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 15 毫秒
1.
Wang SY  Song Y  Xu M  Hao TP  Han QD  Zhang YY 《生理学报》2005,57(4):480-485
为了明确α1-肾上腺素受体(α1-adrenergic receptor,α1-AR)三种亚型在人胚胎肾(human embryonic kidney,HEK)293A细胞株中的分布特点,及其在激动剂作用下在细胞内的定位改变,本研究采用放射配体结合实验、实时荧光共聚焦成像和Western blot方法检测α1-AR三种亚型在细胞中的定位及蛋白质表达的变化。结果发现:(1)α1-AR三种亚型在HEK293A细胞株转染效率相同,均达90%以上。三株细胞的粗制膜上α1B-AR表达量最高,α1D-AR最低,α1A-AR居中,但三者的解离常数(配)相等;(2)在无激动剂作用时,α1A-AR均匀地分布在HEK293A细胞的胞膜和胞浆,α1B-AR主要位于胞膜,而α1D-AR则主要分布在胞浆中:(3)用α1-AR激动剂苯‘肾上腺素(phenylephrine,PE)刺激细胞1h后,α1A-和α1B-AR在胞膜上分布明显减少,而在胞浆中分布增加,其中α1B-AR变化更为显著,α1D-AR的分布在PE作用下无明显变化。以上结果提示,在激动剂作用下,α1-AR二种亚型在HEK293A细胞中的定位特点和分布变化各有不同。  相似文献   

2.
3.
This investigation was performed to assess the importance of interaction in the bindings of selective and nonselective alpha(1)-antagonists to alpha(1)-adrenergic receptor (alpha(1)-AR) subtypes using molecular modeling. The alpha(1)-antagonists used in this study were prazosin, tamsulosin and KMD-3213. Molecular modeling was performed on Octane 2 workstation (Silicon Graphics) using Discover/Insight II software (Molecular Simulations Inc.). Through molecular modeling, possible binding sites for these drugs were suggested to lie between transmembrane domains (TM) 3, 4, 5 and 6 of the alpha(1)-AR subtypes. In prazosin, the 4-amino group, 1-nitrogen atom and two methoxy groups of quinazoline ring possibly interact with the amino acids in TM3, TM5 and TM6 of alpha(1)-ARs. In tamsulosin, amine group of ethanyl amine chain, methoxy group of benzene ring and sulfonamide nitrogen of benzene ring interacts in TM3, TM4 and TM5 of alpha(1)-ARs. In KMD-3213, amine of ethyl amine chain and indoline nitrogen of this compound possibly interact within TM3 and TM5 of alpha(1)-ARs. Amide nitrogen of KMD-3213 also interacts within TM4 of alpha(1A)-AR. The results of the present study suggested that prazosin has similar binding sites in all the alpha(1)-AR subtypes while tamsulosin interacts at higher number of sites with alpha(1D)-subtype than other alpha(1)-AR subtypes. KMD-3213 being an alpha(1A)-AR selective ligand, binds to higher number of sites of alpha(1A) subtype than to other subtypes. All these amino acids are located near the extracellular loop. These findings are consistent with the previous studies that antagonists bind higher in the pocket closer to the extracellular surface unlike agonist binding.  相似文献   

4.
A number of new pyrimido[5,4-b]indole and [1]benzothieno[3,2-d]pyrimidine derivatives were synthesized and evaluated for their binding and functional properties at alpha(1)-adrenergic receptor (alpha(1)-AR) subtypes. They behaved as potent alpha(1)-AR antagonists. In binding experiments, some of them (RC24 and RC23) showed very high affinity for the alpha(1D)-AR subtype.  相似文献   

5.
Activation of alpha(1)-adrenergic receptors influences both the contractile activity and the growth potential of cardiac myocytes. However, the signaling pathways linking activation of specific alpha(1)-adrenergic receptor (AR) subtypes to these physiological responses remain controversial. In the present study, a molecular approach was used to identify conclusively the signaling pathways activated in response to the individual alpha(1A)- and alpha(1B)-AR subtypes in cardiac myocytes. For this purpose, a mutant alpha(1a)-AR subtype (alpha(1a)-S(290/293)-AR) was constructed based on analogy to the previously described constitutively active mutant alpha(1b)-AR subtype (alpha(1b)-S(288-294)-AR). The mutant alpha(1a)-S(290/293)-AR subtype displayed constitutive activity based on four criteria. To introduce the constitutively active alpha(1)-AR subtypes into cardiac myocytes, recombinant Sindbis viruses encoding either the alpha(1a)-S(290/293)-AR or alpha(1b)-S(288-294)-AR subtype were used to infect the whole cell population with >90% efficiency, thereby allowing the biochemical activities of the various signaling pathways to be measured. When expressed at comparable levels, the alpha(1a)-S(290/293)-AR subtype exhibited a significantly elevated basal level as well as agonist-stimulated level of inositol phosphate accumulation, coincident with activation of atrial natriuretic factor-luciferase gene expression. By contrast, the alpha(1b)-S(288-294)-AR subtype displayed a markedly increased serum response element-luciferase gene expression but no activation of atrial natriuretic factor-luciferase gene expression. Taken together, this study provides the first molecular evidence for coupling of the alpha(1a)-AR and the alpha(1b)-AR subtypes to different signaling pathways in cardiac myocytes.  相似文献   

6.
Low-frequency blood pressure oscillations (Mayer waves) are discussed as a marker for sympathetic modulation of vascular tone. However, the factors that determine the frequency response of the vasculature to sympathetic stimuli are not fully understood. Possible mechanisms include functions related to alpha(1)-adrenergic receptors (alpha(1)-AR) and postreceptor processes involved in the vascular contractile response. The purpose of the present study was to examine the hypothesis that expression levels of alpha(1)-AR and their subtype distribution determine velocity and magnitude of alpha(1)-AR-mediated vascular smooth muscle cell (VSMC) contraction. alpha(1A)-, alpha(1B)-, and alpha(1D)-AR subtypes were transfected into VSMCs from rat aorta and characterized immunocytochemically via confocal microscopy. Functional studies in isolated cells were performed using video microscopy. The alpha(1)-AR agonist phenylephrine produced dose-dependent contractions of VSMCs. All transfected groups were more sensitive to phenylephrine compared with controls. Maximal contraction velocity almost doubled in transfected cells. However, no differences in observed parameters were found between the three transfected groups. Contractile properties in response to membrane depolarization with KCl were similar in all groups. In conclusion, alpha(1)-AR density determines velocity and sensitivity of alpha(1)-AR-mediated contraction in VSMCs. alpha(1)-AR subtype distribution does not appear to influence vasoconstriction to sympathetic stimuli.  相似文献   

7.
Expression of alpha(2)-adrenergic receptors (alpha(2)-AR) is very high in fetal rat heart although numbers decline with increasing gestational age. The current experiments were designed to identify the subtypes of alpha(2)-AR expressed and the function of these receptors in fetal cardiac myocytes. Expression of alpha(2)A and alpha(2)C, but not alpha(2)B, was confirmed in the myocyte population by indirect immunofluorescence microscopy with subtype-specific antibodies and by Western blot. Both dexmedetomidine, an alpha(2)-selective agonist, and norepinephrine, increased actin cytoskeleton organization and this increase was blocked by the alpha(2)-selective antagonist, atipamezole. Furthermore, dexmedetomidine inhibited isoproterenol-stimulated cAMP accumulation in isolated fetal rat heart and this was blocked by rauwolscine. Therefore, functional alpha(2)A and alpha(2)B subtypes are present in the fetal rat heart where they may have a role in cardiac development.  相似文献   

8.
Excessive proliferation of vascular wall cells underlies the development of elevated vascular resistance in hypoxic pulmonary hypertension (PH), but the responsible mechanisms remain unclear. Growth-promoting effects of catecholamines may contribute. Hypoxemia causes sympathoexcitation, and prolonged stimulation of alpha(1)-adrenoceptors (alpha(1)-ARs) induces hypertrophy and hyperplasia of arterial smooth muscle cells and adventitial fibroblasts. Catecholamine trophic actions in arteries are enhanced when other conditions favoring growth or remodeling are present, e.g., injury or altered shear stress, in isolated pulmonary arteries from rats with hypoxic PH. The present study examined the hypothesis that catecholamines contribute to pulmonary vascular remodeling in vivo in hypoxic PH. Mice genetically deficient in norepinephrine and epinephrine production [dopamine beta-hydroxylase(-/-) (DBH(-/-))] or alpha(1)-ARs were examined for alterations in PH, cardiac hypertrophy, and vascular remodeling after 21 days exposure to normobaric 0.1 inspired oxygen fraction (Fi(O(2))). A decrease in the lumen area and an increase in the wall thickness of arteries were strongly inhibited in knockout mice (order of extent of inhibition: DBH(-/-) = alpha(1D)-AR(-/-) > alpha(1B)-AR(-/-)). Distal muscularization of small arterioles was also reduced (DBH(-/-) > alpha(1D)-AR(-/-) > alpha(1B)-AR(-/-) mice). Despite these reductions, increases in right ventricular pressure and hypertrophy were not attenuated in DBH(-/-) and alpha(1B)-AR(-/-) mice. However, hematocrit increased more in these mice, possibly as a consequence of impaired cardiovascular activation that occurs during reduction of Fi(O(2)). In contrast, in alpha(1D)-AR(-/-) mice, where hematocrit increased the same as in wild-type mice, right ventricular pressure was reduced. These data suggest that catecholamine stimulation of alpha(1B)- and alpha(1D)-ARs contributes significantly to vascular remodeling in hypoxic PH.  相似文献   

9.
Three distinct subtypes of alpha(1)-adrenergic receptors (alpha(1)A-, alpha(1)B-, and alpha(1)D-AR) play a prominent role in cell growth. However, little is known about subtype-specific effects on cell proliferation. The activation of alpha(1)A- or alpha(1)B-AR inhibits serum-promoted cell proliferation, whereas alpha(1)D-AR activation does not show such an inhibitory effect. Notably, cell-cycle progression was blocked at G(1)/S transition after activation of alpha(1)A/alpha(1)B-AR but not of alpha(1)D-AR. In agreement with the differential cell proliferation effect, cAMP production was increased after activation of alpha(1)A/alpha(1)B-AR but not alpha(1)D-AR, whereas all alpha(1)-AR subtypes are associated with inositol 1,4,5-trisphosphate production and mitogen-activated protein kinase activation in a similar fashion. Furthermore, the serum-induced reduction in the levels of the cyclin-dependent kinase inhibitor, p27(Kip1), was blocked after activation of alpha(1)A/alpha(1)B-AR but not alpha(1)D-AR. These results show that alpha(1)-AR subtypes differentially activate the cAMP/p27(Kip1) pathway and thereby have differential inhibitory effects on cell proliferation. Subtype-dependent effects should be taken into consideration when assessing the physiological response of native cells where alpha(1)-AR subtypes are generally co-expressed.  相似文献   

10.
alpha(1)-Adrenergic receptors (ARs) belong to the large Class I G protein-coupled receptor superfamily and comprise three subtypes (alpha(1A), alpha(1B), and alpha(1D)). Previous work with heterologously expressed C-terminal green fluorescent protein (GFP)-tagged alpha(1)-ARs showed that alpha(1A)- and alpha(1B)-ARs localize to the plasma membrane, whereas alpha(1D)-ARs accumulate intracellularly. We recently showed that alpha(1D)- and alpha(1B)-ARs form heterodimers, whereas alpha(1D)- and alpha(1A)-ARs do not. Here, we examined the role of heterodimerization in regulating alpha(1D)-AR localization using both confocal imaging of GFP- or CFP-tagged alpha(1)-ARs and a luminometer-based surface expression assay in HEK293 cells. Co-expression with alpha(1B)-ARs caused alpha(1D)-ARs to quantitatively translocate to the cell surface, but co-expression with alpha(1A)-ARs did not. Truncation of the alpha(1B)-AR extracellular N terminus or intracellular C terminus had no effect on surface expression of alpha(1D)-ARs, suggesting primary involvement of the hydrophobic core. Co-transfection with an uncoupled mutant alpha(1B)-AR (Delta12alpha(1B)) increased both alpha(1D)-AR surface expression and coupling to norepinephrine-stimulated Ca(2+) mobilization. Finally, GFP-tagged alpha(1D)-ARs were not detected on the cell surface when expressed in rat aortic smooth muscle cells that express no endogenous ARs, but were almost exclusively localized on the surface when expressed in DDT(1)MF-2 cells, which express endogenous alpha(1B)-ARs. These studies demonstrate that alpha(1B)/alpha(1D)-AR heterodimerization controls surface expression and functional coupling of alpha(1D)-ARs, the N- and C-terminal domains are not involved in this interaction, and that alpha(1B)-AR G protein coupling is not required. These observations may be relevant to many other Class I G protein-coupled receptors, where the functional consequences of heterodimerization are still poorly understood.  相似文献   

11.
We have investigated the mechanisms whereby alpha(2B)-adrenergic receptor (alpha(2B)-AR) promotes MAPK activation in a clone of the renal tubular cell line, LLC-PK1, transfected with the rat nonglycosylated alpha(2)-AR gene. Treatment of LLC-PK1-alpha(2B) with UK14304 or dexmedetomidine caused arachidonic acid (AA) release and ERK2 phosphorylation. AA release was abolished by prior treatment of the cells with pertussis toxin, quinacrine, or methyl arachidonyl fluorophosphonate but not by the addition of the MEK inhibitor U0126. The effects of alpha(2)-agonists on MAPK phosphorylation were mimicked by cell exposure to exogenous AA. On the other hand, quinacrine abolished the effects of UK14304, but not of AA, suggesting that AA released through PLA2 is responsible for MAPK activation by alpha(2B)-AR. The effects of alpha(2)-agonists or AA were PKC-independent and were attenuated by indomethacin and nordihydroguaiaretic acid. Treatment with batimastat, CRM 197, or tyrphostin AG1478 suppressed MAPK phosphorylation promoted by alpha(2)-agonist or AA. Furthermore, conditioned culture medium from UK14304-treated LLC-PK1-alpha(2B) induced MAPK phosphorylation in wild-type LLC-PK1. Based on these data, we propose a model whereby activation of MAPK by alpha(2B)-AR is mediated through stimulation of PLA2, AA release, generation of AA derivatives, activation of matrix metalloproteinases, release of heparin-binding EGF-like growth factor, transactivation of epidermal growth factor receptor, and recruitment of Shc. Whether this pathway is particular to alpha(2B)-AR and LLC-PK1 or whether it can be extended to other cell types and/or other G-protein-coupled receptors remains to be established.  相似文献   

12.
Two functional alpha(1)-adrenergic receptor (AR) subtypes (alpha(1A) and alpha(1B)) have been identified in the mouse heart. However, it is unclear whether the third known subtype, alpha(1D)-AR, is also present. To investigate this, we determined whether there were alpha(1)-AR responses in hearts from a novel mouse model lacking alpha(1A)- and alpha(1B)-ARs (double knockout) (ABKO). In Langendorff-perfused hearts, alpha(1)-ARs were stimulated with phenylephrine. For ABKO hearts, phenylephrine reduced left ventricular pressure and coronary flow (to 87 +/- 2% and 86 +/- 4% of initial, respectively, n = 11, P < 0.01). These effects were blocked by prazosin and 8-[2-[4-(2-methoxyphenyl)-1-piperazinyl]-8-azaspirol[4,5]decane-7,9-dione] dihydrochloride, suggesting that alpha(1D)-AR-mediated responses were present. In contrast, right ventricular trabeculae from ABKO hearts did not respond to phenylephrine, suggesting that in ABKO perfused hearts, the effects of phenylephrine were not mediated by direct actions on cardiomyocytes. A novel finding was that alpha(1)-AR stimulation caused positive inotropy in the wild-type mouse heart, in contrast to negative inotropy observed in mouse cardiac muscle strips. We conclude that mouse hearts lacking alpha(1A)- and alpha(1B)-ARs retain functional alpha(1)-AR responses involving decreases of coronary flow and ventricular pressure that reflect alpha(1D)-AR-mediated vasoconstriction. Furthermore, alpha(1)-AR inotropic responses depend critically on the experimental conditions.  相似文献   

13.
14.
15.
Cold constricts cutaneous blood vessels by increasing the reactivity of smooth muscle alpha(2)-adrenergic receptors (alpha(2)-ARs). Experiments were performed to determine the role of alpha(2)-AR subtypes (alpha(2A)-, alpha(2B)-, alpha(2C)-ARs) in this response. Stimulation of alpha(1)-ARs by phenylephrine or alpha(2)-ARs by UK-14,304 caused constriction of isolated mouse tail arteries mounted in a pressurized myograph system. Compared with proximal arteries, distal arteries were more responsive to alpha(2)-AR activation but less responsive to activation of alpha(1)-ARs. Cold augmented constriction to alpha(2)-AR activation in distal arteries but did not affect the response to alpha(1)-AR stimulation or the level of myogenic tone. Western blot analysis demonstrated expression of alpha(2A)- and alpha(2C)-ARs in tail arteries: expression of alpha(2C)-ARs decreased in distal compared with proximal arteries, whereas expression of the glycosylated form of the alpha(2A)-AR increased in distal arteries. At 37 degrees C, alpha(2)-AR-induced vasoconstriction in distal arteries was inhibited by selective blockade of alpha(2A)-ARs (BRL-44408) but not by selective inhibition of alpha(2B)-ARs (ARC-239) or alpha(2C)-ARs (MK-912). In contrast, during cold exposure (28 degrees C), the augmented response to UK-14,304 was inhibited by the alpha(2C)-AR antagonist MK-912, which selectively abolished cold-induced amplification of the response. These experiments indicate that cold-induced amplification of alpha(2)-ARs is mediated by alpha(2C)-ARs that are normally silent in these cutaneous arteries. Blockade of alpha(2C)-ARs may prove an effective treatment for Raynaud's Phenomenon.  相似文献   

16.
The 19-amino acid conopeptide (rho-TIA) was shown previously to antagonize noncompetitively alpha(1B)-adrenergic receptors (ARs). Because this is the first peptide ligand for these receptors, we compared its interactions with the three recombinant human alpha(1)-AR subtypes (alpha(1A), alpha(1B), and alpha(1D)). Radioligand binding assays showed that rho-TIA was 10-fold selective for human alpha(1B)-over alpha(1A)- and alpha(1D)-ARs. As observed with hamster alpha(1B)-ARs, rho-TIA decreased the number of binding sites (B(max)) for human alpha(1B)-ARs without changing affinity (K(D)), and this inhibition was unaffected by the length of incubation but was reversed by washing. However, rho-TIA had opposite effects at human alpha(1A)-ARs and alpha(1D)-ARs, decreasing K(D) without changing B(max), suggesting it acts competitively at these subtypes. rho-TIA reduced maximal NE-stimulated [(3)H]inositol phosphate formation in HEK293 cells expressing human alpha(1B)-ARs but competitively inhibited responses in cells expressing alpha(1A)- or alpha(1D)-ARs. Truncation mutants showed that the amino-terminal domains of alpha(1B)- or alpha(1D)-ARs are not involved in interaction with rho-TIA. Alanine-scanning mutagenesis of rho-TIA showed F18A had an increased selectivity for alpha(1B)-ARs, and F18N also increased subtype selectivity. I8A had a slightly reduced potency at alpha(1B)-ARs and was found to be a competitive, rather than noncompetitive, inhibitor in both radioligand and functional assays. Thus rho-TIA noncompetitively inhibits alpha(1B)-ARs but competitively inhibits the other two subtypes, and this selectivity can be increased by mutation. These differential interactions do not involve the receptor amino termini and are not because of the charged nature of the peptide, and isoleucine 8 is critical for its noncompetitive inhibition at alpha(1B)-ARs.  相似文献   

17.
The activation state of beta-adrenergic receptors (beta-ARs) in vivo is an important determinant of hemodynamic status, cardiac performance, and metabolic rate. In order to achieve homeostasis in vivo, the cellular signals generated by beta-AR activation are integrated with signals from a number of other distinct receptors and signaling pathways. We have utilized genetic knockout models to test directly the role of beta1- and/or beta2-AR expression on these homeostatic control mechanisms. Despite total absence of beta1- and beta2-ARs, the predominant cardiovascular beta-adrenergic subtypes, basal heart rate, blood pressure, and metabolic rate do not differ from wild type controls. However, stimulation of beta-AR function by beta-AR agonists or exercise reveals significant impairments in chronotropic range, vascular reactivity, and metabolic rate. Surprisingly, the blunted chronotropic and metabolic response to exercise seen in beta1/beta2-AR double knockouts fails to impact maximal exercise capacity. Integrating the results from single beta1- and beta2-AR knockouts as well as the beta1-/beta2-AR double knock-out suggest that in the mouse, beta-AR stimulation of cardiac inotropy and chronotropy is mediated almost exclusively by the beta1-AR, whereas vascular relaxation and metabolic rate are controlled by all three beta-ARs (beta1-, beta2-, and beta3-AR). Compensatory alterations in cardiac muscarinic receptor density and vascular beta3-AR responsiveness are also observed in beta1-/beta2-AR double knockouts. In addition to its ability to define beta-AR subtype-specific functions, this genetic approach is also useful in identifying adaptive alterations that serve to maintain critical physiological setpoints such as heart rate, blood pressure, and metabolic rate when cellular signaling mechanisms are perturbed.  相似文献   

18.
To determine the effects of ischemia-reperfusion (I/R) on alpha(1)-adrenergic-receptor (alpha(1)-AR) functions, alpha(1)-AR-mediated contraction, inositol phosphate (IP) accumulation, and alpha(1)-AR-G protein coupling were examined in the tail arteries of anesthetized rats after 60 min of ischemia and 60 min of reperfusion. The contractile response to norepinephrine (NE) was significantly increased after I/R, whereas the contractile response to KCl remained unchanged. This was accompanied by a 69% increase in NE-stimulated IP accumulation. Furthermore, receptor-stimulated coupling of alpha(1a)-AR to G alpha(q/11) proteins was increased, whereas the coupling of alpha(1b)-AR or alpha(1d)-AR to their G proteins was not altered by I/R. These changes in vascular alpha(1)-AR function occurred without concurrent alteration in expression levels of membrane alpha(1)-AR subtypes or in the associated G proteins. These data demonstrate that I/R increases alpha(1a)-AR-G(q/11) protein coupling and alpha(1)-AR-stimulated IP accumulation in the tail artery. The alterations in alpha(1)-AR signaling are associated with and may underlie the enhanced contractile response of the tail artery to adrenergic stimulation after I/R.  相似文献   

19.
Hirasawa A  Awaji T  Xu Z  Shinoura H  Tsujimoto G 《Life sciences》2001,68(19-20):2259-2267
Alpha1-adrenergic receptors (AR) are members of the superfamily of G protein-coupled receptors (GPCRs) which mediate the effects of the sympathetic nervous system. Alpha1-AR comprise a heterogeneous family of three distinct isoforms of alpha1A, alpha1B and alpha1D; however, very little is known about their difference in physiological role or regulation. We have recently observed a subtype-specific differences in subcellular localization of alpha1-ARs; thus, alpha1A-AR predominantly localize intracellularly, while alpha1B-AR on the cell surface. To examine the molecular mechanism for the subtype-specific differences in subcellular localization, we conducted a search for novel proteins that interact with the alpha1B-AR, specifically focusing on the carboxyl-terminal cytoplasmic domain. Using interaction cloning and biochemical techniques, we demonstrate that gC1q-R interacts with alpha1B-AR in vitro and in vivo through the specific site, and that in cells which co-express alpha1B-AR and gC1q-R, the subcellular localization of alpha1B-AR is markedly altered and its expression is down-regulated. These results suggest that gC1q-R plays a role in the regulation of the subcellular localization as well as the function of alpha1B-ARs.  相似文献   

20.
Yuan K  Rhee KS  Park WH  Kim SW  Kim SH 《Peptides》2008,29(7):1207-1215
Sympathetic nervous system and atrial natriuretic peptide (ANP) system play fundamental roles in the regulation of cardiovascular functions. Overactivity of sympathetic nervous system can lead into cardiovascular diseases such as heart failure and hypertension. The present study aimed to define which adrenergic receptors (ARs) affect atrial contractility and ANP release and to determine their modification in renal hypertensive rat atria. An alpha(1)-AR agonist, cirazoline increased ANP release with positive inotropism. These alpha(1)-AR agonist-mediated responses were attenuated by the alpha(1A)-AR antagonist, but not by the alpha(1B)- or alpha(1D)-AR antagonist. An alpha(2)-AR agonist, guanabenz and clonidine increased ANP release with negative inotropism and decreased cAMP level. The order of potency for the increased ANP release was cirazoline>phenylephrine=guanabenz>clonidine. In contrast, a beta-AR agonist, isoproterenol decreased ANP release with positive inotropism and these responses were blocked by the beta(1)-AR antagonist but not by the beta(2)-AR antagonist. The increased cAMP level by isoproterenol was suppressed by pretreatment with both beta(1)- and beta(2)-AR antagonists. In renal hypertensive rat atria, the effects of isoproterenol on atrial contractility, ANP release, and cAMP level were attenuated whereas the effect of cirazoline on ANP release was unaltered. Atrial beta(1)-AR mRNA level but not alpha(1A)-AR mRNA level was decreased in renal hypertensive rats. These findings suggest that alpha(1A)- and beta(1)-AR oppositely regulate atrial ANP release and that atrial beta(1)-AR expression/function is impaired in renal hypertensive rats.  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