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 484 毫秒
1.
2.
Although mesenchymal stem cells (MSCs) are the natural source for bone regeneration, the exact mechanisms governing MSC crosstalk with collagen I have not yet been uncovered. Cell adhesion to collagen I is mostly mediated by three integrin receptors – α1β1, α2β1 and α11β1. Using human MSC (hMSC), we show that α11 subunit exhibited the highest basal expression levels but on osteogenic stimulation, both α2 and α11 integrins were significantly upregulated. To elucidate the possible roles of collagen-binding integrins, we applied short hairpin RNA (shRNA)-mediated knockdown in hMSC and found that α2 or α11 deficiency, but not α1, results in a tremendous reduction of hMSC numbers owing to mitochondrial leakage accompanied by Bcl-2-associated X protein upregulation. In order to clarify the signaling conveyed by the collagen-binding integrins in hMSC, we analyzed the activation of focal adhesion kinase, extracellular signal-regulated protein kinase and serine/threonine protein kinase B (PKB/Akt) kinases and detected significantly reduced Akt phosphorylation only in α2- and α11-shRNA hMSC. Finally, experiments with hMSC from osteoporotic patients revealed a significant downregulation of α2 integrin concomitant with an augmented mitochondrial permeability. In conclusion, our study describes for the first time that disturbance of α2β1- or α11β1-mediated interactions to collagen I results in the cell death of MSCs and urges for further investigations examining the impact of MSCs in bone conditions with abnormal collagen I.  相似文献   

3.
MicroRNA 183 (miR-183) has been reported to inhibit tumor invasiveness and is believed to be involved in the development and function of ciliated neurosensory organs. We have recently found that expression of miR-183 increased after the induction of cellular senescence by exposure to H2O2. To gain insight into the biological roles of miR-183 we investigated two potential novel targets: integrin β1 (ITGB1) and kinesin 2α (KIF2A). miR-183 significantly decreased the expression of ITGB1 and KIF2A measured by Western blot. Targeting of the 3′-untranslated region (3′-UTR) of ITGB1 and KIF2A by miR-183 was confirmed by luciferase assay. Transfection with miR-183 led to a significant decrease in cell invasion and migration capacities of HeLa cells that could be rescued by expression of ITGB1 lacking the 3′-UTR. Although miR-183 had no effects on cell adhesion in HeLa cells, it significantly decreased adhesion to laminin, gelatin, and collagen type I in normal human diploid fibroblasts and human trabecular meshwork cells. These effects were also rescued by expression of ITGB1 lacking the 3′-UTR. Targeting of KIF2A by miR-183 resulted in some increase in the formation of cells with monopolar spindles in HeLa cells but not in human diploid fibroblast or human trabecular meshwork cells. The regulation of ITGB1 expression by miR-183 provides a new mechanism for the anti-metastatic role of miR-183 and suggests that this miRNA could influence the development and function in neurosensory organs, and contribute to functional alterations associated with cellular senescence in human diploid fibroblasts and human trabecular meshwork cells.  相似文献   

4.
The free energy of transfer of nonpolar solutes from water to lipid bilayers is often dominated by a large negative enthalpy rather than the large positive entropy expected from the hydrophobic effect. This common observation has led to the idea that membrane partitioning is driven by the "nonclassical" hydrophobic effect. We examined this phenomenon by characterizing the partitioning of the well-studied peptide melittin using isothermal titration calorimetry (ITC) and circular dichroism (CD). We studied the temperature dependence of the entropic (-TΔS) and enthalpic (ΔH) components of free energy (ΔG) of partitioning of melittin into lipid membranes made of various mixtures of zwitterionic and anionic lipids. We found significant variations of the entropic and enthalpic components with temperature, lipid composition and vesicle size but only small changes in ΔG (entropy-enthalpy compensation). The heat capacity associated with partitioning had a large negative value of about -0.5 kcal mol(-1) K(-1). This hallmark of the hydrophobic effect was found to be independent of lipid composition. The measured heat capacity values were used to calculate the hydrophobic-effect free energy ΔG (hΦ), which we found to dominate melittin partitioning regardless of lipid composition. In the case of anionic membranes, additional free energy comes from coulombic attraction, which is characterized by a small effective peptide charge due to the lack of additivity of hydrophobic and electrostatic interactions in membrane interfaces [Ladokhin and White J Mol Biol 309:543-552, 2001]. Our results suggest that there is no need for a special effect-the nonclassical hydrophobic effect-to describe partitioning into lipid bilayers.  相似文献   

5.
6.
Fibrosis is characterized by elevated transforming growth factor β (TGFβ) signaling, resulting in extracellular matrix accumulation and increased PAI-1 (plasminogen activator inhibitor) expression. PAI-1 induces the internalization of urokinase plasminogen activator/receptor and integrin αvβ3 from the cell surface. Since increased αvβ3 expression correlates with increased TGFβ signaling, we hypothesized that aberrant PAI-1-mediated αvβ3 endocytosis could initiate an autocrine loop of TGFβ activity. We found that in PAI-1 knock-out (KO) mouse embryonic fibroblasts), αvβ3 endocytosis was reduced by ∼75%, leaving αvβ3 in enlarged focal adhesions, similar to wild type cells transfected with PAI-1 small interfering RNA. TGFβ signaling was significantly enhanced in PAI-1 KO cells, as demonstrated by a 3-fold increase in SMAD2/3-containing nuclei and a 2.9-fold increase in TGFβ activity that correlated with an increase in αvβ3 and TGFβ receptor II expression. As expected, PAI-1 KO cells had unregulated plasmin activity, which was only partially responsible for TGFβ activation, as evidenced by a mere 25% reduction in TGFβ activity when plasmin was inhibited. Treatment of cells with an αvβ3-specific cyclic RGD peptide (GpenGRGD) led to a more profound (59%) TGFβ inhibition; a nonspecific RGD peptide (GRGDNP) inhibited TGFβ by only 23%. Human primary fibroblasts were used to confirm that PAI-1 inhibition and β3 overexpression led to an increase in TGFβ activity. Consistent with a fibrotic phenotype, PAI-1 KO cells were constitutively myofibroblasts that had a 1.6-fold increase in collagen deposition over wild type cells. These data suggest that PAI-1-mediated regulation of αvβ3 integrin is critical for the control of TGFβ signaling and the prevention of fibrotic disease.Fibrotic disorders can result from environmental toxins, persistent infection, autoimmune disease, or mechanical injury, leading to the hardening and scarring of tissues. In fibrotic diseases, such as liver cirrhosis, renal fibrosis, and idiopathic lung fibrosis, or in pathological wound healing, such as hypertrophic scarring, scleroderma, and Dupuytren disease, the persistence of myofibroblasts contributes to disease progression by overproduction of extracellular matrix (ECM)2 and by excessive contraction (13). A shift in the balance of growth factors and cytokines that promote ECM deposition and proteases that degrade matrix often contributes to fibrotic disease (4, 5). Plasmin, a broad spectrum protease that is generated from plasminogen by uPA, is one of the proteases that degrades matrix and activates growth factors and other proteases (6). Since uPA activity is inhibited by PAI-1, the overexpression of PAI-1 results in matrix accumulation. For this reason, PAI-1 is a key prognostic marker for fibrotic disease. PAI-1 exerts its inhibitory activity on uPA by stimulating the endocytosis of the cell surface uPA·uPAR complex through the low density lipoprotein receptor-related protein (7). Integrin αvβ3 is also internalized with the uPA·uPAR·low density lipoprotein receptor-related protein complex (8). After endocytosis, uPAR and integrins are recycled back to the cell surface for another round of binding (8, 9). uPAR and αvβ3 promote cellular attachment and spreading, since they are receptors for the extracellular matrix molecule, vitronectin (10). Thus, cycling of the complex is thought to stimulate the attachment and detachment that is necessary for cell migration (8). Consequently, a shift in the expression of any of these components (PAI-1/uPA/uPAR/αvβ3) can result in either aggressive migration, as seen in cancer invasion, or a persistent increase in cell adhesion and cell tension, as seen in myofibroblasts in fibrotic tissue.The family of TGFβ growth factors has been intensively studied for their role in fibrotic wound healing. Up-regulation of TGFβ results in amplified and persistent overproduction of molecules, such as integrins and PAI-1 and other protease inhibitors (e.g. TIMPs) (2, 3). Up-regulated integrins continue the cycle of TGFβ signaling by participating in the sustained activation of TGFβ from its latent form. To date, studies have found that various αv integrins participate in the activation of TGFβ (αvβ3, αvβ5, αvβ6, and αvβ8), but the mechanism differs (1115). Integrins can serve as docking proteins to localize proteases that cleave and activate latent TGFβ in the ECM, or they can directly activate latent TGFβ in a protease-independent manner. Recently, it was discovered that latent TGFβ is also activated by mechanical stress generated from an integrin-mediated interaction between myofibroblasts and the ECM, primarily involving αvβ5. The mechanical stress promotes a conformational change that activates the latent TGFβ complex (15). αv integrins also modulate TGFβ signaling through the binding of αvβ3 to TGFβ receptor II (TGFβRII) in the presence of TGFβ. This interaction was shown to promote a dramatic increase in the proliferation of lung fibroblasts and induce invasion of epithelial breast cancer cells (16, 17).Our data establish a role for the PAI-1-mediated control of αvβ3 expression and support a significant role for αvβ3 in TGFβ signaling. Using PAI-1 KO cells, we tested the hypothesis that the absence of PAI-1 would result in the accumulation of αvβ3 on the cell surface, since PAI-1 promotes the endocytosis of uPA·uPAR·αvβ3. PAI-1-mediated endocytosis of β3 was significantly reduced in the PAI-1 KO cells. Correspondingly, we report that β3 accumulated at the cell surface in enlarged β3-containing focal adhesions. Thus, we explored whether the accumulation of αvβ3 on the cell surface had fibrogenic effects even in the absence of profibrotic PAI-1. Our results demonstrate dramatically increased TGFβ activity and an increase in collagen expression in PAI-1 KO cells. Together, these findings suggest that PAI-1 modulates β3 expression and localization and, in turn, TGFβ signaling. Our data reveal that maintaining precise levels of PAI-1 is a key to preventing fibrosis. Understanding the consequence of regulating PAI-1 activity is critical in light of the many clinical therapies currently under development that target PAI-1 (18, 19).  相似文献   

7.
8.
Endothelin-1 (ET-1) is a potent vasoconstrictor and co-mitogen for vascular smooth muscle and is implicated in pulmonary vascular remodeling and the development of pulmonary arterial hypertension. Vascular smooth muscle is an important source of ET-1. Here we demonstrate synergistic induction of preproET-1 message RNA and release of mature peptide by a combination of tumor necrosis factor α (TNFα) and interferon γ (IFNγ) in primary human pulmonary artery smooth muscle cells. This induction was prevented by pretreatment with the histone acetyltransferase inhibitor anacardic acid. TNFα induced a rapid and prolonged pattern of nuclear factor (NF)-κB p65 subunit activation and binding to the native preproET-1 promoter. In contrast, IFNγ induced a delayed activation of interferon regulatory factor-1 without any effect on NF-κB p65 nuclear localization or consensus DNA binding. However, we found cooperative p65 binding and histone H4 acetylation at distinct κB sites in the preproET-1 promoter after stimulation with both TNFα and IFNγ. This was associated with enhanced recruitment of RNA polymerase II to the ATG start site and read-through of the ET-1 coding region. Understanding such mechanisms is crucial in determining the key control points in ET-1 release. This has particular relevance to developing novel treatments targeted at the inflammatory component of pulmonary vascular remodeling.Endothelin-1 is a 21-amino acid peptide which is known to be both a potent vasoconstrictor and mitogen for vascular smooth muscle (1, 2). It is released as a 38-amino acid precursor (Big ET-12) before cleavage to the mature ET-1 form. As such it has been implicated in the pathogenesis of vascular disease and is particularly associated with pulmonary arterial hypertension (3). Indeed, several endothelin receptor antagonists are now approved for the treatment of pulmonary arterial hypertension (4). However, endothelin receptor antagonists as a class are associated with potentially serious side effects (4), making new treatments aimed at blocking ET-1 synthesis an attractive alternative.Although endothelial cells are thought to be the main source of ET-1 release, several groups including our own have shown that ET-1 can be released from the more numerous vascular smooth muscle cells (510). The vascular pathology observed in pulmonary arterial hypertension is propagated by inflammation, and circulating levels of cytokines including tumor necrosis factor α (TNFα) are elevated in patients with pulmonary arterial hypertension (1115). In many cell types cytokines mediate their biological effects at least in part by the activation of the nuclear factor κB (NF-κB) pathway (16), and a role for NF-κB in pulmonary arterial hypertension has been proposed (17). In addition, we have shown previously that a combination of TNFα and interferon γ (IFNγ) stimulates human pulmonary artery smooth muscle (HPASM) cells to release ET-1 (18). However, the mechanisms underlying this effect are unknown.The preproET-1 promoter region has been shown experimentally to possess binding sites for nuclear factor (NF)-1 and phorbol ester-sensitive c-Fos and c-Jun complexes (19), acute phase reactant regulatory proteins, and binding sites for AP-1 and GATA-2 (2022). In addition, binding sites for interferon regulatory factor-1 (IRF-1) and NF-κB are predicted by Transfac analysis (23). The close proximity of the IRF-1 site and one of the NF-κB sites is characteristic of genes that are regulated by the synergistic action of TNFα and IFNγ, such as interleukin-6 (IL-6) and intercellular adhesion molecule-1 (24, 25), although ET-1 has not previously been recognized in this group.Our aims were, therefore, to investigate the role of NF-κB in ET-1 release by primary HPASM cells. In addition, we were interested in the role of histone acetylation in the epigenetic control of the ET-1 production. Understanding these novel mechanisms will allow a greater understanding of the pathogenesis of vascular remodeling in pulmonary vessels and aid in the development of new treatment strategies aimed at blocking synthesis of ET-1.  相似文献   

9.
10.
Synapses are asymmetric intercellular junctions that mediate neuronal communication. The number, type, and connectivity patterns of synapses determine the formation, maintenance, and function of neural circuitries. The complexity and specificity of synaptogenesis relies upon modulation of adhesive properties, which regulate contact initiation, synapse formation, maturation, and functional plasticity. Disruption of adhesion may result in structural and functional imbalance that may lead to neurodevelopmental diseases, such as autism, or neurodegeneration, such as Alzheimer''s disease. Therefore, understanding the roles of different adhesion protein families in synapse formation is crucial for unraveling the biology of neuronal circuit formation, as well as the pathogenesis of some brain disorders. The present review summarizes some of the knowledge that has been acquired in vertebrate and invertebrate genetic model organisms.Synapses are asymmetric, intercellular junctions that are the basic structural units of neuronal transmission. The correct development of synaptic specializations and the establishment of appropriate connectivity patterns are crucial for the assembly of functional neuronal circuits. Improper synapse formation and function may cause neurodevelopmental disorders, such as mental retardation (MsR) and autism spectrum disorders (ASD) (McAllister 2007; Sudhof 2008), and likely play a role in neurodegenerative disorders, such as Alzheimer''s disease (AD) (Haass and Selkoe 2007).At chemical synapses (reviewed in Sudhof 2004; Zhai and Bellen 2004; Waites et al. 2005; McAllister 2007; Jin and Garner 2008), the presynaptic compartment contains synaptic vesicles (SV), organized in functionally distinct subcellular pools. A subset of SVs docks to the presynaptic membrane around protein-dense release sites, named active zones (AZ). Upon the arrival of an action potential at the terminal, the docked and “primed” SVs fuse with the plasma membrane and release neurotransmitter molecules into the synaptic cleft. Depending on the type of synapse (i.e., excitatory vs. inhibitory synapses), neurotransmitters ultimately activate an appropriate set of postsynaptic receptors that are accurately apposed to the AZ.Synapse formation occurs in several steps (Fig. 1) (reviewed in Eaton and Davis 2003; Goda and Davis 2003; Waites et al. 2005; Garner et al. 2006; Gerrow and El-Husseini 2006; McAllister 2007). Spatiotemporal signals guide axons through heterogeneous cellular environments to contact appropriate postsynaptic targets. At their destination, axonal growth cones initiate synaptogenesis through adhesive interactions with target cells. In the mammalian central nervous system (CNS), immature postsynaptic dendritic spines initially protrude as thin, actin-rich filopodia on the surface of dendrites. Similarly, at the Drosophila neuromuscular junction (NMJ), myopodia develop from the muscles (Ritzenthaler et al. 2000). The stabilization of intercellular contacts and their elaboration into mature, functional synapses involves cytoskeletal arrangements and recruitment of pre- and postsynaptic components to contact sites in spines and boutons. Conversely, retraction of contacts results in synaptic elimination. Both stabilization and retraction sculpt a functional neuronal circuitry.Open in a separate windowFigure 1.(A–C) Different stages of synapse formation. (A) Target selection, (B) Synapse assembly, (C) Synapse maturation and stabilization. (D–F) The role of cell adhesion molecules in synapse formation is exemplified by the paradigm of N-cadherin and catenins in regulation of the morphology and strength of dendritic spine heads. (D) At an early stage the dendritic spines are elongated from motile structures “seeking” their synaptic partners. (E) The contacts between the presynaptic and postsynaptic compartments are stabilized by recruitment of additional cell adhesion molecules. Adhesional interactions activate downstream pathways that remodel the cytoskeleton and organize pre- and postsynaptic apparatuses. (F) Cell adhesion complexes, stabilized by increased synaptic activity, promote the expansion of the dendritic spine head and the maturation/ stabilization of the synapse. Retraction and expansion is dependent on synaptic plasticity.In addition to the plastic nature of synapse formation, the vast heterogeneity of synapses (in terms of target selection, morphology, and type of neurotransmitter released) greatly enhances the complexity of synaptogenesis (reviewed in Craig and Boudin 2001; Craig et al. 2006; Gerrow and El-Husseini 2006). The complexity and specificity of synaptogenesis relies upon the modulation of adhesion between the pre- and postsynaptic components (reviewed in Craig et al. 2006; Gerrow and El-Husseini 2006; Piechotta et al. 2006; Dalva et al. 2007; Shapiro et al. 2007; Yamada and Nelson 2007; Gottmann 2008). Cell adhesive interactions enable cell–cell recognition via extracellular domains and also mediate intracellular signaling cascades that affect synapse morphology and organize scaffolding complexes. Thus, cell adhesion molecules (CAMs) coordinate multiple synaptogenic steps.However, in vitro and in vivo studies of vertebrate CAMs are often at odds with each other. Indeed, there are no examples of mutants for synaptic CAMs that exhibit prominent defects in synapse formation. This apparent “resilience” of synapses is probably caused by functional redundancy or compensatory effects among different CAMs (Piechotta et al. 2006). Hence, studies using simpler organisms less riddled by redundancy, such as Caenorhabditis elegans and Drosophila, have aided in our understanding of the role that these molecules play in organizing synapses.In this survey, we discuss the roles of the best characterized CAM families of proteins involved in synaptogenesis. Our focus is to highlight the complex principles that govern the molecular basis of synapse formation and function from a comparative perspective. We will present results from cell culture studies as well as in vivo analyses in vertebrate systems and refer to invertebrate studies, mainly performed in Drosophila and C. elegans, when they have provided important insights into the role of particular CAM protein families. However, we do not discuss secreted factors, for which we refer the reader to numerous excellent reviews (as for example Washbourne et al. 2004; Salinas 2005; Piechotta et al. 2006; Shapiro et al. 2006; Dalva 2007; Yamada and Nelson 2007; Biederer and Stagi 2008; Salinas and Zou 2008).  相似文献   

11.
We previously found that pigeon IgG possesses unique N-glycan structures that contain the Galα1–4Galβ1–4Galβ1–4GlcNAc sequence at their nonreducing termini. This sequence is most likely produced by putative α1,4- and β1,4-galactosyltransferases (GalTs), which are responsible for the biosynthesis of the Galα1–4Gal and Galβ1–4Gal sequences on the N-glycans, respectively. Because no such glycan structures have been found in mammalian glycoproteins, the biosynthetic enzymes that produce these glycans are likely to have distinct substrate specificities from the known mammalian GalTs. To study these enzymes, we cloned the pigeon liver cDNAs encoding α4GalT and β4GalT by expression cloning and characterized these enzymes using the recombinant proteins. The deduced amino acid sequence of pigeon α4GalT has 58.2% identity to human α4GalT and 68.0 and 66.6% identity to putative α4GalTs from chicken and zebra finch, respectively. Unlike human and putative chicken α4GalTs, which possess globotriosylceramide synthase activity, pigeon α4GalT preferred to catalyze formation of the Galα1–4Gal sequence on glycoproteins. In contrast, the sequence of pigeon β4GalT revealed a type II transmembrane protein consisting of 438 amino acid residues, with no significant homology to the glycosyltransferases so far identified from mammals and chicken. However, hypothetical proteins from zebra finch (78.8% identity), frogs (58.9–60.4%), zebrafish (37.1–43.0%), and spotted green pufferfish (43.3%) were similar to pigeon β4GalT, suggesting that the pigeon β4GalT gene was inherited from the common ancestors of these vertebrates. The sequence analysis revealed that pigeon β4GalT and its homologs form a new family of glycosyltransferases.  相似文献   

12.
Transforming growth factor-β-activated kinase 1 (TAK1) plays an essential role in the tumor necrosis factor α (TNFα)- and interleukin-1β (IL-1β)-induced IκB kinase (IKK)/nuclear factor-κB (NF-κB) and c-Jun N-terminal kinase (JNK)/activator protein 1 (AP-1) activation. Here we report that TNFα and IL-1β induce Lys63-linked TAK1 polyubiquitination at the Lys158 residue within the kinase domain. Tumor necrosis factor receptor-associated factors 2 and 6 (TRAF2 and -6) act as the ubiquitin E3 ligases to mediate Lys63-linked TAK1 polyubiquitination at the Lys158 residue in vivo and in vitro. Lys63-linked TAK1 polyubiquitination at the Lys158 residue is required for TAK1-mediated IKK complex recruitment. Reconstitution of TAK1-deficient mouse embryo fibroblast cells with TAK1 wild type or a TAK1 mutant containing a K158R mutation revealed the importance of this site in TNFα and IL-1β-mediated IKK/NF-κB and JNK/AP-1 activation as well as IL-6 gene expression. Our findings demonstrate that Lys63-linked polyubiquitination of TAK1 at Lys158 is essential for its own kinase activation and its ability to mediate its downstream signal transduction pathways in response to TNFα and IL-1β stimulation.  相似文献   

13.
14.
Transforming growth factor-β (TGF-β) regulates a wide variety of biological activities. It induces potent growth-inhibitory responses in normal cells but promotes migration and invasion of cancer cells. Smads mediate the TGF-β responses. TGF-β binding to the cell surface receptors leads to the phosphorylation of Smad2/3 in their C terminus as well as in the proline-rich linker region. The serine/threonine phosphorylation sites in the linker region are followed by the proline residue. Pin1, a peptidyl-prolyl cis/trans isomerase, recognizes phosphorylated serine/threonine-proline motifs. Here we show that Smad2/3 interacts with Pin1 in a TGF-β-dependent manner. We further show that the phosphorylated threonine 179-proline motif in the Smad3 linker region is the major binding site for Pin1. Although epidermal growth factor also induces phosphorylation of threonine 179 and other residues in the Smad3 linker region the same as TGF-β, Pin1 is unable to bind to the epidermal growth factor-stimulated Smad3. Further analysis suggests that phosphorylation of Smad3 in the C terminus is necessary for the interaction with Pin1. Depletion of Pin1 by small hairpin RNA does not significantly affect TGF-β-induced growth-inhibitory responses and a number of TGF-β/Smad target genes analyzed. In contrast, knockdown of Pin1 in human PC3 prostate cancer cells strongly inhibited TGF-β-mediated migration and invasion. Accordingly, TGF-β induction of N-cadherin, which plays an important role in migration and invasion, is markedly reduced when Pin1 is depleted in PC3 cells. Because Pin1 is overexpressed in many cancers, our findings highlight the importance of Pin1 in TGF-β-induced migration and invasion of cancer cells.  相似文献   

15.
16.
17.
18.
19.
Regulation by the NK and T cell surface receptor CD244 in mice and humans depends both on engagement at the cell surface by CD48 and intracellular interactions with SAP and EAT-2. Relevance to human disease by manipulating CD244 in mouse models is complicated by rodent CD2 also binding CD48. We distinguish between contributions of mouse CD244 and CD2 on engagement of CD48 in a mouse T cell hybridoma. CD2 and CD244 both contribute positively to the immune response as mutation of proline-rich motifs or tyrosine motifs in the tails of CD2 and CD244, respectively, result in a decrease in antigen-specific interleukin-2 production. Inhibitory effects of mouse CD244 are accounted for by competition with CD2 at the cell surface for CD48. In humans CD2 and CD244 are engaged separately at the cell surface but biochemical data suggest a potential conserved intracellular link between the two receptors through FYN kinase. We identify a novel signaling mechanism for CD244 through its potential to recruit phospholipase C-γ1 via the conserved phosphorylated tyrosine motif in the tail of the adaptor protein EAT-2, which we show is important for function.The CD2 family of cell surface receptors is differentially expressed on immune cells (1, 2) and is involved in regulating both innate and adaptive immunity (3). These receptors have related extracellular immunoglobulin superfamily domains and interact either homophilically or heterophilically within the CD2 family (1, 2). The CD2 family contains a subgroup of receptors termed the SLAM family that have a conserved tyrosine signaling motif in their cytoplasmic region TXYXX(I/V) referred to as an immunoreceptor tyrosine-based switch motif (ITSM).2 The SLAM family of receptors include CD244 (2B4), NTB-A (Ly-108), CD319 (CRACC, CS-1), CD150 (SLAM), CD84, and CD229 (Ly-9). Defects in signaling and aberrant expression of these receptors have been implicated in several immunodeficiency and autoimmune disorders in humans and mice (48). Within the SLAM family, CD244 is unusual in that it shares its ligand CD48 with the receptor CD2 in rodents, whereas in humans CD2 has evolved to interact with CD58 (9). The affinity of CD244 for CD48 in rodents is 6–9-fold higher than the still functionally relevant CD2/CD48 interaction (10). CD244 and CD2 have different cytoplasmic regions comprised of tyrosine motifs or proline-rich motifs, respectively.CD244 is predominantly found on NK cells and cytotoxic T cells and primarily characterized as an activating receptor (1115). CD2 is found on the same cells as CD244 but is also expressed on all T cells, both activated and resting, and has an activating or costimulatory function upon engagement of ligand (9). The tyrosine motifs found in the cytoplasmic tail of CD244 have been shown to bind the SH2 domains of cytoplasmic adaptor proteins SAP and EAT-2 and FYN kinase (1618) and are important to its function (5, 1921). In contrast to SH2 interactions of CD244, several SH3 domain-mediated interactions have been reported for the cytoplasmic region of CD2 including CD2AP/CMS, CIN85, FYN, and LCK (2226).The activating function of CD244 was called into question when a study using cells from a CD244 knock-out mouse showed that CD244 had an inhibitory effect as loss of CD244 resulted in enhanced NK killing of target cells (27). This suggested that previous results in mice where positive effects were seen may have been due to blocking CD244 ligand engagement as opposed to cross-linking with antibodies against CD244 (27). This has led to proposals that there are differences in function between mouse and human CD244 as there is more evidence to suggest that human CD244 is a positive regulator enhancing cytotoxicity and cytokine production (13, 15, 28). However, other more recent studies have shown the mouse CD244/CD48 interaction to be important for cytokine production and effector functions such as cytotoxicity against tumor targets in CD244-deficient mice (29). Long and short forms of CD244 have been cloned from mice with the short form being described as activating and the long form inhibitory (27, 30). Only the long form of CD244 is present in humans and is regarded as activating (14).Positive signaling by CD244 has been attributed to the recruitment of SAP (18), which is a signaling adaptor molecule comprised of a single SH2 domain encoded by the SH2D1A gene and has the ability to recruit the kinase FYN by binding its SH3 domain (31, 32). Loss of the SAP/FYN interaction can lead to X-linked lymphoproliferative disease in humans (17). The molecular basis of in vitro inhibitory effects observed with CD244 in mice on ligation with mAb or ligand remains elusive (33). Protein tyrosine and inositol phosphatases have been reported to associate with CD244 (18, 19, 34) but our studies using surface plasmon resonance found them to be very weak and unlikely to bind competitively compared with the SAP family of adaptors or FYN (16). The SAP-related adaptor EAT-2 has been reported to have an active inhibitory effect that is dependent on tyrosine motifs in the tail of EAT-2 (35) but its mechanism is not understood. The only interaction reported for the tail of EAT-2 is with FYN kinase and studies overexpressing EAT-2 in a T cell hybridoma resulted in increased IL-2 production upon antigen stimulation (16).The conservation between mouse and human CD244 cytoplasmic regions and associated adaptors suggests that both function in a similar way. We have explored the main difference between mouse and human CD244, which is the extracellular interaction through CD48 ligation in the mouse. This has revealed that inhibitory effects of CD244 ligation in mice can be due to competition between CD244 and CD2 for CD48. We have also found that the adaptor protein EAT-2 binds PLCγ1 providing a molecular basis for the important role CD244 plays in regulating cellular cytotoxicity (13, 36). We demonstrate that there is a potentially shared signaling mechanism through the FYN kinase that links CD2 and CD244 intracellularly even though in humans CD2 and CD244 no longer share a cell surface ligand.  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