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 46 毫秒
1.
  1. Multiple components of hippocampal glutamate release were examined by study of Ca2+- and K+-evoked hippocampal extracellular glutamate release using an in vivo microdialysis glutamate biosensor in urethane-anaesthetized rats. In addition, the effects of the antiepileptic drugs, carbamazepine (CBZ) and zonisamide (ZNS) perfused through the probe on glutamate release were assessed.
  2. Basal glutamate levels were below detection limits (∼0.1 μM). An increase in extracellular KCl (from 2.7 to 50 and 100 mM) increased extracellular hippocampal glutamate levels to 9.2±1.4 and 20.0±2.6 μM, respectively, calculated from the area under curve (AUC) for 60 min.
  3. This KCl-evoked glutamate release consisted of three components: an initial transient rise, a late gentle rise, and late multiple phasic transient rises.
  4. An increase in or removal of extracellular CaCl2 levels respectively enhanced and reduced the 50 mM KCl-evoked hippocampal glutamate release (AUC for 60 min) from 9.2±1.4 to 12.4±2.1 and 5.8±0.9 μM.
  5. Perfusion with 100 μM CBZ or 1 mM ZNS inhibited both the 50 mM KCl-evoked hippocampal glutamate release (AUC for 60 min) from 9.2±1.4 to 5.5±1.1 and to 5.8±1.3 μM, respectively, as well as the stimulatory effects of Ca2+ on KCl-evoked hippocampal glutamate release.
  6. These results suggest that both CBZ and ZNS may reduce epileptiform events by inhibiting excitatory glutamatergic transmission.
  相似文献   

2.
  1. Intracellular recordings were made in a pontine slice preparation of the rat brain containing the nucleus locus coeruleus (LC). The pressure application of α,β-methylene ATP (α,β-meATP) caused reproducible depolarizations which were depressed by suramin (30 μM) and abolished by suramin (100 μM). Pyridoxal-phosphate-6-azophenyl-2′,4′-disulphonic acid (PPADS; 10, 30 μM) also concentration-dependently inhibited the α,β-meATP-induced depolarization, although with a much slower time-course than suramin. Almost complete inhibition developed with 30 μM PPADS. Reactive blue 2 (30 μM) did not alter the effect of α,β-meATP, while reactive blue 2 (100 μM) slightly depressed it.
  2. Pressure-applied (S)-α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) also depolarized LC neurones. Kynurenic acid (500 μM) depressed and 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX; 50 μM) abolished the response to AMPA. Suramin (100 μM) potentiated the AMPA effect.
  3. Pressure-applied noradrenaline hyperpolarized LC neurones. Suramin (100 μM) did not alter the effect of noradrenaline.
  4. Focal electrical stimulation evoked biphasic synaptic potentials consisting of a fast depolarization (p.s.p.) followed by a slow hyperpolarization (i.p.s.p.). A mixture of D(−)-2-amino-5-phosphonopentanoic acid (AP-5; 50 μM), CNQX (50 μM) and picrotoxin (100 μM) depressed both the p.s.p. and the i.p.s.p. Under these conditions suramin (100 μM) markedly inhibited the p.s.p., but did not alter the i.p.s.p. In the combined presence of AP-5 (50 μM), CNQX (50 μM), picrotoxin (100 μM), strychnine (0.1 μM), tropisetron (0.5 μM) and hexamethonium (100 μM), a high concentration of suramin (300 μM) almost abolished the p.s.p. without changing the i.p.s.p.
  5. In the presence of kynurenic acid (500 μM) and picrotoxin (100 μM), PPADS (30 μM) depressed the p.s.p. Moreover, the application of suramin (100 μM) to the PPADS (30 μM)-containing medium failed to cause any further inhibition. Neither PPADS (30 μM) nor suramin (100 μM) altered the i.p.s.p.
  6. It was concluded that the cell somata of LC neurones are endowed with excitatory P2-purinoceptors. ATP may be released either as the sole transmitter from purinergic neurones terminating at the LC or as a co-transmitter of noradrenaline from recurrent axon collaterals or dendrites of the LC neurones themselves.
  相似文献   

3.
  1. Whole-cell patch-clamp recordings in adult mouse hippocampal slices were used to test the mechanism by which the volatile anesthetic halothane inhibits glutamate receptor-mediated synaptic transmission. Non-N-methyl-D-aspartate (nonNMDA) and NMDA receptor-mediated currents in CA1 pyramidal cells were pharmacologically isolated by bath application of D,L-2-amino-5-phosphonovaleric acid (APV; 100 μM) or 6-cyano-7-nitro-quinoxaline-2,3-dione (CNQX; 5 μM), respectively.
  2. Halothane blocked both nonNMDA and NMDA receptor-mediated excitatory postsynaptic currents (EPSCs) to a similar extent (IC50 values of 0.66 and 0.57 mM, respectively).
  3. Partial blockade of the EPSCs by lowering the extracellular concentration of calcium ([Ca2+]o), but not by application of CNQX (1 μM), was accompanied by an increase in paired-pulse facilitation (PPF). Halothane-induced blockade of the EPSCs also was associated with an increase in PPF.
  4. The effects of halothane on α-amino-3-hydroxy-5-methylisoxazole-4-propionic acid (AMPA) and NMDA receptor-mediated currents induced by agonist iontophoresis, were compared. AMPA-induced currents were blocked with an IC50 of 1.7 mM. NMDA-induced currents were significantly less sensitive to halothane (IC50 of 5.9 mM).
  5. The effect of halothane on iontophoretic AMPA dose-response curves was tested. Halothane suppressed the maximal response to AMPA without affecting its EC50, suggesting a noncompetitive mechanism of inhibition.
  6. All effects of halothane were reversible upon termination of the exposure to the drug.
  7. These data suggest that halothane blocks central glutamatergic synaptic transmission by presynaptically inhibiting glutamate release and postsynaptically blocking the AMPA subtype of glutamate receptors.
  相似文献   

4.
  1. Nitric oxide (NO) synthase activity was studied in slices of human temporal cortex samples obtained in neurosurgery by measuring the conversion of L-[3H]-arginine to L-[3H]-citrulline.
  2. Elevation of extracellular K+ to 20, 35 or 60 mM concentration-dependently augmented L-[3H]-citrulline production. The response to 35 mM KCl was abolished by NG-nitro-L-arginine (100 μM) demonstrating NO synthase specific conversion of L-arginine to L-citrulline. Increasing extracellular MgCl2 concentration up to 10 mM also prevented the K+ (35 mM)-induced NO synthase activation, suggesting the absolute requirement of external calcium ions for enzyme activity.
  3. However, the effect of high K+ (35 mM) on citrulline synthesis was insensitive to the antagonists of ionotropic and metabotropic glutamate receptors dizocilpine (MK-801), 6-nitro-7-sulphamoylbenzo(f)quinoxaline-2-3-dione (NBQX) or L-2-amino-3-phosphonopropionic acid (L-AP3) as well as to the nicotinic receptor antagonist, mecamylamine.
  4. The 35 mM K+ response was insensitive to ω-conotoxin GVIA (1 μM) and nifedipine (100 μM), but could be prevented in part by ω-agatoxin IVA (0.1 and 1 μM). The inhibition caused by 0.1 μM ω-agatoxin IVA (∼30%) was enhanced by adding ω-conotoxin GVIA (1 μM) or nifedipine (100 μM). Further inhibition (up to above 70%) could be observed when the three Ca2+ channel blockers were added together. Similarly, synthetic FTX 3.3 arginine polyamine (sFTX) prevented (50% at 100 μM) the K+-evoked NO synthase activation. This effect of sFTX was further enhanced (up to 70%) by adding 1 μM ω-conotoxin GVIA plus 100 μM nifedipine. No further inhibition could be observed upon addition of MK-801 or/and NBQX.
  5. It was concluded that elevation of extracellular [K+] causes NO synthase activation by external Ca+ entering cells mainly through channels of the P/Q-type. Other Ca2+ channels (L- and N-type) appear to contribute when P/Q-channels are blocked.
  相似文献   

5.
  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   

6.
  1. Nicotine-induced relaxation and release of vasoactive intestinal polypeptide (VIP)- and peptide histidine isoleucine (PHI)-like immunoreactivity (LI) were measured in longitudinal muscle strips from the rat gastric fundus.
  2. Under non-cholinergic conditions (0.3 μM atropine), nicotine (3–300 μM) produced concentration-dependent relaxations of the 5-hydroxytryptamine (3 μM)-precontracted strips. Under non-adrenergic non-cholinergic (NANC) conditions (0.3 μM atropine+1 μM phentolamine+1 μM nadolol), relaxations induced by sub-maximal nicotine concentrations (10 and 30 μM) were significantly smaller, while that produced by the highest concentration used (300 μM) was similar to that seen under non-cholinergic conditions.
  3. Re-exposure to the same nicotine concentration 1 h later induced smaller relaxations, indicating desensitization. The reductions seen in the second responses were proportional to the concentration used.
  4. Under non-cholinergic conditions, the relaxant response to 30 μM nicotine was abolished by hexamethonium (100 μM) and significantly reduced by tetrodotoxin (TTX, 3 μM). The TTX-resistant component was not observed under NANC conditions.
  5. NANC relaxation induced by 30 μM nicotine was significantly reduced by a specific anti-VIP serum (approximately 35% less than that seen with normal rabbit serum).
  6. Nicotine (30–300 μM) caused significant, concentration-dependent increases in the outflow of VIP- and PHI-LI from the strips; these effects were also diminished with re-exposure. The increases in both types of immunoreactivity evoked by nicotine (300 μM) were abolished by hexamethonium (300 μM), TTX (3 μM) and a calcium-free medium.
  7. These findings indicate that VIP and possibly PHI are involved in NANC relaxation of the rat gastric fundus induced by nicotine.
  相似文献   

7.
  1. Nucleotide-induced currents in untreated (proliferating) and lipopolysaccharide (LPS; 100 ng ml−1) treated (non-proliferating) rat microglial cells were recorded by the whole-cell patch-clamp technique. Most experiments were carried out on non-proliferating microglial cells. ATP (100 nM–1 mM), ADP (10 nM–10 mM) and UTP (1 μM–100 mM), but not uridine (100 μM–10 mM) produced a slow outward current at a holding potential of 0 mV. The effect of UTP (1 mM) did not depend on the presence of extracellular Mg2+ (1 mM). The outward current response to UTP (1 mM) was similar in non-proliferating and proliferating microglia.
  2. In non-proliferating microglial cells, the ATP (10 μM)-induced outward current was antagonized by suramin (300 μM) or reactive blue 2 (50 μM), whereas 8-(p-sulphophenyl)-theophylline (8-SPT; 100 μM) was inactive. By contrast, the current induced by UTP (1 mM) was increased by suramin (300 μM) and was not altered by reactive blue 2 (50 μM) or 8-SPT (100 μM).
  3. The current response to UTP (1 mM) disappeared when K+ was replaced in the pipette solution by an equimolar concentration of Cs+ (150 mM). However, the effect of UTP (1 mM) did not change when most Cl was replaced with an equimolar concentration of gluconate (145 mM). The application of 4-aminopyridine (1 mM) or Cs+ (1 mM) to the bath solution failed to alter the UTP (1 mM)-induced current. UTP (1 mM) had almost no effect in a nominally Ca2+-free bath medium, or in the presence of charybdotoxin (0.1 μM); the inclusion of U-73122 (5 μM) or heparin (5 mg ml−1) into the pipette solution also blocked the responses to UTP (1 mM). By contrast, the effect of ATP (10 μM) persisted under these conditions.
  4. I-V relations were determined by delivering fast voltage ramps before and during the application of UTP (1 mM). In the presence of extracellular Cs+ (1 mM) and 4-aminopyridine (1 mM) the UTP-evoked current crossed the zero current level near−75 mV. Omission of Ca2+ from the Cs+ (1 mM)- and 4-aminopyridine (1 mM)-containing bath medium or replacement of K+ by Cs+ (150 mM) in the pipette solution abolished the UTP current.
  5. Replacement of GTP (200 μM) by GDP-β-S (200 μM) in the pipette solution abolished the current evoked by UTP (1 mM).
  6. When the pipette solution contained Cs+ (150 mM) instead of K+ and in addition inositol 1,4,5,-trisphosphate (InsP3; 10 μM), an inward current absolutely dependent on extracellular Ca2+ was activated after the establishment of whole-cell recording conditions. This current had a typical delay, a rather slow time course and did not reverse its amplitude up to 100 mV, as measured by fast voltage ramps.
  7. A rise of the internal free Ca2+ concentration from 0.01 to 0.5 μM on excised inside-out membrane patches produced single channel activity with a reversal potential of 0 mV in a symmetrical K+ solution. The reversal potential was shifted to negative values, when the extracellular K+ concentration was decreased from 144 to 32 mM. By contrast, a decrease of the extracellular Cl concentration from 164 to 38 mM did not change the reversal potential.
  8. Purine and pyrimidine nucleotides act at separate receptors in rat microglial cells. Pyrimidinoceptors activate via a G protein the enzyme phospholipase C with the subsequent release of InsP3. The depletion of the intracellular Ca2+ pool appears to initiate a capacitative entry of Ca+ from the extracellular space. This Ca2+ then activates a Ca2+-dependent K+ current.
  相似文献   

8.
  1. The effects of nifedipine on both levcromakalim-induced membrane currents and unitary currents in pig proximal urethra were investigated by use of patch-clamp techniques (conventional whole-cell configuration and cell-attached patches).
  2. Nifedipine had a voltage-dependent inhibitory effect on voltage-dependent Ba2+ currents at −50 mV (Ki=30.6 nM).
  3. In current-clamp mode, subsequent application of higher concentrations of nifedipine (⩾30 μM) caused a significant depolarization even after the membrane potential had been hyperpolarized to approximately −82 mV by application of 100 μM levcromakalim.
  4. The 100 μM levcromakalim-induced inward current (symmetrical 140 mM K+ conditions, −50 mV) was inhibited by additional application of three different types of Ca antagonists (nifedipine, verapamil and diltiazem, all at 100 μM). In contrast, Bay K 8644 (1 μM) possessed no activating effect on the amplitude of this glibenclamide-sensitive current.
  5. When 100 μM nifedipine was included in the pipette solution during conventional whole-cell recording at −50 mV, application of levcromakalim (100 μM) caused a significant inward membrane current which was suppressed by 5 μM glibenclamide. On the other hand, inclusion of 5 μM glibenclamide in the pipette solution prevented levcromakalim from inducing an inward membrane current.
  6. The levcromakalim-induced K+ channel openings in cell-attached configuration were suppressed by subsequent application of 5 μM glibenclamide but not of 100 μM nifedipine.
  7. These results suggest that in pig proximal urethra, nifedipine inhibits the glibenclamide-sensitive 43 pS K+ channel activity mainly through extracellular blocking actions on the K+ channel itself.
  相似文献   

9.
  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   

10.
  1. Glutamate and other amino acids are the main excitatory neurotransmitters in many brain regions, including the hippocampus, by activating ion channel-coupled glutamate receptors, as well as metabotropic receptors linked to G proteins and second messenger systems. Several conditions which promote the release of glutamate, like frequency stimulation and hypoxia, also lead to an increase in the extracellular levels of the important neuromodulator, adenosine. We studied whether the activation of different subgroups of metabotropic glutamate receptors (mGluR) could modify the known inhibitory effects of a selective adenosine A1 receptor agonist on synaptic transmission in the hippocampus. The experiments were performed on hippocampal slices taken from young (12–14 days old) rats. Stimulation was delivered to the Schaffer collateral/commissural fibres, and evoked field excitatory postsynaptic potentials (fe.p.s.p.) recorded extracellularly from the stratum radiatum in the CA1 area.
  2. The concentration-response curve for the inhibitory effects of the selective adenosine A1 receptor agonist, N6-cyclopentyladenosine (CPA; 2–50 nM), on the fe.p.s.p. slope (EC50=12.5 (9.2–17.3; 95% confidence intervals)) was displaced to the right by the group I mGluR selective agonist, (R,S)-3,5-dihydroxyphenylglycine (DPHG; 10 μM) (EC50=27.2 (21.4–34.5) nM, n=4). The attenuation of the inhibitory effect of CPA (10 nM) on the fe.p.s.p. slope by DHPG (10 μM) was blocked in the presence of the mGluR antagonist (which blocks group I and II mGluR), (R,S)-α-methyl-4-carboxyphenylglycine (MCPG; 500 μM). DHPG (10 μM) itself had an inhibitory effect of 20.1±1.9% (n=4) on the fe.p.s.p. slope.
  3. The concentration-response curves for the inhibitory effects of CPA (2–20 nM) on the fe.p.s.p. slope were not modified either in the presence of the group II mGluR selective agonist, (2S,3S,4S)-α-(carboxycyclopropyl)glycine (L-CCG-I; 1 μM), or in the presence of the non-selective mGluR agonist (which activates both group I and II mGluR), (1S,3R)-1-aminocyclopentyl-1,3-dicarboxylate (ACPD; 100 μM). L-CCG-I had no consistent effects and ACPD (100 μM) decreased by 19.4±1.8% (n=4) the fe.p.s.p. slope.
  4. The concentration-response curve for the inhibitory effects of CPA (2–100 nM) on the fe.p.s.p. slope (EC50=8.2 (6.9–9.6) nM) was displaced to the right by the group III mGluR selective agonist, L-2-amino-4-phosphonobutyrate (L-AP4; 25 μM) (EC50=17.7 (13.1–21.9) nM, n=4). The attenuation of the inhibitory effect of CPA (10 nM) on the fe.p.s.p. slope by L-AP4 (25 μM) was blocked in the presence of the mGluR antagonist (selective for the group III mGluR), (R,S)-α-methyl-4-phosphonophenylglycine (MPPG; 200 μM).
  5. Both the direct effect of DHPG on synaptic transmission and the attenuation of the inhibitory effect of CPA (10 nM) were prevented in the presence of the protein kinase C selective inhibitors, staurosporine (1 μM) or chelerythrine (5 μM), and thus attributed to activation of protein kinase C.
  6. The attenuation by L-AP4 (25 μM) of the inhibitory effect of CPA (10 nM) on the fe.p.s.p. slope was also prevented by the protein kinase C selective inhibitors, staurosporine (1 μM) or chelerythrine (5 μM), and thus attributed to activation of protein kinase C. But this effect seemed to be distinct from the direct effect of L-AP4 (25 μM) on synaptic transmission, which was not modified by the protein kinase C selective inhibitors.
  7. We conclude that agonists of metabotropic glutamate receptors (Groups I and III) are able to attenuate the inhibitory effects of adenosine A1 receptor activation in the hippocampus. This interaction may have pathophysiological relevance in hypoxia, in which there is marked release of both excitatory amino acids and the important endogenous neuroprotective substance, adenosine.
  相似文献   

11.
  1. The pharmacological features of the pre- and postsynaptic metabotropic glutamate receptors (mGluRs) present in the guinea-pig olfactory cortex, were examined in brain slices in vitro by use of a conventional intracellular current clamp/voltage clamp recording technique.
  2. Bath-application of trans-aminocyclopentane-1,3-dicarboxylic acid (trans-ACPD) (50 μM) produced a sustained membrane depolarization, increase in cell excitability and induction of a post-stimulus inward (afterdepolarizing) tail current (IADP) (measured under ‘hybrid'' voltage clamp) similar to those evoked by the muscarinic receptor agonist oxotremorine-M (OXO-M, 2 μM).
  3. L-Glutamate (0.25–1 mM, in the presence of 20 μM 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX) and 100 μM DL-amino-5-phosphono valeric acid (DL-APV)) or the broad spectrum mGluR agonists 1S,3R-aminocyclopentane-1,3-dicarboxylic acid (1S,3R-ACPD, 10 μM), 1S,3S-ACPD (50 μM), ibotenate (Ibo; 25 μM, in the presence of 100 μM DL-APV), the selective mGluR I agonists (S)-3,5-dihydroxyphenylglycine ((S)-3,5-DHPG, 10 μM), (S)-3-hydroxyphenylglycine ((S)-3HPG, 50 μM), or quisqualate (10 μM, in the presence of 20 μM CNQX), but not the mGluR II agonist 2S,1′S,2′S-2-(2′-carboxycyclopropyl)-glycine (L-CCGI, 1 μM) or mGluR III agonist L(+)-2-amino-4-phosphonobutyric acid (L-AP4, 1 mM), were all effective in producing membrane depolarization and inducing a post-stimulus IADP. Unexpectedly, the proposed mGluR II-selective agonist (2S,1′R,2′R,3′R)-2-(2′,3′-dicarboxycyclopropyl)-glycine (DCG-IV, 10 μM, in the presence of 100 μM DL-APV) was also active.
  4. The excitatory effects induced by 10 μM 1S,3R-ACPD were reversibly antagonized by the mGluR I/II antagonist (+)-α-methyl-4-carboxyphenylglycine ((+)-MCPG, 0.5–1 mM), as well as the selective mGluR I antagonists (S)-4-carboxyphenylglycine ((S)-4CPG) and (S)-4-carboxy-3-hydroxyphenyl glycine ((S)-4C3HPG) (both at 1 mM), but not the nonselective mGluR antagonist L(+)-2-amino-3-phosphonopropionic acid (L-AP3, 1 mM) or the selective mGluR III antagonist (S)-α-methyl-L-AP4 (MAP4, 1 mM).
  5. The excitatory postsynaptic potentials (e.p.s.ps), induced by single focal stimulation of cortical excitatory fibre tracts, were markedly reduced by 1S,3R-ACPD or L-AP4 (both at 10 μM), and by the selective mGluR II agonists (mGluR I antagonists) (S)-4CPG or (S)-4C3HPG (both at 1 mM) but not (S)-3,5-DHPG or (S)-3HPG (both at 100 μM).
  6. The inhibitory effects of 1S-3R-ACPD, but not L-AP4, were reversibly blocked by (+)-MCPG (1 mM), whereas those produced by L-AP4, but not 1S,3R-ACPD, were blocked by the selective mGluR III antagonist MAP4 (1 mM).
  7. It is concluded that a group I mGluR is most likely involved in mediating excitatory postsynaptic effects, whereas two distinct mGluRs (e.g. group II and III) might serve as presynaptic inhibitory autoreceptors in the guinea-pig olfactory cortex.
  相似文献   

12.
  1. Inhalation of vanadium compounds, particularly vanadate, is a cause of occupational bronchial asthma. We have now studied the action of vanadate on human isolated bronchus. Vanadate (0.1 μM–3 mM) produced concentration-dependent, well-sustained contraction. Its −logEC50 was 3.74±0.05 (mean±s.e.mean) and its maximal effect was equivalent to 97.5±4.2% of the response to acetylcholine (ACh, 1 mM).
  2. Vanadate (200 μM)-induced contraction of human bronchus was epithelium-independent and was not inhibited by indomethacin (2.8 μM), zileuton (10 μM), a mixture of atropine, mepyramine and phentolamine (each at 1 μM), or by mast cell degranulation with compound 48/80.
  3. Vanadate (200 μM)-induced contraction was unaltered by tissue exposure to verapamil or nifedipine (each 1 μM) or to a Ca2+-free, EGTA (0.1 mM)-containing physiological salt solution (PSS). However, tissue incubation with ryanodine (10 μM) in Ca2+-free, EGTA (0.1 mM)-containing PSS reduced vanadate-induced contraction. A series of vanadate challenges was made in tissues exposed to Ca2+-free EGTA (0.1 mM)-containing PSS with the object of depleting intracellular Ca2+ stores. In such tissues cyclopiazonic acid (CPA; 10 μM) prevented Ca2+-induced recovery of vanadate-induced contraction.
  4. Tissue incubation in K+-rich (80 mM) PSS, K+-free PSS, or PSS containing ouabain (10 μM) did not alter vanadate (200 μM)-induced contraction. Ouabain (10 μM) abolished the K+-induced relaxation of human bronchus bathed in K+-free PSS. This action was not shared by vanadate (200 μM). The tissue content of Na+ was increased and the tissue content of K+ was decreased by ouabain (10 μM). In contrast, vanadate (200 μM) did not alter the tissue content of these ions. Tissue incubation in a Na+-deficient (25 mM) PSS or in PSS containing amiloride (0.1 mM) markedly inhibited the spasmogenic effect of vanadate (200 μM).
  5. Vanadate (200 μM)-induced contractions were markedly reduced by tissue treatment with each of the protein kinase C (PKC) inhibitors H-7 (10 μM), staurosporine (1 μM) and calphostin C (1 μM). Genistein (100 μM), an inhibitor of protein tyrosine kinase, also reduced the response to vanadate.
  6. Vanadate (0.1–3 mM) and ACh (1 μM–3 mM) each increased inositol phosphate accumulation in bronchus. Such responses were unaffected by a Ca2+-free medium either alone or in combination with ryanodine (10 μM).
  7. In human cultured tracheal smooth muscle cells, histamine (100 μM) and vanadate (200 μM) each produced a transient increase in intracellular Ca2+ concentration ([Ca2+]i).
  8. Intracellular microelectrode recording showed that the contractile effect of vanadate (200 μM) in human bronchus was associated with cellular depolarization.
  9. It is concluded that vanadate acts directly on human bronchial smooth muscle, promoting the release of Ca2+ from an intracellular store. The Ca2+ release mechanism involves both the production of inositol phosphate second messengers and inhibition of Ca-ATPase. The activation of PKC plays an important role in mediating vanadate-induced contraction at values of [Ca2+]i that are close to basal.
  相似文献   

13.
  1. This study examined the effects of sodium rhein (0.03–30 μM) on the contractions of the isolated circular muscle of guinea-pig ileum induced by acetylcholine (100 nM), substance P (3 nM) and electrical stimulation (10 Hz for 0.3 s, 100 mA, 0.5 ms pulse duration). The effect of sodium rhein was also evaluated on the ascending excitatory reflex using a partitioned bath (oral and anal compartments). Ascending excitatory enteric nerve pathways were activated by electrical field stimulation (10 Hz for 2 s, 20 mA, 0.5 pulse duration) in the anal compartment and the resulting contraction of the guinea-pig intestinal circular muscle in the oral compartment was recorded.
  2. Sodium rhein (0.3, 3 and 30 μM) significantly potentiated (52±11% at 30 μM) acetylcholine-induced contractions. In the presence of tetrodotoxin (0.6 μM) or ω-conotoxin GVIA (10 nM) sodium rhein (3 and 30 μM) did not enhance, but significantly reduced (49±10% and 44±8%, respectively, at 30 μM) acetylcholine-induced contractions.
  3. Sodium rhein (0.3, 3 and 30 μM) significantly increased (65±11% at 30 μM) substance P-induced contractions. In the presence of tetrodotoxin (0.6 μM), ω-conotoxin GVIA (10 nM) or atropine (0.1 μM), sodium rhein (3 and 30 μM) significantly reduced (50±10%, 55±8% and 46±10%, respectively, at 30 μM) substance P-induced contractions.
  4. NG-nitro-L-arginine methyl ester (L-NAME, 100 μM) abolished the potentiating effect of sodium rhein on acetylcholine and substance P-induced contractions. At the highest concentration (30 μM), sodium rhein, in presence of L-NAME, reduced the acetylcholine (30±6%)- or substance P (36±6%)-induced contractions.
  5. Sodium rhein (30 μM) significantly potentiated (29±9%) the electrically-evoked contractions. L-NAME (100 μM), but not phentolamine, enhanced the effect of sodium rhein. Sodium rhein (30 μM) significantly increased (32±9%) the ascending excitatory reflex when applied in the oral, but not in the anal compartment.
  6. These results indicate that sodium rhein (i) activates excitatory cholinergic nerves on circular smooth muscle presumably through a facilitation of Ca2+ entry through the N-type Ca2+ channel, (ii) has a direct inhibitory effect on circular smooth muscle and (iii) does not affect enteric ascending neuroneural transmission. Nitric oxide could have a modulatory excitatory role on sodium rhein-induced changes of agonist-induced contractions and an inhibitory modulator role on sodium rhein-induced changes of electrically-induced contractions.
  相似文献   

14.
  1. Acetylcholine (ACh) elicits an endothelium-dependent relaxation and hyperpolarization in the absence of nitric oxide (NO) and prostaglandin synthesis in the guinea-pig coronary artery (GPCA). This response has been attributed to a factor termed endothelial-derived hyperpolarizing factor (EDHF). Recently it has been suggested that EDHF may be a cytochrome P450 product of arachidonic acid (AA) i.e., an epoxyeicosatrienoic acid (EET). The present study investigated whether this pathway could account for the response to ACh observed in the GPCA in the presence of 100 μM Nω-nitro-L-arginine and 10 μM indomethacin.
  2. ACh, AA and 11,12-EET each produced concentration-dependent relaxations in arteries contracted with the H1-receptor agonist AEP (2,2-aminoethylpyridine). The AA-induced relaxation was significantly enhanced in the presence of the cyclo-oxygenase/lipoxygenase inhibitor, eicosatetranynoic acid (30 μM).
  3. The cytochrome P450 inhibitors proadifen (10 μM) and clotrimazole (10 μM) inhibited ACh, lemakalim (LEM) and AA-induced relaxation, whereas 17-octadecynoic acid (100 μM) and 7-ethoxyresorufin (10 μM) were without effect on all three vasodilators. Proadifen and clotrimazole also inhibited ACh (1 μM) and LEM (1 μM)-induced hyperpolarization.
  4. The ability of various potassium channel blockers to inhibit relaxation responses elicited with ACh, AA and 11,12-EET was also determined. Iberiotoxin (IBTX; 100 nM) was without effect on responses to ACh but significantly reduced responses to both AA and 11,12-EET. In contrast, 4-aminopyridine (4-AP; 5 mM) significantly reduced response to ACh but not responses to AA and 11,12-EET. Combined IBTX plus (4-AP) inhibited the ACh-induced relaxation to a greater extent than 4-AP alone. Apamin (1 μM), glibenclamide (10 μM) and BaCl2 (50 μM) had no significant effect on responses to ACh, AA and 11,12-EET.
  5. IBTX (100 nM) significantly reduced both 11,12-EET (33 μM) and AA (30 μM) hyperpolarization without affecting the ACh (1 μM)-induced hyperpolarization. In contrast, 4-AP significantly reduced the ACh-induced hyperpolarization without affecting either AA or 11,12-EET-induced hyperpolarizations.
  6. In summary, our results suggest that the coronary endothelium releases a factor upon application of AA which hyperpolarizes the smooth muscle. The similarity of pharmacology between AA and 11,12-EET suggests that this factor is an EET. However, the disparity of pharmacology between responses to ACh versus responses to 11,12-EET do not support the hypothesis that EETs represent the predominant factor which ACh releases from the endothelium that leads to NO- and prostaglandin-independent hyperpolarization and relaxation in the GPCA.
  相似文献   

15.
  1. Previous studies have shown that ciprofloxacin and biphenylacetic acid (BPAA) synergistically inhibit γ-aminobutyric acid (GABA)A receptors. In the present study, we have investigated the actions of these two drugs on other neuronal ligand-gated ion channels.
  2. Agonist-evoked depolarizations were recorded from rat vagus and optic nerves in vitro by use of an extracellular recording technique.
  3. GABA (50 μM)-evoked responses, in the vagus nerve in vitro, were inhibited by bicuculline (0.3–10 μM) and picrotoxin (0.3–10 μM), with IC50 values and 95% confidence intervals (CI) of 1.2 μM (1.1–1.4) and 3.6 μM (3.0–4.3), respectively, and were potentiated by sodium pentobarbitone (30 μM) and diazepam (1 μM) to (mean±s.e.mean) 168±18% and 117±4% of control, respectively. 5-Hydroxytryptamine (5-HT; 0.5 μM)-evoked responses were inhibited by MDL 72222 (1 μM) to 10±4% of control; DMPP (10 μM)-evoked responses were inhibited by hexamethonium (100 μM) to 12±5% of control, and αbMeATP (30 μM)-evoked responses were inhibited by PPADS (10 μM) to 21±5% of control. Together, these data are consistent with activation of GABAA, 5-HT3, nicotinic ACh and P2X receptors, respectively.
  4. Ciprofloxacin (10–3000 μM) inhibited GABAA-mediated responses in the vagus nerve with an IC50 (and 95% CI) of 202 μM (148–275). BPAA (1–1000 μM) had little or no effect on the GABAA-mediated response but concentration-dependently potentiated the effects of ciprofloxacin by up to 33,000 times.
  5. Responses mediated by 5-HT3, nicotinic ACh and P2X receptors in the vagus nerve and strychnine-sensitive glycine receptors in the optic nerve were little or unaffected by ciprofloxacin (100 μM), BPAA (100 μM) or the combination of these drugs (both at 100 μM).
  6. GABA (1 mM)-evoked responses in the optic nerve were inhibited by bicuculline with an IC50 of 3.6 μM (2.8–4.5), a value not significantly different from that determined in the vagus nerve. Ciprofloxacin also inhibited the GABA-evoked response with an IC50 of 334 μM (256–437) and BPAA (100 μM) potentiated these antagonist effects. However, the magnitude of the synergy was 48 times less than that seen in the vagus nerve.
  7. These data indicate that ciprofloxacin and BPAA are selective antagonists of GABAA receptors, an action that may contribute to their excitatory effects in vivo. Additionally, our data suggest that the molecular properties of GABAA receptors in different regions of the CNS influence the extent to which these drugs synergistically inhibit the GABAA receptor.
  相似文献   

16.
  1. The actions of the cannabinoid receptor antagonist, SR 141716A, were examined in rat isolated mesenteric arteries. At concentrations greater than 3 μM, it caused concentration-dependent, but endothelium-independent, relaxations of both methoxamine- and 60 mM KCl-precontracted vessels.
  2. SR 141716A (at 10 μM, but not at 1 μM) inhibited contractions to Ca2+ in methoxamine-stimulated mesenteric arteries previously depleted of intracellular Ca2+ stores. Neither concentration affected the phasic contractions induced by methoxamine in the absence of extracellular Ca2+.
  3. SR 141716A (10 μM) caused a 130 fold rightward shift in the concentration-response curve to levcromakalim, a K+ channel activator, but had no effect at 1 μM.
  4. SR 141716A (10 μM) attenuated relaxations to NS 1619 (which activates large conductance, Ca2+-activated K+ channels; BKCa). The inhibitory effect of SR 141716A on NS 1619 was not significantly different from, and was not additive with, that caused by a selective BKCa inhibitor, iberiotoxin (100 nM). SR 141716A (1 μM) did not effect NS 1619 relaxation.
  5. SR 141716A (10 μM) had no effect on relaxations to the nitric oxide donor S-nitroso-N-acetylpenicillamine, or relaxations to carbachol in the presence of 25 mM KCl.
  6. The results show that, at concentrations of 10 μM and above, SR 141716A causes endothelium-independent vasorelaxation by inhibition of Ca2+ entry. It also inhibits relaxations mediated by K+ channel activation. This suggests that such concentrations of SR 141716A are not appropriate for investigation of cannabinoid receptor-dependent processes.
  相似文献   

17.
18.
  1. The effects of mesaconitine, the main alkaloid contained in Aconiti tuber, were investigated by use of extracellular recordings of stimulus-evoked population spikes and field excitatory postsynaptic potentials (e.p.s.ps) in the CA1 region of rat hippocampal slices.
  2. At a concentration of 10 nM, mesaconitine evoked excitations, which were manifested as an increase in the amplitude of the orthodromic spike and the appearance of multiple spikes following the first postsynaptic spike, without affecting the magnitude of paired-pulse facilitation. The increase in spike amplitude was persistent and was not reversed by up to 90 min of washout. At concentrations of 30 and 100 nM, the alkaloid produced a biphasic effect, that is an excitation followed by an inhibition without having any effect upon the field e.p.s.p. At concentrations above 100 nM, mesaconitine suppressed the orthodromic population spike and the field e.p.s.p.
  3. The excitatory effect was also observed when electrical stimulation was stopped completely during the application of mesaconitine (10 nM) and during the first 15 min of washout.
  4. The enhancement of the population spike and the appearance of multiple spikes induced by mesaconitine (10–100 nM) were blocked by pretreatment with the β-adrenoceptor antagonists propranolol (1 μM) and timolol (1 μM), whereas the inhibitory effect was blocked by the α-adrenoceptor antagonists yohimbine (1 μM) and phentolamine (10 μM). However, when the β-adrenoceptor antagonist timolol was added 10 min after the application of mesaconitine, it failed to block the long-lasting enhancement of the spike amplitude and the appearance of multiple population spikes.
  5. Application of the selective β-adrenoceptor agonist isoprenaline (500 nM) to the hippocampal slices induced an increase in the amplitude of the orthodromic population spike and elicited 2–3 additional spikes. Mesaconitine (10 nM) did not further potentiate this enhancement of the spike amplitude when added after a 15 min pretreatment with isoprenaline.
  6. Perfusion of forskolin, which directly activates adenylate cyclase, enhanced the population spike. Mesaconitine had no additional effect when applied after pretreatment with forskolin.
  7. It is concluded that the excitatory effects evoked by lower concentrations of the plant alkaloid mesaconitine are mediated by stimulation of β-adrenoceptors and the consequent activation of intracellular processes which lead to the long-lasting changes in excitability.
  相似文献   

19.
  1. The aim of the present study was to determine the effect of nitric oxide (NO) on angiotensin-converting enzyme (ACE) activity.
  2. A biochemical study was performed in order to analyse the effect of the NO-donors, SIN-1 and diethylamine/NO (DEA/NO), and of an aqueous solution of nitric oxide on the ACE activity in plasma from 3-month old male Sprague-Dawley rats and on ACE purified from rabbit lung. SIN-1 significantly inhibited the activity of both enzymes in a concentration-dependent way between 1 and 100 μM. DEA/NO inhibited the activity of purified ACE from 0.1 μM to 10 μM and plasma ACE, with a lower potency, between 1 and 100 μM. An aqueous solution of NO (100 and 150 μM) also inhibited significantly the activity of both enzymes. Lineweaver-Burk plots indicated an apparent competitive inhibition of Hip-His-Leu hydrolysis by NO-donors.
  3. Modulation of ACE activity by NO was also assessed in the rat carotid artery by comparing contractions elicited by angiotensin I (AI) and AII. Concentration-response curves to both peptides were performed in arteries with endothelium in the presence of the guanylyl cyclase inhibitor, ODQ (10 μM), and the inhibitor of NO formation, L-NAME (0.1 mM). NO, which is still released from endothelium in the presence of 10 μM ODQ, elicited a significant inhibition of AI contractions at low concentrations (1 and 5 nM). In the absence of endothelium, 1 μM SIN-1 plus 10 μM ODQ, as well as 10 μM DEA/NO plus 10 μM ODQ induced a significant inhibition on AI-induced contractions at 1 and 5 nM and at 1–100 nM, respectively.
  4. In conclusion, we demonstrated that (i) NO and NO-releasing compounds inhibit ACE activity in a concentration-dependent and competitive way and that (ii) NO release from endothelium physiologically reduces conversion of AI to AII.
  相似文献   

20.
  1. We used whole-cell patch clamp to investigate the currents activated by nicorandil in smooth muscle cells isolated from rat small mesenteric arteries, and studied the relaxant effect of nicorandil using myography.
  2. Nicorandil (300 μM) activated currents with near-linear current-voltage relationships and reversal potentials near to the equilibrium potential for K+.
  3. The nicorandil-activated current was blocked by glibenclamide (10 μM), but unaffected by iberiotoxin (100 nM) and the guanylyl cyclase inhibitor LY 83583 (1 μM). During current activation by nicorandil, openings of channels with a unitary conductance of 31 pS were detected.
  4. One hundred μM nicorandil had no effect on currents through Ca2+ channels recorded in response to depolarizing voltage steps using 10 mM Ba2+ as a charge carrier. A small reduction in current amplitude was seen in 300 μM nicorandil, though this was not statistically significant.
  5. In arterial rings contracted with 20 mM K+ Krebs solution containing 200 nM BAYK 8644, nicorandil produced a concentration-dependent relaxation with mean pD2=4.77±0.06. Glibenclamide (10 μM) shifted the curve to the right (pD2=4.32±0.05), as did 60 mM K+. LY 83583 caused a dose-dependent inhibition of the relaxant effect of nicorandil, while LY 83583 and glibenclamide together produced greater inhibition than either alone.
  6. Metabolic inhibition with carbonyl cyanide m-chlorophenyl hydrazone (30 nM), or by reduction of extracellular glucose to 0.5 mM, increased the potency of nicorandil.
  7. We conclude that nicorandil activates KATP channels in these vessels and also acts through guanylyl cyclase to cause vasorelaxation, and that the potency of nicorandil is increased during metabolic inhibition.
  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