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.
2.
Phosphorylation of histone H3 on Ser-10 is regarded as an epigenetic mitotic marker and is tightly correlated with chromosome condensation during both mitosis and meiosis. However, it was also reported that histone H3 Ser-10 phosphorylation occurs when cells are exposed to various death stimuli, suggesting a potential role in the regulation of apoptosis. Here we report that histone H3 Ser-10 phosphorylation is mediated by the pro-apoptotic kinase protein kinase C (PKC) δ during apoptosis. We observed that PKCδ robustly phosphorylates histone H3 on Ser-10 both in vitro and in vivo. Ectopic expression of catalytically active PKCδ efficiently induces condensed chromatin structure in the nucleus. We also discovered that activation of PKCδ is required for histone H3 Ser-10 phosphorylation after treatment with DNA damaging agents during apoptosis. Collectively, these findings suggest that PKCδ is the kinase responsible for histone H3 Ser-10 phosphoryation during apoptosis and thus contributes to chromatin condensation together with other apoptosis-related histone modifications. As a result, histone H3 Ser-10 phosphorylation can be designated a new ‘apoptotic histone code’ mediated by PKCδ.  相似文献   

3.
4.
Activation of the 5-hydroxytryptamine receptor 2B (5-HT2B), a Gq/11 protein-coupled receptor, results in proliferation of various cell types. The 5-HT2B receptor is also expressed on the pacemaker cells of the gastrointestinal tract, the interstitial cells of Cajal (ICC), where activation triggers ICC proliferation. The goal of this study was to characterize the mitogenic signal transduction cascade activated by the 5-HT2B receptor. All of the experiments were performed on mouse small intestine primary cell cultures. Activation of the 5-HT2B receptor by its agonist BW723C86 induced proliferation of ICC. Inhibition of phosphatidylinositol 3-kinase by LY294002 decreased base-line proliferation but had no effect on 5-HT2B receptor-mediated proliferation. Proliferation of ICC through the 5-HT2B receptor was inhibited by the phospholipase C inhibitor U73122 and by the inositol 1,4,5-trisphosphate receptor inhibitor Xestospongin C. Calphostin C, the α, β, γ, and μ protein kinase C (PKC) inhibitor Gö6976, and the α, β, γ, δ, and ζ PKC inhibitor Gö6983 inhibited 5-HT2B receptor-mediated proliferation, indicating the involvement of PKC α, β, or γ. Of all the PKC isoforms blocked by Gö6976, PKCγ and μ mRNAs were found by single-cell PCR to be expressed in ICC. 5-HT2B receptor activation in primary cell cultures obtained from PKCγ−/− mice did not result in a proliferative response, further indicating the requirement for PKCγ in the proliferative response to 5-HT2B receptor activation. The data demonstrate that the 5-HT2B receptor-induced proliferative response of ICC is through phospholipase C, [Ca2+]i, and PKCγ, implicating this PKC isoform in the regulation of cellular proliferation.Tight control of cell proliferation is essential to maintain organ size and function. Proliferation needs to be tightly regulated to maintain a critical mass of a particular cell type while preventing dysplasia or malignancy. Cell proliferation is regulated by a complex interaction between extrinsic and intrinsic factors. Extrinsic factors usually signal through cell surface receptors such as various growth factor receptors. 5-Hydroxytryptamine (5-HT,2 serotonin) is well established as a neurotransmitter and a paracrine factor with over 90% of 5-HT produced by the gastrointestinal tract (1, 2). There is now substantial evidence that, together with these established functions, 5-HT is involved in the control of cell proliferation through various 5-HT receptors, in particular the 5-hydroxytryptamine receptor 2B (5-HT2B (39)). The 5-HT2B receptor is Gq/11 protein-coupled. Activation of the 5-HT2B receptor regulates cardiac function, smooth muscle contractility, vascular physiology, and mood control. Recently it was demonstrated that activation of the 5-HT2B receptor also induces proliferation of neurons, retinal cells (3, 4), hepatocytes (5), osteoblasts (8), and interstitial cells of Cajal (ICC) (9). ICC express the 5-HT2B receptor, and activation by 5-HT induces proliferation of ICC (9). ICC are specialized, mesoderm-derived mesenchymal cells in the gastrointestinal tract. Their best known function is the generation of slow waves (10), but they also conduct and amplify neuronal signals (11, 12), release carbon monoxide to set the intestinal smooth muscle membrane potential gradient (13), and act as mechanosensors (14, 15). Loss of ICC has been associated with pathological conditions such as gastroparesis (1618), infantile pyloric stenosis (19, 20), pseudo-obstruction (21, 22), and slow transit constipation (23), whereas increased proliferation of ICC or their precursors is associated with gastrointestinal stromal tumors (24).The mechanisms by which activation of the 5-HT2B receptor results in increased proliferation are not well understood. In cultured cardiomyocytes, stimulation of the 5-HT2B receptor activated both phosphatidylinositol 3-kinase (PI3′-K)/Akt and ERK1/2/mitogen-activated protein kinase (MAPK) signaling pathways to protect cardiomyocytes from apoptosis (25). On the other hand, the 5-HT2 subfamily of receptors are also known to couple to phospholipase C (PLC) (2628).The objective of this study was to utilize the known expression of the 5-HT2B receptor on ICC to determine whether proliferation through the 5-HT2B receptor required PI3′-K or PLC. This study demonstrates that proliferation mediated by the 5-HT2B receptor requires PLC, intracellular calcium release, and the ERK/MAPK signaling pathway and identifies the PKC isoform activated by the 5-HT2B receptor in ICC as PKCγ.  相似文献   

5.
The pituitary gland dynamically changes its hormone output under various pathophysiological conditions. One of the pathways implicated in the regulatory mechanism of this gland is a dopaminergic system that operates the phosphoinositide (PI) cycle to transmit downstream signal through second messengers. We have previously shown that diacylglycerol kinase β (DGKβ) is coexpressed with dopamine D1 and D2 receptors in medium spiny neurons of the striatum, suggesting a plausible implication of DGKβ in dopaminergic transmission. However, it remains elusive whether DGKβ is involved in the dopaminergic system in the pituitary gland. The aim of this study is to investigate the expression and localization of DGK in the pituitary gland, together with the molecular components involved in the PI signaling cascade, including dopamine receptors, phospholipase C (PLC), and a major downstream molecule, protein kinase C (PKC). Here we show that DGKβ and the dopamine D2 receptor are coexpressed in the intermediate lobe and localize to the plasma membrane side by side. In addition, we reveal that PLCβ4 and PKCα are the subtypes expressed in the intermediate lobe among those families. These findings will substantiate and further extend our understanding of the molecular-anatomical pathway of PI signaling and the functional roles of DGK in the pituitary intermediate lobe. (J Histochem Cytochem 58:119–129, 2010)  相似文献   

6.
Impaired epithelial barrier function plays a crucial role in the pathogenesis of inflammatory bowel disease. Elevated levels of the pro-inflammatory cytokine, interferon-γ (IFNγ), are believed to be prominently involved in the pathogenesis of Crohn disease. Treatment of T84 intestinal epithelial cells with IFNγ severely impairs their barrier properties measured as transepithelial electrical resistance (TER) or permeability and reduces the expression of tight junction proteins such as occludin and zonula occludens-1 (ZO-1). However, little is known about the signaling events that are involved. The cellular energy sensor, AMP-activated protein kinase (AMPK), is activated in response to cellular stress, as occurs during inflammation. The aim of this study was to investigate a possible role for AMPK in mediating IFNγ-induced effects on the intestinal epithelial barrier. We found that IFNγ activates AMPK by phosphorylation, independent of intracellular energy levels. Inhibition of AMPK prevents, at least in part, the IFNγ-induced decrease in TER. Furthermore, AMPK knockdown prevented the increased epithelial permeability, the decreased TER, and the decrease in occludin and ZO-1 caused by IFNγ treatment of T84 cells. However, AMPK activity alone was not sufficient to cause alterations in epithelial barrier function. These data show a novel role for AMPK, in concert with other signals induced by IFNγ, in mediating reduced epithelial barrier function in a cell model of chronic intestinal inflammation. These findings may implicate AMPK in the pathogenesis of chronic intestinal inflammatory conditions, such as inflammatory bowel disease.Inflammatory bowel disease (IBD)2 consists of two major subgroups, ulcerative colitis and Crohn disease (CD). A complex cascade of genetic, immunological, and bacterial factors contributes to IBD pathogenesis (1). In the healthy intestine, the epithelial barrier separates the luminal bacterial microbiota and other aspects of the external environment from cells of the mucosal immune system. In CD in particular, an impaired epithelial barrier (2, 3) leads to increased exposure of the immune system to commensal bacteria. Along with possible genetic defects in bacterial sensing, this might contribute to a dysregulated immune response leading to further epithelial damage and active episodes of IBD (4). Epithelial barrier dysfunction in CD is characterized by alterations in intercellular tight junctions (5), as well as by an excessive loss of water and salt into the lumen. An important immunological marker in CD is the existence of excessively high levels of the pro-inflammatory cytokine, interferon gamma (IFNγ) (6).IFNγ treatment of intestinal epithelial cell monolayers severely compromises their barrier integrity. Most importantly from a functional perspective, IFNγ causes a decrease in transepithelial electrical resistance (TER) and increases epithelial permeability (7, 8). These defects closely resemble observations in CD, where there is a disruption of intercellular tight junctional complexes. This effect is due to disruption of the apical actin cytoskeleton in conjunction with decreased expression, as well as increased internalization, of important tight junction proteins such as occludin and zonula occludens-1 (ZO-1) (811). Conversely, induction of epithelial apoptosis by IFNγ is believed to contribute little to barrier dysfunction (12). IFNγ also induces further alterations in epithelial function that include reduced expression of various ion transporters and associated decreases in epithelial ion transport (13, 14). Despite the influence of IFNγ on a number of epithelial functions, relatively little is known about intracellular signaling mechanisms mediating its effects following receptor activation. Recent studies demonstrated the involvement of phosphatidylinositol 3′-kinase (PI3K) in mediating IFNγ-induced effects on epithelial barrier function (11, 15). However, this is unlikely to be the only regulatory pathway involved. Indeed, increased expression of receptors for tumor necrosis factor core family members, such as the tumor necrosis factor receptor and LIGHT (homologous to lymphotoxin, shows inducible expression and competes with herpes simplex virus glycoprotein D for herpes virus entry mediator (HVEM), a receptor expressed by T lymphocytes), can also occur in response to IFNγ and lead to changes in intestinal barrier function (1618).The effects of IFNγ in intestinal epithelial cells resemble, at least in part, those of the cellular energy sensor, AMP-activated protein kinase (AMPK). Upon activation, AMPK restores intracellular ATP levels by stimulating energy-producing pathways, such as glucose uptake (19) and glycolysis, while inhibiting energy-consuming pathways, such as the synthesis of fatty acids or triglycerides (20, 21). In the intestine, energy-consuming processes include epithelial ion transport, and, indeed, AMPK has been shown to decrease intestinal ATP-consuming ion transport as well as the synthesis of various proteins (22, 23). Moreover, it has previously been demonstrated that ion transport processes are suppressed in intestinal biopsies from IBD patients (2426).AMPK is usually activated in response to cellular stress that depletes intracellular ATP and elevates the AMP:ATP ratio (27, 28). AMPK-activating conditions include oxidative stress (29), hypoxia (30), and hypoglycemia (31). Binding of AMP to AMPK causes an increase in activity of 5-fold or less (32). Further, binding of AMP to AMPK makes AMPK a better substrate for upstream kinase activation, resulting in phosphorylation of the catalytic α-subunit of AMPK on the Thr172 residue and subsequently in a 50- to 100-fold activation of the enzyme (32). A number of upstream kinases for AMPK have been identified, with LKB1 (33, 34) or calmodulin kinase II (3537) being the most important and well studied. However, recent studies also indicate that PI3K can activate AMPK (38, 39).The goal of this study was to determine whether AMPK mediates IFNγ-induced alterations in intestinal epithelial barrier function. We found that IFNγ activates AMPK in intestinal epithelial cells and AMPK inhibition prevents, at least in part, IFNγ-induced barrier dysfunction. Our data indicate a novel role for the cellular energy sensor, AMPK, in the regulation of intestinal epithelial barrier properties in a cell model of chronic inflammation. These findings may have implications for barrier function in the setting of chronic inflammatory processes, such as IBD.  相似文献   

7.
Proper, graded communication between different cell types is essential for normal development and function. In the nervous system, heart, and for some cancer cells, part of this communication requires signaling by soluble and membrane-bound factors produced by the NRG1 gene. We have previously shown that glial-derived neurotrophic factors activate a rapid, localized release of soluble neuregulin from neuronal axons that can, in turn promote proper axoglial development (Esper, R. M., and Loeb, J. A. (2004) J. Neurosci. 24, 6218–6227). Here we elucidate the mechanism of this localized, regulated release by implicating the delta isoform of protein kinase C (PKC). Blocking the PKC delta isoform with either rottlerin, a selective antagonist, or small interference RNA blocks the regulated release of neuregulin from both transfected cells and primary neuronal cultures. PKC activation also leads to the rapid phosphorylation of the pro-NRG1 cytoplasmic tail on serine residues adjacent to the membrane-spanning segment, that, when mutated markedly reduce the rate of NRG1 activity release. These findings implicate this specific PKC isoform as an important factor for the cleavage and neurotrophin-regulated release of soluble NRG1 forms that have important effects in nervous system development and disease.The neuregulins (NRGs)2 are a family of growth and differentiation factors with a broad range of functions during development and in the adult. NRGs are necessary for glial and cardiac development and participate in a wide range of biologic processes ranging from proper formation of peripheral nerves and the neuromuscular junction to tumor growth (29). The NRGs have also been implicated as both potential mediators and therapeutic targets for a number of human diseases including cancer, schizophrenia, and multiple sclerosis (1012). NRGs function as mediators of cell-to-cell communication through a multitude of alternatively spliced isoforms arising from at least four distinct genes that bind to and activate members of the epidermal growth factor receptor family HER-2/3/4 (ErbB-2/3/4) (1319).Although all known isoforms of the NRG1 gene have an epidermal growth factor-like domain sufficient to bind to and activate its receptors (20), products of this gene are divided into three classes based on structurally and functionally different N-terminal regions (21) The type I and II forms have a unique N-terminal, heparin-binding Ig-like domain (2226). This Ig-like domain potentiates the biological activities of soluble NRG1 forms and leads to their highly selective tissue distributions through its affinity for specific cell-surface heparan sulfates (12, 20, 27, 28). These forms are first expressed as transmembrane precursors (pro-NRG1) that undergo proteolytic cleavage to release their soluble ectodomains. The type III NRG1 forms, on the other hand, are not typically released from cells, because their N-terminal domain consists of a cysteine-rich domain that can serve as a membrane tether making this form ideal for juxtacrine signaling. This form has been strongly implicated to be important peripheral nerve myelination (2931).While many of the biological functions of type I/II NRG1 forms are less clear, their ability to be released from axons in the peripheral and central nervous systems in a regulated manner provides the potential for long range cell-cell communication not possible from membrane-bound forms. Studies examining the regulation of type I NRG1 release from neuronal axons have implicated protein kinase C (PKC) as a mediator of NRG1 release from pro-NRG1 in transfected cell lines (32). Subsequent studies in intact neurons found that PKC activation was sufficient to release NRG1 from sensory and motor neuron axons and that NRG1 could also be released by Schwann cell-derived neurotrophic factors, such as BDNF and GDNF (1). Recently, the β-secretase protease BACE1 has been suggested to cleave these NRG1 forms so that when it is knocked out in mice, deficits similar to those seen in NRG1 knockouts are seen (33, 34). These findings suggest that reciprocal communication between NRG1s and neurotrophins could be an important mechanisms for local axoglial communication that is critical for normal peripheral nerve development. Consistently, PKC has been implicated as a key mediator for the electrically mediated release of NRG1 from cultured cerebellar granule cells and pontine nucleus neurons (35).The PKC family consists of 10 serine/threonine kinases isoforms (α, βI, βII, γ, δ, ϵ, ζ, θ, λ, and η) each with a unique cellular distribution, target specificity, mechanism of activation, and function (36). One of these functions promotes the cleavage and release of soluble signaling proteins that are initially synthesized as membrane-spanning precursors. In addition to NRG1, other proteins released upon PKC activation include epidermal growth factor, transforming growth factor-α, amyloid precursor protein, l-selectin, and interleukins (1, 3743). We hypothesize that neurotrophic factors induce the cleavage and release of NRG1 from pro-NRG1 through PKC activation. This hypothesis seems reasonable, because neurotrophin binding to the Trk family of neurotrophin receptor tyrosine kinases, but not the low affinity neurotrophin receptor p75 (44), activates phospholipase Cγ-mediated conversion of membrane-bound phosphatidylinositol bisphosphate to inositol triphosphate and diacylglycerol, which in turn, can activate PKC (4548). Although this can be achieved using phorbol 12-myristate 13-acetate (PMA), a diacylglycerol analog sufficient to activate most PKC isozymes (48), the exact PKC isoform and mechanism by which this occurs is not known. Here, we demonstrate NRG1 is released from cells through direct activation of the PKCδ isoform using siRNA and PKC isoform-specific inhibitors in transfected Chinese hamster ovary (CHO) cells, PC12, and primary neuronal cultures. We further demonstrate that PKC activation induces rapid phosphorylation of the cytoplasmic tail of pro-NRG1 on specific serine residues that are required for efficient NRG1 activity release. These findings provide mechanistic insights into how highly localized, reciprocal signaling occurs along neuronal axons, which has important implications for normal development and disease.  相似文献   

8.
9.
10.
AMP-activated protein kinase (AMPK) is a heterotrimeric complex playing a crucial role in maintaining cellular energy homeostasis. Recently, homodimerization of mammalian AMPK and yeast ortholog SNF1 was shown by us and others. In SNF1, it involved specific hydrophobic residues in the kinase domain αG-helix. Mutation of the corresponding AMPK α-subunit residues (Val-219 and Phe-223) to glutamate reduced the tendency of the kinase to form higher order homo-oligomers, as was determined by the following three independent techniques in vitro: (i) small angle x-ray scattering, (ii) surface plasmon resonance spectroscopy, and (iii) two-dimensional blue native/SDS-PAGE. Recombinant protein as well as AMPK in cell lysates of primary cells revealed distinct complexes of various sizes. In particular, the assembly of very high molecular mass complexes was dependent on both the αG-helix-mediated hydrophobic interactions and kinase activation. In vitro and when overexpressed in double knock-out (α1−/−, α2−/−) mouse embryonic fibroblast cells, activation of mutant AMPK was impaired, indicating a critical role of the αG-helix residues for AMPK activation via its upstream kinases. Also inactivation by protein phosphatase 2Cα was affected in mutant AMPK. Importantly, activation of mutant AMPK by LKB1 was restored by exchanging the corresponding and conserved hydrophobic αG-helix residues of LKB1 (Ile-260 and Phe-264) to positively charged amino acids. These mutations functionally rescued LKB1-dependent activation of mutant AMPK in vitro and in cell culture. Our data suggest a physiological role for the hydrophobic αG-helix residues in homo-oligomerization of heterotrimers and cellular interactions, in particular with upstream kinases, indicating an additional level of AMPK regulation.The maintenance of energy homeostasis is a basic requirement of all living organisms. The AMP-activated protein kinase (AMPK)2 is crucially involved in this essential process by playing a central role in sensing and regulating energy metabolism on the cellular and whole body level (16). AMPK is also participating in several signaling pathways associated with cancer and metabolic diseases, like type 2 diabetes mellitus, obesity, and other metabolic disorders (79).Mammalian AMPK belongs to a highly conserved family of serine/threonine protein kinases with homologs found in all eukaryotic organisms examined (1, 3, 10). Its heterotrimeric structure includes a catalytic α-subunit and regulatory β- and γ-subunits. These subunits exist in different isoforms (α1, α2, β1, β2, γ1, γ2, and γ3) and splice variants (for γ2 and γ3) and can thus assemble to a broad variety of heterotrimeric isoform combinations. The α- and β-subunits possess multiple autophosphorylation sites, which have been implicated in regulation of subcellular localization and kinase activation (1115). The most critical step of AMPK activation, however, is phosphorylation of Thr-172 within the activation segment of the α-subunit kinase domain. At least two AMPK upstream kinases (AMPKKs) have been identified so far, namely the tumor suppressor kinase LKB1 in complex with MO25 and STRAD (16) and Ca2+/calmodulin-dependent protein kinase kinase-2 (CamKK2) (17). Furthermore, the transforming growth factor-β-activated kinase 1 was also shown to activate AMPK using a variety of in vitro approaches (18), but the physiological relevance of these findings remains unclear. Besides direct phosphorylation of Thr-172, AMPK activity is stimulated by the allosteric activator AMP, which can bind to two Bateman domains formed by two pairs of CBS domains within the γ-subunit (1922). Hereby bound AMP not only allosterically stimulates AMPK but also protects Thr-172 from dephosphorylation by protein phosphatase 2Cα (PP2Cα) and thus hinders inactivation of the kinase (19, 22, 23). Consequently, on the cellular level, AMPK is activated upon metabolic stress increasing the AMP/ATP ratio. Furthermore, AMPK activation can also be induced by several chemical compounds, like nucleoside 5-aminoimidazole-4-carboxamide-1-β-d-ribofuranoside (24) and the anti-diabetic drug Metformin (2528). In addition, the small molecule compound A-769662 was recently developed as a direct allosteric activator of AMPK (29, 30).Previous work in our groups proposed a model of AMPK regulation by AMP, which incorporates the major functional features and the latest structural information (31). The latter mainly included truncated core complexes of AMPK from different species (3235). Further valuable structural information is provided by the x-ray structures of the isolated catalytic domains, in particular of the human AMPK α2-subunit (Protein Data Bank code 2H6D) and its yeast ortholog SNF1 (36, 37). The kinase domain of SNF1 is capable of forming homodimers in the protein crystal, as well as in vitro in solution, in a unique way, which has not been observed previously in any other kinase (36). The dimer interface is predominantly formed by hydrophobic interactions of the loop-αG region, also known as subdomain X situated on the large kinase lobe (36, 38, 39), and it mainly involves Ile-257 and Phe-261. Because the T-loop activation segment was buried within the dimer interface, it was suggested that the dimeric state of the SNF1 catalytic domain represents the inactive form of the kinase. Intriguingly, it was shown in our groups by small angle x-ray scattering that AMPK self-organizes in a concentration-dependent manner to form homo-oligomers in solution (31). However, the interface responsible for oligomerization of the AMPK heterotrimer has remained elusive.Here we further investigate the distinct oligomeric states of the AMPK heterotrimer and suggest a possible regulatory function for this process. Most importantly, we provide conclusive evidence for participation of αG-helix residues in the recognition of AMPK by its upstream kinases LKB1 and CamKK2.  相似文献   

11.
12.
13.
Engineered overexpression of protein kinase Cα (PKCα) was previously shown to endow nonmotile MCF-10A human breast cells with aggressive motility. A traceable mutant of PKCα (Abeyweera, T. P., and Rotenberg, S. A. (2007) Biochemistry 46, 2364–2370) revealed that α6-tubulin is phosphorylated in cells expressing traceable PKCα and in vitro by wild type PKCα. Gain-of-function, single site mutations (Ser → Asp) were constructed at each PKC consensus site in α6-tubulin (Ser158, Ser165, Ser241, and Thr337) to simulate phosphorylation. Following expression of each construct in MCF-10A cells, motility assays identified Ser165 as the only site in α6-tubulin whose pseudophosphorylation reproduced the motile behavior engendered by PKCα. Expression of a phosphorylation-resistant mutant (S165N-α6-tubulin) resulted in suppression of MCF-10A cell motility stimulated either by expression of PKCα or by treatment with PKCα-selective activator diacylglycerol-lactone. MCF-10A cells treated with diacylglycerol-lactone showed strong phosphorylation of endogenous α-tubulin that could be blocked when S165N-α6-tubulin was expressed. The S165N mutant also inhibited intrinsically motile human breast tumor cells that express high endogenous PKCα levels (MDA-MB-231 cells) or lack PKCα and other conventional isoforms (MDA-MB-468 cells). Comparison of Myc-tagged wild type α6-tubulin and S165N-α6-tubulin expressed in MDA-MB-468 cells demonstrated that Ser165 is also a major site of phosphorylation for endogenously active, nonconventional PKC isoforms. PKC-stimulated motility of MCF-10A cells was nocodazole-sensitive, thereby implicating microtubule elongation in the mechanism. These findings support a model in which PKC phosphorylates α-tubulin at Ser165, leading to microtubule elongation and motility.  相似文献   

14.
To better study the role of PKCδ in normal function and disease, we developed an ATP analog-specific (AS) PKCδ that is sensitive to specific kinase inhibitors and can be used to identify PKCδ substrates. AS PKCδ showed nearly 200 times higher affinity (Km) and 150 times higher efficiency (kcat/Km) than wild type (WT) PKCδ toward N6-(benzyl)-ATP. AS PKCδ was uniquely inhibited by 1-(tert-butyl)-3-(1-naphthyl)-1H-pyrazolo[3,4-d]pyrimidin-4-amine (1NA-PP1) and 1-(tert-butyl)-3-(2-methylbenzyl)-1H-pyrazolo[3,4-d]pyrimidin-4-amine (2MB-PP1) but not by other 4-amino-5-(4-methylphenyl)-7-(t-butyl)pyrazolo[3,4-d]pyrimidine (PP1) analogs tested, whereas WT PKCδ was insensitive to all PP1 analogs. To understand the mechanisms for specificity and affinity of these analogs, we created in silico WT and AS PKCδ homology models based on the crystal structure of PKCι. N6-(Benzyl)-ATP and ATP showed similar positioning within the purine binding pocket of AS PKCδ, whereas N6-(benzyl)-ATP was displaced from the pocket of WT PKCδ and was unable to interact with the glycine-rich loop that is required for phosphoryl transfer. The adenine rings of 1NA-PP1 and 2MB-PP1 matched the adenine ring of ATP when docked in AS PKCδ, and this interaction prevented the potential interaction of ATP with Lys-378, Glu-428, Leu-430, and Phe-633 residues. 1NA-PP1 failed to effectively dock within WT PKCδ. Other PP1 analogs failed to interact with either AS PKCδ or WT PKCδ. These results provide a structural basis for the ability of AS PKCδ to efficiently and specifically utilize N6-(benzyl)-ATP as a phosphate donor and for its selective inhibition by 1NA-PP1 and 2MB-PP1. Such homology modeling could prove useful in designing molecules to target PKCδ and other kinases to understand their function in cell signaling and to identify unique substrates.  相似文献   

15.
Volatile anesthetics isoflurane possibly improves the ischemic brain injury. However, its molecular actions are still unclear. In ischemia, protein kinase C (PKC)γ and calcium/calmodulin dependent protein kinase II (CaMKII)-α are persistently translocated from cytosol to cell membranes, and diminish these translocation suggested to be neuroprotective. We thus tested a hypothesis that isoflurane inhibits PKCγ and CaMKII-α translocation after ischemic brain insults. C57Bl/6J male mice were made to inhale 1 or 2 MAC isoflurane, after which 3 or 5 min cerebral ischemia was induced by decapitation. The sampled cerebrum cortex was then homogenized and centrifuged into crude synaptosomal fractions (P2), cytosolic fractions (S3), and particulate fractions (P3). CaMKII-α and PKCγ levels of these fractions were analyzed by immunoblotting. PKCγ and CaMKII-α are translocated to synaptic membrane from cytosol by cerebral ischemia, although isoflurane significantly inhibited such translocation. These results may explain in part the cellular and molecular mechanisms of neuroprotective effects of isoflurane.  相似文献   

16.
The novel class of protein kinase C (nPKC) isoform η is expressed in platelets, but not much is known about its activation and function. In this study, we investigated the mechanism of activation and functional implications of nPKCη using pharmacological and gene knock-out approaches. nPKCη was phosphorylated (at Thr-512) in a time- and concentration-dependent manner by 2MeSADP. Pretreatment of platelets with MRS-2179, a P2Y1 receptor antagonist, or YM-254890, a Gq blocker, abolished 2MeSADP-induced phosphorylation of nPKCη. Similarly, ADP failed to activate nPKCη in platelets isolated from P2Y1 and Gq knock-out mice. However, pretreatment of platelets with P2Y12 receptor antagonist, AR-C69331MX did not interfere with ADP-induced nPKCη phosphorylation. In addition, when platelets were activated with 2MeSADP under stirring conditions, although nPKCη was phosphorylated within 30 s by ADP receptors, it was also dephosphorylated by activated integrin αIIbβ3 mediated outside-in signaling. Moreover, in the presence of SC-57101, a αIIbβ3 receptor antagonist, nPKCη dephosphorylation was inhibited. Furthermore, in murine platelets lacking PP1cγ, a catalytic subunit of serine/threonine phosphatase, αIIbβ3 failed to dephosphorylate nPKCη. Thus, we conclude that ADP activates nPKCη via P2Y1 receptor and is subsequently dephosphorylated by PP1γ phosphatase activated by αIIbβ3 integrin. In addition, pretreatment of platelets with η-RACK antagonistic peptides, a specific inhibitor of nPKCη, inhibited ADP-induced thromboxane generation. However, these peptides had no affect on ADP-induced aggregation when thromboxane generation was blocked. In summary, nPKCη positively regulates agonist-induced thromboxane generation with no effects on platelet aggregation.Platelets are the key cellular components in maintaining hemostasis (1). Vascular injury exposes subendothelial collagen that activates platelets to change shape, secrete contents of granules, generate thromboxane, and finally aggregate via activated αIIbβ3 integrin, to prevent further bleeding (2, 3). ADP is a physiological agonist of platelets secreted from dense granules and is involved in feedback activation of platelets and hemostatic plug stabilization (4). It activates two distinct G-protein-coupled receptors (GPCRs) on platelets, P2Y1 and P2Y12, which couple to Gq and Gi, respectively (58). Gq activates phospholipase Cβ (PLCβ), which leads to diacyl glycerol (DAG)2 generation and calcium mobilization (9, 10). On the other hand, Gi is involved in inhibition of cAMP levels and PI 3-kinase activation (4, 6). Synergistic activation of Gq and Gi proteins leads to the activation of the fibrinogen receptor integrin αIIbβ3. Fibrinogen bound to activated integrin αIIbβ3 further initiates feed back signaling (outside-in signaling) in platelets that contributes to the formation of a stable platelet plug (11).Protein kinase Cs (PKCs) are serine/threonine kinases known to regulate various platelet functional responses such as dense granule secretion and integrin αIIbβ3 activation (12, 13). Based on their structure and cofactor requirements, PKCs are divided in to three classes: classical (cofactors: DAG, Ca2+), novel (cofactors: DAG) and atypical (cofactors: PIP3) PKC isoforms (14). All the members of the novel class of PKC isoforms (nPKC), viz. nPKC isoforms δ, θ, η, and ε, are expressed in platelets (15), and they require DAG for activation. Among all the nPKCs, PKCδ (15, 16) and PKCθ (1719) are fairly studied in platelets. Whereas nPKCδ is reported to regulate protease-activated receptor (PAR)-mediated dense granule secretion (15, 20), nPKCθ is activated by outside-in signaling and contributes to platelet spreading on fibrinogen (18). On the other hand, the mechanism of activation and functional role of nPKCη is not addressed as yet.PKCs are cytoplasmic enzymes. The enzyme activity of PKCs is modulated via three mechanisms (14, 21): 1) cofactor binding: upon cell stimulus, cytoplasmic PKCs mobilize to membrane, bind cofactors such as DAG, Ca2+, or PIP3, release autoinhibition, and attain an active conformation exposing catalytic domain of the enzyme. 2) phosphorylations: 3-phosphoinositide-dependent kinase 1 (PDK1) on the membrane phosphorylates conserved threonine residues on activation loop of catalytic domain; this is followed by autophosphorylations of serine/threonine residues on turn motif and hydrophobic region. These series of phosphorylations maintain an active conformation of the enzyme. 3) RACK binding: PKCs in active conformation bind receptors for activated C kinases (RACKs) and are lead to various subcellular locations to access the substrates (22, 23). Although various leading laboratories have elucidated the activation of PKCs, the mechanism of down-regulation of PKCs is not completely understood.The premise of dynamic cell signaling, which involves protein phosphorylations by kinases and dephosphorylations by phosphatases has gained immense attention over recent years. PP1, PP2A, PP2B, PHLPP are a few of the serine/threonine phosphatases reported to date. Among them PP1 and PP2 phosphatases are known to regulate various platelet functional responses (24, 25). Furthermore, PP1c, is the catalytic unit of PP1 known to constitutively associate with αIIb and is activated upon integrin engagement with fibrinogen and subsequent outside-in signaling (26). Among various PP1 isoforms, recently PP1γ is shown to positively regulate platelet functional responses (27). Thus, in this study we investigated if the above-mentioned phosphatases are involved in down-regulation of nPKCη. Furthermore, reports from other cell systems suggest that nPKCη regulates ERK/JNK pathways (28). In platelets ERK is known to regulate agonist induced thromboxane generation (29, 30). Thus, we also investigated if nPKCη regulates ERK phosphorylation and thereby agonist-induced platelet functional responses.In this study, we evaluated the activation of nPKCη downstream of ADP receptors and its inactivation by an integrin-associated phosphatase PP1γ. We also studied if nPKCη regulates functional responses in platelets and found that this isoform regulates ADP-induced thromboxane generation, but not fibrinogen receptor activation in platelets.  相似文献   

17.
18.
The FGF receptors (FGFRs) control a multitude of cellular processes both during development and in the adult through the initiation of signaling cascades that regulate proliferation, survival, and differentiation. Although FGFR tyrosine phosphorylation and the recruitment of Src homology 2 domain proteins have been widely described, we have previously shown that FGFR is also phosphorylated on Ser779 in response to ligand and binds the 14-3-3 family of phosphoserine/threonine-binding adaptor/scaffold proteins. However, whether this receptor phosphoserine mode of signaling is able to regulate specific signaling pathways and biological responses is unclear. Using PC12 pheochromocytoma cells and primary mouse bone marrow stromal cells as models for growth factor-regulated neuronal differentiation, we show that Ser779 in the cytoplasmic domains of FGFR1 and FGFR2 is required for the sustained activation of Ras and ERK but not for other FGFR phosphotyrosine pathways. The regulation of Ras and ERK signaling by Ser779 was critical not only for neuronal differentiation but also for cell survival under limiting growth factor concentrations. PKCϵ can phosphorylate Ser779 in vitro, whereas overexpression of PKCϵ results in constitutive Ser779 phosphorylation and enhanced PC12 cell differentiation. Furthermore, siRNA knockdown of PKCϵ reduces both growth factor-induced Ser779 phosphorylation and neuronal differentiation. Our findings show that in addition to FGFR tyrosine phosphorylation, the phosphorylation of a conserved serine residue, Ser779, can quantitatively control Ras/MAPK signaling to promote specific cellular responses.  相似文献   

19.
The coordinated and physiological behavior of living cells in an organism critically depends on their ability to interact with surrounding cells and with the extracellular space. For this, cells have to interpret incoming stimuli, correctly process the signals, and produce meaningful responses. A major part of such signaling mechanisms is the translation of incoming stimuli into intracellularly understandable signals, usually represented by second messengers or second-messenger systems. Two key second messengers, namely the calcium ion and signaling lipids, albeit extremely different in nature, play an important and often synergistic role in such signaling cascades. In this report, we will shed some light on an entire family of protein kinases, the protein kinases C, that are perfectly designed to exactly decode these two second messengers in all of their properties and convey the signaling content to downstream processes within the cell.Once generated, second messengers relay their information content in a plethora of properties, including time, quantity (i.e., concentration), space (i.e., subcellular distribution), and interestingly into any combination of these three characteristics. Nevertheless, such information is meaningless for the cell unless it has a toolkit of read-out systems that can actually interpret such second-messenger properties and relate them further downstream into complex signaling networks, or directly to effector systems. An important system is the family of protein kinase Cs (PKCs) that can read-out lipid signals alone, or combine the ability to read-out simultaneous lipid and Ca2+ signals. A common denominator of all PKCs is the property to convey signals downstream by phosphorylation of additional signaling partners or effector proteins. We will briefly introduce the PKC subfamilies with particular emphasis on their signaling ability, discuss the important sensing domains, and their properties, before concentrating on sensing details of the subfamily of conventional PKCs and their role in signal integration in greater depth.  相似文献   

20.
Adenosine monophosphate-activated protein (AMP)-activated kinase (AMPK) is a highly conserved kinase that plays a key role in energy homeostasis. Activation of AMPK was shown to reduce inflammation in response to lipolysaccharide in vitro and in vivo. 5-Aminoimidazole-4-carbox-amide-1-β-d-ribofuranoside (AICAR) is intracellularly converted to the AMP analog ZMP, which activates AMPK. Lipoteichoic acid (LTA) is a major component of the cell wall of Gram-positive bacteria that can trigger inflammatory responses. In contrast to lipopolysaccharide, little is known on the effects of AMPK activation in LTA-triggered innate immune responses. Here, we studied the potency of AMPK activation to reduce LTA-induced inflammation in vitro and in lungs in vivo. Activation of AMPK in vitro reduced cytokine production in the alveolar macrophage cell line MH-S. In vivo, AMPK activation reduced LTA-induced neutrophil influx, as well as protein leak and cytokine/chemokine levels in the bronchoalveolar space. In conclusion, AMPK activation inhibits LTA-induced lung inflammation in mice.  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