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.
The beta2-adrenoceptor (beta2AR) fused to the long splice variant of G(s)alpha (G(s)alphaL), but not the beta2AR fused to the short splice variant of G(s)alpha (G(s)alphaS) shows the hallmarks of high constitutive activity, i.e., strong activation of adenylyl cyclase (AC) by GTP and strong inhibition of AC by inverse agonist. These coupling differences are the result of differences in GDP affinity of G(s)alpha splice variants. The aim of this study was to identify experimental variables that differentially affect beta2AR coupling to G(s)alphaS and G(s)alphaL. NaCl substantially reduced agonist-independent AC activation by GTP and inverse agonist inhibition and enhanced agonist stimulation of AC in Sf9 insect cell membranes expressing the beta2AR-G(s)alphaL fusion protein. Salts reduced inverse agonist inhibition and increased agonist stimulation of AC in the order of efficiency NaI approximately KI > NaBr approximately KBr > NaCl approximately LiCl approximately KCl approximately RbCl approximately CsCl approximately choline chloride, indicating that monovalent anions determine salt effects. Salts inhibited guanosine 5'-O-(3-thiotriphosphate)-mediated AC activation by G(salphaL) without beta2AR in the order of efficiency NaI > NaBr > NaCl. NaCl enhanced the affinity of G(s)alphaL for GDP. Salts had much smaller effects on beta2AR ligand regulation of AC in membranes expressing beta2AR-G(s)alphaS than in membranes expressing beta2AR-G(s)alphaL. These data are explained by a model in which anions increase the GDP affinity of G(s)alphaL more efficiently than the GDP affinity of G(s)alphaS, and, thereby, decrease the efficiency of the agonist-free beta2AR and increase the efficiency of the agonist-occupied beta2AR at promoting GDP dissociation from G(s)alphaL. Thus, monovalent anions differentially regulate beta2AR-coupling to G(s)alphaS and G(s)alphaL.  相似文献   

2.
The interaction of an agonist-bound G-protein-coupled receptor (GPCR) with its cognate G-protein initiates a sequence of experimentally quantifiable changes in both the GPCR and G-protein. These include the release of GDP from G(alpha), the formation of a ternary complex between the nucleotide-free G-protein and the GPCR, which has a high affinity for agonist, followed by the binding of GTP to G(alpha), the dissociation of the GPCR/G-protein complex, and the hydrolysis of GTP. The efficacy of an agonist is a measure of its ability to activate this cascade. It has been proposed that efficacy reflects the ability of the agonist to stabilize the active state of the GPCR. We examined a series of beta(2)-adrenoceptor (beta(2)AR) agonists (weak partial agonists to full agonists) for their efficacy at promoting two different steps of the G-protein activation/deactivation cycle: stabilizing the ternary complex (high-affinity, GTP-sensitive agonist binding), and steady-state GTPase activity. We obtained results for the wild-type beta(2)AR and a constitutively active mutant of the beta(2)AR (beta(2)AR(CAM)) using fusion proteins between the GPCRs and G(salpha) to facilitate GPCR/G-protein interactions. There was no correlation between efficacy of ligands in activating GTPase and their ability to stabilize the ternary complex at beta(2)AR(CAM). Our results suggest that the GPCR state that optimally promotes the GDP release and GTP binding is different from the GPCR state that stabilizes the ternary complex. By strongly stabilizing the ternary complex, certain partial agonists may reduce the rate of G-protein turnover relative to a full agonist.  相似文献   

3.
The alpha adrenoceptor properties of the rabbit ear artery and thoracic aorta were assessed using isolated blood vessel rings mounted in tissue baths. ED50 values and dissociation constants for norepinephrine (NE), epinephrine (EPI), isoproterenol (ISO), methoxamine (MET) and clonidine (CLO) were determined by analysis of dose-response data with and without partial inactivation of alpha receptors by phenoxybenzamine. The order of affinity of these agonists for the alpha receptors of ear artery was CLO less than EPI = NE = MET much less than ISO, and that in aorta was EPI less than NE less than CLO less than MET less than ISO. Alpha receptor reserves were larger in ear artery than aorta by factors of 182, 17, 7, 3.8 and 2 for EPI, NE, ISO, MET and CLO, respectively. In contrast, the fraction of receptors occupied at threshold contractile response was larger in aorta than ear artery with every agonist studied except CLO. The maximal response to the partial agonist, CLO, expressed as a fraction of the maximal response to NE in each vessel, was nearly 1.8 times greater in ear artery than in aorta. It is concluded that the alpha receptors of rabbit ear artery are qualitatively different from those of the thoracic aorta. In addition, the larger receptor reserves, smaller fraction of receptors occupied at threshold and the larger maximal response of CLO in ear artery support the hypothesis that alpha receptor density in this vessel is greater than that in the aorta.  相似文献   

4.
Fusion proteins allow for the analysis of receptor/G protein coupling under defined conditions. The beta(2)-adrenoceptor (beta(2)AR) fused to the long splice variant of G(salpha) (G(salphaL)) exhibits a higher apparent constitutive activity than the beta(2)-adrenoceptor fused to the short splice variant of G(salpha) (G(salphaS)). Experimentally, this results in higher efficacy and potency of partial agonists and in higher efficacy of inverse agonists at the beta(2)AR fused to G(salphaL) relative to the beta(2)AR fused to G(salphaS), indicating that the agonist-free beta(2)AR and the beta(2)AR occupied by partial agonists promote GDP dissociation from G(salphaL) more efficiently than from G(salphaS). In fact, the GDP affinity of G(salphaS) fused to the beta(2)AR is higher than the GDP affinity of G(salphaL) fused to the beta(2)AR. We asked the question whether the histamine H(2)-receptor (H(2)R) exhibits similar coupling to G(salpha) splice variants as the beta(2)AR. To address this question, we studied H(2)R-G(salpha) fusion proteins expressed in Sf9 cells. In contrast to beta(2)AR-G(salpha) fusion proteins, the potencies and efficacies of partial agonists and the efficacies of inverse agonists were similar at the H(2)R fused to G(salphaL) and G(salphaS) as assessed by guanosine-5'-O-(3-thio)triphosphate binding and/or steady-state GTPase activity. However, the time course analysis of guanosine-5'-O-(3-thio)triphosphate binding indicated that G(salphaS) fused to the H(2)R possesses a higher GDP-affinity than G(salphaL) fused to the H(2)R. Our data show that the H(2)R fused to G(salphaL) and G(salphaS) possesses similar constitutive activity and is insensitive to differences in GDP affinity of G(salpha) splice variants. Thus, GDP affinity of G proteins does not generally determine constitutive activity of receptors.  相似文献   

5.
It has been known for many years that norepinephrine (NE) is a potent endogenous anticonvulsant, yet there is confusion as to which receptor(s) mediate this effect. This is probably due to multiple factors, including the importance of distinct signaling pathways for different seizure paradigms, a lack of comprehensive pharmacological studies, and difficulty in interpreting existing pharmacological results due to the presence of endogenous NE. We sought to circumvent these problems by testing the anticonvulsant activity of selective agonists for most known adrenoreceptors (ARs) in dopamine beta-hydroxylase knockout (Dbh -/-) mice that lack endogenous NE. Dbh -/- mice are hypersensitive to pentylenetetrazole (PTZ)-induced seizures, demonstrating that endogenous NE inhibits PTZ-induced seizures in the wild type. Pretreatment of Dbh -/- mice with an alpha(1)AR or beta(2)AR, but not an alpha(2)AR or beta(1)AR agonist significantly protected against PTZ-induced seizures. In contrast, only the beta(2)AR agonist showed anticonvulsant activity in heterozygous controls. Furthermore, an alpha(1)AR antagonist exacerbated PTZ-induced seizures in control mice, whereas a beta(2)AR antagonist had no effect. We conclude that activation of the alpha(1)AR is primarily responsible for the anticonvulsant activity of endogenous NE in the murine PTZ model of epilepsy. Endogenous NE probably does not activate the beta(2)AR under these conditions, but exogenous activation of the beta(2)AR produces an anticonvulsant effect.  相似文献   

6.
We recently reported that alpha(2A)-adrenoceptor (AR) desensitization and down-regulation occurs after 24-h treatment with epinephrine (EPI) (0.3 microM) in BE(2)-C cells that express both alpha(2)- and beta(2)-ARs. The same concentration of norepinephrine (NE) has no effect. The effect of EPI is prevented by beta(2)-AR blockade and is associated with an increase in G protein-coupled receptor kinase 3 (GRK3) expression. Because differences in agonist-induced down-regulation of the alpha(2A)-versus alpha(2B)-ARs have been reported, the present study examines the effects of simultaneous activation of alpha(2B)- and beta(2)-ARs on alpha(2B)-AR number and signaling. We studied NG108 cells that naturally express alpha(2B)-ARs, and BN17 cells, NG108 cells transfected to express the human beta(2)-AR. In NG108 cells, alpha(2B)-AR desensitization and down-regulation require treatment with 20 microM EPI or NE; GRK expression was not changed. In BN17 cells expressing beta(2)-ARs, the threshold EPI concentration for alpha(2B)-AR desensitization and down-regulation was reduced to 0.3 microM; 10 microM NE was required for the same effect. Furthermore, 24-h EPI or NE treatments that produced desensitization also resulted in a selective 2-fold up-regulation of GRK3; GRK2 was unchanged. The beta-AR antagonist alprenolol (1 microM) and GRK3 antisense (but not sense) DNA blocked 0.3 microM EPI- and 10 microM NE-induced desensitization and down-regulation of the alpha(2B)-AR as well as GRK3 up-regulation. In conclusion, simultaneous activation of alpha(2B)- and beta(2)-ARs results in a 67-fold decrease in the threshold concentration of EPI required for alpha(2B)-AR down-regulation. This lower threshold for down-regulation is associated with alpha(2B)- and beta(2)-AR dependent up-regulation of GRK3 expression.  相似文献   

7.
BACKGROUND: Because milrinone is a widely used phosphodiesterase-3 (PDE3) inhibitor, it would be of interest to know whether it interacts with beta1- and beta2-adrenergic receptor (AR) agonists in human myocardium. OBJECTIVES: This in vitro study was conducted to test whether milrinone differentially regulates cyclic adenosine-3',5'-monophosphate (cAMP) production and to examine the effect of milrinone on the positive inotropic responses and cAMP production induced by activation of the beta1-AR with norepinephrine (NE) and activation of the beta2-AR with epinephrine (EPI) in human atrial myocardium. METHODS: Right atrial trabeculae were obtained from patients undergoing cardiac surgery for valve repair. Concentration-response curves for inotropic responses mediated through the beta1-AR (NE in the presence of the beta2-blocker ICI 118, 551) and the beta2-AR (EPI in the presence of the beta1-blocker CGP 20712A) were obtained in the absence and presence of milrinone 1 micromol/L. This concentration of milrinone was chosen because it corresponded to its 50% inhibitory concentration as a PDE3 inhibitor and its therapeutic plasma concentration. The production of cAMP induced by exposure to selective beta1- and beta2-AR stimulation was also measured in the absence and presence of milrinone. RESULTS: Right atrial tissue samples were obtained from 12 white patients (7 women, 5 men; mean [SE] age, 64.6 [6.3] years) undergoing cardiac surgery for valve repair (8 mitral, 4 aortic). The presence of milrinone was associated with leftward shifts in the concentration-response curves for both NE and EPI. cAMP production in myocardial tissue samples in the presence of milrinone was increased only with NE induction (mean [SEM], 745.0 [136.7] pmol/g in the absence of milrinone vs 1620.5 [372.3] pmol/g in the presence of milrinone; P < 0.05). CONCLUSIONS: In this preliminary study in human atrial myocardium, milrinone potentiated the contractile responses to both NE and EPI. However, only the effect of NE on tissue levels of cAMP was increased in the presence of milrinone.  相似文献   

8.
beta-Adrenoceptor blocking agents (beta-blockers) that at low concentrations antagonize cardiostimulant effects of catecholamines, but at high concentrations also cause cardiostimulation, have been appearing since the late 1960s. These cardiostimulant beta-blockers, coined non-conventional partial agonists, antagonize the effects of catecholamines through a high-affinity site (beta(1H)AR), but cause cardiostimulation mainly through a low-affinity site (beta(1L)AR) of the myocardial beta(1)-adrenoceptor. The experimental non-conventional partial agonist (-)-CGP12177 increases cardiac L-type Ca(2+) current density and Ca(2+) transients, shortens action potential duration but augments action potential plateau, increases heart rate and force, as well as causes arrhythmic Ca(2+) transients and arrhythmic cardiocyte contractions. Other beta-blockers, which do not cause cardiostimulation, consistently have lower affinity for beta(1L)AR than beta(1H)AR. These sites were verified and the cardiac pharmacology of non-conventional partial agonists confirmed on recombinant beta(1)-adrenoceptors and on beta(1)-adrenoceptors overexpressed into the heart. A targeted mutation of Asp138 to Glu138 virtually abolished the pharmacology of beta(1H)AR but left intact the pharmacology of beta(1L)AR. Non-conventional partial agonists may be beneficial for the treatment of peripheral autonomic neuropathy but probably due to their arrhythmic propensities, may be harmful for the treatment of chronic heart failure.  相似文献   

9.
The effect of 24-h pretreatment with epinephrine (EPI) or norepinephrine (NE) on alpha(2)- and beta-adrenoceptor agonist-induced, cAMP responses and G(i3)alpha-protein expression was studied in primary cultures of rat superior cervical ganglionic (SCG) neurons. SCG neurons, 10 to 12 days in culture, accumulated cAMP when stimulated with the beta-adrenoceptor agonist isoproterenol and the preferential beta(2)-adrenoceptor antagonist ICI 118,551 blocked this response. Similarly, the preferential alpha(2)-adrenoceptor agonist UK14,304 inhibited forskolin-stimulated cAMP accumulation, implying that cultured SCG neurons possess functional alpha(2)- and beta(2)-adrenoceptors. A 24-h treatment of SCG neurons with EPI or NE induced desensitization of the cAMP response to the beta-adrenoceptor agonist isoproterenol. Simultaneously, EPI treatment increased the maximal inhibitory cAMP response to the alpha(2)-adrenoceptor agonist UK14,304 and NE was without effect. Immunoblotting analyses of G(i3)alpha subunits revealed that 24-h EPI but not NE treatment induces a 3- to 4-fold increase in the expression of G(i3)alpha subunits. Furthermore, EPI-induced up-regulation of alpha-subunit expression can be blocked by the preferential beta(2)-adrenoceptor antagonist ICI 118,551 but not by the preferential beta(1)-adrenoceptor antagonist CGP 20712A. Our results suggest that changes in alpha(2)-adrenoceptor responsiveness induced by EPI may involve activation of beta(2)-adrenoceptors that influence the expression of inhibitory G proteins. Thus, primary cultures of sympathetic neurons by possessing functional alpha(2)- and beta-adrenoceptors may be a suitable model system to study the signaling mechanisms of "cross talk" between these adrenoceptor subtypes, which are known to play a central role in cardiovascular function.  相似文献   

10.
The alpha(2C)-adrenergic receptor (alpha(2C)AR) is known to be poorly trafficked to the cell surface when expressed in a variety of cell types. We tested the hypothesis that the surface expression and signaling of alpha(2C)AR might be enhanced by heterodimerization with other G protein-coupled receptors (GPCRs). Cotransfection of alpha(2C)AR with more than 25 related GPCRs revealed that only coexpression with the beta(2)-adrenergic receptor (beta(2)AR) increased the surface localization of alpha(2C)AR in human embryonic kidney-293 cells. Coimmunoprecipitation of alpha(2C)AR with beta(2)AR confirmed a physical interaction between the two receptors. Confocal microscopy studies demonstrated that alpha(2C)AR expressed alone was mainly intracellular, whereas alpha(2C)AR coexpressed with beta(2)AR was predominantly localized to the plasma membrane. Ligand binding studies revealed a significant increase in alpha(2C)AR binding sites upon coexpression with beta(2)AR, with no apparent change in affinity for alpha(2)AR ligands. Functional assays with the alpha(2)AR-specific agonist brimonidine (UK 14,304) revealed that coexpression of beta(2)AR with alpha(2C)AR enhanced alpha(2C)AR-mediated activation of extracellular signal-regulated kinase 1/2. Furthermore, analyses of agonist-promoted receptor endocytosis demonstrated enhanced alpha(2C)AR internalization in response to alpha(2)AR agonists when alpha(2C)AR and beta(2)AR were coexpressed. In addition, substantial cointernalization of alpha(2C)AR in response to betaAR agonists was observed when alpha(2C)AR was coexpressed with beta(2)AR. These data reveal that alpha(2C)AR can interact with beta(2)AR in cells in a manner that regulates alpha(2C)AR surface expression, internalization, and functionality.  相似文献   

11.
Previous studies revealed pharmacological differences between human and guinea pig histamine H(2) receptors (H(2)Rs) with respect to the interaction with guanidine-type agonists. Because H(2)R species variants are structurally very similar, comparative studies are suited to relate different properties of H(2)R species isoforms to few molecular determinants. Therefore, we systematically compared H(2)Rs of human (h), guinea pig (gp), rat (r), and canine (c). Fusion proteins of hH(2)R, gpH(2)R, rH(2)R, and cH(2)R, respectively, and the short splice variant of G(salpha), G(salphaS), were expressed in Sf9 insect cells. In the membrane steady-state GTPase activity assay, cH(2)R-G(salphaS) but neither gpH(2)R-G(salphaS) nor rH(2)R-G(salphaS) showed the hallmarks of increased constitutive activity compared with hH(2)R-G(salphaS), i.e., increased efficacies of partial agonists, increased potencies of agonists with the extent of potency increase being correlated with the corresponding efficacies at hH(2)R-G(salphaS), increased inverse agonist efficacies, and decreased potencies of antagonists. Furthermore, in membranes expressing nonfused H(2)Rs without or together with mammalian G(salphaS) or H(2)R-G(salpha) fusion proteins, the highest basal and GTP-dependent increases in adenylyl cyclase activity were observed for cH(2)R. An example of ligand selectivity is given by metiamide, acting as an inverse agonist at hH(2)R-G(salphaS), gpH(2)R-G(salphaS), and rH(2)R-G(salphaS) in the GTPase assay in contrast to being a weak partial agonist with decreased potency at cH(2)R-G(salphaS). In conclusion, the cH(2)R exhibits increased constitutive activity compared with hH(2)R, gpH(2)R, and rH(2)R, and there is evidence for ligand-specific conformations in H(2)R species isoforms.  相似文献   

12.
We examined the effects of beta2-adrenergic receptor (beta2-AR) agonists on the expression of co-stimulatory molecules on lipopolysaccharide (LPS)-stimulated human peripheral blood mononuclear cells. The study found that beta2-AR agonists inhibited the expression of intercellular adhesion molecule-1 (ICAM-1), CD40 and CD14 on monocytes, and that AR agonist activity was antagonized by the selective beta2-AR antagonist, butoxamine. The selective beta2-AR agonists salbutamol and terbutaline induced a similar co-stimulatory molecule expression pattern. The LPS-induced production of tumour necrosis factor-alpha was inhibited by AR agonists, and this was also antagonized by butoxamine, and mimicked by salbutamol and terbutaline. The AR agonists also inhibited T-cell proliferation through beta2-AR stimulation. This study clearly demonstrated that endogenous catecholamines elicited immunosuppressive effects through beta2-AR stimulation, possibly due to down-regulation of the expression of ICAM-1, CD40 and CD14 on monocytes. These results suggested that the sympathetic nervous system might regulate the T-helper cell balance via the peripheral end-effectors of the stress system.  相似文献   

13.
Compared with cloned, human (h)D(2) receptors (pK(i) = 6.9), the antiparkinsonian agent piribedil showed comparable affinity for halpha(2A)- (7.1) and halpha(2C)- (7.2) adrenoceptors (ARs), whereas its affinity for halpha(2B)-ARs was less marked (6.5). At halpha(2A)- and halpha(2C)-ARs, piribedil antagonized induction of [(35)S]guanosine-5'-O-(3-thio)triphosphate (GTPgammaS) binding by norepinephrine (NE) with pK(b) values of 6.5 and 6.9, respectively. Furthermore, Schild analysis of the actions of piribedil at halpha(2A)-ARs indicated competitive antagonism, yielding a pA(2) of 6.5. At a porcine alpha(2A)-AR-Gi1alpha-Cys351C (wild-type) fusion protein, piribedil competitively abolished (pA(2) = 6.5) GTPase activity induced by epinephrine. However, at a alpha(2A)-AR-Gi1alpha-Cys351I (mutant) fusion protein of amplified sensitivity, although still acting as a competitive antagonist (pA(2) = 6.2) of epinephrine, piribedil itself manifested weak partial agonist properties. Similarly, piribedil weakly induced mitogen-activated protein kinase phosphorylation via wild-type halpha(2A)-ARs, although attenuating its phosphorylation by NE. As demonstrated by functional [(35)S]GTPgammaS autoradiography in rats, piribedil antagonized activation by NE of alpha(2)-ARs in cortex, amygdala, and septum. Antagonist properties were also expressed in a dose-dependent enhancement of the firing rate of adrenergic neurons in locus ceruleus (0.125-4.0 mg/kg i.v.). Furthermore, piribedil (2.5-4.0 mg/kg s.c.) accelerated hippocampal NE synthesis, elevated dialysis levels of NE in hippocampus and frontal cortex, and blocked hypnotic-sedative properties of the alpha(2)-AR agonist xylazine. Finally, piribedil showed only modest affinity for rat alpha(1)-ARs (5.9) and weakly antagonized NE-induced activation of phospholipase C via halpha(1A)-ARs (pK(b) = 5.6). In conclusion, piribedil displays essentially antagonist properties at cloned, human and cerebral, rat alpha(2)-ARs. Blockade of alpha(2)-ARs may, thus, contribute to its clinical antiparkinsonian profile.  相似文献   

14.
The effects of chronic NE administration (100 micrograms/kg/hr s.c., 6 days) on the stimulus-induced overflow of neurotransmitter from the isolated perfused rat kidney were examined. This treatment increased renal NE content and increased the absolute stimulus-induced overflow of NE. The increase in absolute overflow was not simply the result of the increase in renal NE content as fractional overflow was also increased slightly (20%). Alpha adrenoceptor blockade with phentolamine eliminated the NE treatment-induced difference in fractional overflow. However, the dose-response curves to phentolamine and the alpha-2 adrenoceptor agonist UK 14,304 on stimulus-induced overflow from the kidney were not significantly different after NE treatment. Chronic EPI treatment (same dose) produced an 80% increase in fractional stimulus-induced overflow of neurotransmitter but the dose-response curve to UK 14,304 was shifted only slight to the right (3-fold) of the control curve. No influence of prejunctional beta adrenoceptors on stimulus-induced overflow was observed in either the control of the NE-treated group. In conclusion, the data regarding fractional overflow and the effects of the phentolamine suggest that NE treatment produces minimal change whereas EPI treatment produces marked decreases in the influence in prejunctional alpha adrenoceptors. However, although the data with UK 14,304 after NE treatment support this conclusion, the failure of EPI treatment to alter the dose-response curve to UK 14,304 does not. The apparent contradiction of the results with the alpha adrenoceptor agonist in the EPI-treated group as well as the large differences between the effects of NE vs. EPI treatment on stimulus-induced overflow and prejunctional alpha adrenoceptor function cannot be explained at this time.  相似文献   

15.
beta-Adrenergic receptor (AR) agonists have been demonstrated to modulate the production of inflammatory mediators. Recent studies implied that beta 2-AR agonists might be useful for chronic inflammatory diseases caused by interleukin (IL)-18. In the present study, we found that norepinephrine, epinephrine, or isoproterenol down-regulated IL-18 (100 ng/ml)-induced intercellular adhesion molecule (ICAM)-1 expression on monocytes in a dose-dependent manner (10(-8)-10(-4) M), but did not effect B7.1 and B7.2 expression after 24-h incubation. The modulatory effect of these catecholamines on ICAM-1 expression was antagonized by beta 2-AR antagonist, but not by alpha 1-, alpha 2-, or beta 1-AR antagonist. beta 2-AR-selective agonists salbutanol and terbutaline down-regulated IL-18-induced ICAM-1 expression on monocytes, but alpha 1-, alpha 2-, or beta1-AR agonist had no effect. In the same manner, salbutanol and terbutaline as well as norepinephrine, epinephrine, and isoproterenol regulated the IL-18-induced cytokine production, including IL-12, tumor necrosis factor-alpha or interferon-gamma through the stimulation of beta 2-AR. Together with the previous finding that ICAM-1/lymphocyte function-associated antigen-1 interaction plays a crucial role in the IL-18-initiated cytokine network, the present study strongly suggested that the stimulation of beta 2-AR inhibited the IL-18-activated cytokine cascade through the inhibitory effect on ICAM-1 expression, contributing to finding a new method for clinical treatment.  相似文献   

16.
The purpose of this study was to characterize the type of adrenergic receptor(s) involved in both prostaglandin (PG) synthesis and the contractile response elicited by adrenergic receptor agonists in the rabbit aorta. The synthesis of prostacyclin as measured by the production of 6-keto-PGF1 alpha was assessed in vitro after exposing the aortic rings to different adrenergic agonists. Norepinephrine (NE), selective alpha 1 adrenergic receptor agonists methoxamine (MET), phenylephrine (PHE) and cirazoline (CIR) and the alpha 2 adrenergic receptor agonists UK 14304 (UK) and xylazine (XYL), but not the beta adrenergic receptor agonist isoproterenol (ISP), enhanced 6-keto-PGF1 alpha synthesis in a concentration-dependent manner with following order of potency: NE greater than UK 14304 greater than XYL greater than PHE greater than MET greater than CIR. The NE-induced increased in 6-keto-PGF1 alpha synthesis was attenuated by the alpha 1 adrenergic receptor antagonists prazosin (PZ) and corynanthine (COR) and by the alpha 2 adrenergic receptor antagonists rauwolscine (RW) and yohimbine (YOH). MET-induced 6-keto-PGF1 alpha synthesis was reduced by PZ and COR but not by RW. UK-induced 6-keto-PGF1 alpha synthesis was reduced by RW, YOH, and PZ, which also acts as alpha-2B receptor antagonist, but not by COR. In rabbit aortic rings, adrenergic agonists produced contraction with the following order of potency: NE greater than PHE greater than MET greater than CIR greater than UK greater than XYL.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Multiple and paradoxical effects of airway smooth muscle (ASM) 7-transmembrane-spanning receptors activated during asthma, or by treatment with bronchodilators such as beta(2)-adrenergic receptor (beta(2)AR) agonists, indicate extensive receptor crosstalk. We examined the signaling of the prostanoid-EP(1) receptor, since its endogenous agonist prostaglandin E(2) is abundant in the airway, but its functional implications are poorly defined. Activation of EP(1) failed to elicit ASM contraction in mouse trachea via this G(alphaq)-coupled receptor. However, EP(1) activation markedly reduced the bronchodilatory function of beta(2)AR agonist, but not forskolin, indicating an early pathway interaction. Activation of EP(1) reduced beta(2)AR-stimulated cAMP in ASM but did not promote or augment beta(2)AR phosphorylation or alter beta(2)AR trafficking. Bioluminescence resonant energy transfer showed EP(1) and beta(2)AR formed heterodimers, which were further modified by EP(1) agonist. In cell membrane [(35)S]GTPgammaS binding studies, the presence of the EP(1) component of the dimer uncoupled beta(2)AR from G(alphas), an effect accentuated by EP(1) agonist activation. Thus alone, EP(1) does not appear to have a significant direct effect on airway tone but acts as a modulator of the beta(2)AR, altering G(alphas) coupling via steric interactions imposed by the EP(1):beta(2)AR heterodimeric signaling complex and ultimately affecting beta(2)AR-mediated bronchial relaxation. This mechanism may contribute to beta-agonist resistance found in asthma.  相似文献   

18.
Agonists acting at alpha2-adrenergic receptors (alpha2ARs) produce antinociception and synergize with opioids. The alpha2ARs are divided into three subtypes, alpha(2A)AR, alpha(2B)AR, and alpha(2C)AR. Most alpha2AR agonists require alpha(2A)AR activation to produce antinociception and opioid synergy. The same subtype also mediates the side effect of sedation, which limits the clinical utility of these compounds. Identification of a non-alpha(2A)AR-mediated antinociceptive agent would enhance the therapeutic utility of alpha2AR agonists in pain management. Previous studies have suggested that the alpha2AR agonist ST91 [2-(2,6-diethylphenylamino)-2-imidazoline hydrochloride] has a nonsedating, non-alpha(2A)AR mechanism of action. We examined the pharmacology of spinal ST91 and its interaction with the delta-opioid agonist deltorphin II (Tyr-D-Ala-Phe-Glu-Val-Val-Gly amide) in mice lacking either functional alpha(2A)ARs or alpha(2C)ARs. All drugs were administered by direct lumbar puncture, and drug interactions were evaluated using isobolographic analysis. In contrast to the majority of alpha2AR agonists, ST91 potency was only moderately reduced (3-fold) in the absence of the alpha(2A)AR. Studies with the alpha2AR subtype-preferring antagonists BRL-44408 (2-[2H-(1-methyl-1,3-dihydroisoindole)methyl]-4,5-dihydroimidazole maleate) and prazosin [[4-(4-amino-6,7-dimethoxy-quinazolin-2-yl) piperazin-1-yl]-(2-furyl)methanone] and the pan-alpha2AR antagonist SKF-86466 (6-chloro-2,3,4,5-tetrahydro-3-methyl-1-H-3-benzazepine) suggest a shift from alpha(2A)AR to the other alpha2AR subtype(s) in the absence of alpha(2A)AR. Antinociceptive synergy with deltorphin II was preserved in the absence of either alpha(2A)AR or alpha(2C)AR. In conclusion, ST91 activates both alpha(2A)AR and non-alpha(2A)AR subtypes to produce spinal antinociception and opioid synergy. This study confirms that the spinal pharmacology of ST91 differs from that of other alpha2AR agonists and extends those data to include spinal synergy with opioid agonists. The unique profile of ST91 may be advantageous in pain management.  相似文献   

19.
The intrinsic activities of 14 agonists at beta adrenergic receptors were determined in membranes prepared from control L6 muscle cells or from cells preincubated in the presence of isoproterenol for 20 min. Several drugs that were full agonists when studied with membranes from naive cells were partial agonists in membranes prepared from desensitized cells. Prolonged incubation of cells with agonists resulted in a 70 to 80% decrease in the density of beta adrenergic receptors. This effect was observed after incubation of cells with either full or partial agonists, suggesting that the extent of receptor loss is independent of the intrinsic activity of an agonist. The rate of receptor loss was related, however, to the intrinsic activity of an agonist. The rate of receptor loss correlated with the intrinsic activities of agonists in desensitized (r = 0.94) and control (r = 0.87) membranes.  相似文献   

20.
The purpose of the present study was to ascertain the role of adenylate (AC) versus guanylate cyclase (GC) signaling pathways in the internal anal sphincter (IAS) smooth muscle relaxation by beta(1)-, beta(2)-, and beta(3)-adrenoceptor (AR) activation by xamoterol, procaterol, and disodium 5-[(2R)-2-(3-chlorophenyl)-2-hydroxy-ethyl]amino)propyl]-1,3-benzodioxole-2,2-dicarboxylate (CL 316243), respectively. The above-mentioned agonists produced concentration-dependent relaxation of the smooth muscle strips. Both the selective G(i/o)alpha and G(s)alpha antagonists 8,8'-(carbonylbis(imino-3,1-phenylene))bis-(1,3,5-naphthalene trisulfonic acid) (NF 023) and 4,4',4",4"'-(carbonylbis(imino-5,1,3-benzenetriylbis(carbonylimino)))tetrakis-benzene-1,3-disulfonic acid (NF 449), respectively, inhibited the relaxation induced by procaterol. However, only NF 023 inhibited the relaxation induced by xamoterol and CL 316243. 1H-[1,2,4]Oxadiazolo[4,3-a]quinoxalin-1-one, a soluble GC inhibitor, significantly inhibited the relaxation induced by different agonists. In contrast, the selective AC inhibitor [9-(tetrahydro-2'-furyl)adenine] (SQ 22536) inhibited only the relaxation induced by procaterol. (9R,10S,12S)-2,3,9,10,11,12-Hexahydro-10-hydroxy-9-methyl-1-oxo-9,12-epoxy-1H-diindolo[1,2,3-fg: 3',2',1'-kl]pyrrolo[3,4-l][1,6]benzodiazocine-10-carboxylic acid, hexyl ester (KT 5720), a cAMP-dependent protein kinase inhibitor, attenuated the relaxation by procaterol, whereas (9S,10R,12R)-2,3,9,10,11,12, hexahydro-10-methoxy-2,9-dimethyl-1-oxo-9.12-epoxy-1H-diindolo[1,2,3-fg:3',2',1'-kl]pyrrolo[3,4-I][1,6]benzodiazocine-10-carboxylic acid methyl ester (KT 5823), a selective cGMP-dependent protein kinase (PKG) inhibitor, attenuated the relaxation induced by xamoterol and CL 316243. Xamoterol produced significant increase in cGMP levels, whereas only procaterol enhanced the cAMP levels. Western blot analysis confirmed the presence of beta(1), beta(2), and beta(3)-AR subtypes in the IAS. In summary, beta(2)-AR activates both G(s)alpha and G(i/o)alpha-protein subunits and induces relaxation in the rat IAS via both cAMP/cGMP pathways. In contrast, the beta(1)/beta(3)-ARs activation causes the smooth muscle relaxation via G(i/o)alpha-protein subunit/GC/GMP/PKG pathway. These studies are important for the understanding of intracellular mechanisms underlying IAS smooth muscle relaxation and in turn the pathophysiology of certain anorectal motility disorders.  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