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.
Phosducin-like protein (PhLP) is a widely expressed binding partner of the G protein betagamma subunit dimer (Gbetagamma). However, its physiological role is poorly understood. To investigate PhLP function, its cellular expression was blocked using RNA interference, resulting in inhibition of Gbetagamma expression and G protein signaling. This inhibition was caused by an inability of nascent Gbetagamma to form dimers. Phosphorylation of PhLP at serines 18-20 by protein kinase CK2 was required for Gbetagamma formation, while a high-affinity interaction of PhLP with the cytosolic chaperonin complex appeared unnecessary. PhLP bound nascent Gbeta in the absence of Ggamma, and S18-20 phosphorylation was required for Ggamma to associate with the PhLP-Gbeta complex. Once Ggamma bound, PhLP was released. These results suggest a mechanism for Gbetagamma assembly in which PhLP stabilizes the nascent Gbeta polypeptide until Ggamma can associate, resulting in membrane binding of Gbetagamma and release of PhLP to catalyze another round of assembly.  相似文献   

2.
3.
Phosducin-like protein (PhLP) is a widely expressed binding partner of the G protein betagamma subunit complex (Gbetagamma) that has been recently shown to catalyze the formation of the Gbetagamma dimer from its nascent polypeptides. Phosphorylation of PhLP at one or more of three consecutive serines (Ser-18, Ser-19, and Ser-20) is necessary for Gbetagamma dimer formation and is believed to be mediated by the protein kinase CK2. Moreover, several lines of evidence suggest that the cytosolic chaperonin complex (CCT) may work in concert with PhLP in the Gbetagamma-assembly process. The results reported here delineate a mechanism for Gbetagamma assembly in which a stable ternary complex is formed between PhLP, the nascent Gbeta subunit, and CCT that does not include Ggamma. PhLP phosphorylation permits the release of a PhLP x Gbeta intermediate from CCT, allowing Ggamma to associate with Gbeta in this intermediate complex. Subsequent interaction of Gbetagamma with membranes releases PhLP for another round of assembly.  相似文献   

4.
Phosducin-like protein (PhLP) exists in two splice variants PhLP(LONG) (PhLP(L)) and PhLP(SHORT) (PhLP(S)). Whereas PhLP(L) directly inhibits Gbetagamma-stimulated signaling, the G betagamma-inhibitory mechanism of PhLP(S) is not understood. We report here that inhibition of Gbetagamma signaling in intact HEK cells by PhLP(S) was independent of direct Gbetagamma binding; however, PhLP(S) caused down-regulation of Gbeta and Ggamma proteins. The down-regulation was partially suppressed by lactacystine, indicating the involvement of proteasomal degradation. N-terminal fusion of Gbeta or Ggamma with a dye-labeling protein resulted in their stabilization against down-regulation by PhLP(S) but did not lead to a functional rescue. Moreover, in the presence of PhLP(S), stabilized Ggamma subunits did not coprecipitate with stabilized Gbeta subunits, suggesting that PhLP(S) might interfere with Gbetagamma folding. PhLP(S) and several truncated mutants of PhLP(S) interacted with the subunit tailless complex polypeptide-1alpha (TCP-1alpha) of the CCT chaperonin complex, which is involved in protein folding. Knock-down of TCP-1alpha in HEK cells by small interfering RNA also led to down-regulation of Gbetagamma. We therefore conclude that the strong inhibitory action of PhLP(S) on Gbetagamma signaling is the result of a previously unrecognized mechanism of Gbetagamma-regulation, inhibition of Gbetagamma-folding by interference with TCP-1alpha.  相似文献   

5.
Phosducin proteins are known to inhibit G protein-mediated signaling by sequestering Gbetagamma subunits. However, Dictyostelium discoideum cells lacking the phosducin-like protein PhLP1 display defective rather than enhanced G protein signaling. Here we show that green fluorescent protein (GFP)-tagged Gbeta (GFP-Gbeta) and GFP-Ggamma subunits exhibit drastically reduced steady-state levels and are absent from the plasma membrane in phlp1(-) cells. Triton X-114 partitioning suggests that lipid attachment to GFP-Ggamma occurs in wild-type cells but not in phlp1(-) and gbeta(-) cells. Moreover, Gbetagamma dimers could not be detected in vitro in coimmunoprecipitation assays with phlp1(-) cell lysates. Accordingly, in vivo diffusion measurements using fluorescence correlation spectroscopy showed that while GFP-Ggamma proteins are present in a complex in wild-type cells, they are free in phlp1(-) and gbeta(-) cells. Collectively, our data strongly suggest the absence of Gbetagamma dimer formation in Dictyostelium cells lacking PhLP1. We propose that PhLP1 serves as a cochaperone assisting the assembly of Gbeta and Ggamma into a functional Gbetagamma complex. Thus, phosducin family proteins may fulfill hitherto unsuspected biosynthetic functions.  相似文献   

6.
The betagamma subunits of G proteins modulate inwardly rectifying potassium (GIRK) channels through direct interactions. Although GIRK currents are stimulated by mammalian Gbetagamma subunits, we show that they were inhibited by the yeast Gbetagamma (Ste4/Ste18) subunits. A chimera between the yeast and the mammalian Gbeta1 subunits (ymbeta) stimulated or inhibited GIRK currents, depending on whether it was co-expressed with mammalian or yeast Ggamma subunits, respectively. This result underscores the critical functional influence of the Ggamma subunits on the effectiveness of the Gbetagamma complex. A series of chimeras between Ggamma2 and the yeast Ggamma revealed that the C-terminal half of the Ggamma2 subunit is required for channel activation by the Gbetagamma complex. Point mutations of Ggamma2 to the corresponding yeast Ggamma residues identified several amino acids that reduced significantly the ability of Gbetagamma to stimulate channel activity, an effect that was not due to improper association with Gbeta. Most of the identified critical Ggamma residues clustered together, forming an intricate network of interactions with the Gbeta subunit, defining an interaction surface of the Gbetagamma complex with GIRK channels. These results show for the first time a functional role for Ggamma in the effector role of Gbetagamma.  相似文献   

7.
We have used fluorescence resonance energy transfer and co-immunoprecipitation to analyze the interactions among the alpha, beta, and gamma1 subunits of the Arabidopsis heterotrimeric G protein. Using cyan and yellow fluorescent protein fusion constructs, we show that overexpressed Ggamma1 localizes to protoplast membranes, but Gbeta exhibits membrane localization only when the Ggamma1 protein is co-overexpressed. Overexpressed Galpha shows membrane localization unaccompanied by overexpression of either Gbeta or Ggamma1. We detect fluorescence resonance energy transfer between Gbeta and Ggamma1 in the absence of Galpha overexpression and between Galpha and Ggamma1 but only when all three subunits are co-overexpressed. Both Galpha and Gbeta are associated with large macromolecular complexes of approximately 700 kDa in the plasma membrane. Galpha is present in both large complexes and as free Galpha in plasma membranes from wild type plants. In plants homozygous for a null allele of the Gbeta gene, Galpha is associated with smaller complexes in the 200-400-kDa range, indicating that its presence in the large complex depends on association with Gbetagamma. Activation of the Galpha subunit with guanosine 5'-3-O-(thio)triphosphate (GTPgammaS) results in partial dissociation of Galpha from the complex. Hydrogen peroxide (H2O2) promotes extensive dissociation of the Galpha complex but does not interfere with binding of GTPgammaS to purified recombinant Galpha, suggesting that reactive oxygen species affect the stability of the large complex but not the activity of Galpha itself.  相似文献   

8.
9.
Correct placement and orientation of the mitotic spindle is essential for segregation of localized components and positioning of daughter cells. Although these processes are important in many cells, few factors that regulate spindle placement are known. Previous work has shown that GPB-1, the Gbeta subunit of a heterotrimeric G protein, is required for orientation of early cell division axes in C. elegans embryos. Here we show that GOA-1 (a Galphao) and the related GPA-16 are the functionally redundant Galpha subunits and that GPC-2 is the relevant Ggamma subunit that is required for spindle orientation in the early embryo. We show that Galpha and Gbetagamma are involved in controlling distinct microtubule-dependent processes. Gbetagamma is important in regulating migration of the centrosome around the nucleus and hence in orientating the mitotic spindle. Galpha is required for asymmetric spindle positioning in the one-celled embryo.  相似文献   

10.
PLC-epsilon was identified recently as a phosphoinositide-hydrolyzing phospholipase C (PLC) containing catalytic domains (X, Y, and C2) common to all PLC isozymes as well as unique CDC25- and Ras-associating domains. Novel regulation of this PLC isozyme by the Ras oncoprotein and alpha-subunits (Galpha(12)) of heterotrimeric G proteins was illustrated. Sequence analyses of PLC-epsilon revealed previously unrecognized PH and EF-hand domains in the amino terminus. The known interaction of Gbetagamma subunits with the PH domains of other proteins led us to examine the capacity of Gbetagamma to activate PLC-epsilon. Co-expression of Gbeta(1)gamma(2) with PLC-epsilon in COS-7 cells resulted in marked stimulation of phospholipase C activity. Gbeta(2) and Gbeta(4) in combination with Ggamma(1), Ggamma(2), Ggamma(3), or Ggamma(13) also activated PLC-epsilon to levels similar to those observed with Gbeta(1)-containing dimers of these Ggamma-subunits. Gbeta(3) in combination with the same Ggamma-subunits was less active, and Gbeta(5)-containing dimers were essentially inactive. Gbetagamma-promoted activation of PLC-epsilon was blocked by cotransfection with either of two Gbetagamma-interacting proteins, Galpha(i1) or the carboxyl terminus of G protein receptor kinase 2. Pharmacological inhibition of PI3-kinase-gamma had no effect on Gbeta(1)gamma(2)-promoted activation of PLC-epsilon. Similarly, activation of Ras in the action of Gbetagamma is unlikely, because a mutation in the second RA domain of PLC-epsilon that blocks Ras activation of PLC failed to alter the stimulatory activity of Gbeta(1)gamma(2). Taken together, these results reveal the presence of additional functional domains in PLC-epsilon and add a new level of complexity in the regulation of this novel enzyme by heterotrimeric G proteins.  相似文献   

11.
The signal transducing function of Gbeta(5) in brain is unknown. When studied in vitro Gbeta(5) is the only heterotrimeric Gbeta subunit known to interact with both Ggamma subunits and regulators of G protein signaling (RGS) proteins. When tested with Ggamma, Gbeta(5) interacts with other classical components of heterotrimeric G protein signaling pathways such as Galpha and phospholipase C-beta. We recently demonstrated nuclear expression of Gbeta(5) in neurons and brain (Zhang, J. H., Barr, V. A., Mo, Y., Rojkova, A. M., Liu, S., and Simonds, W. F. (2001) J. Biol. Chem. 276, 10284-10289). To gain further insight into the mechanism of Gbeta(5) nuclear localization, we generated a Gbeta(5) mutant deficient in its ability to interact with RGS7 while retaining its ability to bind Ggamma, and we compared its properties to the wild-type Gbeta(5). In HEK-293 cells co-transfection of RGS7 but not Ggamma(2) supported expression in the nuclear fraction of transfected wild-type Gbeta(5). In contrast the Ggamma-preferring Gbeta(5) mutant was not expressed in the HEK-293 cell nuclear fraction with either co-transfectant. The Ggamma-selective Gbeta(5) mutant was also excluded from the cell nucleus of transfected PC12 cells analyzed by laser confocal microscopy. These results define a requirement for RGS protein binding for Gbeta(5) nuclear expression.  相似文献   

12.
13.
The Arabidopsis thaliana heterotrimeric G protein complex is encoded by single canonical Galpha and Gbeta subunit genes and two Ggamma subunit genes (AGG1 and AGG2), raising the possibility that the two potential G protein complexes mediate different cellular processes. Mutants with reduced expression of one or both Ggamma genes revealed specialized roles for each Ggamma subunit. AGG1-deficient mutants, but not AGG2-deficient mutants, showed impaired resistance against necrotrophic pathogens, reduced induction of the plant defensin gene PDF1.2, and decreased sensitivity to methyl jasmonate. By contrast, both AGG1- and AGG2-deficient mutants were hypersensitive to auxin-mediated induction of lateral roots, suggesting that Gbetagamma1 and Gbetagamma2 synergistically inhibit auxin-dependent lateral root initiation. However, the involvement of each Ggamma subunit in this root response differs, with Gbetagamma1 acting within the central cylinder, attenuating acropetally transported auxin signaling, while Gbetagamma2 affects the action of basipetal auxin and graviresponsiveness within the epidermis and/or cortex. This selectivity also operates in the hypocotyl. Selectivity in Gbetagamma signaling was also found in other known AGB1-mediated pathways. agg1 mutants were hypersensitive to glucose and the osmotic agent mannitol during seed germination, while agg2 mutants were only affected by glucose. We show that both Ggamma subunits form functional Gbetagamma dimers and that each provides functional selectivity to the plant heterotrimeric G proteins, revealing a mechanism underlying the complexity of G protein-mediated signaling in plants.  相似文献   

14.
While positive regulation of c-Akt (also known as protein kinase B) by receptor tyrosine kinases is well documented, compounds acting through G protein-coupled receptors can also activate Akt and its downstream targets. We therefore explored the role of G protein subunits in the regulation of Akt in cultured mammalian cells. In HEK-293 and COS-7 cells transiently transfected with beta(2)-adrenergic or m2 muscarinic receptors, respectively, treatment with agonist-induced phosphorylation of Akt at serine 473 as evidenced by phosphoserine-specific immunoblots. This effect was blocked by the phosphatidylinositol-3-OH kinase inhibitor LY294002 and wild-type Galpha(i1), and was not duplicated by co-transfection of the constitutively active Galpha(s)-Q227L or Galpha(i)-Q204L mutant. Co-transfection of Gbeta(1), Gbeta(2) but not Gbeta(5) together with Ggamma(2) activated the kinase when assayed in vitro following immunoprecipitation of the epitope-tagged enzyme. In contrast, constitutively activated G protein subunits representing the four Galpha subfamilies were found unable to activate Akt in either cell line. The latter results are in disagreement with a report by Murga et al. (Murga, C., Laguinge, L., Wetzker, R., Cuadrado, A., and Gutkind, J. S. (1998) J. Biol. Chem. 273, 19080-19085) that described activation of Akt in response to mutationally activated Galpha(q) and Galpha(i) transfection in COS cells. To the contrary, in our experiments Galpha(q)-Q209L inhibited Akt activation resulting from betagamma or mutationally activated H-Ras co-transfection in these cells. In HEK-293 cells Galpha(q)-Q209L transfection inhibited insulin-like growth factor-1 activation of epitope-tagged Akt. In m1 muscarinic receptor transfected HEK-293 cells, carbachol inhibited insulin-like growth factor-1 stimulated phosphorylation at Ser(473) of endogenous Akt in an atropine-reversible fashion. We conclude that G proteins can regulate Akt by two distinct and potentially opposing mechanisms: activation by Gbetagamma heterodimers in a phosphatidylinositol-3-OH kinase-dependent fashion, and inhibition mediated by Galpha(q). This work identifies Akt as a novel point of convergence between disparate signaling pathways.  相似文献   

15.
The Gbeta and Ggamma subunit of the heterotrimeric G proteins form a functional dimer that is stable once assembled in vivo or in vitro. The requirements, mechanism, and specificity of dimer formation are still incompletely understood, but represent important biochemical processes involved in the specificity of cellular signaling through G proteins. Here, seven Gbeta and 12 FLAG-epitope-tagged Ggamma subunits were separately synthesized in vitro using a rabbit reticulocyte lysate expression system. The translation products were combined and dimers isolated by immunoprecipitation. Gbeta1 and Gbeta4 formed dimers with all Ggamma subunit isoforms, generally with Gbeta/Ggamma stoichiometries between 0.2:1 and 0.5:1. Gbeta5, Gbeta5L, and Gbeta3s did not form significant amounts of dimer with any of the gamma subunit isoforms. Gbeta2 and Gbeta3 formed dimers with selected Ggamma isoforms to levels intermediate between that of Gbeta1/Gbeta4 and Gbeta3s/Gbeta5/Gbeta5L. We also expressed selected Gbetagamma in HEK293 cells and measured PLCbeta2 activity. Gbetagamma dimer-dependent increases in IP3 production were seen with most Gbeta1, Gbeta2, and Gbeta5 combinations, indicating functional dimer expression in intact cells. These results define the complete set of G protein betagamma dimers that are formed using a single biochemical assay method and suggest that there are Gbeta isoform-specific factors in rabbit reticulocyte lysates that determine the efficacy of Gbetagamma dimer formation.  相似文献   

16.
Regulator of G-protein signaling 3 (RGS3) enhances the intrinsic rate at which Galpha(i) and Galpha(q) hydrolyze GTP to GDP, thereby limiting the duration in which GTP-Galpha(i) and GTP-Galpha(q) can activate effectors. Since GDP-Galpha subunits rapidly combine with free Gbetagamma subunits to reform inactive heterotrimeric G-proteins, RGS3 and other RGS proteins may also reduce the amount of Gbetagamma subunits available for effector interactions. Although RGS6, RGS7, and RGS11 bind Gbeta(5) in the absence of a Ggamma subunit, RGS proteins are not known to directly influence Gbetagamma signaling. Here we show that RGS3 binds Gbeta(1)gamma(2) subunits and limits their ability to trigger the production of inositol phosphates and the activation of Akt and mitogen-activated protein kinase. Co-expression of RGS3 with Gbeta(1)gamma(2) inhibits Gbeta(1)gamma(2)-induced inositol phosphate production and Akt activation in COS-7 cells and mitogen-activated protein kinase activation in HEK 293 cells. The inhibition of Gbeta(1)gamma(2) signaling does not require an intact RGS domain but depends upon two regions in RGS3 located between acids 313 and 390 and between 391 and 458. Several other RGS proteins do not affect Gbeta(1)gamma(2) signaling in these assays. Consistent with the in vivo results, RGS3 inhibits Gbetagamma-mediated activation of phospholipase Cbeta in vitro. Thus, RGS3 may limit Gbetagamma signaling not only by virtue of its GTPase-activating protein activity for Galpha subunits, but also by directly interfering with the activation of effectors.  相似文献   

17.
18.
One major class of G proteins typically functions as heterotrimeric complexes consisting of Galpha, Gbeta and Ggamma subunits. However, recent work in yeast has identified an atypical Galpha protein, Gpa2p, which functions without cognate Gbetagamma subunits. Two novel kelch repeat protein binding partners of Gpa2p, Krh1p and Krh2p, do not function as alternative Gbeta subunits, as initially thought, but rather as Gpa2p effectors. They directly link Gpa2p to protein kinase A, thus forming an adenylate cyclase bypass pathway that enables inputs other than cellular cAMP concentration to affect protein kinase A activity. Because mammalian protein kinase A expressed in yeast is also subject to control by the same bypass pathway, it is exciting to postulate that a functionally similar mechanism might exist in mammalian cells, and that other Galpha proteins could exhibit similar characteristics to Gpa2p.  相似文献   

19.
G proteins transmit a variety of extracellular signals into intracellular responses. The Galpha and Gbetagamma subunits are both known to regulate effectors. Interestingly, the Galpha subunit also determines subtype specificity of Gbetagamma effector interactions. However, in light of the common paradigm that Galpha and Gbetagamma subunits dissociate during activation, a plausible mechanism of how this subtype specificity is generated was lacking. Using a fluorescence resonance energy transfer (FRET)-based assay developed to directly measure mammalian G protein activation in intact cells, we demonstrate that fluorescent Galpha(i1,2,3), Galpha(z), and Gbeta(1)gamma(2) subunits do not dissociate during activation but rather undergo subunit rearrangement as indicated by an activation-induced increase in FRET. In contrast, fluorescent Galpha(o) subunits exhibited an activation-induced decrease in FRET, reflecting subunit dissociation or, alternatively, a distinct subunit rearrangement. The alpha(B/C)-region within the alpha-helical domain, which is much more conserved within Galpha(i1,2,3) and Galpha(z) as compared with that in Galpha(o), was found to be required for exhibition of an activation-induced increase in FRET between fluorescent Galpha and Gbetagamma subunits. However, the alpha(B/C)-region of Galpha(il) alone was not sufficient to transfer the activation pattern of Galpha(i) to the Galpha(o) subunit. Either residues in the first 91 amino acids or in the C-terminal remainder (amino acids 93-354) of Galpha(il) together with the alpha(B/C)-helical region of Galpha(i1) were needed to transform the Galpha(o)-activation pattern into a Galpha(i1)-type of activation. The discovery of subtype-selective mechanisms of G protein activation illustrates that G protein subfamilies have specific mechanisms of activation that may provide a previously unknown basis for G protein signaling specificity.  相似文献   

20.
In vitro, little specificity is seen for modulation of effectors by different combinations of Gbetagamma subunits from heterotrimeric G proteins. Here, we demonstrate that the coupling of specific combinations of Gbetagamma subunits to different receptors leads to a differential ability to modulate effectors in vivo. We have shown that the beta(1)AR and beta(2)AR can activate homomultimers of the human inwardly rectifying potassium channel Kir 3.2 when coexpressed in Xenopus oocytes, and that this requires a functional mammalian Gs heterotrimer. Modulation was independent of cAMP production, suggesting a membrane-delimited mechanism. To analyze further the importance of different Gbetagamma combinations, we have tested the facilitation of Kir 3.2 activation by betaAR mediated by different Gbetagamma subunits. The subunits tested were Gbeta(1,5) and Ggamma(1,2,7,11). These experiments demonstrated significant variation between the ability of the Gbetagamma combinations to activate the channels after receptor stimulation. This was in marked contrast to the situation in vitro where little specificity for binding of a Kir 3.1 C-terminal GST fusion protein by different Gbetagamma combinations was detected. More importantly, neither receptor, although homologous both structurally and functionally, shared the same preference for Gbetagamma subunits. In the presence of beta(1)AR, Gbeta(5)gamma(1) and Gbeta(5)gamma(11) activated Kir 3.2 to the greatest extent, while for the beta(2)AR, Gbeta(1)gamma(7), Gbeta(1)gamma(11,) and Gbeta(5)gamma(2) produced the greatest responses. Interestingly, no preference was seen in the ability of different Gbetagamma subunits to facilitate receptor-stimulated GTPase activity of the Gsalpha. These results suggest that it is not the receptor/G protein alpha subunit interaction or the Gbetagamma/effector interaction that is altered by Gbetagamma, but rather that the ability of the receptor to interact productively with the Gbetagamma subunit directly and/or the G protein/effector complex is dependent on the specific G protein heterotrimer associated with the receptor.  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