首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 437 毫秒
1.
  1. The effect of protein tyrosine kinase inhibitors on human adenosine A1 receptor-mediated [3H]-inositol phosphate ([3H]-IP) accumulation has been studied in transfected Chinese hamster ovary cells (CHO-A1) cells.
  2. In agreement with our previous studies the selective adenosine A1 receptor agonist N6-cyclopentyladenosine (CPA) stimulated the accumulation of [3H]-IPs in CHO-A1 cells. Pre-treatment with the broad spectrum tyrosine kinase inhibitor genistein (100 μM; 30 min) potentiated the responses elicited by 1 μM (199±17% of control CPA response) and 10 μM CPA (234±15%). Similarly, tyrphostin A47 (100 μM) potentiated the accumulation of [3H]-IPs elicited by 1 μM CPA (280±32%).
  3. Genistein (EC50=13.7±1.2 μM) and tyrphostin A47 (EC50=10.4±3.9 μM) potentiated the [3H]-IP response to 1 μM CPA in a concentration-dependent manner.
  4. Pre-incubation with the inactive analogues of genistein and tyrphostin A47, daidzein (100 μM; 30 min) and tyrphostin A1 (100 μM; 30 min), respectively, had no significant effect on the accumulation of [3H]-IPs elicited by 1 μM CPA.
  5. Genistein (100 μM) had no significant effect on the accumulation of [3H]-IPs produced by the endogenous thrombin receptor (1 u ml−1; 100±10% of control response). In contrast, tyrphostin A47 produced a small augmentation of the thrombin [3H]-IP response (148±13%).
  6. Genistein (100 μM) had no effect on the [3H]-IP response produced by activation of the endogenous Gq-protein coupled CCKA receptor with the sulphated C-terminal octapeptide of cholecystokinin (1 μM CCK-8; 96±6% of control). In contrast, tyrphostin A47 (100 μM) caused a small but significant increase in the response to 1 μM CCK-8 (113±3% of control).
  7. The phosphatidylinositol 3-kinase inhibitor LY 294002 (30 μM) and the MAP kinase kinase inhibitor PD 98059 (50 μM) had no significant effect on the [3H]-IP responses produced by 1 μM CPA and 1 μM CCK-8.
  8. These observations suggest that a tyrosine kinase-dependent pathway may be involved in the regulation of human adenosine A1 receptor mediated [3H]-IP responses in CHO-A1 cells.
  相似文献   

2.
  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   

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

4.
  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   

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

6.
  1. The mitogen-activated protein (MAP) kinase signalling pathway can be activated by a variety of heterotrimeric Gi/Go protein-coupled and Gq/G11 protein-coupled receptors. The aims of the current study were: (i) to investigate whether the Gi/Go protein-coupled adenosine A1 receptor activates the MAP kinase pathway in transfected Chinese hamster ovary cells (CHO-A1) and (ii) to determine whether adenosine A1 receptor activation would modulate the MAP kinase response elicited by the endogenous P2Y2 purinoceptor.
  2. The selective adenosine A1 receptor agonist N6-cyclopentyladenosine (CPA) stimulated time and concentration-dependent increases in MAP kinase activity in CHO-A1 cells (EC50 7.1±0.4 nM). CPA-mediated increases in MAP kinase activity were blocked by PD 98059 (50 μM; 89±4% inhibition), an inhibitor of MAP kinase kinase 1 (MEKI) activation, and by pre-treating cells with pertussis toxin (to block Gi/Go-dependent pathways).
  3. Adenosine A1 receptor-mediated activation of MAP kinase was abolished by pre-treatment with the protein tyrosine inhibitor, genistein (100 μM; 6±10% of control). In contrast, daidzein (100 μM), the inactive analogue of genistein had no significant effect (96±12 of control). MAP kinase responses to CPA (1 μM) were also sensitive to the phosphatidylinositol 3-kinase inhibitors wortmannin (100 nM; 55±8% inhibition) and LY 294002 (30 μM; 40±5% inhibition) but not to the protein kinase C (PKC) inhibitor Ro 31-8220 (10 μM).
  4. Activation of the endogenous P2Y2 purinoceptor with UTP also stimulated time and concentration-dependent increases in MAP kinase activity in CHO-A1 cells (EC50=1.6±0.3 μM). The MAP kinase response to UTP was partially blocked by pertussis toxin (67±3% inhibition) and by the PKC inhibitor Ro 31-8220 (10 μM; 45±5% inhibition), indicating the possible involvement of both Gi/Go protein and Gq protein-dependent pathways in the overall response to UTP.
  5. CPA and UTP stimulated concentration-dependent increases in the phosphorylation state of the 42 kDa and 44 kDa forms of MAP kinase as demonstrated by Western blotting.
  6. Co-activation of CHO-A1 cells with CPA (10 nM) and UTP (1 μM) produced synergistic increases in MAP kinase activity which were not blocked by the PKC inhibitor Ro 31-8220 (10 μM).
  7. Adenosine A1 and P2Y2 purinoceptor activation increased the expression of luciferase in CHO cells transfected with a luciferase reporter gene containing the c-fos promoter. However, co-activating these two receptors produced only additive increases in luciferase expression.
  8. In conclusion, our studies have shown that the transfected adenosine A1 receptor and the endogenous P2Y2 purinoceptor couple to the MAP kinase signalling pathway in CHO-A1 cells. Furthermore, co-stimulation of the adenosine A1 receptor and the P2Y2 purinoceptor produced synergistic increases in MAP kinase activity but not c-fos mediated luciferase expression.
  相似文献   

7.
  1. The aim of this study was to determine the conditions under which the α2-adrenoceptor agonist UK14304 produces vasoconstriction in the porcine isolated ear artery.
  2. UK14304 (0.3 μM) produced a small contraction of porcine isolated ear arteries which was 7.8±3.3% of the response to 60 mM KC1. Similar sized contractions were obtained after precontraction with either 30 nM angiotensin II, or 0.1 μM U46619 (8.2±1.8% and 10.2±2.6% of 60 mM KC1 response, respectively). However, an enhanced α2-adrenoceptor response was uncovered if the tissue was precontracted with U46619, and relaxed back to baseline with 1–2 μM forskolin before the addition of UK14304 (46.9±9.6% of 60 mM KC1 response).
  3. The enhanced responses to UK14304 in the presence of U46619 and forskolin were not inhibited by the α1-adrenoceptor antagonist prazosin (0.1 μM), but were inhibited by the α2-adrenoceptor antagonist rauwolscine (1 μM), indicating that the enhanced responses were mediated via postjunctional α2-adrenoceptors.
  4. In the presence of 0.1 μM U46619 and 1 mM isobutylmethylxanthine (IBMX), 1 μM forskolin produced an increase in [3H]-cyclic AMP levels in porcine isolated ear arteries. Addition of 0.3 μM UK14304 prevented this increase.
  5. The enhanced UK14304 response was dependent upon the agent used to relax the tissue. After relaxation of ear arteries precontracted with 10 nM U46619 and relaxed with forskolin the UK14304 response was 46.9±9.6% of the 60 mM KC1 response, and after relaxation with sodium nitroprusside (SNP) the response was 24.8±3.3%. However, after relaxation of the tissue with levcromakalim the UK14304 response was only 8.2±1.7%, which was not different from the control response in the same tissues (12.2±5.6%). An enhanced contraction was also obtained after relaxation of the tissue with the cyclic AMP analogue dibutyryl cyclic AMP (23.2±1.3%) indicating that at least part of the enhanced response to UK14304 is independent of the ability of the agonist to inhibit cyclic AMP production.
  6. Relaxation of U46619 contracted ear arteries with SNP could be inhibited by the NO-sensitive guanylyl-cyclase inhibitor 1H-[1,2,4] oxadiazolo[4,3-a]quinoxalin-1-one (ODQ) indicating that production of cyclic GMP is necessary for the relaxant effect of SNP. However, ODQ had no effect on the relaxation of tissue by forskolin, suggesting that this compound does not act via production of cyclic GMP. Biochemical studies showed that while forskolin increases the levels of cyclic AMP in the tissues, SNP had no effect on the levels of this cyclic nucleotide.
  7. In conclusion, enhanced contractions to the α2-adrenoceptor agonist UK14304 can be uncovered in porcine isolated ear arteries by precontracting the tissue with U46619, followed by relaxation back to baseline with forskolin, SNP or dibutyryl cyclic AMP before addition of UK14304. There was a greater contractile response to UK14304 after relaxation with forskolin than with SNP or dibutyryl cyclic AMP, suggesting that cyclic AMP-dependent and- independent mechanisms are involved in the enhancement of the UK14304 response.
  相似文献   

8.
  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   

9.
  1. The release of endogenous γ-aminobutyric acid (GABA) and glutamic acid in the human brain has been investigated in synaptosomal preparations from fresh neocortical samples obtained from patients undergoing neurosurgery to reach deeply located tumours.
  2. The basal outflows of GABA and glutamate from superfused synaptosomes were largely increased during depolarization with 15 mM KCl. The K+-evoked overflows of both amino acids were almost totally dependent on the presence of Ca2+ in the superfusion medium.
  3. The GABAB receptor agonist (−)-baclofen (1, 3 or 10 μM) inhibited the overflows of GABA and glutamate in a concentration-dependent manner. The inhibition caused by 10 μM of the agonist ranged from 45–50%.
  4. The effect of three selective GABAB receptor antagonists on the inhibition of the K+-evoked GABA and glutamate overflows elicited by 10 μM (−)-baclofen was investigated. Phaclofen antagonized (by about 50% at 100 μM; almost totally at 300 μM) the effect of (−)-baclofen on GABA overflow but did not modify the inhibition of glutamate release. The effect of (−)-baclofen on the K+-evoked GABA overflow was unaffected by 3-amino-propyl (diethoxymethyl)phosphinic acid (CGP 35348; 10 or 100 μM); however, CGP 35348 (10 or 100 μM) antagonized (−)-baclofen (complete blockade at 100 μM) at the heteroreceptors on glutamatergic terminals. Finally, [3-[[(3,4-dichlorophenyl) methyl]amino]propyl] (diethoxymethyl) phosphinic aid (CGP 52432), 1 μM, blocked the GABAB autoreceptor, but was ineffective at the heteroreceptors. The selectivity of CGP 52432 was lost at 30 μM, as the compound, at this concentration, inhibited completely the (−)-baclofen effect both on GABA and glutamate release.
  5. It is concluded that GABA and glutamate release evoked by depolarization of human neocortex nerve terminals can be affected differentially through pharmacologically distinct GABAB receptors.
  相似文献   

10.
  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   

11.
  1. The effects of adenosine receptor agonists upon phenylephrine-stimulated contractility and [3H]-cyclic adenosine monophosphate ([3H]-cyclic AMP) accumulation in the cauda epididymis of the guinea-pig were investigated. The α1-adrenoceptor agonist, phenylephrine elicited concentration dependent contractile responses from preparations of epididymis. In the absence or presence of the L-type Ca2+ channel blocker, nifedipine (10 μM) the non-selective adenosine receptor agonist, 5′-N-ethylcarboxamido-adenosine (NECA, 1 μM) shifted phenylephrine concentration-response curves to the left (4 and 5 fold respectively). Following the incubation of preparations with pertussis toxin (200 ng ml−1 24 h) NECA shifted phenylephrine concentration-response curves to the right (5.7±0.9 fold).
  2. In the presence of phenylephrine (1 μM), NECA and the A1 adenosine receptor selective agonists, N6-cyclopentyladenosine (CPA) and (2S)-N6-[2-endo-norbornyl]adenosine ((S)-ENBA) elicited concentration-responses dependent contractions from preparations of epididymis (pEC50 values 8.18±0.19, 7.79±0.29 and 8.15±0.43 respectively). The A3 adenosine receptor agonists N6-iodobenzyl-5′-N-methyl-carboxamido adenosine (IBMECA) and N6-2-(4-aminophenyl) ethyladenosine (APNEA) mimicked this effect (but only at concentrations greater than 10 μM). In the presence of 8-cyclopentyl-1,3-dipropylxanthine (DPCPX, 30 nM) CPA concentration-response curves were shifted, in parallel to the right (apparent pKB 8.75±0.88) and the maximal response to NECA was reduced.
  3. In the presence of DPCPX (100 nM) the adenosine agonist NECA and the A2A adenosine receptor selective agonist, CGS 21680 (2-p-(2-carboxyethyl)-phenethylamino-N-ethylcarboxamido adenosine), but not CPA, inhibited phenylephrine (20 μM) stimulated contractions (pIC50 7.15±0.48). This effect of NECA was blocked by xanthine amine congener (XAC, 1 μM) and the A2A adenosine receptor-selective antagonist 4-(2-[7-amino-2-(2-furyl)[1,2,4]triazolo[2,3-a][1,3,5]triazin-5-ylamino]ethyl)phenol (ZM 241385; 30 nM).
  4. (S)-ENBA (in the absence and presence of ZM 241385, 100 nM), but not NECA or CPA inhibited the forskolin (30 μM)-stimulated accumulation of [3H]-cyclic AMP in preparations of the epididymis of the guinea-pig (by 17±6% of control). In the presence of DPCPX (100 nM) NECA and CGS 21680, but not (S)-ENBA, increased the accumulation of [3H]-cyclic AMP in preparations of epididymis (pEC50 values 5.35±0.35 and 6.42±0.40 respectively), the NECA-induced elevation of [3H]-cyclic AMP was antagonised by XAC (apparent pKB 6.88±0.88) and also by the A2A adenosine receptor antagonist, ZM 241385 (apparent pKB 8.60± 0.76).
  5. These studies are consistent with the action of stable adenosine analogues at post-junctional A1 and A2 adenosine receptors in the epididymis of the guinea-pig. A1 Adenosine receptors potentiate α1-adrenoceptor contractility, an effect blocked by pertussis toxin, but which may not be dependent upon an inhibition of adenylyl cyclase. The epididymis of the guinea-pig also contains A2 adenosine receptors, possibly of the A2A subtype, which both inhibit contractility and also stimulate adenylyl cyclase.
  相似文献   

12.
  1. Extracellular single unit recording techniques were used to compare the effects of (-)-3-amino-1-hydroxypyrrolidin-2-one ((–)-HA-966) and (±)-baclofen on the activity of dopamine-containing neurones in 300 μm slices of rat substantia nigra. Electrophysiological data were compared with the outcome of in vitro binding experiments designed to assess the affinity of (–)-HA-966 for γ-aminobutyric acid (GABAB) receptors.
  2. Bath application of (–)-HA-966 produced a concentration-dependent inhibition of dopaminergic neuronal firing (EC50=444.0 μM; 95% confidence interval: 277.6 μM–710.1 μM, n=27) which was fully reversible upon washout from the recording chamber. Although similar effects were observed in response to (±)-baclofen, the direct-acting GABAB receptor agonist proved to be considerably more potent than (–)-HA-966 (EC50=0.54 μM; 95% confidence interval: 0.44 μM–0.66 μM, n=29) in vitro.
  3. Low concentrations of chloral hydrate (10 μM) were without effect on the basal firing rate of nigral dopaminergic neurones but significantly increased the inhibitory effects produced by concomitant application of (–)-HA-966.
  4. The inhibitory effects of (–)-HA-966 were completely reversed in the presence of the GABAB receptor antagonists, CGP-35348 (100 μM) and 2-hydroxysaclofen (500 μM). Bath application of CGP-35348 alone increased basal firing rate. However, the magnitude of the excitation (9.2±0.3%) was not sufficient to account for the ability of the antagonist to reverse fully the inhibitory effects of (–)-HA-966.
  5. (–)-HA-966 (0.1–1.0 mM) produced a concentration-dependent displacement of [H]-GABA from synaptic membranes in the presence of isoguvacine (40 μM). However, the affinity of the drug for GABAB binding sites was significantly less than that of GABA (0.0005 potency ratio) and showed no apparent stereoselectivity.
  6. These results indicate that while (–)-HA-966 appears to act as a direct GABAB receptor agonist in vitro, its affinity for this receptor site is substantially less than that of GABA or baclofen and unlikely to account for the depressant actions of this drug which occur at levels approximately ten fold lower in vivo.
  相似文献   

13.
  1. The effects of adenosine receptor agonists upon both electrically-evoked and phenylephrine-induced contractile responses were investigated in the bisected vas deferens and the cauda epididymis of the guinea-pig. Electrical field-stimulation (10 s trains of pulses at 9 Hz, 0.1 ms duration, supramaximal voltage) elicited biphasic and monophasic contractile responses from preparations of bisected vas deferens and cauda epididymis, respectively; these responses were abolished by tetrodotoxin (300 nM).
  2. In the prostatic half of the vas deferens the A1 selective adenosine receptor agonists, N6-cyclopentyladenosine (CPA) and (2S)-N6-[2-endo-norbornyl]adenosine ((S)-ENBA) and the non-selective A1/A2 adenosine receptor agonist, 5′-N-ethylcarboxamidoadenosine (NECA) inhibited electrically-evoked contractions (pIC50±s.e.mean values 6.15±0.24, 5.99±0.26 and 5.51±0.24, respectively). The responses to CPA were blocked by the A1 adenosine receptor antagonist, 8-cyclopentyl-1,3-dipropylxanthine, DPCPX (100 nM).
  3. In the epididymal half of the vas deferens NECA potentiated (at ⩽100 nM) and inhibited (at ⩾1 μM) electrically-evoked contractions. In the presence of the non-selective α-adrenoceptor antagonist phentolamine (3 μM), the α1-adrenoceptor antagonist, prazosin (100 nM), or at a reduced train length (3 s) NECA inhibited electrically-evoked contractions (pIC50 values 6.05±0.25, 5.97±0.29 and 5.71±0.27, respectively). CPA (at 10 μM) also inhibited electrically-evoked contractions in this half of the vas deferens. In the presence of prazosin (100 nM), CPA also inhibited electrically-evoked contractions (pIC50 6.14±0.67); this effect was antagonized by DPCPX (30 nM, apparent pKB 8.26±0.88). In the presence of the P2 purinoceptor antagonist, suramin (300 μM), CPA (up to 1 μM) potentiated electrically-evoked contractions.
  4. NECA, CPA and APNEA potentiated electrically-evoked contractions in preparations of cauda epididymis (pEC50 values 7.49±0.62, 7.65±0.74 and 5.84±0.86, respectively), the response to CPA was competitively antagonized by DPCPX (100 nM) with an apparent pKB value of 7.64±0.64.
  5. The α1-adrenoceptor agonist phenylephrine elicited concentration-dependent contractile responses from preparations of bisected vas deferens and cauda epididymis. NECA (1 μM) potentiated responses to phenylephrine (⩽1 μM) in the epididymal, but not in the prostatic half of the vas deferens. In preparations of epididymis NECA (1 μM) shifted phenylephrine concentration response curves to the left (4.6 fold). In the presence of a fixed concentration of phenylephrine (1 μM), NECA elicited concentration-dependent contractions of preparations of the epididymal half of the vas deferens and of the epididymis (pEC50 values 7.57±0.54 and 8.08±0.18, respectively). NECA did not potentiate responses to ATP in either the epididymal half of the vas deferens or the epididymis.
  6. These studies are consistent with the action of stable adenosine analogues at prejunctional A1 and postjunctional A1-like adenosine receptors. The prejunctional A1 adenosine receptors only inhibit the electrically-evoked contractions of purinergic origin (an effect predominant in the prostatic half of the vas deferens). At the epididymis, where electrically-evoked contractions are entirely adrenergic, the predominant adenosine receptor agonist effect is a potentiation of α1-adrenoceptor-, but not of ATP-induced contractility.
  相似文献   

14.
  1. γ-Aminobutyric acid (GABA), acting via the GABAA receptor, increased the extracellular acidification rate of rat primary cultured cerebellar granule cells, measured by the Cytosensor microphysiometer.
  2. The optimal conditions for the measurement of GABAA receptor function in cerebellar granule cells by microphysiometry were: cells seeded at 9–12×105 cells/transwell cup and maintained in vitro for 8 days, GABA stimulation performed at 25°C, with a stimulation time of 33 s.
  3. GABA stimulated a concentration-dependent increase in the extracellular acidification rate with an EC50 of 2.0±0.2 μM (mean±s.e.mean, n=7 experiments) and maximal increase (Emax) over basal response of 15.4±1.2%.
  4. The sub-maximal GABA-stimulated increase in acidification rate could be potentiated by the 1,4-benzodiazepine, flunitrazepam (100 nM). The 10 nM GABA response showed the maximal benzodiazepine facilitation (GABA alone, 1.4 μV s−1, GABA+flunitrazepam, 3.8 μV s−1, mean increment over basal, n=7).
  5. The GABA-stimulated increase in acidification rate was inhibited by the GABAA antagonist, bicuculline (100 μM) (90% inhibition at 1 mM GABA).
  6. The results of this study show that activation of GABAA receptors in rat cerebellar granule cells caused an increase in the extracellular acidification rate; an effect which was potentiated by benzodiazepines and inhibited by a GABAA receptor antagonist. This paper defines the conditions and confirms the feasibility of using microphysiometry to investigate GABAA receptor function in primary cultured CNS neurones. The microphysiometer provides a rapid and sensitive technique to investigate the regulation of the GABAA receptor in populations of neurones.
  相似文献   

15.
  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   

16.
  1. In human U373 MG astrocytoma cells agonist-induced increases in intracellular Ca2+ ([Ca2+]i) are rapidly returned towards prestimulated levels. Examination of the effect of histamine and substance P on [Ca2+]i in thapsigargin-treated cells has allowed a mechanism contributing to this effect to be characterized.
  2. Histamine and substance P stimulated [3H]-inositol monophosphate ([3H]-IP1) accumulation in U373 MG cells. Concentration-response curves of [3H]-IP1 accumulation in suspensions of U373 MG cells in HEPES buffer containing 30 mM Li+ yielded best-fit EC50 values of 19.1±1.5 μM for histamine and 5.7±1.3 nM for substance P.
  3. In confluent monolayers of fura-2 loaded U373 MG cells perfusion with 100 μM histamine resulted in a transient 597±50 nM increase in [Ca2+]i. The best-fit EC50 for histamine was 4.6±2.2 μM. The initial, transient, histamine response was often followed by further small transient increases in [Ca2+]i.
  4. Treatment of U373 MG cells with 5 μM thapsigargin, followed by the readdition of 1.8 mM Ca2+ to the perfusion buffer, resulted in a steady-state level of [Ca2+]i 97±5 nM above pretreated levels (measured 400 s after readdition of Ca2+). Perfusion of histamine (100 μM, 100 s) caused a rapid decline in the thapsigargin-induced steady state level of [Ca2+]i. This effect of histamine was normally reversible upon washout. The best-fit EC50 for the histamine response was 0.8±0.2 μM. Substance P (10 nM, 100 s) also caused a reduction in thapsigargin-induced steady-state levels of [Ca2+]i.
  5. Neither 100 μM histamine nor 10 nM substance P inhibited the rate of quench of fura-2 fluorescence by Mn2+ in U373 MG cells pretreated with 5 μM thapsigargin, indicating that the depressant effect on steady-state raised [Ca2+]i was probably not due to a block of Ca2+ entry.
  6. The depressant effect of histamine on [Ca2+]i was blocked by 1 μM mepyramine, and was partially reduced by pre-incubation with 1 μM staurosporine (61±7% reduction) and with Ro 31-8220 (24±10% and 50±6% reduction by 1 and 10 μM Ro 31-8220, respectively). Pre-incubation with H-89 did not alter the depressant effect of histamine.
  7. Neither 1 μM staurosporine nor 10 μM KN-62 inhibited the binding of [3H]-mepyramine to guinea-pig cerebellar membranes, whereas it was reduced by 17±1% and 55±2% by 1 and 10 μM Ro 31-8220, respectively. However, [3H]-IP1 accumulation stimulated by histamine in U373 MG cells was not inhibited by 1 or 10 μM Ro 31-8220 and in 2 out of 3 experiments there was a significant potentiation of the response to histamine with both concentrations of Ro 31-8220. Staurosporine, 1 μM, similarly potentiated the response to 100 μM histamine in 3 out of 4 experiments. KN-62 (10 μM) did not stimulate histamine-induced [3H]-IP1 accumulation.
  8. In HEPES buffer to which no Ca2+ had been added, histamine stimulated a transient 451±107 nM increase in [Ca2+]i. Pretreatment with 1 μM and 10 μM Ro 31-8220 did not significantly alter the initial peak response to histamine, but slowed the rate at which histamine-induced increases in [Ca2+]i were returned to prestimulated levels. Pretreatment with KN-62 had no significant effect on the response to histamine, but consistently inhibited the secondary slower phase of the decline in [Ca2+]i. H-89 did not alter the histamine response.
  9. The effect of histamine in stimulating Ca2+ extrusion was not confined to U373 MG cells, since 100 μM histamine also caused a rapid decrease in steady-state levels of [Ca2+]i in thapsigargin-treated human HeLa cells.
  10. The results indicate that agonists which increase [Ca2+]i via activation of phosphoinositide metabolism can also stimulate a homeostatic mechanism which acts to reduce [Ca2+]i. The balance of the evidence indicates that in U373 MG cells the latter effect most likely involves a PKC-mediated stimulation of a Ca2+-extrusion pump.
  相似文献   

17.
  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   

18.
  1. The mechanism of action of P2 nucleotide receptor agonists that produce endothelium-independent relaxation and the influence of ecto-ATPase activity on this relaxing effect have been investigated in rat portal vein smooth muscle.
  2. At 25°C, ATP, 2-methylthioATP (2-MeSATP) and 2-chloroATP (2-ClATP), dose-dependently inhibited spontaneous contractile activity of endothelium-denuded muscular strips from rat portal vein. The rank order of agonist potency defined from the half-inhibitory concentrations was 2-ClATP (2.7±0.5 μM, n=7)>ATP (12.9±1.1 μM, n=9)⩾2-MeSATP (21.9±4.8 μM, n=4). In the presence of αβ-methylene ATP (αβ-MeATP, 200 μM) which itself produced a transient contractile effect, the relaxing action of ATP and 2-MeSATP was completely abolished and that of 2-ClATP strongly inhibited.
  3. The non-selective P2-receptor antagonist pyridoxalphosphate-6-azophenyl-2′,4′-disulphonic acid (PPADS, 100 μM) did not affect the relaxation induced by ATP, 2-MeSATP, and 2-ClATP.
  4. The A2A-adenosine receptor antagonist ZM 241385 inhibited the ATP-induced relaxation in a concentration-dependent manner (1–100 nM). In the presence of 100 nM ZM 241385, the relaxing effects of 2-MeSATP and 2-ClATP were also inhibited.
  5. ADP, AMP and adenosine also produced concentration-dependent inhibition of spontaneous contractions. The relaxing effects of AMP and adenosine were insensitive to αβ-MeATP (200 μM) but were inhibited by ZM 241385 (100 nM).
  6. Simultaneous measurements of contraction and ecto-ATPase activity estimated by the degradation of [γ-32P]-ATP showed that muscular strips rapidly (10–60 s) hydrolyzed ATP. This ecto-ATPase activity was abolished in the presence of EDTA and was inhibited by 57±11% (n=3) by 200 μM αβ-MeATP.
  7. These results suggest that ATP and other P2-receptor agonists are relaxant in rat portal vein smooth muscle, because ectonucleotidase activity leads to the formation of adenosine which activates A2A-receptors.
  相似文献   

19.
  1. This study was performed to characterize the receptor subtypes involved in the endothelin stimulation of phospholipase D (PLD) in rat cortical astrocytes in primary culture. PLD activity was determined by measuring the formation of [32P]phosphatidylbutanol in [32P]orthophosphate prelabelled cells stimulated in the presence of 25 mM butanol.
  2. The agonists endothelin-1 (ET-1), endothelin-3 (ET-3), sarafotoxin 6c (S6c) and IRL 1620 elicited PLD activation in a concentration-dependent manner. The potencies of ET-1, ET-3 and S6c were similar. The maximal effects evoked by the ETB-preferring agonists, ET-3, S6c and IRL 1620, were significantly lower than the maximal response to the non-selective agonist ET-1.
  3. The response to 1 nM ET-1 was inhibited by increasing concentrations of the ETA receptor antagonist BQ-123 in a biphasic manner. A high potency component of the inhibition curve (24.2±3.5% of the ET-1 response) was defined at low (up to 1 μM) concentrations of BQ-123, yielding an estimated Ki value for BQ-123 of 21.3±2.5 nM. In addition, the presence of 1 μM BQ-123 significantly reduced the maximal response to ET-1 but did not change the pD2 value.
  4. Increasing concentrations of the ETB selective antagonist BQ-788 inhibited the S6c response with a Ki of 17.8±0.8 nM. BQ-788 also inhibited the effect of ET-1, although, in this case, two components were defined, accounting for approximately 50% of the response, and showing Ki values of 20.9±5.1 nM and 439±110 nM, respectively. The ET-1 concentration-response curve was shifted to the right by 1 μM BQ-788, also revealing two components. Only one of them, corresponding to 69.8±4.4% of the response, was sensitive to BQ-788 which showed a Ki value of 28.8±8.9 nM.
  5. Rapid desensitization was achieved by preincubation with ET-1 or S6c. In cells pretreated with S6c neither ET-3 nor S6c activated PLD, but ET-1 still induced approximately 40% of the response shown by non-desensitised cells. This remaining response was insensitive to BQ-788, but fully inhibited by BQ-123.
  6. In conclusion, endothelins activate PLD in rat cortical astrocytes acting through both ETA and ETB receptors, and this response desensitizes rapidly in an apparently homologous fashion. The percentage contribution of ETA and ETB receptors to the ET-1 response was found to be approximately 20% and 80%, respectively, when ETB receptors were not blocked, and 30–50% and 50–70%, respectively, when ETB receptors were inhibited or desensitized. These results may be relevant to the study of a possible role of PLD in the proliferative effects shown by endothelins on cultured and reactive astrocytes.
  相似文献   

20.
  1. The present study has utilized the two electrode voltage-clamp technique to examine the pharmacological profile of a splice variant of the rat orthologue of the 5-hydroxytryptamine type 3A subunit (5-HT3A(b)) heterologously expressed in Xenopus laevis oocytes.
  2. At negative holding potentials, bath applied 5-HT (300 nM–10 μM) evoked a transient, concentration-dependent (EC50=1.1±0.1 μM), inward current. The response reversed in sign at a holding potential of −2.1±1.6 mV.
  3. The response to 5-HT was mimicked by the 5-HT3 receptor selective agonists 2-methyl-5-HT (EC50=4.1±0.2 μM), 1-phenylbiguanide (EC50=3.0±0.1 μM), 3-chlorophenylbiguanide (EC50=140± 10 nM), 3,5-dichlorophenylbiguanide (EC50=14.5±0.4 nM) and 2,5-dichlorophenylbiguanide (EC50= 10.2±0.6 nM). With the exception of 2-methyl-5-HT, all of the agonists tested elicited maximal current responses comparable to those produced by a saturating concentration (10 μM) of 5-HT.
  4. Responses evoked by 5-HT at EC50 were blocked by the 5-HT3 receptor selective antagonist ondansetron (IC50=231±22 pM) and by the less selective agents (+)-tubocurarine (IC50=31.9± 0.01 nM) and cocaine (IC50=2.1±0.2 μM).
  5. The data are discussed in the context of results previously obtained with the human and mouse orthologues of the 5-HT3A subunit. Overall, the study reinforces the conclusion that species differences detected for native 5-HT3 receptors extend to, and appear largely explained by, differences in the properties of homo-oligomeric receptors formed from 5-HT3A subunit orthologues.
  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-23

京公网安备 11010802026262号