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 62 毫秒
1.
  1. The effects of NK-3 receptor agonists on the release of substance P-immunoreactivity (SP-LI) have been investigated using superfused rat spinal cord synaptosomes.
  2. The Ca2+-dependent overflow of SP-LI evoked by 35 mM KCl was concentration-dependently enhanced by senktide (EC50=52 nM; maximal effect=70%) or [MePhe7]NKB (EC50=5.5 nM; maximal effect=125%), both selective agonists at receptors of the NK-3 type.
  3. The potentiation of the SP-LI overflow elicited by 100 nM senktide or [MePhe7]NKB was prevented by the NK-3 receptor antagonist (+)-SR142801. The antagonist halved, at 10 nM, and almost abolished, at 100 nM, the effect of both agonists. The effect of senktide or [MePhe7]NKB was insensitive to antagonists at NK-1 or NK-2 receptors.
  4. Capsaicin (0.1–1 μM) stimulated SP-LI release in a concentration-dependent manner from spinal cord synaptosomes. The SP-LI overflow elicited by 1 μM capsaicin was completely dependent on external Ca2+. Senktide could not affect the capsaicin-evoked release of SP-LI.
  5. Senktide failed to potentiate the K+-evoked overflow of SP-LI from synaptosomes previously exposed for 15 min in superfusion to capsaicin.
  6. The results show that release-enhancing NK-3 receptors are located on axon terminals of capsaicin-sensitive primary afferent neurones in the spinal cord. Antagonists at NK-3 receptors might help controlling pain transmission.
  相似文献   

2.
  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   

3.
  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   

4.
  1. BTS 67 582 (1,1-dimethyl-2-(2-morpholinophenyl)guanidine fumarate) is a novel antidiabetic agent with a short-acting insulin-releasing effect. This study examined its mode of action in the clonal B-cell line BRIN-BD11.
  2. BTS 67 582 increased insulin release from BRIN-BD11 cells in a concentration-dependent manner (10−8 to 10−4M) at both non-stimulating (1.1 mM) and stimulating (16.7 mM) concentrations of glucose.
  3. BTS 67 582 (10−4M) potentiated the insulin-releasing effect of a depolarizing concentration of K+ (30 mM), whereas the K+ channel openers pinacidil (400 μM) and diazoxide (300 μM) inhibited BTS 67 582-induced release.
  4. Suppression of Ca+ channel activity with verapamil (20 μM) reduced the insulin-releasing effect of BTS 67 582 (10−4M).
  5. BTS 67 582 (10−4M) potentiated insulin release induced by amino acids (10 mM), and enhanced the combined stimulant effects of glucose plus either the fatty acid palmitate (10 mM), or agents which raise intracellular cyclic AMP concentrations (25 μM forskolin and 1 mM isobutylmethylxanthine), or the cholinoceptor agonist carbachol (100 μM).
  6. Inhibition of glucose-stimulated insulin release by adrenaline or noradrenaline (10 μM) was partially reversed by BTS 67 582 (10−4M).
  7. These data suggest that the insulin-releasing effect of BTS 67 582 involves regulation of ATP-sensitive K+ channel activity and Ca2+ influx, and that the drug augments the stimulant effects of nutrient insulin secretagogues and agents which enhance adenylate cyclase and phospholipase C. BTS 67 582 may also exert insulin-releasing effects independently of ATP-sensitive K+ channel activity.
  相似文献   

5.
  1. We investigated the contribution of nitric oxide (NO) to inhibitory neuromuscular transmission in murine proximal colon and the possibility that citrulline is recycled to arginine to maintain the supply of substrate for NO synthesis.
  2. Intracellular microelectrode recordings were made from circular smooth muscle cells in the presence of nifedipine and atropine (both 1 μM). Electrical field stimulation (EFS, 0.3–20 Hz) produced inhibitory junction potentials (i.j.ps) composed of an initial transient hyperpolarization (fast component) followed by a slow recovery to resting potential (slow component).
  3. L-Nitro-arginine-methyl ester (L-NAME, 100 μM) selectively abolished the slow component of i.j.ps. The effects of L-NAME were reversed by L-arginine (0.2–2 mM) but not by D-arginine (2 mM). Sodium nitroprusside (an NO donor, 1 μM) reversibly hyperpolarized muscle cells. This suggests that NO mediates the slow component of i.j.ps.
  4. L-Citrulline (0.2 mM) also reversed the effects of L-NAME, and this action was maintained during sustained exposures to L-citrulline (0.2 mM). This may reflect intraneuronal recycling of L-citrulline to L-arginine.
  5. Higher concentrations of L-citrulline (e.g. 2 mM) had time-dependent effects. Brief exposure (15 min) reversed the effects of L-NAME, but during longer exposures (30 min) the effects of L-NAME gradually returned. In the continued presence of L-citrulline, L-arginine (2 mM) readily restored nitrergic transmission, suggesting that during long exposures to high concentrations of L-citrulline, the ability to generate arginine from citrulline was reduced.
  6. Aspartate (2 mM) had no effect on i.j.ps, the effects of L-NAME, or the actions of L-citrulline in the presence of L-NAME. L-Citrulline (0.2–2 mM) alone had no effect on i.j.ps under control conditions.
  7. S-methyl-L-thiocitrulline (10 μM), a novel NOS inhibitor, blocked the slow component of i.j.ps. The effects of this inhibitor were reversed by L-arginine (2 mM), but not by L-citrulline (2 mM).
  8. These results suggest that i.j.ps in the murine colon result from release of multiple inhibitory neurotransmitters. NO mediates a slow component of enteric inhibitory neurotransmission. Recycling of L-citrulline to L-arginine may sustain substrate concentrations in support of NO synthesis and this pathway may be inhibited when concentrations of L-citrulline are elevated.
  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. 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   

8.
  1. In AtT20 mouse corticotroph tumour cells large conductance Ca2+-activated K+-channels (BK-channels) have an essential role in the early glucocorticoid inhibition of adrenocorticotrophin (ACTH) secretion evoked by corticotrophin-releasing factor. The present study examined whether or not BK-channels are also pivotal to glucocorticoid inhibition of normal rat anterior pituitary cells.
  2. A membrane-permeant, non-metabolizable cyclic AMP analogue, 8-(4-Chlorophenylthio)adenosine-3′,5′-cyclic-monophosphate (CPT-cAMP) was used as the primary secretagogue stimulus, as this mimics the increase of intracellular cyclic AMP caused by corticotrophin-releasing factor, but is not subject to the complex Ca2+-dependent regulation of cyclic AMP metabolism that is evident in corticotroph cells.
  3. Experiments in AtT20 cells showed that ACTH secretion stimulated by 1 mM CPT-cAMP was suppressed to 34±1.5% (n=12) of the control stimulus by a maximal dose of 100 nM dexamethasone. The ACTH secretion evoked by the combination of 1 mM CPT-cAMP with either 5 μM (−)BayK8644 (L-type Ca2+-channel activator) or 5 mM TEA (K+-channel blocker) was respectively 69.1±7.6% and 69.3±11.8% of control after 2 h preincubation with 100 nM dexamethasone (P<0.05 vs CPT-cAMP). The ACTH response elicited by 5 μM (−)BayK8644 and 5 mM TEA given together was completely resistant to inhibition by 100 nM dexamethasone. Furthermore, TEA and (−)BayK8644 given together synergistically stimulated ACTH release in combination with 0.1 mM or 1 mM CPT-cAMP, and these ACTH responses were not inhibited by 100 nM dexamethasone.
  4. In primary cultures of rat anterior pituitary cells, TEA (up to 20 mM), charybdotoxin (30 nM) or apamin (100 nM) failed to modify the glucocorticoid inhibition of 0.1 mM CPT-cAMP-induced ACTH release. The combination of 5 mM TEA and 5 μM (−)BayK8644 elicited a small but significant increase in ACTH secretion but did not modify the inhibition of 0.3 mM CPT-cAMP-induced ACTH secretion by 100 nM dexamethasone.
  5. In primary cultures of rat anterior pituitary cells, depolarization of the membrane potential with 40 mM KCl enhanced the ACTH response to CPT-cAMP and markedly reduced the maximal inhibitory effect of dexamethasone to 55±1.2% as well as that of corticosterone to 33±2.1% vs 100±2.5% and 100±1.9% inhibition respectively, when 0.1 mM CPT-cAMP was used alone. Introduction of 5 μM (−)BayK8644 with 40 mM KCl in this system had no additional effect on glucocorticoid inhibition.
  6. No glucocorticoid inhibition of ACTH release to any of the stimuli applied was observed in cells pretreated with the mRNA synthesis inhibitor, 5,6-dichloro-furanosyl-benzimidazole riboside (DRB) (0.1 mM) or the protein synthesis blocker, puromycin (0.1 mM).
  7. In summary, early glucocorticoid inhibition of stimulated ACTH release by cultured rat anterior pituitary cells was dependent on the synthesis of new mRNA and protein. Depolarization of the membrane potential potentiated CPT-cAMP-induced ACTH secretion in AtT20 cells as well as cultured rat corticotrophs and this was associated with a resistance to the early inhibitory effect of glucocorticoids. Glucocorticoid inhibition in rat anterior pituitary corticotrophs was unaltered by TEA, charybdotoxin as well as apamin, and hence it is unlikely to involve predominantly BK-or SK-type Ca2+-activated K+-channels.
  8. These results support the thesis that a prime target of glucocorticoid feedback inhibition in anterior pituitary corticotrophs is the membrane potential and indicate that glucocorticoid-induced proteins regulate the activities of several distinct plasma membrane ion channels.
  相似文献   

9.
  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   

10.
  1. We investigated by intravital microscopy in rats, the in vivo direct effects of theophylline on the diameters of second and third order diaphragm arterioles.
  2. Theophylline (1–100 μM) dilated second and third order diaphragm arterioles significantly, and with an amplitude which was not statistically different from the one obtained with adenosine (1–100 μM). Enprofylline (1–100 μM), a theophylline analogue with poor adenosine-receptor antagonism but with similar or higher phosphodiesterases inhibition properties than theophylline, also dilated diaphragm arterioles, causing however, a significantly smaller dilatation than theophylline.
  3. Neither the A1 adenosine receptor antagonist 8-cyclopentyl-1,3-dipropylxanthine (CPX, 50 nM), nor the A2 adenosine receptor antagonist 3,7-dimethyl-1-proparglyxanthine (DMPX, 10 μM) reduced significantly theophylline-induced arteriolar dilatation.
  4. Theophylline (100 nM) abolished adenosine-induced arteriolar dilatation.
  5. The dilatation induced by theophylline was unchanged by the nitric oxide (NO) synthase inhibitor Nω-nitro-L-arginine (NNA, 300 μM).
  6. Theophylline-induced arteriolar dilatation was abolished by the prostaglandin synthesis inhibitors mefenamic acid or indomethacin (20 μM).
  7. These findings show that theophylline induced a significant dilatation of diaphragm arterioles via the release of prostaglandins.
  相似文献   

11.
  1. The endothelin (ET) receptor subtype that mediates niric oxide (NO)-dependent airway relaxation in tracheal tube preparations precontracted with carbachol and pretreated with indomethacin was investigated. The release of NO induced by ET from guinea-pig trachea using a recently developed porphyrinic microsensor was also measured.
  2. ET-1 (1 pM–100 nM) contracted tracheal tube preparations pretreated with the NO-synthase inhibitor, L-NMMA, and relaxed, in an epithelium-dependent manner, preparations pretreated with the inactive enantiomer D-NMMA. The effect of L-NMMA was reversed by L-Arg, but not by D-Arg.
  3. The selective ETB receptor agonists, IRL 1620 or sarafotoxin S6c, both (1 pM–100 nM) contracted tracheal tube preparations in a similar manner either after treatment with D-NMMA or with L-NMMA. In the presence of the ETA receptor antagonist, FR139317 (10 μM), ET-1 administration resulted in a contraction that was similar after either L-NMMA or D-NMMA. In the presence of the ETB receptor antagonist, BQ788 (1 μM), ET-1 relaxed and contracted tracheas pretreated with D-NMMA and L-NMMA, respectively.
  4. Exposure of tracheal segments to ET-1 (1–1000 nM) caused a concentration-dependent increase in NO release that was reduced by L-NMMA. IRL1620 (1 μM) did not cause any significant NO release. FR139317 (10 μM), but not, BQ788 (1 μM), inhibited the NO release induced by ET-1.
  5. These results demonstrate that in the isolated guinea-pig trachea activation of ETB receptors results in a contractile response, whereas activation of ETA receptors cause both a contraction, and an epithelium-dependent relaxation that is mediated by NO release.
  相似文献   

12.
  1. The dose-related inhibition of the twitch responses of the myenteric plexus-longitudinal muscle preparation of the guinea-pig small intestine by cannabinoid (CB) agonists, (+)-WIN 55212 and CP 55940 during stimulation at 0.1 Hz with supramaximal voltage was confirmed. These agonists inhibited acetylcholine (ACh) release in the presence of physostigmine (7.7 μM) thus indicating a prejunctional site of action.
  2. Inhibition of twitch responses and ACh release by CB agonists was reversed by the CB1-selective cannabinoid receptor antagonist, SR141716A. Dose-response curves to (+)-WIN 55212 and CP 55940 were shifted to the right, with no reduction of maximal response, by pretreatment with SR141716A (31.6–1000 nM), but not its vehicle, Tween 80 (1 μM). However, at very high concentrations (25–400 μM), Tween 80 itself caused a dose-related inhibition of the twitch response which was significantly reduced in the presence of SR141716A (1 μM). The opioid receptor antagonist, naloxone (1 μM) had no significant effect on the inhibition by CP 55940 of the twitch response.
  3. (+)-WIN 55212, CP 55940 and Tween 80 (50 μM) had no effect on responses to exogenous ACh, confirming that their actions were prejunctional. SR141716A (1 μM) did not increase the sensitivity of the longitudinal muscle to either ACh or histamine, but inhibited the responses to high doses of ACh.
  4. The (−)-enantiomer of WIN 55212, was approximately 300 times less active than the (+) enantiomer in inhibiting the twitch response, had no CB1 antagonist activity against the active isomer and did not inhibit the release of ACh in the presence of physostigmine.
  5. The dissociation constant (KD) values for SR 141716A against the inhibitory effect of (+)-WIN 55212 and CP 55940 on the twitch response were 12.07 nM (95% confidence intervals 8.55 and 20.83) and 6.44 nM (95% confidence intervals 4.70 and 10.24), respectively. In experiments in which the release of ACh was inhibited by (+)-WIN 55212, the KD values were 9.21 nM and 10.53 nM at SR141716A concentrations of 31.6 nM and 100 nM, respectively. The KD values for the antagonism by naloxone of the inhibition of the twitch responses and the inhibition of ACh release by normorphine in this preparation were found to be 2.38±0.69 nM and 2.00±0.9 nM, respectively.
  6. During maximal inhibition of ACh release by (+)-WIN 55212, the addition of normorphine (400 nM) caused a further significant decrease in ACh output.
  7. SR141716A alone produced a significant increase in ACh release in both the absence and presence of exogenous cannabinoid drugs, hence we conclude that it has a presynaptic site of action. We also conclude that SR141716A acts either by antagonizing the effect of an endogenous CB1 receptor agonist or by having an inverse agonist effect at these receptors.
  相似文献   

13.
  1. We have investigated the contractile property of cyclosporin A and FK506 in guinea-pig isolated bronchus.
  2. Cyclosporin A (10 μM) failed to significantly attenuate the excitatory non-adrenergic non-cholinergic (eNANC) and cholinergic contractile response (per cent methacholine Emax) induced by electrical field stimulation (EFS). In contrast, eNANC responses were significantly attenuated by both the neurokinin (NK)-1 and (NK)-2 receptor antagonists, N-acetyl-L-tryptophan 3,5-bis (trifluoromethyl)-benzyl and SR48968, respectively.
  3. Cyclosporin A and FK506 caused a concentration-dependent contraction in guinea-pig isolated bronchus, which was significantly attenuated by NK-1 and NK-2 receptor antagonists. The capsaicin receptor antagonist, capsazepine (10 μM) significantly reduced the contractile response to cyclosporin A and capsaicin, but not to FK506.
  4. The N-type calcium channel blocker, ω-Conotoxin (ωCTX: 10 nM), significantly reduced the contractile response to FK506 and the eNANC response following EFS. In contrast, ω-CTX failed to significantly reduce the contractile potency to capsaicin or cyclosporin A.
  5. In bronchial preparations desensitized by repeated application of capsaicin (1 μM), the contractile responses to both cyclosporin A (100 μM) and FK506 (100 μM), were significantly reduced. In contrast, the contractile responses to substance P and neurokinin A (10 μM) were not altered. Furthermore, repeated application of cyclosporin A (100 μM) significantly inhibited the contractile response to capsaicin (1 μM).
  6. The findings from this study would indicate that cyclosporin A and FK506 mediate contraction of guinea-pig isolated bronchus secondary to the release of neuropeptides from airway sensory nerves. However, the release of sensory neuropeptides appears to be mediated via different mechanisms for cyclosporin A and FK506, the former by stimulation of the vanilloid receptor and the latter via opening of N-type calcium channels.
  相似文献   

14.
  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   

15.
  1. The pulmonary vasculature is normally in a low resting state of tone. It has been hypothesized that this basal tone is actively maintained by the continuous release of a vasodilator in the resting state. However, evidence for basal release of nitric oxide (NO) is inconclusive.
  2. We studied the release of NO in arteries from the pulmonary circulation of male Wistar-Kyoto rats by examining the effects of the L-arginine analogue NG-nitro-L-arginine methyl ester (L-NAME) on resting pulmonary arteries and on vessels pre-contracted with prostaglandin F (PGF).
  3. Rats (n=21) were killed by an overdose with pentobarbitone. Pulmonary arteries were dissected (mean internal diameter 459±11 μm) and mounted in a small vessel wire myograph. Resting tensions were set to simulate transmural pressures of 17.5 mmHg.
  4. L-NAME (100 μM) was found to produce a contraction of 0.64±0.09 mN mm−1 in resting pulmonary arteries when added alone to the myograph bath. This contraction was not produced following removal of the endothelium. Vessel contraction to PGF (100 μM) was found to be significantly greater when carried out in the presence of L-NAME (100 μM)–1.37±0.15 mN mm−1 compared with 1.96±0.17 mN mm−1. Dilatation following acetylcholine (ACh) (1 μM) was abolished in the presence of L-NAME (100 μM).
  5. Rat pulmonary artery contraction in response to the addition of L-NAME and the absence of contraction upon removal of the endothelium provides supportive evidence of the active release of nitric oxide for the maintenance of resting tone.
  相似文献   

16.
  1. By use of rabbit isolated perfused intact ears and isolated perfused segments of central and first generation daughter branch ear arteries, we investigated the actions of charybdotoxin (ChTX), a blocker of calcium-activated K+ channels (KCa channels), and Nω-nitro-L-arginine methyl ester (L-NAME) on pressure-flow and diameter-flow relationships.
  2. ChTX (1 nM) induced an upwards shift in the pressure-flow curve in the rabbit intact isolated ear preconstricted with 5-hydroxytryptamine (5-HT; 100 nM) with subsequent administration of L-NAME (100 μM) inducing a further upwards shift. L-NAME itself induced an upwards shift in the pressure-flow curve, but subsequent administration of ChTX was without significant effect.
  3. Microangiographic analysis revealed a tendency of ChTX (1 nM) to decrease vessel diameter in the central ear artery (G0) with little effect on the first two generations of daughter branch arteries (G1 and G2) in the intact ear. Subsequent addition of L-NAME (100 μM) did not significantly further decrease vessel diameter in G0, but did decrease vessel diameter in G1 and G2. L-NAME itself showed a tendency to decrease vessel diameter in G0, G1 and G2 vessels with subsequent addition of ChTX being without significant effect.
  4. In an isolated G0 preparation which was preconstricted with 5-HT (100 nM), ChTX (1 nM) caused an upwards shift in the pressure-flow curve which was augmented by subsequent addition of L-NAME (100 μM). L-NAME (100 μM) itself caused an upwards shift in the pressure-flow curve but subsequent addition of ChTX (1 nM) had no significant effect.
  5. In comparison, in an isolated G1 preparation which was preconstricted with 5-HT (100 nM), ChTX (1 nM) had no significant effect on the pressure-flow curve relative to control, but subsequent addition of L-NAME (100 μM) caused an upwards shift. L-NAME (100 μM) itself induced an upwards shift in the pressure-flow curve with subsequent addition of ChTX (1 nM) being without significant effect.
  6. ChTX (10 pM–10 nM) caused a concentration-dependent increase in perfusion pressure in isolated G0 and G1 preparations at fixed flow rates of 2 ml min−1 and 0.5 ml min−1, respectively. These responses were enhanced in the presence of L-NAME (100 μM) in G1 but not G0 preparations.
  7. We conclude that at 1 nM, ChTX exhibits differential actions on central and daughter branch arteries of the intact ear of the rabbit, which are also apparent in the corresponding arteries when studied in isolation. The action of 1 nM ChTX in G0 vessels may reflect inhibition of either the release or action of nitric oxide as it was blocked in the presence of L-NAME. At higher concentrations of ChTX, there would appear to be a direct constrictor effect on vascular smooth muscle which is apparent in both G0 and G1 vessels. This observed heterogeneity could reflect different distributions of KCa channels between central and daughter branch arteries at either the endothelial or smooth muscle levels, or both.
  相似文献   

17.
  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   

18.
  1. Some cardiovascular disturbances which occur in diabetics are a consequence of alterations in vascular contractility as well as in endothelium-dependent relaxation.
  2. Calcium dobesilate (DOBE) is a drug used in diabetic retinopathy and its mechanism of action is not yet understood.
  3. The aim of this study was to investigate the effects of DOBE on synthesis and release of endothelium-dependent relaxing factor (EDRF) and endothelium-dependent hyperpolarizing factor (EDHF) in rabbit isolated aorta.
  4. Endothelium-dependent relaxation induced by acetylcholine (ACh) (10−8–10−5M) increased in the presence of DOBE 10−5M only when vascular endothelium was kept intact.
  5. NG-nitro-L-arginine methyl ester (L-NAME; 10−8–10−4M progressively decreased the enhancing effect of DOBE on endothelium-dependent relaxation whereas it was progressively increased by L-Arg.
  6. DOBE 10−5M increased in a non-significant manner endothelium-dependent relaxation induced by ACh when the arteries were incubated with both L-NAME 10−4M and indomethacin 10−6M.
  7. DOBE (10−6M and 10−5M) was able to scavenge superoxide anion radicals generated by the hypoxanthine/xanthine oxidase reaction.
  8. These results provide evidence that DOBE is able to affect the vascular disorders associated with diabetes mellitus since it enhances the synthesis of endothelium-dependent relaxing factors.
  相似文献   

19.
  1. The main object of the present study was to determine whether ascorbate, an antioxidant which has been shown to protect nitric oxide (NO) from attack by scavenger molecules, might be released from nitrergically-innervated smooth muscle; ascorbate release from the rat anococcygeus was measured by use of h.p.l.c. with electrochemical detection.
  2. Incubation of rat anococcygeus muscles in normal physiological salt solution (PSS; 30 min) resulted in release of ascorbate into the bathing medium (7.7±0.9 nmol g−1 tissue). This release was increased by 96% when muscles were incubated in high K+ (70 mM) PSS. The resting release of ascorbate was unaffected by tetrodotoxin (TTX; 1 μM), ω-conotoxin GVIA (10 nM) or omission of calcium ions from the PSS (with addition of 0.2 mM EGTA), but all three procedures attenuated the increased release observed under depolarizing conditions. Resting release of ascorbate was unaffected by glutamate (100 μM), aspartate (100 μM), γ-aminobutyric acid (100 μM) or carbachol (50 μM).
  3. A second h.p.l.c. peak, which always preceded the ascorbate peak, was identified as urate. Urate release from the anococcygeus, following 30 min incubation in normal PSS, was 64.6±12.7 nmol g−1 tissue but, unlike ascorbate, urate release was unchanged in high K+ PSS. In functional experiments, urate (100–400 μM) partially protected NO (15 μM)-induced relaxations of the rat anococcygeus from inhibition by 2-(4-carboxyphenyl)-4,4,5,5-tetramethylimidazoline-1-oxyl-3-oxide (carboxy-PTIO; 50 μM), but not from inhibition by hydroquinone or duroquinone (both 100 μM).
  4. Muscles chemically sympathectomized with 6-hydroxydopamine (6-OHDA, 500 μM; 2 h) still exhibited release of ascorbate (2.5±0.4 nmol g−1 tissue) and urate (22.2±2.9 nmol g−1 tissue); in both cases the release was similar to that observed in time-matched control tissues not exposed to 6-OHDA. High K+ PSS produced a TTX-sensitive increase in release of ascorbate, but not urate, from 6-OHDA-treated muscles.
  5. The results demonstrate that significant amounts of ascorbate and urate are released from the rat anococcygeus muscle. Ascorbate, but not urate, release appears to be enhanced by activation of nerves which are resistant to 6-OHDA pretreatment. Since both antioxidants can protect NO from attack by scavenger molecules, their release in nitrergically-innervated tissues may be important for the provision of the correct redox environment to allow NO to fulfill its proposed neurotransmitter role.
  相似文献   

20.
  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   

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