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 15 毫秒
1.
Heterotrimeric G proteins mediate cell growth and differentiation by coupling cell surface receptors to intracellular effector enzymes. The G-protein alpha subunit, Galpha(16), and its murine homologue Galpha(15), are expressed specifically in hematopoietic cells and their expression is highly regulated during differentiation of normal and leukemic cells. In this study, we examined the phosphorylation of Galpha(15)/Galpha(16) and its role in receptor and effector coupling. We observed a PMA-stimulated intact cell phosphorylation of Galpha(15) in COS7 cells transfected with Galpha(15) and protein kinase Calpha (PKCalpha), and phosphorylation of endogenous Galpha(16) in HL60 cells. We also showed that peptides derived from the two G-proteins were phosphorylated in vitro using purified brain PKC. Furthermore, we identified the putative phosphorylation site and showed that mutation or deletion of this PKC phosphorylation site inhibited phospholipase C (PLC) activation. The behavior of double mutants with the constitutively active G-protein mutation (QL-mutant) and mutation in the putative phosphorylation site suggests that the phosphorylation site of Galpha(15/16) is essential for receptor-coupled activation of PLC, but not for direct interaction of the G-protein with PLC-beta.  相似文献   

2.
It has previously been shown that nicotine-evoked dopamine release from rat striatal synaptosomes and nicotine-evoked norepinephrine release from hippocampal synaptosomes are mediated by distinct nicotinic acetylcholine receptor (nAChR) subtypes. In the present study, the functional association of these nicotinic receptors with specific subtypes of voltage-gated calcium channels was examined. Cd(2+) (200 microM), as well as omega-conotoxin MVIIC (5 microM), blocks approximately 85% of nicotine-evoked dopamine release from striatal synaptosomes, indicating a major involvement of calcium channels. Furthermore, the toxin-susceptibility suggests that these calcium channels contain alpha(1A) and/or alpha(1B) subunits. Inhibition of nicotine-evoked dopamine release by conotoxins alpha-MII and omega-GVIA is additive and indicates that presynaptic alpha3beta2 nAChRs are functionally coupled to alpha(1A), but not alpha(1B), calcium channel subtypes. Conversely, insensitivity to alpha-AuIB and sensitivity to omega-MVIIC indicate that non-alpha3beta2/alpha3beta4-containing nAChRs are functionally coupled to alpha(1B)-containing calcium channels. In contrast, Cd(2+) blocks only 65% of nicotine-evoked norepinephrine release from hippocampal synaptosomes, indicating that a substantial fraction of this release occurs through mechanisms not involving calcium channels. This Cd(2+)-insensitive component of release is blocked by alpha-AuIB and therefore appears to be triggered by Ca(2+) flowing directly through the channels of presynaptic alpha3beta4 nAChRs. Thus, these data indicate that different presynaptic termini can have distinctive functional associations of specific nAChRs and voltage-gated calcium channels.  相似文献   

3.
Glycine receptors (GlyRs) can dynamically exchange between synaptic and extrasynaptic locations through lateral diffusion within the plasma membrane. Their accumulation at inhibitory synapses depends on the interaction of the β-subunit of the GlyR with the synaptic scaffold protein gephyrin. An alteration of receptor-gephyrin binding could thus shift the equilibrium between synaptic and extrasynaptic GlyRs and modulate the strength of inhibitory neurotransmission. Using a combination of dynamic imaging and biochemical approaches, we have characterised the molecular mechanism that links the GlyR-gephyrin interaction with GlyR diffusion and synaptic localisation. We have identified a protein kinase C (PKC) phosphorylation site within the cytoplasmic domain of the β-subunit of the GlyR (residue S403) that causes a reduction of the binding affinity between the receptor and gephyrin. In consequence, the receptor's diffusion in the plasma membrane is accelerated and GlyRs accumulate less strongly at synapses. We propose that the regulation of GlyR dynamics by PKC thus contributes to the plasticity of inhibitory synapses and may be involved in maladaptive forms of synaptic plasticity.  相似文献   

4.
J. Neurochem. (2012) 122, 1108-1117. ABSTRACT: Rapamycin-insensitive companion of TOR (Rictor) is a conserved component of target of rapamycin complex 2 (TORC2), a complex implicated in phosphorylation of a number of signal transduction-related kinases, including protein kinase Cs (PKCs) at their 'hydrophobic' site in the carboxy-terminal extension domain. In the marine mollusk, Aplysia californica, an increase in phosphorylation of the novel PKC, Apl II, at the hydrophobic site is associated with a protein synthesis-dependent increase in synaptic strength seen after continuous application of serotonin. To determine if Rictor plays a role in this increase, we cloned the Aplysia ortholog of Rictor (ApRictor). An siRNA-mediated decrease in ApRictor levels in Aplysia sensory neurons led to a decrease in the phosphorylation of PKC Apl II at the hydrophobic site suggesting a role for ApRictor in hydrophobic site phosphorylation. However, over-expression of ApRictor was not sufficient to increase phosphorylation of PKC Apl II. Continuous application of serotonin increased phosphorylation of PKC Apl II at the hydrophobic site in cultured sensory neurons, and this was blocked by Torin, which inhibits both TORC1 and TORC2. Over-expression of ApRictor did not lead to change in the magnitude of serotonin-mediated phosphorylation, but did lead to a small increase in the membrane localization of phosphorylated PKC Apl II. In conclusion, these studies implicate Rictor in phosphorylation of a novel PKC during synaptic plasticity and suggest an additional role for Rictor in regulating the localization of PKCs.  相似文献   

5.
The human neuronal Cys‐loop ligand‐gated ion channel superfamily of ion channels are important determinants of human behavior and the target of many drugs. It is essential for their structural characterization to achieve high‐level expression in a functional state. The aim of this work was to establish stable mammalian cell lines that enable high‐level heterologous production of pure receptors in a state that supports agonist‐induced allosteric conformational changes. In a tetracycline‐inducible stable human embryonic kidney cells (HEK293S) cell line, GABAA receptors containing α1 and β3 subunits could be expressed with specific activities of 29–34 pmol/mg corresponding to 140–170 pmol/plate, the highest expression level reported so far. Comparable figures for serotonin (5‐HT3A) receptors were 49–63 pmol/mg and 245–315 pmol/plate. The expression of 10 nmol of either receptor in suspension in a bioreactor required 0.3–3.0 L. Both receptor constructs had a FLAG epitope inserted at the N‐terminus and could be purified in one step after solubilization using ANTI‐FLAG affinity chromatography with yields of 30–40%. Purified receptors were functional. Binding of the agonist [3H]muscimol to the purified GABAAR was enhanced allosterically by the general anesthetic etomidate, and purified 5‐hydroxytryptamine‐3A receptor supported serotonin‐stimulated cation flux when reconstituted into lipid vesicles.  相似文献   

6.
7.
The protein ER-alpha has been exhaustively characterized in estrogen-sensitive tissues and cell lines. However, little is known regarding the expression and cellular distribution of the newly identified ER-beta protein. We first quantified the specific estradiol binding site content in the estrogen-responsive cell lines MCF-7 (mammary) and SHM (myometrial). In the two cell types, these sites were associated to the expression of both ER-alpha and -beta isoforms. Native ER-beta was visualized to reside inside the nucleus by means of conventional indirect immunofluorescence. The cells expressed ER-beta as a tight approximately 50 kDa triplet when resolved by sodium dodecyl sulfate-polyacrylamide gels (SDS-PAGE) and blotted using antibodies mapping different domains of the cloned ER-beta version. When the cells were subjected to homogenization and differential centrifugation, a substantial proportion of ER-beta immunolabeling was localized at membrane subfractions. ER-beta expression and partitioning was confirmed by Ligand blotting assays using estrogen derivatives coupled to different macromolecular tags. However, ER-alpha was expressed as the major estrogen binding protein in both cell lines. Similar localization experiments were performed on HeLa cells (cervix). Though usually considered ER-negative, this cell line displayed basal significant estrogen binding capacity and co-expression of both ER isoforms. Taken as a whole, the results indicate that ER-beta could be expressed as functional estrogen binding proteins among a dominant population of ER-alpha sites in the cell lines under study.  相似文献   

8.
In Aplysia californica, the serotonin-mediated translocation of protein kinase C (PKC) Apl II to neuronal membranes is important for synaptic plasticity. The orthologue of PKC Apl II, PKC, has been reported to require phosphatidic acid (PA) in conjunction with diacylglycerol (DAG) for translocation. We find that PKC Apl II can be synergistically translocated to membranes by the combination of DAG and PA. We identify a mutation in the C1b domain (arginine 273 to histidine; PKC Apl II-R273H) that removes the effects of exogenous PA. In Aplysia neurons, the inhibition of endogenous PA production by 1-butanol inhibited the physiological translocation of PKC Apl II by serotonin in the cell body and at the synapse but not the translocation of PKC Apl II-R273H. The translocation of PKC Apl II-R273H in the absence of PA was explained by two additional effects of this mutation: (i) the mutation removed C2 domain-mediated inhibition, and (ii) the mutation decreased the concentration of DAG required for PKC Apl II translocation. We present a model in which, under physiological conditions, PA is important to activate the novel PKC Apl II both by synergizing with DAG and removing C2 domain-mediated inhibition.  相似文献   

9.
We have studied the role of different amino acids in the M2 transmembrane domain of the α7 neuronal nicotinic receptor by mutating residues that differ from the ones located at the same positions in other α (α2-α10) or β (β2-β4) subunits. Our aim was to investigate the contribution of these amino acids to the peculiar kinetic and inward rectification properties that differentiate the homomeric α7 receptor from other nicotinic receptors. Mutations of several residues strongly modified receptor function. We found that Thr245 had the most profound effect when mutated to serine, an amino acid present in all heteromeric receptors composed of α and β subunits, by dramatically increasing the maximal current, decreasing the decaying rate of the currents and decreasing receptor rectification. Some mutants also showed altered agonist-binding properties as revealed by shifts in the dose-response curves for acetylcholine. We conclude that residues in the M2 segment and flanking regions contribute to the unusual properties of the α7 receptor, especially to its characteristic fast kinetic behavior and strong inward rectification and furthermore to the potency of agonists.  相似文献   

10.
In the mammalian brain, the α7 nicotinic and NMDA receptor antagonist kynurenic acid is synthesized by irreversible enzymatic transamination of the tryptophan metabolite l-kynurenine. d-kynurenine, too, serves as a bioprecursor of kynurenic acid in several organs including the brain, but the conversion is reportedly catalyzed through oxidative deamination by d-amino acid oxidase. Using brain and liver tissue homogenates from rats and humans, and conventional incubation conditions for kynurenine aminotransferases, we show here that kynurenic acid production from d-kynurenine, like the more efficient kynurenic acid synthesis from l-kynurenine, is blocked by the aminotransferase inhibitor amino-oxyacetic acid. In vivo, focal application of 100 μM d-kynurenine by reverse microdialysis led to a steady rise in extracellular kynurenic acid in the rat striatum, causing a 4-fold elevation after 2 h. Attesting to functional significance, this increase was accompanied by a 36% reduction in extracellular dopamine. Both of these effects were duplicated by perfusion of 2 μM l-kynurenine. Co-infusion of amino-oxyacetic acid (2 mM) significantly attenuated the in vivo effects of d-kynurenine and essentially eliminated the effects of l-kynurenine. Thus, enzymatic transamination accounts in part for kynurenic acid synthesis from d-kynurenine in the brain. These results are discussed with regard to implications for brain physiology and pathology.  相似文献   

11.
Long‐term treatment with nicotine or selective α7 nicotinic acetylcholine receptor (nAChR) agonists increases the number of α7 nAChRs and this up‐regulation may be involved in the mechanism underlying the sustained procognitive effect of these compounds. Here, we investigate the influence of type I and II α7 nAChR positive allosteric modulators (PAMs) on agonist‐induced α7 nAChR up‐regulation. We show that the type II PAMs, PNU‐120596 (10 μM) or TQS (1 and 10 μM), inhibit up‐regulation, as measured by protein levels, induced by the α7 nAChR agonist A‐582941 (10 nM or 10 μM), in SH‐EP1 cells stably expressing human α7 nAChR, whereas the type I PAMs AVL‐3288 or NS1738 do not. Contrarily, neither type I nor II PAMs affect 10 μM nicotine‐induced receptor up‐regulation, suggesting that nicotine and A‐582941 induce up‐regulation through different mechanisms. We further show in vivo that 3 mg/kg PNU‐120596 inhibits up‐regulation of the α7 nAChR induced by 10 mg/kg A‐582941, as measured by [125I]‐bungarotoxin autoradiography, whereas 1 mg/kg AVL‐3288 does not. Given that type II PAMs decrease desensitization of the receptor, whereas type I PAMs do not, these results suggest that receptor desensitization is involved in A‐582941‐induced up‐regulation. Our results are the first to show an in vivo difference between type I and II α7 nAChR PAMs, and demonstrate an agonist‐dependent effect of type II PAMs occurring on a much longer time scale than previously appreciated. Furthermore, our data suggest that nicotine and A‐582941 induce up‐regulation through different mechanisms, and that this confers differential sensitivity to the effects of α7 nAChR PAMs. These results may have implications for the clinical development of α7 nAChR PAMs.  相似文献   

12.
Neuronal nicotinic alpha7 subunits assemble into cell-surface complexes that neither function nor bind alpha-bungarotoxin when expressed in tsA201 cells. Functional alpha-bungarotoxin receptors are expressed if the membrane-spanning and cytoplasmic domains of the alpha7 subunit are replaced by the homologous regions of the serotonin-3 receptor subunit. Bgt-binding surface receptors assembled from chimeric alpha7/serotonin-3 subunits contain subunits in two different conformations as shown by differences in redox state and other features of the subunits. In contrast, alpha7 subunit complexes in the same cell line contain subunits in a single conformation. The appearance of a second alpha7/serotonin-3 subunit conformation coincides with the formation of alpha-bungarotoxin-binding sites and intrasubunit disulfide bonding, apparently within the alpha7 domain of the alpha7/serotonin-3 chimera. In cell lines of neuronal origin that produce functional alpha7 receptors, alpha7 subunits undergo a conformational change similar to alpha7/serotonin-3 subunits. alpha7 subunits, thus, can fold and assemble by two different pathways. Subunits in a single conformation assemble into nonfunctional receptors, or subunits expressed in specialized cells undergo additional processing to produce functional, alpha-bungarotoxin-binding receptors with two alpha7 conformations. Our results suggest that alpha7 subunit diversity can be achieved postranslationally and is required for functional homomeric receptors.  相似文献   

13.
Hemidesmosomes (HDs) are cellular junctions that anchor epithelial cells to the extracellular matrix (ECM) and are associated morphologically with the cytoskeleton. Hemidesmosomal molecular components include two proteins involved in linking intermediate filaments, HD1/plectin and BP230, and two transmembrane proteins, BP180 and the alpha6beta4 integrin, a laminin receptor. In cells lacking BP230 and BP180, HD1/plectin still associates with alpha6beta4 integrin, forming HD-like structures, called type II HDs. In the present study, we used an intestinal epithelial cell line that expresses HD1/plectin and the alpha6beta4 integrin to investigate the regulation of assembly of these proteins in type II HDs. These compounds were found to be clustered at sites of cell-ECM contact and their polarized localization was influenced by either cell confluency or extracellular matrix deposition. Conventional and immunoelectron microscopy showed that HD1/plectin and the beta4 integrin subunit are colocalized in an adhesion structure. Using cytoskeleton-disrupting drugs and confocal microscopy, we demonstrated that type II HDs are made up of numerous individual plaques whose assembly into a cluster requires actin filaments, but not microtubules.  相似文献   

14.
Retinoid acid receptors (RXR-alpha, -beta, -gamma) and Farnesoid X-activated receptor (FXR) expression in the testis of the marbled newt were investigated with special attention to the changes during the annual testicular cycle, using light microscopy immunohistochemistry and Western blot analysis. The annual testicular cycle of the marbled newt (Triturus marmoratus marmoratus) comprises three periods: (a) proliferative period (germ cell proliferation from primordial germ cells to round spermatids, April-June); (b) spermiogenesis period (July-September); and (c) quiescence period (interstitial and follicular cells form the glandular tissue, October-April). In the proliferative period, primordial germ cells and primary spermatogonia immunostained intensely to the three types of RXRs and also to FXR. In the other periods, immunostaining to these antibodies was weak or absent. Secondary spermatogonia stained weakly to the four antibodies in the proliferative period, and only to FXR, also weakly, in the spermiogenesis period. Immunoreactive primary spermatocytes were weakly labeled with the RXR antibodies in the proliferative period. Spermatids and spermatozoa did not stain to any antibody in any period. Follicular cells only immunostained to RXR-gamma and only in the quiescence period when they are forming the glandular tissue, together with the interstitial cells. As follicular cells, interstitial cells only immunostained in the quiescence period; however, they immunoreacted to the three types of RXRs. These findings suggest that in the newt, RXRs and FXR are involved in spermatogenesis control by regulating the proliferation of primordial germ cells and spermatogonia. In addition, RXR-gamma seems to be also involved in the development of the glandular (steroidogenic) tissue.  相似文献   

15.
alpha7 Nicotinic receptors are calcium permeant and provide neuroprotection against many insults. We investigated the roles of intracellular calcium ions and downstream calcium channels in this protection. The alpha7 agonist GTS-21 prevented pheochromocytoma cell death induced by nerve growth factor + serum deprivation over a 3-day interval. This effect was blocked by the intracellular calcium chelator 1,2-bis(2-aminophenoxy)ethane-N,N,N',N'-tetraacetic acid in a manner that did not appear to involve changes in receptor density. 1,2-Bis(2-aminophenoxy)ethane-N,N,N',N'-tetraacetic acid blocked GTS-21-induced protein kinase C activation, a necessary process for protection. The insositol triphosphate calcium-channel blocker xestospongin C and the phospholipases C inhibitor U-73122 blocked protection, ryanodine partially attenuated protection, but the L-type channel antagonist nifedipine had no effect. ERK1/2 but not JNK and p38 were activated by GTS-21, and the ERK phosphorylation inhibitors PD98059 and U0126 blocked protection.  相似文献   

16.
alpha-Conotoxins, from cone snails, and alpha-neurotoxins, from snakes, are competitive inhibitors of nicotinic acetylcholine receptors (nAChRs) that have overlapping binding sites in the ACh binding pocket. These disulphide-rich peptides are used extensively as tools to localize and pharmacologically characterize specific nAChRs subtypes. Recently, a homology model based on the high-resolution structure of an ACh binding protein (AChBP) allowed the three-fingered alpha-neurotoxins to be docked onto the alpha7 nAChR. To investigate if alpha-conotoxins interact with the nAChR in a similar manner, we built homology models of human alpha7 and alpha3beta2 nAChRs, and performed docking simulations of alpha-conotoxins ImI, PnIB, PnIA and MII using the program GOLD. Docking revealed that alpha-conotoxins have a different mode of interaction compared with alpha-neurotoxins, with surprisingly few nAChR residues in common between their overlapping binding sites. These docking experiments show that ImI and PnIB bind to the ACh binding pocket via a small cavity located above the beta9/beta10 hairpin of the (+)alpha7 nAChR subunit. Interestingly, PnIB, PnIA and MII were found to bind in a similar location on alpha7 or alpha3beta2 receptors mostly through hydrophobic interactions, while ImI bound further from the ACh binding pocket, mostly through electrostatic interactions. These findings, which distinguish alpha-conotoxin and alpha-neurotoxin binding modes, have implications for the rational design of selective nAChR antagonists.  相似文献   

17.
Hyaluronan accumulation in the retroorbital connective tissue is one of the pathological features of Graves' ophthalmopathy. Interleukin-1beta (IL-1beta) is known to stimulate hyaluronan synthesis in orbital fibroblasts. In the present study, the intracellular signal transduction pathways involved in this stimulatory effect were investigated in cultured human retroorbital fibroblasts from patients with Graves' ophthalmopathy. IL-1beta-induced hyaluronan synthesis was significantly inhibited by pretreatment of the cells with two protein kinase C (PKC) inhibitors, chlerythrine chloride and H-7. In addition, treatment with phorbol 12-myristate 13-acetate (PMA), a direct PKC activator, also resulted in increased hyaluronan production. IL-1beta- or PMA-stimulated hyaluronan synthesis was blocked by the protein synthesis inhibitor, cycloheximide. Moreover, the intracellular Ca(2+) concentration of the orbital fibroblasts was also involved in the IL-1beta induced transduction pathway, the effect being completely inhibited by BAPTA, an internal calcium chelator. In addition, A23187, a calcium ionophore, increased hyaluronan synthesis in unstimulated cells. These results suggest that the Ca(2+)-dependent PKC signal transduction pathway plays an important role in the IL-1beta-induced hyaluronan synthesis. Moreover, IL-1beta treatment resulted in increased PKC activity and the rapid translocation of PKC betaII from the cytoplasm to the plasma membrane. These results indicate that cytosolic Ca(2+) and PKC betaII are involved in IL-1beta-induced hyaluronan synthesis in cultured orbital fibroblasts from patients with Graves' ophthalmopathy.  相似文献   

18.
Background information. At fertilization in mammalian eggs, the sperm induces a series of Ca2+ oscillations via the production of inositol 1,4,5‐trisphosphate. Increased inositol 1,4,5‐trisphosphate production appears to be triggered by a sperm‐derived PLCζ (phospholipase C‐ζ) that enters the egg after gamete fusion. The specific phosphatidylinositol 4,5‐bisphosphate hydrolytic activity of PLCζ implies that DAG (diacylglycerol) production, and hence PKC (protein kinase C) stimulation, also occurs during mammalian egg fertilization. Fertilization‐mediated increase in PKC activity has been demonstrated; however, its precise role is unclear. Results. We investigated PLCζ‐ and fertilization‐mediated generation of DAG in mouse eggs by monitoring plasma‐membrane translocation of a fluorescent DAG‐specific reporter. Consistent plasma‐membrane DAG formation at fertilization, or after injection of physiological concentrations of PLCζ, was barely detectable. However, when PLCζ is overexpressed in eggs, significant plasma‐membrane DAG production occurs in concert with a series of unexpected secondary high‐frequency Ca2+ oscillations. We show that these secondary Ca2+ oscillations can be mimicked in a variety of situations by the stimulation of PKC and that they can be prevented by PKC inhibition. The way PKC leads to secondary Ca2+ oscillations appears to involve Ca2+ influx and the loading of thapsigargin‐sensitive Ca2+ stores. Conclusions. Our results suggest that overproduction of DAG in PLCζ‐injected eggs can lead to PKC‐mediated Ca2+ influx and subsequent overloading of Ca2+ stores. These results suggest that DAG generation in the plasma membrane of fertilizing mouse eggs is minimized since it can perturb egg Ca2+ homoeostasis via excessive Ca2+ influx.  相似文献   

19.
Phosphorylation of protein kinase Cs (PKCs) by phosphoinositide-dependent kinase I (PDK) is critical for PKC activity. In the nervous system of the marine mollusk Aplysia, there are only two major PKC isoforms, the calcium-activated PKC Apl I and the calcium-independent PKC Apl II, and both PKCs are persistently activated during intermediate memory. We monitored the PDK-dependent phosphorylation of PKC Apl I and PKC Apl II using phosphopeptide antibodies. During persistent activation of PKCs in Aplysia neurons, there is a significant increase in the amount of PDK-phosphorylated PKC Apl II in the particulate fraction but no increase in the amount of PKC Apl I phosphorylated by PDK. PDK phosphorylation of PKCs was not sensitive to inhibitors of phosphatidylinositol 3-kinase, PKC, or expression of a kinase-inactive PDK. Localization of PDK-phosphorylated PKC Apl II using immunocytochemistry revealed an enrichment of phosphorylated PKC Apl II at the plasma membrane. These data suggest that increased PDK phosphorylation of PKC Apl II is important for persistent kinase 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