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 171 毫秒
1.
2.
3.
4.
5.
6.
Calcium (Ca2+) signaling by the pro-inflammatory cytokine interleukin-1 (IL-1) is dependent on focal adhesions, which contain diverse structural and signaling proteins including protein phosphatases. We examined here the role of protein-tyrosine phosphatase (PTP) α in regulating IL-1-induced Ca2+ signaling in fibroblasts. IL-1 promoted recruitment of PTPα to focal adhesions and endoplasmic reticulum (ER) fractions, as well as tyrosine phosphorylation of the ER Ca2+ release channel IP3R. In response to IL-1, catalytically active PTPα was required for Ca2+ release from the ER, Src-dependent phosphorylation of IP3R1 and accumulation of IP3R1 in focal adhesions. In pulldown assays and immunoprecipitations PTPα was required for the association of PTPα with IP3R1 and c-Src, and this association was increased by IL-1. Collectively, these data indicate that PTPα acts as an adaptor to mediate functional links between focal adhesions and the ER that enable IL-1-induced Ca2+ signaling.The interleukin-1 (IL-1)3 family of pro-inflammatory cytokines mediates host responses to infection and injury. Impaired control of IL-1 signaling leads to chronic inflammation and destruction of extracellular matrices (1, 2), as seen in pathological conditions such as pulmonary fibrosis (3), rheumatoid arthritis (4, 5), and periodontitis (6). IL-1 elicits multiple signaling programs, some of which trigger Ca2+ release from the endoplasmic reticulum (ER) as well as expression of multiple cytokines and inflammatory factors including c-Fos and c-Jun (7, 8), and matrix metalloproteinases (9, 10), which mediate extracellular matrix degradation via mitogen-activated protein kinase-regulated pathways (11).In anchorage-dependent cells including fibroblasts and chondrocytes, focal adhesions (FAs) are required for IL-1-induced Ca2+ release from the ER and activation of ERK (1214). FAs are actin-enriched adhesive domains composed of numerous (>50) scaffolding and signaling proteins (1517). Many FA proteins are tyrosine-phosphorylated, including paxillin, focal adhesion kinase, and src family kinases, all of which are crucial for the assembly and disassembly of FAs (1821). Protein-tyrosine phosphorylation plays a central role in regulating many cellular processes including adhesion (22, 23), motility (24), survival (25), and signal transduction (2629). Phosphorylation of proteins by kinases is balanced by protein-tyrosine phosphatases (PTP), which can enhance or attenuate downstream signaling by dephosphorylation of tyrosine residues (3032).PTPs can be divided into two main categories: receptor-like and intracellular PTPs (33). Two receptor-like PTPs have been localized to FA (leukocyte common antigen-related molecule and PTPα). Leukocyte common antigen-related molecule can dephosphorylate and mediate degradation of p130cas, which ultimately leads to cell death (34, 35). PTPα contains a heavily glycosylated extracellular domain, a transmembrane domain, and two intracellular phosphatase domains (33, 36). The amino-terminal domain predominantly mediates catalytic activity, whereas the carboxyl-terminal domain serves a regulatory function (37, 38). PTPα is enriched in FA (23) and is instrumental in regulating FA dynamics (39) via activation of c-Src/Fyn kinases by dephosphorylating the inhibitory carboxyl tyrosine residue, namely Tyr529 (22, 4042) and facilitation of integrin-dependent assembly of Src-FAK and Fyn-FAK complexes that regulate cell motility (43). Although PTPα has been implicated in formation and remodeling of FAs (44, 45), the role of PTPα in FA-dependent signaling is not defined.Ca2+ release from the ER is a critical step in integrin-dependent IL-1 signal transduction and is required for downstream activation of ERK (13, 46). The release of Ca2+ from the ER depends on the inositol 1,4,5-triphosphate receptor (IP3R), which is an IP3-gated Ca2+ channel (47). All of the IP3R subtypes (subtypes 1–3) have been localized to the ER, as well as other the plasma membrane and other endomembranes (4850). Further, IP3R may associate with FAs, enabling the anchorage of the ER to FAs (51, 52). However, the molecule(s) that provide the structural link for this association has not been defined.FA-restricted, IL-1-triggered signal transduction in anchorage-dependent cells may rely on interacting proteins that are enriched in FAs and the ER (53). Here, we examined the possibility that PTPα associates with c-Src and IP3R to functionally link FAs to the ER, thereby enabling IL-1 signal transduction.  相似文献   

7.
8.
9.
A central question in Wnt signaling is the regulation of β-catenin phosphorylation and degradation. Multiple kinases, including CKIα and GSK3, are involved in β-catenin phosphorylation. Protein phosphatases such as PP2A and PP1 have been implicated in the regulation of β-catenin. However, which phosphatase dephosphorylates β-catenin in vivo and how the specificity of β-catenin dephosphorylation is regulated are not clear. In this study, we show that PP2A regulates β-catenin phosphorylation and degradation in vivo. We demonstrate that PP2A is required for Wnt/β-catenin signaling in Drosophila. Moreover, we have identified PR55α as the regulatory subunit of PP2A that controls β-catenin phosphorylation and degradation. PR55α, but not the catalytic subunit, PP2Ac, directly interacts with β-catenin. RNA interference knockdown of PR55α elevates β-catenin phosphorylation and decreases Wnt signaling, whereas overexpressing PR55α enhances Wnt signaling. Taken together, our results suggest that PR55α specifically regulates PP2A-mediated β-catenin dephosphorylation and plays an essential role in Wnt signaling.Wnt/β-catenin signaling plays essential roles in development and tumorigenesis (13). Our previous work found that β-catenin is sequentially phosphorylated by CKIα4 and GSK3 (4), which creates a binding site for β-Trcp (5), leading to degradation via the ubiquitination/proteasome machinery (3). Mutations in β-catenin or APC genes that prevent β-catenin phosphorylation or ubiquitination/degradation lead ultimately to cancer (1, 2).In addition to the involvement of kinases, protein phosphatases, such as PP1, PP2A, and PP2C, are also implicated in Wnt/β-catenin regulation. PP2C and PP1 may regulate dephosphorylation of Axin and play positive roles in Wnt signaling (6, 7). PP2A is a multisubunit enzyme (810); it has been reported to play either positive or negative roles in Wnt signaling likely by targeting different components (1121). Toward the goal of understanding the mechanism of β-catenin phosphorylation, we carried out siRNA screening targeting several major phosphatases, in which we found that PP2A dephosphorylates β-catenin. This is consistent with a recent study where PP2A is shown to dephosphorylate β-catenin in a cell-free system (18).PP2A consists of a catalytic subunit (PP2Ac), a structure subunit (PR65/A), and variable regulatory B subunits (PR/B, PR/B′, PR/B″, or PR/B‴). The substrate specificity of PP2A is thought to be determined by its B subunit (9). By siRNA screening, we further identified that PR55α, a regulatory subunit of PP2A, specifically regulates β-catenin phosphorylation and degradation. Mechanistically, we found that PR55α directly interacts with β-catenin and regulates PP2A-mediated β-catenin dephosphorylation in Wnt signaling.  相似文献   

10.
Intracellular signaling systems of G protein-coupled receptors are well established, but their role in paracrine regulation of adjacent cells is generally considered as a tissue-specific mechanism. We have shown previously that AT1 receptor (AT1R) stimulation leads to diacylglycerol lipase-mediated transactivation of co-expressed CB1Rs in Chinese hamster ovary cells. In the present study we detected a paracrine effect of the endocannabinoid release from Chinese hamster ovary, COS7, and HEK293 cells during the stimulation of AT1 angiotensin receptors by determining CB1 cannabinoid receptor activity with bioluminescence resonance energy transfer-based sensors of G protein activation expressed in separate cells. The angiotensin II-induced, paracrine activation of CB1 receptors was visualized by detecting translocation of green fluorescent protein-tagged β-arrestin2. Mass spectrometry analyses have demonstrated angiotensin II-induced stimulation of 2-arachidonoylglycerol production, whereas no increase of anandamide levels was observed. Stimulation of Gq/11-coupled M1, M3, M5 muscarinic, V1 vasopressin, α1a adrenergic, B2 bradykinin receptors, but not Gi/o-coupled M2 and M4 muscarinic receptors, also led to paracrine transactivation of CB1 receptors. These data suggest that, in addition to their retrograde neurotransmitter role, endocannabinoids have much broader paracrine mediator functions during activation of Gq/11-coupled receptors.Hormones, neurotransmitters, and other chemical mediators acting on G protein-coupled receptors (GPCRs)2 exert their effects on the target cells by stimulating G protein-dependent and independent intracellular signaling pathways (14). Activation of Gq/11 protein-coupled receptors causes phospholipase C activation, which produces inositol-trisphosphate and diacylglycerol from phosphatidylinositol (4,5)-bisphosphate, leading to Ca2+-signal generation and protein kinase C activation. However, the concerted response of tissues to chemical mediators frequently also involves the activation of cells adjacent to the target cells, due to the release of paracrine mediators. A well known example is NO, which can be released from activated endothelial cells to cause relaxation of adjacent vascular smooth muscle cells. Lipid mediators can also act as intercellular messengers. For example, endocannabinoids released from postsynaptic neurons after depolarization act as retrograde transmitters by binding to and stimulating presynaptic cannabinoid receptors, which leads to inhibition of γ-aminobutyric acid release (an event termed depolarization-induced suppression of inhibition, DSI) (57).Cannabinoid receptors were first identified based on their ability to selectively recognize marijuana analogs. To date, two cannabinoid receptors have been identified by molecular cloning, CB1 and CB2 receptors (CB1R and CB2R, respectively) (5, 8, 9), although additional GPCRs have also been proposed to function as cannabinoid receptors (10, 11). Cannabinoid receptors also recognize certain lipids present in animal tissues termed endocannabinoids, such as arachidonylethanolamide (anandamide), 2-arachidonoylglycerol (2-AG), and 2-arachidonoylglyceryl ether (noladin ether) (7, 1216). In adult and fetal neural tissues, the two major endocannabinoids, anandamide and 2-AG, are produced on demand, usually after depolarization of postsynaptic cells or following stimulation of Gq-coupled metabotropic glutamate or muscarinic acetylcholine receptors (7, 12, 1720). Enzymes responsible for 2-AG production and metabolism in tissues are localized to well defined structures at synapses, near the axon terminals of CB1R-expressing cells (5, 7). In contrast, in peripheral tissues baseline levels of endocannabinoid production usually manifest as “endocannabinoid tone,” with poorly understood localization of the various components of the endocannabinoid system. 2-AG levels in brain homogenates and in many peripheral tissues are near its Kd for the CB1R (19), suggesting that function of endocannabinoids may not be limited to localized synaptic signaling.There is mounting evidence that endocannabinoids play important roles in peripheral cardiovascular, inflammatory, intestinal, and metabolic regulation (2124). 2-AG is produced by diacylglycerol-lipase (DAGL) after cleavage of the fatty-acid in the sn-1 position of diacylglycerol (DAG) (19, 25). Phospholipase C activation by Gq/11 protein-coupled receptors produces DAG, which can serve as a substrate for DAGL. Plasma membrane phosphoinositides are enriched in arachidonic acid in the sn-2 position (26), and DAGL is expressed ubiquitously (27), which suggests that phospholipase C-mediated cleavage of polyphosphoinositides may routinely lead to the formation of 2-AG. In accordance with this hypothesis, we have recently shown that angiotensin II- (Ang II)-mediated activation of the Gq/11-coupled AT1 angiotensin receptor (AT1R) leads to DAGL-dependent activation of CB1Rs expressed in Chinese hamster ovary (CHO) cells (28).Here our aim has been to examine the possibility that 2-AG serves as a common paracrine signal generated via activation of Gq/11 protein-coupled, Ca2+-mobilizing receptors. Accordingly, we co-expressed CB1Rs and BRET-based sensors of G protein activation in CHO cells, and used these cells to detect endocannabinoid release from adjacent cells that express AT1R or other Ca2+-mobilizing GPCRs. We have further shown that activation of AT1R by Ang II increases 2-AG levels in CHO cells. These findings suggest that 2-AG is commonly released following activation of Ca2+-mobilizing GPCRs and serves as a paracrine signal to activate CB1R in neighboring cells.  相似文献   

11.
12.
13.
14.
The present study tests the hypothesis that the structure of extracellular domain Loop 2 can markedly affect ethanol sensitivity in glycine receptors (GlyRs) and γ-aminobutyric acid type A receptors (GABAARs). To test this, we mutated Loop 2 in the α1 subunit of GlyRs and in the γ subunit of α1β2γ2GABAARs and measured the sensitivity of wild type and mutant receptors expressed in Xenopus oocytes to agonist, ethanol, and other agents using two-electrode voltage clamp. Replacing Loop 2 of α1GlyR subunits with Loop 2 from the δGABAAR (δL2), but not the γGABAAR subunit, reduced ethanol threshold and increased the degree of ethanol potentiation without altering general receptor function. Similarly, replacing Loop 2 of the γ subunit of GABAARs with δL2 shifted the ethanol threshold from 50 mm in WT to 1 mm in the GABAA γ-δL2 mutant. These findings indicate that the structure of Loop 2 can profoundly affect ethanol sensitivity in GlyRs and GABAARs. The δL2 mutations did not affect GlyR or GABAAR sensitivity, respectively, to Zn2+ or diazepam, which suggests that these δL2-induced changes in ethanol sensitivity do not extend to all allosteric modulators and may be specific for ethanol or ethanol-like agents. To explore molecular mechanisms underlying these results, we threaded the WT and δL2 GlyR sequences onto the x-ray structure of the bacterial Gloeobacter violaceus pentameric ligand-gated ion channel homologue (GLIC). In addition to being the first GlyR model threaded on GLIC, the juxtaposition of the two structures led to a possible mechanistic explanation for the effects of ethanol on GlyR-based on changes in Loop 2 structure.Alcohol abuse and dependence are significant problems in our society, with ∼14 million people in the United States being affected (1, 2). Alcohol causes over 100,000 deaths in the United States, and alcohol-related issues are estimated to cost nearly 200 billion dollars annually (2). To address this, considerable attention has focused on the development of medications to prevent and treat alcohol-related problems (35). The development of such medications would be aided by a clear understanding of the molecular structures on which ethanol acts and how these structures influence receptor sensitivity to ethanol.Ligand-gated ion channels (LGICs)2 have received substantial attention as putative sites of ethanol action that cause its behavioral effects (612). Research in this area has focused on investigating the effects of ethanol on two large superfamilies of LGICs: 1) the Cys-loop superfamily of LGICs (13, 14), whose members include nicotinic acetylcholine, 5-hydroxytryptamine3, γ-aminobutyric acid type A (GABAA), γ-aminobutyric acid type C, and glycine receptors (GlyRs) (10, 11, 1520) and 2) the glutamate superfamily, including N-methyl d-aspartate, α-amino-3-hydroxyisoxazolepropionic acid, and kainate receptors (21, 22). Recent studies have also begun investigating ethanol action in the ATP-gated P2X superfamily of LGICs (2325).A series of studies that employed chimeric and mutagenic strategies combined with sulfhydryl-specific labeling identified key regions within Cys-loop receptors that appear to be initial targets for ethanol action that also can determine the sensitivity of the receptors to ethanol (712, 18, 19, 2630). This work provides several lines of evidence that position 267 and possibly other sites in the transmembrane (TM) domain of GlyRs and homologous sites in GABAARs are targets for ethanol action and that mutations at these sites can influence ethanol sensitivity (8, 9, 26, 31).Growing evidence from GlyRs indicates that ethanol also acts on the extracellular domain. The initial findings came from studies demonstrating that α1GlyRs are more sensitive to ethanol than are α2GlyRs despite the high (∼78%) sequence homology between α1GlyRs and α2GlyRs (32). Further work found that an alanine to serine exchange at position 52 (A52S) in Loop 2 can eliminate the difference in ethanol sensitivity between α1GlyRs and α2GlyRs (18, 20, 33). These studies also demonstrated that mutations at position 52 in α1GlyRS and the homologous position 59 in α2GlyRs controlled the sensitivity of these receptors to a novel mechanistic ethanol antagonist (20). Collectively, these studies suggest that there are multiple sites of ethanol action in α1GlyRs, with one site located in the TM domain (e.g. position 267) and another in the extracellular domain (e.g. position 52).Subsequent studies revealed that the polarity of the residue at position 52 plays a key role in determining the sensitivity of GlyRs to ethanol (20). The findings with polarity in the extracellular domain contrast with the findings at position 267 in the TM domain, where molecular volume, but not polarity, significantly affected ethanol sensitivity (9). Taken together, these findings indicate that the physical-chemical parameters of residues at positions in the extracellular and TM domains that modulate ethanol effects and/or initiate ethanol action in GlyRs are not uniform. Thus, knowledge regarding the physical-chemical properties that control agonist and ethanol sensitivity is key for understanding the relationship between the structure and the actions of ethanol in LGICs (19, 31, 3440).GlyRs and GABAARs, which differ significantly in their sensitivities to ethanol, offer a potential method for identifying the structures that control ethanol sensitivity. For example, α1GlyRs do not reliably respond to ethanol concentrations less than 10 mm (32, 33, 41). Similarly, γ subunit-containing GABAARs (e.g. α1β2γ2), the most predominantly expressed GABAARs in the central nervous system, are insensitive to ethanol concentrations less than 50 mm (42, 43). In contrast, δ subunit-containing GABAARs (e.g. α4β3δ) have been shown to be sensitive to ethanol concentrations as low as 1–3 mm (4451). Sequence alignment of α1GlyR, γGABAAR, and δGABAAR revealed differences between the Loop 2 regions of these receptor subunits. Since prior studies found that mutations of Loop 2 residues can affect ethanol sensitivity (19, 20, 39), the non-conserved residues in Loop 2 of GlyR and GABAAR subunits could provide the physical-chemical and structural bases underlying the differences in ethanol sensitivity between these receptors.The present study tested the hypothesis that the structure of Loop 2 can markedly affect the ethanol sensitivity of GlyRs and GABAARs. To accomplish this, we performed multiple mutations that replaced the Loop 2 region of the α1 subunit in α1GlyRs and the Loop 2 region of the γ subunit of α1β2γ2 GABAARs with corresponding non-conserved residues from the δ subunit of GABAAR and tested the sensitivity of these receptors to ethanol. As predicted, replacing Loop 2 of WT α1GlyRs with the homologous residues from the δGABAAR subunit (δL2), but not the γGABAAR subunit (γL2), markedly increased the sensitivity of the receptor to ethanol. Similarly, replacing the non-conserved residues of the γ subunit of α1β2γ2 GABAARs with δL2 also markedly increased ethanol sensitivity of GABAARs. These findings support the hypothesis and suggest that Loop 2 may play a role in controlling ethanol sensitivity across the Cys-loop superfamily of receptors. The findings also provide the basis for suggesting structure-function relationships in a new molecular model of the GlyR based on the bacterial Gloeobacter violaceus pentameric LGIC homologue (GLIC).  相似文献   

15.
16.
17.
18.
Human cytomegalovirus (HCMV) infection can be fatal to immunocompromised individuals. We have previously reported that gamma interferon and tumor necrosis factor alpha (TNF-α) synergistically inhibit HCMV replication in vitro. Ceramides have been described as second messengers induced by TNF-α. To investigate the mechanisms involved in the inhibition of HCMV by TNF-α, in the present study we have analyzed ceramide production by U373 MG astrocytoma cells and the effects of TNF-α versus ceramides on HCMV replication. Our results show that U373 MG cells did not produce ceramides upon incubation with TNF-α. Moreover, long-chain ceramides induced by treatment with exogenous bacterial sphingomyelinase inhibited HCMV replication in synergy with TNF-α. Surprisingly, short-chain permeant C6-ceramide increased viral replication. Our results show that the anti-HCMV activity of TNF-α is independent of ceramides. In addition, our results suggest that TNF-α and endogenous long-chain ceramides use separate pathways of cell signalling to inhibit HCMV replication, while permeant C6-ceramide appears to activate a third pathway leading to an opposite effect.Human cytomegalovirus (HCMV) infections are well controlled in the immunocompetent host. Cellular immune responses (CD4+ and CD8+ T cells and NK cells) which accompany both acute and latent infections (for a review, see reference 4) are thought to be the main components of this control. HCMV infection during immunosuppression such as in cancer, transplantations, or AIDS results in severe pathology (4). We have previously shown that tumor necrosis factor alpha (TNF-α), in synergy with gamma interferon (IFN-γ), inhibits the replication of HCMV (7). In mice, TNF-α is involved in the clearance of CMV infection (25). TNF-α is a cytokine with multiple effects which is produced by many cell types, including macrophages and CD8+ and CD4+ T lymphocytes (for a review, see reference 40), and is known to possess antiviral effects (20, 47). The molecular mechanisms involved in the signalling by TNF-α depend on the type of receptor, p55 (TNF-R1) or p75 (TNF-R2) (5), to which it binds. Some cells express only one type of TNF-α receptor; however, expression of these receptors is not always mutually exclusive (5). The cytotoxicity of TNF-α has been reported to be mediated by TNF-R1 (38), whose intracellular region carries a death domain which signals for programmed cell death (39). Signalling through TNF-R1 with specific antibodies can also protect Hep-G2 cells from vesicular stomatitis virus-mediated cytopathic effects (48). Ceramide production after TNF-α treatment has been widely reported (16, 19, 31) and has also been shown to depend on signalling through the TNF-R1 receptor (45). In these experiments concerning myeloid cells, TNF-α induced the activation of a sphingomyelinase, which cleaved sphingomyelin to release ceramide and phosphocholine. The production of ceramides can lead to cell apoptosis (11, 14, 23) or cell cycle arrest (13). Induction of apoptosis by TNF-α has been mimicked by exogenous sphingomyelinase and by synthetic, short-chain, permeant ceramides, which suggests that ceramides, as second messengers, are sufficient to induce the cytotoxic effects of TNF-α (11, 23). Acidic and neutral sphingomyelinases activated in different cell compartments may be responsible for the diverse effects of TNF-α (46), with the former being involved in signalling through NF-κB (34) and the latter being involved in signalling through a ceramide-activated protein kinase and phospholipase A2 (46).One of the characteristics of HCMV infection is the increase in the content of intracellular DNA, reported to be of viral (3, 8, 18) or cellular (12, 37) origin. Since TNF-α has been known to display antiproliferative properties and to block cells in the G1 phase (29), we initially tested its effects on the cell cycle of infected cells. Then, based on studies reporting that TNF-α induces ceramides in cells (16, 19, 31) and on a study showing the role of ceramide in cell cycle blockade (13), we originally postulated that ceramide was responsible for the antiviral effect of TNF-α. In the present study, we used astrocytoma cells (U373 MG) as a model for brain cells, which are important targets of HCMV in vivo (22). In contrast to fibroblasts, infected U373 MG cells release smaller quantities of virus particles even though all the cells were infected in our experiments. We believe that the U373 MG model is closer to HCMV infection in vivo. We show that ceramides are not produced by U373 MG cells upon incubation with even high concentrations of TNF-α. In addition, we demonstrate that exogenously added sphingomyelinase induces anti-HCMV effects whereas permeant C6-ceramide increases HCMV proliferation in U373 MG cells. This suggests that lipid second messengers can modulate HCMV infection and that TNF-α and ceramides use distinct signalling pathways in the control of HCMV infection. This is supported by our observation that the protein kinase JNK1 is activated exclusively by TNF-α in U373 MG cells and that TNF-α and exogenous sphingomyelinase act in synergy on HCMV infection.  相似文献   

19.
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