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.
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.
Androgen receptors (AR) play a crucial role in androgen-mediated processes and prostate cancer progression. The pineal hormone melatonin attenuates the androgen-dependent growth of benign and cancer prostate epithelial cells in vitro and may reverse clinical resistance to androgen ablation therapy in patients progressing on gonadotropin releasing hormone (GnRH) analogue. Where along the AR cascade does melatonin act remains to be determined. The effects of melatonin on AR localization, level and activity were assessed using androgen-insensitive prostate carcinoma PC3 cells stably transfected with a wild-type AR-expressing vector (PC3-AR).AR was localized to the PC3-AR cell nucleus in the absence of dihydrotestosterone (DHT). Melatonin caused a robust exclusion of the AR from the cell nucleus to the cytoplasm. The nuclear export inhibitor, leptomycin B prevented this process. The exclusion was selective since melatonin had no such effect on the nuclear localization of estrogen receptors alpha (ERalpha) in these cells.Melatonin also caused nuclear exclusion of the AR in the presence of DHT. In addition, it attenuated androgen induced reporter gene activity in PC3 cells co-transfected with the human AR and AR reporter plasmids. Elevated androgen concentrations counteracted melatonin's effects. Melatonin did not decrease AR level or androgen binding in the cells.The nuclear localization of the AR is a hallmark of its cellular activity. These data point to AR nuclear exclusion as a possible mechanism to attenuate androgen responses in target tissues.  相似文献   

3.
The alpha 2B -adrenergic receptor ( alpha 2B -AR), a member of the G protein-coupled receptor (GPCR) superfamily, was expressed at high levels from Semliki Forest virus (SFV) vectors in mammalian cells. Constructs were engineered by fusing enhanced green fluorescent protein (eGFP) and the SFV capsid to opposite ends of the alpha 2B -AR. The receptor fusions alpha 2B -AR-eGFP and CAP- alpha 2B -AR expressed in CHO-K1 cells generated alpha 2B values of 176 and 122pmol/mg of membrane protein, respectively, and showed similar ligand binding characteristics, alpha 2B -AR subtype-selectivity, and G protein activation as reported for stable expression in CHO-K1 cells. Cryo-electron microscopy and eGFP-based fluorescence indicated the same subcellular receptor distribution. SFV expression is well suited for studies on the pharmacology, biochemistry, and cell biology of GPCRs, and for large-scale recombinant protein production in mammalian suspension culture to generate sufficient receptor quantities for structural biology.  相似文献   

4.
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   

5.
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   

6.
7.
8.
Mice with altered alpha(1)-adrenergic receptor (AR) genes have become important tools in elucidating the subtype-specific functions of the three alpha(1)-AR subtypes because of the lack of sufficiently subtype-selective pharmacological agents. Mice with a deletion (knockout, KO) or an overexpression (transgenic, TG) of the alpha(1A)-, alpha(1B)-, or alpha(1D)-AR subtypes have been generated. The alpha(1)-ARs are the principal mediators of the hypertensive response to alpha(1)-agonists in the cardiovascular system. Studies with these mice indicate that alpha(1A)-AR and alpha(1B)-AR subtypes play an important role in cardiac development and/or function as well as in blood pressure (BP) response to alpha(1)-agonists via vasoconstriction. The alpha(1B)- and alpha(1D)-subtypes also appear to be involved in central nervous system (CNS) processes such as nociceptive responses, modulation of memory consolidation and working memory. The ability to study subtype-specific functions in different mouse strains by altering the same alpha(1)-AR in different ways strengthens the conclusions drawn from these studies. Although these genetic approaches have limitations, they have significantly increased our understanding of the functions of alpha(1)-AR subtypes.  相似文献   

9.
We have characterized the maturation, co- and posttranslational modifications, and functional properties of the alpha(1B)-adrenergic receptor (AR) expressed in different mammalian cells transfected using conventional approaches or the Semliki Forest virus system. We found that the alpha(1B)-AR undergoes N-linked glycosylation as demonstrated by its sensitivity to endoglycosidases and by the effect of tunicamycin on receptor maturation. Pulse-chase labeling experiments in BHK-21 cells demonstrate that the alpha(1B)-AR is synthesized as a 70 kDa core glycosylated precursor that is converted to the 90 kDa mature form of the receptor with a half-time of approximately 2 h. N-Linked glycosylation of the alpha(1B)-AR occurs at four asparagines on the N-terminus of the receptor. Mutations of the N-linked glycosylation sites did not have a significant effect on receptor function or expression. Surprisingly, receptor mutants lacking N-linked glycosylation migrated as heterogeneous bands in SDS-PAGE. Our findings demonstrate that N-linked glycosylation and phosphorylation, but not palmitoylation or O-linked glycosylation, contribute to the structural heterogeneity of the alpha(1B)-AR as it is observed in SDS-PAGE. The modifications found are similar in the different mammalian expression systems explored. Our findings indicate that the Semliki Forest virus system can provide large amounts of functional and fully glycosylated alpha(1B)-AR protein suitable for biochemical and structural studies. The results of this study contribute to elucidate the basic steps involved in the processing of G protein-coupled receptors as well as to optimize strategies for their overexpression.  相似文献   

10.
In this study, we developed a human prostatic epithelial cell line BPH-1-AR stably expressing AR by lentiviral transduction. Characterization by immunoblot and RT-PCR showed that AR was stably expressed in all representative BPH-1-AR clones. Androgen treatment induced a secretory differentiation phenotype in BPH-1-AR cells but suppressed their cell proliferation. Treatments with AR agonists induced transactivation of a transfected PSA-gene promoter reporter in BPH-1-AR cells, whereas this transactivation was suppressed by an AR antagonist flutamide, indicating that the transduced AR in BPH-1-AR cells was functional. Finally, we utilized BPH-1-AR cells to evaluate the androgenic activities and growth effects of five newly developed non-steroidal compounds. Results showed that these compounds showed androgenic activities and growth-inhibitory effects on BPH-1-AR cells. Our results showed that BPH-1-AR cell line would be a valuable in vitro model for the study of androgen-regulated processes in prostatic epithelial cells and identification of compounds with AR-modulating activities.  相似文献   

11.
Many recent evidences indicate that androgen-sensitive prostate cancer cells have a lower malignant phenotype that is in particular characterized by a reduced migration and invasion. We previously demonstrated that expression of androgen receptor (AR) by transfection of the androgen-independent prostate cancer cell line PC3 decreases invasion and adhesion of these cells (PC3-AR) through modulation of alpha6beta4 integrin expression. The treatment with the synthetic androgen R1881 further reduced invasion of the cells without, however, modifying alpha6beta4 expression on the cell surface, suggesting an interference with the invasion process in response to EGF. We investigated whether the presence of the AR could affect EGF receptor (EGFR)-mediated signaling in response to EGF by evaluating autotransphosphorylation of the receptor as well as activation of downstream signalling pathways. Immunoprecipitation studies demonstrated a reduction of EGF-induced tyrosine phosphorylation of EGFR in PC3-AR cells. In addition, EGF-stimulated PI3K activity, a key signalling pathway for invasion of these cells, was decreased in PC3-AR cells and further reduced by treatment with R1881, indicating decreased functionality of EGFR. An interaction between EGFR and AR has been demonstrated by immunoconfocal and co-immunoprecipitation analysis in PC3-AR cells, suggesting a possible interference of AR on EGFR signalling by interaction of the two proteins. In conclusion, our results suggest that the expression of AR by transfection in PC3 cells confers a less malignant phenotype by interfering with EGFR autophosphorylation and signalling in response to EGF leading to invasion through a mechanism involving an interaction between AR and EGFR.  相似文献   

12.
The function and distribution of alpha1-adrenergic receptor (AR) subtypes in prostate cancer cells is well characterized. Previous studies have used RNA localization or low-avidity antibodies in tissue or cell lines to determine the alpha1-AR subtype and suggested that the alpha1A-AR is dominant. Two androgen-insensitive, human metastatic cancer cell lines DU145 and PC3 were used as well as the mouse TRAMP C1-C3 primary and clonal cell lines. The density of alpha1-ARs was determined by saturation binding and the distribution of the different alpha1-AR subtypes was examined by competition-binding experiments. In contrast to previous studies, the major alpha1-AR subtype in DU145, PC3 and all of the TRAMP cell lines is the alpha1B-AR. DU145 cells contained 100% of the alpha1B-AR subtype, whereas PC3 cells were composed of 21% alpha1 A-AR and 79% alpha1B-AR. TRAMP cell lines contained between 66% and 79% of the alpha1B-AR with minor fractions of the other two subtypes. Faster doubling time in the TRAMP cell lines correlated with decreasing alpha 1B-AR and increasing alpha1 A- and alpha1D-AR densities. Transfection with EGFP-tagged alpha1B-ARs revealed that localization was mainly intracellular, but the majority of the receptors translocated to the cell surface after extended preincubation (18 hr) with either agonist or antagonist. Localization was confirmed by ligand-binding studies and inositol phosphate assays where prolonged preincubation with either agonist and/or antagonist increased the density and function of alpha 1-ARs, suggesting that the native receptors were mostly intracellular and nonfunctional. Our studies indicate that alpha1B-ARs are the major alpha1-AR subtype expressed in DU145, PC3, and all TRAMP cell lines, but most of the receptor is localized in intracellular compartments in a nonfunctional state, which can be rescued upon prolonged incubation with any ligand.  相似文献   

13.
A recently introduced extension of video-enhanced light microscopy, called Nanovid microscopy, documents the dynamic reorganization of individual cell surface components on living cells. 40-microns colloidal gold probes coupled to different types of poly-L-lysine label negative cell surface components of PTK2 cells. Evidence is provided that they bind to negative sialic acid residues of glycoproteins, probably through nonspecific electrostatic interactions. The gold probes, coupled to short poly-L-lysine molecules (4 kD) displayed Brownian motion, with a diffusion coefficient in the range 0.1-0.2 micron2/s. A diffusion coefficient in the 0.1 micron2/s range was also observed with 40-nm gold probes coupled to an antibody against the lipid-linked Thy-1 antigen on 3T3 fibroblasts. Diffusion of these probes is largely confined to apparent microdomains of 1-2 microns in size. On the other hand, the gold probes, coupled to long poly-L-lysine molecules (240 kD) molecules and bound to the leading lamella, were driven rearward, toward the boundary between lamelloplasm and perinuclear cytoplasm at a velocity of 0.5-1 micron/min by a directed ATP-dependent mechanism. This uniform motion was inhibited by cytochalasin, suggesting actin microfilament involvement. A similar behavior on MO cells was observed when the antibody-labeled gold served as a marker for the PGP-1 (GP-80) antigen. These results show that Nanovid microscopy, offering the possibility to observe the motion of individual specific cell surface components, provides a new and powerful tool to study the dynamic reorganization of the cell membrane during locomotion and in other biological contexts as well.  相似文献   

14.
15.
Norepinephrine directly induces growth of the vascular wall, which may involve not only proliferation of smooth muscle cells (SMCs) and adventitial fibroblasts (AFBs) but also augmentation of their migration. To test this hypothesis, growth-arrested SMCs and AFBs from rat aorta were exposed to norepinephrine. Norepinephrine caused dose-dependent migration of both cell types that was dependent on chemotaxis. In contrast, platelet-derived growth factor (PDGF)-BB, used as a positive control, stimulated both chemotaxis and chemokinesis. Only alpha(1D)-adrenoceptors (AR) and alpha(2)-AR antagonists inhibited norepinephrine migration of SMCs, whereas norepinephrine migration of AFBs was only inhibited by alpha(1A)-AR and alpha(1B)-AR antagonists; beta-AR blockade was without effect. Norepinephrine and PDGF-BB were additive for AFB, but not SMC, migration. Stimulation of migration was reversed at high norepinephrine concentrations (10 microM); this inhibition was mediated by alpha(2)- and beta-ARs in AFBs but not in SMCs. Thus norepinephrine induces migration of SMCs and AFBs via different alpha-ARs. This action may participate in wall remodeling and norepinephrine potentiation of injury-induced intimal lesion growth.  相似文献   

16.
Receptor density is an important determinant of cellular effector responses to receptor activation. We analysed cytosolic Ca(2+) responses to alpha(2)-adrenergic agents in PC12 cells expressing human alpha(2B)-adrenergic receptors (AR) at two densities (3.8 and 1.3 pmol/mg protein). The efficacy (E(max)) of agonists was greater in cells with higher receptor expression; while the potency (EC(50)) of norepinephrine and oxymetazoline was independent of alpha(2B)-AR levels. Several classical alpha(2)-AR antagonists behaved as either partial or inverse agonists in a receptor density-dependent fashion. No apparent structural similarities were found among the inverse agonists, precluding simple predictions of inverse agonist activity. Transfected PC12 cells expressing alpha(2B)-AR at relatively high density would be a useful approach to screen inverse agonists for this class of receptors. Our results further indicate that receptor density significantly influences the properties of ligands, not only of partial agonists as predicted by classical receptor theory, but also of antagonists and full agonists.  相似文献   

17.
Many G protein-coupled receptors (GPCRs) are internalized from the plasma membrane after agonist exposure. Previously, marked agonist-induced internalization of human alpha2A- and alpha2B-adrenergic receptors (AR) was observed in transfected neuronal rat pheochromocytoma (PC12) cells; alpha2A- and alpha2B-AR were internalized into partly distinct intracellular vesicles (Olli-L?hdesm?ki et al., J. Neurosci. 19, 9281-9288, 1999). In this paper, the extent of alpha2-AR internalization was quantitated in human embryonic kidney (HEK-293) and PC12 cells by combined application of cell surface biotinylation and ELISA methods, which allow measurement of protein trafficking in intact, differentiated and undifferentiated cells. Significant subtype-specific (but not cell type-dependent) trafficking of human alpha2-AR was observed by quantitation and immunocytochemistry. Agonist-induced sequestration of alpha2B-AR was markedly reduced after blocking the formation of clathrin-coated vesicles by hyperosmotic sucrose pretreatment. The sequestration of alpha2A-AR was partly inhibited after sucrose pretreatment but could be further reduced after inhibiting the formation of both clathrin-coated and caveolin vesicles by combined pretreatment with hyperosmotic sucrose and filipin. Differences were also observed in the recycling of alpha2A- and alpha2B-AR. The extent of maximal agonist-induced sequestration in PC12 cells was not directly dependent on relative agonist efficacy.  相似文献   

18.
Sex steroid hormones are important factors in the determination of fat distribution and accumulation. The aim of this study was to investigate the effect of testosterone (T), 17beta-estradiol (17betaE), and progesterone (P) on adrenergic receptor (AR) gene expression in 3T3-L1 preadipocytes and adipocytes and their relation to the proliferation and differentiation processes. Our data clearly show that alpha(2A)-AR was the highest AR subtype expressed in preadipocytes, whereas in mature adipocytes was by far beta(3)-AR. In the differentiation process to adipocytes, alpha(2A)-AR expression was decreased to 0.3-fold (P < 0.01), whereas beta(3)-AR was upregulated 578-fold (P < 0.001) compared with preadipocytes. In addition, the expression of alpha(2A)-AR in preadipocytes was increased upon incubation with T, 17betaE, and P, and a stimulation of proliferation was also observed in 17betaE- and P-treated cells. In mature adipocytes, 17betaE and P enhanced both alpha(2A)- and beta(3)-AR gene expression (although the effects on beta(3)-AR mRNA levels could be more relevant, since beta(3)-AR was the most highly expressed), whereas T only increased alpha(2A)-AR mRNA levels. Leptin and adipocyte fatty acid-binding protein mRNA levels were higher after 17betaE and P treatment, possibly indicating a proadipogenic effect of these hormones. In conclusion, this study indicates that AR gene expression is affected by these hormones in both preadipocytes and adipocytes, which could have potential importance when considering the role of ARs in the mechanisms underlying the sex-related differences in adipose tissue regional distribution.  相似文献   

19.
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   

20.
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   

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