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 390 毫秒
1.
Activation of phospholipase Cbeta (PLCbeta) by G-proteins results in increased intracellular Ca(2+) and activation of protein kinase C. We have previously found that activated PLCbeta-Gbetagamma complex can be rapidly deactivated by Galpha(GDP) subunits without dissociation, which led to the suggestion that Galpha(GDP) binds to PLCbeta-Gbeta gamma and perturbs the activating interaction without significantly affecting the PLCbeta-Gbeta gamma binding energy. Here, we have used high pressure fluorescence spectroscopy to determine the volume change associated with this interaction. Since PLCbeta and G-protein subunits associate on membrane surfaces, we worked under conditions where the membrane surface properties are not expected to change. We also determined the pressure range in which the proteins remain membrane bound: PLCbeta binding was stable throughout the 1-2000 bars range, Gbeta gamma binding was stable only at high membrane concentrations, whereas Galpha(s)(GDP) dissociated from membranes above 1 kbar. High pressure dissociated PLCbeta-Gbeta gamma with a DeltaV = 34 +/- 5 ml/mol. This same volume change is obtained for a peptide derived from Gbeta which also activates PLCbeta. In the presence of Galpha(s)(GDP), the volume change associated with PLCbeta-Gbeta gamma interaction is reduced to 25 +/- 1 ml/mol. These results suggest that activation of PLCbeta by Gbeta gamma is conferred by a small (i.e., 3-15 ml/mol) volume element.  相似文献   

2.
By using purified preparations we show that nanomolar concentrations of Gbetagamma significantly stimulated lipid kinase activity of phosphatidylinositol 3-kinase (PI3K) beta and PI3Kgamma in the presence as well as in the absence of non-catalytic subunits such as p85alpha or p101. Concomitantly, Gbetagamma stimulated autophosphorylation of the catalytic subunit of PI3Kgamma (EC(50), 30 nM; stoichiometry >/=0.6 mol of P(i)/mol of p110gamma), which also occurred in the absence of p101. Surprisingly, we found that p101 affected the lipid substrate preference of PI3Kgamma in its Gbetagamma-stimulated state. With phosphatidylinositol as substrate, p110gamma but not p101/p110gamma was significantly stimulated by Gbetagamma to form PI-3-phosphate (EC(50), 20 nM). The opposite situation was found when PI-4,5-bisphosphate served as substrate. Gbetagamma efficiently and potently (EC(50), 5 nM) activated the p101/p110gamma heterodimer but negligibly stimulated the p110gamma monomer to form PI-3,4,5-trisphosphate. However, this weak stimulatory effect on p110gamma was overcome by excess concentrations of Gbetagamma (EC(50), 100 nM). This finding is in accordance with the in vivo situation, where activated PI3K catalyzes the formation of PI-3,4,5-trisphosphate but not PI-3-phosphate. We conclude that p101 is responsible for PI-4, 5-bisphosphate substrate selectivity of PI3Kgamma by sensitizing p110gamma toward Gbetagamma in the presence of PI-4,5-P(2).  相似文献   

3.
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   

4.
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   

5.
G protein sensitive phosphoinositide 3-kinase gamma (PI3Kgamma) has been characterised as a pleiotropic signalling protein expressing lipid kinase and protein kinase activities. Whereas the regulation of the lipid kinase activity has been investigated in detail, the regulatory features of PI3Kgamma protein kinase activity are unknown. Here we report that Gbetagamma subunits of heterotrimeric G proteins induce a biphasic response of PI3Kgamma autophosphorylation in vitro, which contrasts the regulatory effects of the G proteins on PI3Kgamma lipid kinase activity. In addition to autophosphorylation PI3Kgamma is able to catalyse transphosphorylation of the adapter protein p101 and the protein kinase MEK-1. In the presence of the p101, Gbetagamma affects PI3Kgamma protein kinase activities in a complex manner. In summary, the differential regulatory effects of heterotrimeric G proteins on PI3Kgamma lipid and protein kinase activities in vitro reflect the functional diversity of the enzyme observed in vivo.  相似文献   

6.
Activation of protein kinase C (PKC) can result from stimulation of the receptor-G protein-phospholipase C (PLCbeta) pathway. In turn, phosphorylation of PLCbeta by PKC may play a role in the regulation of receptor-mediated phosphatidylinositide (PI) turnover and intracellular Ca(2+) release. Activation of endogenous PKC by phorbol 12-myristate 13-acetate inhibited both Galpha(q)-coupled (oxytocin and M1 muscarinic) and Galpha(i)-coupled (formyl-Met-Leu-Phe) receptor-stimulated PI turnover by 50-100% in PHM1, HeLa, COSM6, and RBL-2H3 cells expressing PLCbeta(3). Activation of conventional PKCs with thymeleatoxin similarly inhibited oxytocin or formyl-Met-Leu-Phe receptor-stimulated PI turnover. The PKC inhibitory effect was also observed when PLCbeta(3) was stimulated directly by Galpha(q) or Gbetagamma in overexpression assays. PKC phosphorylated PLCbeta(3) at the same predominant site in vivo and in vitro. Peptide sequencing of in vitro phosphorylated recombinant PLCbeta(3) and site-directed mutagenesis identified Ser(1105) as the predominant phosphorylation site. Ser(1105) is also phosphorylated by protein kinase A (PKA; Yue, C., Dodge, K. L., Weber, G., and Sanborn, B. M. (1998) J. Biol. Chem. 273, 18023-18027). Similar to PKA, the inhibition by PKC of Galpha(q)-stimulated PLCbeta(3) activity was completely abolished by mutation of Ser(1105) to Ala. In contrast, mutation of Ser(1105) or Ser(26), another putative phosphorylation target, to Ala had no effect on inhibition of Gbetagamma-stimulated PLCbeta(3) activity by PKC or PKA. These data indicate that PKC and PKA act similarly in that they inhibit Galpha(q)-stimulated PLCbeta(3) as a result of phosphorylation of Ser(1105). Moreover, PKC and PKA both inhibit Gbetagamma-stimulated activity by mechanisms that do not involve Ser(1105).  相似文献   

7.
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   

8.
In vascular smooth muscles, angiotensin II (AII) has been reported to activate phospholipase C (PLC) and phosphatidylinositol 3-kinase (PI3K). We investigated the time-dependent effects of AII on both phosphatidylinositol 3,4,5-trisphosphate (PtdInsP3) and inositol phosphates (InsPs) accumulation in permeabilized microsomes from rat portal vein smooth muscle in comparison with those of noradrenaline (NA). AII stimulated an early production of PtdInsP3 (within 30 s) followed by a delayed production of InsPs (within 3-5 min), in contrast to NA which activated only a fast production of InsPs. The use of pharmacological inhibitors and antibodies raised against the PI3K and PLC isoforms expressed in portal vein smooth muscle showed that AII specifically activated PI3Kgamma and that this isoform was involved in the AII-induced stimulation of InsPs accumulation. NA-induced InsPs accumulation depended on PLCbeta1 activation whereas AII-induced InsPs accumulation depended on PLCgamma1 activation. AII-induced PLCgamma1 activation required both tyrosine kinase and PI3Kgamma since genistein and tyrphostin B48 (inhibitors of tyrosine kinase), LY294002 and wortmannin (inhibitors of PI3K) and anti-PI3Kgamma antibody abolished AII-induced stimulation of InsPs accumulation. Increased tyrosine phosphorylation of PLCgamma1 was only detected for long-lasting applications of AII and was suppressed by genistein. These data indicate that activation of both PI3Kgamma and tyrosine kinase is a prerequisite for AII-induced stimulation of PLCgamma1 in vascular smooth muscle and suggest that the sequential activation of the three enzymes may be responsible for the slow and long-lasting contraction induced by AII.  相似文献   

9.
10.
A preliminary study on the interaction of G protein (guanine triphosphate binding protein) b1g2 subunits and their coupled components in cell signal transduction was conducted in vitro. The insect cell lines, Sf9 (Spodoptera frugiperda) and H5 (Trichoplusia ni) were used to express the recombinant protein Gβ1γ2. The cell membrane containing Gβ1γ2 was isolated through affinity chromatography column with Ni-NTA agarose by FPLC method, and the highly purified protein was obtained. The adenylyl cyclase 2 (AC2) activity assay showed that the purified Gβ1γ2 could significantly stimulate AC2 activity. The interaction of β1γ2 subunits of G protein with the cytoplasmic tail of various mammalian adenylyl cyclases was monitored by BIAcore technology using NTA sensor chip, which relies on the phenomenon of surface plasmon resonance (SPR). The experiments showed the direct binding of Gβ1γ2 to the cytoplasmic tail C2 domain of AC2. The specific binding domain of AC2 with Gβ1γ2 was the same as AC2 activity domain which was stimulated by Gβ1γ2.  相似文献   

11.
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   

12.
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   

13.
We previously identified a 10-amino acid region from the Y domain of phospholipase Cbeta2 (PLCbeta2) that associates with G-protein betagamma subunits (Sankaran, B., Osterhout, J., Wu, D., and Smrcka, A. V. (1998) J. Biol. Chem. 273, 7148-7154). We mapped the site for cross-linking of a synthetic peptide (N20K) corresponding to this Y domain region to Cys(25) within the amino-terminal coiled-coil domain of Gbetagamma (Yoshikawa, D. M., Bresciano, K., Hatwar, M., and Smrcka, A. V. (2001) J. Biol. Chem. 276, 11246-11251). Here, further experiments with a series of variable length cross-linking agents refined the site of N20K binding to within 4.4-6.7 angstroms of Cys(25). A mutant within the amino terminus of the Gbeta subunit, Gbeta(1)(23-27)gamma(2), activated PLCbeta2 more effectively than wild type, with no significant change in the EC(50), indicating that this region is directly involved in the catalytic regulation of PLCbeta2. This mutant was deficient in cross-linking to N20K, suggesting that a binding site for the peptide had been eliminated. Surprisingly, N20K could still inhibit Gbeta(1)(23-27)gamma(2)-dependent activation of PLC, suggesting a second N20K binding site. Competition analysis with a peptide that binds to the Galpha subunit switch II binding surface of Gbetagamma indicates a second N20K binding site at this surface. Furthermore, mutations to the N20K region within the Y-domain of full-length PLCbeta2 inhibited Gbetagamma-dependent regulation of the enzyme, providing further evidence for aGbetagamma binding site within the catalytic domain of PLCbeta2. The data support a model with two modes of PLC binding to Gbetagamma through the catalytic domain, where interactions with the amino-terminal coiled-coil domain are inhibitory, and interactions with the Galpha subunit switch II binding surface are stimulatory.  相似文献   

14.
15.
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   

16.
To investigate the coupling selectivity of G proteins and G protein-coupled receptors (GPCRs), we developed a reconstitution system made up of GPCR and heterotrimeric G proteins on extracellular baculovirus particles (budded virus (BV)). BV released from Sf9 cells infected with a recombinant baculovirus coding for human leukotriene B4 receptor (BLT1) cDNA exhibited a high level of BLT1 expression (27.3 pmol/mg of protein) and specific [3H]leukotriene B4 binding activity (Kd = 3.67 nm). The apparent low affinity of the expressed BLT1 is thought to be due to relative non-availability of the Galphai isoform, which couples to BLT1, in BV. Co-infection of heterotrimeric G protein recombinant viruses led to co-expression of BLT1 and G protein subunits on BV. A guanosine-5'-(beta,gamma-imido)triphosphate-sensitive, high affinity ligand binding was observed in the BLT1 BV co-expressing Galphai1beta1gamma2 (Kd = 0.17 nm). A relatively large amount of high affinity receptor protein was recovered in the co-expressing BV fraction (6.81 pmol/mg of protein). A combination of BLT1 and Galphai1 without Gbeta1gamma2 did not exhibit high affinity ligand binding on BV, indicating the low background environment for the GPCR-G protein coupling in this BV reconstitution system. To test other G proteins for coupling, various Galpha subunits were combinatorially expressed in BV with BLT1 and Gbeta1gamma2. The BLT1 BV co-expressing GalphaoAbeta1gamma2 exhibited a comparably high affinity ligand binding as well as ligand-stimulated guanosine 5'-3-O-(thio)triphosphate binding to Galphai1beta1gamma2. Co-expression of other Galpha isoforms such as Galphas, Galpha11, Galpha14, Galpha16, Galpha12, or Galpha13 did not exhibit any significant effects on ligand binding affinity in this system. These results reveal that BLT1 and coupled trimeric G proteins were functionally reconstituted on BV and that Galphao as well as Galphai couples to BLT1. This expression system should prove highly useful for pharmacological characterization, biosensor chip applications, and also drug discovery directed at highly important targets of the membrane receptor proteins.  相似文献   

17.
Phosphoinositide 3-kinases (PI3Ks) are important regulators of cell migration. The PI3K isoform gamma is primarily expressed in haematopoietic cells, and is activated by G protein-coupled receptors (GPCRs). Here, we investigate the contribution of PI3Kgamma to macrophage responses to chemoattractants, using bone marrow-derived macrophages from wild-type and PI3Kgamma-null mice. We observe that early membrane ruffling induced by MCP-1, which activates a GPCR, or by CSF-1, which activates a tyrosine kinase receptor, is unaltered in PI3Kgamma(-/-) mice, although by 30 min MCP-1-induced cell polarization was strongly reduced in PI3Kgamma(-/-) compared to wild-type macrophages. The migration behaviour of the macrophages was analysed by time-lapse microscopy in Dunn chemotaxis chambers. PI3Kgamma(-/-) macrophages showed reduced migration speed and translocation, and no chemotaxis to MCP-1. Interestingly, there was also a reduction in migration efficiency in PI3Kgamma(-/-) macrophages stimulated with CSF-1 although early CSF-1R signalling was normal. These results indicate that the initial actin reorganization induced by either a GPCR or tyrosine kinase receptor agonist is not dependent on PI3Kgamma, whereas PI3Kgamma is needed for optimal migration of macrophages to either agonist.  相似文献   

18.
Cardiac myocyte contractility is initiated by Ca2+ entry through the voltage-dependent L-type Ca2+ channel (LTCC). To study the effect of Galpha q on the cardiac LTCC, we utilized two transgenic mouse lines that selectively express inducible Galpha q-estrogen receptor hormone-binding domain fusion proteins (Galpha qQ209L-hbER or Galpha qQ209L-AA-hbER) in cardiac myocytes. Both of these proteins inhibit phosphatidylinositol (PI) 3-kinase (PI3K) signaling, but Galpha qQ209L-AA-hbER cannot activate the canonical Galpha q effector phospholipase Cbeta (PLCbeta). L-type Ca2+ current (I(Ca,L)) density measured by whole-cell patch clamping was reduced by more than 50% in myocytes from both Galpha q animals as compared with wild-type cells, suggesting that inhibition of the LTCC by Galpha q does not require PLCbeta. To investigate the role of PI3K in this inhibitory effect, I(Ca,L) was measured in the presence of various phosphoinositides infused through the patch pipette. Infusion of PI 3,4,5-trisphosphate (PI(3,4,5)P3) into wild-type myocytes did not affect I(Ca,L), but it fully restored I(Ca,L) density in both Galpha q transgenic myocytes to wild-type levels. By contrast, PI 4,5-bisphosphate (PI(4,5)P2) or PI 3,5-bisphosphate had no effect. Infusion with p110beta/p85alpha or p110gamma PI3K in the presence of PI(4,5)P2 also restored I(Ca,L) density to wild-type levels. Last, infusion of either PTEN, a PI(3,4,5)P3 phosphatase, or the pleckstrin homology domain of Grp1, which sequesters PI(3,4,5)P3, reduced the peak I(Ca,L) density in wild-type myocytes by approximately 30%. Taken together, these results strongly suggest that the inhibitory effect of Galpha q on the cardiac LTCC is mediated by inhibition of PI3K.  相似文献   

19.
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   

20.
The T-allele of a polymorphism (C825T) in the gene of the G-protein beta3-subunit is associated with a complex phenotype (hypertension, obesity, altered drug responses) and the occurrence of a splice variant termed Gbeta3s which lacks one of the seven WD-domains that compose Gbeta-proteins. Here, we analysed Gbetagamma dimer formation and Galpha activation by Gbeta3s, key functional characteristics of Gbeta-proteins. Cleavage protection assays frequently used to analyse Gbeta1gamma and Gbeta2gamma dimer formation failed for Gbeta3 and Gbeta3s, while in coprecipitation assays, dimerization of Gbeta3 and Gbeta3s with Ggamma5, Ggamma8(c) and Ggamma12 could be demonstrated. Upon expression of Gbeta3s in COS-7 and Sf9 insect cells, binding of GTPgammaS to Galpha-proteins induced by mastoparan-7 and the M(2) muscarinic acetylcholine receptor was facilitated in comparison with cells overexpressing wildtype Gbeta3, as indicated by twofold reduced agonist EC(50) values. Together, these results indicate that Gbeta3s is a biologically active Gbeta-protein that may mediate the enhanced signal transduction observed in cells with the 825T-allele.  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