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 62 毫秒
1.
TrkA receptor signaling is essential for nerve growth factor (NGF)-induced survival and differentiation of sensory neurons. To identify possible effectors or regulators of TrkA signaling, yeast two-hybrid screening was performed using the intracellular domain of TrkA as bait. We identified muc18-1-interacting protein 2 (Mint2) as a novel TrkA-binding protein and found that the phosphotyrosine binding domain of Mint2 interacted with TrkA in a phosphorylation- and ligand-independent fashion. Coimmunoprecipitation assays showed that endogenous TrkA interacted with Mint2 in rat tissue homogenates, and immunohistochemical evidence revealed that Mint2 and TrkA colocalized in rat dorsal root ganglion neurons. Furthermore, Mint2 overexpression inhibited NGF-induced neurite outgrowth in both PC12 and cultured dorsal root ganglion neurons, whereas inhibition of Mint2 expression by RNA interference facilitated NGF-induced neurite outgrowth. Moreover, Mint2 was found to promote the retention of TrkA in the Golgi apparatus and inhibit its surface sorting. Taken together, our data provide evidence that Mint2 is a novel TrkA-regulating protein that affects NGF-induced neurite outgrowth, possibly through a mechanism involving retention of TrkA in the Golgi apparatus.The neurotrophin family member nerve growth factor (NGF)3 is essential for proper development, patterning, and maintenance of nervous systems (1, 2). NGF has two known receptors; TrkA, a single-pass transmembrane receptor-tyrosine kinase that binds selectively to NGF, and p75, a transmembrane glycoprotein that binds all members of the neurotrophin family (3, 4). NGF binding activates the kinase domain of TrkA, leading to autophosphorylation (5). The resulting phosphotyrosines become docking sites for adaptor proteins involved in signal transduction pathways that lead to the activation of Ras, Rac, phosphatidylinositol 3-kinase, phospholipase Cγ, and other effectors (2, 6). Many of these TrkA-interacting adaptor proteins have been identified and include, Grb2, APS, SH2B, fibroblast growth factor receptor substrate 2 (FRS-2), Shc, and human tumor imaginal disc 1 (TID1) (7-10). The identification of these binding partners has contributed greatly to our understanding of the mechanisms underlying the functional diversity of NGF-TrkA signaling.Studies have indicated that the transmission of NGF signaling in neurons involves retrograde transport of NGF-TrkA complexes from the neurite tip to the cell body (11-14). TrkA associates with components of cytoplasmic dynein, and it is thought that vesicular trafficking of neurotrophins occurs via direct interaction of Trk receptors with the dynein motor machinery (14). Furthermore, the atypical protein kinase C-interacting protein, p62, associates with TrkA and plays a novel role in connecting receptor signals with the endosomal signaling network required for mediating TrkA-induced differentiation (15). Recently, the membrane-trafficking protein Pincher has been found to mediate macroendocytosis underlying retrograde signaling by TrkA (16). Despite the progress made to date in understanding Trk complex internalization and trafficking, the mechanisms remain poorly understood.Mint2 (muc18-1-interacting protein 2) belongs to the Mint protein family, which consists of three members, Mint1, Mint2, and Mint3. Mint proteins were first identified as interacting proteins of the synaptic vesicle-docking protein Munc18-1 (17, 18). Mint1 is also sometimes referred to as mLIN-10, as it is the mammalian orthologue of the Caenorhabditis elegans LIN-10 (19). Additionally, Mint1, Mint2, and Mint3 are also referred to as X11α or X11, X11β or X11L (X11-like), and X11γ or X11L2 (X11-like 2), respectively (20). All Mint proteins contain a conserved central phosphotyrosine binding (PTB) domain and two contiguous C-terminal PDZ domains (repeated sequences in the brain-specific protein PSD-95, the Drosophila septate junction protein Discs large, and the epithelial tight junction protein ZO-1) (17, 18, 21). Mint1 and Mint2 are expressed only in neuronal tissue (17), whereas Mint3 is ubiquitously expressed (18). Although the function of Mints proteins is not fully clear, their interactions with the docking and exocytosis factors Mun18 -1 and CASK, ADP-ribosylation factor (Arf) GTPases involved in vesicle budding (22), and other synaptic adaptor proteins, such as neurabin-II/spinophilin (23), tamalin (24), and kalirin-7 (25), all suggest possible roles for Mints in synaptic vesicle docking and exocytosis. Mint proteins have also been implicated in the trafficking and/or processing of β-amyloid precursor protein (β-APP). Through their PTB domains, all three Mints bind to a motif within the cytoplasmic domain of β-APP (21, 26-29), and Mint1 and Mint2 can stabilize β-APP, affect β-APP processing, and inhibit the production and secretion of Aβ (28, 30-32). Although the mechanisms by which Mints inhibit β-APP processing are not yet well known, Mints and their binding partners have emerged as potential therapeutic targets for the treatment of Alzheimer disease.To uncover new TrkA-interacting factors and gain insight into the mechanisms that guide TrkA intracellular trafficking and other aspects of TrkA signaling, we conducted a yeast two-hybrid screen of a brain cDNA library using the intracellular domain of TrkA as bait. The screen identified several candidate TrkA-interacting proteins, one of which was Mint2. Follow-up binding assays showed that the PTB domain of Mint2 alone was necessary and sufficient for mediating the interaction with TrkA. Endogenous Mint2 was also coimmunoprecipitated and colocalized with TrkA in rat DRG tissue. Overexpression and knockdown studies showed that Mint2 could significantly inhibit NGF-induced neurite outgrowth in both TrkA-expressing PC12 cells and DRG neurons. Moreover, Mint2 was found to induce the retention of TrkA in the Golgi apparatus and inhibit its surface sorting. Our results suggest that Mint2 is a novel regulator of TrkA receptor signaling.  相似文献   

2.
3.
Wnt11 signals through both canonical (β-catenin) and non-canonical pathways and is up-regulated during osteoblast differentiation and fracture healing. In these studies, we evaluated the role of Wnt11 during osteoblastogenesis. Wnt11 overexpression in MC3T3E1 pre-osteoblasts increases β-catenin accumulation and promotes bone morphogenetic protein (BMP)-induced expression of alkaline phosphatase and mineralization. Wnt11 dramatically increases expression of the osteoblast-associated genes Dmp1 (dentin matrix protein 1), Phex (phosphate-regulating endopeptidase homolog), and Bsp (bone sialoprotein). Wnt11 also increases expression of Rspo2 (R-spondin 2), a secreted factor known to enhance Wnt signaling. Overexpression of Rspo2 is sufficient for increasing Dmp1, Phex, and Bsp expression and promotes bone morphogenetic protein-induced mineralization. Knockdown of Rspo2 abrogates Wnt11-mediated osteoblast maturation. Antagonism of T-cell factor (Tcf)/β-catenin signaling with dominant negative Tcf blocks Wnt11-mediated expression of Dmp1, Phex, and Rspo2 and decreases mineralization. However, dominant negative Tcf fails to block the osteogenic effects of Rspo2 overexpression. These studies show that Wnt11 signals through β-catenin, activating Rspo2 expression, which is then required for Wnt11-mediated osteoblast maturation.Wnt signaling is a key regulator of osteoblast differentiation and maturation. In mesenchymal stem cell lines, canonical Wnt signaling by Wnt10b enhances osteoblast differentiation (1). Canonical Wnt signaling through β-catenin has also been shown to enhance the chondroinductive and osteoinductive properties of BMP22 (2, 3). During BMP2-induced osteoblast differentiation of mesenchymal stem cell lines, cross-talk between BMP and Wnt pathways converges through the interaction of Smad4 with β-catenin (2).Canonical Wnt signaling is also critical for skeletal development and homeostasis. During limb development, expression of Wnt3a in the apical ectodermal ridge of limb buds maintains cells in a highly proliferative and undifferentiated state (4, 5). Disruption of canonical Wnt signaling in Lrp5/Lrp6 compound knock-out mice results in limb- and digit-patterning defects (6). Wnt signaling is also involved in the maintenance of post-natal bone mass. Gain of function in the Wnt co-receptor Lrp5 leads to increased bone mass, whereas loss of Lrp5 function is associated with decreased bone mass and osteoporosis pseudoglioma syndrome (7, 8). Mice with increased Wnt10b expression have increased trabecular bone, whereas Wnt10b-deficient mice have reduced trabecular bone (9). Similarly, mice nullizygous for the Wnt antagonist sFrp1 have increased trabecular bone accrual throughout adulthood (10).Although canonical Wnt signaling regulates osteoblastogenesis and bone formation, the profile of endogenous Wnts that play a role in osteoblast differentiation and maturation is not well described. During development, Wnt11 is expressed in the perichondrium and in the axial skeleton and sternum (11). Wnt11 expression is increased during glucocorticoid-induced osteogenesis (12), indicating a potential role for Wnt11 in osteoblast differentiation. Interestingly, Wnt11 activates both β-catenin-dependent as well as β-catenin-independent signaling pathways (13). Targeted disruption of Wnt11 results in late embryonic/early post-natal death because of cardiac dysfunction (14). Although these mice have no reported skeletal developmental abnormalities, early lethality obfuscates a detailed examination of post-natal skeletal modeling and remodeling.In murine development, Wnt11 expression overlaps with the expression of R-spondin 2 (Rspo2) in the apical ectodermal ridge (11, 15). R-spondins are a novel family of proteins that share structural features, including two conserved cysteinerich furin-like domains and a thrombospondin type I repeat (16). The four R-spondin family members can activate canonical Wnt signaling (15, 1719). Rspo3 interacts with Frizzled 8 and Lrp6 and enhances Wnt ligand signaling. Rspo1 enhances Wnt signaling by interacting with Lrp6 and inhibiting Dkk-mediated receptor internalization (20). Rspo1 was also shown to potentiate Wnt3a-mediated osteoblast differentiation (21). Rspo2 knock-out mice, which die at birth, have limb patterning defects associated with altered β-catenin signaling (2224). However, the role of Rspo2 in osteoblast differentiation and maturation remains unclear.Herein we report that Wnt11 overexpression in MC3T3E1 pre-osteoblasts activates β-catenin and augments BMP-induced osteoblast maturation and mineralization. Wnt11 increases the expression of Rspo2. Overexpression of Rspo2 in MC3T3E1 is sufficient for augmenting BMP-induced osteoblast maturation and mineralization. Although antagonism of Tcf/β-catenin signaling blocks the osteogenic effects of Wnt11, Rspo2 rescues this block, and knockdown of Rspo2 shows that it is required for Wnt11-mediated osteoblast maturation and mineralization. These studies identify both Wnt11 and Rspo2 as novel mediators of osteoblast maturation and mineralization.  相似文献   

4.
Many G protein-coupled receptors (GPCRs) recycle after agonist-induced endocytosis by a sequence-dependent mechanism, which is distinct from default membrane flow and remains poorly understood. Efficient recycling of the β2-adrenergic receptor (β2AR) requires a C-terminal PDZ (PSD-95/Discs Large/ZO-1) protein-binding determinant (PDZbd), an intact actin cytoskeleton, and is regulated by the endosomal protein Hrs (hepatocyte growth factor-regulated substrate). The PDZbd is thought to link receptors to actin through a series of protein interaction modules present in NHERF/EBP50 (Na+/H+ exchanger 3 regulatory factor/ezrin-binding phosphoprotein of 50 kDa) family and ERM (ezrin/radixin/moesin) family proteins. It is not known, however, if such actin connectivity is sufficient to recapitulate the natural features of sequence-dependent recycling. We addressed this question using a receptor fusion approach based on the sufficiency of the PDZbd to promote recycling when fused to a distinct GPCR, the δ-opioid receptor, which normally recycles inefficiently in HEK293 cells. Modular domains mediating actin connectivity promoted receptor recycling with similarly high efficiency as the PDZbd itself, and recycling promoted by all of the domains was actin-dependent. Regulation of receptor recycling by Hrs, however, was conferred only by the PDZbd and not by downstream interaction modules. These results suggest that actin connectivity is sufficient to mimic the core recycling activity of a GPCR-linked PDZbd but not its cellular regulation.G protein-coupled receptors (GPCRs)2 comprise the largest family of transmembrane signaling receptors expressed in animals and transduce a wide variety of physiological and pharmacological information. While these receptors share a common 7-transmembrane-spanning topology, structural differences between individual GPCR family members confer diverse functional and regulatory properties (1-4). A fundamental mechanism of GPCR regulation involves agonist-induced endocytosis of receptors via clathrin-coated pits (4). Regulated endocytosis can have multiple functional consequences, which are determined in part by the specificity with which internalized receptors traffic via divergent downstream membrane pathways (5-7).Trafficking of internalized GPCRs to lysosomes, a major pathway traversed by the δ-opioid receptor (δOR), contributes to proteolytic down-regulation of receptor number and produces a prolonged attenuation of subsequent cellular responsiveness to agonist (8, 9). Trafficking of internalized GPCRs via a rapid recycling pathway, a major route traversed by the β2-adrenergic receptor (β2AR), restores the complement of functional receptors present on the cell surface and promotes rapid recovery of cellular signaling responsiveness (6, 10, 11). When co-expressed in the same cells, the δOR and β2AR are efficiently sorted between these divergent downstream membrane pathways, highlighting the occurrence of specific molecular sorting of GPCRs after endocytosis (12).Recycling of various integral membrane proteins can occur by default, essentially by bulk membrane flow in the absence of lysosomal sorting determinants (13). There is increasing evidence that various GPCRs, such as the β2AR, require distinct cytoplasmic determinants to recycle efficiently (14). In addition to requiring a cytoplasmic sorting determinant, sequence-dependent recycling of the β2AR differs from default recycling in its dependence on an intact actin cytoskeleton and its regulation by the conserved endosomal sorting protein Hrs (hepatocyte growth factor receptor substrate) (11, 14). Compared with the present knowledge regarding protein complexes that mediate sorting of GPCRs to lysosomes (15, 16), however, relatively little is known about the biochemical basis of sequence-directed recycling or its regulation.The β2AR-derived recycling sequence conforms to a canonical PDZ (PSD-95/Discs Large/ZO-1) protein-binding determinant (henceforth called PDZbd), and PDZ-mediated protein association(s) with this sequence appear to be primarily responsible for its endocytic sorting activity (17-20). Fusion of this sequence to the cytoplasmic tail of the δOR effectively re-routes endocytic trafficking of engineered receptors from lysosomal to recycling pathways, establishing the sufficiency of the PDZbd to function as a transplantable sorting determinant (18). The β2AR-derived PDZbd binds with relatively high specificity to the NHERF/EBP50 family of PDZ proteins (21, 22). A well-established biochemical function of NHERF/EBP50 family proteins is to associate integral membrane proteins with actin-associated cytoskeletal elements. This is achieved through a series of protein-interaction modules linking NHERF/EBP50 family proteins to ERM (ezrin-radixin-moesin) family proteins and, in turn, to actin filaments (23-26). Such indirect actin connectivity is known to mediate other effects on plasma membrane organization and function (23), however, and NHERF/EBP50 family proteins can bind to additional proteins potentially important for endocytic trafficking of receptors (23, 25). Thus it remains unclear if actin connectivity is itself sufficient to promote sequence-directed recycling of GPCRs and, if so, if such connectivity recapitulates the normal cellular regulation of sequence-dependent recycling. In the present study, we took advantage of the modular nature of protein connectivity proposed to mediate β2AR recycling (24, 26), and extended the opioid receptor fusion strategy used successfully for identifying diverse recycling sequences in GPCRs (27-29), to address these fundamental questions.Here we show that the recycling activity of the β2AR-derived PDZbd can be effectively bypassed by linking receptors to ERM family proteins in the absence of the PDZbd itself. Further, we establish that the protein connectivity network can be further simplified by fusing receptors to an interaction module that binds directly to actin filaments. We found that bypassing the PDZ-mediated interaction using either domain is sufficient to mimic the ability of the PDZbd to promote efficient, actin-dependent recycling of receptors. Hrs-dependent regulation, however, which is characteristic of sequence-dependent recycling of wild-type receptors, was recapitulated only by the fused PDZbd and not by the proposed downstream interaction modules. These results support a relatively simple architecture of protein connectivity that is sufficient to mimic the core recycling activity of the β2AR-derived PDZbd, but not its characteristic cellular regulation. Given that an increasing number of GPCRs have been shown to bind PDZ proteins that typically link directly or indirectly to cytoskeletal elements (17, 27, 30-32), the present results also suggest that actin connectivity may represent a common biochemical principle underlying sequence-dependent recycling of various GPCRs.  相似文献   

5.
Galectins have been implicated in T cell homeostasis playing complementary pro-apoptotic roles. Here we show that galectin-8 (Gal-8) is a potent pro-apoptotic agent in Jurkat T cells inducing a complex phospholipase D/phosphatidic acid signaling pathway that has not been reported for any galectin before. Gal-8 increases phosphatidic signaling, which enhances the activity of both ERK1/2 and type 4 phosphodiesterases (PDE4), with a subsequent decrease in basal protein kinase A activity. Strikingly, rolipram inhibition of PDE4 decreases ERK1/2 activity. Thus Gal-8-induced PDE4 activation releases a negative influence of cAMP/protein kinase A on ERK1/2. The resulting strong ERK1/2 activation leads to expression of the death factor Fas ligand and caspase-mediated apoptosis. Several conditions that decrease ERK1/2 activity also decrease apoptosis, such as anti-Fas ligand blocking antibodies. In addition, experiments with freshly isolated human peripheral blood mononuclear cells, previously stimulated with anti-CD3 and anti-CD28, show that Gal-8 is pro-apoptotic on activated T cells, most likely on a subpopulation of them. Anti-Gal-8 autoantibodies from patients with systemic lupus erythematosus block the apoptotic effect of Gal-8. These results implicate Gal-8 as a novel T cell suppressive factor, which can be counterbalanced by function-blocking autoantibodies in autoimmunity.Glycan-binding proteins of the galectin family have been increasingly studied as regulators of the immune response and potential therapeutic agents for autoimmune disorders (1). To date, 15 galectins have been identified and classified according with the structural organization of their distinctive monomeric or dimeric carbohydrate recognition domain for β-galactosides (2, 3). Galectins are secreted by unconventional mechanisms and once outside the cells bind to and cross-link multiple glycoconjugates both at the cell surface and at the extracellular matrix, modulating processes as diverse as cell adhesion, migration, proliferation, differentiation, and apoptosis (410). Several galectins have been involved in T cell homeostasis because of their capability to kill thymocytes, activated T cells, and T cell lines (1116). Pro-apoptotic galectins might contribute to shape the T cell repertoire in the thymus by negative selection, restrict the immune response by eliminating activated T cells at the periphery (1), and help cancer cells to escape the immune system by eliminating cancer-infiltrating T cells (17). They have also a promising therapeutic potential to eliminate abnormally activated T cells and inflammatory cells (1). Studies on the mostly explored galectins, Gal-1, -3, and -9 (14, 15, 1820), as well as in Gal-2 (13), suggest immunosuppressive complementary roles inducing different pathways to apoptosis. Galectin-8 (Gal-8)4 is one of the most widely expressed galectins in human tissues (21, 22) and cancerous cells (23, 24). Depending on the cell context and mode of presentation, either as soluble stimulus or extracellular matrix, Gal-8 can promote cell adhesion, spreading, growth, and apoptosis (6, 7, 9, 10, 22, 25). Its role has been mostly studied in relation to tumor malignancy (23, 24). However, there is some evidence regarding a role for Gal-8 in T cell homeostasis and autoimmune or inflammatory disorders. For instance, the intrathymic expression and pro-apoptotic effect of Gal-8 upon CD4highCD8high thymocytes suggest a role for Gal-8 in shaping the T cell repertoire (16). Gal-8 could also modulate the inflammatory function of neutrophils (26), Moreover Gal-8-blocking agents have been detected in chronic autoimmune disorders (10, 27, 28). In rheumatoid arthritis, Gal-8 has an anti-inflammatory action, promoting apoptosis of synovial fluid cells, but can be counteracted by a specific rheumatoid version of CD44 (CD44vRA) (27). In systemic lupus erythematosus (SLE), a prototypic autoimmune disease, we recently described function-blocking autoantibodies against Gal-8 (10, 28). Thus it is important to define the role of Gal-8 and the influence of anti-Gal-8 autoantibodies in immune cells.In Jurkat T cells, we previously reported that Gal-8 interacts with specific integrins, such as α1β1, α3β1, and α5β1 but not α4β1, and as a matrix protein promotes cell adhesion and asymmetric spreading through activation of the extracellular signal-regulated kinases 1 and 2 (ERK1/2) (10). These early effects occur within 5–30 min. However, ERK1/2 signaling supports long term processes such as T cell survival or death, depending on the moment of the immune response. During T cell activation, ERK1/2 contributes to enhance the expression of interleukin-2 (IL-2) required for T cell clonal expansion (29). It also supports T cell survival against pro-apoptotic Fas ligand (FasL) produced by themselves and by other previously activated T cells (30, 31). Later on, ERK1/2 is required for activation-induced cell death, which controls the extension of the immune response by eliminating recently activated and restimulated T cells (32, 33). In activation-induced cell death, ERK1/2 signaling contributes to enhance the expression of FasL and its receptor Fas/CD95 (32, 33), which constitute a preponderant pro-apoptotic system in T cells (34). Here, we ask whether Gal-8 is able to modulate the intensity of ERK1/2 signaling enough to participate in long term processes involved in T cell homeostasis.The functional integration of ERK1/2 and PKA signaling (35) deserves special attention. cAMP/PKA signaling plays an immunosuppressive role in T cells (36) and is altered in SLE (37). Phosphodiesterases (PDEs) that degrade cAMP release the immunosuppressive action of cAMP/PKA during T cell activation (38, 39). PKA has been described to control the activity of ERK1/2 either positively or negatively in different cells and processes (35). A little explored integration among ERK1/2 and PKA occurs via phosphatidic acid (PA) and PDE signaling. Several stimuli activate phospholipase D (PLD) that hydrolyzes phosphatidylcholine into PA and choline. Such PLD-generated PA plays roles in signaling interacting with a variety of targeting proteins that bear PA-binding domains (40). In this way PA recruits Raf-1 to the plasma membrane (41). It is also converted by phosphatidic acid phosphohydrolase (PAP) activity into diacylglycerol (DAG), which among other functions, recruits and activates the GTPase Ras (42). Both Ras and Raf-1 are upstream elements of the ERK1/2 activation pathway (43). In addition, PA binds to and activates PDEs of the type 4 subfamily (PDE4s) leading to decreased cAMP levels and PKA down-regulation (44). The regulation and role of PA-mediated control of ERK1/2 and PKA remain relatively unknown in T cell homeostasis, because it is also unknown whether galectins stimulate the PLD/PA pathway.Here we found that Gal-8 induces apoptosis in Jurkat T cells by triggering cross-talk between PKA and ERK1/2 pathways mediated by PLD-generated PA. Our results for the first time show that a galectin increases the PA levels, down-regulates the cAMP/PKA system by enhancing rolipram-sensitive PDE activity, and induces an ERK1/2-dependent expression of the pro-apoptotic factor FasL. The enhanced PDE activity induced by Gal-8 is required for the activation of ERK1/2 that finally leads to apoptosis. Gal-8 also induces apoptosis in human peripheral blood mononuclear cells (PBMC), especially after activating T cells with anti-CD3/CD28. Therefore, Gal-8 shares with other galectins the property of killing activated T cells contributing to the T cell homeostasis. The pathway involves a particularly integrated signaling context, engaging PLD/PA, cAMP/PKA, and ERK1/2, which so far has not been reported for galectins. The pro-apoptotic function of Gal-8 also seems to be unique in its susceptibility to inhibition by anti-Gal-8 autoantibodies.  相似文献   

6.
The Notch receptor is critical for proper development where it orchestrates numerous cell fate decisions. The Fringe family of β1,3-N-acetylglucosaminyltransferases are regulators of this pathway. Fringe enzymes add N-acetylglucosamine to O-linked fucose on the epidermal growth factor repeats of Notch. Here we have analyzed the reaction catalyzed by Lunatic Fringe (Lfng) in detail. A mutagenesis strategy for Lfng was guided by a multiple sequence alignment of Fringe proteins and solutions from docking an epidermal growth factor-like O-fucose acceptor substrate onto a homology model of Lfng. We targeted three main areas as follows: residues that could help resolve where the fucose binds, residues in two conserved loops not observed in the published structure of Manic Fringe, and residues predicted to be involved in UDP-N-acetylglucosamine (UDP-GlcNAc) donor specificity. We utilized a kinetic analysis of mutant enzyme activity toward the small molecule acceptor substrate 4-nitrophenyl-α-l-fucopyranoside to judge their effect on Lfng activity. Our results support the positioning of O-fucose in a specific orientation to the catalytic residue. We also found evidence that one loop closes off the active site coincident with, or subsequent to, substrate binding. We propose a mechanism whereby the ordering of this short loop may alter the conformation of the catalytic aspartate. Finally, we identify several residues near the UDP-GlcNAc-binding site, which are specifically permissive toward UDP-GlcNAc utilization.Defects in Notch signaling have been implicated in numerous human diseases, including multiple sclerosis (1), several forms of cancer (2-4), cerebral autosomal dominant arteriopathy with sub-cortical infarcts and leukoencephalopathy (5), and spondylocostal dysostosis (SCD)3 (6-8). The transmembrane Notch signaling receptor is activated by members of the DSL (Delta, Serrate, Lag2) family of ligands (9, 10). In the endoplasmic reticulum, O-linked fucose glycans are added to the epidermal growth factor-like (EGF) repeats of the Notch extracellular domain by protein O-fucosyltransferase 1 (11-13). These O-fucose monosaccharides can be elongated in the Golgi apparatus by three highly conserved β1,3-N-acetylglucosaminyltransferases of the Fringe family (Lunatic (Lfng), Manic (Mfng), and Radical Fringe (Rfng) in mammals) (14-16). The formation of this GlcNAc-β1,3-Fuc-α1, O-serine/threonine disaccharide is necessary and sufficient for subsequent elongation to a tetrasaccharide (15, 19), although elongation past the disaccharide in Drosophila is not yet clear (20, 21). Elongation of O-fucose by Fringe is known to potentiate Notch signaling from Delta ligands and inhibit signaling from Serrate ligands (22). Delta ligands are termed Delta-like (Delta-like1, -2, and -4) in mammals, and the homologs of Serrate are known as Jagged (Jagged1 and -2) in mammals. The effects of Fringe on Drosophila Notch can be recapitulated in Notch ligand in vitro binding assays using purified components, suggesting that the elongation of O-fucose by Fringe alters the binding of Notch to its ligands (21). Although Fringe also appears to alter Notch-ligand interactions in mammals, the effects of elongation of the glycan past the O-fucose monosaccharide is more complicated and appears to be cell type-, receptor-, and ligand-dependent (for a recent review see Ref. 23).The Fringe enzymes catalyze the transfer of GlcNAc from the donor substrate UDP-α-GlcNAc to the acceptor fucose, forming the GlcNAc-β1,3-Fuc disaccharide (14-16). They belong to the GT-A-fold of inverting glycosyltransferases, which includes N-acetylglucosaminyltransferase I and β1,4-galactosyltransferase I (17, 18). The mechanism is presumed to proceed through the abstraction of a proton from the acceptor substrate by a catalytic base (Asp or Glu) in the active site. This creates a nucleophile that attacks the anomeric carbon of the nucleotide-sugar donor, inverting its configuration from α (on the nucleotide sugar) to β (in the product) (24, 25). The enzyme then releases the acceptor substrate modified with a disaccharide and UDP. The Mfng structure (26) leaves little doubt as to the identity of the catalytic residue, which in all likelihood is aspartate 289 in mouse Lfng (we will use numbering for mouse Lunatic Fringe throughout, unless otherwise stated). The structure of Mfng with UDP-GlcNAc soaked into the crystals (26) showed density only for the UDP portion of the nucleotide-sugar donor and no density for two loops flanking either side of the active site. The presence of flexible loops that become ordered upon substrate binding is a common observation with glycosyltransferases in the GT-A fold family (18, 25). Density for the entire donor was observed in the structure of rabbit N-acetylglucosaminyltransferase I (27). In this case, ordering of a previously disordered loop upon UDP-GlcNAc binding may have contributed to increased stability of the donor. In the case of bovine β1,4-galactosyltransferase I, a section of flexible random coil from the apo-structure was observed to change its conformation to α-helical upon donor substrate binding (28). Both loops in Lfng are highly conserved, and we have mutated a number of residues in each to test the hypothesis that they interact with the substrates. The mutagenesis strategy was also guided by docking of an EGF-O-fucose acceptor substrate into the active site of the Lfng model as well as comparison of the Lfng model with a homology model of the β1,3-glucosyltransferase (β3GlcT) that modifies O-fucose on thrombospondin type 1 repeats (29, 30). The β3GlcT is predicted to be a GT-A fold enzyme related to the Fringe family (17, 18, 29).  相似文献   

7.
8.
It has recently been shown that the epithelial Na+ channel (ENaC) is compartmentalized in caveolin-rich lipid rafts and that pharmacological depletion of membrane cholesterol, which disrupts lipid raft formation, decreases the activity of ENaC. Here we show, for the first time, that a signature protein of caveolae, caveolin-1 (Cav-1), down-regulates the activity and membrane surface expression of ENaC. Physical interaction between ENaC and Cav-1 was also confirmed in a coimmunoprecipitation assay. We found that the effect of Cav-1 on ENaC requires the activity of Nedd4-2, a ubiquitin protein ligase of the Nedd4 family, which is known to induce ubiquitination and internalization of ENaC. The effect of Cav-1 on ENaC requires the proline-rich motifs at the C termini of the β- and γ-subunits of ENaC, the binding motifs that mediate interaction with Nedd4-2. Taken together, our data suggest that Cav-1 inhibits the activity of ENaC by decreasing expression of ENaC at the cell membrane via a mechanism that involves the promotion of Nedd4-2-dependent internalization of the channel.Amiloride-sensitive epithelial Na+ channels (ENaC)3 are membrane proteins that are expressed in salt-absorptive epithelia, including the distal collecting tubules of the kidney, the mucosa of the distal colon, the respiratory epithelium, and the excretory ducts of sweat and salivary glands (14). Na+ absorption via ENaC is critical to the normal regulation of Na+ and fluid homeostasis and is important for maintaining blood pressure (5) and the volume of fluid in the respiratory passages (6). Increased ENaC activity has been implicated in the salt-sensitive inherited form of hypertension, Liddle''s syndrome (7), and dehydration of the surface of the airway epithelium in the pathology associated with cystic fibrosis lung disease (8).Expression of ENaC at the cell membrane surface is regulated by the E3 ubiquitin protein ligase, Nedd4-2 (neural precursor cell expressed developmentally down-regulated protein 4) (9). Interaction between the WW domains of Nedd4-2 and the proline-rich PY motifs (PPPXY) on ENaC is essential for Nedd4-2 to exert a negative effect on the channel (10, 11). This interaction leads to ubiquitination-dependent internalization of ENaC (12, 13). Several regulators of ENaC exert their effects on the channel by modulating the action of Nedd4-2. For instance, serum and glucocorticoid-dependent protein kinase (14), protein kinase B (15), and G protein-coupled receptor kinase (16) up-regulate activity of ENaC by inhibiting Nedd4-2. Although the details of cellular mechanisms that underlie internalization of ENaC remain to be elucidated, the physiological significance of Nedd4-dependent internalization of the channel has been well established. For instance, heritable mutations that delete the cytosolic termini of the β-or γ-subunit of ENaC, which contain the proline-rich motifs, are known to cause hyperactivity of ENaC in the kidney (17) and increase cell surface expression of the channel (7, 18).The plasma membranes of most cell types contain lipid raft microdomains that are enriched with glycosphingolipid and cholesterol (19), that have distinctive biophysical properties, and that selectively include or exclude signaling molecules (20). These microdomains promote clustering of an array of integral membrane proteins in the membrane leaflets (21) and may be important for organizing cascades of signaling molecules (22, 23). Processes in which raft microdomains are involved include the intracellular transport of proteins and lipids to the cell membrane (24), the endocytotic retrieval of membrane proteins (25, 26), and signal transduction (27, 28). In addition, segregation of signaling molecules within lipid rafts may facilitate cross-talk between signal transduction pathways (29), a phenomenon that may be important in ensuring rapid and efficient integration of multiple cellular signaling events (30, 31). Of particular interest is the subpopulation of lipid rafts enriched with caveolin proteins. Caveolin-1 (Cav-1), a major caveolin isoform expressed in nonmuscle cells, has been identified as being involved in diverse cellular functions, such as vesicular transport, cholesterol homeostasis, and signal transduction (32). Cav-1 also regulates the activity and membrane expression of ion channels and transporters (28).In epithelia, the majority of lipid rafts exist at the apical membrane surface (22). Pools of ENaC (3336) and several proteins that regulate activity of ENaC, such as Nedd4 (37), protein kinase B (38), protein kinase C (39), Go (40), and the G protein-coupled receptor kinase (41), have been identified in detergent-insoluble and cholesterol-rich membrane fractions from a variety of cell types, consistent with localization of these proteins in lipid rafts. Furthermore, detergent-free buoyant density separation of lipid rafts has revealed the presence of Cav-1 with ENaC in the lipid raft-rich membrane fraction (35). The physiological role of lipid rafts in the regulation of ENaC has been the subject of many recent investigations. Most of these studies used a pharmacological agent, methyl-β-cyclodextrin (MβCD), to promote redistribution of proteins away from the cholesterol-enriched membrane domains. The results were, however, inconclusive. In some studies, MβCD treatment was found to inhibit open probability (42) or cell surface expression of ENaC (35), whereas others found no direct effect of MβCD on the channel (33, 43).Despite a number of studies into the role of lipid rafts on the regulation of ENaC, little is known about the physiological relevance of caveolins to the function of this ion channel. In the present study, we use gene interference and gene expression techniques to determine the role of Cav-1 in the regulation of ENaC activity. We provide evidence of the association of Cav-1 with ENaC and evidence that Cav-1 negatively regulates both activity and abundance of ENaC at the surface of epithelial cells. Importantly, we demonstrate, for the first time, that the mechanism by which Cav-1 regulates activity of ENaC involves the E3 ubiquitin protein ligase, Nedd4-2.  相似文献   

9.
The intestinal epithelium is repetitively deformed by shear, peristalsis, and villous motility. Such repetitive deformation stimulates the proliferation of intestinal epithelial cells on collagen or laminin substrates via ERK, but the upstream mediators of this effect are poorly understood. We hypothesized that the phosphatidylinositol 3-kinase (PI3K)/AKT cascade mediates this mitogenic effect. PI3K, AKT, and glycogen synthase kinase-3β (GSK-3β) were phosphorylated by 10 cycles/min strain at an average 10% deformation, and pharmacologic blockade of these molecules or reduction by small interfering RNA (siRNA) prevented the mitogenic effect of strain in Caco-2 or IEC-6 intestinal epithelial cells. Strain MAPK activation required PI3K but not AKT. AKT isoform-specific siRNA transfection demonstrated that AKT2 but not AKT1 is required for GSK-3β phosphorylation and the strain mitogenic effect. Furthermore, overexpression of AKT1 or an AKT chimera including the PH domain and hinge region of AKT2 and the catalytic domain and C-tail of AKT1 prevented strain activation of GSK-3β, but overexpression of AKT2 or a chimera including the PH domain and hinge region of AKT1 and the catalytic domain and C-tail of AKT2 did not. These data delineate a role for PI3K, AKT2, and GSK-3β in the mitogenic effect of strain. PI3K is required for both ERK and AKT2 activation, whereas AKT2 is sequentially required for GSK-3β. Furthermore, AKT2 specificity requires its catalytic domain and tail region. Manipulating this pathway may prevent mucosal atrophy and maintain the mucosal barrier in conditions such as ileus, sepsis, and prolonged fasting when peristalsis and villous motility are decreased and the mucosal barrier fails.Mechanical forces are part of the normal intestinal epithelial environment. Numerous different forces deform these cells including shear stress from endoluminal chyme, bowel peristalsis, and villous motility (1, 2). During normal bowel function the mucosa is subjected to injury that must be repaired to maintain the mucosal barrier (3, 4). Deformation patterns of the bowel are altered in conditions such as prolonged fasting, post-surgical ileus, and sepsis states, resulting in profoundly reduced mucosal deformation. When such states are prolonged, proliferation slows, the mucosa becomes atrophic, and bacterial translocation may ensue as the mucosal barrier of the gut breaks down (57).In vitro, repetitive deformation is trophic for intestinal epithelial cells (8) cultured on type I or type IV collagen or laminin. Human Caco-2 intestinal epithelial cells (9), non-transformed rat IEC-6 intestinal epithelial cells (10), and primary human intestinal epithelial cells isolated from surgical specimens (11) proliferate more rapidly in response to cyclic strain (12) unless substantial quantities of fibronectin are added to the media or matrix (11) to mimic the acute phase reaction of acute or chronic inflammation and injury. Cyclic strain also stimulates proliferation in HCT 116 colon cancer cells (13) and differentiation of Caco-2 cells cultured on a collagen substrate (9). This phenomenon has also been observed in vivo (14). Thus, repetitive deformation may help to maintain the normal homeostasis of the gut mucosa under non-inflammatory conditions. Previous work in our laboratory has implicated Src, focal adhesion kinase, and the mitogen-activated protein kinase (MAPK)2 extracellular signal-related kinase (ERK) in the mitogenic effect of strain (10). Although p38 is also activated in Caco-2 cells subjected to cyclic strain on a collagen matrix, its activity is not required for the mitogenic effect of strain (12).Although often the PI3K/AKT pathway is thought of as a parallel pathway to the MAPK, this is not always the case. Protein kinase C isoenzymes differentially modulate thrombin effect on MAPK-dependent retinal pigment epithelial cell (RPE) proliferation, and it has been shown that PI3K or AKT inhibition prevented thrombin-induced ERK activation and RPE proliferation (15).PI3K, AKT, and glycogen synthase kinase (GSK), a downstream target of AKT (16), have been implemented in intestinal epithelial cell proliferation in numerous cell systems not involving strain (1719) including uncontrolled proliferation in gastrointestinal cancers (2022). Mechanical forces activate this pathway as well. PI3K and AKT are required for increased extracellular pressure to stimulate colon cancer cell adhesion (23), although the pathway by which pressure stimulates colon cancer cells in suspension differs from the response of adherent intestinal epithelial cells to repetitive deformation (24), and GSK is not involved in this effect.3 Repetitive strain also stimulates vascular endothelial cell proliferation via PI3K and AKT (25, 26), whereas respiratory strain stimulates angiogenic responses via PI3K (27). We, therefore, hypothesized that the PI3K/AKT/GSK axis would be involved in the mitogenic effects of repetitive deformation on a collagen matrix.To test this hypothesis, we used the Flexcell apparatus to rhythmically deform Caco-2 intestinal epithelial cells. IEC-6 cells were used to confirm key results. A frequency of 10 cycles per min was used, which is similar in order of magnitude to the frequency that the intestinal mucosa might be deformed by peristalsis or villous motility in vivo (28, 29). Mechanical forces such as repetitive deformation are likely cell-type and frequency-specific, as different cell types respond to different frequencies. Vascular endothelial cells respond to frequencies of 60–80 cycles/min (25), whereas intestinal epithelial cells may actually decrease proliferation in response to frequencies of 5 cycles/min (30). We characterized PI3K, AKT, and GSK phosphorylation with strain, blocked these molecules pharmacologically or by siRNA, and delineated the specificity of the AKT effect using isozyme-specific siRNA and transfection of AKT1/2 chimeras. We also characterized the interaction of this pathway with the activation of ERK by strain, which has previously been implicated in the mitogenic response (12).  相似文献   

10.
11.
The secreted trefoil factor family 2 (TFF2) protein contributes to the protection of the gastrointestinal mucosa from injury by strengthening and stabilizing mucin gels, stimulating epithelial restitution, and restraining the associated inflammation. Although trefoil factors have been shown to activate signaling pathways, no cell surface receptor has been directly linked to trefoil peptide signaling. Here we demonstrate the ability of TFF2 peptide to activate signaling via the CXCR4 chemokine receptor in cancer cell lines. We found that both mouse and human TFF2 proteins (at ∼0.5 μm) activate Ca2+ signaling in lymphoblastic Jurkat cells that could be abrogated by receptor desensitization (with SDF-1α) or pretreatment with the specific antagonist AMD3100 or an anti-CXCR4 antibody. TFF2 pretreatment of Jurkat cells decreased Ca2+ rise and chemotactic response to SDF-1α. In addition, the CXCR4-negative gastric epithelial cell line AGS became highly responsive to TFF2 treatment upon expression of the CXCR4 receptor. TFF2-induced activation of mitogen-activated protein kinases in gastric and pancreatic cancer cells, KATO III and AsPC-1, respectively, was also dependent on the presence of the CXCR4 receptor. Finally we demonstrate a distinct proliferative effect of TFF2 protein on an AGS gastric cancer cell line that expresses CXCR4. Overall these data identify CXCR4 as a bona fide signaling receptor for TFF2 and suggest a mechanism through which TFF2 may modulate immune and tumorigenic responses in vivo.Trefoil factor 2 (TFF2),2 previously known as spasmolytic polypeptide, is a unique member of the trefoil family that is expressed primarily in gastric mucous neck cells and is up-regulated in the setting of chronic inflammation. Experimental induction of ulceration in the rat stomach leads to rapid up-regulation of TFF2 expression with high levels observed 30 min after ulceration with persistence for up to 10 days (1). TFF2 is secreted into the mucus layer of the gastrointestinal tract of mammals where it stabilizes the mucin gel layer and stimulates migration of epithelial cells (24), suggesting an important role in restitution and in maintenance of the integrity of the gut. Exogenous administration of recombinant TFF2, either orally or intravenously, provides mucosal protection in several rodent models of acute gastric or intestinal injury (5, 6). A TFF2-/- knock-out mouse model has confirmed the importance of TFF2 in the protection of gastrointestinal mucosa against chronic injury (7).It is widely accepted that trefoil factors exert their biological action through a cell surface receptor. This suggestion comes from studies on binding of 125I-labeled TFF2 that demonstrated specific binding sites in the gastric glands, intestine, and colon that could be displaced by non-radioactive TFF2 (6, 810). Structural studies have revealed potential binding sites for receptors for all members of the trefoil factor family (11, 12). In concordance with this hypothesis, several membrane proteins were found to interact with TFF2. First it was shown that recombinant human TFF2 (and TFF3) could bind to a 28-kDa peptide from membrane fractions of rat jejunum and two human adenocarcinoma cell lines, MCF-7 and Colony-29 (13). Later it was found that recombinant TFF3 fused with biotin selectively bound with a 50-kDa protein from the membrane of rat small intestinal cells (14). However, these 28- and 50-kDa proteins were characterized only by their molecular size without further identification. Two TFF2-binding proteins that have been characterized include a 140-kDa protein, the β subunit of the fibronectin receptor, and a 224-kDa protein called muclin (15). Another TFF2-binding protein was isolated by probing two-dimensional blots of mouse stomach with a murine TFF2 fusion protein, leading to the identification of the gastric foveolar protein blottin, a murine homolog of the human peptide TFIZ1(16). Although these three proteins have now been well characterized, none of them has been shown to mediate responses to TFF2, and no activated signaling cascades have been shown.Despite the absence of an identified cell surface receptor for TFF2, there is nevertheless clear evidence that TFF2 and TFF3 rapidly activate signal transduction pathways (17, 18). TFF3 prevents cell death via activation of the serine/threonine kinase AKT in colon cancer cell lines (19). The TFF3 protein also activates STAT3 signaling in human colorectal cancer cells, thus providing cells with invasion potential (20). TFF3 treatment leads to EGF receptor activation and β-catenin phosphorylation in HT-29 cells (21) and to transient phosphorylation of ERK1/2 in oral keratinocytes (22). With respect to TFF2, recombinant peptide enhances the migration of human bronchial epithelial cell line BEAS-2B (4). TFF2 has been shown to induce phosphorylation of c-Jun NH2-terminal kinase (JNK) and ERK1/2. Consistent with this observation, the motogenic effect of TFF2 is significantly inhibited by antagonists of ERK kinases and protein kinase C but not by inhibitors of p38 mitogen-activated protein kinase (MAPK). It is believed that the motogenic effect of trefoil factors and of TFF2 in particular, could contribute to in vivo restitution of gastric epithelium by enhancing cell migration.Although previous studies have suggested that TFF2 functions primarily in cytoprotection, accumulating evidence now suggests that TFF2 may also play a role in the regulation of host immunity. For example, recombinant TFF2 reduces inflammation in rat and mouse models of colitis (23, 24). In addition, TFF2 was detected in rat lymphoid tissues (spleen, lymph nodes, and bone marrow) (25). Recently we and others found TFF2 mRNA expression in primary and secondary lymphopoietic organs (26, 27). These data suggest that TFF2 may play some function in the immune system. In concordance with these findings, we detected an exacerbated inflammatory response to acute injury in TFF2 knock-out animals (27, 28). These observations prompted us to look at the possible function of TFF2 in immune cells. Unexpectedly we found that TFF2 modulates Ca2+ and AKT signaling in lymphoblastic Jurkat cells and that these effects appear to be mediated through the CXCR4 receptor.  相似文献   

12.
The RIG-I-like receptors (RLRs), RIG-I and MDA5, recognize single-stranded RNA with 5′ triphosphates and double-stranded RNA (dsRNA) to initiate innate antiviral immune responses. LGP2, a homolog of RIG-I and MDA5 that lacks signaling capability, regulates the signaling of the RLRs. To establish the structural basis of dsRNA recognition by the RLRs, we have determined the 2.0-Å resolution crystal structure of human LGP2 C-terminal domain bound to an 8-bp dsRNA. Two LGP2 C-terminal domain molecules bind to the termini of dsRNA with minimal contacts between the protein molecules. Gel filtration chromatography and analytical ultracentrifugation demonstrated that LGP2 binds blunt-ended dsRNA of different lengths, forming complexes with 2:1 stoichiometry. dsRNA with protruding termini bind LGP2 and RIG-I weakly and do not stimulate the activation of RIG-I efficiently in cells. Surprisingly, full-length LGP2 containing mutations that abolish dsRNA binding retained the ability to inhibit RIG-I signaling.The innate immune response is the first line of defense against invading pathogens; it is the ubiquitous system of defense against microbial infections (1). Toll-like receptors (TLRs)3 and RIG-I (retinoic acid-inducible gene 1)-like receptors (RLRs) play key roles in innate immune response toward viral infection (2-5). Toll-like receptors TLR3, TLR7, and TLR8 sense viral RNA released in the endosome following phagocytosis of the pathogens (6). RIG-I-like receptors RIG-I and MDA5 detect viral RNA from replicating viruses in infected cells (3, 7, 8). Stimulation of these receptors leads to the induction of type I interferons (IFNs) and other proinflammatory cytokines, conferring antiviral activity to the host cells and activating the acquired immune responses (4, 9).RIG-I discriminates between viral and host RNA through specific recognition of the uncapped 5′-triphosphate of single-stranded RNA (5′ ppp ssRNA) generated by viral RNA polymerases (10, 11). In addition, RIG-I also recognizes double-stranded RNA generated during RNA virus replication (7, 12). Transfection of cells with synthetic double-stranded RNA stimulates the activation of RIG-I (13, 14). Synthetic dsRNA mimics, such as polyinosinic-polycytidylic acid (poly(I·C)), can activate MDA5 when introduced into the cytoplasm of cells. Digestion of poly(I·C) with RNase III transforms poly(I·C) from a ligand for MDA5 into a ligand for RIG-I, suggesting that MDA5 recognizes long dsRNA, whereas RIG-I recognizes short dsRNA (15). Studies of RIG-I and MDA5 knock-out mice confirmed the essential roles of these receptors in antiviral immune responses and demonstrated that they sense different sets of RNA viruses (12, 16).RIG-I and MDA5 contain two caspase recruiting domains (CARDs) at their N termini, a DEX(D/H) box RNA helicase domain, and a C-terminal regulatory or repressor domain (CTD). The helicase domain and the CTD are responsible for viral RNA binding, whereas the CARDs are required for signaling (3, 8). The current model of RIG-I activation suggests that under resting conditions RIG-I is in a suppressed conformation, and viral RNA binding triggers a conformation change that leads to the exposure of the CARDs for the recruitment of the downstream protein IPS-1 (also known as MAVS, Cardif, or VISA) (14, 17). Limited proteolysis of the RIG-I·dsRNA complex showed that RIG-I residues 792-925 of the CTD are involved in dsRNA and 5′ ppp ssRNA binding (14). The CTD of RIG-I overlaps with the C terminus of the previously identified repressor domain (18). The structures of RIG-I and LGP2 (laboratory of genetics and physiology 2) CTD in isolation have been determined by x-ray crystallography and NMR spectroscopy (14, 19, 20). A large, positively charged surface on RIG-I recognizes the 5′ triphosphate group of viral ssRNA (14, 19). RNA binding studies by titrating RIG-I CTD with dsRNA and 5′ ppp ssRNA suggested that overlapping sets of residues on this charged surface are involved in RNA binding (14). Mutagenesis of several positively charged residues on this surface either reduces or disrupts RNA binding by RIG-I, and these mutations also affect the induction of IFN-β in vivo (14, 19). However, the exact nature of how the RLRs recognize viral RNA and how RNA binding activates these receptors remains to be established.LGP2 is a homolog of RIG-I and MDA5 that lacks the CARDs and thus has no signaling capability (21, 22). The expression of LGP2 is inducible by dsRNA or IFN treatment as well as virus infection (21). Overexpression of LGP2 inhibits Sendai virus and Newcastle disease virus signaling (21). When coexpressed with RIG-I, LGP2 can inhibit RIG-I signaling through the interaction of its CTD with the CARD and the helicase domain of RIG-I (18). LGP2 could suppress RIG-I signaling by three possible ways (23): 1) binding RNA with high affinity, thereby sequestering RNA ligands from RIG-I; 2) interacting directly with RIG-I to block the assembly of the signaling complex; and 3) competing with IKKi (IκB kinase ε) in the NF-κB signaling pathway for a common binding site on IPS-1. To elucidate the structural basis of dsRNA recognition by the RLRs, we have crystallized human LGP2 CTD (residues 541-678) bound to an 8-bp double-stranded RNA and determined the structure of the complex at 2.0 Å resolution. The structure revealed that LGP2 CTD binds to the termini of dsRNA. Mutagenesis and functional studies showed that dsRNA binding is likely not required for the inhibition of RIG-I signaling by LGP2.  相似文献   

13.
14.
Rap1 small GTPases interact with Rap1-GTP-interacting adaptor molecule (RIAM), a member of the MRL (Mig-10/RIAM/Lamellipodin) protein family, to promote talin-dependent integrin activation. Here, we show that MRL proteins function as scaffolds that connect the membrane targeting sequences in Ras GTPases to talin, thereby recruiting talin to the plasma membrane and activating integrins. The MRL proteins bound directly to talin via short, N-terminal sequences predicted to form amphipathic helices. RIAM-induced integrin activation required both its capacity to bind to Rap1 and to talin. Moreover, we constructed a minimized 50-residue Rap-RIAM module containing the talin binding site of RIAM joined to the membrane-targeting sequence of Rap1A. This minimized Rap-RIAM module was sufficient to target talin to the plasma membrane and to mediate integrin activation, even in the absence of Rap1 activity. We identified a short talin binding sequence in Lamellipodin (Lpd), another MRL protein; talin binding Lpd sequence joined to a Rap1 membrane-targeting sequence is sufficient to recruit talin and activate integrins. These data establish the mechanism whereby MRL proteins interact with both talin and Ras GTPases to activate integrins.Increased affinity (“activation”) of cellular integrins is central to physiological events such as cell migration, assembly of the extracellular matrix, the immune response, and hemostasis (1). Each integrin comprises a type I transmembrane α and β subunit, each of which has a large extracellular domain, a single transmembrane domain, and a cytoplasmic domain (tail). Talin binds to most integrin β cytoplasmic domains and the binding of talin to the integrin β tail initiates integrin activation (24). A small, PTB-like domain of talin mediates activation via a two-site interaction with integrin β tails (5), and this PTB domain is functionally masked in the intact talin molecule (6). A central question in integrin biology is how the talin-integrin interaction is regulated to control integrin activation; recent work has implicated Ras GTPases as critical signaling modules in this process (7).Ras proteins are small monomeric GTPases that cycle between the GTP-bound active form and the GDP-bound inactive form. Guanine nucleotide exchange factors (GEFs) promote Ras activity by exchanging bound GDP for GTP, whereas GTPase-activating proteins (GAPs)3 enhance the hydrolysis of Ras-bound GTP to GDP (for review, see Ref. 8). The Ras subfamily members Rap1A and Rap1B stimulate integrin activation (9, 10). For example, expression of constitutively active Rap1 activates integrin αMβ2 in macrophage, and inhibition of Rap1 abrogated integrin activation induced by inflammatory agonists (1113). Murine T-cells expressing constitutively active Rap1 manifest enhanced integrin dependent cell adhesion (14). In platelets, Rap1 is rapidly activated by platelet agonists (15, 16). A knock-out of Rap1B (17) or of the Rap1GEF, RasGRP2 (18), resulted in impairment of αIIbβ3-dependent platelet aggregation, highlighting the importance of Rap1 in platelet aggregation in vivo. Thus, Rap1 GTPases play important roles in the activation of several integrins in multiple biological contexts.Several Rap1 effectors have been implicated in integrin activation (1921). Rap1-GTP-interacting adaptor molecule (RIAM) is a Rap1 effector that is a member of the MRL (Mig-10/RIAM/Lamellipodin) family of adaptor proteins (20). RIAM contains Ras association (RA) and pleckstrin homology (PH) domains and proline-rich regions, which are defining features of the MRL protein family. In Jurkat cells, RIAM overexpression induces β1 and β2 integrin-mediated cell adhesion, and RIAM knockdown abolishes Rap1-dependent cell adhesion (20), indicating RIAM is a downstream regulator of Rap1-dependent signaling. RIAM regulates actin dynamics as RIAM expression induces cell spreading; conversely, its depletion reduces cellular F-actin content (20). Whereas RIAM is greatly enriched in hematopoietic cells, Lamellipodin (Lpd) is a paralogue present in fibroblasts and other somatic cells (22).Recently we used forward, reverse, and synthetic genetics to engineer and order an integrin activation pathway in Chinese hamster ovary cells expressing a prototype activable integrin, platelet αIIbβ3. We found that Rap1 induced formation of an “integrin activation complex” containing RIAM and talin (23). Here, we have established the mechanism whereby Ras GTPases cooperate with MRL family proteins, RIAM and Lpd, to regulate integrin activation. We find that MRL proteins function as scaffolds that connect the membrane targeting sequences in Ras GTPases to talin, thereby recruiting talin to integrins at the plasma membrane.  相似文献   

15.
The amyloid precursor protein (APP) is implied both in cell growth and differentiation and in neurodegenerative processes in Alzheimer disease. Regulated proteolysis of APP generates biologically active fragments such as the neuroprotective secreted ectodomain sAPPα and the neurotoxic β-amyloid peptide. Furthermore, it has been suggested that the intact transmembrane APP plays a signaling role, which might be important for both normal synaptic plasticity and neuronal dysfunction in dementia. To understand APP signaling, we tracked single molecules of APP using quantum dots and quantitated APP homodimerization using fluorescence lifetime imaging microscopy for the detection of Förster resonance energy transfer in living neuroblastoma cells. Using selective labeling with synthetic fluorophores, we show that the dimerization of APP is considerably higher at the plasma membrane than in intracellular membranes. Heparan sulfate significantly contributes to the almost complete dimerization of APP at the plasma membrane. Importantly, this technique for the first time structurally defines the initiation of APP signaling by binding of a relevant physiological extracellular ligand; our results indicate APP as receptor for neuroprotective sAPPα, as sAPPα binding disrupts APP dimers, and this disruption of APP dimers by sAPPα is necessary for the protection of neuroblastoma cells against starvation-induced cell death. Only cells expressing reversibly dimerized wild-type, but not covalently dimerized mutant APP are protected by sAPPα. These findings suggest a potentially beneficial effect of increasing sAPPα production or disrupting APP dimers for neuronal survival.The amyloid precursor protein (APP)4 is known both for its important role in the development and plasticity of the nervous system (16) and for its involvement in Alzheimer disease (AD) (7, 8). Despite intensive research efforts, the initial events that lead to the prevalent sporadic, i.e. non-familial, forms of AD are still unclear. Furthermore, although a higher gene dose of APP (9) or the presence of pathological APP mutations is sufficient to induce familial AD (for review, see Ref. 10), the exact pathological mechanism that is triggered by APP is still under debate.Some fragments of APP, such as the β-amyloid peptide (Aβ), are thought to contribute to synaptic dysfunction and neurotoxicity (11, 12). On the other hand, the α-secretase-derived extracellular fragment of APP (sAPPα), which is present at lower levels in AD patients than in controls (13), has been shown to be beneficial for memory function, to possess neuroprotective properties, and to counteract the effects of Aβ (1418).Signaling by transmembrane APP may directly contribute to neurodegeneration in AD (1924); however, the signal transduction pathway for transmembrane APP remains unknown, although several potential regulatory proteins, glycosaminoglycans, and metal ions are known to bind with high affinity to APP and sAPPα (25, 26). The most common form of signal transduction for single-pass transmembrane proteins is the ligand-induced perturbation of a monomer/dimer equilibrium. Indeed, the dimerization of transmembrane APP has been implied several times in the past. Several studies have investigated the effects of presumed dimer-breaking perturbations on biological read-outs, such as the production of Aβ (27, 28), but without directly measuring the APP aggregation state, or have investigated the aggregation state of APP subdomains, often reconstituted in cell-free systems (2732). Dimerization interfaces in both the extracellular and the transmembrane domain have been suggested.In the studies investigating the aggregation state of full-length APP, most of the employed methods, such as chemical cross-linking and co-immunoprecipitation, do not lend themselves readily to a rigorous quantitative analysis of the abundance of potentially instable dimers (31, 33), whereas in other cases the use of chimeras may have influenced the dimerization potential or precluded the search for a natural stimulus (23, 34). The only previously reported direct observation of APP dimerization by Förster resonance energy transfer (FRET) microscopy uses an assay in which the FRET efficiency varies with the level of overexpression (35). Therefore, a concentration-dependent FRET component due to nonspecific stochastic encounters cannot be excluded in this study.Most importantly, as none of the published procedures permitted the selective detection of APP dimers on the surface of live cells, where they would encounter ligands, they could not differentiate between subpopulations of APP. This may be one reason why no natural ligand of APP has ever been shown to signal via modulation of its monomer/dimer equilibrium.Another elusive goal is the identity of the receptor for neuroprotective sAPPα (3639). The ligand-dependent dimerization of sAPPα in solution (40) and its origination from transmembrane APP suggest that APP might serve as receptor for sAPPα, but this binding has never been experimentally shown.  相似文献   

16.
17.
18.
19.
Recent studies have ascribed many non-pumping functions to the Na/K-ATPase. We show here that graded knockdown of cellular Na/K-ATPase α1 subunit produces a parallel decrease in both caveolin-1 and cholesterol in light fractions of LLC-PK1 cell lysates. This observation is further substantiated by imaging analyses, showing redistribution of cholesterol from the plasma membrane to intracellular compartments in the knockdown cells. Moreover, this regulation is confirmed in α1+/– mouse liver. Functionally, the knockdown-induced redistribution appears to affect the cholesterol sensing in the endoplasmic reticulum, because it activates the sterol regulatory element-binding protein pathway and increases expression of hydroxymethylglutaryl-CoA reductase and low density lipoprotein receptor in the liver. Consistently, we detect a modest increase in hepatic cholesterol as well as a reduction in the plasma cholesterol. Mechanistically, α1+/– livers show increases in cellular Src and ERK activity and redistribution of caveolin-1. Although activation of Src is not required in Na/K-ATPase-mediated regulation of cholesterol distribution, the interaction between the Na/K-ATPase and caveolin-1 is important for this regulation. Taken together, our new findings demonstrate a novel function of the Na/K-ATPase in control of the plasma membrane cholesterol distribution. Moreover, the data also suggest that the plasma membrane Na/K-ATPase-caveolin-1 interaction may represent an important sensing mechanism by which the cells regulate the sterol regulatory element-binding protein pathway.The Na/K-ATPase, also called the sodium pump, is an ion transporter that mediates active transport of Na+ and K+ across the plasma membrane by hydrolyzing ATP (1, 2). The functional sodium pump is mainly composed of α and β subunits. The α subunit is the catalytic component of the holoenzyme; it contains both the nucleotide and the cation binding sites (3). So far, four isoforms of α subunit have been discovered, and each one shows a distinct tissue distribution pattern (4, 5). Interestingly, studies during the past few years have uncovered many non-pumping functions of Na/K-ATPase (610). Recently, we have demonstrated that more than half of the Na/K-ATPase may actually perform cellular functions other than ion pumping at least in LLC-PK1 cells (11). Moreover, the non-pumping pool of Na/K-ATPase mainly resides in caveolae and interacts with a variety of proteins such as Src, inositol 1,4,5-trisphosphate receptor, and caveolin-1 (1214). While the interaction between Na/K-ATPase and inositol 1,4,5-trisphosphate receptor facilitates Ca2+ signaling (13) the dynamic association between Na/K-ATPase and Src appears to be an essential step for ouabain to stimulate cellular kinases (15). More recently, we report that the interaction between the Na/K-ATPase and caveolin-1 plays an important role for the membrane trafficking of caveolin-1. Knockdown of the Na/K-ATPase leads to altered subcellular distribution of caveolin-1 and increases the mobility of caveolin-1-containing vesicles (16).Caveolin is a protein marker for caveolae (17). Caveolae are flask-shaped vesicular invaginations of plasma membrane and are enriched in cholesterol, glycosphingolipids, and sphingomyelin (18). There are three genes and six isoforms of caveolin. Caveolin-1 is a 22-kDa protein and is expressed in many types of cells, including epithelial and endothelial cells. In addition to their role in biogenesis of caveolae (19), accumulating evidence has implicated caveolin proteins in cellular cholesterol homeostasis (20). For instance, caveolin-1 directly binds to cholesterol in a 1:1 ratio (21). It was also found to be an integral member of the intracellular cholesterol trafficking machinery between internal membranes and plasma membrane (22, 23). The expression of caveolin-1 appears to be under control of SREBPs,2 the master regulators of intracellular cholesterol level (24). Furthermore, knockout of caveolin-1 significantly affected cholesterol metabolism in mouse embryonic fibroblasts and mouse peritoneal macrophages (25). Because we found that the Na/K-ATPase regulates cellular distribution of caveolin-1, we propose that it may also affect intracellular cholesterol distribution and metabolism. To test our hypothesis, we have investigated whether sodium pump α1 knockdown affects cholesterol distribution and metabolism both in vitro and in vivo. Our results indicate that sodium pump α1 expression level plays a role in the proper distribution of intracellular cholesterol. Down-regulation of sodium pump α1 not only redistributes cholesterol between the plasma membrane and cytosolic compartments, but also alters cholesterol metabolism in mice.  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