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 46 毫秒
1.
2.
3.
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   

4.
5.
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   

6.
7.
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   

8.
9.
10.
Mitochondrial dysregulation is strongly implicated in Parkinson disease. Mutations in PTEN-induced kinase 1 (PINK1) are associated with familial parkinsonism and neuropsychiatric disorders. Although overexpressed PINK1 is neuroprotective, less is known about neuronal responses to loss of PINK1 function. We found that stable knockdown of PINK1 induced mitochondrial fragmentation and autophagy in SH-SY5Y cells, which was reversed by the reintroduction of an RNA interference (RNAi)-resistant plasmid for PINK1. Moreover, stable or transient overexpression of wild-type PINK1 increased mitochondrial interconnectivity and suppressed toxin-induced autophagy/mitophagy. Mitochondrial oxidant production played an essential role in triggering mitochondrial fragmentation and autophagy in PINK1 shRNA lines. Autophagy/mitophagy served a protective role in limiting cell death, and overexpressing Parkin further enhanced this protective mitophagic response. The dominant negative Drp1 mutant inhibited both fission and mitophagy in PINK1-deficient cells. Interestingly, RNAi knockdown of autophagy proteins Atg7 and LC3/Atg8 also decreased mitochondrial fragmentation without affecting oxidative stress, suggesting active involvement of autophagy in morphologic remodeling of mitochondria for clearance. To summarize, loss of PINK1 function elicits oxidative stress and mitochondrial turnover coordinated by the autophagic and fission/fusion machineries. Furthermore, PINK1 and Parkin may cooperate through different mechanisms to maintain mitochondrial homeostasis.Parkinson disease is an age-related neurodegenerative disease that affects ∼1% of the population worldwide. The causes of sporadic cases are unknown, although mitochondrial or oxidative toxins such as 1-methyl-4-phenylpyridinium, 6-hydroxydopamine (6-OHDA),3 and rotenone reproduce features of the disease in animal and cell culture models (1). Abnormalities in mitochondrial respiration and increased oxidative stress are observed in cells and tissues from parkinsonian patients (2, 3), which also exhibit increased mitochondrial autophagy (4). Furthermore, mutations in parkinsonian genes affect oxidative stress response pathways and mitochondrial homeostasis (5). Thus, disruption of mitochondrial homeostasis represents a major factor implicated in the pathogenesis of sporadic and inherited parkinsonian disorders (PD).The PARK6 locus involved in autosomal recessive and early-onset PD encodes for PTEN-induced kinase 1 (PINK1) (6, 7). PINK1 is a cytosolic and mitochondrially localized 581-amino acid serine/threonine kinase that possesses an N-terminal mitochondrial targeting sequence (6, 8). The primary sequence also includes a putative transmembrane domain important for orientation of the PINK1 domain (8), a conserved kinase domain homologous to calcium calmodulin kinases, and a C-terminal domain that regulates autophosphorylation activity (9, 10). Overexpression of wild-type PINK1, but not its PD-associated mutants, protects against several toxic insults in neuronal cells (6, 11, 12). Mitochondrial targeting is necessary for some (13) but not all of the neuroprotective effects of PINK1 (14), implicating involvement of cytoplasmic targets that modulate mitochondrial pathobiology (8). PINK1 catalytic activity is necessary for its neuroprotective role, because a kinase-deficient K219M substitution in the ATP binding pocket of PINK1 abrogates its ability to protect neurons (14). Although PINK1 mutations do not seem to impair mitochondrial targeting, PD-associated mutations differentially destabilize the protein, resulting in loss of neuroprotective activities (13, 15).Recent studies indicate that PINK1 and Parkin interact genetically (3, 16-18) to prevent oxidative stress (19, 20) and regulate mitochondrial morphology (21). Primary cells derived from PINK1 mutant patients exhibit mitochondrial fragmentation with disorganized cristae, recapitulated by RNA interference studies in HeLa cells (3).Mitochondria are degraded by macroautophagy, a process involving sequestration of cytoplasmic cargo into membranous autophagic vacuoles (AVs) for delivery to lysosomes (22, 23). Interestingly, mitochondrial fission accompanies autophagic neurodegeneration elicited by the PD neurotoxin 6-OHDA (24, 25). Moreover, mitochondrial fragmentation and increased autophagy are observed in neurodegenerative diseases including Alzheimer and Parkinson diseases (4, 26-28). Although inclusion of mitochondria in autophagosomes was once believed to be a random process, as observed during starvation, studies involving hypoxia, mitochondrial damage, apoptotic stimuli, or limiting amounts of aerobic substrates in facultative anaerobes support the concept of selective mitochondrial autophagy (mitophagy) (29, 30). In particular, mitochondrially localized kinases may play an important role in models involving oxidative mitochondrial injury (25, 31, 32).Autophagy is involved in the clearance of protein aggregates (33-35) and normal regulation of axonal-synaptic morphology (36). Chronic disruption of lysosomal function results in accumulation of subtly impaired mitochondria with decreased calcium buffering capacity (37), implicating an important role for autophagy in mitochondrial homeostasis (37, 38). Recently, Parkin, which complements the effects of PINK1 deficiency on mitochondrial morphology (3), was found to promote autophagy of depolarized mitochondria (39). Conversely, Beclin 1-independent autophagy/mitophagy contributes to cell death elicited by the PD toxins 1-methyl-4-phenylpyridinium and 6-OHDA (25, 28, 31, 32), causing neurite retraction in cells expressing a PD-linked mutation in leucine-rich repeat kinase 2 (40). Whereas properly regulated autophagy plays a homeostatic and neuroprotective role, excessive or incomplete autophagy creates a condition of “autophagic stress” that can contribute to neurodegeneration (28).As mitochondrial fragmentation (3) and increased mitochondrial autophagy (4) have been described in human cells or tissues of PD patients, we investigated whether or not the engineered loss of PINK1 function could recapitulate these observations in human neuronal cells (SH-SY5Y). Stable knockdown of endogenous PINK1 gave rise to mitochondrial fragmentation and increased autophagy and mitophagy, whereas stable or transient overexpression of PINK1 had the opposite effect. Autophagy/mitophagy was dependent upon increased mitochondrial oxidant production and activation of fission. The data indicate that PINK1 is important for the maintenance of mitochondrial networks, suggesting that coordinated regulation of mitochondrial dynamics and autophagy limits cell death associated with loss of PINK1 function.  相似文献   

11.
Melanoma cells express the chemokine receptor CXCR4 that confers high invasiveness upon binding to its ligand CXCL12. Melanoma cells at initial stages of the disease show reduction or loss of E-cadherin expression, but recovery of its expression is frequently found at advanced phases. We overexpressed E-cadherin in the highly invasive BRO lung metastatic cell melanoma cell line to investigate whether it could influence CXCL12-promoted cell invasion. Overexpression of E-cadherin led to defective invasion of melanoma cells across Matrigel and type I collagen in response to CXCL12. A decrease in individual cell migration directionality toward the chemokine and reduced adhesion accounted for the impaired invasion. A p190RhoGAP-dependent inhibition of RhoA activation was responsible for the impairment in chemokine-stimulated E-cadherin melanoma transfectant invasion. Furthermore, we show that p190RhoGAP and p120ctn associated predominantly on the plasma membrane of cells overexpressing E-cadherin, and that E-cadherin-bound p120ctn contributed to RhoA inactivation by favoring p190RhoGAP-RhoA association. These results suggest that melanoma cells at advanced stages of the disease could have reduced metastatic potency in response to chemotactic stimuli compared with cells lacking E-cadherin, and the results indicate that p190RhoGAP is a central molecule controlling melanoma cell invasion.Cadherins are a family of Ca2+-dependent adhesion molecules that mediate cell-cell contacts and are expressed in most solid tissues providing a tight control of morphogenesis (1, 2). Classical cadherins, such as epithelial (E) cadherin, are found in adherens junctions, forming core protein complexes with β-catenin, α-catenin, and p120 catenin (p120ctn). Both β-catenin and p120ctn directly interact with E-cadherin, whereas α-catenin associates with the complex through its binding to β-catenin, providing a link with the actin cytoskeleton (1, 2). E-cadherin is frequently lost or down-regulated in many human tumors, coincident with morphological epithelial to mesenchymal transition and acquisition of invasiveness (3-6).Although melanoma only accounts for 5% of skin cancers, when metastasis starts, it is responsible for 80% of deaths from skin cancers (7). Melanocytes express E-cadherin (8-10), but melanoma cells at early radial growth phase show a large reduction in the expression of this cadherin, and surprisingly, expression has been reported to be partially recovered by vertical growth phase and metastatic melanoma cells (9, 11, 12).Trafficking of cancer cells from primary tumor sites to intravasation into blood circulation and later to extravasation to colonize distant organs requires tightly regulated directional cues and cell migration and invasion that are mediated by chemokines, growth factors, and adhesion molecules (13). Solid tumor cells express chemokine receptors that provide guidance of these cells to organs where their chemokine ligands are expressed, constituting a homing model resembling the one used by immune cells to exert their immune surveillance functions (14). Most solid cancer cells express CXCR4, a receptor for the chemokine CXCL12 (also called SDF-1), which is expressed in lungs, bone marrow, and liver (15). Expression of CXCR4 in human melanoma has been detected in the vertical growth phase and on regional lymph nodes, which correlated with poor prognosis and increased mortality (16, 17). Previous in vivo experiments have provided evidence supporting a crucial role for CXCR4 in the metastasis of melanoma cells (18).Rho GTPases control the dynamics of the actin cytoskeleton during cell migration (19, 20). The activity of Rho GTPases is tightly regulated by guanine-nucleotide exchange factors (GEFs),4 which stimulate exchange of bound GDP by GTP, and inhibited by GTPase-activating proteins (GAPs), which promote GTP hydrolysis (21, 22), whereas guanine nucleotide dissociation inhibitors (GDIs) appear to mediate blocking of spontaneous activation (23). Therefore, cell migration is finely regulated by the balance between GEF, GAP, and GDI activities on Rho GTPases. Involvement of Rho GTPases in cancer is well documented (reviewed in Ref. 24), providing control of both cell migration and growth. RhoA and RhoC are highly expressed in colon, breast, and lung carcinoma (25, 26), whereas overexpression of RhoC in melanoma leads to enhancement of cell metastasis (27). CXCL12 activates both RhoA and Rac1 in melanoma cells, and both GTPases play key roles during invasion toward this chemokine (28, 29).Given the importance of the CXCL12-CXCR4 axis in melanoma cell invasion and metastasis, in this study we have addressed the question of whether changes in E-cadherin expression on melanoma cells might affect cell invasiveness. We show here that overexpression of E-cadherin leads to impaired melanoma cell invasion to CXCL12, and we provide mechanistic characterization accounting for the decrease in invasion.  相似文献   

12.
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   

13.
Hepatocellular carcinoma (HCC) is one of the most common and aggressive human malignancies. Recombinant tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anti-tumor agent. However, many HCC cells show resistance to TRAIL-induced apoptosis. In this study, we showed that bortezomib, a proteasome inhibitor, overcame TRAIL resistance in HCC cells, including Huh-7, Hep3B, and Sk-Hep1. The combination of bortezomib and TRAIL restored the sensitivity of HCC cells to TRAIL-induced apoptosis. Comparing the molecular change in HCC cells treated with these agents, we found that down-regulation of phospho-Akt (P-Akt) played a key role in mediating TRAIL sensitization of bortezomib. The first evidence was that bortezomib down-regulated P-Akt in a dose- and time-dependent manner in TRAIL-treated HCC cells. Second, LY294002, a PI3K inhibitor, also sensitized resistant HCC cells to TRAIL-induced apoptosis. Third, knocking down Akt1 by small interference RNA also enhanced TRAIL-induced apoptosis in Huh-7 cells. Finally, ectopic expression of mutant Akt (constitutive active) in HCC cells abolished TRAIL sensitization effect of bortezomib. Moreover, okadaic acid, a protein phosphatase 2A (PP2A) inhibitor, reversed down-regulation of P-Akt in bortezomib-treated cells, and PP2A knockdown by small interference RNA also reduced apoptosis induced by the combination of TRAIL and bortezomib, indicating that PP2A may be important in mediating the effect of bortezomib on TRAIL sensitization. Together, bortezomib overcame TRAIL resistance at clinically achievable concentrations in hepatocellular carcinoma cells, and this effect is mediated at least partly via inhibition of the PI3K/Akt pathway.Hepatocellular carcinoma (HCC)2 is currently the fifth most common solid tumor worldwide and the fourth leading cause of cancer-related death. To date, surgery is still the only curative treatment but is only feasible in a small portion of patients (1). Drug treatment is the major therapy for patients with advanced stage disease. Unfortunately, the response rate to traditional chemotherapy for HCC patients is unsatisfactory (1). Novel pharmacological therapy is urgently needed for patients with advanced HCC. In this regard, the approval of sorafenib might open a new era of molecularly targeted therapy in the treatment of HCC patients.Tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a type II transmembrane protein and a member of the TNF family, is a promising anti-tumor agent under clinical investigation (2). TRAIL functions by engaging its receptors expressed on the surface of target cells. Five receptors specific for TRAIL have been identified, including DR4/TRAIL-R1, DR5/TRAIL-R2, DcR1, DcR2, and osteoprotegerin. Among TRAIL receptors, only DR4 and DR5 contain an effective death domain that is essential to formation of death-inducing signaling complex (DISC), a critical step for TRAIL-induced apoptosis. Notably, the trimerization of the death domains recruits an adaptor molecule, Fas-associated protein with death domain (FADD), which subsequently recruits and activates caspase-8. In type I cells, activation of caspase-8 is sufficient to activate caspase-3 to induce apoptosis; however, in another type of cells (type II), the intrinsic mitochondrial pathway is essential for apoptosis characterized by cleavage of Bid and release of cytochrome c from mitochondria, which subsequently activates caspase-9 and caspase-3 (3).Although TRAIL induces apoptosis in malignant cells but sparing normal cells, some tumor cells are resistant to TRAIL-induced apoptosis. Mechanisms responsible for the resistance include receptors and intracellular resistance. Although the cell surface expression of DR4 or DR5 is absolutely required for TRAIL-induced apoptosis, tumor cells expressing these death receptors are not always sensitive to TRAIL due to intracellular mechanisms. For example, the cellular FLICE-inhibitory protein (c-FLIP), a homologue to caspase-8 but without protease activity, has been linked to TRAIL resistance in several studies (4, 5). In addition, inactivation of Bax, a proapoptotic Bcl-2 family protein, resulted in resistance to TRAIL in MMR-deficient tumors (6, 7), and reintroduction of Bax into Bax-deficient cells restored TRAIL sensitivity (8), indicating that the Bcl-2 family plays a critical role in intracellular mechanisms for resistance of TRAIL.Bortezomib, a proteasome inhibitor approved clinically for multiple myeloma and mantle cell lymphoma, has been investigated intensively for many types of cancer (9). Accumulating studies indicate that the combination of bortezomib and TRAIL overcomes the resistance to TRAIL in various types of cancer, including acute myeloid leukemia (4), lymphoma (1013), prostate (1417), colon (15, 18, 19), bladder (14, 16), renal cell carcinoma (20), thyroid (21), ovary (22), non-small cell lung (23, 24), sarcoma (25), and HCC (26, 27). Molecular targets responsible for the sensitizing effect of bortezomib on TRAIL-induced cell death include DR4 (14, 27), DR5 (14, 20, 2223, 28), c-FLIP (4, 11, 2123, 29), NF-κB (12, 24, 30), p21 (16, 21, 25), and p27 (25). In addition, Bcl-2 family also plays a role in the combinational effect of bortezomib and TRAIL, including Bcl-2 (10, 21), Bax (13, 22), Bak (27), Bcl-xL (21), Bik (18), and Bim (15).Recently, we have reported that Akt signaling is a major molecular determinant in bortezomib-induced apoptosis in HCC cells (31). In this study, we demonstrated that bortezomib overcame TRAIL resistance in HCC cells through inhibition of the PI3K/Akt pathway.  相似文献   

14.
15.
Although agonist-dependent endocytosis of G protein-coupled receptors (GPCRs) as a means to modulate receptor signaling has been widely studied, the constitutive endocytosis of GPCRs has received little attention. Here we show that two prototypical class I GPCRs, the β2 adrenergic and M3 muscarinic receptors, enter cells constitutively by clathrin-independent endocytosis and colocalize with markers of this endosomal pathway on recycling tubular endosomes, indicating that these receptors can subsequently recycle back to the plasma membrane (PM). This constitutive endocytosis of these receptors was not blocked by antagonists, indicating that receptor signaling was not required. Interestingly, the G proteins that these receptors couple to, Gαs and Gαq, localized together with their receptors at the plasma membrane and on tubular recycling endosomes. Upon agonist stimulation, Gαs and Gαq remained associated with the PM and these endosomal membranes, whereas β2 and M3 receptors now entered cells via clathrin-dependent endocytosis. Deletion of the third intracellular loop (i3 loop), which is thought to play a role in agonist-dependent endocytosis of the M3 receptor, had no effect on the constitutive internalization of the receptor. Surprisingly, with agonist, the mutated M3 receptor still internalized and accumulated in cells but through clathrin-independent and not clathrin-dependent endocytosis. These findings demonstrate that GPCRs are versatile PM proteins that can utilize different mechanisms of internalization depending upon ligand activation.G protein-coupled receptors (GPCRs)2 belong to a superfamily of seven transmembrane-spanning proteins that respond to a diverse array of sensory and chemical stimuli (14). Activation of GPCRs through the binding of specific agonists induces conformational changes that allow activation of heterotrimeric guanine nucleotide-binding proteins (G proteins) (5, 6). To ensure that the signals are controlled in magnitude and duration, activated GPCRs are rapidly desensitized through phosphorylation carried out by G protein-coupled receptor kinases (GRKs) (7). This facilitates β-arrestin binding and promotes receptor uncoupling from the G protein (8, 9). In addition to its role in GPCRs desensitization, β-arrestins promote the translocation of the receptor to the endocytic machinery involving clathrin and adaptor protein-2 (AP-2), thereby facilitating receptor removal from the plasma membrane (1015). Once internalized, some GPCRs may even continue to signal from endosomes (16).Although GPCR internalization is generally considered to be an agonist-dependent phenomenon, some evidence suggests that GPCRs can be endocytosed even in the absence of agonist, a process known as constitutive internalization (1720). The role of constitutive internalization of GPCRs is not clear. One interesting study on cannabinoid CB1 receptors in neurons has shown that constitutive internalization from the somatodendritic and not axonal membrane is responsible for the overall redistribution of receptors from the somatodentritic to the axonal membrane (17). Another study on the melanocortin MC4 receptor raised the possibility that constitutive endocytosis could be a consequence of the basal activity of the receptor (18).Even less is known about the potential trafficking of the transducer of GPCR signaling, the G protein (21). Generally, the binding of the agonist to the GPCR promotes the exchange of GDP on the Gα protein for GTP and allows the dissociation of the trimeric G protein into Gα-GTP and Gβγ dimer subunits (5, 22). Then, the activated G proteins target different effectors (23, 24). G proteins are localized primarily to the PM where they interact with GPCRs; however, it is not known whether G proteins always remain at the PM or whether they might move into cells along endocytic pathways. Previous work showed that Gαs does not colocalize with β2 receptor on internal compartments after agonist stimulation, but the cellular distribution of Gαs was not examined (25).In general, cargo proteins at the plasma membrane (PM) enter the cell through a variety of endocytic mechanisms that can be divided into two main groups: clathrin-dependent endocytosis (CDE) and clathrin-independent endocytosis (CIE). CDE is used by PM proteins such as the transferrin receptor (TfR) that contain specific cytoplasmic sequences recognized by adaptor proteins allowing a rapid and efficient internalization through clathrin-coated vesicles (26, 27). In contrast, CIE is used by PM proteins that lack adaptor protein binding sequences including cargo proteins such as the major histocompatibility complex class I protein (MHCI), the glycosylphosphatidylinositol-anchored protein CD59, and integrins (2830). In HeLa cells CIE is independent of, and CDE dependent on, clathrin and dynamin and thus the two different endocytic pathways are distinct and well defined (31). After internalization in separate vesicles, MHCI-containing vesicles from CIE and transferrin receptor-containing vesicles from CDE subsequently fuse with the early endosomal compartment that is associated with Rab5 and the early endosomal antigen 1 (EEA1) (32). TfR is recycled back out to the PM in Rab4- and Rab11-dependent processes. In contrast, some MHCI is trafficked on to late endosomes and lysosomes for degradation, and some is recycled back out to the PM along tubular endosomes that lack TfR and emanate from the juxtanuclear area. Recycling of MHCI back to the PM requires the activity of Arf6, Rab22, and Rab11 (33, 34).In this study, we analyzed the trafficking of GPCRs and their G proteins in the presence and absence of agonist in HeLa cells. We examined the trafficking of two prototypical class I GPCRs: the β2 adrenergic receptor (coupled to Gαs) and the M3 acetylcholine muscarinic receptor (coupled to Gαq). We find that β2 and M3 receptors traffic constitutively via CIE, and then, in the presence of agonist, they switch to the CDE pathway. We also examined the role of the third intracellular loop of the M3 receptor in this process. To our knowledge, this study represents the most comprehensive analysis of constitutive trafficking of class I GPCRs and related Gα proteins. We demonstrate that GPCRs are versatile PM cargos that utilize different mechanisms of internalization depending upon ligand activation. Considering the high level of homology between class I GPCRs, this evidence could be applicable to the other members of this family.  相似文献   

16.
17.
It is known that platelet-activating factor (PAF) induces severe endothelial barrier leakiness, but the signaling mechanisms remain unclear. Here, using a wide range of biochemical and morphological approaches applied in both mouse models and cultured endothelial cells, we addressed the mechanisms of PAF-induced disruption of interendothelial junctions (IEJs) and of increased endothelial permeability. The formation of interendothelial gaps filled with filopodia and lamellipodia is the cellular event responsible for the disruption of endothelial barrier. We observed that PAF ligation of its receptor induced the activation of the Rho GTPase Rac1. Following PAF exposure, both Rac1 and its guanine nucleotide exchange factor Tiam1 were found associated with a membrane fraction from which they co-immunoprecipitated with PAF receptor. In the same time frame with Tiam1-Rac1 translocation, the junctional proteins ZO-1 and VE-cadherin were relocated from the IEJs, and formation of numerous interendothelial gaps was recorded. Notably, the response was independent of myosin light chain phosphorylation and thus distinct from other mediators, such as histamine and thrombin. The changes in actin status are driven by the PAF-induced localized actin polymerization as a consequence of Rac1 translocation and activation. Tiam1 was required for the activation of Rac1, actin polymerization, relocation of junctional associated proteins, and disruption of IEJs. Thus, PAF-induced IEJ disruption and increased endothelial permeability requires the activation of a Tiam1-Rac1 signaling module, suggesting a novel therapeutic target against increased vascular permeability associated with inflammatory diseases.The endothelial barrier is made up of endothelial cells (ECs)4 connected to each other by interendothelial junctions (IEJs) consisting of protein complexes organized as tight junctions (TJs) and adherens junctions (AJs). In addition, the focal adhesion complex located at the basal plasma membrane enables firm contact of ECs with the underlying basement membrane and also contributes to the barrier function (1-3). The glycocalyx, the endothelial monolayer, and the basement membrane all together constitute the vascular barrier.The structural integrity of the ECs along with their proper functionality are the two most important factors controlling the tightness of the endothelial barrier. Changes affecting these factors cause loss of barrier restrictiveness and leakiness. Therefore, defining and understanding the cellular and molecular mechanisms controlling these processes is of paramount importance. Increased width of IEJs in response to permeability-increasing mediators (4) regulates the magnitude of transendothelial exchange of fluid and solutes. Disruption of IEJs and the resultant barrier leakiness contribute to the genesis of diverse pathological conditions, such as inflammation (5), metastasis (6, 7), and uncontrolled angiogenesis (8, 9).Accumulated evidence demonstrated that IEJs changes are responsible for increased or decreased vascular permeability, and the generally accepted mechanism responsible for them was the myosin light chain (MLC)-mediated contraction of ECs (5, 10). However, published evidence showed that an increase in vascular permeability could be obtained without a direct involvement of any contractile mechanism (11-16).The main component of the vascular barrier, the ECs, has more than 10% of their total protein represented by actin (17), which under physiological salt concentrations subsists as monomers (G-actin) and assembled into filaments (F-actin). A large number of actin-interacting proteins may modulate the assembly, disassembly, and organization of G-actin and of actin filaments within a given cell type. Similar to the complexity of actin-interacting proteins found in other cell types, the ECs utilize their actin binding proteins to stabilize the endothelial monolayer in order to efficiently function as a selective barrier (11). In undisturbed ECs, the actin microfilaments are organized as different networks with distinctive functional and morphological characteristics: the peripheral filaments also known as peripheral dense band (PDB), the cytoplasmic fibers identified as stress fibers (SF), and the actin from the membrane cytoskeleton (18). The peripheral web, localized immediately under the membrane, is associated with (i) the luminal plasmalemma (on the apical side), (ii) the IEJ complexes on the lateral surfaces, and (iii) the focal adhesion complexes on the abluminal side (the basal part) of polarized ECs. The SF reside inside the endothelial cytoplasm and are believed to be directly connected with the plasmalemma proper on the luminal as well as on the abluminal side of the cell. As described, the endothelial actin cytoskeleton (specifically the SF) seems to be a stable structure helping the cells to remain flat under flow (19). It is also established that the actin fibers participate in correct localization of different junctional complexes while keeping them in place (20). However, it was suggested that the dynamic equilibrium between F- and G-actin might modulate the tightness of endothelial barrier in response to different challenges (13).Mediators effective at nanomolar concentrations or less that disrupt the endothelial barrier and increase vascular permeability include C2 toxin of Clostridium botulinum, vascular permeability factor, better known as vascular endothelial growth factor, and PAF (21). C2 toxin increases endothelial permeability by ribosylating monomeric G-actin at Arg-177 (22). This results in the impairment of actin polymerization (23), followed by rounding of ECs (16) and the disruption of junctional integrity. Vascular permeability factor was shown to open IEJs by redistribution of junctional proteins (24, 25) and by interfering with the equilibrium of actin pools (26). PAF (1-O-alkyl-2-acetyl-sn-glycero-3-phosphocoline), a naturally synthesized phospholipid is active at 10-10 m or less (27). PAF is synthesized by and acts on a variety of cell types, including platelets (28), neutrophils (29), monocytes (30), and ECs (31). PAF-mediated activation of ECs induced cell migration (32), angiogenesis (7), and vascular hyperpermeability (33) secondary to disassembly of IEJs (34). The effects of PAF on the endothelium are initiated through a G protein-coupled receptor (PAF-R) localized at the plasmalemma, in a large endosomal compartment inside the cell (34), and also in the nuclear membrane (35). In ECs, PAF-R was shown to signal through Gαq and downstream activation of phospholipase C isozymes (PLCβ3 and PLCγ1), and via cSrc (32, 36). Studies have shown that PAF challenge induced endothelial actin cytoskeletal rearrangement (37) and marked vascular leakiness (38); however, the signaling pathways have not been elucidated.Therefore, in the present study, we carried out a systematic analysis of PAF-induced morphological and biochemical changes of endothelial barrier in vivo and in cultured ECs. We found that the opening of endothelial barrier and the increased vascular leakiness induced by PAF are the result of a shift in actin pools without involvement of EC contraction, followed by a redistribution of tight junctional associated protein ZO-1 and adherens junctional protein VE-cadherin.  相似文献   

18.
Matrix metalloproteinase-2 (MMP-2) is an important extracellular matrix remodeling enzyme, and it has been involved in different fibrotic disorders. The connective tissue growth factor (CTGF/CCN2), which is increased in these pathologies, induces the production of extracellular matrix proteins. To understand the fibrotic process observed in diverse pathologies, we analyzed the fibroblast response to CTGF when MMP-2 activity is inhibited. CTGF increased fibronectin (FN) amount, MMP-2 mRNA expression, and gelatinase activity in 3T3 cells. When MMP-2 activity was inhibited either by the metalloproteinase inhibitor GM-6001 or in MMP-2-deficient fibroblasts, an increase in the basal amount of FN together with a decrease of its levels in response to CTGF was observed. This paradoxical effect could be explained by the fact that the excess of FN could block the access to other ligands, such as CTGF, to integrins. This effect was emulated in fibroblasts by adding exogenous FN or RGDS peptides or using anti-integrin αV subunit-blocking antibodies. Additionally, in MMP-2-deficient cells CTGF did not induce the formation of stress fibers, focal adhesion sites, and ERK phosphorylation. Anti-integrin αV subunit-blocking antibodies inhibited ERK phosphorylation in control cells. Finally, in MMP-2-deficient cells, FN mRNA expression was not affected by CTGF, but degradation of 125I-FN was increased. These results suggest that expression, regulation, and activity of MMP-2 can play an important role in the initial steps of fibrosis and shows that FN levels can regulate the cellular response to CTGF.Extracellular proteolysis is an essential physiological process that controls the immediate cellular environment and thus plays a key role in cellular behavior and survival (1). The members of the matrix metalloproteinase (MMP)2 family of zinc-dependent endopeptidases are major mediators of extracellular proteolysis by promoting the degradation of extracellular matrix (ECM) components and cell surface-associated proteins (2, 3). Each one of these enzymes is negatively regulated by tissue inhibitors of metalloproteinases (TIMPs) (4) and is secreted as a zymogen (pro-MMPs) that is activated in the extracellular space (57). This mechanism is an important form of regulation of gelatinase activity and in consequence, highly significant for ECM homeostasis. Among the members of the MMP family, the metalloproteinase type 2 (MMP-2 or gelatinase A) is known to be a key player in many physiological and pathological processes, such as cell migration, inflammation, angiogenesis, and fibrosis (811).Fibrotic disorders are typified by excessive connective tissue and ECM deposition that precludes normal healing of different tissues. ECM accumulation can be explained in two ways: increasing expression and deposition of connective tissue proteins and/or decreasing degradation of ECM proteins (12). Transforming growth factor type β, a multifunctional cytokine, is strongly overexpressed, and it is associated to the pathogenesis of these diseases (13, 14). It stimulates the expression of connective tissue growth factor (CTGF/CCN2) (15), a cytokine that is responsible for transforming growth factor type β fibrotic activity (16, 17). The role of CTGF in fibrosis has gained attention in recent years (16, 1822). CTGF overexpression is known to occur in a variety of fibrotic skin disorders (23, 24), renal (25), hepatic (26), and pulmonary fibrosis (27) and in muscles from patients with Duchenne muscular dystrophy (28).On the other hand, several pathologies involving fibrosis show an increase in MMP expression, including gelatinase A. Augmented expression of MMP-2 was found in submucous (29), skin (30), liver (31), and lung fibrosis (32, 33) and dystrophic myotubes from fibrotic muscles of Duchenne muscular dystrophy (34). It has been shown that transforming growth factor type β induces an increase in the amount of MMP-2 in fibroblasts (35) and that CTGF induces MMP-2 expression in cultured renal interstitial fibroblasts (36). The putative role assigned to MMP-2 in fibrotic disorders is related to tissue regeneration because of the capacity of this enzyme to degrade basal lamina (3739). Because MMP-2 expression is up-regulated in these pathologies but still a high ECM deposition is observed, we propose that this accumulation could be explained by a diminution of the MMP-2 enzymatic activity.In this article, we demonstrate that CTGF increases fibronectin (FN) amount, MMP-2 expression, and gelatinase activity in 3T3 fibroblasts. More significantly, we show that MMP-2-deficient cells have an increased basal amount of FN and show a response to CTGF that is opposite to that of control cells. This paradoxical effect could be explained by the increase in the FN amount that blocks the integrins (at least integrins with αV subunit), which can act like CTGF receptors.  相似文献   

19.
20.
The activation of LFA-1 (lymphocyte function-associated antigen) is a critical event for T cell co-stimulation. The mechanism of LFA-1 activation involves both affinity and avidity regulation, but the role of each in T cell activation remains unclear. We have identified antibodies that recognize and block different affinity states of the mouse LFA-1 I-domain. Monoclonal antibody 2D7 preferentially binds to the low affinity conformation, and this specific binding is abolished when LFA-1 is locked in the high affinity conformation. In contrast, M17/4 can bind both the locked high and low affinity forms of LFA-1. Although both 2D7 and M17/4 are blocking antibodies, 2D7 is significantly less potent than M17/4 in blocking LFA-1-mediated adhesion; thus, blocking high affinity LFA-1 is critical for preventing LFA-1-mediated adhesion. Using these reagents, we investigated whether LFA-1 affinity regulation affects T cell activation. We found that blocking high affinity LFA-1 prevents interleukin-2 production and T cell proliferation, demonstrated by TCR cross-linking and antigen-specific stimulation. Furthermore, there is a differential requirement of high affinity LFA-1 in the activation of CD4+ and CD8+ T cells. Although CD4+ T cell activation depends on both high and low affinity LFA-1, only high affinity LFA-1 provides co-stimulation for CD8+ T cell activation. Together, our data demonstrated that the I-domain of LFA-1 changes to the high affinity state in primary T cells, and high affinity LFA-1 is critical for facilitating T cell activation. This implicates LFA-1 activation as a novel regulatory mechanism for the modulation of T cell activation and proliferation.LFA-1 (lymphocyte function-associated antigen), an integrin family member, is important in regulating leukocyte adhesion and T cell activation (1, 2). LFA-1 consists of the αL (CD11a) and β2 (CD18) heterodimer. The ligands for LFA-1, including intercellular adhesion molecule ICAM3-1, ICAM-2, and ICAM-3, are expressed on antigen-presenting cells (APCs), endothelial cells, and lymphocytes (1). Mice that are deficient in LFA-1 have defects in leukocyte adhesion, lymphocyte proliferation, and tumor rejection (35). Blocking LFA-1 with antibodies can prevent inflammation, autoimmunity, organ graft rejection, and graft versus host disease in human and murine models (610).LFA-1 is constitutively expressed on the surface of leukocytes in an inactive state. Activation of LFA-1 is mediated by inside-out signals from the cytoplasm (1, 11). Subsequently, activated LFA-1 binds to the ligands and transduces outside-in signals back into the cytoplasm that result in cell adhesion and activation (12, 13). The activation of LFA-1 is a critical event in the formation of the immunological synapse, which is important for T cell activation (2, 14, 15). The active state of LFA-1 is regulated by chemokines and the T cell receptor (TCR) through Rap1 signaling (16). LFA-1 ligation lowers the activation threshold and affects polarization in CD4+ T cells (17). Moreover, productive LFA-1 engagement facilitates efficient activation of cytotoxic T lymphocytes and initiates a distinct signal essential for the effector function (1820). Thus, LFA-1 activation is essential for the optimal activation of T cells.The mechanism of LFA-1 activation involves both affinity (conformational changes within the molecule) and avidity (receptor clustering) regulation (2123). The I-domain of the LFA-1 αL subunit is the primary ligand-binding site and has been proposed to change conformation, leading to an increased affinity for ligands (2426). The structural basis of the conformational changes in the I-domain of LFA-1 has been extensively characterized (27). Previously, we have demonstrated that the conformation of the LFA-1 I-domain changes from the low affinity to the high affinity state upon activation. By introducing disulfide bonds into the I-domain, LFA-1 can be locked in either the closed or open conformation, which represents the “low affinity” or “high affinity” state, respectively (28, 29). In addition, we identified antibodies that are sensitive to the affinity changes in the I-domain of human LFA-1 and showed that the activation-dependent epitopes are exposed upon activation (30). This study supports the presence of the high affinity conformation upon LFA-1 activation in cell lines. It has been demonstrated recently that therapeutic antagonists, such as statins, inhibit LFA-1 activation and immune responses by locking LFA-1 in the low affinity state (3134). Furthermore, high affinity LFA-1 has been shown to be important for mediating the adhesion of human T cells (35, 36). Thus, the affinity regulation is a critical step in LFA-1 activation.LFA-1 is a molecule of great importance in the immune system, and its activation state influences the outcome of T cell activation. Our previous data using the activating LFA-1 I-domain-specific antibody MEM83 indicate that avidity and affinity of the integrin can be coupled during activation (37). However, whether affinity or avidity regulation of LFA-1 contributes to T cell activation remains controversial (23, 38, 39). Despite the recent progress suggesting that conformational changes represent a key step in the activation of LFA-1, there are considerable gaps to be filled. When LFA-1 is activated, the subsequent outside-in signaling contributes to T cell activation via immunological synapse and LFA-1-dependent signaling. It is critical to determine whether high affinity LFA-1 participates in the outside-in signaling and affects the cellular activation of T cells. Nevertheless, the rapid and dynamic process of LFA-1 activation has hampered further understanding of the role of high affinity LFA-1 in primary T cell activation. The affinity of LFA-1 for ICAM-1 increases up to 10,000-fold within seconds and involves multiple reversible steps (23). In addition, the activation of LFA-1 regulates both adhesion and activation of T cells, two separate yet closely associated cellular functions. When LFA-1 is constitutively expressed in the active state in mice, immune responses are broadly impaired rather than hyperactivated, suggesting the complexity of affinity regulation (40). Therefore, it is difficult to dissect the mechanisms by which high affinity LFA-1 regulates stepwise activation of T cells in the whole animal system.In the present study, we identified antibodies recognizing and blocking different affinity states of mouse LFA-1. These reagents allowed us to determine the role of affinity regulation in T cell activation. We found that blocking high affinity LFA-1 inhibited IL-2 production and proliferation in T cells. Furthermore, there is a differential requirement of high affinity LFA-1 in antigen-specific activation of CD4+ and CD8+ T cells. The activation of CD4+ T cells depends on both high and low affinity LFA-1. For CD8+ T cell activation, only high affinity LFA-1 provides co-stimulation. Thus, affinity regulation of LFA-1 is critical for the activation and proliferation of naive T cells.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号