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 62 毫秒
1.
Recent work has suggested a role for the serine/threonine kinase Akt and IkappaB kinases (IKKs) in nuclear factor (NF)-kappaB activation. In this study, the involvement of these components in NF-kappaB activation through a G protein-coupled pathway was examined using transfected HeLa cells that express the B2-type bradykinin (BK) receptor. The function of IKK2, and to a lesser extent, IKK1, was suggested by BK-induced activation of their kinase activities and by the ability of their dominant negative mutants to inhibit BK-induced NF-kappaB activation. BK-induced NF-kappaB activation and IKK2 activity were markedly inhibited by RGS3T, a regulator of G protein signaling that inhibits Galpha(q), and by two Gbetagamma scavengers. Co-expression of Galpha(q) potentiated BK-induced NF-kappaB activation, whereas co-expression of either an activated Galpha(q)(Q209L) or Gbeta(1)gamma(2) induced IKK2 activity and NF-kappaB activation without BK stimulation. BK-induced NF-kappaB activation was partially blocked by LY294002 and by a dominant negative mutant of phosphoinositide 3-kinase (PI3K), suggesting that PI3K is a downstream effector of Galpha(q) and Gbeta(1)gamma(2) for NF-kappaB activation. Furthermore, BK could activate the PI3K downstream kinase Akt, whereas a catalytically inactive mutant of Akt inhibited BK-induced NF-kappaB activation. Taken together, these findings suggest that BK utilizes a signaling pathway that involves Galpha(q), Gbeta(1)gamma(2), PI3K, Akt, and IKK for NF-kappaB activation.  相似文献   

2.
3.
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   

4.
G-protein coupled receptors (GPCRs) form a ternary complex of agonist, receptor and G-proteins during primary signal transduction at the cell membrane. Downstream signalling is thought to be preceded by the process of dissociation of Galpha and Gbetagamma subunits, thus exposing new surfaces to interact with downstream effectors. We demonstrate here for the first time, the dissociation of heterotrimeric G-protein subunits (i.e., Galpha and Gbetagamma) following agonist-induced GPCR (alpha(2A)-adrenergic receptor; alpha(2A)-AR) activation in a cell-free assay system. alpha(2A)-AR membranes were reconstituted with the G-proteins (+/-hexahistidine-tagged) Galpha(i1) and Gbeta1gamma2 and functional signalling was determined following activation of the reconstituted receptor:G-protein complex with the potent agonist UK-14304, and [35S]GTPgammaS. In the presence of Ni(2+)-coated agarose beads, the activated his-tagged Galpha(i1)his-[35S]GTPgammaS complex was captured on the Ni(2+)-presenting surface. When his-tagged Gbeta1gamma2 (Gbeta1gamma2his) was used with Galpha(i1), the [35S]GTPgammaS-bound Galpha(i1) was not present on the Ni(2+)-coated beads, but rather, it was separated from the beta1gamma2(his)-beads, demonstrating receptor-induced dissociation of Galpha and Gbetagamma subunits. Treatment of the reconstituted alpha(2A)-AR membranes containing Gbeta1gamma2his:Galpha(i1) with imidazole confirmed the specificity for the Ni2+:G-protein surface dissociation of Galpha(i1) from Gbeta1gamma2his. These data demonstrate for the first time, the complete dissociation of the G-protein subunits and extend observations on the role of G-proteins in the assembly and disassembly of the ternary complex in the primary events of GPCR signalling.  相似文献   

5.
A yeast two-hybrid approach was used to discern possible new effectors for the betagamma subunit of heterotrimeric G proteins. Three of the clones isolated are structurally similar to Gbeta, each exhibiting the WD40 repeat motif. Two of these proteins, the receptor for activated C kinase 1 (RACK1) and the dynein intermediate chain, co-immunoprecipitate with Gbetagamma using an anti-Gbeta antibody. The third protein, AAH20044, has no known function; however, sequence analysis indicates that it is a WD40 repeat protein. Further investigation with RACK1 shows that it not only interacts with Gbeta(1)gamma(1) but also unexpectedly with the transducin heterotrimer Galpha(t)beta(1)gamma(1). Galpha(t) alone does not interact, but it must contribute to the interaction because the apparent EC(50) value of RACK1 for Galpha(t)beta(1)gamma(1) is 3-fold greater than that for Gbeta(1)gamma(1) (0.1 versus 0.3 microm). RACK1 is a scaffold that interacts with several proteins, among which are activated betaIIPKC and dynamin-1 (1). betaIIPKC and dynamin-1 compete with Gbeta(1)gamma(1) and Galpha(t)beta(1)gamma(1) for interaction with RACK1. These findings have several implications: 1) that WD40 repeat proteins may interact with each other; 2) that Gbetagamma interacts differently with RACK1 than with its other known effectors; and/or 3) that the G protein-RACK1 complex may constitute a signaling scaffold important for intracellular responses.  相似文献   

6.
Control of cell proliferation depends on intracellular mediators that determine the cellular response to external cues. In neuroendocrine cells, the dopamine D2 receptor short form (D2S receptor) inhibits cell proliferation, whereas in mesenchymal cells the same receptor enhances cell proliferation. Nontransformed BALB/c 3T3 fibroblast cells were stably transfected with the D2S receptor cDNA to study the G proteins that direct D2S signaling to stimulate cell proliferation. Pertussis toxin inactivates G(i) and G(o) proteins and blocks signaling of the D2S receptor in these cells. D2S receptor signaling was reconstituted by individually transfecting pertussis toxin-resistant Galpha(i/o) subunit mutants and measuring D2-induced responses in pertussis toxin-treated cells. This approach identified Galpha(i)2 and Galpha(i)3 as mediators of the D2S receptor-mediated inhibition of forskolin-stimulated adenylyl cyclase activity; Galpha(i)2-mediated D2S-induced stimulation of p42 and p44 mitogen-activated kinase (MAPK) and DNA synthesis, whereas Galpha(i)3 was required for formation of transformed foci. Transfection of toxin-resistant Galpha(i)1 cDNA induced abnormal cell growth independent of D2S receptor activation, while Galpha(o) inhibited dopamine-induced transformation. The role of Gbetagamma subunits was assessed by ectopic expression of the carboxyl-terminal domain of G protein receptor kinase to selectively antagonize Gbetagamma activity. Mobilization of Gbetagamma subunits was required for D2S-induced calcium mobilization, MAPK activation, and DNA synthesis. These findings reveal a remarkable and distinct G protein specificity for D2S receptor-mediated signaling to initiate DNA synthesis (Galpha(i)2 and Gbetagamma) and oncogenic transformation (Galpha(i)3), and they indicate that acute activation of MAPK correlates with enhanced DNA synthesis but not with transformation.  相似文献   

7.
BACKGROUND: Gbeta proteins have traditionally been thought to complex with Ggamma proteins to function as subunits of G protein heterotrimers. The divergent Gbeta(5) protein, however, can bind either Ggamma proteins or regulator of G protein signaling (RGS) proteins that contain a G gamma-like (GGL) domain. RGS proteins inhibit G protein signaling by acting as Galpha GTPase activators. While Gbeta(5) appears to bind RGS proteins in vivo, its association with Ggamma proteins in vivo has not been clearly demonstrated. It is unclear how Gbeta(5) might influence RGS activity. In C. elegans there are exactly two GGL-containing RGS proteins, EGL-10 and EAT-16, and they inhibit Galpha(o) and Galpha(q) signaling, respectively. RESULTS: We knocked out the gene encoding the C. elegans Gbeta(5) ortholog, GPB-2, to determine its physiological roles in G protein signaling. The gpb-2 mutation reduces the functions of EGL-10 and EAT-16 to levels comparable to those found in egl-10 and eat-16 null mutants. gpb-2 knockout animals are viable, and exhibit no obvious defects beyond those that can be attributed to a reduction of EGL-10 or EAT-16 function. GPB-2 protein is nearly absent in eat-16; egl-10 double mutants, and EGL-10 protein is severely diminished in gpb-2 mutants. CONCLUSIONS: Gbeta(5) functions in vivo complexed with GGL-containing RGS proteins. In the absence of Gbeta(5), these RGS proteins have little or no function. The formation of RGS-Gbeta(5) complexes is required for the expression or stability of both the RGS and Gbeta(5) proteins. Appropriate RGS-Gbeta(5) complexes regulate both Galpha(o) and Galpha(q) proteins in vivo.  相似文献   

8.
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   

9.
Menard RE  Mattingly RR 《FEBS letters》2004,556(1-3):187-192
The p21-activated kinase (PAK) family is homologous to the yeast sterile 20 (Ste20) and regulates a wide variety of cellular responses, including cell morphology, proliferation, and survival. In this study we examined the activation of PAK1 by Gbetagamma subunits. Co-transfection of COS7 cells with Gbeta1gamma2 or Gbeta1gamma5 was sufficient to induce agonist-independent activation of PAK1. Expression of dominant/negative Rac, Cdc42, or Ras did not inhibit this Gbetagamma-dependent activation. Wortmannin, which inhibits phosphoinositide 3-kinase (PI3-kinase) activity, and expression of a dominant/negative form of Akt were sufficient to abrogate the activation of PAK1 that was induced by Gbetagamma. These results reveal that stimulation of PAK1 by Gbetagamma can occur via a PI3-kinase and Akt pathway that does not require Rac1 or Cdc42.  相似文献   

10.
Regulator of G-protein signaling (RGS) proteins are GTPase activating proteins (GAPs) of heterotrimeric G-proteins that alter the amplitude and kinetics of receptor-promoted signaling. In this study we defined the G-protein alpha-subunit selectivity of purified Sf9 cell-derived R7 proteins, a subfamily of RGS proteins (RGS6, -7, -9, and -11) containing a Ggamma-like (GGL) domain that mediates dimeric interaction with Gbeta(5). Gbeta(5)/R7 dimers stimulated steady state GTPase activity of Galpha-subunits of the G(i) family, but not of Galpha(q) or Galpha(11), when added to proteoliposomes containing M2 or M1 muscarinic receptor-coupled G-protein heterotrimers. Concentration effect curves of the Gbeta(5)/R7 proteins revealed differences in potencies and efficacies toward Galpha-subunits of the G(i) family. Although all four Gbeta(5)/R7 proteins exhibited similar potencies toward Galpha(o), Gbeta(5)/RGS9 and Gbeta(5)/RGS11 were more potent GAPs of Galpha(i1), Galpha(i2), and Galpha(i3) than were Gbeta(5)/RGS6 and Gbeta(5)/RGS7. The maximal GAP activity exhibited by Gbeta(5)/RGS11 was 2- to 4-fold higher than that of Gbeta(5)/RGS7 and Gbeta(5)/RGS9, with Gbeta(5)/RGS6 exhibiting an intermediate maximal GAP activity. Moreover, the less efficacious Gbeta(5)/RGS7 and Gbeta(5)/RGS9 inhibited Gbeta(5)/RGS11-stimulated GTPase activity of Galpha(o). Therefore, R7 family RGS proteins are G(i) family-selective GAPs with potentially important differences in activities.  相似文献   

11.
In C. elegans, a G(o)/G(q) signaling network regulates locomotion and egg laying [1-8]. Genetic analysis shows that activated Ca(2+)/calmodulin-dependent protein kinase II (CaMKII) is suppressed by perturbations of this network, which include loss of the GOA-1 G(o)alpha, DGK-1 diacylglycerol kinase, EAT-16 G protein gamma subunit-like (GGL)-containing RGS protein, or an unidentified protein encoded by the gene eat-11 [9]. We cloned eat-11 and report that it encodes the Gbeta(5) ortholog GPB-2. Gbeta(5) binds specifically to GGL-containing RGS proteins, and the Gbeta(5)/RGS complex can promote the GTP-hydrolyzing activity of Galpha subunits [10, 11]. However, little is known about how this interaction affects G protein signaling in vivo. In addition to EAT-16, the GGL-containing RGS protein EGL-10 participates in G(o)/G(q) signaling; EGL-10 appears to act as an RGS for the GOA-1 G(o)alpha, while EAT-16 appears to act as an RGS for the EGL-30 G(q)alpha [4, 5]. We have combined behavioral, electrophysiological, and pharmacological approaches to show that GPB-2 is a central member of the G(o)/G(q) network and that GPB-2 may interact with both the EGL-10 and EAT-16 RGS proteins to mediate the opposing activities of G(o)alpha and G(q)alpha. These interactions provide a mechanism for the modulation of behavior by antagonistic G protein networks.  相似文献   

12.
Regulators of G protein signaling (RGS) proteins that contain DEP (disheveled, EGL-10, pleckstrin) and GGL (G protein gamma subunit-like) domains form a subfamily that includes the mammalian RGS proteins RGS6, RGS7, RGS9, and RGS11. We describe the cloning of RGS6 cDNA, the specificity of interaction of RGS6 and RGS7 with G protein beta subunits, and certain biochemical properties of RGS6/beta5 and RGS7/beta5 complexes. After expression in Sf9 cells, complexes of both RGS6 and RGS7 with the Gbeta5 subunit (but not Gbetas 1-4) are found in the cytosol. When purified, these complexes are similar to RGS11/beta5 in that they act as GTPase-activating proteins specifically toward Galpha(o). Unlike conventional G(betagamma) complexes, RGS6/beta5 and RGS7/beta5 do not form heterotrimeric complexes with either Galpha(o)-GDP or Galpha(q)-GDP. Neither RGS6/beta5 nor RGS7/beta5 altered the activity of adenylyl cyclases types I, II, or V, nor were they able to activate either phospholipase C-beta1 or -beta2. However, the RGS/beta5 complexes inhibited beta(1)gamma(2)-mediated activation of phospholipase C-beta2. RGS/beta5 complexes may contribute to the selectivity of signal transduction initiated by receptors coupled to G(i) and G(o) by binding to phospholipase C and stimulating the GTPase activity of Galpha(o).  相似文献   

13.
14.
The GAPs, GEFs, and GDIs of heterotrimeric G-protein alpha subunits   总被引:1,自引:0,他引:1  
The heterotrimeric G-protein alpha subunit has long been considered a bimodal, GTP-hydrolyzing switch controlling the duration of signal transduction by seven-transmembrane domain (7TM) cell-surface receptors. In 1996, we and others identified a superfamily of "regulator of G-protein signaling" (RGS) proteins that accelerate the rate of GTP hydrolysis by Galpha subunits (dubbed GTPase-accelerating protein or "GAP" activity). This discovery resolved the paradox between the rapid physiological timing seen for 7TM receptor signal transduction in vivo and the slow rates of GTP hydrolysis exhibited by purified Galpha subunits in vitro. Here, we review more recent discoveries that have highlighted newly-appreciated roles for RGS proteins beyond mere negative regulators of 7TM signaling. These new roles include the RGS-box-containing, RhoA-specific guanine nucleotide exchange factors (RGS-RhoGEFs) that serve as Galpha effectors to couple 7TM and semaphorin receptor signaling to RhoA activation, the potential for RGS12 to serve as a nexus for signaling from tyrosine kinases and G-proteins of both the Galpha and Ras-superfamilies, the potential for R7-subfamily RGS proteins to couple Galpha subunits to 7TM receptors in the absence of conventional Gbetagamma dimers, and the potential for the conjoint 7TM/RGS-box Arabidopsis protein AtRGS1 to serve as a ligand-operated GAP for the plant Galpha AtGPA1. Moreover, we review the discovery of novel biochemical activities that also impinge on the guanine nucleotide binding and hydrolysis cycle of Galpha subunits: namely, the guanine nucleotide dissociation inhibitor (GDI) activity of the GoLoco motif-containing proteins and the 7TM receptor-independent guanine nucleotide exchange factor (GEF) activity of Ric8/synembryn. Discovery of these novel GAP, GDI, and GEF activities have helped to illuminate a new role for Galpha subunit GDP/GTP cycling required for microtubule force generation and mitotic spindle function in chromosomal segregation.  相似文献   

15.
In mouse periaqueductal gray matter (PAG) membranes, the mu-opioid receptor (MOR) coprecipitated the alpha-subunits of the Gi/o/z/q/11 proteins, the Gbeta1/2 subunits, and the regulator of G-protein signaling RGS9-2 and its partner protein Gbeta5. RGS7 and RGS11 present in this neural structure showed no association with MOR. In vivo intracerebroventricular injection of morphine did not alter MOR immunoreactivity, but 30 min and 3 h after administration, the coprecipitation of Galpha subunits with MORs was reduced by up to 50%. Furthermore, the association between Galpha subunits and RGS9-2 proteins was increased. Twenty-four hours after receiving intracerebroventricular morphine, the Galpha subunits left the RGS9-2 proteins and re-associated with the MORs. However, doses of the opioid able to induce tolerance promoted the stable transfer of Galpha subunits to the RGS9-2 control. This was accompanied by Ser phosphorylation of RGS9-2 proteins, which increased their co-precipitation with 14-3-3 proteins. In the PAG membranes of morphine-desensitized mice, the capacity of the opioid to stimulate G-protein-related guanosine 5'-O-(3-[35S]thiotriphosphate) binding as well as low Km GTPase activity was attenuated. The in vivo knockdown of RGS9-2 expression prevented morphine from altering the association between MORs and G-proteins, and tolerance did not develop. In PAG membranes from RGS9-2 knockdown mice, morphine showed full capacity to activate G-proteins. Thus, the tolerance that develops following an adequate dose of morphine is caused by the stabilization and retention of MOR-activated Galpha subunits by RGS9-2 proteins. This multistep process is initiated by the morphine-induced transfer of MOR-associated Galpha subunits to the RGS9-2 proteins, followed by Ser phosphorylation of the latter and their binding to 14-3-3 proteins. This regulatory mechanism probably precedes the loss of MORs from the cell membrane, which has been observed with other opioid agonists.  相似文献   

16.
17beta-estradiol and 1,25-dihydroxyvitamin D(3)()(calcitriol) rapidly increase (< 5 sec) the concentration of intracellular calcium by mobilizing Ca(2+) from the endoplasmic reticulum and forming inositol 1,4,5-trisphosphate (InsP(3)) and diacylglycerol. Calcitriol increases InsP(3) formation via activation of phospholipase C (PLC)-beta1 linked to a pertussis toxin (PTX)-insensitive G-protein, and estradiol via activation of PLC-beta2 linked to a PTX-sensitive G-protein. Since PLC are effectors of different subunits of various G-proteins, we looked for and identified several G-subunits (Galpha(q/11), Galphas, Galphai, Gbeta and Ggamma) in female rat osteoblasts using Western immunoblotting. The action of calcitriol on InsP(3) formation and Ca(2+) mobilization in Fura-2-loaded confluent osteoblasts involved Galpha(q/11). The membrane effects of estradiol involved Gbetagamma; subunits, and principally Gbeta subunits, but not alpha-subunits. These results may provide additional evidence for membrane receptors of steroid hormones. Since PLC-beta1 is the target effector of Galpha(q/11), whereas PLC-beta2 is only activated by betagamma subunits, this specificity may help to generate membrane receptor-specific responses in vivo.  相似文献   

17.
Chinese hamster embryonic fibroblasts (IIC9 cells) express the Galpha subunits Galphas, Galphai2, Galphai3, Galphao, Galpha(q/11), and Galpha13. Consistent with reports in other cell types, alpha-thrombin stimulates a subset of the expressed G proteins in IIC9 cells, namely Gi2, G13, and Gq as measured by an in vitro membrane [35S]guanosine 5'-O-(3-thio)triphosphate binding assay. Using specific Galpha peptides, which block coupling of G-protein receptors to selective G proteins, as well as dominant negative xanthine nucleotide-binding Galpha mutants, we show that activation of the phosphatidylinositol 3-kinase/Akt pathway is dependent on Gq and Gi2. To examine the role of the two G proteins, we examined the events upstream of PI 3-kinase. The activation of the PI 3-kinase/Akt pathway by alpha-thrombin in IIC9 cells is blocked by the expression of dominant negative Ras and beta-arrestin1 (Phillips-Mason, P. J., Raben, D. M., and Baldassare, J. J. (2000) J. Biol. Chem. 275, 18046-18053, and Goel, R., Phillips-Mason, P. J., Raben, D. M., and Baldassare, J. J. (2002) J. Biol. Chem. 277, 18640-18648), indicating a role for Ras and beta-arrestin1. Interestingly, inhibition of Gi2 and Gq activation blocks Ras activation and beta-arrestin1 membrane translocation, respectively. Furthermore, expression of the Gbetagamma sequestrant, alpha-transducin, inhibits both Ras activation and membrane translocation of beta-arrestin1, suggesting that Gbetagamma dimers from Galphai2 and Galphaq activate different effectors to coordinately regulate the PI 3-kinase/Akt pathway.  相似文献   

18.
RACK1 regulates specific functions of Gbetagamma   总被引:6,自引:0,他引:6  
We showed previously that Gbetagamma interacts with Receptor for Activated C Kinase 1 (RACK1), a protein that not only binds activated protein kinase C (PKC) but also serves as an adaptor/scaffold for many signaling pathways. Here we report that RACK1 does not interact with Galpha subunits or heterotrimeric G proteins but binds free Gbetagamma subunits released from activated heterotrimeric G proteins following the activation of their cognate receptors in vivo. The association with Gbetagamma promotes the translocation of RACK1 from the cytosol to the membrane. Moreover, binding of RACK1 to Gbetagamma results in inhibition of Gbetagamma-mediated activation of phospholipase C beta2 and adenylyl cyclase II. However, RACK1 has no effect on other functions of Gbetagamma, such as activation of the mitogen-activated protein kinase signaling pathway or chemotaxis of HEK293 cells via the chemokine receptor CXCR2. Similarly, RACK1 does not affect signal transduction through the Galpha subunits of G(i), G(s), or G(q). Collectively, these findings suggest a role of RACK1 in regulating specific functions of Gbetagamma.  相似文献   

19.
Accumulating evidence suggests that heterotrimeric G protein activation may not require G protein subunit dissociation. Results presented here provide evidence for a subunit dissociation-independent mechanism for G protein activation by a receptor-independent activator of G protein signaling, AGS8. AGS8 is a member of the AGS group III family of AGS proteins thought to activate G protein signaling primarily through interactions with Gbetagamma subunits. Results are presented demonstrating that AGS8 binds to the effector and alpha subunit binding "hot spot" on Gbetagamma yet does not interfere with Galpha subunit binding to Gbetagamma or phospholipase C beta2 activation. AGS8 stimulates activation of phospholipase C beta2 by heterotrimeric Galphabetagamma and forms a quaternary complex with Galpha(i1), Gbeta(1)gamma(2), and phospholipase C beta2. AGS8 rescued phospholipase C beta binding and regulation by an inactive beta subunit with a mutation in the hot spot (beta(1)(W99A)gamma(2)) that normally prevents binding and activation of phospholipase C beta2. This demonstrates that, in the presence of AGS8, the hot spot is not used for Gbetagamma interactions with phospholipase C beta2. Mutation of an alternate binding site for phospholipase C beta2 in the amino-terminal coiled-coil region of Gbetagamma prevented AGS8-dependent phospholipase C binding and activation. These data implicate a mechanism for AGS8, and potentially other Gbetagamma binding proteins, for directing Gbetagamma signaling through alternative effector activation sites on Gbetagamma in the absence of subunit dissociation.  相似文献   

20.
To define receptor subdomains important for protein interaction and identify components of novel signal transduction complexes for the mu- and delta-opioid receptors (mu-OR, delta-OR), we generated glutathione S-transferase fusion proteins of the carboxyl-termini of the mu-opioid receptor (mu-CT), the delta- (delta-CT), and the third intracellular loop of the delta-opioid receptor (delta-i3L) to search for interactive proteins. Results from pull down experiments have demonstrated for the first time that Gbetagamma complexes, derived from the heterotrimeric Galphatbeta1gamma1, purified Gbeta1gamma1, or Gbeta endogenously present in cell lysates and rat striatal extracts, interact with all mu- and delta-opioid receptor subdomains. On the other hand, the C-terminal peptides of the delta- and the mu-ORs exhibit differential profiles for Galpha subunit binding. Indeed, while mu-CT was unable to bind any form of Galpha, both the delta-CT and the delta-i3L displayed interactive regions for heterotrimeric Galphatbeta1gamma1, inactive Galpha(GDP) and active Galpha(GTPgammaS). Regulators of G protein signaling (RGS) proteins are another class of proteins that can modulate G protein signaling events. We demonstrate for the first time that RGS4 directly interacts with the mu-CT, the delta-CT as well as delta-i3L in a dose dependent manner. Moreover, RGS4 modulates mu-OR signaling and can form stable heterotrimeric complexes with the activated Galpha. Collectively, our data demonstrate that the C-termini of the mu- and delta-ORs provide direct physical scaffolding in which G protein subunits and RGS4 protein can be bound.  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