首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
  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   

2.
  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   

3.
  1. Although stimulation of mouse RAW 264.7 macrophages by UTP elicits a rapid increase in intracellular free Ca2+ ([Ca2+]i), phosphoinositide (PI) turnover, and arachidonic acid (AA) release, the causal relationship between these signalling pathways is still unclear. In the present study, we investigated the involvement of phosphoinositide-dependent phospholipase C (PI-PLC) activation, Ca2+ increase and protein kinase activation in UTP-induced AA release. The effects of stimulating RAW 264.7 cells with thapsigargin, which cannot activate the inositol phosphate (IP) cascade, but results in the release of sequestered Ca2+ and an influx of extracellular Ca2+, was compared with the effects of UTP stimulation to elucidate the multiple regulatory pathways for cPLA2 activation.
  2. In RAW 264.7 cells UTP (100 μM) and thapsigargin (1  μM) caused 2 and 1.2 fold increases, respectively, in [3H]-AA release. The release of [3H]-AA following treatment with UTP and thapsigargin were non-additive, totally abolished in the Ca2+-free buffer, BAPTA (30 μM)-containing buffer or in the presence of the cPLA2 inhibitor MAFP (50 μM), and inhibited by pretreatment of cells with pertussis toxin (100 ng ml−1) or 4-bromophenacyl bromide (100 μM). By contrast, aristolochic acid (an inhibitor of sPLA2) had no effect on UTP and thapsigargin responses.
  3. U73122 (10 μM) and neomycin (3 mM), inhibitors of PI-PLC, inhibited UTP-induced IP formation (88% and 83% inhibition, respectively) and AA release (76% and 58%, respectively), accompanied by a decrease in the [Ca2+]i rise.
  4. Wortmannin attenuated the IP response of UTP in a concentration-dependent manner (over the range 10 nM–3 μM), and reduced the UTP-induced AA release in parallel. RHC 80267 (30 μM), a specific diacylglycerol lipase inhibitor, had no effect on UTP-induced AA release.
  5. Short-term treatment with PMA (1 μM) inhibited the UTP-stimulated accumulation of IP and increase in [Ca2+]i, but had no effect on the release of AA. In contrast, the AA release caused by thapsigargin was increased by PMA.
  6. The role of PKC in UTP- and thapsigargin-mediated AA release was shown by the blockade of these effects by staurosporine (1 μM), Ro 31-8220 (10 μM), Go 6976 (1 μM) and the down-regulation of PKC.
  7. Following treatment of cells with SK&F 96365 (30 μM), thapsigargin-, but not UTP-, induced Ca2+ influx, and the accompanying AA release, were down-regulated.
  8. Neither PD 98059 (100 μM), MEK a inhibitor, nor genistein (100 μM), a tyrosine kinase inhibitor, had any effect on the AA responses induced by UTP and thapsigargin.
  9. We conclude that UTP-induced cPLA2 activity depends on the activation of PI-PLC and the sustained elevation of intracellular Ca2+, which is essential for the activation of cPLA2 by UTP and thapsigargin. The [Ca2+]i-dependent AA release that follows treatment with both stimuli was potentiated by the activity of protein kinase C (PKC). A pertussis toxin-sensitive pathway downstream of the increase in [Ca2+]i was also shown to be involved in AA release.
  相似文献   

4.
  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   

5.
  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   

6.
  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   

7.
  1. This study was aimed at characterizing ATP-induced rises in cytosolic free calcium ion, [Ca2+]i, in a population of rat striatal astrocytes loaded with the fluorescent Ca2+ probe Fura2, by means of fluorescence spectrometry.
  2. ATP triggered a fast and transient elevation of [Ca2+]i in a concentration-dependent manner. The responses of the purine analogues 2-methylthio-ATP (2-meSATP), adenosine-5′-O-(2-thiodiphosphate) (ADPβS), as well as uridine-5′-triphosphate (UTP) resembled that of ATP, while α,β-methylene-ATP (α,β-meATP) and β,γ-methylene-ATP (β,γ-meATP) were totally ineffective.
  3. Suramin (50 μM) had only a minor effect on the ATP response, whereas pyridoxal phosphate-6-azophenyl-2′,4′-disulphonic acid (PPADS) (5 μM) significantly depressed the maximum response.
  4. Extracellular Ca2+ did not contribute to the observed [Ca2+]i rise: removing calcium from the extracellular medium (with 1 mM EGTA) or blocking its influx by means of either Ni2+ (1 mM) or Mn2+ (1 mM) did not modify the nucleotide responses.
  5. Furthermore, after preincubation with 10 μM thapsigargin, the nucleotide-evoked [Ca2+]i increments were completely abolished. In contrast, 10 mM caffeine did not affect the responses, suggesting that thapsigargin-, but not caffeine/ryanodine-sensitive stores are involved.
  6. Both application of the G-protein blocker guanosine-5′-O-(2-thiodiphosphate) (GDPβS) (1 mM) and preincubation with pertussis toxin (PTx) (350 ng ml−1) partially inhibited the nucleotide-mediated responses. Moreover, the phospholipase C (PLC) inhibitor U-73122, but not its inactive stereoisomer U-73343 (5 μM), significantly reduced the ATP-evoked [Ca2+]i rise.
  7. In conclusion, our results suggest that, in rat striatal astrocytes, ATP-elicited elevation of [Ca2+]i is due solely to release from intracellular stores and is mediated by a G-protein-linked P2Y receptor, partially sensitive to PTx and coupled to PLC.
  相似文献   

8.
  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   

9.
  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   

10.
  1. Stellettamide A (ST-A), a novel marine toxin isolated from a marine sponge, inhibited high K+(72.7 mM)-induced contraction in the smooth muscle of guinea-pig taenia coli with an IC50 of 88 μM.
  2. In the taenia permeabilized with Triton X-100, ST-A inhibited Ca2+ (3 and 10 μM)-induced contractions with an IC50 of 46 μM for 3 μM Ca2+ and 105 μM for 10 μM Ca2+. In the permeabilized taenia, calyculin-A (300 nM), a potent inhibitor of type-1 and type-2A phosphatases, induced sustained contraction in the absence of Ca2+. ST-A had no effect on this contraction.
  3. ST-A inhibited Mg2+-ATPase activity in native actomyosin prepared from chicken gizzard with an IC50 of 25 μM.
  4. In a reconstituted smooth muscle contractile system containing calmodulin, myosin light chain (MLC) and MLC kinase, ST-A inhibited MLC phosphorylation with an IC50 of 152 μM. The inhibitory effect of ST-A was antagonized by increasing the concentration of calmodulin.
  5. ST-A inhibited calmodulin activity, assessed by Ca2+/calmodulin-dependent enzymes, (Ca2+-Mg2+)-ATPase of erythrocyte membrane, with an IC50 of 100 μM and phosphodiesterase prepared from bovine cardiac muscle with an IC50 of 52 μM. The inhibitory effect on phosphodiesterase activity was antagonized by increasing the calmodulin concentration.
  6. Interaction between ST-A and calmodulin was demonstrated by instantaneous quenching of the intrinsic tyrosine fluorescence of calmodulin by ST-A (3–300 μM). Similar results were obtained in the presence or absence of Ca2+ suggesting that ST-A binds to calmodulin and that Ca2+ is not essential for the binding of ST-A to calmodulin.
  7. These results suggest that ST-A, isolated from marine metabolites, is a novel inhibitor of calmodulin.
  相似文献   

11.
  1. The binding of modulators of the ATP-sensitive K+ channel (KATP channel) to the murine sulphonylurea receptor, SUR2B, was investigated. SUR2B, a proposed subunit of the vascular KATP channel, was expressed in HEK 293 cells and binding assays were performed in membranes at 37°C using the tritiated KATP channel opener, [3H]-P1075.
  2. Binding of [3H]-P1075 required the presence of Mg2+ and ATP. MgATP activated binding with EC50 values of 10 and 3 μM at free Mg2+ concentrations of 3 μM and 1 mM, respectively. At 1 mM Mg2+, binding was lower than at 3 μM Mg2+.
  3. [3H]-P1075 saturation binding experiments, performed at 3 mM ATP and free Mg2+ concentrations of 3 μM and 1 mM, gave KD values of 1.8 and 3.4 nM and BMAX values of 876 and 698 fmol mg−1, respectively.
  4. In competition experiments, openers inhibited [3H]-P1075 binding with potencies similar to those determined in rings of rat aorta.
  5. Glibenclamide inhibited [3H]-P1075 binding with Ki values of 0.35 and 2.4 μM at 3 μM and 1 mM free Mg2+, respectively. Glibenclamide enhanced the dissociation of the [3H]-P1075-SUR2B complex suggesting a negative allosteric coupling between the binding sites for P1075 and the sulphonylureas.
  6. It is concluded that an MgATP site on SUR2B with μM affinity must be occupied to allow opener binding whereas Mg2+ concentrations ⩾10 μM decrease the affinities for openers and glibenclamide. The properties of the [3H]-P1075 site strongly suggest that SUR2B represents the drug receptor of the openers in vascular smooth muscle.
  相似文献   

12.
  1. We have recently demonstrated the formation of protein-bound dinitrosyl-iron complexes (DNIC) in rat aortic rings exposed to lipopolysaccharide (LPS) and shown that N-acetylcysteine (NAC) can promote vasorelaxation in these arteries, possibly via the release of nitric oxide (NO) as low molecular weight DNIC from these storage sites. The aim of the present study was to investigate further the mechanism of the relaxation induced by NAC in LPS-treated vessels.
  2. In rings incubated with LPS (10 μg ml−1 for 18 h) and precontracted with noradrenaline (NA, 3 μM) plus Nω-nitro-L-arginine methylester (L-NAME, 3 mM), the relaxation evoked by NAC (0.1 to 10 mM) was abolished by 1H-[1,2,4]oxadiazolo[4,3-a]quinoxalin-1-one (ODQ, 1 μM, a selective inhibitor of soluble guanylyl cyclase) but not affected by Rp-8-bromoguanosine 3′5′-cyclic monophosphorothioate (Rp-8BrcGMPS, 60 μM a selective inhibitor of cyclic GMP-dependent protein kinase). Tetrabutylammonium (TBA, 3 mM, as a non selective K+ channels blocker) or elevated concentration of external KCl (25 or 50 mM) significantly attenuated the NAC-induced relaxation. Selective K+ channels blockers (10 μM glibenclamide, 0.1 μM charybdotoxin, 0.5 μM apamin or 3 mM 4-aminopyridine) did not affect the NAC-induced relaxation. The relaxing effect of NAC (10 mM) was not associated with an elevation of guanosine 3′ : 5′ cyclic monophosphate (cyclic GMP) in LPS-treated rings.
  3. In aortic rings precontracted with NA (0.1 μM), low molecular weight DNIC (with thiosulphate as ligand, 1 nM to 10 μM) evoked a concentration-dependent relaxation which was antagonized by ODQ (1 μM) and Rp-8BrcGMPS (150 μM) but not significantly affected by TBA (3 mM) or by the use of KCl (50 mM) as preconstricting agent. The relaxation produced by DNIC (0.1 μM) was associated with an 11 fold increase in aortic cyclic GMP content, which was completely abolished by ODQ (1 μM).
  4. Taken together with our previous data, the main finding of the present study is that the vascular relaxation induced by NAC in LPS-treated aorta, although probably related to NO through an interaction via preformed NO stores, was not mediated by activation of the cyclic GMP pathway. It may involve the activation of TBA-sensitive K+ channels. The differences in the mechanism of relaxation induced by NAC and by exogenous DNIC suggest that the generation of low molecular weight DNIC from protein-bound species does not play a major role in the NAC-induced relaxation observed in LPS-treated rat aorta. In addition, it is suggested that ODQ may display other properties than the inhibition of soluble guanylyl cyclase.
  相似文献   

13.
  1. The mechanisms underlying the midazolam-induced relaxation of the noradrenaline (NA)-contraction were studied by measuring membrane potential, isometric force and intracellular concentration of Ca2+([Ca2+]i) in endothelium-denuded muscle strips from the rabbit mesenteric resistance artery. The actions of midazolam were compared with those of nicardipine, an L-type Ca2+-channel blocker.
  2. Midazolam (30 and 100 μM) did not modify either the resting membrane potential or the membrane depolarization induced by 10 μM NA.
  3. NA (10 μM) produced a phasic, followed by a tonic increase in both [Ca2+]i and force. Midazolam (10–100 μM) did not modify the resting [Ca2+]i, but attenuated the NA-induced phasic and tonic increases in [Ca2+]i and force, in a concentration-dependent manner. In contrast, nicardipine (0.3 μM) attenuated the NA-induced tonic, but not phasic, increases in [Ca2+]i and force.
  4. In Ca2+-free solution containing 2 mM EGTA, NA (10 μM) transiently increased [Ca2+]i and force. Midazolam (10–100 μM), but not nicardipine (0.3 μM), attenuated this NA-induced increase in [Ca2+]i and force, in a concentration-dependent manner. However, midazolam (10 and 30 μM), had no effect on the increases in [Ca2+]i and force induced by 10 mM caffeine.
  5. In ryanodine-treated strips, which have functionally lost the NA-sensitive Ca2+- storage sites, NA slowly increased [Ca2+]i and force. Nicardipine (0.3 μM) did not modify the resting [Ca2+]i but partly attenuated the NA-induced increases in [Ca2+]i and force. In the presence of nicardipine, midazolam (100 μM) lowered the resting [Ca2+]i and further attenuated the remaining NA-induced increases in [Ca2+]i and force.
  6. The [Ca2+]i-force relationship was obtained in ryanodine-treated strips by the application of ascending concentrations of Ca2+ (0.16–2.6 mM) in Ca2+-free solution containing 100 mM K+. NA (10 μM) shifted the [Ca2+]i-force relationship to the left and enhanced the maximum Ca2+-induced force. Under these conditions, whether in the presence or absence of 10 μM NA, midazolam (10 and 30 μM) attenuated the increases in [Ca2+]i and force induced by Ca2+ without changing the [Ca2+]i-force relationship.
  7. It was concluded that, in smooth muscle of the rabbit mesenteric resistance artery, midazolam inhibits the NA-induced contraction through its inhibitory action on NA-induced Ca2+ mobilization. Midazolam attenuates NA-induced Ca2+ influx via its inhibition of both nicardipine-sensitive and -insensitive pathways. Furthermore, midazolam attenuates the NA-induced release of Ca2+ from the storage sites. This effect contributes to the midazolam-induced inhibition of the NA-induced phasic contraction.
  相似文献   

14.
  1. The aim of study was to characterize endothelin (ET)-induced vasodilatation in isolated extrapulmonary rat arteries (EPA) and in intrapulmonary arteries (IPA) preconstricted with 1 μM phenylephrine.
  2. The ET-3 (1 nM–100 nM)- and ET-1 (10 nM–100 nM)-induced transient vasodilatations in EPA were more potent than those in IPA. The vasodilatation induced by ET-3 (100 nM) was larger than that induced by ET-1 (100 nM).
  3. Both the ETB antagonist, BQ788 (3 μM) and or endothelium denudation, but not the ETA antagonist, BQ123 (3 μM), abolished the vasodilatation induced by ET-1 or ET-3 (100 nM each) in EPA and in IPA. The ATP-sensitive K+channel blocker, glibenclamide (20 μM) and the nitric oxide synthase inhibitor, NG-monomethyl-L-arginine (L-NMMA, 1 mM) suppressed the ET-induced vasodilatation in EPA and in IPA.
  4. We conclude that the vasodilatation induced by endothelins is markedly reduced in rat isolated IPA, and suggest that the endothelial ETB-mediated vasodilatation varies depending on rat pulmonary arterial regions. Furthermore, ETB-mediated vasodilatation involves activation of ATP-sensitive K+ channels and of nitric oxide synthase in rat isolated EPA and IPA.
  相似文献   

15.
  1. The rat μ-opioid receptor has recently been cloned, yet its second messenger coupling remains unclear. The endogenous μ-opioid receptor in SH-SY5Y cells couples to phospholipase C (PLC), increases [Ca2+]i and inhibits adenylyl cyclase (AC). We have examined the effects of μ-opioid agonists on inositol(1,4,5)trisphosphate (Ins(1,4,5)P3), [Ca2+]i and adenosine 3′ : 5′-cyclic monophosphate (cyclic AMP) formation in Chinese hamster ovarian (CHO) cells transfected with the cloned μ-opioid receptor.
  2. Opioid receptor binding was assessed with [3H]-diprenorphine ([3H]-DPN) as a radiolabel. Ins(1,4,5)P3 and cyclic AMP were measured by specific radioreceptor assays. [Ca2+]i was measured fluorimetrically with Fura-2.
  3. Scatchard analysis of [3H]-DPN binding revealed that the Bmax varied between passages. Fentanyl (10 pM–1 μM) dose-dependently displaced [3H]-DPN, yielding a curve which had a Hill slope of less than unity (0.6±0.1), and was best fit to a two site model, with pKi values (% of sites) of 9.97±0.4 (27±4.8%) and 7.68±0.07 (73±4.8%). In the presence of GppNHp (100 μM) and Na+ (100 mM), the curve was shifted to the right and became steeper (Hill slope=0.9±0.1) with a pKi value of 6.76±0.04.
  4. Fentanyl (0.1 nM–1 μM) had no effect on basal, but dose-dependently inhibited forskolin (1 μM)-stimulated, cyclic AMP formation (pIC50=7.42±0.23), in a pertussis toxin (PTX; 100 ng ml−1 for 24 h)-sensitive and naloxone-reversible manner (Ki=1.7 nM). Morphine (1 μM) and [D-Ala2, MePhe4, gly(ol)5]-enkephalin (DAMGO, 1 μM) also inhibited forskolin (1 μM)-stimulated cyclic AMP formation, whilst [D-Pen2, D-Pen5], enkephalin (DPDPE, 1 μM) did not.
  5. Fentanyl (0.1 nM–10 μM) caused a naloxone (1 μM)-reversible, dose-dependent stimulation of Ins(1,4,5)P3 formation, with a pEC50 of 7.95±0.15 (n=5). PTX (100 ng ml−1 for 24 h) abolished, whilst Ni2+ (2.5 mM) inhibited (by 52%), the fentanyl-induced Ins(1,4,5)P3 response. Morphine (1 μM) and DAMGO (1 μM), but not DPDPE (1 μM), also stimulated Ins(1,4,5)P3 formation. Fentanyl (1 μM) also caused an increase in [Ca2+]i (80±16.4 nM, n=6), reaching a maximum at 26.8±2.5 s. The increase in [Ca2+]i remained elevated until sampling ended (200 s) and was essentially abolished by the addition of naloxone (1 μM). Pre-incubation with naloxone (1 μM, 3 min) completely abolished fentanyl-induced increases in [Ca2+]i.
  6. In conclusion, the cloned μ-opioid receptor when expressed in CHO cells stimulates PLC and inhibits AC, both effects being mediated by a PTX-sensitive G-protein. In addition, the receptor couples to an increase in [Ca2+]i. These findings are consistent with the previously described effector-second messenger coupling of the endogenous μ-opioid receptor.
  相似文献   

16.
  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   

17.
  1. The effects of the lipoxygenase inhibitor nordihydroguaiaretic acid (NDGA) on the ionic currents of rat carotid body type I cells were investigated by use of whole-cell and outside-out patch clamp techniques.
  2. NDGA (5–50 μM) produced a concentration-dependent inhibition of whole-cell K+ currents at all activating test potentials (holding potential −70 mV). The time-course of the inhibition was also concentration-dependent and the effects of NDGA were only reversible following brief periods of exposure (<2 min). Another lipoxygenase inhibitor, phenidone (5 μM), was without effect on whole-cell K+ currents in carotid body type I cells.
  3. NDGA (5–50 μM) also inhibited whole-cell Ca2+ channel currents (recorded with Ba2+ as charge carrier) in a concentration-dependent manner.
  4. Isolation of voltage-gated K+ channels by use of high [Mg2+] (6 mM), low [Ca2+] (0.1 mM) solutions revealed a direct inhibition of the voltage-sensitive component of the whole-cell K+ current by NDGA (50 μM).
  5. In excised, outside-out patches NDGA (20–50 μM) increased large conductance, Ca2+ activated K+ channel activity approximately 10 fold, an effect which could be reversed by either tetraethylammonium (10 mM) or charybdotoxin (30 nM).
  6. It is concluded that NDGA activates maxi-K+ channels in carotid body type I cells and over the same concentration range inhibits voltage-sensitive K+ and Ca2+ channels. The inhibition of whole cell K+ currents seen is most likely due to a combination of direct inhibition of the voltage-sensitive K+ current and indirect inhibition of maxi-K+ channel activity through blockade of Ca2+ channels.
  相似文献   

18.
  1. We have used whole-cell patch clamping methods to study and characterize the cytolytic P2X7 (P2Z) receptor in the NTW8 mouse microglial cell line.
  2. At room temperature, in an extracellular solution containing 2 mM Ca2+ and 1 mM Mg2+, 2′- and 3′-O-(4-benzoylbenzoyl)-adenosine-5′-triphosphate (Bz-ATP; 300 μM), or ATP (3 mM), evoked peak whole cell inward currents, at a holding potential of −90 mV, of 549±191 and 644±198 pA, respectively. Current-voltage relationships generated with 3 mM ATP reversed at 4.6 mV and did not display strong rectification.
  3. In an extracellular solution containing zero Mg2+ and 500 μM Ca2+ (low divalent solution), brief (0.5 s) application of these agonists elicited larger maximal currents (909±138 and 1818±218 pA, Bz-ATP and ATP, respectively). Longer application of ATP (1 mM for 30 s) produced larger, slowly developing, currents which reached a plateau after approximately 15–20 s and were reversible on washing. Under these conditions, in the presence of ATP, ethidium bromide uptake could be demonstrated. Further applictions of 1 mM ATP produced rapid currents of the same magnitude as those observed during the 30 s application. Subsequent determination of concentration-effect curves to Bz-ATP, ATP and 2-methylthio-ATP yielded EC50 values of 58.3, 298 and 505 μM, respectively. These affects of ATP were antagonized by pyridoxal-phosphate-6-azophenyl- 2′, 4′-disulphonic acid (PPADS; 30 μM) but not suramin (100 μM).
  4. In low divalent solution, repeated application of 1 mM ATP for 1 s produced successively larger currents which reached a plateau, after 8 applications, of 466% of the first application current. PPADS (30 μM) prevented this augmentation, while 5-(N,N-hexamethylene)-amiloride (HMA) (100 μM) accelerated it such that maximal augmentation was observed after only one application of ATP in the presence of HMA. At a bath temperature of 32°C, current augmentation also occurred in normal divalent cation containing solution.
  5. These data demonstrate that mouse microglial NTW8 cells possess a purinoceptor with pharmacological characteristics resembling the P2X7 receptor. We suggest that the current augmentation phenomenon observed reflects formation of the large cytolytic pore characteristic of this receptor. We have demonstrated that pore formation can occur under normal physiological conditions and can be modulated pharmacologically, both positively and negatively.
  相似文献   

19.
  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   

20.
  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   

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