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.
Rac activation is a key step in chemotaxis of hematopoietic cells, which is both positively and negatively regulated by receptors coupled to heterotrimeric G proteins. P-Rex1, a Rac-specific guanine nucleotide exchange factor, is dually activated by phosphatidylinositol (3,4,5)-trisphosphate (PIP(3)) and the Gbetagamma subunits of heterotrimeric G proteins. This study explored the regulation of P-Rex1 by phosphorylation with the cAMP-dependent protein kinase (protein kinase A) in vitro and by G(i)- and G(s)-coupled receptors in HEK293T cells. P-Rex1 isolated from Sf9 and HEK293T cells migrates as two distinct bands that are partially phosphorylated. Phosphorylation of P-Rex1 with protein kinase A (PKA) inhibits the PIP(3)- and Gbetagamma-stimulated P-Rex1 guanine nucleotide exchange activity on Rac. The guanine nucleotide exchange factor activity of three different forms of P-Rex1 (native Sf9, de-phosphorylated, and phosphorylated) was examined in the presence of PIP(3) and varying concentrations of Gbeta(1)gamma(2). Gbeta(1)gamma(2) was 47-fold less potent in activating the phosphorylated form of P-Rex1 compared with the de-phosphorylated form. HEK293T cells expressing P-Rex1 were labeled with (32)P and stimulated with lysophosphatidic acid (LPA) to release Gbetagamma or isoproterenol to activate PKA. Treatment with isoproterenol or S(p)-cAMPS, a potent activator of PKA, increased the incorporation of (32)P into P-Rex1. LPA increased the amount of GTP-bound Rac in the cells and isoproterenol reduced basal levels of GTP-bound Rac and blunted the effect of LPA. Treatment of the cells with S(p)-cAMPS also reduced the levels of GTP-bound Rac. These results outline a novel mechanism for G(s)-linked receptors to regulate the function of P-Rex1 and inhibit its function in cells.  相似文献   

2.
G protein betagamma dimers can be phosphorylated in membranes from various tissues by GTP at a histidine residue in the beta subunit. The phosphate is high energetic and can be transferred onto GDP leading to formation of GTP. Purified Gbetagamma dimers do not display autophosphorylation, indicating the involvement of a separate protein kinase. We therefore enriched the Gbeta-phosphorylating activity present in preparations of the retinal G protein transducin and in partially purified G(i/o) proteins from bovine brain. Immunoblots, autophosphorylation, and enzymatic activity measurements demonstrated enriched nucleoside diphosphate kinase (NDPK) B in both preparations, together with residual Gbetagamma dimers. In the retinal NDPK B-enriched fractions, a Gbeta-specific antiserum co-precipitated phosphorylated NDPK B, and an antiserum against the human NDPK co-precipitated phosphorylated Gbetagamma. In addition, the NDPK-containing fractions from bovine brain reconstituted the phosphorylation of purified Gbetagamma. For identification of the phosphorylated histidine residue, bovine brain Gbetagamma and G(t)betagamma were thiophosphorylated with guanosine 5'-O-(3-[(35)S]thio)triphosphate, followed by digestion with endoproteinase Glu-C and trypsin, separation of the resulting peptides by gel electrophoresis and high pressure liquid chromatography, respectively, and sequencing of the radioactive peptides. The sequence information produced by both methods identified specific labeled fragments of bovine Gbeta(1) that overlapped in the heptapeptide, Leu-Met-Thr-Tyr-Ser-His-Asp (amino acids 261-267). We conclude that NDPK B forms complexes with Gbetagamma dimers and contributes to G protein activation by increasing the high energetic phosphate transfer onto GDP via intermediately phosphorylated His-266 in Gbeta(1) subunits.  相似文献   

3.
Formation of GTP by nucleoside diphosphate kinase (NDPK) can contribute to G protein activation in vitro. To study the effect of NDPK on G protein activity in living cells, the NDPK isoforms A and B were stably expressed in H10 cells, a cell line derived from neonatal rat cardiomyocytes. Overexpression of either NDPK isoform had no effect on cellular GTP and ATP levels, basal cAMP levels, basal adenylyl cyclase activity, and the expression of G(s)alpha and G(i)alpha proteins. However, co-expression of G(s)alpha led to an increase in cAMP synthesis that was largely enhanced by the expression of NDPK B, but not NDPK A, and that was confirmed by direct measurement of adenylyl cyclase activity. Cells expressing an inactive NDPK B mutant (H118N) exhibited a decreased cAMP formation in response to G(s)alpha. Co-immunoprecipitation studies demonstrated a complex formation of the NDPK with Gbetagamma dimers. The overexpression of NDPK B, but not its inactive mutant or NDPK A, increased the phosphorylation of Gbeta subunits. In summary, our data demonstrate a specific NDPK B-mediated activation of a G protein in intact cells, which is apparently caused by formation of NDPK B.Gbetagamma complexes and which appears to contribute to the receptor-independent activation of heterotrimeric G proteins.  相似文献   

4.
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   

5.
Using peptides based on the amino acid sequences surrounding the two histidine residues in histone H4, we have investigated the kinetics of the phosphorylation and dephosphorylation reactions of their histidine residues, when reacted with potassium phosphoramidate, by 1H NMR. We have been able to estimate rate constants for the reactions and have shown that there are differences in the kinetics between the two peptides. The kinetics of hydrolysis of phosphoramidate was measured by 31P NMR and protein histidine phosphatase (PHP) was shown to catalyse the reaction. We have shown that the dephosphorylation of the phosphohistidine of the phosphopeptides is catalysed by PHP. In terms of substrate specificity, there is a small preference for 1-phosphohistidine compared to 3-phosphohistidine, although the rate accelerations for hydrolysis induced by the enzyme were 1100- and 33,333-fold, respectively. The kinetics of both the phosphorylation and dephosphorylation reactions depend on the amino acid sequence surrounding the histidine. PHP shows greater substrate specificity for the peptide whose sequence is similar to that around histidine 18 of histone H4. PHP was unable to catalyse the dephosphorylation of histone H4 that had been phosphorylated with a histone H4 histidine kinase.  相似文献   

6.
RGS5 is a member of regulators of G protein signaling (RGS) proteins that attenuate heterotrimeric G protein signaling by functioning as GTPase-activating proteins (GAPs). We investigated phosphorylation of RGS5 and the resulting change of its function. In 293T cells, transiently expressed RGS5 was phosphorylated by endogenous protein kinases in the basal state. The phosphorylation was enhanced by phorbol 12-myristate 13-acetate (PMA) and endothelin-1 (ET-1), and suppressed by protein kinase C (PKC) inhibitors, H7, calphostin C and staurosporine. These results suggest involvement of PKC in phosphorylation of RGS5. In in vitro experiments, PKC phosphorylated recombinant RGS5 protein at serine residues. RGS5 protein phosphorylated by PKC showed much lower binding capacity for and GAP activity toward Galpha subunits than did the unphosphorylated RGS5. In cells expressing RGS5, the inhibitory effect of RGS5 on ET-1-induced Ca(2+) responses was enhanced by staurosporine. Mass spectrometric analysis of the phosphorylated RGS5 revealed that Ser166 was one of the predominant phosphorylation sites. Substitution of Ser166 by aspartic acid abolished the binding capacity to Galpha subunits and the GAP activity, and markedly reduced the inhibitory effect on ET-1-induced Ca(2+) responses. These results indicate that phosphorylation at Ser166 of RGS5 by PKC causes loss of the function of RGS5 in G protein signaling. Since this serine residue is conserved in RGS domains of many RGS proteins, the phosphorylation at Ser166 by PKC might act as a molecular switch and have functional significance.  相似文献   

7.
Heterotrimeric G proteins play central roles in signal transduction of neurons and other cells. The variety of their alpha-, beta-, and gamma-subunits allows numerous combinations thereby confering specificity to receptor-G-protein-effector interactions. Using antisera against individual G-protein beta-subunits we here present a regional and subcellular distribution of Gbeta1, Gbeta2, and Gbeta5 in rat brain. Immunocytochemical specificity of the subtype-specific antisera is revealed in Sf9 cells infected with various G-protein beta-subunits. Since Gbeta-subunits together with a G-protein gamma-subunit affect signal cascades we include a distribution of the neuron-specific Ggamma2- and Ggamma3-subunits in selected brain areas. Gbeta1, Gbeta2, and Gbeta5 are preferentially distributed in the neuropil of hippocampus, cerebellum and spinal cord. Gbeta2 is highly concentrated in the mossy fibres of dentate gyrus neurons ending in the stratum lucidum of hippocampal CA3-area. High amounts of Gbeta2 also occur in interneurons innervating spinal cord alpha-motoneurons. Gbeta5 is differentially distributed in all brain areas studied. It is found in the pyramidal cells of hippocampal CA1-CA3 as well as in the granule cell layer of dentate gyrus and in some interneurons. In the spinal cord Gbeta5 in contrast to Gbeta2 concentrates around alpha-motoneurons. In cultivated mouse hippocampal and hypothalamic neurons Gbeta2 and Gbeta5 are found in different subcellular compartments. Whereas Gbeta5 is restricted to the perikarya, Gbeta2 is also found in processes and synaptic contacts where it partially colocalizes with the synaptic vesicle protein synaptobrevin. An antiserum recognizing Ggamma2 and Ggamma3 reveals that these subunits are less expressed in hippocampus and cerebellum. Presumably this antiserum specifically recognizes Ggamma2 and Ggamma3 in combinations with certain G alphas and/or Gbetas. The widespread but regionally and cellularly rather different distribution of Gbeta- and Ggamma2/3-subunits suggests that region-specific combinations of G-protein subunits mediate signal transduction in the central nervous system. The different subcellular distribution of Gbeta-subunits in cultivated neurons reflects that observed in tissue where Gbeta5 and Gbeta2 associate preferentially with the perikarya and the neuropil, respectively, and suggests an additional association of Gbeta2 with secretory vesicles.  相似文献   

8.
Lee YI  Kim SY  Cho CH  Seo M  Cho DH  Kwak SJ  Juhnn YS 《FEBS letters》2003,555(2):329-334
Individual cell types express a characteristic balance between heterotrimeric G protein alpha and betagamma subunits, but little is known about the regulatory mechanism. We systemically examined the regulatory mechanism in CHO cells. We found that expression of Galphas, Galphai2, and Galphaq proteins increased in direct proportion to the increase of Gbeta1gamma2 overexpressed transiently. Expression of Gbeta protein also increased following overexpression of Galphas, Galphai2, and Galphaq. The Gbetagamma overexpression stimulated degradation of Gbeta in contrast to reduction of Galphas degradation. We conclude that coordinate expression of the G protein subunits involves regulation of protein degradation via proteasome in CHO cells.  相似文献   

9.
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   

10.
All dividing cells entering the M phase of the cell cycle undergo the transient activation of an M-phase-specific histone H1 kinase which was recently shown to be constituted of at least two subunits, p34cdc2 and cyclincdc13. The DNA-binding high-mobility-group (HMG) proteins 1, 2, 14, 17, I, Y and an HMG-like protein, P1, were investigated as potential substrates of H1 kinase. Among these HMG proteins, P1 and HMG I and Y are excellent substrates of the M-phase-specific kinase obtained from both meiotic starfish oocytes and mitotic sea urchin eggs. Anticyclin immunoprecipitates, extracts purified on specific p34cdc2-binding p13suc1-Sepharose and affinity-purified H1 kinase display strong HMG I, Y and P1 phosphorylating activities, demonstrating that the p34cdc2/cyclincdc13 complex is the active kinase phosphorylating these HMG proteins. HMG I and P1 phosphorylation is competitively inhibited by a peptide mimicking the consensus phosphorylation sequence of H1 kinase. HMG I, Y and P1 all possess the consensus sequence for phosphorylation by the p34cdc2/cyclincdc13 kinase (Ser/Thr-Pro-Xaa-Lys/Arg). HMG I is phosphorylated in vivo at M phase on the same sites phosphorylated in vitro by H1 kinase. P1 is phosphorylated by H1 kinase on sites different from the sites of phosphorylation by casein kinase II. The three thermolytic phosphopeptides of P1 phosphorylated in vitro by purified H1 kinase are all present in thermolytic peptide maps of P1 phosphorylated in vivo in proliferating HeLa cells. These phosphopeptides are absent in nonproliferating cells. These results demonstrate that the DNA-binding proteins HMG I, Y and P1 are natural substrates for the M-phase-specific protein kinase. The phosphorylation of these proteins by p34cdc2/cyclincdc13 may represent a crucial event in the intense chromatin condensation occurring as cells transit from the G2 to the M phase of the cell cycle.  相似文献   

11.
All eukaryotic cells sense extracellular stimuli and activate intracellular signaling cascades via G protein-coupled receptors (GPCR) and associated heterotrimeric G proteins. The Saccharomyces cerevisiae GPCR Gpr1 and associated Galpha subunit Gpa2 sense extracellular carbon sources (including glucose) to govern filamentous growth. In contrast to conventional Galpha subunits, Gpa2 forms an atypical G protein complex with the kelch repeat Gbeta mimic proteins Gpb1 and Gpb2. Gpb1/2 negatively regulate cAMP signaling by inhibiting Gpa2 and an as yet unidentified target. Here we show that Gpa2 requires lipid modifications of its N-terminus for membrane localization but association with the Gpr1 receptor or Gpb1/2 subunits is dispensable for membrane targeting. Instead, Gpa2 promotes membrane localization of its associated Gbeta mimic subunit Gpb2. We also show that the Gpa2 N-terminus binds both to Gpb2 and to the C-terminal tail of the Gpr1 receptor and that Gpb1/2 binding interferes with Gpr1 receptor coupling to Gpa2. Our studies invoke novel mechanisms involving GPCR-G protein modules that may be conserved in multicellular eukaryotes.  相似文献   

12.
The role that Gbeta(5) regulator of G protein signaling (RGS) complexes play in signal transduction in brain remains unknown. The subcellular localization of Gbeta(5) and RGS7 was examined in rat PC12 pheochromocytoma cells and mouse brain. Both nuclear and cytosolic localization of Gbeta(5) and RGS7 was evident in PC12 cells by immunocytochemical staining. Subcellular fractionation of PC12 cells demonstrated Gbeta(5) immunoreactivity in the membrane, cytosolic, and nuclear fractions. Analysis by limited proteolysis confirmed the identity of Gbeta(5) in the nuclear fraction. Subcellular fractionation of mouse brain demonstrated Gbeta(5) and RGS7 but not Ggamma(2/3) immunoreactivity in the nuclear fraction. RGS7 and Gbeta(5) were tightly complexed in the brain nuclear extract as evidenced by their coimmunoprecipitation with anti-RGS7 antibodies. Chimeric protein constructs containing green fluorescent protein fused to wild-type Gbeta(5) but not green fluorescent fusion proteins with Gbeta(1) or a mutant Gbeta(5) impaired in its ability to bind to RGS7 demonstrated nuclear localization in transfected PC12 cells. These findings suggest that Gbeta(5) undergoes nuclear translocation in neurons via an RGS-dependent mechanism. The novel intracellular distribution of Gbeta(5).RGS protein complexes suggests a potential role in neurons communicating between classical heterotrimeric G protein subunits and/or their effectors at the plasma membrane and the cell nucleus.  相似文献   

13.
BACKGROUND: Studies show that high phosphotidylinositol 3,4,5-trisphosphate (PIP(3)) promotes cytoskeletal rearrangements and alters cell motility and chemotaxis, possibly through activation of protein kinase Bs (PKBs). However, chemotaxis can still occur in the absence of PIP(3), and the identities of the PIP(3)-independent pathways remain unknown. RESULTS: Here, we outline a PIP(3)-independent pathway linking temporal and spatial activation of PKBs by Tor complex 2 (TorC2) to the chemotactic response. Within seconds of stimulating Dictyostelium cells with chemoattractant, two PKB homologs, PKBA and PKBR1, mediate transient phosphorylation of at least eight proteins, including Talin, PI4P 5-kinase, two Ras GEFs, and a RhoGap. Surprisingly, all of the substrates are phosphorylated with normal kinetics in cells lacking PI 3-kinase activity. Cells deficient in TorC2 or PKB activity show reduced phosphorylation of the endogenous substrates and are impaired in chemotaxis. The PKBs are activated through phosphorylation of their hydrophobic motifs via TorC2 and subsequent phosphorylation of their activation loops. These chemoattractant-inducible events are restricted to the cell's leading edge even in the absence of PIP(3). Activation of TorC2 depends on heterotrimeric G protein function and intermediate G proteins, including Ras GTPases. CONCLUSIONS: The data lead to a model where cytosolic TorC2, encountering locally activated small G protein(s) at the leading edge of the cell, becomes activated and phosphorylates PKBs. These in turn phosphorylate a series of signaling and cytoskeletal proteins, thereby regulating directed migration.  相似文献   

14.
Serine/threonine protein kinases of the Ste20p/PAK family are highly conserved from yeast to man. These protein kinases have been implicated in the signaling from heterotrimeric G proteins to mitogen-activated protein (MAP) kinase cascades and to cytoskeletal components such as myosin-I. In the yeast Saccharomyces cerevisiae, Ste20p is involved in transmitting the mating-pheromone signal from the betagamma-subunits of a heterotrimeric G protein to a downstream MAP kinase cascade. We have previously shown that binding of the G-protein beta-subunit (Gbeta) to a short binding site in the non-catalytic carboxy-terminal region of Ste20p is essential fortransmitting the pheromone signal. In this study, we searched protein sequence databases for sequences that are similar to the Gbeta binding site in Ste20p. We identified a sequence motif with the consensus sequence S S L phi P L I/V x phi phi beta (x: any residue; phi: A, I, L, S, or T; beta: basic residues) that is solely present in members of Ste20p/PAK family protein kinases. We propose that this sequence motif, which we have designated GBB (Gbeta binding) motif, is specifically responsible for binding of Gbeta to Ste20p/PAK protein kinases in response to activation of heterotrimeric G protein coupled receptors. Thus, the GBB motif is a novel type of signaling domain that serves to link protein kinases of the Ste20p/PAK family to G protein coupled receptors.  相似文献   

15.
Chen H  Michel T 《Biochemistry》2006,45(26):8023-8033
Activation of insulin receptors stimulates the phosphoinositide 3-kinase (PI3-K)/Akt signaling pathway in vascular endothelial cells. Heterotrimeric G proteins appear to modulate some of the cellular responses that are initiated by receptor tyrosine kinases, but the roles of specific G protein subunits in signaling are less clearly defined. We found that insulin treatment of cultured bovine aortic endothelial cells (BAEC) activates the alpha isoform of PI3-K (PI3-Kalpha) and discovered that purified G protein Gbeta1gamma2 inhibits PI3-Kalpha enzyme activity. Transfection of BAEC with a duplex siRNA targeting bovine Gbeta1 leads to a 90% knockdown in Gbeta1 protein levels, with no effect on expression of other G protein subunits. siRNA-mediated Gbeta1 knockdown markedly and specifically potentiates insulin-dependent activation of kinase Akt, likely reflecting the removal of the inhibitory effect of Gbetagamma on PI3-Kalpha activity. Insulin-induced tyrosine phosphorylation of insulin receptors is unaffected by Gbeta1 siRNA. By contrast, Gbeta1 knockdown leads to a significant decrease in the level of serine phosphorylation of the insulin receptor substrate IRS-1. We explored the effects of siRNA on several serine/threonine protein kinases that have been implicated in insulin signaling. Gbeta1 siRNA significantly attenuates phosphorylation of the 70 kDa ribosomal protein S6 kinase (p70S6K) in the basal state and following insulin treatment. We also found that IGF-1-initiated activation of Akt is significantly enhanced after siRNA-mediated Gbeta1 knockdown, while IGF-1-induced p70S6K activation is markedly suppressed following transfection of Gbeta1 siRNA. We propose that Gbeta1 participates in the activation of p70S6K, which in turn promotes the serine phosphorylation and inhibition of IRS-1. Taken together, these studies suggest that Gbeta1 plays an important role in insulin and IGF-1 signaling in endothelial cells, both by inhibiting the activity of PI3-Kalpha and by stimulating pathways that lead to activation of protein kinase p70S6K and to the serine phosphorylation of IRS-1.  相似文献   

16.
In vertebrate photoreceptors, photoexcited rhodopsin interacts with the G protein transducin, causing it to bind GTP and stimulate the enzyme cGMP phosphodiesterase. The rapid termination of the active state of this pathway is dependent upon a photoreceptor-specific regulator of G protein signaling RGS9-1 that serves as a GTPase activating protein (GAP) for transducin. Here, we show that, in preparations of photoreceptor outer segments (OS), RGS9-1 is readily phosphorylated by an endogenous Ser/Thr protein kinase. Protein kinase C and MAP kinase inhibitors reduced labeling by about 30%, while CDK5 and CaMK II inhibitors had no effect. cAMP-dependent protein kinase (PKA) inhibitor H89 reduced RGS9-1 labeling by more than 90%, while dibutyryl-cAMP stimulated it 3-fold, implicating PKA as the major kinase responsible for RGS9-1 phosphorylation in OS. RGS9-1 belongs to an RGS subfamily also including RGS6, RGS7, and RGS11, which exist as heterodimers with the G protein beta subunit Gbeta5. Phosphorylated RGS9-1 remains associated with Gbeta5L, a photoreceptor-specific splice form, which itself was not phosphorylated. RGS9-1 immunoprecipitated from OS was in vitro phosphorylated by exogenous PKA. The PKA catalytic subunit could also phosphorylate recombinant RGS9-1, and mutational analysis localized phosphorylation sites to Ser(427) and Ser(428). Substitution of these residues for Glu, to mimic phosphorylation, resulted in a reduction of the GAP activity of RGS9-1. In OS, RGS9-1 phosphorylation required the presence of free Ca(2+) ions and was inhibited by light, suggesting that RGS9-1 phosphorylation could be one of the mechanisms mediating a stronger photoresponse in dark-adapted cells.  相似文献   

17.
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   

18.
Reversible phosphorylation is the most widespread posttranslational protein modification, playing regulatory role in almost every aspect of cell life. The majority of protein phosphorylation research has been focused on serine, threonine and tyrosine that form acid-stable phosphomonoesters. However, protein histidine, arginine and lysine residues also may undergo phosphorylation to yield acid-labile phosphoramidates, most often remaining undetected in conventional studies of protein phosphorylation. It has become increasingly evident that acid-labile protein phosphorylations play important roles in signal transduction and other regulatory processes. Beside acting as high-energy intermediates in the transfer of the phosphoryl group from donor to acceptor molecules, phosphohistidines have been found so far in histone H4, heterotrimeric G proteins, ion channel KCa3.1, annexin 1, P-selectin and myelin basic protein, as well as in recombinant thymidylate synthase expressed in bacterial cells. Phosphoarginines occur in histone H3, myelin basic protein and capsidic protein VP12 of granulosis virus, whereas phospholysine in histone H1. This overview of the current knowledge on phosphorylation of protein basic amino-acid residues takes into consideration its proved or possible roles in cell functioning. Specific requirements of studies on acid-labile protein phosphorylation are also indicated.  相似文献   

19.
After stimulation with agonist, G protein-coupled receptors (GPCRs) activate G proteins and become phosphorylated by G protein-coupled receptor kinases (GRKs), and most of them translocate cytosolic arrestin proteins to the cytoplasmic membrane. Agonist-activated GPCRs are specifically phosphorylated by GRKs and are targeted for endocytosis by arrestin proteins, suggesting a connection between GPCR conformational changes and interaction with GRKs and arrestins. Previously, we showed that by substitution of histidine for residues at the cytoplasmic side of helix 3 (H3) and helix 6 (H6) of the parathyroid hormone (PTH) receptor (PTHR), a zinc metal ion-binding site is engineered that prevents PTH-stimulated G(s) activation (Sheikh, S. P., Vilardaga, J.-P., Baranski, T. J., Lichtarge, O., Iiri, T., Meng, E. C., Nissenson, R. A., and Bourne, H. R. (1999) J. Biol. Chem. 274, 17033-17041). These data suggest that relative movements between H3 and H6 are critical for G(s) activation. Does this molecular event play a similar role in activation of GRK and arrestin and in PTHR-mediated G(q) activation? To answer this question, we utilized the two previously described mutant forms of PTHR, H401 and H402, which contain a naturally present histidine residue at position 301 in H3 and a second substituted histidine residue at positions 401 and 402 in H6, respectively. Both mutant receptors showed inhibition of PTH-stimulated inositol phosphate and cAMP generation in the presence of increasing concentrations of Zn(II). However, the mutants showed no Zn(II)-dependent impairment of phosphorylation by GRK-2. Likewise, the mutants were indistinguishable from wild-type PTHR in the ability to translocate beta-arrestins/green fluorescent protein to the cell membrane and were also not affected by sensitivity to Zn(II). These results suggest that agonist-mediated phosphorylation and internalization of PTHR require conformational switches of the receptor distinct from the cAMP and inositol phosphate signaling state. Furthermore, PTHR sequestration does not appear to require G protein activation.  相似文献   

20.
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