also compared as modulators of [3H]-flunitrazepam (FNZ) binding to synaptic membranes prepared from rat whole brain homogenate.
  2. In rat cuneate nucleus slices, the drugs shifted muscimol log dose–response lines to the left in an approximately parallel fashion with the result that 200 μM chlormethiazole potentiated muscimol responses by 0.567±0.037 log unit (mean±s.e.mean, n=4) while loreclezole gave a maximal potentiation at 10 μM of only 0.121±0.037 (n=6) log unit and 0.071±0.039 (n=22) at 50 μM.
  3. While 50 μM chlormethiazole and 30 μM pentobarbitone showed no significant interactions between each other when potentiating muscimol responses in combination, 50 μM loreclezole in combination with either chlormethiazole or pentobarbitone attenuated their potentiating effects, possibly by inducing desensitization of GABAA receptors.
  4. In the [3H]-FNZ binding studies on well-washed membranes, loreclezole enhanced binding to a maximum of 47.3±2.83% of control (mean±s.e.mean, n=3) at 300 μM. Scatchard analysis revealed no change in Bmax but a decrease in KD for [3H]-FNZ from 3.9±0.29 nM to 2.7±0.10 nM (mean±s.e.mean, n=4) in the presence of 100 μM loreclezole. In contrast, 100 μM chlormethiazole caused no potentiation. A small component of the enhancement by loreclezole could be blocked by 100 μM bicuculline and could also be blocked by 100 μM chlormethiazole. It seems likely that the effects on [3H]-FNZ binding are due predominantly to direct actions of the drugs on the GABAA receptor and are separate from the GABA-potentiating effects.
  5. The results indicate distinctly different profiles of action for loreclezole, chlormethiazole and pentobarbitone on GABAA receptors.
  相似文献   

14.
  1. Cumulative concentration-response curves (CRC) to prostaglandin E1 (PGE1), PGE2, PGD2 and PGF (0.01–30 μM) and to the thromboxane A2 (TXA2) receptor agonist U-46619 (0.01–30 μM) were constructed in human isolated detrusor muscle strips both in basal conditions and during electrical field stimulation.
  2. All the agonists tested contracted the detrusor muscle. The rank order of agonist potency was: PGF>U-46619>PGE2 whereas weak contractile responses were obtained with PGD2 and PGE1. Any of the agonists tested was able to induce a clear plateau of response even at 30 μM.
  3. The selective TXA2 antagonist, GR 32191B (vapiprost), antagonized U-46619-induced contractions with an apparent pKB value of 8.27±0.12 (n=4 for each antagonist concentration). GR 32191B (0.3 μM) did not antagonize the contractile responses to PGF and it was a non-surmountable antagonist of PGE2 (apparent pKB of 7.09±0.04; n=5). The EP receptor antagonist AH 6809 at 10 μM shifted to the right the CRC to U-46619 (apparent pKB value of 5.88±0.04; n=4).
  4. Electrical field stimulation (20 Hz, 70 V, pulse width 0.1 ms, trains of 5 s every 60 s) elicited contractions fully sensitive to TTX (0.3 μM) and atropine (1 μM). U-46619 (0.01–3 μM) potentiated the twitch contraction in a dose-dependent manner and this effect was competitively antagonized by GR 32191B with an estimated pKB of 8.54±0.14 (n=4 for each antagonist concentration). PGF in the range 0.01–10 μM (n=7), but not PGE2 and PGE1 (n=3 for each), also potentiated the twitch contraction of detrusor muscle strips (23.5±0.3% of KCl 100 mM-induced contraction) but this potentiation was unaffected by 0.3 μM GR 32191B (n=5).
  5. Cumulative additions of U-46619 (0.01–30 μM) were without effect on contractions induced by direct smooth muscle excitation (20 Hz, 40 V, 6 ms pulse width, trains of 2 s every 60 s, in the presence of TTX 1 μM; n=3). Moreover, pretreatment of the tissue with 0.3 μM U-46619 did not potentiate the smooth muscle response to 7 μM bethanecol (n=2).
  6. We concluded that TXA2 can induce direct contraction of human isolated urinary bladder through the classical TXA2 receptor. Prostanoid receptors, fully activated by PGE2 and PGF are also present. All these receptors are probably located post-junctionally. The rank order of agonist potency and the fact that GR32191B, but not AH6809, antagonized responses to PGE2 seem to indicate the presence of a new EP receptor subtype. Moreover, we suggest the presence of prejuctional TXA2 and FP receptors, potentiating acetylcholine release from cholinergic nerve terminals.
  相似文献   

15.
  1. γ-Aminobutyric acid (GABA) and trans-4-aminocrotonic acid (TACA) have been shown to activate GABAC receptors. In this study, a range of C2, C3, C4 and N-substituted GABA and TACA analogues were examined for activity at GABAC receptors.
  2. The effects of these compounds were examined by use of electrophysiological recording from Xenopus oocytes expressing the human ρ1 subunit of GABAC receptors with the two-electrode voltage-clamp method.
  3. trans-4-Amino-2-fluorobut-2-enoic acid was found to be a potent agonist (KD=2.43 μM). In contrast, trans-4-amino-2-methylbut-2-enoic acid was found to be a moderately potent antagonist (IC50=31.0 μM and KB=45.5 μM). These observations highlight the possibility that subtle structural substitutions may change an agonist into an antagonist.
  4. 4-Amino-2-methylbutanoic acid (KD=189 μM), 4-amino-2-methylenebutanoic acid (KD=182 μM) and 4-amino-2-chlorobutanoic acid (KD=285 μM) were weak partial agonists. The intrinsic activities of these compounds were 12.1%, 4.4% and 5.2% of the maximal response of GABA, respectively. These compounds more effectively blocked the effects of the agonist, GABA, giving rise to KB values of 53 μM and 101 μM, respectively.
  5. The sulphinic acid analogue of GABA, homohypotaurine, was found to be a potent partial agonist (KD=4.59 μM, intrinsic activity 69%).
  6. It was concluded that substitution of a methyl or a halo group in the C2 position of GABA or TACA is tolerated at GABAC receptors. However, there was dramatic loss of activity when these groups were substituted at the C3, C4 and nitrogen positions of GABA and TACA.
  7. Molecular modelling studies on a range of active and inactive compounds indicated that the agonist/competitive antagonist binding site of the GABAC receptor may be smaller than that of the GABAA and GABAB receptors. It is suggested that only compounds that can attain relatively flat conformations may bind to the GABAC receptor agonist/competitive antagonist binding site.
  相似文献   

16.
  1. The site(s) at which diadenosine 5′,5′′′-P1,P4-tetraphosphate (AP4A) and diadenosine 5′, 5′′′-P1,P5-pentaphosphate (AP5A) act to evoke contraction of the guinea-pig isolated vas deferens was studied by use of a series of P2-receptor antagonists and the ecto-ATPase inhibitor 6-N,N-diethyl-D-β,γ-dibromomethyleneATP (ARL 67156).
  2. Pyridoxalphosphate-6-azophenyl-2′,4′-disulphonic acid (PPADS) (300 nM–30 μM), suramin (3–100 μM) and pyridoxal-5′-phosphate (P-5-P) (3–1000 μM) inhibited contractions evoked by equi-effective concentrations of AP5A (3 μM), AP4A (30 μM) and α,β-methyleneATP (α,β-meATP) (1 μM), in a concentration-dependent manner and abolished them at the highest concentrations used.
  3. PPADS was more potent than suramin, which in turn was more potent than P-5-P. PPADS inhibited AP5A, AP4A and α,β-meATP with similar IC50 values. No significant difference was found between IC50 values for suramin against α,β-meATP and AP5A or α,β-meATP and AP4A, but suramin was more than 2.5 times more potent against AP4A than AP5A. P-5-P showed the same pattern of antagonism.
  4. Desensitization of the P2X1-receptor by α,β-meATP abolished contractions evoked by AP5A (3 μM) and AP4A (30 μM), but had no effect on those elicited by noradrenaline (100 μM).
  5. ARL 67156 (100 μM) reversibly potentiated contractions evoked by AP4A (30 μM) by 61%, but caused a small, significant decrease in the mean response to AP5A (3 μM).
  6. It is concluded that AP4A and AP5A act at the P2X1-receptor, or a site similar to the P2X1-receptor, to evoke contraction of the guinea-pig isolated vas deferens. Furthermore, the potency of AP4A, but not AP5A, appears to be inhibited by an ecto-enzyme which is sensitive to ARL 67156.
  相似文献   

17.
  1. A high density of receptors for somatostatin (SRIF) exists in the anterior cingulate cortex but their function is unknown. Whole-cell patch clamp recordings were made from visualized deep layer pyramidal cells of the rat anterior cingulate cortex contained in isolated brain slices to investigate the putative effects of SRIF and to identify the receptor subtype(s) involved.
  2. SRIF (1–1000 nM) produced a concentration-dependent outward current which was associated with an increased membrane conductance, was sensitive to Ba2+ (300 μM–1 mM), and was absent in the presence of a maximal concentration of the GABAB receptor agonist, baclofen (100 μM). These observations suggest the outward current was carried by K+ ions.
  3. SRIF analogues also elicited outward currents with a rank potency order of (EC50, nM): octreotide (1.8)>BIM-23027 (3.7)>SRIF (20)=L-362,855 (20). BIM-23056 was without agonist or antagonist activity. Responses to L-362,855 were unlike those to the other agonists since they were sustained for the duration of the application.
  4. The sst2 receptor antagonist, L-Tyr8Cyanamid 154806 (1 μM), had no effect alone but partially reversed responses to submaximal concentrations of SRIF (100 nM, 44±6% reversal) and L-362,855 (100 nM, 70±6% reversal) and fully reversed the response to BIM-23027 (10 nM). In contrast, L-Tyr8Cyanamid 154806 did not antagonize the response to baclofen (10 μM).
  5. We conclude that SRIF activates a K+ conductance in anterior cingulate pyramidal neurones via an action predominantly at sst2 receptors.
  相似文献   

18.
  1. The β-adrenoceptor agonist, isoprenaline, inhibited the IgE-mediated release of histamine from human lung mast cells (HLMC) in a dose-dependent manner. Maximal inhibitory effects were obtained with 0.1 μM isoprenaline. However, the inhibition of histamine release from HLMC by isoprenaline (0.1 μM) was highly variable ranging from 33 to 97% inhibition (mean, 59±3%, n=27).
  2. Long-term (24 h) incubation of HLMC with isoprenaline led to a subsequent reduction in the ability of a second exposure of isoprenaline to inhibit IgE-mediated histamine release from HLMC. The impairment in the ability of isoprenaline (0.1 μM) to inhibit histamine release following desensitizing conditions (1 μM isoprenaline for 24 h) was highly variable amongst HLMC preparations ranging from essentially negligible levels of desensitization in some preparations to complete abrogation of the inhibitory response in others (mean, 65±6% desensitization, n=27).
  3. The ability of HLMC to recover from desensitization was investigated. Following desensitizing conditions (1 μM isoprenaline for 24 h), HLMC were washed and incubated for 24 h in buffer and the effectiveness of isoprenaline (0.1 μM) to inhibit IgE-mediated histamine release from HLMC was assessed. The extent of recovery was highly variable with some HLMC preparations failing to recover and others displaying a complete restoration of responsiveness to isoprenaline (mean, 40±6% recovery, n=23).
  4. The effects of the glucocorticoid, dexamethasone, were also investigated. Long-term (24–72 h) treatments with dexamethasone (0.1 μM) had no effect on IgE-mediated histamine release from HLMC. Additionally, long-term (24–72 h) treatments with dexamethasone (0.1 μM) had no effect on the effectiveness of isoprenaline to inhibit histamine release. However, long-term (24–72 h) treatments with dexamethasone (0.1 μM) protected against the functional desensitization induced by incubation (24 h) of HLMC with isoprenaline (1 μM). The protective effect was time-dependent and pretreatment of HLMC with dexamethasone for either 24, 48 or 72 h prevented desensitization by either 15±7, 19±5 or 51±10%, respectively (n=5–7).
  5. HLMC preparations which were relatively refractory to isoprenaline even after withdrawal (24 h) from desensitizing conditions responded more effectively to isoprenaline (0.1 μM) if dexamethasone (0.1 μM) was also included during the recovery period (19±9% recovery after 24 h in buffer; 50±8% recovery after 24 h with dexamethasone, n=5).
  6. These data indicate that the responses of different HLMC preparations to isoprenaline, the susceptibility of HLMC to desensitization and the ability of HLMC to recover from desensitizing conditions varies markedly. Dexamethasone, which itself has no direct effects on IgE-mediated histamine release from HLMC, protected HLMC from the functional desensitization to β-adrenoceptor agonists. Because β2-adrenoceptor agonists and glucocorticoids are important in the therapeutic management of asthma and as the HLMC is probably important in certain types of asthma, these findings may have wider clinical implications.
  相似文献   

19.
  1. γ-Aminobutyric acid (GABA), acting via the GABAA receptor, increased the extracellular acidification rate of rat primary cultured cerebellar granule cells, measured by the Cytosensor microphysiometer.
  2. The optimal conditions for the measurement of GABAA receptor function in cerebellar granule cells by microphysiometry were: cells seeded at 9–12×105 cells/transwell cup and maintained in vitro for 8 days, GABA stimulation performed at 25°C, with a stimulation time of 33 s.
  3. GABA stimulated a concentration-dependent increase in the extracellular acidification rate with an EC50 of 2.0±0.2 μM (mean±s.e.mean, n=7 experiments) and maximal increase (Emax) over basal response of 15.4±1.2%.
  4. The sub-maximal GABA-stimulated increase in acidification rate could be potentiated by the 1,4-benzodiazepine, flunitrazepam (100 nM). The 10 nM GABA response showed the maximal benzodiazepine facilitation (GABA alone, 1.4 μV s−1, GABA+flunitrazepam, 3.8 μV s−1, mean increment over basal, n=7).
  5. The GABA-stimulated increase in acidification rate was inhibited by the GABAA antagonist, bicuculline (100 μM) (90% inhibition at 1 mM GABA).
  6. The results of this study show that activation of GABAA receptors in rat cerebellar granule cells caused an increase in the extracellular acidification rate; an effect which was potentiated by benzodiazepines and inhibited by a GABAA receptor antagonist. This paper defines the conditions and confirms the feasibility of using microphysiometry to investigate GABAA receptor function in primary cultured CNS neurones. The microphysiometer provides a rapid and sensitive technique to investigate the regulation of the GABAA receptor in populations of neurones.
  相似文献   

20.
  1. We investigated the inhibitory effects of the cysteinyl leukotriene (CysLT1) receptor antagonists, pranlukast and zafirlukast, on 35SO4 labelled mucus output, in vitro, in guinea-pig trachea, induced by leukotriene D4 (LTD4) or by antigen challenge of sensitized animals. Agonists and antagonists were administered mucosally, except in selected comparative experiments where drugs were administered both mucosally and serosally to assess the influence of the epithelium on evoked-secretion.
  2. LTD4 increased 35SO4 output in a concentration-related manner with a maximal increase of 23 fold above controls at 100 μM and an approximate EC50 of 2 μM. Combined mucosal and serosal addition of LTD4 did not significantly affect the secretory response compared with mucosal addition alone. Neither LTC4 nor LTE4 (10 μM each) affected 35SO4 output. Pranlukast or zafirlukast significantly inhibited 10 μM LTD4-evoked 35SO4 output in a concentration-dependent fashion, with maximal inhibitions of 83% at 10 μM pranlukast and 78% at 10 μM zafirlukast, and IC50 values of 0.3 μM for pranlukast and 0.6 μM for zafirlukast. Combined mucosal and serosal administration of the antagonists (5 μM each) gave degrees of inhibition of mucosal-serosal 10 μM LTD4-evoked 35SO4 output similar to those of the drugs given mucosally. Pranlukast (0.5 μM) caused a parallel rightward shift of the LTD4 concentration-response curve with a pKB of 7. Pranlukast did not inhibit ATP-induced 35SO4 output.
  3. Ovalbumin (10–500 μg ml−1) challenge of tracheae from guinea-pigs actively sensitized with ovalbumin caused a concentration-related increase in 35SO4 output with a maximal increase of 20 fold above vehicle controls at 200 μg ml−1. The combination of the antihistamines pyrilamine and cimetidine (0.1 mM each) did not inhibit ovalbumin-induced 35SO4 output in sensitized guinea-pigs. Neither mucosal (10 μM or 100 μM) nor mucosal-serosal (100 μM) histamine had any significant effect on 35SO4 output.
  4. Pranlukast or zafirlukast (5 μM each) significantly suppressed ovalbumin-induced secretion in tracheae from sensitized guinea-pigs by 70% and 65%, respectively.
  5. We conclude that LTD4 or ovalbumin challenge of sensitized animals provokes mucus secretion from guinea-pig trachea in vitro and this effect is inhibited by the CysLT1 receptor antagonists pranlukast and zafirlukast. These antagonists may be beneficial in the treatment of allergic airway diseases in which mucus hypersecretion is a clinical symptom, for example asthma and allergic rhinitis.
  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